Abstract

Artificial transmembrane ion transporters are attractive in the context of drug discovery since, in principle, they may be used as novel therapeutics by promoting apoptosis (programmed cell death), interfering with autophagic processes, or inducing an antibiotic response via disrupting cellular ion homeostasis or dissipating intracellular pH gradients. Considerable effort has been devoted into the design of artificial transporters with selective ion recognition and transport. However, only limited progress has been made in terms of translating fundamental research involving transport mechanisms into systems that function in vitro. In this review, we summarize recent progress in the area of artificial ion transporters with a focus on their application as potential therapeutics. The associated discussion of accomplishments and remaining challenges is designed to hasten the day when research on artificial transmembrane ion transporters is translated into bona fide clinical benefits. Various artificial transmembrane transporters, designed to function through mobile carrier or channel mechanisms, have been developed in the past decade. With the aid of structural manipulation and by employing either discrete chemical entities or self-assembled nanostructures, progress has been made in achieving the selective recognition and transmembrane transport of key ions. The ability to perturb intracellular pH or disrupt intracellular ion homeostasis makes transmembrane ion transporters of interest as potential therapeutics that might see use as cancer treatments or as antibacterial agents. In this review, recent progress in the area of artificial transmembrane ion transporter research is summarized with an emphasis on applications involving anticancer research and antibiotic applications. The examples chosen for highlights are meant to be illustrative of key themes involving synthetic ion transport rather than comprehensive. Nevertheless, it is anticipated that this review will provide a useful entry point for the general reader and set the stage for further progress in the area. Various artificial transmembrane transporters, designed to function through mobile carrier or channel mechanisms, have been developed in the past decade. With the aid of structural manipulation and by employing either discrete chemical entities or self-assembled nanostructures, progress has been made in achieving the selective recognition and transmembrane transport of key ions. The ability to perturb intracellular pH or disrupt intracellular ion homeostasis makes transmembrane ion transporters of interest as potential therapeutics that might see use as cancer treatments or as antibacterial agents. In this review, recent progress in the area of artificial transmembrane ion transporter research is summarized with an emphasis on applications involving anticancer research and antibiotic applications. The examples chosen for highlights are meant to be illustrative of key themes involving synthetic ion transport rather than comprehensive. Nevertheless, it is anticipated that this review will provide a useful entry point for the general reader and set the stage for further progress in the area. Cellular membranes serve as barriers for the exchange of most substrates into and out of cells. However, the exchange of solutes between cells and their extracellular milieu is crucial for maintaining the normal operation of biological systems.1Szostak J.W. Bartel D.P. Luisi P.L. Synthesizing life.Nature. 2001; 409: 387-390Google Scholar, 2Dubyak G.R. Ion homeostasis, channels, and transporters: an update on cellular mechanisms.Adv. Physiol. Educ. 2004; 28: 143-154Google Scholar, 3Bröer S. Amino acid transport across mammalian intestinal and renal epithelia.Physiol. Rev. 2008; 88: 249-286Google Scholar Small neutral compounds, such as water and carbon dioxide, can cross the phospholipid membranes relatively easily by diffusion. On the other hand, the transmembrane exchange of charged species, including most physiologically important anions and cations, relies on the assistance of proteins that function as either ion carriers or transmembrane ion channels.4Gamper N. Shapiro M.S. Regulation of ion transport proteins by membrane phosphoinositides.Nat. Rev. Neurosci. 