Subscriber access provided by Kaohsiung Medical University
Energy, Environmental, and Catalysis Applications
Shape-Controlled Synthesis of Metal-Organic Frameworks with Adjustable Fenton-like Catalytic Activity Jiayi Liu, Xuning Li, Biao Liu, Chunxiao Zhao, Zhichong Kuang, Ruisheng Hu, Bin Liu, Zhimin Ao, and Junhu Wang ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b12686 • Publication Date (Web): 12 Oct 2018 Downloaded from http://pubs.acs.org on October 12, 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 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Shape-Controlled Synthesis of Metal-Organic Frameworks with Adjustable Fenton-like Catalytic Activity Jiayi Liua,e†, Xuning Lia,c†, Biao Liub†, Chunxiao Zhaoa,d, Zhichong Kuanga,e, Ruisheng Hud, Bin Liuc,*, Zhimin Aob,*, and Junhu Wanga,* aMössbauer
Effect Data Center, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China Key Laboratory of Environmental Catalysis and Pollution Control, Institute of Environmental Health and Pollution Control, School of Environmental Science and Engineering, Guangdong University of Technology, Guangzhou 510006, China cSchool of Chemical and Biomedical Engineering, Nanyang Technological University, 62 Nanyang Drive, Singapore 637459, Singapore dCollege of Chemistry and Chemical Engineering, Inner Mongolia University, Hohhot 010021, China eUniversity of Chinese Academy of Sciences, Beijing 100049, China †These authors contributed equally to this work. Supporting Information Placeholder bGuangzhou
ABSTRACT: Controllable synthesis of metal−organic frameworks (MOFs) with well-defined morphology, composition and size is of great importance towards understanding their structure-property relationship in various applications. Herein, we demonstrate a general strategy to modulate the relative growth rate of the secondary building units (SBUs) along different crystal facets for the synthesis of Fe-Co, Mn0.5Fe0.5-Co and Mn-Co Prussian blue analogues (PBAs) with tunable morphologies. The same growth rate of SBUs along the {100}, {110} and {111} surface at 0 ºC results in the formation of spherical PBA particles, while the lowest growth rate of SBUs along the {100} surface resulting from the highest surface energy with increasing reaction temperature induces the formation of PBA cubes. Fenton reaction was used as the model reaction to probe the structurecatalytic activity relation for the as-synthesized catalysts. The cubic Fe-Co PBA was found to exhibit the best catalytic performance with reaction rate constant six times higher than that of the spherical counterpart. Via density functional theory (DFT) calculations, the abundant enclosed {100} facets in cubic Fe-Co PBA were identified to have the highest surface energy and favor high Fenton reaction activity. Keywords: secondary building units, shape-controlled, Prussian blue analogue, DFT calculation, Fenton reaction.
INTRODUCTION Metal-organic frameworks (MOFs) are a class of porous materials assembled from metal ions and bridging organic ligands, which have attracted immense attention over recent years.[1-3] Prussian blue analogues (PBAs) constitute a subclass of MOFs and show wide applications in catalysis, adsorption, gas storage and etc.[4-7] Shape-controlled synthesis of PBAs enables their properties to be tuned by tailoring the surface atomic structure and channel orientation.[8, 9] To date, PBAs with well-defined morphologies have been synthesized through various methods including reverse micro-emulsion,[10] chemical etching,[11] coordination modulation,[12] and etc, among which, the gradual transformation in shapes provides insights into the structure-activity relationship.[13, 14] The morphology
and size of PBAs could be tuned by changing pH of the growth solution, concentration of precursor as well as the type and amount of additives (e.g., PVP or transition metal ions).[15, 16] However, these methods generally led to more or less difference in composition or size, which poses challenges to studying the relation between morphology and activity. In this work, we report a facile method to synthesize a series of Fe3[Co(CN)6]2 (Fe-Co PBA) with finely tunable shapes without influencing the size and composition. The morphology modulation relies on the discovery that growth temperature is able to influence the growth rate of the secondary building units (SBUs) of PBA along different crystal directions. The shapes of the Fe-Co PBAs can be changed from
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 2 of 8
spheres, truncated spheres, truncated cubes to cubes with increase in growth temperature from 0 to 85 °C. The as-obtained Fe-Co PBAs were studied as the Fenton catalyst for bisphenol (BPA) degradation to investigate their shape-dependent catalytic activities. Degradation efficiency of BPA was achieved with more than 85% on Fe-Co PBA cubes, which is about 6 times higher than that on Fe-Co PBA spheres. Via density functional theory (DFT) calculations, BET measurement, X-ray photoelectron spectroscopy and Mössbauer spectroscopy, it is identified that the {100} facets on cubic Fe-Co PBA favor high Fenton reaction activity.
