首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 109 毫秒
1.
Root-to-shoot translocation and shoot homeostasis of potassium (K) determine nutrient balance, growth, and stress tolerance of vascular plants. To maintain the cation-anion balance, xylem loading of K+ in the roots relies on the concomitant loading of counteranions, like nitrate (NO3). However, the coregulation of these loading steps is unclear. Here, we show that the bidirectional, low-affinity Nitrate Transporter1 (NRT1)/Peptide Transporter (PTR) family member NPF7.3/NRT1.5 is important for the NO3-dependent K+ translocation in Arabidopsis (Arabidopsis thaliana). Lack of NPF7.3/NRT1.5 resulted in K deficiency in shoots under low NO3 nutrition, whereas the root elemental composition was unchanged. Gene expression data corroborated K deficiency in the nrt1.5-5 shoot, whereas the root responded with a differential expression of genes involved in cation-anion balance. A grafting experiment confirmed that the presence of NPF7.3/NRT1.5 in the root is a prerequisite for proper root-to-shoot translocation of K+ under low NO3 supply. Because the depolarization-activated Stelar K+ Outward Rectifier (SKOR) has previously been described as a major contributor for root-to-shoot translocation of K+ in Arabidopsis, we addressed the hypothesis that NPF7.3/NRT1.5-mediated NO3 translocation might affect xylem loading and root-to-shoot K+ translocation through SKOR. Indeed, growth of nrt1.5-5 and skor-2 single and double mutants under different K/NO3 regimes revealed that both proteins contribute to K+ translocation from root to shoot. SKOR activity dominates under high NO3 and low K+ supply, whereas NPF7.3/NRT1.5 is required under low NO3 availability. This study unravels nutritional conditions as a critical factor for the joint activity of SKOR and NPF7.3/NRT1.5 for shoot K homeostasis.The macronutrient potassium (K) is essential for plant growth and development because of its crucial roles in various cellular processes (i.e. regulation of enzyme activities), stabilization of protein synthesis, and neutralization of negative charges. In addition, it is a major component of the cation-anion balance and osmoregulation in plants, thereby influencing cellular turgor, xylem and phloem transport, pH homeostasis, and the setting of membrane potentials (Maathuis, 2009; Marschner, 2012; Sharma et al., 2013). K+ uptake and distribution in Arabidopsis (Arabidopsis thaliana) are accomplished by a total of 71 membrane proteins that have been assigned to five gene families: the Shaker and Tandem-Pore K+ channels (now also including the inward-rectifier K-like (Kir-like) channels), the K+ uptake permeases (KUP/HAK/KT), the K+ transporter (HKT) family, and the cation proton antiporters (CPA; Gierth and Mäser, 2007; Gomez-Porras et al., 2012; Sharma et al., 2013).Root xylem loading is a key step for the delivery of nutrients to the shoot (Poirier et al., 1991; Engels and Marschner, 1992a; Gaymard et al., 1998; Takano et al., 2002; Park et al., 2008). Root-to-shoot translocation of K+ is mediated by the voltage-dependent Shaker family K+ channel Stelar K+ Outward Rectifier (SKOR). The gene is primarily expressed in pericycle and root xylem parenchyma cells, and it is down-regulated upon K shortage and in response to treatments with the phytohormones abscisic acid, cytokinin, and auxin. Such gene expression changes are thought to control K+ secretion into the xylem sap and K+ reallocation through the phloem to adjust root K+ transport activity to K+ availability and shoot demand (Pilot et al., 2003). SKOR is activated upon membrane depolarization, and it is in a closed state when the driving force for K+ is inwardly directed. It elicits outward K+ currents, facilitating the release of the cation from the cells into the xylem. The voltage dependency of the channel is modulated by the external K+ concentration to minimize the risk of an undesired K+ influx under high K+ availability (Johansson et al., 2006). Root-to-shoot K+ transfer was strongly reduced in the knockout mutant skor-1, resulting in a decreased shoot K content, whereas the root K content remained unaffected (Gaymard et al., 1998).Root xylem loading is subject to the maintenance of a cation-anion balance, and nitrate (NO3) is the quantitatively most important anion counterbalancing xylem loading of K+ (Engels and Marschner, 1993). Members of the Nitrate Transporter1 (NRT1)/Peptide Transporter (PTR) transporter family (NPF) play a prominent role in NO3 uptake and allocation in Arabidopsis (summarized in Krouk et al., 2010; Wang et al., 2012; and Léran et al., 2014). Two of them have recently been reported to control xylem NO3 loading and unloading. The low-affinity, pH-dependent bidirectional NO3 transporter NPF7.3/NRT1.5 (subsequently termed NRT1.5) mediates NO3 efflux from pericycle cells to the xylem vessels, whereas the low-affinity influx protein NPF7.2/NRT1.8 removes NO3 from the xylem sap and transfers it into xylem parenchyma cells (Lin et al., 2008; Li et al., 2010; Chen et al., 2012). Accordingly, the expression of both genes is oppositely regulated under various stress conditions (Li et al., 2010). In nrt1.5 mutants, NRT1.8 expression is increased, which is thought to enhance NO3 reallocation to the root (Chen et al., 2012).The NRT1.5 gene is mainly expressed in root pericycle cells close to the xylem, and the protein localizes to the plasma membrane. In nrt1.5 mutants, less NO3 is transported from the root to the shoot, and the NO3 concentration in the xylem sap is reduced. However, root-to-shoot NO3 transport is not completely abolished in these mutants, indicating the existence of additional xylem-loading activities for NO3 (Lin et al., 2008; Wang et al., 2012). The recent observation that NPF6.3/NRT1.1/CHL1 and NPF6.2/NRT1.4 are also capable of mediating bidirectional NO3 transport in Xenopus laevis oocytes might indicate that more NPF family members are contributing to xylem loading with NO3 (Léran et al., 2013).Electrophysiological studies with NRT1.5-expressing X. laevis oocytes revealed that NO3 excited an inward current at pH 5.5, which would be expected for a proton-coupled nitrate transporter with a proton to nitrate ratio larger than one (Lin et al., 2008). The inward currents elicited by exposure to nitrate were pH dependent, and Lin et al. (2008) observed that NRT1.5 can also facilitate nitrate efflux when the oocytes were incubated at pH 7.4. Lin et al. (2008) concluded that NRT1.5 can transport nitrate in both directions, presumably through a proton-coupled mechanism. Interestingly, a K+ gradient was not sufficient to drive NRT1.5-mediated NO3 export. However, the determination of root and shoot cation concentrations in the nrt1.5-1 mutant revealed that the amount of K+ translocated to the shoot was reduced when NO3 but not NH4+ was supplied as the N source. Therefore, Lin et al. (2008) suggested a regulatory loop between NO3 and K+ at the xylem loading step.A close relationship between these two nutrients concerning uptake, translocation, recycling, and reduction (of NO3) has been described in physiological studies since the 1960s (e.g. Ben Zioni et al., 1971; Blevins et al., 1978; Barneix and Breteler, 1985), but only recently, common components in the NO3 and K+ uptake pathways were identified and led to the first ideas of how such a cross talk might be coordinated on the molecular level. The uptake activity of the K+ channel AKT1 as well as the affinity of the NO3 transporter NPF6.3/NRT1.1/CHL1 are both modulated by the activity of CALCINEURIN B-LIKE PROTEIN-INTERACTING PROTEIN KINASE23 (CIPK23), which itself is regulated by CALCINEURIN B-LIKE PROTEIN9 (CBL9) under both deficiencies (Xu et al., 2006; Ho et al., 2009). Yet, the details of this interaction in root K+ uptake, the (regulation of) xylem loading with K+ and NO3, and the involvement of SKOR and NRT1.5 in this process are unknown.In this study, we approached this problem by investigating the molecular and physiological responses of Arabidopsis wild-type (Columbia-0 [Col-0]), nrt1.5, and skor transfer DNA (T-DNA) insertion lines to varying NO3 and K+ regimes. The nrt1.5 mutant developed an early senescence phenotype under low NO3 nutrition, which could be attributed to a reduced K+ translocation to the shoot. The assessment of nrt1.5 and skor single- and double-knockout lines disclosed an interplay of the two proteins in the NO3-dependent control of shoot K homeostasis. The presented data indicate that SKOR mediates K+ root-to-shoot translocation under high NO3 and low K+ availability, whereas NRT1.5 is important for K+ translocation under low NO3 availability, irrespective of the K+ supply.  相似文献   

2.
3.
Salinity is a growing issue worldwide, with nearly 30% of arable land predicted to be lost due to soil salinity in the next 30 years. Many grass crops that are vital to sustain the world’s caloric intake are salt sensitive. Studying mechanisms of salt tolerance in halophytic grasses, plants that thrive in salt conditions, may be an effective approach to ultimately improve salt-sensitive grass crops. Seashore paspalum (Paspalum vaginatum) is a halophytic Panicoid grass able to grow in salt concentrations near that of seawater. Despite its widespread cultivation as a sustainable turfgrass, the mechanism underlying its ability to retain high Na+ concentrations in photosynthetic tissue while maintaining growth remains unknown. We examined the leaf structure and ion content in P. vaginatum ‘HI10’, which shows increased growth under saline conditions, and Paspalum distichum ‘Spence’, which shows reduced growth under salt, to better understand the superior salt tolerance of cv HI10. A striking difference between cv HI10 and cv Spence was the high steady-state level of K+ in cv HI10. Imaging further showed that the adaxial surface of both cv HI10 and cv Spence contained dense costal ridges of papillae. However, these unicellular extensions of the epidermis were significantly larger in cv HI10 than in cv Spence. The cv HI10 papillae were shown to act as Na+ sinks when plants were grown under saline conditions. We provide evidence that leaf papillae function as specialized structures for Na+ sequestration in P. vaginatum, illustrating a possible path for biotechnological improvement of salt-sensitive Panicoid crops with analogous leaf structures.

About 20% of irrigated land is considered saline, with the amount of saline soils increasing worldwide (Mayak et al., 2004). This is due to increased irrigation in agricultural fields necessitated by more frequent droughts due to climate change. This trend is alarming due to the high salt sensitivity of most crop species that we rely on for vital resources. Yield reduction in crops in saline soils amounts to losses on the order of 12 to 27.3 billion U.S. dollars annually (Qadir et al., 2014). Thus, the improvement of salt tolerance in plants will become key in the coming decades. Breeding salt-tolerant crops is a cost-effective approach to improve growth in saline soils. Although much work has focused on breeding salt-tolerant species, progress in this area has been slow due to the complex genetic and physiological nature of the salt response. Furthermore, most research has been conducted on glycophytic model systems that are salt sensitive (Munns and Gilliham, 2015). Unraveling the salt-tolerance mechanisms in halophytes, species that can complete their life cycle in 200 mm salt concentrations, and transferring these pathways into glycophytes is therefore of great interest (Rajalakshmi and Parida, 2012; Roy and Chakraborty, 2014).Both glycophytes and halophytes have evolved a multitude of salt-tolerance mechanisms, including sodium (Na+) exclusion, sequestration, and secretion; osmolyte production; ion homeostasis; and reactive oxygen species (ROS) detoxification (Meng et al., 2018). Often, mechanisms present in glycophytes, such as osmolyte production and Na+ exclusion, are utilized in halophytes at higher efficiency (Wyn Jones and Storey, 1981; Grieve and Maas, 1984). However, halophytes also use mechanisms that are absent in glycophytes. Salt sequestration and secretion via salt glands is a halophyte-specific mechanism of coping with salt (Flowers and Colmer, 2008). Salt glands are found in over 50 species in 14 angiosperm families with four subtypes: epidermal bladder cells, complex multicellular glands, bicellular glands, and unicellular glands (Dassanayake and Larkin, 2017). The Poales order contains ∼8% of all halophytes (Flowers et al., 2010) and has therefore been the focus of much salt-gland-focused work (Ceccoli et al., 2015). As salt tolerance has independently evolved >70 times in grass lineages (Bennett et al., 2013), studying these salt sequestering/secreting structures in grasses is an excellent approach to better understand salt tolerance mechanisms in halophytes.Most structural and physiological work on salt glands in grasses has been conducted in the Chloridoideae and Oryzoideae subfamilies. Grasses carry either unicellular or bicellular glands, often referred to as glandular trichomes or microhairs, on the leaf surface (Dassanayake and Larkin, 2017). Microhairs have been observed on the leaf surface in all grass subfamilies except the Pooideae, and have evolved diverse functions including the sequestration or secretion of substances such as callose and heavy metals (Burke et al., 2000; Ceccoli et al., 2015). Unicellular structures on the adaxial leaf side able to secrete salt are only found in the Oryzoideae wild rice species Porteresia coarctata (Flowers et al., 1990; Sengupta and Majumder, 2009). Salt glands in the Chloridoideae are bicellular, consisting of a cap cell and a lower basal cell, both of which are dense in cytoplasm and mitochondria (Ceccoli et al., 2015). The cuticle is thickened above the cap cell in some species, forming a cuticular chamber used for storing secreted salts (Amarasinghe and Watson, 1988). In the Panicoideae, a few cases of Na+ secretion have been reported (McWhorter et al., 1995; Ramadan and Flowers, 2004), but to date, no sequestration structures have been identified.The Panicoideae subfamily includes the agronomically important food crops maize (Zea mays) and sorghum (Sorghum bicolor) in addition to the biofuel grasses miscanthus (Miscanthus sinensis), switchgrass (Panicum virgatum), and sugarcane (Saccharum officinarum). One of the most salt-tolerant species in the Panicoideae is the halophyte seashore paspalum (Paspalum vaginatum). It is cultivated as a turfgrass worldwide and derives its popularity from its ability to be irrigated with brackish water. P. vaginatum can survive in salt concentrations near that of seawater (Uddin et al., 2012) and uses osmolyte production, ion homeostasis, and Na+ exclusion to cope with salt stress (Peacock and Dudeck, 1985; Lee et al., 2008; Guo et al., 2016). However, its ability to maintain growth while accumulating high levels of Na+ in leaf tissue remains perplexing.Here, we studied the leaf structure and Na+ sequestration in ‘HI10’, a P. vaginatum cultivar, and ‘Spence’, a Paspalum distichum cultivar. P. vaginatum and P. distichum are closely related (and possibly the same species; Eudy et al., 2017), and constitute group “Disticha” in the tribe Paspaleae. P. distichum is less salt tolerant than P. vaginatum and is typically found in freshwater habitats (Leithead et al., 1971). P. vaginatum and P. distichum therefore represent a useful species pair to study salt tolerance. Furthermore, their salt responses can be compared with those of sorghum, a Panicoid glycophyte. Our main research objective was to identify the phenotypic and physiological factors that contribute to the differential tolerance to salt stress of the two Paspalum spp. cultivars and sorghum ‘BTx623’. We show that both Paspalum species contain dense rows of translucent papillae on the adaxial surface. The papillae are unicellular protrusions from epidermal cells and are much larger in cv HI10 than in cv Spence. We further demonstrate that the papillae sequester Na+ under salt stress. This study thus provides evidence of Na+ sequestration in specialized leaf-borne organs within the Panicoideae.  相似文献   

4.
