Early stage growth of rutile titania mesocrystals

two synthetic strategies to prepare rutile TiO2 mesocrystals: 1) ... Page 1 of 7. ACS Paragon Plus Environment. Crystal Growth & Design. 1. 2. 3. 4. 5...
1 downloads 0 Views 819KB Size
Subscriber access provided by Technical University of Munich University Library

Communication

Early stage growth of rutile titania mesocrystals Hanglong Wu, Yueke Yang, Yang Ou, Bin Lu, Jun Li, Wentao Yuan, Yong Wang, and Ze Zhang Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.8b00028 • Publication Date (Web): 03 Jul 2018 Downloaded from http://pubs.acs.org on July 7, 2018

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

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

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

Crystal Growth & Design

Early stage growth of rutile titania mesocrystals Hanglong Wu‡†, Yueke Yang‡, Yang Ou, Bin Lu, Jun Li, Wentao Yuan, Yong Wang*, Ze Zhang Center of Electron Microscopy and State Key Laboratory of Silicon Materials, School of Materials Science and Engineering, Zhejiang University, Hangzhou 310027, China.

Supporting Information Placeholder ABSTRACT: Exploring the crystallization process of metal oxide mesocrystals has attracted enormous recent attention. However, due to the lack of insight into the behaviors of short-lived species at initial growth stages, there is currently an understanding gap in the underlying growth mechanism. Here, a combination of a pre-seeded hydrothermal method and transmission electron microscopy allows us to witness the rapid crystallization by particle attachement (CPA) in the early growth stage of capsuleshaped rutile titania mesocrystals. The presence of the embryonic form of nanocapsules, the slight misalignment of the primary particles and most importantly, the atomic interface between primary particles during attachment strongly indicate the existence of CPA mechanism. Furthermore, we rationalize our findings in terms of the free energy landscapes that govern nonclassical formation path-ways. Our study provides a practical approach to explore the formation mechanism of fast-growing crystals at their initial growth stages.

An in-depth understanding of crystallization offers guidance for structure and property control during material synthesis1-10. The crystal formation in solution are usually described by classical nucleation theory (CNT)11 and the terrace-ledge-kink model of crystal growth12, 13, which are both based on monomer-bymonomer addition of atoms or molecules. In recent years, many alternative pathways to CNT have been shown including aggregational processes of nanomeric building blocks such as prenucleation crystalline14 and amorphous nanoparticles15, 16. Such crystallization mechanisms could not be explained by CNT and the classical growth theory, and have been termed non-classical crystallization or crystallization by particle attachment (CPA)17. CPA is known to be a prevalent growth mechanism at the early stage of crystal growth, particularly in mesocrystals13, 18-22. Yet it is difficult to study the pre-formed primary precursors due to the short life of those mesocrystal intermediates13, and whether or not these precursors undergo CPA is still under debate in some material systems17, for example, rutile TiO25, 23. Rutile TiO2 mesocrystals with desired and tunable architectures are one of the promising candidates for a wide range of applications, such as photocatalysis24 and batteries25, 26, due to their biological and chemical inertness, structural stability during Li insertion/extraction27-29. To date, various morphologies have already been reported in this filed, including nanorods25, 30, microspheres27, 31 and wulff-shaped nanoparticles32, 33. There are mainly two synthetic strategies to prepare rutile TiO2 mesocrystals: 1) phase transformation from precursor hydrogen titanate to rutile TiO225, 26; 2) hydrolysis of Ti-containing precursor at low temper-

ature for long reaction time assisted by organic or biologic molecules27, 34-37. Both methods are under acidic conditions and allow us to fabricate rutile mesocrystals with nanorod subunits. However, there still exists a huge understanding gap in the underlying rutile titania crystallization mechanisms. For example, most studies that discuss crystallization through CPA, infer the mechanism from observed morphologies and defects in the final structures30, 36-38. In fact, such features can be misleading and they alone can’t prove the particle-based growth process17. Dislocations and ‘finger-like structures’ in rutile TiO2 appear to be ubiquitous under lower temperature conditions39-42, which can also form through classical Oswald ripening (OR) mechanism due to Coulombic and steric hindrance39. Lan et al. were able to identify rod-like subunits to TiO2 mesocrystals by sampling a long (24 h) hydrothermal synthesis at various times35. However, this approach can only be applied to CPA processes which undergo slow conversion and there currently has been no information reported on the atomic interface between primary rutile nanoparticles in such fastgrowing systems which are necessary to fully understand and confirm a CPA mechanism.

