ArticlePDF Available

Proteasome: a Nanomachinery of Creative Destruction

Authors:

Abstract and Figures

In the middle of the 20th century, it was postulated that degradation of intracellular proteins is a stochastic process. More than fifty years of intense studies have finally proven that protein degradation is a very complex and tightly regulated in time and space process that plays an incredibly important role in the vast majority of metabolic pathways. Degradation of more than a half of intracellular proteins is controlled by a hierarchically aligned and evolutionarily perfect system consisting of many components, the main ones being ubiquitin ligases and proteasomes, together referred to as the ubiquitin–proteasome system (UPS). The UPS includes more than 1000 individual components, and most of them are critical for the cell functioning and survival. In addition to the well-known signaling functions of ubiquitination, such as modification of substrates for proteasomal degradation and DNA repair, polyubiquitin (polyUb) chains are involved in other important cellular processes, e.g., cell cycle regulation, immunity, protein degradation in mitochondria, and even mRNA stability. This incredible variety of ubiquitination functions is related to the ubiquitin ability to form branching chains through the ε-amino group of any of seven lysine residues in its sequence. Deubiquitination is accomplished by proteins of the deubiquitinating enzyme family. The second main component of the UPS is proteasome, a multisubunit proteinase complex that, in addition to the degradation of functionally exhausted and damaged proteins, regulates many important cellular processes through controlled degradation of substrates, for example, transcription factors and cyclins. In addition to the ubiquitin-dependent-mediated degradation, there is also ubiquitin-independent degradation, when the proteolytic signal is either an intrinsic protein sequence or shuttle molecule. Protein hydrolysis is a critically important cellular function; therefore, any abnormalities in this process lead to systemic impairments further transforming into serious diseases, such as diabetes, malignant transformation, and neurodegenerative disorders (multiple sclerosis, Alzheimer’s disease, Parkinson’s disease, Creutzfeldt–Jakob disease and Huntington’s disease). In this review, we discuss the mechanisms that orchestrate all components of the UPS, as well as the plurality of the fine-tuning pathways of proteasomal degradation.
Content may be subject to copyright.
Intracellular concentration of polypeptides is deter
mined by the rates of their synthesis and degradation.
Cellular protein homeostasis, or proteostasis, is main
tained by a complex adaptive system that controls the
entire protein life cycle: synthesis, folding, maintenance
of correct conformation and necessary amount, subcellu
lar localization, and degradation. For example, during
translation, aberrant RNA molecules are recognized and
their polypeptide products undergo degradation [13].
Most proteins fold into the predetermined threedimen
sional structures encoded in their amino acid sequences,
which is necessary for proper functioning of these pro
teins. Some proteins retain conformational mobility,
while others are partially or even completely unordered,
which increases the risk of their misfolding or even aggre
gation [4]. In addition, misfolded proteins may appear
because of genetic mutations or ribosomal errors, or when
folded proteins lose their structure due to denaturation,
aging, or chemical modifications. Therefore, the main
problem of proteostasis is not only regulation of intracel
lular protein concentrations but also protection of cells
from the harmful effects of abnormal and potentially
toxic misfolded or damaged proteins that can seriously
disrupt cell metabolism.
ISSN 00062979, Biochemistry (Moscow), 2019, Vol. 84, Suppl. 1, pp. S159S192. © Pleiades Publishing, Ltd., 2019.
Russian Text © A. A. Kudriaeva, A. A. Belogurov, 2019, published in Uspekhi Biologicheskoi Khimii, 2019, Vol. 59, pp. 323392.
REVIEW
S159
Abbreviations: ALS, autophagic–lysosomal system; IFN, inter
feron; MBP, myelin basic protein; MHC, major histocompati
bility complex; ODC, ornithine decarboxylase; Ub, ubiquitin;
UPS, ubiquitin–proteasome system.
* To whom correspondence should be addressed.
Proteasome: a Nanomachinery of Creative Destruction
A. A. Kudriaeva1,a* and A. A. Belogurov1,2,b*
1Shemyakin
Ovchinnikov Institute of Bioorganic Chemistry, Russian Academy of Sciences, 117997 Moscow, Russia
2Lomonosov Moscow State University, 119991 Moscow, Russia
aemail: anna.kudriaeva@gmail.com
bemail: belogurov@mx.ibch.ru
Received August 10, 2018
Revised September 7, 2018
Accepted September 7, 2018
Abstract—In the middle of the 20th century, it was postulated that degradation of intracellular proteins is a stochastic
process. More than fifty years of intense studies have finally proven that protein degradation is a very complex and tightly
regulated in time and space process that plays an incredibly important role in the vast majority of metabolic pathways.
Degradation of more than a half of intracellular proteins is controlled by a hierarchically aligned and evolutionarily perfect
system consisting of many components, the main ones being ubiquitin ligases and proteasomes, together referred to as the
ubiquitin–proteasome system (UPS). The UPS includes more than 1000 individual components, and most of them are crit
ical for the cell functioning and survival. In addition to the wellknown signaling functions of ubiquitination, such as mod
ification of substrates for proteasomal degradation and DNA repair, polyubiquitin (polyUb) chains are involved in other
important cellular processes, e.g., cell cycle regulation, immunity, protein degradation in mitochondria, and even mRNA
stability. This incredible variety of ubiquitination functions is related to the ubiquitin ability to form branching chains
through the εamino group of any of seven lysine residues in its sequence. Deubiquitination is accomplished by proteins of
the deubiquitinating enzyme family. The second main component of the UPS is proteasome, a multisubunit proteinase
complex that, in addition to the degradation of functionally exhausted and damaged proteins, regulates many important cel
lular processes through controlled degradation of substrates, for example, transcription factors and cyclins. In addition to
the ubiquitindependentmediated degradation, there is also ubiquitinindependent degradation, when the proteolytic sig
nal is either an intrinsic protein sequence or shuttle molecule. Protein hydrolysis is a critically important cellular function;
therefore, any abnormalities in this process lead to systemic impairments further transforming into serious diseases, such as
diabetes, malignant transformation, and neurodegenerative disorders (multiple sclerosis, Alzheimer’s disease, Parkinson’s
disease, Creutzfeldt–Jakob disease and Huntington’s disease). In this review, we discuss the mechanisms that orchestrate all
components of the UPS, as well as the plurality of the finetuning pathways of proteasomal degradation.
DOI: 10.1134/S0006297919140104
Keywords: proteasome, protein degradation, ubiquitindependent proteolysis, ubiquitinindependent proteolysis, ubiquitin
S160 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
Proteins that have lost their functional conformation
undergo controlled degradation (including cotransla
tional degradation); aggregated protein molecules are
destroyed via different cellular mechanisms [57]. In
many cases, these processes involve different classes of
molecular chaperones and their regulators that promote
proper protein folding and maintain proteins in a soluble
state. Besides contributing to the protein folding, chaper
ones participate in the degradation of misfolded proteins
[8]. To summarize the aforementioned, proteostasis is
essential for cell metabolism, organelle biogenesis, and
cell adaptation to stress that greatly affects cell viability
[9].
Currently, the autophagic–lysosomal system (ALS)
and the ubiquitin–proteasome system (UPS) are consid
ered as the main proteolytic pathways in the cell,
although extralysosomal proteases, primarily caspases
and calpains, are also involved in protein hydrolysis. The
studies on the crucial role of protein degradation systems
were recognized by three Nobel prizes: in 1974, bio
chemist Christian de Duve was awarded the Nobel Prize
in Medicine and Physiology for his contribution to the
discovery of the structural and functional organization of
the cell (discovery of lysosomes); in 2004, the Nobel Prize
in Chemistry was awarded jointly to Aaron Ciechanover,
Avram Hershko, and Irwin Rose for the discovery of ubiq
uitinmediated protein degradation; and in 2016,
Yoshinori Ohsumi received the Nobel Prize in Medicine
and Physiology for discovery of the autophagy mecha
nisms.
UPS is the main component in the degradation of
shortlived, misfolded, damaged, and obsolete proteins.
UPS actively participates in cell signal transduction, tran
scription, cell cycle progression, proliferation, and apop
tosis [10]. ALS recognizes and removes large and poten
tially dangerous cellular components, such as protein
aggregates and nonfunctional or superfluous organelles.
Its functioning is an important adaptive mechanism for
fighting various types of cell stress, such as nutrient
deficit, hypoxia, oxidative stress, etc. [11]. UPS and ALS
are responsible for approximately 8090 and 1020% of
cellular proteolysis, respectively, although these ratios
may vary depending on physiological conditions and cell
type [12, 13]. In the cytosol, UPS and ALS act simulta
neously; they share some components and constantly
interact with each other [14]. Ubiquitination is used as a
degradation signal for tagging substrates in both systems.
Protein degrons (protein sequences involved in the regu
lation of protein degradation) modified with ubiquitin
(Ub) are recognized and bound by proteins containing
Ubbinding domains (UBDs) [15]. UBDs decode the
polyUb code (linkage type and chain length) and deliver
ubiquitinated substrates to the proteasome (RAD23 and
UBQLN), autophagosomal vacuoles [p62 and BRCA1
(NBR1)], or targets not involved in proteolytic processes,
e.g., in the cases of DNA damage and inflammation [16].
In this review, we will discuss the mechanisms of
UPS functioning and regulation. Because protein hydro
lysis is a continuously occurring process, it has an impor
tant physiological significance. Its impairment leads to
multiple disturbances in the normal cell functioning and,
ultimately, to serious disorders, e.g., diabetes, malignant
transformation, and neurodegenerative diseases, such as
multiple sclerosis, Alzheimer’s disease, Parkinson’s dis
ease, Creutzfeldt–Jakob disease, and Huntington’s dis
ease [17]. By now, it has become obvious that degradation
of intracellular proteins is a specific process. The stability
of an intracellular protein varies widely, depending on
both the properties of the protein itself and its environ
ment. In the middle of the 20th century, protein degrada
tion was postulated as probabilistic and uncontrollable,
but now it is commonly accepted that protein breakdown
is a very complex, tightly controlled in time and space
process that plays an essential role in most metabolic
pathways.
UBIQUITIN–PROTEASOME SYSTEM
Proteasome
Proteasome is a multisubunit proteinase complex
responsible for the degradation of most intracellular pro
teins. In addition to the breakdown of obsolete and dam
aged proteins, proteasomes regulate many important cel
lular processes via controlled degradation of transcription
factors, cell cycle regulators, enzymes, etc. Eukaryotic
proteasome contains three catalytic subunits: β1, β2, and
β5, each having its own substrate specificity. Besides, var
ious cell types contain proteasome variants with different
activity and specificity profiles; the functions of these
proteasomes have not been fully understood yet. Many
inhibitors of proteasomal proteolytic activity are current
ly used as drugs in the treatment of hematologic malig
nancies; potential application of proteasome inhibitors in
the treatment of other diseases, including autoimmune
disorders, is being actively studied.
The proteasomal degradation signal represents a
chain of Ub molecules covalently bound to the target pro
tein. The signal specificity is determined by the length
and structure of the polyUb chains that are recognized by
receptors of the proteasome regulatory particles. It has
become evident that the Ubmediated modification sys
tem is far from simple, because at least seven functional
lysine residues in the Ub sequence can be involved in the
polyUb chain growth. After polypeptide substrate binds
to the proteasome via the Ub tag, its hydrolysis starts at
the unstructured initiation site, one of the elements of the
protein degradation signal.
In addition to the Ubdependent degradation, there
is also Ubindependent degradation. In this case, the
degradation signal is either a specific motif in the protein
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S161
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
sequence or some auxiliary molecule. Currently, about
two dozen proteins are known to be hydrolyzed without
preliminary modification with Ub. For example, some
protooncoproteins and oncosuppressive proteins are
Ubindependent proteasomal substrates, whose degrada
tion might have a tumorigenic effect. Identification of
Ubindependent degradation mechanisms suggests the
existence of multiple protein degradation pathways.
Proteasome structure and function. Most proteins in
the cytoplasm and nucleus of eukaryotic cells are
hydrolyzed by proteasome. Proteasome is a 2.5MDa
multisubunit protein complex that controls the content of
more than a hundred regulatory proteins and destroys
misfolded and damaged proteins [18]. Proteasomes are
present in all eukaryotic cells, as well as in some bacterial
species, which confirms their high significance for nor
mal cell activity [19]. Just as protein synthesis is regulat
ed at various levels, an equally complex system exists to
control protein hydrolysis by the proteasome. That is why
proteasomes are incredibly selective. In principle, the
proteasome is most likely able to hydrolyze any cellular
protein modified by Ub; at the same time, the risk of acci
dental degradation of other components of cellular pro
teome is excluded.
The proteasome complex consists of two functional
and structural parts. The name proteasome refers to differ
ent types of particles with different molecular weights.
The 20S proteasome, which is the catalytic core of the
proteasome, is capable of binding many regulatory parti
cles, such as the 19S regulator, PA28αβ, PA28γ, PA200,
Emc29, PI31, etc., thereby forming various proteasome
types known as the 26S proteasome (19S–20S), 30S pro
teasome (19S–20S–19S), hybrid proteasome (19S–20S–
PA28), PA28 proteasome (PA28–20S–PA28), etc. [20].
Free 20S proteasome does not hydrolyze structured pro
teins, but cleaves small peptides and unfolded protein
molecules [21]. Evaluation of the relative content of all
proteasome complexes in HeLa cells using immunopre
cipitation and immunoblotting revealed that about 40%
of all 20S proteasomes were in the free form, while the
remaining fraction was associated with PA28 and 19S
regulators in various combinations [22]. Higher levels of
free 20S proteasome were found in U947 cells treated
with γinterferon (IFNγ) [23]. Electronic cryotomogra
phy provided information on the number of 26S and 30S
proteasomes in neurons lacking the PA28 regulator; it
was found that a quarter of all proteasomes were 30S pro
teasomes, while the rest were 26S proteasomes [24].
Given specific functions of many cell types in mammals
and the fact that some cell types lack certain proteasomal
regulatory particles, it is highly likely that each cell type
contains its own individual pool of proteasomal complex
es [25]. Thus, the 20S proteasomal complex associated
with the neuronal plasma membrane through GPM6 gly
coproteins was recently found [26]. This complex can
degrade intracellular proteins into biologically active
extracellular peptides that induce calcium signaling via
NMDA receptors.
20S proteasome. According to the Xray diffraction
data, the 20S proteasome consists of four heptameric
rings forming a hollow cylinder approximately 1517 nm
in length and 11 to 12 nm in diameter [27]. Each ring
contains seven subunits with molecular masses from 20 to
35 kDa; the molecular weight of the total complex is
~750 kDa. The outer rings are composed of αsubunits,
and the inner ones are composed of βsubunits. Four
stacked rings form three internal chambers with a diame
ter of ~5 nm [28]. The volume of the central proteolytic
chamber is approximately 84 nm3; it can accommodate
proteins up to 70 kDa, but the entrance to the chamber is
restricted by a small width of the entrance opening [29].
The presence of two copies of 14 different subunits is a
common feature of all eukaryotic 20S proteasomes.
Proteasomes with a simpler composition but with the
same basic architecture (one type of αsubunits and one
type of βsubunits) were found in archaea (Thermoplasma
acidophilum) and some bacteria (Rhodococcus erythropolis
and Mycobacterium tuberculosis) [30]. However, the com
position of such proteasomes may vary: one αsubunit
type and two βsubunit types, two αsubunit types and
two βsubunit types, or two αsubunit types and one β
subunit type. Among the species that synthesize two β
subunit types, one of the two βsubunits is believed to be
inactive in Sulfolobus, Pyrobaculum, and Aeropyrum.
The central chamber of the eukaryotic 20S protea
some contains six catalytic sites formed by two β1, two
β2, and two β5 subunits. Therefore, the proteasome has
three main types of proteolytic activity: trypsinlike
(cleavage after positively charged amino acid residues),
chymotrypsinlike activity (cleavage after aromatic
amino acid residues), and caspaselike (cleavage after
negatively charged residues). Since the proteasome also
possesses some additional activities, it can cleave the
polypeptide chains between almost any amino acid
residues. The catalytic mechanism of the peptide bond
hydrolysis involves an attack of the carbonyl carbon atom
by the hydroxyl of the Nterminal threonine of the pro
teasome catalytic subunit followed by hydroxyl reactiva
tion with a water molecule (Fig. 1).
A distinctive feature of large intracellular proteases,
such as the proteasome or tripeptidyl peptidase II
(TPPII), is the active site location in the internal cavity
formed by the complex subunits (compartmentalization).
Isolation of the active sites from the cellular environment
prevents undesirable degradation of cellular proteins [31].
The proteasome assembly from subunits is a complex
multistep process. In cells, βsubunits are synthesized as
inactive precursors with longer Nterminal sequences
compared to the mature βsubunits. The Nterminal pep
tides are signals for protein factors involved in the protea
some assembly and also prevent premature activation of
βsubunits. During the assembly, relatively stable inter
S162 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
mediates, preproteasomes, are formed first that consist of
one αring and a partially assembled ring composed of
the βprecursors and auxiliary factors. Then, the two
intermediates join to form the 20S proteasome. The N
terminal propeptides are autocatalytically cleaved off,
releasing the threonine residues of the catalytic sites and
simultaneously sequestering them from the cellular envi
ronment [32]. The entrance to the 20S proteasome prote
olytic chamber is blocked by the Nterminal hydrophobic
regions of αsubunits, which also prevents spontaneous
hydrolysis of proteins. It was shown by atomic force
microscopy that the gates to the catalytic chamber are in
a dynamic equilibrium between open and closed states,
with the equilibrium being shifted towards the closed state
[33]. Therefore, the spatial configuration of the 20S pro
teasome and a small size of the openings leading into the
catalytic chamber prevent degradation of folded proteins.
αSubunits interact with the 19S regulatory complex
and PA28 and PA200 activators that induce conforma
tional changes in αsubunits leading to opening of the
entrance to the catalytic chamber [34]. It was found that
the 20S proteasome can be activated by certain low
molecularweight compounds [35]. In addition, deletion
of the Nterminal portion of the α3subunit (α3ΔN)
leads to the activation of the proteasome in mammalian
cells. Because the proteasomal degradation is the main
mechanism that regulates the level of toxic aggregated
proteins [36], an increase in the proteasomal activity due
to the opening of the proteolytic chamber may prevent
the toxicity of such proteins and suppress associated
development of proteotoxic disorders (e.g., Alzheimer’s
disease) [37]. For example, cells expressing α3ΔN protea
somes have a lower level of tau proteins and their aggre
gates [38].
proteasome
proteasome
proteasome
proteasome
1. Nucleophilic attack
2. Acylenzyme
3. Hydrolysis
of acylenzyme
4. Reaction product
removal
Fig. 1. Catalytic mechanism of the peptide bond cleavage by the proteasome. The mechanism of catalysis was revealed by Xray diffraction,
and the catalytically important residue was identified by mutagenesis. During the reaction, a proton of the threonine hydroxyl group is trans
ferred to the amino group of the same amino acid, and the water molecule in the active site acts as a base. Similar mechanism underlies activ
ity of three other enzymes: penicillin acylase (Nterminal serine in the active site), glutamine amidotransferase (Nterminal cysteine in the
active site), and aspartyl glucosaminidase (Nterminal threonine in the active site).
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S163
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
In addition to the proteolytic activity, the 20S pro
teasome (especially, the α5subunit) exhibits the RNase
activity in the presence of divalent cations (not required
for the proteolytic activity) [39]. The RNase activity of α5
subunit strongly correlates with the extent of phosphory
lation of α6 and α7 subunits and markedly increases dur
ing erythroid differentiation and programmed cell death
[39]. Most αsubunits (except α3) are also known to
exhibit RNase activity against p53 mRNA in vitro [40].
In mammals, five additional subunits, β1i, β2i, β5i,
β5t, and α4s, have been found [41]. The genes encoding
β1i and β5i are located in the MHC (major histocompati
bility complex) class II gene region on chromosome 6, and
their transcription is induced by the proinflammatory
cytokines, e.g., INFγ[42]. Subunits β1i, β2i, and β5i are
stably expressed in spleen and hematopoietic professional
antigenpresenting cells due to continuous activation of
the INFγinducible promoter of the β1i and β5i genes via
binding of the unphosphorylated transcription factor Stat1
dimer to the IRF1 protein [43]. These subunits replace the
constitutive subunits of the 20S proteasome and change its
proteolytic specificity. The 20S proteasome containing
β1i, β2i, and β5i subunits is called immunoproteasome; it
lacks the caspaselike activity, but displays elevated
trypsinlike and chymotrypsinlike activities compared to
the constitutive proteasome. A set of antigenic peptides
produced by the immunoproteasome differs from a set of
peptides produced by the constitutive proteasome [44, 45].
Due to the difference in the substrate specificity between
the constitutive proteasome and the immunoproteasome,
the antigenic peptides produced by the immunoprotea
some have hydrophobic Ctermini and therefore, are more
structurally suitable for loading onto the MHC class I mol
ecules. Cells containing the immunoproteasome present
antigens on their surfaces more efficiently [42].
There are cases of the socalled proteasomemediat
ed peptide splicing, when the two peptides (proteasomal
cleavage products) that are normally not adjacent in the
original protein, are ligated by the proteasome. This addi
tional proteasome activity contributes to the expansion of
the pool of epitopes presented on MHC class I. Peptide
splicing can occur by either transpeptidation or conden
sation [46]. The mechanism of peptide splicing by the
proteasome is not completely understood yet. It is only
known that splicing does not occur spontaneously, but
requires certain conditions to be met, e.g., one of the
peptides should contain lysine or arginine at the Ctermi
nus [47]. Splicing of peptides from the same protein is
more likely than splicing of peptides from different pro
teins [48]. There is evidence that peptide splicing is not a
rare event, as previously assumed, and spliced peptides
account for one third of the peptidome diversity generat
ed by the proteasome for the presentation on MHC class
I molecules [49].
In many nonimmune cells, expression of protea
some immune subunits depends on interferons, TNFα,
and liposaccharides [42]. However, their biosynthesis can
also be induced by less specific physiological triggers –
aging and environmental stress factors, such as heat
shock. Therefore, in addition to the generation of epi
topes for MHC class I, immunoproteasomes are involved
in other cell processes, rapid elimination of damaged pro
teins after oxidative stress, cell proliferation control, and
cytokine production [50]. Hence, immunoproteasomes
can either aggravate and alleviate the course of various
diseases, such as viral infections, colitis, myocarditis, and
diabetes.
Although IFNγinducible αsubunits are unknown,
there is an alternative α4 subunit that was designated as
α4s, because it was localized exclusively to spermatocytes
differentiated from male germ cells [51]. Also, the β5t
subunit expressed exclusively in the cortical thymic
epithelial cells was found. This subunit replaces the β5i
subunit, leading to the β1iβ2iβ5t configuration called
the thymoproteasome. The mechanism of the β5t expres
sion regulation is still unknown [52]. The chymotrypsin
like activity of thymoproteasomes is lower than that of
constitutive proteasomes and immunoproteasomes,
which is related to the hydrophilic nature of the substrate
binding pocket in β5t [53]; in addition, these proteasome
types differ in their susceptibility to proteasome inhibitors
[54]. It is believed that the thymoproteasome increases
the repertoire of peptides for the positive selection of T
cells during their development in the thymus [55] (Fig. 2).
26S proteasome. The 26S proteasome is the largest
and most complex member of the ancient superfamily of
ATPdependent proteases [56]. These proteases are char
acterized by the presence of the AAA ATPase (ATPase
associated with diverse cellular activities) ring responsible
for the substrate unfolding and translocation through a
narrow channel into the proteasome proteolytic chamber.
The ATPase ring converts chemical energy of ATP
hydrolysis into mechanical force for substrate unfolding.
Most often, the term 26S or 30S proteasome refers to the
20S proteasome associated with one or two, respectively,
PA700 (protein activator) or 19S regulatory complexes
containing many additional specialized subunits in addi
tion to the ATPase ring.
The regulatory particles of the proteasome are
responsible for the binding, deubiquitination, unfolding,
and transfer of the substrates to the proteasome catalytic
chamber and for opening of the channel in the αsubunit
ring. The 19S particle contains at least 19 subunits with a
total molecular mass of about 1 MDa. This regulatory
complex may be divided into two subcomplexes called the
lid and the base. The base consists of nine subunits (regu
latory particles); of these, six are homologous AAA
ATPases (Rpt16) and three are nonATPases (Rpn1,
Rpn2, and Rpn13). Rpt16 form a heterohexameric ring
that directly contacts the αsubunit ring of the 20S pro
teasome [57]. Rpn1 and Rpn2 are the two largest struc
tural proteasomal subunits. The central parts of these sub
S164 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
units consist of 11 αhelical repeats that presumably form
the frame for the accommodation of the lid subunits and
the substrates. Rpn13 and Rpn10 can bind Ub and act as
receptors of ubiquitinated substrates [58, 59]. Recent
studies have also assigned Rpn1 to the Ubbinding sub
units [60]. In addition to the resident proteasome sub
units, the 19S particle is associated with a large number of
proteins indirectly involved in the degradation process.
Some of these proteins (Rad23, Dsk2, and Ddi1 in yeast)
contain Ublike (UBL) and Ubassociated (UBA)
domains and function as alternative Ub receptors. The lid
consists of nine different Rpn subunits: Rpn3, Rpn59,
11, 12, and Sem1 (Rpn15). Rpn11 is a deubiquitinating
enzyme in the proteasome [61]. The lid is structurally
similar to the COP9 signalosome and eIF3 complexes of
translation initiation factors [62]; it plays an important
role in the stabilization of the entire 26S proteasome
complex and also integrates and coordinates the func
tioning of different proteasomal components by allosteric
regulation [63]. Binding of a ubiquitinated substrate caus
es many structural changes in the 19S regulatory com
plex. The most noticeable changes revealed by the cryo
electron microscopy include expansion of the channel in
the ATPase ring, as well as its alignment with the entrance
to the 20S proteasome complex. Similar structural
changes also occur upon binding of the nonhydrolyzed
adenosine triphosphate analogue ATPγS, that “freezes”
the enzymatic complex in a state that the proteasome
temporarily acquires after binding ATP [64]. Biochemical
and structural studies have revealed that the entrance to
the channel leading to the 20S proteasome proteolytic
chamber opens when the Cterminal HbYX motifs of
three ATPase subunits (Rpt2, Rpt3, and Rpt5) of the 19S
regulatory complex bind to the lysine residues in the
intersubunit pockets in the 20S proteasome outer ring
formed by the αsubunits. This interaction occurs after
Positive selection
Positive
selection
Golgi apparatus Golgi apparatus
thymo
proteasome
immuno
proteasome
peptides peptides
cTECs β5tdeficient cTECs
TAP TAP
MHC I MHC I
ER
Fig. 2. The unique catalytic subunit β5t. Together with β1i and β2i, β5t is forms the thymoproteasome with reduced chymotrypsinlike activ
ity. This proteasome type plays an important role in the positive selection of MHC Ispecific T cells, which significantly expands the final
repertoire of cytotoxic lymphocytes. cTEC, cortical thyroid epithelial cell; ER, endoplasmic reticulum; MHC I, major histocompatibility
complex class I; TAP, the transporter associated with antigen processing.
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S165
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
ATP binds to the ATPase ring subunits and triggers dis
placement of the Ntermini of the αsubunits, which
opens the entrance to the proteolytic chamber for the
substrate. The Cterminal parts of the remaining three
ATPase subunits lacking the HbYX motifs support the
association between the regulatory particle and the 20S
proteasome subunits.
During the last decade, highresolution electron
microscopy combined with Xray crystallography of indi
vidual proteasome subunits and generation of homolo
gous models have been used to determine the structures of
the 19S regulatory particle and the entire 26S proteasome
at the nearatomic resolution [65, 66] (Fig. 3).
According to the latest data, six ATPase subunits
(Rpt16) form a ring that anchors to the 20S particle.
Each ATPase consists of three domains: the extended α
helical domain, followed by the oligonucleotide and
oligosaccharidebinding domain (OB domain), and the
AAA+ domain [67] containing elements necessary for
ATP binding and hydrolysis. All six OB domains, togeth
er with the AAA+ domains, form two concentric rings.
The Cterminal tails of the AAA+ domains are located in
recesses on the 20S proteasome surface, opening the
entrance to the catalytic chamber (AAAring). The N
terminal αhelices of the AAA+ domains form three pairs
of coiled coils that constitute the outer border of the
cylindrical part of the 26S proteasome (Nring) (Fig. 4).
The ATPase ring accounts for about 2/3 of the total vol
ume of the 19S regulatory particle and provides a link
between the Ub receptors and the ATPase motor [65, 66].
The structural subunit Rpn1 is attached laterally to
the outer side of the ATPase ring, near the 20S subunit;
Rpn2 is attached to the top on the opposite side. Ub
receptors Rpn10 and Rpn13 are located at the protea
some periphery, where they can bind the substrates.
Rpn13 is associated with Rpn2 and situated at the maxi
mal distance from the proteasome center. A quarter turn
away from Rpn13, a little closer to the ATPase ring, there
is Rpn10 attached to Rpn11 deubiquitinase. Therefore,
the active sites are located along the proteasome central
Fig. 3. Proteasome structure. According to the Xray structural analysis (PDB 5GJR), the 20S proteasome is a hollow cylinder of 1517 nm
in length and 1112 nm in diameter, consisting of four rings lying on top of each other, with two identical peripheral rings formed by αsub
units (lilac), and two identical central rings formed by catalytic βsubunits (violet). Each of the rings consists of seven 2035 kDa subunits.
PA700 (also called the 19S regulatory complex) consists of 19 different protein subunits and contains two basic elements in its structure: the
lower (base) element consists of six subunits differing in structure and having ATPase activity (Rpt16, blue), Rpn2 structural subunit, and
three Ubbinding subunits, Rpn1, Rpn10 and Rpn13 (yellow; Rpn13 is not shown due to low resolution). PA700 binds directly to the αring
of the catalytic core and provides unfolding of the polypeptide chain of protein substrates. The energy of ATP hydrolysis is spent on the
polypeptide chain unfolding and its translocation into the catalytic chamber. The upper element consists of nine nonATPase subunits: Rpn3,
59, 1112 (gray). Rpn11 (orange) scans the substrate polypeptide chain and removes the polyUb chain.
26S proteasome
ββring
ααring
ATPase ring
S166 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
axis, on the way into the proteolytic chamber, according
to the order of substrate conversion. Lid structural sub
units (Rpn3, Rpn5, Rpn6, Rpn7, Rpn9, and Rpn12)
form a Ushaped frame connecting most of the proteaso
mal elements: Rpn2, Rpn11, ATPase ring, and αring of
the catalytic particle. The lid plays an important role in
the stabilization of the entire 26S proteasome complex
[68]; it also integrates and coordinates the functions of
different proteasome parts through allosteric regulation
[63, 69].
The amount of 20S proteasomes in the cell can be
increased via 26S proteasome disassembly into its compo
nents: 20S proteasome and 19S regulator. Several studies
have demonstrated that such disassembly actually takes
place under oxidative stress, when there is a need to
increase the efficiency of damaged proteins degradation
[70]. Apparently, the disassembly process may involve dif
ferent proteins. In mammalian cells, Hsp70 chaperone is
involved in the stabilization of the 19S regulator after its
dissociation from the 20S proteasome, as well as in the
reassembly of functional 26S proteasomes after oxidative
stress [71]. It was also found that low cellular levels of
Hsp90 cause almost complete disassembly of yeast 26S
proteasome and, correspondingly, an increase in the
amount of free 20S proteasomes [44]. It was shown
recently that the contents of 26S and 20S proteasomes
depend on the cell metabolic state. For example, low
NADH/NAD+ratio destabilizes the 26S proteasome
complex, leading to the emergence of free 20S protea
somes [72]. Similarly, a decrease in the ATP content
affects the 26S to 20S proteasome ratio in cells [73].
Substrate hydrolysis by the proteasome is associated
with a significant conformational reorganization of the
Rpt1Rpt6 ATPase subunits. Due to the cascade hydroly
sis of ATP in the substratebinding pockets, the Rpt16
subunits turn like a propeller and align in the horizontal
plane. This global structural reorganization leads to the
alignment of the channel formed by the ATPases and the
axis of the 20S catalytic core. Another important conse
quence of the ATPase rotation is displacement of Rpn11
deubiquitinase to a position directly above the catalytic
chamber entrance, which greatly facilitates scanning of
the polypeptide chain and subsequent cleavage of conju
gated Ub.
Posttranslational modifications of the proteasome.
Recent studies have demonstrated that posttranslational
modifications, such as phosphorylation, Nacetylation,
ubiquitination, myristoylation, glycosylation, ribosyla
ααhelical regions
Nring
AAAring
lid
anchoring Ctermini
coiledcoil
OB domain
AAA+ domain
Fig. 4. Structure of the ATPase ring of the 26S proteasome. Rpt16 subunits with ATPase activity are shown in blue, and the deubiquitinase
subunit Rpn11 is shown in orange. Each ATPase consists of three domains: the extended αhelical region, the oligonucleotide and oligosac
charidebinding domain (OB domain), and the AAA+ domain.
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S167
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
tion, and proteolytic cleavage of proteasomal subunits,
regulate proteasome assembly and localization and affect
its the proteolytic function [74]. Using modern proteom
ic techniques, about 110 co and posttranslational mod
ifications of the yeast proteasome have been identified,
but only a small part of them have been functionally char
acterized [75, 76]. The number, types, and sites of modi
fications are significantly different in the yeast, human,
and mouse proteasomes despite their high amino acid
homology [76]. Currently known functional modifica
tions of the proteasome are often divided into three types.
Modifications of the first type affect proteasome assembly
and stability (e.g., ADPribosylation of the PI31 factor
facilitates proteasome assembly [77]). Modifications of
the second type directly affect the proteolytic function of
the proteasome. Thus, glycosylation of the ATPase Rpt2
inhibits proteasome by reducing the ATPase ring activity
[78]. Modifications of the third type control proteasome
localization (e.g., Nmyristoylation of Rpt2 in yeast
[79]).
Phosphorylation. Almost all 26S proteasome sub
units are phosphorylated [76]. Phosphorylation of the
20S proteasome αsubunits primarily affects proteasome
stability. For example, phosphorylation of the α7 subunit
promotes association between the 19S regulatory complex
and the 20S proteasome with the formation of the 26S
proteasome. In the presence of INFγ, α7 subunit phos
phorylation is partially suppressed, which leads to the
destabilization of the 26S proteasome complex and,
according to [80], replacement of 19S complex with the
PA28 regulatory complex, although this fact has not been
confirmed in other studies [81]. Phosphorylation of the
regulatory complex subunits predominantly affect the
access of the substrate to the proteolytic chamber. For
example, phosphorylation of the Rpt6 ATPase subunit by
protein kinase A (PKA) or Ca2+/calmodulindependent
protein kinase II (CaMKII) [8284] increases the rate of
substrate degradation. PKA also phosphorylates the non
ATPase Rpn6 subunit, resulting in the acceleration of
substrate hydrolysis in vitro [85].
Ubiquitination. Ubiquitination of proteasome sub
units can be induced by proteasome inhibitors and pro
teotoxic stress [86, 87]. Under these conditions, five E3
ligases (Ube3a/E6AP, Ube3c/Hul5, Rnf181, Huwe1, and
Ubr4) mediating ubiquitination of approximately 14 sub
units were found to associate with the proteasome.
Multiple ubiquitination leads to the elimination of inac
tive proteasomes by autophagy (the socalled pro
teaphagy). However, modification of the Ub receptor
Rpn13, a specific substrate of E3 ligase Ube3c/Hul5,
does not result in proteaphagy. Ubiquitination of lysines
K21 and K34 located in the Nterminal fragment of the
Ubbinding Pru domain of Rpn13 reduces the ability of
Rpn13 to interact with ubiquitinated proteins, thereby
preventing their degradation [87]. Interestingly, partial
inhibition of the proteasome causes specific ubiquitina
tion of Rpn13 both in vivo and in vitro, which suggests that
this modification prevents the binding of ubiquitinated
substrates when the proteasome proteolytic functions are
impaired, although the physiological significance of this
mechanism is not yet completely understood. Another Ub
receptor, Rpn10, was found to be monoubiquitinated.
Monoubiquitination of Ub receptors often blocks their
ability to bind ubiquitinated substrates due to the
intramolecular interaction between the UBD and Ub [88].
This is also true for the Rpn10 subunit [74], whose ability
to recognize proteasomal substrates markedly decreases
upon ubiquitination. While proteotoxic stress upregulates
Rpn13 ubiquitination, the amount of monoubiquitinated
Rpn10 decreases under stress conditions.
Alternative regulators. The PA28 regulator (also
known as REG or 11S) is a 180kDa heptameric ring
shaped complex. It binds to one or both sides of the 20S
proteasome in an ATPindependent manner and signifi
cantly increases its ability to hydrolyze short peptide sub
strates, but not proteins or ubiquitinconjugated proteins.
However, there have been a growing number of studies
that claim that the 20S proteasome with the PA28 regula
tors can also degrade proteins. In addition, PA28 can bind
to the free end of the asymmetric 26S proteasome
(19S–20S) to form a hybrid proteasome (19S–20S–
PA28) that hydrolyzes tri and tetrapeptides at a higher
rate compared to the 26S proteasome [89].
PA28
αβ
.In mammals, PA28 consists of two homol
ogous subunits, PA28α(REGαor PSME1) and PA28β
(REGβor PSME2), whose expression is induced by
INFγ[90]. Professional antigenpresenting cells normal
ly express high amounts of PA28αβ, which is consistent
with the potential involvement of this complex in the
presentation of antigens on the MHC class I molecules
[91]. In addition, this complex is involved in the mainte
nance of protein homeostasis during oxidative stress [92],
regulation of cell growth, and apoptosis. PA28αand
PA28αβ were found to promote presentation of some, but
not all, antigens on MHC class I. Cells lacking this regu
latory complex have a lower ability to present certain
antigens. INFγinduced coordinated expression of
PA28αand PA28β, as well as β1i, β2i, and β5i protea
some immune subunits, leads to the formation of the
PA28αβ–20S immunoproteasomes in vitro and in vivo.
However, PA28αβ was also found in cells, tissues, and
organs lacking immunoproteasomes, e.g., red blood cells
and muscles. PA28αβ also facilitates the release of pep
tides that cannot be presented on MHC class I [93],
which indicates that this complex can function as a regu
lator of autoimmune processes by reducing excessive
cytotoxic response to autoantigens. In addition, PA28αβ
stimulates the degradation by the 20S proteasome of oxi
dized and misfolded proteins, regardless of their ubiquiti
nation [92, 9496].
PA28
γ
.There is also the third subunit – PA28γ
(REGγ, PSME3, or Ki antigen) that forms homohep
S168 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
tamers in the nucleus [97]. It is not INFγinducible and
not involved in the antigen presentation on the MHC
class I molecules. It was found that the PA28γand
PA28αβ differ in their activation properties and peptide
cleavage patterns. This difference may be associated with
the tendency of these regulators to bind to different pro
teasome types. Although it is still unclear how PA28γrec
ognizes substrates, it can hydrolyze short peptides as well
as intact, unstructured, or intrinsically disordered intra
cellular proteins via the Ubindependent mechanism. A
number of substrates hydrolyzed by the PA28γprotea
some have been identified; these include SRC3 [98], hep
atitis C virus core protein 3 [99], and inhibitors of cyclin
dependent kinases p21, p16, and p14 [94, 100]. PA28γ
directly interacts with the SRC3 domain of histone acety
lase (HAT) [98] and p21 [94] in vitro. PA28γknockout
mice are characterized by retarded growth, defects in
cellspecific mitosis, and reduced body weight in adults
[101, 102]. Therefore, PA28γcan be associated with the
metabolism and pathogenesis of various diseases via pro
moting Ubindependent substrate degradation.
PA200/Blm10. Another alternative regulator, the
PA200 protein encoded by the Psme4 gene, was discov
ered in rabbit reticulocyte lysate [103]. Its ortholog
Blm10 was found in yeast [104], nematodes, and plants
[105]. The PA200 subunit, a 250kDa highly conserved
protein located in the nucleus, can attached to one or
both ends of the 20S proteasome. In vitro, the proteasome
with this regulator hydrolyzes only short peptides or
unfolded proteins and is involved in the regulation of glu
tamine/glutamate homeostasis in tumor cells [106].
PA200 is present in all mammalian tissues, but it is most
abundant in the testes. It is believed that PA200 facilitates
the maintenance of protein homeostasis in mitochondria
[107]. This hypothesis is supported by the data obtained
in experiments with PA200 knockout mice. Male mice
deficient by PA200 were found to be infertile due to the
impaired spermatogenesis [108]. Further investigation
demonstrated that this regulator also contributes to the
chromosomal stability [109] and participates in the ATP
and Ubindependent degradation of acetylated histones
in somatic cells in response to DNA damage [110]. PA200
and Blm10 contain the bromodomain (BRD)like
regions that recognize acetylation. The BRDlike regions
of PA200 and Blm10 are able to specifically bind acety
lated core histones [110]. Male mice with the
PA200/PA28γdouble knockout are completely infertile,
with pronounced defects in the sperm motility due to the
decreased proteasomal activity [111]. Based on the
impairment of normal spermatogenesis in PA200defi
cient mice, PA200mediated degradation of histones is an
important component in the formation of spermatozoa.
Yeast lacking Blm10 are hypersensitive to DNAdamag
ing agents and exhibit a decreased respiratory capacity
[107]. The lack of Blm10 also results in the elevated sen
sitivity to oxidative stress, increased mitochondrial fis
sion, and reduced degradation of the fission protein
Dnm1 [112]. Blm10 proteasomes also mediate degrada
tion of the ribosomeassociated transcription factor Sfp1
in response to nutrient depletion, thereby affecting ribo
somal biosynthesis [113]. Therefore, the PA200/Blm10
proteasomes can regulate cellular metabolism by enhanc
ing proteasomal activity in the absence of Ub.
PI31. PI31, a 30 kDa prolinerich protein, was first
described as an inhibitor of proteasomal activity. It com
petes with the 19S (or PA28) regulatory particle for the
binding to the 20S proteasome [114]. A study conducted
in Drosophila melanogaster revealed that PI31 in a com
plex with the E3 ubiquitin ligase Nutcracker regulates
proteasome functions, promotes the activity of 26S pro
teasome, and suppresses the activity of free 20S protea
some. However, PI31 overexpression in mouse embryon
ic cells does not affect proteasomemediated hydrolysis
[115]. Instead, PI31 located on the nuclear or endoplas
mic reticulum membranes can selectively inhibit matura
tion of the immunoproteasome precursor complexes,
which in turn reduces antigen presentation by the MHC
class I molecules on the surface of INFγtreated mouse
embryonic cells.
Ecm29. Ecm29 is a 205kDa large protein capable of
binding to the 20S proteasome and regulating its function
by several mechanisms. Direct inhibition of proteasome
activity partially by inhibiting ATPase activity of the 19S
regulatory particle was shown in yeast [116]. On the other
hand, activation of the yeast proteasome by Ecm29 was
also described. Ecm29 was found to promote proteasome
assembly by stabilizing the intermediate 20S–19S com
plex, in which maturation of the 20S proteasome is
delayed due to a temporary lack of some βsubunits [117].
Other studies demonstrated that Ecm29 binds to the 19S
complex in response to the oxidative stress and causes the
26S proteasome disassembly [118]. It was suggested that
the Ecm29dependent disassembly of the 26S proteasome
increases the amount of the 20S proteasome, which
allows the cells to manage large amounts of oxidized pro
teins. In mammalian cells, Ecm29 (encoded by the
KIAA0368 gene) facilitates proteasome dissociation under
oxidative stress. Ecm29 is also associated with various
molecular motors and endosomal components, which
might be related to its ability to transfer 26S proteasomes
to various cell compartments, e.g., endoplasmic reticu
lum and centrosome.
Proteaphagy. The term proteaphagy is used to refer to
proteasome utilization by autophagy [119]. In
Arabidopsis, proteaphagy can is stimulated by nitrogen
starvation (i.e., conditions promoting nonselective
autophagy), while chemical or genetic inhibition of the
proteasome cause selective removal of nonfunctional
proteasomes. Selective elimination of proteasomes is
accompanied by extensive ubiquitination of proteasome
subunits, which promotes their recognition by the
extraproteasomal Rpn10 that acts as a classic selective
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S169
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
autophagosome receptor by simultaneously binding to
the Ubmodified subunits via the ubiquitininteracting
motif (UIM) and to the autophagosomal membrane pro
tein ATG8 via the ATG8interacting motif (AIM) [119].
In yeast, most proteasomes localize to the nucleus,
whereas autophagy is limited to the cytosol. Furthermore,
yeast Rpn10 does not contain the AIM. Similar to
Arabidopsis, yeast have two proteaphagy pathways that
respond either to the nitrogen starvation or inactivation of
proteasome particles, respectively [120, 121]. Similar to
Rpn10, the Cue5 protein acting as a selective proteaphagy
receptor was identified [120].
Mechanisms of Substrate Recognition by the Proteasome
To be hydrolyzed by the proteasome, a protein
should contain the degradation signal. In the most gener
al case, the degradation signal includes two components:
(i) a site for the recognition and binding by the protea
some and (ii) a site initiating protein unfolding and fur
ther translocation into the catalytic chamber of the 20S
proteasome [122]. In most substrates, the role of the
recognition signal belongs to the polyUb chain.
Ubdependent proteolysis. Ubiquitination system.
Ubiquitin (Ub) is an 8.5kDa signaling protein present in
all eukaryotic cells. Posttranslational modification of
proteins with Ub (ubiquitination) regulates numerous
cellular processes, such as protein degradation, sorting,
localization, and activation and repression of protein syn
thesis. Up to 14 different families of Ub and Ublike pro
teins are known that differ in the amino acid sequences
but have the same characteristic spatial structure.
Ub is attached to the target protein via sequential
actions of several enzymes. Covalent attachment of Ub to
the substrates is catalyzed by three enzymes: E1 (activat
ing enzyme), E2 (conjugating enzyme), and E3 ubiquitin
ligase. The E1 enzyme activates Ub in the twostep ATP
dependent reaction by forming the highenergy E1Ub
thioester complex. Next, activated Ub is transferred to
the Ubconjugating enzyme E2. E3 ubiquitin ligases
include two enzyme types. RING domainbearing E3 lig
ases bind to E2 and to the substrate and bring them to a
distance sufficient for the E2catalyzed Ub transfer to the
substrate; HECT domainbearing ubiquitin ligases inde
pendently catalyze Ub transfer to the amino group of the
target protein. In some cases, E4 ubiquitin ligases that
catalyze elongation of the Ub chain are distinguished in a
separate class.
The specificity of substrate ubiquitination is achieved
due to the hierarchy of the ubiquitination system. In
mammalian cells, there are only two types of Ubactivat
ing E1 enzymes (Uba1 and Uba6 [123]), ~30 conjugating
E2 enzymes, and ~600 E3 ubiquitin ligases. E1 enzymes
activate Ub for all types of E2 enzymes, most of which
interact with several E3 ubiquitin ligases. Usually, E3
enzymes transfer Ub to various substrates with similar or
identical recognition motifs. At the same time, specific
E3 ligases can interact with more than one E2 enzyme,
and some substrates can be recognized by more than one
E3 ligase. Therefore, the ubiquitination system hierarchy
is a complex network distinct from the pyramid (Fig. 5).
There are many ways of protein modification with
Ub, including monoubiquitination (of one or several
residues) and polyubiquitination with Ub chains different
in the linkage type and length. In the Ub molecule, seven
εamino groups of lysine residues K6, K11, K27, K29,
K33, K48, and K63 can form isopeptide bonds. The first
Ub molecule is usually attached to the substrate lysine
residue through the Cterminal glycine residue (G76).
Further chain growth occurs due to formation of isopep
tide bonds between lysines of the already attached Ub and
the Cterminal glycine residue of a new Ub molecule.
Also, Ub can be directly bound to the Nterminal
methionine residue of the substrate, followed by linear
attachment of Ub molecules. The structure of polyUb
chains is extremely diverse. PolyUb chains can be homo
geneous (with Ub molecules connected via lysine residues
in a strictly defined position) and heterogeneous (com
bine different linkage types); in turn, heterogeneous
chains can branch through ubiquitination at several sites
simultaneously [125]. In addition, serine, threonine, and
lysine residues of Ub can be phosphorylated and acetylat
ed [126]. The biological meaning of the ubiquitination
signal depends on the linkage type and polyUb chain
length. The ubiquitination signals are removed by deubiq
uitinating enzymes (DUBs) that cleave off the polyUb
chains from the substrates [127].
Ub is not a simple switch that triggers on/off the
degradation of protein substrates, but rather a finely
tuned signal that can determine the order in which pro
teins involved in the regulatory pathway will be
hydrolyzed. For example, cell advancement through the
cell cycle requires degradation of the regulatory proteins
in a correct order. The timing of protein hydrolysis can be
determined by the ubiquitination time; many E3 ubiqui
tin ligases recognize substrates only when their recogni
tion site is phosphorylated [128]. The order of degrada
tion might be also controlled by the polyUb chain type.
Thus, during the cell cycle, regulatory proteins with
longer polyUb chains are hydrolyzed faster than proteins
modified with shorter polyUb chains [129].
PolyUb chains. Thousands of proteins are ubiquiti
nated in cells, but almost half of them are not degraded by
the proteasome [130]. So it remains unclear how the cell
distinguishes between different Ub signals (Fig. 6). It is
commonly believed that the proteasome recognizes K48
linked polyUb chains containing at least four Ub mole
cules. Protein modification by a single Ub molecule,
polyUb chains linked through other lysine residues (e.g.,
K63), and linear polyubiquitin chains are not associated
with proteasomal degradation, but play a role in various
S170 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
cell processes, such as chromatin structure regulation,
membrane transport, and signal transduction. However,
this is not entirely true, as polyUb chains form by binding
via K63 [131] and monoubiquitinated proteins [132] can
be recognized by the proteasome. The binding affinity of
the purified proteasome for the K63linked polyUb
chains is almost the same as for the K48linked polyUb
chains [133]; therefore, the recognition of different
polyUb chains may involve some auxiliary proteins or
chain topology.
Recent advances in proteomic technologies have
provided a map of the Ub proteome [134]. By combining
quantitative proteomics and antidiglycine (diGly) anti
bodies recognizing ubiquitinated peptides in trypsin
hydrolysates, it was revealed that a large number of pro
teins (~5000) in cultured human cells are ubiquitinated
[130]. Several ubiquitination sites were found in ~60%
proteins with about ~4% proteins containing more than
10 modified sites; ~58% of the quantified sites increased
in abundance when the cells were treated with a protea
some inhibitor.
Using isotopelabeled Ub, it was demonstrated that
ubiquitinated substrates exist mainly in the monoubiqui
tinated form [135]. Therefore, proteomic studies have
shown that a significant number of endogenous protea
some substrates can by mono or polyubiquitinated at
several sites rather than modified with a single polyUb
chain. The chains formed through K48 quickly accumu
late in the cells treated with a proteasome inhibitor, while
chains formed through other residues (K6, K11, K27,
K29, K33, and to a lesser extent K63) also became more
abundant, which indicates involvement of such chains in
proteasomal degradation [136].
In line with the in vivo observations, recently report
ed in vitro studies have shown that purified proteasomes
can recognize a very wide range of polyUb chain topolo
gies in different substrates. A singlemolecule approach
was used to demonstrate that the proteasome efficiently
hydrolyzes cyclin B1 modified with several short polyUb
chains [137, 138]. It was found using a reconstituted sys
tem that the large multiprotein E3 ligase APC/C rapidly
modifies substrates with mono, di, and triUb at sever
signaling ubiquitination
ubiquitin precursors
Ub ligase cascade
free ubiquitin pool
proteasome
autophagy
Fig. 5. Schematic representation of the UPS. Ub (yellow ovals) is synthesized as four precursor proteins that are further processed by special
ized deubiquitinating enzymes – ubiquitin isopeptidases (DUBs). The UPS involves tree types of ubiquitin ligase (E1, E2, and E3; see the
text); it is highly specific and selective due to its hierarchical complexity. Ub is conjugated to substrates (S) as a monomer or the polyUb chain
linked through internal lysine residues. The polyUb chain is elongated by E3 ubiquitin ligases or relatively recently discovered E4 ubiquitin lig
ases. There is a dynamic equilibrium between ubiquitination and Ub removal by ubiquitin isopeptidase; the optimal polyUb chain length is
about six Ub residues per substrate molecule, according to the modern concept [124]. Next, the ubiquitinated substrate binds to the protea
some subunits Rpn10, Rpn13, and Rpn1 either directly or with the involvement of shuttle proteins of the UBLUBA family; there may also be
a variant of specific autophagy. Some Ub molecules enter the proteolytic chamber together with the substrate, where they get degraded. In most
cases, resident proteasomal deubiquitinase Rpn11 successfully removes the entire polyUb chain, which is then cleaved into monomers for recy
cle.
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S171
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
al sites [139]. In addition, cyclin B1 ubiquitinated with
chains of a certain length was produced and then used to
monitor the average dwell time on the proteasome by the
total internal reflection fluorescent microscopy (TIRF)
of individual molecules. The substrate binding increased
exponentially for the first three Ub units in any configu
ration (several monoUbs or short polyUb chains with four
to nine Ub molecules arranged linearly). Among poten
tial configurations with four Ub molecules, substrates
modified with two K48linked diUb chains were degrad
ed more efficiently than those modified with one K48
linked tetraUb chain. These findings were confirmed in a
later study that used a model GFP substrate fused to an
unstructured region [140]. At the same time, K48linked
polyUb chains are not always associated with substrate
degradation. In yeast, transcription activators Met4 [141]
and Cdc34 [142] can be ubiquitinated with long K48
linked polyUb chains, but they are not hydrolyzed by the
proteasome. Several short chains linked through K11,
K27, and K63 also contribute to the substrate binding to
the proteasome, which suggests that the proteasome does
not distinguish among the chain types in the case of sub
strate multiple ubiquitination. It is likely that the local
concentration of Ub units, but not the type of linkage, is
important for the substrate recognition by the protea
some.
In cells, APC/C with two different E2 ligases,
UBE2C and UBE2S, can synthesize branched ubiquitin
chains with K11 and K48linkages [143]. APC/C sub
strates modified with branched K11/K48linked chains
are degraded more efficiently than those containing
homotypic chains formed through either K48 or K11
linkages [143, 144]. In addition, K11linked chains are
involved in the processes unrelated to proteasomal degra
dation. Thus, branched K11/K63linked chains promote
internalization of MHC class I molecules by endocytosis
[145]. There is evidence that in yeast, K11linked polyUb
chains are as abundant (28% of all ubiquitin chains) [136]
as the K48linked chains; however, they are much less
common in unsynchronized mammalian cells (25% of
all ubiquitin chains) [135, 146]. As found recently, cells
contain a significant amount of branched K48/K63
linked polyUb chains [147]. The number of these chains
increases after cell treatment with a proteasome inhibitor,
thereby indicating that branched chains of this type are
proteasome degradation signals. It was also demonstrated
that the amount of K11/K48linked chains increases dur
ing proteotoxic stress caused by the inhibition of the pro
teasome, HSP70, or HSP90. In some cell types, includ
ing differentiated neurons, aggregates of newly synthe
sized and misfolded proteins are modified with the
K11/K48linked polyUb chains, which suggests the func
tion of these chains as markers for substrate degradation
by the proteasome. In addition, it was shown that patho
logical 73Qhuntingtin (HTT), but not benign 23QHTT,
is modified with the K11/K48linked chains and under
goes rapid proteasomal degradation. However, K11/K48
linked chains are redistributed during the proteotoxic
stress leading to the emergence of a large amount of
unfolded proteins; they modify newly synthetized
unstructured proteins, which in turn results in the accu
mulation of unmodified 73QHTT [148].
In yeast cells, K63linked polyUb chains account for
~16% of all linkages in yeast cells and, thus, rank third in
abundance after K48 and K11linked chains. Earlier, a
simple ubiquitination system was developed in [149] that
uses the HECTtype E3 ligase Rsp5 and a degron with the
PY motif. The Sic1 protein containing the PY motif
(Sic1PY) was overubiquitinated with long K63linked
chains under the action of Rsp5 ligase; the ubiquitinated
Sic1PY was rapidly degraded by the proteasome. This sys
tem has been widely used to analyze the proteasomal
function [60, 87]; however, as mentioned above, K63
linked chains do not participate in the proteasomal degra
dation in cells. One of the explanations for this phenom
enon may be the effect of UBDcontaining proteins that
Fig. 6. Different types of polyUb chains and their modifications. Ubiquitinconjugated substrates may be divided into three main types:
monoubiquitinated, polymonoubiquitinated, and polyubiquitinated. In addition, there is homotypic ubiquitination, when each polyUb
chain contains only one linkage type, and heterotypic ubiquitination, when Ub molecules can be connected through two or more different
types of linkages, thereby creating a mixed chain. Also, Ub can be modified at two or more sites, forming branched chains. In addition, Ub
can be conjugated with ubiquitinlike modifiers, such as SUMO, NEDD8, and ISG15; it can also undergo posttranslational modifications
(PTM), such as phosphorylation and acetylation.
Substrate
monoUb polymonoUb Homotypic chains Heterotypic chains Additional modifications
mixed branched Ublike proteins PTM
Signal complexity
S172 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
specifically recognize the K63links in vivo. In addition,
polyUb chains formed via K63 residues on endogenous
substrates are too short to be recognized by proteasomal
subunits or adaptor proteins.
Linear polyUb chains bind to the proteasome less
efficiently than chains with the K48links [150]. The only
known E3 ligase capable of assembling linear polyUb
chains is the linear Ub chain assembly complex (LUBAC)
that consists of three proteins: the catalytic RNF31 sub
unit (HOIP) [151], RBCK1 (HOIL1L), and SIPL1
(SHARPIN) [152154]. OTULIN (FAM105B or
GUMBY) is the main deubiquitinating enzyme (DUB)
with a high specificity for linear polyUb chains [155].
Linear polyUb chains are not sufficient for protein degra
dation by the proteasome in vitro [140]; however, they
were found to act as degradation signals in yeast and mam
malian cells [156, 157] (target proteins were either artifi
cially modified with such chains, or natural proteasomal
targets were used, such as protein kinase C and TRIM25
[158]). Recently, ATR, BRAP, LGALS7, PLAA, SEPT2,
HDAC6, VDAC1, and TRAF6 proteins were found to be
modified by linear polyUb chains. For example, the mod
ification of TRAF6 is necessary for proper functioning of
the IL1βdependent NFκB signaling pathway [159].
Interestingly, the type of ubiquitination required for
proteasomal degradation correlates with the size and
structural features of the substrate. Thus, monoubiquiti
nation is sufficient to induce proteasomal degradation of
small (<150 amino acids) structurally disordered proteins
in reticulocyte lysates [132, 160]. Recent proteomic study
using yeast and mammalian cells in which all Ub was
replaced by modified Ub lacking lysine residues and inca
pable of forming polyUb chains [161] showed that many
proteins (~25% in yeast and ~50% in human cells) were
degraded due to single or multiple monoubiquitination.
Furthermore, the degraded substrates shared some com
mon features; in particular, most of them were small in
comparison with polyubiquitinated proteins. Monoubi
quitinated substrates are more common among the pro
teins involved in carbohydrate transport and oxidative
stress, which may indicate that they are monoubiquitinat
ed by specific E3. In another study [162], 1392 ubiquitina
tion sites were identified in 794 proteins in HEK293T
cells. The ubiquitination sites were evenly distributed
throughout all identified proteins without significant pref
erence to the N or Cterminus. No obvious correlation
was found between the number of ubiquitination sites and
protein length. Also, 506 of the 794 identified proteins
were modified by a single lysine residue, while the remain
ing 288 proteins contained several Ubconjugated lysine
residues. Chaperones HSP701 and HSC71 had 22 and 15
ubiquitinated lysine residues, respectively; however, the
biological significance of this extensive ubiquitination
remains a mystery.
While physiological functions of homotypic polyUb
chains have been characterized, heterotypic chains
remain poorly understood [163]. In addition to the
K11/K48linked chains, four other heterotypic chain
types have been described: K29/K48, K11/K63,
K48/K63, and M1/K63. The existence and physiological
role of other heterotypic combinations still remain an
open question. Previous studies demonstrated that the
presence of both linkage types in the branched K11/K48
linked chains facilitates their recognition by proteasomal
UBDs and, thereby, ensures priority access of the modi
fied substrates to the proteasome. However, signal ampli
fication due to the polyUb chain elongation is rather
unlikely; to the contrary, the binding of a long chain over
a large area might be perceived by the UBDcontaining
receptor as a decrease in the overall signal level.
Deubiquitination system. PolyUb chains can be elon
gated or shortened even in the proteasomebound sub
strates due to the activity of E3 and E4 ubiquitin ligases
and deubiquitinating enzymes (DUBs). All currently
known DUBs are cysteine proteases or metalloproteases
that specifically hydrolyze the isopeptide bond immedi
ately after the Cterminal Ub residue (Gly76). About 100
DUBs have been found in mammalian cells; at least four
of them are associated with the proteasome (Rpn11,
Ubp6, UCH37, Doa4). The isopeptidase activity of free
DUB is rather weak, but it increases when the enzyme
binds to the proteasome. DUBs associated with the pro
teasome can allosterically affect certain proteasomal
functions, such as ATPase activity, opening of the prote
olytic chamber entrance, and substrate degradation.
These enzymes remove or gradually shorten polyUb
chains on the substrate, thereby limiting its dwell time on
the proteasome. Consequently, a protein that is difficult
to hydrolyze due to its compact folding or poorly accessi
ble initiation sites will be uncoupled from the proteasome
within a certain period of time after failed attempts to
degrade it, thereby freeing up the space for the next sub
strate molecule. In this regard, inhibitors of proteasome
associated DUBs, e.g., Ubp6/Usp14, are very promising
as potential agents for treatment of neurodegenerative
diseases because they improve the ability of proteasome to
hydrolyze degradationresistant substrates presumably
due to a longer time of their association with the protea
some [164]. Interestingly, inhibition of Ubp6/Usp14 also
accelerates proteolysis of oxidized proteins and enhances
cell resistance to oxidative stress. On the contrary,
inhibitors of other DUBs can cause accumulation of
ubiquitinated proteins [165], which makes them potential
drugs for the treatment of malignant tumors. In this case,
the biological effect of such inhibitors is generally similar
to that of proteasome inhibitors that have already been
successfully used in the treatment of multiple myeloma
[166].
Determining the detailed structure of the 19S regu
latory particle by the cryoelectron microscopy allowed to
suggest the mechanism of DUB activation upon binding
to the proteasome, effect of DUBs on the 19S complex
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S173
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
conformation, and their role in the substrate degrada
tion.
Rpn11. The most important DUB is the proteasomal
regulatory particle lid subunit RPN11 belonging to the
JAMM family of Zn2+ metalloproteinases. [61]. Rpn11
contains the catalytic JAMM motif that has also been
found in seven other human proteins, such as the Csn5
subunit of the COP9 signalosome, AMSH, AMSHLP,
the BRCC36 subunit of BRISC, MPND, and MYSM1
[167169]. The conservative JAMM domain contains the
EXnHS/THX7SXXD sequence, in which His and Asp
residues are fixed by coordination with the Zn2+ ion,
while the fourth coordination position is occupied by a
water molecule that forms the hydrogen bond with the
conserved Glu residue. Zn2+ ion acts as a Lewis acid and
enhances the nucleophilicity of the bound water, enabling
the nucleophilic attack on the isopeptide bond [168, 170].
Rpn11 is essential for proteasome activity and cell viabil
ity [171]. Mutations His109Ala and His111Ala in the cat
alytic site do not disrupt the assembly or conformation of
the proteasome, but largely inhibit the proteasomal func
tions and are lethal in yeast [61]. Rpn11 is located just
above the entrance to the ATPase ring [65]; therefore,
substrate deubiquitination occurs before Ub loses its con
formation. It is believed that due to this location, Rpn11
is uncapable of cleaving bonds between internal Ub mol
ecules in the polyUb chain because of steric hindrances
[172]. Deubiquitinase Ubp6 is nonstoichiometrically
associated with the proteasome and cleaves the bonds
between the Ub molecules in the polyUb chains [173].
Due to the presence of MPN (Mpr1/Pad1 Nterminal)
domain, Rpn11 forms a heterodimeric complex with
Rpn8 [174, 175]. The crystal structure of the
Rpn11–Rpn8 heterodimer revealed that the Insert1 loop
in Rpn11 is capable of blocking the access to the deubiq
uitinase active site, but the same loop also participates in
the isopeptide bond cleavage. Rpn5 stabilizes the inactive
state of Rpn11, thereby preventing its deubiquitinating
activity until Rpn11 is incorporated in the proteasome
[176]. Ub binding induces conformational rearrange
ments in Rpn11, and the Insert1 loop transits from the
closed inactive state to the βhairpin state, thereby open
ing the access to the active site. This transition is a rate
limiting step in deubiquitination and significantly accel
erates mechanical translocation of the substrate to the
ATPase ring. Therefore, rapid removal of Ub from the
captured substrates reduces the risk for Ub to be
hydrolyzed together with the substrate [177].
Ubp6/Usp14. Yeast deubiquitinase Ubp6 and its
highly conserved mammalian homologue Usp14 are asso
ciated with the 19S regulatory complex lid [178]. Ubp6
was initially identified as proteasomal DUB by using the
proteasome inhibitor ubiquitin vinyl sulfone (UbVS, Ub
derivative modified at the Cterminus) that binds to the
catalytic cysteine residue of this deubiquitinase [179].
Removal of Ubp6 from the proteasome accelerates degra
dation of model substrates in vitro [180] and is not lethal
in Saccharomyces cerevisiae [181]. However, this removal
slows down cell growth due to the elevated degradation of
proteasomal substrates and depletion of free Ub [182].
Ubp6knockout embryonic fibroblasts display an upregu
lated deubiquitinating activity of Rpn11, as well as an
increased number of Rpn13 and Uch37 deubiquitinase
subunits in the proteasome. In addition, proteasomes
lacking Ubp6 hydrolyze nonubiquitinated proteins more
actively [183]. In the absence of ubiquitinated substrates,
Ubp6 suppresses basal hydrolysis of ATP and nonubiqui
tinated substrates. Inactive proteasomes do not hydrolyze
ATP; the probability of the nonspecific proteolysis
decreases, while the specificity of hydrolysis of ubiquiti
nated substrates increases. Interaction with the protea
some activates Ubp6 about 300fold [182]. In the pres
ence of the nonhydrolyzed analogue ATPγS, Ubp6 shifts
the conformation of the 19S regulatory particle to that
similar to the substratebound state (s3), which leads to a
2fold increase in the efficiency of cleavage of the sub
strate UbAMC (Ub7amino4methylcoumarin)
[184]. Before the substrate is captured by the proteasome,
the Nterminal UBL domain of Ubp6 binds to Rpn1,
while its catalytic USP domain can take differently posi
tions. Ubp6 interaction with the ATPase ring stimulates
its deubiquitinating activity, probably by changing the
conformation of two surface loops (BL1 and BL223). The
activity of Ubp6 further increases upon the capture of the
ubiquitinated substrate [184, 185]. Ubbound Ubp6
maintains the coaxially aligned state of the Nring,
AAA+ ring, and 20S proteasome [184]. Therefore, ubiq
uitinbound Ubp6 stimulates the ATPase activity, pro
motes the opening of the entrance to the 20S proteasome,
and facilitates inhibition of the substrate capture by desta
bilizing the s1 (substrateacceptor) state or preventing the
proteasome from returning to this conformational state
[182, 184, 185]. Taking these facts together, Ubp6 plays a
key role in the allosteric regulation of the proteasome,
partly through its binding to Ub. It was also was shown
that Ubp6containing proteasomes are able to recognize
substrates with multiple ubiquitination, while protea
somes lacking Ubp6 do not exhibit this activity. Ubp6 and
Rpn11 cleave K1 and K63linked chains with equal effi
ciency that increases with the chain elongation. In the
case of K48linked chains, an inverse dependence is
observed. Ubp6 removes K48linked chains more effi
ciently than Rpn11 [186]; the cleavage occurs only when
more than one chain is attached to the substrate [173].
Ubp6 cannot remove the closest to the substrate Ub mol
ecule regardless of the substrate positioning probably due
to steric hinderances. If the chain bypassed by Ubp6 is
short, then the substrate hydrolysis is suppressed; if the
chain length is sufficient, then the proteasomesubstrate
interaction will be preserved and the chain will be
removed by Rpn11 after initiation of the substrate
translocation [173]. Therefore, Ubp6 is most likely
S174 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
involved in the removal of the excessive Ub molecules
from the substrate [173].
Uch37. Cysteinedependent deubiquitinase Uch37
was identified as a subunit of the 19S regulatory complex
in Homo sapiens and D. melanogaster. This enzyme was
also found in Schizosaccharomyces pombe (Uch2), but not
in S. cerevisiae. Like Ubp6, Uch37 is activated by binding
to the proteasome. The crossover loop of the Uch37
active site blocks the catalytic cysteine [187], but interac
tion with the Cterminal deubiquitinase adaptor domain
(DEUBAD) of Rpn13 stabilizes the loop in one state,
which increases protein affinity for Ub [188, 189]. In the
extraproteasomal Rpn13, the DEUBAD interacts
intramolecularly with the Pru domain and decreases the
affinity of this receptor subunit for Ub. This interaction is
disrupted by Rpn13 binding to the proteasome, which
makes the above domains accessible for Ub binding and
facilitates Uch37 activation [190]. These facts suggest
that Rpn13 activates Uch37 only in the proteasome com
plex [188, 189]. In the proteasome, Uch37 is able to
cleave distal K48, K6, and K11linked polyUb chains.
It was suggested that Uch37 can edit ubiquitinated sub
strates and facilitate removal of insufficiently ubiquitinat
ed and very slowly hydrolyzed ubiquitinated proteins
from the proteasome [191]. In addition, Uch37 can
remove regulatory Ub from proteasome subunits [86],
cleave polyUb chains, and vacate Ubbinding proteasome
receptors, thus allowing continuous substrate loading
[192]. Despite a large distance between this deubiquiti
nase and the ATPase ring [193], Uch37 also stimulates
the gate opening and the activity of proteasomal ATPases,
which suggests that Uch37, like Ubp6, is able to affect the
proteasome conformational state [194]. These data
demonstrate close allosteric relationships between all
proteasome subunits. The yeast protein Doa4 is another
DUB associated with the 19S regulator, but this associa
tion is less strong than proteasome interactions with
Ubp6 and Uch37 [195].
E3 ubiquitin ligases were shown to be capable of
binding to the proteasome [196]. In particular, E3 ubiqui
tin ligase Hul5 is associated with Ubp6 in the 19S regula
tory complex and regulates the protein degradation rate
by counteracting the activity of Ubp6, i.e., by catalyzing
elongation of polyUb chains [197]. It is important to note
that Ub attachment to the substrate directly in the pro
teasome makes substrate degradation more processive
because it allows to avoid formation of partially degraded
protein fragments [198] through reubiquitination as the
proteasome moves along the polypeptide chain of extend
ed proteins [199].
Ublike proteins. Some Ublike (UBL) proteins,
sometimes called Ublike modifiers (ULMs), are struc
turally similar to Ub and can be conjugated to other pro
teins in the reactions catalyzed by enzymes that are simi
lar (but not identical) to the E1E3 enzymes of the ubiq
uitination system. ULM functions include a wide variety
of processes, such as autophagy, nuclear protein transfer,
alignment/segregation of replicated chromosomes, DNA
repair, and various signaling pathways. Each member of
this family has a βgrasp fold consisting of a fivestranded
βsheet that is partially wrapped around the central α
helix [200]. UBL proteins are divided into two classes:
type I proteins modify the substrates (SUMO, NEDD8,
ATG8, ATG12, URM1, UFM1, FAT10, and ISG15 fam
ily proteins) [201], while type II UBLs are usually com
ponents of multidomain proteins – many of them are
found in some E1 activating enzymes, E3 ligases, and
Ub/Ubl proteases. For example, Hub1 and Esc2 proteins
comprising the autonomous Ubl domain may also be
considered as type II UBLs because they do not form
conjugates with the substrates [202, 203].
SUMO. Modification of proteins through attach
ment of the small ULM SUMO (sumoylation) plays an
essential role in the regulation of various cellular process
es, such as nuclear transport, transcription, DNA repair
and replication, apoptosis, stabilization of protein mole
cules, and cell advancement through the cell cycle [204].
Like ubiquitination, SUMO conjugation to the substrate
occurs through the isopeptide bond formation between
the Cterminal glycine residue in the SUMO molecule
and εamino group of a lysine residue in the substrate
molecule. Pathogens and stress (e.g., heat, oxidative
stress, insufficient blood supply) lead to the global
changes in the sumoylation pattern. Impaired sumoyla
tion and disorders in the sumoylation regulation con
tribute to the development of serious disorders, such as
malignant neoplasms and heart failure [205].
FAT10. Covalent modification by the Ublike FAT10
protein (human leukocyte antigenF adjacent transcript
10) facilitates Ubindependent proteasomal degradation
of proteins [206]. Like other ULMs, FAT10 is bound to
the substrates through the isopeptide bond formation by a
cascade of E1 and E2 enzymes, in which UBA6 (also
called UBE1L2, E1L2, or MOP4) activates FAT10 and
USE1 (UBA6specific E2) conjugates it to the substrates
[207, 208]. FAT10 is involved in several important cellu
lar processes, including apoptosis and NFκB activation.
Basal expression of FAT10 is most pronounced in the
immune system organs, such as thymus, liver, lymph
nodes, and spleen. FAT10 expression can be induced by
proinflammatory cytokines IFNγand TNF; its expres
sion is increased in dendritic cells during their maturation
[206]. Mice lacking FAT10 are viable and fertile, which
indicates a weak dependence of constitutive functions on
FAT10 [209]. However, lymphocytes of FAT10deficient
mice are more prone to spontaneous apoptotic death;
furthermore, FAT10knockout mice demonstrate a high
sensitivity to endotoxins [210]. According to some data,
FAT10 itself can undergo Ubindependent hydrolysis by
the proteasome due to its ability to interact with the VWA
domain of the Ubbinding subunit Rpn10 [206].
However, an alternative hypothesis exists that FAT10 is
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S175
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
degraded by the proteasome with the involvement of Ub
[211].
Ub receptors. Ubiquitinated substrates are recog
nized by Ub receptors before undergoing proteasomal
degradation. The Ub receptors can be classified accord
ing to their association with the proteasome: components
of the 19S regulatory complex and nonproteasomal pro
teins that bind ubiquitinated substrates and deliver them
to the proteasome.
A) Proteasomal Ub receptors. Three subunits of the
19S regulatory complex were shown to bind ubiquitinated
substrates: Rpn13 [59], Rpn1 [60], and Rpn10 [212].
Rpt5 [213] and Rpn15 [214] are also considered as poten
tial Ub receptors because they have been demonstrated to
bind Ub; however, whether these proteins recognize ubiq
uitinated substrates targeted to proteasomal degradation
still remains unclear.
Ub receptors are located at a distance of 7080 Å
from the entrance to the Nring of ATPases, which
approximately corresponds to four Ub molecules in the
polyUb chain. The process of substrate translocation into
the proteolytic chamber starts through the Nring. The
arrangement of Ub receptors at the proteasome complex
periphery provides the proteasome with the ability to cap
ture substrates with diverse geometry of polyUb chains
and folded domains for their more efficient binding,
unfolding, and deubiquitination.
Rpn10. Rpn10 (S5a/PSMD4 in mammals) was the
first identified Ub receptor. It binds polyUb chains and
UBL domains of substrates through the UIMs [215, 216],
while its affinity for free Ub is much lower. Rpn10 is a
unique Ub receptor because it is able to function both in
the proteasomeassociated and free state, as it has been
shown for D. melanogaster, S. cerevisiae, and Arabidopsis
thaliana [217]. Rsp5dependent monoubiquitination
mediates Rpn10 dissociation from the proteasome and
leads to its accumulation in the cytoplasm [74]. Free
Rpn10 has been suggested to act as an adaptor (shuttle)
protein that binds ubiquitinated substrates, thereby
increasing proteasomal processivity and presumably
compensating for a low diffusion capacity of this large
proteolytic complex [218]. Proteasomebound Rpn10
captures polyubiquitinated substrates, while its free form
exhibits high affinity for the UBL domains of Dsk2 [217]
and Rad23/hHR23 proteins [219]. Monoubiquitination
of Rpn10 regulates its ability to bind substrates, because it
promotes intramolecular interactions that reduce the
ability of the Rpn10 UIM to bind ubiquitinated proteins
[220]. Rpn10 contains the Nterminal von Willebrand A
(VWA) domain that facilitates Rpn10 binding to the pro
teasome and promotes degradation of some ubiquitinated
substrates [221]. Rpn10 also has two UIM domains
(UIM1 and UIM2) that enable Rpn10 to bind K48 and
K63branched polyUb chains [222]. The yeast ortholog
of Rpn10 has only one UIM that recognizes K48linked
polyUb chains [223]. Disruption of Ub binding to the Ub
receptor does not cause the death of S. cerevisiae cells,
but results in the accumulation of polyubiquitinated pro
teins [59, 60]. In higher eukaryotes, deletion of the Rpn10
UIM was found to be lethal in Mus musculus. Rpn13
deficient mice are viable, but have tissuespecific defects
in proteasome functions [224, 225], which probably indi
cates the existence of specific substrates of some Ub
receptors.
Rpn13. Rpn13 was initially identified as a subunit of
the 19S regulatory complex in yeast. In mammals, it was
first discovered as a 46kDa membrane glycoprotein
(called ADRM1/GP110), whose expression was induced
by IFNγduring cell adhesion [226]. ADRM1/GP110 was
later identified as a subunit of the 19S regulatory particle
and its homology to the yeast Rpn13 was demonstrated
[227]. Rpn13 binds ubiquitinated substrates through its
Nterminal pleckstrinlike domain [58]. Another impor
tant region of Rpn13 is the KEKE motif of the Ctermi
nal DEUBAD that binds and activates Uch37 deubiquiti
nase. Together, they function as an editing complex that
removes extra Ub monomers and thereby limits the length
of polyUb chains in ubiquitinated substrates [228] to the
length optimal for substrate association with the protea
some in order to ensure efficient substrate degradation.
Rpn13 and Uch37 exert a synergistic effect on the degra
dation of ubiquitinated substrates, such as the inhibitor of
NFκBα(IκBα) and inducible nitric oxide synthase
(iNOS) [229]. Mice deficient by Rpn13 are viable, but
they are sterile and display decreased proteasomal activi
ty in the testes and brain. In addition, Rpn13 knockout
leads to an increase in the body fat and abnormal levels of
the growth hormone and follicular stimulating hormone
[225]. Rpn13 is associated with the degradation of liver
kinase B1 (LKB1), a key regulator of autophagy [230],
which may explain how Rpn13 is involved in the regula
tion of fat deposits degraded by autophagy [231]. The lack
of Rpn13 reduces recruitment of Parkin to the 26S pro
teasome and delays the autophagosomal clearance of
mitochondrial proteins [232], thereby indicating that
Rpn13 might be involved in the regulation of mitophagy.
Overexpression of Rpn13 promotes the growth of cancer
cells and their migration, whereas inhibition of Rpn13
induces apoptosis and suppresses tumor progression
[233]. Elevated Rpn13 levels are associated with multiple
myeloma [234] and other types of malignant tumors,
including ovarian [235], colorectal [236], and gastric
[237] cancers. Two lowmolecularweight Rpn13
inhibitors, RA190 [238] and KDT11 [239], were devel
oped for the treatment of malignant neoplasms. Unlike
proteasome inhibitors that target proteasome catalytic
subunits, bisbenzylidene piperidone RA190 inhibits
Rpn13dependent functions. It covalently binds to cys
teine 88 of the Pru domain in Rpn13 and disrupts inter
domain interactions in the protein [238]. RA190 can
cause apoptosis associated with the endoplasmic reticu
lum stress in multiple myeloma cells. It is important to
S176 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
note that application of this inhibitor may help to avoid
serious clinical complications that accompany the use of
Bortezomib and Carfilzomib [240]. KDT11 is struc
turally different from RA190 and interacts with a specific
Rpn13 region with medium affinity (KD ~2 μM). This
inhibitor exhibits a synergistic effect with Bortezomib in
the combination therapy of multiple myeloma [239].
Rpn1. Rpn1 was identified as a Ub receptor only
recently [60]. It contains two proteins binding sites, T1
and T2. UBLUBA proteins and polyUb chains bind to
T1, while the UBL domain of Ubp6 binds to the T2 site
[60, 241]. This binding facilitates Rpn1 association with
its substrates, adaptor proteins, and DUBs followed by
substrate binding. Rpn1 is a part of the regulatory particle
and interacts with the 20S proteasome [65]. Rpn1 and
Rpn2 form αhelical toroidal solenoids that serving as
substraterecruiting sites. Rpn1 contains nine segments
known as leucinerich repeats (LRRs) [242]. Rpn1, like
other Ub receptors, is able to interact with the adaptor
proteins, including Ddi1 and Ubp6. For example, bind
ing of Ddi1 to the proteasome depends directly on the
modification of D517 residue in the LRR1 domain of
Rpn1 [243].
B) Nonproteasomal Ub receptors. In addition to the
stoichiometric proteasomal Ub receptors, the nonpro
teasomal adaptor proteins UBLUBA can act as Ub
receptors. At present, there are three known families of
yeast mediator proteins that have homologues in higher
eukaryotes and act through a similar mechanism: Rad23,
Dsk2, and Ddi1. Each of these proteins contains the UBL
domain interacting with the proteasome, as well as one or
two ubiquitinassociated (UBA) domains that bind
polyUb chains. This UBLmediated interaction occurs
by binding to the Rpn1, Rpn13, or Rpn10 subunits. The
role of shuttle proteins in proteasomal degradation is con
troversial: depending on the concentration, they can
either accelerate and inhibit substrate hydrolysis.
Overexpression of Dsk2 inhibits proteolysis and has a
cytotoxic effect [217] that is reduced by binding of the
UIM domain of nonproteasomal Rpn10 to the UBL
domain of Dsk2.
Ubiquilins, eukaryotic orthologs of yeast Dsk2, are a
family of four UBL proteins that function as adaptors.
Ubiquilins facilitate degradation of damaged proteins
after oxidative stress. Mutations in ubiquilin impairing its
ability to bind to Rpn10 increase the amount of ubiquiti
nated proteins in the cell, which in turn leads to the pro
tein aggregate formation that may be associated with the
pathogenesis of some neurodegenerative diseases (e.g.,
amyotrophic lateral sclerosis, Huntington’s disease, and
Alzheimer’s disease [244]).
Rad23 contains two UBA domains: the central
UBA1 domain and the Cterminal UBA2 domain that
bind mono and polyubiquitinated substrates with differ
ent affinities. UBA1 binds K63linked polyUb chains
with higher affinity than K48linked chains, while the
UBA2 domain preferably binds K48linked chains [245].
In addition, ubiquitinated proteins associated with Rad23
are protected from subsequent modifications (e.g., elon
gation) of their polyUb ubiquitin chains, as well as from
deubiquitination. This stabilizing effect is supposed to
provide an efficient substrate delivery to the proteasome
[246]. Rad23 also participates in the endoplasmic reticu
lumassociated degradation (ERAD) through association
between its Rad4binding domain and deglycosylase
Png1 with the formation of a complex mediating protea
somal degradation of specific endoplasmic reticulum pro
teins [247].
Despite the fact that UBLUBA proteins directly
interact with the proteasome, they themselves are not
hydrolyzed by it. The stability of Dsk2 and Rad23 is
explained by the presence of the Cterminal UBA
domains protecting unstructured initiation sites in these
proteins [248, 249]. According to some data, the unstruc
tured initiation site of Rad23 cannot initiate hydrolysis,
regardless of the presence of the UBA domain, which may
indicate that proteasome has certain preferences for the
amino acid composition of the unstructured sites [142].
The p62 protein (sequestosome 1) is an adaptor pro
tein [250] that binds ubiquitinated substrates with its C
terminal UBA domain and attaches to the proteasomal
Rpt1 and Rpn10 subunits with its Nterminal PB1136
domain, thereby facilitating the delivery of protein sub
strates (e.g., tau protein) for proteasomal degradation
[251]. p62 also acts as a Ub receptor in autophagy by
directly binding to the LC3 protein, a known mediator of
autophagosome formation [252]. The role of p62 as a Ub
receptor in both proteasomal and autophagymediated
degradation of ubiquitinated proteins is also confirmed by
the fact that downregulation of endogenous p62 leads to
the accumulation of ubiquitinated proteins.
p97/VCP/Cdc48p. In addition to proteasomal Ub
receptors and UBAUBL adaptor proteins, there are
other proteins that can bind ubiquitinated substrates and
recruit them for proteasomal degradation. The p97 pro
tein [also known as valosincontaining protein (VCP) in
mammals and Cdc48p in S. cerevisiae] is an evolutionar
ily conserved ATPase present in all eukaryotes and
archaebacteria. p97/VCP/Cdc48p belongs to a large
ATPase family called AAA+ ATPases. Enzymes of this
family often act as chaperones promoting protein fold
ing/unfolding. p97 is also involved in DNA synthesis and
repair, membrane fusion, mitotic spindle disassembly,
autophagy, and proteasomal degradation. Given an
important role of this enzyme in protein quality control,
p97 mutations can cause some neurodegenerative dis
eases [253].
p97/Cdc48p has two AAA ATPase domains (D1 and
D2, respectively). A short linker (D1D2 linker) connects
the two ATPase domains, and another linker (ND1 link
er) connects D1 to a large Nterminal domain. A short
(~40amino acid) region is attached to the Cend of the
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S177
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
D2 domain. Interactions of p97/Cdc48p with partner
proteins are mainly mediated by the Nterminal domain,
although some proteins bind to p97/Cdc48p at its Cter
minus. Six Cdc48 monomers form a double ring sur
rounding the central pore. Although D1 and D2 domains
are homologous in both their sequence and structure,
they have different functions. For example, only the D1
domain, but not the D2 domain, is required for the p97
hexamer assembly.
In mammalian cells, p97 is located mainly in the
cytoplasm in fractions associated with the membranes of
subcellular organelles, such as the endoplasmic reticu
lum, Golgi apparatus, mitochondria, and endosomes.
Apparently, its perimembrane localization is mediated by
some membrane receptors that have not been identified
yet. p97/Cdc48p is also present in the nucleus, where it is
involved in quality control of nuclear proteins [254]. p97
is one of the most abundant proteins in eukaryotic cells.
In humans, expression of the p97 mRNA is moderately
upregulated in certain cancers; the expression level of p97
mRNA to some extent correlates with the sensitivity of
cancer cells to the p97 inhibitor currently considered as a
potential agent for treatment of malignant transforma
tions [255].
The mechanism of Cdc48/p97 action is poorly
understood, despite its crucial role in many cellular
processes. The best known p97/Cdc48 substrates are con
jugated to polyUb chains and degraded by the 26S pro
teasome. Accordingly, many cofactors and adaptors of
p97/Cdc48 are capable of recognizing Ub conjugates
[256]. The association of Cdc48 with ubiquitinated pro
teins is believed to be achieved by intermediaries (e.g.,
ubiquitin ligase) able to recognize both p97 and Ub. p97
can bind to Ub ligases and deubiquitinases that edit the
polyUb chains by making them suitable for the substrate
to be recognized by the proteasome or, on the contrary, to
avoid hydrolysis. Finally, the p97associated Ub ligase
recruits an adaptor protein that delivers the substrates to
the proteasome. Therefore, p97 dictates the fate of pro
teins and plays the key role in Ubdependent degradation
[257].
Initiation of substrate degradation. The proteasome
recognizes and binds substrates based on the presence of
the Ub tag, but the degradation itself is initiated at the
unstructured initiation site in the substrate [122]. After
substrate binding to the proteasome, the latter initiates
the unfolding of the polypeptide chain starting from the
initiation site and then translocates the substrate into the
proteolytic chamber, where protein degradation occurs.
Bacterial proteases of the AAA+ family recognize
degrons in their substrates due to the loops located in the
center of the ring of ATPase subunits. It is likely that the
proteasome recognizes the initiation sites in a similar way
[258]. In the proteasome, equivalent loops surround the
degradation channel, being at a distance of 3060 Åfrom
the entrance to the proteolytic channel. The channel
entrance is too narrow for the folded proteins to pass
through it; therefore, the unstructured initiation site
should be at least 2030 amino acid residues in length to
be within the reach of the ATPase loops. This length of
the unstructured site is consistent with the results of in
vitro degradation experiments using model proteasome
substrates. These experiments demonstrated that protein
hydrolysis by purified yeast proteasomes occurred faster
when the protein substrate contained an unstructured tail
of approximately 30 amino acids [259]. Therefore, the
loops of the AAA+ domains of the ATPase ring most like
ly function as “lobes”, initiating the substrate unfolding
and pushing it into the proteolytic chamber.
The need for unstructured initiation sites is reflected
in the global protein stability profile. At least 30% eukary
otic proteins contain unstructured regions that are neces
sary for these proteins to perform their functions in cells
[260]. There is bioinformatic evidence that proteins con
taining unstructured regions have on average shorter half
lives than proteins lacking these regions [261, 262]; how
ever, this correlation should be confirmed by more exper
imental data, since a number of other studies failed to
find it [263265]. Furthermore, it was found that the sites
of substrate ubiquitination are predominantly located in
unstructured protein regions [266, 267]. Even when the
protein lacks an unstructured region, ubiquitination itself
can cause local protein unfolding near the ubiquitinated
residue, which can create the initiation site for the pro
teasome [268].
The Ub tag and the initiation site do not have to be
in the same polypeptide chain. They can function sym
batically while being located in different polypeptide
chains of the protein complex, so that the ubiquitinated
subunit can facilitate degradation of its unubiquitinated
partner [269]. In this case, the ubiquitinated subunit acts
as an adaptor that binds to the proteasome and targets the
associated protein to proteolysis. Presumably, UBLUBA
proteins function in a similar way, acting as nonstoichio
metric Ub receptors for the proteasome [178].
On the other hand, the proteasome is able to
rearrange protein complexes by hydrolyzing only ubiqui
tinated subunit and leaving other proteins in the complex
intact [270, 271]. This rearrangement is important for
many regulatory processes. For example, during the cell
cycle in yeast, the proteasome removes the cyclin
dependent kinase inhibitor Sic1 from its complex with
cyclin and cyclindependent kinase by hydrolyzing Sic1
exclusively [272]. Shortly thereafter, cyclin is ubiquitinat
ed and then hydrolyzed, leaving the intact but inactive
kinase [273]. The most likely site for the degradation ini
tiation is the unstructured substrate region located closer
to the entrance to the proteasome proteolytic chamber.
Indeed, it was experimentally demonstrated that the ini
tiation site should be located at some distance from the
Ub tag to enable protein hydrolysis by the proteasome,
since substrate degradation requires simultaneous binding
S178 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
of the Ub tag and the substrate unstructured site by the
proteasomal ATPase ring [259].
Processing. Although most substrates are completely
digested by the proteasome, some of them undergo limit
ed proteolysis. Proteasomes are able to recognize the so
called stop signals encoded in the amino acid sequences
of proteins, such as ~60300amino acid glycinealanine
repeats (GArs) found in the EBNA1 protein of the
Epstein–Barr virus. EBNA1 is recognized by the protea
some with subsequent initiation of degradation; however,
its translocation through the proteolytic channel stops
when the GAr region reaches the ATPase ring. Aromatic
hydrophobic loops of the ATPase subunits are incapable
to properly interact with glycine and alanine residues due
to their small size, which causes polypeptide chain slip
ping followed by interruption of substrate translocation
and formation of partially processed products [274]. The
blockade of proteolysis helps the Epstein–Barr virus to
avoid presentation of EBNA1 antigenic peptides to MHC
class I and, therefore, to prevent immune response.
Another wellstudied example of limited proteolysis
is processing of the NFκB transcription factor p50 sub
unit from its p105 precursor. NFκB can exist as a homo
or heterodimer, with the p50–p65 complex being its most
common form. The p105 precursor can inhibit the NF
κB activity by forming dimers with proteins of the NFκB
family. The precursor blocks the Nterminal nuclear
localization signal of NFκB, thereby keeping the protein
in the cytoplasm. Activation of the NFκB signaling
pathway triggers the Ubdependent processing of p105 by
the proteasome, which leads to the hydrolysis of the C
terminal fragment of p105 and release of the nuclear
localization signal. After hydrolysis, p50 dimerizes with
p65 to form active NFκB complex and translocates to
the nucleus, where it regulates transcription. The block
ade of the proteasome during p105 processing into p50
occurs with the involvement of the stop signal consisting
of a glycinerich sequence followed by the tightly folded
Rel homology domain [198]. Similar mechanisms of lim
ited degradation were described for other transcription
factors, such as Spt23 and Mga2 in yeast and Cubitus
interruptus (Ci) in D. melanogaster. However, p105 can
also undergo complete hydrolysis, which is difficult to
explain in terms of the above concept alone. Recently, it
was shown that in addition to the previously known ubiq
uitinligase βTrCP, p105 can be modified by the alterna
tive KPC1 ubiquitin ligase [275]. The fate of p105 was
found to be determined by the type of ubiquitin ligase
involved in its modification, which suggests that these
enzymes perform different types of ubiquitination, and,
as a consequence, different signals are generated.
Ubindependent proteolysis. Most cellular proteins
are degraded by the proteasome with the involvement of
Ub; however, a number of proteins can be degraded with
out ubiquitination [276]. In the latter case, the substrate
should have alternative ways for association with the pro
teasome, such as auxiliary molecules or specific protein
regions (degrons). Degrons are minimal structural ele
ments that mediate interaction between protein substrates
and the proteasome resulting in protein degradation.
Despite the importance of these elements in proteostasis,
the number of known Ubindependent degrons is very
small.
The first discovered and most studied example of
Ubindependent proteasomal substrate is ornithine
decarboxylase (ODC), a protein involved in the biogene
sis of polyamines. Regulation of the polyamine levels is
important for proper cell proliferation, and increased
polyamine concentrations are associated with various
pathological processes, including carcinogenesis.
Therefore, the intracellular levels of ODC should be
strictly controlled. The optimal content of polyamines in
eukaryotic cells is maintained by a negative feedback
mechanism. High spermine and spermidine concentra
tions lead to the increased expression of the antizyme
(AZ) that replaces one of the ODC molecule in the cat
alytically active homodimeric complex, thereby inhibit
ing the enzymatic activity of ODC. In addition, AZ bind
ing changes the conformation of ODC, which results in
the exposure of its unstructured Cterminal fragment and
significantly acceleration of ODC hydrolysis [277, 278].
ODC degradation is carried out by the 26S proteasome,
but does not require Ub involvement. The 37amino acid
Cterminal fragment of ODC contains the proteasome
binding site and the unstructured hydrolysis initiation site
[277, 278]. The attachment of this ODC fragment to pro
teins that are not proteasomal substrates facilitates their
degradation via the Ubindependent pathway.
One possible explanation for the Ubindependent
degradation is that unstructured protein regions bind
closely enough to the ATPase ring loops, so that Ub is not
required for the association between the substrate and the
proteasome. This mechanism can be considered as a vari
ant of a common proteasomal degron that lacks one of
the components, namely, the ubiquitin tag, and resembles
degrons in archaea and bacteria [281].
The mechanisms of Ubindependent hydrolysis are
poorly understood. It is possible that proteins cleaved by
the isolated 20S particles in the absence of ATP in vitro
[282] are hydrolyzed by either the 20S proteasome acti
vated by alternative regulatory complexes [100] or even
the 26S proteasome under natural conditions [283].
Proteins belonging to Ubindependent proteasome sub
strates have long unstructured regions. Moreover, two
alternative proteasomal degradation mechanisms, Ub
dependent and Ubindependent, are not mutually exclu
sive; different fractions of the same protein can be
degraded via either of these two pathways.
Most substrates of the 20S proteasome are proteins
whose structure is partially or completely disordered due
to aging, mutations, or oxidation [284]. Native proteins
that contain large unstructured regions (>30 amino acids)
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S179
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
or intrinsically disordered regions (IDRs), as well as pro
teins with a completely unstructured sequence called
intrinsically disordered proteins (IDPs) [285], are also
degraded by the 20S proteasome. The latter group of sub
strates mostly includes key regulatory and signaling pro
teins that promote cell advancement through the cell
cycle and participate in the cell growth regulation and
carcinogenesis. Obviously, the content of these proteins
in the cells should be strictly controlled, because changes
in their concentration may lead to the development of
various diseases [286].
Currently, several proteins are known that are
hydrolyzed by the 20S proteasome due to their binding
directly to it. The IκBαprotein was demonstrated to
interact with the PSMA2/α2 subunit of the 20S protea
some due to the presence of certain repeated fragments in
the IκBαsequence [287], which presumably leads to the
Ubindependent hydrolysis of this protein. Recently, cal
cineurin was demonstrated to interact with PSMA2/α2
and to promote degradation of IκBαvia the Ubdepend
ent pathway [288]. The PSMA4/α3 subunit interacts with
the F protein of the hepatitis C virus, thereby facilitating
Ubindependent hydrolysis of this protein [289]. The
PSMA7/α4 subunit interacts with the REGα/βregulato
ry subunits (PA28α/β), as demonstrated by using a two
hybrid yeast system, as well as by inhibiting proteasome
activation by the hepatitis B virus polypeptide X that
binds directly to the PSMA7/α4 subunit [290]. The C
terminal part of PSMA7/α4 specifically interacts with the
Nterminal region of Rab7 and participates in the cargo
transport at the late stages of endocytosis; however, this
interaction does not facilitate Rab7 degradation [291].
Parkin (E3 ligase involved in the pathogenesis of
Parkinson’s disease) interacts with the Cterminal region
of PSMA7/α4 via its Cterminal IBRRING domain and
can function as an auxiliary protein in proteasomal
hydrolysis of substrates [292]. Also, PSMA7/α4 interacts
with the nucleotidebinding oligomerization domain
containing protein 1 (NOD1), thus promoting NOD1
degradation by the proteasome [293]. PSMA3/α7 also
interacts with the REGα/β(PA28α/β) particles that
mediate proteasome activation together with PSMA1/α1
and PSMA7/α4. Interaction of the p21WAF1/CIP1 C
terminus with PSMA3/α7 facilitates p21 degradation by
the Ubindependent mechanism [294]. Several proteins
were also shown to act as mediators to promote p21
hydrolysis by the proteasome. For example, MDM2 (E3
ubiquitin ligase) does not ubiquitinate p21, but rather
binds to this protein, thereby promoting its binding to the
PSMA3/α7 proteasome subunit [295]. The binding of
p21 to the REGγ(PA28γ) regulatory subunit also medi
ates p21 degradation by the proteasome [94]. The SRC3/
AIB1 protein is a steroid receptor coactivator that inter
acts directly with the PSMA3/α7 subunit [296] or binds
to REGγ(PA28γ) for the proteasomal degradation [98].
MDM2 binds to PSMA3/α7 and promotes interaction
between the retinoblastoma (Rb) protein and this sub
unit, which leads to the Ubindependent degradation of
Rb [297]. The Nterminal region (amino acids 160) of
αsynuclein, a protein involved in Parkinson’s disease,
interacts with the Cterminal region of PSMA3/α7,
thereby ensuring αsynuclein degradation [298].
PSMB6/β1 binds directly to the p27Kip1 protein and
promotes its proteasomal degradation [299]. The Smad1
protein undergoes hydrolysis by the proteasome, both by
the Ubdependent mechanism and without preliminary
modification with Ub [300] via binding PSMB4/β7 and
antizyme (AZ).
The list of cellular proteins whose degradation does
not require preliminary ubiquitination is constantly
expanding. Proteins that have already been described in
detail [283, 301] include ornithine decarboxylase (ODC),
p21, p53 whose degradation is inhibited by NAD(P)H
quinone oxidoreductase 1 (NQO1), cFos whose hydrol
ysis is also inhibited by NQO1 [302], Fra1 that directly
interacts with the 19S proteasome regulatory particle,
TBP1 containing a Cterminal Ubindependent degron
[303], Rb protein, αsynuclein, HIF1α, SRC3/AIB1
transcriptional coactivator, IκBα, Ybinding protein 1
(YB1), thymidylate synthase (TS), and tau protein
involved in the pathogenesis of Alzheimer’s disease.
Recently, myelin basic protein (MBP), which is one
of the main autoantigens in multiple sclerosis, was found
to undergo proteasomal hydrolysis without preliminary
modification with Ub [304, 305]. Furthermore, the num
ber of proteasome immune subunits was shown to
increase in the central nervous system in mice developing
experimental autoimmune encephalomyelitis (EAE). In
this case, the β1i subunit was localized mainly to oligo
dendrocytes (MBPexpressing cells), and β5i occurred
mainly in cytotoxic lymphocytes penetrating the central
nervous system through the damaged bloodbrain barrier
[306] (Fig. 7). Given the wellknown facts about the
mechanism of proteasome functioning, it may be suggest
ed that MBP binds either to the 19S regulatory particle or
to one of the alternative regulators due to its high positive
charge [307], because the 19S complex subunits are most
ly neutral or negatively charged. In addition, recent stud
ies have demonstrated that the immunoproteasome
hydrolyzes unstructured proteins, such as MBP and his
tones [308], at a higher rate compared to the constitutive
proteasome.
The list of Ubindependent proteasomal substrates
includes proteins that undergo Ubindependent protea
somal hydrolysis due to the interaction with alternative
proteasome activators, mainly REGγ(PA28γ) and
PA200/Blm10. In addition to p21, REGγ(PA28γ) is also
involved in the proteasomal degradation of other cell
cycle regulators, such as p16 (INK4A) and p19 (Arf) [94].
Activationinduced cytidine deaminase (AID), which is
responsible for the initiation of immunoglobulin gene
diversification in activated B lymphocytes, undergoes
S180 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
Ubindependent proteolysis also due to the interaction
with REGγ(PA28γ). The PA200/Blm10 activator binds
through its Cterminal YYX motif to the 20S proteasome
and activates degradation of the tau protein in vitro [309].
More recently, PA200/Blm10 was shown to facilitate Ub
independent degradation of acetylated histones [110].
Some attempts have been made to define the mini
mum requirements for a protein substrate hydrolyzed via
the Ubindependent mechanism [310312]; however, no
generalized mechanism for this process exists so far. To
elucidate this mechanism, it is necessary to identify spe
cific and nonspecific interactions with proteasomal sub
units that mediate Ubindependent substrate degrada
tion. One of the critical problems is identification of pro
teasome subunits in the 20S, 19S, or alternative regulato
ry particles that specifically interact with proteins
hydrolyzed by the Ubindependent mechanism. At the
same time, there is a hypothesis that substrates can enter
the proteolytic chamber through the lateral surface of the
20S subunit via the free space between the α and βsub
units, thereby bypassing the end channels blocked by N
terminal peptides of the αsubunit [27].
Several mechanisms that prevent partially or com
pletely the degradation of partially or completely unstruc
tured substrates have been discovered, e.g., interaction
with the socalled nanny proteins that mask the unstruc
Fig. 7. The role of immunoproteasomes in the development of autoimmune neurodegeneration and promising approaches in targeted ther
apy of this disease. The proteasome is a major proteolytic complex that generates peptides presented on the cell surface by the MHC class
I. The lack of the ubiquitination system control in hydrolysis of the myelin basic protein (MBP) means that the qualitative and quantitative
spectrum of MBP peptides presented on the oligodendrocyte surface is almost completely determined by the proteasome catalytic subunits.
We found that in the brain of immunized mice with experimental autoimmune encephalomyelitis (EAE; an animal model of multiple scle
rosis), the constitutive proteasome was largely replaced with the PA28αβ immunoproteasome under action of IFNγ; the β1i immune sub
unit was preferably localized to oligodendrocytes. An increased content of the immunoproteasome in the brain of EAE mice elevated the
amount of some pathogenic MBP peptides, including the ENPVVHFF peptide that is a fragment of the encephalitogenic MBP region.
Activated CD8+T cells specific for this peptide efficiently lysed oligodendrocytes treated with IFNγ. The specific inhibitor of the β1i
immune subunit (IPSI001) selectively affected the immunoproteasome in vitro and suppressed the of EAE development in experimental
animals. These findings suggest a relationship between the Ubindependent proteolytic cleavage of MBP by the immunoproteasome and
development of multiple sclerosis, as well as indicate the potential of specific immunoproteasome inhibitors as drugs. ER, endoplasmic
reticulum; MHC I, major histocompatibility complex class I; TAP, the transporter associated with antigen processing; TCR, Tcell recep
tor; PAD, peptidylarginine deiminase; CTL, cytotoxic T lymphocyte; β2M, beta2 microglobulin.
immunoproteasome
Oligodendrocyte ACTIVATION
cytotoxicity
Golgi apparatus
TCR
MHC I
myelin
constitutive
proteasome
nucleus
cytoplasm ER
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S181
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
tured regions [313] or oxidationmediated structure sta
bilization [298]. However, the extent of occurrence of
these mechanisms is the subject of further research.
Inhibition of proteasome activity or large proteaso
mal load lead to the activation of autophagy in most cell
types, which clearly indicates a functional relationship
between the two degradation systems. By now, the shift
from proteasomal degradation to autophagy has been well
studied. Such shift provides a compensation for a reduced
proteolytic capacity of the proteasome and eliminates the
risk of accumulation of potentially toxic protein aggre
gates. So far, there is no reliable evidence for the opposite
shift from autophagy to the proteasome that would be
activated by the autophagy impairment. Autophagy inhi
bition caused by the nutrient deficiency in malignant
colon cells increases proteasome proteolytic activity and
expression of proteasomal subunits [314]. However, other
studies revealed no increase in the proteasome activity
upon inhibition of lysosomal functions [315]. Autophagy
inhibition in HeLa cells is accompanied by reduction in
the protein degradation by the UPS due to stabilization of
the p62 protein that, in its basic state, is hydrolyzed
together with the autophagosomal load. Accumulated
p62 sequesters ubiquitinated proteins thus delaying their
translocation to the proteasome; however, this protein
does not affect the proteasome activity [316]. In addition,
the compensatory function of the proteasome in the case
of impaired autophagy is unlikely due to a significant size
of most autophagosome substrates [317].
Apart from the compensatory function of autophagy
in the case of impaired proteasome activity, both systems
share several pathways involved in the coordination
between their activities in proteostasis and organelle
homeostasis. Both UPS and autophagy can affect each
other due to the mutual control of expression levels of
their key components. For example, during the oxidative
stress, the autophagy receptor p62 mediates autophagoso
mal degradation of the E3 ligase Keap1 and degradation of
whole proteasomes by proteaphagy. LC3 can undergo pro
teasomal degradation by the 20S proteasome, but can also
avoid proteolysis using the autophagy receptor p62 [14].
Depending on the cell state, autophagy and proteasomes
can have common substrates and common regulatory fac
tors. Several main regulators of both pathways were shown
to physically interact with each other, thereby providing
mutual control and coordination of the activities [14].
The most significant common feature of both prote
olytic systems is the use of Ub as a degradation signal.
Interestingly, even E3 ligases that have pronounced sub
strate specificity do not always modify substrates exclu
sively for one degradation system. For example, Parkin
E3 ligase, which plays a central role in mitophagy, medi
ates proteasomal degradation of some of its mitochondri
al substrates, while another subset of its substrates under
goes autophagosomal degradation.
Therefore, UPS and ALS together form a machinery
able to monitor the critical states of the cells, to prevent
toxicity resulting from protein misfolding, and to clean
the cells from superfluous proteins and organelles. These
two systems synergistically maintain cellular proteostasis;
moreover, the ALS is able to compensate for the UPS
dysfunction. During the past decades, considerable
efforts have been made to understand the molecular
mechanisms underlying the UPS and ALS. Investigation
of biochemical pathways connecting these two systems
may help to develop new drugs that would enhance cell
sensitivity to intracellular degradation inhibitors in the
anticancer therapy and increase the effectiveness of mis
folded protein degradation in the treatment of neurode
generative diseases. Further investigations of the UPS and
ALS molecular mechanisms and relationships between
them will not only ensure the progress in basic science but
also promote active development of translational medi
cine.
Funding
The work was supported by the Russian Science
Foundation (project 141400585P).
REFERENCES
1. Schubert, U., Anton, L. C., Gibbs, J., Norbury, C. C.,
Yewdell, J. W., and Bennink, J. R. (2000) Rapid degrada
tion of a large fraction of newly synthesized proteins by pro
teasomes, Nature, 404, 770774.
2. Joazeiro, C. A. P. (2017) Ribosomal stalling during trans
lation: providing substrates for ribosomeassociated pro
tein quality control, Annu. Rev. Cell Dev. Biol., 33, 343
368.
3. Brandman, O., and Hegde, R. S. (2016) Ribosomeassoci
ated protein quality control, Nat. Struct. Mol. Biol., 23, 7
15.
4. Chiti, F., and Dobson, C. M. (2017) Protein misfolding,
amyloid formation, and human disease: a summary of
progress over the last decade, Annu. Rev. Biochem., 86, 1
42.
5. Lamark, T., Svenning, S., and Johansen, T. (2017)
Regulation of selective autophagy: the p62/SQSTM1 para
digm, Essays Biochem., 61, 609624.
6. Kaganovich, D., Kopito, R., and Frydman, J. (2008)
Misfolded proteins partition between two distinct quality
control compartments, Nature, 454, 10881095.
7. Balchin, D., HayerHartl, M., and Hartl, F. U. (2016) In
vivo aspects of protein folding and quality control, Science,
353, aac4354.
8. McClellan, A. J., Tam, S., Kaganovich, D., and Frydman,
J. (2005) Protein quality control: chaperones culling cor
rupt conformations, Nat. Cell Biol., 7, 736741.
9. Labbadia, J., and Morimoto, R. I. (2015) The biology of
proteostasis in aging and disease, Annu. Rev. Biochem., 84,
130.
S182 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
10. Bard, J. A. M., Goodall, E. A., Greene, E. R., Jonsson, E.,
Dong, K. C., and Martin, A. (2018) Structure and function
of the 26S proteasome, Annu. Rev. Biochem., 87, 697724.
11. Lamb, C. A., Yoshimori, T., and Tooze, S. A. (2013) The
autophagosome: origins unknown, biogenesis complex,
Nat. Rev. Mol. Cell Biol., 14, 759774.
12. Hershko, A., and Ciechanover, A. (1998) The ubiquitin sys
tem, Annu. Rev. Biochem., 67, 425479.
13. Kwon, Y. T., and Ciechanover, A. (2017) The ubiquitin code
in the ubiquitin–proteasome system and autophagy, Trends
Biochem. Sci., 42, 873886.
14. Dikic, I. (2017) Proteasomal and autophagic degradation
systems, Annu. Rev. Biochem., 86, 193224.
15. Scott, D., Oldham, N. J., Strachan, J., Searle, M. S., and
Layfield, R. (2015) Ubiquitinbinding domains: mecha
nisms of ubiquitin recognition and use as tools to investi
gate ubiquitinmodified proteomes, Proteomics, 15, 844
861.
16. Grice, G. L., and Nathan, J. A. (2016) The recognition of
ubiquitinated proteins by the proteasome, Cell. Mol. Life
Sci., 73, 34973506.
17. Hartl, F. U. (2017) Protein misfolding diseases, Annu. Rev.
Biochem., 86, 16.
18. Rock, K. L., Gramm, C., Rothstein, L., Clark, K., Stein,
R., Dick, L., Hwang, D., and Goldberg, A. L. (1994)
Inhibitors of the proteasome block the degradation of most
cell proteins and the generation of peptides presented on
MHC class I molecules, Cell, 78, 761771.
19. MaupinFurlow, J., Gil, M., and Karadzic, I. (2004)
Proteasomes: perspectives from the Archaea, Front. Biosci.,
1, 17431758.
20. Dahlmann, B. (2016) Mammalian proteasome subtypes:
their diversity in structure and function, Arch. Biochem.
Biophys., 591, 132140.
21. DeMartino, G. N., and Slaughter, C. A. (1999) The pro
teasome, a novel protease regulated by multiple mecha
nisms, J. Biol. Chem., 274, 2212322126.
22. Tanahashi, N., Murakami, Y., Minami, Y., Shimbara, N.,
Hendil, K. B., and Tanaka, K. (2000) Hybrid proteasomes.
Induction by interferongamma and contribution to ATP
dependent proteolysis, J. Biol. Chem., 275, 1433614345.
23. Fabre, B., Lambour, T., Delobel, J., Amalric, F.,
Monsarrat, B., BurletSchiltz, O., and Bousquet
Dubouch, M.P. (2013) Subcellular distribution and
dynamics of active proteasome complexes unraveled by a
workflow combining in vivo complex crosslinking and
quantitative proteomics, Mol. Cell. Proteomics, 12, 687699.
24. Asano, S., Fukuda, Y., Beck, F., Aufderheide, A., Forster,
F., Danev, R., and Baumeister, W. (2015) A molecular cen
sus of 26S proteasomes in intact neurons, Science,347,
439442.
25. Rechsteiner, M., and Hill, C. P. (2005) Mobilizing the pro
teolytic machine: cell biological roles of proteasome activa
tors and inhibitors, Trends Cell Biol., 15, 2733.
26. Ramachandran, K. V., and Margolis, S. S. (2017) A mam
malian nervoussystemspecific plasma membrane protea
some complex that modulates neuronal function, Nat.
Struct. Mol. Biol., 24, 419430.
27. Groll, M., Ditzel, L., Lowe, J., Stock, D., Bochtler, M.,
Bartunik, H. D., and Huber, R. (1997) Structure of 20S
proteasome from yeast at 2.4 Åresolution, Nature, 386,
463471.
28. Groll, M., and Huber, R. (2004) Inhibitors of the eukaryot
ic 20S proteasome core particle: a structural approach,
Biochim. Biophys. Acta,1695, 3344.
29. Groll, M., Bochtler, M., Brandstetter, H., Clausen, T., and
Huber, R. (2005) Molecular machines for protein degrada
tion, Chembiochem, 6, 222256.
30. Zuhl, F., Tamura, T., Dolenc, I., Cejka, Z., Nagy, I., De
Mot, R., and Baumeister, W. (1997) Subunit topology of
the Rhodococcus proteasome, FEBS Lett.,400, 8390.
31. Baumeister, W., Walz, J., Zuhl, F., and Seemuller, E. (1998)
The proteasome: paradigm of a selfcompartmentalizing
protease, Cell, 92, 367380.
32. Kunjappu, M. J., and Hochstrasser, M. (2014) Assembly of
the 20S proteasome, Biochim. Biophys. Acta,1843, 212.
33. Osmulski, P. A., Hochstrasser, M., and Gaczynska, M.
(2009) A tetrahedral transition state at the active sites of the
20S proteasome is coupled to opening of the alpharing
channel, Structure,17, 11371147.
34. Rabl, J., Smith, D. M., Yu, Y., Chang, S.C., Goldberg, A.
L., and Cheng, Y. (2008) Mechanism of gate opening in the
20S proteasome by the proteasomal ATPases, Mol. Cell, 30,
360368.
35. Jones, C. L., Njomen, E., Sjogren, B., Dexheimer, T. S.,
and Tepe, J. J. (2017) Small molecule enhancement of 20S
proteasome activity targets intrinsically disordered pro
teins, ACS Chem. Biol., 12, 22402247.
36. Goldberg, A. L. (2003) Protein degradation and protection
against misfolded or damaged proteins, Nature,426, 895
899.
37. Keck, S., Nitsch, R., Grune, T., and Ullrich, O. (2003)
Proteasome inhibition by paired helical filamenttau in
brains of patients with Alzheimer’s disease, J. Neurochem.,
85, 115122.
38. Choi, W. H., de Poot, S. A. H., Lee, J. H., Kim, J. H., Han,
D. H., Kim, Y. K., Finley, D., and Lee, M. J. (2016) Open
gate mutants of the mammalian proteasome show
enhanced ubiquitinconjugate degradation, Nat. Commun.,
7, 10963.
39. Kulichkova, V. A., Fedorova, O. A., Tsimokha, A. S.,
Moiseeva, T. N., Bottril, A., Lezina, L., Gauze, L. N.,
Konstantinova, I. M., Mittenberg, A. G., and Barlev, N.
(2010) 26S proteasome exhibits endoribonuclease activity
controlled by extracellular stimuli, Cell Cycle,9, 840849.
40. Mittenberg, A. G., Moiseeva, T. N., Kuzyk, V. O., and
Barlev, N. A. (2016) Regulation of endoribonuclease activ
ity of alphatype proteasome subunits in proery
throleukemia K562 upon hemininduced differentiation,
Protein J.,35, 1723.
41. Kloetzel, P. M. (2004) Generation of major histocompati
bility complex class I antigens: functional interplay between
proteasomes and TPPII, Nat. Immunol., 5, 661669.
42. Sijts, E. J. A. M., and Kloetzel, P.M. M. (2011) The role
of the proteasome in the generation of MHC class I ligands
and immune responses, Cell Mol. Life Sci., 68, 14911502.
43. ChatterjeeKishore, M., Wright, K. L., Ting, J. P., and
Stark, G. R. (2000) How Stat1 mediates constitutive gene
expression: a complex of unphosphorylated Stat1 and IRF1
supports transcription of the LMP2 gene, EMBO J.,19,
41114122.
44. Chapiro, J., Claverol, S., Piette, F., Ma, W., Stroobant, V.,
Guillaume, B., Gairin, J.E., Morel, S., BurletSchiltz, O.,
Monsarrat, B., Boon, T., and van den Eynde, B. J. (2006)
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S183
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
Destructive cleavage of antigenic peptides either by the
immunoproteasome or by the standard proteasome results
in differential antigen presentation, J. Immunol., 176,
10531061.
45. Chen, W., Norbury, C., and Cho, Y. (2001)
Immunoproteasomes shape immunodominance hierar
chies of antiviral CD8+T cells at the levels of T cell reper
toire and presentation of viral antigens, J. Exp. Med., 193,
13191326.
46. Ebstein, F., TextorisTaube, K., Keller, C., Golnik, R.,
Vigneron, N., Van Den Eynde, B. J., SchulerThurner, B.,
Schadendorf, D., Lorenz, F. K. M., Uckert, W., Urban, S.,
Lehmann, A., AlbrechtKoepke, N., Janek, K., Henklein,
P., Niewienda, A., Kloetzel, P. M., and Mishto, M. (2016)
Proteasomes generate spliced epitopes by two different
mechanisms and as efficiently as nonspliced epitopes, Sci.
Rep.,6, 24032.
47. Berkers, C. R., de Jong, A., Schuurman, K. G.,
Linnemann, C., Geenevasen, J. A. J., Schumacher, T. N.
M., Rodenko, B., and Ovaa, H. (2015) Peptide splicing in
the proteasome creates a novel type of antigen with an
isopeptide linkage, J. Immunol., 195, 40754084.
48. Berkers, C. R., de Jong, A., Schuurman, K. G.,
Linnemann, C., Meiring, H. D., Janssen, L., Neefjes, J. J.,
Schumacher, T. N. M., Rodenko, B., and Ovaa, H. (2015)
Definition of proteasomal peptide splicing rules for high
efficiency spliced peptide presentation by MHC class I
molecules, J. Immunol., 195, 40854095.
49. Liepe, J., Marino, F., Sidney, J., Jeko, A., Bunting, D. E.,
Sette, A., Kloetzel, P. M., Stumpf, M. P. H., Heck, A. J. R.,
and Mishto, M. (2016) A large fraction of HLA class I lig
ands are proteasomegenerated spliced peptides, Science,
354, 354358.
50. Qureshi, N., Morrison, D. C., and Reis, J. (2012)
Proteasome protease mediated regulation of cytokine
induction and inflammation, Biochim. Biophys. Acta Mol.
Cell Res., 1823, 20872093.
51. Uechi, H., Hamazaki, J., and Murata, S. (2014)
Characterization of the testisspecific proteasome subunit
α4s in mammals, J. Biol. Chem.,289, 1236512374.
52. Tomaru, U., Ishizu, A., Murata, S., Miyatake, Y., Suzuki, S.,
Takahashi, S., Kazamaki, T., Ohara, J., Baba, T., Iwasaki, S.,
Fugo, K., Otsuka, N., Tanaka, K., and Kasahara, M. (2009)
Exclusive expression of proteasome subunit 5t in the human
thymic cortex, Blood, 113, 51865191.
53. Murata, S., Takahama, Y., and Tanaka, K. (2008)
Thymoproteasome: probable role in generating positively
selecting peptides, Curr. Opin. Immunol.,20, 192196.
54. Florea, B. I., Verdoes, M., Li, N., van der Linden, W. A.,
Geurink, P. P., van den Elst, H., Hofmann, T., de Ru, A.,
van Veelen, P. A., Tanaka, K., Sasaki, K., Murata, S., den
Dulk, H., Brouwer, J., Ossendorp, F. A., Kisselev, A. F.,
and Overkleeft, H. S. (2010) Activitybased profiling
reveals reactivity of the murine thymoproteasomespecific
subunit β5t, Chem. Biol., 17, 795801.
55. Murata, S., Sasaki, K., Kishimoto, T., Niwa, S., Hayashi,
H., Takahama, Y., and Tanaka, K. (2007) Regulation of
CD8+T cell development by thymusspecific proteasomes,
Science, 316, 13491353.
56. Sauer, R. T., and Baker, T. A. (2011) AAA+ proteases: ATP
fueled machines of protein destruction, Annu. Rev.
Biochem.,80, 587612.
57. Tomko, R. J., Funakoshi, M., Schneider, K., Wang, J., and
Hochstrasser, M. (2010) Heterohexameric ring arrange
ment of the eukaryotic proteasomal ATPases: implications
for proteasome structure and assembly, Mol. Cell, 38, 393
403.
58. Schreiner, P., Chen, X., Husnjak, K., Randles, L., Zhang,
N., Elsasser, S., Finley, D., Dikic, I., Walters, K. J., and
Groll, M. (2008) Ubiquitin docking at the proteasome
through a novel pleckstrinhomology domain interaction,
Nature,453, 548552.
59. Husnjak, K., Elsasser, S., Zhang, N., Chen, X., Randles,
L., Shi, Y., Hofmann, K., Walters, K. J., Finley, D., and
Dikic, I. (2008) Proteasome subunit Rpn13 is a novel ubiq
uitin receptor, Nature,453, 481488.
60. Shi, Y., Chen, X., Elsasser, S., Stocks, B. B., Tian, G., Lee,
B.H., Shi, Y., Zhang, N., de Poot, S. A. H., Tuebing, F.,
Sun, S., Vannoy, J., Tarasov, S. G., Engen, J. R., Finley, D.,
and Walters, K. J. (2016) Rpn1 provides adjacent receptor
sites for substrate binding and deubiquitination by the pro
teasome, Science,351, aad9421.
61. Verma, R., Aravind, L., Oania, R., McDonald, W. H.,
Yates, J. R., Koonin, E. V., and Deshaies, R. J. (2002) Role
of Rpn11 metalloprotease in deubiquitination and degrada
tion by the 26S proteasome, Science, 298, 611615.
62. Glickman, M. H., Rubin, D. M., Coux, O., Wefes, I.,
Pfeifer, G., Cjeka, Z., Baumeister, W., Fried, V. A., and
Finley, D. (1998) A subcomplex of the proteasome regula
tory particle required for ubiquitinconjugate degradation
and related to the COP9signalosome and eIF3, Cell, 94,
615623.
63. Matyskiela, M. E., Lander, G. C., and Martin, A. (2013)
Conformational switching of the 26S proteasome enables
substrate degradation, Nat. Struct. Mol. Biol.,20, 781788.
64. Sledz, P., Unverdorben, P., Beck, F., Pfeifer, G.,
Schweitzer, A., Forster, F., and Baumeister, W. (2013)
Structure of the 26S proteasome with ATPγS bound pro
vides insights into the mechanism of nucleotidedependent
substrate translocation, Proc. Natl. Acad. Sci. USA,110,
72647269.
65. Lander, G. C., Estrin, E., Matyskiela, M. E., Bashore, C.,
Nogales, E., and Martin, A. (2012) Complete subunit
architecture of the proteasome regulatory particle, Nature,
482, 186191.
66. Schweitzer, A., Aufderheide, A., Rudack, T., Beck, F.,
Pfeifer, G., Plitzko, J. M., Sakata, E., Schulten, K.,
Forster, F., and Baumeister, W. (2016) Structure of the
human 26S proteasome at a resolution of 3.9 Å, Proc. Natl.
Acad. Sci. USA,113, 78167821.
67. Djuranovic, S., Hartmann, M. D., Habeck, M., Ursinus,
A., Zwickl, P., Martin, J., Lupas, A. N., and Zeth, K.
(2009) Structure and activity of the Nterminal substrate
recognition domains in proteasomal ATPases, Mol. Cell,
34, 580590.
68. Pathare, G. R., Nagy, I., Bohn, S., Unverdorben, P.,
Hubert, A., Korner, R., Nickell, S., Lasker, K., Sali, A.,
Tamura, T., Nishioka, T., Forster, F., Baumeister, W., and
Bracher, A. (2012) The proteasomal subunit Rpn6 is a
molecular clamp holding the core and regulatory subcom
plexes together, Proc. Natl. Acad. Sci. USA,109, 149154.
69. BechOtschir, D., Helfrich, A., Enenkel, C., Consiglieri,
G., Seeger, M., Holzhutter, H.G., Dahlmann, B., and
Kloetzel, P.M. (2009) Polyubiquitin substrates allosteri
S184 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
cally activate their own degradation by the 26S proteasome,
Nat. Struct. Mol. Biol.,16, 219225.
70. Wang, X., Yen, J., Kaiser, P., and Huang, L. (2010)
Regulation of the 26S proteasome complex during oxidative
stress, Sci. Signal., 3, ra88.
71. Grune, T., Catalgol, B., Licht, A., Ermak, G., Pickering,
A. M., Ngo, J. K., and Davies, K. J. A. (2011) HSP70
mediates dissociation and reassociation of the 26S protea
some during adaptation to oxidative stress, Free Radic. Biol.
Med., 51, 13551364.
72. Tsvetkov, P., Myers, N., Eliav, R., Adamovich, Y., Hagai,
T., Adler, J., Navon, A., and Shaul, Y. (2014) NADH binds
and stabilizes the 26S proteasomes independent of ATP, J.
Biol. Chem.,289, 1127211281.
73. Huang, Q., Wang, H., Perry, S. W., and FigueiredoPereira,
M. E. (2013) Negative regulation of 26S proteasome stabil
ity via calpainmediated cleavage of Rpn10 subunit upon
mitochondrial dysfunction in neurons, J. Biol. Chem.,288,
1216112174.
74. Isasa, M., Katz, E. J., Kim, W., Yugo, V., Gonzalez, S.,
Kirkpatrick, D. S., Thomson, T. M., Finley, D., Gygi, S. P.,
and Crosas, B. (2010) Monoubiquitination of Rpn10 regu
lates substrate recruitment to the proteasome, Mol. Cell, 38,
733745.
75. Kikuchi, J., Iwafune, Y., Akiyama, T., Okayama, A.,
Nakamura, H., Arakawa, N., Kimura, Y., and Hirano, H.
(2010) Co and posttranslational modifications of the 26S
proteasome in yeast, Proteomics, 10, 27692779.
76. Hirano, H., Kimura, Y., and Kimura, A. (2016) Biological
significance of co and posttranslational modifications of
the yeast 26S proteasome, J. Proteomics,134, 3746.
77. ChoPark, P. F., and Steller, H. (2013) Proteasome regula
tion by ADPribosylation, Cell, 153, 614627.
78. Zhang, F., Su, K., Yang, X., Bowe, D. B., Paterson, A. J.,
and Kudlow, J. E. (2003) OGlcNAc modification is an
endogenous inhibitor of the proteasome, Cell, 115, 715
725.
79. Kimura, A., Kato, Y., and Hirano, H. (2012) Nmyristoy
lation of the Rpt2 subunit regulates intracellular localiza
tion of the yeast 26S proteasome, Biochemistry,51, 8856
8866.
80. Bose, S., Stratford, F. L. L., Broadfoot, K. I., Mason, G. G.
F., and Rivett, A. J. (2004) Phosphorylation of 20S protea
some alpha subunit C8 (alpha7) stabilizes the 26S protea
some and plays a role in the regulation of proteasome com
plexes by gammainterferon, Biochem. J., 378, 177184.
81. Schmidtke, G., Schregle, R., Alvarez, G., Huber, E. M.,
and Groettrup, M. (2017) The 20S immunoproteasome
and constitutive proteasome bind with the same affinity to
PA28αβ and equally degrade FAT10, Mol. Immunol., doi:
10.1016/j.molimm.2017.11.030.
82. Djakovic, S. N., MarquezLona, E. M., Jakawich, S. K.,
Wright, R., Chu, C., Sutton, M. A., and Patrick, G. N.
(2012) Phosphorylation of Rpt6 regulates synaptic strength
in hippocampal neurons, J. Neurosci.,32, 51265131.
83. Djakovic, S. N., Schwarz, L. A., Barylko, B., DeMartino,
G. N., and Patrick, G. N. (2009) Regulation of the protea
some by neuronal activity and calcium/calmodulin
dependent protein kinase II, J. Biol. Chem.,284, 26655
26665.
84. Jarome, T., Kwapis, J. L., Ruenzel, W. L., and Helmstetter, F.
J. (2013) CaMKII, but not protein kinase A, regulates Rpt6
phosphorylation and proteasome activity during the forma
tion of longterm memories, Front. Behav. Neurosci.,7, 115.
85. Lokireddy, S., Kukushkin, N. V., and Goldberg, A. L.
(2015) cAMPinduced phosphorylation of 26S protea
somes on Rpn6/PSMD11 enhances their activity and the
degradation of misfolded proteins, Proc. Natl. Acad. Sci.
USA,112, 71767185.
86. Jacobson, A. D., MacFadden, A., Wu, Z., Peng, J., and
Liu, C.W. (2014) Autoregulation of the 26S proteasome by
in situ ubiquitination, Mol. Biol. Cell, 25, 18241835.
87. Besche, H. C., Sha, Z., Kukushkin, N. V., Peth, A., Hock,
E.M., Kim, W., Gygi, S., Gutierrez, J. A., Liao, H., Dick,
L., and Goldberg, A. L. (2014) Autoubiquitination of the
26S proteasome on Rpn13 regulates breakdown of ubiqui
tin conjugates, EMBO J.,33, 11591176.
88. Hoeller, D., Crosetto, N., Blagoev, B., Raiborg, C.,
Tikkanen, R., Wagner, S., Kowanetz, K., Breitling, R.,
Mann, M., Stenmark, H., and Dikic, I. (2006) Regulation
of ubiquitinbinding proteins by monoubiquitination, Nat.
Cell Biol., 8, 163169.
89. Cascio, P., Call, M., Petre, B. M., Walz, T., and Goldberg,
A. L. (2002) Properties of the hybrid form of the 26S pro
teasome containing both 19S and PA28 complexes, EMBO
J.,21, 26362645.
90. Tanahashi, N., Yokota, K., Ahn, J. Y., Chung, C. H.,
Fujiwara, T., Takahashi, E., DeMartino, G. N., Slaughter,
C. A., Toyonaga, T., Yamamura, K., Shimbara, N., and
Tanaka, K. (1997) Molecular properties of the proteasome
activator PA28 family proteins and gammainterferon reg
ulation, Genes Cells, 2, 195211.
91. Ossendorp, F., Fu, N., Camps, M., Granucci, F., Gobin, S.
J. P., van den Elsen, P. J., Schuurhuis, D., Adema, G. J.,
Lipford, G. B., Chiba, T., Sijts, A., Kloetzel, P.M.,
RicciardiCastagnoli, P., and Melief, C. J. M. (2005)
Differential expression regulation of the alpha and beta
subunits of the PA28 proteasome activator in mature den
dritic cells, J. Immunol.,174, 78157822.
92. Li, J., Powell, S. R., and Wang, X. (2011) Enhancement of
proteasome function by PA28αoverexpression protects
against oxidative stress, FASEB J., 25, 883893.
93. Raule, M., Cerruti, F., Benaroudj, N., Migotti, R.,
Kikuchi, J., Bachi, A., Navon, A., Dittmar, G., and
Cascio, P. (2014) PA28αβ reduces size and increases
hydrophilicity of 20S immunoproteasome peptide prod
ucts, Chem. Biol.,21, 470480.
94. Li, X., Amazit, L., Long, W., Lonard, D. M., Monaco, J.
J., and O’Malley, B. W. (2007) Ubiquitin and ATPinde
pendent proteolytic turnover of p21 by the REGgamma
proteasome pathway, Mol. Cell, 26, 831842.
95. Pickering, A. M., and Davies, K. J. A. (2012) Differential
roles of proteasome and immunoproteasome regulators
Pa28αβ, Pa28γand Pa200 in the degradation of oxidized
proteins, Arch. Biochem. Biophys.,523, 181190.
96. Pickering, A. M., Koop, A. L., Teoh, C. Y., Ermak, G.,
Grune, T., and Davies, K. J. A. (2010) The immunoprotea
some, the 20S proteasome and the PA28αβ proteasome
regulator are oxidativestressadaptive proteolytic com
plexes, Biochem. J., 432, 585595.
97. Masson, P., Andersson, O., Petersen, U. M., and Young, P.
(2001) Identification and characterization of a Drosophila
nuclear proteasome regulator. A homolog of human 11S
REGgamma (PA28gamma), J. Biol. Chem.,276, 13831390.
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S185
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
98. Li, X., Lonard, D. M., Jung, S. Y., Malovannaya, A.,
Feng, Q., Qin, J., Tsai, S. Y., Tsai, M.J., and O’Malley, B.
W. (2006) The SRC3/AIB1 coactivator is degraded in a
ubiquitin and ATPindependent manner by the REGγ
proteasome, Cell, 124, 381392.
99. Moriishi, K., Okabayashi, T., Nakai, K., Moriya, K.,
Koike, K., Murata, S., Chiba, T., Tanaka, K., Suzuki, R.,
Suzuki, T., Miyamura, T., and Matsuura, Y. (2003)
Proteasome activator PA28gammadependent nuclear
retention and degradation of hepatitis C virus core protein,
J. Virol., 77, 1023710249.
100. Chen, X., Barton, L. F., Chi, Y., Clurman, B. E., and
Roberts, J. M. (2007) Ubiquitinindependent degradation
of cellcycle inhibitors by the REGγproteasome, Mol.
Cell, 26, 843852.
101. Murata, S., Kawahara, H., Tohma, S., Yamamoto, K.,
Kasahara, M., Nabeshima, Y., Tanaka, K., and Chiba, T.
(1999) Growth retardation in mice lacking the proteasome
activator PA28gamma, J. Biol. Chem.,274, 3821138215.
102. Barton, L. F., Runnels, H. A., Schell, T. D., Cho, Y.,
Gibbons, R., Tevethia, S. S., Deepe, G. S., and Monaco,
J. J. (2004) Immune defects in 28kDa proteasome activa
tor gammadeficient mice, J. Immunol.,172, 39483954.
103. Ustrell, V., Hoffman, L., Pratt, G., and Rechsteiner, M.
(2002) PA200, a nuclear proteasome activator involved in
DNA repair, EMBO J.,21, 35163525.
104. Doherty, K., Pramanik, A., Pride, L., Lukose, J., and
Wood Moore, C. (2004) Expression of the expanded
YFL007w ORF and assignment of the gene name BLM10,
Yeast, 21, 10211023.
105. Verma, R., Aravind, L., Oania, R., McDonald, W. H.,
Yates, J. R., Koonin, E. V., and Deshaies, R. J. (2002)
Role of Rpn11 metalloprotease in deubiquitination and
degradation by the 26S proteasome, Science, 298, 611615.
106. Blickwedehl, J., McEvoy, S., Wong, I., Kousis, P.,
Clements, J., Elliott, R., Cresswell, P., Liang, P., and
Bangia, N. (2007) Proteasomes and proteasome activator
200 kDa (PA200) accumulate on chromatin in response to
ionizing radiation, Radiat. Res., 167, 663674.
107. SadreBazzaz, K., Whitby, F. G., Robinson, H., Formosa,
T., and Hill, C. P. (2010) Structure of a Blm10 complex
reveals common mechanisms for proteasome binding and
gate opening, Mol. Cell, 37, 728735.
108. Khor, B., Bredemeyer, A. L., Huang, C.Y., Turnbull, I.
R., Evans, R., Maggi, L. B., White, J. M., Walker, L. M.,
Carnes, K., Hess, R. A., and Sleckman, B. P. (2006)
Proteasome activator PA200 is required for normal sper
matogenesis, Mol. Cell Biol., 26, 29993007.
109. Blickwedehl, J., Agarwal, M., Seong, C., Pandita, R. K.,
Melendy, T., Sung, P., Pandita, T. K., and Bangia, N.
(2008) Role for proteasome activator PA200 and postglu
tamyl proteasome activity in genomic stability, Proc. Natl.
Acad. Sci. USA,105, 1616516170.
110. Qian, M.X., Pang, Y., Liu, C. H., Haratake, K., Du, B.
Y., Ji, D.Y., Wang, G.F., Zhu, Q.Q., Song, W., Yu, Y.,
Zhang, X.X., Huang, H.T., Miao, S., Chen, L.B.,
Zhang, Z.H., Liang, Y.N., Liu, S., Cha, H., Yang, D.,
Zhai, Y., Komatsu, T., Tsuruta, F., Li, H., Cao, C., Li, W.,
Li, G.H., Cheng, Y., Chiba, T., Wang, L., Goldberg, A.
L., Shen, Y., and Qiu, X.B. (2013) Acetylationmediated
proteasomal degradation of core histones during DNA
repair and spermatogenesis, Cell, 153, 10121024.
111. Huang, L., Haratake, K., Miyahara, H., and Chiba, T.
(2016) Proteasome activators, PA28γand PA200, play
indispensable roles in male fertility, Sci. Rep., 6, 23171.
112. Tar, K., Dange, T., Yang, C., Yao, Y., Bulteau, A.L.,
Salcedo, E. F., Braigen, S., Bouillaud, F., Finley, D., and
Schmidt, M. (2014) Proteasomes associated with the
Blm10 activator protein antagonize mitochondrial fission
through degradation of the fission protein Dnm1, J. Biol.
Chem.,289, 1214512156.
113. Lopez, A. D., Tar, K., Krugel, U., Dange, T., Ros, I. G.,
and Schmidt, M. (2011) Proteasomal degradation of Sfp1
contributes to the repression of ribosome biogenesis during
starvation and is mediated by the proteasome activator
Blm10, Mol. Biol. Cell, 22, 528540.
114. Zaiss, D. M., Standera, S., Holzhutter, H., Kloetzel, P.,
and Sijts, A. J. (1999) The proteasome inhibitor PI31 com
petes with PA28 for binding to 20S proteasomes, FEBS
Lett.,457, 333338.
115. Zaiss, D. M. W., Standera, S., Kloetzel, P.M., and Sijts,
A. J. A. M. (2002) PI31 is a modulator of proteasome for
mation and antigen processing, Proc. Natl. Acad. Sci. USA,
99, 1434414349.
116. De La MotaPeynado, A., Lee, S. Y.C., Pierce, B. M.,
Wani, P., Singh, C. R., and Roelofs, J. (2013) The protea
someassociated protein Ecm29 inhibits proteasomal
ATPase activity and in vivo protein degradation by the pro
teasome, J. Biol. Chem.,288, 2946729481.
117. Lehmann, A., Niewienda, A., Jechow, K., Janek, K., and
Enenkel, C. (2010) Ecm29 fulfils quality control functions
in proteasome assembly, Mol. Cell, 38, 879888.
118. Park, S., Kim, W., Tian, G., Gygi, S. P., and Finley, D.
(2011) Structural defects in the regulatory particle–
core particle interface of the proteasome induce a novel
proteasome stress response, J. Biol. Chem.,286, 36652
36666.
119. Marshall, R. S., Li, F., Gemperline, D. C., Book, A. J.,
and Vierstra, R. D. (2015) Autophagic degradation of the
26S proteasome is mediated by the dual ATG8/ubiquitin
receptor RPN10 in arabidopsis, Mol. Cell, 58, 1053
1066.
120. Marshall, R. S., McLoughlin, F., and Vierstra, R. D.
(2016) Autophagic turnover of inactive 26S proteasomes in
yeast is directed by the ubiquitin receptor Cue5 and the
Hsp42 chaperone, Cell Rep., 16, 17171732.
121. Waite, K. A., MotaPeynado, A. D.L., Vontz, G., and
Roelofs, J. (2016) Starvation induces proteasome
autophagy with different pathways for core and regulatory
particles, J. Biol. Chem.,291, 32393253.
122. Prakash, S., Tian, L., Ratliff, K. S., Lehotzky, R. E., and
Matouschek, A. (2004) An unstructured initiation site is
required for efficient proteasomemediated degradation,
Nat. Struct. Mol. Biol.,11, 830837.
123. Schulman, B. A., and Wade Harper, J. (2009) Ubiquitin
like protein activation by E1 enzymes: the apex for down
stream signalling pathways, Nat. Rev. Mol. Cell Biol., 10,
319331.
124. Pierce, N. W., Kleiger, G., Shan, S., and Deshaies, R. J.
(2009) Detection of sequential polyubiquitylation on a
millisecond timescale, Nature,462, 615619.
125. Ciechanover, A., and Stanhill, A. (2014) The complexity of
recognition of ubiquitinated substrates by the 26S protea
some, Biochim. Biophys. Acta,1843, 8696.
S186 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
126. Herhaus, L., and Dikic, I. (2015) Expanding the ubiquitin
code through posttranslational modification, EMBO
Rep., 16, 10711083.
127. Komander, D., Clague, M. J., and Urbe, S. (2009)
Breaking the chains: structure and function of the deubiq
uitinases, Nat. Rev. Mol. Cell Biol., 10, 550563.
128. Deshaies, R. J., and Joazeiro, C. A. P. (2009) RING
domain E3 ubiquitin ligases, Annu. Rev. Biochem.,78, 399
434.
129. Rape, M., Reddy, S. K., and Kirschner, M. W. (2006) The
processivity of multiubiquitination by the APC determines
the order of substrate degradation, Cell, 124, 89103.
130. Kim, W., Bennett, E. J., Huttlin, E. L., Guo, A., Li, J.,
Possemato, A., Sowa, M. E., Rad, R., Rush, J., Comb, M.
J., Harper, J. W., and Gygi, S. P. (2011) Systematic and
quantitative assessment of the ubiquitinmodified pro
teome, Mol. Cell, 44, 325340.
131. Saeki, Y., Kudo, T., Sone, T., Kikuchi, Y., Yokosawa, H.,
Tohe, A., and Tanaka, K. (2009) Lysine 63linked polyu
biquitin chain may serve as a targeting signal for the 26S
proteasome, EMBO J.,28, 359371.
132. Shabek, N., HermanBachinsky, Y., Buchsbaum, S.,
Lewinson, O., HajYahya, M., Hejjaoui, M., Lashuel, H.
A., Sommer, T., Brik, A., and Ciechanover, A. (2012) The
size of the proteasomal substrate determines whether its
degradation will be mediated by mono or polyubiquityla
tion, Mol. Cell, 48, 8797.
133. Nathan, J. A., Kim, H. T., Ting, L., Gygi, S. P., and
Goldberg, A. L. (2013) Why do cellular proteins linked to
K63polyubiquitin chains not associate with proteasomes?
EMBO J.,32, 552565.
134. Ordureau, A., Munch, C., and Harper, J. W. (2015)
Quantifying ubiquitin signaling, Mol. Cell, 58, 660676.
135. Kaiser, S. E., Riley, B. E., Shaler, T. A., Trevino, R. S.,
Becker, C. H., Schulman, H., and Kopito, R. R. (2011)
Protein standard absolute quantification (PSAQ) method
for the measurement of cellular ubiquitin pools, Nat.
Methods., 8, 691696.
136. Xu, P., Duong, D. M., Seyfried, N. T., Cheng, D., Xie, Y.,
Robert, J., Rush, J., Hochstrasser, M., Finley, D., and
Peng, J. (2009) Quantitative proteomics reveals the func
tion of unconventional ubiquitin chains in proteasomal
degradation, Cell, 137, 133145.
137. Lu, Y., Lee, B.H., King, R. W., Finley, D., and Kirschner,
M. W. (2015) Substrate degradation by the proteasome: a
singlemolecule kinetic analysis, Science, 348, 1250834.
138. Lu, Y., Wang, W., and Kirschner, M. W. (2015) Specificity
of the anaphasepromoting complex: a singlemolecule
study, Science,348, 1248737.
139. Kirkpatrick, D. S., Hathaway, N. A., Hanna, J., Elsasser,
S., Rush, J., Finley, D., King, R. W., and Gygi, S. P. (2006)
Quantitative analysis of in vitro ubiquitinated cyclin B1
reveals complex chain topology, Nat. Cell Biol., 8, 700
710.
140. MartinezFonts, K., and Matouschek, A. (2016) A rapid
and versatile method for generating proteins with defined
ubiquitin chains, Biochemistry,55, 18981908.
141. Flick, K., Ouni, I., Wohlschlegel, J. A., Capati, C.,
McDonald, W. H., Yates, J. R., and Kaiser, P. (2004)
Proteolysisindependent regulation of the transcription
factor Met4 by a single Lys 48linked ubiquitin chain, Nat.
Cell Biol., 6, 634641.
142. Fishbain, S., Inobe, T., Israeli, E., Chavali, S., Yu, H.,
Kago, G., Babu, M. M., and Matouschek, A. (2015)
Sequence composition of disordered regions finetunes
protein halflife, Nat. Struct. Mol. Biol.,22, 214221.
143. Meyer, H.J., and Rape, M. (2014) Enhanced protein
degradation by branched ubiquitin chains, Cell, 157, 910
921.
144. Grice, G. L., Lobb, I. T., Weekes, M. P., Gygi, S. P.,
Antrobus, R., and Nathan, J. A. (2015) The proteasome
distinguishes between heterotypic and homotypic lysine
11linked polyubiquitin chains, Cell Rep., 12, 545553.
145. Boname, J. M., Thomas, M., Stagg, H. R., Xu, P., Peng,
J., and Lehner, P. J. (2010) Efficient internalization of
MHC I requires lysine11 and lysine63 mixed linkage
polyubiquitin chains, Traffic, 11, 210220.
146. Dammer, E. B., Na, C. H., Xu, P., Seyfried, N. T., Duong,
D. M., Cheng, D., Gearing, M., Rees, H., Lah, J. J.,
Levey, A. I., Rush, J., and Peng, J. (2011) Polyubiquitin
linkage profiles in three models of proteolytic stress suggest
the etiology of alzheimer disease, J. Biol. Chem.,286,
1045710465.
147. Ohtake, F., Saeki, Y., Ishido, S., Kanno, J., and Tanaka,
K. (2016) The K48K63 branched ubiquitin chain regu
lates NFκB signaling, Mol. Cell, 64, 251266.
148. Yau, R. G., Doerner, K., Castellanos, E. R., Haakonsen, D.
L., Werner, A., Wang, N., Yang, X. W., MartinezMartin,
N., Matsumoto, M. L., Dixit, V. M., and Rape, M. (2017)
Assembly and function of heterotypic ubiquitin chains in
cellcycle and protein quality control, Cell, 171, 918933.
149. Saeki, Y., Isono, E., and TohE, A. (2005) Preparation of
ubiquitinated substrates by the PY motifinsertion method
for monitoring 26S proteasome activity, Methods Enzymol.,
399, 215227.
150. Thrower, J. S., Hoffman, L., Rechsteiner, M., and Pickart,
C. M. (2000) Recognition of the polyubiquitin proteolytic
signal, EMBO J.,19, 94102.
151. Stieglitz, B., Rana, R. R., Koliopoulos, M. G., Morris
Davies, A. C., Schaeffer, V., Christodoulou, E., Howell,
S., Brown, N. R., Dikic, I., and Rittinger, K. (2013)
Structural basis for ligasespecific conjugation of linear
ubiquitin chains by HOIP, Nature,503, 422426.
152. Ikeda, F., Deribe, Y. L., Skanland, S. S., Stieglitz, B.,
Grabbe, C., FranzWachtel, M., van Wijk, S. J. L.,
Goswami, P., Nagy, V., Terzic, J., Tokunaga, F.,
Androulidaki, A., Nakagawa, T., Pasparakis, M., Iwai, K.,
Sundberg, J. P., Schaefer, L., Rittinger, K., Macek, B., and
Dikic, I. (2011) SHARPIN forms a linear ubiquitin ligase
complex regulating NFκB activity and apoptosis, Nature,
471, 637641.
153. Gerlach, B., Cordier, S. M., Schmukle, A. C., Emmerich,
C. H., Rieser, E., Haas, T. L., Webb, A. I., Rickard, J. A.,
Anderton, H., Wong, W. W.L., Nachbur, U., Gangoda,
L., Warnken, U., Purcell, A. W., Silke, J., and Walczak, H.
(2011) Linear ubiquitination prevents inflammation and
regulates immune signalling, Nature,471, 591596.
154. Tokunaga, F., Nakagawa, T., Nakahara, M., Saeki, Y.,
Taniguchi, M., Sakata, S., Tanaka, K., Nakano, H., and
Iwai, K. (2011) SHARPIN is a component of the NFκB
activating linear ubiquitin chain assembly complex,
Nature,471, 633636.
155. Rivkin, E., Almeida, S. M., Ceccarelli, D. F., Juang, Y.
C., MacLean, T. A., Srikumar, T., Huang, H., Dunham,
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S187
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
W. H., Fukumura, R., Xie, G., Gondo, Y., Raught, B.,
Gingras, A.C., Sicheri, F., and Cordes, S. P. (2013) The
linear ubiquitinspecific deubiquitinase gumby regulates
angiogenesis, Nature,498, 318324.
156. Kirisako, T., Kamei, K., Murata, S., Kato, M., Fukumoto,
H., Kanie, M., Sano, S., Tokunaga, F., Tanaka, K., and
Iwai, K. (2006) A ubiquitin ligase complex assembles lin
ear polyubiquitin chains, EMBO J.,25, 48774887.
157. Zhao, S., and Ulrich, H. D. (2010) Distinct consequences
of posttranslational modification by linear versus K63
linked polyubiquitin chains, Proc. Natl. Acad. Sci. USA,
107, 77047709.
158. Inn, K.S., Gack, M. U., Tokunaga, F., Shi, M., Wong,
L.Y., Iwai, K., and Jung, J. U. (2011) Linear ubiquitin
assembly complex negatively regulates RIGI and
TRIM25mediated type I interferon induction, Mol. Cell,
41, 354365.
159. Kliza, K., Taumer, C., Pinzuti, I., FranzWachtel, M.,
Kunzelmann, S., Stieglitz, B., Macek, B., and Husnjak,
K. (2017) Internally tagged ubiquitin: a tool to identify lin
ear polyubiquitinmodified proteins by mass spectrometry,
Nat. Methods, 14, 504512.
160. KravtsovaIvantsiv, Y., Cohen, S., and Ciechanover, A.
(2009) Modification by single ubiquitin moieties rather than
polyubiquitination is sufficient for proteasomal processing
of the p105 NFkappaB precursor, Mol. Cell, 33, 496504.
161. Braten, O., Livneh, I., Ziv, T., Admon, A., Kehat, I.,
Caspi, L. H., Gonen, H., Bercovich, B., Godzik, A.,
Jahandideh, S., Jaroszewski, L., Sommer, T., Kwon, Y. T.,
Guharoy, M., Tompa, P., and Ciechanover, A. (2016)
Numerous proteins with unique characteristics are degrad
ed by the 26S proteasome following monoubiquitination,
Proc. Natl. Acad. Sci. USA,113, 46394647.
162. Oshikawa, K., Matsumoto, M., Oyamada, K., and
Nakayama, K. I. (2012) Proteomewide identification of
ubiquitylation sites by conjugation of engineered lysine
less ubiquitin, J. Proteome Res., 11, 796807.
163. Swatek, K. N., and Komander, D. (2016) Ubiquitin mod
ifications, Nat. Publ. Gr.,26, 399422.
164. Lee, B.H., Lee, M. J., Park, S., Oh, D.C., Elsasser, S.,
Chen, P.C., Gartner, C., Dimova, N., Hanna, J., Gygi, S.
P., Wilson, S. M., King, R. W., and Finley, D. (2010)
Enhancement of proteasome activity by a smallmolecule
inhibitor of USP14, Nature,467, 179184.
165. D’Arcy, P., Brnjic, S., Olofsson, M. H., Fryknas, M.,
Lindsten, K., De Cesare, M., Perego, P., Sadeghi, B.,
Hassan, M., Larsson, R., and Linder, S. (2011) Inhibition
of proteasome deubiquitinating activity as a new cancer
therapy, Nat. Med., 17, 16361640.
166. Kisselev, A. F., van der Linden, W. A., and Overkleeft, H.
S. (2012) Proteasome inhibitors: an expanding army
attacking a unique target, Chem. Biol.,19, 99115.
167. McCullough, J., Clague, M. J., and Urbe, S. (2004)
AMSH is an endosomeassociated ubiquitin isopeptidase,
J. Cell Biol., 166, 487492.
168. Ambroggio, X. I., Rees, D. C., and Deshaies, R. J. (2004)
JAMM: a metalloproteaselike zinc site in the proteasome
and signalosome, PLoS Biol., 2, 113119.
169. Cope, G. A., Suh, G. S. B., Aravind, L., Schwarz, S. E.,
Zipursky, S. L., Koonin, E. V., and Deshaies, R. J. (2002)
Role of predicted metalloprotease motif of Jab1/Csn5 in
cleavage of Nedd8 from Cul1, Science, 298, 608611.
170. Shrestha, R. K., Ronau, J. A., Davies, C. W., Guenette, R.
G., Strieter, E. R., Paul, L. N., and Das, C. (2014) Insights
into the mechanism of deubiquitination by JAMM deubiq
uitinases from cocrystal structures of the enzyme with the
substrate and product, Biochemistry,53, 31993217.
171. Rinaldi, T., Pick, E., Gambadoro, A., Zilli, S., Maytal
Kivity, V., Frontali, L., and Glickman, M. H. (2004)
Participation of the proteasomal lid subunit Rpn11 in
mitochondrial morphology and function is mapped to a
distinct Cterminal domain, Biochem. J., 381, 275285.
172. Worden, E. J., Padovani, C., and Martin, A. (2014)
Structure of the Rpn11–Rpn8 dimer reveals mechanisms
of substrate deubiquitination during proteasomal degrada
tion, Nat. Struct. Mol. Biol.,21, 220227.
173. Lee, B.H., Lu, Y., Prado, M. A., Shi, Y., Tian, G., Sun,
S., Elsasser, S., Gygi, S. P., King, R. W., and Finley, D.
(2016) USP14 deubiquitinates proteasomebound sub
strates that are ubiquitinated at multiple sites, Nature,532,
398401.
174. Worden, E. J., Padovani, C., and Martin, A. (2014)
Structure of the Rpn11–Rpn8 dimer reveals mechanisms
of substrate deubiquitination during proteasomal degrada
tion, Nat. Struct. Mol. Biol.,21, 220227.
175. Pathare, G. R., Nagy, I., Sledz, P., Anderson, D. J., Zhou,
H.J., Pardon, E., Steyaert, J., Forster, F., Bracher, A., and
Baumeister, W. (2014) Crystal structure of the proteasomal
deubiquitylation module Rpn8–Rpn11, Proc. Natl. Acad.
Sci. USA,111, 29842989.
176. Dambacher, C. M., Worden, E. J., Herzik, M. A., Martin,
A., and Lander, G. C. (2016) Atomic structure of the 26S
proteasome lid reveals the mechanism of deubiquitinase
inhibition, Elife, 5, e13027.
177. Worden, E. J., Dong, K. C., and Martin, A. (2017) An
AAA motordriven mechanical switch in Rpn11 controls
deubiquitination at the 26S proteasome, Mol. Cell, 67,
799811.
178. Finley, D. (2009) Recognition and processing of ubiqui
tin–protein conjugates by the proteasome, Annu. Rev.
Biochem.,78, 477513.
179. Borodovsky, A., Kessler, B. M., Casagrande, R.,
Overkleeft, H. S., Wilkinson, K. D., and Ploegh, H. L.
(2001) A novel active sitedirected probe specific for deu
biquitylating enzymes reveals proteasome association of
USP14, EMBO J.,20, 51875196.
180. Hanna, J., Hathaway, N. A., Tone, Y., Crosas, B., Elsasser,
S., Kirkpatrick, D. S., Leggett, D. S., Gygi, S. P., King, R.
W., and Finley, D. (2006) Deubiquitinating enzyme Ubp6
functions noncatalytically to delay proteasomal degrada
tion, Cell, 127, 99111.
181. Guterman, A., and Glickman, M. H. (2004)
Complementary roles for Rpn11 and Ubp6 in deubiquiti
nation and proteolysis by the proteasome, J. Biol. Chem.,
279, 17291738.
182. Leggett, D. S., Hanna, J., Borodovsky, A., Crosas, B.,
Schmidt, M., Baker, R. T., Walz, T., Ploegh, H., and
Finley, D. (2002) Multiple associated proteins regulate
proteasome structure and function, Mol. Cell, 10, 495
507.
183. Kim, H. T., and Goldberg, A. L. (2017) The deubiquiti
nating enzyme Usp14 allosterically inhibits multiple pro
teasomal activities and ubiquitinindependent proteolysis,
J. Biol. Chem.,292, 98309839.
S188 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
184. Bashore, C., Dambacher, C. M., Goodall, E. A.,
Matyskiela, M. E., Lander, G. C., and Martin, A. (2015)
Ubp6 deubiquitinase controls conformational dynamics
and substrate degradation of the 26S proteasome, Nat.
Struct. Mol. Biol.,22, 712719.
185. Aufderheide, A., Beck, F., Stengel, F., Hartwig, M.,
Schweitzer, A., Pfeifer, G., Goldberg, A. L., Sakata, E.,
Baumeister, W., and Forster, F. (2015) Structural charac
terization of the interaction of Ubp6 with the 26S protea
some, Proc. Natl. Acad. Sci. USA,112, 86268631.
186. Mansour, W., Nakasone, M. A., von Delbruck, M., Yu,
Z., Krutauz, D., Reis, N., Kleifeld, O., Sommer, T.,
Fushman, D., and Glickman, M. H. (2015) Disassembly
of Lys11 and mixed linkage polyubiquitin conjugates
provides insights into function of proteasomal deubiqui
tinases Rpn11 and Ubp6, J. Biol. Chem.,290, 4688
4704.
187. Burgie, S. E., Bingman, C. A., Soni, A. B., and Phillips,
G. N., Jr. (2012) Structural characterization of human
Uch37, Proteins, 80, 649654.
188. VanderLinden, R. T., Hemmis, C. W., Schmitt, B., Ndoja,
A., Whitby, F. G., Robinson, H., Cohen, R. E., Yao, T.,
and Hill, C. P. (2015) Structural basis for the activation
and inhibition of the UCH37 deubiquitylase, Mol. Cell, 57,
901911.
189. Sahtoe, D. D., van Dijk, W. J., El Oualid, F., Ekkebus, R.,
Ovaa, H., and Sixma, T. K. (2015) Mechanism of UCH
L5 activation and inhibition by DEUBAD domains in
RPN13 and INO80G, Mol. Cell, 57, 887900.
190. Chen, X., Lee, B.H., Finley, D., and Walters, K. J. (2010)
Structure of proteasome ubiquitin receptor hRpn13 and its
activation by the scaffolding protein hRpn2, Mol. Cell, 38,
404415.
191. Lam, Y. A., Xu, W., DeMartino, G. N., and Cohen, R. E.
(1997) Editing of ubiquitin conjugates by an isopeptidase
in the 26S proteasome, Nature,385, 737740.
192. Zhang, N.Y., Jacobson, A. D., MacFadden, A., and Liu,
C.W. (2011) Ubiquitin chain trimming recycles the sub
strate binding sites of the 26S proteasome and promotes
degradation of lysine 48linked polyubiquitin conjugates,
J. Biol. Chem.,286, 2554025546.
193. Chen, S., Wu, J., Lu, Y., Ma, Y.B., Lee, B.H., Yu, Z.,
Ouyang, Q., Finley, D. J., Kirschner, M. W., and Mao, Y.
(2016) Structural basis for dynamic regulation of the
human 26S proteasome, Proc. Natl. Acad. Sci. USA,113,
1299112996.
194. Peth, A., Kukushkin, N., Bosse, M., and Goldberg, A. L.
(2013) Ubiquitinated proteins activate the proteasomal
ATPases by binding to Usp14 or Uch37 homologs, J. Biol.
Chem.,288, 77817790.
195. Swaminathan, S., Amerik, A. Y., and Hochstrasser, M.
(1999) The Doa4 deubiquitinating enzyme is required for
ubiquitin homeostasis in yeast, Mol. Biol. Cell, 10, 2583
2594.
196. Varshavsky, A. (2012) The ubiquitin system, an immense
realm, Annu. Rev. Biochem.,81, 167176.
197. Crosas, B., Hanna, J., Kirkpatrick, D. S., Zhang, D. P.,
Tone, Y., Hathaway, N. A., Buecker, C., Leggett, D. S.,
Schmidt, M., King, R. W., Gygi, S. P. P., and Finley, D.
(2006) Ubiquitin chains are remodeled at the proteasome
by opposing ubiquitin ligase and deubiquitinating activi
ties, Cell, 127, 14011413.
198. Tian, L., Holmgren, R. A., and Matouschek, A. (2005) A
conserved processing mechanism regulates the activity of
transcription factors Cubitus interruptus and NFκB, Nat.
Struct. Mol. Biol.,12, 10451053.
199. Aviram, S., and Kornitzer, D. (2010) The ubiquitin ligase
Hul5 promotes proteasomal processivity, Mol. Cell Biol.,
30, 985994.
200. Cappadocia, L., and Lima, C. D. (2018) Ubiquitinlike
protein conjugation: structures, chemistry, and mecha
nism, Chem. Rev., 118, 889918.
201. Jentsch, S., and Pyrowolakis, G. (2000) Ubiquitin and its
kin: how close are the family ties? Trends Cell Biol., 10,
335342.
202. Luders, J., Pyrowolakis, G., and Jentsch, S. (2003) The
ubiquitinlike protein HUB1 forms SDSresistant com
plexes with cellular proteins in the absence of ATP, EMBO
Rep., 4, 11691174.
203. Novatchkova, M., Bachmair, A., Eisenhaber, B., and
Eisenhaber, F. (2005) Proteins with two SUMOlike
domains in chromatinassociated complexes: the RENi
(Rad60Esc2NIP45) family, BMC Bioinformatics, 6,
22.
204. Eifler, K., and Vertegaal, A. C. O. (2015) SUMOylation
mediated regulation of cell cycle progression and cancer,
Trends Biochem. Sci., 40, 779793.
205. Flotho, A., and Melchior, F. (2013) Sumoylation: a regu
latory protein modification in health and disease, Annu.
Rev. Biochem.,82, 357385.
206. Schmidtke, G., Aichem, A., and Groettrup, M. (2014)
FAT10ylation as a signal for proteasomal degradation,
Biochim. Biophys. Acta,1843, 97102.
207. Aichem, A., Pelzer, C., Lukasiak, S., Kalveram, B.,
Sheppard, P. W., Rani, N., Schmidtke, G., and Groettrup,
M. (2010) USE1 is a bispecific conjugating enzyme for
ubiquitin and FAT10, which FAT10ylates itself in cis, Nat.
Commun., 1, 110.
208. Pelzer, C., and Groettrup, M. (2010) FAT10: activated by
UBA6 and functioning in protein degradation, Subcell
Biochem., 54, 238246.
209. Canaan, A., DeFuria, J., Perelman, E., Schultz, V., Seay,
M., Tuck, D., Flavell, R. A., Snyder, M. P., Obin, M. S.,
and Weissman, S. M. (2014) Extended lifespan and
reduced adiposity in mice lacking the FAT10 gene, Proc.
Natl. Acad. Sci. USA,111, 53135318.
210. Canaan, A., Yu, X., Booth, C. J., Lian, J., Lazar, I.,
Gamfi, S. L., Castille, K., Kohya, N., Nakayama, Y., Liu,
Y.C., Eynon, E., Flavell, R., and Weissman, S. M. (2006)
FAT10/diubiquitinlike proteindeficient mice exhibit
minimal phenotypic differences, Mol. Cell Biol.,26, 5180
5189.
211. Buchsbaum, S., Bercovich, B., and Ciechanover, A. (2012)
FAT10 is a proteasomal degradation signal that is itself reg
ulated by ubiquitination, Mol. Biol. Cell, 23, 225232.
212. Fu, H., Sadis, S., Rubin, D. M., Glickman, M., van
Nocker, S., Finley, D., and Vierstra, R. D. (1998)
Multiubiquitin chain binding and protein degradation are
mediated by distinct domains within the 26S proteasome
subunit Mcb1, J. Biol. Chem.,273, 19701981.
213. Lam, Y. A., Lawson, T. G., Velayutham, M., Zweier, J. L.,
and Pickart, C. M. (2002) A proteasomal ATPase subunit
recognizes the polyubiquitin degradation signal, Nature,
416, 763767.
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S189
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
214. Paraskevopoulos, K., Kriegenburg, F., Tatham, M. H.,
Rosner, H. I., Medina, B., Larsen, I. B., Brandstrup, R.,
Hardwick, K. G., Hay, R. T., Kragelund, B. B.,
HartmannPetersen, R., and Gordon, C. (2014) Dss1 is a
26S proteasome ubiquitin receptor, Mol. Cell, 56, 453461.
215. Deveraux, Q., Ustrell, V., Pickart, C., and Rechsteiner, M.
(1994) A 26S protease subunit that binds ubiquitin conju
gates, J. Biol. Chem.,269, 70597061.
216. Hamazaki, J., Hirayama, S., and Murata, S. (2015)
Redundant roles of Rpn10 and Rpn13 in recognition of
ubiquitinated proteins and cellular homeostasis, PLOS
Genet., 11, e1005401.
217. Matiuhin, Y., Kirkpatrick, D. S., Ziv, I., Kim, W.,
Dakshinamurthy, A., Kleifeld, O., Gygi, S. P., Reis, N.,
and Glickman, M. H. (2008) Extraproteasomal Rpn10
restricts access of the polyubiquitinbinding protein Dsk2
to proteasome, Mol. Cell, 32, 415425.
218. KerenKaplan, T., Zeev Peters, L., LevinKravets, O.,
Attali, I., Kleifeld, O., Shohat, N., Artzi, S., Zucker, O.,
Pilzer, I., Reis, N., Glickman, M. H., BenAroya, S., and
Prag, G. (2016) Structure of ubiquitylatedRpn10 provides
insight into its autoregulation mechanism, Nat. Commun.,
7, 12960.
219. Kang, Y., Vossler, R. A., DiazMartinez, L. A., Winter, N.
S., Clarke, D. J., and Walters, K. J. (2006) UBL/UBA
ubiquitin receptor proteins bind a common tetraubiquitin
chain, J. Mol. Biol.,356, 10271035.
220. Woelk, T., Oldrini, B., Maspero, E., Confalonieri, S.,
Cavallaro, E., Di Fiore, P. P., and Polo, S. (2006)
Molecular mechanisms of coupled monoubiquitination,
Nat. Cell Biol., 8, 12461254.
221. Verma, R., Oania, R., Graumann, J., and Deshaies, R. J.
(2004) Multiubiquitin chain receptors define a layer of
substrate selectivity in the ubiquitin–proteasome system,
Cell, 118, 99110.
222. Wang, Q., Young, P., and Walters, K. J. (2005) Structure of
S5a bound to monoubiquitin provides a model for polyu
biquitin recognition, J. Mol. Biol.,348, 727739.
223. Van Nocker, S., Sadis, S., Rubin, D. M., Glickman, M.,
Fu, H., Coux, O., Wefes, I., Finley, D., and Vierstra, R. D.
(1996) The multiubiquitinchainbinding protein Mcb1 is
a component of the 26S proteasome in Saccharomyces
cerevisiae and plays a nonessential, substratespecific role
in protein turnover, Mol. Cell. Biol., 16, 60206028.
224. Hamazaki, J., Sasaki, K., Kawahara, H., Hisanaga, S.I.,
Tanaka, K., and Murata, S. (2007) Rpn10mediated
degradation of ubiquitinated proteins is essential for mouse
development, Mol. Cell. Biol.,27, 66296638.
225. AlShami, A., Jhaver, K. G., Vogel, P., Wilkins, C.,
Humphries, J., Davis, J. J., Xu, N., Potter, D. G.,
Gerhardt, B., Mullinax, R., Shirley, C. R., Anderson, S.
J., and Oravecz, T. (2010) Regulators of the proteasome
pathway, Uch37 and Rpn13, play distinct roles in mouse
development, PLoS One, 5, e13654.
226. Simins, A. B., Weighardt, H., Weidner, K. M., Weidle, U.
H., and Holzmann, B. (1999) Functional cloning of
ARM1, an adhesionregulating molecule upregulated in
metastatic tumor cells, Clin. Exp. Metastasis, 17, 641648.
227. Hamazaki, J., Iemura, S., Natsume, T., Yashiroda, H.,
Tanaka, K., and Murata, S. (2006) A novel proteasome
interacting protein recruits the deubiquitinating enzyme
UCH37 to 26S proteasomes, EMBO J.,25, 45244536.
228. Yao, T., Song, L., Xu, W., DeMartino, G. N., Florens, L.,
Swanson, S. K., Washburn, M. P., Conaway, R. C.,
Conaway, J. W., and Cohen, R. E. (2006) Proteasome
recruitment and activation of the Uch37 deubiquitinating
enzyme by Adrm1, Nat. Cell Biol., 8, 9941002.
229. Mazumdar, T., Gorgun, F. M., Sha, Y., Tyryshkin, A.,
Zeng, S., HartmannPetersen, R., Jorgensen, J. P.,
Hendil, K. B., and Eissa, N. T. (2010) Regulation of NF
kappaB activity and inducible nitric oxide synthase by reg
ulatory particle nonATPase subunit 13 (Rpn13), Proc.
Natl. Acad. Sci. USA,107, 1385413859.
230. Huang, Y., and Ratovitski, E. A. (2010) Phospho
ΔNp63α/Rpn13dependent regulation of LKB1 degrada
tion modulates autophagy in cancer cells, Aging (Albany.
NY), 2, 959968.
231. Sabio, G., Das, M., Mora, A., Zhang, Z., Jun, J. Y., Ko,
H. J., Barrett, T., Kim, J. K., and Davis, R. J. (2008) A
stress signaling pathway in adipose tissue regulates hepatic
insulin resistance, Science, 322, 15391543.
232. Aguileta, M. A., Korac, J., Durcan, T. M., Trempe, J.F.,
Haber, M., Gehring, K., Elsasser, S., Waidmann, O., Fon,
E. A., and Husnjak, K. (2015) The E3 ubiquitin ligase
parkin is recruited to the 26S proteasome via the proteaso
mal ubiquitin receptor Rpn13, J. Biol. Chem.,290, 7492
7505.
233. Miao, X., Miao, X., Wen, Y., Hu, J., Dai, W., and Yin, B.
(2012) A possible connection between adhesion regulating
molecule 1 overexpression and nuclear factor kappa B
activity in hepatocarcinogenesis, Oncol. Rep., 28, 283290.
234. Song, Y., Ray, A., Li, S., Das, D. S., Tai, Y. T., Carrasco,
R. D., Chauhan, D., and Anderson, K. C. (2016)
Targeting proteasome ubiquitin receptor Rpn13 in multi
ple myeloma, Leukemia, 30, 18771886.
235. Fejzo, M., Anderson, L., von Euw, E., Kalous, O.,
Avliyakulov, N., Haykinson, M., Konecny, G., Finn, R.,
and Slamon, D. (2013) Amplification target ADRM1: role
as an oncogene and therapeutic target for ovarian cancer,
Int. J. Mol. Sci., 14, 30943109.
236. Carvalho, B., Postma, C., Mongera, S., Hopmans, E.,
Diskin, S., van de Wiel, M. A., van Criekinge, W., Thas,
O., Matthai, A., Cuesta, M. A., Droste, J. S. T. S.,
Craanen, M., Schrock, E., Ylstra, B., and Meijer, G. A.
(2009) Multiple putative oncogenes at the chromosome
20q amplicon contribute to colorectal adenoma to carci
noma progression, Gut, 58, 7989.
237. Jang, S. H., Park, J. W., Kim, H. R., Seong, J. K., and
Kim, H. K. (2014) ADRM1 gene amplification is a candi
date driver for metastatic gastric cancers, Clin. Exp.
Metastasis, 31, 727733.
238. Anchoori, R. K., Karanam, B., Peng, S., Wang, J. W.,
Jiang, R., Tanno, T., Orlowski, R. Z., Matsui, W., Zhao,
M., Rudek, M. A., Hung, C., Chen, X., Walters, K. J., and
Roden, R. B. S. (2013) A bisbenzylidine piperidone tar
geting proteasome ubiquitin receptor RPN13/ADRM1 as
a therapy for cancer, Cancer Cell, 24, 791805.
239. Trader, D. J., Simanski, S., and Kodadek, T. (2015) A
reversible and highly selective inhibitor of the proteasomal
ubiquitin receptor Rpn13 is toxic to multiple myeloma
cells, J. Am. Chem. Soc., 137, 63126319.
240. Khan, M. L., and Stewart, A. K. (2011) Carfilzomib: a
novel secondgeneration proteasome inhibitor, Futur.
Oncol., 7, 607612.
S190 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
241. Chen, X., Randles, L., Shi, K., Tarasov, S. G., Aihara, H.,
and Walters, K. J. (2016) Structures of Rpn1T1:Rad23 and
hRpn13:hPLIC2 reveal distinct binding mechanisms
between substrate receptors and shuttle factors of the pro
teasome, Structure,24, 12571270.
242. Saeki, Y., Sone, T., Tohe, A., and Yokosawa, H. (2002)
Identification of ubiquitinlike proteinbinding subunits of
the 26S proteasome, Biochem. Biophys. Res. Commun.,
296, 813819.
243. Gomez, T. A., Kolawa, N., Gee, M., Sweredoski, M. J.,
and Deshaies, R. J. (2011) Identification of a functional
docking site in the Rpn1 LRR domain for the UBAUBL
domain protein Ddi1, BMC Biol.,9, 33.
244. Haapasalo, A., Viswanathan, J., Kurkinen, K. M.,
Bertram, L., Soininen, H., Dantuma, N. P., Tanzi, R. E.,
and Hiltunen, M. (2011) Involvement of ubiquilin1 tran
script variants in protein degradation and accumulation,
Commun. Integr. Biol., 4, 428432.
245. Raasi, S., Varadan, R., Fushman, D., and Pickart, C. M.
(2005) Diverse polyubiquitin interaction properties of
ubiquitinassociated domains, Nat. Struct. Mol. Biol.,12,
708714.
246. Richly, H., Rape, M., Braun, S., Rumpf, S., Hoege, C.,
and Jentsch, S. (2005) A series of ubiquitin binding factors
connects CDC48/p97 to substrate multiubiquitylation and
proteasomal targeting, Cell, 120, 7384.
247. Kim, I., Ahn, J., Liu, C., Tanabe, K., Apodaca, J., Suzuki,
T., and Rao, H. (2006) The Png1–Rad23 complex regu
lates glycoprotein turnover, J. Cell Biol., 172, 211219.
248. Heessen, S., Masucci, M. G., and Dantuma, N. P. (2005)
The UBA2 domain functions as an intrinsic stabilization
signal that protects rad23 from proteasomal degradation,
Mol. Cell, 18, 225235.
249. Heinen, C., Acs, K., Hoogstraten, D., and Dantuma, N.
P. (2011) Cterminal UBA domains protect ubiquitin
receptors by preventing initiation of protein degradation,
Nat. Commun.,2, 191.
250. Vadlamudi, R. K., Joung, I., Strominger, J. L., and Shin,
J. (1996) p62, a phosphotyrosineindependent ligand of
the SH2 domain of p56lck, belongs to a new class of
ubiquitinbinding proteins, J. Biol. Chem.,271, 20235
20237.
251. Seibenhener, M. L., Babu, J. R., Geetha, T., Wong, H. C.,
Krishna, N. R., and Wooten, M. W. (2004) Sequestosome
1/p62 is a polyubiquitin chain binding protein involved in
ubiquitin proteasome degradation, Mol. Cell. Biol.,24,
80558068.
252. Pankiv, S., Clausen, T. H., Lamark, T., Brech, A., Bruun,
J.A., Outzen, H., Overvatn, A., Bjorkoy, G., and
Johansen, T. (2007) p62/SQSTM1 binds directly to
Atg8/LC3 to facilitate degradation of ubiquitinated pro
tein aggregates by autophagy, J. Biol. Chem.,282, 24131
24145.
253. Johnson, J. O., Mandrioli, J., Benatar, M., Abramzon, Y.,
Van Deerlin, V. M., Trojanowski, J. Q., Gibbs, J. R.,
Brunetti, M., Gronka, S., Wuu, J., Ding, J., McCluskey,
L., MartinezLage, M., Falcone, D., Hernandez, D. G.,
Arepalli, S., Chong, S., Schymick, J. C., Rothstein, J.,
Landi, F., Wang, Y.D., Calvo, A., Mora, G., Sabatelli,
M., Monsurro, M. R., Battistini, S., Salvi, F., Spataro, R.,
Sola, P., Borghero, G., Galassi, G., Scholz, S. W., Taylor,
J. P., Restagno, G., Chio, A., Traynor, B. J., and Traynor,
B. J. (2010) Exome sequencing reveals VCP mutations as a
cause of familial ALS, Neuron, 68, 857864.
254. Madeo, F., Schlauer, J., Zischka, H., Mecke, D., and
Frohlich, K. U. (1998) Tyrosine phosphorylation regulates
cell cycledependent nuclear localization of Cdc48p, Mol.
Biol. Cell, 9, 131141.
255. Anderson, D. J., Le Moigne, R., Djakovic, S., Kumar, B.,
Rice, J., Wong, S., Wang, J., Yao, B., Valle, E., Kiss von
Soly, S., Madriaga, A., Soriano, F., Menon, M.K., Wu,
Z. Y., Kampmann, M., Chen, Y., Weissman, J. S., Aftab,
B. T., Yakes, F. M., Shawver, L., Zhou, H.J., Wustrow,
D., and Rolfe, M. (2015) Targeting the AAA ATPase p97
as an approach to treat cancer through disruption of pro
tein homeostasis, Cancer Cell, 28, 653665.
256. Ye, Y. (2006) Diverse functions with a common regulator:
ubiquitin takes command of an AAA ATPase, J. Struct.
Biol., 156, 2940.
257. Meyer, H., Bug, M., and Bremer, S. (2012) Emerging
functions of the VCP/p97 AAAATPase in the ubiquitin
system, Nat. Cell Biol., 14, 117123.
258. Beckwith, R., Estrin, E., Worden, E. J., and Martin, A.
(2013) Reconstitution of the 26S proteasome reveals func
tional asymmetries in its AAA+ unfoldase, Nat. Struct.
Mol. Biol.,20, 11641172.
259. Inobe, T., Fishbain, S., Prakash, S., and Matouschek, A.
(2011) Defining the geometry of the twocomponent pro
teasome degron, Nat. Chem. Biol.,7, 161167.
260. Babu, M. M., van der Lee, R., de Groot, N. S., and
Gsponer, J. (2011) Intrinsically disordered proteins: regu
lation and disease, Curr. Opin. Struct. Biol., 21, 432440.
261. Gsponer, J., Futschik, M. E., Teichmann, S. A., and Babu,
M. M. (2008) Tight regulation of unstructured proteins:
from transcript synthesis to protein degradation, Science,
322, 13651368.
262. Tompa, P., Prilusky, J., Silman, I., and Sussman, J. L.
(2008) Structural disorder serves as a weak signal for intra
cellular protein degradation, Proteins Struct. Funct.
Bioinformatics, 71, 903909.
263. Yen, H.C. S., Xu, Q., Chou, D. M., Zhao, Z., and
Elledge, S. J. (2008) Global protein stability profiling in
mammalian cells, Science, 322, 918923.
264. Yen, H.C. S., and Elledge, S. J. (2008) Identification of
SCF ubiquitin ligase substrates by global protein stability
profiling, Science,322, 923929.
265. Suskiewicz, M. J., Sussman, J. L., Silman, I., and Shaul,
Y. (2011) Contextdependent resistance to proteolysis of
intrinsically disordered proteins, Protein Sci., 20, 1285
1297.
266. Radivojac, P., Vacic, V., Haynes, C., Cocklin, R. R.,
Mohan, A., Heyen, J. W., Goebl, M. G., and Iakoucheva,
L. M. (2010) Identification, analysis, and prediction of
protein ubiquitination sites, Proteins Struct. Funct.
Bioinformatics, 78, 365380.
267. Hagai, T., Azia, A., TothPetroczy, A., and Levy, Y. (2011)
Intrinsic disorder in ubiquitination substrates, J. Mol.
Biol.,412, 319324.
268. Hagai, T., and Levy, Y. (2010) Ubiquitin not only serves as
a tag but also assists degradation by inducing protein
unfolding, Proc. Natl. Acad. Sci. USA,107, 20012006.
269. Prakash, S., Inobe, T., Hatch, A. J., and Matouschek, A.
(2009) Substrate selection by the proteasome during degra
dation of protein complexes, Nat. Chem. Biol.,5, 2936.
PROTEASOME: A NANOMACHINERY OF CREATIVE DESTRUCTION S191
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
270. Johnson, E. S., Gonda, D. K., and Varshavsky, A. (1990)
Cistrans recognition and subunitspecific degradation of
shortlived proteins, Nature,346, 287291.
271. Hochstrasser, M., and Varshavsky, A. (1990) In vivo degra
dation of a transcriptional regulator: the yeast alpha 2
repressor, Cell, 61, 697708.
272. Verma, R., McDonald, H., Yates, J. R., and Deshaies, R.
J. (2001) Selective degradation of ubiquitinated Sic1 by
purified 26S proteasome yields active S phase cyclinCdk,
Mol. Cell, 8, 439448.
273. Nasmyth, K., Shirayama, M., Toth, A., and Galova, M.
(1999) APC(Cdc20) promotes exit from mitosis by
destroying the anaphase inhibitor Pds1 and cyclin Clb5,
Nature,402, 203207.
274. Hoyt, M. A., Zich, J., Takeuchi, J., Zhang, M., Govaerts,
C., and Coffino, P. (2006) Glycine–alanine repeats impair
proper substrate unfolding by the proteasome, EMBO J.,
25, 17201729.
275. KravtsovaIvantsiv, Y., Shomer, I., CohenKaplan, V.,
Snijder, B., SupertiFurga, G., Gonen, H., Sommer, T.,
Ziv, T., Admon, A., Naroditsky, I., Jbara, M., Brik, A.,
Pikarsky, E., Kwon, Y. T., Doweck, I., and Ciechanover, A.
(2015) KPC1mediated ubiquitination and proteasomal
processing of NFκB1 p105 to p50 restricts tumor growth,
Cell, 161, 333347.
276. Baugh, J. M., Viktorova, E. G., and Pilipenko, E. V.
(2009) Proteasomes can degrade a significant proportion
of cellular proteins independent of ubiquitination, J. Mol.
Biol.,386, 814827.
277. Li, X., and Coffino, P. (1993) Degradation of ornithine
decarboxylase: exposure of the Cterminal target by a
polyamineinducible inhibitory protein, Mol. Cell. Biol.,
13, 23772383.
278. Wu, D., Kaan, H. Y. K., Zheng, X., Tang, X., He, Y.,
Vanessa Tan, Q., Zhang, N., and Song, H. (2015)
Structural basis of ornithine decarboxylase inactivation
and accelerated degradation by polyamine sensor
Antizyme1, Sci. Rep., 5, 14738.
279. Bercovich, Z., RosenbergHasson, Y., Ciechanover, A.,
and Kahana, C. (1989) Degradation of ornithine decar
boxylase in reticulocyte lysate is ATPdependent but ubiq
uitinindependent, J. Biol. Chem.,264, 1594915952.
280. Takeuchi, J., Chen, H., and Coffino, P. (2007) Proteasome
substrate degradation requires association plus extended
peptide, EMBO J.,26, 123131.
281. Bertolaet, B. L., Clarke, D. J., Wolff, M., Watson, M. H.,
Henze, M., Divita, G., and Reed, S. I. (2001) UBA
domains of DNA damageinducible proteins interact with
ubiquitin, Nat. Struct. Biol., 8, 417422.
282. Orlowski, M., and Wilk, S. (2000) Catalytic activities of
the 20S proteasome, a multicatalytic proteinase complex,
Arch. Biochem. Biophys.,383, 116.
283. Erales, J., and Coffino, P. (2014) Ubiquitinindependent
proteasomal degradation, Biochim. Biophys. Acta,1843,
216221.
284. Pickering, A. M., and Davies, K. J. A. (2012) Degradation
of damaged proteins – the main function of the 20S pro
teasome, Prog. Mol. Biol. Transl. Sci., 109, 227248.
285. Van der Lee, R., Buljan, M., Lang, B., Weatheritt, R. J.,
Daughdrill, G. W., Dunker, A. K., Fuxreiter, M., Gough,
J., Gsponer, J., Jones, D. T., Kim, P. M., Kriwacki, R. W.,
Oldfield, C. J., Pappu, R. V., Tompa, P., Uversky, V. N.,
Wright, P. E., and Babu, M. M. (2014) Classification of
intrinsically disordered regions and proteins, Chem. Rev.,
114, 65896631.
286. Aiken, C. T., Kaake, R. M., Wang, X., and Huang, L.
(2011) Oxidative stressmediated regulation of proteasome
complexes, Mol. Cell. Proteomics,10, R110.006924.
287. AlvarezCastelao, B., and Castano, J. G. (2005)
Mechanism of direct degradation of IκBαby 20S protea
some, FEBS Lett., 579, 47974802.
288. Zhang, W., and Wei, Q. (2011) Calcineurin stimulates the
expression of inflammatory factors in RAW 264.7 cells by
interacting with proteasome subunit alpha type 6,
Biochem. Biophys. Res. Commun., 407, 668673.
289. Yuksek, K., Chen, W.L., Chien, D., and Ou, J.H. (2009)
Ubiquitinindependent degradation of hepatitis C virus F
protein, J. Virol.,83, 612621.
290. Stohwasser, R., Holzhutter, H.G., Lehmann, U.,
Henklein, P., and Kloetzel, P.M. (2003) Hepatitis B virus
HBx peptide 116138 and proteasome activator PA28
compete for binding to the proteasome α4/MC6 subunit,
Biol. Chem.,384, 3949.
291. Dong, J., Chen, W., Welford, A., and WandingerNess, A.
(2004) The proteasome αsubunit XAPC7 interacts specif
ically with Rab7 and late endosomes, J. Biol. Chem.,279,
2133421342.
292. Dachsel, J. C., Lucking, C. B., Deeg, S., Schultz, E.,
Lalowski, M., Casademunt, E., Corti, O., Hampe, C.,
Patenge, N., Vaupel, K., Yamamoto, A., Dichgans, M.,
Brice, A., Wanker, E. E., Kahle, P. J., and Gasser, T. (2005)
Parkin interacts with the proteasome subunit α4, FEBS
Lett.,579, 39133919.
293. Yang, L., Tang, Z., Zhang, H., Kou, W., Lu, Z., Li, X., Li,
Q., and Miao, Z. (2013) PSMA7 directly interacts with
NOD1 and regulates its function, Cell. Physiol. Biochem.,
31, 952959.
294. Touitou, R., Richardson, J., Bose, S., Nakanishi, M.,
Rivett, J., and Allday, M. J. (2001) A degradation signal
located in the Cterminus of p21WAF1/CIP1 is a binding
site for the C8 alphasubunit of the 20S proteasome,
EMBO J.,20, 23672375.
295. Zhang, Z., Wang, H., Li, M., Agrawal, S., Chen, X., and
Zhang, R. (2004) MDM2 is a negative regulator of
p21WAF1/CIP1, independent of p53, J. Biol. Chem.,279,
1600016006.
296. Yi, P., Feng, Q., Amazit, L., Lonard, D. M., Tsai, S. Y.,
Tsai, M.J., and O’Malley, B. W. (2008) Atypical protein
kinase C regulates dual pathways for degradation of the
oncogenic coactivator SRC3/AIB1, Mol. Cell, 29, 465
476.
297. Sdek, P., Ying, H., Chang, D. L. F., Qiu, W., Zheng, H.,
Touitou, R., Allday, M. J., and Jim Xiao, Z.X. (2005)
MDM2 promotes proteasomedependent ubiquitininde
pendent degradation of retinoblastoma protein, Mol. Cell,
20, 699708.
298. AlvarezCastelao, B., Goethals, M., Vandekerckhove, J.,
and Castano, J. G. (2014) Mechanism of cleavage of
alphasynuclein by the 20S proteasome and modulation of
its degradation by the RedOx state of the Nterminal
methionines, Biochim. Biophys. Acta Mol. Cell Res., 1843,
352365.
299. Yuan, F., Ma, Y., You, P., Lin, W., Lu, H., Yu, Y., Wang,
X., Jiang, J., Yang, P., Ma, Q., and Tao, T. (2013) A novel
S192 KUDRIAEVA, BELOGUROV
BIOCHEMISTRY (Moscow) Vol. 84 Suppl. 1 2019
role of proteasomal β1 subunit in tumorigenesis, Biosci.
Rep., 33, 555565.
300. Gruendler, C., Lin, Y., Farley, J., and Wang, T. (2001)
Proteasomal degradation of Smad1 induced by bone
morphogenetic proteins, J. Biol. Chem.,276, 46533
46543.
301. JarielEncontre, I., Bossis, G., and Piechaczyk, M. (2008)
Ubiquitinindependent degradation of proteins by the pro
teasome, Biochim. Biophys. Acta,1786, 153177.
302. Adler, J., Reuven, N., Kahana, C., and Shaul, Y. (2010) c
Fos proteasomal degradation is activated by a default
mechanism, and its regulation by NAD(P)H:quinone oxi
doreductase 1 determines cFos serum response kinetics,
Mol. Cell. Biol., 30, 37673778.
303. Pakay, J. L., Diesch, J., Gilan, O., Yip, Y.Y., Sayan, E.,
Kolch, W., Mariadason, J. M., Hannan, R. D.,
Tulchinsky, E., and Dhillon, A. S. (2012) A 19S proteaso
mal subunit cooperates with an ERK MAPKregulated
degron to regulate accumulation of Fra1 in tumour cells,
Oncogene, 31, 18171824.
304. Kuzina, E. S., Chernolovskaya, E. L., Kudriaeva, A. A.,
Zenkova, M. A., Knorre, V. D., Surina, E. A.,
Ponomarenko, N. A., Bobik, T. V., Smirnov, I. V.,
Bacheva, A. V., Belogurov, A. A., Gabibov, A. G., and
Vlasov, V. V. (2013) Immunoproteasome enhances intra
cellular proteolysis of myelin basic protein, Dokl. Biochem.
Biophys.,453, 300303.
305. Belogurov, A., Kudriaeva, A., Kuzina, E., Smirnov, I.,
Bobik, T., Ponomarenko, N., KravtsovaIvantsiv, Y.,
Ciechanover, A., and Gabibov, A. (2014) Multiple sclero
sis autoantigen myelin basic protein escapes control by
ubiquitination during proteasomal degradation, J. Biol.
Chem.,289, 1775817766.
306. Belogurov, A., Kuzina, E., Kudriaeva, A., Kononikhin, A.,
Kovalchuk, S., Surina, Y., Smirnov, I., Lomakin, Y.,
Bacheva, A., Stepanov, A., Karpova, Y., Lyupina, Y.,
Kharybin, O., Melamed, D., Ponomarenko, N., Sharova,
N., Nikolaev, E., and Gabibov, A. (2015) Ubiquitininde
pendent proteosomal degradation of myelin basic protein
contributes to development of neurodegenerative autoim
munity, FASEB J., 29, 19011913.
307. Kuzina, E., Kudriaeva, A., Smirnov, I., Dubina, M. V.,
Gabibov, A., and Belogurov, A. (2014) Glatiramer acetate
and nanny proteins restrict access of the multiple sclerosis
autoantigen myelin basic protein to the 26S proteasome,
Biomed. Res. Int., 2014, 926394.
308. Raule, M., Cerruti, F., and Cascio, P. (2014) Enhanced
rate of degradation of basic proteins by 26S immunopro
teasomes, Biochim. Biophys. Acta,1843, 19421947.
309. Dange, T., Smith, D., Noy, T., Rommel, P. C., Jurzitza,
L., Cordero, R. J. B., Legendre, A., Finley, D., Goldberg,
A. L., and Schmidt, M. (2011) Blm10 protein promotes
proteasomal substrate turnover by an active gating mecha
nism, J. Biol. Chem.,286, 4283042839.
310. Melo, S. P., Barbour, K. W., and Berger, F. G. (2011)
Cooperation between an intrinsically disordered region
and a helical segment is required for ubiquitinindepend
ent degradation by the proteasome, J. Biol. Chem.,286,
3655936567.
311. Singh Gautam, A. K., Balakrishnan, S., and Venkatraman,
P. (2012) Direct ubiquitin independent recognition and
degradation of a folded protein by the eukaryotic protea
somes – origin of intrinsic degradation signals, PLoS One,
7, e34864.
312. Alfassy, O. S., Cohen, I., Reiss, Y., Tirosh, B., and Ravid,
T. (2013) Placing a disrupted degradation motif at the C
terminus of proteasome substrates attenuates degradation
without impairing ubiquitylation, J. Biol. Chem.,288,
1264512653.
313. Tsvetkov, P., Reuven, N., and Shaul, Y. (2009) The nanny
model for IDPs, Nat. Chem. Biol.,5, 778781.
314. Wang, X. J., Yu, J., Wong, S. H., Cheng, A. S., Chan, F.
K., Ng, S. S., Cho, C. H., Sung, J. J., and Wu, W. K.
(2013) A novel crosstalk between two major protein degra
dation systems, Autophagy, 9, 15001508.
315. Qiao, L., and Zhang, J. (2009) Inhibition of lysosomal
functions reduces proteasomal activity, Neurosci. Lett.,
456, 1519.
316. Korolchuk, V. I., Mansilla, A., Menzies, F. M., and
Rubinsztein, D. C. (2009) Autophagy inhibition compro
mises degradation of ubiquitinproteasome pathway sub
strates, Mol. Cell, 33, 517527.
317. Gao, Z., Gammoh, N., Wong, P.M., Erdjument
Bromage, H., Tempst, P., and Jiang, X. (2010) Processing
of autophagic protein LC3 by the 20S proteasome,
Autophagy, 6, 126137.
... The presence of basic amino acid residues in the P1 position is common in peptides generated during the classical secretory pathway, such as neuropeptides and hormones [27]. The proteasome is well known for its role in generating antigenic fragments during adaptive immunity [28,29]. However, it has been suggested that the proteasome also produces antimicrobial peptides as part of the innate immune response, as shown with the intermediate filament keratin 6a (K6a), which is constitutively processed into antimicrobial fragments in corneal epithelial cells [30]. ...
Article
Full-text available
The seahorse is a marine teleost fish member of the Syngnathidae family that displays a complex variety of morphological and reproductive behavior innovations and has been recognized for its medicinal importance. In the Brazilian ichthyofauna, the seahorse Hippocampus reidi is among the three fish species most used by the population in traditional medicine. In this study, a protocol was performed based on fast heat inactivation of proteases plus liquid chromatography coupled to mass spectrometry to identify native peptides in gills of seahorse H. reidi. The MS/MS spectra obtained from gills allowed the identification of 1080 peptides, of which 1013 peptides were present in all samples and 67 peptide sequences were identified in an additional LC-MS/MS run from an alkylated and reduced pool of samples. The majority of peptides were fragments of the internal region of the amino acid sequence of the precursor proteins (67%), and N- and C-terminal represented 18% and 15%, respectively. Many peptide sequences presented ribosomal proteins, histones and hemoglobin as precursor proteins. In addition, peptide fragments from moronecidin-like protein, described with antimicrobial activity, were found in all gill samples of H. reidi. The identified sequences may reveal new bioactive peptides.
... The ubiquitin-proteasome system and the autophagy-lysosome system are essential intracellular machinery [1,2]. The proteasome is assembled from a catalytic core particle, also termed the 20S proteasome, which is capped by different regulatory particles. ...
Article
Full-text available
Proteasomes exist in mammalian cells in multiple combinatorial variants due to the diverse regulatory particles and exchange of catalytic subunits. Here, using biotin carboxyl carrier domain of transcarboxylase from Propionibacterium shermanii fused with different proteasome subunits of catalytic and regulatory particles, we report comprehensive characterization of highly homogenous one-step purified human constitutive and immune 20S and 26S/30S proteasomes. Hydrolysis of a multiple sclerosis (MS) autoantigen, myelin basic protein (MBP), by engineered human proteasomes with different catalytic phenotypes, revealed that peptides which may be directly loaded on the HLA class I molecules are produced mainly by immunoproteasomes. We detected at least five MBP immunodominant core regions, namely, LPRHRDTGIL, SLPQKSHGR, QDENPVVHFF, KGRGLSLSRF and GYGGRASDY. All peptides, except QDENPVVHFF, which originates from the encephalitogenic MBP part, were associated with HLA I alleles considered to increase MS risk. Prediction of the affinity of HLA class I to this peptide demonstrated that MS-protective HLA-A*44 and -B*35 molecules are high-affinity binders, whereas MS-associated HLA-A*23, -A*24, -A*26 and -B*51 molecules tend to have moderate to low affinity. The HLA-A*44 molecules may bind QDENPVVHFF and its deamidated form in several registers with unprecedently high affinity, probably linking its distinct protective phenotype with thymic depletion of the repertoire of autoreactive cytotoxic T cells or induction of CD8+ regulatory T cells, specific to the encephalitogenic MBP peptide.
... For example, Calnan et al. demonstrated that FoxO3 methylation by the SET domain containing lysine methyltransferase 9 (Set9) at Lys270 and Lys271 resulted in the reduction of the DNA-binding activity and transactivation of FoxO3 [53]. Ubiquitination is a reversible PTM whose primary function is to mediate protein degradation via the ubiquitin-proteasome pathway [54]. In this way, ubiquitin-proteasome-mediated protein degradation plays a crucial role in regulating several cellular processes, such as cell cycle progression, transcriptional regulation, and DNA repair [55]. ...
Article
Full-text available
Oxidative stress is a major cause of morbidity and mortality in human health and disease. In this review, we focus on the Forkhead Box (Fox) subclass O3 (FoxO3), an extensively studied transcription factor that plays a pleiotropic role in a wide range of physiological and pathological processes by regulating multiple gene regulatory networks involved in the modulation of numerous aspects of cellular metabolism, including fuel metabolism, cell death, and stress resistance. This review will also focus on regulatory mechanisms of FoxO3 expression and activity, such as crucial post-translational modifications and non-coding RNAs. Moreover, this work discusses and evidences some pathways to how this transcription factor and reactive oxygen species regulate each other, which may lead to the pathogenesis of various types of diseases. Therefore, in addition to being a promising therapeutic target, the FoxO3-regulated signaling pathways can also be used as reliable diagnostic and prognostic biomarkers and indicators for drug responsiveness.
Article
Genome stability is critical for normal functioning of cells and depends on accuracy of DNA replication, chromosome segregation, and DNA repair. Cellular defense mechanisms against DNA damage are important for preventing the development of cancer and aging. The E3 ubiquitin ligase RNF168 of the RING superfamily is an essential component of the complex responsible for the ubiquitination of H2A/H2A.X histones near DNA double-strand breaks, which is a key step in attracting repair factors to the injury site. In this study, we unequivocally showed that RNF168 does not have ability to directly distinguish the architecture of polyubiquitin chains, except for tropism of its two ubiquitin-binding domains UDM1/2 to K63 ubiquitin chains. Analysis of the intracellular chromatosomal environment of full-length RNF168 and its domains by ligand-induced bioluminescence resonance energy transfer (BRET) revealed that the C-terminal part of UDM1 is associated with K63 ubiquitin chains; RING and the N-terminal part of UDM2 are sterically close to K63- and K48- ubiquitin chains, while the C-terminal part of UDM1 is colocalized with all possible ubiquitin variants. Our observations together with the available structural data suggest that the C-terminal part of UDM1 binds K63 polyubiquitin chains on linker histone H1; RING and the N-terminal part of UDM2 are located in the central part of the nucleosome and sterically close to H1 and K48-ubiquitinated alternative substrates of RNF168, such as JMJD2A/B demethylases, while the C-terminal part of UDM1 is in the region of an activated ubiquitin residue associated with E2 ubiquitin ligase, engaged by RNF168.
Article
Today, genetic studies yield quite a large amount of information about a person, which, in many cases, allows predicting the risks of certain diseases. This gives grounds to believe that such testing can also be applied in the field of manned spaceflights in order to identify candidates best adapted to specific risks. The article examines publications on genetic polymorphisms and their effects on the carrier phenotype, namely, on such manifestations that are of interest in the context of risks arising during long-term space flights. Specific genes are listed and examples of allelic variants are given. Publications describing new molecular methods of monitoring human health are also considered, biomarkers that can be used for research in the interests of regular examination of active astronauts are identified.
Article
Genome stability is critical for normal functioning of cells, it depends on accuracy of DNA replication, chromosome segregation, and DNA repair. Cellular defense mechanisms against DNA damage are important for preventing cancer development and aging. The E3 ubiquitin ligase RNF168 of the RING superfamily is an essential component of the complex responsible for ubiquitination of the H2A/H2A.X histones near DNA double-strand breaks, which is a key step in attracting repair factors to the damage site. In this study, we unequivocally showed that RNF168 does not have the ability to directly distinguish architecture of polyubiquitin chains, except for the tropism of its two ubiquitin-binding domains UDM1/2 to K63 ubiquitin chains. Analysis of intracellular chromatosomal environment of the full-length RNF168 and its domains using the ligand-induced bioluminescence resonance energy transfer (BRET) revealed that the C-terminal part of UDM1 is associated with the K63 ubiquitin chains; RING and the N-terminal part of UDM2 are sterically close to the K63- and K48-ubiquitin chains, while the C-terminal part of UDM1 is co-localized with all possible ubiquitin variants. Our observations together with the available structural data suggest that the C-terminal part of UDM1 binds the K63 polyubiquitin chains on the linker histone H1; RING and the N-terminal part of UDM2 are located in the central part of nucleosome and sterically close to H1 and K48-ubiquitinated alternative substrates of RNF168, such as JMJD2A/B demethylases, while the C-terminal part of UDM1 is in the region of activated ubiquitin residue associated with E2 ubiquitin ligase, engaged by RNF168.
Article
Introduction:Tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is considered to be an effective apoptosis inducer due to its selectivity for tumor cells. However, many cancer cells, especially metastatic cancer cells, often exhibit resistance to TRAIL because their apoptotic pathway is impaired or their pro-survival pathway is overactivated. TRAIL-resistance is the main obstacle to current TRAIL therapy. Nowadays, ceramide analogues represent a new class of potential anticancer agents. Therefore, we hypothesized that disrupting pro-survival signaling with ceramide analogues would increase TRAIL-mediated apoptosis. Methods: MTT assay and Flow cytometry were conducted to evaluate the synergistic effect of ceramide analogue 5cc to TRAIL in metastatic colon cancer cells. Western blot was used to detect signaling proteins affected by 5cc. RNA interference was performed to analyzed the effects of specific gene on 5cc-enhanced apoptosis. Results:Ceramide analogue 5cc markedly enhanced TRAIL-induced apoptosis evidenced by increased PI/Annexin V double-positive cells and PARP cleavage in SW620 and LS411N cells. At the molecular level, 5cc significantly reduced the expression of anti-apoptotic protein X-linked inhibitor of apoptosis protein (XIAP) through the activation of C-Jun n-terminal kinase (JNK) pathway which is critically involved in sensitizing tumor cells to TRAIL/5cc combination. JNK-silenced cells exhibited a significant reversal of TRAIL/5cc-mediated apoptosis. Conclusion:Our data demonstrated that ceramide analogue 5cc overcomes TRAIL-resistance by enhancing JNK activation and repressing XIAP expression in metastatic colon cancer cells.
Article
Full-text available
Peptides and proteins have been found to possess an inherent tendency to convert from their native functional states into intractable amyloid aggregates. This phenomenon is associated with a range of increasingly common human disorders, including Alzheimer and Parkinson diseases, type II diabetes, and a number of systemic amyloidoses. In this review, we describe this field of science with particular reference to the advances that have been made over the last decade in our understanding of its fundamental nature and consequences. We list the proteins that are known to be deposited as amyloid or other type of aggregates in human tissues and the disorders with which they are associated, as well as the proteins that exploit the amyloid motif to play specific functional roles in humans. In addition, we summarize the genetic factors that have provided insight into the mechanisms of disease onset. We describe recent advances in our knowledge of the structures of amyloid fibrils and their oligomeric precursors and of the mechanisms by which they are formed and proliferate to generate cellular dysfunction. We show evidence that a complex proteostasis network actively combats protein aggregation and that such an efficient system can fail in some circumstances and give rise to disease. Finally, we anticipate the development of novel therapeutic strategies with which to prevent or treat these highly debilitating and currently incurable conditions. Expected final online publication date for the Annual Review of Biochemistry Volume 86 is June 20, 2017. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
As the endpoint for the ubiquitin-proteasome system, the 26S proteasome is the principal proteolytic machine responsible for regulated protein degradation in eukaryotic cells. The proteasome's cellular functions range from general protein homeostasis and stress response to the control of vital processes such as cell division and signal transduction. To reliably process all the proteins presented to it in the complex cellular environment, the proteasome must combine high promiscuity with exceptional substrate selectivity. Recent structural and biochemical studies have shed new light on the many steps involved in proteasomal substrate processing, including recognition, deubiquitination, and ATP-driven translocation and unfolding. In addition, these studies revealed a complex conformational landscape that ensures proper substrate selection before the proteasome commits to processive degradation. These advances in our understanding of the proteasome's intricate machinery set the stage for future studies on how the proteasome functions as a major regulator of the eukaryotic proteome. Expected final online publication date for the Annual Review of Biochemistry Volume 87 is June 20, 2018. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
In selective autophagy, cytoplasmic components are selected and tagged before being sequestered into an autophagosome by means of selective autophagy receptors such as p62/SQSTM1. In this review, we discuss how selective autophagy is regulated. An important level of regulation is the selection of proteins or organelles for degradation. Components selected for degradation are tagged, often with ubiquitin, to facilitate recognition by autophagy receptors. Another level of regulation is represented by the autophagy receptors themselves. For p62, its ability to co-aggregate with ubiquitinated substrates is strongly induced by post-translational modifications (PTMs). The transcription of p62 is also markedly increased during conditions in which selective autophagy substrates accumulate. For other autophagy receptors, the LC3-interacting region (LIR) motif is regulated by PTMs, inhibiting or stimulating the interaction with ATG8 family proteins. ATG8 proteins are also regulated by PTMs. Regulation of the capacity of the core autophagy machinery also affects selective autophagy. Importantly, autophagy receptors can induce local recruitment and activation of ULK1/2 and PI3KC3 complexes at the site of cargo sequestration.
Article
The 20S immunoproteasome (IP) is an interferon(IFN)-γ - and tumor necrosis factor (TNF) -inducible variant of the 20S constitutive proteasome (CP) in which all its peptidolytically active subunits β1, β2, and β5 are replaced by their cytokine inducible homologues β1i (LMP2), β2i (MECL-1), and β5i (LMP7). These subunit replacements alter the cleavage specificity of the proteasome and the spectrum of proteasome-generated peptide ligands of MHC class I molecules. In addition to antigen processing, the IP has recently been shown to serve unique functions in the generation of pro-inflammatory T helper cell subtypes and cytokines as well as in the pathogenesis of autoimmune diseases, but the mechanistic involvement of the IP in these processes has remained elusive. In this study we investigated whether the IP differs from the CP in the interaction with two IFN-γ/TNF inducible factors: the 11S proteasome regulator PA28αβ and the ubiquitin-like modifier FAT10 (ubiquitin D). Using thermophoresis, we determined the affinity of PA28αβ for the CP and IP to be 12.2nM +/- 2.8nM and 15.3nM +/- 2.7nM, respectively, which is virtually identical. Also the activation of the peptidolytic activities of the IP and CP by PA28αβ did not differ. For FAT10 we determined the degradation kinetics in cycloheximide chase experiments in cells expressing almost exclusively IP or CP as well as in IFN-γ stimulated and unstimulated cells and found no differences between the degradation rates. Taken together, we conclude that neither differences in the binding strength to, nor activation by PA28αβ, nor a difference in the rate of FAT10-mediated degradation can account for distinct functional capabilities of the IP as compared to the CP.
Article
Posttranslational modification with ubiquitin chains controls cell fate in all eukaryotes. Depending on the connectivity between subunits, different ubiquitin chain types trigger distinct outputs, as seen with K48- and K63-linked conjugates that drive protein degradation or complex assembly, respectively. Recent biochemical analyses also suggested roles for mixed or branched ubiquitin chains, yet without a method to monitor endogenous conjugates, the physiological significance of heterotypic polymers remained poorly understood. Here, we engineered a bispecific antibody to detect K11/K48-linked chains and identified mitotic regulators, misfolded nascent polypeptides, and pathological Huntingtin variants as their endogenous substrates. We show that K11/K48-linked chains are synthesized and processed by essential ubiquitin ligases and effectors that are mutated across neurodegenerative diseases; accordingly, these conjugates promote rapid proteasomal clearance of aggregation-prone proteins. By revealing key roles of K11/K48-linked chains in cell-cycle and quality control, we establish heterotypic ubiquitin conjugates as important carriers of biological information.
Article
The conjugation of the 76 amino acid protein ubiquitin to other proteins can alter the metabolic stability or non-proteolytic functions of the substrate. Once attached to a substrate (monoubiquitination), ubiquitin can itself be ubiquitinated on any of its seven lysine (Lys) residues or its N-terminal methionine (Met1). A single ubiquitin polymer may contain mixed linkages and/or two or more branches. In addition, ubiquitin can be conjugated with ubiquitin-like modifiers such as SUMO or small molecules such as phosphate. The diverse ways to assemble ubiquitin chains provide countless means to modulate biological processes. We overview here the complexity of the ubiquitin code, with an emphasis on the emerging role of linkage-specific degradation signals (degrons) in the ubiquitin-proteasome system (UPS) and the autophagy-lysosome system (hereafter autophagy).
Article
Poly-ubiquitin chains direct protein substrates to the 26S proteasome, where they are removed by the deubiquitinase Rpn11 during ATP-dependent substrate degradation. Rapid deubiquitination is required for efficient degradation but must be restricted to committed substrates that are engaged with the ATPase motor to prevent premature ubiquitin chain removal and substrate escape. Here we reveal the ubiquitin-bound structure of Rpn11 from S. cerevisiae and the mechanisms for mechanochemical coupling of substrate degradation and deubiquitination. Ubiquitin binding induces a conformational switch of Rpn11's Insert-1 loop from an inactive closed state to an active β hairpin. This switch is rate-limiting for deubiquitination and strongly accelerated by mechanical substrate translocation into the AAA+ motor. Deubiquitination by Rpn11 and ubiquitin unfolding by the ATPases are in direct competition. The AAA+ motor-driven acceleration of Rpn11 is therefore important to ensure that poly-ubiquitin chains are removed only from committed substrates and fast enough to prevent their co-degradation.
Article
The 20S proteasome is the main protease for degradation of oxidatively damaged and intrinsically disordered proteins. When accumulation of disordered or oxidatively damaged proteins exceed proper clearance in neurons, imbalanced pathway signaling or aggregation occurs, which have been implicated in the pathogenesis of several neurological disor-ders. Screening of the NIH Clinical Collection and Prestwick libraries identified the neuroleptic agent chlorpromazine as a lead agent capable of enhancing 20S proteasome activity. Chemical manipulation of chlorpromazine abrogated its D2R receptor binding affinity while retaining its ability to enhance 20S mediated proteolysis at low micromolar concentrations. The resulting small molecule enhancers of 20S proteasome activity induced the degradation of intrinsically disordered protein, alpha-synuclein but not structured proteins. These small molecules enhancers of 20S proteasome can serve as leads to explore the therapeutic potential of 20S activation, or as new tools to provide insight into the yet unclear mechanics of 20S-gate regulation.
Article
Cells of all organisms survey problems during translation elongation, which may happen as a consequence of mRNA aberrations, inefficient decoding, or other sources. In eukaryotes, ribosome-associated quality control (RQC) senses elongation-stalled ribosomes and promotes dissociation of ribosomal subunits. This so-called ribosomal rescue releases the mRNA for degradation and allows 40S subunits to be recycled for new rounds of translation. However, the nascent polypeptide chains remain linked to tRNA and associated with the rescued 60S subunits. As a final critical step in this pathway, the Ltn1/Listerin E3 ligase subunit of the RQC complex (RQCc) ubiquitylates the nascent chain, which promotes clearance of the 60S subunit while simultaneously marking the nascent chain for elimination. Here we review the ribosomal stalling and rescue steps upstream of the RQCc, where one witnesses intersection with cellular machineries implicated in translation elongation, translation termination, ribosomal subunit recycling, and mRNA quality control. We emphasize both recent progress and future directions in this area, as well as examples linking ribosomal rescue with the production of Ltn1-RQCc substrates. Expected final online publication date for the Annual Review of Cell and Developmental Biology Volume 33 is October 6, 2017. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.