E. coli is one of the most common species of bacteria colonizing humans and animals. The singularity of E. coli’s genus and species underestimates its multifaceted nature, which is represented by different strains, each with different combinations of distinct virulence factors. In fact, several E. coli pathotypes, or hybrid strains, may be associated with both subclinical infection and a range of clinical conditions, including enteric, urinary, and systemic infections. E. coli may also express DNA-damaging toxins that could impact cancer development. This review summarizes the different E. coli pathotypes in the context of their history, hosts, clinical signs, epidemiology, and control. The pathotypic characterization of E. coli in the context of disease in different animals, including humans, provides comparative and One Health perspectives that will guide future clinical and research investigations of E. coli infections.

Escherichia coli (E. coli) is the most common bacterial model used in research and biotechnology. It is an important cause of morbidity and mortality in humans and animals worldwide, and animal hosts can be involved in the epidemiology of infections.240,367,373,452,727 The adaptive and versatile nature of E. coli argues that ongoing studies should receive a high priority in the context of One Health involving humans, animals, and the environment.240,315,343,727 Two of the 3 E. coli pathogens associated with death in children with moderate-to-severe diarrhea in Asia and Africa are classified into 2 E. coli pathogenic groups (also known as pathotypes or pathovars): enterotoxigenic E. coli (ETEC) and enteropathogenic E. coli (EPEC).367 In global epidemiologic studies, ETEC and EPEC rank among the deadliest causes of foodborne diarrheal illness and are important pathogens for increasing disability adjusted life years.355,382,570 Furthermore, in humans, E. coli is one of the top-ten organisms involved in coinfections, which generally have deleterious effects on health.270ETEC is also an important etiologic agent of diarrhea in the agricultural setting.183 E. coli-associated extraintestinal infections, some of which may be antibiotic-resistant, have a tremendous impact on human and animal health. These infections have a major economic impact on the poultry, swine, and dairy industries.70,151,168,681,694,781,797 The pervasive nature of E. coli, and its capacity to induce disease have driven global research efforts to understand, prevent, and treat these devastating diseases. Animal models for the study of E. coli infections have been useful for pathogenesis elucidation and development of intervention strategies; these include zebrafish, rats, mice, Syrian hamsters, guinea pigs, rabbits, pigs, and nonhuman primates.27,72,101,232,238,347,476,489,493,566,693,713,744,754 Experiments involving human volunteers have also been important for the study of infectious doses associated with E. coli-induced disease and of the role of virulence determinants in disease causation.129,176,365,400,497,702,703 E. coli strains (or their lipopolysaccharide) have also been used for experimental induction of sepsis in animals; the strains used for these studies, considered EPEC, are not typically involved in systemic disease.140,205,216,274,575,782This article provides an overview of selected topics related to E. coli, a common aerobic/facultative anaerobic gastrointestinal organism of humans and animals.14,277,432,477,716 In addition, we briefly review: history, definition, pathogenesis, prototype (archetype or reference) strains, and features of the epidemiology and control of specific pathotypes. Furthermore, we describe cases attributed to different E. coli pathotypes in a range of animal hosts. The review of scientific and historical events regarding the discovery and characterization of the different E. coli pathotypes will increase clinical awareness of E. coli, which is too often regarded merely as a commensal organism, as a possible primary or co- etiologic agent during clinical investigations. As Will and Ariel Durant write in The Lessons of History: “The present is the past rolled up for action, and the past is the present unrolled for understanding”.  相似文献   

5.
6.
7.
PSI is an essential component of the photosynthetic apparatus of oxygenic photosynthesis. While most of its subunits are conserved, recent data have shown that the arrangement of the light-harvesting complexes I (LHCIs) differs substantially in different organisms. Here we studied the PSI-LHCI supercomplex of Botryococccus braunii, a colonial green alga with potential for lipid and sugar production, using functional analysis and single-particle electron microscopy of the isolated PSI-LHCI supercomplexes complemented by time-resolved fluorescence spectroscopy in vivo. We established that the largest purified PSI-LHCI supercomplex contains 10 LHCIs (∼240 chlorophylls). However, electron microscopy showed heterogeneity in the particles and a total of 13 unique binding sites for the LHCIs around the PSI core. Time-resolved fluorescence spectroscopy indicated that the PSI antenna size in vivo is even larger than that of the purified complex. Based on the comparison of the known PSI structures, we propose that PSI in B. braunii can bind LHCIs at all known positions surrounding the core. This organization maximizes the antenna size while maintaining fast excitation energy transfer, and thus high trapping efficiency, within the complex.

The multisubunit-pigment-protein complex PSI is an essential component of the electron transport chain in oxygenic photosynthetic organisms. It utilizes solar energy in the form of visible light to transfer electrons from plastocyanin to ferredoxin.PSI consists of a core complex composed of 12 to 14 proteins, which contains the reaction center (RC) and ∼100 chlorophylls (Chls), and a peripheral antenna system, which enlarges the absorption cross section of the core and differs in different organisms (Mazor et al., 2017; Iwai et al., 2018; Pi et al., 2018; Suga et al., 2019; for reviews, see Croce and van Amerongen, 2020; Suga and Shen, 2020). For the antenna system, cyanobacteria use water-soluble phycobilisomes; green algae, mosses, and plants use membrane-embedded light-harvesting complexes (LHCs); and red algae contain both phycobilisomes and LHCs (Busch and Hippler, 2011). In the core complex, PsaA and PsaB, the subunits that bind the RC Chls, are highly conserved, while the small subunits PsaK, PsaL, PsaM, PsaN, and PsaF have undergone substantial changes in their amino acid sequences during the evolution from cyanobacteria to vascular plants (Grotjohann and Fromme, 2013). The appearance of the core subunits PsaH and PsaG and the change of the PSI supramolecular organization from trimer/tetramer to monomer are associated with the evolution of LHCs in green algae and land plants (Busch and Hippler, 2011; Watanabe et al., 2014).A characteristic of the PSI complexes conserved through evolution is the presence of “red” forms, i.e. Chls that are lower in energy than the RC (Croce and van Amerongen, 2013). These forms extend the spectral range of PSI beyond that of PSII and contribute significantly to light harvesting in a dense canopy or algae mat, which is enriched in far-red light (Rivadossi et al., 1999). The red forms slow down the energy migration to the RC by introducing uphill transfer steps, but they have little effect on the PSI quantum efficiency, which remains ∼1 (Gobets et al., 2001; Jennings et al., 2003; Engelmann et al., 2006; Wientjes et al., 2011). In addition to their role in light-harvesting, the red forms were suggested to be important for photoprotection (Carbonera et al., 2005).Two types of LHCs can act as PSI antennae in green algae, mosses, and plants: (1) PSI-specific (e.g. LHCI; Croce et al., 2002; Mozzo et al., 2010), Lhcb9 in Physcomitrella patens (Iwai et al., 2018), and Tidi in Dunaliela salina (Varsano et al., 2006); and (2) promiscuous antennae (i.e. complexes that can serve both PSI and PSII; Kyle et al., 1983; Wientjes et al., 2013a; Drop et al., 2014; Pietrzykowska et al., 2014).PSI-specific antenna proteins vary in type and number between algae, mosses, and plants. For example, the genomes of several green algae contain a larger number of lhca genes than those of vascular plants (Neilson and Durnford, 2010). The PSI-LHCI complex of plants includes only four Lhcas (Lhca1–Lhc4), which are present in all conditions analyzed so far (Ballottari et al., 2007; Wientjes et al., 2009; Mazor et al., 2017), while in algae and mosses, 8 to 10 Lhcas bind to the PSI core (Drop et al., 2011; Iwai et al., 2018; Pinnola et al., 2018; Kubota-Kawai et al., 2019; Suga et al., 2019). Moreover, some PSI-specific antennae are either only expressed, or differently expressed, under certain environmental conditions (Moseley et al., 2002; Varsano et al., 2006; Swingley et al., 2010; Iwai and Yokono, 2017), contributing to the variability of the PSI antenna size in algae and mosses.The colonial green alga Botryococcus braunii (Trebouxiophyceae) is found worldwide throughout different climate zones and has been targeted for the production of hydrocarbons and sugars (Metzger and Largeau, 2005; Eroglu et al., 2011; Tasić et al., 2016). Here, we have purified and characterized PSI from an industrially relevant strain isolated from a mountain lake in Portugal (Gouveia et al., 2017). This B. braunii strain forms colonies, and since the light intensity inside the colony is low, it is expected that PSI in this strain has a large antenna size (van den Berg et al., 2019). We provide evidence that B. braunii PSI differs from that of closely related organisms through the particular organization of its antenna. The structural and functional characterization of B. braunii PSI highlights a large flexibility of PSI and its antennae throughout the green lineage.  相似文献   

8.
9.
10.
11.
The evolution of Na+-selective four-domain voltage-gated channels (4D-Navs) in animals allowed rapid Na+-dependent electrical excitability, and enabled the development of sophisticated systems for rapid and long-range signaling. While bacteria encode single-domain Na+-selective voltage-gated channels (BacNav), they typically exhibit much slower kinetics than 4D-Navs, and are not thought to have crossed the prokaryote–eukaryote boundary. As such, the capacity for rapid Na+-selective signaling is considered to be confined to certain animal taxa, and absent from photosynthetic eukaryotes. Certainly, in land plants, such as the Venus flytrap (Dionaea muscipula) where fast electrical excitability has been described, this is most likely based on fast anion channels. Here, we report a unique class of eukaryotic Na+-selective, single-domain channels (EukCatBs) that are present primarily in haptophyte algae, including the ecologically important calcifying coccolithophores, Emiliania huxleyi and Scyphosphaera apsteinii. The EukCatB channels exhibit very rapid voltage-dependent activation and inactivation kinetics, and isoform-specific sensitivity to the highly selective 4D-Nav blocker tetrodotoxin. The results demonstrate that the capacity for rapid Na+-based signaling in eukaryotes is not restricted to animals or to the presence of 4D-Navs. The EukCatB channels therefore represent an independent evolution of fast Na+-based electrical signaling in eukaryotes that likely contribute to sophisticated cellular control mechanisms operating on very short time scales in unicellular algae.

Electrical signals trigger rapid physiological events that underpin an array of fundamental processes in eukaryotes, from contractile amoeboid locomotion (Bingley and Thompson, 1962), to the action potentials of mammalian nerve and muscle cells (Hodgkin and Huxley, 1952). These events are mediated by voltage-gated ion channels (Brunet and Arendt, 2015). In excitable animal cells, Ca2+- or Na+-selective members of the four-domain voltage-gated cation channel family (4D-Cav/Nav) underpin well-characterized signaling processes (Catterall et al., 2017). The 4D-Cav/Nav family is broadly distributed across eukaryotes, contributing to signaling processes associated with motility in some unicellular protist and microalgal species (Fujiu et al., 2009; Lodh et al., 2016), although these channels are absent from land plants (Edel et al., 2017). It is likely that the ancestral 4D-Cav/Nav channel was Ca2+-permeable, with Na+-selective channels arising later within the animal lineage (Moran et al., 2015). This shift in ion selectivity represented an important innovation in animals, allowing rapid voltage-driven electrical excitability to be decoupled from intracellular Ca2+ signaling processes (Moran et al., 2015).Na+-selective voltage-gated channels have not been described in other eukaryotes, although a large family of Na+-selective channels (BacNav) is present in prokaryotes (Ren et al., 2001; Koishi et al., 2004). BacNav are single-domain channels that form homotetramers, resembling the four-domain architecture of 4D-Cav/Nav. Studies of BacNav channels have provided considerable insight into the mechanisms of gating and selectivity in voltage-dependent ion channels (Payandeh et al., 2012; Zhang et al., 2012). The range of activation and inactivation kinetics of native BacNav are generally slower than observed for 4D-Nav, suggesting that the concatenation and subsequent differentiation of individual pore-forming subunits may have enabled 4D-Nav to develop specific properties such as fast inactivation, which is mediated by the conserved intracellular Ile–Phe–Met linker between domains III and IV (Fig. 1A; Irie et al., 2010; Catterall et al., 2017).Open in a separate windowFigure 1.EukCatBs represent a novel class of single-domain channels. A, Schematic diagram of a single-domain EukCatB channel. The voltage-sensing module (S1–S4, blue), including conserved positively charged (++) residues of segment (S4) that responds to changes in membrane potential, is shown. The pore module (S5–S6, red) is also indicated, including the SF motif (Ren et al., 2001). The structure of a 4D-Nav (showing the SF of rat 4D-Nav1.4 with canonical “DEKA” locus of Na+-selective 4D-Nav1s) is also displayed (right). The Ile–Phe–Met motif of the fast inactivation gate is indicated (West et al., 1992) B, Maximum likelihood phylogenetic tree of single-domain, voltage-gated channels including BacNav and the three distinct classes of EukCat channels (EukCatA–C). Representatives of the specialized family of single-domain Ca2+ channels identified in mammalian sperm (CatSpers) are also included. SF for each sequence is shown (right). “Position 0” of the high-field–strength site that is known to be important in determining Na+ selectivity (Payandeh et al., 2011), is colored red. Channel sequences selected for functional characterization in this study are shown in bold. EukCatA sequences previously characterized (Helliwell et al., 2019) are also indicated, as is NaChBac channel from B. halodurans (Ren et al., 2001). Maximum likelihood bootstrap values (>70) and Bayesian posterior probabilities (>0.95) are above and below nodes, respectively. Scanning electron micrographs of coccolithophores E. huxleyi (scale bar = 2 μm) and S. apsteinii, (scale bar = 10 μm) are shown.We recently identified several classes of ion channel (EukCats) bearing similarity to BacNav in the genomes of eukaryotic phytoplankton. Characterization of EukCatAs found in marine diatoms demonstrated that these voltage-gated channels are nonselective (exhibiting permeability to both Na+ and Ca2+) and play a role in depolarization-activated Ca2+ signaling (Helliwell et al., 2019). Two other distinct classes of single-domain channels (EukCatBs and EukCatCs) were also identified that remain uncharacterized. These channels are present in haptophytes, pelagophytes, and cryptophytes (EukCatBs), as well as dinoflagellates (EukCatCs; Helliwell et al., 2019). Although there is a degree of sequence similarity between the distinct EukCat clades, the relationships between clades are not well resolved, and there is not clear support for a monophyletic origin of EukCats. The diverse classes of EukCats may therefore exhibit significant functional differences. Characterization of these different classes of eukaryote single-domain channels is thus vital to our understanding of eukaryote ion channel structure, function, and evolution, and to our gaining insight into eukaryote membrane physiology more broadly.Notably, EukCatB channels were found in ecologically important coccolithophores, a group of unicellular haptophyte algae that represent major primary producers in marine ecosystems. Coccolithophores are characterized by their ability to produce a cell covering of ornate calcium carbonate platelets (coccoliths; Fig. 1B; Taylor et al., 2017). The calcification process plays an important role in global carbon cycling, with the sinking of coccoliths representing a major flux of carbon to the deep ocean. Patch-clamp studies of coccolithophores indicate several unusual aspects of membrane physiology, such as an inwardly rectifying Cl conductance and a large outward H+ conductance at positive membrane potentials, which may relate to the increased requirement for pH homeostasis associated with intracellular calcification. Here we report that EukCatB channels from two coccolithophore species (Emiliania huxleyi and Scyphosphaera apsteinii) act as very fast Na+-selective voltage-gated channels that exhibit many similarities to the 4D-Navs, which underpin neuronal signaling in animals. Thus, our findings demonstrate that the capacity for rapid Na+-based signaling has evolved in certain photosynthetic eukaryotes, contrary to previous widely held thinking.  相似文献   

12.
13.
14.
15.
Recent studies have shown that loss of pollen-S function in S4′ pollen from sweet cherry (Prunus avium) is associated with a mutation in an S haplotype-specific F-box4 (SFB4) gene. However, how this mutation leads to self-compatibility is unclear. Here, we examined this mechanism by analyzing several self-compatible sweet cherry varieties. We determined that mutated SFB4 (SFB4ʹ) in S4′ pollen (pollen harboring the SFB4ʹ gene) is approximately 6 kD shorter than wild-type SFB4 due to a premature termination caused by a four-nucleotide deletion. SFB4′ did not interact with S-RNase. However, a protein in S4′ pollen ubiquitinated S-RNase, resulting in its degradation via the 26S proteasome pathway, indicating that factors in S4′ pollen other than SFB4 participate in S-RNase recognition and degradation. To identify these factors, we used S4-RNase as a bait to screen S4′ pollen proteins. Our screen identified the protein encoded by S4-SLFL2, a low-polymorphic gene that is closely linked to the S-locus. Further investigations indicate that SLFL2 ubiquitinates S-RNase, leading to its degradation. Subcellular localization analysis showed that SFB4 is primarily localized to the pollen tube tip, whereas SLFL2 is not. When S4-SLFL2 expression was suppressed by antisense oligonucleotide treatment in wild-type pollen tubes, pollen still had the capacity to ubiquitinate S-RNase; however, this ubiquitin-labeled S-RNase was not degraded via the 26S proteasome pathway, suggesting that SFB4 does not participate in the degradation of S-RNase. When SFB4 loses its function, S4-SLFL2 might mediate the ubiquitination and degradation of S-RNase, which is consistent with the self-compatibility of S4′ pollen.

In sweet cherry (Prunus avium), self-incompatibility is mainly controlled by the S-locus, which is located at the end of chromosome 6 (Akagi et al., 2016; Shirasawa et al., 2017). Although the vast majority of sweet cherry varieties show self-incompatibility, some self-compatible varieties have been identified, most of which resulted from the use of x-ray mutagenesis and continuous cross-breeding (Ushijima et al., 2004; Sonneveld et al., 2005). At present, naturally occurring self-compatible varieties are rare (Marchese et al., 2007; Wünsch et al., 2010; Ono et al., 2018). X-ray-induced mutations that have given rise to self-compatibility include a 4-bp deletion (TTAT) in the gene encoding an SFB4′ (S-locus F-box 4′) protein, located in the S-locus and regarded as the dominant pollen factor in self-incompatibility. This mutation is present in the first identified self-compatible sweet cherry variety, ‘Stellar’, as well as in a series of its self-compatible descendants, including ‘Lapins’, ‘Yanyang’, and ‘Sweet heart’ (Lapins, 1971; Ushijima et al., 2004). Deletion of SFB3 and a large fragment insertion in SFB5 have also been identified in other self-compatible sweet cherry varieties (Sonneveld et al., 2005; Marchese et al., 2007). Additionally, a mutation not linked to the S-locus (linked instead to the M-locus) could also cause self-compatibility in sweet cherry and closely related species such as apricot (Prunus armeniaca; Wünsch et al., 2010; Zuriaga et al., 2013; Muñoz-Sanz et al., 2017; Ono et al., 2018). Much of the self-compatibility in Prunus species seems to be closely linked to mutation of SFB in the S-locus (Zhu et al., 2004; Muñoz-Espinoza et al., 2017); however, the mechanism of how this mutation of SFB causes self-compatibility is unknown.The gene composition of the S-locus in sweet cherry differs from that of other gametophytic self-incompatible species, such as apple (Malus domestica), pear (Pyrus spp.), and petunia (Petunia spp.). In sweet cherry, in addition to a single S-RNase gene, the S-locus contains one SFB gene, which has a high level of allelic polymorphism, and three SLFL (S-locus F-box-like) genes with low levels of, or no, allelic polymorphism (Ushijima et al., 2004; Matsumoto et al., 2008). By contrast, the apple, pear, and petunia S-locus usually contains one S-RNase and 16 to 20 F-box genes (Kakui et al., 2011; Okada et al., 2011, 2013; Minamikawa et al., 2014; Williams et al., 2014a; Yuan et al., 2014; Kubo et al., 2015; Pratas et al., 2018). The F-box gene, named SFBB (S-locus F-box brother) in apple and pear and SLF (S-locus F-box) in petunia, exhibits higher sequence similarity with SLFL than with SFB from sweet cherry (Matsumoto et al., 2008; Tao and Iezzoni, 2010). The protein encoded by SLF in the petunia S-locus is thought to be part of an SCF (Skp, Cullin, F-box)-containing complex that recognizes nonself S-RNase and degrades it through the ubiquitin pathway (Kubo et al., 2010; Zhao et al., 2010; Chen et al., 2012; Entani et al., 2014; Li et al., 2014, 2016, 2017; Sun et al., 2018). In sweet cherry, a number of reports have described the expression and protein interactions of SFB, SLFL, Skp1, and Cullin (Ushijima et al., 2004; Matsumoto et al., 2012); however, only a few reports have examined the relationship between SFB/SLFL and S-RNase (Matsumoto and Tao, 2016, 2019), and none has investigated whether the SFB/SLFL proteins participate in the ubiquitin labeling of S-RNase.Although the function of SFB4 and SLFL in self-compatibility is unknown, the observation that S4′ pollen tubes grow in sweet cherry pistils that harbor the same S alleles led us to speculate that S4′ pollen might inhibit the toxicity of self S-RNase. In petunia, the results of several studies have suggested that pollen tubes inhibit self S-RNase when an SLF gene from another S-locus haplotype is expressed (Sijacic et al., 2004; Kubo et al., 2010; Williams et al., 2014b; Sun et al., 2018). For example, when SLF2 from the S7 haplotype is heterologously expressed in pollen harboring the S9 or S11 haplotype, the S9 or S11 pollen acquire the capacity to inhibit self S-RNase and break down self-incompatibility (Kubo et al., 2010). The SLF2 protein in petunia has been proposed to ubiquitinate S9-RNase and S11-RNase and lead to its degradation through the 26S proteasome pathway (Entani et al., 2014). If SFB/SLFL in sweet cherry have a similar function, the S4′ pollen would not be expected to inhibit self S4-RNase, prompting the suggestion that the functions of SFB/SLFL in sweet cherry and SLF in petunia vary (Tao and Iezzoni, 2010; Matsumoto et al., 2012).In this study, we used sweet cherry to investigate how S4′ pollen inhibits S-RNase and causes self-compatibility, focusing on the question of whether the SFB/SLFL protein can ubiquitinate S-RNase, resulting in its degradation.  相似文献   

16.
17.
18.
19.
Pannexin 1 (Panx1) plays a decisive role in multiple physiological and pathological settings, including oxygen delivery to tissues, mucociliary clearance in airways, sepsis, neuropathic pain, and epilepsy. It is widely accepted that Panx1 exerts its role in the context of purinergic signaling by providing a transmembrane pathway for ATP. However, under certain conditions, Panx1 can also act as a highly selective membrane channel for chloride ions without ATP permeability. A recent flurry of publications has provided structural information about the Panx1 channel. However, while these structures are consistent with a chloride selective channel, none show a conformation with strong support for the ATP release function of Panx1. In this Viewpoint, we critically assess the existing evidence for the function and structure of the Panx1 channel and conclude that the structure corresponding to the ATP permeation pathway is yet to be determined. We also list a set of additional topics needing attention and propose ways to attain the large-pore, ATP-permeable conformation of the Panx1 channel.

IntroductionInitially, the pannexin field got off to an inauspicious start. Because the three proteins of the pannexin family (Panx1, Panx2, and Panx3) were discovered on the basis of their limited sequence homology to the invertebrate innexin gap junction proteins, it was assumed that they, too, would form gap junctions (Panchin et al., 2000; Bruzzone et al., 2003). However, notwithstanding some disputed science, it soon became clear that the gap junction function of pannexins is not realized (Dahl and Locovei, 2006; Huang et al., 2007; Sosinsky et al., 2011). Gap junction formation by pannexins is moreover prevented by glycosylation of the Panx1 protein (Boassa et al., 2007; Penuela et al., 2007; Boassa et al., 2008). It is conceivable, however, that the nonglycosylated form of Panx1 could allow the docking of oligomers in apposing membranes to each other. Indeed, such docking was observed in cryo-EM preparations of a mutant Panx1, N255A (Ruan et al., 2020). However, because glycosylation of the protein is required for membrane trafficking (Boassa et al., 2007; Penuela et al., 2007), it is unlikely that this process occurs in vivo. This notion is supported by the observation that a glycosylation mutant of Panx1 failed to lead to junctional conductance between paired oocytes, while expression of wtPanx1 followed by glycosidase F treatment before pairing of the cells resulted in gap junction formation, albeit at a much reduced rate as compared with the rate observed with connexins (Boassa et al., 2008). Furthermore, the gap junction function also is unlikely because of theoretical reasons and experimental findings, including expression in single cells, such as erythrocytes (Locovei et al., 2006a), and exclusive expression in the apical membrane of polarized epithelial cells (Ransford et al., 2009; Hanner et al., 2012; Shum et al., 2019). Of the three pannexins, only Panx1 has been studied in detail, and a clear channel function is well established for it. Therefore, this Viewpoint will exclusively deal with Panx1.It is generally assumed that an important function of Panx1 is to form ATP release channels. For example, as of this writing, a PubMed search with “pannexin” and “ATP release” as search terms yielded 287 publications. However, the biophysical evidence for this function is sparse, and the supporting evidence is largely correlative and/or based on pharmacological or genetic interference with ATP release. The ATP release function of Panx1 channels has been contested based on evidence that the channels are highly selective for Cl and the lack of detectable ATP release under the experimental conditions, i.e., activation by voltage, used in these studies (Ma et al., 2012; Romanov et al., 2012).It was long held that the oligomeric state of the Panx1 channel is that of a hexamer (Penuela et al., 2013; Dahl, 2015; Chiu et al., 2018). Recently, in a <6-mo period, six research groups independently published similar cryo-EM structures of the Panx1 membrane channel (Deng et al., 2020; Jin et al., 2020; Michalski et al., 2020; Mou et al., 2020; Qu et al., 2020; Ruan et al., 2020). All these papers challenge the view of the oligomeric state of the channel. Instead of the hexameric arrangement of identical Panx1 subunits, the channel appears to be formed by a homomeric heptamer.It is uncontested that the Panx1 channel can operate in a conformation in which the channel exhibits a unitary conductance of <100 pS. However, it has also been reported that, under certain experimental conditions, the channel can exhibit several subconductances with very rare sojourns to a maximal conductance of ∼500 pS (Dahl 2015).Activation of the Panx1 channel can be obtained by various stimuli, some physiological or pathological, others in the form of experimental tools. The majority of the physiological stimuli involve ligands, including ATP, glutamate, α adrenergic agonists, and bradykinin, binding to their cognitive receptors and leading to opening of Panx1 channels and “secondary” ATP release (Dahl 2015, 2018). A nonreversible activation of the channel is induced by cleavage with caspase 3 or 8 (Chekeni et al., 2010), leading to cell death. Experimental stimulation can be obtained in some cells but not in others by increasing the extracellular K+ concentration (Silverman et al., 2009; Wang and Dahl 2018; Chiu et al., 2018).Here, we present a point of view on these contested issues in the Panx1 field and attempt to reconcile apparently contradictory data or their interpretations. In addition, we ask to what extent the new structural data support the functional data.The structure of Panx1 channels as revealed by cryo-EMThe excitement for the long-awaited Panx1 structure was rewarded with not one but six publications, all appearing within the first half of 2020 and using cryo-EM to image the channel (Deng et al., 2020; Jin et al., 2020; Michalski et al., 2020; Mou et al., 2020; Qu et al., 2020; Ruan et al., 2020). Five of the six Panx1 publications showed human Panx1 structures (Deng et al., 2020; Jin et al., 2020; Michalski et al., 2020; Mou et al., 2020; Qu et al., 2020; Ruan et al., 2020), and two showed frog Panx1 structures (Deng et al., 2020; Michalski et al., 2020). To stabilize the frog Panx1 structure, thus improving the resolution for cryo-EM, Michalski et al. (2020) truncated the C terminus by 71 amino acids and removed 24 amino acids from the intracellular loop between transmembrane (TM) helices 2 and 3; this truncation was the version reported. The other publications reported full-length Panx1 structures, although one used a truncated Panx1 variant (Ruan et al., 2020) and another used a mutation (Jin et al., 2020) as their principal references. Because the C terminus–truncated structure and the (combined) WT structure had the highest overall quality of the several Panx1 structures presented by Ruan et al. (2020), it was used for de novo model building, as the reference structure for analyses, and for discussion. For the same reason, Jin et al. (2020) used the double mutation D376E/D379E, which eliminated the caspase-cleavage site, for their de novo model building. The overall architecture and dimensions of the human and frog Panx1 structures were similar, and the extracellular domains (ECDs) and transmembrane domains (TMDs) were nearly identical, except that the N-terminal helix (NTH) of the frog Panx1 is positioned on the intracellular side and not within the TMD as reported in human Panx1. These structures are becoming an important and critical tool to guide functional experiments and open a new venue to perform molecular dynamics simulations.The most novel finding shared by all the Panx1 structures is that instead of the hexamer hitherto believed to represent the native state of the Panx1 channel, the cryo-EM data show unequivocally a homo-heptamer arranged around a central symmetry axis that constitutes the principal permeation pathway. Panx1’s structural envelope adopts an inverted pail shape in the view parallel to the membrane. The heptameric channel is ∼110 Å long and ∼100 Å wide, with a flat ECD protruding ∼35 Å above the cell membrane and an intracellular domain (ICD) extending ∼35 Å into the cytoplasm with the TMD in between (Fig. 1). Both the N and C termini reside on the cytoplasmic side of the channel. Each TMD has four membrane-spanning helices per protomer. The TMD pore is predominantly lined by hydrophobic amino acids. The NTH is short and lines the TMD pore (human Panx1) with the NTH of one subunit interacting with the TMD of the adjacent subunit. Thus, it has been proposed that the NTH helps to maintain a rigid TMD pore (Ruan et al., 2020). Both the extracellular and intracellular pore entrances are lined with positively charged amino acids, making them favorable for negatively charged cargos such as Cl and ATP to enter and leave the pore.Open in a separate windowFigure 1.Structural model of Panx1. (A) Cartoon representation of Panx1 WT (PDB accession no. 6wbk) stabilized with detergent. The approximate position of the membrane is indicated by the bars. One protomer of the heptameric assembly is indicated in orange. The lipids resolved in the structure are colored in magenta. It is notable that many of the resolved lipids are found at the interface between subunits. CTH, C-terminal helix; CLH, cytoplasmic linker helix. (B) Overlay of different Panx1 cryo-EM–based structures from different groups (PDB accession no. 6lto = gold, WT hPanx1; PDB accession no. 6wbk = blue, hPanx1 ΔC terminus, ΔN terminus; PDB accession no. 6v6d = lilac, WT hPanx1). The structures diverge very little in the ECD and the membrane domain. However, the variation is larger in the ICD. The approximate positions of F54 and I58 are indicated. (C) Comparison of the extracellular pore. Key residues are annotated using their one-letter abbreviation. All structures are shown as backbone traces except W74, which is part of the constriction ring of the pore. The narrowest constriction is indicated by the arrow.A cap structure is formed on the extracellular side from seven ECDs, one from each subunit. The ECDs are organized into the pore’s most constricted site, thus defining the maximum size of permeable molecules and establishing the extracellular entrance to the TMD. Each ECD contains one helix cross-linked through a disulfide bond to a three-stranded antiparallel β-sheet. The N-terminal end of each subunit helix protrudes inward toward the central pore axis contributing to the constriction site. The key residues involved in the gating of Panx1 are part of the W74-R75-D81 inter-subunit triplet (Fig. 1) from the N-terminal ends of the ECD helices and are proposed to provide rigidity to the extracellular entrance containing positive charges (Deng et al., 2020). This ring lining the wall of the constriction defines a hydrated pore diameter of ∼9 Å (Fig. 1) and is the crux of the argument that the extracellular constriction site is the effective selectivity filter discriminating cargoes on the basis of their charge and size. By contrast, the narrowest point of the pore in the TMD has a hydrated diameter of ∼13 Å.Other inter-subunit interactions that stabilize the pore arrangement in the TMD are (1) the F67-Y79 aromatic–aromatic and Q266-T252 hydrogen bonding in the ECD, (2) the TM helix TM1-TM1 and TM2-TM4 interfaces between adjacent subunits, (3) the N-terminal loop from the neighboring subunit extending into the TMD pore lining, and (4) the resolvable residues of C-terminal helix 3 association with the cytoplasmic linker helices 1 and 2 from the adjacent subunit (Fig. 1; Deng et al., 2020; Michalski et al., 2020). In addition, membrane lipids occupy the crevice between TM3 and an adjacent TM4 at the edge of the channel. Lipid head groups associate with positively charged residues from the N-terminal end of TM3 (K214) and the C-terminal domain of TM4 (R302 and K303; Deng et al., 2020). It may be that the lipid environment influences the assembly and possibly the activation of the Panx1 channel.The pore substantially widens toward the intracellular side. The ICDs extend away from the pore axis, producing a voluminous intracellular vestibule. Two helices in the ICD connect to two TMD helices to form a compact assembly that provides rigidity to the vestibule. The vestibule is slightly larger in the frog Panx1 structure, but the pore architecture is essentially the same as in the human version. The C-terminal segment consisting of ∼70 amino acids, including the caspase cleavage site located immediately before the C terminus, was not visualized in any of the published cryo-EM structures, suggesting that it is intrinsically flexible. C-terminal cleavage removes its autoinhibitory effect without inducing an overall conformational change to the pore (Jin et al., 2020).A novel feature presented by Ruan et al. (2020) is seven narrow side tunnels in the upper ICD running perpendicular to and ending at the pore, like a T-intersection. This tunnel network was proposed to allow passage of small anions because each tunnel contains several positively charged and other polar residues along its length. The flexible NTH-TM1 linker would function as a gate. They hypothesized that these side tunnels facilitate Cl flux when the pore is blocked by the C terminus residing in the ICD vestibule. Movement or cleavage of the C terminus then would induce the ATP-permeable large-pore conformation. Further structural work is needed to determine whether the C terminus actually resides in the ICD vestibule. One complication with this tunnel hypothesis is that each tunnel has two constriction sites formed by R29 and A33, with diameters of 5.8 and 4.0 Å, respectively, meaning that partial dehydration of a Cl ion is likely required for it to pass through the tunnel. The possibility of anion tunnels in the Panx1 structure raises the question: could there be multiple open (active) conformations, and does “open” mean open to ATP and Cl ion or Cl ion only? Three of the six papers (Michalski et al., 2020; Qu et al., 2020) assume that the Panx1 structure is in the closed or inactive conformation, whereas the other three report an open or active conformation. These discrepancies should be resolved with further work. Interestingly, the overlay of three of these structures, one presumably open and two closed, shows nearly identical conformations of the extracellular constriction of the pore (Fig. 1 C).The structures of Panx1 truncation and mutant variants, as well as its docked inhibitor, carbenoxolone (CBX), illuminate the roles of distinct structural elements. For example, cryo-EM revealed that CBX sits atop the ECD and plugs the W74-R75-D81 ring (Jin et al., 2020; Ruan et al., 2020), supporting pore blocking as its mechanism of Panx1 inhibition. Second, removal of the NTH did not much change the shape of the TMD. This truncation still generated a CBX-sensitive, voltage-dependent current similar to the full-length Panx1 current (Ruan et al., 2020), suggesting that the NTH may play a role in the assembly of the heptamer from the Panx1 protomers rather than being an essential element in the pore conduction pathway. Third, the Panx1 structures obtained in the presence of Ca2+ and K+ were indistinguishable from the untreated Panx1 structure, suggesting that neither Ca2+ nor K+ is likely to directly activate the Panx1 channel (Ruan et al., 2020). Fourth, the glycosylation-deficient mutant (N255A) generated gap junction structures as well as hemichannels as determined by cryo-EM (Ruan et al., 2020). The N255A gap junction is formed by two hemichannels docked by the interaction of the extracellular linker 2 on apposing ECDs. Ruan et al. (2020) noted that this Panx1 gap junction likely is not a normal physiological conformation.Comparing the Panx1 structure with other ATP-permeant pore structures offers a blueprint for dissecting pore properties of Panx1. The oligomeric configuration of the ATP-permeant channels does not appear to correlate with pore diameter, in particular when comparing the narrowest extent of the pore (Table 1). Like Panx1, CALHM1 was originally thought to adopt a hexameric configuration (Siebert et al., 2013) until the cryo-EM structure revealed an octameric configuration (Syrjanen et al., 2020). It should be noted that not all connexin channels transport ATP. For example, Cx26, Cx30, Cx43, and Cx46 transport ATP, but Cx32 does not (Harris, 2001; Hansen et al., 2014). An interesting comparison can be made between Panx1 and human Cx31.3. Cx31.3 has a pore with a constriction site diameter of ∼8 Å, similar to Panx1’s constriction of ∼9 Å, and also selectively transports Cl and ATP (Lee et al., 2020). One difference between the two structures is that the constriction is in the ICD’s NTH for Cx31.3 and in the ECD for Panx1. Lee et al. (2020) pointed out that because Cx31.3’s pore constriction at the cytoplasmic entrance is smaller than the effective hydrated diameter of ATP (∼12 Å; Sabirov and Okada, 2004), the reported conformation would not allow ATP to pass through the pore without substantial conformational changes of the NTH. The same argument could be applied to Panx1’s ECD.Table 1.Comparison of oligomeric configuration with pore constriction diameter
ATP-permeant channelOligomeric configurationPore constriction diameter (Å)Reference
VDACMonomer (β barrel)25Colombini, 2012
Cx26Hexamer14 (open), 6 (closed)Maeda et al., 2009; Nielsen et al., 2012
Cx31.3Hexamer8 (closed?)Lee et al., 2020
Cx43Hexamer25 (open), 18 (closed)Nielsen et al., 2012
Cx46Hexamer14Myers et al., 2018; Flores et al., 2020
VRACHexamer12–14 (open), 2–5 (closed)Deneka et al., 2018; Kasuya et al., 2018; Kefauver et al., 2018; Osei-Owusu et al., 2018
Panx1Heptamer9Deng et al., 2020; Jin et al., 2020; Michalski et al., 2020; Mou et al., 2020; Qu et al., 2020; Ruan et al., 2020
Innexin6Octamer18Oshima et al., 2016
CALHM1Octamer19.5Syrjanen et al., 2020
CALHM2Undecamer50 (open), 23 (inhibited)Choi et al., 2019
CALHM4Decamer, undecamer20 (without current), 30 (without current)Drożdżyk et al., 2020
Open in a separate windowIn a comparison of structural motifs, Panx1’s protomer shares a motif of four TMD helices with connexins, innexins, CALHM1 and 2, and volume-regulated anion channel (VRAC) protomers. They also share ECD topological similarities, but differences are manifested as well. For example, Panx1 and innexin6 have their ECD2 nestled between the inner and outer lobes of ECD1. CALHM1 and 2 and VRAC do not, and the outer lobe of ECD1 is absent in Cx46 (Myers et al., 2018; Deng et al., 2020; Flores et al., 2020; Qu et al., 2020). In Panx1 and VRAC, residues in the ECD1 helix form the pore constriction site, whereas in Cx46 and innexin6, the NTH forms the constriction in the middle of the TMD. Also, Panx1, innexin6, and CALHM2 have two ECD disulfide bonds, whereas Cx46 has three. The NTH of Panx1 is unique among ATP-permeant channels because it resides in the TMD pore close to the ECD, whereas the NTH in connexins and innexins is in the TMD closer to the ICD (Maeda et al., 2009; Myers et al., 2018; Flores et al., 2020), and the NTH in VRAC is in the ICD near the membrane face (Deneka et al., 2018; Kasuya et al., 2018; Kefauver et al., 2018). In Panx1, connexins, innexins, and VRAC, TM1–TM4 are arranged anticlockwise as viewed from intracellular side, with TM1 nestled between TM2 and TM4, but the CALHM2 arrangement is clockwise and TM1 is attached to TM3. Finally, even though the ICD helix 1 is long, it is not involved in inter-subunit interactions, whereas in CALHM2, the equivalent cytoplasmic helix 1 is nearly parallel to the membrane and forms extensive inter-subunit interactions (Choi et al., 2019; Syrjanen et al., 2020).Biophysical properties and function of Panx1 channelsWhile the Panx1 structures obtained in six different laboratories are remarkably congruent, including the structures from different species, this is not the case for several functional aspects. It is now generally accepted that, contrary to original beliefs, pannexins do not form gap junctions despite their classification as “gap junction proteins” (Sosinsky et al., 2011). Instead, Panx1 is well documented to form a plasma membrane channel, which allows the exchange of solutes between the cytoplasm and the extracellular space.The first demonstration of Panx1 channel function involved the activation of the channel by positive membrane potentials in excess of +20 mV (Bruzzone et al., 2003). Robust Panx1-mediated membrane currents induced by positive membrane voltage were subsequently reported by various laboratories (Pelegrin and Surprenant, 2006; Iglesias et al., 2008; Iglesias et al., 2009b; Qiu and Dahl, 2009; Bunse et al., 2009; Prochnow et al., 2009; Silverman et al., 2009; Bunse et al., 2011; Gründken et al., 2011; Ma et al., 2012; Romanov et al., 2012; Zhan et al., 2012; Nomura et al., 2017). In oocytes expressing rodent Panx1, the currents are in the microampere range, and in these as in mammalian cells, the currents exhibit strong outward rectification.Weak Panx1 channel activation by positive membrane voltage was described, which could be boosted substantially by the insertion of a glycine-serine motif immediately after the first methionine (Michalski et al., 2018). Human and mouse Panx1 behaved similarly in these experiments, while mammalian cell lines (HEK and CHO) yielded different current densities for both pannexins. Frog Panx1 was subsequently found to exhibit similar voltage-sensitive properties (Michalski et al., 2020). In contrast, complete absence of voltage activation has been reported for human and frog Panx1, while mouse Panx1 was found to be weakly voltage sensitive (Chekeni et al., 2010; Chiu et al., 2018; Narahari et al., 2021).The first physiological role ascribed to Panx1 was that of a permeation pathway for ATP, allowing the efflux of ATP to act as an external signal to cells via activation of purinergic receptors (Bao et al., 2004). Subsequently, hundreds of publications affirmed the ATP release functions of Panx1 in multiple cell types based mainly on pharmacological and/or genetic interference. However, the role of Panx1 as an ATP-release channel has been challenged by data obtained with biophysical measurements. It has been shown that the voltage-activated Panx1 channel was highly selective for chloride ions and did not exhibit detectable ATP release (Ma et al., 2012; Romanov et al., 2012; Wang et al., 2014). On the other hand, several groups reported ATP release and/or activation of Panx1 currents after exposure of cells to high extracellular potassium ion (K+) concentrations (Silverman et al., 2009; Santiago et al., 2011; Heinrich et al., 2012; Suadicani et al., 2012; Michalski and Kawate, 2016; Qu et al., 2020). However, a lack of K+-mediated activation of Panx1 has also been reported (Chiu et al., 2018; Nielsen et al., 2020).Neither high positive membrane potentials nor high extracellular K+ concentrations can be considered physiological stimuli. A third stimulus for Panx1 activity, cleavage of C-terminal amino acids by caspase, occurs exclusively in apoptotic cells (Chekeni et al., 2010). A correlation of caspase cleavage with ATP release has been established (Chekeni et al., 2010; Boyd-Tressler et al., 2014; Imamura et al., 2020). However, whether caspase cleavage of Panx1 per se causes ATP release remains to be determined. The similarity of the electrophysiological profiles of the voltage-activated and the caspase-cleaved Panx1 channels (Chiu et al., 2017; Wang and Dahl, 2018) raises the question whether caspase by itself is sufficient for initiating ATP release.However, if we accept that physiological ATP release mediated by Panx1 occurs in a reversible fashion, as proposed in >200 publications, there must be an as-yet-unidentified gating mechanisms, since the activation mechanisms presently documented are either irreversible or unphysiological. Erythrocytes, for example, release ATP reversibly in a low oxygen environment but also in response to shear stress (Locovei et al., 2006a; Sridharan et al., 2010; Forsyth et al., 2011; Cinar et al., 2015; Zhang et al., 2018), two unrelated stimuli converging on the same ATP release channel Panx1.While it appears that the Panx1 field is riddled with controversies, the following paragraphs show that this is not necessarily so. It can safely be assumed that most, if not all, published data are correct. It may take only a little tuning of the interpretations to get a more coherent picture of the Panx1 field.Evidence for ATP release mediated by Panx1 channelsBased on the biophysical properties of the channel formed by Panx1, its function as an ATP release channel was proposed soon after its discovery (Bao et al., 2004; Dahl and Locovei, 2006; Locovei et al., 2006a). Ever since, supporting evidence has been accumulated by a large number of researchers, as documented in several reviews (Dahl and Locovei, 2006; Scemes et al., 2007; MacVicar and Thompson, 2010; Sosinsky et al., 2011; Penuela et al., 2013; Dahl, 2015). Key evidence includes the localization of Panx1 expression matching the site of ATP release at the apical membrane of polarized epithelial cells such as in the airways or renal tubules (Ransford et al., 2009; Hanner et al., 2012). Interestingly, in polarized cells releasing ATP at the basolateral membrane, this function is exerted by another release channel, CALHM (Taruno et al., 2013; Kashio et al., 2019). Additional evidence for the involvement of Panx1 in ATP release is that knockout of Panx1 expression attenuates ATP release (Iglesias et al., 2009a; Qiu et al., 2011; Qu et al., 2011; Suadicani et al., 2012). Furthermore, mutations of Panx1 and chemical modification of the protein affect ATP release (Wang and Dahl, 2010; Qiu et al., 2012).Evidence that Panx1 channels can be Cl selective and lack ATP permeabilityThe ATP release function has been challenged, however, also based surprisingly on sound biophysical evidence (Ma et al., 2012; Romanov et al., 2012; Wang et al., 2014; Wang and Dahl, 2018). No ATP release was observed in voltage-clamped, Panx1-expressing HEK cells over the voltage range of −50 to +80 mV, despite robust CBX-sensitive currents at positive potentials (Romanov et al., 2012). Similarly, oocytes expressing Panx1 did not release ATP when the membrane potential was clamped at −60, 0, or +40 mV. However, CBX-sensitive ATP release was observed at the same holding potentials in the presence of high extracellular K+ concentrations (Fig. 2 A; Wang et al., 2014). Instead, the exclusively voltage-activated Panx1 channel was found to be highly selective for Cl. Substitution of extracellular Cl by larger anions resulted in attenuation of the membrane currents and a large shift of the reversal potential from negative to positive membrane potentials in both HEK cells and Xenopus laevis oocytes (Fig. 3, A, B, and D; Ma et al., 2012; Romanov et al., 2012; Chiu et al., 2014; Nomura et al., 2017; Li et al., 2018; Wang and Dahl, 2018). Thus, the Panx1 channel is highly selective for Cl and lacks ATP permeability under the experimental conditions used in the referenced studies. Furthermore, in measuring the permeability of the Panx1 channel, it was found that substitution of extracellular Na+ by larger cations changed neither the reversal potential nor the amplitude of the currents in response to voltage steps or voltage ramps, indicating lack of cation permeability (Ma et al., 2012; Wang et al., 2018; Michalski et al., 2020).Open in a separate windowFigure 2.ATP release by oocytes expressing Panx1. (A) Oocytes expressing WT Panx1. All colored bars represent data obtained under voltage-clamp conditions; white bars are data from unclamped cells. ATP in the medium was measured as luciferase luminescence. ATP release from Panx1-expressing oocytes induced by potassium gluconate (KGlu) without and with holding the membrane potential under voltage-clamp conditions at −60, 0, or +40 mV, was determined 20 min after initiating the stimulus. Voltage-clamp conditions are indicated with teal lines below the graph and teal bars in the graph. The presence of 150 mM KGlu (K+ label) is indicated, and the data are displayed as hatched bars. Data are shown as means ± SD; n = 5 for each measurement. Adapted from Wang et al., 2014. (B) Oocytes expressing the truncation mutant Panx1Δ378. Oocytes were not voltage clamped (white bar) or were clamped at −60 or 0 mV (red bars). The cells were exposed to oocyte Ringer solution or to a solution containing 85 mM KGlu (K+ label, hatched bars) as indicated, for 10 min. An aliquot of the supernatant was analyzed for the presence of ATP with the luciferase/luciferin assay. Means ± SE; n as indicated above each bar. Adapted from Wang and Dahl, 2018. Because oocytes expressing Panx1Δ378 have a shortened life span (Jackson et al., 2014), measurements had to be taken in a short time window after injection of mRNA at 80× lower concentration than wtPanx1. Thus, ATP release data cannot be compared between wtPanx1- and Panx1Δ378-expressing cells.Open in a separate windowFigure 3.Current–voltage relations of mouse WT Panx1 and Panx1Δ378. (A) Replacing extracellular Cl by gluconate in a voltage step protocol applied to HEK cells expressing Panx1 exogenously attenuated the membrane currents and shifted the reversal potential to a positive potential. Adapted with permission from Journal of Cell Science (Romanov et al., 2012). (B) Similarly, in oocytes expressing Panx1 exogenously, replacement of Cl by gluconate (NaGluc) resulted in a shift of the reversal potential to positive values and an attenuation of the currents induced by a voltage ramp from −100 to +100 mV. (C) Voltage ramp–induced currents of wtPanx1-expressing oocytes in KCl solution (black trace) were substantially larger than the currents in uninjected control cells under identical conditions (gray trace). Replacing Cl by gluconate in the bath solution of the same oocyte resulted in attenuation of the currents and a shift of the reversal potential to more negative potential (green trace). (D) Quantitative analysis of reversal potentials after anion replacement shows a shift from positive to negative potentials with K+ as the extracellular cation. Means ± SE; n = 5 (wtPanx1). (E) Voltage ramp–induced membrane currents of WT Panx1 channels (green trace) and of channels formed by the truncation mutant Panx1Δ378, where C-terminal amino acids after aspartate 378 are deleted. Uninjected oocytes served as control (magenta trace). Because Panx1Δ378-expressing cells had a short life span, mRNA was injected 80× diluted as compared with WT, and measurements were performed in a 24-h window. Thus, the current amplitudes of WT Panx1 and Panx1Δ378 cannot be compared. In contrast to WT Panx1 channels, those formed by Panx1Δ378 were active over a wide voltage range, yet currents through both types of channels reversed at the same membrane potential. (F) Similar to WT Panx1, currents through Panx1Δ378 channels reversed at negative membrane potential with extracellular chloride (blue trace) and at positive potential with gluconate solution (red trace). (G) Similar to WT Panx1 channels, replacement of chloride by gluconate ions resulted in the currents reversing at negative potentials when Panx1Δ378-expressing cells were exposed to high extracellular K+. (H) Quantitative analysis of reversal potentials of membrane currents carried by Panx1Δ378 after anion replacement shows a shift from positive to negative potentials with K+ as the extracellular cation. Means ± SE; n = 4. (B–H) Data from Wang and Dahl (2018).Lack of cation permeability, however, appears to be restricted to the voltage-activated Panx1 channel. The only direct measure of ATP permeability presently available indicates that the ATP-permeable conformation of the Panx1 channel also is permeable to cations (Bao et al., 2004; Locovei et al., 2006a). In an isolated membrane patch subjected to a K+-ATP gradient, the channel currents reversed neither at the K+ nor the ATP2− reversal potential, but in between the two (closer to the K+ equilibrium potential). Furthermore, uptake of positively charged dyes is considered to be an acceptable surrogate for channel-mediated ATP release (Dahl, 2015; Johnson et al., 2016; Chiu et al., 2018). Thus, under conditions of ATP release, the channel does not discriminate on the basis of charge. However, since dye uptake can be mediated by various other mechanisms, including other membrane channels such as gasdermin D (de Vasconcelos et al., 2019) or CALHM (Siebert et al., 2013), dye uptake per se cannot be used as evidence for the involvement of Panx1.Can Panx1 be both a Cl-selective and a nonselective channel?The Panx1 channel was activated by voltage when it was shown to have high selectivity for Cl and to lack both cation and ATP permeability (Romanov et al., 2012; Ma et al., 2012; Wang et al., 2014; Wang and Dahl 2018). This activation was unphysiological, however, for the membrane potential had to be at high positive potentials for Panx1-mediated currents to be detectable. In contrast, all the evidence supporting the ATP release function involved alternative activation mechanisms. Several stimuli, including oxygen deprivation and mechanical stress, trigger channel-mediated ATP release through Panx1 channels (Locovei et al., 2006a; Thompson et al., 2006) This “primary” ATP release is complemented by a “secondary” ATP release (Dahl, 2018), in which a ligand binding to its cognizant receptor opens Panx1 and releases ATP. This process initially explained ATP-induced ATP release (Locovei et al., 2006b) and has now been demonstrated to apply to many other ligands and receptors, including NMDA, angiotensin II, and α adrenergic receptors (Thompson et al., 2008; Billaud et al., 2011; Murali et al., 2014). The function of secondary ATP release is to boost the cytoplasmic signal, which in all known cases is [Ca2+]i. This amplification is achieved by ATP binding to P2Y receptors, triggering further Ca2+ release from intracellular stores.We propose that the Panx1 channel has at least two distinct open conformations with different permeabilities. When activated by high positive voltage or when truncated at the caspase cleavage site, the channel is highly selective for Cl and has no cation or ATP permeability. In contrast, it has been inferred that when opened in physiological settings by various stimuli, the channel is permeant to ATP and allows the TM flux of cationic and anionic dyes. The latter permeability properties can experimentally be induced by high extracellular K+ concentrations, as verified by changes in the reversal potential and high rate of ATP release. However, since activation by extreme voltage or by K+ is unphysiological, it seems prudent to reexamine with biophysical measurements the permeability properties of the Panx1 channel when activated physiologically, i.e., by ligand binding to receptors known to activate Panx1.Activation of Panx1 by extracellular K+ATP release in a variety of cell types can be stimulated by raising the extracellular K+ concentration; based on pharmacological interference, the release has been attributed to Panx1 channels (Bao et al., 2004; Qiu and Dahl, 2009; Silverman et al., 2009; Santiago et al., 2011; Heinrich et al., 2012; Suadicani et al., 2012; Wang et al., 2013). The ATP release shown in one of the published cryo-EM studies also used K+ to stimulate ATP release (Qu et al., 2020). Since an increase in extracellular K+ depolarizes the plasma membrane, it is logical to assume that depolarization is the trigger for the release (Heinrich et al., 2012). However, the shift in membrane potential induced by K+ should not be sufficient to bring the Panx1 channel into the voltage range where it is fully active. Indeed K+-induced Panx1-mediated currents were observed under voltage-clamp conditions over a wide voltage range (Silverman et al., 2009). The slow time course of the K+-induced Panx1 activation also speaks against depolarization as the critical step. Substitution of several amino acids in the extracellular loops of Panx1 by alanine modulated the response (Wang et al., 2018). Among the substitutions attenuating the K+ effect were amino acids W74 and R75, which are part of the external constriction observed in all published cryo-EM structures. Alanine substitution of amino acid D241 in the second extracellular loop resulted in amplification of the K+ response. Although the alanine substitution effects are suggestive for a direct action of K+ on Panx1 through binding and/or gating, the very slow activation of Panx1-mediated currents indicates that the mechanism responsible could be more complex. Additional evidence for K+-activation of Panx1 channels comes from the observation that currents carried by Panx1 were inhibited in a cysteine replacement mutant (Panx1T62C, C426S) by the thiol reagent 2-(trimethylammonium)ethyl methanethiosulfonate regardless of whether voltage or K+ served as stimulus (Wang et al., 2014).It has been shown that unpaired connexons also exhibit sensitivity to extracellular K+. For example, replacement of extracellular Na+ with K+ resulted in a reversible >10-fold potentiation of Cx50 hemichannel currents (Srinivas et al., 2006). As observed for Panx1 channels, the effect of K+ on Cx50 channels was observed under voltage-clamp conditions, excluding depolarization as a mechanism.It appears that the K+-induced activation of Panx1 may not work in all cell types. For example, Panx1 currents were induced in CHO cells by K+ (Michalski and Kawate, 2016), while HEK cells expressing Panx1 exogenously did not respond to increased extracellular K+ with an induction of CBX-sensitive currents (Chiu et al., 2018). The discrepancy in cell responses to K+ could be due to the involvement of additional factors, which could be facilitating in some cell types or inhibitory in others. The observation of a lack of K+ effect on the cryo-EM structure is consistent with such a scenario (Ruan et al., 2020). However, the solution of this problem could be much more trivial. In the experiments where K+ failed to activate Panx1 in cells, the extracellular solution contained millimole concentrations of Ca2+ (Chiu et al., 2018; Nielsen et al., 2020). Because divalent cations, including Ca2+, interfere with the K+ effect on Panx1 (Wang et al., 2018), a response would not be expected under these conditions.Whatever the detailed mechanism of the Panx1 channel activation by K+ may be, the Cl selectivity of the channel is not evident in high extracellular K+ (Fig. 3, C and D). Replacement of extracellular Cl by gluconate resulted in Panx1-mediated currents, which reversed at negative potentials instead of the reversal at positive potentials observed in the absence of the K+ stimulus. In high extracellular K+, the Panx1 channels were active over the whole voltage range from −100 to +100 mV, yielding membrane currents with amplitudes exceeding the endogenous oocyte currents by far (see also Fig. 3 C; Silverman et al., 2009; Wang et al., 2018).ATP release studies under voltage clamp conditions have shown that Panx1 channel activity induced by membrane depolarization does not result in ATP release (Romanov et al., 2012; Wang et al., 2014). However, ATP release was detected when the extracellular solution contained high K+ concentrations at negative, zero, or positive membrane potentials (Fig. 2 A).The extracellular [K+] required to activate Panx1 channels exceeds by far the physiological concentration of the ion. However, like the activation by voltage, the K+ stimulus is an experimentally convenient tool. More importantly, under pathological conditions, such as epilepsy, stroke, or central nervous system trauma, the extracellular [K+] can exceed 50 mM (Somjen, 1979; Gidö et al., 1997; Sick et al., 1998), thereby activating Panx1. Moreover, at these concentrations, K+ interferes with the negative feedback of ATP on Panx1 channel function (Qiu and Dahl, 2009; Qiu et al., 2012; Jackson et al., 2014). Alanine scanning mutagenesis suggests partially overlapping amino acids in the extracellular loops being involved in both activation of Panx1 channels by K+ and inhibition of the channel by ATP (Qiu et al., 2012; Wang et al., 2018).Activation of Panx1 by caspase cleavageA completely different activation for Panx1 channels has been discovered in apoptotic cells and involves the cleavage of 48 (mouse) or 47 (human) C-terminal amino acids by caspase (Chekeni et al., 2010). Certainly, this is an irreversible process associated with cell death. Nevertheless, this mechanism has found general acceptance as a physiological function of Panx1, despite the fact that in normal cells ATP release is reversible and consequently cannot involve cleavage of the C terminus. Because a deletion mutant mimicking the caspase cleaved Panx1 has no cytoplasmic plug of the sort observed in the WT Panx1 structure, this has been said to provide a structural basis for the cleaved Panx1’s ATP permeability (Ruan et al., 2020).With caspase cleavage, the channel becomes constitutively active, with a high open probability not only at positive potentials, but also at potentials as low as −100 mV (Chekeni et al., 2010; Jackson et al., 2014; Chiu et al., 2017; Chiu et al., 2018; Wang and Dahl, 2018). Fig. 3 E shows the current–voltage relationship of WT Panx1 channels and that of Panx1Δ378, a truncation mutant mimicking the caspase-cleaved Panx1 (Wang and Dahl, 2018). Panx1Δ378 was active over the whole voltage range of −100 to + 100 mV. Thus, caspase cleavage has a profound effect on channel gating by increasing the open probability at negative membrane potentials. In their original report, Chekeni et al. (2010) reported that apoptosis, ATP release, and caspase cleavage of Panx1 were correlated. However, it was not clear whether the anti-Fas stimulus for apoptosis itself caused the large pore conformation, before or while the channel was rendered constitutively active by the cleavage, which sealed the apoptotic fate of the cells.Subsequent data on the caspase-cleaved Panx1 channel published by the same group raised doubts about the ATP permeability of the truncated channel in the absence of another stimulus. They reported that the cleaved channel has the same selectivity for Cl as the exclusively voltage-activated channel (Chiu et al., 2014). The voltage-activated channel, however, has been reported not to be ATP or cation permeable (Ma et al., 2012; Romanov et al., 2012; Wang et al., 2014). Because neither cleavage nor truncation at the caspase site affected the reversal potential (Chiu et al., 2017), it can be expected that the selectivity of the caspase-cleaved channel did not change either compared with the voltage-activated WT Panx1 channel.The lack of effect on the reversal potential is particularly puzzling in the experiments involving concatemers of the Panx1 protein. The concatemers were designed to allow sequential removal of C-terminal tails in a hexameric channel. As expected, this resulted in a “quantal” activation of the channel, with each cleavage of a protomer leading to an incremental increase in single-channel conductance (Chiu et al., 2017). The data also show a continuous increase in current density in whole-cell recordings as more caspase sites within the concatemer were used. Yet, the ATP release was discontinuous, requiring removal of more than two C termini. Thus, there was selective (Cl?) current in the absence of ATP release. However, this change in selectivity was not reflected in the voltage ramps applied to the channels with variable numbers of cleaved C termini; all currents reversed at exactly the same potential. In these experiments, collecting periods of 4 to 8 h were used to determine ATP released into the supernatant, which are orders of magnitude longer than typically used in ATP release studies, attenuating the confidence level for interpretation of the measured values. During this time period, many events of cell division and apoptosis could have been the source of extracellular ATP. In contrast, ATP release by airway epithelial cells was detectable within seconds after a sudden osmotic stimulus, was reversible, and was attenuated by the Panx1 inhibitors CBX or probenecid (Ransford et al., 2009). Similarly, ATP release by Schwann cells induced by hypotonicity was detectable within 2 min after the stimulus, occurred in the absence of the cytoplasmic marker lactate dehydrogenase, and was blocked by the same Panx1 inhibitors (Wei et al., 2021). Furthermore, ATP release by oocytes expressing WT Panx1 or Panx1Δ378 was detectable in the unstirred supernatant as early as 5 min after a K+ stimulus (Wang and Dahl, 2018). Even in apoptotic cells, ATP release and that of metabolites was observed in a shorter time scale when a complex signaling chain was initiated by anti-Fas or UV treatment (Chekeni et al., 2010; Medina et al., 2020). Intriguingly, different sets of metabolites were released depending on the apoptosis stimulus. UV radiation and anti-Fas stimulation had only a small subset of metabolites, including ATP, in common, while the majority of metabolites released in response to the two stimuli did not overlap (Medina et al., 2020). This observation suggests that different signaling modalities modify the selectivity of the Panx1 channel or that other release mechanisms were involved concurrently.Both voltage- and caspase-activated channels have different unitary conductances at negative potentials (∼15 pS) as compared with positive potentials (∼90 pS; Ma et al., 2012; Romanov et al., 2012; Chiu et al., 2017). Since physiological ATP release occurs at the normal or slightly depolarized membrane potential, i.e., at negative potentials, the ∼15-pS conformation rather than the ∼90-pS channel would need to be responsible for ATP release. Permeability measurements with Panx1 truncated at the caspase cleavage site (Panx1Δ378) and involving extracellular ion replacement confirmed the Cl selectivity and lack of cation permeability like the WT Panx1 channel (Fig. 3, F and H; Wang and Dahl, 2018). However, in the presence of extracellular K+, Cl selectivity was attenuated (Fig. 3, G and H), while cation permeability and ATP permeability became apparent just as in WT Panx1 channels (Wang and Dahl, 2018).As observed for WT Panx1 channels, Panx1Δ378 channels did not release ATP under voltage-clamp conditions despite robust membrane currents (Wang and Dahl, 2018). Extracellular K+ not only boosted the membrane currents of Panx1Δ378 channels but also induced ATP release with the membrane potential clamped at −60 or 0 mV (Fig. 2 B; Wang and Dahl, 2018).Recently, data obtained with purified Panx1 protein in lipid bilayers and liposomes were interpreted in support of the view that caspase cleavage is sufficient to induce the ATP-permeable conformation of the Panx1 channel (Narahari et al., 2021). Indeed, the data show properties expected from Panx1 channels in the ATP-permeable conformation, including cation permeability and flux of ATP and of dyes with a preference of anionic over cationic dyes. However, the data also show that the caspase-activated Panx1 channel in lipid bilayers is distinct from the caspase-activated Panx1 channel in mammalian cells. In bilayers, three populations of channels were observed. One group of channels exhibited a single channel conductance of 100 pS, which is similar to the one observed in mammalian cells. The second group has been shown to dwell mainly in a 189-pS state. The third group did not require caspase cleavage and was observed only transiently after insertion of the protein into the lipid bilayer. The channels in this group exhibited variable conductance states with a maximal conductance of ∼400 pS.Another distinction of the Panx1 channel in lipid bilayers from the channel in mammalian cells is the lack of rectification. In mammalian cells, the caspase-cleaved Panx1 channel exhibits a 12-pS conductance at negative membrane potentials and a conductance of 96 pS at positive potentials (Chiu et al., 2017). In contrast, in lipid bilayers the voltage–current relationship was linear over a voltage range of −200 to +200 mV (Narahari et al., 2021). Thus, the disposition of the purified and caspase-cleaved Panx1 channel to allow the flux of ATP and other molecules is much more favorable in lipids than in mammalian cells. A channel with a conductance of ∼200 pS at the resting membrane potential or a depolarized potential consistent with physiological events is not the same as the 12-pS channel in mammalian cells. In the lipid bilayer, the channel is not only exposed to a different lipid environment, but it is also missing the protein environment prevailing in intact cells. It will be intriguing to see whether lipids or proteins prevent the transition from the 12-pS to the ∼200-pS channel conformation and how this is regulated in cells. Overall, the data obtained with purified Panx1 in an exclusive lipid environment support the hypothesis of more than one open conformation of the Panx1 channel.Thus, although caspase cleavage of Panx1 may not be sufficient for ATP release, this process does not interfere with other stimuli for the release. For example, in oocytes expressing Panx1 truncated at the caspase cleavage site, the channel was constitutively active, yielding membrane currents, which were boosted further by extracellular K+. Furthermore, ATP release via the truncated Panx1 was observed only with the additional K+ stimulus (Wang and Dahl, 2018). The ATP release observed in apoptotic/pyroptotic cells, therefore, may be the consequence of other stimuli. This release, however, plays a key role in the attraction of macrophages or microglia to injured or sick cells, a process working in vertebrate and invertebrate organisms (Chekeni et al., 2010; Samuels et al., 2010; Samuels et al., 2013). In the invertebrate nervous system, the ATP signal, rather than being a “find me” signal (Chekeni et al., 2010), appears to serve as a mobilization signal for microglial cells (Duan et al., 2009; Dahl and Muller, 2014).What is the oligomeric state of the Panx1 channel?Until recently, it was held that Panx1 oligomerizes to form a homomeric hexameric channel (Penuela et al., 2013; Chiu et al., 2018). This conclusion was based on cross-linking studies (Boassa et al., 2007; Chiu et al., 2017; Epp et al., 2019) and EM with negative staining of the isolated protein (Wang et al., 2014; Chiu et al., 2017). Acceptance of these findings was facilitated because a hexameric oligomeric state of Panx1 matched that of connexin “hemichannels” (the half of gap junction channels in one of the two opposing plasma membranes). Cross-linking studies in general suffer from both under- and overestimation of protein interactions. If the cross-linking agent is applied in very low concentrations and/or for an insufficiently long period, native interactions may not be captured fully, and using high concentrations for excessive periods of time can result in fortuitous linkages, resulting in overestimation of the oligomeric state. Interestingly, in one of the cross-linking studies on Panx1 (Epp et al., 2019), monomeric Panx1 predominated, followed by intermediate oligomers (dimers and trimers), while the apparent hexameric form was barely detectable. In this setting, heptamers, even if present, would have escaped detection because of an insufficient amount of protein in that oligomeric state. These observations, although consistent with incomplete cross-linking, also could indicate that the majority of Panx1 molecules are not assembled to a functional channel. Instead, the intermediate oligomers may only acutely form a functional channel upon a still-mysterious stimulus. Given the existence of dimers and trimers, the functional channel may be a hexamer (2 + 2 + 2 or 3 + 3), but also could be a heptamer (2 + 2 + 3) or higher oligomer.While the visualization of single-particle Panx1 channels by negative stain appears to support the hexameric state of the channel, the imposed sixfold symmetry used in both studies (Wang et al., 2014; Chiu et al., 2017) could have biased the outcome. However, the support for the hexameric arrangement of the Panx1 channel culminated in the use of concatemers of Panx1 forcing the formation of hexamers. Not only were these concatemers shown to form functional channels, but these channels exhibited biophysical properties essentially identical to those of WT Panx1 channels (Chiu et al., 2017). Thus, the channel formed by WT Panx1 ought to have the same oligomeric state as the one formed by the concatemer. It is conceivable that instead of forming the expected hexamer, one protomer of a concatemer would incorporate into the channel to form a heptamer with other concatemers. However, Chiu et al. (2017) emphasized the hexameric state with Western blots and the exact number of quantal steps predicted for a hexamer. However, like Panx1, CALHM1 was originally thought to adopt a hexameric configuration based on concatemers (Siebert et al., 2013) until the cryo-EM structure revealed an octameric configuration (Syrjanen et al., 2020).The oligomeric state of the Panx1 channel has been challenged in six publications, all appearing within the first half of 2020 and using cryo-EM to image the channel (Deng et al., 2020; Jin et al., 2020; Michalski et al., 2020; Mou et al., 2020; Qu et al., 2020; Ruan et al., 2020). Instead of the hexamer hitherto believed to represent the native state of the Panx1 channel, the cryo-EM data show unequivocally a homo-heptamer forming the Panx1 channel. However, it would be desirable to have the heptameric arrangement be supported by independent methods. So far, the heptameric assembly supports the closed and the small, Cl-selective conformations of the Panx1 channel, but not the large-pore conformation. It needs to be seen whether the ATP-permeable Panx1 channel has the same heptameric arrangement or whether higher-order oligomers need to form to render the channel ATP permeable.Do cryo-EM structures support an ATP release function of Panx1?In the Panx1 structure publications (Deng et al., 2020; Jin et al., 2020; Michalski et al., 2020; Mou et al., 2020; Qu et al., 2020; Ruan et al., 2020), the reported narrowest width of the pore, or narrowest exclusion width of the closed conformation for the Panx1 structures, is remarkably uniform—8–9.4 Å. It is worth noting that the extracellular constriction has the same dimensions for all the reported structures, irrespective of whether the C or N terminus is cleaved off (Fig. 1 C shows three overlaid structures). For Panx1 to function as an ATP release channel, the pore must be at least the size of an ATP molecule with an effective hydrated diameter of ATP of ∼12 Å (Sabirov and Okada, 2004) and a hydrodynamic diameter of ATP of 15.4 Å (Rostovtseva and Bezrukov, 1998). Indeed, a surface model of ATP is a poor fit in the cryo-EM map (Fig. 4 A). As seen in a side view of the pore, the ATP molecule resides below the external constriction. The diameter of ATP (mol wt 507) exceeds that of the constriction, making a passage unlikely. As shown in Table 1, other ATP release channels exhibit a considerably wider pore. The situation for the tracer molecule Yo-Pro (mol wt 375) appears to be somewhat more favorable (Fig. 4 B). It is conceivable that, despite the tight fit, thermal fluctuations of the pore and the tracer molecule could enable the Yo-Pro molecule to squeeze through the pore at a low rate.Open in a separate windowFigure 4.Space-filling model of Panx1 without the C terminus. Related to PDB accession no. 6wbg. (A) Top: Top view of the extracellular pore. The space-filling model of ATP (PubChem accession no. 5957, mol wt 507.2) is shown in cyan. Bottom: Side view of the extracellular pore showing the W 74 ring above the ATP. (B) Top: The same model as in A, top view of the extracellular pore. In magenta is the space-filling model of Yo-Pro-1 (PubChem accession no. 6913121, mol wt 375.5). Bottom: Side views of the extracellular pore with a model for Yo-Pro (magenta) placed at the level of the external constriction.For reference, the minimal pore diameter of other ATP release channels is ≥14 Å (Table 1). The CALHM 1 channel, for example, is an octamer with a pore diameter of 19.5 Å at its narrowest part. Interestingly, CALHM 1 originally also was considered to be a hexamer (Ma et al., 2016), but cryo-EM data show that the channels formed by various CALHM isoforms are all higher-order oligomers, including octamers, decamers, and undecamers (Choi et al., 2019; Drożdżyk et al., 2020; Syrjanen et al., 2020). CALHM 4 even has been shown to exhibit two oligomeric states with different pore diameters of 20 and 30 Å (Drożdżyk et al., 2020).The heptameric Panx1 channel with an 8–9.4-Å constriction is a poor candidate to represent a channel conformation with high ATP permeability. It does, however, provide a sound structural basis for a Cl-selective channel, as it is observed when the channel is activated by voltage or caspase cleavage (Ma et al., 2012; Romanov et al., 2012; Chiu et al., 2014; Wang and Dahl, 2018; Michalski et al., 2020). The external restriction has a width to accommodate Cl to pass through the channel when it is unobstructed by other channel components. Furthermore, the presence of positively charged amino acids could be the basis for anion selectivity. Indeed, mutations of R75 support this notion. The anion selectivity was preserved if R75 was changed to K, while a change to A not only diminished the Cl selectivity but also promoted cation permeability (Michalski et al., 2020). Because overlay of the structures of WT Panx1 and the truncated version mimicking the caspase-cleaved Panx1 did not reveal changes in the external constriction (Mou et al., 2020), it can be expected that both exhibit the same exclusion limits and Cl selectivity (Fig. 1).Taken at face value, the presently available Panx1 channel structures do not support an ATP-release function of Panx1. Abundant functional data, on the other hand, are consistent with such a function. Therefore, one may consider that Panx1 exhibits an additional conformation not captured in the cryo-EM studies so far. How could a structure of the ATP-permeable Panx1 conformation be obtained? To boost the chance to observe the large-pore conformation of the Panx1 channel, it may be advantageous to activate Panx1 through ligand binding to G protein–coupled receptors, known to activate Panx1 in the process of secondary ATP release (Locovei et al., 2006b; Billaud et al., 2012; Murali et al., 2014; Dahl, 2018). For example, there is ample evidence that activation of Panx1 through P2X7 and NMDA receptors involves the action of the Src gene product on Panx1, likely by tyrosine phosphorylation of the Panx1 protein itself (Weilinger et al., 2012; DeLalio et al., 2019; Lohman et al., 2019).To test whether Panx1 phosphorylation per se is sufficient to drive the Panx1 channel to the open conformation, a Panx1 channel in an excised patch could be exposed to the active form of Src in the presence of ATP. If successful, the cryo-EM sample could be treated likewise to enhance the chance to capture the ATP-permeable conformation.Other soluble potential activators of Panx1 include K+ and Ca2+. Although micromolar concentrations of Ca2+ were observed to activate large channels in inside-out membrane patches excised from oocytes expressing Panx1 (Locovei et al., 2006b), activation of Panx1 by Ca2+ was not observed in HEK cells (Ma et al., 2009). In contrast, secondary ATP release in carotid body type II cells was found to be attenuated by the Ca2+ chelator BAPTA (1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid; Murali et al., 2014), indicating Ca2+ involvement in Panx1 activation. Furthermore, it has been proposed that the activation of Panx1 by the mechanosensitive Piezo channels involves cytoplasmic [Ca2+] (Cinar et al., 2015; Wang et al., 2016; Diem et al., 2020). Although K+ and Ca2+ failed to alter the cryo-EM structure (Ruan et al., 2020), it is not clear whether all possible iterations of the experimental approach have been exhausted.Alternatively, the Panx1-P2X7 complex is particularly interesting because these proteins interact physically and functionally (Pelegrin and Surprenant, 2006; Locovei et al., 2007; Iglesias et al., 2008; Silverman et al., 2009). Cryo-EM of the complex of Panx1 and P2X7 with and without ATP not only may allow the capture of the large pore conformation but may also yield clues about the gating mechanism. In both approaches, the ATP concentration has to be carefully chosen, since high concentrations of ATP were found to inhibit Panx1-mediated currents (Qiu and Dahl, 2009; Qiu et al., 2012).Contribution of the C terminus to the pore structureNone of the cryo-EM structures published so far has resolved the C terminus. Only one paper shows an amorphous mass at the cytoplasmic entry to the channel (Ruan et al., 2020). It is noteworthy that all cryo-EM structures presently available were obtained with C terminus–tagged Panx1. It is conceivable that the C terminus was misfolded because of the tags and consequently escaped resolution. Resolution of the C terminus is important, since functional data strongly indicate a decisive role of the C terminus in channel function, in terms of both gating and permeability (Chekeni et al., 2010; Chiu et al., 2014; Jackson et al., 2014). In the voltage-activated and Cl-selective channel, the terminal cysteine moiety is reactive to externally applied thiol reagents, as indicated by attenuation of Panx1-mediated membrane currents (Wang and Dahl, 2010). Furthermore, an analysis of pore-lining moieties by the substituted cysteine accessibility method showed that cysteine replacements of several amino acids in the C terminus exhibited reactivity to thiol reagents (Wang and Dahl, 2010), indicating that this terminus protrudes from the cytoplasmic site far into the channel permeation pathway. Further support for this conclusion comes from the observation that the endogenous C-terminal cysteine C426 interacted with Panx1F54C via a disulfide bond (Sandilos et al., 2012). The F54C position was identified as a pore lining moiety together with other positions in the first TM segment and the first extracellular loop of Panx1 in the substituted cysteine accessibility method analysis (Wang and Dahl, 2010). F54 is located in the proximity of I58 (Fig. 1 B), which is part of a second constriction (13 Å in diameter) of the permeation pathway (Deng et al., 2020). This observation suggests that the WT Panx1 cysteine C426 would be located a few angstroms from the extracellular surface in the voltage-activated channel.Given that a large stretch of C-terminal amino acids fills the pore from the intracellular vestibule up to close to the extracellular restriction, one would expect that removal of the C terminus by caspase cleavage should affect the unitary conductance of the Panx1 channel to some extent. However, the WT Panx1 and truncated or caspase-cleaved Panx1 channels exhibit similar unitary conductances (Ma et al., 2012; Romanov et al., 2012; Chiu et al., 2014; Wang et al., 2014; Chiu et al., 2017)Removal of the C terminus by caspase cleavage in apoptotic cells or genetically by placing a stop codon at the cleavage site renders the channel constitutively active, with sojourns to open and closed states (Chekeni et al., 2010; Chiu et al., 2014; Chiu et al., 2017). However, since nonapoptotic ATP release is reversible in most cells, cleavage cannot be an obligate step to transform the Cl-selective and cation-impermeant conformation to the large-pore conformation, which allows the flux of ATP and is also cation permeant. Thus, for nonapoptotic ATP release, instead of being cleaved, the C terminus has to undergo a major rearrangement to move out of the permeation pathway for adopting the ATP-permeable large-pore conformation. Indeed, in the large-pore conformation, the terminal cysteine was not further reactive to externally applied thiol reagents, while engineered cysteines at the external pore entry remained reactive (Wang et al., 2014). Energetically, the most favorable scenario would be a swinging out of the intact C terminus into a gap between TM helices created by a gating mechanism for the physiological, i.e., reversible activation of the large pore conformation.An overlay of truncated and WT Panx1 cryo-EM structures did not reveal differences in the extracellular pore structure (Fig. 1; Mou et al., 2020). Thus, cleavage of the C terminus of Panx1 may not be sufficient to alter the pore conformation, which is consistent with identical selectivities of channels formed by WT and truncated Panx1. Consequently, the structure of the truncated Panx1 channel cannot account for ATP permeability of the full-length channel.Ruan et al. (2020) suggested an intriguing hypothesis for the existence of different selectivities of the Panx1 channel. According to this hypothesis, the pathway for selective Cl flux involves a side tunnel, while the main pore is blocked by the C terminus residing in the intracellular vestibule. Movement or cleavage of the C terminus then would provide the ATP-permeable large-pore conformation. However, the functional data listed above, including thiol reaction of the terminal cysteine (Wang and Dahl, 2010) and cross-linking of it with an engineered cysteine in a position close to the external constriction (Sandilos et al., 2012), indicate that the C terminus extends deep into the channel pore and thus are not consistent with a role of a side tunnel. Furthermore, the reported restriction at the extracellular end of the pore still would exclude ATP molecules.At present, the available information about the C terminus is restricted to functional data, since none of the published cryo-EM structures has this part of the protein resolved. The functional observations listed above suggest that the C terminus reaches from the intracellular site deep into the channel pore, probably occluding it in the closed state (Fig. 5 A). It can be hypothesized that at positive potential, the C terminus is rearranged (Fig. 5 B) so that the C-terminal cysteine moiety can react with extracellularly applied thiol reagents. The rearrangement of the C terminus allows mainly Cl to pass through the external restriction depicted in the cryo-EM structures, which imposes Cl selectivity. Alternatively, the C terminus is cleaved by caspase with the same consequences, plus the channel becoming constitutively active at negative and positive potentials (Fig. 5 C). Physiological stimuli, such as mechanical stress, directly or via Piezo (Cinar et al., 2015; Diem et al., 2020), initiate a gating mechanism, which widens the external constriction to accommodate the passage of ATP and render the channel nonselective for ions. In addition, the C terminus undergoes a conformational change, so that the terminal cysteine becomes inaccessible to thiol reagents (or a reaction does not affect membrane currents in this conformation). Extracellular K+ can mimic the effect of the physiological stimuli (Fig. 5 D). Cleavage of the C terminus then solidifies the apoptotic fate of the cell (Fig. 5 E). Alternatively, the caspase-cleaved Panx1 channel can be “hyperactivated” to render the Cl-selective conformation to the nonselective large-pore conformation (Fig. 5, C–E).Open in a separate windowFigure 5.Activation mechanisms of Panx1 channels. (A) Based on the presently available structural and functional data, we hypothesize that the closed channel exhibits the external restriction of ∼9 Å and that the C termini reach deep into the pore, occluding it. (B) Depolarization dislocates the C termini, so that the channel pore diameter equals or exceeds that of the external restriction and exposes the terminal cysteine (yellow C) to thiol reagents. The yellow dot indicates the position of Panx1F54 which, when mutated to a cysteine, can be disulfide bonded with the terminal cysteine (Sandilos et al., 2012). In this configuration, the external restriction limits entry by size and selects for chloride (orange dot) over other ions. (C) Cleavage of Panx1 at position 378 removes the “pore gate,” rendering the channel constitutively active. However, as indicated by the published cryo-EM data, caspase cleavage does not affect the external constriction, leaving a chloride-selective channel. (D) Various physiological stimuli initiate a gating mechanism at the external restriction, widening it to accept ATP (green dot) and other molecules in this size range. (E) This mechanism also may move charged amino acids within the external constriction, with the consequence that charge selectivity is attenuated. In the case of K+ stimulation, the terminal cysteine is no further reactive to thiol reagents, while engineered cysteines at the external end of the pore still are. Whether this terminal cysteine concealment also applies to the physiological stimuli remains to be determined. Cleavage of the C terminus by caspase “super stimulates” the channel and irreversibly seals the apoptotic fate of the cell.In theory, an alternative mechanism to switch the Cl-selective channel to an ATP-permeable conformation would be a change of the oligomeric state of the channel. Such a mechanism may operate in CALHM 4 channels, for which cryo-EM data suggest the coexistence of decameric and undecameric channels, which exhibit different pore diameters (Drożdżyk et al., 2020). Furthermore, other large-pore channels such as gasdermin appear to have variable oligomeric states (Mulvihill et al., 2018). No experimental evidence for such a mechanism operating for Panx1 channels is presently in existence, though.Problems to be solvedThe cryo-EM structures of the Panx1 channel published in rapid succession have raised a number of crucial questions. While they are consistent with the chloride-selective conformation of Panx1, they do not support an ATP-release function of this channel despite the abundance of evidence for such a role from functional data. The following points need to be addressed in future studies.Although not likely, it is possible that the heptameric organization of the Panx1 channel could be a consequence of the procedures involved in cryo-EM. It would be desirable to have this oligomeric state verified by independent methods.The published cryo-EM structures exhibit a restriction with a diameter of ∼9 Å, which would allow the passage of Cl but would be too small for the permeation of ATP. Considering, that innexin channels (innexons) are octamers and CALHM channels can be octamers, decamers, or undecamers (Choi et al., 2019; Drożdżyk et al., 2020; Syrjanen et al., 2020), one should not a priori rule out that the large-pore conformation is a higher-order oligomer.For structural support of the ATP-release function of Panx1, a large-pore conformation is required with a pore dimension fitting the effective hydrated diameter of ATP of ∼12 Å (Sabirov and Okada, 2004) or a hydrodynamic diameter of ∼15.4 Å (Rostovtseva and Bezrukov, 1998). The structures published so far fall short of that.Further support for the ATP-release function of Panx1 could come from cryo-EM structures of Panx1 in the presence of ATP. ATP molecules located in a properly sized pore would represent the strongest evidence for the ATP-release function of this protein. Furthermore, extracellular ATP triggers a negative feedback mechanism inhibiting the channel (Qiu and Dahl, 2009; Qiu et al., 2012). This process involves the critical extracellular loop amino acids R75 and W74. Thus, ATP may be found docked to the pore entry.Because functional data strongly suggest that the C terminus of Panx1 can protrude deep into the channel pore, any functional conclusions from cryo-EM structures may be premature until the detailed C-terminal structure is resolved.Because several receptors associate (at least functionally) with Panx1 to induce secondary ATP release, a cryo-EM structure of receptor-Panx1 complexes may not only reveal the type of physical interaction but may also enhance the chance of capturing the large-pore conformation by including the respective receptor ligands.The puzzling observation that the forced hexameric channel has the same biophysical properties as the presumably heptameric WT channel needs a plausible explanation.The channels formed by WT Panx1 and by caspase-cleaved or truncated Panx1 have been reported to have the same unitary conductances. Why does removal of a “pore plug” (the C terminus) not affect unitary conductance of the channel?It is accepted that the voltage-activated Panx1 channel has a unitary conductance of 50–90 pS at positive potentials (Ma et al., 2012; Romanov et al., 2012; Wang et al., 2014; Chiu et al., 2018). However, what is the “unitary” conductance of the Panx1 channel in its large-pore conformation? Several laboratories have reported a ∼500-pS conductance for this channel conformation (Bao et al., 2004; Thompson et al., 2006; Kienitz et al., 2011; Orellana et al., 2011; Kurtenbach et al., 2013; Wang et al., 2014). Typically, in most recordings this large unitary conductance is a rare occurrence since the channel mainly dwells in one of the several subconductance states for extended periods of time. Thus, the term unitary can be misleading, and “maximal” conductance describes the situation better. However, a conductance of <100 pS at high positive membrane potentials and ∼15 pS at negative potentials has been proposed to be correlated with ATP release (Chiu et al., 2017). It is unclear, however, whether this lower conductance is actually a stable subconductance state or a record of the chloride-selective conformation. The latter is suggested by the similarity to the exclusively voltage-activated channel (Ma et al., 2012; Romanov et al., 2012; Wang et al., 2014). Reconstitution of the Panx1 protein into lipid bilayers yielded channel conductances of ∼1,000 pS (Mou et al., 2020). Whether this value is truly a single-channel conductance or whether the bilayer contained at least two active 500-pS channels remains to be determined.In summary, the structural models from cryo-EM are consistent with each other and with electrophysiological recordings that describe a Panx1 channel that is selectively permeable to Cl. However, the models’ small predicted pore sizes are insufficient to explain the considerable body of data showing that Panx1 is permeable to ATP and cations both physiologically and in apoptotic cells. Our previous experiments have suggested that Panx1 may work in more than one conformation, and this Viewpoint proposes ways to find a structural evidence for this hypothesis.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号