Figure 1. (a) and (b) FE-SEM images of the products synthesized at 180 oC for 75 min before calcination. (c) High resolution SEM images of an individual rutile TiO2 nanocapsule. (d) XRD pattern of the as-synthesized rutile TiO2 nanocapsules. In our work, we use a pre-seeded treatment with the combination of transmission electron microscopy (TEM) to explore the growth mechanism of a fast growing rutile TiO2 nanocapsules at the initial stage. We prove that the mesocrystals form by the oriented self-assembly of mesoporous nanorod clusters. More specifically, those particle aggregation traces such as the presence of embryonic form of the nanocapsules, the slight misalignment of the prima-

ACS Paragon Plus Environment

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

ry particles and particularly the atomic interface during attachment provide strong evidence of CPA. The synthesis of rutile TiO2 nanocapsules was achieved by a modified hydrothermal treatment as reported by Wu et al.,43 in a quinary solution system consisting of titanium butoxide (TBOT), water, hydrochloric acid (HCl), normal butanol (CH3(CH2)3OH) and hydrofluoric acid (HF). Figure 1a displays a lowmagnification scanning electron microscopy (SEM) image of the products after heating to 180 oC for 75 min before calcination. The image shows large-scale formation of uniform nanocapsules predominantly ca. 1.2 μm in length and ca. 400 nm in width. The product were calcinated to remove residual organics, after which they displayed a similar morphology and size (Figure S1). However, the surfaces are not smooth and many porous-looking “holes” are present. The related X-ray diffraction (XRD) patterns shown in Figure 1d confirm all the products are tetragonalstructured rutile TiO2 (JCPDS No. 21-1276), although the weak characteristic peaks indicate the poor crystalline nature of these rutile nanocapsules. Moreover, as no additional diffraction peaks are observed, the products appear to be of high crystal purity. In our last report43, we obtained rutile crystals with curved surfaces with 93.8% rutile and 6.2 % anatase. Here we provide a solution

to obtain the pure rutile phase by simply increasing the precursor concentration. It is also worthwhile to note that the crystallinity can be greatly enhanced after calcination and the sharp diffraction peaks in Figure S1 also suggest the preferred [001] growth direction.

Figure 2. TEM micrograph of the rutile TiO2 capsule. (b) TEM micrograph of the image area in (a). (c) HR-TEM micrographs of the image area in (b). Inset: a FFT micrograph of (c).

Figure 3. (a) FE-SEM images of the products synthesized at 180 oC for: (a) 50 min, (b) 60 min and (c) 75 min with the pre-seeded treatment. TEM (d) and HRTEM (e) micrographs of TiO2 nanoparticles at 50 min. Insets in (b) and (e) exhibit a freestanding particle and a FFT micrograph from the image area in (e) respectively. (f) An inversed fast Fourier transform (IFFT) micrograph of the left part of (e). Figure 2a-c show TEM and HRTEM images of the TiO2 capsules at 180 oC for 75 min. It is obvious that the TiO2 capsules grow along the preferred [001] direction (Figure 2a). The selected area electron diffraction (SAED) pattern inset corresponding to [1 0] zone axis of the rutile phase indicates the single-crystal nature of the TiO2 particles, although slightly elongated diffraction spots suggest small misorientation exists deviating from perfect alignment among nanorods.4, 13, 39 This may be a result of lattice orientation imperfections among the nanorods during crystal growth. Furthermore, Figure 2b-c shows the tip of a single TiO2 nanocapsule consists of numerous nanorods growing preferentially along the [001] direction. The fast Fourier transform (FFT) pattern of the tiny nanorods at the tip demonstrates the poor crystallinity of them, which is accordance with the result of XRD.