RESULTS AND DISCUSSION Fe-Co PBA with well-defined morphologies were synthesized by reacting K3[Co(CN)6] with FeCl2 in aqueous solution of PVP at temperatures from 0 to 85 °C. As shown in Figures 1a-d, with increasing the reaction temperature from 0 to 85 oC, the shape of the resultant Fe-Co PBA gradually changes from microspheres with no well-defined exposed facets to microcubes with six exposed {100} facets. Figures 1f-i display the EDS elemental mappings of the Fe-Co PBA microcubes synthesized at 85 oC, which show homogeneous distribution of the Fe, Co, C and N species. The atomic ratio of Fe/Co for various Fe-Co PBAs were measured to be around 1.5 (Table S1). Figure S1 compares the XRD patterns of all Fe-Co PBA samples, as shown, the identical diffraction peaks are well indexed to the face-centered cubic Fe3[Co(CN)6]2 (JCPDS No. 89-3736).[17] The strong and sharp diffraction peaks indicate high crystallinity. The thermal behavior of as-prepared Fe3[Co(CN)6]2 was studied by TGA as shown in Figure S2a, in which, all samples show similar stages of weight losses. The weight loss of around 20% in the first stage (below 150 oC) is due to the loss of crystalline water. The weight loss in the second (150-280 oC) could be ascribed to the removal of coordinated water molecules, while the the third stage (over 280 oC) to the decomposition of organic ligands (C≡N).[18] The specific surface areas of Fe-Co PBA synthesized at 0, 25, 55, 85 oC were determined to be 350, 252, 274 and 268 m2 g-1, respectively (Figure S2b and Table S1). Among different Fe-Co PBAs, the Fe-Co PBA microspheres synthesized at 0 oC display the largest surface area. Data obtained from dynamic light scattering (DLS) measurements were shown in Figure S3 and Table S2. The sizes of Fe-Co PBA synthesized under 0 oC, 25 oC, 55 oC, 85 oC were determined to be 955±144.6 nm, 1202±265.6 nm, 1473±270.8 nm, and 1425±398.6 nm, respectively, showing an increasing trend with temperature increase. During synthesis of Fe-Co PBA, a small portion of FeII would be inevitably oxidized into FeIII.[19] To investigate the content of FeII and FeIII species in the as-
Figure 1. SEM images of Fe-Co PBA synthesized at (a) 0 ℃, (b) 25 ℃, (c) 55 ℃, and (d) 85 ℃. (e) Bright-field TEM image and the corresponding EDS mappings of (f) Fe, (g) Co, (h) C, and (i) N.
