ISSN : 0970 - 020X, ONLINE ISSN : 2231-5039
     FacebookTwitterLinkedinMendeley

Complex of Tris(Phenanthroline)Cobalt(II)Trifluoroacetate: Characterisation and Powder XRD Analysis

Isana Supiah Yosephine Louise, Savetsilla Nabila and Kristian Handoyo Sugiyarto*

Department of Chemistry Education, Yogyakarta State University, Jl. Colombo No.1., Yogyakarta 55281, Indonesia

Corresponding Author E-mail: sugiyarto@uny.ac.id

DOI : http://dx.doi.org/10.13005/ojc/350504

Article Publishing History
Article Received on : 13-09-2019
Article Accepted on : 03-10-2019
Article Published : 12 Oct 2019
Article Metrics
ABSTRACT:

The powder complex of tris(phenanthroline)cobalt(II) trifluoroacetate (TFA) has been prepared by interaction of cobalt(II) nitrate, phenanthroline (phen) in aqueous solution with drops of ethanol, and an excess of saturated aqueous solution of sodium trifluoroacetate, whereupon the yellowish powder was produced. AAS measurment for the metal content, equivalent conductance (1:2), and TGA-DTA of the powder suggest the corresponding formula of [Co(phen)3](TFA)2.5H2O. The efffective magnetic moment of 4.9 BM indicates the high-spin nature of this complex which corresponds to three unpaired electrons in the electronic configuration of Co(II) with significantly higher than the spin only value due to the orbital contribution. UV-Vis spectrum of the complex reveals the d-d spin-allowed transition bands as well as the MLCT and intraligand band. The estimating transition energy ratio of 2.04 for ν2/ν1, is clearly in the range for octahedral configuration. The infrared spectral property indicates the main mode of vibrations for the functional groups of ligand phen and TFA, and thus supports strongly the formula proposed for this complex. The corresponding powder XRD was then refined using Rietica-Le Bail method and found to be fit as triclinic crystal system with space group of PĪ.

KEYWORDS:

Cobalt(II); Rietica; Phen; Trifluoroacetate

Download this article as: 

Copy the following to cite this article:

Louise I. S. Y. , Nabila S, Sugiyarto K. H. Complex of Tris(phenanthroline)cobalt(II) Trifluoroacetate: Characterisation and Powder XRD Analysis. Orient J Chem 2019;35(5).


Copy the following to cite this URL:

Louise I. S. Y. , Nabila S, Sugiyarto K. H. Complex of Tris(phenanthroline)cobalt(II) Trifluoroacetate: Characterisation and Powder XRD Analysis. Orient J Chem 2019;35(5). Available from: https://bit.ly/36hzfi4


Introduction

Crystal structure of tris-bidentate-cobalt(II) complexes have been well known, particularly those with N-donor atoms in six-membered ring system, bipyridine (bipy)1,2, and 1,10-phenanthroline (phen)3. Structural studies through single crystal XRD analyses of tris(phen)cobalt(II) with several counter anions, the perchlorate (ClO4)4, squarate (C4O4)5, tetrathionate (S4O6)6, triiodide (I3)7 and the trichloroacetate (CCl3COO)8 have been well refined. The point is that the cell parameters for the cationic complex in those complexes are different one to another even though the symmetry and space group of the crystal may or may not the same class. Therefore, discussing the cell parameters through other simple method by powder X-ray diffraction may be relevant. Quite recently, the cell parameters of some complexes have been estimated from the corresponding powder-XRD by applying Le Bail method9-14, which is usually applied for powder metal oxides15-18. While in the absence of single crystals due to unsuccessful on producing them are quite often, the powder XRD offers an alternative way to estimate the cell parameters, and probably the position of atoms and thus bond lengths as well as bond angles might be depicted in one occation.

