A Cationic Coordination Polymer and Its Orange II Anion-Exchanged

On the other hand, organic dyes are also notorious pollutants in water.(40-49) In comparison to oxidation and biological ..... This is because when th...
0 downloads 4 Views 3MB Size
Article pubs.acs.org/IC

A Cationic Coordination Polymer and Its Orange II Anion-Exchanged Products: Isolation, Structural Characterization, Photocurrent Responses, and Dielectric Properties Dan Liu,† Fei-Fan Lang,‡ Xuan Zhou,† Zhi-Gang Ren,*,† David James Young,§ and Jian-Ping Lang*,† †

College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, People’s Republic of China ‡ Department of Chemistry, University of Sheffield, Brook Hill, Sheffield S3 7HF, U.K. § Faculty of Science, Health, Education and Engineering, University of the Sunshine Coast, Maroochydore DC, Queensland 4558, Australia S Supporting Information *

ABSTRACT: Solvothermal reactions of AgNO3 with N1,N4bis(5-fluoropyridin-3-yl)succinamide (bfps) in MeCN afforded the one-dimensional cationic coordination polymer {[Ag(bfps)]NO3}n (1). Upon treatment of 1 with the anionic azo dye orange II (NaOII) in aqueous solution, the NO3− anions of 1 could be gradually exchanged by the OII− anions via an anion-exchange process. The resulting OII anionexchanged products {[Ag(bfps)](NO3)0.85(OII)0.15}n (2) and {[Ag(bfps)](NO3)0.1(OII)0.9}n (3) were formed by different molar ratios of 1 and the newly formed phase “{[Ag(bfps)](OII)}n” (4), confirmed by PXRD patterns. Relative to those of the precursors 1 and NaOII, complexes 2 and 3 demonstrated enlarged photocurrent responses and reduced dielectric constants and dielectric losses, which could be correlated with the OII− contents in their structures. Complex 3 acquired a stable anodic photocurrent of 12.06 μA, which was 4.9 times higher than that of 1. The dielectric constant (εr = 4.2) and dielectric loss (0.002) of 3 were nearly frequency independent in the range from 1 to 106 Hz. The results provide an interesting insight into the rational assembly of CP-dye complexes and their tunable optoelectronic applications.



INTRODUCTION Research on organic−inorganic hybrid materials has been an intriguing topic over the past decades because of their diphasic structures1−3 and the associated applications in optics,4 thermotics,5,6 electronics,7−9 mechanics,10 biology,11 etc.12−17 Coordination polymers (CPs) are attracting considerable attention because the structures of these compounds can be determined by X-ray crystallography and readily controlled by adjusting the metal centers and bridging ligands.26,27 This flexibility seems less apparent in more conventional inorganic materials such as metal nanoparticles,18,19 metal oxides/ sulfides,20,21 silica,22 quantum dots,23 polyoxometalates,24 and graphene derivatives.25 Now many research groups have employed CPs as substrates to composite with the aforementioned inorganic materials, and also with organic polymers, enzymes, organic dyes, metalloporphyrins, biomolecules, and other functional molecules,28,29 for potential applications in catalysis, sorption, chemical sensors, etc.30−33 Among functional organic species, organic dyes are well-known for capturing light and have been employed for the enhancement of light absorption or emission in, for example, dye-sensitized solar cells34,35 and chemiluminescent sensors. Wu36 and Lan37 © 2017 American Chemical Society

reported the composites Rho@MOF (Rho = Rhodamine B; MOF = metal organic framework) and Rh6@MOF (Rh6 = Rhodamine 6G) could be used to detect volatile organic compounds. Upon encapsulation of a luminescent dye C460 (7-diethylamino-4-methylcoumarin) into a terbium-based MOF, Cui and co-workers employed this material as a highly sensitive luminescent thermometer.38 To date, only composites of dyes and CdS nanoclusters have been investigated for their electronic properties such as photocurrent response and dielectric performance,39 and those CPs-based composites are very limited. On the other hand, organic dyes are also notorious pollutants in water.40−49 In comparison to oxidation and biological treatment, the uptake of dyes from wastewater is comparatively fast, cheap, and easy to perform.49,50 CPs have proven to be potential candidates for the removal of dyes via physical/ chemical sorption,51−55 employing a variety of nonbonded interactions such as electrostatic attraction, hydrogen bonding, π−π stacking, and ion exchange.56−61 Among these interReceived: August 8, 2017 Published: October 2, 2017 12542

