Next Article in Journal
Cytotoxic Compounds Derived from Marine Sponges. A Review (2010–2012)
Next Article in Special Issue
On the Morphology of Group II Metal Fluoride Nanocrystals at Finite Temperature and Partial Pressure of HF
Previous Article in Journal
Identification of Larvicidal Constituents of the Essential Oil of Echinops grijsii Roots against the Three Species of Mosquitoes
Previous Article in Special Issue
Hydrodechlorination of Tetrachloromethane over Palladium Catalysts Supported on Mixed MgF2-MgO Carriers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal Fluorides, Metal Chlorides and Halogenated Metal Oxides as Lewis Acidic Heterogeneous Catalysts. Providing Some Context for Nanostructured Metal Fluorides

School of Chemistry, University of Glasgow, G12 8QQ Glasgow, UK
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(2), 201; https://doi.org/10.3390/molecules22020201
Submission received: 1 December 2016 / Revised: 13 January 2017 / Accepted: 17 January 2017 / Published: 28 January 2017
(This article belongs to the Special Issue Nano-sized Metal Fluorides: Novel Approaches to Lewis Acid Catalysts)

Abstract

:
Aspects of the chemistry of selected metal fluorides, which are pertinent to their real or potential use as Lewis acidic, heterogeneous catalysts, are reviewed. Particular attention is paid to β-aluminum trifluoride, aluminum chlorofluoride and aluminas γ and η, whose surfaces become partially fluorinated or chlorinated, through pre-treatment with halogenating reagents or during a catalytic reaction. In these cases, direct comparisons with nanostructured metal fluorides are possible. In the second part of the review, attention is directed to iron(III) and copper(II) metal chlorides, whose Lewis acidity and potential redox function have had important catalytic implications in large-scale chlorohydrocarbons chemistry. Recent work, which highlights the complexity of reactions that can occur in the presence of supported copper(II) chloride as an oxychlorination catalyst, is featured. Although direct comparisons with nanostructured fluorides are not currently possible, the work could be relevant to possible future catalytic developments in nanostructured materials.

1. Introduction

Close-packed, solid metal fluorides have, for the most part, relatively small surface areas, although in some cases the surface metal cations have significant Lewis acidity due to their highly electronegative fluoride anion nearest-neighbors. For heterogeneous catalysis, partially fluorinated oxide surfaces have often been preferred, in view of their significantly increased surface areas without great loss in Lewis acidity.
The development of nanostructured metal fluorides, which has been rapid over the past 10–15 years [1], has resulted in renewed interest in the possible catalytic applications of these high-surface-area metal fluorides and it is timely to consider how they compare as catalysts with materials prepared by more conventional routes. Catalytic studies are described explicitly in other contributions to this themed collection; here, in the first part of the review, we describe some surface properties of what might be termed ‘competitor catalysts’. The emphasis is on various forms of aluminum fluoride and γ-alumina, which have been fluorinated using various fluorinating agents. Although fluorinated chromia is also an obvious comparator, its catalytic activity is not dealt with here, since we have very recently compiled a detailed account of this topic [2].