Salt of trifluoroasetate (TFA) has been known as strongly ionic nature due to highly electronegative fluorine atoms in the trifluoromethyl group.19 The coordinating agent of oxygen atoms of acetate group to the transition metal ion has been known20, however, it seems to diminish particularly when competing with other chelating bidentate ligand, bipyridine and phenanthroline, in an octahedral configuration. Indeed, complexes of [M(phen)3](TFA)2 where M = Ni(II), Cu(II), and Mn(II)12-14 have been characterized to be ionic.

Therefore, preparation of a complex containing cobalt(II), 1,10-phenanthroline and TFA should be of interest to confirm the role of TFA as a ligand or counter anion as well as structural study of the powder-XRD.

Materials and Methods

Materials

The common reagents, cobalt(II) nitrate, 1,10-phenanthroline, sodium trifluoroacetate (NaTFA), ammonium nitrate, calcium chloride, calcium nitrate, and aluminium nitrate were purchased from Sigma-Aldrich, and used as they were recieved.

Preparation of tris-bipyridinecobalt(II) complex

A mixture of Co(NO3)2 (0.1 mmol) and phenanthroline (0.32 mmol) in about 15 mL aqueous solution with 2-3 drops of ethanol was well stirred and warmed till solution become clear. To this solution, an excess of saturated aqueous solution of CF3COONa (0.5 mmol in 5 mL) was added. The resultant was concentrated on warming to about 8-10 mL, whereupon the yellow-orange powder was settled down on cooling while scratching. This solid was then filtered, rinsed with a minimum of cold water, and then dried in exposure. Preparation of this complex was seperately performed three times to confirm the reproducibility of this sample.

Instruments and Physical Measurements

Magnetic susceptibility for powder samples were measured at room temperature only using magnetic susceptibility balance (MSB) of Auto Sherwood Scientific 10169 model, calibrated with MnCl2. The molar magnetic susceptibility data were then corrected for diamagnetism according to Pascal’s constant.21 The effective magnetic moment (μeff) was then calculated from the corrected molar magnetic susceptibility by the formula, μeff = 2.828 √ (χM.T) BM.

For the solid sample, the electronic spectrum was recorded on Pharmaspec UV 1700 spectrophotometer by spreading the powder on a 2×2 cm particular thin glass pasted with ethanol. The fitting was then set in the cell holder and the spectrum was recorded at 300-1000 nm. While the infrared spectra of the powder sample and sodium TFA were recorded on an infrared Spectrophotometer of FTIR Shimadzu Prestige 21 model at 600-4000 cm-1.

Metal content in the complex was estimated using an Atomic Absorption Spectrophotometer of Shimadzu AA-6650 model, while the ionic property of complex was estimated by comparing the data of electrical (equivalent) conductance with those of known ionic solutions recorded in the same conductometer calibrated with an aqueous solution of 1 M potassium chloride at 250C.

The mass loss of hydrated molecule of water and further decomposition of complex was performed on Diamond Perkin Elmer Instrument, and simultaneous TGA-DTA were obtained by a NETZSCH STA 409C/CO thermal analyzer model with the rate of 10°C/mins.

Scanning Electron Microscopy with Energy Dispersive X-Ray (SEM-DEX) of images for the complex were recorded on JEOL JED-2300 model to confirm the crystalinity as well as the presence of main elements in the sample.

Powder X-Ray Diffraction of the complex was recorded using a Rigaku Miniflex Benchtop Diffractometer, CuKα (λ=1.5406 Å). The powder was spread on the glass plate which was then placed on the cell holder. The data of reflections was recorded in scan mode at 10-80 degree of 2θ with interval of 0.02 and rate of 10. The obtained diffractogram was then refined by Rietica program of Le Bail method (10-80 degree of 2θ) within 75 cycles.

Results and Discussion

Conductance, TGA-DTA and Chemical Formula of the Complex

Interaction of light-pink cobalt(II) ions and colorless phenanthroline molecules in a mixture that produces a light yelowish color seems almost certainly a cationic complex, [Co(phen)n]2+. The addition of anionic TFA in excess should exert the precipitation of a complex containing TFA. The data of electrical equivalent conductance for this complex together with some known ionic simple compounds in aqueous solution as shown in Table 1 suggests that the corresponding value is clearly in the range of ionic compounds with three ions per molecule, and thus the possible empirical formula of [Co(phen)n](CF3COO)2.xH2O might be reasonably proposed for the complex. This data also confirms strongly the highly ionic property of the complex containing Co(II)-phen-TFA.