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry

refluxed for 6 h and then reduced to 5 mL in vacuo. Cold water (100 mL) was added to form a light yellow precipitate, which was then collected by filtration, washed with water, and dried in vacuo. Yield: 1.58 g (80% based on succinyl dichloride). Anal. Calcd for C14H12F2N4O2: C, 54.90; H, 3.95; N, 18.29. Found: C, 55.04; H, 4.16; N, 18.62. IR (KBr disk): 3286 (s), 3035 (m), 1657 (vs), 1597 (m), 1537 (s), 1467 (m), 1427 (m), 1333 (m), 1289 (m), 1195 (m), 1163 (m), 1022 (w), 989 (w), 973 (w), 878 (m), 701 (m), 673 (w), 535 (w) cm−1. 1H NMR (400 MHz, DMSO-d6, 298 K, TMS): δ 10.54 (s, 2H, −NH−), 8.52 (s, 2H, −Py), 8.25 (d, J = 1.5 Hz, 2H, −Py), 8.05 (d, J = 11.4 Hz, 2H, −Py), 2.73 (s, 4H, −CH2−). 13C NMR (151 MHz, DMSO-d6): δ 171.35 (−CO−), 158.73 (d, J = 252.9 Hz, −Py), 137.06 (d, J = 6.2 Hz, −Py), 136.56 (d, J = 2.8 Hz, −Py), 131.41 (d, J = 22.8 Hz, −Py), 112.54 (d, J = 22.7 Hz, −Py), 30.79 (−CH2−). HRMS: calcd for (C14H12F2N4O2 + H+) 307.1007, found 307.1010 (100%). Preparation of {[Ag(bfps)]NO3}n (1). In a Pyrex glass tube (15 cm in length, 7 mm in inner diameter) were placed AgNO3 (17 mg, 0.1 mmol), bfps (30 mg, 0.1 mmol), and MeCN (14 mL). The tube was sealed and heated in an oven at 120 °C for 16 h and then cooled to room temperature at a rate of 5 °C h−1 to form pale yellow crystals of 1, which were collected by filtration, washed with ethanol, and dried in air. Yield: 29 mg (62% based on Ag). Anal. Calcd for C14H12AgF2N5O5: C, 35.28; H, 2.52; N, 14.70. Found: C, 35.44; H, 2.73; N, 14.84. IR (KBr disk): 3286 (s), 3081 (m), 1701 (s), 1597 (m), 1543 (s), 1471 (s), 1423 (s), 1384 (vs), 1320 (s), 1299 (m), 1250 (w), 1173 (s), 1024 (w), 995 (w), 874 (m), 694 (m), 535 (w) cm−1. Preparation of {[Ag(bfps)](NO3)0.85(OII)0.15}n (2). To an aqueous NaOII (0.09 mmol) solution (40 mL) was added compound 1 (48 mg, 0.1 mmol). The mixture was stirred for 15 min. The resulting solid 2 was separated by centrifugation, washed with water, and dried in vacuo. Yield: 50 mg (97% based on Ag). Anal. Calcd for C16.4H13.65AgF2N5.15O5.15S0.15: C, 38.18; H, 2.67; N, 13.98; S, 0.93; Ag, 20.90. Found: C, 38.02; H, 2.78; N, 13.80; S, 0.92; Ag, 20.50. IR (KBr disk): 3288 (m), 3080 (m), 1702 (s), 1598 (s), 1548 (s), 1507 (w), 1472 (m), 1423 (m), 1384 (vs), 1321 (m), 1298 (m), 1207 (m), 1174 (s), 1119 (m), 1035 (w), 1004 (w), 873 (w), 833 (w), 701 (w), 696 (w), 644 (w), 535 (w) cm−1. Preparation of {[Ag(bfps)](NO3)0.1(OII)0.9}n (3). Complex 3 was prepared by a method similar to that used for the isolation of 2, with 45 min of stirring. Yield: 69 mg (96% based on Ag). Anal. Calcd for C28.4H21.9AgF2N5.9O5.9S0.9: C, 47.71; H, 3.09; N, 11.56; S, 4.04; Ag, 15.09. Found: C, 47.71; H, 3.73; N, 12.07; S, 4.16; Ag, 14.86. IR (KBr disk): 3269 (s), 3079 (w), 1705 (s), 1618 (m), 1598 (s), 1552 (s), 1507 (s), 1473 (s), 1453 (w), 1423 (s), 1384 (w), 1332 (m), 1294 (m), 1230 (w), 1208 (s), 1178 (s), 1119 (s), 1034 (m), 1005 (w), 987 (w), 873 (w), 834 (m), 759 (m), 697 (s), 644 (m), 596 (w), 535 (w) cm−1. X-ray Structure Determinations. Single crystals of bfps and 1 suitable for X-ray analysis were obtained directly from the above preparations. Data were obtained on an Agilent Xcalibur CCD X-ray diffractometer by using graphite-monochromated Mo Kα (λ = 0.71073 Å) radiation at 223 K. The program CrysAlisPro (Agilent Technologies, Ver. 1.171.36.32, 2013) was used for the integration of reflection data and the refinement of unit cell parameters by using all observed reflections. An absorption correction (multiscan) was performed using the program SADABS,67−69 and the reflection data were corrected for Lorentz and polarization effects. Both structures were solved by the direct method using SHELXS-2016 and refined on F2 by full-matrix least squares using the SHELXL-2016 program.70 All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were located at the geometrically idealized positions and also constrained to ride on their parent atoms. A summary of the key crystallographic data for bfps and 1 is provided in Table 1.

actions, ion exchange has arguably the greatest potential for maximal loading because of the possibility of replacing all counterions in the structure. For example, three anionic MOFs including [(CH 3 ) 2 NH 2 ] 6 [M(H 2 O) 6 ] 3 {M 6 (η 6 -TATAT)4(H2O)12}·xH2O (M = Co2+, Ni2+; TATAT = 5,5′,5″((1,3,5-triazine-2,4,6-triyl)tris(azanediyl))triisophthalic acid)62 and [(CH 3 )NH 2 ] 5 [Cd 3 . 5 (TDPAT) 2 (H 2 O) 2 ]·20H 2 O (H6TDPAT = 2,4,6-tris(3,5-dicarboxylphenylamino)-1,3,5-triazine)63 could uptake cationic dyes such as methylene blue (MB) with maximum capacities of 708, 725, and 1323 mg g−1, respectively, via exchanging [M(H2O)6]2+ or [(CH3)2NH2]+ cations with the cationic part of MB. Recently, we found that the one-dimensional (1D) anionic coordination polymer [(H2L)0.5][Zn(1,2,3-BTA)(H2O)]·2H2O (L = N,N′-bis(3-pyridyl)succinamide, 1,2,3-H3BTA = 1,2,3benzenetricarboxylic acid) could strongly adsorb the dye Congo Red (CR) with a capacity of 1129 mg g−1 by hydrogen-bonding interactions among H2L2+, the anionic chain, and CR anions. The resulting CP-CR composite exhibited an evidently enhanced photocurrent response.64 In addition, a family of 1D iodoplumbate complexes of 4cyanopyridinium and N,N′-dialkyl-4,4′-bipyridium, among which the anionic [PbI3]nn− chains were enclosed into channels assembled by the pyridinium cations, showed nearly frequency independent low dielectric constants (low k) in the range of 102−106 Hz with low dielectric losses.65 In this paper, we report the syntheses of N1,N4-bis(5-fluoropyridin-3-yl)succinamide (bfps, Chart 1) and its 1D Ag(I)-based coordination polymer Chart 1. Schematic Representation of bfps and Orange II (NaOII)

{[Ag(bfps)]NO3}n (1), isolated from the solvothermal reaction of AgNO3 with bfps. When 1 was treated with the anionic azo dye orange II (NaOII, Chart 1), the OII− anions could be gradually exchanged into the crystal lattice of 1 via an anionexchange process. We isolated two the time-dependent OII anion-exchanged products {[Ag(bfps)](NO3)0.85(OII)0.15}n (2) and {[Ag(bfps)](NO3)0.1(OII)0.9}n (3), which were generated by different molar ratios of 1 and the newly formed phase “{[Ag(bfps)](OII)}n” (4). Both products exhibited enhanced photocurrent responses and improved dielectric properties relative to those of 1 and NaOII. Described below are their synthesis, structural characterization, photocurrent responses, and dielectric properties.