2007; 8: 921-934Google Scholar, 5Wang X. Zhang X. Dong X.P. Samie M. Li X. Cheng X. Goschka A. Shen D. Zhou Y. Harlow J. et al.TPC proteins are phosphoinositide-activated sodium-selective ion channels in endosomes and lysosomes.Cell. 2012; 151: 372-383Google Scholar, 6Guo J. Zeng W. Chen Q. Lee C. Chen L. Yang Y. Cang C. Ren D. Jiang Y. Structure of the voltage-gated two-pore channel TPC1 from Arabidopsis thaliana.Nature. 2016; 531: 196-201Google Scholar In healthy cells, cooperative feedback between various ion channels serves to maintain precisely the ion gradients across biological membranes and, as a result, desired intracellular ion concentrations crucial for normal biological processes.7Sheppard D.N. Rich D.P. Ostedgaard L.S. Gregory R.J. Smith A.E. Welsh M.J. Mutations in CFTR associated with mild-disease-form Cl- channels with altered pore properties.Nature. 1993; 362: 160-164Google Scholar, 8Choi J.Y. Muallem D. Kiselyov K. Lee M.G. Thomas P.J. Muallem S. Aberrant CFTR-dependent HCO3− transport in mutations associated with cystic fibrosis.Nature. 2001; 410: 94-97Google Scholar, 9Vaughan-Jones R.D. Spitzer K.W. Swietach P. Intracellular pH regulation in heart.J. Mol. Cell. Cardiol. 2009; 46: 318-331Google Scholar Failure to maintain ion homeostasis is typically manifest in terms of severe pathological consequences. For instance, dysfunction of Cl− channels is associated with cystic fibrosis, a disease correlated with higher-than-normal intracellular chloride anion concentrations.10Gadsby D.C. Vergani P. Csanády L. The ABC protein turned chloride channel whose failure causes cystic fibrosis.Nature. 2006; 440: 477-483Google Scholar In principle, cystic fibrosis and other disorders traced to disruptions in ion transport could be controlled by treating patients with functioning chloride anion channels. However, the limited availability, poor stability, and structural complexity of natural ion channels have so far precluded their use in these and other practical applications. Artificial transmembrane transporters, simpler synthetic molecular systems that mimic the salient properties of natural ion channels, could provide a means to overcome these limitations (so-called channel replacement therapy).11Ghadiri M.R. Granja J.R. Buehler L.K. Artificial transmembrane ion channels from self-assembling peptide nanotubes.Nature. 1994; 369: 301-304Google Scholar, 12Sakai N. Matile S. Synthetic ion channels.Langmuir. 2013; 29: 9031-9040Google Scholar, 13Gokel G.W. Negin S. Synthetic ion channels: from pores to biological applications.Acc. Chem. Res. 2013; 46: 2824-2833Google Scholar, 14Fyles T.M. Synthetic ion channels in bilayer membranes.Chem. Soc. Rev. 2007; 36: 335-347Google Scholar, 15Barboiu M. Gilles A. From natural to bioassisted and biomimetic artificial water channel systems.Acc. Chem. Res. 2013; 46: 2814-2823Google Scholar, 16Montenegro J. Ghadiri M.R. Granja J.R. Ion channel models based on self-assembling cyclic peptide nanotubes.Acc. Chem. Res. 2013; 46: 2955-2965Google Scholar Appropriate molecular design might allow selective recognition and adaptable ion transport to be achieved for a range of ionic targets, including such physiologically important species as chloride, potassium, and sodium. In this context, both discrete molecular systems (so-called unimers) and self-assembled nanostructures have been the subject of study.17Carmichael V.E. Dutton P.J. Fyles T.M. James T.D. Swan J.A. Zojaji M. Biomimetic ion transport: a functional model of a unimolecular ion channel.J. Am. Chem. Soc. 1989; 111: 767-769Google Scholar, 18Baumeister B. Sakai N. Matile S. Giant artificial ion channels formed by self-assembled, cationic rigid-rod β-barrels.Angew. Chem. Int. Ed. 2000; 39: 1955-1958Google Scholar, 19Gokel G.W. Murillo O. Synthetic organic chemical models for transmembrane channels.Acc. Chem. Res. 1996; 29: 425-432Google Scholar, 20Li X. Wu Y.D. Yang D. α-Aminoxy acids: new possibilities from foldamers to anion receptors and channels.Acc. Chem. Res. 2008; 41: 1428-1438Google Scholar, 21Fyles T.M. How do amphiphiles form ion-conducting channels in membranes? Lessons from linear oligoesters.Acc. Chem. Res. 2013; 46: 2847-2855Google Scholar, 22Sisson A.L. Shah M.R. Bhosale S. Matile S. Synthetic ion channels and pores (2004-2005).Chem. Soc. Rev. 2006; 35: 1269-1286Google Scholar In addition to targeting ion transport-related disorders, artificial ion carriers and channels (collectively referred to as transporters) show promise for inducing apoptosis in cancer cells by perturbing the intracellular pH or disrupting cellular ion homeostasis.23Davis J.T. Gale P.A. Quesada R. Advances in anion transport and supramolecular medicinal chemistry.Chem. Soc. Rev. 2020; 49: 6056-6086Google Scholar In this review we summarize progress made in the construction of synthetic ion transporters (channels and carriers) with a focus on their potential use as anticancer agents. A limited treatment of antibiotic opportunities afforded by synthetic ion transporters is also included. Over the past decades, enormous progress has been made in terms of understanding cancer. Nevertheless, cancer remains the second largest cause of death worldwide. Currently, systemic treatments, including anticancer drugs (chemotherapy, hormone, and biological therapies), are the modalities of choice in many instances. However, the indiscriminate destruction of normal cells, the toxicity of conventional chemotherapeutic drugs, as well as the development of multi-drug resistance, provide an incentive to seek new treatments based on new mechanisms of action. Artificial transmembrane ion transporters, which can induce apoptosis (so-called programmed cell death) in cancer cells by changing the intracellular pH or disrupting intracellular ion homeostasis, could provide one such alternative approach.24Hosogi S. Kusuzaki K. Inui T. Wang X. Marunaka Y. Cytosolic chloride ion is a key factor in lysosomal acidification and function of autophagy in human gastric cancer cell.J. Cell. Mol. Med. 2014; 18: 1124-1133Google Scholar, 25Busschaert N. Park S.H. Baek K.H. Choi Y.P. Park J. Howe E.N.W. Hiscock J.R. Karagiannidis L.E. Marques I. Félix V. et al.A synthetic ion transporter that disrupts autophagy and induces apoptosis by perturbing cellular chloride concentrations.Nat. Chem. 2017; 9: 667-675Google Scholar, 26Smith B.A. Daschbach M.M. Gammon S.T. Xiao S. Chapman S.E. Hudson C. et al.In vivo cell death mediated by synthetic ion channels.Chem. Commun. 2011; 47: 7977-7979Google Scholar, 27Park S.H. Park S.H. Howe E.N.W. Hyun J.Y. Chen L.J. Hwang I. Vargas-Zuñiga G. Busschaert N. Gale P.A. Sessler J.L. Shin I. Determinants of ion-transporter cancer cell death.Chem. 2019; 5: 2079-2098Google Scholar, 28Rodilla A.M. Korrodi-Gregório L. Hernando E. Manuel-Manresa P. Quesada R. Pérez-Tomás R. Soto-Cerrato V. Synthetic tambjamine analogues induce mitochondrial swelling and lysosomal dysfunction leading to autophagy blockade and necrotic cell death in lung cancer.Biochem. Pharmacol. 2017; 126: 23-33Google Scholar, 29Kim D.H. Rozhkova E.A. Ulasov I.V. Bader S.D. Rajh T. Lesniak M.S. Novosad V. Biofunctionalized magnetic-vortex microdiscs for targeted cancer-cell destruction.Nat. Mater. 2010; 9: 165-171Google Scholar, 30Alfonso I. Quesada R. Biological activity of synthetic ionophores: ion transporters as prospective drugs?.Chem. Sci. 2013; 4: 3009-3019Google Scholar Moreover, it is possible that by combining ion transporters with other therapeutic modalities, drug resistance could be effectively overcome.31Li H. Valkenier H. Thorne A.G. Dias C.M. Cooper J.A. Kieffer M. Busschaert N. Gale P.A. Sheppard D.N. Davis A.P. Anion carriers as potential treatments for cystic fibrosis: transport in cystic fibrosis cells, and additivity to channel-targeting drugs.Chem. Sci. 2019; 10: 9663-9672Google Scholar It is also important to appreciate that the therapeutic benefits associated with perturbing ion homeostasis are not limited to cancer. Bacteria are also attractive targets. As the result of their ability to induce an ion imbalance and to disrupt bacterial membranes, several artificial transmembrane ion transporters have shown promise as antibacterial motifs in recent years.32Leevy W.M. Donato G.M. Ferdani R. Goldman W.E. Schlesinger P.H. Gokel G.W. Synthetic hydraphile channels of appropriate length kill Escherichia coli.J. Am. Chem. Soc. 2002; 124: 9022-9023Google Scholar,33Brea R.J. Reiriz C. Granja J.R. Towards functional bionanomaterials based on self-assembling cyclic peptide nanotubes.Chem. Soc. Rev. 2010; 39: 1448-1456Google Scholar For instance, polyether ionophores, complex natural products that can transport cations across biological membranes, often display antimicrobial activity, a finding that has inspired the design and synthesis of artificial ion transporters as antibacterial agents.34Kevin II, D.A. Meujo D.A. Hamann M.T. Polyether ionophores: broad-spectrum and promising biologically active molecules for the control of drug-resistant bacteria and parasites.Expert Opin. Drug Discov. 2009; 4: 109-146Google Scholar Since a variety of ion transporters have been reported to date, the discussion will be classified according to the type of ions involved, with particular emphasis being given to systems that show promise as transporters of chloride, potassium, and sodium. This focus reflects not only the physiological importance of these ions but also the fact that they have been the target of most artificial ion transport development work reported to date. Additionally, the examples chosen in this review represent a selection of anion transporters reported in the literature that have been studied most extensively in biological systems. To provide the readers with an introductory guideline to this field, a brief summary of synthetic ion transporters is given in this section. The term ion carrier applies to a transporter that shuttles across the bilayer and, in doing so, brings about the translocation of ions (Figure 1). In contrast, ion channels are scaffolds that can mediate ion transport across membrane bilayers without undergoing appreciable motion and without disturbing the lipid bilayer supra-structures. Synthetic ion channels include both unimolecular channels and self-assembled channels. Unimolecular synthetic ion channels are typically macrocyclic molecules that are long enough to insert into lipid bilayer membranes and serve as guides for transmembrane ion transport. An advantage of unimolecular channels is their stability, reflecting the fact that they comprise single molecules that remain intact inside lipid bilayers. However, the synthesis of unimolecular channels is often difficult and time consuming. Therefore, the majority of synthetic ion channels reported to date have been created in situ via the controlled self-assembly of monomers within the lipid bilayers. However, constructing channels with stable structures and selective binding sites via the rational self-assembly of smaller building blocks remains a challenge. To differentiate an ion channel from a carrier-type ion transporter, black lipid membrane (BLM) experiments in planar bilayers are typically performed (Figure 2A).35Matile S. Sakai N. The characterization of synthetic ion channels and pores.in: Schalley C.A. Analytical Methods in Supramolecular Chemistry. Second Edition. Wiley-VCH, Weinheim, Germany2012: 711-742Google Scholar Dynamic “open-closed” conductance is generally taken as evidence of ion channel formation (Figure 2B). Moreover, some key parameters of the ion channels could be inferred from these studies, such as the conductance g, the lifetime τ1, and the open probability Po. The construction of so-called Hill plots is another tool used to distinguish carrier- from channel-based transport mechanisms.36Litvinchuk S. Bollot G. Mareda J. Som A. Ronan D. Shah M.R. Perrottet P. Sakai N. Matile S. Thermodynamic and kinetic stability of synthetic multifunctional rigid-rod β-barrel pores: evidence for supramolecular catalysis.J. Am. Chem. Soc. 2004; 126: 10067-10075Google Scholar The slope of a Hill plot is equal to the Hill coefficient for the interactions between the agent in question and the membrane under study. A Hill plot with a slope greater than 1.0 is taken as an indication that there is a synergistic concentration dependence (channel mechanism), whereas a slope less than 1.0 indicates competitive concentration dependence (carrier mechanism) (Figure 3A). These limiting regimes are generally interpreted in terms of channel and carrier mechanisms, respectively. However, it is to be appreciated that Hill analyses of synthetic ion channels are not fully developed, and appropriate caution must be exercised in terms of drawing concrete mechanistic conclusions.Figure 3(A) Typical Hill plot for synthetic ion transporters. Reprinted with permission from Matile and Sakai,35Matile S. Sakai N. The characterization of synthetic ion channels and pores.in: Schalley C.A. Analytical Methods in Supramolecular Chemistry. Second Edition. Wiley-VCH, Weinheim, Germany2012: 711-742Google Scholar copyright 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.Show full caption(B) Standard configuration of the so-called HPTS (8-hydroxypyrene-1,3,6-trisulfonic acid) assay. Reprinted with permission from Wu et al.,37Wu X. Howe E.N.W. Gale P.A. Supramolecular transmembrane anion transport: new assays and insights.Acc. Chem. Res. 2018; 51: 1870-1879Google Scholar copyright 2018 American Chemical Society.View Large Image Figure ViewerDownload Hi-res image Download (PPT) (B) Standard configuration of the so-called HPTS (8-hydroxypyrene-1,3,6-trisulfonic acid) assay. Reprinted with permission from Wu et al.,37Wu X. Howe E.N.W. Gale P.A. Supramolecular transmembrane anion transport: new assays and insights.Acc. Chem. Res. 2018; 51: 1870-1879Google Scholar copyright 2018 American Chemical Society. A synthetic ion transporter involved in mediating the transport of only a single ion or molecule across a lipid membrane is termed as uniporter, and the process is referred to as uniport. Transporters that move two molecules or ions in the same direction across the membrane are called symporters. If two molecules are moved in opposite directions across the bilayer, the process is called antiport. Transporters involved in the movement of specific ions are called ionophores. If the action of a transporter in moving ions across a membrane results in a net change in charge, the process is described as electrogenic; if there is no net change in charge, the transport is described as electroneutral. The primary method used to characterize putative ion transporters under cell-free conditions is the liposomal assay.37Wu X. Howe E.N.W. Gale P.A. Supramolecular transmembrane anion transport: new assays and insights.Acc. Chem. Res. 2018; 51: 1870-1879Google Scholar In general, fluorescence spectroscopy is used to monitor fluorophore-containing large unilamellar vesicles (LUVs). Depending on the design of the setup and the fluorescent probes employed, such analyses can provide insights into the influx or efflux of various types of ions. Often, LUVs are prepared from 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) and then encapsulated with either the pH-sensitive ratiometric fluorescent dye 8-hydroxypyrene-1,3,6-trisulfonic acid (HPTS) or the halide-sensitive fluorophore lucigenin (Figure 3B). The fluorescence intensities of HPTS and lucigenin are calibrated to the pH and chloride concentrations, respectively. This allows the optical-based monitoring of H+ or Cl− transport processes. Other methods for monitoring changes in ion concentrations, such as NMR spectroscopy and ion-selective electrodes, have also been used in liposomal transport assays. To test whether a synthetic compound can act as an ion transporter in cells, its ability to induce cytosolic ion concentration changes needs to be evaluated. Taking a potential chloride transporter as an example, changes in Cl− concentration in the intracellular matrix are typically monitored using a chloride-selective fluorescent probe, such as N-(ethoxycarbonylmethyl)-6-methoxyquinolinium bromide (MQAE), which undergoes fluorescence quenching in the presence of chloride ions. Therefore, upon preincubation of cells with an effective transporter or synthetic channel, quenching of MQAE fluorescence would be expected due to a mediated increase in the intracellular Cl− concentration.38Verkman A.S. Sellers M.C. Chao A.C. Leung T. Ketcham R. Synthesis and characterization of improved chloride-sensitive fluorescent indicators for biological applications.Anal. Biochem. 1989; 178: 355-361Google Scholar A related strategy involves the use of Fischer rat thyroid epithelial (FRT) cells that express a mutant yellow fluorescent protein (YFP) whose fluorescence is sensitively quenched by chloride ions.39Jayaraman S. Haggie P. Wachter R.M. Remington S.J. Verkman A.S. Mechanism and cellular applications of a green fluorescent protein-based halide sensor.J. Biol. Chem. 2000; 275: 6047-6050Google Scholar Based on the level of fluorescence quenching, the extent of Cl− transport activity in cells can be estimated. FRT cells expressing a mutant YFP have the advantage that changes in the cellular chloride ion concentration induced by low levels of endogenous Cl− channels/carriers can be detected. Thus, engineered cells provide useful tools for determining whether a putative synthetic transporter promotes chloride anion influx. An increase in cytosolic chloride concentrations can induce cell death. Thus, the effect of a synthetic chloride anion transporter on cell viability can be assessed using a standard MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) assay.40Van Meerloo J. Kaspers G.J. Cloos J. Cell sensitivity assays: the MTT assay.Methods Mol. Biol. 2011; 731: 237-245Google Scholar Apoptosis in cells can be determined by treating cells with a mixture of fluorescein-labeled annexin V and propidium iodide (PI).41Williams D.R. Ko S.K. Park S. Lee M.R. Shin I. An apoptosis-inducing small molecule that binds to heat shock protein 70.Angew. Chem. Int. Ed. 2008; 47: 7466-7469Google Scholar Transporters with apoptosis-inducing activity will give rise to positive annexin V binding and PI uptake. Because the loss of mitochondrial membrane potential is a hallmark of apoptosis, this event can be determined using a membrane-potential-sensitive probe JC-1. Cells undergoing apoptosis display an increase in the JC-1 green fluorescence and a decrease in red fluorescence.42Salvioli S. Ardizzoni A. Franceschi C. Cossarizza A. JC-1, but not DIOC6(3) or rhodamine 123, is a reliable fluorescent probe to assess ΔΨ changes in intact cells: implications for studies on mitochondrial functionality during apoptosis.FEBS Lett. 1997; 411: 77-82Google Scholar The observation of DNA fragmentation in cells provides additional evidence for apoptosis.43Hengartner M.O. The biochemistry of apoptosis.Nature. 2000; 407: 770-776Google Scholar DNA fragmentation can be monitored using DNA electrophoresis, nucleus image analysis after staining with a DNA-binding fluorescent dye or a terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) assay. Taken in concert, these systematic studies can provide robust support for the conclusion that a putative synthetic Cl− transporter induces apoptosis. Selective and regulated transmembrane ion exchange across lipid bilayers mediated by ion transporters plays a crucial role in various biological processes. Transmembrane ion channels are essential for cell proliferation and appear to play a role in the development of cancer. This was initially noted in the case of potassium channels but has been suggested for other cation channels and chloride channels.44Kunzelmann K. Ion channels and cancer.J. Membr. Biol. 2005; 205: 159-173Google Scholar Thus, over the past decades, there has been considerable effort devoted to creating artificial ion transporters that can help understand and mimic the functions of natural ion channels. In fact, a number of synthetic ion transporters, designed to promote the through-membrane transport of specific ions, such as chloride, sodium, and potassium, have been developed to date.45Benke B.P. Aich P. Kim Y. Kim K.L. Rohman M.R. Hong S. Hwang I.C. Lee E.H. Roh J.H. Kim K. Iodide-selective synthetic ion channels based on shape-persistent organic cages.J. Am. Chem. Soc. 2017; 139: 7432-7435Google Scholar, 46Lang C. Li W. Dong Z. Zhang X. Yang F. Yang B. et al.Biomimetic transmembrane channels with high stability and transporting efficiency from helically folded macromolecules.Angew. Chem. Int. Ed. 2016; 55: 9723-9727Google Scholar, 47Burns J.R. Seifert A. Fertig N. Howorka S. A biomimetic DNA-based channel for the ligand-controlled transport of charged molecular cargo across a biological membrane.Nat. Nanotechnol. 2016; 11: 152-156Google Scholar, 48Haynes C.J.E. Zhu J. Chimerel C. Hernández-Ainsa S. Riddell I.A. Ronson T.K. et al.Blockable Zn10L15 ion channels through subcomponent self-assembly.Angew. Chem. Int. Ed. 2017; 56: 15388-15392Google Scholar Self-assembled nanostructures created from molecular subunits through noncovalent interactions, such as hydrogen bonding, π-π donor-acceptor interactions, metal coordination, etc., have also been widely created as ion transporters.49Talukdar P. Bollot G. Mareda J. Sakai N. Matile S. Synthetic ion channels with rigid-rod π-stack architecture that open in response to charge-transfer complex formation.J. Am. Chem. Soc. 2005; 127: 6528-6529Google Scholar, 50Danial M. Tran C.M.N. Jolliffe K.A. Perrier S. Thermal gating in lipid membranes using thermoresponsive cyclic peptide–polymer conjugates.J. Am. Chem. Soc. 2014; 136: 8018-8026Google Scholar, 51Jung M. Kim H. Baek K. Kim K. Synthetic ion channel based on metal–organic polyhedra.Angew. Chem. Int. Ed. 2008; 47: 5755-5757Google Scholar In most cases, classic ion-binding motifs, including crown ethers and so-called cation- or anion-π slides, have been incorporated into synthetic ion transporters.52Pfeifer J.R. Reiss P. Koert U. Crown ether–gramicidin hybrid ion channels: dehydration-assisted ion selectivity.Angew. Chem. Int. Ed. 2006; 45: 501-504Google Scholar, 53Kumpf R.A. Dougherty D.A. A mechanism for ion selectivity in potassium channels: computational studies of cation-pi interactions.Science. 1993; 261: 1708-1710Google Scholar, 54Vargas Jentzsch A. Hennig A. Mareda J. Matile S. Synthetic ion transporters that work with anion−π interactions, halogen bonds, and anion–macrodipole interactions.Acc. Chem. Res. 2013; 46: 2791-2800Google Scholar, 55Behera H. Madhavan N. Anion-selective cholesterol decorated macrocyclic transmembrane ion carriers.J. Am. Chem. Soc. 2017; 139: 12919-12922Google Scholar, 56Yuan L. Shen J. Ye R. Chen F. Zeng H. Structurally simple trimesic amides as highly selective anion channels.Chem. Commun. 2019; 55: 4797-4800Google Scholar Unimolecular transmembrane channels and carriers have also been exploited based on a variety of acyclic and macrocyclic hosts, such as squaramides, pillararenes, calixarenes, cucurbiturils, and calixpyrroles.57Busschaert N. Kirby I.L. Young S. Coles S.J. Horton P.N. Light M.E. et al.Squaramides as potent transmembrane anion transporters.Angew. Chem. Int. Ed. 2012; 51: 4426-4430Google Scholar, 58Si W. Xin P. Li Z.T. Hou J.L. Tubular unimolecular transmembrane channels: construction strategy and transport activities.Acc. Chem. Res. 2015; 48: 1612-1619Google Scholar, 59Kaucher M.S. Harrell W.A. Davis J.T. A unimolecular G-quadruplex that functions as a synthetic transmembrane Na+ transporter.J. Am. Chem. Soc. 2006; 128: 38-39Google Scholar, 60Su G. Zhang M. Si W. Li Z.T. Hou J.L. Directional potassium transport through a unimolecular peptide channel.Angew. Chem. Int. Ed. 2016; 55: 14678-14682Google Scholar, 61Jeon Y.J. Kim H. Jon S. Selvapalam N. Oh D.H. Seo I. et al.Artificial ion channel formed by cucurbit[n]uril derivatives with a carbonyl group fringed portal reminiscent of the selectivity filter of K+ channels.J. Am. Chem. Soc. 2004; 126: 15944-15945Google Scholar, 62Kim D.S. Sessler J.L. Calix[4]pyrroles: versatile molecular containers with ion transport, recognition, and molecular switching functions.Chem. Soc. Rev. 2015; 44: 532-546Google Scholar Typical intracellular and extracellular chloride concentrations are 5–15 mM and 110 mM, respectively.63Sonawane N.D. Thiagarajah J.R. Verkman A.S. Chloride concentration in endosomes measured using a ratioable fluorescent Cl− indicator: evidence for chloride accumulation during acidification.J. Biol. Chem. 2002; 277: 5506-5513Google Scholar Naturally occurring chloride (Cl−) ion transporters (e.g., channels) that can mediate the transmembrane transport of Cl− have been reported to play significant roles in important biological processes, including maintaining ion homeostasis, regulating intracellular pH; they are also important in the context of controlling electrical excitability. Artificial chloride transporters that circumvent the associated exchange equilibrium can disrupt critical pH gradients, perturb ion homeostasis, and induce apoptosis of cells.64Saha T. Hossain M.S. Saha D. Lahiri M. Talukdar P. Chloride-mediated apoptosis-inducing activity of bis (sulfonamide) anionophores.J. Am. Chem. Soc. 2016; 138: 7558-7567Google Scholar In the case of cancer, t

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call