prepared Fe-Co PBAs, we conducted 57Fe Mössbauer spectroscopy (Figure S4). The fitting model for the spectra is the same as our previous reports to a similar PBA complex.[17, 19] The spectra were fitted with four doublets, of which three with IS = 1.1 mm s-1 are assignable to FeII in various coordination environments and the one with IS = 0.33 mm s-1 belongs to high spin FeIII. As shown in Table S3, the ratio of FeII in Fe-Co PBA synthesized at 0, 25, 55, and 85 oC are 71.1%, 74.1%, 66.7%, and 73.4%, respectively, which match well with the XPS (Figure S5 and Table S4). Considering that oxidation of FeII may affect crystal growth and thus formation of product shapes. FeCl3 and FeCl2 mixture were used to prepare FeII0.5FeIII0.5Co at 0 oC and 55 oC. As shown in Figure S6, the obtained products with addition of half FeIII present the same shape as those synthesized from FeII, indicating that valence state of Fe may have little impact on the shapes of the final products. To investigate the morphology evolution, time-dependent SEM images were captured during the synthesis of Fe-Co PBA microcubes at 85 oC. It is well known that the final shape of a crystal is mainly determined by the coexistence of slow and fast rate of different growth facets. As crystal grows, it gradually shapes its morphology surrounded by the facets with slow growth rates. Here we define the simplified SBU of [FeCo(CN)6]n as a standard growth unit,[22-25] based
ACS Paragon Plus Environment
2
Page 3 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
Figure 2. (a) Schematic illustration showing the evolution of Fe-Co-85. (b) Mechanistic scheme showing FeCo PBA crystal formation at different reaction temperatures.
on which, a possible mechanism was proposed to explain the formation process of Fe-Co PBA microcubes. As shown in Figure 2a, the final microcubes of Fe-Co PBA compose of six square {100} facets, whereas the truncated microcubes obtained after 2 to 3 min of reaction are surrounded by six square {100} and eight triangle {111} facets, which are regarded as an intermediate stage between a cube and an octahedron. Because the nucleus was too sparse to be collected at the stage of nucleation, the earlier morphology could not be captured. However, based on the research performed by Cravillon et. al., who reported the observation of very small particles as early as 30 s during nucleation of ZIF-8,[26] it is reasonable to deduce that the PBA microcubes were developed from nucleus of small particles. This rapid nucleation process was followed by a relatively slower crystal growth. During growth, the {111} and {100} facets grew slower than other facets, leading to evolution of truncated microcubes. Finally, the {111} facets slowly diminished with time, resulting in the formation of Fe-Co PBA microcubes, due to the slower growth rate along the direction than that along the direction at 85 oC. At 0 oC, the growth rates along various directions keep nearly the same, leading to formation of spherical particles. As growth temperature increases, the growth rate along different directions will be accelerated but at different degrees. For example, when growth temperature increased to 55 oC, the growth rate along the direction became the fastest, which resulted in exposing the {111} and {100} facets of Fe-Co PBA, forming truncated microcubes. Further increase in growth temperature made the growth rate along the direction faster than that along the direction, eventually leading to the evolution of microcubes. Based on the above analysis, Figure 2b gives a
schematic summary of the morphology evolution process at various growth temperatures.
Figure 3. SEM images of (a) Mn0.5Fe0.5-Co-0, (b) Mn0.5Fe0.5-Co-25, (c) Mn0.5Fe0.5-Co-55, (d) Mn0.5Fe0.5-Co-85, (e) Mn-Co-0, and (f) Mn-Co-25.
To help understanding the growth behavior, surface energy of various Fe-Co PBA facets including (100), (110), and (111) were calculated by DFT method (Figure S7 and Table S5). As shown in Figure S5, (100) surface has 2 C and 2 N unsaturated coordination atoms, which correspond 4 broken C≡N bonds, (110) surface has 2 Co and 2 Fe unsaturated coordination atoms, corresponding 4 Co–C bonds and 4 Fe–N bonds broken, while (111) surface has 4 unsaturated coordination Fe atoms, corresponding 12 Fe–N bonds broken. Based on the energies of broken bonds, it can be predicted that {100} facets have the highest surface energy, then {111} facets, while the {110} facets have the lowest surface energy. From DFT calculations, it is clear to see that the (100) surface displays the highest surface energy followed by (111) and (110) surfaces (Table S5), which matches well with the experimental observation that spherical PBA particles were
ACS Paragon Plus Environment
3
ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 4 of 8
Figure 4. (a) Removal efficiency of BPA. (b) The corresponding rate constant. (c) Comparison of catalytic performance, specific surface area, FeII content and percentage of the {100} facet. (d) EPR spectra. (Black line: no catalysts; red line: Fe-Co-85 added). Reaction conditions: [BPA] = 20 mg L−1, [H2O2] = 2 mM, catalyst = 0.2 g L−1, T = 298 K, initial solution pH = 6.0.