EXPERIMENTAL SECTION

General Procedures. Some analytical instruments related to the work are the same as those employed in our previous articles.65,66 Elemental analyses for C, H, N, and S were performed on a Carlo-Erba CHNO-S microanalyzer or a Elementar Vario MICRO cube elemental analyzer. Ion chromatography (IC) was measured on a Metrohm 861 Advanced Compact IC system. The photocurrent response data were collected with a CHI 630E analyzer. Preparation of N1,N4-Bis(5-fluoropyridin-3-yl)succinamide (bfps). To a DMF solution (50 mL) of 5-fluoropyridin-3-amine (1.5 g, 13.4 mmol) were added succinyl dichloride (1.00 g, 6.45 mmol) and triethylamine (1.31 g, 12.9 mmol). The resulting mixture was



RESULTS AND DISCUSSION Synthesis and Structural Characterization. Ligand bfps was prepared in two steps via a standard acylation reaction.71 It

12543

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry

was insoluble in water, CHCl3, and CH2Cl2 but soluble in polar organic solvents such as MeOH, EtOH, acetone, MeCN, DMF, and DMSO. In its IR spectrum, stretching vibrations at 3286 cm−1 (N−H), 1657 (CO), 3035, 1537, 1467, 1427 (Py), and 1195 cm−1 (C−F) were observed. Its 1H NMR and 13C NMR spectra in DMSO-d6 (Figures S1 and S2 in the Supporting Information) showed their corresponding signals at 8.52−8.05 and 158.73−112.54 ppm for the −Py groups, 10.54 ppm for the −NH− group, 2.73 and 30.79 ppm for the −CH2− group, and 171.35 ppm for the −CO− group. The crystal structure of bfps was confirmed by single-crystal X-ray diffraction (Figure S3 in the Supporting Information). Compound 1 was isolated as colorless crystals from the solvothermal reaction of AgNO3 and equimolar bfps in MeCN. It was relatively air and moisture stable and insoluble in water and common organic solvents such as CHCl3, CH2Cl2, CH3CN, CH3OH, and benzene. Its elemental analyses were compatible with its chemical formula. Stretching vibrations at 3286 cm−1 (N−H), 1701 (CO), 3081, 1543, 1471, 1423 (Py), and 1173 (C−F) cm−1 were observed in its IR spectrum. The characteristic absorption at 1384 cm−1 for the NO3− anion was also located. According to the TGA curve of 1 (Figure S4

Table 1. Crystal Data and Structure Refinement Parameters for bfps and 1 empirical formula fw cryst syst space group a/Å b/Å c/Å β/deg V/Å3 ρcalc/g cm−3 Z μ/mm−1 F(000) R1a wR2b GOFc

bfps

1

C14H12F2N4O2 306.28 monoclinic P21/c 5.0345(6) 12.7583(14) 11.6106(14) 111.747(10) 692.69(15) 1.468 2 0.120 316 0.0620 0.1946 1.056

C14H12F2N5O5Ag 476.16 monoclinic P2/c 10.4293(6) 7.9200(4) 10.2901(5) 108.956(6) 803.87(8) 1.967 2 1.317 472 0.0242 0.0485 1.009

R1 = ∑||Fo| − |Fc||/∑|Fo|. bwR2 = {∑w(Fo2 − Fc2)2/∑w(Fo2)2}1/2. GOF = {∑w((Fo2 − Fc2)2)/(n − p)}1/2, where n = number of reflections and p = total number of parameters refined.

a c

Figure 1. (a) View of a section of the 1D chain in 1 with a labeling scheme and 50% thermal ellipsoids. (b) View of the 2D network composed by 1D chains {[Ag(bfps)]NO3}n via the H-bonding interactions in 1 extending approximately along the ac plane. (c) View of a pair of parallel 2D Hbonded sheets in 1. All H atoms are omitted for clarity. Symmetry codes: (A) −x, −y, −z; (B) 1 − x, y, 1.5 − z; (C) 1 − x, −y, 1 − z. Selected bond lengths (Å) and angles (deg): Ag1−N1 2.1688(18), Ag1···O2 2.696(2), N3−O2 1.251(2), N3−O3 1.245(4); N1−Ag1−N1B 159.87(11), O2−N3− O2B 119.1(3), O2−N3−O3 120.46(16). 12544

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry in the Supporting Information), it was stable up to 225 °C and then two weight losses appeared, which correlated to the elimination of bfps at 227−395 °C (64.0%, calcd 64.3%) and NO3− at 395−445 °C (14.0%, calcd 13.0%). The residual species was assumed to be metal silver (22.0%, calcd 22.7%). The observed PXRD patterns of 1 matched well with those simulated from its single-crystal X-ray diffraction data (Figure S5 in the Supporting Information). Compound 1 crystallizes in the monoclinic space group P2/c, and its asymmetric unit contains half a [Ag(bfps)] moiety and half a NO3− anion (Figure 1). Ag1 is coordinated with two N atoms of two pyridyl groups from two bfps ligands and two O atoms from one NO3−, forming a seesaw-shaped coordination geometry. The Ag1−N1 bond length (2.1688(18) Å) is slightly longer than that in {[Ag2(L)2](NO3)2·(H2O)2.5}n (2.155(2) Å, L = N,N′-bis(3-pyridyl)succinamide)72 but shorter than those in {Ag(L1)(NO3)}n (2.203(2) Å, L1 = 1,4-bis(2-pyridyl)buta1,3-diyne)73 and {Ag(py)2(NO3)}n (2.180(2) Å).74 Each bfps bridges two Ag atoms to yield a zigzag 1D chain extending along the [2,0,3] direction. The NO3− anion is appended to the cationic 1D chain by two weak Ag···O interactions (Ag1···O2 and Ag1···O2B, 2.696(2) Å), which are between those of {[Ag 2 (L) 2 ](NO 3 ) 2 ·(H 2 O) 2.5 } n (2.423(2) Å) and {Ag(py)2(NO3)}n (2.886(2) Å). Such a 1D chain is further linked to its equivalent chains by intermolecular H-bonding interactions (N2−H2A···O2C; symmetry code for A −x, −y, −z and for C 1 − x, − y, 1 − z) to form a 2D H-bonding wavelike network with the contact between two sheets being 7.9200(4) Å (Figure 1b,c). Anion Exchange between 1 and NaOII. Well-ground crystalline solid 1 (0.1 mmol) was mixed with a 40 mL aqueous solution of NaOII (0.09 mmol) with continuous stirring. The solution was sampled at 5 min intervals to measure the concentrations of the residual OII− anion along with the UV− vis spectra and the PXRD patterns of the resulting solids. With the prolonging of time, the solid-state UV−vis spectra of the resulting solids (Figure S6 in the Supporting Information) showed clearly enhanced absorption peaks at ∼440 nm related to that of NaOII. The UV−vis spectra of the aqueous solution (Figure 2a) indicated that after 45 min no OII− was left in the solution. Attempts to use more NaOII (0.10−0.12 mmol) and longer times (up to 6 h) did not lead to more exchange