Table 1: Electrical equivalent conductance of the complex and some known salts

Compounds

Equivalent conductance(Λc) Ω-1cm2mol-1 Ratio of cation/anion

Number of ions

NH4NO3

128.174 1 : 1 2
CaCl2 196.898 1 : 2

3

Ca(NO3)2

229.292 2 : 1 3
Al(NO3)3 510.054 3 : 1

4

Co(phen)3(CF3COO3)2.5H2O

198.341 2 : 1

3

The coordination number (n) for the empirical formula of this complex could be then initially estimated on the basis of metal content obtained from atomic aborption spectral data which was found to be 6.42%, the calculated metal content being 6.43% for n =3 and x = 5 (but 6.37% for n =3 and x = 5.5 and 6.50% for n = 3 and x = 4.5). This was then confirmed further by the loss of mass for each composing entity as indicated in TGA-DTA graphs shown in Fig. 1 and in Table 2. The graph consists of 5 stages of decomposision with the total marking figures of 89.529%, leaving residual percentage of 10.471%. Unfortunately, the last stage of decomposition shows clearly that TG curve lasts at about only 10.00% starting at 5000C, which corresponds to metal oxide, Co2O(cal. 9.06%)5. This means that marking figure of 8.565% for the first stage of mass loss seems to be slightly corrected becoming about 9.036%. This is due to loss of 5 molecules of H2O (calculated 9.82%). Without specifying the detailed compounds of decompositions, the next stage of 24.748% at 200-3000C may correspond to loss of TFA (cal. 24.68%) as also observed in other complexes containing TFA12-14,22, and the next stage of 56.216% at 300-5000C5,11,13,23 seems due to loss of phenanthroline (cal. 59.04%).

 Figure 1: The TGA-DTA of [Co(phen)3](CF3COO)2·5H2O at 30-10000C Figure 1: The TGA-DTA of [Co(phen)3](CF3COO)2·5H2O at 30-10000C

Click here to view figure

 

Table 2: Composing entity contained in [Co(phen)3](CF3COO)2.5H2O

Type

Co 5H2O (CF3COO)2 (phen)3 ½ Co2O3
Calculated 6.43 9.82 24.68 59.04

9.06

Found

6.42 9.036*) 24.748 56.216 10.00*)
Method AAS TGA TGA TGA

TGA

Error (%)

0.00 8.67 0.27 4.78

0.70

*) corrected figures; when the total 2nd -5th stage is 80.964%, for the residue to be 10.00%

(see the graph in Fig.1), the corrected figure for 1st stage is then 9.036%.

Magnetic Property

The magnetic susceptibility data obtained on measurement (Table 3) for the powder complex were then calculated to the magnetic moment on the basis of the formula proposed, and the result of the three separated samples are 4.89-4.90 BM, being similar to that found in the bromide counter anion24, but slightly lower (4.3 BM) in the squarate counter anion5. These are definitely higher than that of spin only value for three unpaired electrons (3.87 BM) in high-spin 3d7 configuration, due to the contribution to the moment which is quite commonly observed in Co(II) coplexes as predicted from the triply ground term of 4T1g in octahedral configuration25,26. Thus, phenanthroline produces clearly an octahedral weak ligand field surrounding Co(II).

