Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Structural Biology 215 (2023) 107959

Contents lists available at ScienceDirect

Journal of Structural Biology


journal homepage: www.elsevier.com/locate/yjsbi

A review of the approaches used to solve sub-100 kDa membrane proteins


by cryo-electron microscopy
Peter J. Harrison a, b, 1, Tereza Vecerkova a, b, c, 1, Daniel K. Clare b, *, Andrew Quigley a, *
a
Membrane Protein Laboratory, Diamond Light Source, Research Complex at Harwell, Didcot, UK
b
Electron Bio-Imaging Centre, Diamond Light Source, Didcot, UK
c
School of Medicine, Medical Sciences and Nutrition, University of Aberdeen, Aberdeen, UK

A R T I C L E I N F O A B S T R A C T

Keywords: Membrane proteins (MPs) are essential components of all biological membranes, contributing to key cellular
Atomic resolution functions that include signalling, molecular transport and energy metabolism. Consequently, MPs are important
cryo-EM biomedical targets for therapeutics discovery. Despite hardware and software developments in cryo-electron
Membrane protein
microscopy, as well as MP sample preparation, MPs smaller than 100 kDa remain difficult to study structur­
Protein purification
Structural biology
ally. Significant investment is required to overcome low levels of naturally abundant protein, MP hydrophobicity
as well as conformational and compositional instability. Here we have reviewed the sample preparation ap­
proaches that have been taken to successfully express, purify and prepare small MPs for analysis by cryo-EM
(those with a total solved molecular weight of under 100 kDa), as well as examining the differing approaches
towards data processing and ultimately obtaining a structural solution. We highlight common challenges at each
stage in the process as well as strategies that have been developed to overcome these issues. Finally, we discuss
future directions and opportunities for the study of sub-100 kDa membrane proteins by cryo-EM.

1. Introduction structural genomics, sample preparation and X-ray crystallography.


With the beginning of the electron microscopy ‘resolution revolution’ in
The human proteome comprises of approximately 20,000 unique 2011, there has been systematic shift toward single-particle analysis
proteins, of which around a quarter encode for MPs (Almén et al., 2009; (SPA) using cryo-EM for determining the atomic resolution structure of
Uhlén et al., 2015). Not all proteins in the human proteome can be MPs (Fig. 1B). Fewer than 10% of MP structures were solved using X-ray
classified as a suitable pharmaceutical target but of those proteins that crystallography in 2022 compared to 85% of MP structures in 2011. The
are, the membrane proteome is disproportionally represented with 50%- cryo-EM resolution revolution has been primarily driven by advances in
60% of all FDA approved small-molecule pharmaceuticals targeting a microscope, camera and data processing technology, such as the intro­
human MP (Santos et al., 2017). The first structure of a MP was deter­ duction of direct electron detection cameras, in combination with
mined at 7 Å by electron diffraction and published in 1975 (Henderson improved processing software. With the development of computational
and Unwin, 1975), but it took another 10 years for the first high- algorithms to address the protein folding problem such as AlphaFold
resolution structure to be determined (Deisenhofer et al., 1985). Since (Jumper et al., 2021), these experimentally derived structures have been
the first MP structures were published there has been an exponential an invaluable tool in the understanding of global structural biology.
increase in the number of available structures deposited in the Protein Despite the incidence and importance of MPs, they account for only
Data Bank (PDB). As of December 2022, 1534 unique structures have 14% of unique depositions in the PDB (1534 unique structures in the
been deposited according to the ‘Membrane Proteins of Known 3D ‘Membrane Proteins of Known 3D Structure’ database compared to
Structure’ database. These structures have been solved using either X- 10,712 unique structures in the PDB (December 2022)). MPs are noto­
ray crystallography, cryo-EM, electron diffraction or nuclear magnetic riously difficult to study as they are intrinsically more unstable (espe­
resonance spectroscopy (Fig. 1A). The exponential increase in high cially outside of the membrane), express in lower yields and they must
resolution structures of MPs has been driven by developments in be removed from their native membrane-based environment. Typically,

* Corresponding authors.
E-mail addresses: daniel.clare@diamond.ac.uk (D.K. Clare), andrew.quigley@diamond.ac.uk (A. Quigley).
1
Contributed equally.

https://doi.org/10.1016/j.jsb.2023.107959
Received 20 December 2022; Received in revised form 7 March 2023; Accepted 28 March 2023
Available online 31 March 2023
1047-8477/© 2023 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

this is achieved using detergents, but detergent free systems are Sample type, preparation, data acquisition approaches and data
becoming increasingly popular. Cryo-EM can alleviate some of the issues processing choices all require optimisation to determine an atomic res­
of MP production, especially the need for large quantities of material, olution structure by cryo-EM. Making rational choices at each step in the
but challenges remain. For cryo-EM, particle size and conformational process is sometimes opaque and we aim to provide an analysis of the
stability are major determinants of success. It is unsurprising therefore, choices made to solve sub-100 kDa MPs. Our aim is not to provide a
that the first high-resolution single-particle cryo-EM structures of MPs recipe for a successful structural campaign, but to identify trends or
were of large multimeric proteins such as TRPV1 (Cao et al., 2013; Liao common threads running through these published structures. For ease of
et al., 2013). Despite the benefits of single-particle cryo-EM, sub-100 analysis, we have used 1/resolution (Å− 1) to analyse structures. A brief
kDa MPs are still challenging targets due to low contrast and signal-to- explanation of our analysis methodology is provided in the Supple­
noise as well as a lack of features outside the membrane. Of the mentary Information section.
13,200 structures solved by cryo-EM only 1159 are under 100 kDa in
size, and of these only 190 are MPs (based on PDB query searches in 2. Protein family & species origin
December 2022). For the purposes of this review, we have defined a sub-
100 kDa MP as a MP resolved at less than 5 Å resolution, with a total There are 77 published unique sub-100 kDa MP studies at a resolu­
structural weight that is less than 100 kDa and published in a peer tion of below 5 Å. The transporter family contributes the largest fraction
reviewed journal. A description of the search query parameters can be of these unique studies (39), followed by receptors (17), enzymes (11),
found in the Supplementary Information. Sub-100 kDa MPs with a total channels (6) as well as others (Fig. 1C). In 32 out of a total of 77 studies
solved molecular weight of above 100 kDa (for example by attachment of sub-100 kDa MPs structures, apo and ligand-bound states were re­
to a megabody or fab fragment) are not covered by this review, but we ported, thus there are 128 structures deposited from 77 published
provide a brief discussion of these approaches. Based on these param­ studies. There is no statistical difference (p > 0.05) in the final resolution
eters we identified 128 unique MP structures in the PDB belonging to 64 achieved by different classes of proteins (SI Fig. 1A). The bias towards
target proteins. The data we have collated is available as a Supple­ the numbers of deposited transporter structures may be explained by the
mentary File. availability of ligands, that may potentially stabilise conformations as

Fig. 1. Analysis of deposited MP structures in the Protein Data Bank. A: Membrane protein PDB depositions (based upon the ‘Membrane Proteins of Known 3D
Structure’ database) since the first atomic resolution MP structure was solved in 1985 by X-ray crystallography. There has been an exponential increase over the last
20 years. In particular, the number of structures solved by cryo-EM since 2011, whilst there is a small contribution from electron diffraction (ED) and Nuclear
Magnetic Resonance spectroscopy. B: Depositions of sub-100 kDa MP structures resolved to less than 5 Å since the first structure was deposited in 2017. A total of 77
unique MPs and a total of 128 conformations have been deposited to date (December 2022) C: Proportion of solved and published sub-100 kDa MP cryo-EM
structures by protein family. D: Proportion of protein species origin represented in sub- 100 kDa MPs solved by cryo-EM.

2
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

well as the ability to select different conformational and oligomeric 0.03) to a His tag (average Å− 1 = 0.30 ± 0.04) (p = 0.000973). More­
states by cryo-EM allowing for higher-level sample heterogeneity over, there is also a statistically significant difference between the FLAG
compared to X-ray crystallography. tag (average Å− 1 = 0.298 ± 0.04) and the strep tag (p = 0.031). There
Structures of sub-100 kDa MPs have been solved from all domains of was no statistical difference for any other comparison (SI Fig. 3A). This
life, but the vast majority are of animal origin with human sources ac­ observation may be due to the fact that Strep-based resin is often used to
counting for 70 structures, and other mammalian sources a further 22 purify difficult to obtain or low yielding membrane proteins, so may
out of a total of 128 structures. Far fewer sub-100 kDa MPs originate indicate that these proteins are intrinsically less stable. The most pop­
from bacteria with 18 reported structures. Archaea and viral MPs ac­ ular immobilised metal-ion affinity chromatographic (IMAC) resins
count for 6 each and plants and fungi 3 each (Fig. 1D). The shift of focus were nickel-based. 27 studies reported using nickel-nitrilotriacetic acid
to structures of higher order eukaryotic proteins is partly due to (Ni-NTA) and a further 3 used unspecified nickel-based resins (SI
improved methods for protein production in mammalian systems, a Fig. 3B). Five studies used cobalt-based resins. Other resins reported
reduction in production costs over the past decade and a focus on human include Streptactin (14), Flag M1 (9) and M2 resins (7), and anti-GFP
pharmaceutical targets. There is no statistically significant difference (p conjugated affinity resins (9). Four other resins, Tse3, an immobilised
> 0.05) in final resolution achieved between proteins of human, animal DARPin, G1 and Streptavidin resins, were used in single cases (SI
or bacterial origin (SI Fig. 1B). Fig. 3B).