between NO3− and OII−. Thus, 1 g of 1 consumed 662 mg of NaOII. This exchange process was monitored by measuring the residual concentrations of other inorganic ions in solution by ion chromatography (IC) for NO3− and inductively coupled plasma emission spectroscopy (ICP) for Na+ and Ag+. During the process, c(Na+) remained nearly unchanged (2.47−2.55 mmol/L) and no Ag+ was detected. However, c(OII−) was decreased from 2.25 mmol/L (0 min) to 1.91 mmol/L (15 min) and 0.02 mmol/L (45 min), respectively, while c(NO3−) was gradually increased from 0 (0 min) to 0.38 (15 min) and 2.24 (45 min) mmol/L. These results revealed that, as compound 1 contains the cationic 1D [Ag(bfps)]nn+ chain and weakly interacting NO3−, the aforementioned process might go through an anion exchange between NO3− and OII−. Control experiments using other dyes (methyl orange (MO), orange I (OI) and orange IV (OIV), methylene blue (MB), malachite green (MG), and methyl violet (MV)) (Figure S7 in the Supporting Information) and activated carbon (Figure S8 in the Supporting Information) were performed. Under the same experimental conditions, compound 1 could rapidly adsorb the anionic dyes MO, OI, and OIV but was almost inactive in adsorbing cationic dyes such as MB, MG, and MV, which was consistent with the aforementioned anion exchange mechanism. On the other hand, 1 g of activated carbon could adsorb approximately 264 mg of NaOII within 30 min, and no anion exchange was found to proceed. Monitoring changes in the concentrations of the residual OII− (Figure 2b) in solution revealed that the anion exchange was relatively slow during the first 15 min but became accelerated after this time period. Figure 3 shows the simulated and observed PXRD patterns of solids separated from mixing 1 with NaOII at 5 min intervals. Some new peaks (e.g., 2θ = 7.1, 7.8, 19.4, 21.8, 25.2°) gradually appeared and remained unchanged until 45 min. The powder patterns were then indexed and fitted by Pawley refinement75 using the TOPAS-

Figure 2. (a) Absorption spectra of NaOII solutions and (b) different concentrations of OII− after mixing with 1 at different time intervals.

Figure 3. Simulated and experimental PXRD patterns of NaOII and solids obtained from mixing 1 with NaOII at 5 min intervals. 12545

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry

1423, and 1173 cm−1 in 1 were almost unshifted in 2 and 3, indicating the presence of the similar coordination environments of bfps and Ag(I) and the similar [Ag(bfps)]nn+ chain in 2 and 3. With the increase of OII− in 2 and 3, the intensity of the NO3− stretching vibration at 1384 cm−1 in the IR spectrum of 1 gradually went down while those of the −SO3− (in OII−) stretching vibrations at 1210 and 1037 cm−1 were gradually enhanced. Therefore, the aforementioned process may more likely arise from the introduction of the larger anion OII− into the lattice of 1 via anion exchange, which may understandably enlarge the separation between the [Ag(bfps)]nn+ chains. This assumption was in agreement with the reduced dielectric εr values caused by the decreasing density, discussed later in this paper. To this end, we suppose that the above process may adopt a surface-exchange step followed by an internal-exchange step. As the PXRD patterns showed (Figure 3), during the first 15 min, the major portion of 1 remained unchanged while minor exchange between NO3− and OII− took place mainly on the surface of 1 to yield a new phase (compound 4). Only 15% of NO3− in 1 was exchanged into the solution. As time went on, the anion exchange went further into the lattice of 1 and more and more NO3− anions were exchanged by larger OII− anions, which squeezed into the 1D [Ag(bfps)]nn+ chains. Such a replacement understandably dilated the distance between the chains, accelerated the anion exchange rate, and increased the amount of the exchanged NO3−. Thus, in the following 30 min, another 75% of NO3− in 1 was exchanged into the solution and no more anion exchange could proceed. Complex 3 was readily separated by centrifugation and flushed with saturated NaNO3 aqueous solution (20 mL). Only 0.14% of OII− was eluted (calculated by UV−vis spectrophotometry on the eluent), suggesting that the interactions of the cationic [Ag(bfps)]nn+ chain with OII− anions in 3 were strong and 3 was stable to leaching. However, if the eluent was changed to a saturated solution of NaNO3 in EtOH and H2O (100 mL, 1/1 v/v), a total of 34% of OII− could be eluted (calculated by UV−vis spectrophotometry on the eluent) but further flushing could not increase the amount of the eluted OII−. Intriguingly, such an anion exchange between NO3− and OII− was unfinished and 10% of NO3− anions in 1 still remained. The reason was probably attributed to the fact that each OII− anion has a much larger size than each NO3− anion. When an increasing amount of compound 4 forms around solid 1, it may block the exchange pathways to the remaining NO3− anions. Although this process was incomplete, its driving force was probably due to the presence of strong electrostatic interactions, π−π interactions, and hydrogen bonds between the cationic [Ag(bfps)]nn+ chain and the OII− anions. Photocurrent Responses. The electrochemical properties and photocurrent responses of bfps and products 1−3 were examined using the method detailed in our previous work.77 Cyclic voltammetry in water (Figure S14 in the Supporting Information) showed no peak for bfps and reversible oxidation and reduction peaks at 0.63/0.15 V for 1, 0.48/−0.10 V for 2, and 0.56/0.10 V for 3 without light, which may be ascribed to the Ag(I)/Ag(0) couple. The oxidation state for Ag in 1−3 was +1, which indicated that they were stable during the anodic process. Thus, we investigated the anodic current that caused the H2O → O2 reaction. Under a Xe light (150 W, 40 mW cm−2), the anodic currents for 1−3 were all enhanced at potentials superior to their oxidation peaks. Considering that the dark current became conspicuous at 1.0 V, a bias potential

Academic program.76 Pawley fitting of the PXRD patterns of the resulting solid isolated in the first 15 min (Figure S9 in the Supporting Information) indicated the existence of two phases: one for 1 (major) and the other for a newly emerged phase (minor), tentatively assigned as “{[Ag(bfps)](OII)}n” (4). On the basis of the intensities which did not belong to those of 1, we indexed the rough unit cell of compound 4 as follows: orthorhombic, P2221, a = 12.14(3) Å, b = 11.15(4) Å, c = 15.87 (2) Å, V = 2148.5(8) Å3. Similar procedures were applied for the PXRD data for complex 3 (at 45 min), resulting in a more reliable unit cell as follows: orthorhombic, P2221, a = 12.08(1) Å, b = 11.036(9) Å, c = 15.65(1) Å, V = 2087.1(11) Å3 (Figure S10 in the Supporting Information). We thus designated the two end products as complexes 2 and 3 and further examined their structural and spectral aspects. Elemental analyses of both products (C, H, N, S by combustion and Ag by ICP) confirmed the formulas for 2 and 3, respectively. Both formulas matched quite well with the measured concentrations of the residual NO3− and OII− after 15 and 45 min. The color of 1 gradually darkened when more OII− was exchanged into its lattice, thereby yielding 2 and 3 (Figure 4). SEM images (Figure S11 in the Supporting

Figure 4. Photos of products 1−3 and NaOII.