synthesized at low temperature, while PBA microcubes were formed at high temperature. Additionally, by partially or completely replacing Fe with Mn, Mn0.5Fe0.5-Co and Mn-Co PBAs with tunable morphologies via changing growth temperature could be prepared (Figure 3), indicating the universal rule of tuning the growth rate SBUs for shape-controlled synthesis of PBA. To gain insights into the relation between catalytic performance and morphology, Fenton catalytic activity was probed on Fe-Co PBA synthesized at various temperatures. Figure 4a shows the removal profiles of BPA in the presence of different Fe-Co PBAs. Less than 20% of BPA could be degraded after 6 min of reaction with Fe-Co-0 and Fe-Co-25, while as high as 85% of BPA could be removed for the case of Fe-Co-85 as the catalyst. These results suggest that Fe-Co PBA microcubes are highly active towards activation of H2O2 for BPA degradation. A first-order kinetic model was used to determine the apparent rate constant (k) (Figure S8). As displayed in Figure 4b, the value of k is increased from 0.05 to 0.3 min-1 with increasing percentage of the exposed {100} facet in Fe-Co PBA. Moreover, as shown in Table S6, the catalytic performance of Fe-Co-85 is much higher than most of the recent reported catalysts and even comparable with homogeneous of Fe2+.
Specific surface area and content of FeII have been reported to influence the Fenton performance of FeCo PBA.[27, 28] To understand the underlying origin of different catalytic activities, the specific surface area and content of FeII for the four Fe-Co PBAs were compared as shown in Figure 4c. As can be seen, Fe-Co-0 shows the largest BET surface area, but has the lowest Fenton catalytic activity, while Fe-Co-85 exhibits the best performance in Fenton reaction, but only has moderate specific surface area and FeII content. Therefore, factors other than specific surface area and content of FeII should make Fe-Co-85 a superior Fenton catalyst. By further comparing the difference among the four Fe-Co PBAs, the catalytic activity seems to exhibit the same trend as the percentage of exposed {100} facet. By controlling the centrifugal rate, larger size of Fe-Co PBA nanocubes were collected (Figure S9), while showing poor degradation rate of less than 50% after 6 min. Further support that the Fenton-like catalytic activity of PBA is most probably related to {100} facets. EPR measurements were performed with Fe-Co-85 as the catalyst to verify the active species responsible for BPA degradation. As shown in Figure 4d and Figure S13, the characteristic signals are very similar with our previous reports, suggesting that 1O2 is the direct active species.[27, 29]
ACS Paragon Plus Environment
4
Page 5 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces
To further investigate the facet dependent Fenton activity, DFT calculations were performed to study the adsorption of H2O2 on various Fe-Co PBA surfaces. Figure 5 shows the optimized structure for the adsorption of a H2O2 molecule on the (100) [panel (a)] and (111) [panel (b)] surface of Fe-Co PBA. The adsorbed H2O2 molecule on (100) or (111) surface can be automatically decomposed to two hydroxyl radicals, indicating the high activity for HO• generation on both surfaces. Furthermore, it is shown that the two hydroxyl groups formed on the (100) surface are adsorbed on two different carbon atoms, while those formed on the (111) surface are just adsorbed on one iron atom. To better understand the interaction be-
Figure 5. Atomic configuration of H2O2 on the (100) [panel (a)] and (111) surface [panel (b)] of Fe-Co PBA. .
