ISSN 0006-2979, Biochemistry (Moscow), 2025, Vol. 90, No. 11, pp. 1521-1535 © The Author(s) 2025. This article is an open access publication.
Russian Text © The Author(s), 2025, published in Biokhimiya, 2025, Vol. 90, No. 11, pp. 1621-1637.
1521
REVIEW
Methodological Toolbox
for Identifying and Studying Micropeptides:
From Genome to Function
Aleksandr I. Lavrov
1,2
, Nikita M. Shepelev
1,3
, Olga A. Dontsova
1,3
,
and Maria P. Rubtsova
1,3,a
*
1
Lomonosov Moscow State University, Faculty of Chemistry,
119991 Moscow, Russia
2
Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University,
119992 Moscow, Russia
3
Shemyakin–Ovchinnikov Institute of Bioorganic Chemistry, Russian Academy of Sciences,
117997 Moscow, Russia
a
e-mail: mprubtsova@gmail.com
Received July 22, 2025
Revised October 2, 2025
Accepted October 6, 2025
AbstractMicropeptides encoded by small open reading frames (sORFs) represent a novel, actively studied
class of functional molecules regulating key cellular processes. Studying micropeptides is complicated by
methodological challenges, in particular, their small size, low cellular abundance, and difficulty in generating
specific antibodies. The review systematizes modern approaches to the identification and functional character-
ization of micropeptides. The main strategies for their discovery include the use of bioinformatic algorithms,
global translation analysis via ribosome profiling, direct detection using mass spectrometry-based proteomics,
and phenotypic screenings. The methods for confirming the functions of micropeptides and elucidating mo-
lecular mechanisms of their action genetic knockouts, affinity tagging for visualization, and investigation of
protein-protein interactions. The review discusses key challenges and future prospects in the field, emphasiz-
ing the importance of an integrated multi-omics approach for the comprehensive micropeptidome mapping.
DOI: 10.1134/S0006297925602242
Keywords: micropeptides, small open reading frames, long non-coding RNAs, mass spectrometry, Ribo-Seq,
affinity labeling, co-immunoprecipitation
* To whom correspondence should be addressed.
INTRODUCTION
Any extended nucleotide sequence, whether nat-
ural or random, contains numerous open reading
frames (ORFs), as random nucleotide combinations
form various start and stop codons [1]. Historically,
large genome annotation consortia have focused ex-
clusively on identifying protein-coding ORFs and ig-
nored small ORFs (sORFs) encoding peptides shorter
than 100 amino acid residues [2]. For a long time,
sORFs had been considered an untranslated genomic
“noise,” incapable of encoding stable and functional
peptides.
This conception has changed dramatically with
the emergence of high-throughput sequencing tech-
nologies, particularly, ribosome profiling (Ribo-Seq),
which enables global mapping of translated RNA
regions at a single-nucleotide resolution [3]. Com-
bined with the proteomic data [4, 5], these studies
have provided compelling evidence for the trans-
lation of thousands of sORFs in various organisms,
including humans [6,  7]. Furthermore, function-
al studies have shown that many sORF translation
products – micropeptides – are involved in the reg-
ulation of fundamental processes, such as signaling,
metabolism, homeostasis, muscle activity, DNA repair,
and immune response [8-11]. Moreover, micropep-
tides have been found to influence the development
LAVROV et al.1522
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
of various pathologies, including cancer. In other
words, micropeptides represent the “dark matter” of
the proteome and possess a significant regulatory po-
tential [12].
Due to some properties of micropeptides, includ-
ing their low cellular abundance, potential instabili-
ty, and difficulty in generating specific antibodies, the
search for these compounds and their characteriza-
tion are associated with methodological challenges,
which necessitated adaptation of existing methods
and development of new approaches for the reliable
identification, confirmation, and functional analysis
of micropeptides. The goal of this review is to provide
a comprehensive and systematic analysis of modern
methodological arsenal used in the studies of micro-
peptides and to discuss the prospects for new devel-
opments in this field.
MICROPEPTIDES
sORFs are short nucleotide sequences (no longer
than 100 codons from the start codon to the stop co-
don) in DNA or RNA. A significant number of sORFs
are located within protein-coding genes, in particu-
lar, in the 5′-untranslated regions (5′-UTRs) of mRNA
(the so-called upstream ORFs that often play a reg-
ulatory role in the translation of the main coding
sequence), 3′-untranslated regions (3′-UTRs) of mRNA
(downstream ORFs), or in the coding sequences in
alternative reading frames. Additionally, many sORFs
have been identified in transcripts initially classified
as non-coding RNAs, including long non-coding RNAs
(lncRNAs), primary microRNA transcripts, and circu-
lar RNAs [13].
sORFs can serve as templates for the synthesis of
short proteins, called micropeptides or microproteins.
Over two decades, numerous micropeptides involved
in embryogenesis [8,  9], metabolism [10,  11], and DNA
repair [14] have been described. Some micropeptides
are known to promote carcinogenesis, while others
act as tumor suppressors [15,  16]. Micropeptides with
the neuroprotective properties that inhibited the de-
velopment of the Alzheimers disease have also been
identified [17].
For example, physiologically active micropep-
tide DWORF functions as an activator of the calcium
ATPase SERCA responsible for transporting Ca
2+
ions
from the cytoplasm into the sarcoplasmic reticulum (a
process necessary for muscle relaxation). Overexpres-
sion of DWORF in the cardiac muscle increases SERCA
activity, improves myocardial contractility, and affects
calcium homeostasis [18, 19]. Conversely, the micro-
peptides myoregulin (MLN), phospholamban (PLN),
and sarcolipin (SLN) inhibit SERCA and suppress its
activity [20, 21].
METHODS
FOR MICROPEPTIDE DISCOVERY
The search for and identification of function-
al micropeptides is a multi-step process that begins
with a large-scale screening of genome and tran-
scriptome for potential candidates. Historically, three
main approaches have been developed: bioinformatic
sequence analysis, global translation mapping by ri-
bosome profiling, and direct peptide detection using
mass spectrometry-based proteomics. In the last de-
cade, functional phenotypic screening has been added
to this list, allowing direct identification of micropep-
tides involved in specific cellular processes (Fig. 1).
BIOINFORMATIC APPROACHES
The bioinformatic search for sORFs is a non-triv-
ial task. The early methods of genome analysis used
to predict coding ORFs have set the minimal protein
product length as 100 amino acid residues. This re-
sulted in the loss of information about functional
peptides shorter than this threshold, despite known
existence of such molecules [22].
Firstly, the bioinformatic search for sORFs based
on the evolutionary sequence conservation across
different species [23]. An example of this approach
is the use of the Ka/Ks metric, which reflects the ra-
tio of non-synonymous to synonymous substitutions
in codons in the nucleotide sequence alignment for
different species and indicates whether the sequence
is under selective pressure [24]. More advanced meth-
ods have been developed later, e.g., PhyloCSF, which
uses codon substitution frequencies in both coding
and non-coding genomic regions [25].
To improve the reliability of predictions, other
features indicating the evolutionary conservation of
a sequence should be considered as well, such as the
absence of frameshift-causing insertions or deletions,
a decrease in the sequence conservatism at the edges
of coding regions, and others [26]. However, for short
sequences, the statistical significance of such analysis
is low [27]. Moreover, many sORFs are species-spe-
cific or arise de  novo in the genome, so a functional
peptide might not exhibit the sequence conservatism
and will be missed in analysis [6, 13].
The identification of sORFs is also possible
through the analysis of known features of coding se-
quences, such as their codon composition, GC content,
and others. One of the early algorithms involved iden-
tification of coding sequences by comparing the codon
frequencies in coding and non-coding genomic regions
[22, 28]. Another approach used analysis of six math-
ematical metrics of DNA sequence to assess the cod-
ing potential of sORFs [29]. Machine learning-based
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1523
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig.  1. Main approaches to the large-scale micropeptide discovery. Bioinformatic approaches include analysis of nu-
cleotide sequences to predict the coding potential of sORFs. The key methods are assessment of evolutionary conser-
vation across species, evaluation of sequence properties (e.g., codon composition), and application of machine learning
algorithms to integrate various features. Ribo-Seq (ribosome profiling) is an experimental method that maps all actively
translated regions of the transcriptome. It is based on ribosome stalling, followed by enzymatic cleavage of mRNA and
isolation and high-throughput sequencing of ribosome-protected fragments (footprints) in order to detect sORF transla-
tion in vivo. Mass spectrometry enables direct detection of sORF translation products and can be employed as either bot-
tom-up strategy, in which proteins are first digested by proteases into peptides before analysis, or top-down strategy
that implies analysis of intact micropeptides. Both strategies require careful sample preparation, including cell lysis, frac-
tionation for the enrichment with low-molecular-weight proteins, and analysis by liquid chromatography- tandem mass
spectrometry (LC-MS/MS). Phenotypic screenings aim to identify functional micropeptides. The loss-of-function screen-
ings use guide RNA libraries to knock out sORFs using CRISPR/Cas9, followed by the analysis of changes in the cell phe-
notype (e.g., cell proliferation rate). The gain-of-function screenings use sORF libraries to identify micropeptides whose
overexpression induces a specific cellular response.
analysis, which has become a common research tool
in recent years, allows identification of complex,
non-linear patterns in large datasets, as well as in-
tegration of various features of analyzed sequences,
including their evolutionary conservatism, predicted
structure, and other calculated metrics, to build high-
ly accurate predictive models [30, 31].
Most modern studies on the identification of cod-
ing sORFs use the same approaches for the analysis of
high-throughput sequencing data – ribosome profiling
(Ribo-Seq) and RNA sequencing (RNA-Seq), which sig-
nificantly increases the reliability of obtained results
[32, 33].
RIBOSOME PROFILING
High-throughput ribosome profiling was first
proposed in 2009 [3]. In the classic version of the
method, cells are treated with cycloheximide, which
binds to the E-site of the 60S ribosomal subunit and
inhibits translocation by stopping it at the translation
elongation stage [34]. The cells are then lysed and
treated with nucleases to cleave RNA, while regions
protected by stalled ribosomes are preserved. These
fragments, which are approximately 30 nucleotides
long, are called ribosome footprints. They are puri-
fied and sequenced using high-throughput methods.
LAVROV et al.1524
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
The resulting sequences are mapped onto the tran-
scriptome assembled based on the RNA-Seq data to
obtain a ribosome profile, i.e., the positions of ribo-
somes on mRNAs at the moment of translation ar-
rest with a single-nucleotide resolution [3]. The main
advantage of the approach is its ability to reflect the
translation levels of specific mRNAs at a given time,
thus allowing to study rapid changes in gene expres-
sion in response to external factors or during specific
cellular processes [35, 36].
Despite its broad capabilities, RNA-Seq has sever-
al limitations. In particular, the experimental results
depend significantly on the sample quality. The most
common problem is co-isolation of large ribonucleo-
protein complexes and non-coding RNAs along with
80S ribosomes, leading to false signals [37]. Also,
the percentage of footprint reads from the obtained
library is limited due to the contamination with ri-
bosomal RNA. Another fundamental problem is dis-
tortion of ribosome profiles because of the action of
certain antibiotics [38].
Other difficulties are associated with the analysis
of sORFs. Because of the sORF length, it is sometimes
impossible to accurately determine the translation ini-
tiation site if several potential start codons are located
nearby [36]. Another problem arises when sORFs over-
lap with the main ORF or are located within it [38].
To determine the exact position of start codons,
researchers use translation initiation inhibitors. These
compounds do not disrupt elongation or translation
termination, but stalls ribosomes mostly in the trans-
lation initiation regions, so ribosome profiles provide
accurate information about the position of ribosome
[20]. The first antibiotic used to map translation ini-
tiation sites was harringtonine [39] that binds to the
A-site of free 60S ribosomal subunit. After formation
of the 80S ribosome during translation initiation, har-
ringtonine blocks the transfer of methionine from the
initiator tRNA to the aminoacyl-tRNA located in the
A-site, leading to ribosome stalling at the translation
initiation site [40]. Importantly, harringtonine does not
bind to the 60S subunit in the content of the 80S ribo-
some and does not affect translation elongation and
termination; hence, the obtained footprints show po-
sitions of translation initiation sites [39, 40]. However,
when using harringtonine, some identified footprints
were found for the region located downstream of the
start codon, which, in certain cases, failed to provide
a sufficiently narrow peak for the initiation site map-
ping. A more accurate identification of start codons
can be achieved by using the GTI-Seq (Global Trans-
lation Initiation Sequencing) method, which employs
lactimidomycin to stall translation initiation. Lactimi-
domycin blocks the empty E-site of the 80S ribosome
and causes ribosome to stall immediately after trans-
lation initiation [41]. In a subsequent modification of
the method, QTI-Seq (Quantitative Translation Initia-
tion Sequencing), lysed cells are treated with lactim-
idomycin for a short period of time, followed by the
addition of puromycin to causes the dissociation of
elongating ribosomes. This approach allows to achieve
the maximum coverage of translation initiation sites,
while reducing the noise from elongating ribosomes
and artifacts associated with prolonged incubation of
cells with lactimidomycin [42].
Although ribosome profiling can identify translat-
ed regions, it cannot establish whether the identified
sORFs are coding, regulatory, or non-functional [43].
An indirect evidence of coding potential can be ob-
tained through bioinformatic analysis of candidate
sequences [44, 45]. The existence of an encoded pep-
tide in the cell can be proven by mass spectrometry
methods [46, 47].
MASS SPECTROMETRY
Mass spectrometry-based proteomic analysis al-
lows direct detection of sORF translation products.
As in classical proteomics, either peptide fragments
after digestion with proteases (bottom-up approach)
or intact proteins (top-down approach) can be ana-
lyzed. In both cases, identification of micropeptides
faces two main obstacles. First, it requires the use
of special sample preparation procedures to enrich
the sample with micropeptides because of their small
size, low stability, and low cellular abundance. Sec-
ond, identification of previously unannotated micro-
peptides requires the use of special reference data-
bases, as databases based on Ensembl, Ref-Seq, or
UniProt do not contain most potential micropeptides.
To prevent the degradation of micropeptides,
cell lysates are heated to 95°C to inactivate proteas-
es [5]. In some cases, this procedure is followed by
precipitation of large proteins with trichloroacetic
acid [48]. The resulting samples are fractionated by
polyacrylamide gel electrophoresis (PAGE) [49,  50],
size-exclusion chromatography [48], or reverse-phase
chromatography [51] to enriched them with low-mo-
lecular-weight proteins. More advanced methods of
preliminary fractionation have been developed, such
as GELFrEE (Gel-Eluted Liquid Fraction Entrapment
Electrophoresis, i.e., separation on a column contain-
ing polyacrylamide gel) [52, 53], ERLIC (Electrostatic
Repulsion Hydrophilic Interaction Chromatography,
i.e., chromatography based on electrostatic repulsion
and hydrophilic interaction) [54], and others. Depend-
ing on the selected approach, the proteins are then
either subjected to proteolysis (most often, with tryp-
sin) for the bottom-up analysis or directly analyzed by
liquid chromatography combined with tandem mass
spectrometry (LC-MS/MS).
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1525
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
The identification of micropeptides is based on
matching experimental mass spectra with the theoret-
ical ones. However, construction of databases poses a
problem, as in  silico translation of all reading frames
in the genome or transcriptome covers almost all po-
tential micropeptides, but the size of the resulting da-
tabase is tens of times larger than that of standard
databases, leading to a high computational complex-
ity and high probability of false identifications [55].
Therefore, limited databases based on RNA-seq and
Ribo-Seq data are used [5,  56], which, however, does
not exclude the appearance of false-positive results.
In order to confirm an identified micropeptide its
isotopically labeled synthetic analog, which has an
identical mass spectrum that is shifted relative to the
spectrum of the validated peptide, is added in the re-
action [57, 58].
Micropeptides can be studied using the top-down
proteomics approach, which is particularly useful
for analyzing various proteoforms of micropeptides
formed by alternative splicing [59] and post-transla-
tional modifications [57]. For example, identification
of micropeptides presented by the major histocom-
patibility complex (MHC-I) allows to detect intact pep-
tides. A database of translated sequences obtained by
the ribosome profiling of cancer cells, was used to
identify thousands of previously unannotated pep-
tides by mass spectrometry [60].
PHENOTYPIC SCREENINGS
In addition to the analysis of sequences and
translation data, functional sORFs can be discovered
through phenotypic screening, which can be divided
into two approaches: loss-of-function screening and
gain-of-function screening.
CRISPR/Cas9 system has become the main tool for
the loss-of-function screening. In these experiments,
single-guide RNAs (sgRNAs) target Cas9 nuclease to
specific sORF regions due to a 20-nucleotide guide
sequence in the sgRNA and the presence of the pro-
tospacer adjacent motif (PAM) sequence immediate-
ly downstream of the targeted DNA region. Double-
strand breaks caused by Cas9 are repaired mostly
through the non-homologous end joining (NHEJ), lead-
ing to the insertions or deletions of nucleotides that
can cause frameshifts and appearance of premature
stop codons, thus preventing the biosynthesis of a
functional micropeptide [61].
Genome-wide and targeted screenings use the
sgRNA libraries targeting a large number of sORFs.
The libraries are typically delivered to the cells by
lentiviral vectors, allowing inactivation of thousands
of genomic loci in a single experiment [62]. Cells with
specific phenotypes resulting from the sORF knockout
are selected from the general population, and the cas-
settes containing the sgRNA sequences are sequenced
to identify the sORFs whose knockout caused chang-
es in the cell phenotype. Averaging the results across
multiple sgRNAs for a specific sORF allows to account
for the potential off-target effects of CRISPR/Cas9.
The phenotypes observed after CRISPR-mediated
inactivation of sORFs can range from changes in cell
growth, viability, morphology, signaling pathways, and
drug resistance to interactions with other molecules.
A specific phenotype used in the screening is chosen
in each particular experiment [62].
So far, CRISPR/Cas9-mediated phenotype-based
screenings have not yet been commonly accepted as
a method for identification of functional micropep-
tides. However, several functional micropeptides have
been discovered in proliferation screenings of librar-
ies based on Ribo-Seq data [63-65].
An alternative approach to the phenotype analysis
involves exogenous expression of constructs encoding
peptides from a pre-compiled library and cloned into
plasmid or lentiviral vectors [66]. The advantage of
this method is that it allows to analyze any sequence
without requiring the presence of specific motifs (as
in the case of CRISPR/Cas9). However, this approach
has the same drawbacks as the commonly used over-
expression of proteins for studying their functions.
CONFIRMATION OF EXISTENCE
AND CHARACTERIZATION OF MICROPEPTIDES
Although the methods for investigating micropep-
tides are similar to those used for conventional pro-
teins, the overall approaches to studying these classes
of molecules differ. Unlike “classical” proteins, most
micropeptides do not contain potential structural
domains, have a small size, and are present in cells
at low concentrations, which requires adjustment of
standard approaches. Moreover, if a micropeptide is
encoded in an lncRNA, it may be necessary to distin-
guish between the functional roles of the transcript
and the peptide encoded within it. The studies of
micropeptides include three main stages: candidate
selection, confirmation of its existence, and determi-
nation of its function.
There are two main approaches for the search
of potential candidates. The first is the analysis of
data obtained by high-throughput methods (see the
first part of the review). Typically, these methods
provide only indirect evidence of the micropeptide
existence, so they have to be supplemented by mass
spectrometry and Ribo-Seq data [46]. A completely
different approach is a manual search for sORFs and
annotation of the transcripts of interest. For exam-
ple, the micropeptide MIEF1-mp was found during the
LAVROV et al.1526
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig.  2. Main strategies for studying micropeptides. The diagram shows key methods used to confirm the existence and
determine the structure, location, function, and molecular partners of a selected micropeptide candidate. The structure of
the micropeptide is determined by analysis of its secondary and tertiary structures. Thus, circular dichroism spectrosco-
py is used to assess elements of the secondary structure (α-helices, β-sheets). Nuclear magnetic resonance (NMR) allows
to determine the three-dimensional structure of small peptides in solution, while X-ray crystallography and cryo-elec-
tron microscopy are more often used to study the complexes of micropeptides with their partner proteins. Affinity la-
beling is a central approach in the studies of micropeptides. Fusing a micropeptide with short affinity tags (FLAG, HA,
6xHis) or fluorescent proteins (GFP, mCherry) allows its visualization in the cells by fluorescence microscopy, detection
by Western blotting, and isolation of protein complexes by co-immunoprecipitation (Co-IP). An alternative to affinity la-
beling is generation of specific antibodies to native micropeptides. The studies of micropeptide interactions includes sev-
eral approaches. Chemical labeling of synthetic micropeptides (e.g., with biotin) allows to search for the interacting pro-
teins in vitro in pull-down experiments, whereas weak or transient interactions in vivo can be investigated by proximity
labeling methods using BirA ligase (BioID) or ascorbate peroxidase (APEX), which covalently label neighboring proteins.
analysis of the 5′-UTR sequence of the MIEF1 gene
mRNA that encodes a described protein product [67].
Another example of this approach is the discovery of
the cancer-associated peptide PACMP in the analysis
of differential expression of lncRNAs in breast cancer
cells [68].
The second stage is confirmation of the micro-
peptide presence in the cells, as the protein product
of sORF translation may be unstable. In some cases,
it might be necessary to perform RACE (Rapid Am-
plification of cDNA Ends) to confirm the sequence of
the target transcript in the cell [69]. Direct evidence
of the micropeptide biosynthesis can be obtained by
mass spectrometry. For example, the presence of the
APPLE peptide was confirmed by mass spectrometry
after immunoprecipitation of the endogenously la-
beled peptide [46]. However, in most cases, the confir-
mation of translation of the studied sORF and biosyn-
thesis of a stable protein product can be achieved in
a simpler experiment by using exogenous expression
of the micropeptide fused with an affinity tag or a
fluorescent protein [70, 71].
The next step after confirmation of the micro-
peptide biosynthesis is elucidation of its function. The
methods used to study micropeptides are almost the
same as those used for conventional proteins and
include affinity labeling, Western blotting, immuno-
cytochemistry (ICC), proximity labeling, etc. (Fig.  2),
although most of these classical methods have to be
modified due to the small size of micropeptides.
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1527
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Fig.  3. The scheme of studies of the oncogenic micropeptide APPLE as an example of a comprehensive approach to mi-
cropeptide characterization. The diagram illustrates the three main stages of the study. I. Candidate selection: integrative
analysis of high-throughput Ribo-Seq, mass spectrometry, and RNA-Seq data led to the identification of the lncRNA ASH1L-AS1
as a potential source of the micropeptide. II. Confirmation of the micropeptide existence: APPLE biosynthesis in cells was
confirmed through two independent methods. (1)Exogenous expression: the coding sequence of APPLE was fused with var-
ious affinity and fluorescent tags (GFP, FLAG, HA), and the expression products were detected by fluorescence microscopy
and Western blotting. (2)Endogenous tagging: the FLAG tag was inserted into the native gene locus using the CRISPR/Cas9
genome editing system, which enabled isolation of the endogenous peptide via immunoprecipitation. III. Elucidation of
micropeptide function by functional analysis. (1) Studying the effect of peptide expression on the cell phenotype. Inactiva-
tion of APPLE expression led to cell apoptosis and proliferation arrest, while its overexpression promoted cell proliferation,
indicating the oncogenic role of APPLE. (2) Determination of APPLE intracellular location: immunofluorescence analysis re-
vealed that APPLE localizes to the endoplasmic reticulum. (3)Search for the APPLE protein partners. Co-IP followed by mass
spectrometry identified PABPC1 and eIF4G as micropeptide interacting partners. Subsequent studies confirmed that APPLE
is a part of the cap-binding complex and enhances translation initiation, thus promoting the development of malignancies.
The transcript of the lncRNA ASH1L-AS1 as a
potential source of the functional APPLE micropep-
tide, which promotes development of hematopoietic
malignancies, was first identified by integration of Ri-
bo-Seq, mass spectrometry, and RNA-Seq data (Fig.3).
The biosynthesis of APPLE was confirmed by multiple
methods, including generation of specific antibodies
against the micropeptide and endogenous tagging
with the CRISPR/Cas9 system, followed by detection
by Western blotting, immunoprecipitation, and mass
spectrometry analysis.
The functional role of APPLE in oncogenesis
was established using knockdown and overexpres-
sion experiments, which demonstrated its effects on
proliferation, differentiation, and apoptosis of acute
myeloid leukemia cells, both in  vitro and in in vivo
(in a mouse model). A critical step in this study was
the phenotypic complementation experiment on the
background of the APPLE inactivation, which con-
firmed that the observed effects were specifical-
ly linked to the micropeptide rather than lncRNA
ASH1L-AS1 [46].
The information on the methods used to char-
acterize micropeptides is summarized in Online Re-
source 1.
KNOCKOUT AND KNOCKDOWN
To confirm the biological function of micropep-
tides, researchers can generate the cells in which the
corresponding encoding sequences are knocked-out
with genome editing systems (CRISPR/Cas9 or CRISPR-
Cas12a). Analysis of phenotypic changes in these cells
allows for assessing the role of investigated peptide.
However, in some cases, selecting guide RNAs may
be impossible due to the requirement for a PAM mo-
tif. The commonly used phenotypic markers are cell
proliferation rate, changes in the cell cycle, and oth-
ers [72]. An additional confirmation of the biological
LAVROV et al.1528
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
significance of a micropeptide is the phenotype res-
toration through its exogenous expression in the
cells where this micropeptide had been the knocked
out [68].
A much less labor-intensive method for suppress-
ing expression of a target micropeptide is RNA in-
terference (RNAi)-mediated knockdown. Thus, small
interfering RNAs were used in the studies of micro-
peptides translated from lncRNAs [46, 73]. However,
in the case of sORFs encoded by mRNAs, RNAi-me-
diated knockdown will lead to the degradation of
the entire mRNA molecule, and consequently, the
observed effects may be associated with a decrease
in the expression of the main translation product of
this mRNA. In such cases, 2′-O-methyl antisense RNA
oligonucleotides complementary to the region around
the sORF start codon can be used to suppress the
biosynthesis of the micropeptide but not of the main
protein product [74].
AFFINITY TAGGING OF MICROPEPTIDES
Another important method in studying micropep-
tides is introduction of affinity tags. The small size of
micropeptides imposes restrictions on the choice of
the tag, as for many of them, the size of the intro-
duced sequence is comparable to the size of the pep-
tide itself. Moreover, the charge of the tag also plays
a significant role. A case of incorrect choice the tag
can disrupt the native localization and structure of the
micropeptide or its interaction with partner proteins.
The most common tags used with micropeptides
are small protein sequences, such as HA [75], FLAG
[11,  46], 6xHis [70]. In some studies, GFP or another
fluorescent protein are used. Fluorescent proteins can
be easily detected by microscopy and Western blotting
[10,  70], but there is a risk that the functions of the
tagged micropeptide in the cells can be disrupted.
Endogenous tagging by CRISPR/Cas9 is also used
to confirm the existence of micropeptides in the cells.
Insertion of affinity tag-encoding sequences directly
into the genomic locus allows for the detection of
micropeptides expressed from the native promoters,
which excludes artifacts associated with the overex-
pression of micropeptides and is a reliable method
for validating their biosynthesis.
The methods of affinity tagging of micropeptides
are discussed in detail in a recent review [76].
WESTERN BLOTTING
Western blotting is a standard method for detec-
tion and semi-quantitative determination of proteins.
However, the small size of micropeptides limits the
number of highly antigenic suitable epitopes, making
it difficult to generate antibodies against micropep-
tides. Nevertheless, for some peptides (over 50 amino
acids in length), the antibodies were successfully ob-
tained and used for Western blotting [70-78]. In most
cases, micropeptides are labeled with various affinity
tags (see above) for further detection [10, 70, 79]. It is
important to note that the content of native or endog-
enously tagged micropeptides in the cells can be very
low, making their detection extremely challenging.
Another limitation is the resolving power of
polyacrylamide gels used for the separation of pro-
teins from cell lysates. To increase the resolution of
low-molecular-weight proteins, researchers use the
Tris-tricine buffer system [80]. For example, it was
employed for the detection of the FLAG-tagged HOXB-
AS3 micropeptide (7  kDa) and native form of myoreg-
ulin (10  kDa) [71, 81]. However, small proteins weak-
ly bind to polyvinylidene fluoride and nitrocellulose
membranes used for the transfer and can be easily
lost during multiple membrane washes. In such cases,
alternative techniques can be applied, such as fixa-
tion on the membrane by crosslinking with blocking
proteins using formaldehyde or glutaraldehyde [82].
DETERMINATION OF MICROPEPTIDE
INTRACELLULAR LOCATION
The location of a micropeptide can help in elu-
cidating its protein partners and biological function.
A micropeptide located in the nucleus is likely to
interact with nuclear proteins, such as transcription
factors or chromatin-organizing proteins.
The most common method for determining the
location of micropeptides is immunocytochemistry,
which relies on the use of antibodies with a high spec-
ificity for the target antigen. The procedure typically
begins with cell fixation to preserve their morphology
and prevent degradation of cellular components. The
cells are then permeabilized to allow antibody access
into the cell. Generally, primary antibodies used in
the analysis are unlabeled and visualized using sec-
ondary antibodies conjugated with a fluorophore.
Similar to Western blotting, this method requires
antibodies against the target micropeptide, which can
be an obstacle. For example, antibodies against the
native micropeptide MP31 were used to determine its
mitochondrial location [83]. More often, immunocyto-
chemical staining is performed for the exogenously
expressed tagged micropeptide [45, 72, 79].
An alternative, widely used approach is the fu-
sion of micropeptides with fluorescent proteins (GFP,
mCherry, etc.), which allows their direct visualization
in the cells without the use of antibodies. However,
it may disrupt the properties of the micropeptide
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1529
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
because of the large size of the tag. To overcome this
limitation, split fluorescent proteins can be used, when
fluorescent protein, such as GFP, is divided into two
non-functional fragments – large fragment GFP1-10
and small GFP11, consisting of 16 amino acids and
used to tag the peptide of interest. The complementa-
tion of these fragments restores the tertiary structure
of the protein and its fluorescence [84, 85]. This system
was used to determine the location of the PIGBOS mi-
cropeptide, when the peptide tagged with three GFP11
repeats was co-expressed with GFP1-10 [77].
Chemical fluorescent tags have the least impact
on the native localization of micropeptides. However,
such experiments can be technically difficult due to
the need for the chemical synthesis of the micropep-
tide for its chemical labeling with a fluorescent ligand
(e.g., FITC). This labeling method was used to deter-
mine the location of the MP155 micropeptide [78].
The MicroID method is another tagging-based
method for identification of new micropeptides with
specific localization. This technique (a modification
of the BioID method, see below) uses biotin ligases
targeted to specific cellular compartments. After iso-
lation of covalent complexes, those containing ligase
crosslinked with low-molecular-weight proteins are
selected by fractionation and the bound micropep-
tides are identified by mass spectrometry [86].
IDENTIFICATION OF PROTEIN PARTNERS
Identification of protein partners of micropep-
tides is critically important for understanding the bi-
ological mechanisms involving these micropeptides.
The most widely used approach for studying protein–
protein interactions in vivo is Co-IP, in which cells ex-
pressing a micropeptide labeled with an affinity tag
(most commonly, FLAG or HA) are lysed under mild
conditions that preserve existing protein complexes.
Antibodies specific to the affinity tag are added to
the cell lysate to bind the tagged micropeptide along
with all proteins that interact with it. The resulting
immune complexes are isolated from the solution,
usually with agarose or magnetic beads coated with
antibody-binding proteins  A or G. After thorough
washing to remove non-specifically bound proteins,
the captured complexes are eluted and identified,
most often by immunoprecipitation mass spectrome-
try (IP-MS) or Western blotting. This method was used
to demonstrate the interaction of the CYREN micro-
peptide with Ku70/80 and other proteins involved in
DNA repair [14]. The main requirement for Co-IP is a
sufficient stability of protein complexes to withstand
the lysis and washing procedures. However, the bind-
ing of a micropeptide to its protein partners is often
not strong enough.
The pull-down is a method similar to Co-IP, ex-
cept that the “bait” is typically a purified, tagged mi-
cropeptide immobilized on a solid phase. This can be
achieved through exogenous expression (e.g., with the
GST or 6xHis tag) or chemical synthesis, often with
the addition of the biotin tag. The immobilized bait is
incubated with the cell lysate or a solution of purified
proteins. Proteins that bind to the bait are retained
on the carrier. The carrier is washed, and the pro-
teins are then eluted and analyzed. For example, a
pull-down experiment with biotinylated P155 peptide
immobilized on streptavidin beads revealed the in-
teraction between this peptide and HSC70 chaperone
[78]. Unlike Co-IP, pull-down is convenient for con-
firming direct protein interactions in vitro and does
not require specific antibodies. However, the results
may not take into account the influence of the cel-
lular context or the post-translational modifications
necessary for the interactions in vivo.
To detect weaker or transient interactions in vivo,
proximity labeling methods, such as BioID and APEX,
have been developed [87,  88]. These methods are
based on the expression of a chimeric protein con-
sisting of the target micropeptide fused with a special
enzyme. The enzyme is activated by the addition of
specific substrates that generate short-lived reactive
molecules (usually, biotin derivatives), which cova-
lently bind to proteins in a close proximity (within a
few nanometers) to the chimeric protein. The BioID
method uses the mutant biotin ligase BirA*, which,
in the presence of biotin and ATP, generates acti-
vated biotin-AMP that reacts with lysine residues of
neighboring proteins  [89]. In the APEX method, ascor-
bate peroxidase in the presence of biotin-phenol and
a short pulse of hydrogen peroxide rapidly (within
minutes) generates the biotin-phenoxyl radical that la-
bels tyrosine residues in a close proximity to the en-
zyme. After the labeling reaction, the cells are lysed,
biotinylated proteins are isolated using streptavidin
and identified by mass spectrometry [89]. The APEX
labeling method was used to search for the protein
partners of the mitochondrial peptide MIEF1 [79].
APEX labeling can also provide information on the
subcellular localization of micropeptides and their
functional environment, as demonstrated for C11orf98
in the nucleolus [90].
DETERMINATION
OF MICROPEPTIDE STRUCTURE
X-ray crystallography is rarely used for deter-
mining the three-dimensional structure of micro-
peptides, as it requires crystallization of proteins
or peptides into an ordered three-dimensional lat-
tice to obtain a diffraction pattern suitable for the
LAVROV et al.1530
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
structure resolution. Because of relatively small size,
micropeptides often do not form stable and well-or-
dered crystals suitable for obtaining high-resolution
diffraction patterns. Also, most micropeptides do
not have their own function but act as modulators
of protein partners and structural units of protein
complexes, so the crystal structure of the peptide
itself is often uninformative. However, X-ray crystal-
lography can be used to determine the structure of
peptide complexes with protein partners, for exam-
ple, the structure of phospholamban complex with
SERCA [91].
The presence of a stable secondary structure can
indicate the micropeptide stability in the cells and its
biological function. A relatively simple method for de-
termining the secondary structure of proteins is cir-
cular dichroism (CD) spectroscopy, which is based on
the differential absorption of circularly polarized light
by chiral molecules. Proteins, which are composed of
chiral amino acids, are optically active and exhibit
characteristic CD spectra. While this method does not
directly determine the tertiary structure, it is sensi-
tive to the presence of certain secondary structure
elements in the protein and can be used to elucidate
the types of secondary structures present in a micro-
peptide [75]. This approach was used to describe the
interaction of the CYREN peptide with the Ku70/Ku80
protein complex [14].
A more informative method for resolving the
tertiary structure of peptides is nuclear magnetic
resonance (NMR). Most micropeptides have small
molecular weights (up to 11 kDa) and, therefore, are
good subjects for the structure determination by NMR
[92]. This method was used for the characterization of
the DWORF structure and revealed a proline-induced
bend necessary for the activation of SERCA [19].
CONCLUSION
The studies of micropeptides encoded by sORFs
expand our understanding of the genome coding
potential and complexity of the proteome. Over the
past 10-15 years, a large body of evidence has been
accumulated regarding the functional role of micro-
peptides. As a result, sORFs are no longer considered
as non-functional sequences, as it has become clear
that they provide a new level in the regulation of
cellular processes. This paradigm shift has become
possible due to the prior development of molecular
biology methods. Ribosome profiling has revealed the
scale of sORF translation, while improvement in sam-
ple preparation for mass spectrometry allowed direct
detection of their peptide products, and CRISPR/Cas9-
based genome editing provided a powerful tool for
the functional studies.
Despite a significant progress, detection of micro-
peptides is still associated with certain challenges. It is
important to reliably differentiate sORFs that encode
micropeptides from those translated but non-function-
al or those that act as translational regulatory ele-
ments, which may require a deeper investigation of
their evolutionary origin and analysis by molecular
biology methods. The confirmation of the existence
of micropeptides and characterization of their prop-
erties remain a non-trivial task due to small size of
these molecules and limitations it imposes on the
use of classical molecular biology methods. Current
strategies used for micropeptide discovery may need
certain improvements. For example, the detection and
quantification of short-lived micropeptides require
further increase in the sensitivity of mass spectrom-
etry methods, optimization of sample preparation pro-
tocols, and generation of sufficiently comprehensive
but non-redundant reference databases.
Nevertheless, a substantial amount of data on mi-
cropeptides has already been accumulated. Future re-
search will likely involve more comprehensive integra-
tion of genomics, transcriptomics, and proteomics data.
This multi-omics approach will open new horizons in
studying the role of micropeptides and their potential
applications in disease diagnostics and therapy.
Abbreviations
Co-IP co-immunoprecipitation
GFP green fluorescent protein
Ribo-Seq ribosome profiling
RNA-seq RNA sequencing
sgRNA single-guide RNA
sORF small open reading frame
Supplementary information
The online version contains supplementary material
available at https://doi.org/10.1134/S0006297925602242.
Contributions
A.I.L. wrote the text of the article; O.A.D., M.P.R., and
N.M.Sh. edited the manuscript.
Funding
This work was supported by the Russian Science
Foundation (project no. 23-14-00058) and as part of
the State Assignment to the Lomonosov Moscow State
University (registration no.121031300037-7).
Ethics approval and consent to participate
This work does not contain any studies involving hu-
man and animal subjects.
Conflict of interest
The authors of this work declare that they have
noconflicts of interest.
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1531
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
Open access
This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits
use, sharing, adaptation, distribution, and reproduc-
tion in any medium or format, as long as you give
appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons li-
cense, and indicate if changes were made. The images
or other third party material in this article are includ-
ed in the article’s Creative Commons license, unless
indicated otherwise in a credit line to the material.
If material is not included in the article’s Creative
Commons license and your intended use is not permit-
ted by statutory regulation or exceeds the permitted
use, you will need to obtain permission directly from
the copyright holder. To view a copy of this license,
visit http://creativecommons.org/licenses/by/4.0/.
REFERENCES
1. Basrai, M.A., Hieter, P., and Boeke, J.D. (1997) Small
open reading frames: beautiful needles in the hay-
stack, Genome Res., 7, 768-771, https://doi.org/10.1101/
gr.7.8.768.
2. Harrison, P.M. (2002) A question of size: the eukary-
otic proteome and the problems in defining it, Nu-
cleic Acids Res., 30, 1083-1090, https://doi.org/10.1093/
nar/30.5.1083.
3. Ingolia, N. T., Ghaemmaghami, S., Newman, J. R. S.,
and Weissman, J. S. (2009) Genome-wide analysis in
vivo of translation with nucleotide resolution using
ribosome profiling, Science, 324, 218-223, https://
doi.org/10.1126/science.1168978.
4. Chong, C., Müller, M., Pak, H., Harnett, D., Huber, F.,
Grun, D., Leleu, M., Auger, A., Arnaud, M., Steven-
son, B. J., Michaux, J., Bilic, I., Hirsekorn, A., Calvi-
ello, L., Simó-Riudalbas, L., Planet, E., Lubiński, J.,
Bryśkiewicz, M., Wiznerowicz, M., Xenarios, I.,
Zhang,L., Trono,D., Harari,A., Ohler,U., Coukos,G.,
and Bassani-Sternberg, M. (2020) Integrated pro-
teogenomic deep sequencing and analytics accu-
rately identify non-canonical peptides in tumor im-
munopeptidomes, Nat. Commun., 11, 1293, https://
doi.org/10.1038/s41467-020-14968-9.
5. Slavoff, S. A., Mitchell, A. J., Schwaid, A. G., Cabili,
M.N., Ma,J., Levin, J.Z., Karger, A. D., Budnik, B.A.,
Rinn, J. L., and Saghatelian, A. (2013) Peptidom-
ic discovery of short open reading frame-encoded
peptides in human cells, Nat. Chem. Biol., 9, 59-64,
https://doi.org/10.1038/nchembio.1120.
6. Sandmann, C.-L., Schulz, J. F., Ruiz-Orera, J.,
Kirchner, M., Ziehm, M., Adami, E., Marczenke, M.,
Christ, A., Liebe, N., Greiner, J., Schoenenberger, A.,
Muecke, M. B., Liang, N., Moritz, R. L., Sun, Z.,
Deutsch, E.W., Gotthardt,M., Mudge, J.M., Prensner,
J. R., Willnow, T. E., Mertins, P., Van Heesch, S., and
Hubner, N. (2023) Evolutionary origins and interac-
tomes of human, young microproteins and small
peptides translated from short open reading frames,
Mol. Cell, 83, 994-1011.e18, https://doi.org/10.1016/
j.molcel.2023.01.023.
7. Cao,X., Sun,S., and Xing,J. (2024) A massive proteog-
enomic screen identifies thousands of novel peptides
from the human “dark” proteome, Mol. Cell. Pro-
teomics, 23, 100719, https://doi.org/10.1016/j.mcpro.
2024.100719.
8. Pauli,A., Norris, M.L., Valen,E., Chew, G.-L., Gagnon,
J.A., Zimmerman,S., Mitchell,A., Ma,J., Dubrulle,J.,
Reyon,D., Tsai, S.Q., Joung, J.K., Saghatelian,A., and
Schier, A.F. (2014) Toddler: an embryonic signal that
promotes cell movement via apelin receptors, Science,
343, 1248636, https://doi.org/10.1126/science.1248636.
9. Kondo, T., Hashimoto, Y., Kato, K., Inagaki, S.,
Hayashi, S., and Kageyama, Y. (2007) Small peptide
regulators of actin-based cell morphogenesis encoded
by a polycistronic mRNA, Nat. Cell. Biol., 9, 660-665,
https://doi.org/10.1038/ncb1595.
10. Chugunova,A., Loseva,E., Mazin,P., Mitina,A., Nava-
layeu,T., Bilan,D., Vishnyakova,P., Marey,M., Golovi-
na, A., Serebryakova, M., Pletnev, P., Rubtsova, M.,
Mair,W., Vanyushkina,A., Khaitovich,P., Belousov,V.,
Vysokikh, M., Sergiev, P., and Dontsova, O. (2019)
LINC00116 codes for a mitochondrial peptide linking
respiration and lipid metabolism, Proc. Natl. Acad.
Sci. USA, 116, 4940-4945, https://doi.org/10.1073/
pnas.1809105116.
11. Matsumoto,A., Pasut,A., Matsumoto,M., Yamashita,R.,
Fung, J., Monteleone, E., Saghatelian, A., Nakayama,
K.I., Clohessy, J.G., and Pandolfi, P.P. (2017) mTORC1
and muscle regeneration are regulated by the
LINC00961-encoded SPAR polypeptide, Nature, 541,
228-232, https://doi.org/10.1038/nature21034.
12. Dong,X., Zhang,K., Xun,C., Chu,T., Liang,S., Zeng,Y.,
and Liu,Z. (2023) Small Open reading frame-encoded
micro-peptides: an emerging protein world, Int. J. Mol.
Sci., 24, 10562, https://doi.org/10.3390/ijms241310562.
13. Couso, J.-P., and Patraquim, P. (2017) Classification
and function of small open reading frames, Nat. Rev.
Mol. Cell Biol., 18, 575-589, https://doi.org/10.1038/
nrm.2017.58.
14. Xie,L., Bowman, M.E., Louie, G.V., Zhang,C., Ardeja-
ni, M.S., Huang,X., Chu,Q., Donaldson, C.J., Vaughan,
J.M., Shan,H., Powers, E.T., Kelly, J.W., Lyumkis,D.,
Noel, J. P., and Saghatelian, A. (2023) Biochemistry
and protein interactions of the CYREN microprotein,
Biochemistry, 62, 3050-3060, https://doi.org/10.1021/
acs.biochem.3c00397.
15. Zhu, K.-G., Yang,J., Zhu,Y., Zhu,Q., Pan,W., Deng,S.,
He, Y., Zuo, D., Wang, P., Han, Y., and Zhang, H.-Y.
(2023) The microprotein encoded by exosomal
lncAKR1C2 promotes gastric cancer lymph node
LAVROV et al.1532
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
metastasis by regulating fatty acid metabolism, Cell
Death Dis., 14, 708, https://doi.org/10.1038/s41419-
023-06220-1.
16. Zheng,W., Guo,Y., Zhang,G., Bai,J., Song,Y., Song,X.,
Zhu, Q., Bao, X., Wu, G., and Zhang, C. (2023) Pep-
tide encoded by lncRNA BVES-AS1 promotes cell
viability, migration, and invasion in colorectal can-
cer cells via the SRC/mTOR signaling pathway, PLoS
One, 18, e0287133, https://doi.org/10.1371/journal.
pone.0287133.
17. Guo, B., Zhai, D., Cabezas,E., Welsh, K., Nouraini, S.,
Satterthwait, A.C., and Reed, J.C. (2003) Humanin pep-
tide suppresses apoptosis by interfering with Bax ac-
tivation, Nature, 423, 456-461, https://doi.org/10.1038/
nature01627.
18. Makarewich, C. A., Munir, A. Z., Schiattarella, G. G.,
Bezprozvannaya, S., Raguimova, O. N., Cho, E. E.,
Vidal, A. H., Robia, S. L., Bassel-Duby, R., and Olson,
E.N. (2018) The DWORF micropeptide enhances con-
tractility and prevents heart failure in a mouse mod-
el of dilated cardiomyopathy, Elife, 7, e38319, https://
doi.org/10.7554/eLife.38319.
19. Reddy, U. V., Weber, D. K., Wang, S., Larsen, E. K.,
Gopinath, T., De Simone, A., Robia, S., and Veglia, G.
(2022) A kink in DWORF helical structure controls the
activation of the sarcoplasmic reticulum Ca
2+
- ATPase,
Structure, 30, 360-370, https://doi.org/10.1016/j.str.
2021.11.003.
20. Tonkin,J., and Rosenthal,N. (2015) One small step for
muscle: a new micropeptide regulates performance,
Cell Metab., 21, 515-516, https://doi.org/10.1016/
j.cmet.2015.03.013.
21. Anderson, D.M., Makarewich, C.A., Anderson, K.M.,
Shelton, J. M., Bezprozvannaya, S., Bassel-Duby, R.,
and Olson, E.N. (2016) Widespread control of calcium
signaling by a family of SERCA-inhibiting micropep-
tides, Sci. Signal., 9, aaj1460, https://doi.org/10.1126/
scisignal.aaj1460.
22. Frith, M. C., Forrest, A. R., Nourbakhsh, E., Pang,
K. C., Kai, C., Kawai, J., Carninci, P., Hayashizaki, Y.,
Bailey, T. L., and Grimmond, S. M. (2006) The abun-
dance of short proteins in the mammalian proteome,
PLoS Genet., 2, e52, https://doi.org/10.1371/journal.
pgen.0020052.
23. Kute, P.M., Soukarieh,O., Tjeldnes,H., Trégouët, D.-A.,
and Valen,E. (2022) Small open reading frames, how to
find them and determine their function, Front. Genet.,
12, 796060, https://doi.org/10.3389/fgene.2021.796060.
24. Hurst,L. (2002) The Ka/Ks ratio: diagnosing the form
of sequence evolution, Trends Genet., 18, 486-487,
https://doi.org/10.1016/s0168-9525(02)02722-1.
25. Lin, M.F., Jungreis,I., and Kellis,M. (2011) PhyloCSF:
a comparative genomics method to distinguish pro-
tein coding and non-coding regions, Bioinformatics,
27, i275-i282, https://doi.org/10.1093/bioinformatics/
btr209.
26. Mackowiak, S. D., Zauber, H., Bielow, C., Thiel, D.,
Kutz, K., Calviello, L., Mastrobuoni, G., Rajewsky, N.,
Kempa, S., Selbach, M., and Obermayer, B. (2015)
Extensive identification and analysis of conserved
small ORFs in animals, Genome Biol., 16, 179, https://
doi.org/10.1186/s13059-015-0742-x.
27. Ladoukakis,E., Pereira,V., Magny, E.G., Eyre-Walker,A.,
and Couso, J. P. (2011) Hundreds of putatively func-
tional small open reading frames in Drosophila, Ge-
nome Biol., 12, R118, https://doi.org/10.1186/gb-2011-
12-11-r118.
28. Badger, J. H., and Olsen, G. J. (1999) CRITICA: coding
region identification tool invoking comparative analy-
sis, Mol. Biol. Evol., 16, 512-524, https://doi.org/10.1093/
oxfordjournals.molbev.a026133.
29. Kong, L., Zhang, Y., Ye, Z.-Q., Liu, X.-Q., Zhao, S.-Q.,
Wei,L., and Gao,G. (2007) CPC: assess the protein-cod-
ing potential of transcripts using sequence features
and support vector machine, Nucleic Acids Res., 35,
W345-W349, https://doi.org/10.1093/nar/gkm391.
30. Chen, Z., Meng, J., Zhao, S., Yin, C., and Luan, Y.
(2023) sORFPred: a method based on comprehen-
sive features and ensemble learning to predict the
sORFs in plant LncRNAs, Interdiscip. Sci. Comput.
Life Sci., 15, 189-201, https://doi.org/10.1007/s12539-
023-00552-4.
31. Zhao,S., Meng,J., Kang,Q., and Luan,Y. (2022) Iden-
tifying LncRNA-encoded short peptides using opti-
mized hybrid features and ensemble learning, Trans.
Comput. Biol. and Bioinf., 19, 2873-2881, https://
doi.org/10.1109/TCBB.2021.3104288.
32. Raj, A., Wang, S. H., Shim, H., Harpak, A., Li, Y. I.,
Engelmann,B., Stephens,M., Gilad,Y., and Pritchard,
J. K. (2016) Thousands of novel translated open
reading frames in humans inferred by ribosome
footprint profiling, Elife, 5, e13328, https://doi.org/
10.7554/eLife.13328.
33. Bazzini, A.A., Johnstone, T.G., Christiano,R., Macko-
wiak, S.D., Obermayer,B., Fleming, E.S., Vejnar, C.E.,
Lee, M. T., Rajewsky, N., Walther, T. C., and Giraldez,
A.J. (2014) Identification of small ORFs in vertebrates
using ribosome footprinting and evolutionary conser-
vation, EMBO J., 33, 981-993, https://doi.org/10.1002/
embj.201488411.
34. Schneider-Poetsch, T., Ju, J., Eyler, D. E., Dang, Y.,
Bhat, S., Merrick, W. C., Green, R., Shen, B., and
Liu, J. O. (2010) Inhibition of eukaryotic translation
elongation by cycloheximide and lactimidomycin,
Nat. Chem. Biol., 6, 209-217, https://doi.org/10.1038/
nchembio.304.
35. Brar, G. A., and Weissman, J.S. (2015) Ribosome pro-
filing reveals the what, when, where and how of pro-
tein synthesis, Nat. Rev. Mol. Cell. Biol., 16, 651-664,
https://doi.org/10.1038/nrm4069.
36. Ingolia, N. T., Brar, G. A., Stern-Ginossar, N., Harris,
M.S., Talhouarne, G.J. S., Jackson, S. E., Wills, M. R.,
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1533
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
and Weissman, J. S. (2014) Ribosome profiling re-
veals pervasive translation outside of annotat-
ed protein-coding genes, Cell Rep., 8, 1365-1379,
https://doi.org/10.1016/j.celrep.2014.07.045.
37. Bartholomäus, A., Del Campo, C., and Ignatova, Z.
(2016) Mapping the non-standardized biases of ri-
bosome profiling, Biol. Chem., 397, 23-35, https://
doi.org/10.1515/hsz-2015-0197.
38. Martinez, T. F., Chu, Q., Donaldson, C., Tan, D.,
Shokhirev, M. N., and Saghatelian, A. (2020) Ac-
curate annotation of human protein-coding small
open reading frames, Nat. Chem. Biol., 16, 458-468,
https://doi.org/10.1038/s41589-019-0425-0.
39. Ingolia, N.T., Lareau, L.F., and Weissman, J.S. (2011)
Ribosome profiling of mouse embryonic stem cells
reveals the complexity and dynamics of mammalian
proteomes, Cell, 147, 789-802, https://doi.org/10.1016/
j.cell.2011.10.002.
40. Fresno, M., Jiménez, A., and Vázquez, D. (1977) In-
hibition of translation in eukaryotic systems by
harringtonine, Eur. J. Biochem., 72, 323-330, https://
doi.org/10.1111/j.1432-1033.1977.tb11256.x.
41. Lee,S., Liu,B., Lee,S., Huang, S.-X., Shen,B., and Qian,
S.-B. (2012) Global mapping of translation initiation
sites in mammalian cells at single-nucleotide reso-
lution, Proc. Natl. Acad. Sci. USA, 109, E2424-E2432,
https://doi.org/10.1073/pnas.1207846109.
42. Gao, X., Wan, J., Liu, B., Ma, M., Shen, B., and Qian,
S.-B. (2015) Quantitative profiling of initiating ri-
bosomes in vivo, Nat. Methods, 12, 147-153, https://
doi.org/10.1038/nmeth.3208.
43. Mudge, J. M., Ruiz-Orera, J., Prensner, J. R., Brunet,
M. A., Calvet, F., Jungreis, I., Gonzalez, J. M.,
Magrane,M., Martinez, T. F., Schulz, J. F., Yang, Y. T.,
Albà, M. M., Aspden, J. L., Baranov, P. V., Bazzini,
A.A., Bruford,E., Martin, M.J., Calviello,L., Carvunis,
A.-R., Chen, J., Couso, J. P., Deutsch, E. W., Flicek, P.,
Frankish, A., Gerstein, M., Hubner, N., Ingolia, N. T.,
Kellis, M., Menschaert, G., Moritz, R. L., Ohler, U.,
Roucou, X., Saghatelian, A., Weissman, J. S., and Van
Heesch,S. (2022) Standardized annotation of translat-
ed open reading frames, Nat. Biotechnol., 40, 994-999,
https://doi.org/10.1038/s41587-022-01369-0.
44. Spealman, P., Naik, A. W., May, G. E., Kuersten, S.,
Freeberg, L., Murphy, R. F., and McManus, J. (2018)
Conserved non-AUG uORFs revealed by a novel re-
gression analysis of ribosome profiling data, Genome
Res., 28, 214-222, https://doi.org/10.1101/gr.221507.117.
45. Quaife, N. M., Chothani, S., Schulz, J. F., Lindberg,
E. L., Vanezis, K., Adami, E., O’Fee, K., Greiner, J.,
Litviňuková, M., Van Heesch, S., Whiffin, N.,
Hubner, N., Schafer,S., Rackham,O., Cook, S.A., and
Barton, P. J. R. (2023) LINC01013 is a determinant of
fibroblast activation and encodes a novel fibroblast-
activating micropeptide, J. Cardiovasc. Trans. Res.,
16, 77-85, https://doi.org/10.1007/s12265-022-10288-z.
46. Sun, L., Wang, W., Han, C., Huang, W., Sun, Y.,
Fang,K., Zeng, Z., Yang,Q., Pan,Q., Chen,T., Luo, X.,
and Chen, Y. (2021) The oncomicropeptide APPLE
promotes hematopoietic malignancy by enhancing
translation initiation, Mol. Cell, 81, 4493-4508.e9,
https://doi.org/10.1016/j.molcel.2021.08.033.
47. Van Heesch,S., Witte,F., Schneider-Lunitz,V., Schulz,
J. F., Adami, E., Faber, A. B., Kirchner, M., Maatz, H.,
Hubner, N., et al. (2019) The translational landscape
of the human heart, Cell, 178, 242-260.e29, https://
doi.org/10.1016/j.cell.2019.05.010.10.
48. Wang, S., Tian, L., Liu, H., Li, X., Zhang, J., Chen, X.,
Jia,X., Zheng,X., Wu,S., Chen,Y., Yan,J., and Wu,L.
(2020) Large-scale discovery of non-conventional pep-
tides in maize and arabidopsis through an integrated
peptidogenomic pipeline, Mol. Plant, 13, 1078-1093,
https://doi.org/10.1016/j.molp.2020.05.012.
49. Cardon, T., Hervé, F., Delcourt, V., Roucou, X.,
Salzet,M., Franck,J., and Fournier,I. (2020) Optimized
sample preparation workflow for improved identifi-
cation of ghost proteins, Anal. Chem., 92, 1122-1129,
https://doi.org/10.1021/acs.analchem.9b04188.
50. Ma,J., Ward, C.C., Jungreis,I., Slavoff, S.A., Schwaid,
A.G., Neveu, J., Budnik, B. A., Kellis,M., and Saghat-
elian, A. (2014) Discovery of human sORF-encoded
polypeptides (SEPs) in cell lines and tissue, J. Pro-
teome Res., 13, 1757-1765, https://doi.org/10.1021/
pr401280w.
51. Ma, J., Diedrich, J. K., Jungreis, I., Donaldson, C.,
Vaughan,J., Kellis,M., Yates, J.R., and Saghatelian,A.
(2016) Improved identification and analysis of
small open reading frame encoded polypeptides,
Anal. Chem., 88, 3967-3975, https://doi.org/10.1021/
acs.analchem.6b00191.
52. Cassidy,L., Helbig, A.O., Kaulich, P.T., Weidenbach,K.,
Schmitz, R.A., and Tholey,A. (2021) Multidimension-
al separation schemes enhance the identification
and molecular characterization of low molecular
weight proteomes and short open reading frame-
encoded peptides in top-down proteomics, J. Pro-
teomics, 230, 103988, https://doi.org/10.1016/j.jprot.
2020.103988.
53. Tran, J.C., and Doucette, A.A. (2008) Gel-eluted liquid
fraction entrapment electrophoresis: an electropho-
retic method for broad molecular weight range pro-
teome separation, Anal. Chem., 80, 1568-1573, https://
doi.org/10.1021/ac702197w.
54. Branca, R. M. M., Orre, L. M., Johansson, H. J.,
Granholm,V., Huss, M., Pérez-Bercoff,Å., Forshed, J.,
Käll, L., and Lehtiö, J. (2014) HiRIEF LC-MS enables
deep proteome coverage and unbiased proteog-
enomics, Nat. Methods, 11, 59-62, https://doi.org/
10.1038/nmeth.2732.
55. Nesvizhskii, A. I. (2014) Proteogenomics: concepts,
applications and computational strategies, Nat. Meth-
ods, 11, 1114-1125, https://doi.org/10.1038/nmeth.3144.
LAVROV et al.1534
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
56. Fijalkowski, I., Peeters, M. K. R., and Van Damme, P.
(2021) Small protein enrichment improves proteom-
ics detection of sORF encoded polypeptides, Front.
Genet., 12, 713400, https://doi.org/10.3389/fgene.
2021.713400.
57. Cassidy,L., Kaulich, P.T., and Tholey,A. (2023) Proteo-
forms expand the world of microproteins and short
open reading frame-encoded peptides, iScience, 26,
106069, https://doi.org/10.1016/j.isci.2023.106069.
58. Khitun, A., and Slavoff, S. A. (2019) Proteomic detec-
tion and validation of translated small open reading
frames, Curr. Protoc. Chem. Biol., 11, e77, https://
doi.org/10.1002/cpch.7.
59. Slavoff, S. A., Heo, J., Budnik, B. A., Hanakahi, L. A.,
and Saghatelian,A. (2014) A human short open read-
ing frame (sORF)-encoded polypeptide that stimulates
DNA end joining, J. Biol. Chem., 289, 10950-10957,
https://doi.org/10.1074/jbc.C113.533968.
60. Ouspenskaia, T., Law, T., Clauser, K. R., Klaeger, S.,
Sarkizova,S., Aguet,F., Li,B., Christian,E., Knisbacher,
B.A., Le, P.M., Hartigan, C.R., Keshishian,H., Apffel,A.,
Oliveira, G., Zhang, W., Chen, S., Chow, Y. T., Ji, Z.,
Jungreis, I., Shukla, S. A., Justesen, S., Bachireddy, P.,
Kellis, M., Getz, G., Hacohen, N., Keskin, D. B., Carr,
S. A., Wu, C. J., and Regev, A. (2022) Unannotated
proteins expand the MHC-I-restricted immunopepti-
dome in cancer, Nat. Biotechnol, 40, 209-217, https://
doi.org/10.1038/s41587-021-01021-3.
61. Wang, J.Y., and Doudna, J. A. (2023) CRISPR technol-
ogy: a decade of genome editing is only the begin-
ning, Science, 379, eadd8643, https://doi.org/10.1126/
science.add8643.
62. Shalem, O., Sanjana, N. E., and Zhang, F. (2015)
High-throughput functional genomics using CRISPR-
Cas9, Nat. Rev. Genet., 16, 299-311, https://doi.org/
10.1038/nrg3899.
63. Chen,J., Brunner, A.-D., Cogan, J.Z., Nuñez, J.K., Fields,
A. P., Adamson, B., Itzhak, D. N., Li, J. Y., Mann, M.,
Leonetti, M. D., and Weissman, J. S. (2020) Perva-
sive functional translation of noncanonical human
open reading frames, Science, 367, 1140-1146, https://
doi.org/10.1126/science.aay0262.
64. Prensner, J. R., Enache, O. M., Luria, V., Krug, K.,
Clauser, K. R., Dempster, J. M., Karger, A., Wang, L.,
Stumbraite, K., Wang, V. M., Botta, G., Lyons, N. J.,
Goodale, A., Kalani, Z., Fritchman, B., Brown, A.,
Alan, D., Green, T., Yang, X., Jaffe, J. D., Roth, J. A.,
Piccioni, F., Kirschner, M. W., Ji, Z., Root, D. E., and
Golub, T. R. (2021) Noncanonical open reading
frames encode functional proteins essential for can-
cer cell survival, Nat. Biotechnol., 39, 697-704, https://
doi.org/10.1038/s41587-020-00806-2.
65. Schlesinger, D., Dirks, C., Navarro, C., Lafranchi, L.,
Spinner, A., Raja, G. L., Mun-Sum Tong, G., Eirich, J.,
Martinez, T. F., and Elsässer, S. J. (2025) A large-
scale sORF screen identifies putative microproteins
involved in cancer cell fitness, iScience, 28, 111884,
https://doi.org/10.1016/j.isci.2025.111884.
66. Lafranchi,L., Spinner,A., Hornisch,M., Schlesinger,D.,
Luzon, C. N., Brinkenstråhle, L., Shao, R., Piazza, I.,
and Elsässer, S. J. (2024) Pooled overexpression
screening identifies PIPPI as a novel micropro-
tein involved in the ER stress response, bioRxiv,
https://doi.org/10.1101/2024.12.08.627409.
67. Delcourt, V., Brunelle, M., Roy, A. V., Jacques, J.-F.,
Salzet,M., Fournier,I., and Roucou,X. (2018) The pro-
tein coded by a short open reading frame, not by the
annotated coding sequence, is the main gene product
of the dual-coding gene MIEF1, Mol. Cell. Proteom., 17,
2402-2411, https://doi.org/10.1074/mcp.RA118.000593.
68. Zhang, C., Zhou, B., Gu, F., Liu, H., Wu, H., Yao, F.,
Zheng,H., Fu,H., Chong,W., Cai,S., Huang,M., Ma,X.,
Guo, Z., Li, T., Deng, W., Zheng, M., Ji, Q., Zhao, Y.,
Ma, Y., Wang, Q.-E., Tang, T.-S., and Guo, C. (2022)
Micropeptide PACMP inhibition elicits synthetic le-
thal effects by decreasing CtIP and poly(ADP-ribo-
syl)ation, Mol. Cell, 82, 1297-1312.e8, https://doi.org
/10.1016/j.molcel.2022.01.020.
69. Zhang,H., Liao,Z., Wang,W., Liu,Y., Zhu,H., Liang,H.,
Zhang,B., and Chen, X. (2023) A micropeptide JunBP
regulated by TGF-β promotes hepatocellular car-
cinoma metastasis, Oncogene, 42, 113-123, https://
doi.org/10.1038/s41388-022-02518-0.
70. Guo, B., Wu, S., Zhu, X., Zhang, L., Deng, J., Li, F.,
Wang,Y., Zhang,S., Wu,R., Lu,J., and Zhou,Y. (2020)
Micropeptide CIP 2A-BP encoded by LINC 00665 in-
hibits triple-negative breast cancer progression,
EMBO J., 39, e102190, https://doi.org/10.15252/embj.
2019102190.
71. Huang, J.-Z., Chen, M., Chen, D., Gao, X.-C., Zhu, S.,
Huang, H., Hu, M., Zhu, H., and Yan, G.-R. (2017) A
peptide encoded by a putative lncRNA HOXB-AS3 sup-
presses colon cancer growth, Mol. Cell, 68, 171-184.e6,
https://doi.org/10.1016/j.molcel.2017.09.015.
72. Rocha, A. L., Pai, V., Perkins, G., Chang, T., Ma, J., De
Souza, E. V., Chu, Q., Vaughan, J. M., Diedrich, J. K.,
Ellisman, M. H., and Saghatelian, A. (2024) An inner
mitochondrial membrane microprotein from the
SLC35A4 upstream ORF regulates cellular metabo-
lism, J. Mol. Biol., 436, 168559, https://doi.org/10.1016/
j.jmb.2024.168559.
73. Konina, D., Sparber, P., Viakhireva, I., Filatova, A.,
and Skoblov, M. (2021) Investigation of LINC00493/
SMIM26 gene suggests its dual functioning at mRNA
and protein level, Int. J. Mol. Sci. Artic. J. Mol. Sci., 22,
8477, https://doi.org/10.3390/ijms22168477.
74. Liang, X., Shen, W., Sun, H., Migawa, M. T., Vickers,
T. A., and Crooke, S. T. (2016) Translation efficien-
cy of mRNAs is increased by antisense oligonucle-
otides targeting upstream open reading frames,
Nat. Biotechnol., 34, 875-880, https://doi.org/10.1038/
nbt.3589.
METHODS FOR IDENTIFYING AND STUDYING MICROPEPTIDES 1535
BIOCHEMISTRY (Moscow) Vol. 90 No. 11 2025
75. Anderson, D. M., Anderson, K. M., Chang, C.-L.,
Makarewich, C. A., Nelson, B. R., McAnally, J. R.,
Kasaragod,P., Shelton, J. M., Liou, J., Bassel-Duby, R.,
and Olson, E.N. (2015) A micropeptide encoded by a
putative long noncoding RNA regulates muscle per-
formance, Cell, 160, 595-606, https://doi.org/10.1016/
j.cell.2015.01.009.
76. Ryskina, A. M., Kudriaeva, A. A., and Belogurov,
A. A. (2024) Microproteins tracking: when size does
really matter, Rev. Adv. Chem., 14, 305-319, https://
doi.org/10.1134/S2634827624600324.
77. Chu, Q., Martinez, T. F., Novak, S. W., Donaldson,
C. J., Tan, D., Vaughan, J. M., Chang, T., Diedrich,
J. K., Andrade, L., Kim, A., Zhang, T., Manor, U., and
Saghatelian,A. (2019) Regulation of the ER stress re-
sponse by a mitochondrial microprotein, Nat. Com-
mun., 10, 12816, https://doi.org/10.1038/s41467-019-
12816-z.
78. Niu,L., Lou,F., Sun,Y., Sun,L., Cai,X., Liu,Z., Zhou,H.,
Wang,H., Wang,Z., Bai,J., Yin,Q., Zhang,J., Chen,L.,
Peng,D., Xu,Z., Gao,Y., Tang,S., Fan,L., and Wang,H.
(2020) A micropeptide encoded by lncRNA MIR155HG
suppresses autoimmune inflammation via modu-
lating antigen presentation, Sci. Adv., 6, eaaz2059,
https://doi.org/10.1126/sciadv.aaz2059.
79. Rathore,A., Chu,Q., Tan,D., Martinez, T.F., Donaldson,
C. J., Diedrich, J. K., Yates, J. R., and Saghatelian, A.
(2018) MIEF1 microprotein regulates mitochondri-
al translation, Biochemistry, 57, 5564-5575, https://
doi.org/10.1021/acs.biochem.8b00726.
80. Schägger,H. (2006) Tricine-SDS-PAGE, Nat. Protoc., 1,
16-22, https://doi.org/10.1038/nprot.2006.4.
81. Stein, C. S., Jadiya, P., Zhang, X., McLendon, J. M.,
Abouassaly, G. M., Witmer, N. H., Anderson, E. J.,
Elrod, J. W., and Boudreau, R. L. (2018) Mitoregulin:
a lncRNA-encoded microprotein that supports mito-
chondrial supercomplexes and respiratory efficien-
cy, Cell Rep., 23, 3710-3720.e8, https://doi.org/10.1016/
j.celrep.2018.06.002.
82. Okita, N., Higami, Y., Fukai, F., Kobayashi, M., Mi-
tarai, M., Sekiya, T., and Sasaki, T. (2017) Modified
Western blotting for insulin and other diabetes-as-
sociated peptide hormones, Sci. Rep., 7, 6949, https://
doi.org/10.1038/s41598-017-04456-4.
83. Huang,N., Li,F., Zhang,M., Zhou,H., Chen,Z., Ma,X.,
Yang,L., Wu,X., Zhong,J., Xiao,F., Yang,X., Zhao,K.,
Li,X., Xia,X., Liu,Z., Gao,S., and Zhang,N. (2021) An
Upstream Open Reading Frame in Phosphatase and
Tensin Homolog Encodes a Circuit Breaker of Lac-
tate Metabolism, Cell Metab., 33, 128-144.e9, https://
doi.org/10.1016/j.cmet.2020.12.008.
84. Cabantous,S., Terwilliger, T.C., and Waldo, G.S. (2005)
Protein tagging and detection with engineered self-
assembling fragments of green fluorescent protein,
Nat. Biotechnol., 23, 102-107, https://doi.org/10.1038/
nbt1044.
85. Kamiyama, D., Sekine, S., Barsi-Rhyne, B., Hu, J.,
Chen, B., Gilbert, L. A., Ishikawa, H., Leonetti, M. D.,
Marshall, W.F., Weissman, J.S., and Huang,B. (2016)
Versatile protein tagging in cells with split fluo-
rescent protein, Mol. Cell, 82, 2900-2911.e7, https://
doi.org/10.1038/ncomms11046.
86. Na, Z., Dai, X., Zheng, S.-J., Bryant, C. J., Loh, K. H.,
Su, H., Luo, Y., Buhagiar, A. F., Cao, X., Baserga, S. J.,
Chen, S., and Slavoff, S. A. (2022) Mapping subcellu-
lar localizations of unannotated microproteins and
alternative proteins with MicroID, Mol. Cell, 82, 2900-
2911.e7, https://doi.org/10.1016/j.molcel.2022.06.035.
87. Nguyen, T. M. T., Kim, J., Doan, T. T., Lee, M.-W., and
Lee,M. (2020) APEX proximity labeling as a versatile
tool for biological research, Biochemistry, 59, 260-269,
https://doi.org/10.1021/acs.biochem.9b00791.
88. Roux, K.J., Kim, D.I., Burke,B., and May, D.G. (2018)
BioID: A Screen for Protein-Protein Interactions,
Curr. Protoc. Protein Sci., 91, 19.23.1-19.23.15, https://
doi.org/10.1002/cpps.51.
89. Bosch, J. A., Chen, C., and Perrimon, N. (2021) Prox-
imity-dependent labeling methods for proteomic
profiling in living cells: an update, Wiley Interdis-
cip. Rev. Dev. Biol., 10, e392, https://doi.org/10.1002/
wdev.392.
90. Chu, Q., Rathore, A., Diedrich, J. K., Donaldson, C. J.,
Yates, J.R., and Saghatelian,A. (2017) Identification of
microprotein-protein interactions via APEX tagging,
Biochemistry, 56, 3299-3306, https://doi.org/10.1021/
acs.biochem.7b00265.
91. Singh, D. R., Dalton, M. P., Cho, E. E., Pribadi, M. P.,
Zak, T. J., Šeflová, J., Makarewich, C. A., Olson, E. N.,
and Robia, S. L. (2019) Newly discovered micropep-
tide regulators of SERCA form oligomers but bind to
the pump as monomers, J. Mol. Biol., 431, 4429-4443,
https://doi.org/10.1016/j.jmb.2019.07.037.
92. Hu,Y., Cheng,K., He,L., Zhang,X., Jiang,B., Jiang,L.,
Li, C., Wang, G., Yang, Y., and Liu, M. (2021) NMR-
based methods for protein analysis, Anal. Chem.,
93, 1866-1879, https://doi.org/10.1021/acs.analchem.
0c03830.
Publishers Note. Pleiades Publishing remains
neutral with regard to jurisdictional claims in published
maps and institutional affiliations. AI tools may have
been used in the translation or editing of this article.