Preparation, Crystal Structures, and Properties of a Series of

Jul 11, 2018 - ... with the lanthanide ions in two kinds of coordination geometries, i.e., eightfold square antiprism and ninefold monocapped square a...
0 downloads 0 Views 3MB Size
Article Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

pubs.acs.org/crystal

Preparation, Crystal Structures, and Properties of a Series of Crystalline Tetra(4-sulfonatophenyl)porphyrinato Histidine 4f-3d Porphyrinic Compounds Wen-Tong Chen,*,†,‡ Zhuan-Xia Zhang,† Long-Zhen Lin,† Yan Sui,† Dong-Sheng Liu,† and Hua-Long Chen†

Downloaded via NORTHWESTERN UNIV on July 11, 2018 at 17:06:24 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



Institute of Applied Chemistry, Jiangxi Province Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Jinggangshan University, 343009, Ji’an, Jiangxi, P. R. China ‡ Key Laboratory of Jiangxi Province for Persistant Pollutants Control and Resources Recycle, Nanchang Hangkong University, 330000, Nanchang, Jiangxi, China S Supporting Information *

ABSTRACT: A series of novel 4f-3d crystalline porphyrinic compounds, {[Co(TPPS)] 2 [Ln(Histidine)(H 2 O)][Ln(H3O)3]}n·nH2O (Ln = Sm (1), Eu (2), Dy (3); TPPS = 5,10,15,20-tetra(4-sulfonatophenyl)porphyrinato) with TPPS and histidine as mixed ligands, have been prepared and characterized by single-crystal X-ray diffraction technique. Complexes 1−3 are isomorphous and characteristic of a threedimensional (3D) framework with the lanthanide ions in two kinds of coordination geometries, i.e., eightfold square antiprism and ninefold monocapped square antiprism. The porphyrin macrocycles adopt a saddle-distorted nonplanar conformation with an embedded cobalt ion binding to four pyrrole nitrogen atoms. As revealed by the photoluminescence measurement, they exhibit blue emission. Variable-temperature magnetic susceptibility reveals that complexes 1−3 are antiferromagnetic. Their FTIR, CV, DPV, and UV/vis results are also studied in detail.



INTRODUCTION Porphyrins are natural bioactive substances, and they have important roles in green leaves and red blood cells. Due to their common characteristics of redox, intensive optical absorption and emission, as well as long lifetime of the singlet and triplet states,1−5 porphyrins may also provide many useful applications in medicines, catalysts, dyes, solar cells, organic light-emitting diodes, and so on.6−9 It is well-known that amino acids are also natural bioactive substance and over 500 kinds of amino acids have been found in nature. Amino acids possess both amido and carboxylic groups which enable them to coordinate to metal ions to prepare coordination complexes. A large number of coordination complexes containing amino acids have so far been reported.10−13However, lanthanide complexes containing amino acids are still rare,14,15 especially lanthanide complexes containing both amino acids and porphyrins as ligands. The physicochemical performance of porphyrinic compounds can be easily adjusted by outer substituent groups as well as the coordination of metal ions at the center of the porphyrinic macrocycle. Furthermore, the insertion of a metal atom at the center and the modification of the periphery of a porphyrin with different functional groups can provide rich synthetic elements for preparing porphyrinic materials. There© XXXX American Chemical Society

fore, porphyrins nowadays attract a lot of attention and have become a research hotspot in material science. At present, many kinds of porphyrins such as meso-tetra(4carboxyphenyl)porphyrin (TCPP),16,17 meso-tetrakis(4-methoxyphenyl)-21H, 23H-porphyrin (TMeOPP),18−20 mesotetraphenyl-21H,23H-porphyrin(TPP),21−23 meso-tetrakis(4aminophenyl)porphyrin (TAPP),24,25 and 5-(3-hydroxyphenyl)-10,15,20-triphenylporphyrin (TPPOH)26−28 have been applied to synthesis of novel materials. However, most of them are water insoluble and only dissolved in different organic solvents. In contrast, 5,10,15,20-tetra(4sulfonatophenyl)porphyrinato (TPPS) is water-soluble, which enables it to react with metal salts in an eco-friendly way, i.e., in water. Moreover, from the perspective of crystal engineering, TPPS is a versatile ligand because it features a lot of coordination sites (the porphyrinic macrocycle center and 12 Osulfonic group atoms). So, TPPS can possibly bind to both transition metal atoms and lanthanide atoms at the center and the periphery, respectively, to achieve a high-dimension coordination complex. To the best of our knowledge, among Received: May 30, 2018 Revised: June 25, 2018

A

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Table 1. Crystal Parameters of Complexes 1, 2, and 3 compounds

1

2

3

formula Fw color crystal size/mm3 crystal system space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z 2θmax (deg) reflections collected independent, observed reflections (Rint) dcalcd. (g/cm3) μ(mm−1) T (K) F(000) R1, wR2 S largest and mean Δ/σ Δρ(max, min) (e/Å3)

C94H69Co2N11O31S8Sm2 2523.64 purple 0.24 0.12 0.10 triclinic P1̅ 15.187(5) 19.471(6) 19.569(6) 68.422(4) 84.741(4) 79.439(3) 5288(3) 2 50 31791 17732, 6813 (0.0307) 1.585 1.642 293(2) 2528 0.0836, 0.2104 0.965 0.002, 0 2.182, −1.240

C94H69Co2Eu2N11O31S8 2526.86 purple 0.22 0.13 0.11 triclinic P1̅ 15.162(6) 19.466(6) 19.573(6) 68.395(4) 84.898(4) 79.243(4) 5276(3) 2 50 33053 17641, 7253 (0.0295) 1.591 1.722 293(2) 2532 0.0807, 0.2398 0.945 0.004, 0 2.616, −1.647