Information) revealed that the resulting products became more crystalline with the continuation of anion exchange, which was consistent with the higher crystal system (orthorhombic) deduced by Pawley refinement of their PXRD patterns. Notably, in their XPS spectra (Figure S12a in the Supporting Information), 1−3 exhibited the same peaks at 368.28 and 374.30 eV attributed to Ag 3d5/2 and Ag 3d3/2, respectively, suggesting that the Ag(I) centers in these complexes hold the same oxidation state of +1 and the same coordination surroundings. In addition, the S 2p peaks at 168.22 eV in 2 and 3 were almost the same but were gradually enhanced, implying that the SO3− group of OII− remained intact with Ag(I) and also could be correlated to the growing amounts of OII− in 2 and 3 (Figure S12b). In their IR spectra (Figure S13 in the Supporting Information), the peaks at 1701, 1543, 1471, 12546

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry of 0.85 V (vs SCE) was applied for the photocurrent measurements. The photocurrent responses of 1−3 were recorded using a manual shutter which blocked Xe light irradiation with a typical interval of 20 s. The PXRD analyses of samples scraped from the ITO electrode (Figure S15 in the Supporting Information) indicated that 1−3 remained stable during the photocurrent experiments. Products 1−3 produced stable anodic photocurrents of 2.46 μA (1), 4.46 μA (2), and 12.06 μA (3), respectively (Figure 5). The photocurrents of the OII− anion-

Figure 5. Photocurrent responses of ITO electrodes coated with 1 (black), 2 (red), and 3 (green). Conditions: bias 0.85 V vs SCE, [Na2SO4] = 0.1 mol/L.

exchanged products 2 and 3 were 1.8 times and 4.9 times higher than that of 1. This was probably due to the increased content of OII− anions, which acquired more light adsorption and greater electron transportation, thereby resulting in larger anodic currents. Dielectric Properties. The dielectric behaviors of 1−3 and NaOII were examined at variable frequencies (1 to 106 Hz) at 298 K. The dielectric constant (permittivity, εr) of 3 was found to be nearly frequency-independent from 1 to 106 Hz with an estimated εr value of 4.2 (3) (Figure 6a). However, the εr values of 1, 2, and NaOII were found to gradually decrease with increasing frequency (1 to 102 Hz), and then became nearly frequency independent in the range from 102 to 106 Hz with estimated εr values of 5.7 (1), 5.2 (2), and 4.0 (NaOII). The estimated εr values for 3 and NaOII were slightly lower than that of SiO2 (4.3).78 The dielectric constant was mainly determined by the density and the total polarizability of the materials.79,80 Thus, we assumed that the εr value order of 1 > 2 > 3 > NaOII was correlated to their reduced densities and reduced polarizabilities caused by the increasing amounts of OII− exchanged into the lattice of 1. Correspondingly, the dielectric loss (εi/εr) values of NaOII and 1−3 exhibited the order NaOII > 1 > 2 > 3 within the range of 1 to 102 Hz (Figure 6b). This trend was related to the presence of small Na+ ions in NaOII and the decreasing NO3− percentage from 1 to 2 to 3. For 1−3, fewer NO3− anions mean more OII− anions entered into their lattices, which might lead to the retardation of the rapidly rotating dipoles and thus the reduction of dielectric loss in the order 1 > 2 > 3. In the higher frequency range from 102 to 106 Hz, their εi/εr values gradually approach zero. Notably, complex 3 exhibited a frequencyindependent, steady low dielectric loss (0.002) over the whole frequency range examined. This is because when the OII−

Figure 6. (a) Frequency dependence of dielectric constants (permittivity, εr) of 1-3 and NaOII at 298 K. (b) Frequency dependence of dielectric loss (εi/εr) of 1−3 and NaOII at 298 K. The εr and εi values are the real and imaginary parts of permittivity, respectively.

anions were sandwiched between the [Ag(bfps)]nn+ chains, both anions and cations are large and hard to rotate freely. This phenomenon was observed in previous reports.65,81 For example, enclosing anionic [PbI3]nn− chains into pyridinium cation channels64 and exchanging chlorides and water molecules with larger CH3COO− anions and CH3CN solvates in [Ni(H2bbim)3]Cl2·2H2O78 could lead to the evident reduction in their dielectric constants and dielectric losses.



CONCLUSIONS In this work, we prepared the ligand bfps and its 1D cationic coordination polymer 1 from the solvothermal reaction of AgNO3 and bfps. Compound 1 consists of 1D cationic [Ag(bfps)]nn+ chains and NO3− counteranions. Further treatment of 1 with NaOII at 15 and 45 min yielded the two OII anion-exchanged products 2 and 3, respectively. The above process may adopt a surface-exchange step followed by an internal-exchange step. From an analysis of their PXRD patterns, as the exchange proceeded, compound 1 was gradually converted into a new phase with a formula tentatively assigned as “{[Ag(bfps)]OII}n” (4), thereby resulting in the formation of products 2 and 3 with different molar ratios of the two 12547

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry phases 1 and 4. Owing to the increased content of OII− anions, which could capture more light adsorption and greater electron transportation, the photocurrent responses of 2 and 3 were 1.8 and 4.9 times larger than that of 1, respectively. Relative to those of the precursor 1, the dielectric constants and dielectric losses of 2 and 3 were significantly reduced, which could be correlated with the OII− contents in their structures. Complex 3 possessed a low dielectric constant (εr = 4.2) and dielectric loss (0.002) that were nearly frequency independent in the range from 1 to 106 Hz. The synthetic strategy of 2 and 3 and their excellent performances in photocurrent responses and dielectric properties demonstrated herein may provide a convenient pathway for multifunctional optoelectronic materials by combining ionic CPs with counterionic dyes.