tween the surface and H2O2 for H2O2 activation, Table S7 compares the adsorption energy of H2O2 (Eads) as well as the total atomic charge of H2O2 molecule (Q) on the two surfaces. The adsorption on both surfaces is quite strong with Eads of -7.77 and -5.84 eV for (100) and (111) surface, respectively, agreeing well with the OH formation and chemically adsorbed as shown in Figure 5. Comparing the two surfaces, the adsorption of H2O2 on the (100) surface is stronger. On the (100) surface, one H2O2 molecule loses 0.15 e and transfers to two C atoms, while one H2O2 molecule obtains 0.68 e from one Fe atom on the (111) surface. The hydroxyl pairs produced on the (100) surface are adsorbed separately on two different carbon atoms, while they are adsorbed on the same iron atom on the (111) surface. Hence, the generation of hydroxyl radical becomes easier on the (100) surface of Fe-Co PBA than that on the (111) surface due to the separated hydroxyl
groups already existed and easier to detach from the surface with weaker adsorption energy. To give a better insight into the process of H2O2 activation on Fe-Co PBA nanocubes, a series of characterizations on the used catalysts were carried out. The result of XRD shows the crystallinity and main phase remained after reaction. SEM of Fe-Co PBA before and after reaction were shown in Figure S10. A layer of new phase was observed to cover on the surface of the nanocubes. High-resolution XPS spectra of Fe 2p2/3 suggests the oxidation of FeII with a new peak appeared at binding energy of 713.64 eV, which could be assigned to FeIII in Fe-O species. This can be further confirmed by XPS spectra of O 1s. The appearance of the new peak at 529.82 eV further confirming the formation of new Fe-O species after reaction. The Fenton reaction intermediates were identified by LC-MS to elucidate the degradation pathway of BPA.[30-33] Seven intermediates were detected as shown in Figure S11. At first, the nucleophilic attack of phenolic radical by water molecule occurs, resulting in mono- (C, m/z =244) or multi-hydroxylation (D) of the aromatic rings. The hydroxylated BPA then undergoes dehydration, forming quinone (E, m/z =242) and ring-opened intermediates like carboxylic compounds (F, m/z =324; G, m/z =266). Sequential decarboxylation and esterification of the formed carboxylic compounds leads to the formation of esters with higher molecular weight (G, m/z = 432;). The degradation pathway of BPA was proposed as shown in Figure S12. Based on these results, the mechanism of H2O2 activation on Fe-Co PBA can be deduced as the following: first, FeII coordinated with N activates H2O2 to generate OH·(Eq. 1) with itself being oxidized to FeIII. Subsequently, the oxidized FeIII is reduced back to FeII by H2O2 to form·OOH (Eq. 2). Finally, 1O2 is produced from ·OOH and OH·, acting as the active species to degrade BPA (Eq. 3 and 4). The easier detachment of hydroxyl groups from the (100) surface of Fe-Co PBA with weaker adsorption energy contributes to the optimal catalytic performance of Fe-Co-85. FeII + H2 O2 → FeIII + OH ∙ +OH−
(1)
FeIII + H2 O2 → FeII + OOH ∙ +H+
(2)
OOH ∙ +OH ∙→ OH− + 1O2∙
(3)
1O2∙ +BPA
(4)
→ intermediates → CO2 + H2 O
CONCLUSIONS In summary, we have developed a general strategy to synthesize Fe-Co PBA with well-defined morphology by controllably modulating the relative growth rate of the secondary building units (SBUs) along different crystallographic orientations. With increase in
ACS Paragon Plus Environment
5
ACS Applied Materials & Interfaces 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
synthesis temperature from 0 to 85 oC, the Fe-Co PBA crystals evolve from microspheres to microcubes. When applied as catalyst for H2O2 activation in Fenton reaction, the Fe-Co PBA microcubes exhibit the highest catalytic activity. Via DFT calculations, the (100) surface of Fe-Co PBA microcubes is demonstrated to favor the activation of H2O2 to generate the active 1O2 species for BPA degradation.