3. Sample preparation 3.2. Detergents

Without doubt, the provision of usable yields of high-quality protein To study MP function and structure, it is often necessary to extract
is key to any high-resolution structure determination. The careful design MPs from membranes into encapsulation agents that mimic lipid envi­
of expression constructs must consider the incorporation of promotor ronment to maintain protein integrity and stability. MPs were tradi­
elements, reporter and purification tags as well as enzyme cleavage sites tionally solubilised using detergents which are still the most popular
that may be used to remove these tags. The target gene sequence may choice for cryo-EM studies of sub-100 kDa MPs. Selection of the most
require codon optimisation, specific mutations, or insertions introduced suitable detergent can be time consuming, but there are now numerous
to stabilise the MP. The choice of host expression system as well as published screening protocols available for optimal detergent selection
purification strategy will have been made prior to construct design un­ (Desuzinges Mandon et al., 2017; Kotov et al., 2019; Vergis et al., 2010).
less using expression vectors that contain multiple host promoter ele­ Given the breadth of detergents available, there is surprisingly little
ments or vectors with multiple purification tags (3.1 Construct Design, variation in the detergents used for solubilisation of sub-100 kDa MPs
Expression and Purification). Once expressed, decisions on the use of (Fig. 2A). The popularity of n-Dodecyl-β-D-maltoside (DDM) has per­
detergent (3.2 Detergents) or non-detergent (3.3 Non-detergent encap­ sisted from crystallographic studies, due to its low critical micelle con­
sulation systems) solubilisation and stabilisation approaches are centration (CMC), mild nature, and price compared to other detergents.
required and sometimes excess detergent may need to be removed (3.4 DDM was used in 47 out of 77 studies demonstrating the robustness of
Homogenous samples and removing excess detergent). Finally, it may be DDM, its suitability for variety of different proteins and its use as a good
necessary to solve the structures as part of a complex (4.0 Binders, ‘first choice’ detergent. However, it may be necessary to screen a range
Complexes and Symmetry). These parameters and strategies for sample of detergents before performing cryo-EM analysis. Lauryl Maltose Neo­
optimisation have been well reviewed elsewhere (Birch et al., 2020; pentyl Glycol (LMNG) is another notable detergent with its use reported
Kesidis et al., 2020). in 22 studies.
Detergents are the most popular encapsulation agents for cryo-EM
3.1. Construct design, expression and purification when purified MPs are applied to a cryo-EM grid (Fig. 2B). The popu­
larity of LMNG is close to that of DDM (11 and 12 out of 44 studies
Our analysis of sub-100 kDa MP structures reveals that most proteins respectively) among the proteins solved in detergent micelles (Fig. 2C).
are expressed in a phylogenetically closely related organism, presum­ This may be due to the low CMC and slow off rate of LMNG, which
ably to ensure close-to-native protein is produced (SI Table 1). Human means LMNG solubilised proteins are stable even at sub-CMC concen­
expression systems are used for the majority of proteins of human origin trations (Hauer et al., 2015), which can help to reduce the detergent
(25 out of 44 human proteins), with the remainder expressed in insect background in micrographs. Digitonin (5) and its synthetic derivative
cell lines (Sf9 and Trichoplusia ni). Seven sub-100 kDa MPs mammalian glycol-diosgenin (GDN) (7 in GDN and 7 in GDN/LMNG) are also
proteins were reportedly expressed in human based hosts with a further commonly used during cryo-EM. Given the differences in the biophysical
four MPs expressed in insect cells and four MPs expressed in fungal/ properties of DDM, DM (n-Decyl-β-Maltoside), LMNG, digitonin and
yeast hosts. There is no effect on the final resolution (p > 0.05) when GDN, it is not possible to say that sub-100 kDa proteins have a prefer­
expressing human or mammalian proteins in either human or insect cells ence for one over the other (SI Fig. 4A and SI Fig. 4B). There is a clear
lines (SI Fig. 1C). The majority (14 out of 17) prokaryotic proteins are supremacy of LMNG and DDM over other detergents and detergent
expressed in Escherichia coli, unless a more closely related expression combinations, which may be biased by the fact they are common ‘first
system is available, such as Mycobacterium smegmatis (3 out of 17) for choice’ detergents.
tuberculosis targets. The Saccharomyces cerevisiae expression system has Protein extraction by detergents from native membrane removes
been used for expression of proteins from fungi (1), plant (1), yeast (2), much of the surrounding lipid environment, which may severely impact
and a few mammalian proteins (3). protein integrity. Cholesteryl hemisuccinate (CHS) is commonly used
The poly-histidine-tag (his-tag) is the most popular purification tag with detergents (43 out of 77 studies), as it is a more soluble derivative
accounting for 36 out of 77 sub-100 kDa MP studies. FLAG and Strep of cholesterol. Cholesterol is a major component of membranes across
based tags are the second most popular accounting for 16 and 15 MPs higher eukaryotic species, and so its addition during purification is a
respectively. Green fluorescent protein (GFP) has been used as a puri­ rational step for proteins from species whose membranes natively
fication tag in 9 cases, purified with nanobody-conjugated resins. Out of contain cholesterol. CHS can help stabilise proteins, and so for many
74 constructs (construct choice not always mentioned in methods), 41 projects its addition may well be an essential component of the purifi­
contained more than one purification tag (SI Fig. 1A), but only 7 pro­ cation process to maintain conformational stability (Drew et al., 2008;
tocols utilised more than one tag for affinity chromatography. There is a Meury et al., 2014). CHS was supplemented during purifications of
small (0.4 Å) statistically significant difference in the final resolution mammalian, viral and plant proteins, in fact 74.1% of all mammalian
achieved when comparing the use of a Strep tag (average Å− 1 = 0.26 ± proteins had CHS added during solubilisation.

3
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

Fig. 2. Encapsulation agents used for solubilisation and imaging by cryo-EM A: Distribution of detergents used for solubilisation. DDM and LMNG are by far the
most popular solubilisation agents used in 69 out of 77 solubilisation protocols. B: Encapsulation system used when imaging by cryo-EM. Detergents (43 out of 77)
and discs (26) are the most common encapsulation agents for cryo-EM of sub-100 kDa MPs, amphipols (8) are preferentially used for non chaperone-MP complexes
(see section 4.0) and oligomeric structures, and SMALP system was used only in 1 study. C: Detergent used when imaging by cryo-EM. Structures solved in detergent
micelles are most commonly in DDM, LMNG, GDN, or their combinations. D: Disc system used when imaging by cryo-EM. Out of the wide variety of nanodiscs
available, nanodiscs of smaller sized diameters, mainly MSP1D1 and MSP1E3D1 are used for cryo-EM analysis. E: Amphipol used when imaging by cryo-EM.
Amphipols are used mostly for cryo-EM of multimeric proteins, with PMAL-C8 used in 6 out of 8 preparations. F-I: Representation of sub-100 kDa in encapsula­
tion agents: ZnT8 (PDB: 6XPD) in a detergent micelle, GLUT4 (PDB: 6X12) in a nanodisc, Signal peptidase complex paralog A (PDB: 7P2P) in an amphipol, PBP1
(PDB: 7LQ6) in a SMALP system.

3.3. Non-detergent encapsulation systems example of a co-polymer-based encapsulation system, capable of direct
solubilisation of MPs with native lipids from membranes (Jamshad et al.,
Although detergents are excellent at solubilising MPs and preserving 2011; Knowles et al., 2009). SMALP solutions do not need to be sup­
structural features, they can make image processing difficult due to an plemented, like detergents, during the purification processes reducing
empty detergent micelle background and reduced conformational sta­ the costs and potentially the background noise in cryo-EM images
bility of MPs in detergents over other non-detergent-based encapsula­ (Pollock et al., 2018). Where Salipro or SMALPs have been used for the
tion systems. Nanoscale phospholipid bilayer systems such as nanodiscs encapsulation of sub-100 kDa MPs, the proteins were reconstituted from
and co-polymer systems provide an alternative to detergents. Some MPs a detergent extraction.
(GLUT4, PepT1, TMEM120A and the hepatic sodium / bile acid MSP1D1 (~9.7 nm) (Denisov et al., 2004) and MSP1E3D1 (~12 nm)
cotransporter) have been solved both in detergents and in a nanodisc. (Puthenveetil and Vinogradova, 2013) are the most commonly used
There has been a rapid expansion in the reported number of non- nanodiscs for sub-100 kDa MPs with 10 and 8 reported studies respec­
detergent systems available for MP encapsulation. Nanodiscs are the tively, while MSP1E3 (~12.9 nm) (Denisov et al., 2004) has been only
most popular non-detergent system with 26 studies reporting their use reported four times. Recently two structures were solved using Saposin
(Fig. 2D). Other reported non-detergent encapsulation systems include A (adaptable size) (Flayhan et al., 2018) as the encapsulation agent (Xu
amphipols, reported in 8 studies, and recently the use of a styrene maleic et al., 2022; Zhang et al., 2022) and one structure has been solved using
acid lipid particle (SMALP) was reported (Caveney et al., 2021). the nanodisc NW9 (Niu et al., 2022) (~9 nm) (Nasr et al., 2017). There is
Nanodiscs are disc-shaped particles composed of lipids and an one example of a sub-100 kDa MP (TWIK1) that has been solved both in
amphipathic protein or co-polymer scaffold. Various types and sizes of MSP1D1 and MSP1E3D1 (Turney et al., 2022). We do not observe that
nanodisc have been used to study sub-100 kDa MPs allowing a tailored smaller sub-100 kDa MPs tend towards smaller MSP sizes, and vice
approach to the individual protein. The first protein based nanodiscs versa. In addition, for sub-100 kDa MPs solved in discs, there is no
were synthetically developed from Apolipoprotein A and termed mem­ correlation between protein size and resolution achieved (SI Fig. 4C).
brane scaffold proteins (MSPs) (Bayburt et al., 2002). Subsequently However, nanodisc selection is often made based on the predicted
other protein-based systems have been developed including peptidiscs number of transmembrane helices, and it is for this reason that we
(Carlson et al., 2018) and Salipro (saposin-lipoporotein) (Drulyte et al., suspect larger MSP sizes are not represented here (Ritchie et al., 2009).
2023; Popovic et al., 2012). Both the peptidisc and Salipro encapsulation Lipids are an essential component of the nanodisc system helping to
systems can directly solubilise MPs (Carlson et al., 2018; Lloris-Garcerá maintain the MP in an environment more similar to the native mem­
et al., 2020). Styrene-maleic acid lipid particles (SMALPs) are an brane. The most common lipid is 1-palmitoyl-2-oleoyl-glycero-3-