(4) Gomez-Romero, P. Hybrid Organic-Inorganic Materials-in Search of Synergic Activity. Adv. Mater. 2001, 13, 163−174. (5) Gilman, J. W.; Jackson, C. L.; Morgan, A. B.; Harris, R., Jr. Flammability Properties of Polymer-Layered-Silicate Nanocomposites. Polypropylene and Polystyrene Nanocomposites. Chem. Mater. 2000, 12, 1866−1873. (6) Gilman, J. W. Flammability and Thermal Stability Studies of Polymer Layered-Silicate (Clay) Nanocomposites. Appl. Clay Sci. 1999, 15, 31−49. (7) Innis, P. C.; Norris, I. D.; Kane-Maguire, L. A. P.; Wallace, G. G. Electrochemical Formation of Chiral Polyaniline Colloids Codoped with (+)- or (−)-10-Camphorsulfonic Acid and Polystyrene Sulfonate. Macromolecules 1998, 31, 6521−6528. (8) Du, H. J.; Zhang, W. L.; Wang, C. H.; Niu, Y. Y.; Hou, H. W. A New Nanocrystalline Inorganic-Organic Hybrid Exhibiting Semiconducting Properties and Applications. Dalton Trans. 2016, 45, 2624−2628. (9) Xie, J. L.; Guo, C. X.; Li, C. M. Construction of One-Dimensional Nanostructures on Graphene for Efficient Energy Conversion and Storage. Energy Environ. Sci. 2014, 7, 2559−2579. (10) Okada, A.; Usuki, A. The Chemistry of Polymer-Clay Hybrids. Mater. Sci. Eng., C 1995, 3, 109−115. (11) Chung, J. H.; Kim, Y. K.; Kim, K. H.; Kwon, T. Y.; Vaezmomeni, S. Z.; Samiei, M.; Aghazadeh, M.; Davaran, S.; Mahkam, M.; Asadi, G.; Akbarzadeh, A. Synthesis, Characterization, Biocompatibility of Hydroxyapatite-Natural Polymers Nanocomposites for Dentistry Applications. Artif. Cells, Nanomed., Biotechnol. 2016, 44, 277−284. (12) Lee, D. W.; Yoo, B. R. Advanced Silica/Polymer Composites: Materials and Applications. J. Ind. Eng. Chem. 2016, 38, 1−12. (13) Fan, D. W.; Hao, J. C.; Wei, Q. Assembly of PolyoxometalateBased Composite Materials. J. Inorg. Organomet. Polym. Mater. 2012, 22, 301−306. (14) Unterlass, M. M. Green Synthesis of Inorganic-Organic Hybrid Materials: State of the Art and Future Perspectives. Eur. J. Inorg. Chem. 2016, 2016, 1135−1156. (15) Genovese, M.; Lian, K. Polyoxometalate Modified InorganicOrganic Nanocomposite Materials for Energy Storage Applications: A Review. Curr. Opin. Solid State Mater. Sci. 2015, 19, 126−137. (16) Wang, F. F.; Jia, S. Y.; Li, D. L.; Yu, B.; Zhang, L. W.; Liu, Y.; Han, X.; Zhang, R.; Wu, S. H. Self-Template Synthesis of CuO@ Cu3(BTC)2 Composite and its Application in Cumene Oxidation. Mater. Lett. 2016, 164, 72−75. (17) Kayan, A. Synthesis, Characterization, and Application of Hybrid Inorganic-Organic Composites (K/Na)ZrSi(R)Ox. J. Inorg. Organomet. Polym. Mater. 2016, 26, 640−647. (18) Castro, Y.; Ferrari, B.; Moreno, R.; Durán, A. Coatings Produced by Electrophoretic Deposition from Nano-Particulate Silica Sol-Gel Suspensions. Surf. Coat. Technol. 2004, 182, 199−203. (19) Sun, S. H.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Monodisperse FePt Nanoparticles and Ferromagnetic FePt Nanocrystal Superlattices. Science 2000, 287, 1989−1992. (20) Karlsson, H. L.; Cronholm, P.; Gustafsson, J.; Möller, L. Copper Oxide Nanoparticles are Highly Toxic: A Comparison between Metal Oxide Nanoparticles and Carbon Nanotubes. Chem. Res. Toxicol. 2008, 21, 1726−1732. (21) Ozin, G. A.; Asefa, T.; MacLachlan, M. J.; Coombs, N. Periodic Mesoporous Organosilicas with Organic Groups inside the Channel Walls. Nature 1999, 402, 867−871. (22) Hoffmann, F.; Cornelius, M.; Morell, J.; Fröba, M. Silica-Based Mesoporous Organic-Inorganic Hybrid Materials. Angew. Chem., Int. Ed. 2006, 45, 3216−3251. (23) Coe, S.; Woo, W. K.; Bawendi, M.; Bulović, V. Electroluminescence from Single Monolayers of Nanocrystals in Molecular Organic Devices. Nature 2002, 420, 800−803. (24) Dolbecq, A.; Mialane, P.; Lisnard, L.; Marrot, J.; Sécheresse, F. Hybrid Organic-Inorganic 1D and 2D Frameworks with ε-keggin Polyoxomolybdates as Building Blocks. Chem. - Eur. J. 2003, 9, 2914− 2920.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b02034. Additional information regarding 1H NMR and 13C NMR spectra of bfps, TGA data, IR spectra, SEM images, solid-state UV−vis absorption spectra, XPS and UV−vis spectra, cyclic voltammograms, and PXRD patterns (PDF) Accession Codes

CCDC 1554123−1554124 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*Z.-G.R.: e-mail, [email protected]; tel, 86-51265880328. *J.-P.L.: e-mail, [email protected]; fax, 86-512-65880328; tel, 86-512-65882865. ORCID

Jian-Ping Lang: 0000-0003-2942-7385 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank the National Natural Science Foundation of China (21373142, 21531006, 21671144, and 21773163) for financial support. J.-P.L. also appreciates financial support from the Priority Academic Program Development of Jiangsu Higher Education Institutions, the “SooChow Scholar” Program of Soochow University, and the Project of Scientific and Technologic Infrastructure of Suzhou (SZS201708).



REFERENCES

(1) Kickelbick, G. Concepts for the Incorporation of Inorganic Building Blocks into Organic Polymers on a Nanoscale. Prog. Polym. Sci. 2003, 28, 83−114. (2) Judeinstein, P.; Sanchez, C. Hybrid Organic-Inorganic Materials: a Land of Multidisciplinarity. J. Mater. Chem. 1996, 6, 511−525. (3) Kamat, P. V. Photophysical, Photochemical and Photocatalytic Aspects of Metal Nanoparticles. J. Phys. Chem. B 2002, 106, 7729− 7744. 12548

