Next Article in Journal
The Ros/MucR Zinc-Finger Protein Family in Bacteria: Structure and Functions
Next Article in Special Issue
Non-Noble-Metal Mono and Bimetallic Composites for Efficient Electrocatalysis of Phosphine Oxide and Acetylene C-H/P-H Coupling under Mild Conditions
Previous Article in Journal
The Influence of Acitretin on Brain Lipidomics in Adolescent Mice—Implications for Pediatric and Adolescent Dermatological Therapy
Previous Article in Special Issue
Complexes of Sodium Pectate with Nickel for Hydrogen Oxidation and Oxygen Reduction in Proton-Exchange Membrane Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Conductive Mediators in Oxidation Based on Ferrocene Functionalized Phosphonium Ionic Liquids

by
Vadim V. Ermolaev
1,
Liliya R. Kadyrgulova
1,
Mikhail N. Khrizanforov
1,2,*,
Tatiana P. Gerasimova
1,
Gulnaz R. Baembitova
1,2,
Anna A. Lazareva
1,2 and
Vasili A. Miluykov
1
1
Arbuzov Institute of Organic and Physical Chemistry, FRC Kazan Scientific Center, Russian Academy of Sciences, 8 Arbuzov Street, 420088 Kazan, Russia
2
A.M. Butlerov Chemistry Institute, Kazan Federal University, Kremlevskaya Str. 18, 420008 Kazan, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(24), 15534; https://doi.org/10.3390/ijms232415534
Submission received: 11 November 2022 / Revised: 30 November 2022 / Accepted: 5 December 2022 / Published: 8 December 2022

Abstract

:
Herein, the synthesis of ferrocene-containing salts is presented. Acylation of ferrocene (Fc) according to the Friedel–Crafts method led to ω-bromoacyl ferrocenes. The ω-bromoacyl ferrocenes were subsequently introduced to quaternization reaction with tri-tert-butyl phosphine, which resulted in phosphonium salts. Obtained phosphonium salts were characterized by physical methods. The electrochemical properties of phosphonium salts were studied by cyclic voltammetry (CV). It was found that the replacement of n-butyl fragments at the phosphorus atom by tert-butyl leads to a more anodic potential shift. In contrast to isolobal structures Fc-C(O)(CH2)nP+(n-Bu)3X and Fc-(CH2)n+1P+(n-Bu)3X, the CV curves of Fc-C(O)(CH2)nP+(t-Bu)3X and Fc-(CH2)n+1P+(t-Bu)3X did not show a large discrepancy between forward and reverse currents. The transformation of the C=O groups to CH2 fragments had a significant effect on the electrochemical properties of ferrocene salts, the oxidation potential of which is close to that of pure ferrocene.

1. Introduction

In recent years, a significant number of reports has been devoted to materials that respond to external influences [1]. These materials find applications in the most advanced fields, such as sensors, electronic and optical technologies, biomedicine, molecular machines, and biomimetics [2]. Ionic liquids (ILs) are attractive components for creating such materials due to the fact that their chemical and physical properties can be easily tuned according to the desired application by carefully selecting cations and anions [3,4]. This certainly makes them excellent candidates as building blocks for “smart” materials [5]. A comparison of ammonium and phosphonium ionic liquids shows the advantages of the latter, since such compounds have higher thermal and chemical stability, as well as increased biological activity [6,7,8].
At the same time, the addition of different individual molecular synthons to ionic liquids can bring new desired properties. For example, it was shown that ferrocene-containing ionic liquids demonstrated important electrochemical, magnetic, and other properties of ILs [9,10,11,12].
Despite the fact that the ferrocene was discovered almost seventy years ago, the interest of chemists has not faded away [13]. Ferrocene itself and ferrocene-containing compounds find wide application [14]. Thus, it has been shown that ferrocenyl-substituted compounds, such as ferroceneacetic acid and ferrocenemethanol, can act as mediators of ascorbic acid oxidation, i.e., reducing the reaction potential [15].
The unique electrochemical properties of ferrocene allow for the design of electrochromic devices with enhanced properties. A fast response, high contrast, and lower voltage were achieved for the composition of viologen and redox-active ionic liquid containing a ferrocenyl fragment [16,17]. This combined material does not require any additive that is commonly used in the creation of electrochromic devices, such as solvents, supporting electrolytes, and sacrificial agents. On the other hand, the ferrocene fragment can influence the biological activity. Increased activity against Escherichia coli and Staphylococcus aureus bacteria has been demonstrated in imidazolium ionic liquids and IL membranes. This is explained by the formation of active oxygen particles caused by the presence of the redox-active Fc-moiety [18]. The universality of ferrocene chemistry is proved by using core-shell magnetic nanoparticles decorated with ferrocene-containing ILs as a nickel catalyst support in the synthesis of various pyrans’ derivatives in a one-pot manner under ultrasonic conditions [19,20]. A large amount of work was completed by Garib and Hirsch. A series of several new families of ferrocene-containing imidazolium salts were synthesized by etherification of ferrocene methanol, acylation of ferrocene, and amide coupling of ferrocene carboxylic acid [21]. These new redox-active ionic liquids were fully characterized.
At the same time, Fc-containing phosphonium ionic liquids are presented much less often in the literature. One of the first publications devoted to the electrochemical properties of ferrocene-functionalized phosphonium IL was reported in 2011 [22]. Ferrocene fragment was linked to tri-n-butyl phosphine moiety through C8 alkyl or acyl spacer. The authors discovered that Fc0/+ self-exchange reactions were responsible for electron transfer as opposed to physical diffusion of the phosphonium species. Later, Sundermeyer described phosphonium salt where the Fc-fragment was connected to the P-atom directly [23]. Several routes to achieve the desired product were evaluated.
We demonstrated that steric hindrance at phosphorus atom could play a key role in different aspects of phosphonium ionic liquids’ application [24,25,26]. Due to unique physical properties such as high conductivity, a very wide electrochemical window, stability, and reproducibility, phosphonium IL with sterically hindered cation was successfully applied in an electrochemical study of insoluble compounds [27,28]. Recently, we reported the synthesis and physical properties of the series of the ferrocenyl-containing sterically hindered phosphonium salts based on di(tert-butyl)ferrocenyl phosphine [29,30]. In development of this concept, we extended the series of ferrocene-containing phosphonium salts, where the ferrocene fragment and the phosphorus atom are linked by a spacer of different length and nature. Analysis of voltammograms of the obtained compounds revealed correlations between their structures and electrochemical properties.