Table 3: Magnetic data of [Co(phen)3](CF3COO)2.5H2O at 292K

Sample

χM’  x 10-6 (cgs) Magnetic Moment (BM)
1 10240

4.89

2

10310 4.90
3 10225

4.89

 

UV-VIS Electronic Spectrum

The electronic spectrum of the powder complex shown in Fig. 2 reveals the d-d transitions. The first broad absorption band of the lowest energy centred at about 930 nm (10750 cm-1) is attributed to the spin-allowed transition, ν1: 4T1g ® 4T2g (F), while the relatively sharp shoulder at 455 nm (21980 cm-1) should be assigned as ν2: 4T1g ® 4T1g(P), leading to the transition energy ratio ν21 = 2.04. This ratio is in the range 1.9-2.2 for octahedral configuration.26 The broad shoulder centred at a lower energy, 675 nm (14800 cm-1), might be then due to spin forbidden transitions, 4T1g ® 2T1g;2T2g(G)5,27. When a two-electron transition in this complex is considered in near UV region, the ν3: 4T1g ® 4A2g(F) should be masked by the high intensity of metal-ligand charge transfer which appears sharply at 335 nm (29850 cm-1)24. Another sharp band at 257 nm (38900 cm-1) is then to be π ® π* intra ligand transition5. For comparison, the diffuse reflectance spectrum of [Co(phen)3](BF4)2 might be considered27. It shows somewhat different spectral pattern in terms of position of band peaks and in assigning the transitions which result in a slightly lower transition energy ratio, ν21 = 1.93. It is noted that its magnetic moment was also reported considerably lower, 4.03 BM in this instance.

 Figure 2: Electronic spectrum of powder  [Co(phen)3](CF3COO)2·5H2O Figure 2: Electronic spectrum of powder  [Co(phen)3](CF3COO)2·5H2O

Click here to view figure

Infrared Spectrum

The infrared spectra of phen and the tris(phenanthroline)cobalt(II) complex have been known and reported to shift slightly in vibration modes due to chelation in the complex.24 The infrared spectra of the complex prepared in this work and of the TFA anion were overlaid so as to assign straight forward as shown in Fig. 3. For the complex (Fig. 3A-red) a broad band at about 3400 cm-1 might be due to -OH stretching of H2O molecule as observed at 3417 cm-1 according to Tao et al.3, but at 3512 cm-1 reported by Yeşilel el al.5. Another medium band at 3080 cm-1 is due to aromatic C-H bond as reported by Hamdani et al.28 at 3100 cm-1 and Zang et al.29 at 3064 cm-1. Meanwhile, aromatic C=C is assigned at 1517 cm-1, as observed by others5,28,24 at 1516 cm-1, 1518 cm-1, and 1520 cm-1. C-N mode of vibration is assigned at 1582 cm-1 as in Hamdani et al.28 but it was reported at 1627 cm-1 in Ramírez-Delgado et al.27. In the case of TFA, a very strong mode at 1682 cm-1 might be due to C=O stretching30 as compared to that of sodium TFA (Fig. 3B-black). The bands at 867 and 800 cm-1 are to be asymmetry- symmetry deformation of CF3, respectively.30

 Figure 3: Infrared Spectra of powder  [Co(phen)3](CF3COO)2·5H2O (A, red) and CF3COONa (B, black Figure 3: Infrared Spectra of powder  [Co(phen)3](CF3COO)2·5H2O (A, red) and CF3COONa (B, black

Click here to view figure

Therefore, the infrared data strongly suggest that the complex of sample should contain the chelation of phenanthroline ligand to cobalt(II) and (uncoordinated) TFA anion.  Thus, all physical measurements of the complex support the proposed formula of the complex.

SEM-EDX

SEM images of the powder is depicted in Fig. 4 (a-c), and the related energy dispersive X-ray (EDX) analysis is shown in Fig. 4d. These signify crystalinity of the powder, and the presence of all elements composing the complex (except the hydrogen atom), C, N, F, O, and Co. It should be noted that the percentage ratio of the number of atoms does not represent the empirical formula of the complex, since it is not possible to have homogenous powder as indicated by the SEM images and its EDX.

 Figure 4: SEM images of [Co(phen)3](CF3COO)2·5H2O at magnification of 100x (a), 1000x (b), and  5000x (c) and its EDX Figure 4: SEM images of [Co(phen)3](CF3COO)2·5H2O at magnification of 100x (a), 1000x (b), and  5000x (c) and its EDX

Click here to view figure

Powder X-Ray Diffaction and Structural Analysis

The single crystals of cation [Co(phen)3]2+ with particularly various anions have been reported with cell parameters as shown in Table 4. They are [Co(phen)3] X, where X = (H3btec)(H2btec)0.5.DMF.6H2O3, (S4O6).7H2O6, [C4O4].8H2O5, (I3)27, (ClO4)2·H2O4, and (CCl3COO)28. They were found as triclinic-PĪ space group for the first two3,6,  monoclinic- P21/c for the second two5,7, and monoclinic-C2/c for the last two4,8, respectively. In fact, no cell parameters are found to be very close one to another although they are in the same symmetry and space group.

From those facts, the powder X-ray diffraction of [Co(phen)3](CF3COO)2.5H2O was  then recorded and shown in Fig. 5 together with its refinement. The triclinic symmetry of space group PĪ was found considerably fit for this complex in this work. The red full line of the refinement according to Le Bail method of Rietica program (at 10-80 degree of 2 theta) is  almost coincident with all the black signs (+) of the experimentally observed data. With the blue bar-lines corresponding to the position of peaks for the symmetry model, the difference between the observed data and the refinement was found nearly flat line (green). This is confirmed by the relative-low figures of merit of Rp = 2.01, Rwp = 3.11, Rexp = 1.91, and Gof  = 2.646, with the derived Bragg R-Factor of 0.06. The cell parameters were found to be a = 12,3983 Å, b = 13.4523 Å, c = 14.3230 Å, α = 76.3580o; β = 67.3296o,γ = 70.76120 and V = 2064.9050 Å3, as shown in Table 4, together with known single crystal data for comparison. Thus, the symmetry-space group is quite similarly fit as to the tetrathionate, (S4O6).7H2O6 with cell parameters being quite close. Surprisingly, it is considerably different from the trichloroacetate8, though both counter anions should adopt the same tetrahedral structure.

 Figure 5: Diffractogram of [Co(phen)3](CF3COO)2·5H2O (black sign +, a), and the refined triclinic space group of PĪ model Figure 5: Diffractogram of [Co(phen)3](CF3COO)2·5H2O (black sign +, a), and the refined triclinic space group of PĪ model 

Click here to view figure

 

Table 4: Detailed cell parameters of [Co(phen)3] X*, 3-8 (*due to Le Bail method of Rietica program)

X*, 3-8

(CF3COO)2·5H2O(This work)* (S4O6)·7H2O6 (H3btec)(H2btec)0.5·DMF·6H2O3 [C4O4]

  • 8H2O5
(I3)2 7 (ClO4)2

  • 8H2O4

(CCl3COO)28

Symmetry

Triclinic Triclinic Triclinic Monoclinic Monoclinic Monoclinic

Monoclinic

Space Group

P21/c P21/c C2/c

C2/c

a ( )

12.3983 12.4441 11.8123(14) 10.9832(5) 10.4187 (5) 36.366(12)

18.367 (4)

b ( )

13.4523 13.1924 13.0356(16) 21.4569(12) 29.565 (1) 15.854(2)

10.753 (2)

c ( )

14.3230 14.2385 17.575(2) 17.0649(7) 12.9299 (6) 12.341(1)

19.020 (4)

V ( )

2064.9050

2018.77 2609.5(5) 3980.9(3) 3975.8 (3) 6938 (2)

3688.2 (13)

76.3580

75.968

91.461(2)

β (°)

67.3296

66.748 101.347(2) 98.159(3) 93.395 (10) 102.83 (2)

100.94 (3)

70.7612

71.562 99.830(2)

Z

2 2 2 8 4 8

4

Figures of merit

Rp

2.01

Rwp

3.11

Rexp

1.91

GOF

2.46

1.029

0.918

Bragg R-Factor

0.06

 

