Next Article in Journal
Thermal Treatment of Melt-Spun Fibers Based on High Density PolyEthylene and Lignin
Previous Article in Journal
MoS2 Decorated Carbon Nanofibers as Efficient and Durable Electrocatalyst for Hydrogen Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Acyclic Arylamine-Based Ionophores as Potentiometric Sensors for Zn2+ and Ni2+ Ions

School of Chemistry and Biochemistry, Thapar University, Patiala 147004, India
*
Author to whom correspondence should be addressed.
Submission received: 21 September 2017 / Revised: 27 October 2017 / Accepted: 28 October 2017 / Published: 7 November 2017

Abstract

:
Two receptor molecules N-(2-nitrophenyl)benzene-1,2-diamine (DPA) and N,N-bis(2-nitrophenyl)benzene-1,2-diamine (TPA) are proposed as Zn2+ and Ni2+-selective electrodes, respectively. The two electrodes respond to Zn2+ and Ni2+ ions with the detection limits of 1.3 × 10−6 M and 2.8 × 10−6 M, respectively. Both the electrodes have a life time of four months and respond within 15 s and 20 s, respectively, for Zn2+ and Ni2+ over a wide pH range (3–9). The electrodes show very good selectivity towards the primary ions in presence of some alkali, alkaline earth, and transition metal ions.

Graphical Abstract

1. Introduction

The design and synthesis of artificial ionophores for selective recognition of ions is a field of significant interest due to their key role in major biological and chemical processes [1]. A large number of receptors based on pyrroles, lactams, aliphatic amines, hydrazines and urea have been reported where the amine group (-NH-) acts as the anion or cation binding site [2,3,4,5,6]. Selectivity of the ionphore to act as acation sensor depends upon availability of the electron pair on nitrogen and the pre-organisation of the cavity size, while anion sensing is governed by chemical interactions [2,3,4]. Further, receptor properties of an ionophore are governed by a delicate balance between rigidity, required for high selectivity, and flexibility, for a sufficiently fast ion exchange. Macrocyclicarylamines offer advantages of both rigidity and flexibility. The rigidity is due to its planer aromatic ring and the flexibility due to rotation around amine linkages. Thus, two new ionophores, diphenyl amine (DPA) and triphenyl amine (TPA), were designed and synthesized to explore their sensing properties by potentiometric methods. Potentiometry offers an advantage over other electrochemical methods in terms of cost, sensitivity, and selectivity [5,6,7]. The results indicated strong affinity for Zn2+ ions in case of DPA, while TPA was selective towards Ni2+ ions.
Determination of both zinc and nickel in trace amounts is important in biological, pharmaceutical, and environmental monitoring. The role of zinc as an essential element (8 to 15 mg/day in humans) for immune functioning, DNA synthesis, and enzyme activity is well known [8,9,10]. Also, extensive use of both the metals in pharmaceutical, electroplating, and paint industries increases its concentration in water, air, and soil [11]. Similarly, deficiency as well as excess of nickel is detrimental to animal and plant systems [12,13,14,15]. It has been known that the presence of nickel ions beyond a tolerable limit in drinking water (2.5 mg/mL) can cause disorders of the central nervous system, asthma, dermatitis, and cancer [16,17].
There are a number of reports in literature where analytical techniques such as atomic absorption spectrometry, voltammetry, atomic emission, and UV-Vis spectroscopy have been used for the detection of trace amounts of zinc [18,19,20,21,22,23,24,25,26,27,28,29,30] and nickel ions [31,32,33,34,35]. However, potentiometry, due to its inexpensive, wide, and linear range, is a fast and a simple procedure that has an advantage. A number of ions selective membranes (ISMs) have been reported with various compounds for the detection of both zinc and nickel. This includes [Bzo2Me2Ph2(16)hexaeneN4] [36], PME-6,7:14,15-Bzo2-10,11-(4-methylbenzene)-[15]-6,8,12,14-tetraene-9,12-N2-1,5-O2 [37], Dibenzo-24-crown-8 [26], 3-[(2-Furylmethylene)amino]-2-thioxo-1,3-thiazolidin-4-one [38], N,N-Bis(acetylacetone)ethylene diamine [25] for zinc ions and porphyrins, crown ethers, cyclams, amines, and Schiff bases-based molecules for nickel ions [33,34,39,40,41,42,43].
In this work, we report the use of two acyclic arylamine-based ionophores, N-(2-nitrophenyl)benzene-1,2-diamine (DPA) and N,N-bis(2-nitrophenyl)benzene-1,2-diamine (TPA), for preparing highly selective Zn2+ and Ni2+ membrane electrodes, respectively. To the best of our knowledge, this is the first report where diphenylamines have been reported as potentiometric sensors for Zn2+ and Ni2+ ions.