2. Results and Discussion

In this study, the synthesis of ferrocene acyl phosphonium salts is presented. In the first step, ω-bromoacyl chlorides were obtained from corresponding carboxylic acid and thionyl chloride. The product was isolated by distillation under reduced pressure and then introduced in the reaction with ferrocene in the presence of Lewis acid (Scheme 1).
This synthesis was performed according to Hirsch’s report; however, the synthesis was performed without Zn powder. This allows obtaining the single desired product that was purified by column chromatography on silica.
At the next stage, the quaternization reaction of compounds 2ad with tri-tert-butyl-phosphine was carried out under inert atmosphere in absolute MeCN at the boiling point of MeCN. The reaction was monitored by 31P NMR. After the reaction, the finished product was washed several times with petroleum ether and diethyl ether (Scheme 2).
For further electrochemical study, the bromide-ion was changed to BF4 by simple anion exchange reaction in a minimal amount of water with subsequent extraction of salt with CH2Cl2. The organic layer was collected and washed several times with water and dried over Na2SO4. After filtration, the solvent was removed in vacuum, resulting in dark orange-brown-colored honey-like amorphous substances.
The obtained salts were characterized with the NMR 1H, 13C, 31P (Table 1). In the phosphorus spectrum, only one singlet at 49–50 ppm characteristic for tri-tert-butyl phosphonium salts [31] was observed.
At the same time, anion exchange leads to changes in the chemical shifts of protons P–CH2. They are not of the same nature due to the fact that they are exposed to different factors (location of the ferrocene fragment, carbonyl group, and phosphorus atom). The move of the chemical shift to high fields when Br was replaced with BF4 in the case of salts 3a4a can be explained by the proximity of the carbonyl group. The same movement of the chemical shift for 3c4c is possibly connected with interaction of carbonyl group trough space. Pairs 3b4b and 3d4d behave in the same was as like salts, with a normal chain as the fourth substituent.
Thermo-gravimetric analysis of obtained salts was performed (see Supplementary Materials). Only one salt, 4a, possesses phase transformation at 167 °C (Figure 1). This is probably connected with the fact that the salt with the short spacer forms some sort of ordered structure. All other salts did not demonstrate phase transition due to its honey-like amorphous state. It was found that all tetrafluoroborates are more thermally stable (up to 200 °C) in contrast to their bromides’ analogues (about 150 °C).
After thermal decomposition began, jitter occurred on all DSC graphs. Most likely it is caused by the vibration of the sample holder lid induced by intense exhaust gases.
It was interesting to obtain phosphonium salts containing an alkyl chain as a linker. To study the mutual influence of the carbonyl moiety on the cyclopentadienyl and phosphonium moieties, we carried out a reduction reaction using tert-butylamine-borane in the presence of Lewis acid (Scheme 3).
The reaction mixture was treated with water and products were extracted with CH2Cl2. Concentrated solution was placed in chromatographic column and the isolated ferrocenyl bromoalkanes were introduced in a quaternization reaction with tri-tret-butyl phosphine (Scheme 4). Anion replacement was carried out by interaction of the halides with a twofold excess of sodium tetrafluoroborate or bis(trifluoromethanesulfonyl)imide lithium. By employing one of such salts, we carried out a reversible electrochemical change in the color of the solution.
The electrochemical properties of phosphonium salts 4ad and 8ad were studied by cyclic voltammetry and compiled with data from the literature analogue (Figure 2; Table 1). It was found that the replacement of n-Bu fragments at the phosphorus atom with t-Bu led to a more anodic potential shift (Table 2).
The observed effect can be described by the fact that, in addition to the acceptor properties of the acyl group, on the electronic effect in the case of salts with t-Bu, phosphorus, due to the spatial effect, pulls away a part of the electron density of the cyclopentadiene ring.
R.W. Murray et al. [22] also noted that the Fc-C(O)(CH2)nP(n-Bu)3X and Fc-C(CH2)n+1P(n-Bu)3X salts have weak ionic conductivity. On the CV curves, a large discrepancy between the forward and reverse currents was observed. Such an effect is not observed for salts 4ad and 8ad. On the contrary, these salts can be used as an electrolyte.
It was noted earlier that, based on previous works, the phosphonium cation with normal alkyl substituents has a low diffusion coefficient (10−8–10−9 cm2/s), which is associated with the interaction of the chains with the medium. At the same time, the sterically hindered phosphonium cation described in this work is more compact, which is reflected in the high D value (Table 2).
When iron is oxidized (to ferrocenium form), a characteristic color change from yellow to dark green is observed; when re-reduced, the system reversibly passes into its original form (Figure S1). The color change is controlled by the difference between forward and reverse current.
Table 2. Electrochemical data for the redox properties of 4ad and 8ad.
Table 2. Electrochemical data for the redox properties of 4ad and 8ad.
Compound1/2Eox,
V vs. Ag/AgCl
ΔEPEAK, VDCV, cm2/s
Ferrocene0.410.061.30 × 10−7 [32]
4a0.770.061.72 × 10−5
4b0.710.061.71 × 10−5
4c0.710.061.71 × 10−5
The analog of 4c (Fc-C(O)(CH2)5P(n-Bu)3PF6) [22]0.600.306.40 × 10−9
4d0.710.061.68 × 10−5
8a0.460.061.64 × 10−5
8b0.440.061.65 × 10−5
8c0.430.061.62 × 10−5
The analog of 8c (Fc(CH2)6P(n-Bu)3PF6) [22]0.250.301.53 × 10−8
8d0.420.061.58 × 10−5
The diffusion coefficients DCV were calculated based on the electrochemical data [33].
At CVs for all salts, the ΔEPEAK does not exceed 60 mV. The transformation of the C=O groups to CH2 fragments has a significant effect on the electrochemical properties of ferrocene salts, the oxidation potential of which is close to that of pure ferrocene. An increase in the alkyl chain in the composition of ferrocene salts does not lead to a significant potential shift (when going from C3 to C11, the potential shift remains 50 mV) (averaged measurements of 10 experiments for each substance).
The generated ferrocenium can be used as an oxidizing agent (or oxidation mediator) for, for example, ascorbic acid. When ascorbic acid is added to a solution of oxidized ferrocenyl-phosphonium salt, a change in the color of the solution to the original yellow color is observed, which characterizes the implementation of the oxidation reaction (Figure 3).
The color changes can be detected by UV/Vis spectroscopy. The initial yellow salt 7a absorbs at 434 nm (black spectrum at Figure 4) that corresponds to the Fc-moiety. After oxidation, an additional band at 463 nm appears (green spectrum at Figure 4). The addition of ascorbic acid to neutral 7a leads to the appearance of its own band at 241 nm (red spectrum at Figure 4). However, in the spectrum of reaction mixture, the intensity of this band significantly decreases due to its oxidation (blue spectrum at Figure 4), whereas the spectral pattern at 350–550 nm is close to the observed for oxidized form of 7a.
This, in turn, demonstrates rather mild oxidation conditions, which makes ferrocenyl-phosphonium salts convenient electrolytes and mediators at the same time. On the CV curves (Figure 5), with the addition of increasing amounts of ascorbic acid, an increase in current is observed corresponding to the transition of Fe (II) to Fe (III).
The reaction mixture and initial compounds were analyzed by ESI-MS. In the spectrum of the reaction mixture, we did not observe any peak of the starting ascorbic acid. Signals of oxidative degradation products are recorded, which are comparable to the literature data [34]. According to the spectrum, we observed products of both partial and complete oxidative degradation (Figure S2).
Thus, given the high ionic conductivity, the obtained salts can be used as oxidation mediators, where the use of additional electrolytes is not required.

3. Materials and Methods

3.1. NMR Experiments

NMR spectra were recorded with multi-nuclear spectrometer Bruker AVANCE-400 (BRUKER BioSpin GMBH am Silberstreifen, D-76287, Rheinstetten, Germany) (400.1 MHz (1H), 100.6 MHz (13C) and 162.0 MHz (31P)). Chemical shifts were given in parts per million relative to SiMe4 (1H, internal solvent) and 85% H3PO4 (31P, external).

3.2. TG-DSC

Thermogravimetric analysis was performed on the NETZSCH STA 449F1 (NETZSCH, Selb, Germany) with a heating rate 10 K per minute up to 400 °C in an argon atmosphere.

3.3. Mass-Spectra

Mass Electrospray ionization mass spectrometry (ESI-MS) was performed on the AmazonX mass spectrometer (Bruker Daltonik GmbH, Bremen, Germany). The measurements were carried out in the positive/negative ion detection mode in the m/z range from 100 to 1000. The voltage on the capillary was 140 V. Data were processed using the DataAnalysis 4.0 program (Bruker Daltonik GmbH, Bremen, Germany).

3.4. Electrochemical Measurements

Electrochemical measurements were taken on a BASi Epsilon EClipse electrochemical analyzer (West Lafayette, IN, USA). The program concerned Epsilon-ECUSB-V200 waves at potential scan rate t = 100 mV∙s−1 in CH3CN/without an electrolyte (except clear ascorbic acid—for ascorbic acid, a 0.1 M solution of Bu4NBF4 was used) at 295 K. A glassy carbon working electrode (ð = 3 mm) embedded in Teflon and Pt wire as counter electrode was used in the electrochemical cell. Before each measurement, the surface of the working electrode was mechanically polished.

3.5. Reagents and Research Subjects

All the work related to the preparation of the starting substrates, as well as the synthesis and the workup of products, was carried out in an inert atmosphere using the standard Schlenk apparatus. Tri-tert-butylphosphine (Dal-Chem, Nizhny Novgorod, Russia), halogenated acids (Sigma-Aldrich, St. Louis, MO, USA), anhydrous aluminum chloride (Alfa Aesar, Heysham, Lancashire, UK), Thionyl Chlride (Himreaktiv, Nizhny Novgorod, Russia), Ferrocene (Thermo Fisher (Kandel) GmbH, Kandel, Germany). All solvents and purchased reagents were absolute by the appropriate methods, mainly by distillation in an inert atmosphere. Detailed characterization of the phosphonium salts described in Supplementary Materials (Figures S3–S86).

3.6. General Procedure

3.6.1. General Procedure for the Synthesis of ω-Bromoalkanoyl Chloride 1a1d (Scheme S1)

At room temperature, halogenated acid (1 eq.) and SOCl2 (4.2 eq.) were added with a stirrer to a 50 mL Schlenk vessel and gradually warmed to reflux. The apparatus should be connected to a bubbler and an exhaust gas absorbing device. The reaction mixture was refluxed within 4 h and then was stirred within 12 h at room temperature. The crude product was purified by vacuum distillation.

3.6.2. General Procedure for the Synthesis of 2a2d

Under an inert gas atmosphere, the ω-bromoalkanoyl chloride (1 eq.) and anhydrous AlCl3 (1 eq.) were added to dry CH2Cl2 (50 mL) at 0 °C. The mixture was stirred for ten minutes before the addition of ferrocene (1.2 eq.) at the same temperature. The appearance of a purple color signaled the beginning of the acylation process. Stirring was continued after addition overnight while the mixture was allowed to warm up to room temperature. The mixture was again cooled to 0 °C by an ice bath, and water (40 mL) was carefully added under an inert gas atmosphere to quench the reaction. After quenching, more H2O (50 mL) was added and the mixture was stirred for 30 min at 0 °C. The aqueous phase was extracted four times with CH2Cl2 (4 × 60 mL). The combined organic extracts were washed sequentially with sat. NaHCO3 (2 × 100 mL), H2O (100 mL), and brine (100 mL). The organic phase was dried over MgSO4, concentrated, and the residue purified by column chromatography (SiO2; petroleum ether/ethyl acetate). All products were dried in vacuo after purification.

3.6.3. General Procedure for the Synthesis of 3a3d

Bromoalkanoyl ferrocene was dissolved in dry MeCN (20 mL) and an equivalent amount of tri-tert-butyl phosphine was added. The reaction mixture was stirred at 82 °C for 6 h. The solvent was removed in vacuo and the product was stirred with 3 portions of 20 mL of petroleum ether and 3 portions of 20 mL of diethyl ether. After each stirring, the solvent was removed with a cannula with a filter nozzle, and the solvent residues were removed in vacuo.

3.6.4. General Procedure for the Synthesis of 4a4d

Salts 3ad were dissolved in 10 mL of water, and a twofold excess of sodium tetrafluoroborate solution in water was added to them. The reaction mixture was stirred for 12 h. The precipitate was filtered, and the solvent residues were removed in vacuo. The salt was dissolved in 20 mL of methylene chloride and washed with distilled water. The organic phase was separated and dried over magnesium sulfate. The solvent was evaporated, and the salt was dried under vacuum at 40 °C for 8 h.

3.6.5. General Procedure for the Synthesis of 5a5d

Salts 3ad were dissolved in 10 mL of water, a one and a half excess of lithium bis(trifluoromethanesulfonyl)imide solution in water was added to them. The reaction mixture was stirred for 1 h. The product was extracted with CH2Cl2. The extract was washed with distilled water. The organic phase was separated and dried over magnesium sulfate. The solvent was evaporated, and the salt was dried under vacuum at 50 °C for 8 h.

3.6.6. General Procedure for the Synthesis of 6a6d

The tert-Butylamine borane (6 eq.) was added to a suspension of anhydrous AlCl3 (3 eq.) in dry CH2Cl2 (40 mL) at 0 °C under an inert gas atmosphere. The resulting mixture was stirred at the same temperature for 1 h. The solution of bromoalkanoyl ferrocene (1 eq.) in CH2Cl2 was added dropwise. The resulting solution was stirred overnight. The initial purple color of the solution changed to brown during the reduction process. This indicated the end of the reaction. The reaction mixture was cooled to 0 °C and water (20 mL) was added dropwise to remove excess of the reducing agents. More H2O (20 mL) was added and the solution was stirred for 30 min. The phases were separated and the aqueous phase was extracted with CH2Cl2 until the organic extract was colorless. The combined organic extracts were sequentially washed with sat. NaHCO3 (2 × 20 mL), H2O (20 mL), and brine (20 mL). The organic phase was dried over MgSO4, and all products were dried in vacuo after purification.

3.6.7. General Procedure for the Synthesis of 7a7d

Tri-tert-butyl phosphine was dissolved in dry DMF (5 mL) and an equivalent amount of bromoalkyl ferrocene was added. The reaction mixture was stirred at 55 °C for 5–20 h. The solvent was removed in vacuo and the product was stirred with 3 portions of 20 mL of petroleum ether and 3 portions of 20 mL of diethyl ether. The precipitate was filtered out and the solvent residues were removed in vacuo.

3.6.8. General Procedure for the Synthesis of 8a8d

Salts 6ad were dissolved in 10 mL of EtOH, and a one and a half excess of lithium bis(trifluoromethanesulfonyl)imide solution in EtOH was added to them. The reaction mixture was stirred for 1 h. The product was extracted with CH2Cl2. The extract was washed with distilled water. The organic phase was separated and dried over magnesium sulfate. The solvent was evaporated, and the salt was dried under vacuum at 50 °C for 8 h.

4. Conclusions

In this study, the synthesis of ferrocene containing phosphonium salts with sterically hindered cation was described. Obtained salts were characterized with several physical methods which are listed as follows: NMR 1H; 13C; 31P; ESI-MS; TG-DSC; IR.
It was found by cyclic voltammetry that the replacement of n-Bu fragments at the phosphorus atom with t-Bu leads to a more anodic potential shift. For the obtained phosphonium salts on CV, the discrepancy between forward and reverse currents does not exceed 60 mV, in contrast to isolobal structures with n-alkyl fragments. It was established that the phosphonium cation with t-Bu substituents has a high diffusion coefficient, which is associated with a more compact structure. In this regard, salts can be simultaneously used as mediators and electrolytes.
We suppose that the ferrocene moiety is responsible for the cation stabilization and facilitating the charge transfer. Altogether, this helps to define the trends in fundamental research on new materials for electrochemical application.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232415534/s1.

Author Contributions

Conceptualization, V.V.E. and M.N.K., methodology, V.V.E., M.N.K., T.P.G., L.R.K., G.R.B. and V.A.M.; formal analysis, V.V.E., L.R.K., G.R.B., A.A.L., M.N.K. and T.P.G.; investigation, V.V.E., T.P.G., L.R.K., G.R.B., A.A.L. and M.N.K.; resources, V.V.E. and M.N.K.; data curation, V.V.E. and M.N.K.; writing—original draft preparation, V.V.E., M.N.K. and T.P.G.; writing—review and editing, V.V.E., M.N.K., T.P.G. and V.A.M.; visualization, V.V.E. and M.N.K.; supervision, V.A.M.; project administration, M.N.K.; funding acquisition, V.V.E. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation and the Cabinet of Ministers of the Republic of Tatarstan within the framework of the scientific project No. 22-23-20196.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are contained within the article or are available upon request from the corresponding author, Mikhail N. Khrizanforov.

Acknowledgments

The measurements have been carried out using the equipment of Distributed Spectral-Analytical Center of Shared Facilities for Study of Structure, Composition and Properties of Substances and Materials of FRC Kazan Scientific Center of RAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wieszczycka, K.; Staszak, K.; Woźniak-Budych, M.J.; Litowczenko, J.; Maciejewska, B.M.; Jurga, S. Surface functionalization–The way for advanced applications of smart materials. Coord. Chem. Rev. 2021, 436, 213846. [Google Scholar] [CrossRef]
  2. Mrinalini, M.; Prasanthkumar, S. Recent Advances on Stimuli-Responsive Smart Materials and their Applications. ChemPlusChem 2019, 84, 1103–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Welton, T. Ionic liquids: A brief history. Biophys. Rev. 2018, 10, 691–706. [Google Scholar] [CrossRef] [Green Version]
  4. Egorova, K.S.; Gordeev, E.G.; Ananikov, V.P. Biological Activity of Ionic Liquids and Their Application in Pharmaceutics and Medicine. Chem. Rev. 2017, 117, 7132–7189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Cui, J.; Li, Y.; Chen, D.; Zhan, T.-G.; Zhang, K.-D. Ionic Liquid-Based Stimuli-Responsive Functional Materials. Adv. Funct. Mater. 2020, 30, 2005522. [Google Scholar] [CrossRef]
  6. Khazalpour, S.; Yarie, M.; Kianpour, E.; Amani, A.; Asadabadi, S.; Seyf, J.Y.; Rezaeivala, M.; Azizian, S.; Zolfigol, M.A. Applications of phosphonium-based ionic liquids in chemical processes. J. Iran. Chem. Soc. 2020, 17, 1775–1917. [Google Scholar] [CrossRef]
  7. Zhang, H.; Liu, L.; Hou, P.; Pan, H.; Fu, S. Polyisocyanide Quaternary Ammonium Salts with Exceptionally Star-Shaped Structure for Enhanced Antibacterial Properties. Polymers 2022, 14, 1737. [Google Scholar] [CrossRef]
  8. Metelytsia, L.O.; Hodyna, D.M.; Semenyuta, I.V.; Kovalishyn, V.V.; Rogalsky, S.P.; Derevianko, K.Y.; Tetko, I.V. Theoretical and Experimental Studies of Phosphonium Ionic Liquids as Potential Antibacterials of MDR Acinetobacter baumannii. Antibiotics 2022, 11, 491. [Google Scholar] [CrossRef]
  9. Rochefort, D. Enabling new electrochemical methods with redox-active ionic liquids. Cur. Opin. Electrochem. 2019, 15, 125–132. [Google Scholar] [CrossRef]
  10. Gholamhosseini-Nazari, M.; Esmati, S.; Safa, K.D.; Khataee, A.; Teimuri-Mofrad, R. Synthesis and application of novel 1,2,3-triazolylferrocene-containing ionic liquid supported on Fe3O4 nanocatalyst in the synthesis of new pyran-substituted Betti bases: Novel nanocatalyst for the synthesis of pyran substituted Betti bases. Appl. Organomet. Chem. 2019, 33, e4701. [Google Scholar] [CrossRef]
  11. Arkhipova, E.A.; Ivanov, A.S.; Maslakov, K.I.; Kupreenko, S.Y.; Levin, M.M.; Iurenkov, M.V.; Lyssenko, K.A.; Savilov, S.V.; Aldoshin, S.M. Ferrocene based ionic liquid: Synthesis, structure, transport properties and mechanism of thermal degradation. J. Mol. Liq. 2022, 355, 118933. [Google Scholar] [CrossRef]
  12. Shen, G.; Shen, Y. Label-Free Electrochemical Immunosensor Based on β-Cyclodextrin-Functionalized Helical Carbon Nanotube and Ionic Liquid Containing Ferrocene and Aldehyde Groups. ACS Omega 2019, 4, 20252–20256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Astruc, D. Why is Ferrocene so Exceptional? Eur. J. Inorg. Chem. 2017, 1, 6–29. [Google Scholar] [CrossRef]
  14. Elisabeth, S.P. Ferrocenes: Compounds, Properties and Applications; Nova Science Publishers: Hauppauge, NY, USA, 2011; ISBN 978-1-61761-880-2. [Google Scholar]
  15. Pournaghi-Azar, M.H.; Ojani, R. Catalytic oxidation of ascorbic acid by some ferrocene derivative mediators at the glassy carbon electrode. Application to the voltammetric resolution of ascorbic acid and dopamine in the same sample. Talanta 1995, 42, 1839–1848. [Google Scholar] [CrossRef] [PubMed]
  16. Tahara, H.; Baba, R.; Iwanaga, K.; Sagara, T.; Murakami, H. Electrochromism of a bipolar reversible redox-active ferrocene–viologen linked ionic liquid. Chem. Commun. 2017, 53, 2455–2458. [Google Scholar] [CrossRef] [Green Version]
  17. Tahara, H.; Uranaka, K.; Hirano, M.; Ikeda, T.; Sagara, T.; Murakami, H. Electrochromism of Ferrocene- and Viologen-Based Redox-Active Ionic Liquids Composite. ACS Appl. Mater. Interfaces 2019, 11, 1–6. [Google Scholar] [CrossRef]
  18. Zhang, T.; Guo, J.; Ding, Y.; Mao, H.; Yan, F. Redox-responsive ferrocene-containing poly(ionic liquid)s for antibacterial applications. Sci. China Chem. 2018, 62, 95–104. [Google Scholar] [CrossRef]
  19. Teimuri-Mofrad, R.; Esmati, S.; Rabiei, M.; Gholamhosseini-Nazari, M. Efficient synthesis of new pyrano [3,2-b] pyran derivatives via Fe3O4@SiO2-IL-Fc catalyzed three-component reaction. Heterocycl. Commun. 2017, 23, 439–444. [Google Scholar] [CrossRef]
  20. Mohammadi, R.; Esmati, S.; Gholamhosseini-Nazari, M.; Teimuri-Mofrad, R.J. Novel ferrocene substituted benzimidazolium based ionic liquid immobilized on magnetite as an efficient nano-catalyst for the synthesis of pyran derivatives. Mol. Liq. 2019, 275, 523–534. [Google Scholar] [CrossRef]
  21. Gharib, B.; Hirsch, A. Synthesis and Characterization of New Ferrocene-Containing Ionic Liquids. Eur. J. Org. Chem. 2014, 19, 4123–4136. [Google Scholar] [CrossRef]
  22. Weaver, J.E.F.; Breadner, D.; Deng, F.; Ramjee, B.; Ragogna, P.J.; Murray, R.W. Electrochemistry of Ferrocene-Functionalized Phosphonium Ionic Liquids. J. Phys. Chem. C 2011, 115, 19379–19385. [Google Scholar] [CrossRef]
  23. Kübler, P.; Sundermeyer, J. Ferrocenyl-phosphonium ionic liquids—Synthesis, characterisation and electrochemistry. Dalton Trans. 2014, 43, 3750–3766. [Google Scholar] [CrossRef]
  24. Arkhipova, D.M.; Ermolaev, V.V.; Miluykov, V.A.; Valeeva, F.G.; Gaynanova, G.A.; Zakharova, L.Y.; Minyaev, M.E.; Ananikov, V.P. Tri-tert-butyl(n-alkyl)phosphonium Ionic Liquids: Structure, Properties and Application as Hybrid Catalyst Nanomaterials. Sustainability 2021, 13, 9862. [Google Scholar] [CrossRef]
  25. Ermolaev, V.V.; Arkhipova, D.M.; Miluykov, V.A.; Lyubina, A.P.; Amerhanova, S.K.; Kulik, N.V.; Voloshina, A.D.; Ananikov, V.P. Sterically Hindered Quaternary Phosphonium Salts (QPSs): Antimicrobial Activity and Hemolytic and Cytotoxic Properties. Int. J. Mol. Sci. 2022, 23, 86. [Google Scholar] [CrossRef] [PubMed]
  26. Arkhipova, D.M.; Ermolaev, V.V.; Milyukov, V.A.; Valeeva, F.G.; Gaynanova, G.A.; Zakharova, L.Y. Micellar nanocontainers based on sterically hindered cationic phosphonium amphiphiles. Rus. Chem. Bull. 2022, 71, 804–811. [Google Scholar] [CrossRef]
  27. Khrizanforov, M.N.; Arkhipova, D.M.; Shekurov, R.P.; Gerasimova, T.P.; Ermolaev, V.V.; Islamov, D.R.; Miluykov, V.A.; Kataeva, O.N.; Khrizanforova, V.V.; Sinyashin, O.G.; et al. Novel paste electrodes based on phosphonium salt room temperature ionic liquids for studying the redox properties of insoluble compounds. J. Solid State Electrochem. 2015, 19, 2883–2890. [Google Scholar] [CrossRef]
  28. Khrizanforov, M.N.; Shekurov, R.P.; Ermolaev, V.V.; Arkhipova, D.M.; Miluykov, V.A.; Kataeva, O.N.; Khrizanforova, V.V.; Budnikova, Y.H. Novel phosphonium salt for paste electrode to study the redox properties of insoluble compounds. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 1611–1612. [Google Scholar] [CrossRef]
  29. Ermolaev, V.; Gerasimova, T.; Kadyrgulova, L.; Shekurov, R.; Dolengovski, E.; Kononov, A.; Miluykov, V.; Sinyashin, O.; Katsyuba, S.; Budnikova, Y.; et al. Ferrocene-Containing Sterically Hindered Phosphonium Salts. Molecules 2018, 23, 2773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Kadyrgulova, L.; Ermolaev, V.; Gilmanova, L.; Miluykov, V. Novel sterically hindered ferrocenyl-phosphonium salt. Phosphorus Sulfur Silicon Relat. Elem. 2019, 194, 989–991. [Google Scholar] [CrossRef]
  31. Arkhipova, D.M.; Ermolaev, V.V.; Miluykov, V.A.; Gubaidullin, A.T.; Islamov, D.R.; Kataeva, O.N.; Ananikov, V.P. Sterically Hindered Phosphonium Salts: Structure, Properties and Palladium Nanoparticle Stabilization. Nanomaterials 2020, 10, 2457. [Google Scholar] [CrossRef]
  32. Eisele, S.; Schwarz, M.; Speiser, B.; Tittel, C. Diffusion coefficient of ferrocene in 1-butyl-3-methylimidazolium tetrafluoroborate—Concentration dependence and solvent purity. Electrochim. Acta 2006, 51, 5304–5306. [Google Scholar] [CrossRef]
  33. Zanello, P.; Nervi, C.; De Biani, F.F. Inorganic electrochemistry: Theory, practice and application. R. Soc. Chem. 2019. [Google Scholar] [CrossRef]
  34. Cimino, P.; Troiani, A.; Pepi, F.; Garzoli, S.; Salvitti, C.; Di Rienzo, B.; Barone, V.; Ricci, A. From ascorbic acid to furan derivatives: The gas phase acid catalyzed degradation of vitamin C. Phys. Chem. Chem. Phys. 2018, 20, 17132–17140. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. ω-bromoacyl ferrocene synthesis.
Scheme 1. ω-bromoacyl ferrocene synthesis.
Ijms 23 15534 sch001
Scheme 2. Sterically hindered phosphonium salts synthesis (3ad) and anion exchange reaction (4ad, 5ad).
Scheme 2. Sterically hindered phosphonium salts synthesis (3ad) and anion exchange reaction (4ad, 5ad).
Ijms 23 15534 sch002
Figure 1. TG-DSC plot for 4a.
Figure 1. TG-DSC plot for 4a.
Ijms 23 15534 g001
Scheme 3. Reduction of bromoalkanoyl ferrocene.
Scheme 3. Reduction of bromoalkanoyl ferrocene.
Ijms 23 15534 sch003
Scheme 4. Synthesis tri-tert-butyl(3-ferrocenylalkyl) phosphonium bromides and anion exchange reaction.
Scheme 4. Synthesis tri-tert-butyl(3-ferrocenylalkyl) phosphonium bromides and anion exchange reaction.
Ijms 23 15534 sch004
Figure 2. Cyclic voltammograms of phosphonium salts 4ad and 8ad in CH3CN on glassy carbon (GC) electrode (0.5 mM complex). The potentials vs. Ag/AgCl. Scan rate = 100 mV s−1, room temperature.
Figure 2. Cyclic voltammograms of phosphonium salts 4ad and 8ad in CH3CN on glassy carbon (GC) electrode (0.5 mM complex). The potentials vs. Ag/AgCl. Scan rate = 100 mV s−1, room temperature.
Ijms 23 15534 g002
Figure 3. (a) Photo of oxidized salt 7a WITHOUT an additional electrolyte in CH3CN; (b) Photo of oxidized salt 7a in CH3CN by addition of 1 eq. ascorbic acid; (c) Photo of oxidized salt 7a in CH3CN by addition of 1 eq. ascorbic acid after 1 h.
Figure 3. (a) Photo of oxidized salt 7a WITHOUT an additional electrolyte in CH3CN; (b) Photo of oxidized salt 7a in CH3CN by addition of 1 eq. ascorbic acid; (c) Photo of oxidized salt 7a in CH3CN by addition of 1 eq. ascorbic acid after 1 h.
Ijms 23 15534 g003
Figure 4. UV-Vis spectra of 7a (black), oxidized form of 7a (green), mixture of 7a and ascorbic acid as prepared (red), and reaction mixture of oxidized form of 7a and ascorbic acid (blue).
Figure 4. UV-Vis spectra of 7a (black), oxidized form of 7a (green), mixture of 7a and ascorbic acid as prepared (red), and reaction mixture of oxidized form of 7a and ascorbic acid (blue).
Ijms 23 15534 g004
Figure 5. Cyclic voltammograms of phosphonium salt 7a (indicated on the chart as Fc*) in CH3CN on glassy carbon (GC) electrode (0.5 mM complex) in the presence of increasing amounts of ascorbic acid. The solid curve shows the CV of AA oxidation. The potentials vs. Ag/AgCl. Scan rate = 100 mV s−1, room temperature.
Figure 5. Cyclic voltammograms of phosphonium salt 7a (indicated on the chart as Fc*) in CH3CN on glassy carbon (GC) electrode (0.5 mM complex) in the presence of increasing amounts of ascorbic acid. The solid curve shows the CV of AA oxidation. The potentials vs. Ag/AgCl. Scan rate = 100 mV s−1, room temperature.
Ijms 23 15534 g005
Table 1. 1H NMR chemical shifts of salts.
Table 1. 1H NMR chemical shifts of salts.
Compoundδ(C(O)CH2), ppmδ(P-CH2), ppm
3a3.593.24
4a3.502.92
3b2.842.04
4b2.942.42
3c2.852.85
4c2.842.41
3d2.711.94
4d2.712.35
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ermolaev, V.V.; Kadyrgulova, L.R.; Khrizanforov, M.N.; Gerasimova, T.P.; Baembitova, G.R.; Lazareva, A.A.; Miluykov, V.A. Conductive Mediators in Oxidation Based on Ferrocene Functionalized Phosphonium Ionic Liquids. Int. J. Mol. Sci. 2022, 23, 15534. https://doi.org/10.3390/ijms232415534

AMA Style

Ermolaev VV, Kadyrgulova LR, Khrizanforov MN, Gerasimova TP, Baembitova GR, Lazareva AA, Miluykov VA. Conductive Mediators in Oxidation Based on Ferrocene Functionalized Phosphonium Ionic Liquids. International Journal of Molecular Sciences. 2022; 23(24):15534. https://doi.org/10.3390/ijms232415534

Chicago/Turabian Style

Ermolaev, Vadim V., Liliya R. Kadyrgulova, Mikhail N. Khrizanforov, Tatiana P. Gerasimova, Gulnaz R. Baembitova, Anna A. Lazareva, and Vasili A. Miluykov. 2022. "Conductive Mediators in Oxidation Based on Ferrocene Functionalized Phosphonium Ionic Liquids" International Journal of Molecular Sciences 23, no. 24: 15534. https://doi.org/10.3390/ijms232415534

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