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry (25) Stankovich, S.; Dikin, D. A.; Dommett, G. H. B.; Kohlhaas, K. M.; Zimney, E. J.; Stach, E. A.; Piner, R. D.; Nguyen, S. T.; Ruoff, R. S. Graphene-Based Composite Materials. Nature 2006, 442, 282−286. (26) Batten, S. R. Coordination Polymers. Curr. Opin. Solid State Mater. Sci. 2001, 5, 107−114. (27) Janiak, C. Engineering Coordination Polymers towards Applications. Dalton Trans. 2003, 14, 2781−2804. (28) Zhu, Q. L.; Xu, Q. Metal-Organic Framework Composites. Chem. Soc. Rev. 2014, 43, 5468−5512. (29) Cui, Z. P.; Qi, J.; Xu, X. X.; Liu, L.; Wang, Y. Photocatalytic Activity of PANI Loaded Coordination Polymer Composite Materials: Photoresponse Region Extension and Quantum Yields Enhancement via the Loading of PANI Nanofibers on Surface of Coordination Polymer. J. Solid State Chem. 2013, 205, 142−148. (30) Czaja, A. U.; Trukhan, N.; Müller, U. Industrial Applications of Metal-Organic Frameworks. Chem. Soc. Rev. 2009, 38, 1284−1293. (31) Furukawa, H.; Cordova, K. E.; O'Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. (32) Kitagawa, S.; Kitaura, R.; Noro, S. Functional Porous Coordination Polymers. Angew. Chem., Int. Ed. 2004, 43, 2334−2375. (33) Dou, Y. B.; Zhou, J.; Zhou, A.; Li, J. R.; Nie, Z. R. Visible-Light Responsive MOF Encapsulation of Noble-Metal-Sensitized Semiconductors for High-Performance Photoelectrochemical Water Splitting. J. Mater. Chem. A 2017, 5, 19491. (34) Mishra, A.; Fischer, M. K. R.; Bäuerle, P. Metal-Free Organic Dyes for Dye-Sensitized Solar Cells: From Structure: Property Relationships to Design Rules. Angew. Chem., Int. Ed. 2009, 48, 2474−2499. (35) Law, M.; Greene, L. E.; Johnson, J. C.; Saykally, R.; Yang, P. Nanowire Dye-Sensitized Solar Cells. Nat. Mater. 2005, 4, 455−459. (36) Zheng, J. P.; Ou, S.; Zhao, M.; Wu, C. D. A Highly Sensitive Luminescent Dye@MOF Composite for Probing Different Volatile Organic Compounds. ChemPlusChem 2016, 81, 758. (37) Xie, W.; He, W. W.; Li, S. L.; Shao, K. Z.; Su, Z. M.; Lan, Y. Q. An Anionic Interpenetrated Zeolite-Like Metal-Organic Framework Composite as a Tunable Dual-Emission Luminescent Switch for Detecting Volatile Organic Molecules. Chem. - Eur. J. 2016, 22, 17298. (38) Cui, Y. J.; Song, R. J.; Yu, J. C.; Liu, M.; Wang, Z. Q.; Wu, C. D.; Yang, Y.; Wang, Z. Y.; Chen, B. L.; Qian, G. D. Dual-Emitting MOF⊃Dye Composite for Ratiometric Temperature Sensing. Adv. Mater. 2015, 27, 1420−1425. (39) Liu, Y.; Lin, Q. P.; Zhang, Q. C.; Bu, X. H.; Feng, P. Y. VisibleLight-Driven, Tunable, Photoelectrochemical Performance of a Series of Metal-Chelate, Dye-Organized, Crystalline, CdS Nanoclusters. Chem. - Eur. J. 2014, 20, 8297−8301. (40) Ncibi, M. C.; Mahjoub, B.; Seffen, M. Kinetic and Equilibrium Studies of Methylene Blue Biosorption by Posidonia Oceanica (L.) Fibres. J. Hazard. Mater. 2007, 139, 280−285. (41) Lv, L. L.; Yang, J.; Zhang, H. M.; Liu, Y. Y.; Ma, J. F. Metal-Ion Exchange, Small-Molecule Sensing, Selective Dye Adsorption, and Reversible Iodine Uptake of Three Coordination Polymers Constructed by a New Resorcin[4]arene-Based Tetracarboxylate. Inorg. Chem. 2015, 54, 1744−1755. (42) Liu, A. R.; Zhou, W.; Shen, K. L.; Liu, J.; Zhang, X. X. One-Pot Hydrothermal Synthesis of Hematite-Reduced Graphene Oxide Composites for Efficient Removal of Malachite Green from Aqueous Solution. RSC Adv. 2015, 5, 17336. (43) Long, J. R.; Yaghi, O. M. The Pervasive Chemistry of MetalOrganic Frameworks. Chem. Soc. Rev. 2009, 38, 1213−1214. (44) Zhang, J.; Gondal, M. A.; Wei, W.; Zhang, T. N.; Xu, Q. Y.; Shen, K. Preparation of Room Temperature Ferromagnetic BiFeO3 and its Application as an Highly Efficient Magnetic Separable Adsorbent for Removal of Rhodamine B from Aqueous Solution. J. Alloys Compd. 2012, 530, 107−110. (45) Luo, P.; Zhang, B.; Zhao, Y. F.; Wang, J. H.; Zhang, H. Q.; Liu, J. D. Removal of Methylene Blue from Aqueous Solutions by Adsorption onto Chemically Activated Halloysite Nanotubes. Korean J. Chem. Eng. 2011, 28, 800−807.

(46) Liu, H. Y.; Wu, H.; Ma, J. F.; Song, S. Y.; Yang, J.; Liu, Y. Y.; Su, Z. M. Structural Study of Silver(I) Sulfonate Complexes with Pyrazine Derivatives. Inorg. Chem. 2007, 46, 7299−7311. (47) Lu, W. G.; Wei, Z. W.; Gu, Z. Y.; Liu, T. F.; Park, J.; Park, J.; Tian, J.; Zhang, M. W.; Zhang, Q.; Gentle, T.; Bosch, M.; Zhou, H. C. Tuning the Structure and Function of Metal-Organic Frameworks via Linker Design. Chem. Soc. Rev. 2014, 43, 5561−5593. (48) Xie, Y. B.; Yang, H.; Wang, Z. U.; Liu, Y. Y.; Zhou, H. C.; Li, J. R. Unusual Preservation of Polyhedral Molecular Building Units in a Metal-Organic Framework with Evident Desymmetrization in Ligand Design. Chem. Commun. 2014, 50, 563−565. (49) Crini, G. Non-Conventional Low-Cost Adsorbents for Dye Removal: A Review. Bioresour. Technol. 2006, 97, 1061−1085. (50) Guo, H. X.; Lin, F.; Chen, J. H.; Li, F. M.; Weng, W. MetalOrganic Framework MIL-125(Ti) for Efficient Adsorptive Removal of Rhodamine B from Aqueous Solution. Appl. Organomet. Chem. 2015, 29, 12−19. (51) Du, P. Y.; Li, H.; Fu, X.; Gu, W.; Liu, X. A 1D Anionic Lanthanide Coordination Polymer as an Adsorbent Material for the Selective Uptake of Cationic Dyes from Aqueous Solutions. Dalton Trans. 2015, 44, 13752−13759. (52) Lin, S.; Song, Z. L.; Che, G. B.; Ren, A.; Li, P.; Liu, C. B.; Zhang, J. S. Adsorption Behavior of Metal-Organic Frameworks for Methylene Blue from Aqueous Solution. Microporous Mesoporous Mater. 2014, 193, 27−34. (53) Tan, Y. X.; Zhang, Y.; He, Y. P.; Zheng, Y. J.; Zhang, J. Multifunctional Anionic MOF Material for Dye Enrichment and Selective Sorption of C2 Hydrocarbons over Methane via Ag+Exchange. Inorg. Chem. 2014, 53, 12973−12976. (54) Haque, E.; Jun, J. W.; Jhung, S. H. Adsorptive Removal of Methyl Orange and Methylene Blue from Aqueous Solution with a Metal-Organic Framework Material, Iron Terephthalate (MOF-235). J. Hazard. Mater. 2011, 185, 507−511. (55) Haque, E.; Lee, J. E.; Jang, I. T.; Hwang, Y. K.; Chang, J. S.; Jegal, J.; Jhung, S. H. Adsorptive Removal of Methyl Orange from Aqueous Solution with Metal-Organic Frameworks, Porous Chromium-Benzenedicarboxylates. J. Hazard. Mater. 2010, 181, 535−542. (56) Gao, C. Y.; Yang, Y.; Ai, J.; Tian, H. R.; Li, L. J.; Yang, W. T.; Dang, S.; Sun, Z. M. A Multifunctional MnII Phosphonate for Rapid Separation of Methyl Orange and Electron-Transfer Photochromism. Chem. - Eur. J. 2016, 22, 11652−11659. (57) Xiao, L.; Xiong, Y.; Wen, Z. H.; Tian, S. H. Adsorption Behaviour of One-Dimensional Coordination Polymers [Cu(bipy)X]n (X = 2Cl− and SO42−) toward Acid Orange 7 and Methyl Orange. RSC Adv. 2015, 5, 61593−61600. (58) Yan, W.; Han, L. J.; Jia, H. L.; Shen, K.; Wang, T.; Zheng, H. G. Three Highly Stable Cobalt MOFs Based on ″Y″-Shaped Carboxylic Acid: Synthesis and Absorption of Anionic Dyes. Inorg. Chem. 2016, 55, 8816−8821. (59) Zhang, Q.; Yu, J. C.; Cai, J. F.; Song, R. J.; Cui, Y. J.; Yang, Y.; Chen, B. L.; Qian, G. D. A Porous Metal-Organic Framework with -COOH Groups for Highly Efficient Pollutant Removal. Chem. Commun. 2014, 50, 14455−14458. (60) He, Y. C.; Yang, J.; Kan, W. Q.; Zhang, H. M.; Liu, Y. Y.; Ma, J. F. A New Microporous Anionic Metal-Organic Framework as a Platform for Highly Selective Adsorption and Separation of Organic Dyes. J. Mater. Chem. A 2015, 3, 1675−1681. (61) Qin, J. S.; Zhang, S. R.; Du, D. Y.; Shen, P.; Bao, S. J.; Lan, Y. Q.; Su, Z. M. A Microporous Anionic Metal-Organic Framework for Sensing Luminescence of Lanthanide(III) Ions and Selective Absorption of Dyes by Ionic Exchange. Chem. - Eur. J. 2014, 20, 5625−5630. (62) Zhu, Z. F.; Bai, Y. L.; Zhang, L. L.; Sun, D. F.; Fang, J. H.; Zhu, S. R. Two Nanocage Anionic Metal-Organic Frameworks with Rht Topology and {[M(H2O)6]6}12+ Charge Aggregation for Rapid and Selective Adsorption of Cationic Dyes. Chem. Commun. 2014, 50, 14674−14677. (63) Wang, X. S.; Liang, J.; Li, L.; Lin, Z. J.; Bag, P. P.; Gao, S. Y.; Huang, Y. B.; Cao, R. An Anion Metal-Organic Framework with Lewis 12549