C94H69Co2Dy2N11O31S8 2547.94 purple 0.20 0.12 0.10 triclinic P1̅ 15.082(4) 19.370(3) 19.508(2) 68.463(2) 84.2320(10) 78.636(2) 5194.5(16) 2 50 32650 17386, 9897 (0.0360) 1.629 1.980 123.15 2544 0.0676, 0.1549 0.985 0.005, 0 1.751, −1.100

pulse voltammetry) were carried out at room temperature on a BAS 100 W electrochemical analyzer in a deaerated solvent containing tetra-n-butylammonium hexafluorophosphate (TBAPF6; 0.10 M) as a supporting electrolyte. A conventional three-electrode cell was applied with a carbon working electrode and a platinum wire as a counter electrode. The measured potentials were recorded with respect to the Ag/AgNO3 (0.01 M). All electrochemical measurements were conducted under an atmospheric pressure of argon. Syntheses. SmCl3·6H2O (A.R., Ji’Nan Henghua), EuCl3·6H2O (A.R., Ji’Nan Henghua), DyCl3·6H2O (A.R., Ji’Nan Henghua), CoCl2·6H2O (A.R., Ji’Nan Henghua), TPPS (>85.0%, TCI), and Lhistidine (>99.0%, TCI) were obtained commercially and directly used without further purification. The synthesis of {[Co(TPPS)]2[Sm(Histidine)(H2O)][Sm(H3O)3]}n·nH2O (1): Complex 1 was prepared by loading SmCl3· 6H2O (1 mmol, 365 mg), CoCl2·6H2O (1 mmol, 238 mg), TPPS (1 mmol, 940 mg), L-histidine (1 mmol, 155 mg), and 10 mL distilled water into a Teflon-lined stainless steel vessel (25 mL); then, the vessel was heated to 473 K and held for 10 days. After the vessel was cooled slowly down to room temperature, purple block-like crystals suitable for X-ray analysis were collected. Yield: 26% (based on samarium). Anal. Calcd for C94H69Co2N11O31S8Sm2: C 44.79, H 2.64, N 6.11. Found: C 43.83, H 2.67, N 6.25. Fourier transform IR (KBr, cm−1): 3425(s), 3100(w), 2862(w), 2366(w), 1939(w), 1848(w), 1689(m), 1599(s), 1557(m), 1497(m), 1455(m), 1395(m), 1346(s), 1236(s), 1171(s), 1126(vs), 1047(s), 1006(vs), 854(m), 809(s), 741(s), 718(m), 669(w), and 639(s). The synthesis of {[Co(TPPS)] 2 [Eu(Histidine)(H 2 O)][Eu(H3O)3]}n·nH2O (2): Complex 2 was prepared according to the same prcedure of 1, except for using EuCl3·6H2O (1 mmol, 367 mg) instead of SmCl3·6H2O. Yield: 22% (based on europium). Anal. Calcd for C94H69Co2Eu2N11O31S8: C 44.73, H 2.64, N 6.10. Found: C 44.29, H 2.70, N 6.34. Fourier transform IR (KBr, cm−1): 3436(s), 3103(w), 2854(w), 1943(w), 1845(w), 1690(w), 1598(m), 1556(w), 1495(w), 1452(w), 1395(m), 1347(m), 1239(s), 1175(s), 1127(vs), 1051(s), 1006(vs), 856(w), 809(m), 744(s), 717(m), and 638(s). The synthesis of {[Co(TPPS)]2[Dy(Histidine)(H2O)][Dy(H3O)3]}n·nH2O (3): Complex 3 was prepared according to the

the large amount of porphyrinic coordination complexes, those containing TPPS are relatively rare.29 Moreover, lanthanide TPPS complexes are especially rare.30 Therefore, it is desirable to investigate the crystal structure and physicochemical behavior of lanthanide complexes containing both amino acid and TPPS as mixed ligands. Considering the fact that a large amount of amino acids and porphyrins with various functional groups can offer very versatile synthetic elements for preparing different materials, our group in recent years has mainly aimed at the study of the crystal structure and physicochemical behavior of lanthanide complexes containing both amino acid and TPPS as mixed ligands. In the present work, we report the hydrothermal syntheses, crystal structures, and photophysical and electrochemical properties of a series of novel 4f-3d crystalline porphyrinic compounds, {[Co(TPPS)] 2 [Ln(Histidine)(H2O)][Ln(H3O)3]}n·nH2O (Ln = Sm (1), Eu (2), Dy (3); TPPS = 5,10,15,20-tetra(4-sulfonatophenyl)porphyrinato).



EXPERIMENTAL SECTION

Measurements. In order to check the bulk-phase purity of the samples, elemental analysis for carbon, hydrogen, and nitrogen was performed on an ELEMENTAR VARIO EL CUBE microanalyzer. The FT-IR spectra were measured on a Thermo Fisher Nicolet 6700 FT−IR spectrophotometer in the frequency span of 4000−400 cm−1 with the KBr pellet technique. The UV−vis absorption spectra with solution sample were carried out at room temperature on a computercontrolled PerkinElmer Lambda900 UV−vis spectrometer in the wavelength span of 190−1100 nm. Photoluminescence spectra were carried out on an Edinburgh Instruments FLS980 UV/vis/NIR fluorescence spectrometer. Variable-temperature magnetic susceptibility and field dependence magnetization measurements of the title complexes with polycrystalline samples were performed on a MPMS Quantum Design SQUID magnetometer. All of the magnetic data were corrected for diamagnetism estimated from Pascal’s constants. Measurements of CV (cyclic voltammetry) and DPV (differential B

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 1. (a) ORTEP drawing with 30% thermal ellipsoids and (b) wire representation shows the saddle-distorted conformation of 1. Hydrogen atoms were omitted for clarity. procedure for 1, except for using DyCl3·6H2O (1 mmol, 377 mg) instead of SmCl3·6H2O. Yield: 27% (based on dysprosium). Anal. Calcd for C94H69Co2Dy2N11O31S8: C 44.36, H 2.61, N 6.05. Found: C 44.15, H 2.65, N 6.09. Fourier transform IR (KBr, cm−1): 3417(s), 3099(w), 2857(w), 2367(w), 1937(w), 1847(w), 1693(m), 1650(m), 1597(m), 1558(m), 1494(w), 1459(w), 1398(m), 1348(m), 1236(s), 1165(s), 1128(vs), 1042(s), 1003(vs), 859(w), 810(m), 744(s), 718(w), 667(w), and 638(s). X-ray Crystallographic Studies. The diffraction data was collected on a Rigaku Mercury CCD X-ray diffractometer with a carefully selected single crystal of complexes 1, 2, and 3. The X-ray source is graphite monochromated Mo Kα radiation with λ being 0.71073 Å. The reduction and empirical absorption correction of the diffraction data were carried out with the CrystalClear software.31 The crystal structures are solved by means of the direct method and the Siemens SHELXTL V5 software32 and refined with a full-matrix leastsquares refinement on F2. All non-hydrogen atoms were located on the difference Fourier maps and applied anisotropic refinement. The hydrogen atoms, except for those of the water molecules, were theoretically attached to their parent atoms and included in the structural factor calculations with assigned isotropic thermal parameters. The crystal data as well as the details of the data collection and refinement are given in Table 1, while the selected bond distances and angles are listed in Table S1. CCDC-1815115 (1), 1815116 (2), and 1815119 (3) contains the supplementary crystallographic data for this paper. These data can be obtained free

of charge from The Cambridge Crystallographic Data Center via http://www.ccdc.cam.ac.uk/data_request/cif.



RESULTS AND DISCUSSION Crystal Structures. {[Co(TPPS)]2[Sm(Histidine)(H2O)][Sm(H3O)3]}n·nH2O (1). Single-crystal X-ray diffraction analysis discovers that complexes 1−3 are isomorphous and, therefore, only the crystal structure of complex 1 is discussed here in detail. The crystal structure of complex 1 is presented as an ORTEP drawing in Figure 1. The crystal structure of complex 1 is composed of neutral {[Co(TPPS)]2[Sm(Histidine)(H2O)][Sm(H3O)3]}n framework and lattice water molecules. Complexes 1−3 are crystallized in the space group of P1̅ of the triclinic system with two formular units in each cell. The divalent cobalt ions adopt a distorted square geometry without an axial ligation and locate at the centers of the porphyrin macrocycles. The Sm1 ion coordinates to eight oxygen atoms of which five are from five sulfonic groups, two are from two histidine ligands, and one is from one coordinating water molecule, resulting in a square antiprism. Differently, the Sm2 ion binds to nine oxygen atoms of which five are from four sulfonic groups, one is from one histidine ligand, and three are from three coordinating water molecules, yielding a ninefold monocapped square antiprism. The bond distances from the cobalt ions to the pyrrole nitrogen atoms are in the range of C

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

1.946(6)−1.992(6) Å in complex 1, which are normal and comparable with those reported in the Cambridge Structural Database.33 The bond lengths of Sm−O locate in the range of 2.325(5)−2.707(8) Å which are normal and comparable with those documented for related species in the references.34 In complex 1 the bond angles of O−Sm−O are in a wide span of 56.5(2)−145.08(19)°, while those of N−Co−N are in the range 88.6(2)−176.8(2)°. The porphyrinic macrocycles exhibit a saddle-distorted conformation with four pyrrole rings slightly distorted in an alternate mode either upward or downward relative to the average plane of the saddle-like porphyrinic macrocycle. The deviation of the atoms of the 24member porphyrinic macrocycle is from −0.397 to 0.440 Å. The displacement of the pyrrole nitrogen atoms is in the range between −0.071 and 0.071 Å apart from their average N4 plane. As for complex 1, the dihedral angles between the pyrrole rings distorted in the same direction are 16.50°, 19.15°, 19.27°, and 22.60°, respectively, while those between the neighboring pyrrole rings are 13.73°, 15.37°, 8.87°, 12.68°, 12.64°, 16.12°, 12.78°, and 17.20°, respectively. With respect to the N4 plane that probably represent the average porphyrinic macrocycle plane, the twisted angles of the phenyl rings are 61.19°, 76.26°, 59.06°, 70.06°, 65.34°, 88.73°, 57.29°, and 75.79°. Interestingly, the neighboring Sm1 ions are linked by two histidine molecules, while the neighboring Sm2 ions are linked by two sulfonic groups. The histidine molecule acts as a μ2-bridging ligand to interconnect two neighboring samarium ions with the distance of the closest Sm···Sm being 4.8056(9) Å. Each Coporphyrin connects to five samarium ions. However, the Sm1 ion links to five Co-porphyrins and the Sm2 ion links to four Co-porphyrins. The Co-porphyrins are interlinked by the samarium ions to complete a 3-D framework, as presented in Figure 2. Additionally, in complex 1 there are abundant N− H···O, C−H···N, and C−H···O interactions. Steady-State Absorption and Emission Studies. To the best of our knowledge, metalloporphyrinic complexes can usually display two types of absorption bands, i.e., intensive B band (namely, Soret band) locating at about 400 nm with the value of ε (namely, molar absorption coefficient) being about 105 M−1 cm−1 and some weak Q bands residing at about 500− 650 nm with the ε values being about 103−104 M−1 cm−1. The UV−vis absorption spectra for complexes 1−3 and free base TPPS are presented in Figure 3. The B band of free base TPPS is found at 434 nm, while its four Q bands are found at 515, 552, 598, and 645 nm, respectively. The B bands are 432, 433, and 432 nm for complexes 1−3, respectively, which is slightly blue-shifted by 1−2 nm as compared with the counterpart of free base TPPS. Complexes 1−3 have only one Q-band which is located at 549 nm. The difference of the Q-band number between complexes 1−3 and free base TPPS can be attributed to the increase of the molecular symmetry that is originated from the metalation of the porphyrinic macrocycle center of the free base TPPS. The absorption coefficients of the B band for metalloporphyrinic complexes 1−3 are around 105 M−1 cm−1, which is in accordance with the conclusion drawn by Gouterman.35 Generally porphyrins and porphyrinic derivatives in the solid state cannot exhibit photoluminescence emission due to socalled concentration quenching effect.36 However, they can usually display photoluminescence emission when they are dissolved into solvent. It is well-known that lanthanide complexes (especially those containing samarium, europium,

Figure 2. Packing diagram in wire representation of 1 with the dashed lines representing hydrogen bonding interactions. Geometry of represented hydrogen bonding interactions: N9−H9B···O5W with dN···O = 2.796(10) Å and ∠(DHA) = 103°; N9−H9B···O20(2 − x, 2 − y, −2 − z) with dN···O = 2.875(10) Å and ∠(DHA) = 121°; C26− H26A···N3(1 − x, 3 − y, −4 − z) with dC···N = 3.378(10) Å and ∠(DHA) = 141°; C29−H29A···O15(−1 + x, y, −1 + z) with dC···O = 3.433(10) Å and ∠(DHA) = 168°; C72−H72A···N5(2 − x, 2 − y, −2 − z) with dC···N = 3.324(10) Å and ∠(DHA) = 141°; C87−H87A··· O5W(2 − x, 2 − y, −2 − z) with dC···O = 3.387(10) Å and ∠(DHA) = 153°; C91−H91B···O5(2 − x, 2 − y, −2 − z) with dC···O = 3.249(13) Å and ∠(DHA) = 141°.

terbium, and dysprosium) can usually show interesting photoluminescent behavior because they have rich f-orbital configurations. Therefore, the title complexes are supposed to display photoluminescent behaviors. Based on these considerations and in order to validate this point, in this work, we measured the photoluminescent properties of complexes 1−3 and free base TPPS in solution at room temperature and the results are presented in Figure 4. The effective energy absorption is almost located in the wavelength range of 300−400 nm for complexes 1−3, while that for free base TPPS is resided at the wavelength range of 400−470 nm. With regard to free base TPPS, its photoluminescence emission spectrum displays a band at 679 nm. However, the photoluminescence spectra of complexes 1−3 exhibit only one emission band at 467, 469, and 465 nm for complexes 1−3, respectively. Obviously the photoluminescence spectra of complexes 1−3 are not same as that of free base TPPS. To the best of our knowledge, the characteristic emission bands of lanthanide ions are located at around 567, 602, and 650 nm for Sm3+ (4G5/2 → 6HJ (J = 5/2, 7/2, 9/2)),37 ∼579, 592, 612, and 653 nm for Eu3+ (5D0 → 7FJ (J = 1, 2, 3, 4)),38 and ∼575 nm for Dy3+ (4F9/2 → 6H13/2).39 As a result, it is clear that the photoluminescence spectra of complexes 1−3 do not result from the lanthanide ions. So, we finally measured the photoluminescence emission spectrum of histidine. As shown in the inset of Figure 4, the photoluminescence emission spectrum of histidine shows one emission band at 464 nm, which is very similar to that of complexes 1−3. Therefore, the photoluminescence spectra of complexes 1−3 result from the histidine ligand. Magnetic Behaviors. The variable-temperature (between 2 and 300 K) magnetic susceptibility for complexes 1−3 has been carried out with crystalline samples under an applied magnetic field of 5000 Oe. The diagrams of χM vs T and μeff vs T of complex 1 are given in Figure 5a, of which the χM is the D

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 3. (a) UV−vis absorption of 1, 2, and 3 in deaerated DMSO at room temperature. (b) UV−vis absorption of free base TPPS.

Figure 5. Thermal dependence of χM and μeff for 1 (a), 2 (b), and 3 (c). The red lines correspond to the theoretical line based on the Curie−Weiss model.

the existence of antiferromagnetic behavior in complex 1. Magnetic behaviors of lanthanide porphyrins with antiferromagnetic interactions have been reported previously.40 The curves of χM vs T and μeff vs T for complex 2 are presented in Figure 5b, of which the χM means the magnetic susceptibility for per Eu2Co2 unit (Eu3+ ion: S = 1 with ground state 7F1 and low-spin Co2+ ions: S = 1/2). The μeff at 300 K is 2.34 μB, which is smaller than the expected value of 2.45 μB for noninteracting free ions per Eu2Co2 unit. As the temperature is decreased from 300 to 2 K, the μeff slowly decreases to a value of 0.83 μB at 2 K. The data of χM vs T were analyzed by using

Figure 4. Photoluminescence spectra of 1, 2, 3, and free base TPPS (Inset: histidine).

magnetic susceptibility for per Sm2Co2 unit (Sm3+ ion: S = 5/2 with ground state 6H5/2 and low-spin Co2+ ions: S = 1/2). The μeff at 300 K is 2.68 μB, which is close to the expected value of 2.72 μB for noninteracting free ions per Sm2Co2 unit. When the temperature is lowered from 300 to 2 K, the μeff slowly decreases to a value of 2.16 μB at 2 K. The data of χM vs T were analyzed using the well-known Curie−Weiss model to yield C = 0.59 K and a negative Weiss constant θ = −0.06 K, indicating E