2. Aluminum Fluorides and Halogenated Aluminas

Two of the important aluminum fluorides, which can be regarded as precursors to nanoscale metal fluorides, are β-aluminum trifluoride and aluminum chlorofluoride (ACF). The former contains a six-coordinate AlIII but, unlike α-AlF3, which is close-packed, has the more open hexagonal tungsten bronze (HTB) structure [3]. This phase behaves as a solid Lewis acid and has significant catalytic behavior [4]; a model for the coordinatively unsaturated surface AlIII species has been deduced [5]. In contrast, ACF was developed from industrial research in fluorocarbon chemistry [6]. Judged by its chemical behavior, its Lewis acidity is comparable to that of liquid antimony pentafluoride, a benchmark in the field. Synthesized by halogen exchange between solid aluminum trichloride and various chlorofluorocarbons, hydrochlorofluorocarbons or hexafluoropropene, the reactions do not result in complete replacement of Cl by F; the stoichiometry of the solid product is AlClxF3−x, with x being in the range 0.05–0.3. Surface areas are in the range 100–150 m2·g−1. The structure of the amorphous solid has been deduced at the atomic level from a number of spectroscopic techniques [7]. Chlorine is distributed throughout the solid, rather than in a discrete AlCl3 phase, and is possibly bridged to three AlIII centers. There appear to be three different types of six-coordinated AlIII.
The defect spinel oxide γ-alumina is an ideal base material to investigate the effects of fluorination upon an oxide surface. The surface properties depend on both the reagent employed and the conditions used, as illustrated in Table 1.
These fluorinated aluminas are not supported reagents, since the fluorine is incorporated within the surface, and in some cases, possibly in the bulk to some extent. They contrast therefore with materials, such as supported boron trifluoride [9], which have been used widely as solid Lewis acids. Two of the possible surface species that have been suggested are shown in Figure 1.
The uptake of F by γ-alumina is slow when CHF3 is used. The process is initiated at the surface of oxide particles, subsequently being incorporated into sub-surface regions. The reaction between sulfur tetrafluoride and γ-alumina is essentially a hydrolysis; the course of the reaction has been followed using the radiotracers fluorine-18 and sulfur-35. The hydrolysis involves the replacement of surface hydroxyls by F and the partial replacement of bridging Al-O-Al moieties by Al-F, with the concomitant formation of OSF2 and SO2. These are not strongly adsorbed. The Lewis acid sites have disordered F/O environments rather than being fully fluorinated. When the SF4 treatment is performed under static conditions, nominally at room temperature, the retention of the co-product, HF, appears to be complete. Consequently, γ-alumina fluorinated by this method possesses Brønsted acidity in addition to the expected Lewis acidity.
A most powerful method to probe the Lewis acidity of fluorinated surfaces is FTIR spectroscopy using Lewis base probe molecules such as pyridine (py) or carbon monoxide, and 2-methyl-substituted py as a probe where Brønsted surface hydroxyl groups are present. Studies using CO at low temperatures can be particularly powerful, since different types of surface sites can be detected and compared. The aluminum fluorides mentioned above and various oxofluorides of different structural types have been studied in this way and the results compared with nanoscale aluminum fluorides; accounts of the methodology and some illustrative results are to be found in [10,11].
Useful data can be obtained in addition, using a radiolabeled probe molecule to investigate a fluorinated surface. Experiments using anhydrous hydrogen chloride and its precursors, labeled with chlorine-36, are described below. In addition, a radiotracer can provide important mechanistic details.
Two examples are provided by the behavior of the trichlorotrifluoroethane isomers at β-AlF3 and the fluorinated γ-alumina surfaces. Isomerization of CCl2FCClF2 to the thermodynamically preferred isomer CCl3CF3 does not involve any formation of Al-Cl surface bonds, as indicated by the labeling of CCl2FCClF2 with [36Cl]. This indicates that the isomerization proceeds by an intra- rather than by an inter-molecular mechanism, as shown in Figure 2a. Isomerization is followed by a dismutation reaction giving a mixture of CCl2FCF3 and CCl3CClF2, which, from the lack of the incorporation of [36Cl] into the solid surface, is believed to involve a halogen exchange between two adsorbed CCl3CF3 molecules, as shown in Figure 2b [12].
These observations provide indirect evidence for the presence of strong Lewis acid surface sites on β-AlF3 and fluorinated γ-aluminas, sufficiently strong to allow C–F bonds, unexpectedly, to behave as Lewis bases.

3. Anhydrous Hydrogen Chloride as a Surface Probe