EXPERIMENTAL METHODS See Supporting Information, Text S1.
ASSOCIATED CONTENT Supporting Infromation The supplementary data associated with this article is available free of charge via the Internet at http://pubs.acs.org. Methods, characterization of Fe-Co PBA synthesized in different temperature before and after reaction; Detailed data of XRD, TG, BET, DLS, Mössbauer and XPS; kinetic linearization for catalysis degradation of BPA and the intermediate product shown by LC-MS; SEM of Fe-Co PBA with half FeIII and Fe-Co PBA collected under different centrifugal speed.
AUTHOR INFORMATION Corresponding Author Email:
[email protected],
[email protected],
[email protected] Notes: The authors declare no competing financial interests.
ACKNOWLEDGMENTS This work was supported by National Natural Science Foundation of China (21607029 and 21777033), Science and Technology Program of Guangdong Province (2017B020216003), Science and Technology Program of Guangzhou City (201707010359), “1000 plan” for young professionals' program of China, and “100 talents” program of Guangdong University of Technology.
REFERENCES [1] Kitagawa S, Matsuda R. Chemistry of Coordination Space of Porous Coordination Polymers [J]. Coord. Chem. Rev., 2007, 251 (21): 24902509. [2] Kitagawa S, Kitaura R, Noro S-I. Functional Porous Coordination Polymers [J]. Angew. Chem. int. Ed., 2004, 43 (18): 2334-2375. [3] Bae Y-S, Snurr R Q. Development and Evaluation of Porous Materials for Carbon Dioxide Separation and Capture [J]. Angew. Chem. int. Ed., 2011, 50 (49): 11586-11596. [4] Peeters A, Valvekens P, Ameloot R, Sankar G, Kirschhock C E A, De Vos D E. Zn-Co Double Metal Cyanides as Heterogeneous Catalysts for Hydroamination: A Structure-Activity Relationship [J]. ACS Catal., 2013, 3 (4): 597-607. [5] Li X, Wang J, Rykov A I, Sharma V K, Wei H, Jin C, Liu X, Li M, Yu S, Sun C, Dionysiou D D. Prussian Blue/Tio2 Nanocomposites as A Heterogeneous Photo-Fenton Catalyst for Degradation of Organic Pollutants in Water [J]. Catal. Sci. Technol., 2015, 5 (1): 504-514. [6] Takahashi A, Tanaka H, Parajuli D, Nakamura T, Minami K, Sugiyama Y, Hakuta Y, Ohkoshi S, Kawamoto T. Historical Pigment Exhibiting Ammonia Gas Capture Beyond Standard Adsorbents with Adsorption Sites of Two Kinds [J]. J. Am. Chem. Soc., 2016, 138 (20): 63766379. [7] Kaye S S, Long J R. Hydrogen Storage in the Dehydrated Prussian Blue Analogues M3[Co(CN)6]2 (M = Mn, Fe, Co, Ni, Cu, Zn) [J]. J. Am.