2. Results and Discussion

2.1. Potentiometric Studies

2.1.1. Determination of Binding Constants

Preliminary studies to determine the binding affinity of the ionophores with different metal ions were done potentiometrically using sandwich membrane method. A poly vinylchloride (PVC) membrane consists of plasticizer, o-nitrophenyloctylether (o-NPOE), lipophilic salt, and sodium tetraphenylborate (NaTPB), and a required amount of the ionophore was pressed against a blank membrane (without ionophore) for each ionophore, as shown in Figure 1. Binding constants for both the ionophores were calculated by measuring EMF using a known procedure [44,45] and following equation where LT, RT, and ZI are the concentration of ionophore, lipophilic salt, and charge carried by primary ion, respectively. Expressions R, T, and F carry their usual meanings and n represents stoichiometry of the metal ion-ionophore complex.
βILn = (LT − nRT/ZI)−nexp(EMZIF/RT)
The values calculated for Fe2+, Co2+, Ca2+, Mn2+, Cu2+, and Zn2+ for their affinity with DPA were 5.3, 4.7, 4.1, 3.7, 5.8, and 7.4, respectively. Thus Zn2+ ions displayed highest affinity towards DPA. Similarly, values calculated with TPA indicated more stable complex in case of Ni2+ ions. Formation constant values with Ca2+, Cu2+, Ni2+, Mg2+, Zn2+, and Co2+ ions were found to be 3.4, 6.2, 8.3, 3.8, 4.8, and 3.6, respectively.

2.1.2. Membrane Optimization

It is well known that response, mechanical strength, and life time of any liquid membrane electrode is significantly dependent upon its composition [1,46,47,48,49]. While PVC provides mechanical strength to the membrane and is inert to the chemical interactions, the o-NPOE used as a plasticizer serves as a membrane solvent with reduced viscosity and ensures high mobility of the membrane constituents. The plasticizer also helps in improving selectivity and is compatible with other membrane components. The role of lipophilic reagent (NaTPB, in our case) is to introduce Donnan exclusion of ion in the membrane, i.e., to introduce permselectivity in the membrane so that no counter ions are co-extracted from the solution to the membrane phase. The presence of sufficient amount of non-exchangeable lipophilic ions in the membrane ensures that the concentration of primary ions is much higher than the co-extracted amount. Optimization of the ratio of ionophore: NaTPB is very critical to ensure the best response of the electrode. The amount of lipophilic ions was carefully monitored to get the optimized ratio of ionophore: lipophilic salt. Thus, based on preliminary experiments (Figure 2) and values of the binding constants for both DPA and TPA, membranes were optmized to get lowest detection limit with closest Nernstian response.
Table 1 shows different compositions of DPA-based membrane and its response to Zn2+ ions. Figure 3 shows that any change in the ionophore less or greater than 3% leads to non-Nernstian responses. Variation in the plasticizer, lipophilic salt, and PVC within the permissible limit gave membrane electrode no. 3, having composition PVC: o-NPOE: DPA: NaTPB as 33:61:3:3 (w/w) as the optmium composition with closest Nernstian response. This composition was used for further studies in case of DPA.
Similarly, Table 2 shows different compositions tried for TPA-based membrane and its response towards Ni2+ ions. Here again, membranes with 3% of ionophore gave closest Nernstian response, having slope in the range of 25–30 mV/decade (membrane electrodes no. 9–11). It can be observed (Table 2) that any deviation from 3% of ionophore leads to significant divergence in the Nernstian response of the membrane. Thus, membrane electrode no. 9, having composition of TPA: PVC: o-NPOE: NaTPB as 3:32:62:3 (w/w), was selected for further studies as Ni2+ selective electrode. The response of blank membranes (membrane electrode no. 1 and 2) towards Ni2+ ions was insignificant.

2.1.3. Calibration Curve, Response Time, and Life Time of the Electrode

The effect of internal solution concentration was studied on the working of membrane electrodes to know the best performing concentration, as per International Union of Pure and Applied Chemistry (IUPAC) recommendations [50]. Three internal solution concentrations in the range 1 × 10−3 M to 1 × 10−1 M were used with best response observed in case of 1 × 10−1 M concentration. This concentration of the internal solution was used to draw a calibration curve for the Zn2+ selective electrode (Membrane no. 3, Table 1) over a concentration range of 10−8 M to 10−1 M of the primary ion (Zn2+). Figure 4 shows the calibration curve with workable concentration range of 1 × 10−6 M to 1 × 10−1 M and a Nernstian slope of 30 mV/decade. The lower detection limit obtained was 1.3 × 10−6 M for the Zn2+ ions, as calculated according to IUPAC recommendation [48]. Similarly, the calibration curve drawn for the Ni2+—selective electrode containing triphenylamine as an ionophore gave best response with membrane electrode no. 9 (Table 2) having a slope of 30 mV/decade. (Figure 5). Workable concentration range in this case was 1 × 10−5 M to 1 × 10−1 M and the lower detection limit 2.8 × 10−6 M of Ni2+. Response time was measured by recording EMF of the electrode as a function of time till a stable reading was achieved.
Response time:This is an important parameter to evaluate an electrode’s working efficiency. The steady response that in case of DPA was found within 15 s after addition of Zn2+ ions, while nickel-selective, TPA-based electrode stabilized in 20 s. Both the electrodes could be used for 4 months without observing any considerable change when stored in 0.1 M respective ion solutions.

2.1.4. Effect of pH

In order to evaluate the behavior of electrodes in different working conditions, their EMF was measured over entire pH range. Figure 6a shows the response of the DPA ionophore for the zinc ions (10−2 M and 10−3 M) in the pH range 1 to 12. It can be observed that DPA has a steady response over wide pH range of 3.6–9.3. Similarly, the response of TPA remained unaffected in the pH range of 4.2–9.2 when different concentrations of Ni2+ (10−1 M and 10−2 M) were used (Figure 6b). A decrease in the potential at lower pH in case of TPA can be attributed to protonation of the donor NH groups that become unavailable to complex with respective metal ions. Furthermore, the electrodes may be exhibiting interference due to hydrogen ions. At elevated pH, the coordinated metal gets hydrolysed and thus electrode response deviates.

