Abstract
Membrane proteins are challenging targets to functionally and structurally characterize. An enduring bottleneck in their study is the reliable production of sufficient yields of stable protein. Here, we evaluate all eukaryotic membrane protein production experiments that have supported the deposition of a high-resolution structure. We focused on the most common yeast host systems, Saccharomyces cerevisiae and Pichia pastoris. The first high-resolution structure of a membrane protein produced in yeast was described in 1999 and today there are 186 structures of α-helical membrane proteins, representing 101 unique proteins from 37 families. Homologous and heterologous production are equally common in S. cerevisiae, while heterologous production dominates in P. pastoris, especially of human proteins, which represent about one-third of the total. Investigating protein engineering approaches (78 proteins from seven families) demonstrated that the majority contained a polyhistidine tag for purification, typically at the C-terminus of the protein. Codon optimization and truncation of hydrophilic extensions were also common approaches to improve yields. We conclude that yeast remains a useful production host for the study of α-helical membrane proteins.
Keywords: membrane protein production, Saccharomyces cerevisiae, Pichia pastoris
Although the majority of eukaryotic MEMBRANE PROTEIN structures are DERIVED FROM PROTEINS produced in HEK293 and insect cells, the authors show here the importance of yeast as a production host and its role as an essential player in the production of eukaryotic membrane proteins for structural and functional analysis.
Introduction
From 1998 to 2002, two of us had the privilege of working in Stefan Hohmann’s laboratory as one of his PhD students (Kristina Hedfalk) and one of his postdoctoral research fellows (Roslyn Bill). Work under his guidance on the yeast aquaporin Fps1, resulted in the first publication (Bill et al. 2001) of many together over the next few years (Tamas et al. 2003, Elbing et al. 2004, Hedfalk et al. 2004, Karlgren et al. 2004, 2005, Otterstedt et al. 2004, Henricsson et al. 2005, Pettersson et al. 2006). True to his generous and supportive nature, Stefan encouraged Roslyn to be the sole corresponding author on that first article and he encouraged us both to pursue our independent research interests. One of those was the development of yeast as a host for the production of recombinant membrane proteins (Bill 2001, Oberg et al. 2009, Bill et al. 2011, Oberg and Hedfalk 2013), which is a line of investigation that we have continued to collaborate upon (Nyblom et al. 2007, Oberg et al. 2011, Bill and Hedfalk 2021).
The successful production of recombinant proteins plays a critical role in academic and industrial research. A wide range of host organisms exist for this purpose, including several yeast species that have proved to be effective and cost-efficient for the production of soluble and membrane proteins. The production of recombinant membrane proteins has facilitated the elucidation of transport mechanisms, substrate binding, and the effect of new molecules as drug leads (Gulezian et al. 2021). In common with Escherichia coli, yeasts can grow on inexpensive chemically defined media and cultivation technologies are available that produce high biomass (and hence high protein) yields. In common with mammalian cells, yeasts are eukaryotes. The production of biopharmaceuticals in yeast is also approved for human use (Gerngross 2004, Wang et al. 2017). The low fidelity of carbohydrate modifications is a known disadvantage of using yeast, a property, i.e. most pronounced for Saccharomyces cerevisiae (Vieira Gomes et al. 2018). However, yeast strains with humanized glycosylation pathways have been reported (De Wachter et al. 2021).
Stefan catalyzed our shared love of yeast as an experimental tool and our enduring fascination with membrane proteins. In this review, in his memory, and together with some of our students, we examine the use of S. cerevisiae and Pichia pastoris for production of membrane proteins of eukaryotic origin, for subsequent structural and functional characterization. Thus, α-helical transmembrane proteins are the focus of this mini-review.
Yeast is an attractive host for resolving new membrane protein structures
From 1999 to 2022, 190 high-resolution α-helical membrane protein structures were derived from proteins produced in yeast. Of those, 100 proteins were extracted from S. cerevisiae, 86 from P. pastoris, and 4 from Schizosaccharomyces pombe (Ago et al. 2007, Wang et al. 2019a, Deng et al. 2021). In S. cerevisiae, 57% of the 100 proteins were homologous, either extracted from the native membrane or recombinantly produced, and 43% were produced by heterologous recombinant protein production (Table S1, Supporting Information). In P. pastoris, only 2% of the 86 proteins were from the homologous host while almost all (98%) had an origin other than yeast (Table S2, Supporting Information).
Figure 1 shows that the use of yeast as a recombinant host has become more common with time, a development that follows the overall upward trend of membrane protein production using any host system. The first structure derived from a protein produced in yeast was reported in 1999 when we were working in Stefan’s laboratory: ATP synthase extracted from its native source S. cerevisiae (Stock et al. 1999). The first protein structure of a protein produced in P. pastoris was reported in 2005: recombinant rat Kv1.2 voltage-gated potassium channel (Long et al. 2005). From 2009 onwards, there has been an upward trajectory in the deposition of unique structures derived from S. cerevisiae and P. pastoris (Fig. 1A). Notably, production of homologous and heterologous proteins has been equally popular in S. cerevisiae, while there is a clear dominance of production of heterologous proteins in P. pastoris. The exception is P. pastoris Aqy1, for which two structures have been reported (Fischer et al. 2009, Eriksson et al. 2013; Fig. 1B).
Figure 1.
High resolution membrane protein structures derived from proteins produced in yeast from 1999 to date; unique structures derived from yeast-produced proteins (homologous and heterologous) (A) presented as a percentage of all structures (derived from proteins produced in all hosts; gray dashed line) and (B) with dotted lines showing the linear regression of the use of P. pastoris (green) and S. cerevisiae (yellow), respectively.
For proteins produced in S. cerevisiae, there is an even distribution of structures resolved by crystallization and X-ray diffraction compared with Electron Microscopy (EM) analysis (45% and 55%, respectively). For P. pastoris, the majority of all structures result from X-ray analysis (74%) compared with EM analysis (24%). There is one NMR structure of a yeast-derived protein: human AQP1 produced in P. pastoris (Dingwell et al. 2019; Table S1 and S2, Supporting Information).
In the following sections, we review the typical approaches taken using S. cerevisiae and P. pastoris to facilitate these types of structural investigations.
Cloning and growth in yeast
Cloning
Good construct design is key to achieving high yields of functional recombinant membrane protein (Fig. 2). Codon optimization of the gene sequence for optimal expression, the selection of tags (and cleavage sites) and the truncation of the corresponding protein sequence are all considerations in designing an effective construct (Vieira Gomes et al. 2018). Most yeast plasmids can be manipulated in E. coli to facilitate the necessary molecular biology prior to transformation, which is why they are known as shuttle vectors (Fig. 2). Three main types of shuttle vectors are used: integrative, centromeric, or episomal plasmids, which differ in several aspects including integration (or not) into the genome and plasmid copy number (Gnugge and Rudolf 2017). The polyhistidine tag is by far the most commonly used in yeast expression experiments (Table S3, Supporting Information). Sometimes, several tags are used in tandem including a GFP tag, which is enables protein yields to be monitored through the measurement of GFP fluorescence (Drew et al. 2008). Tags are most often placed at the C-terminus rather than the N-terminus of the recombinant protein, and a cleavage site is typically inserted to enable tag removal, especially with longer polyhistidine-tags or when several tags are present which might alter protein function. TEV protease is most often used for tag cleavage due to its high site specificity, but thrombin is also very common (Kapust and Waugh 2000). In some cases, the protein of interest is truncated at a terminus, which can increase yields, presumably due to stabilization of the folded protein (Bill et al. 2011). It is common practice to remove flexible termini to improve homogeneity for crystal formation (Mooij et al. 2009). Codon usage is organism dependent meaning that codons that may be abundant in the source organism of the protein target are rare in yeast. As such, optimizing the gene sequence according to the codons present in the heterologous host can facilitate higher recombinant protein yields (Table S3, Supporting Information). While codon optimized genes adapted for a certain host are available commercially, a higher translation efficiency can be achieved by a more complex optimization approach, which also includes the condition under which heterologous genes are being expressed (Lanza et al. 2014). Codon optimization is widely used for membrane proteins produced in P. pastoris (Table S3, Supporting Information; see also Fig. 5 below).
Figure 2.
From construct design to cell fractionation in yeast. In S. cerevisiae, the gene of interest may be produced from a constitutive (e.g. PGK) or an inducible (e.g. GAL) promoter, while the very strong and methanol inducible AOX1 promoter is commonly used in P. pastoris. The plasmid for expression in S. cerevisiae is selected for using media lacking one essential amino acid or nucleobase and the gene for its production is instead part of the expression plasmid. In P. pastoris, the plasmid for expression is selected by adding Zeocin to the media, a selection, i.e. applied in both E. coli and P. pastoris. Linearization and integration into the genome after transformation is the more common procedure in P. pastoris, where high production clones are evaluated in total cell extracts using immunoblots and a suitable antibody. Independent of yeast species, it is important to verify production and correct membrane localization before scaled-up growth. The main difference between S. cerevisiae and P. pastoris is that the latter has a higher demand for oxygen, especially when grown on methanol, hence it is preferably cultured using more tightly controlled conditions in fermenters. For further purification of the protein of interest, cells are collected by centrifugation, broken, and the fraction of interest is collected by differential centrifugation.
Figure 5.
Protein engineering approaches for α-helical membrane protein family members heterologously produced in S. cerevisiae and P. pastoris, respectively. (A) Length of the polyhistidine-tag. (B) Location of the polyhistidine-tag. (C) Choice of protease cleavage site between the protein sequence and the tag.
Growth
In optimizing growth conditions, the type of promoter used (inducible or constitutive) needs to be considered as well as the influence of media components, the presence or absence of a stabilizing ligand, the pH and oxygen levels in the culture and its temperature (Liu et al. 2013). For yeast, a common procedure is to start with smaller cultures and upscale to larger cultures once evidence of expression has been found. For cultures with constitutive expression, there is a correlation between the specific growth rate and the produced protein (Liu et al. 2013). Fed-batch cultures are typically performed in a fermenter when protein expression is under the control of an inducible promoter (Fig. 2). Normally, S. cerevisiae will first be grown in Synthetic Defined (SD) medium, which contains glucose as the main source of carbon, but lacks the amino acid or nucleobase used to maintain selection of the episomal plasmid. Protein expression is commonly under the control of the GAL1 or GAL10 promoter in S. cerevisiae, which is repressed by the MIG1 repressor while glucose is present in the medium. This enables maximum cell growth (Kang et al. 2005) prior to the galactose addition that induces gene expression (Vieira Gomes et al. 2018). Pichia pastoris has a rapid growth rate and can be cultured to very high cell-densities in fermenters, which allows high protein yields. Expression in P. pastoris is typically under the control of strong and tightly regulated promoters (Fig. 2). The alcohol oxidase promoter (PAOX1) is induced by methanol, but repressed by glucose, glycerol, and ethanol. Constitutive promoters may also be used, such as the glyceraldehyde-3-phosphate promoter (GAP), which achieves comparable expression levels in presence of glucose as with AOX1 (Rabert et al. 2013, Ahmad et al. 2014). Strains for large-scale production are most commonly derived from P. pastoris CBS7435, as they are not limited by patent protection or materials ownership policies. Nevertheless, there are some modified strains that may be more useful depending on the protein of interest. These modifications include auxotrophy, protease-deficiency, or glyco-engineering (Ahmad et al. 2014).
Membrane protein production in S. cerevisiae
Saccharomyces cerevisiae is a useful host for a wide range of membrane proteins
Members of 26 α-helical membrane protein families have been produced in S. cerevisiae (Table 1).
Table 1.
Protein families produced in S. cerevisiae.
Protein family | % | References |
---|---|---|
Potassium, sodium, and protein ion-selective channels | 3 | Kintzer and Stroud (2016), Kintzer et al. (2018), Dickinson et al. (2022) |
Transient receptor potential channels | 7 | Huynh et al. (2016), Hughes et al. (2018, 2019), Dosey et al. (2019), Pumroy et al. (2019), Conde et al. (2021), Ahmed et al. (2022) |
Aquaporins and glyceroporin channels | 2 | Gotfryd et al. (2018), van den Berg et al. (2021) |
Other ion channels | 1 | Wang et al. (2019b) |
SWEET and semisweet transporters | 2 | Brauer et al. (2019), Gerondopoulos et al. (2021) |
AAA-ATPase membrane translocators | 1 | Kater et al. (2020) |
Intramembrane proteases | 1 | Pryor et al. (2013) |
ABC-transporters | 3 | Kodan et al. (2019), Bickers et al. (2021), Harris et al. (2021) |
Solute carrier transporters | 4 | Coudray et al. (2017), Parker and Newstead (2017), Parker et al. (2019), Tsai et al. (2020) |
Oligosaccharyltransferases | 3 | Bai et al. (2018), Wild et al. (2018), Neuhaus et al. (2021) |
Protein O-mannosyl transferases | 1 | Bai et al. (2019a) |
Oxidases | 2 | Ma et al. (2004), Son et al. (2008) |
Antiporters | 4 | Ruprecht et al. (2014, 2019), Winklemann et al. (2020), Matsuoka et al. (2022) |
P-type ATPases | 10 | Clausen et al. (2013), Bai et al. (2019b, 2020a, 2021), Timcenko et al. (2019, 2021), Geurts et al. (2020), McKenna et al. (2020), Li et al. (2021a), Zhao et al. (2021) |
Permease channels | 1 | van den Berg et al. (2016) |
Major facilitator superfamily transporters | 7 | Pedersen et al. (2013), Parker and Newstead (2014), Kapoor et al. (2016), Paulsen et al. (2019), Qureshi et al. (2020), Bavnhoj et al. (2021), Custodio et al. (2021) |
Membrane-integral pyrophosphatases | 5 | Kellosalo et al. (2012), Lin et al. (2012), Li et al. (2016), Tsai et al. (2019) |
Transmembrane protein 16 family proteins | 4 | Brunner et al. (2014), Falzone et al. (2019), Kalienkova et al. (2019), Khelashvili et al. (2019) |
Nucleobase-cation-symport-2 family proteins | 1 | Alguel et al. (2016) |
Sec, translocase, and insertase proteins | 11 | Schoebel et al. (2017), Itskanov and Park (2019), Wu et al. (2019, 2020), Bai et al. (2020b), Matoba et al. (2020), Miller-Vedam et al. (2020), Itskanov et al. (2021), Weng et al. (2021), Zhang et al. (2021) |
Cation antiporter family | 1 | Waight et al. (2013) |
Sterol-sensing domain proteins | 1 | Winkler et al. (2019) |
Vacuolar ATPase | 7 | Oot et al. (2012, 2016), Mazhab-Jafari et al. (2016), Roh et al. (2018, 2020), Vasanthakumar et al. (2019) |
F-type ATPase | 9 | Stock et al. (1999), Kabaleeswaran et al. (2006), Dautant et al. (2010), Symersky et al. (2012a,b), Robinson et al. (2013), Guo et al. (2017), Srivastava et al. (2018), Luo et al. (2020) |
Electron transport chain complexes | 5 | Hunte et al. (2000), Lange et al. (2001), Palsdottir et al. (2003), Lancaster et al. (2007), Solmaz and Hunte (2008) |
Electron transport chain supercomplexes | 4 | Hartley et al. (2019, 2020), Rathore et al. (2019), Berndtsson et al. (2020) |
Overall, diverse membrane proteins have been produced in S. cerevisiae with Sec, Translocase, and Insertase proteins; P-type ATPases and F-type ATPase being most common (Fig. 3A).
Figure 3.
Production of α-helical membrane proteins in S. cerevisiae. (A) Membrane protein families for which high-resolution structures have been derived from protein produced in S. cerevisiae. (B) The origin of recombinant proteins for which a high-resolution structure has been derived from S. cerevisiae membranes.
For some protein families, like AAA-ATPase membrane translocators (Kater et al. 2020), intramembrane proteases (Pryor et al. 2013), oligosaccharyltransferases (Bai et al. 2018, Wild et al. 2018, Neuhaus et al. 2021), protein O-mannosyl transferases (Bai et al. 2019a), cation antiporter family (Waight et al. 2013), sterol-sensing domain proteins (Winkler et al. 2019), vacuolar ATPase (Oot et al. 2012, 2016, Mazhab-Jafari et al. 2016, Roh et al. 2018, 2020, Vasanthakumar et al. 2019), F-type ATPase (Stock et al. 1999, Kabaleeswaran et al. 2006, Dautant et al. 2010, Symersky et al. 2012a, 2012b, Robinson et al. 2013, Guo et al. 2017, Srivastava et al. 2018, Luo et al. 2020), electron transport chain complexes (Hunte et al. 2000, Lange et al. 2001, Palsdottir et al. 2003, Lancaster et al. 2007, Solmaz and Hunte 2008), and electron transport chain supercomplexes (Hartley et al. 2019, 2020, Rathore et al. 2019, Berndtsson et al. 2020), only homologous production (native or recombinant) has been used with many structures reported over several years. There are also some protein families that are only produced recombinantly using S. cerevisiae as a heterologous host; potassium, sodium, and protein ion-selective channels (Kintzer and Stroud 2016, Kintzer et al. 2018, Dickinson et al. 2022), aquaporins and glyceroporin channels (Gotfryd et al. 2018, van den Berg et al. 2021), other ion channels (Wang et al. 2019a), SWEET and semisweet transporters (Brauer et al. 2019, Gerondopoulos et al. 2021), oxidases (Ma et al. 2004, Son et al. 2008), permease channels (van den Berg et al. 2021), major facilitator superfamily transporters (Pedersen et al. 2013, Parker and Newstead 2014, Kapoor et al. 2016, Paulsen et al. 2019, Qureshi et al. 2020, Bavnhoj et al. 2021, Custodio et al. 2021), membrane-integral pyrophosphatases (Kellosalo et al. 2012, Lin et al. 2012, Li et al. 2016, Tsai et al. 2019), and transmembrane protein 16 family proteins (Brunner et al. 2014, Falzone et al. 2019, Kalienkova et al. 2019, Khelashvili et al. 2019; Table S1, Supporting Information). Overall, 58% of all α-helical membrane proteins produced in S. cerevisiae rely on homologous production (Fig. 3B).
Saccharomyces cerevisiae is commonly used for homologous production of native yeast proteins
Proteins from 11 different sources have been recombinantly produced in S. cerevisiae (Fig. 3B). A total of three bacterial, one parasite, and 11 plant and algal proteins have been produced. Mammalian targets together represent 19 proteins of which human proteins are represented by six targets. Fungi constitute 11 targets with the majority of targets being from yeast (58 proteins; Fig. 3B). Of these 58 proteins, 57 originate from S. cerevisiae and one from S. pombe (Matoba et al. 2020; Table S1, Supporting Information). Overall, this means that the majority of α-helical membrane protein structures from protein produced in S. cerevisiae are from the homologous expression (Table 3).
Table 3.
Comparison of the two most widely used yeast hosts for production of α-helical membrane proteins; S. cerevisiae and P. pastoris, respectively.
Structures | Homologous | Heterologous | Heterologous/homologous | Unique proteins | Unique proteins/structures | Protein families | Average number of proteins from each family | Protein families/total structures | Sources | Sources/total structures | |
---|---|---|---|---|---|---|---|---|---|---|---|
S. cerevisiae | 100 | 57 | 43 | 0.75 | 52 | 0.52 | 26 | 3.8 | 0.26 | 11 | 0.11 |
P. pastoris | 86 | 2 | 84 | 42 | 48 | 0.56 | 20 | 4.3 | 0.23 | 13 | 0.15 |
Membrane protein production in P. pastoris
Pichia pastoris is a successful host for synthesizing channel proteins
Proteins produced in P. pastoris represent 20 protein families (Table 2).
Table 2.
Protein families produced in P. pastoris.
Protein family | % | References |
---|---|---|
Potassium, sodium, and protein ion-selective channels | 33 | Long et al. (2005, 2007), Tao et al. (2009, 2010), Chen et al. (2010), Hansen et al. (2011), Whorton and MacKinnon (2011, 2013), Brohawn et al. (2012, 2013, 2014, 2019), Miller and Long (2012), Lolicato et al. (2014, 2017, 2020), Guo et al. (2016, 2017), Lee et al. (2016a), Yang et al. (2016), Pau et al. (2017), Matthies et al. (2018), Geng et al. (2020), Li et al. (2020), Niu et al. (2020), Pope et al. (2020), Zangerl-Plessl et al. (2020), Rietmeijer et al. (2021) |
Transient receptor potential channels | 2 | Deng et al. (2018, 2020a) |
Aquaporins and glyceroporin channels | 14 | Tornroth-Horsefield et al. (2006), Horsefield et al. (2008), Fischer et al. (2009), Ho et al. (2009), Nyblom et al. (2009), Eriksson et al. (2013), Frick et al. (2014), Kirscht et al. (2016), Dingwell et al. (2019), Lieske et al. (2019), de Mare et al. (2020), Wang et al. (2020) |
Other ion channels | 8 | Dickson et al. (2014), Vaisey et al. (2016), Miller et al. (2019), Ren et al. (2019), Deng et al. (2020b, 2021) |
SWEET and semisweet transporters | 1 | Tao et al. (2015) |
AAA-ATPase membrane translocators | 1 | Tang et al. (2020) |
Intramembrane proteases | 1 | Liu et al. (2020) |
ABC-transporters | 12 | Aller et al. (2009), Jin et al. (2012), Kodan et al. (2014), Li et al. (2014), Szewczyk et al. (2015), Lee et al. (2016b), Nicklisch et al. (2016), Oldham et al. (2016), Le et al. (2020), Barbieri et al. (2021) |
Solute carrier transporters | 5 | Garaeva et al. (2018, 2019), Ahuja and Whorton (2019), Garibsingh et al. (2021) |
Novel membrane proteins | 1 | Liu et al. (2020) |
Tetraspanins | 1 | Yang et al. (2020) |
Rhodopsins | 1 | Li et al. (2021b) |
Mechanosensitive channels | 2 | Maity et al. (2019), Deng et al. (2020c) |
Calcium ion-selective channels | 5 | Hou et al. (2012, 2018, 2020), Baradaran et al. (2018) |
G-protein coupled receptors | 3 | Shimamura et al. (2011), Hino et al. (2012), White et al. (2018) |
S-acyltransferases | 1 | Rana et al. (2018) |
Methyltrantransferases | 1 | Diver et al. (2018) |
Multidrug efflux transporters | 1 | Tanaka et al. (2017) |
Membrane-associated proteins in eicosanoid and gluthathione metabolism | 3 | Molina et al. (2007), Niegowski et al. (2014), Thulasingam et al. (2021) |
Oxidoreductases | 2 | Liu et al. (2021) |
Overall, several membrane protein families are represented, although there are slightly fewer than those produced in S. cerevisiae (Table 3). For P. pastoris, there is a clear dominance of certain targets where potassium, sodium, and protein ion-selective channels constitute one-third of all proteins (Fig. 4A). In comparison to S. cerevisiae, most structures from P. pastoris-derived protein are from recombinant production, the endogenous aquaporin Aqy1 being the only exception. Overall, P. pastoris is a suitable host system for recombinant production of α-helical membrane proteins coming from a wide range of protein families.
Figure 4.
Production of α-helical membrane proteins in P. pastoris. (A) Protein families successfully produced in the yeast P. pastoris with a resulting high-resolution structure. (B) Origin of proteins for which a high-resolution structure has been derived after extraction from the P. pastoris membrane.
Pichia pastoris is a successful host for recombinant production of human proteins
Proteins from 13 different eukaryotic sources have been produced in P. pastoris, which is slightly less than for S. cerevisiae (Table 3). Human membrane proteins represent more than one-third of all proteins produced in P. pastoris (Fig. 4B) and are represented in more than half of all protein families (Table S2, Supporting Information). Taken together, P. pastoris is a suitable host for α-helical proteins from different sources, especially of human origin. Indeed, for the yeast P. pastoris the ratio between heterologous and homologous production is 42, as compared to 0.75 in S. cerevisiae (Table 3).
Discussion and conclusion
The use of yeasts as expression hosts for α-helical membrane proteins is increasing year-on-year (Fig. 1). Saccharomyces cerevisiae, but not P. pastoris, is widely used for the extraction of homologous proteins. Both systems are used to produce a wide variety of recombinant membrane proteins and the number of unique proteins produced in each host are comparable (Table 3). In comparison, slightly more protein families have been produced in S. cerevisiae compared to P. pastoris (Tables S1 and S2, Supporting Information). The number of sources for the proteins produced are also similar between the two systems.
A total of 78 proteins from seven protein families have been heterologously produced in S. cerevisiae and P. pastoris (Table S3, Supporting Information). Of these, 16 (21%) were produced in S. cerevisiae and 62 (79%) in P. pastoris (Table 3). The majority (79%) of these 78 proteins contained a polyhistidine tag for purification, but the length of the tag varied between 4 and 10 residues (Fig. 5A). The location of the histidine tag was typically at the C-terminus of the protein (Fig. 5B). In 53% of cases, a cleavage site was present between the protein sequence and the tag; the PreScission Cleavage Site (PSC) was most widely used (Fig. 5C). Other commonly applied engineering principles were truncation of hydrophilic termini as well as codon optimization, which were used in 36% and 35% of cases, respectively (Table S3, Supporting Information).
The purpose of construct engineering is to improve production yields as well as to aid subsequent purification, commonly for structural evaluation (Scott et al. 2013; Fig. 6). These approaches are general and not specific for a given host system (Kesidis et al. 2020). Codon optimization, however, is applied to match the codons of the host system, which improves the translation process, and it is also done to improve the stability of the mRNA template (Hanson and Coller 2018). Truncation of hydrophilic termini is commonly done to improve crystal formation (Hedfalk 2013). The start point for purification is the membrane fraction isolated by differential centrifugation (Fig. 6). A suitable detergent is necessary, i.e. both efficient in terms of solubilization and keeps the membrane protein in its functional fold (Kotov et al. 2019). The most common purification procedure is based on two steps, where the first step is generally Ni-chromatography using the polyhistidine-tag (Table S3, Supporting Information) and the second is size-exclusion chromatography. Alternatively, ion-exchange chromatography may be used instead of or in combination with Ni-chromatography. If the production level is relatively high, two purification steps are usually sufficient, but other methods and additional steps will sometimes be necessary to achieve the required purity for subsequent evaluation. To succeed with crystallography, a monodisperse peak is desirable, being an indication of a homologous protein population (Drew et al. 2008). Also, a traditional SDS-PAGE gel is informative regarding size and stability of the purified protein product (Bill and Hedfalk 2021).
Figure 6.
Typical procedure for extraction and purification of the membrane protein of interest from the yeast membrane. Several purification steps may be required.
Despite the availability of more complex host systems, such as mammalian or insect cell lines, P. pastoris and S. cerevisiae are still regarded as powerful hosts (Cereghino and Cregg 2000, Rabert et al. 2013, Ahmad et al. 2014). Their genomes are fully sequenced, and they can perform complex translational and post-translational processes. The use of yeast is particularly relevant for the expression of eukaryotic proteins since post-translational modifications are important for proper folding and cannot always be achieved in prokaryotic systems such as bacteria. Furthermore, yeast is usually cheap to work with, easy to genetically modify and can result in high protein yields. Notably, S. cerevisiae is an important biotechnological host, producing 20% of all biopharmaceuticals (Nielsen 2013). In comparison, P. pastoris has been used as a protein production system for more than 25 years. Recent improvements include replacing methanol as inducer or improving protein secretion (Ahmad et al. 2014). A significant advantage of this yeast is that, there is no endotoxin, bacteriophage, or human pathogen contamination (Rabert et al. 2013).
The number of new membrane protein structures is increasing daily (Fig. 1). New emerging technologies (Pandey et al. 2016), combined with ongoing improvements in host systems have allowed us to obtain many more eukaryotic membrane protein structures than were available two decades ago when we worked with Stefan. Although the majority of eukaryotic MEMBRANE PROTEIN structures are DERIVED FROM PROTEINS produced in HEK293 and insect cells (Kesidis et al. 2020), we show here the importance of yeast as a production host and its role as an essential player in the production of eukaryotic membrane proteins for structural and functional analysis.
Supplementary Material
Acknowledgments
The authors would like to acknowledge Professor Jens B Nielsen, Chalmers University of Technology, Sweden, for the invitation to write this min-review to honor Professor Stefan Hohmann.
Contributor Information
Antonio Carlesso, Department of Chemistry and Molecular Biology, University of Gothenburg, Göteborg, Sweden; Università della Svizzera Italiana (USI), Faculty of Biomedical Sciences, Euler Institute, Via G. Buffi 13, CH-6900 Lugano, Switzerland; Department of Pharmacology, University of Gothenburg, Göteborg, Sweden.
Raquel Delgado, Department of Chemistry and Molecular Biology, University of Gothenburg, Göteborg, Sweden.
Oriol Ruiz Isant, Department of Chemistry and Molecular Biology, University of Gothenburg, Göteborg, Sweden.
Owens Uwangue, Department of Chemistry and Molecular Biology, University of Gothenburg, Göteborg, Sweden.
Dylan Valli, Department of Chemistry, Uppsala University, Uppsala, Sweden.
Roslyn M Bill, College of Health and Life Sciences, Aston University, Aston Triangle, Birmingham B4 7ET, United Kingdom.
Kristina Hedfalk, Department of Chemistry and Molecular Biology, University of Gothenburg, Göteborg, Sweden.
Conflict of interest statement
None declared.
Funding
This work was supported by the Swedish Research Council (RFI—Protein Production Sweden to K.H.). Elisabeth and Alfred Ahlqvists Foundation (A.C.) are also gratefully acknowledged for the funding. The authors have no conflict of interest to declare.
References
- Ago H, Kanaoka Y, Irikura Det al. Crystal structure of a human membrane protein involved in cysteinyl leukotriene biosynthesis. Nature. 2007;448:609–12. [DOI] [PubMed] [Google Scholar]
- Ahmad M, Hirz M, Pichler Het al. Protein expression in Pichia pastoris: recent achievements and perspectives for heterologous protein production. Appl Microbiol Biotechnol. 2014;98:5301–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahmed T, Nisler CR, Fluck EC 3rdet al. Structure of the ancient TRPY1 channel from Saccharomyces cerevisiae reveals mechanisms of modulation by lipids and calcium. Structure. 2022;30:139–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahuja S, Whorton MR. Structural basis for mammalian nucleotide sugar transport. Elife. 2019;8:e45221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alguel Y, Amillis S, Leung Jet al. Structure of eukaryotic purine/H(+) symporter UapA suggests a role for homodimerization in transport activity. Nat Commun. 2016;7:11336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aller SG, Yu J, Ward Aet al. Structure of P-glycoprotein reveals a molecular basis for poly-specific drug binding. Science. 2009;323:1718–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai L, Jain BK, You Qet al. Structural basis of the P4B ATPase lipid flippase activity. Nat Commun. 2021;12:5963. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai L, Kovach A, You Qet al. Autoinhibition and activation mechanisms of the eukaryotic lipid flippase Drs2p-Cdc50p. Nat Commun. 2019a;10:4142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai L, Kovach A, You Qet al. Structure of the eukaryotic protein O-mannosyltransferase Pmt1-Pmt2 complex. Nat Struct Mol Biol. 2019b;26:704–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai L, Wang T, Zhao Get al. The atomic structure of a eukaryotic oligosaccharyltransferase complex. Nature. 2018;555:328–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai L, You Q, Feng Xet al. Structure of the ER membrane complex, a transmembrane-domain insertase. Nature. 2020a;584:475–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai L, You Q, Jain BKet al. Transport mechanism of P4 ATPase phosphatidylcholine flippases. Elife. 2020b;9:e62163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baradaran R, Wang C, Siliciano AFet al. Cryo-EM structures of fungal and metazoan mitochondrial calcium uniporters. Nature. 2018;559:580–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barbieri A, Thonghin N, Shafi Tet al. Structure of ABCB1/P-Glycoprotein in the presence of the CFTR potentiator ivacaftor. Membranes. 2021;11:923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bavnhoj L, Paulsen PA, Flores-Canales JCet al. Molecular mechanism of sugar transport in plants unveiled by structures of glucose/H(+) symporter STP10. Nat Plants. 2021;7:1409–19. [DOI] [PubMed] [Google Scholar]
- Berndtsson J, Aufschnaiter A, Rathore Set al. Respiratory supercomplexes enhance electron transport by decreasing cytochrome c diffusion distance. EMBO Rep. 2020;21:e51015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bickers SC, Benlekbir S, Rubinstein JLet al. Structure of Ycf1p reveals the transmembrane domain TMD0 and the regulatory region of ABCC transporters. Proc Natl Acad Sci. 2021;118:e2025853118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bill RM, Hedfalk K, Karlgren Set al. Analysis of the pore of the unusual major intrinsic protein channel, yeast Fps1p. J Biol Chem. 2001;276:36543–9. [DOI] [PubMed] [Google Scholar]
- Bill RM, Hedfalk K. Aquaporins - expression, purification and characterization. Biochim Biophys Acta Biomemb. 2021;1863:183650. [DOI] [PubMed] [Google Scholar]
- Bill RM, Henderson PJ, Iwata Set al. Overcoming barriers to membrane protein structure determination. Nat Biotechnol. 2011;29:335–40. [DOI] [PubMed] [Google Scholar]
- Bill RM. Yeast–a panacea for the structure-function analysis of membrane proteins?. Curr Genet. 2001;40:157–71. [DOI] [PubMed] [Google Scholar]
- Brauer P, Parker JL, Gerondopoulos Aet al. Structural basis for pH-dependent retrieval of ER proteins from the Golgi by the KDEL receptor. Science. 2019;363:1103–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brohawn SG, Campbell EB, MacKinnon R. Domain-swapped chain connectivity and gated membrane access in a Fab-mediated crystal of the human TRAAK K+ channel. Proc Natl Acad Sci. 2013;110:2129–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brohawn SG, Campbell EB, MacKinnon R. Physical mechanism for gating and mechanosensitivity of the human TRAAK K+ channel. Nature. 2014;516:126–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brohawn SG, del Marmol J, MacKinnon R. Crystal structure of the human K2P TRAAK, a lipid- and mechano-sensitive K+ ion channel. Science. 2012;335:436–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brohawn SG, Wang W, Handler Aet al. The mechanosensitive ion channel TRAAK is localized to the mammalian node of Ranvier. Elife. 2019;8:e50403. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brunner JD, Lim NK, Schenck Set al. X-ray structure of a calcium-activated TMEM16 lipid scramblase. Nature. 2014;516:207–12. [DOI] [PubMed] [Google Scholar]
- Cereghino JL, Cregg JM. Heterologous protein expression in the methylotrophic yeast Pichia pastoris. FEMS Microbiol Rev. 2000;24:45–66. [DOI] [PubMed] [Google Scholar]
- Chen X, Wang Q, Ni Fet al. Structure of the full-length shaker potassium channel Kv1.2 by normal-mode-based X-ray crystallographic refinement. Proc Natl Acad Sci. 2010;107:11352–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clausen JD, Bublitz M, Arnou Bet al. SERCA mutant E309Q binds two Ca(2+) ions but adopts a catalytically incompetent conformation. EMBO J. 2013;32:3231–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Conde J, Pumroy RA, Baker Cet al. Allosteric antagonist modulation of TRPV2 by piperlongumine impairs glioblastoma progression. ACS Centr Sci. 2021;7:868–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Coudray N, S LS, Lasala Ret al. Structure of the SLC4 transporter Bor1p in an inward-facing conformation. Protein Sci. 2017;26:130–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Custodio TF, Paulsen PA, Frain KMet al. Structural comparison of GLUT1 to GLUT3 reveal transport regulation mechanism in sugar porter family. Life Sci Alliance. 2021;4:e202000858. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dautant A, Velours J, Giraud MF. Crystal structure of the Mg.ADP-inhibited state of the yeast F1c10-ATP synthase. J Biol Chem. 2010;285:29502–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Mare SW, Venskutonyte R, Eltschkner Set al. Structural basis for glycerol efflux and selectivity of human aquaporin 7. Structure. 2020;28:215–22. [DOI] [PubMed] [Google Scholar]
- De Wachter C, Van Landuyt L, Callewaert N. Engineering of yeast glycoprotein expression. Adv Biochem Eng Biotechnol. 2021;175:93–135. [DOI] [PubMed] [Google Scholar]
- Deng Z, He Z, Maksaev Get al. Cryo-EM structures of the ATP release channel Pannexin 1. Nat Struct Mol Biol. 2020a;27:373–81. [DOI] [PubMed] [Google Scholar]
- Deng Z, Maksaev G, Rau Met al. Gating of human TRPV3 in a lipid bilayer. Nat Struct Mol Biol. 2020b;27:635–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng Z, Maksaev G, Schlegel AMet al. Structural mechanism for gating of a eukaryotic mechanosensitive channel of small conductance. Nat Commun. 2020c;11:3690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng Z, Paknejad N, Maksaev Get al. Cryo-EM and X-ray structures of TRPV4 reveal insight into ion permeation and gating mechanisms. Nat Struct Mol Biol. 2018;25:252–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng Z, Zhao Y, Feng Jet al. Cryo-EM structure of a proton-activated chloride channel TMEM206. Sci Adv. 2021;7:eabe5983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dickinson MS, Lu J, Gupta Met al. Molecular basis of multistep voltage activation in plant two-pore channel 1. Proc Natl Acad Sci. 2022;119:e2110936119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dickson VK, Pedi L, Long SB. Structure and insights into the function of a Ca(2+)-activated Cl(-) channel. Nature. 2014;516:213–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dingwell DA, Brown LS, Ladizhansky V. Structure of the functionally important extracellular loop C of human aquaporin 1 obtained by solid-state NMR under nearly physiological conditions. J Phys Chem B. 2019;123:7700–10. [DOI] [PubMed] [Google Scholar]
- Diver MM, Pedi L, Koide Aet al. Atomic structure of the eukaryotic intramembrane RAS methyltransferase ICMT. Nature. 2018;553:526–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dosey TL, Wang Z, Fan Get al. Structures of TRPV2 in distinct conformations provide insight into role of the pore turret. Nat Struct Mol Biol. 2019;26:40–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drew D, Newstead S, Sonoda Yet al. GFP-based optimization scheme for the overexpression and purification of eukaryotic membrane proteins in Saccharomyces cerevisiae. Nat Protoc. 2008;3:784–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elbing K, Larsson C, Bill RMet al. Role of hexose transport in control of glycolytic flux in Saccharomyces cerevisiae. Appl Environ Microbiol. 2004;70:5323–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eriksson UK, Fischer G, Friemann Ret al. Subangstrom resolution X-ray structure details aquaporin-water interactions. Science. 2013;340:1346–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Falzone ME, Rheinberger J, Lee BCet al. Structural basis of Ca(2+)-dependent activation and lipid transport by a TMEM16 scramblase. Elife. 2019;8:e43229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fischer G, Kosinska-Eriksson U, Aponte-Santamaria Cet al. Crystal structure of a yeast aquaporin at 1.15 angstrom reveals a novel gating mechanism. PLoS Biol. 2009;7:e1000130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Frick A, Eriksson UK, de Mattia Fet al. X-ray structure of human aquaporin 2 and its implications for nephrogenic diabetes insipidus and trafficking. Proc Natl Acad Sci. 2014;111:6305–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garaeva AA, Guskov A, Slotboom DJet al. A one-gate elevator mechanism for the human neutral amino acid transporter ASCT2. Nat Commun. 2019;10:3427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garaeva AA, Oostergetel GT, Gati Cet al. Cryo-EM structure of the human neutral amino acid transporter ASCT2. Nat Struct Mol Biol. 2018;25:515–21. [DOI] [PubMed] [Google Scholar]
- Garibsingh RA, Ndaru E, Garaeva AAet al. Rational design of ASCT2 inhibitors using an integrated experimental-computational approach. Proc Natl Acad Sci. 2021;118:e2104093118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geng Y, Deng Z, Zhang Get al. Coupling of Ca(2+) and voltage activation in BK channels through the alphaB helix/voltage sensor interface. Proc Natl Acad Sci. 2020;117:14512–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gerngross TU. Advances in the production of human therapeutic proteins in yeasts and filamentous fungi. Nat Biotechnol. 2004;22:1409–14. [DOI] [PubMed] [Google Scholar]
- Gerondopoulos A, Brauer P, Sobajima Tet al. A signal capture and proofreading mechanism for the KDEL-receptor explains selectivity and dynamic range in ER retrieval. Elife. 2021;10:e68380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geurts MMG, Clausen JD, Arnou Bet al. The SERCA residue glu340 mediates interdomain communication that guides Ca(2+) transport. Proc Natl Acad Sci. 2020;117:31114–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gnugge R, Rudolf F. Saccharomyces cerevisiae shuttle vectors. Yeast. 2017;34:205–21. [DOI] [PubMed] [Google Scholar]
- Gotfryd K, Mosca AF, Missel JWet al. Human adipose glycerol flux is regulated by a pH gate in AQP10. Nat Commun. 2018;9:4749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gulezian E, Crivello C, Bednenko Jet al. Membrane protein production and formulation for drug discovery. Trends Pharmacol Sci. 2021;42:657–74. [DOI] [PubMed] [Google Scholar]
- Guo H, Bueler SA, Rubinstein JL. Atomic model for the dimeric FO region of mitochondrial ATP synthase. Science. 2017;358:936–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo J, Zeng W, Chen Qet al. Structure of the voltage-gated two-pore channel TPC1 from Arabidopsis thaliana. Nature. 2016;531:196–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo J, Zeng W, Jiang Y. Tuning the ion selectivity of two-pore channels. Proc Natl Acad Sci. 2017;114:1009–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hansen SB, Tao X, MacKinnon R. Structural basis of PIP2 activation of the classical inward rectifier K+ channel Kir2.2. Nature. 2011;477:495–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hanson G, Coller J. Codon optimality, bias and usage in translation and mRNA decay. Nat Rev Mol Cell Biol. 2018;19:20–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harris A, Wagner M, Du Det al. Structure and efflux mechanism of the yeast pleiotropic drug resistance transporter Pdr5. Nat Commun. 2021;12:5254. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hartley AM, Lukoyanova N, Zhang Yet al. Structure of yeast cytochrome c oxidase in a supercomplex with cytochrome bc1. Nat Struct Mol Biol. 2019;26:78–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hartley AM, Meunier B, Pinotsis Net al. Rcf2 revealed in cryo-EM structures of hypoxic isoforms of mature mitochondrial III-IV supercomplexes. Proc Natl Acad Sci. 2020;117:9329–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hedfalk K, Bill RM, Mullins JGet al. A regulatory domain in the C-terminal extension of the yeast glycerol channel Fps1p. J Biol Chem. 2004;279:14954–60. [DOI] [PubMed] [Google Scholar]
- Hedfalk K. Further advances in the production of membrane proteins in Pichia pastoris. Bioengineered. 2013;4:363–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henricsson C, de Jesus Ferreira MC, Hedfalk Ket al. Engineering of a novel Saccharomyces cerevisiae wine strain with a respiratory phenotype at high external glucose concentrations. Appl Environ Microbiol. 2005;71:6185–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hino T, Arakawa T, Iwanari Het al. G-protein-coupled receptor inactivation by an allosteric inverse-agonist antibody. Nature. 2012;482:237–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ho JD, Yeh R, Sandstrom Aet al. Crystal structure of human aquaporin 4 at 1.8 A and its mechanism of conductance. Proc Natl Acad Sci. 2009;106:7437–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horsefield R, Norden K, Fellert Met al. High-resolution x-ray structure of human aquaporin 5. Proc Natl Acad Sci. 2008;105:13327–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hou X, Burstein SR, Long SB. Structures reveal opening of the store-operated calcium channel Orai. Elife. 2018;7:e36758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hou X, Outhwaite IR, Pedi Let al. Cryo-EM structure of the calcium release-activated calcium channel Orai in an open conformation. Elife. 2020;9:e62772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hou X, Pedi L, Diver MMet al. Crystal structure of the calcium release-activated calcium channel Orai. Science. 2012;338:1308–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hughes TE, Del Rosario JS, Kapoor Aet al. Structure-based characterization of novel TRPV5 inhibitors. Elife. 2019;8:e49572. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hughes TET, Lodowski DT, Huynh KWet al. Structural basis of TRPV5 channel inhibition by econazole revealed by cryo-EM. Nat Struct Mol Biol. 2018;25:53–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hunte C, Koepke J, Lange Cet al. Structure at 2.3 A resolution of the cytochrome bc(1) complex from the yeast Saccharomyces cerevisiae co-crystallized with an antibody Fv fragment. Structure. 2000;8:669–84. [DOI] [PubMed] [Google Scholar]
- Huynh KW, Cohen MR, Jiang Jet al. Structure of the full-length TRPV2 channel by cryo-EM. Nat Commun. 2016;7:11130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Itskanov S, Kuo KM, Gumbart JCet al. Stepwise gating of the Sec61 protein-conducting channel by Sec63 and Sec62. Nat Struct Mol Biol. 2021;28:162–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Itskanov S, Park E. Structure of the posttranslational Sec protein-translocation channel complex from yeast. Science. 2019;363:84–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin MS, Oldham ML, Zhang Qet al. Crystal structure of the multidrug transporter P-glycoprotein from Caenorhabditis elegans. Nature. 2012;490:566–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kabaleeswaran V, Puri N, Walker JEet al. Novel features of the rotary catalytic mechanism revealed in the structure of yeast F1 ATPase. EMBO J. 2006;25:5433–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kalienkova V, Clerico Mosina V, Bryner Let al. Stepwise activation mechanism of the scramblase nhTMEM16 revealed by cryo-EM. Elife. 2019;8:e44364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kang HA, Kang WK, Go SMet al. Characteristics of Saccharomyces cerevisiae gal1 Delta and gal1 Delta hxk2 Delta mutants expressing recombinant proteins from the GAL promoter. Biotechnol Bioeng. 2005;89:619–29. [DOI] [PubMed] [Google Scholar]
- Kapoor K, Finer-Moore JS, Pedersen BPet al. Mechanism of inhibition of human glucose transporter GLUT1 is conserved between cytochalasin B and phenylalanine amides. Proc Natl Acad Sci. 2016;113:4711–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kapust RB, Waugh DS. Controlled intracellular processing of fusion proteins by TEV protease. Protein Expr Purif. 2000;19:312–8. [DOI] [PubMed] [Google Scholar]
- Karlgren S, Filipsson C, Mullins JGet al. Identification of residues controlling transport through the yeast aquaglyceroporin Fps1 using a genetic screen. Eur J Biochem. 2004;271:771–9. [DOI] [PubMed] [Google Scholar]
- Karlgren S, Pettersson N, Nordlander Bet al. Conditional osmotic stress in yeast: a system to study transport through aquaglyceroporins and osmostress signaling. J Biol Chem. 2005;280:7186–93. [DOI] [PubMed] [Google Scholar]
- Kater L, Wagener N, Berninghausen Oet al. Structure of the Bcs1 AAA-ATPase suggests an airlock-like translocation mechanism for folded proteins. Nat Struct Mol Biol. 2020;27:142–9. [DOI] [PubMed] [Google Scholar]
- Kellosalo J, Kajander T, Kogan Ket al. The structure and catalytic cycle of a sodium-pumping pyrophosphatase. Science. 2012;337:473–6. [DOI] [PubMed] [Google Scholar]
- Kesidis A, Depping P, Lode Aet al. Expression of eukaryotic membrane proteins in eukaryotic and prokaryotic hosts. Methods. 2020;180:3–18. [DOI] [PubMed] [Google Scholar]
- Khelashvili G, Falzone ME, Cheng Xet al. Dynamic modulation of the lipid translocation groove generates a conductive ion channel in Ca(2+)-bound nhTMEM16. Nat Commun. 2019;10:4972. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kintzer AF, Green EM, Dominik PKet al. Structural basis for activation of voltage sensor domains in an ion channel TPC1. Proc Natl Acad Sci. 2018;115:E9095–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kintzer AF, Stroud RM. Structure, inhibition and regulation of two-pore channel TPC1 from Arabidopsis thaliana. Nature. 2016;531:258–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirscht A, Kaptan SS, Bienert GPet al. Crystal structure of an ammonia-permeable aquaporin. PLoS Biol. 2016;14:e1002411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kodan A, Yamaguchi T, Nakatsu Tet al. Inward- and outward-facing X-ray crystal structures of homodimeric P-glycoprotein CmABCB1. Nat Commun. 2019;10:88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kodan A, Yamaguchi T, Nakatsu Tet al. Structural basis for gating mechanisms of a eukaryotic P-glycoprotein homolog. Proc Natl Acad Sci. 2014;111:4049–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kotov V, Bartels K, Veith Ket al. High-throughput stability screening for detergent-solubilized membrane proteins. Sci Rep. 2019;9:10379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lancaster CR, Hunte C, Kelley J 3rdet al. A comparison of stigmatellin conformations, free and bound to the photosynthetic reaction center and the cytochrome bc1 complex. J Mol Biol. 2007;368:197–208. [DOI] [PubMed] [Google Scholar]
- Lange C, Nett JH, Trumpower BLet al. Specific roles of protein-phospholipid interactions in the yeast cytochrome bc1 complex structure. EMBO J. 2001;20:6591–600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lanza AM, Curran KA, Rey LGet al. A condition-specific codon optimization approach for improved heterologous gene expression in Saccharomyces cerevisiae. BMC Syst Biol. 2014;8:33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Le CA, Harvey DS, Aller SG. Structural definition of polyspecific compensatory ligand recognition by P-glycoprotein. IUCrJ. 2020;7:663–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee JY, Kinch LN, Borek DMet al. Crystal structure of the human sterol transporter ABCG5/ABCG8. Nature. 2016a;533:561–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee SJ, Ren F, Zangerl-Plessl EMet al. Structural basis of control of inward rectifier Kir2 channel gating by bulk anionic phospholipids. J Gen Physiol. 2016b;148:227–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li B, Rietmeijer RA, Brohawn SG. Structural basis for pH gating of the two-pore domain K(+) channel TASK2. Nature. 2020;586:457–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li H, Huang CY, Govorunova EGet al. The crystal structure of bromide-bound Gt ACR1 reveals a pre-activated state in the transmembrane anion tunnel. Elife. 2021a;10:e65903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J, Jaimes KF, Aller SG. Refined structures of mouse P-glycoprotein. Protein Sci. 2014;23:34–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li KM, Wilkinson C, Kellosalo Jet al. Membrane pyrophosphatases from Thermotoga maritima and Vigna radiata suggest a conserved coupling mechanism. Nat Commun. 2016;7:13596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li P, Wang K, Salustros Net al. Structure and transport mechanism of P5B-ATPases. Nat Commun. 2021b;12:3973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lieske J, Cerv M, Kreida Set al. On-chip crystallization for serial crystallography experiments and on-chip ligand-binding studies. IUCrJ. 2019;6:714–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lin SM, Tsai JY, Hsiao CDet al. Crystal structure of a membrane-embedded H+-translocating pyrophosphatase. Nature. 2012;484:399–403. [DOI] [PubMed] [Google Scholar]
- Liu S, Li S, Shen Get al. Structural basis of antagonizing the vitamin K catalytic cycle for anticoagulation. Science. 2021;371:eabc5667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu S, Li S, Yang Yet al. Termini restraining of small membrane proteins enables structure determination at near-atomic resolution. Sci Adv. 2020;6:eabe3717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Z, Hou J, Martinez JLet al. Correlation of cell growth and heterologous protein production by Saccharomyces cerevisiae. Appl Microbiol Biotechnol. 2013;97:8955–62. [DOI] [PubMed] [Google Scholar]
- Lolicato M, Arrigoni C, Mori Tet al. K2P2.1 (TREK-1)-activator complexes reveal a cryptic selectivity filter binding site. Nature. 2017;547:364–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lolicato M, Natale AM, Abderemane-Ali Fet al. K2P channel C-type gating involves asymmetric selectivity filter order-disorder transitions. Sci Adv. 2020;6:eabc9174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lolicato M, Riegelhaupt PM, Arrigoni Cet al. Transmembrane helix straightening and buckling underlies activation of mechanosensitive and thermosensitive K(2P) channels. Neuron. 2014;84:1198–212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Long SB, Campbell EB, Mackinnon R. Crystal structure of a mammalian voltage-dependent shaker family K+ channel. Science. 2005;309:897–903. [DOI] [PubMed] [Google Scholar]
- Long SB, Tao X, Campbell EBet al. Atomic structure of a voltage-dependent K+ channel in a lipid membrane-like environment. Nature. 2007;450:376–82. [DOI] [PubMed] [Google Scholar]
- Luo M, Zhou W, Patel Het al. Bedaquiline inhibits the yeast and human mitochondrial ATP synthases. Commun Biol. 2020;3:452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma J, Yoshimura M, Yamashita Eet al. Structure of rat monoamine oxidase a and its specific recognitions for substrates and inhibitors. J Mol Biol. 2004;338:103–14. [DOI] [PubMed] [Google Scholar]
- Maity K, Heumann JM, McGrath APet al. Cryo-EM structure of OSCA1.2 from Oryza sativa elucidates the mechanical basis of potential membrane hyperosmolality gating. Proc Natl Acad Sci. 2019;116:14309–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matoba K, Kotani T, Tsutsumi Aet al. Atg9 is a lipid scramblase that mediates autophagosomal membrane expansion. Nat Struct Mol Biol. 2020;27:1185–93. [DOI] [PubMed] [Google Scholar]
- Matsuoka R, Fudim R, Jung Set al. Structure, mechanism and lipid-mediated remodeling of the mammalian Na(+)/H(+) exchanger NHA2. Nat Struct Mol Biol. 2022;29:108–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matthies D, Bae C, Toombes GEet al. Single-particle cryo-EM structure of a voltage-activated potassium channel in lipid nanodiscs. Elife. 2018;7:e37558. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mazhab-Jafari MT, Rohou A, Schmidt Cet al. Atomic model for the membrane-embedded VO motor of a eukaryotic V-ATPase. Nature. 2016;539:118–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McKenna MJ, Sim SI, Ordureau Aet al. The endoplasmic reticulum P5A-ATPase is a transmembrane helix dislocase. Science. 2020;369:eabc5809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller AN, Long SB. Crystal structure of the human two-pore domain potassium channel K2P1. Science. 2012;335:432–6. [DOI] [PubMed] [Google Scholar]
- Miller AN, Vaisey G, Long SB. Molecular mechanisms of gating in the calcium-activated chloride channel bestrophin. Elife. 2019;8:e43231. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller-Vedam LE, Brauning B, Popova KDet al. Structural and mechanistic basis of the EMC-dependent biogenesis of distinct transmembrane clients. Elife. 2020;9:e62611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Molina DM, Wetterholm A, Kohl Aet al. Structural basis for synthesis of inflammatory mediators by human leukotriene C4 synthase. Nature. 2007;448:613–6. [DOI] [PubMed] [Google Scholar]
- Mooij WT, Mitsiki E, Perrakis A. ProteinCCD: enabling the design of protein truncation constructs for expression and crystallization experiments. Nucleic Acids Res. 2009;37:W402–405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neuhaus JD, Wild R, Eyring Jet al. Functional analysis of Ost3p and Ost6p containing yeast oligosaccharyltransferases. Glycobiology. 2021;31:1604–15. [DOI] [PubMed] [Google Scholar]
- Nicklisch SC, Rees SD, McGrath APet al. Global marine pollutants inhibit P-glycoprotein: environmental levels, inhibitory effects, and cocrystal structure. Sci Adv. 2016;2:e1600001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niegowski D, Kleinschmidt T, Olsson Uet al. Crystal structures of leukotriene C4 synthase in complex with product analogs: implications for the enzyme mechanism. J Biol Chem. 2014;289:5199–207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nielsen J. Production of biopharmaceutical proteins by yeast: advances through metabolic engineering. Bioengineered. 2013;4:207–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niu Y, Tao X, Touhara KKet al. Cryo-EM analysis of PIP2 regulation in mammalian GIRK channels. Elife. 2020;9:e60552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nyblom M, Frick A, Wang Yet al. Structural and functional analysis of SoPIP2;1 mutants adds insight into plant aquaporin gating. J Mol Biol. 2009;387:653–68. [DOI] [PubMed] [Google Scholar]
- Nyblom M, Oberg F, Lindkvist-Petersson Ket al. Exceptional overproduction of a functional human membrane protein. Protein Expr Purif. 2007;56:110–20. [DOI] [PubMed] [Google Scholar]
- Oberg F, Ekvall M, Nyblom Met al. Insight into factors directing high production of eukaryotic membrane proteins; production of 13 human AQPs in Pichia pastoris. Mol Membr Biol. 2009;26:215–27. [DOI] [PubMed] [Google Scholar]
- Oberg F, Hedfalk K. Recombinant production of the human aquaporins in the yeast Pichia pastoris (invited review). Mol Membr Biol. 2013;30:15–31. [DOI] [PubMed] [Google Scholar]
- Oberg F, Sjohamn J, Conner MTet al. Improving recombinant eukaryotic membrane protein yields in Pichia pastoris: the importance of codon optimization and clone selection. Mol Membr Biol. 2011;28:398–411. [DOI] [PubMed] [Google Scholar]
- Oldham ML, Grigorieff N, Chen J. Structure of the transporter associated with antigen processing trapped by herpes simplex virus. Elife. 2016;5:e21829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oot RA, Huang LS, Berry EAet al. Crystal structure of the yeast vacuolar ATPase heterotrimeric EGC(head) peripheral stalk complex. Structure. 2012;20:1881–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oot RA, Kane PM, Berry EAet al. Crystal structure of yeast V1-ATPase in the autoinhibited state. EMBO J. 2016;35:1694–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Otterstedt K, Larsson C, Bill RMet al. Switching the mode of metabolism in the yeast Saccharomyces cerevisiae. EMBO Rep. 2004;5:532–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Palsdottir H, Lojero CG, Trumpower BLet al. Structure of the yeast cytochrome bc1 complex with a hydroxyquinone anion Qo site inhibitor bound. J Biol Chem. 2003;278:31303–11. [DOI] [PubMed] [Google Scholar]
- Pandey A, Shin K, Patterson REet al. Current strategies for protein production and purification enabling membrane protein structural biology. Biochem Cell Biol. 2016;94:507–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parker JL, Corey RA, Stansfeld PJet al. Structural basis for substrate specificity and regulation of nucleotide sugar transporters in the lipid bilayer. Nat Commun. 2019;10:4657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parker JL, Newstead S. Molecular basis of nitrate uptake by the plant nitrate transporter NRT1.1. Nature. 2014;507:68–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parker JL, Newstead S. Structural basis of nucleotide sugar transport across the Golgi membrane. Nature. 2017;551:521–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pau V, Zhou Y, Ramu Yet al. Crystal structure of an inactivated mutant mammalian voltage-gated K(+) channel. Nat Struct Mol Biol. 2017;24:857–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paulsen PA, Custodio TF, Pedersen BP. Crystal structure of the plant symporter STP10 illuminates sugar uptake mechanism in monosaccharide transporter superfamily. Nat Commun. 2019;10:407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pedersen BP, Kumar H, Waight ABet al. Crystal structure of a eukaryotic phosphate transporter. Nature. 2013;496:533–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pettersson N, Hagstrom J, Bill RMet al. Expression of heterologous aquaporins for functional analysis in Saccharomyces cerevisiae. Curr Genet. 2006;50:247–55. [DOI] [PubMed] [Google Scholar]
- Pope L, Lolicato M, Minor DL Jr. Polynuclear ruthenium amines inhibit K2P channels via a “Finger in the dam” mechanism. Cell Chem Biol. 2020;27:511–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pryor EE Jr, Horanyi PS, Clark KMet al. Structure of the integral membrane protein CAAX protease Ste24p. Science. 2013;339:1600–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pumroy RA, Samanta A, Liu Yet al. Molecular mechanism of TRPV2 channel modulation by cannabidiol. Elife. 2019;8:e48792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qureshi AA, Suades A, Matsuoka Ret al. The molecular basis for sugar import in malaria parasites. Nature. 2020;578:321–5. [DOI] [PubMed] [Google Scholar]
- Rabert C, Weinacker D, Pessoa A Jret al. Recombinants proteins for industrial uses: utilization of Pichia pastoris expression system. Brazil J Microbiol. 2013;44:351–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rana MS, Kumar P, Lee CJet al. Fatty acyl recognition and transfer by an integral membrane S-acyltransferase. Science. 2018;359:eaao6326. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rathore S, Berndtsson J, Marin-Buera Let al. Cryo-EM structure of the yeast respiratory supercomplex. Nat Struct Mol Biol. 2019;26:50–7. [DOI] [PubMed] [Google Scholar]
- Ren F, Logeman BL, Zhang Xet al. X-ray structures of the high-affinity copper transporter ctr1. Nat Commun. 2019;10:1386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rietmeijer RA, Sorum B, Li Bet al. Physical basis for distinct basal and mechanically gated activity of the human K(+) channel TRAAK. Neuron. 2021;109:2902–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Robinson GC, Bason JV, Montgomery MGet al. The structure of F(1)-ATPase from Saccharomyces cerevisiae inhibited by its regulatory protein IF(1). Open Biol. 2013;3:120164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roh SH, Shekhar M, Pintilie Get al. Cryo-EM and MD infer water-mediated proton transport and autoinhibition mechanisms of Vo complex. Sci Adv. 2020;6:eabb9605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roh SH, Stam NJ, Hryc CFet al. The 3.5-A CryoEM structure of nanodisc-reconstituted yeast vacuolar ATPase Vo proton channel. Mol Cell. 2018;69:993–1004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruprecht JJ, Hellawell AM, Harding Met al. Structures of yeast mitochondrial ADP/ATP carriers support a domain-based alternating-access transport mechanism. Proc Natl Acad Sci. 2014;111:E426–434. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruprecht JJ, King MS, Zogg Tet al. The molecular mechanism of transport by the mitochondrial ADP/ATP carrier. Cell. 2019;176:435–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schoebel S, Mi W, Stein Aet al. Cryo-EM structure of the protein-conducting ERAD channel Hrd1 in complex with Hrd3. Nature. 2017;548:352–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scott DJ, Kummer L, Tremmel Det al. Stabilizing membrane proteins through protein engineering. Curr Opin Chem Biol. 2013;17:427–35. [DOI] [PubMed] [Google Scholar]
- Shimamura T, Shiroishi M, Weyand Set al. Structure of the human histamine H1 receptor complex with doxepin. Nature. 2011;475:65–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Solmaz SR, Hunte C. Structure of complex III with bound cytochrome c in reduced state and definition of a minimal core interface for electron transfer. J Biol Chem. 2008;283:17542–9. [DOI] [PubMed] [Google Scholar]
- Son SY, Ma J, Kondou Yet al. Structure of human monoamine oxidase a at 2.2-A resolution: the control of opening the entry for substrates/inhibitors. Proc Natl Acad Sci. 2008;105:5739–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Srivastava AP, Luo M, Zhou Wet al. High-resolution cryo-EM analysis of the yeast ATP synthase in a lipid membrane. Science. 2018;360:eaas9699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stock D, Leslie AG, Walker JE. Molecular architecture of the rotary motor in ATP synthase. Science. 1999;286:1700–5. [DOI] [PubMed] [Google Scholar]
- Symersky J, Osowski D, Walters DEet al. Oligomycin frames a common drug-binding site in the ATP synthase. Proc Natl Acad Sci. 2012a;109:13961–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Symersky J, Pagadala V, Osowski Det al. Structure of the c(10) ring of the yeast mitochondrial ATP synthase in the open conformation. Nat Struct Mol Biol. 2012b;19:485–91., S481. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Szewczyk P, Tao H, McGrath APet al. Snapshots of ligand entry, malleable binding and induced helical movement in P-glycoprotein. Acta Crystallogr Sect D Biol Crystallogr. 2015;71:732–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tamas MJ, Karlgren S, Bill RMet al. A short regulatory domain restricts glycerol transport through yeast Fps1p. J Biol Chem. 2003;278:6337–45. [DOI] [PubMed] [Google Scholar]
- Tanaka Y, Iwaki S, Tsukazaki T. Crystal structure of a plant multidrug and toxic compound extrusion family protein. Structure. 2017;25:1455–60. [DOI] [PubMed] [Google Scholar]
- Tang WK, Borgnia MJ, Hsu ALet al. Structures of AAA protein translocase Bcs1 suggest translocation mechanism of a folded protein. Nat Struct Mol Biol. 2020;27:202–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tao X, Avalos JL, Chen Jet al. Crystal structure of the eukaryotic strong inward-rectifier K+ channel Kir2.2 at 3.1 A resolution. Science. 2009;326:1668–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tao X, Lee A, Limapichat Wet al. A gating charge transfer center in voltage sensors. Science. 2010;328:67–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tao Y, Cheung LS, Li Set al. Structure of a eukaryotic SWEET transporter in a homotrimeric complex. Nature. 2015;527:259–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thulasingam M, Orellana L, Nji Eet al. Crystal structures of human MGST2 reveal synchronized conformational changes regulating catalysis. Nat Commun. 2021;12:1728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Timcenko M, Dieudonne T, Montigny Cet al. Structural basis of substrate-independent phosphorylation in a P4-ATPase lipid flippase. J Mol Biol. 2021;433:167062. [DOI] [PubMed] [Google Scholar]
- Timcenko M, Lyons JA, Januliene Det al. Structure and autoregulation of a P4-ATPase lipid flippase. Nature. 2019;571:366–70. [DOI] [PubMed] [Google Scholar]
- Tornroth-Horsefield S, Wang Y, Hedfalk Ket al. Structural mechanism of plant aquaporin gating. Nature. 2006;439:688–94. [DOI] [PubMed] [Google Scholar]
- Tsai JY, Chu CH, Lin MGet al. Structure of the sodium-dependent phosphate transporter reveals insights into human solute carrier SLC20. Sci Adv. 2020;6:eabb4024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsai JY, Tang KZ, Li KMet al. Roles of the hydrophobic gate and exit channel in Vigna radiata pyrophosphatase ion translocation. J Mol Biol. 2019;431:1619–32. [DOI] [PubMed] [Google Scholar]
- Vaisey G, Miller AN, Long SB. Distinct regions that control ion selectivity and calcium-dependent activation in the bestrophin ion channel. Proc Natl Acad Sci. 2016;113:E7399–408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- van den Berg B, Chembath A, Jefferies Det al. Structural basis for Mep2 ammonium transceptor activation by phosphorylation. Nat Commun. 2016;7:11337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- van den Berg B, Pedebos C, Bolla JRet al. Structural basis for silicic acid uptake by higher plants. J Mol Biol. 2021;433:167226. [DOI] [PubMed] [Google Scholar]
- Vasanthakumar T, Bueler SA, Wu Det al. Structural comparison of the vacuolar and Golgi V-ATPases from Saccharomyces cerevisiae. Proc Natl Acad Sci. 2019;116:7272–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vieira Gomes AM, Souza Carmo T, Carvalho LSet al. Comparison of yeasts as hosts for recombinant protein production. Microorganisms. 2018;6:38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Waight AB, Pedersen BP, Schlessinger Aet al. Structural basis for alternating access of a eukaryotic calcium/proton exchanger. Nature. 2013;499:107–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang G, Huang M, Nielsen J. Exploring the potential of Saccharomyces cerevisiae for biopharmaceutical protein production. Curr Opin Biotechnol. 2017;48:77–84. [DOI] [PubMed] [Google Scholar]
- Wang H, Schoebel S, Schmitz Fet al. Characterization of aquaporin-driven hydrogen peroxide transport. Biochim Biophys Acta Biomemb. 2020;1862:183065. [DOI] [PubMed] [Google Scholar]
- Wang K, Preisler SS, Zhang Let al. Structure of the human Clc-1 chloride channel. PLoS Biol. 2019a;17:e3000218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang XH, Su M, Gao Fet al. Structural basis for activity of TRIC counter-ion channels in calcium release. Proc Natl Acad Sci. 2019b;116:4238–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weng TH, Steinchen W, Beatrix Bet al. Architecture of the active post-translational Sec translocon. EMBO J. 2021;40:e105643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- White KL, Eddy MT, Gao ZGet al. Structural connection between activation microswitch and allosteric sodium site in GPCR signaling. Structure. 2018;26:259–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whorton MR, MacKinnon R. Crystal structure of the mammalian GIRK2 K+ channel and gating regulation by g proteins, PIP2, and sodium. Cell. 2011;147:199–208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whorton MR, MacKinnon R. X-ray structure of the mammalian GIRK2-betagamma G-protein complex. Nature. 2013;498:190–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wild R, Kowal J, Eyring Jet al. Structure of the yeast oligosaccharyltransferase complex gives insight into eukaryotic N-glycosylation. Science. 2018;359:545–50. [DOI] [PubMed] [Google Scholar]
- Winklemann I, Matsuoka R, Meier PFet al. Structure and elevator mechanism of the mammalian sodium/proton exchanger NHE9. EMBO J. 2020;39:e105908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Winkler MBL, Kidmose RT, Szomek Met al. Structural insight into eukaryotic sterol transport through Niemann-pick type C proteins. Cell. 2019;179:485–97. [DOI] [PubMed] [Google Scholar]
- Wu X, Cabanos C, Rapoport TA. Structure of the post-translational protein translocation machinery of the ER membrane. Nature. 2019;566:136–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu X, Siggel M, Ovchinnikov Set al. Structural basis of ER-associated protein degradation mediated by the Hrd1 ubiquitin ligase complex. Science. 2020;368:eaaz2449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang H, Hu M, Guo Jet al. Pore architecture of TRIC channels and insights into their gating mechanism. Nature. 2016;538:537–41. [DOI] [PubMed] [Google Scholar]
- Yang Y, Liu XR, Greenberg ZJet al. Open conformation of tetraspanins shapes interaction partner networks on cell membranes. EMBO J. 2020;39:e105246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zangerl-Plessl EM, Lee SJ, Maksaev Get al. Atomistic basis of opening and conduction in mammalian inward rectifier potassium (Kir2.2) channels. J Gen Physiol. 2020;152:e201912422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y, Ou X, Wang Xet al. Structure of the mitochondrial TIM22 complex from yeast. Cell Res. 2021;31:366–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao P, Zhao C, Chen Det al. Structure and activation mechanism of the hexameric plasma membrane H(+)-ATPase. Nat Commun. 2021;12:6439. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.