Page 6 of 8
Chem. Soc., 2005, 127 (18): 6506-6507. [8] Mao Y, Su B, Cao W, Li J, Ying Y, Ying W, Hou Y, Sun L, Peng X. Specific Oriented Metal-Organic Framework Membranes and their Facet-Tuned Separation Performance [J]. ACS Appl Mater interfaces, 2014, 6 (18): 15676-15685. [9] Hu L, Zhang P, Chen Q W, Yan N, Mei J Y. Prussian Blue Analogue Mn3[Co(CN)6]2.Nh2o Porous Nanocubes: Large-Scale Synthesis and their CO2 Storage Properties [J]. Dalton Trans, 2011, 40 (20): 55575562. [10] Mchale R, Ghasdian N, Liu Y, Ward M B, Hondow N S, Wang H, Miao Y, Brydson R, Wang X. Prussian Blue Coordination Polymer Nanobox Synthesis Using Miniemulsion Periphery Polymerization (MEPP) [J]. Chem. Commun., 2010, 46 (25): 4574-4576. [11] Hu M, Belik A A, Imura M, Yamauchi Y. Tailored Design of Multiple Nanoarchitectures in Metal-Cyanide Hybrid Coordination Polymers [J]. J. Am. Chem. Soc., 2013, 135 (1): 384-391. [12] Hu L, Zhang P, Chen Q W, Zhong H, Hu X Y, Zheng X R, Wang Y, Yan N. Morphology-Controllable Synthesis of Metal Organic Framework Cd3[Co(CN)6]2·Nh2o Nanostructures for Hydrogen Storage Applications [J]. Cryst. Growth Des., 2012, 12 (5): 2257-2264. [13] Chen J, Lim B, Lee E P, Xia Y. Shape-Controlled Synthesis of Platinum Nanocrystals for Catalytic and Electrocatalytic Applications [J]. Nano today, 2009, 4 (1): 81-95. [14] Ji W, Li M, Zhang G, Wang P. Controlled Synthesis of Bi25FeO40 with Different Morphologies: Growth Mechanism and Enhanced PhotoFenton Catalytic Properties [J]. Dalton Trans, 2017, 46 (11):1058610593. [15] Ming H, torad N L K, Chiang Y-D, Wu K C W, Yamauchi Y. Size- and Shape-Controlled Synthesis of Prussian Blue Nanoparticles by A Polyvinylpyrrolidone-assisted Crystallization Process [J]. Crystengcomm, 2012, 14 (10): 3387. [16] Li X, Yuan L, Wang J, Jiang L, Rykov A I, Nagy D L, Bogdan C, Ahmed M A, Zhu K, Sun G, Yang W. A "Copolymer-Co-Morphology" Conception for Shape-Controlled Synthesis of Prussian Blue Analogues and as-Derived Spinel Oxides [J]. Nanoscale, 2016, 8 (4): 2333-2342. [17] Liu J, Li X, Rykov A I, Fan Q, Xu W, Cong W, Jin C, Tang H, Zhu K, Ganeshraja A S, Ge R, Wang X, Wang J. Zinc-Modulated Fe–Co Prussian Blue Analogues with Well-Controlled Morphologies for the Efficient Sorption of Cesium [J]. J. Mater. Chem. A, 2017, 5 (7): 3284-3292. [18] Li X, Wang Z, Zhang B, Rykov A I, Ahmed M A, Wang J. Fexco3−Xo4 Nanocages Derived From Nanoscale Metal–Organic Frameworks for Removal of Bisphenol A by Activation of Peroxymonosulfate [J]. Appl. Catal. B-Environ., 2016, 181: 788-799. [19] Li X, Liu J, Rykov A I, Han H, Jin C, Liu X, Wang J. Excellent PhotoFenton Catalysts of Fe–Co Prussian Blue Analogues and their Reaction Mechanism Study [J]. Appl. Catal. B-Environ., 2015, 179: 196-205. [20] Grosvenor A P, Kobe B A, Biesinger M C, Mcintyre N S. investigation of Multiplet Splitting of Fe 2p XPS Spectra and Bonding in Iron Compounds [J]. Surf. interface Anal., 2004, 36 (12): 1564-1574. [21] Berrettoni M, Ciabocco M, Fantauzzi M, Giorgetti M, Rossi A, Caponetti E. Physicochemical Characterization of