2.1.5. Effect of Interfering Ions on Potentiometric Selectivity of Electrodes

The behavior of both the electrodes was studied in the presence of interfering ions in order to ascertain their selectivity. This is expressed as the potentiometric selectivity coefficient l o g K M n + , B P o t and was determined by IUPAC-recommended, fixed-interference method (FIM) [50]. Addition of a fixed concentration of interfering ions (1.0 × 10−2 M) to a varying concentration (1 × 10−8 M to 1 × 10−1 M) of the target ion showed hardly any interference from the other ions. The selectivity coefficients values so determined for DPA and TPA electrodes have been compiled in Table 3 and Table 4, respectively.
From the logK values obtained for both DPA and TPA electrodes, it can be proposed that electrodes have high selectivity for Zn2+ and Ni2+ ions, respectively, over other ions and can be conveniently used for the determination of their respective ions.

2.1.6. Applications

Applicability of both DPA and TPA membranes was evaluated as indicator electrodes in the potentiometric titrations with 1 × 10−2 M solutions of Zn2+ and Ni2+ ion, respectively, by titrating against Ethylene Diamine Tetra Acetic Acid (EDTA) solutions. Figure 7a shows an inflection point corresponding to 1:1 stoichiometry for Zn2+-EDTA complex in case of DPA membrane electrode. Similarly, Figure 7b shows decrease in the EMF as an increasing volume of EDTA is added to fixed volume of 1 × 10−2 M Ni2+ ions. Here again, titration curve shows sharp inflexion point corresponding to 1:1 stoichiometry of the complex. Therefore, both the membranes can be used as indicator electrodes for the potentiometric determination of equivalent point.

2.1.7. Comparison of the Proposed Sensors with Other Electrodes

Performance of the membrane electrodes has been compared with the earlier literature reports for the lower detection limit, slope of calibration curve, and pH range. Table 5 and Table 6 show that both DPA and TPA electrodes are more close to the Nernst Equation and thus determine the concentration of the target species closer to the actual values.

2.2. Theoretical Studies

In an attempt to better understand the nature of ionophores and their interaction with metal ions, theoretical studies were carried out using density functional theory method (DFT). B3LYP function with 6–31 G basis set was used for quantum chemical calculations of ionophores, independently and along with metal ions on Gaussian 03W programme.
Quantum chemical calculations carried out for energy optimization of DPA independently and along with other meal ions showed highest energy decrease (1778 a.u.) with Zn2+ ions. The decrease in the energy with Mn2+, Fe2+, Ni2+, and Cu2+ was 1150, 1263, 1507, and 1639 a.u., respectively. Thus, theoretical results were consistent with the experimental outcome. A look at the minimum energy configuration of DPA (Figure 8a) shows presence of a pseudo-cavity constituted by the three nitrogens. Two of these nitrogens contribute from amino and one from nitro groups. In the presence of metal ions, the nitrogen atoms align themselves in a way to fit metal in the cavity. The best fit alignment is obtained in case of Zn2+ ions with minimum energy (Figure 8b).
Similar calculations done with TPA show a decrease in the energy decreases from −1213 a.u. to −2721 a.u after addition of Ni2+, thus indicating the formation of highly stable TPA—Ni2+ complex. Figure 9 shows the three dimensional structure of TPA, in which two amine nitrogens along with one of the nitro groups are involved in composing a pseudo cavity for the trapping of metal ion. The nitro groups on the terminal phenyl ring were observed to be trans to each other. However, in the presence of Ni2+ ions, the trans nitro group flips to take cis position, where all the four nitrogens are involved in complexation with metal ion.

3. Materials and Methods

3.1. Materials and Instrumentation

Reagents for the synthesis of DPA and TPA were purchased from LOBA Chemie (Mumbai, India) and Sd Fine (Mumbai, India) and used as such. All solvents were distilled before column chromatography. For TLC, precoated plates were used, and for column chromatography, silica gel (mesh 100–120) was used. Characterizaion of the synthesized ionophores was done using 1H NMR and 13C NMR spectra that were recorded on a Bruker AVANCE—400 MHz NMR spectrometer in CDCl3 with SiMe4 as the internal standard. Membranes were prepared using high molecular weight poly vinylchloride (PVC), o-nitrophenyloctylether (o-NPOE), and sodium tetraphenylborate, tetrahydrofuran (THF) that were purchased from Aldrich (India) and used as received. All metal salts used were of AR grade and used as nitrates by dissolving in double-distilled water. Solutions were prepared by serial dilutions and the potentials measured with an Equip-Tronics (EQ-602, accuracy, ±0.1 mV) potentiometer (Mumbai, India).

3.2. Synthesis of Ionophores

Ionophores DPA and TPA were synthesized by “Modified Ullmann reaction” [51], which involves a copper mediated reaction between an aryl halide and an amine to get the corresponding arylamine. The two ionophores were then tested for binding with different metal ions by potentiometric and theoretical studies.