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550

Article

Inorganic Chemistry Basic Sites-Rich Toward Charge-Exclusive Cationic Dyes Separation and Size-Selective Catalytic Reaction. Inorg. Chem. 2016, 55, 2641− 2649. (64) Wu, B.; Zhang, W. H.; Ren, Z. G.; Lang, J. P. A 1D Anionic Coordination Polymer Showing Superior Congo Red Sorption and its Dye Composite Exhibiting Remarkably Enhanced Photocurrent Response. Chem. Commun. 2015, 51, 14893−14896. (65) Chen, Y.; Yang, Z.; Guo, C. X.; Ni, C. Y.; Li, H. X.; Ren, Z. G.; Lang, J. P. Using Alcohols as Alkylation Reagents for 4Cyanopyridinium and N,N′-Dialkyl-4,4′-Bipyridinium and their OneDimensional Iodoplumbates. CrystEngComm 2011, 13, 243−250. (66) Chen, M. M.; Chen, L.; Li, H. X.; Brammer, L.; Lang, J. P. Highly Selective Detection of Hg2+ and MeHgI by Di-pyridin-2-yl-[4(2-pyridin-4-yl-vinyl)-phenyl]-amine and its Zinc Coordination Polymer. Inorg. Chem. Front. 2016, 3, 1297−1305. (67) Sheldrick, G. M. SADABS Empirical Absorption Correction Program; University of Göttingen, Göttingen, Germany; based on the method of Blessing. (68) Krause, L.; Herbst-Irmer, R.; Sheldrick, G. M.; Stalke, D. Comparison of Silver and Molybdenum Microfocus X-ray Sources for Single-Crystal Structure Determination. J. Appl. Crystallogr. 2015, 48, 3−10. (69) Blessing, R. H. An Empirical Correction for Absorption Anisotropy. Acta Crystallogr., Sect. A: Found. Crystallogr. 1995, 51, 33−38. (70) Sheldrick, G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (71) Sarkar, M.; Biraha, K. Amide-to-Amide Hydrogen Bonds in the Presence of a Pyridine Functionality: Crystal Structures of Bis(pyridinecarboxamido)alkanes. Cryst. Growth Des. 2006, 6, 202−208. (72) Muthu, S.; Yip, J. H. K.; Vittal, J. J. Coordination Polymers of d10 Metals and N,N′-bis(3-pyridinecarboxamide)-1,2-ethane. J. Chem. Soc., Dalton Trans. 2001, 24, 3577−3584. (73) Richardson, C.; Steel, P. J. A Ribbon-Like Metallopolymer from 1,4-bis (2-pyridyl) buta- 1,3-diyne and Silver(I) Nitrate. Inorg. Chem. Commun. 1998, 1, 260−262. (74) Bowmaker, G. A.; Effendy; Lim, K. C.; Skelton, B. W.; Sukarianingsih, D.; White, A. H. Syntheses, Structures and Vibrational Spectroscopy of Some 1:2 and 1:3 Adducts of Silver(I) Oxyanion Salts with Pyridine and Piperidine Bases Containing Non-Coordinating 2(,6)-substituents. Inorg. Chim. Acta 2005, 358, 4342−4370. (75) Pawley, G. S. Unit-Cell Refinement from Powder Diffraction Scans. J. Appl. Crystallogr. 1981, 14, 357−361. (76) Coelho, A. A. TPOAS-Academic V4.1; Coelho Software, Oct 2007. (77) Wang, J. F.; Liu, S. Y.; Liu, C. Y.; Ren, Z. G.; Lang, J. P. Silver(I) Complexes with a P-N Hybrid Ligand and Oxyanions: Synthesis, Structures, Photocatalysis and Photocurrent Responses. Dalton Trans. 2016, 45, 9294−9306. (78) Volksen, W.; Miller, R. D.; Dubois, G. Low Dielectric Constant Materials. Chem. Rev. 2010, 110, 56−110. (79) Fröhlich, H. Theory of Dielectrics: Dielectric Constant and Dielectric Loss; Clarendon Press: Oxford, U.K., 1958. (80) Böttcher, C. J. F.; van Belle, O. C.; Bordewijk, P.; Rip, A. Theory of Electric Polarization; Elsevier Scientific: Amsterdam, 1978. (81) Mendiratta, S.; Usman, M.; Luo, T. T.; Lee, S. F.; Lin, Y. C.; Lu, K. L. Guest Dependent Dielectric Properties of Nickel(II)-Based Supramolecular Networks. CrystEngComm 2014, 16, 6309−6315.

12550

DOI: 10.1021/acs.inorgchem.7b02034 Inorg. Chem. 2017, 56, 12542−12550