Acidity of Carboxylic Acid Ligands Influences the Formation of VO2(A) and VO2(B) Nanocrystals under Solvothermal ConditionsClick to copy article linkArticle link copied!
- Brittney A. BeidelmanBrittney A. BeidelmanDepartment of Chemistry, University of Rochester, Rochester, New York 14627, United StatesMore by Brittney A. Beidelman
- Xiaotian ZhangXiaotian ZhangDepartment of Chemistry, University of Rochester, Rochester, New York 14627, United StatesMore by Xiaotian Zhang
- Ellen M. MatsonEllen M. MatsonDepartment of Chemistry, University of Rochester, Rochester, New York 14627, United StatesMore by Ellen M. Matson
- Kathryn E. Knowles*Kathryn E. Knowles*Email: [email protected]Department of Chemistry, University of Rochester, Rochester, New York 14627, United StatesMore by Kathryn E. Knowles
Abstract
Vanadium dioxide (VO2) can adopt many different crystal structures at ambient temperature and pressure, each with different, and often desirable, electronic, optical, and chemical properties. Understanding how to control which crystal phase forms under various reaction conditions is therefore crucial to developing VO2 for various applications. This paper describes the impact of ligand acidity on the formation of VO2 nanocrystals from the solvothermal reaction of vanadyl acetylacetonate (VO(acac)2) with stoichiometric amounts of water. Carboxylic acids examined herein favor the formation of the monoclinic VO2(B) phase over the tetragonal VO2(A) phase as the concentration of water in the reaction increases. However, the threshold concentration of water required to obtain phase-pure VO2(B) nanocrystals increases as the pKa of the carboxylic acid decreases. We also observe that increasing the concentration of VO(acac)2 or the concentration of acid while keeping the concentration of water constant favors the formation of VO2(A). Single-crystal electron diffraction measurements enable the identification of vanadyl carboxylate species formed in reactions that do not contain enough water to promote the formation of VO2. Increasing the length of the carbon chain on aliphatic carboxylic acids did not impact the phase of VO2 nanocrystals obtained but did result in a change from nanorod to nanoplatelet morphology. These results suggest that inhibiting the rate of hydrolysis of the VO(acac)2 precursor either by decreasing the ratio of water to VO(acac)2 or by increasing the fraction of water molecules that are protonated favors the formation of VO2(A) over VO2(B).
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Introduction
Results and Discussion
Conclusions
Experimental Methods
General Considerations
Safety Considerations
Synthesis of VO2 Nanocrystals in the Presence of Different Carboxylic Acids
Synthesis of VO2 Nanocrystals with Various Concentrations of Precursor, Water, and Acid
Nanocrystal Characterization
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnanoscienceau.3c00014.
Tables reporting the amounts of all reagents used in each reaction reported here, powder X-ray diffraction spectra, TEM, and SEM images of nanocrystalline products, and details of the characterization of VO(benzoate)2 and VO(4-nitrobenzoate)2 by single-crystal electron diffraction (PDF)
Crystallographic data for VO(benozate)2 (CIF)
Crystallographic data for VO(4-nitrobenzoate)2 (CIF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
This work was supported financially by a Negative Emissions Sciences Scialog grant funded by the Alfred P. Sloan foundation (G-2021-14157). E.M.M. is the recipient of a Camille Dreyfus Teacher-Scholar Award, which has also supported this work. Dr. Akihito Yamano and Dr. Sho Ito of Rigaku Corp. in Japan performed the electron diffraction experiments on the nanocrystals to obtain crystallographic data for the VO(benzoate)2 and VO(4-nitrobenzoate)2 structures. Dr. Lee Daniels and Dr. Joseph Ferrara of Rigaku Americas Corp. in the USA performed the structural solving analysis for these new structures. We also acknowledge William W. Brennessel from the University of Rochester for his assistance in collecting the powder X-ray diffraction data.
References
This article references 41 other publications.
- 1Wu, C.; Feng, F.; Xie, Y. Design of vanadium oxide structures with controllable electrical properties for energy applications. Chem. Soc. Rev. 2013, 42, 5157– 5183, DOI: 10.1039/c3cs35508jGoogle Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXot1ersr8%253D&md5=ddc30e955f7e900d2efa35cb65c64809Design of vanadium oxide structures with controllable electrical properties for energy applicationsWu, Changzheng; Feng, Feng; Xie, YiChemical Society Reviews (2013), 42 (12), 5157-5183CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The elec. properties of inorg. materials has been a long-standing pursued research topic, and successfully controlling the elec. property of an inorg. material has attracted significant attention for a wide range of energy-related applications, covering energy storage, energy conversion and energy utilization. During the few past decades, vanadium oxides have been studied to gain a clear picture of how microstructural characteristics generating the e-e correlations influence the electronic structure of a material, through which the charge concn., elec. cond. as well as the metal-insulator transition (MIT), etc., can be precisely controlled, giving promising signs for constructing energy-related devices. In this review, we present an extensive review of the engineering of the microstructures of vanadium oxides with control of their elec. properties, and with attempts to rationally construct energy-related devices, such as aq. lithium ion batteries, supercapacitors for energy storage, and thermoelec. generators for energy conversion. Furthermore, the MIT performance of vanadium oxides has also seen tremendous advantages for the applications of "smart windows" and magnetocaloric refrigerators for energy utilization. Collectively, progresses to date suggest that in vanadium oxide systems, the elec. properties, including elec. cond., carrier concns., and the MIT performance, were all strongly dependent on the microstructural characteristics at the at. scale, which have presented extensive promising energy applications covering energy storage, energy conversion and energy utilization.
- 2Li, M.; Magdassi, S.; Gao, Y.; Long, Y. Hydrothermal Synthesis of VO2 Polymorphs: Advantages, Challenges and Prospects for the Application of Energy Efficient Smart Windows. Small 2017, 13, 1701147, DOI: 10.1002/smll.201701147Google ScholarThere is no corresponding record for this reference.
- 3Zhang, Y.; Fan, M.; Liu, X.; Xie, G.; Li, H.; Huang, C. Synthesis of VO2(A) nanobelts by the transformation of VO2(B) under the hydrothermal treatment and its optical switching properties. Solid State Commun. 2012, 152, 253– 256, DOI: 10.1016/j.ssc.2011.11.036Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XpsFOjsA%253D%253D&md5=e99134ae24630224d8694450f8f372fdSynthesis of VO2(A) nanobelts by the transformation of VO2(B) under the hydrothermal treatment and its optical switching propertiesZhang, Yifu; Fan, Meijuan; Liu, Xinghai; Xie, Guangyong; Li, Houbin; Huang, ChiSolid State Communications (2012), 152 (4), 253-256CODEN: SSCOA4; ISSN:0038-1098. (Elsevier Ltd.)VO2(A) nanobelts had been successfully synthesized by the transformation of VO2(B) using H2O as the solvent under the hydrothermal approach at 280 °C for 48 h. Some parameters, such as the reaction temp. and time, had been briefly discussed to reveal the transition from VO2(B) to VO2(A). It was found that H2O played a crucial role in the transition from VO2(B) to VO2(A). The phase transition of VO2(A) nanobelts was at 162 °C. The optical switching properties of VO2(A) were studied by the variable-temp. IR spectra for the first time. In addn., VO2(A) nanobelts were calcined at 700 °C for 2 h under a high purity Ar (99.999%) atm. to obtain VO2(M) which exhibited a strong crystallog. transition at around 65 °C.
- 4Liu, P.; Zhu, K.; Gao, Y.; Wu, Q.; Liu, J.; Qiu, J.; Gu, Q.; Zheng, H. Ultra-long VO2(A) nanorods using the high-temperature mixing method under hydrothermal conditions: synthesis, evolution and thermochromic properties. CrystEngComm 2013, 15, 2753– 2760, DOI: 10.1039/c3ce27085hGoogle Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjslKgs78%253D&md5=24b8f349248f47c43ba3f776251a7173Ultra-long VO2 (A) nanorods using the high-temperature mixing method under hydrothermal conditions: synthesis, evolution and thermochromic propertiesLiu, Pengcheng; Zhu, Kongjun; Gao, Yanfeng; Wu, Qingliu; Liu, Jinsong; Qiu, Jinhao; Gu, Qilin; Zheng, HongjuanCrystEngComm (2013), 15 (14), 2753-2760CODEN: CRECF4; ISSN:1466-8033. (Royal Society of Chemistry)Well-crystd., ultra-long VO2 (A) nanorods were prepd. using a facile high-temp. mixing method (HTMM) under hydrothermal conditions. The as-obtained products were characterized by x-ray diffraction, field-emission SEM, transmission electron microscopy, XPS, UV-Vis-NIR, DSC, and FTIR. The effect of W doping on the phase-transition properties of VO2(A) was also studied. The optimal condition of VO2(A) using the HTMM is hydrothermal treatment at 240° for 48 h with a 3:1 molar ratio of the reducing agent to the vanadium source. The reason why the polymorphic forms of VO2 show different colors is that the light in the visible region reflected by the samples is different. The phase-transition temp. of the pure VO2 (A) is 154.75°, which is significantly lower than the 162 °C reported previously. When a small amt. of W is doped, VO2(A) will be transformed into other polymorphic forms, which indicates that the crystal structure of VO2(A) is highly sensitive to limited doping. Importantly, the as-obtained pure VO2(A) shows good thermochromic properties and optical-switching characters. A crystal growth mechanism for VO2(A), oriented-attachment-exfoliation-recrystn.-oriented-attachment, is proposed and described in detail.
- 5Subba Reddy, C. V.; Walker, E. H., Jr.; Wicker, S. A., Sr.; Williams, Q. L.; Kalluru, R. R. Synthesis of VO2(B) nanorods for Li battery application. Curr. Appl. Phys. 2009, 9, 1195– 1198, DOI: 10.1016/j.cap.2009.01.012Google ScholarThere is no corresponding record for this reference.
- 6Soltane, L.; Sediri, F. Rod-like nanocrystalline B-VO2: Hydrothermal synthesis, characterization and electrochemical properties. Mater. Res. Bull. 2014, 53, 79– 83, DOI: 10.1016/j.materresbull.2014.01.046Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXmsFCmtLk%253D&md5=8b9e591eaf47af46b4d92ff3dd4afc0aRod-like nanocrystalline B-VO2: Hydrothermal synthesis, characterization and electrochemical propertiesSoltane, L.; Sediri, F.Materials Research Bulletin (2014), 53 (), 79-83CODEN: MRBUAC; ISSN:0025-5408. (Elsevier Ltd.)Rod-like nanocryst. VO2(B) has been successfully synthesized via a simple hydrothermal process by using V2O5 as vanadium source and 4-butylaniline H3C-(CH2)3-(C6H4)-NH2 as reducing and structure directing agent. The compds. were analyzed through X-ray diffraction (XRD), SEM (SEM), Fourier transform IR spectroscopy (FTIR), Raman spectroscopy and UV-visible spectroscopy. The VO2(B) nanorods are up to several micrometers in length and about 80 nm in thickness with a large optical band gap of ∼2.709 eV. Thin films of VO2(B) nanorods deposited on ITO substrates were electrochem. characterized by cyclic voltammetry. The voltammograms show reversible redox behavior with charge-discharge cycling process corresponding to the reversible lithium intercalation/deintercalation into the crystal lattice. The av. coulombic efficiency to this redox processes is upper 98% during the electrochem. measurements.
- 7Yin, H.; Yu, K.; Zhang, Z.; Zhu, Z. Morphology-control of VO2(B) nanostructures in hydrothermal synthesis and their field emission properties. Appl. Surf. Sci. 2011, 257, 8840– 8845, DOI: 10.1016/j.apsusc.2011.04.079Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXos1egsr0%253D&md5=8e63026251f96878a79d78084f070814Morphology-control of VO2 (B) nanostructures in hydrothermal synthesis and their field emission propertiesYin, Haihong; Yu, Ke; Zhang, Zhengli; Zhu, ZiqiangApplied Surface Science (2011), 257 (21), 8840-8845CODEN: ASUSEE; ISSN:0169-4332. (Elsevier B.V.)VO2 (B) nanostructures were synthesized via a facile hydrothermal process using V2O5 as source material and oxalic acid as reductant. Three nanostructures of nanorods, nanocarambolas and nanobundles were found existing in the products, and a continuous changing of morphol. was found in the synthesis process, during which the proportion of these three types of nanostructures can be adjusted by altering the concns. of oxalic acid. The microstructures were evaluated using X-ray diffraction, SEM and TEM. Field emission (FE) properties of these three types of nanostructures showed that the nanobundles have the best field emission performance with a turn-on field of ∼1.4 V/μm and a threshold field of ∼5.38 V/μm. These characteristics make VO2 (B) nanostructures a competitive cathode material in field emission devices.
- 8Ganganagappa, N.; Siddaramanna, A. One step synthesis of monoclinic VO2 (B) bundles of nanorods: Cathode for Li ion battery. Mater. Charact. 2012, 68, 58– 62, DOI: 10.1016/j.matchar.2012.03.010Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XntFyksLY%253D&md5=1aa9780a7480b9a28d9680b529a2636eOne step synthesis of monoclinic VO2 (B) bundles of nanorods: Cathode for Li ion batteryGanganagappa, Nagaraju; Siddaramanna, AshokaMaterials Characterization (2012), 68 (), 58-62CODEN: MACHEX; ISSN:1044-5803. (Elsevier)One of the metastable phases of vanadium dioxide, VO2(B) bundles of nanorods and microspheres have been synthesized through a simple hydrothermal method by dispersing V2O5 in aq. quinol. The obtained products were characterized by X-ray diffraction (XRD), Fourier transform IR (FTIR) spectroscopy, SEM (SEM) and electrochem. discharge-charge test for lithium battery. It was found that the morphologies of the obtained VO2(B) can be tuned by manipulating the relative amt. of quinol. The electrochem. test found that the bundles of nanorods exhibit an initial discharge capacity of 171 mAh g- 1 and its almost stabilized capacity was reached to 108 mAh g- 1 after 47 cycles at a c.d. of 0.1 mA g- 1. The formation mechanism of the VO2(B) bundles of nanorods and microspheres was also discussed.
- 9Zhang, Y. VO2(B) conversion to VO2(A) and VO2(M) and their oxidation resistance and optical switching properties. Mater. Sci.-Pol. 2016, 34, 169– 176, DOI: 10.1515/msp-2016-0023Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XnvFWgtLg%253D&md5=824fffaabc1ee70900b734e7cb8ed72dVO2(B) conversion to VO2(A) and VO2(M) and their oxidation resistance and optical switching propertiesZhang, YifuMaterials Science-Poland (2016), 34 (1), 169-176CODEN: MSACDV; ISSN:2083-134X. (Walter de Gruyter GmbH)Vanadium dioxide VO2 has been paid in recent years increasing attention because of its various applications, however, its oxidn. resistance properties in air atm. have rarely been reported. Herein, VO2(B) nanobelts were transformed into VO2(A) and VO2(M) nanobelts by hydrothermal route and calcination treatment, resp. Then, we comparatively studied the oxidn. resistance properties of VO2(B), VO2(A) and VO2(M) nanobelts in air atm. by thermo-gravimetric anal. and DTA (TGA/DTA). It was found that the nanobelts had good thermal stability and oxidn. resistance below 341°C, 408°C and 465°C in air, resp., indicating that they were stable in air at room temp. The fierce oxidn. of the nanobelts occurred at 426, 507 and 645°C, resp. The results showed that the VO2(M) nanobelts had the best thermal stability and oxidn. resistance among the others. Furthermore, the phase transition temps. and optical switching properties of VO2(A) and VO2(M) were studied by differential scanning calorimetry (DSC) and variable temp. IR spectra. It was found that the VO2(A) and VO2(M) nanobelts had outstanding thermochromic character and optical switching properties.
- 10Sediri, F.; Gharbi, N. Controlled hydrothermal synthesis of VO2(B) nanobelts. Mater. Lett. 2009, 63, 15– 18, DOI: 10.1016/j.matlet.2008.08.022Google ScholarThere is no corresponding record for this reference.
- 11Wang, Q.; Xu, J.; Zhang, W.; Mao, M.; Wei, Z.; Wang, L.; Cui, C.; Zhu, Y.; Ma, J. Research progress on vanadium-based cathode materials for sodium ion batteries. J. Mater. Chem. A 2018, 6, 8815– 8838, DOI: 10.1039/c8ta01627eGoogle Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXntlCnsrs%253D&md5=32772400483c5ac72c28bae062b824d9Research progress on vanadium-based cathode materials for sodium ion batteriesWang, Qinghong; Xu, Jiantie; Zhang, Wenchao; Mao, Minglei; Wei, Zengxi; Wang, Lei; Cui, Chunyu; Zhu, Yuxuan; Ma, JianminJournal of Materials Chemistry A: Materials for Energy and Sustainability (2018), 6 (19), 8815-8838CODEN: JMCAET; ISSN:2050-7496. (Royal Society of Chemistry)A review is presented. Sodium ion batteries (SIBs) have attracted increasing attention as one of the most promising candidates for cost-effective, high-energy rechargeable batteries. Owing to their high theor. capacity and energy d., and rich electrochem. interaction with Na+ (V2+-V5+), a large no. of vanadium(V)-based cathode materials, including vanadium oxides (e.g., V2O5 and VO2), vanadium bronzes (e.g., NaxVO2, NaV3O8, NaV6O15 and δ-NH4V4O10), V-based phosphates (e.g., Na3V2(PO4)3, VOPO4, NaVOPO4, Na7V3(P2O7)4 and Na2(VO)P2O7) and F-contg. V-based polyanions (e.g., NaVPO4F, Na3V2(PO4)2F3 and Na3(VOx)2(PO4)2F3-2x), have been explored for SIBs. In this review, we mainly summarize the basic structures, modified/optimized structures, synthetic methods and morphol. control of V-based cathode materials for SIBs. Addnl., major drawbacks, emerging challenges and some perspectives on the development of V-based cathode materials for SIBs are also discussed.
- 12Zhang, L.; Yao, J.; Guo, Y.; Xia, F.; Cui, Y.; Liu, B.; Gao, Y. VO2(A) Nanorods: One-pot Synthesis, Formation Mechanism and Thermal Transformation to VO2(M). Ceram. Int. 2018, 44, 19301– 19306, DOI: 10.1016/j.ceramint.2018.07.157Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhsVartb3L&md5=b980416b5b434be3ddf18b028717cc1cVO2(A) nanorods: One-pot synthesis, formation mechanism and thermal transformation to VO2(M)Zhang, Liangmiao; Yao, Jianing; Guo, Yunfeng; Xia, Fang; Cui, Yuanyuan; Liu, Bin; Gao, YanfengCeramics International (2018), 44 (16), 19301-19306CODEN: CINNDH; ISSN:0272-8842. (Elsevier Ltd.)The monoclinic VO2(M) has promising applications in intelligent devices but its prepn. still requires improvement to permit cost-effective mass prodn. In this work, we report a 2-stage approach for producing VO2(M) nanorods by (1) hydrothermal redn. of vanadium pentoxide by sodium bisulfate at 220°C to form VO2(A), and (2) subsequent thermal activated phase transformation of VO2(A) to VO2(M) at 350-450°C in vacuum. The obtained VO2(M) nanorods showed a reversible phase transition temp. at about 62.5°C and a narrow thermal hysteresis width of 10°C. The mechanism of the hydrothermal redn. was studied by combined ex situ microscopic and diffraction characterization of cooled samples as well as in situ PXRD expts., in which the hydrothermal synthesis was monitored in real time by time-resolved diffraction datasets. It was found that the hydrothermal synthesis of VO2(A) is a 4-step process: (1) redn. of V2O5 to form VO2(B) nanoparticles, (2) oriented attachment of VO2(B) nanoparticles along the [110] direction, (3) formation of VO2(B) nanorods as a results of oriented attachments, and (4) hydrothermal transformation of the metastable intermediate VO2(B) nanorods to VO2(A) nanorods. This clear understanding of the mechanism will help the further optimization of synthesis temp. and time for prepg. VO2(A). This method provides a low temp. thermal treatment alternative and hence helps the redn. of cost for the prodn. of VO2(M), bring the mass application of VO2(M) one step closer.
- 13Popuri, S.; Artemenko, A.; Labrugere, C.; Miclau, M.; Villesuzanne, A.; Pollet, M. VO2(A) Reinvestigation of crystal structure, phase transition and crystal growth mechanisms. J. Solid State Chem. 2014, 213, 79– 86, DOI: 10.1016/j.jssc.2014.01.037Google ScholarThere is no corresponding record for this reference.
- 14Ji, S.; Zhang, F.; Jin, P. Selective formation of VO2(A) or VO2(R) polymorph by controlling the hydrothermal pressure. J. Solid State Chem. 2011, 184, 2285– 2292, DOI: 10.1016/j.jssc.2011.06.029Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXpt1Cms7Y%253D&md5=7be35329c5ffd268cc056cfdda465286Selective formation of VO2(A) or VO2(R) polymorph by controlling the hydrothermal pressureJi, Shidong; Zhang, Feng; Jin, PingJournal of Solid State Chemistry (2011), 184 (8), 2285-2292CODEN: JSSCBI; ISSN:0022-4596. (Elsevier B.V.)VO2(A) usually occurs during the prepn. of VO2 polymorphs. This leads to an ambiguous understanding of the transformation between VO2 polymorphs. The calcn. of the ground state energies for different VO2 polymorphs indicated that there is only a small energy gap between VO2(A) and VO2(R), which suggests that the transformation from VO2(A) to VO2(R) should be pressure sensitive. This hypothesis was verified during the synthesizing of VO2 polymorphs by reducing V2O5 with oxalic acid through hydrothermal treatment process. Selective formation of pure phase VO2(A) or VO2(R) was achieved by controlling the hydrothermal pressure through varying the filling ratio at 270°. It was found that a filling ratio >0.5 favors the formation of pure VO2(R) while a filling ratio ≤0.4 results in the formation of VO2(A). Based on our expts., VO2(B) nanobelts always formed first and then transformed to VO2(A) by an assembling process at increased temp. or extended reaction time. Under further higher pressure, the VO2(A) transformed spontaneously to VO2(R) initialized from the vol. shrinkage due to the formation of denser VO2(R).
- 15Cheng, X. H.; Xu, H. F.; Wang, Z. Z.; Zhu, K. R.; Li, G.; Jin, S. Synthesis, characterization and formation mechanism of metastable phase VO2(A) nanorods. Mater. Res. Bull. 2013, 48, 3383– 3388, DOI: 10.1016/j.materresbull.2013.05.016Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXovVCms7c%253D&md5=a5f9bc3aa90d57928080e541829afa04Synthesis, characterization and formation mechanism of metastable phase VO2(A) nanorodsCheng, X. H.; Xu, H. F.; Wang, Z. Z.; Zhu, K. R.; Li, G.; Jin, ShaoweiMaterials Research Bulletin (2013), 48 (9), 3383-3388CODEN: MRBUAC; ISSN:0025-5408. (Elsevier Ltd.)Pure phase VO2(A) nanorods were synthesized via the redn. of V2O5 by oxalic acid during the hydrothermal treatment. Two sets of samples were prepd. by varying both system temp. and reaction time under a filling ratio of 0.40 for observing the formation and evolution of VO2(A) nanorods. Structures were characterized by X-ray diffraction, scanning and transmission electron microscopies, resp. It was found that VO2(B) was firstly formed and then transformed into VO2(A) as the increasing system temp. or extending reaction time. An assembling and following crystal adjustment was proposed for explanation the formation process of VO2(A) from VO2(B). For VO2(A) nanorods, the phase transition temp. of 169.7°C was higher than that of the VO2(A) bulk, it might be ascribed to the lower crystallinity or nonstoichiometry in VO2(A) nanorods. VO2 nanostructures with controllable phases and properties should find their promising applications in a single VO2 nanodevice.
- 16Zhang, L.; Xia, F.; Song, Z.; Webster, N. A. S.; Luo, H.; Gao, Y. Synthesis and formation mechanism of VO2(A) nanoplates with intrinsic peroxidase-like activity. RSC Adv. 2015, 5, 61371– 61379, DOI: 10.1039/c5ra11014aGoogle Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFChtLrL&md5=f3ad20d4327c286dc3fbafce611c073aSynthesis and formation mechanism of VO2(A) nanoplates with intrinsic peroxidase-like activityZhang, Liangmiao; Xia, Fang; Song, Zhengdong; Webster, Nathan A. S.; Luo, Hongjie; Gao, YanfengRSC Advances (2015), 5 (75), 61371-61379CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Monocryst. VO2(A) nanoplates were synthesized via a 1-pot hydrothermal process. In situ powder x-ray diffraction was used to monitor the hydrothermal synthesis and VO2(A) nucleates and grows directly from soln. after the complete hydrolysis of a 2.0M VO(acac)2 precursor soln., rather than involving a previously reported intermediate phase VO2(B). A hydrating-exfoliating-splitting mechanism was established to explain the formation of the nanoplate architecture. The synthesized VO2(A) nanoplates showed outstanding peroxidase-like activity and hence are a promising candidate for artificial peroxidase.
- 17Zhang, S.; Shang, B.; Yang, J.; Yan, W.; Wei, S.; Xie, Y. From VO2(B) to VO2(A) nanobelts: first hydrothermal transformation, spectroscopic study and first principles calculation. Phys. Chem. Chem. Phys. 2011, 13, 15873– 15881, DOI: 10.1039/c1cp20838aGoogle Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtVGqu7bN&md5=39cdf6a5903c9c1ccf91631e97f28cb5From VO2 (B) to VO2 (A) nanobelts: first hydrothermal transformation, spectroscopic study and first principles calculationZhang, Shudong; Shang, Bo; Yang, Jinlong; Yan, Wensheng; Wei, Shiqiang; Xie, YiPhysical Chemistry Chemical Physics (2011), 13 (35), 15873-15881CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The phase transition process from VO2 (B) to VO2 (A) was first obsd. through a mild hydrothermal approach, using hybrid d. functional theory (DFT) calcns. and crystallog. VO2 topol. anal. All theor. analyses reveal that VO2 (A) is a thermodynamically stable phase and has a lower formation energy compared with the metastable VO2 (B). For the first time, X-ray absorption spectroscopy (XAS) of the V L-edge and O K-edge was performed on different VO2 phases, and the differences in the electronic structure of the two polymorphic forms provide further exptl. evidence of the more stable VO2 (A). Consequently, transformation from VO2 (B) to VO2 (A) is much easier to be realized from a dynamical point of view. Notably, the transformation of VO2 (B) into VO2 (A) show the sequence VO2 (B)-high-temp. VO2 (AH) phase-low-temp. VO2 (A) phase, which was achieved by hydrothermal treatment, resp. Also, an alternative synthesis route was proposed based on the above hydrothermal transformation, and VO2 (A) was successfully prepd. via the simple one-step hydrothermal method by hydrolysis of VO(acac)2 (acac = acetylacetonate). Therefore, VO2 nanostructures with controlled phase compns. can be obtained in high yields. Through elucidating the structural evolution in the crystallog. shear mechanism, we can easily guide the design of other metal oxide nanostructures with controllable phases.
- 18Zhang, S.; Li, Y.; Wu, C.; Zheng, F.; Xie, Y. Novel Flowerlike Metastable Vanadium Dioxide (B) Micronanostructures: Facile Synthesis and Application in Aqueous Lithium Ion Batteries. J. Phys. Chem. C 2009, 113, 15058– 15067, DOI: 10.1021/jp903312hGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXovFKlt74%253D&md5=853e81161aac174de5b76d5e68fc5045Novel Flowerlike Metastable Vanadium Dioxide (B) Micronanostructures: Facile Synthesis and Application in Aqueous Lithium Ion BatteriesZhang, Shudong; Li, Yingmei; Wu, Changzheng; Zheng, Fei; Xie, YiJournal of Physical Chemistry C (2009), 113 (33), 15058-15067CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Novel flowerlike VO2(B) micro-nanostructures assembled by single-cryst. nanosheets were synthesized via a hydrothermal route using polymer polyvinyl pyrrolidone (PVP, K30) as capping reagent. Detailed proofs indicated that the process of crystal growth was dominated by a nucleation and growth, self-assembly, and then Ostwald ripening growth mechanism. The flowerlike micro-nanostructured VO2(B) was applied as the active material in aq. Li-ion batteries which showed improved electrochem. properties with the 1st discharge capacity reaching 74.9 mA-h/g, a good value for aq. Li ion battery systems in light of previous reports (usually <65 mA-h/g). Corresponding VO2(B) nanostructures with better crystallinity were obtained by calcining the precursor of flowerlike VO2(B) structures. The post-treated flowerlike VO2(B) electrode shows better electrochem. properties with the 1st discharge capacity reaching 81.3 mA-h/g, which is higher than that of the flowerlike VO2(B) sample before annealing. The electrochem. intercalation and deintercalation properties of VO2(B) nanobelts and carambola-like VO2(B) structures with Li+ were also studied. The unique flowerlike structure plays a role in the morphol. requirement to serve as transport paths for Li ions in aq. Li ion batteries. The morphol. and crystallinity of the synthesized products had an influence on the electrochem. intercalation and deintercalation properties with Li ions.
- 19Walton, R. I. Subcritical solvothermal synthesis of condensed inorganic materials. Chem. Soc. Rev. 2002, 31, 230– 238, DOI: 10.1039/b105762fGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XkvVCmt7o%253D&md5=ac8805ccaff49f722b4c472ee85726daSubcritical solvothermal synthesis of condensed inorganic materialsWalton, Richard I.Chemical Society Reviews (2002), 31 (4), 230-238CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The solvothermal method has recently been extended from zeolite synthesis to the formation of condensed inorg. solids, which find uses in diverse areas due to properties such as ionic-cond., solid-state magnetism, giant magnetoresistance, low thermal expansion and ferroelectricity. This offers specific advantages over the traditional ceramic synthetic routes to inorg. solids and these are highlighted with examples from the recent literature, and the efforts focussed on detg. the formation mechanism of solids from the heterogeneous mixts. used in solvothermal procedures are discussed.
- 20Yu, W.; Li, S.; Huang, C. Phase evolution and crystal growth of VO2 nanostructures under hydrothermal reactions. RSC Adv. 2016, 6, 7113– 7120, DOI: 10.1039/c5ra23898fGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XlvV2lsQ%253D%253D&md5=38c86bd180df72c567bacbc3a3b8057dPhase evolution and crystal growth of VO2 nanostructures under hydrothermal reactionsYu, Weilai; Li, Shuai; Huang, ChiRSC Advances (2016), 6 (9), 7113-7120CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)The phase evolution and crystal growth of VO2 nanostructures against reaction time in a high-pressure V2O5-oxalic acid hydrothermal system were systematically investigated. It was found that the rather thin VO2 (B) nanobelts were first obtained, then stacked to form large belt-like structures and subsequently phase transformed into VO2 (A), based on an oriented attachment-recrystn. mechanism. The large VO2 (A) belt-like structures could further assemble into novel "snowflake" VO2 (M) microcrystals with even bigger sizes and nearly well-defined six-fold symmetry. Due to the Ostwald ripening effect regarding crystal size discrepancy, the VO2 (M) phase could further grow at the cost of the gradual dissoln. of VO2 (A) and full elimination of VO2 (B). The phase evolution from VO2 (B) first to VO2 (A) and then to VO2 (M), is actually a step-by-step thermodynamically downhill process, owing to the gradual relaxation of structural tension within the VO2 crystal lattice. Thus, our investigation, for the first time, demonstrated the feasibility of the well-known Ostwald's step rules towards the phase evolution process of VO2 and could provide unprecedented new insight to promote understanding of the synthesis and properties of vanadium oxide compds.
- 21Beidelman, B. A.; Zhang, X.; Sanchez-Lievanos, K. R.; Selino, A. V.; Matson, E. M.; Knowles, K. E. Influence of water concentration on the solvothermal synthesis of VO2(B) nanocrystals. CrystEngComm 2022, 24, 6009– 6017, DOI: 10.1039/d2ce00813kGoogle Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XitV2ks73O&md5=9d7d4808dd7ab07a57b23fb3525febaaInfluence of water concentration on the solvothermal synthesis of VO2(B) nanocrystalsBeidelman, Brittney A.; Zhang, Xiaotian; Sanchez-Lievanos, Karla R.; Selino, Annabel V.; Matson, Ellen M.; Knowles, Kathryn E.CrystEngComm (2022), 24 (34), 6009-6017CODEN: CRECF4; ISSN:1466-8033. (Royal Society of Chemistry)Nanocrystals of VO2(B) have attracted significant attention for their promising performance in electrochem. energy storage applications. Phase-purity and nanocrystal morphol. are both crucial factors for the performance of these materials, but given the large no. of crystal polymorphs available to VO2, achieving simultaneous control over crystal phase and nanocrystal size is a significant synthetic challenge. This paper describes the impact of water concn. on the synthesis of VO2(B) nanocrystals via solvothermal reaction of a mol. V(IV) precursor in toluene. Using controlled, stoichiometric amts. of water (20 equiv per vanadium center) enables access to short nanorods of VO2(B) whose lengths follow a Gaussian distribution with an av. and std. deviation of 110 ± 30 nm. Decreasing the amt. of water present to two or four equiv. results in formation of VO2(A) nanocrystals and eliminating it entirely results in no reaction. Increasing the amt. of water to more than 20 equiv increases the av. length of the VO2(B) nanorods and causes the distribution in rod lengths to evolve from Gaussian to lognormal. This evolution in the size distribution is consistent with changes in a model Gaussian distribution obsd. upon simulated end-to-end oriented attachment events. These results demonstrate that control over the concn. of water is a useful strategy for tuning the morphol. and crystal phase of VO2 nanocrystals.
- 22Serjeant, E. P.; Dempsey, B. Ionisation Constants of Organic Acids in Aqueous Solution; Pergamon Press: New York, New York, 1979; Vol. 23.Google ScholarThere is no corresponding record for this reference.
- 23Haynes, W. H. CRC Handbook of Chemistry and Physics, 91st ed.; CRC Press Inc.: Boca Raton, FL, 2010–2011.Google ScholarThere is no corresponding record for this reference.
- 24Dean, J. A. Lange’s Handbook of Chemistry, 13th ed.; McGraw-Hill Book Co: New York, NY, 1985.Google ScholarThere is no corresponding record for this reference.
- 25Muckerman, J. T.; Skone, J. H.; Ning, M.; Wasada-Tsutsui, Y. Toward the accurate calculation of pKa values in water and acetonitrile. Biochim. Biophys. Acta, Bioenerg. 2013, 1827, 882– 891, DOI: 10.1016/j.bbabio.2013.03.011Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXmsFCjtbk%253D&md5=c6e56760baf2f86e8940fcd72bb08af1Toward the accurate calculation of pKa values in water and acetonitrileMuckerman, James T.; Skone, Jonathan H.; Ning, Ming; Wasada-Tsutsui, YukoBiochimica et Biophysica Acta, Bioenergetics (2013), 1827 (8-9), 882-891CODEN: BBBEB4; ISSN:0005-2728. (Elsevier B. V.)We present a simple approach for the calcn. of accurate pKa values in water and acetonitrile based on the straightforward calcn. of the gas-phase abs. free energies of the acid and conjugate base with use of only a continuum solvation model to obtain the corresponding soln.-phase free energies. Most of the error in such an approach arises from inaccurate differential solvation free energies of the acid and conjugate base which is removed in our approach using a correction based on the realization that the gas-phase acidities have only a small systematic error relative to the dominant systematic error in the differential solvation. The methodol. is outlined in the context of the calcn. of a set of neutral acids with water as the solvent for a reasonably accurate electronic structure level of theory (DFT), basis set, and implicit solvation model. It is then applied to the comparison of results for three different hybrid d. functionals to illustrate the insensitivity to the functional. Finally, the approach is applied to the comparison of results for sets of neutral acids and protonated amine cationic acids in both aq. (water) and nonaq. (acetonitrile) solvents. The methodol. is shown to generally predict the pKa values for all the cases investigated to within 1 pH unit so long as the differential solvation error is larger than the systematic error in the gas-phase acidity calcns. Such an approach is rather general and does not have addnl. complications that would arise in a cluster-continuum method, thus giving it strength as a simple high-throughput means to calc. abs. pKa values. This article is part of a Special Issue entitled: Metals in Bioenergetics and Biomimetics Systems.
- 26Weeks, C.; Song, Y.; Suzuki, M.; Chernova, N. A.; Zavalij, P. Y.; Whittingham, M. S. The one dimensional chain structures of vanadyl glycolate and vanadyl acetate. J. Mater. Chem. 2003, 13, 1420– 1423, DOI: 10.1039/b208100hGoogle Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXjvVCgsLc%253D&md5=b570fd9cea9efed13b78c01cc79cd069The one dimensional chain structures of vanadyl glycolate and vanadyl acetateWeeks, Curtis; Song, Yanning; Suzuki, Masatsugu; Chernova, Natasha A.; Zavalij, Peter Y.; Whittingham, M. StanleyJournal of Materials Chemistry (2003), 13 (6), 1420-1423CODEN: JMACEP; ISSN:0959-9428. (Royal Society of Chemistry)The solvothermal reaction, at 200°, of vanadium pentoxide and lithium hydroxide in acetic acid or ethylene glycol gives vanadyl acetate and vanadyl glycolate resp. The structure of the acetate contains vanadium in octahedral coordination whereas the glycolate contains VO5 square pyramids. The VO6 octahedra in the acetate, VO(CH3COO)2, are joined through the vanadyl groups, giving a rather long V:O bond of 1.684(7) Å and a short trans V-O bond of 2.131(7) Å, and by bridging acetate groups. The vanadium atoms interact along the ···V:O···V:O··· chain giving 1-dimensional antiferromagnetic behavior. In contrast in the glycolate, the apical V:O bond is shorter, 1.58(1) Å, and the square pyramids share edges in a two up-two down fashion to give chains VO(OCH2CH2O). Magnetic susceptibility of vanadyl glycolate is consistent with an isolated spin dimers model.
- 27Casey, A. T.; Thackeray, J. R. The preparation and magnetic properties of oxovanadium(IV) acetate. Aust. J. Chem. 1969, 22, 2549– 2553, DOI: 10.1071/ch9692549Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXjslWmsg%253D%253D&md5=6c034e42b8006146026c3a6ed3ef4aa6Preparation and magnetic properties of oxovandium(IV) acetateCasey, Allan T.; Thackery, J. R.Australian Journal of Chemistry (1969), 22 (12), 2549-53CODEN: AJCHAS; ISSN:0004-9425.The variation of the magnetic susceptibility of oxovanadium(IV) acetate with temp. may be fitted to the Ising model of an infinite linear antifer romagnetic chain with exchange energy of -166 cm-1. On the basis of this and ir evidence, a structure for the compd. is proposed.
- 28Dakternieks, D. R.; Harris, C. M.; Milham, P. J.; Morris, B. S.; Sinn, E. Magnetism and structure of polymeric oxovanadium(IV) acetate and other polymeric oxovanadium(IV) complexes. Inorg. Nucl. Chem. Lett. 1969, 5, 97– 100, DOI: 10.1016/0020-1650(69)80177-7Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXhtF2mt7s%253D&md5=a63282881ea3c3e5c17a11c0144f2955Magnetism and structure of polymeric oxovanadium(IV) acetate and other polymeric oxovanadium(VI) complexesDakternieks, D. R.; Harris, Clive Melville; Milham, P. J.; Morris, Benjamin S.; Sinn, EkkehardInorganic and Nuclear Chemistry Letters (1969), 5 (2), 97-100CODEN: INUCAF; ISSN:0020-1650.From magnetic and other evidence, it is concluded that oxovanadium(IV) acetate and other carboxylate complexes consist of polymeric structures. The V:O stretching absorption at 895 cm.-1 gives a force const. k of 5.8 millidynes/A.
- 29Ikekwere, P. O.; Adeniyi, A. A. Oxovanadium(IV) Complexes of Some Aromatic Carboxylic Acids. Synth. React. Inorg. Met.-Org. Chem. 1989, 19, 87– 100, DOI: 10.1080/00945718908048053Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXktFKltbs%253D&md5=ab83fb438d173fba7cf52badac767214Oxovanadium(IV) complexes of some aromatic carboxylic acidsIkekwere, P. O.; Adeniyi, A. A.Synthesis and Reactivity in Inorganic and Metal-Organic Chemistry (1989), 19 (1), 87-100CODEN: SRIMCN; ISSN:0094-5714.VOL2.nH2O (L = BzOH and its 2-, 3- and 4-cyano-, 3-cyano-4-methyl-, 2-, 3- and 4-methyl-, 2,5- and 3,5-dimethyl-, 2,4,6-trimethyl-, 2-, 3- and 4-nitro derivs.) were prepd. and characterized by their magnetic and spectral properties. The complexes had magnetic moments much lower than the spin-only value of 1.73 μB at room temp. indicating strong magnetic interactions. Their Nujol mull spectra showed 3 main electronic absorption bands typical of tetragonally distorted octahedral oxovanadium(IV) compds. The IR vibrational spectra indicate that the carboxylate groups are bridging. The complexes are carboxylate-bridged polymers with V:O...V:O interaction.
- 30Puri, M.; Sharma, R. D.; Verma, R. D. Trifluoroacetates of Vanadium(III) and Oxovanadium(IV) and (V) Preparation and Characterization. Synth. React. Inorg. Met.-Org. Chem. 1981, 11, 539– 546, DOI: 10.1080/00945718108055998Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38Xht1aitLs%253D&md5=dfdabce00d2634d5c7519bd8d3f3a35dTrifluoroacetates of vanadium(III) and oxovanadium(IV) and (V). Preparation and characterizationPuri, Madhu; Sharma, R. D.; Verma, Rajendar D.Synthesis and Reactivity in Inorganic and Metal-Organic Chemistry (1981), 11 (6), 539-46CODEN: SRIMCN; ISSN:0094-5714.V(O2CCF3)3 (I), VO(O2CCF3)2 (II), and VO2(O2CCF3) were prepd. by treating VCl3, VOCl2, and VOCl3, resp., with CF3CO2H. I and II were also prepd. by heating V(OAc)3 and VO(OAc)2 with CF3CO2H. The IR spectra, thermal degrdn., magnetic susceptibilities, and reflectance spectra of the compds. were detd.
- 31Lorenzotti, A.; Leonesi, D.; Cingolani, A.; Di Bernardo, P. Vanadyl(IV)─Acetate complexes in aqueous solution. J. Inorg. Nucl. Chem. 1981, 43, 737– 738, DOI: 10.1016/0022-1902(81)80213-8Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXktlWht78%253D&md5=d53355c3f23bd1243e57ac91ab797f5bVanadyl(IV)-acetate complexes in aqueous solutionLorenzotti, Adriana; Leonesi, Dante; Cingolani, Augusto; Di Bernardo, PlinioJournal of Inorganic and Nuclear Chemistry (1981), 43 (4), 737-8CODEN: JINCAO; ISSN:0022-1902.The formation of vanadyl(IV) complexes with acetates were studied by potentiometric titrn. methods; the stability consts. of mononuclear complexes, [VO(CH3COO)]+ and [VO(CH3COO)2], were calcd. to be 72.2 and 12.6 M-1, resp., in 1M aq. perchlorate at 25°. A comparison was made with the literature values for the corresponding neptunyl(IV) complexes.
- 32Di Bernardo, P.; Tomat, G.; Zanonato, P.; Portanova, R.; Tolazzi, M. Thermodynamics of vanadyl(IV)─carboxylate complex formation in aqueous solution. Inorg. Chim. Acta 1988, 145, 285– 288, DOI: 10.1016/s0020-1693(00)83971-7Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXltFOjur8%253D&md5=f08fc3a701e03e24092175d6e4861700Thermodynamics of vanadyl(IV)-carboxylate complex formation in aqueous solutionDi Bernardo, Plinio; Tomat, Giuliana; Zanonato, Pierluigi; Portanova, Roberto; Tolazzi, MarilenaInorganica Chimica Acta (1988), 145 (2), 285-8CODEN: ICHAA3; ISSN:0020-1693.The stability consts. and the heats of formation of vanadyl(IV)-acetate, -glycolate, and -glycine complexes were detd. in aq. soln. by potentiometric and calorimetric measurements. In the pH range where the protolytic equil. of VO2+ is negligible, acetate forms 2 mononuclear complexes, glycolate 3, while glycine reacts in its zwitterionic form. The stabilities of the glycolate complexes are considerably higher than the acetate complexes, in spite of its lower basicity, indicating that complex formation involves the coordination of the OH group. The enthalpy changes are pos. except for glycolate where a small neg. value is found. For all systems, the entropy changes are pos. and therefore favorable to the complex formation.
- 33Mehio, N.; Ivanov, A. S.; Ladshaw, A. P.; Dai, S.; Bryantsev, V. S. Theoretical Study of Oxovanadium(IV) Complexation with Formamidoximate: Implications for the Design of Uranyl-Selective Adsorbents. Ind. Eng. Chem. Res. 2016, 55, 4231– 4240, DOI: 10.1021/acs.iecr.5b03398Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvVyit7jL&md5=ffaf2049955182832f1549baa63c4f99Theoretical Study of Oxovanadium(IV) Complexation with Formamidoximate: Implications for the Design of Uranyl-Selective AdsorbentsMehio, Nada; Ivanov, Alexander S.; Ladshaw, Austin P.; Dai, Sheng; Bryantsev, Vyacheslav S.Industrial & Engineering Chemistry Research (2016), 55 (15), 4231-4240CODEN: IECRED; ISSN:0888-5885. (American Chemical Society)Poly(acrylamidoxime) fibers are the current state-of-the-art adsorbent for mining uranium from seawater. However, the competition between uranyl (UO22+) and vanadium ions poses a challenge to mining on the industrial scale. In this work, we employ d. functional theory and coupled-cluster methods in the restricted formalism to investigate potential binding motifs of the oxovanadium(IV) ion (VO2+) with the formamidoximate ligand. Consistent with exptl. extended X-ray absorption fine structure data, the hydrated six-coordinate complex is predicted to be preferred over the hydrated five-coordinate complex. Our investigation of formamidoximate-VO2+ complexes universally identified the most stable binding motif formed by chelating a tautomerically rearranged imino hydroxylamine via the imino nitrogen and hydroxylamine oxygen. The alternative binding motifs for amidoxime chelation via a nonrearranged tautomer and η2 coordination are found to be ∼11 kcal/mol less stable. Natural bond orbital anal. was performed to understand the nature of the interactions in the VO2+ complexes. The difference in the most stable VO2+ and UO22+ binding conformation has important implications for the design of more selective UO22+ ligands.
- 34Winkler, J. R.; Gray, H. B. Electronic Structures of Oxo-Metal Ions. In Molecular Electronic Structures of Transition Metal Complexes I. Structure and Bonding; Mingos, D. M. P., Day, P., Dahl, J. P., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2012; vol. 142, pp 17– 28.Google ScholarThere is no corresponding record for this reference.
- 35Krakowiak, J.; Lundberg, D.; Persson, I. A Coordination Chemistry Study of Hydrated and Solvated Cationic Vanadium Ions in Oxidation States +III, +IV, and +V in Solution and Solid State. Inorg. Chem. 2012, 51, 9598– 9609, DOI: 10.1021/ic300202fGoogle Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtlSks7nL&md5=e69f2a1d2daac25be6f4a51cedebef0bA Coordination Chemistry Study of Hydrated and Solvated Cationic Vanadium Ions in Oxidation States +III, +IV, and +V in Solution and Solid StateKrakowiak, Joanna; Lundberg, Daniel; Persson, IngmarInorganic Chemistry (2012), 51 (18), 9598-9609CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The coordination chem. of hydrated and solvated vanadium(III), oxovanadium(IV), and dioxovanadium(V) ions in the oxygen-donor solvents water, DMSO , and N,N'-dimethylpropyleneurea (DMPU) was studied in soln. by extended x-ray absorption fine structure (EXAFS) and large-angle X-ray scattering (LAXS) and in the solid state by single-crystal X-ray diffraction and EXAFS. The hydrated vanadium(III) ion has a regular octahedral configuration with a mean V-O bond distance of 1.99 Å. In the hydrated and DMSO-solvated oxovanadium(IV) ions, vanadium binds strongly to an oxo group at ∼1.6 Å. The solvent mol. trans to the oxo group is very weakly bound, at ∼2.2 Å, while the remaining four solvent mols., with a mean V-O bond distance of 2.0 Å, form a plane slightly below the vanadium atom; the mean O=V-Operp bond angle is ∼98°. In the DMPU-solvated oxovanadium(IV) ion, the space-demanding properties of the DMPU mol. leave no solvent mol. in the trans position to the oxo group, which reduces the coordination no. to 5. The O=V-O bond angle is consequently much larger, 107°, and the mean V=O and V-O bond distances decrease to 1.58 and 1.97 Å, resp. The hydrated and DMSO-solvated dioxovanadium(V) ions display a very distorted octahedral configuration with the oxo groups in the cis position with a mean V=O bond distance of 1.6 Å and a O=V=O bond angle of ∼105°. The solvent mols. trans to the oxo groups are weakly bound, at ∼2.2 Å, while the remaining two have bond distances of 2.02 Å. The exptl. studies of the coordination chem. of hydrated and solvated vanadium(III,IV,V) ions are complemented by summarizing previously reported crystal structures to yield a comprehensive description of the coordination chem. of vanadium with oxygen-donor ligands.
- 36Rumble, J. R. CRC Handbook of Chemistry and Physics, Internet Edition 2018, 99th ed.; CRC Press/Taylor & Francis: Boca Raton, FL, 2018.Google ScholarThere is no corresponding record for this reference.
- 37Guo, Y.; Surblys, D.; Matsubara, H.; Kawagoe, Y.; Ohara, T. Molecular Dynamics Study on the Effect of Long-Chain Surfactant Adsorption on Interfacial Heat Transfer between a Polymer Liquid and Silica Surface. J. Phys. Chem. C 2020, 124, 27558– 27570, DOI: 10.1021/acs.jpcc.0c08940Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXisV2hu73K&md5=b84d8894327b9cc141f818c12772aef2Molecular dynamics study on effect of long-chain surfactant adsorption on interfacial heat transfer between polymer liquid and silica surfaceGuo, Yuting; Surblys, Donatas; Matsubara, Hiroki; Kawagoe, Yoshiaki; Ohara, TakuJournal of Physical Chemistry C (2020), 124 (50), 27558-27570CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)The addn. of surfactants to polymer-based thermal interface materials applied to improve the heat dissipation efficiency of chip surfaces in contact has attracted attention for the microelectronic processing technol. In the present study, the mechanism by which a long-chain surfactant affects heat transfer across the interface between solid surface and polymer liq. was investigated by non-equil. mol. dynamics simulation. We constructed a system where tetracosane was used as a solvent and contained alc. mols. as a surfactant, and they were placed between two flat silica surfaces under a thermal gradient. The effect of the hydrophilicity of silica surface, the concn. of the surfactant, and chain length of the surfactant on silica-liq. interfacial thermal resistance Rb were examd. Alc. surfactant mols. preferred to adsorb onto the hydrophilic silica (Si-OH) surface due to hydrogen bonding between alc. and silanol hydroxyl groups. It was found that Rb reduced not only with the adsorption amt. of alc. mols. but also with the chain length of alc. The van der Waals interaction contribution was dominant for solid-liq. and liq.-liq. heat conduction near the interface. The hydroxyl terminals of alc. mols. were vertically adsorbed onto the Si-OH surface due to hydrogen bonds, which produced a heat path from silanols to the hydroxyl groups of alc. Furthermore, heat was also exchanged between alc. hydroxyl and alkyl groups via intramol. interaction and between the alc. alkyl groups and nearby solvent mols. via van der Waals (vdW) intermol. interaction. This resulted in an efficient heat path from solid surface silanols to liq. bulk. As the alc. chain length increased without changing the no. of adsorbed alc. mols., the heat transfer through this heat path increased, which led to a decrease in Rb. These results provided insight toward the guiding principle for the mol. design of complex surfactants to enhance the interfacial heat transfer.
- 38Uchaker, E.; Gu, M.; Zhou, N.; Li, Y.; Wang, C.; Cao, G. Enhanced Intercalation Dynamics and Stability of Engineered Micro/Nano-Structured Electrode Materials: Vanadium Oxide Mesocrystals. Small 2013, 9, 3880– 3886, DOI: 10.1002/smll.201203187Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXntF2ksLk%253D&md5=6c44f9f3abed96d1ede41c768401b5a9Enhanced Intercalation Dynamics and Stability of Engineered Micro/Nano-Structured Electrode Materials: Vanadium Oxide MesocrystalsUchaker, Evan; Gu, Meng; Zhou, Nan; Li, Yanwei; Wang, Chongmin; Cao, GuozhongSmall (2013), 9 (22), 3880-3886CODEN: SMALBC; ISSN:1613-6810. (Wiley-VCH Verlag GmbH & Co. KGaA)An additive and template free process is developed for the facile synthesis of VO2(B) mesocrystals via the solvothermal reaction of oxalic acid and vanadium pentoxide. The six-armed star architectures are composed of stacked nanosheets homoepitaxially oriented along the [100] crystallog. register with respect to one another, as confirmed by means of selected area electron diffraction and electron microscopy. It is proposed that the mesocrystal formation mechanism proceeds through classical as well as non-classical crystn. processes, and is possibly facilitated or promoted by the presence of a reducing/chelating agent. The synthesized VO2(B) mesocrystals are tested as a cathodic electrode material for lithium-ion batteries, and show good capacity at discharge rates ranging from 150-1500 mA g-1 and a cyclic stability of 195 mA h g-1 over fifty cycles. The superb electrochem. performance of the VO2(B) mesocrystals is attributed to the porous and oriented superstructure that ensures large surface area for redox reaction and short diffusion distances. The mesocryst. structure ensures that all the surfaces are in intimate contact with the electrolyte, and that lithium-ion intercalation occurs uniformly throughout the entire electrode. The exposed (100) facets also lead to fast lithium intercalation, and the homoepitaxial stacking of nanosheets offers a strong inner-sheet binding force that leads to better accommodation of the strain induced during cycling, thus circumventing the capacity fading issues typically assocd. with VO2(B) electrodes.
- 39Zeininger, L.; Portilla, L.; Halik, M.; Hirsch, A. Quantitative Determination and Comparison of the Surface Binding of Phosphonic Acid, Carboxylic Acid, and Catechol Ligands on TiO2 Nanoparticles. Chem.─Eur. J. 2016, 22, 13506– 13512, DOI: 10.1002/chem.201601920Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xht1WmtLbL&md5=9c49d50096a1908a791fdd7cc4bf7991Quantitative Determination and Comparison of the Surface Binding of Phosphonic Acid, Carboxylic Acid, and Catechol Ligands on TiO2 NanoparticlesZeininger, Lukas; Portilla, Luis; Halik, Marcus; Hirsch, AndreasChemistry - A European Journal (2016), 22 (38), 13506-13512CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The adsorption, desorption, co-adsorption, and exchange of phosphonic acid, carboxylic acid, and catechol derivs. on the surface of titanium oxide (anatase) nanoparticles are investigated. Thermogravimetric anal. provides a facile and fast-track quant. detn. of the wet-chem. monolayer adsorption consts. and grafting densities of ten adsorbates, all under neutral pH conditions. This characterization protocol allows straightforward quantification of the relevant thermodn. data of ligand adsorption and a comparison of ligand adsorption strengths. The reported procedure is proposed as a universal tool and it should be applicable to many other colloidal metal oxide materials. Moreover, the detd. values for the adsorption consts. and the monolayer grafting densities provide a toolbox for the assessment of the adsorbates' behavior in desorption, exchange, and co-adsorption equil. This versatile evaluation procedure will help to identify optimal monolayer-surface combinations and to evaluate crit. parameters, such as monolayer robustness, ligand exchange rates, or targeted mixed assembly of functionalities.
- 40De Roo, J.; Justo, Y.; De Keukeleere, K.; Van den Broeck, F.; Martins, J. C.; Van Driessche, I.; Hens, Z. Carboxylic-Acid-Passivated Metal Oxide Nanocrystals: Ligand Exchange Characteristics of a New Binding Motif. Angew. Chem., Int. Ed. 2015, 54, 6488– 6491, DOI: 10.1002/anie.201500965Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXmtlGqsr4%253D&md5=a55ec98a8761a97e2f1900720b8f7aa5Carboxylic-Acid-Passivated Metal Oxide Nanocrystals: Ligand Exchange Characteristics of a New Binding MotifDe Roo, Jonathan; Justo, Yolanda; De Keukeleere, Katrien; Van den Broeck, Freya; Martins, Jose C.; Van Driessche, Isabel; Hens, ZegerAngewandte Chemie, International Edition (2015), 54 (22), 6488-6491CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Studying sterically stabilized HfO2 and ZrO2 NCs using 1H soln. NMR and IR spectroscopy as well as elemental anal., this paper demonstrates the reversible exchange of initial oleic acid ligands for octylamine and self-adsorption of oleic acid at NC surfaces. Both processes are incompatible with an X-type binding motif of carboxylic acids as reported for sulfide and selenide NCs. This behavior stems from the dissociative adsorption of carboxylic acids at the oxide surface. Both proton and carboxylate moieties must be regarded as X-type ligands yielding a combined X2 binding motif that allows for self-adsorption and exchange for L-type ligands.
- 41Calatayud, D. G.; Rodríguez, M.; Jardiel, T. Controlling the morphology of TiO2 nanocrystals with different capping agents. Bol. Soc. Esp. Ceram. Vidrio 2015, 54, 159– 165, DOI: 10.1016/j.bsecv.2015.07.001Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1Snu77O&md5=691f7e4efcd1878d15164aa9c21d983dControlling the morphology of TiO2 nanocrystals with different capping agentsCalatayud, David G.; Rodriguez, Monica; Jardiel, TeresaBoletin de la Sociedad Espanola de Ceramica y Vidrio (2015), 54 (4), 159-165CODEN: BSCVB9; ISSN:0366-3175. (Sociedad Espanola de Ceramica y Vidrio)This paper provides direct evidence to support the role of capping agents in controlling the evolution of TiO2 seeds into nanocrystals with a specific shape. Starting with Ti(OBut)4 and using oleid acid, oleylamine, dioleamide, 11-aminoundecanoic acid, arginine, trifluroacetic acid or HF as capping agents, mainly TiO2 truncated octahedrons enclosed by {1 0 1} and {0 0 1} facets were obtained. We could also selectively obtain square, rods and rounded rhombic-shaped nanoparticles by growing of {0 1 0} facets by adding oleic acid and oleylamine in ratio 6:4, resp., while all other parameters were kept the same. This research not only offers new insights into the role played by a capping agent in shape-controlled synthesis but also provides, a versatile approach to controlling the shape of metal oxide nanocrystals.
Cited By
This article is cited by 1 publications.
- Emma J. Endres, Jeremy R. Bairan Espano, Alexandra Koziel, Antony R. Peng, Andrey A. Shults, Janet E. Macdonald. Controlling Phase in Colloidal Synthesis. ACS Nanoscience Au 2024, 4
(3)
, 158-175. https://doi.org/10.1021/acsnanoscienceau.3c00057
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
References
This article references 41 other publications.
- 1Wu, C.; Feng, F.; Xie, Y. Design of vanadium oxide structures with controllable electrical properties for energy applications. Chem. Soc. Rev. 2013, 42, 5157– 5183, DOI: 10.1039/c3cs35508j1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXot1ersr8%253D&md5=ddc30e955f7e900d2efa35cb65c64809Design of vanadium oxide structures with controllable electrical properties for energy applicationsWu, Changzheng; Feng, Feng; Xie, YiChemical Society Reviews (2013), 42 (12), 5157-5183CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The elec. properties of inorg. materials has been a long-standing pursued research topic, and successfully controlling the elec. property of an inorg. material has attracted significant attention for a wide range of energy-related applications, covering energy storage, energy conversion and energy utilization. During the few past decades, vanadium oxides have been studied to gain a clear picture of how microstructural characteristics generating the e-e correlations influence the electronic structure of a material, through which the charge concn., elec. cond. as well as the metal-insulator transition (MIT), etc., can be precisely controlled, giving promising signs for constructing energy-related devices. In this review, we present an extensive review of the engineering of the microstructures of vanadium oxides with control of their elec. properties, and with attempts to rationally construct energy-related devices, such as aq. lithium ion batteries, supercapacitors for energy storage, and thermoelec. generators for energy conversion. Furthermore, the MIT performance of vanadium oxides has also seen tremendous advantages for the applications of "smart windows" and magnetocaloric refrigerators for energy utilization. Collectively, progresses to date suggest that in vanadium oxide systems, the elec. properties, including elec. cond., carrier concns., and the MIT performance, were all strongly dependent on the microstructural characteristics at the at. scale, which have presented extensive promising energy applications covering energy storage, energy conversion and energy utilization.
- 2Li, M.; Magdassi, S.; Gao, Y.; Long, Y. Hydrothermal Synthesis of VO2 Polymorphs: Advantages, Challenges and Prospects for the Application of Energy Efficient Smart Windows. Small 2017, 13, 1701147, DOI: 10.1002/smll.201701147There is no corresponding record for this reference.
- 3Zhang, Y.; Fan, M.; Liu, X.; Xie, G.; Li, H.; Huang, C. Synthesis of VO2(A) nanobelts by the transformation of VO2(B) under the hydrothermal treatment and its optical switching properties. Solid State Commun. 2012, 152, 253– 256, DOI: 10.1016/j.ssc.2011.11.0363https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XpsFOjsA%253D%253D&md5=e99134ae24630224d8694450f8f372fdSynthesis of VO2(A) nanobelts by the transformation of VO2(B) under the hydrothermal treatment and its optical switching propertiesZhang, Yifu; Fan, Meijuan; Liu, Xinghai; Xie, Guangyong; Li, Houbin; Huang, ChiSolid State Communications (2012), 152 (4), 253-256CODEN: SSCOA4; ISSN:0038-1098. (Elsevier Ltd.)VO2(A) nanobelts had been successfully synthesized by the transformation of VO2(B) using H2O as the solvent under the hydrothermal approach at 280 °C for 48 h. Some parameters, such as the reaction temp. and time, had been briefly discussed to reveal the transition from VO2(B) to VO2(A). It was found that H2O played a crucial role in the transition from VO2(B) to VO2(A). The phase transition of VO2(A) nanobelts was at 162 °C. The optical switching properties of VO2(A) were studied by the variable-temp. IR spectra for the first time. In addn., VO2(A) nanobelts were calcined at 700 °C for 2 h under a high purity Ar (99.999%) atm. to obtain VO2(M) which exhibited a strong crystallog. transition at around 65 °C.
- 4Liu, P.; Zhu, K.; Gao, Y.; Wu, Q.; Liu, J.; Qiu, J.; Gu, Q.; Zheng, H. Ultra-long VO2(A) nanorods using the high-temperature mixing method under hydrothermal conditions: synthesis, evolution and thermochromic properties. CrystEngComm 2013, 15, 2753– 2760, DOI: 10.1039/c3ce27085h4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjslKgs78%253D&md5=24b8f349248f47c43ba3f776251a7173Ultra-long VO2 (A) nanorods using the high-temperature mixing method under hydrothermal conditions: synthesis, evolution and thermochromic propertiesLiu, Pengcheng; Zhu, Kongjun; Gao, Yanfeng; Wu, Qingliu; Liu, Jinsong; Qiu, Jinhao; Gu, Qilin; Zheng, HongjuanCrystEngComm (2013), 15 (14), 2753-2760CODEN: CRECF4; ISSN:1466-8033. (Royal Society of Chemistry)Well-crystd., ultra-long VO2 (A) nanorods were prepd. using a facile high-temp. mixing method (HTMM) under hydrothermal conditions. The as-obtained products were characterized by x-ray diffraction, field-emission SEM, transmission electron microscopy, XPS, UV-Vis-NIR, DSC, and FTIR. The effect of W doping on the phase-transition properties of VO2(A) was also studied. The optimal condition of VO2(A) using the HTMM is hydrothermal treatment at 240° for 48 h with a 3:1 molar ratio of the reducing agent to the vanadium source. The reason why the polymorphic forms of VO2 show different colors is that the light in the visible region reflected by the samples is different. The phase-transition temp. of the pure VO2 (A) is 154.75°, which is significantly lower than the 162 °C reported previously. When a small amt. of W is doped, VO2(A) will be transformed into other polymorphic forms, which indicates that the crystal structure of VO2(A) is highly sensitive to limited doping. Importantly, the as-obtained pure VO2(A) shows good thermochromic properties and optical-switching characters. A crystal growth mechanism for VO2(A), oriented-attachment-exfoliation-recrystn.-oriented-attachment, is proposed and described in detail.
- 5Subba Reddy, C. V.; Walker, E. H., Jr.; Wicker, S. A., Sr.; Williams, Q. L.; Kalluru, R. R. Synthesis of VO2(B) nanorods for Li battery application. Curr. Appl. Phys. 2009, 9, 1195– 1198, DOI: 10.1016/j.cap.2009.01.012There is no corresponding record for this reference.
- 6Soltane, L.; Sediri, F. Rod-like nanocrystalline B-VO2: Hydrothermal synthesis, characterization and electrochemical properties. Mater. Res. Bull. 2014, 53, 79– 83, DOI: 10.1016/j.materresbull.2014.01.0466https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXmsFCmtLk%253D&md5=8b9e591eaf47af46b4d92ff3dd4afc0aRod-like nanocrystalline B-VO2: Hydrothermal synthesis, characterization and electrochemical propertiesSoltane, L.; Sediri, F.Materials Research Bulletin (2014), 53 (), 79-83CODEN: MRBUAC; ISSN:0025-5408. (Elsevier Ltd.)Rod-like nanocryst. VO2(B) has been successfully synthesized via a simple hydrothermal process by using V2O5 as vanadium source and 4-butylaniline H3C-(CH2)3-(C6H4)-NH2 as reducing and structure directing agent. The compds. were analyzed through X-ray diffraction (XRD), SEM (SEM), Fourier transform IR spectroscopy (FTIR), Raman spectroscopy and UV-visible spectroscopy. The VO2(B) nanorods are up to several micrometers in length and about 80 nm in thickness with a large optical band gap of ∼2.709 eV. Thin films of VO2(B) nanorods deposited on ITO substrates were electrochem. characterized by cyclic voltammetry. The voltammograms show reversible redox behavior with charge-discharge cycling process corresponding to the reversible lithium intercalation/deintercalation into the crystal lattice. The av. coulombic efficiency to this redox processes is upper 98% during the electrochem. measurements.
- 7Yin, H.; Yu, K.; Zhang, Z.; Zhu, Z. Morphology-control of VO2(B) nanostructures in hydrothermal synthesis and their field emission properties. Appl. Surf. Sci. 2011, 257, 8840– 8845, DOI: 10.1016/j.apsusc.2011.04.0797https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXos1egsr0%253D&md5=8e63026251f96878a79d78084f070814Morphology-control of VO2 (B) nanostructures in hydrothermal synthesis and their field emission propertiesYin, Haihong; Yu, Ke; Zhang, Zhengli; Zhu, ZiqiangApplied Surface Science (2011), 257 (21), 8840-8845CODEN: ASUSEE; ISSN:0169-4332. (Elsevier B.V.)VO2 (B) nanostructures were synthesized via a facile hydrothermal process using V2O5 as source material and oxalic acid as reductant. Three nanostructures of nanorods, nanocarambolas and nanobundles were found existing in the products, and a continuous changing of morphol. was found in the synthesis process, during which the proportion of these three types of nanostructures can be adjusted by altering the concns. of oxalic acid. The microstructures were evaluated using X-ray diffraction, SEM and TEM. Field emission (FE) properties of these three types of nanostructures showed that the nanobundles have the best field emission performance with a turn-on field of ∼1.4 V/μm and a threshold field of ∼5.38 V/μm. These characteristics make VO2 (B) nanostructures a competitive cathode material in field emission devices.
- 8Ganganagappa, N.; Siddaramanna, A. One step synthesis of monoclinic VO2 (B) bundles of nanorods: Cathode for Li ion battery. Mater. Charact. 2012, 68, 58– 62, DOI: 10.1016/j.matchar.2012.03.0108https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XntFyksLY%253D&md5=1aa9780a7480b9a28d9680b529a2636eOne step synthesis of monoclinic VO2 (B) bundles of nanorods: Cathode for Li ion batteryGanganagappa, Nagaraju; Siddaramanna, AshokaMaterials Characterization (2012), 68 (), 58-62CODEN: MACHEX; ISSN:1044-5803. (Elsevier)One of the metastable phases of vanadium dioxide, VO2(B) bundles of nanorods and microspheres have been synthesized through a simple hydrothermal method by dispersing V2O5 in aq. quinol. The obtained products were characterized by X-ray diffraction (XRD), Fourier transform IR (FTIR) spectroscopy, SEM (SEM) and electrochem. discharge-charge test for lithium battery. It was found that the morphologies of the obtained VO2(B) can be tuned by manipulating the relative amt. of quinol. The electrochem. test found that the bundles of nanorods exhibit an initial discharge capacity of 171 mAh g- 1 and its almost stabilized capacity was reached to 108 mAh g- 1 after 47 cycles at a c.d. of 0.1 mA g- 1. The formation mechanism of the VO2(B) bundles of nanorods and microspheres was also discussed.
- 9Zhang, Y. VO2(B) conversion to VO2(A) and VO2(M) and their oxidation resistance and optical switching properties. Mater. Sci.-Pol. 2016, 34, 169– 176, DOI: 10.1515/msp-2016-00239https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XnvFWgtLg%253D&md5=824fffaabc1ee70900b734e7cb8ed72dVO2(B) conversion to VO2(A) and VO2(M) and their oxidation resistance and optical switching propertiesZhang, YifuMaterials Science-Poland (2016), 34 (1), 169-176CODEN: MSACDV; ISSN:2083-134X. (Walter de Gruyter GmbH)Vanadium dioxide VO2 has been paid in recent years increasing attention because of its various applications, however, its oxidn. resistance properties in air atm. have rarely been reported. Herein, VO2(B) nanobelts were transformed into VO2(A) and VO2(M) nanobelts by hydrothermal route and calcination treatment, resp. Then, we comparatively studied the oxidn. resistance properties of VO2(B), VO2(A) and VO2(M) nanobelts in air atm. by thermo-gravimetric anal. and DTA (TGA/DTA). It was found that the nanobelts had good thermal stability and oxidn. resistance below 341°C, 408°C and 465°C in air, resp., indicating that they were stable in air at room temp. The fierce oxidn. of the nanobelts occurred at 426, 507 and 645°C, resp. The results showed that the VO2(M) nanobelts had the best thermal stability and oxidn. resistance among the others. Furthermore, the phase transition temps. and optical switching properties of VO2(A) and VO2(M) were studied by differential scanning calorimetry (DSC) and variable temp. IR spectra. It was found that the VO2(A) and VO2(M) nanobelts had outstanding thermochromic character and optical switching properties.
- 10Sediri, F.; Gharbi, N. Controlled hydrothermal synthesis of VO2(B) nanobelts. Mater. Lett. 2009, 63, 15– 18, DOI: 10.1016/j.matlet.2008.08.022There is no corresponding record for this reference.
- 11Wang, Q.; Xu, J.; Zhang, W.; Mao, M.; Wei, Z.; Wang, L.; Cui, C.; Zhu, Y.; Ma, J. Research progress on vanadium-based cathode materials for sodium ion batteries. J. Mater. Chem. A 2018, 6, 8815– 8838, DOI: 10.1039/c8ta01627e11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXntlCnsrs%253D&md5=32772400483c5ac72c28bae062b824d9Research progress on vanadium-based cathode materials for sodium ion batteriesWang, Qinghong; Xu, Jiantie; Zhang, Wenchao; Mao, Minglei; Wei, Zengxi; Wang, Lei; Cui, Chunyu; Zhu, Yuxuan; Ma, JianminJournal of Materials Chemistry A: Materials for Energy and Sustainability (2018), 6 (19), 8815-8838CODEN: JMCAET; ISSN:2050-7496. (Royal Society of Chemistry)A review is presented. Sodium ion batteries (SIBs) have attracted increasing attention as one of the most promising candidates for cost-effective, high-energy rechargeable batteries. Owing to their high theor. capacity and energy d., and rich electrochem. interaction with Na+ (V2+-V5+), a large no. of vanadium(V)-based cathode materials, including vanadium oxides (e.g., V2O5 and VO2), vanadium bronzes (e.g., NaxVO2, NaV3O8, NaV6O15 and δ-NH4V4O10), V-based phosphates (e.g., Na3V2(PO4)3, VOPO4, NaVOPO4, Na7V3(P2O7)4 and Na2(VO)P2O7) and F-contg. V-based polyanions (e.g., NaVPO4F, Na3V2(PO4)2F3 and Na3(VOx)2(PO4)2F3-2x), have been explored for SIBs. In this review, we mainly summarize the basic structures, modified/optimized structures, synthetic methods and morphol. control of V-based cathode materials for SIBs. Addnl., major drawbacks, emerging challenges and some perspectives on the development of V-based cathode materials for SIBs are also discussed.
- 12Zhang, L.; Yao, J.; Guo, Y.; Xia, F.; Cui, Y.; Liu, B.; Gao, Y. VO2(A) Nanorods: One-pot Synthesis, Formation Mechanism and Thermal Transformation to VO2(M). Ceram. Int. 2018, 44, 19301– 19306, DOI: 10.1016/j.ceramint.2018.07.15712https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhsVartb3L&md5=b980416b5b434be3ddf18b028717cc1cVO2(A) nanorods: One-pot synthesis, formation mechanism and thermal transformation to VO2(M)Zhang, Liangmiao; Yao, Jianing; Guo, Yunfeng; Xia, Fang; Cui, Yuanyuan; Liu, Bin; Gao, YanfengCeramics International (2018), 44 (16), 19301-19306CODEN: CINNDH; ISSN:0272-8842. (Elsevier Ltd.)The monoclinic VO2(M) has promising applications in intelligent devices but its prepn. still requires improvement to permit cost-effective mass prodn. In this work, we report a 2-stage approach for producing VO2(M) nanorods by (1) hydrothermal redn. of vanadium pentoxide by sodium bisulfate at 220°C to form VO2(A), and (2) subsequent thermal activated phase transformation of VO2(A) to VO2(M) at 350-450°C in vacuum. The obtained VO2(M) nanorods showed a reversible phase transition temp. at about 62.5°C and a narrow thermal hysteresis width of 10°C. The mechanism of the hydrothermal redn. was studied by combined ex situ microscopic and diffraction characterization of cooled samples as well as in situ PXRD expts., in which the hydrothermal synthesis was monitored in real time by time-resolved diffraction datasets. It was found that the hydrothermal synthesis of VO2(A) is a 4-step process: (1) redn. of V2O5 to form VO2(B) nanoparticles, (2) oriented attachment of VO2(B) nanoparticles along the [110] direction, (3) formation of VO2(B) nanorods as a results of oriented attachments, and (4) hydrothermal transformation of the metastable intermediate VO2(B) nanorods to VO2(A) nanorods. This clear understanding of the mechanism will help the further optimization of synthesis temp. and time for prepg. VO2(A). This method provides a low temp. thermal treatment alternative and hence helps the redn. of cost for the prodn. of VO2(M), bring the mass application of VO2(M) one step closer.
- 13Popuri, S.; Artemenko, A.; Labrugere, C.; Miclau, M.; Villesuzanne, A.; Pollet, M. VO2(A) Reinvestigation of crystal structure, phase transition and crystal growth mechanisms. J. Solid State Chem. 2014, 213, 79– 86, DOI: 10.1016/j.jssc.2014.01.037There is no corresponding record for this reference.
- 14Ji, S.; Zhang, F.; Jin, P. Selective formation of VO2(A) or VO2(R) polymorph by controlling the hydrothermal pressure. J. Solid State Chem. 2011, 184, 2285– 2292, DOI: 10.1016/j.jssc.2011.06.02914https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXpt1Cms7Y%253D&md5=7be35329c5ffd268cc056cfdda465286Selective formation of VO2(A) or VO2(R) polymorph by controlling the hydrothermal pressureJi, Shidong; Zhang, Feng; Jin, PingJournal of Solid State Chemistry (2011), 184 (8), 2285-2292CODEN: JSSCBI; ISSN:0022-4596. (Elsevier B.V.)VO2(A) usually occurs during the prepn. of VO2 polymorphs. This leads to an ambiguous understanding of the transformation between VO2 polymorphs. The calcn. of the ground state energies for different VO2 polymorphs indicated that there is only a small energy gap between VO2(A) and VO2(R), which suggests that the transformation from VO2(A) to VO2(R) should be pressure sensitive. This hypothesis was verified during the synthesizing of VO2 polymorphs by reducing V2O5 with oxalic acid through hydrothermal treatment process. Selective formation of pure phase VO2(A) or VO2(R) was achieved by controlling the hydrothermal pressure through varying the filling ratio at 270°. It was found that a filling ratio >0.5 favors the formation of pure VO2(R) while a filling ratio ≤0.4 results in the formation of VO2(A). Based on our expts., VO2(B) nanobelts always formed first and then transformed to VO2(A) by an assembling process at increased temp. or extended reaction time. Under further higher pressure, the VO2(A) transformed spontaneously to VO2(R) initialized from the vol. shrinkage due to the formation of denser VO2(R).
- 15Cheng, X. H.; Xu, H. F.; Wang, Z. Z.; Zhu, K. R.; Li, G.; Jin, S. Synthesis, characterization and formation mechanism of metastable phase VO2(A) nanorods. Mater. Res. Bull. 2013, 48, 3383– 3388, DOI: 10.1016/j.materresbull.2013.05.01615https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXovVCms7c%253D&md5=a5f9bc3aa90d57928080e541829afa04Synthesis, characterization and formation mechanism of metastable phase VO2(A) nanorodsCheng, X. H.; Xu, H. F.; Wang, Z. Z.; Zhu, K. R.; Li, G.; Jin, ShaoweiMaterials Research Bulletin (2013), 48 (9), 3383-3388CODEN: MRBUAC; ISSN:0025-5408. (Elsevier Ltd.)Pure phase VO2(A) nanorods were synthesized via the redn. of V2O5 by oxalic acid during the hydrothermal treatment. Two sets of samples were prepd. by varying both system temp. and reaction time under a filling ratio of 0.40 for observing the formation and evolution of VO2(A) nanorods. Structures were characterized by X-ray diffraction, scanning and transmission electron microscopies, resp. It was found that VO2(B) was firstly formed and then transformed into VO2(A) as the increasing system temp. or extending reaction time. An assembling and following crystal adjustment was proposed for explanation the formation process of VO2(A) from VO2(B). For VO2(A) nanorods, the phase transition temp. of 169.7°C was higher than that of the VO2(A) bulk, it might be ascribed to the lower crystallinity or nonstoichiometry in VO2(A) nanorods. VO2 nanostructures with controllable phases and properties should find their promising applications in a single VO2 nanodevice.
- 16Zhang, L.; Xia, F.; Song, Z.; Webster, N. A. S.; Luo, H.; Gao, Y. Synthesis and formation mechanism of VO2(A) nanoplates with intrinsic peroxidase-like activity. RSC Adv. 2015, 5, 61371– 61379, DOI: 10.1039/c5ra11014a16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtFChtLrL&md5=f3ad20d4327c286dc3fbafce611c073aSynthesis and formation mechanism of VO2(A) nanoplates with intrinsic peroxidase-like activityZhang, Liangmiao; Xia, Fang; Song, Zhengdong; Webster, Nathan A. S.; Luo, Hongjie; Gao, YanfengRSC Advances (2015), 5 (75), 61371-61379CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Monocryst. VO2(A) nanoplates were synthesized via a 1-pot hydrothermal process. In situ powder x-ray diffraction was used to monitor the hydrothermal synthesis and VO2(A) nucleates and grows directly from soln. after the complete hydrolysis of a 2.0M VO(acac)2 precursor soln., rather than involving a previously reported intermediate phase VO2(B). A hydrating-exfoliating-splitting mechanism was established to explain the formation of the nanoplate architecture. The synthesized VO2(A) nanoplates showed outstanding peroxidase-like activity and hence are a promising candidate for artificial peroxidase.
- 17Zhang, S.; Shang, B.; Yang, J.; Yan, W.; Wei, S.; Xie, Y. From VO2(B) to VO2(A) nanobelts: first hydrothermal transformation, spectroscopic study and first principles calculation. Phys. Chem. Chem. Phys. 2011, 13, 15873– 15881, DOI: 10.1039/c1cp20838a17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtVGqu7bN&md5=39cdf6a5903c9c1ccf91631e97f28cb5From VO2 (B) to VO2 (A) nanobelts: first hydrothermal transformation, spectroscopic study and first principles calculationZhang, Shudong; Shang, Bo; Yang, Jinlong; Yan, Wensheng; Wei, Shiqiang; Xie, YiPhysical Chemistry Chemical Physics (2011), 13 (35), 15873-15881CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)The phase transition process from VO2 (B) to VO2 (A) was first obsd. through a mild hydrothermal approach, using hybrid d. functional theory (DFT) calcns. and crystallog. VO2 topol. anal. All theor. analyses reveal that VO2 (A) is a thermodynamically stable phase and has a lower formation energy compared with the metastable VO2 (B). For the first time, X-ray absorption spectroscopy (XAS) of the V L-edge and O K-edge was performed on different VO2 phases, and the differences in the electronic structure of the two polymorphic forms provide further exptl. evidence of the more stable VO2 (A). Consequently, transformation from VO2 (B) to VO2 (A) is much easier to be realized from a dynamical point of view. Notably, the transformation of VO2 (B) into VO2 (A) show the sequence VO2 (B)-high-temp. VO2 (AH) phase-low-temp. VO2 (A) phase, which was achieved by hydrothermal treatment, resp. Also, an alternative synthesis route was proposed based on the above hydrothermal transformation, and VO2 (A) was successfully prepd. via the simple one-step hydrothermal method by hydrolysis of VO(acac)2 (acac = acetylacetonate). Therefore, VO2 nanostructures with controlled phase compns. can be obtained in high yields. Through elucidating the structural evolution in the crystallog. shear mechanism, we can easily guide the design of other metal oxide nanostructures with controllable phases.
- 18Zhang, S.; Li, Y.; Wu, C.; Zheng, F.; Xie, Y. Novel Flowerlike Metastable Vanadium Dioxide (B) Micronanostructures: Facile Synthesis and Application in Aqueous Lithium Ion Batteries. J. Phys. Chem. C 2009, 113, 15058– 15067, DOI: 10.1021/jp903312h18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXovFKlt74%253D&md5=853e81161aac174de5b76d5e68fc5045Novel Flowerlike Metastable Vanadium Dioxide (B) Micronanostructures: Facile Synthesis and Application in Aqueous Lithium Ion BatteriesZhang, Shudong; Li, Yingmei; Wu, Changzheng; Zheng, Fei; Xie, YiJournal of Physical Chemistry C (2009), 113 (33), 15058-15067CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)Novel flowerlike VO2(B) micro-nanostructures assembled by single-cryst. nanosheets were synthesized via a hydrothermal route using polymer polyvinyl pyrrolidone (PVP, K30) as capping reagent. Detailed proofs indicated that the process of crystal growth was dominated by a nucleation and growth, self-assembly, and then Ostwald ripening growth mechanism. The flowerlike micro-nanostructured VO2(B) was applied as the active material in aq. Li-ion batteries which showed improved electrochem. properties with the 1st discharge capacity reaching 74.9 mA-h/g, a good value for aq. Li ion battery systems in light of previous reports (usually <65 mA-h/g). Corresponding VO2(B) nanostructures with better crystallinity were obtained by calcining the precursor of flowerlike VO2(B) structures. The post-treated flowerlike VO2(B) electrode shows better electrochem. properties with the 1st discharge capacity reaching 81.3 mA-h/g, which is higher than that of the flowerlike VO2(B) sample before annealing. The electrochem. intercalation and deintercalation properties of VO2(B) nanobelts and carambola-like VO2(B) structures with Li+ were also studied. The unique flowerlike structure plays a role in the morphol. requirement to serve as transport paths for Li ions in aq. Li ion batteries. The morphol. and crystallinity of the synthesized products had an influence on the electrochem. intercalation and deintercalation properties with Li ions.
- 19Walton, R. I. Subcritical solvothermal synthesis of condensed inorganic materials. Chem. Soc. Rev. 2002, 31, 230– 238, DOI: 10.1039/b105762f19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XkvVCmt7o%253D&md5=ac8805ccaff49f722b4c472ee85726daSubcritical solvothermal synthesis of condensed inorganic materialsWalton, Richard I.Chemical Society Reviews (2002), 31 (4), 230-238CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review. The solvothermal method has recently been extended from zeolite synthesis to the formation of condensed inorg. solids, which find uses in diverse areas due to properties such as ionic-cond., solid-state magnetism, giant magnetoresistance, low thermal expansion and ferroelectricity. This offers specific advantages over the traditional ceramic synthetic routes to inorg. solids and these are highlighted with examples from the recent literature, and the efforts focussed on detg. the formation mechanism of solids from the heterogeneous mixts. used in solvothermal procedures are discussed.
- 20Yu, W.; Li, S.; Huang, C. Phase evolution and crystal growth of VO2 nanostructures under hydrothermal reactions. RSC Adv. 2016, 6, 7113– 7120, DOI: 10.1039/c5ra23898f20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XlvV2lsQ%253D%253D&md5=38c86bd180df72c567bacbc3a3b8057dPhase evolution and crystal growth of VO2 nanostructures under hydrothermal reactionsYu, Weilai; Li, Shuai; Huang, ChiRSC Advances (2016), 6 (9), 7113-7120CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)The phase evolution and crystal growth of VO2 nanostructures against reaction time in a high-pressure V2O5-oxalic acid hydrothermal system were systematically investigated. It was found that the rather thin VO2 (B) nanobelts were first obtained, then stacked to form large belt-like structures and subsequently phase transformed into VO2 (A), based on an oriented attachment-recrystn. mechanism. The large VO2 (A) belt-like structures could further assemble into novel "snowflake" VO2 (M) microcrystals with even bigger sizes and nearly well-defined six-fold symmetry. Due to the Ostwald ripening effect regarding crystal size discrepancy, the VO2 (M) phase could further grow at the cost of the gradual dissoln. of VO2 (A) and full elimination of VO2 (B). The phase evolution from VO2 (B) first to VO2 (A) and then to VO2 (M), is actually a step-by-step thermodynamically downhill process, owing to the gradual relaxation of structural tension within the VO2 crystal lattice. Thus, our investigation, for the first time, demonstrated the feasibility of the well-known Ostwald's step rules towards the phase evolution process of VO2 and could provide unprecedented new insight to promote understanding of the synthesis and properties of vanadium oxide compds.
- 21Beidelman, B. A.; Zhang, X.; Sanchez-Lievanos, K. R.; Selino, A. V.; Matson, E. M.; Knowles, K. E. Influence of water concentration on the solvothermal synthesis of VO2(B) nanocrystals. CrystEngComm 2022, 24, 6009– 6017, DOI: 10.1039/d2ce00813k21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XitV2ks73O&md5=9d7d4808dd7ab07a57b23fb3525febaaInfluence of water concentration on the solvothermal synthesis of VO2(B) nanocrystalsBeidelman, Brittney A.; Zhang, Xiaotian; Sanchez-Lievanos, Karla R.; Selino, Annabel V.; Matson, Ellen M.; Knowles, Kathryn E.CrystEngComm (2022), 24 (34), 6009-6017CODEN: CRECF4; ISSN:1466-8033. (Royal Society of Chemistry)Nanocrystals of VO2(B) have attracted significant attention for their promising performance in electrochem. energy storage applications. Phase-purity and nanocrystal morphol. are both crucial factors for the performance of these materials, but given the large no. of crystal polymorphs available to VO2, achieving simultaneous control over crystal phase and nanocrystal size is a significant synthetic challenge. This paper describes the impact of water concn. on the synthesis of VO2(B) nanocrystals via solvothermal reaction of a mol. V(IV) precursor in toluene. Using controlled, stoichiometric amts. of water (20 equiv per vanadium center) enables access to short nanorods of VO2(B) whose lengths follow a Gaussian distribution with an av. and std. deviation of 110 ± 30 nm. Decreasing the amt. of water present to two or four equiv. results in formation of VO2(A) nanocrystals and eliminating it entirely results in no reaction. Increasing the amt. of water to more than 20 equiv increases the av. length of the VO2(B) nanorods and causes the distribution in rod lengths to evolve from Gaussian to lognormal. This evolution in the size distribution is consistent with changes in a model Gaussian distribution obsd. upon simulated end-to-end oriented attachment events. These results demonstrate that control over the concn. of water is a useful strategy for tuning the morphol. and crystal phase of VO2 nanocrystals.
- 22Serjeant, E. P.; Dempsey, B. Ionisation Constants of Organic Acids in Aqueous Solution; Pergamon Press: New York, New York, 1979; Vol. 23.There is no corresponding record for this reference.
- 23Haynes, W. H. CRC Handbook of Chemistry and Physics, 91st ed.; CRC Press Inc.: Boca Raton, FL, 2010–2011.There is no corresponding record for this reference.
- 24Dean, J. A. Lange’s Handbook of Chemistry, 13th ed.; McGraw-Hill Book Co: New York, NY, 1985.There is no corresponding record for this reference.
- 25Muckerman, J. T.; Skone, J. H.; Ning, M.; Wasada-Tsutsui, Y. Toward the accurate calculation of pKa values in water and acetonitrile. Biochim. Biophys. Acta, Bioenerg. 2013, 1827, 882– 891, DOI: 10.1016/j.bbabio.2013.03.01125https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXmsFCjtbk%253D&md5=c6e56760baf2f86e8940fcd72bb08af1Toward the accurate calculation of pKa values in water and acetonitrileMuckerman, James T.; Skone, Jonathan H.; Ning, Ming; Wasada-Tsutsui, YukoBiochimica et Biophysica Acta, Bioenergetics (2013), 1827 (8-9), 882-891CODEN: BBBEB4; ISSN:0005-2728. (Elsevier B. V.)We present a simple approach for the calcn. of accurate pKa values in water and acetonitrile based on the straightforward calcn. of the gas-phase abs. free energies of the acid and conjugate base with use of only a continuum solvation model to obtain the corresponding soln.-phase free energies. Most of the error in such an approach arises from inaccurate differential solvation free energies of the acid and conjugate base which is removed in our approach using a correction based on the realization that the gas-phase acidities have only a small systematic error relative to the dominant systematic error in the differential solvation. The methodol. is outlined in the context of the calcn. of a set of neutral acids with water as the solvent for a reasonably accurate electronic structure level of theory (DFT), basis set, and implicit solvation model. It is then applied to the comparison of results for three different hybrid d. functionals to illustrate the insensitivity to the functional. Finally, the approach is applied to the comparison of results for sets of neutral acids and protonated amine cationic acids in both aq. (water) and nonaq. (acetonitrile) solvents. The methodol. is shown to generally predict the pKa values for all the cases investigated to within 1 pH unit so long as the differential solvation error is larger than the systematic error in the gas-phase acidity calcns. Such an approach is rather general and does not have addnl. complications that would arise in a cluster-continuum method, thus giving it strength as a simple high-throughput means to calc. abs. pKa values. This article is part of a Special Issue entitled: Metals in Bioenergetics and Biomimetics Systems.
- 26Weeks, C.; Song, Y.; Suzuki, M.; Chernova, N. A.; Zavalij, P. Y.; Whittingham, M. S. The one dimensional chain structures of vanadyl glycolate and vanadyl acetate. J. Mater. Chem. 2003, 13, 1420– 1423, DOI: 10.1039/b208100h26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXjvVCgsLc%253D&md5=b570fd9cea9efed13b78c01cc79cd069The one dimensional chain structures of vanadyl glycolate and vanadyl acetateWeeks, Curtis; Song, Yanning; Suzuki, Masatsugu; Chernova, Natasha A.; Zavalij, Peter Y.; Whittingham, M. StanleyJournal of Materials Chemistry (2003), 13 (6), 1420-1423CODEN: JMACEP; ISSN:0959-9428. (Royal Society of Chemistry)The solvothermal reaction, at 200°, of vanadium pentoxide and lithium hydroxide in acetic acid or ethylene glycol gives vanadyl acetate and vanadyl glycolate resp. The structure of the acetate contains vanadium in octahedral coordination whereas the glycolate contains VO5 square pyramids. The VO6 octahedra in the acetate, VO(CH3COO)2, are joined through the vanadyl groups, giving a rather long V:O bond of 1.684(7) Å and a short trans V-O bond of 2.131(7) Å, and by bridging acetate groups. The vanadium atoms interact along the ···V:O···V:O··· chain giving 1-dimensional antiferromagnetic behavior. In contrast in the glycolate, the apical V:O bond is shorter, 1.58(1) Å, and the square pyramids share edges in a two up-two down fashion to give chains VO(OCH2CH2O). Magnetic susceptibility of vanadyl glycolate is consistent with an isolated spin dimers model.
- 27Casey, A. T.; Thackeray, J. R. The preparation and magnetic properties of oxovanadium(IV) acetate. Aust. J. Chem. 1969, 22, 2549– 2553, DOI: 10.1071/ch969254927https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXjslWmsg%253D%253D&md5=6c034e42b8006146026c3a6ed3ef4aa6Preparation and magnetic properties of oxovandium(IV) acetateCasey, Allan T.; Thackery, J. R.Australian Journal of Chemistry (1969), 22 (12), 2549-53CODEN: AJCHAS; ISSN:0004-9425.The variation of the magnetic susceptibility of oxovanadium(IV) acetate with temp. may be fitted to the Ising model of an infinite linear antifer romagnetic chain with exchange energy of -166 cm-1. On the basis of this and ir evidence, a structure for the compd. is proposed.
- 28Dakternieks, D. R.; Harris, C. M.; Milham, P. J.; Morris, B. S.; Sinn, E. Magnetism and structure of polymeric oxovanadium(IV) acetate and other polymeric oxovanadium(IV) complexes. Inorg. Nucl. Chem. Lett. 1969, 5, 97– 100, DOI: 10.1016/0020-1650(69)80177-728https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXhtF2mt7s%253D&md5=a63282881ea3c3e5c17a11c0144f2955Magnetism and structure of polymeric oxovanadium(IV) acetate and other polymeric oxovanadium(VI) complexesDakternieks, D. R.; Harris, Clive Melville; Milham, P. J.; Morris, Benjamin S.; Sinn, EkkehardInorganic and Nuclear Chemistry Letters (1969), 5 (2), 97-100CODEN: INUCAF; ISSN:0020-1650.From magnetic and other evidence, it is concluded that oxovanadium(IV) acetate and other carboxylate complexes consist of polymeric structures. The V:O stretching absorption at 895 cm.-1 gives a force const. k of 5.8 millidynes/A.
- 29Ikekwere, P. O.; Adeniyi, A. A. Oxovanadium(IV) Complexes of Some Aromatic Carboxylic Acids. Synth. React. Inorg. Met.-Org. Chem. 1989, 19, 87– 100, DOI: 10.1080/0094571890804805329https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXktFKltbs%253D&md5=ab83fb438d173fba7cf52badac767214Oxovanadium(IV) complexes of some aromatic carboxylic acidsIkekwere, P. O.; Adeniyi, A. A.Synthesis and Reactivity in Inorganic and Metal-Organic Chemistry (1989), 19 (1), 87-100CODEN: SRIMCN; ISSN:0094-5714.VOL2.nH2O (L = BzOH and its 2-, 3- and 4-cyano-, 3-cyano-4-methyl-, 2-, 3- and 4-methyl-, 2,5- and 3,5-dimethyl-, 2,4,6-trimethyl-, 2-, 3- and 4-nitro derivs.) were prepd. and characterized by their magnetic and spectral properties. The complexes had magnetic moments much lower than the spin-only value of 1.73 μB at room temp. indicating strong magnetic interactions. Their Nujol mull spectra showed 3 main electronic absorption bands typical of tetragonally distorted octahedral oxovanadium(IV) compds. The IR vibrational spectra indicate that the carboxylate groups are bridging. The complexes are carboxylate-bridged polymers with V:O...V:O interaction.
- 30Puri, M.; Sharma, R. D.; Verma, R. D. Trifluoroacetates of Vanadium(III) and Oxovanadium(IV) and (V) Preparation and Characterization. Synth. React. Inorg. Met.-Org. Chem. 1981, 11, 539– 546, DOI: 10.1080/0094571810805599830https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38Xht1aitLs%253D&md5=dfdabce00d2634d5c7519bd8d3f3a35dTrifluoroacetates of vanadium(III) and oxovanadium(IV) and (V). Preparation and characterizationPuri, Madhu; Sharma, R. D.; Verma, Rajendar D.Synthesis and Reactivity in Inorganic and Metal-Organic Chemistry (1981), 11 (6), 539-46CODEN: SRIMCN; ISSN:0094-5714.V(O2CCF3)3 (I), VO(O2CCF3)2 (II), and VO2(O2CCF3) were prepd. by treating VCl3, VOCl2, and VOCl3, resp., with CF3CO2H. I and II were also prepd. by heating V(OAc)3 and VO(OAc)2 with CF3CO2H. The IR spectra, thermal degrdn., magnetic susceptibilities, and reflectance spectra of the compds. were detd.
- 31Lorenzotti, A.; Leonesi, D.; Cingolani, A.; Di Bernardo, P. Vanadyl(IV)─Acetate complexes in aqueous solution. J. Inorg. Nucl. Chem. 1981, 43, 737– 738, DOI: 10.1016/0022-1902(81)80213-831https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXktlWht78%253D&md5=d53355c3f23bd1243e57ac91ab797f5bVanadyl(IV)-acetate complexes in aqueous solutionLorenzotti, Adriana; Leonesi, Dante; Cingolani, Augusto; Di Bernardo, PlinioJournal of Inorganic and Nuclear Chemistry (1981), 43 (4), 737-8CODEN: JINCAO; ISSN:0022-1902.The formation of vanadyl(IV) complexes with acetates were studied by potentiometric titrn. methods; the stability consts. of mononuclear complexes, [VO(CH3COO)]+ and [VO(CH3COO)2], were calcd. to be 72.2 and 12.6 M-1, resp., in 1M aq. perchlorate at 25°. A comparison was made with the literature values for the corresponding neptunyl(IV) complexes.
- 32Di Bernardo, P.; Tomat, G.; Zanonato, P.; Portanova, R.; Tolazzi, M. Thermodynamics of vanadyl(IV)─carboxylate complex formation in aqueous solution. Inorg. Chim. Acta 1988, 145, 285– 288, DOI: 10.1016/s0020-1693(00)83971-732https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1cXltFOjur8%253D&md5=f08fc3a701e03e24092175d6e4861700Thermodynamics of vanadyl(IV)-carboxylate complex formation in aqueous solutionDi Bernardo, Plinio; Tomat, Giuliana; Zanonato, Pierluigi; Portanova, Roberto; Tolazzi, MarilenaInorganica Chimica Acta (1988), 145 (2), 285-8CODEN: ICHAA3; ISSN:0020-1693.The stability consts. and the heats of formation of vanadyl(IV)-acetate, -glycolate, and -glycine complexes were detd. in aq. soln. by potentiometric and calorimetric measurements. In the pH range where the protolytic equil. of VO2+ is negligible, acetate forms 2 mononuclear complexes, glycolate 3, while glycine reacts in its zwitterionic form. The stabilities of the glycolate complexes are considerably higher than the acetate complexes, in spite of its lower basicity, indicating that complex formation involves the coordination of the OH group. The enthalpy changes are pos. except for glycolate where a small neg. value is found. For all systems, the entropy changes are pos. and therefore favorable to the complex formation.
- 33Mehio, N.; Ivanov, A. S.; Ladshaw, A. P.; Dai, S.; Bryantsev, V. S. Theoretical Study of Oxovanadium(IV) Complexation with Formamidoximate: Implications for the Design of Uranyl-Selective Adsorbents. Ind. Eng. Chem. Res. 2016, 55, 4231– 4240, DOI: 10.1021/acs.iecr.5b0339833https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvVyit7jL&md5=ffaf2049955182832f1549baa63c4f99Theoretical Study of Oxovanadium(IV) Complexation with Formamidoximate: Implications for the Design of Uranyl-Selective AdsorbentsMehio, Nada; Ivanov, Alexander S.; Ladshaw, Austin P.; Dai, Sheng; Bryantsev, Vyacheslav S.Industrial & Engineering Chemistry Research (2016), 55 (15), 4231-4240CODEN: IECRED; ISSN:0888-5885. (American Chemical Society)Poly(acrylamidoxime) fibers are the current state-of-the-art adsorbent for mining uranium from seawater. However, the competition between uranyl (UO22+) and vanadium ions poses a challenge to mining on the industrial scale. In this work, we employ d. functional theory and coupled-cluster methods in the restricted formalism to investigate potential binding motifs of the oxovanadium(IV) ion (VO2+) with the formamidoximate ligand. Consistent with exptl. extended X-ray absorption fine structure data, the hydrated six-coordinate complex is predicted to be preferred over the hydrated five-coordinate complex. Our investigation of formamidoximate-VO2+ complexes universally identified the most stable binding motif formed by chelating a tautomerically rearranged imino hydroxylamine via the imino nitrogen and hydroxylamine oxygen. The alternative binding motifs for amidoxime chelation via a nonrearranged tautomer and η2 coordination are found to be ∼11 kcal/mol less stable. Natural bond orbital anal. was performed to understand the nature of the interactions in the VO2+ complexes. The difference in the most stable VO2+ and UO22+ binding conformation has important implications for the design of more selective UO22+ ligands.
- 34Winkler, J. R.; Gray, H. B. Electronic Structures of Oxo-Metal Ions. In Molecular Electronic Structures of Transition Metal Complexes I. Structure and Bonding; Mingos, D. M. P., Day, P., Dahl, J. P., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2012; vol. 142, pp 17– 28.There is no corresponding record for this reference.
- 35Krakowiak, J.; Lundberg, D.; Persson, I. A Coordination Chemistry Study of Hydrated and Solvated Cationic Vanadium Ions in Oxidation States +III, +IV, and +V in Solution and Solid State. Inorg. Chem. 2012, 51, 9598– 9609, DOI: 10.1021/ic300202f35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhtlSks7nL&md5=e69f2a1d2daac25be6f4a51cedebef0bA Coordination Chemistry Study of Hydrated and Solvated Cationic Vanadium Ions in Oxidation States +III, +IV, and +V in Solution and Solid StateKrakowiak, Joanna; Lundberg, Daniel; Persson, IngmarInorganic Chemistry (2012), 51 (18), 9598-9609CODEN: INOCAJ; ISSN:0020-1669. (American Chemical Society)The coordination chem. of hydrated and solvated vanadium(III), oxovanadium(IV), and dioxovanadium(V) ions in the oxygen-donor solvents water, DMSO , and N,N'-dimethylpropyleneurea (DMPU) was studied in soln. by extended x-ray absorption fine structure (EXAFS) and large-angle X-ray scattering (LAXS) and in the solid state by single-crystal X-ray diffraction and EXAFS. The hydrated vanadium(III) ion has a regular octahedral configuration with a mean V-O bond distance of 1.99 Å. In the hydrated and DMSO-solvated oxovanadium(IV) ions, vanadium binds strongly to an oxo group at ∼1.6 Å. The solvent mol. trans to the oxo group is very weakly bound, at ∼2.2 Å, while the remaining four solvent mols., with a mean V-O bond distance of 2.0 Å, form a plane slightly below the vanadium atom; the mean O=V-Operp bond angle is ∼98°. In the DMPU-solvated oxovanadium(IV) ion, the space-demanding properties of the DMPU mol. leave no solvent mol. in the trans position to the oxo group, which reduces the coordination no. to 5. The O=V-O bond angle is consequently much larger, 107°, and the mean V=O and V-O bond distances decrease to 1.58 and 1.97 Å, resp. The hydrated and DMSO-solvated dioxovanadium(V) ions display a very distorted octahedral configuration with the oxo groups in the cis position with a mean V=O bond distance of 1.6 Å and a O=V=O bond angle of ∼105°. The solvent mols. trans to the oxo groups are weakly bound, at ∼2.2 Å, while the remaining two have bond distances of 2.02 Å. The exptl. studies of the coordination chem. of hydrated and solvated vanadium(III,IV,V) ions are complemented by summarizing previously reported crystal structures to yield a comprehensive description of the coordination chem. of vanadium with oxygen-donor ligands.
- 36Rumble, J. R. CRC Handbook of Chemistry and Physics, Internet Edition 2018, 99th ed.; CRC Press/Taylor & Francis: Boca Raton, FL, 2018.There is no corresponding record for this reference.
- 37Guo, Y.; Surblys, D.; Matsubara, H.; Kawagoe, Y.; Ohara, T. Molecular Dynamics Study on the Effect of Long-Chain Surfactant Adsorption on Interfacial Heat Transfer between a Polymer Liquid and Silica Surface. J. Phys. Chem. C 2020, 124, 27558– 27570, DOI: 10.1021/acs.jpcc.0c0894037https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXisV2hu73K&md5=b84d8894327b9cc141f818c12772aef2Molecular dynamics study on effect of long-chain surfactant adsorption on interfacial heat transfer between polymer liquid and silica surfaceGuo, Yuting; Surblys, Donatas; Matsubara, Hiroki; Kawagoe, Yoshiaki; Ohara, TakuJournal of Physical Chemistry C (2020), 124 (50), 27558-27570CODEN: JPCCCK; ISSN:1932-7447. (American Chemical Society)The addn. of surfactants to polymer-based thermal interface materials applied to improve the heat dissipation efficiency of chip surfaces in contact has attracted attention for the microelectronic processing technol. In the present study, the mechanism by which a long-chain surfactant affects heat transfer across the interface between solid surface and polymer liq. was investigated by non-equil. mol. dynamics simulation. We constructed a system where tetracosane was used as a solvent and contained alc. mols. as a surfactant, and they were placed between two flat silica surfaces under a thermal gradient. The effect of the hydrophilicity of silica surface, the concn. of the surfactant, and chain length of the surfactant on silica-liq. interfacial thermal resistance Rb were examd. Alc. surfactant mols. preferred to adsorb onto the hydrophilic silica (Si-OH) surface due to hydrogen bonding between alc. and silanol hydroxyl groups. It was found that Rb reduced not only with the adsorption amt. of alc. mols. but also with the chain length of alc. The van der Waals interaction contribution was dominant for solid-liq. and liq.-liq. heat conduction near the interface. The hydroxyl terminals of alc. mols. were vertically adsorbed onto the Si-OH surface due to hydrogen bonds, which produced a heat path from silanols to the hydroxyl groups of alc. Furthermore, heat was also exchanged between alc. hydroxyl and alkyl groups via intramol. interaction and between the alc. alkyl groups and nearby solvent mols. via van der Waals (vdW) intermol. interaction. This resulted in an efficient heat path from solid surface silanols to liq. bulk. As the alc. chain length increased without changing the no. of adsorbed alc. mols., the heat transfer through this heat path increased, which led to a decrease in Rb. These results provided insight toward the guiding principle for the mol. design of complex surfactants to enhance the interfacial heat transfer.
- 38Uchaker, E.; Gu, M.; Zhou, N.; Li, Y.; Wang, C.; Cao, G. Enhanced Intercalation Dynamics and Stability of Engineered Micro/Nano-Structured Electrode Materials: Vanadium Oxide Mesocrystals. Small 2013, 9, 3880– 3886, DOI: 10.1002/smll.20120318738https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXntF2ksLk%253D&md5=6c44f9f3abed96d1ede41c768401b5a9Enhanced Intercalation Dynamics and Stability of Engineered Micro/Nano-Structured Electrode Materials: Vanadium Oxide MesocrystalsUchaker, Evan; Gu, Meng; Zhou, Nan; Li, Yanwei; Wang, Chongmin; Cao, GuozhongSmall (2013), 9 (22), 3880-3886CODEN: SMALBC; ISSN:1613-6810. (Wiley-VCH Verlag GmbH & Co. KGaA)An additive and template free process is developed for the facile synthesis of VO2(B) mesocrystals via the solvothermal reaction of oxalic acid and vanadium pentoxide. The six-armed star architectures are composed of stacked nanosheets homoepitaxially oriented along the [100] crystallog. register with respect to one another, as confirmed by means of selected area electron diffraction and electron microscopy. It is proposed that the mesocrystal formation mechanism proceeds through classical as well as non-classical crystn. processes, and is possibly facilitated or promoted by the presence of a reducing/chelating agent. The synthesized VO2(B) mesocrystals are tested as a cathodic electrode material for lithium-ion batteries, and show good capacity at discharge rates ranging from 150-1500 mA g-1 and a cyclic stability of 195 mA h g-1 over fifty cycles. The superb electrochem. performance of the VO2(B) mesocrystals is attributed to the porous and oriented superstructure that ensures large surface area for redox reaction and short diffusion distances. The mesocryst. structure ensures that all the surfaces are in intimate contact with the electrolyte, and that lithium-ion intercalation occurs uniformly throughout the entire electrode. The exposed (100) facets also lead to fast lithium intercalation, and the homoepitaxial stacking of nanosheets offers a strong inner-sheet binding force that leads to better accommodation of the strain induced during cycling, thus circumventing the capacity fading issues typically assocd. with VO2(B) electrodes.
- 39Zeininger, L.; Portilla, L.; Halik, M.; Hirsch, A. Quantitative Determination and Comparison of the Surface Binding of Phosphonic Acid, Carboxylic Acid, and Catechol Ligands on TiO2 Nanoparticles. Chem.─Eur. J. 2016, 22, 13506– 13512, DOI: 10.1002/chem.20160192039https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xht1WmtLbL&md5=9c49d50096a1908a791fdd7cc4bf7991Quantitative Determination and Comparison of the Surface Binding of Phosphonic Acid, Carboxylic Acid, and Catechol Ligands on TiO2 NanoparticlesZeininger, Lukas; Portilla, Luis; Halik, Marcus; Hirsch, AndreasChemistry - A European Journal (2016), 22 (38), 13506-13512CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)The adsorption, desorption, co-adsorption, and exchange of phosphonic acid, carboxylic acid, and catechol derivs. on the surface of titanium oxide (anatase) nanoparticles are investigated. Thermogravimetric anal. provides a facile and fast-track quant. detn. of the wet-chem. monolayer adsorption consts. and grafting densities of ten adsorbates, all under neutral pH conditions. This characterization protocol allows straightforward quantification of the relevant thermodn. data of ligand adsorption and a comparison of ligand adsorption strengths. The reported procedure is proposed as a universal tool and it should be applicable to many other colloidal metal oxide materials. Moreover, the detd. values for the adsorption consts. and the monolayer grafting densities provide a toolbox for the assessment of the adsorbates' behavior in desorption, exchange, and co-adsorption equil. This versatile evaluation procedure will help to identify optimal monolayer-surface combinations and to evaluate crit. parameters, such as monolayer robustness, ligand exchange rates, or targeted mixed assembly of functionalities.
- 40De Roo, J.; Justo, Y.; De Keukeleere, K.; Van den Broeck, F.; Martins, J. C.; Van Driessche, I.; Hens, Z. Carboxylic-Acid-Passivated Metal Oxide Nanocrystals: Ligand Exchange Characteristics of a New Binding Motif. Angew. Chem., Int. Ed. 2015, 54, 6488– 6491, DOI: 10.1002/anie.20150096540https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXmtlGqsr4%253D&md5=a55ec98a8761a97e2f1900720b8f7aa5Carboxylic-Acid-Passivated Metal Oxide Nanocrystals: Ligand Exchange Characteristics of a New Binding MotifDe Roo, Jonathan; Justo, Yolanda; De Keukeleere, Katrien; Van den Broeck, Freya; Martins, Jose C.; Van Driessche, Isabel; Hens, ZegerAngewandte Chemie, International Edition (2015), 54 (22), 6488-6491CODEN: ACIEF5; ISSN:1433-7851. (Wiley-VCH Verlag GmbH & Co. KGaA)Studying sterically stabilized HfO2 and ZrO2 NCs using 1H soln. NMR and IR spectroscopy as well as elemental anal., this paper demonstrates the reversible exchange of initial oleic acid ligands for octylamine and self-adsorption of oleic acid at NC surfaces. Both processes are incompatible with an X-type binding motif of carboxylic acids as reported for sulfide and selenide NCs. This behavior stems from the dissociative adsorption of carboxylic acids at the oxide surface. Both proton and carboxylate moieties must be regarded as X-type ligands yielding a combined X2 binding motif that allows for self-adsorption and exchange for L-type ligands.
- 41Calatayud, D. G.; Rodríguez, M.; Jardiel, T. Controlling the morphology of TiO2 nanocrystals with different capping agents. Bol. Soc. Esp. Ceram. Vidrio 2015, 54, 159– 165, DOI: 10.1016/j.bsecv.2015.07.00141https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1Snu77O&md5=691f7e4efcd1878d15164aa9c21d983dControlling the morphology of TiO2 nanocrystals with different capping agentsCalatayud, David G.; Rodriguez, Monica; Jardiel, TeresaBoletin de la Sociedad Espanola de Ceramica y Vidrio (2015), 54 (4), 159-165CODEN: BSCVB9; ISSN:0366-3175. (Sociedad Espanola de Ceramica y Vidrio)This paper provides direct evidence to support the role of capping agents in controlling the evolution of TiO2 seeds into nanocrystals with a specific shape. Starting with Ti(OBut)4 and using oleid acid, oleylamine, dioleamide, 11-aminoundecanoic acid, arginine, trifluroacetic acid or HF as capping agents, mainly TiO2 truncated octahedrons enclosed by {1 0 1} and {0 0 1} facets were obtained. We could also selectively obtain square, rods and rounded rhombic-shaped nanoparticles by growing of {0 1 0} facets by adding oleic acid and oleylamine in ratio 6:4, resp., while all other parameters were kept the same. This research not only offers new insights into the role played by a capping agent in shape-controlled synthesis but also provides, a versatile approach to controlling the shape of metal oxide nanocrystals.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnanoscienceau.3c00014.
Tables reporting the amounts of all reagents used in each reaction reported here, powder X-ray diffraction spectra, TEM, and SEM images of nanocrystalline products, and details of the characterization of VO(benzoate)2 and VO(4-nitrobenzoate)2 by single-crystal electron diffraction (PDF)
Crystallographic data for VO(benozate)2 (CIF)
Crystallographic data for VO(4-nitrobenzoate)2 (CIF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.