Metal Hexacyanometallate–Tio2composite Materials [J]. RSC Adv., 2015, 5 (45): 35435-35447. [22] Hu L, Mei J Y, Chen Q W, Zhang P, Yan N. Magnetically Separable Prussian Blue Analogue Mn-3 Co(CN)(6) (2)Center Dot Nh(2)O Porous Nanocubes as Excellent Absorbents for Heavy Metal Ions [J]. Nanoscale, 2011, 3 (10): 4270-4274. [23] Lee K J, Kim T-H, Kim T K, Lee J H, Song H-K, Moon H R. Preparation of Co3O4electrode Materials with Different Microstructures via Pseudomorphic Conversion of Co-Based Metal–Organic Frameworks [J]. J. Mater. Chem. A, 2014, 2 (35): 14393. [24] Tranchemontagne D J, Mendoza-Cortes J L, O'keeffe M, Yaghi O M. Secondary Building Units, Nets and Bonding in the Chemistry of MetalOrganic Frameworks [J]. Chem. Soc. Rev., 2009, 38 (5): 1257-1283. [25] Eddaoudi M, Moler D B, Li H, Chen B, Reineke T M, O'keeffe M, Yaghi O M. Modular Chemistry: Secondary Building Units as A Basis for the Design of Highly Porous and Robust Metal−Organic Carboxylate Frameworks [J]. Acc. Chem. Res., 2001, 34 (4): 319-330. [26] Cravillon J, Schroder C A, Nayuk R, Gummel J, Huber K, Wiebcke M. Fast Nucleation and Growth of ZIF-8 Nanocrystals Monitored by TimeResolved in Situ Small-Angle and Wide-Angle X-Ray Scattering [J]. Angew. Chem. int. Ed. Engl., 2011, 50 (35): 8067-8071. [27] Wang Y, Sun H, Ang H M, Tadé M O, Wang S. 3D-Hierarchically Structured Mno2 for Catalytic Oxidation of Phenol Solutions by Activation of Peroxymonosulfate: Structure Dependence and Mechanism [J]. Appl. Catal. B-Environ., 2015, 164: 159-167.
ACS Paragon Plus Environment
6
Page 7 of 8 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Applied Materials & Interfaces [28] Li X, Rykov A I, Wang J. Hydrazine Drastically Promoted Fenton Oxidation of Bisphenol A Catalysed by A Feiii–Co Prussian Blue Analogue [J]. Catal. Commun., 2016, 77: 32-36. [29] Saputra E, Muhammad S, Sun H, Ang H-M, Tadé M O, Wang S. Shape-Controlled Activation of Peroxymonosulfate by Single Crystal ΑMn2O3 for Catalytic Phenol Degradation in Aqueous Solution [J]. Appl. Catal. B-Environ., 2014, 154-155: 246-251. [30] Du J, Bao J, Liu Y, Ling H, Zheng H, Kim S H, Dionysiou D D. Efficient Activation of Peroxymonosulfate by Magnetic Mn-MGO for Degradation of Bisphenol A [J]. J. Hazard. Mater., 2016, 320: 150-159. [31] Sharma J, Mishra I M, Kumar V. Mechanistic Study of Photo-Oxidation of Bisphenol-A (BPA) with Hydrogen Peroxide (H2O2) and Sodium
Persulfate (SPS) [J]. J. Environ. Manage., 2016, 166: 12-22. [32] Sharma J, Mishra I M, Dionysiou D D, Kumar V. Oxidative Removal of Bisphenol A by UV-C/Peroxymonosulfate (PMS): Kinetics, Influence of Co-Existing Chemicals and Degradation Pathway [J]. Chem. Eng. J., 2015, 276: 193-204. [33] Li W, Wu P-X, Zhu Y, Huang Z-J, Lu Y-H, Li Y-W, Dang Z, Zhu N-W. Catalytic Degradation of Bisphenol A by Comnal Mixed Metal Oxides Catalyzed Peroxymonosulfate: Performance and Mechanism [J]. Chem. Eng. J., 2015, 279: 93-102.
ACS Paragon Plus Environment
7
ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 8
Table of Contents
ACS Paragon Plus Environment
8