4
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

phosphocholine (POPC) (supplemented in 13 out of 26 nanodisc


reconstitution studies) either alone, or more commonly in combination
with other lipids. Even though there is no correlation between nanodisc
type and lipids used, the most common nanodisc lipid composition is a
2:1:1 ratio of DOPE:POPS:POPC (6 reconstitutions). Other less
commonly used lipids were E. coli, brain and liver total lipid extract,
based on the species and expression system to closely represent given
native lipid composition.
Amphipathic polymers (amphipols) (Tribet et al., 1996) are poly­
mers with a hydrophilic backbone that makes contact with the buffer
solution, and hydrophobic side chains which closely interact with
transmembrane domains of MPs. Since their introduction, novel
amphipols with different properties have been developed, such as non-
ionic amphipols (NaPol) (Prata et al., 2000; Sharma et al., 2012), or
zwitterionic PMALs (Nagy et al., 2001). Out of eight MP structures that
have used amphipols, two used the original A8-35 polymer while six
studies report the use of the zwitterionic PMALs. Six of the studies that
used amphipols were of protein complexes, suggesting their potential
benefits for multimeric proteins (Fig. 2E).
The versatility of available encapsulation systems (Fig. 2F-I) allows
accomodation of different protein characteristics and requirements.
There is no significant difference (p > 0.05) in final resolution achieved
for different encapsulation systems (detergent Å− 1 = 0.29 ± 0.04, disc
Å− 1 = 0.29 ± 0.044 and amphipol Å− 1 = 0.26 ± 0.06 (Fig. 3A). Within
detergents there is no variation in the final resolution (Fig. 3B). Within
nanodiscs, there are only sufficient data to compare MSP1D1 and
MSP1E3D1, and there is no statistically significant difference in reso­
lution obtained between them.

3.4. Homogenous samples and removing excess detergent

During the purification process, detergent exchange is often under­


taken as the solubilisation detergent is not always the most appropriate
for cryo-EM analysis due to stability, yield and cost of the original
detergent. Out of 44 studies which solved structures in detergent, the
extraction detergent was exchanged, or at least supplemented, in 24
preparations. Exchanges usually took place during size exclusion chro­
matography (SEC). The most common SEC resin for sub-100 kDa MPs is
Superdex 200 Increase as reported in 34 studies. Superdex 200 was re­
ported 16 times while Superose 6 Increase resin and Superose 6 resin
were reported 14 and 8 time respectively (SI Fig. 5). SEC Columns such
as Sepax SRT-C SEC-300 (1), Biorad650 (1) and ENrich SEC 650 (2) were
used to a lesser extent. Column selection is often a laboratory-based
Fig. 3. Encapsulation agent versus the final resolution achieved by each
equipment selection factor rather than inherent (un)suitability for MP
sub-100 kDa protein structure. A: Comparison of encapsulation agent type
purification; however, selection of SEC media can affect the stability of a with the final resolution achieved. The final resolution achieved is not affected
MP during SEC especially since columns such as SRT-C may offer by the encapsulation system used (p > 0.05, SMALP (1) structure were excluded
reduced detergent and hydrophobic protein adsorption. Hi-Trap ion from the analysis). B: Comparison of detergent type with the final resolution
exchange columns were used only in 2 out of 77 studies as a purification achieved. Detergent selection does not affect the final resolution achieved by
step. Techniques such as gradient centrifugation can also be used to sub-100 kDa proteins solved in detergent micelles (p > 0.05, DDM/LMNG (3)
remove excess detergent before vitrification (Hauer et al., 2015). and digitonin/sodium cholate (1) structures were excluded from the analysis).

4. Binders, complexes & symmetry reduces the intrinsic flexibility and heterogeneity of a sample, making
downstream data processing easier. For MPs, it can often be difficult to
Cryo-EM data processing of sub-100 kDa MPs is challenging and resolve the transmembrane helices as they can be occluded by the
crucially dependent on the quality of the primary data. Identification of detergent micelle or encapsulation agent. This can be problematic for
particles and image alignment of sub-100 kDa MPs can be improved sub-100 kDa MPs which do not have any soluble domains. Megabodies
through the use of native protein ligands to form larger complexes. (nanobodies with large soluble domains) are frequently also used as they
These MP complex structures are important for understanding protein give a large increase in the mass of the target. This mass increase is often
function, interactions and cellular pathways and have the additional to over 100 kDa in size, and so proteins solved with bound megabodies
benefit of increasing the size and mass of the target which can help with are beyond the scope of this analysis.
data processing. Additionally, chaperone-MP complexes can be formed Of the 128 sub-100 kDa MP structures that we identified, 30 were
where the protein of interest is bound to nanobody or Fab fragments. monomeric structures, 28 were part of a homeric complex and 70 were
The addition of nanobodies or Fab fragments has two benefits for part of a heteromeric complex. Structural chaperone-MP complexes,
structural determination of small proteins; adding mass to the MP target defined as a MP bound to either a nanobody, sybody (laboratory evolved
(Uchański et al., 2020; Uchański et al., 2021) thus aiding particle nanobody) or Fab fragment consisted of 54 of the 70 heteromeric
picking and potentially locking proteins in a fixed conformation. This

5
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

structures (Fig. 4A). Breaking down the structural chaperone-MP com­


plex classification further revealed the use of Fab fragments, nanobodies
and sybodies 33, 18 and 3 times respectively. 15 structures out of the 70
heteromeric structures were in complex with either another MP or a
soluble protein that was not a structural chaperone. One further struc­
ture was a combination of a heteromeric MP complex and a structural
chaperone (7TDN, which was solved in complex with a Fab). The
remaining 28 sub-100 kDa homomeric MP structures consisted of 26
homodimeric and 2 homotrimeric structures (Fig. 4A). There is no sta­
tistical difference (p > 0.05) between the final resolution achieved be­
tween monomeric, homodimeric and homotrimeric structures. It is
worth nothing, that two proteins have been solved with and without
nanobodies. PepT2 achieved a Å− 1 of 0.29 (7NQK) in the presence of
nanobody and 0.24 (7PMY) in the absence of a nanobody (both in DDM/
CHS). Similarly, the folate carrier/transporter achieved Å− 1 of 0.30,
0.26 and 0.28 (7TX6, 7TX7 & 8DEP, in digitonin) and 0.31 and 0.3
(7BC6 & 7BC7, in DDM/CHS) without a nanobody. A third protein, the
hepatic sodium/bile acid cotransporter, has been solved with both a Fab
(7VAD, 7VAE, 7VAF, 7VAG & 7WSI, Å− 1 of 0.3, 0.29, 0.28, 0.32, 0.30
and 0.30) and a nanobody (7FCI, Å− 1 of 0.27). There are not enough
data to draw any conclusions as to whether nanobodies or other binders
improve resolution (Fig. 4A).
As previously mentioned, soluble domains can be beneficial for
image alignment. Of the 70 heteromeric complex structures, only 4 did
not have a soluble domain (and thus were fully embedded within the
membrane) (Fig. 4B). Of the 58 monomeric or homomeric structures, 17
did not have a soluble domain. For most structures, a small-molecule or
metal ion binder was reported. There are insufficient data to make a
statistical comparison as to whether this affects the final resolution
achieved. Small molecules may have the same effect as nanobodies by
reducing conformational flexibility. Interestingly, there is no correlation
between the size of the solved structure and the resolution achieved
(Fig. 4C).
From our analysis of sub-100 kDa MPs there is no significant dif­
ference (p > 0.05) in the final resolution achieved for structures solved
in C1 symmetry (100 out of 130 structures, Å− 1 = 0.29 ± 0.04) or C2
symmetry (28 of 130 structures, Å− 1 = 0.28 ± 0.04) symmetry (SI
Fig. 6A) (all but two structures analysed were solved in either of these
two space groups). As expected, structures in C2 symmetry had on
average fewer particles in their final reconstructions than those in C1
(on average 180,816 ± 158,424 particles versus 305,706 ± 254,798
particles) (p = 0.018) (SI Fig. 6B). The two structures in C3 were solved
with even fewer particles (87,865), but there are too few structures to
make an accurate comparison.

5. Grid preparation approaches

For SPA, most grids (~95%) are prepared using a ThermoFisher


Scientific Vitrobot (122 out of 128 solely used a Vitrobot), wherein the
protein sample is applied to a grid, blotted between two sheets of filter
paper, and plunged into liquid ethane. The entire process can take up­
wards of 15 seconds, which increases the chances of proteins denaturing
or adopting a preferred orientation on the grid. A handful of researchers Fig. 4. Soluble domains, chaperone-MP complexes, non chaperone-MP
have used either a Leica GP2 or Gatan Cryoplunge, the former uses complexes and protein size versus reported final resolution. A: Compari­
single-sided blotting and is the preferred option for blotting cellular son of the final resolution achieved for monomeric MP structures, homodimeric
samples. All three instruments blot away excess sample before rapidly MP structures, non chaperone-MP heteromeric complexes and chaperone-MP
plunging into liquid ethane to vitrify the sample in the grid holes. As heteromeric complexes (Fab, nanobody or sybody). B: Comparison of the
such, it is difficult to critique new developments in plunge freezing de­ final resolution achieved for heteromeric complexes (chaperone-MP and non
vices (Weissenberger et al., 2021). chaperone-MP), monomeric and homomeric (homodimeric and homotrimeric)
MPs with and without soluble domains. Most MPs solved either have a soluble
Grid choice is an important factor to consider. EM grids are generally
domain or a protein ligand which provides a soluble domain. There is no sta­
made of either copper or gold supports, with a film (usually amorphous
tistically significant difference between the final resolution achieved. C: Rela­
carbon) on top (Thompson et al., 2016). The film contains holes in which tionship between protein size and the achieved final resolution of the structure.
the protein sits. For SPA, small hole sizes are considered to be preferable, There is no correlation between the two.
as this reduces ice buckling and sample movement (Naydenova et al.,
2020). Moreover, smaller hole sizes may prevent over blotting of thin ice
films in the middle of the hole which can lead to clustering at the hole

6
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

edge, preferred orientation and sample denaturation (Kampjut et al., beam induced sample motion and improves the distribution of the
2021). 88% of reported structures are solved using 1.2/1.3 hole grids particles on the grid (Rice et al., 2018). These factors should aid struc­
with the remaining studies using 2/1, 2/2 and 0.6/1 holes (Fig. 5A). Due tural determination.
to the dominance of 1.2/1.3 grids it is not possible to draw any con­ Sample concentration is an important consideration when making
clusions in comparison to other hole sizes beyond the popularity of 1.2/ grids as it is necessary to ensure that there are enough particles in the
1.3 grids. holes, whilst not overcrowding the grid. Ultimately the best concen­
Traditionally grid supports are made of copper due to its high ther­ tration to use is sample dependent. Where a value is given (for a range,
mal conductivity. Gold grids were introduced primarily for cellular to­ average value was taken), most proteins are vitrified within a range of 4
mography as gold is non-toxic to cells. Recently, UltrAUfoil grids where to 7 mg mL− 1 (51 out of 115), followed closely by 1 to 4 mg mL− 1 (41 out
the holey carbon film has been replaced by a thin holey gold film were of 115). Small proteins with lower molecular weight will have more
developed (Russo and Passmore, 2014; Russo and Passmore, 2016). Our molecules than of a larger protein at a given concentration (in mg mL− 1),
analysis, with the caveat that fewer studies report the use of UltrAUfoil so it is unsurprising that nine structures report a concentration below 1
grids, indicates that there is no significant difference in final resolution mg mL− 1, although seven structures resulted from MPs frozen at con­
achieved when comparing carbon on copper grids with carbon on gold centrations above 10 mg mL− 1 (SI Fig. 7). GPCRs are often frozen at high
or UltrAUfoil, or when comparing UltrAUfoil grids with carbon on concentrations which gives densely packed particles in the foil holes
copper grids (Fig. 5B). A comparison of UltrAUfoils against carbon on (Danev et al., 2021), saturating the carbon film and forcing particles into
gold grids indicates a statistically significant difference (p = 0.028) in the holes.
the final resolution achieved. There are however two statistical outliers
7RXH (0.43 Å− 1) & 7UL2 (0.42 Å− 1) and we note that these two data­ 6. Data collection approaches
points achieved the highest resolution of any structures in our sub-100
kDa membrane protein dataset. These high resolution datapoints show When collecting single-particle data for high-resolution analysis of
the potential gains to be had from improvements to EM technology and any protein, there are several considerations to be made including the
processing. The gold film, coupled with the gold grid support, reduces microscope acceleration voltage, the detector and the dose. For sub-100
kDa MPs it is essential to maximise the signal-to-noise ratio. Our analysis
considers acceleration voltage, detector choice, pixel size, collection
settings (super-resolution, Correlated Double Sampling (CDS)), energy
filter slit width, total dose and number of fractions. SPA data collection
is performed using different software packages such as EPU (Thermo­
Fisher), SerialEM (Cheng et al., 2016; Mastronarde, 2005; Tan et al.,
2015), Leginon (Suloway et al., 2005), AutoEMation (Lei and Frank,
2005) or Latitude (Gatan). SerialEM is the most used (56%), automated,
SPA data collection software for sub-100 kDa MPs (SI Fig. 8).
86% of near-atomic resolution, sub-100 kDa, MP cryo-EM structures
were solved using data collected on a 300 keV microscope. 300 keV
microscopes are widely considered the best approach for obtaining high-
resolution data due to greater sample penetration, decreased inelastic
scattering and decreased sample charging relative to 200 keV in­
struments (Wu et al., 2020). However, at lower acceleration voltages,
the elastic cross-section is greater (2.01-fold greater at 100 kV vs. 300
keV) (Peet et al., 2019). This increases the contrast, at the expense of
radiation damage which only increases 1.57-fold at 100 keV compared
to 300 keV. As such, imaging smaller objects at lower acceleration
voltages should be preferable but is not routine as detectors are opti­
mised to work at 300 keV.
Pixel size dictates the maximum final resolution that is possible in a
reconstruction. For ease of analysis we have grouped pixel sizes into 5
groups (~0.5 Å/px, ~0.65 Å/px, ~0.83 Å/px, ~1.1 Å/px and ~ 1.3 Å/
px). High-resolution data collections typically use a pixel size of ~ 1 Å/
px (Feathers et al., 2021). In our analysis of 128 sub-100 kDa MP
structures, a pixel size of ~ 0.83 Å/px was most frequently reported (61
of 128 structures) (SI Fig. 9). Pixel sizes of ~ 1.1 Å/px were the next
most common (46 of 128 structures). The highest resolution recorded for
a sub-100 kDa MP in our dataset is 2.3 Å (7RXH) (at a pixel size of 1.06
Å/px) (Falzone et al., 2022) and from the structures we analysed, an
average final resolution of 3.52 ± 0.50 Å was obtained. There is a pos­
itive correlation between the fraction of Nyquist achieved (Pearson’s r
= 0.78) for a structure and the pixel size data was collected at (data
collected at a larger pixel size is solved closer to the Nyquist frequency)
(SI Fig. 10A). However, data collected at smaller pixel sizes does not
achieve a higher final resolution than those collected at larger pixel sizes
Fig. 5. Grid choice for sub-100 kDa MPs. A: Distribution of grid hole sizes
(SI Fig. 10B) (Pearson’s r = 0.024). This is an interesting observation, as
used for sub-100 kDa MPs. A hole size of 1.2/1.3 is by far the most common. B:
Distribution of the final resolution achieved using different types of grid foils. one might expect the improvement in the detective quantum efficiency
Carbon on Gold grids are the most popular choice. There is a statistically sig­ (DQE) that imaging at higher magnifications affords to be beneficial. As
nificant difference in final resolution achieved when comparing UltrAUfoil such, it can be argued that there is less to gain from collecting at smaller
grids with carbon on gold grids. There is no significant difference in the final pixel sizes and that a pixel size of ~ 1.35 Å/px could be used to maximise
resolution achieved for any other comparison of grid supports and foils. the throughput of data collection.

7
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

Most structures in our analysis were solved using data collected with
either a Gatan K2 (28 of 128) or K3 (93 of 128) direct detection camera.
The K3 camera has an improved frame frate in comparison to the K2,
which reduces coincidence loss and improves detector performance. Our
analysis shows that structures solved with a K3 camera achieve a better
final resolution than those solved with a K2 camera (p = 0.001) (SI
Fig. 11). This is unexpected since there is no significant difference in the
DQE between the K2 and K3 cameras. It should be noted that the K2 is an
older camera, so structures solved with a K3 benefit from improvements
in data processing. The biggest advantage of the K3 is data collection
throughput. A K3 detector is capable of in excess of 1000 movies per
hour (Sheng et al., 2022). There are not enough data to make a com­
parison between Falcon 3 and Falcon 4 cameras. Interestingly, our
analysis suggests that more data is not necessarily linked to an
improvement in resolution. The average number of final particles is
276,801 (with a standard deviation of ± 241,346 which reflects the
variation in the number of final particles used in reconstructions) from
an average of 10,124 ± 8081 movies. We see a slight positive correlation
between the number of particles in the final reconstruction (Pearson’s r
= 0.31) (Fig. 6A). There is no significant correlation between the initial
number of particles picked (Pearson’s r = 0.22) or the number of movies
collected (Pearson’s r = -0.11) and the resolution achieved in the final
reconstruction (Fig. 6B and Fig. 6C). We see a weak positive correlation
(Pearson’s r = 0.24) between the number of movies collected and the
number of final particles in the reconstruction (SI Fig. 12), suggesting
that more movies does not result in higher resolution structures, and
further emphasises the need for high quality sample preparation.
Energy filtered transmission electron microscopy (EFTEM) removes
inelastically scattered electrons, which improves contrast and signal to
noise (Langmore and Smith, 1992; Schröder et al., 1990; Yonekura et al.,
2006; Zhu et al., 1997). This should be of significant benefit to sub-100
kDa MPs. It should be noted that the greatest effect of energy filtering is
for thicker samples (Yonekura et al., 2006), and thus for small objects in
thin ice the amplitude contrast enhancement will be less pronounced.
Where a slit width was reported (62 cases), 73% used a slit width of 20
eV (SI Fig. 13). Slit widths of 10 eV, 15 eV and 25 eV were also reported.
The current width of energy filter slits used is dictated by their stability
as drift of the zero-loss peak can result in a loss of electron transmission
to the detector. As new energy filters become available that have greater
stability, smaller slit widths will become more accessible, thus the signal
to noise achievable should improve further.
Many reported structures have been solved using ‘super-resolution’
mode. Super-resolution is a detector feature which allows a single
electron event to be sub-localised within a quadrant of a single pixel
(Booth, 2012; Li et al., 2013). Reconstructions may subsequently go
beyond the physical Nyquist limit dictated by the pixel size. From our
analysis, there is no significant difference in final resolution achieved for
structures in normal resolution versus in super-resolution, indicating
that there is less of a benefit for sub-100 kDa MPs (SI Fig. 14). A small
handful of structures were collected in CDS mode (Sun et al., 2021). CDS
mode, samples each pixel’s voltage twice per frame (at the start and at
the end), with the initial value subtracted from the final value, which
increases ‘the detectability’ of an electron, improving signal to noise
which in turn should result in higher-resolution reconstructions, espe­
cially for small proteins which are often difficult to see. Unfortunately,
in our dataset there were only three collections done in CDS, therefore
we cannot make any inferences on the effectiveness of CDS for sub-100
kDa MPs. It will be interesting to see if this feature becomes more readily
Fig. 6. Initial and final number of particles and number of movies versus
used. One disadvantage of CDS is the reduction in relative throughput,
final resolution A: Relationship between the final number of particles in each
as due to the two read per frame (750 fps versus 1500), lower electron
reconstruction, B: the initial number of particles picked and C: the number of
counts must be used to avoid coincidence loss. movies collected versus the final resolution achieved for each structure. There is
Total dose applied to a sample for imaging is an important consid­ a low correlation between either of the variables and the final resolu­
eration. The electron dose experienced by a sample is a function of the tion achieved.
sample itself as well as the energy, fluence and rate at which electrons
are delivered. In the field of electron microscopy electron dose and
electron fluence are often used interchangeably (Ilett et al., 2020) and

8
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

thus dose is usually recorded as the number of electrons per Å2 (e/ Å2). structures we analysed at some point used either Relion or CryoSPARC.
This is effectively a measure of the ‘dose over vacuum’. A higher total In 41 of the 128 structures analysis here, a mixture of both software
dose should give better signal at lower spatial frequencies, increases platforms were used. Bayesian polishing (Scheres, 2012) of the final map
contrast and gives a better signal-to-noise in the micrographs, thus is often performed in Relion. From our analysis of solved sub-100 kDa
making it easier to see small objects. Higher electron doses come at the MP structures we do not find that there are definitive trends towards a
expense of beam induced sample damage with high-resolution sample particular software type or feature. It is worth noting that the majority of
information lost first (Cheng et al., 2015; Grant and Grigorieff, 2015). the processing pipelines use either CryoSPARC or Relion, but there may
Beam induced damage is alleviated through dose-weighting of the mo­ still be some advantages to trying other software such as Sphire (Au -
tion corrected micrographs (Grant and Grigorieff, 2015; Zheng et al., Moriya et al., 2017) or cisTEM (Grant et al., 2018) for cases where other
2017). The average dose applied is 56.3 ± 10.4 e/Å2 (SI Fig. 15A) and software struggles.
there is no significant correlation between applied dose and the final
resolution achieved (Pearson’s r = 0.18) (SI Fig. 15B). 8. Further developments
As discussed, a key issue with imaging small proteins in the TEM is
that of contrast. Smaller samples have less scattering material, reducing At the time of writing this review there are a handful of MP structures
contrast, and making the identification of particles much harder. In in the region of ~ 44 kDa (excluding the size of the encapsulation agent)
microscopy, phase plates are used to introduce additional phase shift to deposited in the PDB, demonstrating that the limits of single particle
the scattered or un-scattered electrons depending on the type of phase cryo-EM are being continually pushed. Future developments in micro­
plate (or photons in the case of light microscopy). This additional phase scope, detector and image processing software are likely to decrease the
shift increases the phase contrast in the image making it easier to lower limit of solvable structures both in SPA and cryoET even further.
identify and subsequently align your protein of interest (Wang and Fan, Future developments in microscope architecture have been nicely
2019). In biological cryo-EM, the most commonly used phase plate is the summarised in a recent review by Russo et al. (Russo et al., 2022). Im­
Volta phase plate (VPP) (Danev et al., 2017; Danev et al., 2014). Despite aging at liquid helium temperatures potentially offers benefits in
its promise, the VPP is not routinely used for SPA in Cryo-EM. Of the decreased radiation damage and improved signal to noise (as a higher
structures we examined, only three studies (6 structures in total) used a electron dose can be applied to the sample) (Naydenova et al., 2022;
VPP (Shang et al., 2019; Xue et al., 2020; Zhao et al., 2019). The highest Pfeil-Gardiner et al., 2019; Russo et al., 2022). However, due to
resolution of these structures was reported at 3.8 Å. Practically, the VPP increased sample movement in liquid helium cooled samples, these
is difficult to align and operate, has limited reproducibility and makes gains are currently out of reach (although improvements in image pro­
fitting and correction of CTF harder (Wang and Fan, 2019). These dif­ cessing software are likely to help offset this). Chromatic aberration
ficulties, coupled with the fact the phase plate film absorbs approxi­ correctors which can focus inelastically scattered electrons and laser
mately 10% of electrons thus leading to signal loss, help explain their phase plates could both help to increase signal and contrast in cryo-EM
limited use. This may change in the coming years when the next gen­ images (Dickerson et al., 2022; Schwartz et al., 2019). As discussed,
eration of laser phase plates come online. imaging at 100 keV could also have benefits for small MPs, if suitable
detectors become available. It remains to be seen whether these im­
7. Data processing approaches provements in microscope technology lead us to a concurrent increase in
resolution.
A final part of the pipeline for structural determination by cryo-EM is ‘Artificial intelligence’ or deep learning techniques are already being
data processing. Software choice for motion correction and CTF esti­ applied to the image processing pipeline (Chung et al., 2022). Currently
mation is a personal choice. From our analysis, motion correction was they are mostly used for particle picking, with deep learning software
most commonly performed in MotionCorr2 (99 out of 128 structures) such as crYOLO, Topaz and Warp (Tegunov and Cramer, 2019) already
(Zheng et al., 2017). However, other motion correction software such as commonly used. Tools such as CryoDRGN (Zhong et al., 2021) and
Unblur (Grant and Grigorieff, 2015), and implementations in Relion and 3DFlex (Punjani and Fleet, 2022) are also available to aid with hetero­
CryoSPARC are also commonly used (Rubinstein and Brubaker, 2015; geneity in the 3D reconstruction phase and DeepEMhancer (Sanchez-
Zivanov et al., 2018). Similarly, CTF correction most commonly uses Garcia et al., 2021) are aiding with map sharpening. Given the
either GCTF (39 out of 128 structures) (Zhang, 2016) or CTFFIND (60 increasing complexity of cryo-SPA data sets due to intrinsic protein
out of 128 structures) (Rohou and Grigorieff, 2015) or an implantation flexibility and complexes rather than individual proteins becoming
in Relion or CryoSPARC (Zivanov et al., 2020). targets, deep learning tools for model building, classification and
Following motion correction and CTF estimation, particles must be refinement will become crucial.
picked from the micrographs. Generally picking software can be divided For SPA cryo-EM of sub-100 kDa MPs, new polymer reagents such as
into 3 discrete categories: Template pickers, neural networks (where a SMALPs or Salipro may aid in the solubilisation stage of sample prepa­
small subset of picked particles is used to train a model such as crYOLO ration with maintaining near-native lipid environment. With the
(Wagner et al., 2019) or Topaz (Bepler et al., 2020)) and Gaussian development of large nanodiscs and peptidiscs (allowing adaptability to
pickers such as DoG picker (Voss et al., 2009). Often with small objects it protein size) (Angiulli et al., 2020) the ability to capture larger com­
is difficult to distinguish particles on a micrograph before class aver­ plexes will open up the possibility to image large membrane complexes.
aging, so it is essential to ensure particle picking is as accurate as As technology improves, we will move to a situation where more MPs of
possible and that particles, in several different orientations, are selected smaller size can be imaged in cellulo by cryo-electron tomography,
from the micrographs. Typically, a small set of several thousand parti­ (species smaller than 500 kDa are considered small for CryoET) (Turk
cles are picked either manually or using software such as Blob Picker. and Baumeister, 2020). Similarly, developments in labelling and func­
Selected particles are then used to train a model for automated picking. tionalisation of proteins will be crucial for this. Moreover, development
Alternatively, 2D classes generated from the initial picking can be used of time-resolved EM techniques will allow structural dynamics to be
as templates for the autopicking software. From our analysis we have viewed at atomic resolution (Kelly et al., 2022).
observed that there are multiple software options and a high variability
in approaches taken. It is not possible to quantify which software 9. Conclusions
package or approach is the most successful.
2D classification, initial model generation, 3D classification, 3D Our analysis has provided some interesting insights. For example,
refinement and polishing are generally performed using either Relion or there is no correlation between sub-100 kDa MP size and resolution
CryoSPARC (Punjani et al., 2017; Zivanov et al., 2018). In fact, all of the achieved, but there is a weak positive correlation between the final

9
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

number of particles and the resolution achieved. Most sub-100 kDa MPs provided by Diamond Light Source and the Research Complex at Har­
solved by cryo-EM are encapsulated in detergent with DDM the most well. We also acknowledge the Diamond Light Source for access and the
popular. Carbon on gold grids are by far the most popular for sub- 100 support of the cryo-EM facilities at the UK national electron Bio-Imaging
kDa MPs while 1.2/1.3 grids are almost exclusively used. Although the Centre, funded by the Wellcome Trust, MRC and BBSRC. Graphical ab­
basics of data processing are fundamentally the same, many different stract and Fig. 2 were created with BioRender.com.
approaches are taken for particle picking, 2D and 3D classification as
well as refinement. Appendix A. Supplementary data
For our analysis we have the final resolution reported for each
structure (Å− 1). We have also performed a parallel analysis (see Sup­ Supplementary data to this article can be found online at https://doi.
plementary Information) using ‘fraction of Nyquist’ (FoN), where the org/10.1016/j.jsb.2023.107959.
final resolution is divided by the Nyquist frequency of the imaging
conditions. The purpose of this was to try to observe how close structures References
were to achieving their theoretical highest resolution based on the pixel
size used. However, there are issues with this metric as one would expect Almén, M.S., Nordström, K.J.V., Fredriksson, R., Schiöth, H.B., 2009. Mapping the
human membrane proteome: a majority of the human membrane proteins can be
that collections at smaller pixel sizes will achieve better resolutions by classified according to function and evolutionary origin. BMC Biol. 7, 50.
virtue of the imaging conditions. Interestingly, both metrics produces Angiulli, G., Dhupar, H.S., Suzuki, H., Wason, I.S., Duong Van Hoa, F., Walz, T., 2020.
identical conclusions in 12 out of 17 cases (see SI). Both of these metrics New approach for membrane protein reconstitution into peptidiscs and basis for
their adaptability to different proteins. eLife 9, e53530.
are dependent on consistent resolution estimation from processing Au - Moriya, T., Au - Saur, M., Au - Stabrin, M., Au - Merino, F., Au - Voicu, H., Au -
software. A better metric may well be the B-factor, as described by Huang, Z., Au - Penczek, P.A., Au - Raunser, S., Au - Gatsogiannis, C., 2017. High-
Henderson and Rosenthal (Rosenthal and Henderson, 2003). The B- resolution Single Particle Analysis from Electron Cryo-microscopy Images Using
SPHIRE. JoVE, e55448.
factor can be used to estimate the number of particles needed to produce Bayburt, T.H., Grinkova, Y.V., Sligar, S.G., 2002. Self-Assembly of Discoidal
a reconstruction at any resolution and thus provides estimates the Phospholipid Bilayer Nanoparticles with Membrane Scaffold Proteins. Nano Lett. 2,
experimental error for a specific experiment. As such, the B-factor takes 853–856.
Bepler, T., Kelley, K., Noble, A.J., Berger, B., 2020. Topaz-Denoise: general deep
into account multiple factors including sample preparation, data
denoising models for cryoEM and cryoET. Nat. Commun. 11, 5208.
collection parameters and image processing. Thus the B-factor provides Birch, J., Cheruvara, H., Gamage, N., Harrison, P.J., Lithgo, R., Quigley, A., 2020.
a better description of the overall experiment (Nakane et al., 2020). Changes in Membrane Protein Structural Biology. Biology 9, 401.
However, B-factors are not always reported, which currently makes it Booth, C., 2012. K2: A Super-Resolution Electron Counting Direct Detection Camera for
Cryo-EM. Microsc. Microanal. 18, 78–79.
difficult to use this as a metric and we therefore strongly encourage the Cao, E., Liao, M., Cheng, Y., Julius, D., 2013. TRPV1 structures in distinct conformations
reporting Bfactors with data collection metrics. reveal activation mechanisms. Nature 504, 113–118.
Fundamentally, data collection and image processing are limited Carlson, M.L., Young, J.W., Zhao, Z., Fabre, L., Jun, D., Li, J., Li, J., Dhupar, H.S.,
Wason, I., Mills, A.T., Beatty, J.T., Klassen, J.S., Rouiller, I., Duong, F., 2018. The
unless a protein of high-quality is supplied. Given the variety of different Peptidisc, a simple method for stabilizing membrane proteins in detergent-free
purification methods, encapsulation agents and binding partners solution. eLife 7, e34085.
(nanobodies, antibodies) used, there is not a single encapsulation agent Caveney, N.A., Workman, S.D., Yan, R., Atkinson, C.E., Yu, Z., Strynadka, N.C.J., 2021.
CryoEM structure of the antibacterial target PBP1b at 3.3 Å resolution. Nat.
that works for every protein, or system achieving the best possible res­ Commun. 12, 2775.
olution. As such, it is important to expend time and effort on sample Cheng, A., Tan, Y.Z., Dandey, V.P., Potter, C.S., Carragher, B., 2016. Chapter Four -
optimisation before collecting data. Strategies for Automated CryoEM Data Collection Using Direct Detectors, p. 87-102,
in: R. A. Crowther, (Ed.), Methods in Enzymology, Academic Press.
Cheng, Y., Grigorieff, N., Penczek, P.A., Walz, T., 2015. A Primer to Single-Particle Cryo-
CRediT authorship contribution statement Electron Microscopy. Cell 161, 438–449.
Chung, J.M., Durie, C.L., Lee, J., 2022. Artificial Intelligence in Cryo-Electron
Microscopy. Life 12, 1267.
Peter J. Harrison: Conceptualization, Data curation, Formal anal­
Danev, R., Buijsse, B., Khoshouei, M., Plitzko, J.M., Baumeister, W., 2014. Volta potential
ysis, Writing – original draft. Tereza Vecerkova: Formal analysis, phase plate for in-focus phase contrast transmission electron microscopy. Proc. Natl.
Writing – original draft. Daniel K. Clare: Conceptualization, Funding Acad. Sci. 111, 15635–15640.
acquisition, Supervision, Writing – original draft. Andrew Quigley: Danev, R., Tegunov, D., Baumeister, W., 2017. Using the Volta phase plate with defocus
for cryo-EM single particle analysis. eLife 6, e23006.
Conceptualization, Funding acquisition, Supervision, Writing – original Danev, R., Belousoff, M., Liang, Y.-L., Zhang, X., Eisenstein, F., Wootten, D., Sexton, P.M.,
draft. 2021. Routine sub-2.5 Å cryo-EM structure determination of GPCRs. Nature.
Communications 12, 4333.
Deisenhofer, J., Epp, O., Miki, K., Huber, R., Michel, H., 1985. Structure of the protein
Declaration of Competing Interest subunits in the photosynthetic reaction centre of Rhodopseudomonas viridis at 3Å
resolution. Nature 318, 618–624.
The authors declare the following financial interests/personal re­ Denisov, I.G., Grinkova, Y.V., Lazarides, A.A., Sligar, S.G., 2004. Directed Self-Assembly
of Monodisperse Phospholipid Bilayer Nanodiscs with Controlled Size. J. Am. Chem.
lationships which may be considered as potential competing interests: Soc. 126, 3477–3487.
Andrew Quigley reports financial support was provided by Wellcome Desuzinges Mandon, E., Agez, M., Pellegrin, R., Igonet, S., Jawhari, A., 2017. Novel
Trust. systematic detergent screening method for membrane proteins solubilization. Anal.
Biochem. 517, 40–49.
Dickerson, J.L., Lu, P.-H., Hristov, D., Dunin-Borkowski, R.E., Russo, C.J., 2022. Imaging
Data availability biological macromolecules in thick specimens: The role of inelastic scattering in
cryoEM. Ultramicroscopy 237, 113510.
Drew, D., Newstead, S., Sonoda, Y., Kim, H., von Heijne, G., Iwata, S., 2008. GFP-based
The data used in this paper is freely available in the Protein Data optimization scheme for the overexpression and purification of eukaryotic
Bank. A curated version of the results from our search query is made membrane proteins in Saccharomyces cerevisiae. Nat. Protoc. 3, 784–798.
available in the supplementary materials. Drulyte, I., Gutgsell, A.R., Lloris-Garcerá, P., Liss, M., Geschwindner, S., Radjainia, M.,
Frauenfeld, J., Löving, R., 2023. Direct cell extraction of membrane proteins for
structure–function analysis. Sci. Rep. 13, 1420.
Acknowledgements Falzone, M.E., Feng, Z., Alvarenga, O.E., Pan, Y., Lee, B., Cheng, X., Fortea, E.,
Scheuring, S., Accardi, A., 2022. TMEM16 scramblases thin the membrane to enable
lipid scrambling. Nat. Commun. 13, 2604.
We would like to thank colleagues in eBIC, MPL and MX at Diamond
Feathers, J.R., Spoth, K.A., Fromme, J.C., 2021. Experimental evaluation of super-
Light Source for helpful discussions and comments. We also thank resolution imaging and magnification choice in single-particle cryo-EM. Journal of
Daniel Hatton (Diamond Light Source) for helpful discussions. The Structural Biology: X 5, 100047.
Membrane Protein Laboratory is funded by grants from the Wellcome
Trust (202892/Z/16/Z and 223727/Z/21/Z) with additional support

10
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

Flayhan, A., Mertens, H.D.T., Ural-Blimke, Y., Martinez Molledo, M., Svergun, D.I., Pfeil-Gardiner, O., Mills, D.J., Vonck, J., Kuehlbrandt, W., 2019. A comparative study of
Löw, C., 2018. Saposin Lipid Nanoparticles: A Highly Versatile and Modular Tool for single-particle cryo-EM with liquid-nitrogen and liquid-helium cooling. IUCrJ 6,
Membrane Protein Research. Structure 26, 345–355.e345. 1099–1105.
Grant, T., Grigorieff, N., 2015. Measuring the optimal exposure for single particle cryo- Pollock, N.L., Lee, S.C., Patel, J.H., Gulamhussein, A.A., Rothnie, A.J., 2018. Structure
EM using a 2.6 Å reconstruction of rotavirus VP6. eLife 4, e06980. and function of membrane proteins encapsulated in a polymer-bound lipid bilayer.
Grant, T., Rohou, A., Grigorieff, N., 2018. cisTEM, user-friendly software for single- Biochimica et Biophysica Acta (BBA) -. Biomembranes 1860, 809–817.
particle image processing. eLife 7, e35383. Popovic, K., Holyoake, J., Pomès, R., Privé, G.G., 2012. Structure of saposin A lipoprotein
Hauer, F., Gerle, C., Fischer, N., Oshima, A., Shinzawa-Itoh, K., Shimada, S., discs. Proceedings of the National Academy of Sciences 109, 2908-2912.
Yokoyama, K., Fujiyoshi, Y., Stark, H., 2015. GraDeR: Membrane Protein Complex Prata, C., Giusti, F., Gohon, Y., Pucci, B., Popot, J.-L., Tribet, C., 2000. Nonionic
Preparation for Single-Particle Cryo-EM. Structure 23, 1769–1775. amphiphilic polymers derived from Tris(hydroxymethyl)-acrylamidomethane keep
Henderson, R., Unwin, P.N.T., 1975. Three-dimensional model of purple membrane membrane proteins soluble and native in the absence of detergent. Biopolymers 56,
obtained by electron microscopy. Nature 257, 28–32. 77–84.
Ilett, M., S’ari, M., Freeman, H., Aslam, Z., Koniuch, N., Afzali, M., Cattle, J., Hooley, R., Punjani, A., Fleet, D., 2022. 3D Flexible Refinement: Structure and Motion of Flexible
Roncal-Herrero, T., Collins, S.M., Hondow, N., Brown, A., Brydson, R., 2020. Proteins from Cryo-EM. Microscopy Microanaly. 28, 1218-1218.
Analysis of complex, beam-sensitive materials by transmission electron microscopy Punjani, A., Rubinstein, J.L., Fleet, D.J., Brubaker, M.A., 2017. cryoSPARC: algorithms
and associated techniques. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 378, for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296.
20190601. Puthenveetil, R., Vinogradova, O., 2013. Optimization of the design and preparation of
Jamshad, M., Lin, Y.-P., Knowles, T.J., Parslow, R.A., Harris, C., Wheatley, M., Poyner, D. nanoscale phospholipid bilayers for its application to solution NMR. Proteins:
R., Bill, R.M., Thomas, O.R.T., Overduin, M., Dafforn, T.R., 2011. Surfactant-free Structure. Function, and Bioinformatics 81, 1222–1231.
purification of membrane proteins with intact native membrane environment. Rice, W.J., Cheng, A., Noble, A.J., Eng, E.T., Kim, L.Y., Carragher, B., Potter, C.S., 2018.
Biochem. Soc. Trans. 39, 813–818. Routine determination of ice thickness for cryo-EM grids. J. Struct. Biol. 204, 38–44.
Jumper, J., Evans, R., Pritzel, A., Green, T., Figurnov, M., Ronneberger, O., Ritchie, T.K., Grinkova, Y.V., Bayburt, T.H., Denisov, I.G., Zolnerciks, J.K., Atkins, W.M.,
Tunyasuvunakool, K., Bates, R., Žídek, A., Potapenko, A., Bridgland, A., Meyer, C., Sligar, S.G., 2009. Chapter Eleven - Reconstitution of Membrane Proteins in
Kohl, S.A.A., Ballard, A.J., Cowie, A., Romera-Paredes, B., Nikolov, S., Jain, R., Phospholipid Bilayer Nanodiscs, p. 211-231, in: N. Düzgünes, (Ed.), Methods in
Adler, J., Back, T., Petersen, S., Reiman, D., Clancy, E., Zielinski, M., Steinegger, M., Enzymology, Academic Press.
Pacholska, M., Berghammer, T., Bodenstein, S., Silver, D., Vinyals, O., Senior, A.W., Rohou, A., Grigorieff, N., 2015. CTFFIND4: Fast and accurate defocus estimation from
Kavukcuoglu, K., Kohli, P., Hassabis, D., 2021. Highly accurate protein structure electron micrographs. J. Struct. Biol. 192, 216–221.
prediction with AlphaFold. Nature 596, 583–589. Rosenthal, P.B., Henderson, R., 2003. Optimal Determination of Particle Orientation,
Kampjut, D., Steiner, J., Sazanov, L.A., 2021. Cryo-EM grid optimization for membrane Absolute Hand, and Contrast Loss in Single-particle Electron Cryomicroscopy.
proteins. iScience 24, 102139. J. Mol. Biol. 333, 721–745.
Kelly, D.F., DiCecco, L.-A., Jonaid, G.M., Dearnaley, W.J., Spilman, M.S., Gray, J.L., Rubinstein, J.L., Brubaker, M.A., 2015. Alignment of cryo-EM movies of individual
Dressel-Dukes, M.J., 2022. Liquid-EM goes viral – visualizing structure and particles by optimization of image translations. J. Struct. Biol. 192, 188–195.
dynamics. Curr. Opin. Struct. Biol. 75, 102426. Russo, C.J., Passmore, L.A., 2014. Ultrastable gold substrates for electron
Kesidis, A., Depping, P., Lodé, A., Vaitsopoulou, A., Bill, R.M., Goddard, A.D., Rothnie, A. cryomicroscopy. Science 346, 1377–1380.
J., 2020. Expression of eukaryotic membrane proteins in eukaryotic and prokaryotic Russo, C.J., Passmore, L.A., 2016. Ultrastable gold substrates: Properties of a support for
hosts. Methods 180, 3–18. high-resolution electron cryomicroscopy of biological specimens. J. Struct. Biol. 193,
Knowles, T.J., Finka, R., Smith, C., Lin, Y.-P., Dafforn, T., Overduin, M., 2009. Membrane 33–44.
Proteins Solubilized Intact in Lipid Containing Nanoparticles Bounded by Styrene Russo, C.J., Dickerson, J.L., Naydenova, K., 2022. Cryomicroscopy in situ: what is the
Maleic Acid Copolymer. J. Am. Chem. Soc. 131, 7484–7485. smallest molecule that can be directly identified without labels in a cell? Faraday
Kotov, V., Bartels, K., Veith, K., Josts, I., Subhramanyam, U.K.T., Günther, C., Labahn, J., Discuss. 240, 277–302.
Marlovits, T.C., Moraes, I., Tidow, H., Löw, C., Garcia-Alai, M.M., 2019. High- Sanchez-Garcia, R., Gomez-Blanco, J., Cuervo, A., Carazo, J.M., Sorzano, C.O.S.,
throughput stability screening for detergent-solubilized membrane proteins. Sci. Vargas, J., 2021. DeepEMhancer: a deep learning solution for cryo-EM volume post-
Rep. 9, 10379. processing. Communications Biology 4, 874.
Langmore, J.P., Smith, M.F., 1992. Quantitative energy-filtered electron microscopy of Santos, R., Ursu, O., Gaulton, A., Bento, A.P., Donadi, R.S., Bologa, C.G., Karlsson, A., Al-
biological molecules in ice. Ultramicroscopy 46, 349–373. Lazikani, B., Hersey, A., Oprea, T.I., Overington, J.P., 2017. A comprehensive map of
Lei, J., Frank, J., 2005. Automated acquisition of cryo-electron micrographs for single molecular drug targets. Nat. Rev. Drug Discov. 16, 19–34.
particle reconstruction on an FEI Tecnai electron microscope. J. Struct. Biol. 150, Scheres, S.H.W., 2012. RELION: Implementation of a Bayesian approach to cryo-EM
69–80. structure determination. J. Struct. Biol. 180, 519–530.
Li, X., Zheng, S.Q., Egami, K., Agard, D.A., Cheng, Y., 2013. Influence of electron dose Schröder, R.R., Hofmann, W., Ménétret, J.-F., 1990. Zero-loss energy filtering as
rate on electron counting images recorded with the K2 camera. J. Struct. Biol. 184, improved imaging mode in cryoelectronmicroscopy of frozen-hydrated specimens.
251–260. J. Struct. Biol. 105, 28–34.
Liao, M., Cao, E., Julius, D., Cheng, Y., 2013. Structure of the TRPV1 ion channel Schwartz, O., Axelrod, J.J., Campbell, S.L., Turnbaugh, C., Glaeser, R.M., Müller, H.,
determined by electron cryo-microscopy. Nature 504, 107–112. 2019. Laser phase plate for transmission electron microscopy. Nat. Methods 16,
Lloris-Garcerá, P., Klinter, S., Chen, L., Skynner, M.J., Löving, R., Frauenfeld, J., 2020. 1016–1020.
DirectMX – One-Step Reconstitution of Membrane Proteins From Crude Cell Shang, G., Zhang, C., Chen, Z.J., Bai, X.-C., Zhang, X., 2019. Cryo-EM structures of STING
Membranes Into Salipro Nanoparticles. Front. Bioeng. Biotechnol. 8. reveal its mechanism of activation by cyclic GMP–AMP. Nature 567, 389–393.
Mastronarde, D.N., 2005. Automated electron microscope tomography using robust Sharma, K.S., Durand, G., Gabel, F., Bazzacco, P., Le Bon, C., Billon-Denis, E., Catoire, L.
prediction of specimen movements. J. Struct. Biol. 152, 36–51. J., Popot, J.-L., Ebel, C., Pucci, B., 2012. Non-Ionic Amphiphilic Homopolymers:
Meury, M., Costa, M., Harder, D., Stauffer, M., Jeckelmann, J.-M., Brühlmann, B., Synthesis, Solution Properties, and Biochemical Validation. Langmuir 28,
Rosell, A., Ilgü, H., Kovar, K., Palacín, M., Fotiadis, D., 2014. Detergent-Induced 4625–4639.
Stabilization and Improved 3D Map of the Human Heteromeric Amino Acid Sheng, Y., Harrison, P.J., Vogirala, V., Yang, Z., Strain-Damerell, C., Frosio, T., Himes, B.
Transporter 4F2hc-LAT2. PLoS One 9, e109882. A., Siebert, C.A., Zhang, P., Clare, D.K., 2022. Application of super-resolution and
Nagy, J.K., Kuhn Hoffmann, A., Keyes, M.H., Gray, D.N., Oxenoid, K., Sanders, C.R., correlative double sampling in cryo-electron microscopy. Faraday Discuss. 240,
2001. Use of amphipathic polymers to deliver a membrane protein to lipid bilayers. 261–276.
FEBS Lett. 501, 115–120. Suloway, C., Pulokas, J., Fellmann, D., Cheng, A., Guerra, F., Quispe, J., Stagg, S.,
Nakane, T., Kotecha, A., Sente, A., McMullan, G., Masiulis, S., Brown, P.M.G.E., Potter, C.S., Carragher, B., 2005. Automated molecular microscopy: The new
Grigoras, I.T., Malinauskaite, L., Malinauskas, T., Miehling, J., Uchański, T., Yu, L., Leginon system. J. Struct. Biol. 151, 41–60.
Karia, D., Pechnikova, E.V., de Jong, E., Keizer, J., Bischoff, M., McCormack, J., Sun, M., Azumaya, C.M., Tse, E., Bulkley, D.P., Harrington, M.B., Gilbert, G., Frost, A.,
Tiemeijer, P., Hardwick, S.W., Chirgadze, D.Y., Murshudov, G., Aricescu, A.R., Southworth, D., Verba, K.A., Cheng, Y., Agard, D.A., 2021. Practical considerations
Scheres, S.H.W., 2020. Single-particle cryo-EM at atomic resolution. Nature 587, for using K3 cameras in CDS mode for high-resolution and high-throughput single
152–156. particle cryo-EM. J. Struct. Biol. 213, 107745.
Nasr, M.L., Baptista, D., Strauss, M., Sun, Z.-Y.-J., Grigoriu, S., Huser, S., Plückthun, A., Tan, Y.Z., Cheng, A., Potter, C.S., Carragher, B., 2015. Automated data collection in
Hagn, F., Walz, T., Hogle, J.M., Wagner, G., 2017. Covalently circularized nanodiscs single particle electron microscopy. Microscopy 65, 43–56.
for studying membrane proteins and viral entry. Nat. Methods 14, 49–52. Tegunov, D., Cramer, P., 2019. Real-time cryo-electron microscopy data preprocessing
Naydenova, K., Jia, P., Russo, C.J., 2020. Cryo-EM with sub–1 Å specimen movement. with Warp. Nat. Methods 16, 1146–1152.
Science 370, 223–226. Thompson, R.F., Walker, M., Siebert, C.A., Muench, S.P., Ranson, N.A., 2016. An
Naydenova, K., Kamegawa, A., Peet, M.J., Henderson, R., Fujiyoshi, Y., Russo, C.J., 2022. introduction to sample preparation and imaging by cryo-electron microscopy for
On the reduction in the effects of radiation damage to two-dimensional crystals of structural biology. Methods 100, 3–15.
organic and biological molecules at liquid-helium temperature. Ultramicroscopy Tribet, C., Audebert, R., Popot, J.-L., 1996. Amphipols: Polymers that keep membrane
237, 113512. proteins soluble in aqueous solutions. Proc. Natl. Acad. Sci. 93, 15047–15050.
Niu, Y., Liu, R., Guan, C., Zhang, Y., Chen, Z., Hoerer, S., Nar, H., Chen, L., 2022. Turk, M., Baumeister, W., 2020. The promise and the challenges of cryo-electron
Structural basis of inhibition of the human SGLT2–MAP17 glucose transporter. tomography. FEBS Lett. 594, 3243–3261.
Nature 601, 280–284. Turney, T.S., Li, V., Brohawn, S.G., 2022. Structural Basis for pH-gating of the K+
Peet, M.J., Henderson, R., Russo, C.J., 2019. The energy dependence of contrast and channel TWIK1 at the selectivity filter. Nat. Commun. 13, 3232.
damage in electron cryomicroscopy of biological molecules. Ultramicroscopy 203, Uchański, T., Pardon, E., Steyaert, J., 2020. Nanobodies to study protein conformational
125–131. states. Curr. Opin. Struct. Biol. 60, 117–123.

11
P.J. Harrison et al. Journal of Structural Biology 215 (2023) 107959

Uchański, T., Masiulis, S., Fischer, B., Kalichuk, V., López-Sánchez, U., Zarkadas, E., Ye, S., 2022. Embigin facilitates monocarboxylate transporter 1 localization to the
Weckener, M., Sente, A., Ward, P., Wohlkönig, A., Zögg, T., Remaut, H., Naismith, J. plasma membrane and transition to a decoupling state. Cell Rep. 40.
H., Nury, H., Vranken, W., Aricescu, A.R., Pardon, E., Steyaert, J., 2021. Megabodies Xue, J., Xie, T., Zeng, W., Jiang, Y., Bai, X.-C., 2020. Cryo-EM structures of human ZnT8
expand the nanobody toolkit for protein structure determination by single-particle in both outward- and inward-facing conformations. eLife 9, e58823.
cryo-EM. Nat. Methods 18, 60–68. Yonekura, K., Braunfeld, M.B., Maki-Yonekura, S., Agard, D.A., 2006. Electron energy
Uhlén, M., Fagerberg, L., Hallström, B.M., Lindskog, C., Oksvold, P., Mardinoglu, A., filtering significantly improves amplitude contrast of frozen-hydrated protein at
Sivertsson, Å., Kampf, C., Sjöstedt, E., Asplund, A., Olsson, I., Edlund, K., 300kV. J. Struct. Biol. 156, 524–536.
Lundberg, E., Navani, S., Szigyarto, C.-A.-K., Odeberg, J., Djureinovic, D., Zhang, K., 2016. Gctf: Real-time CTF determination and correction. J. Struct. Biol. 193,
Takanen, J.O., Hober, S., Alm, T., Edqvist, P.-H., Berling, H., Tegel, H., Mulder, J., 1–12.
Rockberg, J., Nilsson, P., Schwenk, J.M., Hamsten, M., von Feilitzen, K., Zhang, K., Wu, H., Hoppe, N., Manglik, A., Cheng, Y., 2022. Fusion protein strategies for
Forsberg, M., Persson, L., Johansson, F., Zwahlen, M., von Heijne, G., Nielsen, J., cryo-EM study of G protein-coupled receptors. Nat. Commun. 13, 4366.
Pontén, F., 2015. Tissue-based map of the human proteome. Science 347, 1260419. Zhao, D.Y., Pöge, M., Morizumi, T., Gulati, S., Van Eps, N., Zhang, J., Miszta, P.,
Vergis, J.M., Purdy, M.D., Wiener, M.C., 2010. A high-throughput differential filtration Filipek, S., Mahamid, J., Plitzko, J.M., Baumeister, W., Ernst, O.P., Palczewski, K.,
assay to screen and select detergents for membrane proteins. Anal. Biochem. 407, 2019. Cryo-EM structure of the native rhodopsin dimer in nanodiscs. J. Biol. Chem.
1–11. 294, 14215–14230.
Voss, N.R., Yoshioka, C.K., Radermacher, M., Potter, C.S., Carragher, B., 2009. DoG Zheng, S.Q., Palovcak, E., Armache, J.-P., Verba, K.A., Cheng, Y., Agard, D.A., 2017.
Picker and TiltPicker: Software tools to facilitate particle selection in single particle MotionCor2: anisotropic correction of beam-induced motion for improved cryo-
electron microscopy. J. Struct. Biol. 166, 205–213. electron microscopy. Nat. Methods 14, 331–332.
Wagner, T., Merino, F., Stabrin, M., Moriya, T., Antoni, C., Apelbaum, A., Hagel, P., Zhong, E.D., Bepler, T., Berger, B., Davis, J.H., 2021. CryoDRGN: reconstruction of
Sitsel, O., Raisch, T., Prumbaum, D., Quentin, D., Roderer, D., Tacke, S., Siebolds, B., heterogeneous cryo-EM structures using neural networks. Nat. Methods 18,
Schubert, E., Shaikh, T.R., Lill, P., Gatsogiannis, C., Raunser, S., 2019. SPHIRE- 176–185.
crYOLO is a fast and accurate fully automated particle picker for cryo-EM. Zhu, J., Penczek, P.A., Schröder, R., Frank, J., 1997. Three-Dimensional Reconstruction
Communications Biology 2, 218. with Contrast Transfer Function Correction from Energy-Filtered Cryoelectron
Wang, H.-W., Fan, X., 2019. Challenges and opportunities in cryo-EM with phase plate. Micrographs: Procedure and Application to the 70SEscherichia coliRibosome.
Curr. Opin. Struct. Biol. 58, 175–182. J. Struct. Biol. 118, 197–219.
Weissenberger, G., Henderikx, R.J.M., Peters, P.J., 2021. Understanding the invisible Zivanov, J., Nakane, T., Forsberg, B.O., Kimanius, D., Hagen, W.J.H., Lindahl, E.,
hands of sample preparation for cryo-EM. Nat. Methods 18, 463–471. Scheres, S.H.W., 2018. New tools for automated high-resolution cryo-EM structure
Wu, M., Lander, G.C., Herzik, M.A., 2020. Sub-2 Angstrom resolution structure determination in RELION-3. eLife 7, e42166.
determination using single-particle cryo-EM at 200 keV. Journal of Structural Zivanov, J., Nakane, T., Scheres, S.H.W., 2020. Estimation of high-order aberrations and
Biology: X 4, 100020. anisotropic magnification from cryo-EM data sets in RELION-3.1. IUCrJ 7, 253–267.
Xu, B., Zhang, M., Zhang, B., Chi, W., Ma, X., Zhang, W., Dong, M., Sheng, L., Zhang, Y.,
Jiao, W., Shan, Y., Chang, W., Wang, P., Wen, S., Pei, D., Chen, L., Zhang, X., Yan, H.,

12

You might also like