3.3. Potentiometric Studies

3.3.1. Preparation of Electrode

Membrane components, PVC, ionophore, ortho—nitrophenyloctylether, and sodium tetraphenylborate, as per Table 1 and Table 2, were dissolved in 2–3 mL of tetrahydrofuran (THF) [52]. The homogeneous mixture obtained after complete dissolution was poured in a glass ring of diameter 20 mm, placed on dust free glass plate, and allowed to evaporate for 24 h at room temperature. To obtain membranes with similar characteristics, viscosity of the casting solution and the rate of solvent evaporation were controlled so that thickness and morphology of the membranes remain almost unchanged. Transparent membranes so obtained were then removed from the glass ring and circular pieces of 1.25 cm diameter were cut and glued on the ground end of pyrex glass tube with araldite. Membranes were conditioned with 0.1 M metal salt solutions for 24 h.

3.3.2. EMF Measurement

The ratio of different membrane ingredients and the contact time were optimized to provide membranes, which resulted in reproducible and stable potentials. The activities of metal ions were calculated by using modified form of the Debye-Huckel equation [53]. All EMF measurements were carried out with the following assembly:
SCE||Internal Solution (0.1M Mn+)|PVC Membrane|Test solution Mn+||SCE
The concentrations of test solution were varied from 1 × 10−8 M to 1 × 10−1 M. The potential measurements were carried out at 25 ± 0.1 °C with a digital Equip-Tronics potentiometer.

3.4. Quantum Chemical Calculations

The quantum chemical study was carried out by density functional theory (DFT) method using Gauss View 4.1.2 software (Gaussian Inc., Wallingford, CT, USA). B3LYP method with 6–31 G basis set was employed for the optimization of ionophores, independently and along with metal ions.

4. Conclusions

Two ion-selective PVC membrane electrodes incorporating acyclic arylamines DPA and TPA as ionophores were used to determine Zn2+ and Ni2+, respectively. The nernstian slope of 30 mV/decade of activity of respective ions was displayed by both the electrodes with detection limits of the order of 10−6 M. The electrodes responded very fast within 15 s and 20 s, respectively, for Zn(II) and Ni(II) ions. The electrodes can be used for 4 months over a wide pH range of 3–9. The potentiometric sensors were successfully used as indicator electrodes in potentiometric titration of Zn2+ and Ni2+ with EDTA.

Acknowledgments

Authors are thankful to Director, Thapar Institute of Engineering & Technology University, Patiala, for providing research facilities.