All the analyses above indicate the as-grown nanocapsules could be mesocrystals, but to confirm the CPA mechanism, more aggregation traces should be provided particularly at early growth stages. Therefore, we developed a pre-seeded synthetic strategy to monitor the whole growth process: The supernatant after the first hydrothermal treatment was collected and heated to 180 oC for

ACS Paragon Plus Environment

Page 2 of 7

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

Crystal Growth & Design

Figure 4. HRTEM images of the 20-nm-in-size primary nanoparticles. different reaction times. As a result of the "seeds" in the solution, we could collect the products prior to nanocapsules. It is found that large amounts of mesoporous rutile nanocrystals (less than 100 nm in length in the majority), first appear after 50 minutes (Figure 3a). After 60 minutes TiO2 capsules were observed ranging from 500-800 nm in length (Figure 3b), in the presence of the nanorod clusters. In comparison, in the control experiment without the pre-seeded process, we couldn’t collect anything even after centrifugations at early growth stages (< 1 h), although the capsule-shaped mesocrystals appeared at 1 h (Figure S2). Thus, the pre-seeded treatment enables us to directly acquire the shortlived primary particles at early growth stages. Intriguingly, after 1 hour and 15 minutes, we obtained nanocapsules with a similar size (Figure 3c) as those without pre-seeded treatment after the same time, but the yields were several times larger. Prolonging

the reaction duration resulted in the nanocapsules branching in the middle section and finally forming ca. 8.5-μm-in-diameter microspheres at 3 h (Figure S3). Representative images in Figure 3d-f show the TEM and HRTEM images of nanoparticles prepared with a reaction time of 50 min. As is illustrated in Figure 3d, the length of these nanoparticles is typically tens of nanometers and Figure 3e shows the lattice fringes whose spacings are consistent with rutile. The FFT inset can be indexed to the [1 1] zone axis of rutile TiO2. More importantly, according to the HRTEM analysis, these nanoparticles are nanorod clusters, i.e. nanorod mesocrystals, which seem to be assembled in the same crystallographic [001] direction. The white dotted lines in the IFFT micrograph of Figure 3f displays the outline of two freestanding nanorods at the atomic scale. The slighted elongated diffraction spots in the FFT also indicate the misorientation in the growth process during oriented aggregation. Penn et al. reported similar nanorod structures and claimed that the mesoporous rutile crystals were formed by oriented attachment of anatase nanoparticles under low particle solubility conditions via Cryo-TEM analysis44. However, in our case, there is no evidence of the anatase phase. In fact some ca. 20-nm-in-size rutile nanoparticles were found among the nanorod clusters (Figure 4), and the agglomerate of such primary particles (Figure S4) indicates that these nanoparticles could evolve into nanorod clusters via CPA mechanism, due to the presence of defects inside these nanoparticles44, 45. This finding is different from previous studies that show rutile crystals transformed from anatase or other intermediates44, 46, particular in the existence of HF, as H-F bonds are well known to stabilize the anatase phase.

Figure 5. (a) TEM image of a large nanorod cluster agglomerate surrounded by small ones. (b) Electron diffraction pattern obtained from the yellow dashed circular area in (a). (c)FFT pattern recorded from the red dashed rectangular area. (d) HRTEM image of the interface between particle 1 and 3. Inset I and II: enlarged TEM images of adjacent particle 2 and 3 in (a), respectively. Inset III and IV: FFT patterns of particle 3 and 1. The length of the initial rutile TiO2 capsules was at least 500 nm, at 60 min. The absence of smaller size nanocapsules could be explained in two ways: 1) OR: the nanorod clusters dissolve because of thermodynamic instability, and the large capsules grow quickly by consuming the dissolved species in solution. 2) CPA: a rapid oriented self-aggregation of primary clusters within 10 minutes due to a lower activation energy barrier. Although Oswald ripening and CPA can exist simultaneously and compete with each other, in our case, we believe CPA dominates the crystallization in the early stages. We infer this from the aforementioned poor crystallinity of these TiO2 mesocrystals before calci-

nation and more importantly the presence of intermediate agglomerates among the nanorod clusters and the primary particle interfaces at atomic scale. This is illustrated in Figure 5 and Figure S5. Figure 5a shows a large aggregate, i.e., particle 1 (ca. 650 nm in length) surrounded by many adjacent comparatively smaller nanorod clusters, including particle 2 and 3, that share similar crystallographic orientation with respective to the particle 1. This suggests that they might have undergone continuous rotation, interaction, and self-adjusted their orientations14, 47. The SAED pattern (Figure 5b) confirms that the aggregate is single-crystallike, as the spots are elongated. That is to say, this aggre

ACS Paragon Plus Environment

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

Figure 6. Illustration of the formation mechanism of the capsule-shaped rutile TiO2 mesocrystals. gate is already in an early form of the capsule-shaped mesocrystals. Furthermore, insets I and II clearly depict the interfaces between the neighboring primary nanoparticles (particle 2 and 3) and the large particle (particle 1), indicating that these nanoparticles loosely attach with the large one. It appears that more time is needed to adjust their orientation and fuse. The interface is determined by the IFFT patterns recorded from the corresponding particles. Part of the enlarged interfacial area at atomic resolution shows the grain boundary, delineated by a black dashed line (inset II Figure 5d). The FFT patterns (inset III and IV) demonstrates the single crystalline nature of both particle 2 and 3, and that they are both aligned along the [010] zone axis regardless of the slight misalignment of 3.5˚ shown in the FFT pattern (Figure 5c). The small mismatch angle here is also a strong evidence of oriented aggregation behaviors. Furthermore, a slight misalignment at the time of attachment was reported to lead to defect formation at the interface, and the defects would be translated laterally across the interface eliminating any trace of the interparticle boundary according to Li's observation14. In this sense, the small particles such as particle 3 in the vicinity of larger one will eventually fuse across a mismatched interface after post-attachment behaviors. Interestingly, in addition to those nanorod clusters, we also found that those ca. 20 nm primary nanoparticles can also have the possibility to directly attach to the large mesocrystal and fuse (Figure S5). However, as only several primary particles were observed among the nanorod clusters, more work need to be done to better understand the behaviors of these metastable intermediates before they evolve into nanorod clusters. According to the above observations, we speculate that these capsule-shaped rutile TiO2 mesocrystals are not formed via classical crystallization, but undergoing a non-classical CPA mechanism. The formation mechanism is illustrated in Figure 6. First, mesoporous rutile nanorod clusters appear as primary particles which rapidly aggregate to become nearly aligned agglomerates through Brownian motion14 and form the embryonic capsuleshaped rutile mesocrystals in 10 mins. At the same time, Ostwald ripening (OR) occurs where the agglomerate surfaces recrystallize and fuse, and some remaining clusters undergo dissolution and reprecipitation on the mesocrystal until they are fully exhausted. The key question in this mechanism is why the primary clusters choose CPA rather than OR in the early growth stages. This can be understood by considering the interplay of free-energy landscapes and reaction dynamics17. In the view of dynamics, it is known that a transition from monomer-based to particle based pathway occurs when the supersaturation become high enough and the free energy barrier is relatively small and comparable with ( : Boltzmann constant, : temperathe thermal energy, ture). In this case as the Ti precursor concentration is much higher than our previous report43, it is logical to assume it is related to increased supersaturation. Furthermore, when the free-energy landscape presents local minima, we can expect to observe as-

sembly pathway involving thermodynamically metastable intermediates as dissolution back to individual complexes is unfavorable. In our case this would indicate that the primary clusters are metastable species. However, due to their high surface area and high supersaturation, rapid CPA becomes favored. In summary, we unravel the CPA mechanism of the TiO2 mesocrystal nanocapsules in the early growth stage with a combination of a simple pre-seeded synthetic strategy and TEM. The key experimental indicators of CPA are the presence of the early form of the nanocapsules, the slight misalignment of the primary particles and particularly the atomic interface during attachment. Furthermore we rationalize our findings in terms of the free energy landscapes that govern non-classical formation pathways. Moreover, our study provides a practical approach to explore the formation mechanism of fast-growing crystals at their initial growth stages.

Supporting Information is available Supporting Information Available: Detailed synthesis method and characterization data (other SEM, TEM images and XRD patterns of as-prepared products) (PDF). This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]

ORCID Hanglong Wu: 0000-0002-8042-9952

Present Addresses †Laboratory of Materials and Interface Chemistry and Center of Multiscale Electron Microscopy, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P.O. Box 513, 5600 MB, Eindhoven, The Netherlands

Author Contributions ‡These authors contributed equally.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT We acknowledge the support of National Science Foundation of China (51390474, 91645103, 11234011, 11327901). The authors thank Dr. Chenghua Sun of Swinburne University of Technology and Dr. Lingqing Dong from Zhejiang University for helpful discussions.

ACS Paragon Plus Environment

Page 4 of 7

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

Crystal Growth & Design

REFERENCES (1) Zhuang, Z.; Huang, F.; Lin, Z.; Zhang, H., AggregationInduced Fast Crystal Growth of SnO2 Nanocrystals. J. Am. Chem. Soc. 2012, 134, 16228-16234. (2) Puntes, V. F.; Krishnan, K. M.; Alivisatos, A. P., Colloidal nanocrystal shape and size control: The case of cobalt. Science 2001, 291, 2115-2117. (3) Peng, X. G.; Wickham, J.; Alivisatos, A. P., Kinetics of II-VI and III-V colloidal semiconductor nanocrystal growth: "Focusing" of size distributions. J. Am. Chem. Soc. 1998, 120, 5343-5344. (4) Ye, J.; Liu, W.; Cai, J.; Chen, S.; Zhao, X.; Zhou, H.; Qi, L., Nanoporous Anatase TiO2 Mesocrystals: Additive-Free Synthesis, Remarkable Crystalline-Phase Stability, and Improved Lithium Insertion Behavior. J. Am. Chem. Soc. 2011, 133, 933940. (5) Li, D.; Soberanis, F.; Fu, J.; Hou, W.; Wu, J.; Kisailus, D., Growth mechanism of highly branched titanium dioxide nanowires via oriented attachment. Cryst. Growth Des. 2013, 13, 422-428. (6) Kashyap, S.; Woehl, T. J.; Liu, X. P.; Mallapragada, S. K.; Prozorov, T., Nucleation of Iron Oxide Nanoparticles Mediated by Mms6 Protein in Situ. Acs Nano 2014, 8, 9097-9106. (7) Yang, S.; Yang, B. X.; Wu, L.; Li, Y. H.; Liu, P.; Zhao, H.; Yu, Y. Y.; Gong, X. Q.; Yang, H. G., Titania single crystals with a curved surface. Nat. Commun. 2014, 5, 5355. (8) Wang, J. Y.; Yang, N. L.; Tang, H. J.; Dong, Z. H.; Jin, Q.; Yang, M.; Kisailus, D.; Zhao, H. J.; Tang, Z. Y.; Wang, D., Accurate Control of Multishelled Co3O4 Hollow Microspheres as High-Performance Anode Materials in Lithium-Ion Batteries. Angew. Chem., Int. Ed. 2013, 52, 6417-6420. (9) Fang, J. X.; Zhang, L. L.; Li, J.; Lu, L.; Ma, C. S.; Cheng, S. D.; Li, Z. Y.; Xiong, Q. H.; You, H. J., A general softenveloping strategy in the templating synthesis of mesoporous metal nanostructures. Nat. Commun. 2018, 9. (10) Zhang, L. L.; Li, J.; You, H. J.; Ma, C. S.; Lan, S.; Wu, Z. D.; Zeng, J. R.; Tian, F.; Fang, J. X., In Situ Probing of the Particle-Mediated Mechanism of WO3-Networked Structures Grown inside Confined Mesoporous Channels. Small 2018, 14. (11) Kashchiev, D., Thermodynamically consistent description of the work to form a nucleus of any size. J. Chem. Phys. 2003, 118, 1837-1851. (12) Burton, W. K.; Cabrera, N.; Frank, F. C., The growth of crystals and the equilibrium structure of their surfaces. Philos. Trans. R. Soc., A 1951, 243, 299-358. (13) Cölfen, H.; Antonietti, M., Mesocrystals and nonclassical crystallization. ed.; John Wiley & Sons: 2008. (14) Li, D.; Nielsen, M. H.; Lee, J. R. I.; Frandsen, C.; Banfield, J. F.; De Yoreo, J. J., Direction-Specific Interactions Control Crystal Growth by Oriented Attachment. Science 2012, 336, 1014-1018. (15) Beniash, E.; Metzler, R. A.; Lam, R. S. K.; Gilbert, P., Transient amorphous calcium phosphate in forming enamel. J. Struct. Biol. 2009, 166, 133-143. (16) Pouget, E. M.; Bomans, P. H.; Goos, J. A.; Frederik, P. M.; Sommerdijk, N. A., The initial stages of template-controlled CaCO3 formation revealed by cryo-TEM. Science 2009, 323, 1455-1458. (17) De Yoreo, J. J.; Gilbert, P.; Sommerdijk, N.; Penn, R. L.; Whitelam, S.; Joester, D.; Zhang, H. Z.; Rimer, J. D.; Navrotsky, A.; Banfield, J. F.; Wallace, A. F.; Michel, F. M.; Meldrum, F. C.; Colfen, H.; Dove, P. M., Crystallization by particle attachment in synthetic, biogenic, and geologic environments. Science 2015, 349. (18) Cölfen, H.; Antonietti, M., Mesocrystals: Inorganic superstructures made by highly parallel crystallization and

controlled alignment. Angew. Chem., Int. Ed. 2005, 44, 55765591. (19) Fang, J. X.; Leufke, P. M.; Kruk, R.; Wang, D.; Scherer, T.; Hahn, H., External electric field driven 3D ordering architecture of silver (I) oxide meso-superstructures. Nano Today 2010, 5, 175-182. (20) Gebauer, D.; Cölfen, H., Prenucleation clusters and non-classical nucleation. Nano Today 2011, 6, 564-584. (21) You, H. J.; Fang, J. X., Particle-mediated nucleation and growth of solution-synthesized metal nanocrystals: A new story beyond the LaMer curve. Nano Today 2016, 11, 145-167. (22) Fang, J. X.; Ding, B. J.; Gleiter, H., Mesocrystals: Syntheses in metals and applications. Chem. Soc. Rev. 2011, 40, 5347-5360. (23) Jia, B.; Gao, L., Growth of well-defined cubic hematite single crystals: Oriented aggregation and Ostwald ripening. Cryst. Growth Des. 2008, 8, 1372-1376. (24) Aoyama, Y.; Oaki, Y.; Ise, R.; Imai, H., Mesocrystal nanosheet of rutile TiO2 and its reaction selectivity as a photocatalyst. CrystEngComm 2012, 14, 1405-1411. (25) Hong, Z.; Wei, M.; Lan, T.; Jiang, L.; Cao, G., Additive-free synthesis of unique TiO2 mesocrystals with enhanced lithium-ion intercalation properties. Energy Environ. Sci. 2012, 5, 5408-5413. (26) Hong, Z. S.; Wei, M. D.; Lan, T. B.; Cao, G. Z., Selfassembled nanoporous rutile TiO2 mesocrystals with tunable morphologies for high rate lithium-ion batteries. Nano Energy 2012, 1, 466-471. (27) Liu, S.-J.; Gong, J.-Y.; Hu, B.; Yu, S.-H., Mesocrystals of Rutile TiO2: Mesoscale Transformation, Crystallization, and Growth by a Biologic Molecules-Assisted Hydrothermal Process. Cryst. Growth Des. 2009, 9, 203-209. (28) Hoffmann, M. R.; Martin, S. T.; Choi, W.; Bahnemann, D. W., Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69-96. (29) Koudriachova, M. V.; Harrison, N. M.; de Leeuw, S. W., Diffusion of Li-ions in rutile. An ab initio study. Solid State Ionics 2003, 157, 35-38. (30) Cai, J.; Ye, J.; Chen, S.; Zhao, X.; Zhang, D.; Chen, S.; Ma, Y.; Jin, S.; Qi, L., Self-cleaning, broadband and quasiomnidirectional antireflective structures based on mesocrystalline rutile TiO2 nanorod arrays. Energy Environ. Sci. 2012, 5, 75757581. (31) Chen, F.; Cao, F.; Li, H.; Bian, Z., Exploring the important role of nanocrystals orientation in TiO2 superstructure on photocatalytic performances. Langmuir 2015, 31, 3494-9. (32) Hong, Z.; Wei, M.; Lan, T.; Cao, G., Self-assembled nanoporous rutile TiO2 mesocrystals with tunable morphologies for high rate lithium-ion batteries. Nano Energy 2012, 1, 466-471. (33) Da Silva, R. O.; Goncalves, R. H.; Stroppa, D. G.; Ramirez, A. J.; Leite, E. R., Synthesis of recrystallized anatase TiO2 mesocrystals with Wulff shape assisted by oriented attachment. Nanoscale 2011, 3, 1910-1916. (34) Lan, T. B.; Zhang, W. F.; Wu, N. L.; Wei, M. D., NbDoped Rutile TiO2 Mesocrystals with Enhanced Lithium Storage Properties for Lithium Ion Battery. Chem. - Eur. J. 2017, 23, 5059-5065. (35) Lan, T. B.; Wang, T.; Zhang, W. F.; Wu, N. L.; Wei, M. D., Rutile TiO2 mesocrystals with tunable subunits as a long-term cycling performance anode for sodium-ion batteries. J. Alloys Compd. 2017, 699, 455-462. (36) Lan, T. B.; Qiu, H. Y.; Xie, F. Y.; Yang, J.; Wei, M. D., Rutile TiO2 Mesocrystals/Reduced Graphene Oxide with HighRate and Long-Term Performance for Lithium-Ion Batteries. Sci. Rep. 2015, 5. (37) Hong, Z. S.; Zhou, K. D.; Zhang, J. W.; Huang, Z. G.; Wei, M. D., Facile synthesis of rutile TiO2 mesocrystals with

ACS Paragon Plus Environment

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

enhanced sodium storage properties. J. Mater. Chem. A 2015, 3, 17412-17416. (38) Zhang, D.; Li, G.; Wang, F.; Jimmy, C. Y., Green synthesis of a self-assembled rutile mesocrystalline photocatalyst. CrystEngComm 2010, 12, 1759-1763. (39) Wisnet, A.; Betzler, S. B.; Zucker, R. V.; Dorman, J. A.; Wagatha, P.; Matich, S.; Okunishi, E.; Schmidt-Mende, L.; Scheu, C., Model for Hydrothermal Growth of Rutile Wires and the Associated Development of Defect Structures. Cryst. Growth Des. 2014, 14, 4658-4663. (40) Wisnet, A.; Bader, K.; Betzler, S. B.; Handloser, M.; Ehrenreich, P.; Pfadler, T.; Weickert, J.; Hartschuh, A.; SchmidtMende, L.; Scheu, C.; Dorman, J. A., Defeating Loss Mechanisms in 1D TiO2-Based Hybrid Solar Cells. Adv. Funct. Mater. 2015, 25, 2601-2608. (41) Liu, L.; Qian, J.; Li, B.; Cui, Y.; Zhou, X.; Guo, X.; Ding, W., Fabrication of rutile TiO2 tapered nanotubes with rectangular cross-sections via anisotropic corrosion route. Chem. Commun. 2010, 46, 2402-2404. (42) Cha, S. I.; Hwang, K. H.; Kim, Y. H.; Yun, M. J.; Seo, S. H.; Shin, Y. J.; Moon, J. H.; Lee, D. Y., Crystal splitting and

enhanced photocatalytic behavior of TiO2 rutile nano-belts induced by dislocations. Nanoscale 2013, 5, 753-758. (43) Wu, H.; Li, H.; Li, J.; Lu, B.; Yang, Y.; Yuan, W.; Wang, Y.; Zhang, Z., Controllable synthesis of rutile titania with novel curved surfaces. CrystEngComm 2015, 17, 7254-7257. (44) Sabyrov, K.; Yuwono, V. M.; Penn, R. L. In Synthesis of Nanoporous Rutile Nanocrystals under Mild Conditions, MRS Proceedings, 2015; Cambridge Univ Press: 2015; pp mrsf141721-e02-06. (45) Ribeiro, C.; Vila, C.; Stroppa, D. B.; Mastelaro, V. R.; Bettini, J.; Longo, E.; Leite, E. R., Anisotropic growth of oxide nanocrystals: insights into the rutile TiO2 phase. J. Phys. Chem. C 2007, 111, 5871-5875. (46) Zhu, S. C.; Xie, S. H.; Liu, Z. P., Nature of Rutile Nuclei in Anatase-to-Rutile Phase Transition. J. Am. Chem. Soc. 2015, 137, 11532-11539. (47) Zhan, H.; Chen, Z.-G.; Zhuang, J.; Yang, X.; Wu, Q.; Jiang, X.; Liang, C.; Wu, M.; Zou, J., Correlation between Multiple Growth Stages and Photocatalysis of SrTiO3 Nanocrystals. J. Phys. Chem. C 2015, 119, 3530-3537.

ACS Paragon Plus Environment

Page 6 of 7

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

Crystal Growth & Design

For Table of Contents Use Only

Title: Early stage growth of rutile titania mesocrystals

Author list: Hanglong Wu‡†, Yueke Yang‡, Yang Ou, Bin Lu, Jun Li, Wentao Yuan, Yong Wang*, Ze Zhang

TOC graphic

Synopsis

A combination of a pre-seeded synthetic strategy and TEM revealed the rapid crystallization by particle attachement (CPA) in the early growth stage of rutile titania mesocrystals. The key experimental indicators of CPA are the presence of the early form of the nanocapsules, the slight misalignment of the primary particles and particularly the atomic interface during attachment.

7 ACS Paragon Plus Environment