It is not intuitively obvious that anhydrous HCl might be a useful probe molecule which can be used to compare the surface behavior of different fluorinated surfaces, but this has proven to be the case, particularly for comparisons among nanoscale aluminum fluorides and their conventionally prepared counterparts. The genesis of the approach was a fortuitous observation of reactions that occur when 1,1,1-trichloroethane is exposed to an SF4-fluorinated (static conditions, see Table 1) γ-alumina at room temperature. The reactions involved are shown in Figure 3 [9].
The surface-catalyzed reactions shown in Figure 3 are, respectively, a Lewis acid catalyzed dehydrochlorination of CH3CCl3 and an oligomerization, accompanied by a partial dehydrochlorination, of adsorbed CH2=CCl2; both have their equivalent reactions using the archetypal Lewis acid, solid aluminum chloride. The presence of retained HF on the surface probably accounts for the fluorination step. This can be made catalytic if the purple, oligomeric supported layer is exposed to a mixture of HF and CH3CCl3. Although similar behavior is observed when BF3 is used to fluorinate γ-alumina (cf. Figure 1), it occurs to a far smaller extent, reflecting, no doubt, the smaller fluorine content of the surface [9].
Dehydrochlorination occurs also when t-butyl chloride is exposed to a fluorinated Lewis acidic surface at room temperature, although the chemistry is not as dramatic visually as that derived from CH3CCl3. It is a useful reaction, however, since ButCl is readily labeled with [36Cl] and the resulting fate of the H36Cl produced can be monitored from its radioactivity [12].
The experimental procedures for quantifying β activity from the long-lived chlorine-36 isotope by Geiger-Müller counting have been well documented in the literature [10,11,12] and are not described here. Although the behavior observed when [36Cl]-ButCl is exposed to SF4-fluorinated γ-alumina, β-AlF3 and aluminum chlorofluoride (ACF) is similar, there are differences in the details [12]. In particular, the outcomes with SF4-fluorinated γ-alumina appear to be dominated by oligomerization reactions that cover the surface, whereas strongly and weakly surface-bound H36Cl are observed with ACF. More controversially, a significant fraction of H36Cl appears to be retained in the bulk ACF, from which it can be removed by prolonged pumping. The observations using β-AlF3 are similar; in this case, however, HCl interacts with residual H2O located in the HTB hexagonal channels and is desorbed as H2O·HCl. The behaviors of ACF and β-AlF3 towards H36Cl added directly replicate those described above.
The behavior of nanoscale aluminum(III) fluoride, HS-AlF3, towards H36Cl parallels that of ACF; weakly and strongly surface-bound H36Cl and material retained in the bulk are all observed [13]. Dehydrochlorination of ButCl is rapid, and HCl is observed immediately in the vapor; the analogous reaction over β-AlF3 is far slower. Nanoscale magnesium fluoride, HS-MgF2, behaves in an identical way, indicating that unlike rutile MgF2, HS-MgF2 can behave as a Lewis acid. The composite material, 15 mol % FeF3 in HS-MgF2, in its behavior towards [36Cl]-ButCl, appears to utilize Lewis acid sites based on FeIII in addition to those involving MgII [13].
As is evident from the summary given above, the level of detail available about the interactions between HCl and fluorinated surfaces is limited, largely by the corrosive nature of the materials involved. More detailed descriptions become possible for the interactions between anhydrous HCl, which of course has its own corrosion problems, and transitional aluminas such as γ- and η-alumina. FTIR spectroscopy, utilizing temperature-programmed techniques and, more recently, inelastic neutron scattering, has been widely used. Together with a knowledge of the molecular environments of the various surface Lewis acid sites on the aluminas, it is possible to construct realistic hypotheses for events that take place at a chlorinated surface [2,14].
An illustration of what can be achieved under favourable circumstances is the FTIR diffuse reflectance spectra as a function of temperature for a saturated, chemisorbed overlayer of HCl at the surface of η-alumina, as shown in Figure 4 [14].
For a detailed interpretation of the spectra, the reader is directed to Reference [14]; however, in summary, there is good evidence for weakly adsorbed molecular HCl and dissociatively adsorbed species in which HCl has interacted with Al-O species to form Al-Cl and OH groups, or has replaced a surface Al-OH by Al-Cl with the formation of H2O. This type of information is very important in producing realistic sequences of surface events. A very good example is the large-scale selective synthesis of methyl chloride from methanol and HCl, using an η-alumina heterogeneous catalyst [2]. The full sequence of events involving HCl and methanol at the surface is shown schematically in Figure 5 [15].

4. Iron(III) and Copper(II) Chlorides as Chlorination Catalysts: Redox and Lewis Acid Possibilities

From a historical standpoint, iron(III) and copper(II) chlorides are two of the most important metal halides in heterogeneous catalysis, being used in a variety of situations, both in laboratory-based studies and in large-scale processes, where catalysis is an important factor.
Anhydrous iron(III) chloride and its hexa-hydrate feature widely in contemporary organic syntheses, since, because of their Lewis acid properties, they can initiate or catalyze a variety of organic transformations [17,18,19,20,21,22,23,24,25,26]. Because of their redox properties arising from the 1e, FeIII/FeII redox transformation, they are also the basis of a useful polymerization process [27,28,29,30,31], for example the polymerization of thiophene. Anhydrous FeCl3 has been widely reported as a Lewis acid catalyst for chlorination and hydrochlorination of a variety of hydrocarbons [32,33,34], particularly in large-scale process reactions such as the chlorination of ethene and the hydrochlorination of vinyl chloride [32,33]. Reaction pathways suggested for these types of reactions were traditionally based on polar intermediates but these are probably unrealistic based on current thinking.
Reactions carried out in mild steel that involve Cl2 may be problematic, since surface chlorination can occur, leading to a chloride surface that behaves like solid FeCl3. As a result, the unwanted chlorination of hydrochlorocarbons can occur; this may be accompanied by oligomerization of unsaturated species [35,36,37]. This type of behavior in the 1,1,2,2-tetrachloroethane/Cl2 system was observed in a recent laboratory study of reactions carried out in stainless steel or Pyrex [36,37]. A rather different example of over-chlorination occurred in large-scale reactions of Cl2 with CH2=CH2 and led to unwanted 1,1,2-trichloroethane in the product, 1,2-dichloroethane. The over-chlorination is believed to be catalyzed by the presence of small quantities of FeCl3 formed adventitiously, either in solution or at the reactor wall. Considerable effort has been made to inhibit over-chlorination in large-scale processes; for example, by means of FeCl3 removal by complexation with various Lewis bases. However, fundamental studies targeted to the identification of molecular species present have apparently never been reported, and in the reports of inhibition by chloride ion–forming tetrachloroferrate(III) anions, differences exist as to the optimum stoichiometry to be used. Because anhydrous FeCl3 is extremely hygroscopic, its surface may be complexed by trace water, present even in purified CH2ClCH2Cl. Molecular species such as FeCl3(OH2), so formed, are soluble in a hydrochlorocarbon solvent and the resulting solid/solution distribution of chloroferrate(III) species is not simple [38].
The ability of copper(II) chloride, either as a component of a melt or supported on an oxide, to chlorinate a wide variety of saturated and unsaturated hydrocarbons has been known for many years [39,40]; the behavior of pumice-supported CuCl2 towards olefins suggests that CuCl2 is the chlorinating agent rather than adsorbed Cl* [40]. From a heterogeneous catalytic standpoint, one of the most important uses of CuCl2 is as the catalyst for the Deacon reaction, a process for the conversion of HCl to Cl2. There is renewed interest in this sequence of reactions, shown in Equations (1)–(4) below, in the context of the utilization of waste HCl. A scheme for the integration of processes producing and using HCl has the aim of “closing the chlorine cycle”; it includes the steps: chlorination, dehydrochlorination and oxychlorination [41].
2CuIICl2 → 2CuICl + Cl2
2CuICl + 1/2O2 → CuII2OCl2
CuII2OCl2 + 2HCl → 2CuIICl2 + H2O
Overall
2HCl + 1/2O2 → H2O + Cl2
Since a Deacon, or oxychlorination, catalyst typically comprises, in addition to CuCl2, a support, generally a high-surface-area oxide, and a promoter, such as an ionic halide, for example KCl, Equations (1)–(3) are likely to be an approximation of the surface species present. Furthermore, in a process environment, at the typical operating temperatures used, the species may well be liquid. One of the most detailed studies made to date under laboratory conditions involved CuCl2 supported on γ-alumina. The catalyst was used in the oxychlorination of ethene; the product, CH2ClCH2Cl, underwent dehydrochlorination to CHCl=CH2, so the study was very relevant to PVC production [42,43,44,45,46,47,48,49]. Physical methods, particularly surface science–based together with reactor-based experiments, indicate that at least three CuII-containing phases are present and that further mixed CuII/Group I phases can be formed as well.
Chlorination or oxychlorination of CH2ClCH2Cl in the presence of a CuCl2/KCl catalyst supported on the clay mineral, attapulgite, leads to a complex mixture of C2 chlorohydrocarbons and chlorinated olefins; the products depend on the exact conditions used but careful control can result in CHCl=Cl2 or CCl2=CCl2 being the predominant product. The reaction scheme deduced is shown in Figure 6 [36,37].
There is, inevitably, a trade-off to be made between high conversions and the loss of materials through oligomerization, particularly of CHCl=CCl2. Product distributions are the result of competition between chlorination, facilitated by the CuCl2 catalyst, and dehydrochlorination. The latter for CHCl2CHCl2 is initiated by Cl*, whereas dehydrochlorination of CHCl2CCl3 occurs at Lewis sites on the catalyst.
The CuII catalyst supports the Deacon reaction to convert the co-product, HCl, to the reactant Cl2; however, the reaction is slow compared to the main organo-chlorine transformations.
Phosgene is prepared on a large scale usually by the reaction between carbon monoxide and Cl2 over a carbon catalyst. An oxychlorination route has been developed, however, on a laboratory scale using a catalyst comprising CuCl2/KCl supported on silica gel [50,51,52]. The challenge in this case is to prepare OCCl2 continuously with good conversion and without hydrolysis. This has been achieved by employing a three-stage arrangement corresponding to Equations (1)–(4), viz. Equations (5)–(8).
2CuCl2 + CO → 2CuCl + COCl2
2CuCl + 1/2O2 → Cu2OCl2
Cu2OCl2 + 2HCl → 2CuCl2 + H2O
Overall:
CO + 1/2O2 + 2HCl → COCl2 + H2O

5. Conclusions

The two parts of this short review have different objectives. In the first part, we sought to show that selected metal fluorides prepared in conventional and unconventional ways have many similarities in their Lewis acidity, using a variety of probe molecules, including the very weak Lewis base anhydrous HCl. In the second part, some aspects of FeCl3 and, particularly, CuCl2, chemistry demonstrated the complexity in their catalytic behavior. The challenge for those who are synthesizing nanoscale halides is to investigate whether there is a parallel chemistry to be uncovered.

Acknowledgments

The authors’ studies have benefitted enormously from talented students and post-doctoral fellows over many years. Particularly in the second part, we have been very fortunate to collaborate with colleagues based in Runcorn, UK. Our thanks go to all.

Conflicts of Interest

The authors declare no conflict of interest.

References and Notes

  1. Tressaud, A. Functionalised Inorganic Fluorides; Wiley: Chichester, UK, 2010. [Google Scholar]
  2. Lennon, D.; Winfield, J.M. Heterogeneous Catalysts used for Large-Scale Syntheses of Selected Catalytic Applications; Parvulescu, V.I., Kemnitz, E., Eds.; Elsevier: Amsterdam, The Netherlands; Oxford, UK; Boston, MA, USA, 2016; Chapter 7; pp. 193–217. [Google Scholar]
  3. Le Bail, A.; Jacaboni, C.; Leblanc, M.; De Pape, R.; Duroy, H.; Fourquet, J.L. Crystal Structure of the Metastable form of Aluminium Trifluoride βAlF3 and the Gallium and Indium Homologues. J. Solid State Chem. 1988, 77, 96–101. [Google Scholar]
  4. Hess, A.; Kemnitz, E. Characterisation of Catalytically Active Sites on Aluminium Oxides, Hydroxyfluorides and Fluorides in Correlation with their Catalytic Behaviour. J. Catal. 1994, 149, 449–457, and references therein. [Google Scholar]
  5. Kemnitz, E.; Winfield, J.M. Fluoride Catalysts: Their Application to Heterogeneous Catalytic Fluorination and Related Processes. In Advanced Inorganic Fluorides: Synthesis, Characterization and Applications; Nakajima, T., Žemva, B., Tressaud, A., Eds.; Elsevier Science S.A.: Lausanne, Switzerland, 2000; Chapter 12; pp. 367–401. [Google Scholar]
  6. Krespan, C.G.; Petrov, V.A. The Chemistry of Highly Fluorinated Carbocations. Chem. Rev. 1996, 96, 3269–3301, and references therein. [Google Scholar]
  7. Krahl, T.; Kemnitz, E. The Very Strong Lewis Acids Aluminium Chlorofluoride (ACF) and Bromofluoride (ABF)—Synthesis, Structure and Lewis Acidity. J. Fluor. Chem. 2006, 127, 663–678, and references therein. [Google Scholar]
  8. Kemnitz, E.; Skapin, T.; Winfield, J.M. Preparation of Fluorinated γ-Alumina. In Efficient Preparations of Fluorine Compounds, 1st ed.; Roesky, H.W., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2013; Chapter 67; pp. 442–446. [Google Scholar]
  9. Klapötke, T.M.; McMonagle, F.; Spence, R.R.; Winfield, J.M. γ-Alumina-Supported Boron Trifluoride: Catalysis, Radiotracer Studies and Computations. J. Fluor. Chem. 2006, 127, 1446–1453. [Google Scholar]
  10. Vimont, A.; Daturi, M.; Winfield, J.M. Investigation of Surface Acidity using a Range of Probe Molecules. In Functionalised Inorganic Fluorides; Tressaud, A., Ed.; Wiley: Chichester, UK, 2010; Chapter 4; pp. 101–139. [Google Scholar]
  11. Winfield, J.M. Investigating Acidity of Metal Fluoride Surfaces by Spectroscopic and Chemical Methods. J. Fluor. Chem. 2009, 130, 1069–1079. [Google Scholar]
  12. Nickkho-Amiry, M.; Winfield, J.M. Investigation of Fluorinated Surfaces by means of Radio-Labelled Probe Molecules. J. Fluor. Chem. 2007, 128, 344–352. [Google Scholar]
  13. Nickkho-Amiry, M.; Eltanany, G.; Wüttke, S.; Rüdiger, S.; Kemnitz, E.; Winfield, J.M. A Comparative Study of Surface Activity in the Amorphous, High Surface Area Solids, Aluminium Fluoride, Magnesium fluoride and Magnesium Fluoride containing Iron(III) or Aluminium(III) Fluoride. J. Fluor. Chem. 2008, 129, 366–375. [Google Scholar]
  14. McInroy, A.R.; Lundie, D.T.; Winfield, J.M.; Dudman, C.C.; Jones, P.; Parker, S.F.; Lennon, D. The Interaction of Alumina with HCl: An Infrared Spectroscopy, Temperature-Programmed Desorption and Inelastic Neutron Scattering Study. Catal. Today 2006, 114, 403–411. [Google Scholar]
  15. McInroy, A.L.; Winfield, J.M.; Dudman, C.C.; Jones, P.; Lennon, D. The Development of a New Generation of Methyl Chloride Synthesis Catalysts. Faraday Discuss. 2016, 188, 467–499. [Google Scholar]
  16. Lundie, D.T.; McInroy, A.R.; Marshall, R.; Winfield, J.M.; Jones, P.; Dudman, C.C.; Parker, S.F.; Mitchell, C.; Lennon, D. Improved Description of the Surface Acidity of η-Alumina. J. Phys. Chem. B 2005, 109, 11592–11601. [Google Scholar]
  17. Zhang, X.; Fan, X.; Niu, H.; Wang, J. An Ionic Liquid As a Recyclable Medium for the Green Preparation of α,α′-Bis(substituted benzylidine)cycloalkanones Catalysed by FeCl3·6H2O. Green Chem. 2003, 5, 267–269. [Google Scholar]
  18. Kabalka, G.W.; Yao, M.-L.; Borella, S.; Goins, L.K. Iron Trichloride Mediated Allylation of Lithium Alkoxides through an Unusual Carbon-Oxygen Bond Cleavage. Organometallics 2007, 26, 4112–4114. [Google Scholar]
  19. Xa, W.-W.; Lin, M.; Huang, Y.; Chen, L.; Zhan, Z.-P. Iron Trichloride Catalysed One-Pot Synthesis of Substituted Oxazoles from Propargylic Acetates and Amides under Microwave Irradiation. Lett. Org. Chem. 2009, 6, 664–665. [Google Scholar]
  20. Cantagrei, G.; De Canné-Carnavalet, B.; Meyer, C.; Cossy, J. IronTrichloride Promoted Cyclisation of O-Alkynylaryl Isocyanates. Synthesis of 3-(Chloromethylene)Oxindoles. Org. Lett. 2009, 11, 8584–8586. [Google Scholar]
  21. Schröder, K.; Enthaler, S.; Bitterlich, B.; Schulz, T.; Spannenberg, A.; Tse, M.K.; Junge, K.; Beller, M. Design of and Mechanistic Studies on a Biomimetic Iron-Imidazole Catalyst System for Epoxidation of Olefins with Hydrogen Peroxide. Chem. Eur. J. 2009, 15, 5471–5481. [Google Scholar]
  22. Buchwald, S.L.; Bolm, C. On the Role of Metal Contaminants in Catalyses with FeCl3. Angew. Chem. Int. Ed. 2009, 48, 5586–5587. [Google Scholar]
  23. Schröder, K.; Junge, K.; Spannenberg, A.; Beller, M. Design of a Bio-Inspired Imidazole-Based Iron Catalyst for Epoxidation of Olefins: Mechanistic Insights. Catal. Today 2010, 157, 364–370. [Google Scholar]
  24. Su, L.; Lei, C.-Y.; Fan, W.-Y.; Liu, L.-X. FeCl3-Mediated Reaction of Alkynols with Iodine: An Efficient and Convenient Route to Vinyl Iodides. Synth. Commun. 2011, 41, 1200–1207. [Google Scholar]
  25. Blevins, D.W.; Yao, M.-L.; Yong, L. Iron Trichloride Promoted Hydrolysis of Potassium Organotrifluoroborates. Tetrahedron Lett. 2011, 52, 6534–6536. [Google Scholar]
  26. Yeh, M.-C.P.; Fang, C.-W.; Lin, H.-H. Facile Synthesis of Azaspirocycles via Iron Trichloride Promoted Cyclisation/Chlorination of Cyclic 8-Aryl-5-aza-5-tosyl-2-en-7-yn-1-ols. Org. Lett. 2012, 14, 1830–1833. [Google Scholar]
  27. Della Casa, C.; Bertinelli, F.; Costa Bizzarri, P.; Salatelli, E. Properties of a Hydroxydecyl-Functionalized Polythiophene Synthesized by the Iron Trichloride Route. Adv. Mater. 1995, 7, 1005–1007. [Google Scholar]
  28. Tanaka, K.; Mihara, T.; Koide, N. Synthesis and Physical Properties of Regioregular Poly(3-Alkoxy-4-Methylthiophene)s. Polym. J. 2004, 36, 628–663. [Google Scholar]
  29. Svoboda, J.; Bláha, M.; Sedláček, J.; Vohlídal, H.; Balcar, H.; Mav-Golež, I.; Žigon, M. New Approaches to the Synthesis of Pure Conjugated Polymers. Acta Chim. Slov. 2006, 53, 407–416. [Google Scholar]
  30. Cloutet, E.; Mumtaz, M.; Cramail, H. Synthesis of PEDOT Latexes by Dispersion Polymerization in Aqueous Media. Mater. Sci. Eng. C 2009, 29, 377–382. [Google Scholar]
  31. Lanzi, M.; Paganin, L.; Errani, F. Synthesis, Characterization and Photovoltaic Properties of a New Thiophene-Based Double-Cable Polymer with Pendent Fullerene Group. Polymer 2012, 53, 2134–2145. [Google Scholar]
  32. Reed, D.J. Chlorocarbons and Chlorohydrocarbons. In Kirk-Othmer Encyclopaedia of Chemical Technology, 4th ed.; Kroschwitz, J.I., Howe-Grant, M., Eds.; John Wiley & Sons: New York, NY, USA, 1993; Volume 5, pp. 1017–1028. [Google Scholar]
  33. Weissermel, K.; Arpe, H.-J. Industrial Organic Chemistry, 4th ed.; Wiley-VCH: Weinheim, Germany, 2003; pp. 449–450. [Google Scholar]
  34. Delaude, L.; Lazlo, P. Aromatic Chlorination of Toluene and of Anisole Using Clay-Supported FeCl3 and m-Chloroperbenzoic Acid. A Biomimetic Approach. Catal. Lett. 1990, 5, 35–44. [Google Scholar]
  35. Kawaguchi, T.; Suzuki, Y.; Saito, R. Trichloroethylene and Perchloroethylene Coproduction by Ethylene Process. Ind. Eng. Chem. 1970, 62, 36–41. [Google Scholar]
  36. Sutherland, I.W.; Hamilton, N.G.; Dudman, C.C.; Jones, P.; Lennon, D.; Winfield, J.M. Chlorination and Dehydrochlorination Reactions Relevant to the Manufacture of Trichloroethene and Tetrachloroethene. Part 1. Reaction Pathways. Appl. Catal. A Gen. 2011, 399, 1–11. [Google Scholar]
  37. Sutherland, I.W.; Hamilton, N.G.; Dudman, C.C.; Jones, P.; Lennon, D.; Winfield, J.M. Chlorination Reactions Relevant to the Manufacture of Trichloroethene and Tetrachloroethene. Part 2. Effects of Chlorine Supply. Appl. Catal. A Gen. 2014, 471, 149–156. [Google Scholar]
  38. Taylor, D.S.C. Catalysis in the System 1,2-Dichloroethane, Iron(III) Chloride. Ph.D. Thesis, University of Glasgow, Glasgow, UK, 2002. [Google Scholar]
  39. Fontana, C.M.; Gorin, E.; Kidder, G.A.; Meredith, C.S. Ternary System, Cuprous Chloride-Cupric Chloride-Potassium Chloride and its Equilibrium Chlorine Pressures. Ind. Eng. Chem. 1952, 44, 363–368. [Google Scholar]
  40. Arganbright, R.P.; Yates, W.F. Chlorinations with Cupric Chloride. J. Org. Chem. 1962, 27, 1205–1208. [Google Scholar]
  41. Cavani, F. Catalytic Selective Oxidation: The Forefront in the Challenge for a more Sustainable Chemical Industry. Catal. Today 2010, 157, 8–15. [Google Scholar]
  42. Leofanti, G.; Padovan, M.; Garilli, M.; Carmello, D.; Zecchina, A.; Spoto, G.; Bordiga, S.; Turnes Palomino, G.; Lamberti, C. Alumina-Supported Copper Chloride 1. Characterization of freshly Prepared Catalyst. J. Catal. 2000, 189, 91–104. [Google Scholar]
  43. Leofanti, G.; Padovan, M.; Garilli, M.; Carmello, D.; Marra, G.L.; Zecchina, A.; Spoto, G.; Bordiga, S.; Lamberti, C. Alumina-Supported Copper Chloride 2. Effect of Aging and Thermal Treatments. J. Catal. 2000, 189, 105–116. [Google Scholar]
  44. Leofanti, G.; Marsella, A.; Cremaschi, B.; Garilli, M.; Zecchina, A.; Spoto, G.; Bordiga, S.; Fisicaro, P.; Berlier, G.; Prestipino, C.; et al. Alumina-Supported Copper Chloride 3. Effect of Exposure to Ethylene. J. Catal. 2001, 202, 279–295. [Google Scholar]
  45. Leofanti, G.; Marsella, A.; Cremaschi, B.; Garilli, M.; Zecchina, A.; Spoto, G.; Bordiga, S.; Fisicaro, P.; Prestipino, C.; Villain, F.; et al. Alumina-Supported Copper Chloride 4. Effect of Exposure to O2 and HCl. J. Catal. 2002, 205, 375–381. [Google Scholar]
  46. Lamberti, C.; Prestipino, C.; Bonino, F.; Capello, L.; Bordiga, S.; Spoto, G.; Zecchina, A.; Diaz Moreno, S.; Cremaschi, B.; Garilli, M.; et al. The Chemistry of the Oxychlorination Catalyst: An In Situ, Time-Resolved XANES Study. Angew. Chem. Int. Ed. 2002, 41, 2341–2344. [Google Scholar]
  47. Prestipino, C.; Bordiga, S.; Lamberti, C.; Vidotto, S.; Garilli, M.; Cremaschi, B.; Marsella, A.; Leofanti, G.; Fisicaro, P.; Spoto, G.; et al. Structural Determination of Copper Species on the Alumina-Supported Copper Chloride Catalyst: A Detailed EXAFS Study. J. Phys. Chem. B 2003, 107, 5022–5030. [Google Scholar]
  48. Muddada, N.B.; Olsbye, U.; Caccialupi, L.; Cavani, F.; Leofanti, G.; Gianolio, D.; Bordiga, S.; Lamberti, C. Influence of Additives in Defining the Active Phase of the Ethylene Oxychlorination Catalyst. Phys. Chem. Chem. Phys. 2010, 12, 5605–5618. [Google Scholar]
  49. Muddada, N.B.; Olsbye, U.; Leofanti, G.; Gianolio, D.; Bonino, F.; Bordiga, S.; Fuglerud, T.; Vidotto, S.; Marsella, A.; Lamberti, C. Quantification of Copper Phases, their Reducibility and Dispersion in Doped-CuCl2/Al2O3 Catalysts for Ethylene Oxychlorination. Dalton Trans. 2010, 39, 8437–8449. [Google Scholar]
  50. Zhang, T.; Troll, C.; Rieger, B.; Kintrup, J.; Schlüter, O.F.-K.; Weber, R. Reaction Kinetics of Oxychlorination of Carbon Monoxide to Phosgene based on Copper(II) Chloride. Appl. Catal. A Gen. 2009, 357, 51–57. [Google Scholar]
  51. Zhang, T.; Troll, C.; Rieger, B.; Kintrup, J.; Schlüter, O.F.-K.; Weber, R. Composition Optimisation of Silica-Supported Copper(II) Chloride Substance for Phosgene Production. Appl. Catal. A Gen. 2009, 365, 20–27. [Google Scholar]
  52. Zhang, T.; Troll, C.; Rieger, B.; Kintrup, J.; Schlüter, O.F.-K.; Weber, R. Oxychlorination of CO to Phosgene in a Three-Step Reaction Cycle and Corresponding Catalytic Mechanism. J. Catal. 2010, 270, 76–85. [Google Scholar]
Figure 1. Possible surface species derived from BF3 and oxides. Redrawn from Reference [9].
Figure 1. Possible surface species derived from BF3 and oxides. Redrawn from Reference [9].
Molecules 22 00201 g001
Figure 2. Schematic representation of possible adsorbed states for (a) CCl2FCClF2 prior to isomerization to CCl3CF3; and (b) CCl3CF3 prior to its dismutation. Redrawn from Reference [12].
Figure 2. Schematic representation of possible adsorbed states for (a) CCl2FCClF2 prior to isomerization to CCl3CF3; and (b) CCl3CF3 prior to its dismutation. Redrawn from Reference [12].
Molecules 22 00201 g002
Figure 3. Behavior of CH3CCl3 on SF4-fluorinated γ-alumina. Redrawn from Reference [9].
Figure 3. Behavior of CH3CCl3 on SF4-fluorinated γ-alumina. Redrawn from Reference [9].
Molecules 22 00201 g003
Figure 4. Diffuse reflectance spectra as a function of temperature of a saturated chemisorbed overlayer of HCl: (a) 4000–1200 cm−1 and (b) the free hydroxyl region; (i) Saturation spectrum dosed at 293 K. The sample was then progressively warmed to (ii) 423; (iii) 523; and (iv) 623 K. A flow of He was continually passed over the sample while heating to progressively higher temperatures. The sample was held at the desorption temperature for 15 min and then allowed to cool to room temperature, where the spectrum was acquired. All spectra are background subtracted, where the spectrum of the clean activated catalyst has been subtracted from the measured spectrum. Reproduced with permission from Reference [14].
Figure 4. Diffuse reflectance spectra as a function of temperature of a saturated chemisorbed overlayer of HCl: (a) 4000–1200 cm−1 and (b) the free hydroxyl region; (i) Saturation spectrum dosed at 293 K. The sample was then progressively warmed to (ii) 423; (iii) 523; and (iv) 623 K. A flow of He was continually passed over the sample while heating to progressively higher temperatures. The sample was held at the desorption temperature for 15 min and then allowed to cool to room temperature, where the spectrum was acquired. All spectra are background subtracted, where the spectrum of the clean activated catalyst has been subtracted from the measured spectrum. Reproduced with permission from Reference [14].
Molecules 22 00201 g004
Figure 5. A schematic representation of the site-selective formation of methyl chloride over η-alumina. The red circles represent strong/medium-strong Lewis acid sites; the blue circles represent medium-weak Lewis acid sites; the red/blue shaded circles represent strong/medium-strong/medium-weak Lewis acid sites. The definition and form of these sites are described in D. T. Lundie et al. J. Phys. Chem. B 2005, 109, 11592–11601 [16]. Reproduced with permission from Reference [15].
Figure 5. A schematic representation of the site-selective formation of methyl chloride over η-alumina. The red circles represent strong/medium-strong Lewis acid sites; the blue circles represent medium-weak Lewis acid sites; the red/blue shaded circles represent strong/medium-strong/medium-weak Lewis acid sites. The definition and form of these sites are described in D. T. Lundie et al. J. Phys. Chem. B 2005, 109, 11592–11601 [16]. Reproduced with permission from Reference [15].
Molecules 22 00201 g005
Figure 6. The dehydrochlorination (1, 4) and chlorination (2, 3) processes that connect CHCl2CHCl2 with CHCl=CCl2 and CCl2=CCl2. Possible oligomer formation (5, 6). Reproduced with permission from Reference [36].
Figure 6. The dehydrochlorination (1, 4) and chlorination (2, 3) processes that connect CHCl2CHCl2 with CHCl=CCl2 and CCl2=CCl2. Possible oligomer formation (5, 6). Reproduced with permission from Reference [36].
Molecules 22 00201 g006
Table 1. Some properties of fluorinated γ-alumina [8].
Table 1. Some properties of fluorinated γ-alumina [8].
Fluorination Conditions aBET Area (m2·g−1)Fluorine Content (%)Other Properties
SF4 static conditions, nominally room temperature, 2 h, procedure repeated ×280–90 bca. 22 cSome Brønsted sites also
SF4 20% in N2, flow conditions, 523 K, 2 h6747.1γ-alumina present; possibly an amorphous phase also
CHF3 20% in N2, then 100%, flow conditions, 623 K, total time 5 h34 d58.4α- and β-AlF3 present; possibly an amorphous phase
a In all cases samples were calcined before use; b Before fluorination 110 m2·g−1; c Calculated from a radiotracer study; d Before fluorination 240 m2·g−1.

Share and Cite

MDPI and ACS Style

Lennon, D.; Winfield, J.M. Metal Fluorides, Metal Chlorides and Halogenated Metal Oxides as Lewis Acidic Heterogeneous Catalysts. Providing Some Context for Nanostructured Metal Fluorides. Molecules 2017, 22, 201. https://doi.org/10.3390/molecules22020201

AMA Style

Lennon D, Winfield JM. Metal Fluorides, Metal Chlorides and Halogenated Metal Oxides as Lewis Acidic Heterogeneous Catalysts. Providing Some Context for Nanostructured Metal Fluorides. Molecules. 2017; 22(2):201. https://doi.org/10.3390/molecules22020201

Chicago/Turabian Style

Lennon, David, and John M. Winfield. 2017. "Metal Fluorides, Metal Chlorides and Halogenated Metal Oxides as Lewis Acidic Heterogeneous Catalysts. Providing Some Context for Nanostructured Metal Fluorides" Molecules 22, no. 2: 201. https://doi.org/10.3390/molecules22020201

Article Metrics

Back to TopTop