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

the Curie−Weiss law to yield C = 0.09 K and a negative Weiss constant θ = −0.02 K, suggesting the existence of an antiferromagnetic interaction in complex 2. The figure of χM vs T and μeff vs T for complex 3 are depicted in Figure 5c, of which the χM means the magnetic susceptibility for per Dy2Co2 unit (Dy3+ ion: S = 13/2 with ground-state 6H13/2 and low-spin Co2+ ions: S = 1/2). The μeff at 300 K is 15.11 μB, which is slightly smaller than the expected value of 15.19 μB for noninteracting free ions per Dy2Co2 unit. As the temperature is decreased from 300 to 2 K, the μeff slowly decreases to a value of 12.56 μB at 2 K. The data of χM vs T were analyzed by using the Curie−Weiss model to give C = 21.08 K and a negative Weiss constant θ = −0.15 K, indicating the existence of antiferromagnetic behavior in complex 3. The essence of the antiferromagnetic-like behavior of complexes 1−3 remains to be clarified, but it is probably ascribed to the progressive thermal depopulation of the Stark components of the Ln3+ ions,41 anisotropy of the Co2+ ions, and a possible interaction between the Ln3+ and Co2+ ions. As for the Ln3+ ion, its 4f orbitals are well-shielded by outer shells of the 5s and 5p orbitals, so the interaction between the Ln3+ and Co2+ ions is supposed to be very small and negligible. We also measured the field dependence of the magnetization for complexes 1−3 under 2 K, as shown in Figure 6. The magnetization of complex 1 increases slowly with the increased field (Figure 6a). The M vs H diagram reveals that the magnetization of complex 1 is still not saturated up to 80 000 Oe with a value of about 0.06 Nβ. The M vs H curve reveals that complex 1 has a small coercive field of around 35 Oe and a small remnant magnetization of about 1.1 × 10−4 Nβ. The magnetization of complex 2 slowly increases with the increased field, as given in Figure 6b. The M vs H curve of complex 2 shows that the magnetization is still not saturated up to 80 000 Oe with a value of ∼0.03 Nβ. The M vs H curve reveals that complex 2 possesses a very small coercive field of about 8 Oe and a small remnant magnetization of about 1.0 × 10−5 Nβ. The magnetization of complex 3 first increases abruptly with the increased field and then slowly (Figure 6c). The M vs H curve of complex 3 reveals that the magnetization is almost saturated at 80 000 Oe with a value of about 0.97 Nβ. The M vs H curve reveals that complex 3 features a small coercive field of around 32 Oe and a small remnant magnetization of about 4 × 10−3 Nβ. Such small coercive field and remnant magnetization are normal and comparable with those found in the known compounds.42,43 Electrochemical Performance. The electrochemical measurements were conducted by using cyclic voltammetry (CV) and differential pulse voltammetry (DPV) techniques with a glassy carbon electrode in the solution of DMF containing tetra-n-butylammonium hexafluorophosphate (TBAPF6; 0.10 M) under room temperature. According to the conclusion drawn by Kadish and his colleagues,44 the factors which affect the metalloporphyrin redox potentials mainly include the species of supporting electrolytes/solvents and the nature of porphyrin itself. The latter one dominantly refers to different substituents and metal ions as well as the planarity of the porphyrinic macrocycles. Affected by different factors, metalloporphyrins can exhibit different redox potentials that is probably varied up to 1.0 V or even more. Electrochemical performance of some porphyrins has been documented previously.45−48As given in Figure 7a, the slow sweep CV of complex 1 displays one quasi-reversible wave with E1/2 = −0.75 V, accompanied by a reductive peak at −1.36 V.

Figure 6. Magnetization vs H of 1 (a), 2 (b), and 3 (c).

They are close to those of DPV (−0.73 V, −1.37 V). The electrochemical HOMO−LUMO gap of complex 1 is determined by 0.72 V. The CV of complex 2 is characteristic of one quasi-reversible redox couple with E1/2 = −1.36 V (Figure 7b). The electrochemical HOMO−LUMO gap of complex 2 is determined by 0.56 V. As for complex 3, the slow sweep CV shows one quasi-reversible wave with E1/2 being F

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

good agreement with the lanthanide contraction order. It is an arithmetic progression with the D-value between two neighboring values being 0.16 V. The obvious difference in the electrochemical behaviors among complexes 1−3 is probably attributed to various metal ions, validating that different metal ions could affect the electrochemical behaviors of porphyrinic complexes. Therefore, metalation plays a vital role in adjusting electrochemical behaviors. The electrochemical performance of complexes 1−3 is similar to the cases found in the references reported previously.49,50



CONCLUSION In conclusion, we have reported a series of novel 4f-3d crystalline porphyrinic compounds with TPPS and histidine as mixed ligands. The title complexes are isomorphous and characterized by a 3D framework with the lanthanide ions in two kinds of coordination geometries. The porphyrin macrocycles show a saddle-distorted nonplanar conformation with an embedded cobalt ion binding to four pyrrole nitrogen atoms. They exhibit photoluminescence emission in the blue region. Variable-temperature magnetic susceptibility reveals that they are antiferromagnetic. Further investigations on the relationship between structures and properties for novel 3d-4f crystalline porphyrinic compounds are in progress in our lab.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b00824. Diagrams of TPPS and histidine; FT-IR spectra for 1, 2, and 3; Selected bond lengths and bond angles (DOC) Accession Codes

CCDC 1815115−1815116 and 1815119 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 Author

*E-mail: [email protected]. Fax: (+86)-796-8100490. ORCID

Wen-Tong Chen: 0000-0002-3486-1875 Yan Sui: 0000-0002-3936-4026 Notes Figure 7. Red, CV; and green, DPV for 1 (a), 2 (b), and 3 (c) in TBAPF6 and DMF. Scan rate = 100 mV s−1.

The authors declare no competing financial interest.

−1.38 V, accompanied by a reductive peak at −1.04 V (Figure 7c). The CV value −1.38 V is close to that of DPV (−1.46 V). The electrochemical HOMO−LUMO gap of complex 3 is determined by 0.40 V. The first oxidation half-wave potentials E1/2 for complexes 1, 2, and 3 are −0.75, −1.36, and −1.38 V, respectively, indicating that complex 3 is easier to be oxidized than 1 and 2. It is noteworthy that the electrochemical HOMO−LUMO gaps of complexes 1−3 increase in the order 1 (Sm, 0.72 V) > 2 (Eu, 0.56 V) > 3 (Dy, 0.40 V), which is in

We thank the financial support of the NSF of China (21361013), Jiangxi Provincial Department of Education’s Item of Science and Technology (GJJ170637), the open foundation (20180008, 20150019) of the State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, and the open foundation (ST201522007) of the Key Laboratory of Jiangxi Province for Persistant Pollutants Control and Resources Recycle (Nanchang Hangkong University).



G

ACKNOWLEDGMENTS

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design



Article

Crystal Structures, Supramolecular Motifs and Properties. Chem. Asian J. 2011, 6 (6), 1416. (21) Schneider, J.; Kollhoff, F.; Schindler, T.; Bichlmaier, S.; Bernardi, J.; Unruh, T.; Libuda, J.; Berger, T.; Diwald, O. Adsorption, Ordering, and Metalation of Porphyrins on MgO Nanocube Surfaces: The Directional Role of Carboxylic Anchoring Groups. J. Phys. Chem. C 2016, 120 (47), 26879. (22) Zhai, X.; Alexander, D.; Derosa, P.; Garno, J. C. DistanceDependent Measurements of the Conductance of Porphyrin Nanorods Studied with Conductive Probe Atomic Force Microscopy. Langmuir 2017, 33 (5), 1132. (23) El-Denglawey, A. Illumination effect on the structural and optical properties of nano meso nickel (II) tetraphenyl-21H, 23Hporphyrin films induces new two hours photo bleached optical sensor. J. Lumin. 2018, 194, 381. (24) Uddin, S. M. N.; Nagao, Y. Multilayer Growth of PorphyrinBased Polyurea Thin Film Using Solution-Based Molecular Layer Deposition Technique. Langmuir 2017, 33 (44), 12777. (25) Bhunia, S.; Das, S. K.; Jana, R.; Peter, S. C.; Bhattacharya, S.; Addicoat, M.; Bhaumik, A.; Pradhan, A. Electrochemical StimuliDriven Facile Metal-Free Hydrogen Evolution from PyrenePorphyrin-Based Crystalline Covalent Organic Framework. ACS Appl. Mater. Interfaces 2017, 9 (28), 23843. (26) Pinto, S. M. A.; Neves, A. C. B.; Calvete, M. J. F.; Abreu, A. R.; Rosado, M. T. S.; Costa, T.; Burrows, H. D.; Pereira, M. M. Metalloporphyrin triads: Synthesis and photochemical characterization. J. Photochem. Photobiol., A 2012, 242, 59. (27) Pinto, S. M. A.; Lourenco, M. A. O.; Calvete, M. J. F.; Abreu, A. R.; Rosado, M. T. S.; Burrows, H. D.; Pereira, M. M. Synthesis of New Metalloporphyrin Triads: Efficient and Versatile Tripod Optical Sensor for the Detection of Amines. Inorg. Chem. 2011, 50 (17), 7916. (28) Wahadoszamen, Md.; Yamamura, T.; Momotake, A.; Nishimura, Y.; Arai, T. High binding affinity of DABCO with porphyrin in a porphyrin-cis-stilbene-porphyrin triad. Heterocycles 2009, 79, 331. (29) (a) Tian, Y.; Busani, T.; Uyeda, G. H.; Martin, K. E.; van Swol, F.; Medforth, C. J.; Montano, G. A.; Shelnutt, J. A. Hierarchical cooperative binary ionic porphyrin nanocomposites. Chem. Commun. 2012, 48 (40), 4863. (b) Eskelsen, J. R.; Phillips, K. J.; Hipps, K. W.; Mazur, U. Hyperbranched crystalline nanostructure produced from ionic π-conjugated molecules. Chem. Commun. 2015, 51 (13), 2663. (c) Adinehnia, M.; Mazur, U.; Hipps, K. W. Predicting the Size Distribution in Crystallization of TSPP:TMPyP Binary Porphyrin Nanostructures in a Batch Desupersaturation Experiment. Cryst. Growth Des. 2014, 14 (12), 6599. (d) Adinehnia, M.; Borders, B.; Ruf, M.; Chilukuri, B.; Hipps, K. W.; Mazur, U. Comprehensive structurefunction correlation of photoactive ionic π-conjugated supermolecular assemblies: an experimental and computational study. J. Mater. Chem. C 2016, 4 (43), 10223. (30) Demel, J.; Kubat, P.; Millange, F.; Marrot, J.; Cisarova, I.; Lang, K. Lanthanide-Porphyrin Hybrids: from Layered Structures to MetalOrganic Frameworks with Photophysical Properties. Inorg. Chem. 2013, 52 (5), 2779. (31) CrystalClear v 1.35; Rigaku Corporation, ,2002. (32) SHELXTL v 5 Reference Manual; Siemens Energy & Automation Inc.: Madison, Wisconsin, USA, 1994. (33) (a) Zhang, Z.; Zhang, L.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M. J. Template-Directed Synthesis of Nets Based upon Octahemioctahedral Cages That Encapsulate Catalytically Active Metalloporphyrins. J. Am. Chem. Soc. 2012, 134, 928. (b) Mathura, S.; de Sousa, A. S.; Fernandes, M. A.; Marques, H. M. 5,10,15,20-Tetra-pphenylsulfonylporphyrinatocobalt(III), a water-soluble Co(III) porphyrin. Inorg. Chim. Acta 2012, 392, 108. (34) (a) Visinescu, D.; Toma, L. M.; Fabelo, O.; Ruiz-Perez, C.; Lloret, F.; Julve, M. Low-Dimensional 3d-4f Complexes Assembled by Low-Spin [FeIII(phen)(CN)4]− Anions. Inorg. Chem. 2013, 52, 1525. (b) Song, X.-Q.; Xing, D.-Y.; Lei, Y.-K.; Zhao, M.-M.; Cheng, G.-Q.; Wang, X.-R.; Peng, Y.-Q. Lanthanide coordination polymers

REFERENCES

(1) Hoarau, M.; Hureau, C.; Gras, E.; Faller, P. Coordination complexes and biomolecules: A wise wedding for catalysis upgrade. Coord. Chem. Rev. 2016, 308, 445. (2) Howarth, A. J.; Majewski, M. B.; Wolf, M. O. Photophysical properties and applications of coordination complexes incorporating pyrene. Coord. Chem. Rev. 2015, 282−283, 139. (3) So, Y. M.; Leung, W. H. Recent advances in the coordination chemistry of cerium(IV) complexes. Coord. Chem. Rev. 2017, 340, 172. (4) Mokhtarzadeh, C. C.; Moore, C. E.; Rheingold, A. L.; Figueroa, J. S. Terminal Iron Carbyne Complexes Derived from Arrested CO2 Reductive Disproportionation. Angew. Chem., Int. Ed. 2017, 56, 10894. (5) Zhang, P.; Suárez, J. M.; Driant, T.; Derat, E.; Zhang, Y.; Ménand, M.; Roland, S.; Sollogoub, M. Cyclodextrin Cavity-Induced Mechanistic Switch in Copper-Catalyzed Hydroboration. Angew. Chem., Int. Ed. 2017, 56, 10821. (6) Wang, Y.; Liu, Y.; Li, G.; Hao, J. Porphyrin-based honeycomb films and their antibacterial activity. Langmuir 2014, 30, 6419. (7) Giribabu, L.; Reeta, P. S.; Kanaparthi, R. K.; Srikanth, M.; Soujanya, Y. Bis(porphyrin) anthraqu- inone triads: synthesis, spectroscopy, and photochemistry. J. Phys. Chem. A 2013, 117, 2944. (8) Fuhrhop, J. H. Porphyrin assemblies and their scaffolds. Langmuir 2014, 30, 1. (9) Haumesser, J.; Gisselbrecht, J. P.; Karmazin-Brelot, L.; Bailly, C.; Weiss, J.; Ruppert, R. Synthesis and Electrochemical Studies of Porphyrin Dimers Linked by Metallocarbenes. Organometallics 2014, 33, 4923. (10) Zhang, J.; Zhang, F.; Wang, L.; Du, J.; Wang, S.; Li, S. Synthesis, characterization, and cytotoxicity of complexes of platinum(II) with 2,2′-bipyridine and N-benzoyl-L-amino acid dianion. J. Coord. Chem. 2012, 65, 2159. (11) Li, L. J.; Fu, B.; Qiao, Y.; Wang, C.; Huang, Y. Y.; Liu, C. C.; Tian, C.; Du, J. L. Synthesis, characterization and cytotoxicity studies of platinum(II) complexes with reduced amino acid ester Schiff-bases as ligands. Inorg. Chim. Acta 2014, 419, 135. (12) Li, Z. Z.; Du, L.; Zhang, X. Z.; Li, Z. L.; Li, L.; Yang, J.; Zhao, Q. H. Two 2D grid-based Co(II) amino acid Schiff base complexes with left- and right-handed helical chains: Structures and magnetism. Inorg. Chem. Commun. 2014, 45, 20. (13) Yang, E.; Wang, L.; Wang, F.; Lin, Q.; Kang, Y.; Zhang, J. Zeolitic metal-organic frameworks based on amino acid. Inorg. Chem. 2014, 53, 10027. (14) Peterson, K. L.; Dang, J. V.; Weitz, E. A.; Lewandowski, C.; Pierre, V. C. Effect of Lanthanide Complex Structure on Cell Viability and Association. Inorg. Chem. 2014, 53, 6013. (15) Wang, X. H.; Chang, H. J.; Xie, J.; Zhao, B. Z.; Liu, B. T.; Xu, S. L.; Pei, W. B.; Ren, N.; Huang, L.; Huang, W. Recent developments in lanthanide-based luminescent probes. Coord. Chem. Rev. 2014, 273, 201. (16) Portz, A.; Gebhardt, C. R.; Duerr, M. Real-Time Investigation of the H/D Exchange Kinetics of Porphyrins and Oligopeptides by Means of Neutral Cluster-Induced Desorption/Ionization Mass Spectrometry. J. Phys. Chem. B 2017, 121 (49), 11031. (17) Amayuelas, E.; Fidalgo-Marijuan, A.; Bazan, B.; Urtiaga, M. K.; Barandika, G.; Arriortua, M. I. Highly thermally stable heterogeneous catalysts: study of 0D and 3D porphyrinic MOFs. CrystEngComm 2017, 19 (48), 7244. (18) Chen, W.; El-Khouly, M. E.; Fukuzumi, S. Saddle Distortion of a Sterically Unhindered Porphyrin Ring in a Copper Porphyrin with Electron-Donating Substituents. Inorg. Chem. 2011, 50 (2), 671. (19) Lee, S.; Kwak, D.-H.; Han, S.-B.; Lee, Y.-W.; Lee, J.-Y.; Choi, I.A.; Park, H.-S.; Park, J.-Y.; Park, K.-W. Bimodal Porous Iron/ Nitrogen-Doped Highly Crystalline Carbon Nanostructure as a Cathode Catalyst for the Oxygen Reduction Reaction in an Acid Medium. ACS Catal. 2016, 6 (8), 5095. (20) Chen, W.; Suenobu, T.; Fukuzumi, S. A Vanadium Porphyrin with Temperature-Dependent Phase Transformation: Synthesis, H

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

constructed by a new semirigid bridging salicylamide ligand: Synthesis, supramolecular structure and luminescence properties. Inorg. Chim. Acta 2013, 404, 113. (35) Gouterman, M. In The Porphyrins; Dolphin, D., Ed.; Academic Press: New York, 1978; Vol. 3, Chapter 1 and references therein. (36) Hurst, J. K.; Lee, L. Y. C.; Graetzel, M. Photoredox behavior of zinc(II) porphyrins in vesicle assemblies. J. Am. Chem. Soc. 1983, 105 (24), 7048. (37) (a) Vij, A.; Gautam, S.; Kumar, V.; Brajpuriya, R.; Kumar, R.; Singh, N.; Chae, K. H. X-ray absorption spectroscopy and photoluminescence study of rare earth ions doped strontium sulphide phosphors. Appl. Surf. Sci. 2013, 264, 237. (b) Li, Z.-M.; Deng, L.-G.; Zhao, S.-C.; Zhang, S.-Q.; Guo, C.-Y.; Liang, J.-Y.; Yue, H.; Wan, C.Y. Photoluminescence properties of a new orange-red-emitting Sm3+La3SbO7 phosphor. Luminescence 2016, 31 (2), 462. (c) Zhang, S.; He, M.; Yu, X.; Zhang, W. Enhanced photoluminescense of Gd2O3:Sm3+ nanocrystals by Ag+ doping. Henan Daxue Xuebao 2008, 38 (4), 349. (38) Martinez, E.; Garcia, M.; Ramos-Brito, F.; Alvarez-Fregoso, O.; Lopez, S.; Granados, S.; Chavez-Ramirez, J.; Martinez M, R.; Falcony, C. Characterization of Al2O3:Eu3+ luminescent coatings prepared by spray pyrolysis technique. Phys. Status Solidi B 2000, 220 (1), 677. (39) (a) Yu, C.; Yu, M.; Li, C.; Zhang, C.; Yang, P.; Lin, J. Spindlelike Lanthanide Orthovanadate Nanoparticles: Facile Synthesis by Ultrasonic Irradiation, Characterization, and Luminescent Properties. Cryst. Growth Des. 2009, 9 (2), 783. (b) Katkova, M. A.; Ilichev, V. A.; Konev, A. N.; Pestova, I. I.; Fukin, G. K.; Bochkarev, M. N. 2Mercaptobenzothiazolate complexes of rare earth metals and their electroluminescent properties. Org. Electron. 2009, 10 (4), 623. (40) Yi, X.-G.; Huang, J.-G.; Hu, R.-H.; Luo, Z.-G.; Pei, Y.-P.; Chen, W.-T. A series of thermal stable lanthanide porphyrins with a condensed three-dimensional porous open framework: gas adsorption and magnetic properties. J. Porphyrins Phthalocyanines 2015, 19, 1072. (41) Zhao, J.; Li, H.; Li, Y.; Li, C.; Wang, Z.; Chen, L. Rectangle versus Square Oxalate-Connective Tetralanthanide Cluster Anchored in Lacunary Lindqvist Isopolytungstates: Syntheses, Structures, and Properties. Cryst. Growth Des. 2014, 14 (11), 5495. (42) Nagaraja, C. M.; Maji, T. K.; Rao, C. N. R. Synthesis, structure and magnetic properties of two organically-templated coordination polymers, {[EDAH2][M1M2F2(SO4)2(H2O)2]}n (M1 = M2 = NiII and M1 = CoII, M2 = NiII). Inorg. Chim. Acta 2012, 389, 85. (43) Liu, T.; Wang, B.-W.; Chen, Y.-H.; Wang, Z.-M.; Gao, S. Syntheses, structures and magnetic properties of two mixed-valent disc-like hepta-nuclear compounds of [FeIIFeIII6(tea)6](ClO4)2 and [MnII3MnIII4(nmdea)6(N3)6]·CH3OH (tea = N(CH2CH2O)33‑, nmdea = CH3N(CH2CH2O)22‑. Z. Anorg. Allg. Chem. 2008, 634, 778. (44) Kadish, K. M.; Van Caemelbecke, E.; Royal, G. The Porphyrin Handbook, Kadish, K. M.; Smith, K. M.; Guilard, R., Eds.; Academic Press: San Diego, CA, 2000; Vol. 8, Chapter 55. (45) Kadish, K. M.; Morrison, M. M. Solvent and substituent effects on the redox reactions of para-substituted tetraphenylporphyrin. J. Am. Chem. Soc. 1976, 98, 3326. (46) Vecchi, A.; Erickson, N. R.; Sabin, J. R.; Floris, B.; Conte, V.; Venanzi, M.; Galloni, P.; Nemykin, V. N. Electronic Properties of Mono-Substituted Tetraferrocenyl Porphyrins in Solution and on a Gold Surface: Assessment of the Influencing Factors for Photoelectrochemical Applications. Chem. - Eur. J. 2015, 21 (1), 269. (47) Grabowska, I.; Zborowska, M.; Nguyen, N. T.; Dehaen, W.; Stulz, E.; Wood, J. W.; Radecka, H.; Radecki, J. Deposition of Ni(II)Porphyrin Monolayer on the Gold Electrode via Azide-Alkyne Click − Coupling and its Electrochemical Characterization. Int. J. Electrochem. Sci. 2014, 9 (11), 5948. (48) Monsu Scolaro, L.; Plutino, M. R.; Romeo, A.; Romeo, R.; Ricciardi, G.; Belviso, S.; Albinati, A. Platinum(II) complexes bearing 1,1′-bis(diphenylphosphino)ferrocene as building blocks for functionalized redox active porphyrins. Dalton T. 2006, 21, 2551. (49) Pei, Y.; Huang, J.; Hu, R.; Yang, Y.; Zhou, J.; Chen, W. Metalinduced property variation for a series of lanthanide-vanadium porphyrins. J. Porphyrins Phthalocyanines 2015, 19, 811.

(50) Chen, W.; Huang, J.; Luo, Q.; Xu, Y.; Fu, H. A novel terbiumcobalt tetra(4-sulfonatophenyl) porphyrin: synthesis, structure and photophysical and electrochemical properties. J. Porphyrins Phthalocyanines 2015, 19, 154.

I

DOI: 10.1021/acs.cgd.8b00824 Cryst. Growth Des. XXXX, XXX, XXX−XXX