Conclusion

The powder complex of [Co(phen)3](CF3COO)2·5H2O has been successfully prepared. The equivalent conductance confirms the ionic nature of this complex and the magnetic moment supports strongly the electronic configuration of high-spin Co(II) with three unpaired electrons, indicating an octahedral weak ligand field of tris-phenantroline ligand surrounding Co(II). The electronic spectral property of this complex reveals the d-d transitions with metal-ligand charge transfer (MLCT) and intraligand transition band. The infrared spectral property indicates clearly the main mode of vibrations of the ligand phen and TFA within the complex which support strongly formula of the complex. The corresponding powder X-ray diffraction which was analysed using Le Bail method of Rietica program suggests that the complex adopt triclinic symmetry of space group PĪ.

Declaration Conflict of Interest

The authors declare no conflict of interest in this writing.

References

  1. Szalada, D. J.; Creutz, C.; Mahajan, D.; Sutin, N., Chem. 1983, 22, 2372-2379. https://doi.org/10.1021/ic00159a007
  2. Yao, J.-C.; Ma, L.-F.; Yao, F.-J. Ζ. Kristallogr. , 2005, 220, 483-484. https://www.degruyter.com/downloadpdf/j/ncrs.2005.220.issue-3/ncrs.2005.220.3.483/ncrs.2005.220.3.483.pdf
  3. Tao, B.; Xia, H.; Zhu, Y.-F.; Wang, X. J. Inorg. Chem., 2012, 57(6), 822–826. DOI: 10.1134/S0036023612060058
  4. Boys, D.; Escobar, C.; Wittke, O. Acta Cryst., 1984, C40, 1359-1362. DOI:10.1107/s0108270184007964
  5. Yeşilel, O. Z.; Ölmez, H.; Yılan, O.; Paşaoğlu, H.; Büyükgüngör, O. Naturforsch., 2006, 61b, 1094-1100. http://www.znaturforsch.com/ab/v61b/61b1094.pdf
  6. Sharma, Raj. Pal.; Singh, A.; Venugopalan, P.; Brandao, P.; Felix, V. Polyhedron, 2012, 40, 175-184. https://doi.org/10.1016/j.poly.2012.03.049
  7. Tershansy, M. A.; Goforth, A. M.; Smith, M. D.; Peterson Jr., L.-R.; Loye, H.-C. Acta Cryst., 2005, E61, m1680-m1681. DOI: org/10.1107/S160053680502388
  8. Li, L.-M.; Li, Y.-F.; Liu, L.; Zhang, Z.-H. Acta Cryst., 2011, E67, m973 DOI:10.1107/S160053681102410X
  9. Sugiyarto, K. H.; Saputra, H. W.; Permanasari, L.; Kusumawardani, C. AIP Conference Proceedings ICEMS-IPCSM., 2017, 1847, 040006
  10. Kusumawardani, C.;Permanasari, L.; Fatonah, S. D.; Sugiyarto, K. H. J. Chem. 2017, 33(6), 2841-2847. DOI: http://dx.doi.org/10.13005/ojc/330617
  11. Kusumawardani, C.; Kainastiti, F.; Sugiyarto, K. H. Chiang Mai J. Sci., 2018, 45(4), 1944-1952. http://epg.science.cmu.ac.th/ejournal/
  12. Sutrisno, H.; Kusumawardani, C.; Rananggana, R. Y;. Sugiyarto, K. H. Chiang Mai J. Sci., 2018, 45(7), 2768-2778. http://epg.science.cmu.ac.th/ejournal/
  13. Sugiyarto, K. H.; Kusumawardani, C.; Wigati, ; Sutrisno, H. Orient. J. Chem., 2019, 35(1), 325-331. DOI : http://dx.doi.org/10.13005/ojc/350139
  14. Sugiyarto, K. H.; Kusumawardani, C.; Sutrisno, ; Wibowo, M. W. A. Orient. J. Chem., 2018, 34(2), 735-742. http://dx.doi.org/10.13005/ojc/340216
  15. Prijamboedi, Z. B.; Nugroho, A. A.; Ismunandar, Chin. Chem. Soc., 2009, 56(6), 1108-1111. DOI: 10.1002/jccs.200900160
  16. Purwaningsih, S. Y.; Rosidah, N.; Zainuri, M.; Triwikantoro, T.; Pratapa, S.; Darminto, D. Phys. Conf. Ser., 2019,1153, 012070 (1-5). https://iopscience.iop.org/issue/1742-6596/1153/1
  17. Lüdtke, T.; Weber, D.; Schmidt, A.; Müller, A.; Reimann, C.; Becker, N.; Bredow, T.; Dronskowski, R.; Ressler, T.; Lerch, M. Kristallogr. Cryst. Mater., 2017, 232,1-3. https://core.ac.uk/download/pdf/143954851.pdf
  18. Aughterson, R. D.; Lumpkin, G. R.; Reyes, M. de los; Sharma, N.; Ling, C. D.; Gault B.; Smith, K. L.; Avdeev, M.; Cairney, J. M. Solid State Chem. 2014, 213, 182-192. http://dx.doi.org/10.1016/j.jssc.2014.02.029
  19. Trifluoroacetic Acid. Retrieved on April 2019 from: http://www.commonorganicchemistry.com/Common_Reagents/Trifluoroacetic_Acid/Trifluoroacetic_Acid.htm
  20. Tokareva, A. O.; Tereshchenko, D. S.; Boltalin, A. I.; Troyanov, S. I. J. Coord. Chem., 2006, 32(9), 663-668. DOI: 10.1134/s1070328406090077
  21. Figgis, B. N. and Lewis, J., Modern Coordination Chemistry, Edited by Lewis, J., and Wilkins, R. G., Interscience: New York, 1960, 400.
  22. Eloussi, H.; Farjas, J.; Roura, P.; Ricart, S.; Puig, T.; Obradors, X.; Dammak, M. Thin Solid Films, 2013, 545, 200-204. DOI: 10.1016/j.tsf.2013.07.082
  23. Chandraleka, S.; Chandramohan, G. J. Pure Appl. Chem. , 2014, 8(10), 162-175. DOI: 10.5897/ajpac2014.0592
  24. Thomas, N. C.; Pringle, K.; Deacon, G. B. Chem. Edu. 1989, 66(6), 516-517. https://datapdf.com/cobaltii-and-cobaltiii-coordination-compounds-journal-of.html
  25. Mabbs, F. E. and Machin, D. J. Magnetism and Transition Metal Complexes, Capman and Hall Ltd.: London, 1973, 21
  26. Lever, A. B. P. Inorganic Electronic Spectrocopy, Elsevier Publishing Company: Amsterdam, 1968, 320.
  27. Ramírez-Delgado, V.; Cruz-Ramirez, M.; Hernández-Ayala, L.; Reyes-Vidal, Y.; Patakfalvi, R.; García-Ramos, J. C.; Tenorio, F. J.; Ruiz-Azuara, L.; Ortiz-Frade, L. Mex. Chem. Soc., 2015, 59(4), 282-293. https://www.researchgate.net/publication/303821083
  28. Hamdani, H. E. L.; Amane, M. E. L.; Atmani, Z.; Haddad, M. Mater. Environ. Sci., 2016, 7(9), 3100-3109. https://www.jmaterenvironsci.com/Document/vol7/vol7_N9/324-JMES-2434-El%20Amane.pdf.
  29. Zang, J.; Wang, L. X.; Zang, L.; Chen, Y.; Zang, Q. T. Rare Metals, 2013, 32(6). 599-604. DOI: 10.1007/s12598-013-0168-0
  30. Tas, E.; Kilic, A.; Durgun, M.; Küpecik, L.; Yilmaz, I.; Arslan, S. Spectrochim. Acta A Mol. Biomol. Spectrosc., 2010, 75(2), 811-818. DOI: 10.1016/j.saa.2009.12.002

 


Creative Commons License
This work is licensed under a Creative Commons Attribution 4.0 International License.