Author Contributions

Harpreet Kaur performed the experiments, analyzed the data, created the figures, and wrote the manuscript. Susheel K. Mittal and Manmohan Chhibber supervised the project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bühlmann, P.; Pretsch, E.; Bakker, E. Carrier-based ion-selective electrodes and bulk optodes. 2. Ionophores for potentiometric and optical sensors. Chem. Rev. 1998, 98, 1593–1688. [Google Scholar] [CrossRef] [PubMed]
  2. Pankratova, N.; Cuartero, M.; Jowett, L.A.; Howe, E.N.; Gale, P.A.; Bakker, E.; Crespo, G.A. Fluorinated tripodal receptors for potentiometric chloride detection in biological fluids. Biosens. Bioelectron. 2018, 99, 70–76. [Google Scholar] [CrossRef] [PubMed]
  3. Cuartero, M.; Más-Montoya, M.; García, M.S.; Curiel, D.; Ortuño, J.A. New carbazolo [1,2-a] carbazole derivative as ionophore for anion-selective electrodes: Remarkable recognition towards dicarboxylate anions. Talanta 2014, 123, 200–206. [Google Scholar] [CrossRef] [PubMed]
  4. Cuartero, M.; Ortuño, J.A.; García, M.S.; Sánchez, G.; Más-Montoya, M.; Curiel, D. Benzodipyrrole derivates as new ionophores for anion-selective electrodes: Improving potentiometric selectivity towards divalent anions. Talanta 2011, 85, 1876–1881. [Google Scholar] [CrossRef] [PubMed]
  5. Abraham, A.A.; Rezayi, M.; Manan, N.S.; Narimani, L.; Rosli, A.N.B.; Alias, Y. A novel potentiometric sensor based on 1,2-Bis (N′-benzoylthioureido) benzene and reduced graphene oxide for determination of lead (II) cation in raw milk. Electrochim. Acta 2015, 165, 221–231. [Google Scholar] [CrossRef]
  6. Kopylovich, M.N.; Mahmudov, K.T.; Pombeiro, A.J. Poly (vinyl) chloride membrane copper-selective electrode based on 1-phenyl-2-(2-hydroxyphenylhydrazo) butane-1,3-dione. J. Hazard. Mater. 2011, 186, 1154–1162. [Google Scholar] [CrossRef] [PubMed]
  7. Ceresa, A.; Radu, A.; Peper, S.; Bakker, E.; Pretsch, E. Rational design of potentiometric trace level ion sensors. A Ag+-selective electrode with a 100 ppt detection limit. Anal. Chem. 2002, 74, 4027–4036. [Google Scholar] [CrossRef] [PubMed]
  8. Hosseini, M.; Abkenar, S.D.; Ganjali, M.R.; Faridbod, F. Determination of zinc (II) ions in waste water samples by a novel zinc sensor based on a new synthesized Schiff's base. Mater. Sci. Eng. C 2011, 31, 428–433. [Google Scholar] [CrossRef]
  9. Berg, J.M.; Shi, Y. The galvanization of biology: A growing appreciation for the roles of zinc. Science 1996, 271, 1081. [Google Scholar] [CrossRef] [PubMed]
  10. Kouremenou-Dona, E.; Dona, A.; Papoutsis, J.; Spiliopoulou, C. Copper and zinc concentrations in serum of healthy Greek adults. Sci. Total Environ. 2006, 359, 76–81. [Google Scholar] [CrossRef] [PubMed]
  11. Gupta, V.K.; Goyal, R.N.; Al Khayat, M.; Kumar, P.; Bachheti, N. A new Zn (II)-selective potentiometric sensor based on 4-tert-butylcalix [4] arene in PVC matrix. Talanta 2006, 69, 1149–1155. [Google Scholar] [CrossRef] [PubMed]
  12. István, P.; Benton, J.J. The Hand Book of Trace Elements; St. Lucie Press: St. Lucie, FL, USA, 1997; Volume 124, ISBN 1-88415-34-4. [Google Scholar]
  13. Wuana, R.A.; Okieimen, F.E. Heavy metals in contaminated soils: A review of sources, chemistry, risks and best available strategies for remediation. ISRN Ecol. 2011, 1–20. [Google Scholar] [CrossRef]
  14. Igbum, O.G.; Asemave, K.; Nwadinigwe, C.A.; Akaasah, N.Y. Heavy Metal Analysis of Blends of Methyl Esters Obtained from Four Virgin Tropical Seed Oils. Int. J. Sci. Res. 2013, 2, 463–467. [Google Scholar]
  15. Barreto, W.J.; Barreto, S.R.G.; Scarminio, I.S.; Ishikawa, D.N.; Soares, M.D.F.; Proença, M.V.B.D. Determination of Ni (II) in metal alloys by spectrophotometry UV-Vis using dopasemiquinone. Quím. Nova 2010, 33, 109–113. [Google Scholar] [CrossRef]
  16. Bohn, H.L.; McNeal, B.L.; O’ Connor, A.G. Soil Chemistry, 2nd ed.; John Willey & Sons: New York, NY, USA, 1985. [Google Scholar]
  17. Hernberg, S.; Nikkanen, J.; Mellin, G.; Lilius, H. δ-aminolevulinic acid dehydrase as a measure of lead exposure. Arch. Environ. Health 1970, 21, 140–145. [Google Scholar] [CrossRef] [PubMed]
  18. Wilhartitz, P.; Dreer, S.; Krismer, R.; Bobleter, O. High performance ultra trace analysis in molybdenum and tungsten accomplished by on-line coupling of ion chromatography with simultaneous ICP-AES. Microchim. Acta 1997, 125, 45–52. [Google Scholar] [CrossRef]
  19. Kaur, P.; Kaur, S.; Mahajan, A.; Singh, K. Highly selective colorimetric sensor for Zn2+ based on hetarylazo derivative. Inorg. Chem. Commun. 2008, 11, 626–629. [Google Scholar] [CrossRef]
  20. Gupta, N.R.; Mittal, S.; Kumar, S.; Kumar, S.A. Potentiometric studies of N,N′-Bis(2-dimethylaminoethyl)-N,N′-dimethyl-9,10 anthracene dimethanamine as a chemical sensing material for Zn (II) ions. Mater. Sci. Eng. C 2008, 28, 1025–1030. [Google Scholar] [CrossRef]
  21. Li, Q.; Zhao, X.; Lv, Q.; Liu, G. The determination of zinc in water by flame atomic absorption spectrometry after its separation and preconcentration by malachite green loaded microcrystalline triphenylmethane. Sep. Purif. Technol. 2007, 55, 76–81. [Google Scholar] [CrossRef]
  22. Lu, X.; Wang, Z.; Geng, Z.; Kang, J.; Gao, J. 2,4,6-tri (3,5-Dimethylpyrazoyl)-1,3,5-triazine modified carbon paste electrode for trace Cobalt (II) determination by differential pulse anodic stripping voltammetry. Talanta 2000, 52, 411–416. [Google Scholar] [CrossRef]
  23. Fakhari, A.R.; Shamsipur, M.; Ghanbari, K. Zn (II)-selective membrane electrode based on tetra (2-aminophenyl) porphyrin. Anal. Chim. Acta 2002, 460, 177–183. [Google Scholar] [CrossRef]
  24. Shamsipur, M.; Rouhani, S.; Ganjali, M.R.; Sharghi, H.; Eshghi, H. Zinc-selective membrane potentiometric sensor based on a recently synthesized benzo-substituted macrocyclic diamide. Sens. Actuators B 1999, 59, 30–34. [Google Scholar] [CrossRef]
  25. Gupta, V.K.; Agarwal, S.; Jakob, A.; Lang, H. A zinc-selective electrode based on N,N′-bis(acetylacetone) ethylenediimine. Sens. Actuators B 2006, 114, 812–818. [Google Scholar] [CrossRef]
  26. Gupta, V.K.; Al Khayat, M.; Minocha, A.K.; Kumar, P. Zinc (II)-selective sensors based on dibenzo-24-crown-8 in PVC matrix. Anal. Chim. Acta 2005, 532, 153–158. [Google Scholar] [CrossRef]
  27. Gupta, V.K.; Kumar, A.; Mangla, R. Protoporphyrin IX dimethyl ester as active material in PVC matrix membranes for the fabrication of zinc (II) selective sensor. Sens. Actuators B 2001, 76, 617–623. [Google Scholar] [CrossRef]
  28. Gupta, V.K.; Jain, A.K.; Mangla, R.; Kumar, P. A new Zn2+-selective sensor based on 5,10,15,20-tetraphenyl-21H,23H-porphine in PVC Matrix. Electroanalysis 2001, 13, 1036–1040. [Google Scholar] [CrossRef]
  29. Gupta, V.K. A PVC-based 12-crown-4 membrane potentiometric sensor for zinc (II) ions. Sens. Actuators B 1999, 55, 195–200. [Google Scholar] [CrossRef]
  30. Jain, A.K.; Sondhi, S.M.; Rajvanshi, S. A PVC based hematoporphyrin IX membrane potentiometric sensor for zinc (II). Electroanalysis 2002, 14, 293–296. [Google Scholar] [CrossRef]
  31. Rao, G.N.; Srivastava, S.; Srivastava, S.K.; Singh, M. Chelating ion-exchange resin membrane sensor for nickel (II) ions. Talanta 1996, 43, 1821–1825. [Google Scholar] [CrossRef]
  32. Gupta, V.K.; Jain, A.K.; Singh, L.P.; Khurana, U. Porphyrins as carrier in PVC based membrane potentiometric sensors for nickel (II). Anal. Chim. Acta 1997, 355, 33–41. [Google Scholar] [CrossRef]
  33. Jain, A.K.; Gupta, V.K.; Singh, R.D.; Khurana, U.; Singh, L.P. Nickel (II)-selective sensors based on heterogeneous membranes of macrocyclic compounds. Sens. Actuators B 1997, 40, 15–20. [Google Scholar] [CrossRef]
  34. Gupta, V.K.; Prasad, R.; Kumar, P.; Mangla, R. New nickel (II) selective potentiometric sensor based on 5,7,12,14-tetramethyldibenzotetraazaannulene in a poly (vinyl chloride) matrix. Anal. Chim. Acta 2000, 420, 19–27. [Google Scholar] [CrossRef]
  35. Mazloum, M.; Niassary, M.S.; Amini, M.K. Pentacyclooctaaza as a neutral carrier in coated-wire ion-selective electrode for nickel (II). Sens. Actuators B 2002, 82, 259–264. [Google Scholar] [CrossRef]
  36. Singh, A.K.; Jain, A.K.; Saxena, P.; Mehtab, S. Zn (II)-Selective Membrane Electrode Based on Tetraazamacrocycle [Bzo2Me2Ph2(16)hexaeneN4]. Electroanalysis 2006, 18, 1186–1192. [Google Scholar] [CrossRef]
  37. Singh, P.; Singh, A.K.; Jain, A.K. Electrochemical sensors for the determination of Zn2+ ions based on pendant armed macrocyclic ligand. Electrochim. Acta 2011, 56, 5386–5395. [Google Scholar] [CrossRef]
  38. Ganjali, M.R.; Zamani, H.A.; Norouzi, P.; Adib, M.; Rezapour, M.; Aceedy, M. Zn2+ PVC-based membrane sensor based on 3-[(2-furylmethylene)amino]-2-thioxo-1,3-thiazolidin-4-one. Bull. Korean Chem. Soc. 2005, 26, 579–584. [Google Scholar] [CrossRef]
  39. Jain, A.K.; Gupta, V.K.; Ganeshpure, P.A.; Raisoni, J.R. Ni (II)-selective ion sensors of salen type Schiff base chelates. Anal. Chim. Acta 2005, 553, 177–184. [Google Scholar] [CrossRef]
  40. Kumar, K.G.; Poduval, R.; Augustine, P.; John, S.; Saraswathyamma, B. A PVC plasticized sensor for Ni (II) ion based on a simple ethylenediamine derivative. Anal. Sci. 2006, 22, 1333–1337. [Google Scholar] [CrossRef] [PubMed]
  41. Singh, L.P.; Bhatnagar, J.M. PVC based selective sensors for Ni2+ ions using carboxylated and methylated porphine. Sensors 2003, 3, 393–403. [Google Scholar] [CrossRef]
  42. Mousavi, M.F.; Alizadeh, N.; Shamsipur, M.; Zohari, N. A new PVC-based 1,10-dibenzyl-1,10-diaza-18-crown-6 selective electrode for detecting nickel (II) ion. Sens. Actuators B 2000, 66, 98–100. [Google Scholar] [CrossRef]
  43. Gupta, V.K.; Prasad, R.; Kumar, A. Dibenzocyclamnickel (II) as ionophore in PVC-matrix for Ni2+-selective sensor. Sensors 2000, 2, 384–396. [Google Scholar] [CrossRef]
  44. Qin, Y.; Mi, Y.; Bakker, E. Determination of complex formation constants of 18 neutral alkali and alkaline earth metal ionophores in poly (vinyl chloride) sensing membranes plasticized with bis (2-ethylhexyl) sebacate and o-nitrophenyloctylether. Anal. Chim. Acta 2000, 421, 207–220. [Google Scholar] [CrossRef]
  45. Mi, Y.; Bakker, E. Determination of complex formation constants of lipophilic neutral ionophores in solvent polymeric membranes with segmented sandwich membranes. Anal. Chem. 1999, 71, 5279–5287. [Google Scholar] [CrossRef] [PubMed]
  46. Bakker, E. Selectivity of liquid membrane ion-selective electrodes. Electroanalysis 1997, 9, 7–12. [Google Scholar] [CrossRef]
  47. Bereczki, R.; Takács, B.; Langmaier, J.; Neely, M.; Gyurcsányi, R.E.; Tóth, K.; Nagy, G.; Lindner, E. How to assess the limits of ion-selective electrodes: Method for the determination of the ultimate span, response range, and selectivity coefficients of neutral carrier-based cation selective electrodes. Anal. Chem. 2006, 78, 942–950. [Google Scholar] [CrossRef] [PubMed]
  48. Buck, R.P.; Lindner, E. Recommendations for nomenclature of ionselective electrodes (IUPAC Recommendations 1994). Pure Appl. Chem. 1994, 66, 2527–2536. [Google Scholar] [CrossRef]
  49. Lindner, E.; Umezawa, Y. Performance evaluation criteria for preparation and measurement of macro-and microfabricated ion-selective electrodes (IUPAC Technical Report). Pure Appl. Chem. 2008, 80, 85–104. [Google Scholar] [CrossRef]
  50. Umezawa, Y.; Bühlmann, P.; Umezawa, K.; Tohda, K.; Amemiya, S. Potentiometric selectivity coefficients of ion-selective electrodes. Part I. Inorganic cations (technical report). Pure Appl. Chem. 2000, 72, 1851–2082. [Google Scholar] [CrossRef]
  51. Ma, D.; Cai, Q.; Zhang, H. Mild method for Ullmann coupling reaction of amines and aryl halides. Org. Lett. 2003, 5, 2453–2455. [Google Scholar] [CrossRef] [PubMed]
  52. Craggs, A.; Moody, G.J.; Thomas, J.D.R. PVC matrix membrane ion-selective electrodes. Construction and laboratory experiments. J. Chem. Educ. 1974, 51, 541. [Google Scholar] [CrossRef]
  53. Kielland, J. Individual activity coefficients of ions in aqueous solutions. J. Am. Chem. Soc. 1937, 59, 1675–1678. [Google Scholar] [CrossRef]
Figure 1. Structure of ionophores.
Figure 1. Structure of ionophores.
Carbon 03 00034 g001
Figure 2. Trends of electromotive force (EMF) responses as a function of concentration by N,N-bis(2-nitrophenyl)benzene-1,2-diamine (TPA) towards different metal ions.
Figure 2. Trends of electromotive force (EMF) responses as a function of concentration by N,N-bis(2-nitrophenyl)benzene-1,2-diamine (TPA) towards different metal ions.
Carbon 03 00034 g002
Figure 3. Response of N-(2-nitrophenyl)benzene-1,2-diamine (DPA) towards different membrane compositions given in Table 1.
Figure 3. Response of N-(2-nitrophenyl)benzene-1,2-diamine (DPA) towards different membrane compositions given in Table 1.
Carbon 03 00034 g003
Figure 4. Potentiometric characteristics calibration curve for Zn2+ selective sensor based on DPA.
Figure 4. Potentiometric characteristics calibration curve for Zn2+ selective sensor based on DPA.
Carbon 03 00034 g004
Figure 5. Potentiometric characteristics calibration curve for Ni2+ selective sensor based on TPA.
Figure 5. Potentiometric characteristics calibration curve for Ni2+ selective sensor based on TPA.
Carbon 03 00034 g005
Figure 6. Effect of pH on the potential response of (a) Zn2+ISE based on DPA and (b) Ni2+ ISE based on TPA.
Figure 6. Effect of pH on the potential response of (a) Zn2+ISE based on DPA and (b) Ni2+ ISE based on TPA.
Carbon 03 00034 g006
Figure 7. Potentiometric titration of (a) Zn(II) (25 mL of 1 × 10−2 M) with EDTA (1 × 10−1 M) using Zn(II)-selective electrode based on DPA and (b) Ni(II) (20 mL of 1 × 10−2 M) with EDTA(1 × 10−1 M) using Ni(II)-selective electrode based on TPA.
Figure 7. Potentiometric titration of (a) Zn(II) (25 mL of 1 × 10−2 M) with EDTA (1 × 10−1 M) using Zn(II)-selective electrode based on DPA and (b) Ni(II) (20 mL of 1 × 10−2 M) with EDTA(1 × 10−1 M) using Ni(II)-selective electrode based on TPA.
Carbon 03 00034 g007
Figure 8. DFT optimized structures of DPA (a) and DPA-Zn2+ complex (b). White, red, blue, grey, and purple spheres refer to H, O, N, C, and Zn2+, respectively.
Figure 8. DFT optimized structures of DPA (a) and DPA-Zn2+ complex (b). White, red, blue, grey, and purple spheres refer to H, O, N, C, and Zn2+, respectively.
Carbon 03 00034 g008
Figure 9. DFT optimized structures of TPA (a) and TPA-Ni2+ complex (b). White, red, dark blue, grey and light blue spheres refer to H, O, N, C, and Ni2+, respectively.
Figure 9. DFT optimized structures of TPA (a) and TPA-Ni2+ complex (b). White, red, dark blue, grey and light blue spheres refer to H, O, N, C, and Ni2+, respectively.
Carbon 03 00034 g009
Table 1. Optimization of membrane components for Zn2+ selective electrode based on diphenylamine derivative.
Table 1. Optimization of membrane components for Zn2+ selective electrode based on diphenylamine derivative.
Membrane Electrode No.PVC (Wt %)Plasticizer (Wt %)Ionophore (Wt %)NaTPB (Wt %)Slope, mV/DecadeDetection Limit, (M) (Zn2+)
1336313201.0 × 10−5
2336421218.2 × 10−5
3336133301.3 × 10−6
4346123213.2 × 10−5
5336043366.3 × 10−5
6326053331.2 × 10−5
7325963203.6 × 10−5
Table 2. Optimization of membrane components for Ni2+ ISE based on triphenylamine derivative.
Table 2. Optimization of membrane components for Ni2+ ISE based on triphenylamine derivative.
Membrane Electrode No.Ionophore (% w/w)PVC (% w/w)NPOE (% w/w)NaTPB (% w/w)Slope, mV/decadeDetection Limit, (M) (Ni2+)
1-3367---
2-33652--
313365114.32.4 × 10−4
413265215.27.1 × 10−4
523364120.86.8 × 10−4
623265119.82.7 × 10−4
723264221.28.2 × 10−5
823363221.87.6 × 10−5
933262330.02.8 × 10−6
1033361326.59.2 × 10−6
1133363125.81.2 × 10−5
1243360337.04.3 × 10−6
1343261335.23.8 × 10−5
1453261234.01.2 × 10−5
1553260333.28.7 × 10−5
Table 3. Potentiometric selectivity coefficient values for DPA electrode by fixed interference method.
Table 3. Potentiometric selectivity coefficient values for DPA electrode by fixed interference method.
Metal Ionlog KZn2+,BMetal Ionlog KZn2+,B
K+−1.4Ni2+−2.0
Ca2+−2.7Cu2+−2.5
Na+−1.9Sr2+−2.3
Mg2+−2.1Ba2+−3.1
Pb2+−2.4Cd2+−1.9
Fe2+−2.8Co2+−2.7
Table 4. Potentiometric selectivity coefficient values for TPA electrode by fixed interference method.
Table 4. Potentiometric selectivity coefficient values for TPA electrode by fixed interference method.
Metal Ionlog KNi2+,BMetal Ionlog KNi2+,B
Na+−1.0Pb2+−2.6
Sr2+−3.1Fe2+−3.2
Cu2+−1.9Ca2+−3.1
Cd2+−2.9Ba2+−3.5
Mg2+−2.8Zn2+−2.1
Co2+−3.1Mn2+−3.2
Table 5. Comparison of proposed DPA-based zinc selective electrode with the reported electrodes.
Table 5. Comparison of proposed DPA-based zinc selective electrode with the reported electrodes.
S. No.IonophoreDetection Limit (M)pH RangeSlope (mV/Decade)Ref
1[Bzo2Me2Ph2(16)hexaeneN4]2.24 × 10−62.5–8.528.5[36]
26,7:14,15-Bzo2-10,11-(4-methylbenzene)-[15]-6,8,12,14-tetraene-9,12-N2-1,5-O23.3 × 10−73.0–8.028.8[37]
3Dibenzo-24-crown-89.2 × 10−54.8–6.229.0[26]
43-[(2-Furylmethylene)amino]-2-thioxo-1,3-thiazolidin-4-one8.5 × 10−73.0–7.029.3[38]
5N,N-Bis(acetylacetone)ethylenediamine8.9 × 10−53.2–7.130.0[25]
6This work (DPA)1.3 × 10−63.6–9.330.0-
Table 6. Comparison of proposed TPA-based Ni2+ selective electrode with the reported electrodes.
Table 6. Comparison of proposed TPA-based Ni2+ selective electrode with the reported electrodes.
S. No.IonophoreDetection Limit (M)pH RangeSlope (mV/decade)Ref
1Hexamethyl-1,4,8,11-tetraazacyclotetradeca-4,11-diene diperchlorate1.0 × 10−51.7–5.4Near-Nernstian[33]
2[Me4Bzo2TAA]7.9 × 10−62.7–7.630[34]
3N-(2-hydroxybenzylidene)-N-(2-picolyl)ethylenediamine3.2 × 10−62.2–5.929[39]
4N1,N2-bis((naphthalen-1-yl)-methylene)ethane-1,2-diamine1.3 × 10−63.6–7.429.9[40]
52,3,7,8,12,13,17,18-octamethyl-21H, 23H-porphine1.0 × 10−52–729[41]
61,10-dibenzyl-1,10-diaza-18-crown-62.0 × 10−54–8Nernstian[42]
75,7,8,14-tetramethyldibenzo[b,i]-1,4,8,11-tetrazacyclo tetradecane7.0 × 10−62–7.629.8[43]
8This work (TPA)2.8 × 10−64.2–9.230.0-

Share and Cite

MDPI and ACS Style

Kaur, H.; Chhibber, M.; Mittal, S.K. Acyclic Arylamine-Based Ionophores as Potentiometric Sensors for Zn2+ and Ni2+ Ions. C 2017, 3, 34. https://doi.org/10.3390/c3040034

AMA Style

Kaur H, Chhibber M, Mittal SK. Acyclic Arylamine-Based Ionophores as Potentiometric Sensors for Zn2+ and Ni2+ Ions. C. 2017; 3(4):34. https://doi.org/10.3390/c3040034

Chicago/Turabian Style

Kaur, Harpreet, Manmohan Chhibber, and Susheel K. Mittal. 2017. "Acyclic Arylamine-Based Ionophores as Potentiometric Sensors for Zn2+ and Ni2+ Ions" C 3, no. 4: 34. https://doi.org/10.3390/c3040034

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop