Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Jun 1.
Published in final edited form as: Methods. 2024 Apr 10;226:35–48. doi: 10.1016/j.ymeth.2024.03.011

Exploring GPCR conformational dynamics using single-molecule fluorescence

Eugene Agyemang 1, Alyssa N Gonneville 2, Sriram Tiruvadi-Krishnan 2, Rajan Lamichhane 1,2,*
PMCID: PMC11098685  NIHMSID: NIHMS1988114  PMID: 38604413

Abstract

G protein-coupled receptors (GPCRs) are membrane proteins that transmit specific external stimuli into cells by changing their conformation. This conformational change allows them to couple and activate G-proteins to initiate signal transduction. A critical challenge in studying and inferring these structural dynamics arises from the complexity of the cellular environment, including the presence of various endogenous factors. Due to the recent advances in cell-expression systems, membrane-protein purification techniques, and labeling approaches, it is now possible to study the structural dynamics of GPCRs at a single-molecule level both in vitro and in live cells. In this review, we discuss state-of-the-art techniques and strategies for expressing, purifying, and labeling GPCRs in the context of single-molecule research. We also highlight four recent studies that demonstrate the applications of single-molecule microscopy in revealing the dynamics of GPCRs. These techniques are also useful as complementary methods to verify the results obtained from other structural biology tools like cryo-electron microscopy and x-ray crystallography.

Keywords: GPCR dynamics, single-molecule, TIRF microscopy, FRET, nanodiscs, co-polymers

Graphical Abstract

graphic file with name nihms-1988114-f0001.jpg

1. Introduction

G protein-coupled receptors (GPCRs) are a large group of membrane proteins in eukaryotic cells that are involved in the sensation of light, taste, and smell [1]. GPCRs mediate physiological and pathological processes like pain transmission and cardiovascular regulation and are implicated in several diseases, including Alzheimer’s, schizophrenia, and cancer [2-5]. They have a conserved structure of seven transmembrane (TM) helices linked by three intracellular and three extracellular loops [6]. Human GPCRs are classified into four different groups based on their sequence homology and functional similarity. Class A GPCRs (rhodopsin-like receptors) constitute the largest and most extensively studied group of mammalian GPCRs. The other groups include class B (secretin and adhesion family), class C (glutamate family), and class F (frizzled family) [6, 7]. GPCRs mediate cellular responses by interacting with various external stimuli, including photons and neurotransmitters, at their orthosteric binding sites. These binding sites are the main molecular targets that drive the activation and signaling properties of GPCRs [6].

GPCRs initiate cellular signaling cascades upon stimulation with ligands by coupling to cytoplasmic transducer proteins like heterotrimeric G proteins and arrestins (Figure 1) [8]. G protein-dependent signaling begins with the binding of an agonist to the orthosteric binding site of the GPCR. This binding results in conformational changes within the TM helices, which primes the receptor for G protein binding and activation. The activation of G proteins promotes the exchange of GDP for GTP at the G protein’s α subunit, causing it to dissociate from the Gβγ subunits. This process then stimulates the production of second messengers such as cyclic adenosine monophosphate (cAMP), which triggers several physiological changes at the cellular level [1, 9]. Alternatively, GPCR signaling through arrestins is triggered by the phosphorylation of the receptor’s C terminal tail by G protein-coupled receptor kinases (GRKs). The recruited arrestins block G-protein-mediated signaling, leading to receptor desensitization and internalization (Figure 1) [1].

Figure 1. GPCR signaling at the plasma membrane.

Figure 1.

(1) Inactive GPCRs form transient pre-coupled complexes with inactive heterotrimeric G proteins through random collisions at the cell surface (Refer to Mafi et al. (2022) [9]). Upon agonist binding, the GPCR undergoes conformational rearrangement of the TM helices, activating the receptor and stabilizing the receptor-G protein complex in a ligand-specific manner. The activated complex then initiates both G protein-dependent and independent signaling pathways. (2) In this example, guanine nucleotide exchange, mediated by the agonist-bound GPCR (β2AR) initiates G protein-dependent signaling by activating the Gα subunit of the G protein and causing the dissociation of the Gβγ subunit. The activated Gα subunit subsequently transduces downstream signals by activating various effector proteins such as adenylyl cyclase, resulting in diverse cellular responses. (3) Additionally, β-arrestins mediate G protein-independent signaling. During this process, the agonist-activated receptor undergoes phosphorylation of the C-terminal tail by GRKs, inhibiting G protein coupling and facilitating β-arrestin recruitment to the active phosphorylated receptor, β-arrestin coupling then initiates several downstream signaling events, including receptor desensitization and trafficking. Figure created using BioRender.com.

The development and progress of biophysical methods have provided valuable insights that have transformed our understanding of the dynamic processes in GPCRs. Various biophysical techniques, such as X-ray crystallography [10-13] and cryogenic-electron microscopy (cryo-EM) [14-18] have been employed to capture static snapshots of receptor conformations at high resolution. Additionally, studies using nuclear magnetic resonance (NMR) [19-22], single-molecule fluorescence (SMF) [23-26], and molecular dynamics (MD) simulations [27-30] have shown that GPCRs exhibit structural flexibility and undergo transitions between multiple conformational states during receptor activation, providing a clear picture of the complexity of GPCR activity. Of these techniques, SMF has emerged as a powerful complementary tool to other biophysical approaches, as it allows for real-time observation of GPCR structural dynamics and their interactions with other signaling molecules [31]. Moreover, SMF requires minimal structural modifications and can be performed in the context of complex native environments.

In this review, we will discuss the strategies and methods successfully implemented for preparing GPCRs for single-molecule studies. Our focus will be mainly on the expression systems, purification approaches, and labeling strategies, as well as the methods available for the structural and functional characterization of GPCRs at the single-molecule level. Special emphasis will be placed on single-molecule total internal reflection fluorescence (TIRF) microscopy, a high-precision method for determining distance changes between fluorescent probes. We will also highlight its applications and limitations. The techniques discussed in this review will provide researchers with an overview of the pipeline for conducting single-molecule research on GPCRs.

2. Expression strategies for GPCR production

Isolating pure and adequate amounts of GPCRs for analysis is challenging due to their nature as integral membrane proteins. A diverse range of heterologous expression systems have been extensively explored to facilitate GPCR production. The selection of an appropriate expression system depends on several factors, such as the GPCR of interest, expression yield for downstream applications, and the specific requirements for achieving correct folding, post-translational modifications, and integration into the cell membrane [32].

GPCRs like rhodopsin can be obtained in significant quantities from natural sources due to their abundance in the eye's retina [33]. However, most GPCRs are naturally produced in small amounts within cell membranes, requiring heterologous expression systems to generate larger quantities, particularly for structural and functional characterization. Commonly employed expression systems include mammalian, insect, and yeast cells (Table 1) [34]). Beyond these, alternative systems have been successfully used to achieve GPCR expression. For instance, bacterial cells, specifically Escherichia coli, have successfully been used to express a water-soluble variant of the mu-opioid receptor [35, 36]. Additionally, an E. coli-based cell-free system has proven effective for expressing the human neuropeptide Y4 (NPY) receptor and human melatonin 1B (MTN) receptor [37]. Notably, the protozoan eukaryotic system LEXSY (Leishmania tarentolae) has been used to achieve remarkably high-yield expression of a stable and functional human A2A adenosine receptor (A2AAR) [38]. These diverse systems contribute significantly to our understanding of GPCR structure and function and facilitate the development of high throughput screening methods for therapeutic improvement and novel drug discovery.

Table 1.

Recombinant human GPCRs derived from different expression systems.

GPCR Class Expression
system
Cell line Vector References
β2-adrenergic receptor (β2AR) A Insect cell Sf9 pFastBac [24]
Glucagon receptor (GCGR) B Insect cell Sf9 pFastBac1 [61]
Metabotropic glutamate receptor 5 (mGluR5) C Insect cell Sf21 pFastBac [62]
Corticotropin-releasing factor receptor (CRF1) B Insect cell High Five N/A [63]
μ opioid receptor (μOR) A Yeast cell S. cerevisiae pBB161 [64]
A2A adenosine receptor (A2AAR) A Yeast cell P. pastoris pPIC9K [23]
Histamine H1 receptor (H1R) A Yeast cell P. pastoris pPIC9K (SMD1163) [65]
Metabotropic glutamate receptor 2 (mGluR2) C Mammalian cell CHO (stable) pcDNA5/FRT/TOIRES [66]
Glucagon-like peptide-1 receptor (GLP1R) B Mammalian cell HEK293 pCDNA3 [67]

Abbreviations: CHO - Chinese hamster ovary; HEK293 - Human embryonic kidney 293

2.1. Mammalian cell expression

Mammalian cells represent one of the most widely used eukaryotic expression systems for GPCR production. They possess the necessary cellular components to correctly translate, fold, modify, and incorporate proteins into the cell membrane. The intrinsic machinery within these cells enables the processing and maturation of proteins, ensuring their functional integrity and appropriate localization within the cellular environment. This capability of mammalian cells to perform these essential tasks makes them a good choice for producing GPCRs. Despite being optimal expression systems, several concerns have been raised over the production of GPCRs using mammalian cells. The most common issues identified include improper protein folding, incomplete protein insertion into the cell membrane, overloading the post-translational machinery, and cell toxicity, which results in reduced expression and yield [39].

Two general approaches are used when it comes to expressing GPCRs in mammalian cells: transient and stable expression. In transient expression, cells are transfected with the gene encoding the GPCR of interest either with a plasmid DNA complexed with cationic compounds like lipofectamine, enabling membrane insertion, or with a recombinant virus such as the vaccinia virus (for review, see [40, 41]) and the Semliki Forest Virus (SFV) (for review see [42]). With this approach, the gene of interest is not integrated into the host cell’s genome. Regardless of the transfection efficiency, GPCR expression is short-lived in transiently transfected cells since the cells are only viable for a few days [41]. For instance, transient expression is typically sufficient in the single-molecule studies of GPCRs, where only small amounts of pure, active protein are required. On the other hand, stable expression is recommended if consistently high quantities of functional protein are necessary for high throughput screening or drug discovery assays. Stable expression requires creating stable cell lines and results in the integration of the gene of interest into the genome of the host cell [39]. When conducting long-term experiments, using stable cell lines can be advantageous as they provide consistent expression levels that can be easily replicated. However, it is important to note that generating stable cell lines can be a time-consuming and expensive process.

2.2. Yeast cell expression

Yeast cells like Pichia pastoris [43-47], Schizosaccharomyces pombe [48-50], and Saccharomyces cerevisiae [51, 52] are often used to express GPCRs. P. pastoris is the most preferred choice due to its ability to produce higher levels of functional GPCRs. It has even been successfully employed to produce different constructs of the human A2A adenosine receptor (A2AAR) for single-molecule studies (Table 1) [23]). On the other hand, S. cerevisiae is better suited for cloning and screening protein constructs [52].

Compared to other eukaryotic expression systems, yeast cells are easy to manipulate and can grow to high cell densities in simple and inexpensive media. They can also perform most post-translational modifications, including disulfide bond formation and glycosylation, which are sometimes crucial for the proper folding and activity of GPCRs [53]. It is worth noting that although yeast cells can perform PTMs, their glycosylation patterns differ from those in mammalian and insect cells. Additionally, the lipid composition of yeast membranes differs from that of mammalian cells, with ergosterol being the dominant sterol instead of cholesterol. Since cholesterol is essential for modulating and maintaining the stability of most GPCRs, yeast cells have been engineered to produce cholesterol, increasing the expression of functional human GPCRs [54].

2.3. Insect cell expression

Insect cells are commonly used in structural studies of GPCRs because they provide high yields - milligram amounts of pure protein [39]. They are also ideal expression systems for single-molecule applications that require even smaller amounts of pure protein. GPCRs undergo heavy post-translational modifications, and insect cells can perform essential post-translational modifications, making them an excellent alternative expression system. To ensure efficient expression and membrane localization of GPCRs in insect cells, signal peptide sequences are typically attached to their N-termini [55-57]. However, like yeast cells, insect cells possess distinct membrane lipid compositions and glycosylation patterns that differ from mammalian cells [58].

Insect cells like Spodoptera frugiperda (American fall armyworm; Sf9 and Sf21 cell lines) and Trichoplusia ni (High Five) have been successfully used to express GPCRs (Table 1). Among these cells, Sf9 and Sf21 insect cell lines are preferred due to their susceptibility to baculoviral infection and replication [59]. To express GPCRs in Sf9 cells, for instance, a recombinant form of the Autographa californica baculovirus is used. Insect cells are typically infected with this virus, which leads to the expression of the viral protein polyhedrin at high levels. However, the gene for polyhedrin can be replaced with a GPCR gene since its expression is not necessary to produce a functional virus [60]. The polyhedrin promoter is a late promoter that only expresses at the end of the infection cycle, specifically 8-24 hours post-infection, resulting in the expression of high levels of the target protein [41].

3. Solubilization and purification of structurally intact GPCRs

To study the structural dynamics of GPCRs using single-molecule techniques, they must be purified to maintain their structural and dynamic integrity. As integral membrane proteins, it is beneficial to purify GPCRs in a native-like or native membrane environment for structural studies [68]. Therefore, in vitro studies require solubilization methods that utilize either detergent micelles or lipid nanodiscs to shield their hydrophobic regions, preventing precipitation since these proteins are insoluble in aqueous solutions [68]. The most commonly used methods are reviewed below.

3.1. Detergent micelles

The solubilization of membrane proteins requires a disruption of the lipid bilayer for protein removal without immediate reconstitution back into the membrane [69]. Amphiphilic detergent molecules can disrupt the membrane and mimic phospholipid properties to maintain the stability of the membrane protein [69]. A wide variety of detergents commercially available for the extraction of membrane proteins vary by charge and chain length [69]. Some of the most common detergents used for studies of GPCRs are dodecyl maltose (DDM) and lauryl maltose neopentylglycol (LMNG), with cholesteryl hemisuccinate (CHS) to mimic a more membrane-like environment [70]. DDM-CHS is an ideal detergent for solubilization and purification of GPCRs that have been thermostabilized, while LMNG has been shown to stabilize the less-stable GPCRs [70]. The use of detergents for the solubilization of GPCRs has been incredibly useful in structure determinations.

The process of solubilizing GPCRs into detergent micelles is straightforward upon generating membrane fractions from the expression system of choice. The process involves the resuspension and homogenization of the membrane fraction using a solubilization buffer that contains the detergent of interest [71]. After resuspension, the insoluble material is separated from the solubilized portion through ultracentrifugation, followed by affinity purification using affinity tags (e.g., FLAG or poly-histidine tags) [71, 72].

In single-molecule studies, detergent micelles offer an advantage in that they are quick and effective at solubilizing and maintaining the stability of the receptor after purification. However, they do not accurately represent GPCR dynamic studies due to the lack of native lipids. Lipids are known regulators of many membrane proteins, including GPCRs, and could ultimately play a role in the dynamics and activity of GPCRs (Figure 2A).

Figure 2. Schematic showing the solubilization and purification of GPCRs.

Figure 2.

(A) Stepwise process of the solubilization of GPCRs in detergent micelles (e.g. DDM or LMNG) and the purification of the GPCR through IMAC. (B) Reconstitution of GPCRs into lipid nanodiscs purified from detergent micelles, lipids dissolved in detergents, and membrane scaffold proteins. After reconstitution, the detergent is removed using bio-beads and empty nanodiscs are removed with another purification step. (C) Stepwise process of the solubilization of membrane proteins using co-polymers (e.g.: SMA, DIBMA, or AASTY) and the purification of GPCRs in native membrane nanoparticles through IMAC. Figure created using BioRender.com.

3.2. Protein nanodiscs

Unlike detergent micelles, the reconstitution of GPCRs into lipid nanodiscs provides insights into how particular lipids affect the structure and dynamics of these receptors. After solubilization of GPCRs using detergent micelles, they can be reconstituted into nanodiscs using desired lipid compositions, membrane scaffold proteins (MSPs), and bio-beads to remove the detergent [71, 73, 74]. The MSPs (human serum apolipoprotein A-I) are amphipathic helical proteins that encircle and stabilize the hydrophobic acyl chains of lipid bilayers upon self-assembly with lipids to form nanodiscs [75, 76]. Nanodiscs offer an advantage over detergent micelles by providing a native-like environment for the receptor. This allows for controlled manipulation of the lipid composition to observe how different lipids and lipid compositions affect GPCR activity [47]. Recent NMR studies of the class A human A2A adenosine receptor in lipid nanodiscs have been used to show how anionic lipids and cholesterol allosterically influence the activity of the receptor, leading to a more active state in which TM6 moves outward to allow for G-protein coupling [47, 77]. Nanodiscs also offer the ability to alter lipid ratios to study how the different lipid-GPCR interactions affect receptor dynamics. In addition, controlling the lipid composition within nanodiscs is advantageous for single-molecule experiments. For instance, biotin-conjugated lipids can be incorporated into nanodiscs to aid in the immobilization of individual molecules, as opposed to needing an additional biotin-conjugated antibody[71, 73]. Alternatively, biotinylated MSPs can be used for protein nanodisc immobilization. Although the lipid composition within nanodiscs provides a native-like membrane environment, it is essential to note that this environment is not native. Native lipid compositions could affect the dynamics observed during single-molecule experiments (Figure 2B).

3.3. Native membrane protein nanodiscs

Co-polymers like styrene-maleic acid (SMA) and diisobutylene maleic acid (DIBMA) are amphipathic compounds that can spontaneously insert themselves into lipid membranes to encapsulate membrane proteins as a means of solubilization into lipid particles [78-81]. Once native membrane nanodiscs are formed, they are easier to work with than detergent micelles, as it is not necessary to supplement buffers with additives like detergents at any stage of protein purification and characterization. These co-polymers offer an advantage in maintaining the native lipid environment around the receptor. SMA creates lipid nanoparticles (SMALPs) around 9-12 nm in diameter, while DIBMA can make lipid nanoparticles (DIBMALPs) around 25 nm in diameter [79-82]. Both SMALPs and DIBMALPs effectively solubilize GPCRs that retain their structural integrity and functionality [81]. Depending on the size of the nanodiscs, the choice of co-polymers may vary, as the disc size could potentially alter receptor dynamics [79]. Smaller discs may constrict the dynamics of GPCRs, confounding the interpretation of single-molecule data (Figure 2C).

Although the use of co-polymers for solubilization of membrane proteins has made great strides in purification, there are some other limitations. SMA and DIBMA are sensitive to divalent metal ions (e.g., Mg2+ and Ca2+) and pH. Below a pH of 6.5, SMA is insoluble [83]. These properties can complicate experiments as the result would affect the charge of the outward-facing maleic acid group [81, 84]. Another limitation of these co-polymers is their light-absorbing properties due to the styrene group absorbing at 260 nm, which overlaps with UV protein absorption at 280 nm. The absorption of the styrene group could interfere with UV absorption and light scattering assays. [81,84] Despite the disadvantages of some co-polymers such as SMA and DIBMA, the field continues developing new co-polymers for membrane protein solubilization. Overall, using co-polymers for GPCR solubilization allows the receptors to retain their native lipid environment, which is a significant advantage when studying the dynamic nature of GPCRs using single-molecule approaches.

4. Site-specific labeling of GPCRs

To study the dynamics of GPCRs using fluorescence microscopy, regions of the receptor that are exposed outside the lipid bilayer are fluorescently labeled with fluorescent probes. If the structure of the extracellular or intracellular loops has been resolved, then these regions can be fluorescently labeled. However, if the structure of these loops is unknown, then the most commonly labeled regions for GPCR research are the ends of the TM helices or the N- or C-terminus regions. The labeling strategy for studying GPCRs varies depending on whether it is an in vitro or live cell study [31]. The currently available labeling approaches in live cells allow for the study of receptor dynamics [66], interactions with cytosolic proteins like G proteins [85], oligomerization [86], and hetero-dimerization [87] of GPCRs. On the other hand, single-molecule level dynamics of TM helices [24, 25, 73], extracellular domains [88], and C-terminal tails [89] are studied with purified GPCRs, as detailed in the previous section.

Alternatively, GPCRs can be directly labeled by fusing the receptor with a fluorescent protein (e.g., Green Fluorescence Protein (GFP)) connected by a linker peptide [90, 91]. This approach eliminates the need for additional dyes and allows fluorescence can be detected in live cells after protein expression. However, due to the large size of the fluorescent proteins and their propensity to form homo-oligomers, they may influence the dynamics of the receptor. As a result, this method is usually used for studying protein-protein interactions [91]. Regardless of the labeling strategy used, additional validation methods are needed to ensure that the modifications made to the receptor do not alter its function. These methods include detecting intracellular activation by assaying for cAMP levels [92], G protein activation [93], and ligand binding assays [94, 95]. The most used GPCR labeling strategies for in vitro and live cell studies are detailed here:

4.1. Maleimide-based click chemistry

For maleimide-based labeling of GPCRs with a fluorescent dye, an amino acid at the region of interest is modified to cysteine. The thiol group of the exposed cysteine is conjugated to a maleimide-containing fluorescent dye via a thiol-maleimide reaction (click chemistry). This labeling method does not affect the cysteine residues that form disulfide bonds and those buried within the lipid bilayer (Figure 3A), as the thiol moieties are inaccessible. However, other cysteines with the exposed thiol group must be modified to different amino acids to avoid non-specific labeling. This labeling strategy is widely used for labeling GPCRs in vitro but requires extensive knowledge of the protein’s structural information to ensure that amino acid modifications do not affect its function [23, 89].

Figure 3. Labeling strategies for GPCRs.

Figure 3.

(A) A purified GPCR in an MSP nanodisc is shown with the intracellular region of the transmembrane helix to be labeled highlighted. The exposed cysteine is labeled with a fluorescent dye via a thiol-maleimide reaction. (B) The incorporation of an unnatural amino acid (UAA) with an azide moiety in the purified GPCR can be used for labeling with strain-promoted azide-alkyne cyclo-addition (SpAAC) reaction. (C) The tetra-cysteine (TC) moiety in the exposed transmembrane helix can be labeled with a non-fluorescent fluorescein arsenical hairpin binder-ethane dithiol (FlAsH-EdT2) molecule, which becomes fluorescent after coordinating with the sulfide groups of the TC-motif. (D) Illustration of the steps involved in the fluorescent labeling of GPCRs expressed on live cell membranes. Different protein tags and their corresponding fluorescent ligands that can be used for labeling are shown. Figure created using BioRender.com.

4.2. Unnatural amino acid-based click chemistry

A novel labeling approach involves using genetic code expansion to incorporate unnatural amino acids such as p-azido-L-phenyl alanine (azF) at the desired site [96]. This is achieved by inserting the low-abundance amber stop codon, TGA, in the gene that encodes for a GPCR. The expressed protein will then contain azF, which can be tagged with a fluorescent dye using the strain-promoted azide-alkyne cyclo-addition (SpAAc) method (Figure 3B). This labeling strategy has been successfully implemented for labeling GPCRs in both in vitro and in-cell studies and has minimal impact on protein structure [97, 98]. However, protein expression may require extensive optimization, and the amount of expressed protein may be lower than other labeling strategies. Moreover, only a limited number of expression systems can be used for this method (e.g., bacteria [99], yeast [100], or mammalian cells [101]).

4.3. Tetra-cysteine peptide motifs

An alternative labeling method uses short tetra-cysteine peptide motifs (CCXXCC, where X denotes any amino acid), which can be conjugated to a small fluorescein derivative called fluorescein arsenical hairpin binder-ethane dithiol (FlAsH-EDT). Rhodamine-based ReAsH can also be used in place of FlAsH. FlAsH-EDT is non-fluorescent by itself and becomes fluorescent on binding to the tetra-cysteine motif [61] (Figure 3C). Flanking amino acid sequences around the tetra-cysteine motif have been shown to improve the binding efficiency of the FlAsH-EDT molecule (e.g., FLNCCPGCCMEP and HRWCCPGCCKTF). This method has been implemented to label GPCRs in vitro [89], in live cells [102], and in intracellular proteins like arrestins [103, 104]. It is not necessary to know the full structure of the protein to implement this labeling method, however, it must be noted that the exposed cysteine-rich regions in the receptor may interact non-specifically with the FlAsH-EDT [105].

4.4. SNAP-tag and CLIP-tag

SNAP-tag is a 182-amino acid long protein tag (Mw. ~19.4 kDa) that can be fused to a GPCR using recombinant DNA technology with a flexible linker of 6 to 12 amino acids. The linker peptide sequence is crucial to the dynamics of the tagged GPCR. The SNAP-tag is a genetically modified version of the human alkyl-guanine transferase (AGT) enzyme. It forms a covalent bond with O6-benzylguanine derivatives, which can then be attached to a fluorescent dye (Figure 3D) [106]. This allows a wide variety of dyes with unique photophysical characteristics to be used for live-cell microscopy. This method is the most common way of labeling GPCRs in live cells [107-109], compared to the fusion of fluorescent tags to proteins. Another modified AGT enzyme, CLIP-tag, binds to fluorescent O6-benzycytosine derivatives [110]. Typically, the N-terminus of GPCRs, facing the extracellular side of the cell, is used as the fusion site for these protein tags [109, 111].

4.5. HALO-tag

The HALO-tag is a type of self-labeling protein tag composed of 293 amino acids (Mw. ~33 kDa) and can be used to label GPCRs. It is a modified haloalkane dehalogenase enzyme derived from bacteria. It forms a strong covalent interaction with chloroalkanes, which can be attached to a fluorescent dye (Figure 3D) [112]. The binding reaction is stable, rapid, and irreversible under physiological conditions. Due to its bacterial origin, it does not undergo any non-specific interactions inside the cell [113]. However, because it is large, it requires a more extended, flexible linker peptide to fuse with GPCRs and reduce any interference it may cause with the functional dynamics of the receptor [113]. Chloroalkanes tagged with cell membrane-permeable dyes are often used to study the intracellular interactions of GPCRs within the cell, as chloroalkanes can easily cross the cell membrane [109, 114].

5. Total internal reflection fluorescence (TIRF) microscopy

Single-molecule TIRF microscopy uses the evanescent field generated from a laser excitation to excite immobilized samples. TIRF has several advantages: (i) it excites the fluorophores only at a penetration depth of ~100-150 nm, thereby reducing background noise, and (ii) it minimizes photobleaching of fluorophores in solution or photo-toxicity in cells by minimal excitation [115]. Two types of TIRF microscopes are typically used for single-molecule studies: Prism-based TIRF (pTIRF) (Figure 4C) and Objective-based TIRF (oTIRF) (Figure 4D). Both microscope setups provide a high signal-to-noise ratio (SNR), with the prism-based TIRF giving a better SNR due to the possibility of using higher laser powers. Since the pTIRF microscopes use open lasers, they are not commercially available; instead, they are usually custom-built by research labs. For a step-by-step guide on building a pTIRF setup for single-molecule microscopy, refer to [116-118] for more information. Also, it must be noted that pTIRF setups use expensive quartz slides instead of regular glass slides, which are used for oTIRF. The smTIRF movies are recorded by either a scientific Complementary Metal-Oxide-Semiconductor (sCMOS) camera [119] or an Electron-Magnified Charge-Coupled Device (EMCCD) camera [88]. These detectors are sensitive enough to detect a single fluorophore molecule at ~200 nm spatial resolution and ~20 - 30 ms temporal resolution, depending on the imaging conditions [120]. The TIRF movies are usually recorded until the fluorophores are bleached. In the case of live-cell microscopy, the imaging is performed until the cell viability is affected. (Figure 4C).

Figure 4. Illustration of TIRF microscopy.

Figure 4.

(A) Illustration of the sample preparation method for in vitro TIRF microscopy. (B) Sample preparation methods in a dish or microscope chamber for in vivo TIRF microscopy. (C) Schematic of a prism-based TIRF microscopy setup highlighting different components as well as the excitation path and an evanescent field. (D) Schematic of objective-based TIRF microscopy components and the excitation path with an evanescent field. (E) In vitro sample immobilization strategy using streptavidin and biotinylated antibody. (F) An example smFRET image of GPCR molecules with no ligand added in vitro with donor fluorescence spots (left) and acceptor fluorescence spots (right). (G) An example of an in vivo smTIRF image of a live cell with GPCR occurring as bright spots when activated (This figure was reproduced from the article Grimes et al. [108] published in Cell journal with permission. Figure created using BioRender.com.

5.1. Sample preparation and protein immobilization

One significant advantage to a single-molecule TIRF study is that it allows the monitoring of individual GPCR molecules and their dynamic behavior over time [23]. However, to visualize individual molecules, the receptors must be immobilized on microscope slides [121]. There are variations in immobilization techniques, but most involve the passivation of quartz slides with a mixture of biotin-PEG and unmodified PEG to prevent non-specific interactions with the slide while allowing biotin to serve as an anchor [117, 122]. The biotin on the PEG can then interact with neutravidin or streptavidin, acting as a bridge between biotin-conjugated antibodies specific for a tag on the receptor, or biotin-conjugated lipids in nanodiscs [23, 71, 117].

To prepare samples for single-molecule TIRF experiments, flow channels must be created on quartz or glass slides that have been drilled with two holes [71, 117]. The channel is prepared using double-sided tape, grease, and a glass coverslip. The samples are then injected through the hole on one end of the slide and allowed to flow out through the other end of the channel [71, 117] (Figures 4A and 4E).

5.2. Sample preparation for live cell TIRF microscopy

The development of novel DNA transfection methods in recent years has made it possible to study the dynamics of labeled GPCRs in live cell membranes. This is achieved by transfecting adherent cells with plasmids expressing GPCRs fused with fluorescent tags [66] or tetra-cysteine motifs [102]. The cells are then labeled with fluorescent dyes and imaged under a microscope. Since the over-expression of GPCRs on the membrane can give high background fluorescence, the receptors are either moderately expressed or minimally labeled to study their dynamics at the single-molecule level. The cells can then be grown directly in a dish with a coverslip for short-time imaging [123] or in special incubation chambers for long-time imaging with a TIRF microscope [107, 109] (Figures 4B and 4G).

5.3. smTIRF techniques

The SMF raw data obtained from smTIRF microscopy can be very informative. The oligomerization of GPCRs can be studied from the step-photobleaching patterns observed in fluorescence intensity traces [66, 86, 87]. The number of bleaching steps corresponds to the number of oligomers of the GPCRs localized in a region [86]. Similarly, based on the localization and tracking of the labeled GPCR spots typically on live cell membranes, the receptor dynamics during the activation process can be determined [85, 109]. Also, labeling GPCRs and their interacting proteins like G-proteins [85] and β-arrestins [109] with different fluorophores enables studies of the colocalization and dynamics between the proteins in vitro and live cells.

Further biophysical techniques have been developed by harnessing the photo-physics of novel fluorophores. For instance, the quantum yield of cyanine dyes can vary depending on the environment [124]. With this discovery, researchers have utilized this photophysical property to study the structural dynamics of biomolecules [125] through a technique called single-molecule protein-induced fluorescence enhancement (smPIFE). Recent studies have employed this technique to reveal previously unknown states and molecular kinetics during the activation process of GPCRs [23, 24, 43, 73]. Since only a single dye is required for this technique, it is easy to implement and is versatile for studying GPCRs.

Another important fluorescence microscopy technique that is widely used for studying the structural dynamics of GPCRs both in vitro and in live cells is single-molecule fluorescence resonance energy transfer (smFRET) [117]. This technique requires labeling GPCR samples with a unique pair of fluorescent dyes commonly known as a ‘FRET-pair’. In the ‘FRET-pair,’ the ‘donor-dye’ is excited by the laser beam, and the emitted fluorescence from the donor is absorbed by the ‘acceptor-dye’ depending on its proximity to the donor. This technique acts as a ‘molecular ruler’ [117] that allows variations in the emission intensities from both the donor and acceptor dyes to be detected, revealing the proximity dynamics of the labeled regions.

smFRET can be harnessed to reveal the dynamics of different regions of GPCRs, such as the TM-helices [126, 127], extracellular domains (ECDs) [128], and C-terminal domains [129] during the activation process. This technique has also been used to investigate the kinetics of GPCR dimerization [66, 107] and ligand binding [108]. This method can determine previously unknown hidden states in GPCR activation, which can aid in developing novel biosensors for drug screening (Figure 4F).

6. Applications

Single-molecule techniques have been applied to study class A, B, and C GPCRs in micelles, native-like, and native membrane environments [23, 88, 108]. Here, we will review some of the studies used to showcase the wide range of experimental techniques that can be conducted to study these receptors at the single molecule level, both in live cells and in vitro.

6.1. Sequential A2AAR TM dynamics in a native-like milieu

A recent application of single-molecule fluorescence experiments on a class A GPCR involved investigations on the conformational dynamics of the human A2A adenosine receptor (A2AAR) in native-like lipid environments. A2AAR has been the focus of many NMR spectroscopic studies and has become of great interest for single molecule studies. A receptor variant was created with a single cysteine (A289C) positioned on the intracellular surface of helix 7 [23]. (Figure 5). This variant was created for chemical conjugation with the environment-sensitive fluorescent dye Cy3. The receptor was expressed in Pichia pastoris and later solubilized and reconstituted into lipid nanodiscs with a molar ratio of 65:30:5 POPC, POPS, and biotinylated POPE, respectively [23, 71].

Figure 5. Sequential A2AAR TM dynamics in a native-like milieu.

Figure 5.

(A) Schematic representation of A2AAR-nanodisc immobilized to a PEGylated slide via biotin-streptavidin interactions. (B) Single-molecule time trajectories of apo- and agonist-bound A2AAR. Red lines indicate hidden Markov fitting of the raw traces (Black). (C) Representative transition density plots (TDPs) obtained from single–molecule fluorescence intensity trajectories in apo- and agonist-bound A2AAR. (D) Model of A2AAR TM helix VII dynamics during receptor activation. The figure is prepared from Wei et al. [23] with permission from Structure journal.

The nanodiscs containing the receptor labeled with Cy3 were immobilized on a microscope slide and observed using a TIRF microscope [23]. In these experiments, the conformational dynamics of A2AAR were observed in three different conditions - an unliganded (apo) condition, an antagonist-bound (ZM241385) condition, and an agonist-bound (NECA) condition. For the apo- and antagonist-bound A2AAR, there were predominately two intensity states observed - ~59% for state 1 and ~41% for state 2, with a small population transitioning to a higher intensity state, state 3. However, the agonist-bound A2AAR showed significant changes from the apo- and antagonist-bound receptors. The population of state 1 decreased (49%), state 2 increased (39%), and a third state emerged (12%) [23]. Each condition’s fluorescence intensity time trace showed a reversible temporal ordering of the fluorescence emission states. It was observed that the receptor does not transition directly from its inactive state to its active state, i.e., state 1 to state 3, but transitions through an intermediate state, state 2. Here, the single-molecule study has revealed new information upon receptor activation that had not been shown in previous crystal and cryo-EM structures [23].

6.2. Single-molecule visualization of glucagon receptor extracellular domain dynamics

Single-molecule studies of the class B glucagon receptor (GCGR) have recently been published, showing the dynamics of the extracellular domain (ECD) at the molecular level [88]. For class B GPCRs, two models have been proposed on the activation mechanism of peptide ligands that target and bind to the ECD and transmembrane domain binding pocket [130, 131]. Here, the authors showed the dynamics of the ECD of GCGR in the presence of the peptide ligand glucagon. For these experiments, GCGR with a C-terminal hexa-histidine tag (6xHis) and FLAG tag were expressed in HEK293T cells and solubilized into DDM-CHS detergent micelles. The receptor was then purified through Ni2+-NTA affinity chromatography and site-specifically labeled through chemical conjugation at C287 and C28 with a FRET donor (Alexa Fluor 555) and acceptor (Alexa Fluor 647) pair [88].

In the single-molecule FRET experiment, fluorophore-labeled GCGR in detergent micelles were immobilized on a microscope slide via biotin-streptavidin interactions using biotinylated anti-FLAG antibodies (Figure 6). The donor fluorophore was excited with a 532 nm laser, and emissions from both the donor and acceptor fluorophores were recorded in the presence and absence of glucagon (agonist) and MK0893 (antagonist) [88]. The FRET states of individual molecules were estimated from the donor and acceptor emissions and used to generate population FRET distributions of the molecules. In the absence of glucagon, two distinct states were observed. Based on the distance between the fluorophores, these states corresponded to the open and closed conformations of the ECD of GCGR. Upon the addition of glucagon, additional populations appeared on the FRET distributions of the population between the open and closed state FRET efficiencies, with lower FRET values. This data suggested that the ECD remains dynamic in the presence of glucagon, with multiple open states [88, 132]. Single-molecule experiments have proven helpful in understanding ECD dynamics and ligand recognition for class B GPCR activation and signaling.

Figure 6. Single-molecule visualization of glucagon receptor extracellular domain dynamics.

Figure 6.

The receptor (GCGR) was labeled with two different fluorophores – one on the extracellular face of the TM domain and the other on the ECD – and imaged using TIRF microscopy to monitor the dynamics of the ECD upon ligand binding. (A) Schematic of the prism TIRF-based smFRET setup used for imaging individual glucagon receptors. (B) Representative single-molecule fluorescence trajectory fora glucagon bound receptor. (C) Proposed model of ligand-induced GCGR ECD dynamics based on single-molecule fluorescence data. The figure is assembled from Liu et al. [87] with permission from JBC journal.

6.3. Dimerization dynamics of the glutamate receptor

In ensemble FRET experiments, mechanistic information like receptor conformations and transition states cannot be delineated for GPCRs. However, for class C GPCRs, smFRET has recently been applied to investigate the activation of metabotropic glutamate receptor 2 (mGluR2) receptors upon homodimerization and to study the effects of the cysteine-rich domain (CRD) that bridges the transmembrane domain of the GPCR with the extracellular Venus flytrap domain (VFT) [108]. This work used HEK293T cells to express GPCR with azi-CRD, an allosteric modulator of mGluR2 containing a C-terminal FLAG-tag. The cells were then labeled via azide-alkyne click chemistry with the FRET donor (Cy3 alkyne) and FRET acceptor (Cy5 alkyne). After cell lysis, the lysate was applied to the biotin-PEG surface of the slide and pulled down using a biotinylated anti-FLAG antibody, a method called single-molecule pull-down (SiMPull) [107, 108, 133].

In this study, they observed how negative allosteric modulators that bind to the transmembrane domain act to prevent glutamate-dependent activation of the receptor using MNI-137 [108, 134]. Through smFRET experiments, they found that intermediate and saturating concentrations of glutamate and MNI-137 had little effect on the active conformation of mGluR2 but increased the intermediate states 1 and 2 [108]. Transition density plots derived from Hidden Markov Model analysis of fluorescence time traces showed that saturating glutamate conditions resulted in transitions between state 2 and the active conformations of the CRD. However, in the presence of saturating glutamate conditions and MNI-137, the transitions occurred primarily between intermediate states 1 and 2, with very few transitioning to the active conformation of the CRD [108]. Thus, this study provides insights into how negative allosteric modulators influence the activation of class C GPCR homodimers at the single-molecule level (Figure 7).

Figure 7. Dimerization dynamics of the glutamate receptor.

Figure 7.

(A) Click chemistry labeling with an unnatural amino acid and immobilization of the glutamate receptor. (B) smTIRF image of the donor (left) and acceptor (right) molecules. Blue circles indicate selected single molecules of labeled glutamate receptors. (C) Population FRET histograms showing the effect of the modulator MNI137 (top: no modulator and bottom: 5 μM MNI137 added) on the dimerization dynamics of the glutamate receptor. (D) Transition density plots of the no modulator (top) and the 5 μM MNI137 (bottom) conditions. The figure is reproduced from the article Liauw et al. [107] with permission from eLife journal.

Additionally, it is worth noting that as a future direction, the application of TIRF-based single-molecule photobleaching techniques such as Native-nanoBleach [135] could prove to be very useful for studying how GPCRs are organized at nanoscale resolution. By directly visualizing and quantifying GPCR oligomerization in its native environment, this technique could potentially provide valuable insights into the oligomeric states of GPCRs under different ligand-binding conditions, mutations, or interactions with other proteins. Ultimately, this could lead to a deeper understanding of GPCR function and help identify new therapeutic approaches.

6.4. Single-molecule Tracking of GPCRs

In addition to using single-molecule imaging to study receptor conformations and dynamics, SMF also serves as a powerful tool for studying receptor interactions at the cellular level [109, 119]. It has been found that the fluidity of the cell membrane can affect the dynamic behavior of GPCRs. Recently, single-molecule tracking techniques have been applied to observe how β2 adrenergic receptor (β2AR) - β-arrestin2 (βArr2) interactions occur in the plasma membrane of live cells at a temporal resolution of ~30 ms and a spatial resolution of ~20 nm [109]. For these experiments, Chinese hamster ovary (CHO) cells, which have no detectable β2AR expression, were transiently transfected with SNAP-β2AR and Halo-βArr2 constructs and were labeled with saturating concentrations of the organic fluorophores - Alexa Fluor 647 and Janelia Fluor 549. Additionally, the clathrin-coated pits (CCPs) of the cell were also visualized through the transfection of GFP-tagged clathrin light chains. The cells were imaged using fast multicolor TIRF microscopy with single particle tracking [109].

The results of this single-molecule particle tracking study showed that βArr2 translocated to the plasma membrane before interacting with activated GPCR molecules. The diffusion of β2AR and βArr2 was heterogeneous. Upon analysis, they were observed to be freely diffusing, confined outside CCPs, and trapped inside CCPs [109]. The addition of isoproterenol, a β2AR full agonist, increased β2AR and βArr2 confinement to CCPs. The study showed that βArr2 spontaneously associates with the membrane, and the accumulation of β2AR and βArr2 to the CCPs is agonist-dependent [109]. The findings from the study demonstrate the usefulness of single-molecule particle tracking in understanding GPCR desensitization in live cells (Figure 8).

Figure 8. Single-molecule dynamics of β2AR and β-arrestins (βArr2) in live cells:

Figure 8.

(A) Illustration depicting the labeling strategy used to image β2AR and βArr2. (B) Objective-TIRF image of β2AR (green) and βArr2 (pink) tracks in live cells colocalized with tracks of clathrin-coated pits (CCPs). (C) Population proportions of β2AR (top) and βArr2 (bottom) exhibiting ‘freely diffusing’, ‘confined outside CCP’ and ‘trapped inside CCP’ dynamics. (D) Model showing the sequence of dynamic events observed during the arrestin-based desensitization of β2AR. The figure is adapted from the article Grimes et al. [108] with permission from Cell journal.

7. Concluding remarks

Single-molecule research on GPCRs will continue to provide valuable information on the real-time dynamics of these receptors and their response to different types of external signals, including ligands. Such studies can improve our understanding of the dynamics of GPCRs in their native environment, enabling the development of assays for high-throughput drug screening and drug development. The single-molecule microscopy techniques discussed in this context can also complement structural studies of the inactive and active states of GPCRs. Furthermore, SMF experiments can help identify key intermediate states that cannot be resolved using X-ray crystallography and cryo-EM methods.

Highlights.

  • Explores different approaches for expressing, purifying, and labeling GPCRs both in vitro and in live cell environments.

  • Discusses specific sample preparation methods for single-molecule fluorescence microscopy.

  • Examines the unique challenges involved in each stage of sample preparation for single-molecule GPCR experiments.

  • Highlights recent applications of single-molecule TIRF microscopy in investigating different classes of human GPCRs.

Acknowledgments

We thank Ms. Shushu Wei and Dr. Ting Liu for their discussions in writing this review article.

Funding

This work was supported by the National Institutes of Health R35GM142946 (R.L), University of Tennessee, Knoxville, and the National Institutes of Health T32 training Grants GM142621 (A.N.G.).

Funding Source

All sources of funding should also be acknowledged, and you should declare any involvement of study sponsors in the study design; collection, analysis and interpretation of data; the writing of the manuscript; the decision to submit the manuscript for publication. If the study sponsors had no such involvement, this should be stated.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Declaration of competing interest

The authors declare no competing interests.

Methods

Conflict of Interest Policy

Declarations

Methods requires that all authors sign a declaration of conflicting interests. If you have nothing to declare in any of these categories, then this should be stated.

Conflict of Interest

A conflicting interest exists when professional judgment concerning a primary interest (such as patient’s welfare or the validity of research) may be influenced by a secondary interest (such as financial gain or personal rivalry). It may arise for the authors when they have financial interest that may influence their interpretation of their results or those of others. Examples of potential conflicts of interest include employment, consultancies, stock ownership, honoraria, paid expert testimony, patent applications/registrations, and grants or other funding.

Please state any competing interests

None

11. References

  • [1].Weis WI, Kobilka BK, The Molecular Basis of G Protein–Coupled Receptor Activation, Annual Review of Biochemistry 87(1) (2018) 897–919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Morris AJ, Malbon CC, Physiological Regulation of G Protein-Linked Signaling, Physiological Reviews 79(4) (1999) 1373–1430. [DOI] [PubMed] [Google Scholar]
  • [3].Zhao J, Deng Y, Jiang Z, Qing H, G Protein-Coupled Receptors (GPCRs) in Alzheimer’s Disease: A Focus on BACE1 Related GPCRs, Frontiers in Aging Neuroscience 8 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].Boczek T, Mackiewicz J, Sobolczyk M, Wawrzyniak J, Lisek M, Ferenc B, Guo F, Zylinska L, The Role of G Protein-Coupled Receptors (GPCRs) and Calcium Signaling in Schizophrenia. Focus on GPCRs Activated by Neurotransmitters and Chemokines, Cells 10(5) (2021) 1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [5].Lappano R, Maggiolini M, G protein-coupled receptors: novel targets for drug discovery in cancer, Nature Reviews Drug Discovery 10(1) (2011) 47–60. [DOI] [PubMed] [Google Scholar]
  • [6].Stevens RC, Cherezov V, Katritch V, Abagyan R, Kuhn P, Rosen H, Wüthrich K, The GPCR Network: a large-scale collaboration to determine human GPCR structure and function, Nature Reviews Drug Discovery 12(1) (2013) 25–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Bjarnadóttir TK, Gloriam DE, Hellstrand SH, Kristiansson H, Fredriksson R, Schiöth HB, Comprehensive repertoire and phylogenetic analysis of the G protein-coupled receptors in human and mouse, Genomics 88(3) (2006) 263–273. [DOI] [PubMed] [Google Scholar]
  • [8].Wootten D, Christopoulos A, Marti-Solano M, Babu MM, Sexton PM, Mechanisms of signalling and biased agonism in G protein-coupled receptors, Nature Reviews Molecular Cell Biology 19(10) (2018) 638–653. [DOI] [PubMed] [Google Scholar]
  • [9].Mafi A, Kim S-K, Goddard WA, The mechanism for ligand activation of the GPCR–G protein complex, Proceedings of the National Academy of Sciences 119(18) (2022) e2110085119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Manglik A, Kruse AC, Kobilka TS, Thian FS, Mathiesen JM, Sunahara RK, Pardo L, Weis WI, Kobilka BK, Granier S, Crystal structure of the μ-opioid receptor bound to a morphinan antagonist, Nature 485(7398) (2012) 321–326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [11].Cherezov V, Rosenbaum DM, Hanson MA, Rasmussen SGF, Thian FS, Kobilka TS, Choi H-J, Kuhn P, Weis WI, Kobilka BK, Stevens RC, High-Resolution Crystal Structure of an Engineered Human β2-Adrenergic G Protein-Coupled Receptor, Science 318(5854) (2007) 1258–1265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Jaakola V-P, Griffith MT, Hanson MA, Cherezov V, Chien EYT, Lane JR, IJzerman AP, Stevens RC, The 2.6 Angstrom Crystal Structure of a Human A2A Adenosine Receptor Bound to an Antagonist, Science 322(5905) (2008) 1211–1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [13].Tsai C-J, Pamula F, Nehmé R, Mühle J, Weinert T, Flock T, Nogly P, Edwards PC, Carpenter B, Gruhl T, Ma P, Deupi X, Standfuss J, Tate CG, Schertler GFX, Crystal structure of rhodopsin in complex with a mini-Go sheds light on the principles of G protein selectivity, Science Advances 4(9) (2018) eaat7052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [14].Gao Y, Robertson MJ, Rahman SN, Seven AB, Zhang C, Meyerowitz JG, Panova O, Hannan FM, Thakker RV, Bräuner-Osborne H, Mathiesen JM, Skiniotis G, Asymmetric activation of the calcium-sensing receptor homodimer, Nature 595(7867) (2021)455–459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].García-Nafría J, Nehmé R, Edwards PC, Tate CG, Cryo-EM structure of the serotonin 5-HT1B receptor coupled to heterotrimeric Go, Nature 558(7711) (2018) 620–623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Lee Y, Warne T, Nehmé R, Pandey S, Dwivedi-Agnihotri H, Chaturvedi M, Edwards PC, García-Nafría J, Leslie AGW, Shukla AK, Tate CG, Molecular basis of β-arrestin coupling to formoterol-bound β1-adrenoceptor, Nature 583(7818) (2020) 862–866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Kato HE, Zhang Y, Hu H, Suomivuori C-M, Kadji FMN, Aoki J, Krishna Kumar K, Fonseca R, Hilger D, Huang W, Latorraca NR, Inoue A, Dror RO, Kobilka BK, Skiniotis G, Conformational transitions of a neurotensin receptor 1–Gi1 complex, Nature 572(7767) (2019) 80–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [18].Seven AB, Barros-Álvarez X, de Lapeyrière M, Papasergi-Scott MM, Robertson MJ, Zhang C, Nwokonko RM, Gao Y, Meyerowitz JG, Rocher J-P, Schelshorn D, Kobilka BK, Mathiesen JM, Skiniotis G, G-protein activation by a metabotropic glutamate receptor, Nature 595(7867) (2021) 450–454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Casiraghi M, Point E, Pozza A, Moncoq K, Banères J-L, Catoire LJ, NMR analysis of GPCR conformational landscapes and dynamics, Molecular and Cellular Endocrinology 484 (2019) 69–77. [DOI] [PubMed] [Google Scholar]
  • [20].Wu F-J, Williams LM, Abdul-Ridha A, Gunatilaka A, Vaid TM, Kocan M, Whitehead AR, Griffin MDW, Bathgate RAD, Scott DJ, Gooley PR, Probing the correlation between ligand efficacy and conformational diversity at the α1A-adrenoreceptor reveals allosteric coupling of its microswitches, Journal of Biological Chemistry 295(21) (2020) 7404–7417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Ye L, Van Eps N, Zimmer M, Ernst OP, Scott Prosser R, Activation of the A2A adenosine G-protein-coupled receptor by conformational selection, Nature 533(7602) (2016) 265–268. [DOI] [PubMed] [Google Scholar]
  • [22].Sušac L, Eddy MT, Didenko T, Stevens RC, Wüthrich K, A2A adenosine receptor functional states characterized by 19F-NMR, Proceedings of the National Academy of Sciences 115(50) (2018) 12733–12738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [23].Wei S, Thakur N, Ray AP, Jin B, Obeng S, McCurdy CR, McMahon LR, Gutiérrez-de-Terán H, Eddy MT, Lamichhane R, Slow conformational dynamics of the human A2A adenosine receptor are temporally ordered, Structure 30(3) (2022) 329–337.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [24].Lamichhane R, Liu JJ, White KL, Katritch V, Stevens RC, Wüthrich K, Millar DP, Biased Signaling of the G-Protein-Coupled Receptor β2AR Is Governed by Conformational Exchange Kinetics, Structure 28(3) (2020) 371–377.e3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Gregorio GG, Masureel M, Hilger D, Terry DS, Juette M, Zhao H, Zhou Z, Perez-Aguilar JM, Hauge M, Mathiasen S, Javitch JA, Weinstein H, Kobilka BK, Blanchard SC, Single-molecule analysis of ligand efficacy in β2AR–G-protein activation, Nature 547(7661) (2017) 68–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Fernandes DD, Neale C, Gomes G-NW, Li Y, Malik A, Pandey A, Orazietti AP, Wang X, Ye L, Scott Prosser R, Gradinaru CC, Ligand modulation of the conformational dynamics of the A2A adenosine receptor revealed by single-molecule fluorescence, Scientific Reports 11(1) (2021) 5910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Dror RO, Arlow DH, Maragakis P, Mildorf TJ, Pan AC, Xu H, Borhani DW, Shaw DE, Activation mechanism of the β2-adrenergic receptor, Proceedings of the National Academy of Sciences 108(46) (2011) 18684–18689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [28].Nygaard R, Zou Y, Dror Ron O., Mildorf Thomas J., Arlow Daniel H., Manglik A, Pan Albert C., Liu Corey W., Fung Juan J., Bokoch Michael P., Thian Foon S., Kobilka Tong S., Shaw David E., Mueller L, Prosser RS, Kobilka Brian K., The Dynamic Process of β2-Adrenergic Receptor Activation, Cell 152(3) (2013) 532–542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Avlani VA, Gregory KJ, Morton CJ, Parker MW, Sexton PM, Christopoulos A, Critical Role for the Second Extracellular Loop in the Binding of Both Orthosteric and Allosteric G Protein-coupled Receptor Ligands*, Journal of Biological Chemistry 282(35) (2007) 25677–25686. [DOI] [PubMed] [Google Scholar]
  • [30].Huber T, Botelho AV, Beyer K, Brown MF, Membrane Model for the G-Protein-Coupled Receptor Rhodopsin: Hydrophobic Interface and Dynamical Structure, Biophysical Journal 86(4) (2004) 2078–2100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Tian H, Fürstenberg A, Huber T, Labeling and Single-Molecule Methods To Monitor G Protein-Coupled Receptor Dynamics, Chemical Reviews 117(1) (2017) 186–245. [DOI] [PubMed] [Google Scholar]
  • [32].Bertheleme N, Singh S, Dowell S, Byrne B, Chapter Seven - Heterologous Expression of G-Protein-Coupled Receptors in Yeast, in: Shukla AK (Ed.), Methods Enzymol, Academic Press; 2015, pp. 141–164. [DOI] [PubMed] [Google Scholar]
  • [33].Hofmann KP, Scheerer P, Hildebrand PW, Choe H-W, Park JH, Heck M, Ernst OP, A G protein-coupled receptor at work: the rhodopsin model, Trends in Biochemical Sciences 34(11) (2009) 540–552. [DOI] [PubMed] [Google Scholar]
  • [34].Rasmussen SGF, DeVree BT, Zou Y, Kruse AC, Chung KY, Kobilka TS, Thian FS, Chae PS, Pardon E, Calinski D, Mathiesen JM, Shah STA, Lyons JA, Caffrey M, Gellman SH, Steyaert J, Skiniotis G, Weis WI, Sunahara RK, Kobilka BK, Crystal structure of the β2 adrenergic receptor–Gs protein complex, Nature 477(7366) (2011)549–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [35].Perez-Aguilar JM, Xi J, Matsunaga F, Cui X, Selling B, Saven JG, Liu R, A Computationally Designed Water-Soluble Variant of a G-Protein-Coupled Receptor: The Human Mu Opioid Receptor, PLOS ONE 8(6) (2013) e66009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [36].Xi J, Yang N, Perez-Aguilar JM, Selling B, Grothusen JR, Lamichhane R, Saven JG, Liu R, Novel variants of engineered water soluble mu opioid receptors with extensive mutations and removal of cysteines, Proteins: Structure, Function, and Bioinformatics 89(10) (2021) 1386–1393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Klammt C, Schwarz D, Eifler N, Engel A, Piehler J, Haase W, Hahn S, Dötsch V, Bernhard F, Cell-free production of G protein-coupled receptors for functional and structural studies, Journal of Structural Biology 158(3) (2007) 482–493. [DOI] [PubMed] [Google Scholar]
  • [38].Luginina A, Maslov I, Khorn P, Volkov O, Khnykin A, Kuzmichev P, Shevtsov M, Belousov A, Kapranov I, Dashevskii D, Kornilov D, Bestsennaia E, Hofkens J, Hendrix J, Gensch T, Cherezov V, Ivanovich V, Mishin A, Borshchevskiy V, Functional GPCR Expression in Eukaryotic LEXSY System, Journal of Molecular Biology 435(23) (2023) 168310. [DOI] [PubMed] [Google Scholar]
  • [39].Milić D, Veprintsev DB, Large-scale production and protein engineering of G protein-coupled receptors for structural studies, Frontiers in Pharmacology 6 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [40].Moss B, Vaccinia Virus: A Tool for Research and Vaccine Development, Science 252(5013) (1991) 1662–1667. [DOI] [PubMed] [Google Scholar]
  • [41].Sarramegna V, Talmont F, Demange P, Milon A, Heterologous expression of G-protein-coupled receptors: comparison of expression systems from the standpointof large-scale production and purification, Cellular and Molecular Life Sciences CMLS 60(8) (2003) 1529–1546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [42].Liljeström P, Garoff H, A New Generation of Animal Cell Expression Vectors Based on the Semliki Forest Virus Replicon, Bio/Technology 9(12) (1991) 1356–1361. [DOI] [PubMed] [Google Scholar]
  • [43].Wei S, Pour NG, Tiruvadi-Krishnan S, Ray AP, Thakur N, Eddy MT, Lamichhane R, Single-molecule visualization of human A2A adenosine receptor activation by a G protein and constitutively activating mutations, Communications Biology 6(1) (2023) 1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Shiroishi M, Kobayashi T, Ogasawara S, Tsujimoto H, Ikeda-Suno C, Iwata S, Shimamura T, Production of the stable human histamine H1 receptor in Pichia pastoris for structural determination, Methods 55(4) (2011) 281–286. [DOI] [PubMed] [Google Scholar]
  • [45].Yurugi-Kobayashi T, Asada H, Shiroishi M, Shimamura T, Funamoto S, Katsuta N, Ito K, Sugawara T, Tokuda N, Tsujimoto H, Murata T, Nomura N, Haga K, Haga T, Iwata S, Kobayashi T, Comparison of functional non-glycosylated GPCRs expression in Pichia pastoris, Biochemical and Biophysical Research Communications 380(2) (2009) 271–276. [DOI] [PubMed] [Google Scholar]
  • [46].Hino T, Arakawa T, Iwanari H, Yurugi-Kobayashi T, Ikeda-Suno C, Nakada-Nakura Y, Kusano-Arai O, Weyand S, Shimamura T, Nomura N, Cameron AD, Kobayashi T, Hamakubo T, Iwata S, Murata T, G-protein-coupled receptor inactivation by an allosteric inverse-agonist antibody, Nature 482(7384) (2012) 237–240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [47].Thakur N, Ray AP, Sharp L, Jin B, Duong A, Pour NG, Obeng S, Wijesekara AV, Gao Z-G, McCurdy CR, Jacobson KA, Lyman E, Eddy MT, Anionic phospholipids control mechanisms of GPCR-G protein recognition, Nature Communications 14(1) (2023) 794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Ficca AG, Testa L, Tocchini Valentini GP, The human β 2-adrenergic receptor expressed in Schizosaccharomyces pombe retains its pharmacological properties, FEBS Letters 377(2) (1995) 140–144. [DOI] [PubMed] [Google Scholar]
  • [49].Croft W, Hill C, McCann E, Bond M, Esparza-Franco M, Bennett J, Rand D, Davey J, Ladds G, A Physiologically Required G Protein-coupled Receptor (GPCR)-Regulator of G Protein Signaling (RGS) Interaction That Compartmentalizes RGS Activity*, Journal of Biological Chemistry 288(38) (2013) 27327–27342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [50].Chung K-S, Kim D-U, Ryoo S-W, Kang E-J, Won M, Kim L, Jang Y-J, Maeng P-J, Kim S-C, Yoo H-S, Hoe K-L, Functional over-expression of the Stm1 protein, a G-protein-coupled receptor, in Schizosaccharomyces pombe, Biotechnology Letters 25(3) (2003) 267–272. [DOI] [PubMed] [Google Scholar]
  • [51].Weston C, Poyner D, Patel V, Dowell S, Ladds G, Investigating G protein signalling bias at the glucagon-like peptide-1 receptor in yeast, Br J Pharmacol 171(15) (2014) 3651–3665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [52].Shiroishi M, Tsujimoto H, Makyio H, Asada H, Yurugi-Kobayashi T, Shimamura T, Murata T, Nomura N, Haga T, Iwata S, Kobayashi T, Platform for the rapid construction and evaluation of GPCRs for crystallography in Saccharomyces cerevisiae, Microbial Cell Factories 11(1) (2012) 78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Wiseman DN, Otchere A, Patel JH, Uddin R, Pollock NL, Routledge SJ, Rothnie AJ, Slack C, Poyner DR, Bill RM, Goddard AD, Expression and purification of recombinant G protein-coupled receptors: A review, Protein Expression and Purification 167 (2020) 105524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [54].Hirz M, Richter G, Leitner E, Wriessnegger T, Pichler H, A novel cholesterol-producing Pichia pastoris strain is an ideal host for functional expression of human Na,K-ATPase α3β1 isoform, Applied Microbiology and Biotechnology 97(21) (2013) 9465–9478. [DOI] [PubMed] [Google Scholar]
  • [55].Wang Y, Zhuang Y, DiBerto JF, Zhou XE, Schmitz GP, Yuan Q, Jain MK, Liu W, Melcher K, Jiang Y, Roth BL, Xu HE, Structures of the entire human opioid receptor family, Cell 186(2) (2023) 413–427.e17. [DOI] [PubMed] [Google Scholar]
  • [56].Du Y, Due NM, Rasmussen SGF, Hilger D, Kubiak X, Wang L, Bohon J, Kim HR, Wegrecki M, Asuru A, Jeong KM, Lee J, Chance MR, Lodowski DT, Kobilka BK, Chung KY, Assembly of a GPCR-G Protein Complex, Cell 177(5) (2019) 1232–1242.e11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [57].Rosenbaum DM, Cherezov V, Hanson MA, Rasmussen SGF, Thian FS, Kobilka TS, Choi H-J, Yao X-J, Weis WI, Stevens RC, Kobilka BK, GPCR Engineering Yields High-Resolution Structural Insights into β2-Adrenergic Receptor Function, Science 318(5854) (2007) 1266–1273. [DOI] [PubMed] [Google Scholar]
  • [58].Contreras-Gómez A, Sánchez-Mirón A, García-Camacho F, Molina-Grima E, Chisti Y, Protein production using the baculovirus-insect cell expression system, Biotechnology Progress 30(1) (2014) 1–18. [DOI] [PubMed] [Google Scholar]
  • [59].Saarenpää T, Jaakola V-P, Goldman A, Chapter Nine - Baculovirus-Mediated Expression of GPCRs in Insect Cells, in: Shukla AK (Ed.), Methods Enzymol, Academic Press; 2015, pp. 185–218. [DOI] [PubMed] [Google Scholar]
  • [60].Smith GE, Fraser MJ, Summers MD, Molecular Engineering of the Autographa californica Nuclear Polyhedrosis Virus Genome: Deletion Mutations Within the Polyhedrin Gene, Journal of Virology 46(2) (1983) 584–593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [61].Zhang H, Qiao A, Yang L, Van Eps N, Frederiksen KS, Yang D, Dai A, Cai X, Zhang H, Yi C, Cao C, He L, Yang H, Lau J, Ernst OP, Hanson MA, Stevens RC, Wang M-W, Reedtz-Runge S, Jiang H, Zhao Q, Wu B, Structure of the glucagon receptor in complex with a glucagon analogue, Nature 553(7686) (2018) 106–110. [DOI] [PubMed] [Google Scholar]
  • [62].Doré AS, Okrasa K, Patel JC, Serrano-Vega M, Bennett K, Cooke RM, Errey JC, Jazayeri A, Khan S, Tehan B, Weir M, Wiggin GR, Marshall FH, Structure of class C GPCR metabotropic glutamate receptor 5 transmembrane domain, Nature 511(7511) (2014) 557–562. [DOI] [PubMed] [Google Scholar]
  • [63].Hollenstein K, Kean J, Bortolato A, Cheng RKY, Doré AS, Jazayeri A, Cooke RM, Weir M, Marshall FH, Structure of class B GPCR corticotropin-releasing factor receptor 1, Nature 499(7459) (2013) 438–443. [DOI] [PubMed] [Google Scholar]
  • [64].Bean BDM, Mulvihill CJ, Garge RK, Boutz DR, Rousseau O, Floyd BM, Cheney W, Gardner EC, Ellington AD, Marcotte EM, Gollihar JD, Whiteway M, Martin VJJ, Functional expression of opioid receptors and other human GPCRs in yeast engineered to produce human sterols, Nature Communications 13(1) (2022) 2882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Shimamura T, Shiroishi M, Weyand S, Tsujimoto H, Winter G, Katritch V, Abagyan R, Cherezov V, Liu W, Han GW, Kobayashi T, Stevens RC, Iwata S, Structure of the human histamine H1 receptor complex with doxepin, Nature 475(7354) (2011)65–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [66].Asher WB, Geggier P, Holsey MD, Gilmore GT, Pati AK, Meszaros J, Terry DS, Mathiasen S, Kaliszewski MJ, McCauley MD, Govindaraju A, Zhou Z, Harikumar KG, Jaqaman K, Miller LJ, Smith AW, Blanchard SC, Javitch JA, Single-molecule FRET imaging of GPCR dimers in living cells, Nature Methods 18(4) (2021) 397–405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [67].Anton SE, Kayser C, Maiellaro I, Nemec K, Möller J, Koschinski A, Zaccolo M, Annibale P, Falcke M, Lohse MJ, Bock A, Receptor-associated independent cAMP nanodomains mediate spatiotemporal specificity of GPCR signaling, Cell 185(7) (2022) 1130–1142.e11. [DOI] [PubMed] [Google Scholar]
  • [68].Seddon AM, Curnow P, Booth PJ, Membrane proteins, lipids and detergents: not just a soap opera, Biochimica et Biophysica Acta (BBA) - Biomembranes 1666(1) (2004) 105–117. [DOI] [PubMed] [Google Scholar]
  • [69].Arachea BT, Sun Z, Potente N, Malik R, Isailovic D, Viola RE, Detergent selection for enhanced extraction of membrane proteins, Protein Expression and Purification 86(1) (2012) 12–20. [DOI] [PubMed] [Google Scholar]
  • [70].Lee S, Ghosh S, Jana S, Robertson N, Tate CG, Vaidehi N, How Do Branched Detergents Stabilize GPCRs in Micelles?, Biochemistry 59(23) (2020) 2125–2134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Thakur N, Wei S, Ray AP, Lamichhane R, Eddy MT, Production of human A2AAR in lipid nanodiscs for 19F-NMR and single-molecule fluorescence spectroscopy, STAR Protocols 3(3) (2022) 101535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [72].Jamshad M, Charlton J, Lin Y-P, Routledge Sarah J., Bawa Z, Knowles Timothy J., Overduin M, Dekker N, Dafforn Tim R., Bill Roslyn M., Poyner David R., Wheatley M, G-protein coupled receptor solubilization and purification for biophysical analysis and functional studies, in the total absence of detergent, Bioscience Reports 35(2) (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Lamichhane R, Liu JJ, Pljevaljcic G, White KL, van der Schans E, Katritch V, Stevens RC, Wüthrich K, Millar DP, Single-molecule view of basal activity and activation mechanisms of the G protein-coupled receptor β2AR, Proceedings of the National Academy of Sciences 112(46) (2015) 14254–14259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [74].Ritchie TK, Grinkova YV, Bayburt TH, Denisov IG, Zolnerciks JK, Atkins WM, Sligar SG, Chapter Eleven - Reconstitution of Membrane Proteins in Phospholipid Bilayer Nanodiscs, in: Düzgünes N (Ed.), Methods Enzymol, Academic Press; 2009, pp. 211–231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [75].Dominik Pawel K., Borowska Marta T., Dalmas O, Kim Sangwoo S., Perozo E, Keenan Robert J., Kossiakoff Anthony A., Conformational Chaperones for Structural Studies of Membrane Proteins Using Antibody Phage Display with Nanodiscs, Structure 24(2) (2016) 300–309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [76].Bayburt TH, Sligar SG, Membrane protein assembly into Nanodiscs, FEBS Letters 584(9) (2010) 1721–1727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [77].Ray AP, Thakur N, Pour NG, Eddy MT, Dual mechanisms of cholesterol-GPCR interactions that depend on membrane phospholipid composition, Structure 31(7) (2023) 836–847.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Tedesco D, Maj M, Malarczyk P, Cingolani A, Zaffagnini M, Wnorowski A, Czapiński J, Benelli T, Mazzoni R, Bartolini M, Jóźwiak K, Application of the SMALP technology to the isolation of GPCRs from low-yielding cell lines, Biochimica et Biophysica Acta (BBA) - Biomembranes 1863(9) (2021) 183641. [DOI] [PubMed] [Google Scholar]
  • [79].Harwood CR, Sykes DA, Hoare BL, Heydenreich FM, Uddin R, Poyner DR, Briddon SJ, Veprintsev DB, Functional solubilization of the β2-adrenoceptor using diisobutylene maleic acid, iScience 24(12) (2021) 103362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [80].Wheatley M, Charlton J, Jamshad M, Routledge Sarah J., Bailey S, La-Borde Penelope J., Azam Maria T., Logan Richard T., Bill Roslyn M., Dafforn Tim R., Poyner David R., GPCR–styrene maleic acid lipid particles (GPCR–SMALPs): their nature and potential, Biochemical Society Transactions 44(2) (2016) 619–623. [DOI] [PubMed] [Google Scholar]
  • [81].Gulamhussein AA, Uddin R, Tighe BJ, Poyner DR, Rothnie AJ, A comparison of SMA (styrene maleic acid) and DIBMA (di-isobutylene maleic acid) for membrane protein purification, Biochimica et Biophysica Acta (BBA) - Biomembranes 1862(7) (2020) 183281. [DOI] [PubMed] [Google Scholar]
  • [82].Grime RL, Goulding J, Uddin R, Stoddart LA, Hill SJ, Poyner DR, Briddon SJ, Wheatley M, Single molecule binding of a ligand to a G-protein-coupled receptor in real time using fluorescence correlation spectroscopy, rendered possible by nano-encapsulation in styrene maleic acid lipid particles, Nanoscale 12(21) (2020) 11518–11525. [DOI] [PubMed] [Google Scholar]
  • [83].Routledge SJ, Jamshad M, Little HA, Lin Y-P, Simms J, Thakker A, Spickett CM, Bill RM, Dafforn TR, Poyner DR, Wheatley M, Ligand-induced conformational changes in a SMALP-encapsulated GPCR, Biochimica et Biophysica Acta (BBA) - Biomembranes 1862(6) (2020) 183235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].Simon KS, Pollock NL, Lee SC, Membrane protein nanoparticles: the shape of things to come, Biochemical Society Transactions 46(6) (2018) 1495–1504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Sungkaworn T, Jobin M-L, Burnecki K, Weron A, Lohse MJ, Calebiro D, Single-molecule imaging reveals receptor–G protein interactions at cell surface hot spots, Nature 550(7677) (2017) 543–547. [DOI] [PubMed] [Google Scholar]
  • [86].Kasai RS, Ito SV, Awane RM, Fujiwara TK, Kusumi A, The Class-A GPCR Dopamine D2 Receptor Forms Transient Dimers Stabilized by Agonists: Detection by Single-Molecule Tracking, Cell Biochemistry and Biophysics 76(1) (2018) 29–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Vasudevan L, Borroto-Escuela DO, Huysentruyt J, Fuxe K, Saini DK, Stove C, Heterodimerization of Mu Opioid Receptor Protomer with Dopamine D2 Receptor Modulates Agonist-Induced Internalization of Mu Opioid Receptor, Biomolecules 9(8) (2019) 368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [88].Liu T, Khanal S, Hertslet GD, Lamichhane R, Single-molecule analysis reveals that a glucagon-bound extracellular domain of the glucagon receptor is dynamic, Journal of Biological Chemistry 299(9) (2023) 105160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].Granier S, Kim S, Shafer AM, Ratnala VRP, Fung JJ, Zare RN, Kobilka B, Structure and Conformational Changes in the C-terminal Domain of the β2-Adrenoceptor: INSIGHTS FROM FLUORESCENCE RESONANCE ENERGY TRANSFER STUDIES*, Journal of Biological Chemistry 282(18) (2007) 13895–13905. [DOI] [PubMed] [Google Scholar]
  • [90].Milstein JN, Nino DF, Zhou X, Gradinaru CC, Single-molecule counting applied to the study of GPCR oligomerization, Biophysical Journal 121(17) (2022) 3175–3187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [91].Gormal RS, Padmanabhan P, Kasula R, Bademosi AT, Coakley S, Giacomotto J, Blum A, Joensuu M, Wallis TP, Lo HP, Budnar S, Rae J, Ferguson C, Bastiani M, Thomas WG, Pardon E, Steyaert J, Yap AS, Goodhill GJ, Hilliard MA, Parton RG, Meunier FA, Modular transient nanoclustering of activated β2-adrenergic receptors revealed by single-molecule tracking of conformation-specific nanobodies, Proceedings of the National Academy of Sciences 117(48) (2020) 30476–30487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Vasudevan NT, Chapter 3 - cAMP assays in GPCR drug discovery, in: Shukla AK (Ed.), Methods in Cell Biology, Academic Press; 2017, pp. 51–57. [DOI] [PubMed] [Google Scholar]
  • [93].Salahpour A, Espinoza S, Masri B, Lam V, Barak L, Gainetdinov R, BRET biosensors to study GPCR biology, pharmacology, and signal transduction, Frontiers in Endocrinology 3 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Flanagan CA, Chapter 10 - GPCR-radioligand binding assays, in: Shukla AK (Ed.), Methods in Cell Biology, Academic Press; 2016, pp. 191–215. [DOI] [PubMed] [Google Scholar]
  • [95].Casadó-Anguera V, Moreno E, Mallol J, Ferré S, Canela EI, Cortés A, Casadó V, Reinterpreting anomalous competitive binding experiments within G protein-coupled receptor homodimers using a dimer receptor model, Pharmacological Research 139 (2019) 337–347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Chin JW, Santoro SW, Martin AB, King DS, Wang L, Schultz PG, Addition of p-Azido-l-phenylalanine to the Genetic Code of Escherichia coli, Journal of the American Chemical Society 124(31) (2002) 9026–9027. [DOI] [PubMed] [Google Scholar]
  • [97].Ye S, Köhrer C, Huber T, Kazmi M, Sachdev P, Yan ECY, Bhagat A, RajBhandary UL, Sakmar TP, Site-specific Incorporation of Keto Amino Acids into Functional G Protein-coupled Receptors Using Unnatural Amino Acid Mutagenesis*, Journal of Biological Chemistry 283(3) (2008) 1525–1533. [DOI] [PubMed] [Google Scholar]
  • [98].Shah UH, Toneatti R, Gaitonde SA, Shin JM, González-Maeso J, Site-Specific Incorporation of Genetically Encoded Photo-Crosslinkers Locates the Heteromeric Interface of a GPCR Complex in Living Cells, Cell Chemical Biology 27(10) (2020) 1308–1317.e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [99].Ovaa H, wals k., Unnatural amino acid incorporation in E. coli: current and future applications in the design of therapeutic proteins, Frontiers in Chemistry 2 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Tir N, Heistinger L, Grünwald-Gruber C, Jakob LA, Dickgiesser S, Rasche N, Mattanovich D, From strain engineering to process development: monoclonal antibody production with an unnatural amino acid in Pichia pastoris, Microbial Cell Factories 21(1) (2022) 157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Coin I, Katritch V, Sun T, Xiang Z, Siu Fai Y., Beyermann M, Stevens Raymond C., Wang L, Genetically Encoded Chemical Probes in Cells Reveal the Binding Path of Urocortin-I to CRF Class B GPCR, Cell 155(6) (2013) 1258–1269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Hoffmann C, Gaietta G, Bünemann M, Adams SR, Oberdorff-Maass S, Behr B, Vilardaga J-P, Tsien RY, Ellisman MH, Lohse MJ, A FlAsH-based FRET approach to determine G protein–coupled receptor activation in living cells, Nature Methods 2(3) (2005) 171–176. [DOI] [PubMed] [Google Scholar]
  • [103].Han M-J, He Q.-t., Yang M, Chen C, Yao Y, Liu X, Wang Y, Zhu Z.-l., Zhu K.-k., Qu C, Yang F, Hu C, Guo X, Zhang D, Chen C, Sun J.-p., Wang J, Single-molecule FRET and conformational analysis of beta-arrestin-1 through genetic code expansion and a Se-click reaction, Chemical Science 12(26) (2021) 9114–9123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Han M-J, He Q.-t., Yang M, Chen C, Yao Y, Liu X, Wang Y, Zhu Z.-l., Zhu K.-k., Qu C, Yang F, Hu C, Guo X, Zhang D, Chen C, Sun J.-p., Wang J, Correction: Single-molecule FRET and conformational analysis of beta-arrestin-1 through genetic code expansion and a Se-click reaction, Chemical Science 12(28) (2021) 9851–9851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [105].Stroffekova K, Proenza C, Beam KG, The protein-labeling reagent FLASH-EDT2 binds not only to CCXXCC motifs but also non-specifically to endogenous cysteine-rich proteins, Pflügers Archiv 442(6) (2001) 859–866. [DOI] [PubMed] [Google Scholar]
  • [106].Keppler A, Gendreizig S, Gronemeyer T, Pick H, Vogel H, Johnsson K, A general method for the covalent labeling of fusion proteins with small molecules in vivo, Nature Biotechnology 21(1) (2003) 86–89. [DOI] [PubMed] [Google Scholar]
  • [107].Liauw BW-H, Afsari HS, Vafabakhsh R, Conformational rearrangement during activation of a metabotropic glutamate receptor, Nature Chemical Biology 17(3) (2021) 291–297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [108].Liauw BW-H, Foroutan A, Schamber MR, Lu W, Samareh Afsari H, Vafabakhsh R, Conformational fingerprinting of allosteric modulators in metabotropic glutamate receptor 2, eLife 11 (2022) e78982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Grimes J, Koszegi Z, Lanoiselée Y, Miljus T, O’Brien SL, Stepniewski TM, Medel-Lacruz B, Baidya M, Makarova M, Mistry R, Goulding J, Drube J, Hoffmann C, Owen DM, Shukla AK, Selent J, Hill SJ, Calebiro D, Plasma membrane preassociation drives β-arrestin coupling to receptors and activation, Cell 186(10) (2023) 2238–2255.e20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Gautier A, Juillerat A, Heinis C, Corrêa IR, Kindermann M, Beaufils F, Johnsson K, An Engineered Protein Tag for Multiprotein Labeling in Living Cells, Chemistry & Biology 15(2) (2008) 128–136. [DOI] [PubMed] [Google Scholar]
  • [111].Ward RJ, Pediani JD, Milligan G, Ligand-induced internalization of the orexin OX1 and cannabinoid CB1 receptors assessed via N-terminal SNAP and CLIP-tagging, Br J Pharmacol 162(6) (2011) 1439–1452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [112].Los GV, Encell LP, McDougall MG, Hartzell DD, Karassina N, Zimprich C, Wood MG, Learish R, Ohana RF, Urh M, Simpson D, Mendez J, Zimmerman K, Otto P, Vidugiris G, Zhu J, Darzins A, Klaubert DH, Bulleit RF, Wood KV, HaloTag: A Novel Protein Labeling Technology for Cell Imaging and Protein Analysis, ACS Chemical Biology 3(6) (2008) 373–382. [DOI] [PubMed] [Google Scholar]
  • [113].Chen W, Younis MH, Zhao Z, Cai W, Recent biomedical advances enabled by HaloTag technology, Biocell 46(8) (2022) 1789–1801. [PMC free article] [PubMed] [Google Scholar]
  • [114].Lesiak L, Zhou X, Fang Y, Zhao J, Beck JR, Stains CI, Imaging GPCR internalization using near-infrared Nebraska red-based reagents, Organic & Biomolecular Chemistry 18(13) (2020) 2459–2467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [115].Kudalkar EM, Davis TN, Asbury CL, Single-Molecule Total Internal Reflection Fluorescence Microscopy, Cold Spring Harbor Protocols 2016(5) (2016) pdb.top077800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Gibbs DR, Kaur A, Megalathan A, Sapkota K, Dhakal S, Build Your Own Microscope: Step-By-Step Guide for Building a Prism-Based TIRF Microscope, Methods and Protocols 1(4) (2018) 40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Roy R, Hohng S, Ha T, A practical guide to single-molecule FRET, Nature Methods 5(6) (2008) 507–516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Zhao R, Rueda D, RNA folding dynamics by single-molecule fluorescence resonance energy transfer, Methods 49(2) (2009) 112–117. [DOI] [PubMed] [Google Scholar]
  • [119].Asher WB, Terry DS, Gregorio GGA, Kahsai AW, Borgia A, Xie B, Modak A, Zhu Y, Jang W, Govindaraju A, Huang L-Y, Inoue A, Lambert NA, Gurevich VV, Shi L, Lefkowitz RJ, Blanchard SC, Javitch JA, GPCR-mediated β-arrestin activation deconvoluted with single-molecule precision, Cell 185(10) (2022) 1661–1675.e16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [120].Möckl L, Moerner WE, Super-resolution Microscopy with Single Molecules in Biology and Beyond–Essentials, Current Trends, and Future Challenges, Journal of the American Chemical Society 142(42) (2020) 17828–17844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Choi UB, Weninger KR, Bowen ME, Immobilization of Proteins for Single-Molecule Fluorescence Resonance Energy Transfer Measurements of Conformation and Dynamics, in: Uversky VN, Dunker AK (Eds.), Intrinsically Disordered Protein Analysis: Volume 2, Methods and Experimental Tools, Springer New York, New York, NY, 2012, pp. 3–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [122].Lamichhane R, Solem A, Black W, Rueda D, Single-molecule FRET of protein–nucleic acid and protein–protein complexes: Surface passivation and immobilization, Methods 52(2) (2010) 192–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Tany R, Goto Y, Kondo Y, Aoki K, Quantitative live-cell imaging of GPCR downstream signaling dynamics, Biochemical Journal 479(8) (2022) 883–900. [DOI] [PubMed] [Google Scholar]
  • [124].Aramendia PF, Negri RM, Roman ES, Temperature Dependence of Fluorescence and Photoisomerization in Symmetric Carbocyanines.Influence of Medium Viscosity and Molecular Structure, The Journal of Physical Chemistry 98(12) (1994) 3165–3173. [Google Scholar]
  • [125].Hwang H, Myong S, Protein induced fluorescence enhancement (PIFE) for probing protein–nucleic acid interactions, Chemical Society Reviews 43(4) (2014) 1221–1229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [126].Bockenhauer S, Fürstenberg A, Yao XJ, Kobilka BK, Moerner WE, Conformational Dynamics of Single G Protein-Coupled Receptors in Solution, The Journal of Physical Chemistry B 115(45) (2011) 13328–13338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Cao C, Tan Q, Xu C, He L, Yang L, Zhou Y, Zhou Y, Qiao A, Lu M, Yi C, Han GW, Wang X, Li X, Yang H, Rao Z, Jiang H, Zhao Y, Liu J, Stevens RC, Zhao Q, Zhang XC, Wu B, Structural basis for signal recognition and transduction by platelet-activating-factor receptor, Nature Structural & Molecular Biology 25(6) (2018) 488–495. [DOI] [PubMed] [Google Scholar]
  • [128].Castro M, Nikolaev VO, Palm D, Lohse MJ, Vilardaga J-P, Turn-on switch in parathyroid hormone receptor by a two-step parathyroid hormone binding mechanism, Proceedings of the National Academy of Sciences 102(44) (2005) 16084–16089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [129].Rahmeh R, Damian M, Cottet M, Orcel H, Mendre C, Durroux T, Sharma KS, Durand G, Pucci B, Trinquet E, Zwier JM, Deupi X, Bron P, Banères J-L, Mouillac B, Granier S, Structural insights into biased G protein-coupled receptor signaling revealed by fluorescence spectroscopy, Proceedings of the National Academy of Sciences 109(17) (2012)6733–6738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Parthier C, Reedtz-Runge S, Rudolph R, Stubbs MT, Passing the baton in class B GPCRs: peptide hormone activation via helix induction?, Trends in Biochemical Sciences 34(6) (2009) 303–310. [DOI] [PubMed] [Google Scholar]
  • [131].Hollenstein K, de Graaf C, Bortolato A, Wang M-W, Marshall FH, Stevens RC, Insights into the structure of class B GPCRs, Trends Pharmacol Sci 35(1) (2014) 12–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [132].Wu F, Yang L, Hang K, Laursen M, Wu L, Han GW, Ren Q, Roed NK, Lin G, Hanson MA, Jiang H, Wang M-W, Reedtz-Runge S, Song G, Stevens RC, Full-length human GLP-1 receptor structure without orthosteric ligands, Nature Communications 11(1) (2020) 1272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [133].Jain A, Liu R, Ramani B, Arauz E, Ishitsuka Y, Ragunathan K, Park J, Chen J, Xiang YK, Ha T, Probing cellular protein complexes using single-molecule pull-down, Nature 473(7348) (2011) 484–488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Kamondanai H, Herve Da C, Yi N, Ashley EB, Qingwei L, Colleen MN, Gilles DT, Conn PJ, A Novel Family of Potent Negative Allosteric Modulators of Group II Metabotropic Glutamate Receptors, Journal of Pharmacology and Experimental Therapeutics 322(1) (2007) 254. [DOI] [PubMed] [Google Scholar]
  • [135].Walker G, Brown C, Ge X, Kumar S, Muzumdar MD, Gupta K, Bhattacharyya M, Oligomeric organization of membrane proteins from native membranes at nanoscale spatial and single-molecule resolution, Nature Nanotechnology 19(1) (2024) 85–94. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES