ArticlePDF Available

Structural Organization of the 19S Proteasome Lid: Insights from MS of Intact Complexes

PLOS
PLOS Biology
Authors:
  • Rosetta Design Group LLC

Abstract and Figures

The 26S proteasome contains a 19S regulatory particle that selects and unfolds ubiquitinated substrates for degradation in the 20S catalytic particle. To date there are no high-resolution structures of the 19S assembly, nor of the lid or base subcomplexes that constitute the 19S. Mass spectra of the intact lid complex from Saccharomyces cerevisiae show that eight of the nine subunits are present stoichiometrically and that a stable tetrameric subcomplex forms in solution. Application of tandem mass spectrometry to the intact lid complex reveals the subunit architecture, while the coupling of a cross-linking approach identifies further interaction partners. Taking together our results with previous analyses we are able to construct a comprehensive interaction map. In summary, our findings allow us to identify a scaffold for the assembly of the particle and to propose a regulatory mechanism that prevents exposure of the active site until assembly is complete. More generally, the results highlight the potential of mass spectrometry to add crucial insight into the structural organization of an endogenous, wild-type complex.
Content may be subject to copyright.
Structural Organization of the 19S Proteasome
Lid: Insights from MS of Intact Complexes
Michal Sharon
1
, Thomas Taverner
1
, Xavier I. Ambroggio
2
, Raymond J. Deshaies
2
, Carol V. Robinson
1*
1 Department of Chemistry, University of Cambridge, Cambridge, United Kingdom, 2 Division of Biology, California Institute of Techno logy, Pasadena, California, United
States of America
The 26S proteasome contains a 19S regulatory particle that selects and unfolds ubiquitinated substrates for
degradation in the 20S catalytic particle. To date there are no high-resolution structures of the 19S assembly, nor of
the lid or base subcomplexes that constitute the 19S. Mass spectra of the intact lid complex from Saccharomyces
cerevisiae show that eight of the nine subunits are present stoichiometrically and that a stable tetrameric subcomplex
forms in solution. Application of tandem mass spectrometry to the intact lid complex reveals the subunit architecture,
while the coupling of a cross-linking approach identifies further interaction partners. Taking together our results with
previous analyses we are able to construct a comprehensive interaction map. In summary, our findings allow us to
identify a scaffold for the assembly of the particle and to propose a regulatory mechanism that prevents exposure of
the active site until assembly is complete. More generally, the results highlight the potential of mass spectrometry to
add crucial insight into the structural organization of an endogenous, wild-type complex.
Citation: Sharon M, Taverner T, Ambroggio XI, Deshaies RJ, Robinson CV (2006) Structural organization of the 19S proteasom e lid: Insights from MS of intact complexes. PLoS
Biol 4(8): e267. DOI: 10.1371/journal.pbio.0040267
Introduction
In eukaryotes, the ubiquitin–proteasome pathway is essen-
tial for eliminating damaged or misfolded proteins and for the
degradation of short-lived regulatory proteins [1,2]. The two
major subcomplexes of the 26S proteasome are the 20S
proteolytic core particle (;700 kDa) and the 19S regulatory
particle (;900 kDa). The 19S is attached at either or both ends
of the 20S [3]. Selection of proteasomal substrates by the 19S
regulatory particle occurs via the recognition of polyubiquitin
chains bound to proteins that are destined to be degraded [4].
After selection of the labeled substrate, the polyubiquitin
chain is then removed by the deubiquitinating enzymes Rpn11
[5,6] and Ubp6 [7]. The substrate is then unfolded and
translocated into the 20S channel, where it is degraded.
Nineteen different subunits have been identified in the 19S
proteasome from yeast as well as mammals [8–10]. This
regulatory particle can be dissociated further into two
subcomplexes, the base that binds directly to the 20S and a
peripheral lid [11]. The base consists of six AAA-ATPase
subunits, Rpt1–Rpt6, and four non-ATPase subunits, Rpn1,
Rpn2, Rpn10, and Rpn13. The ATPases in the base are
required for unfolding of substrate proteins [12] and channel
opening [13] before translocation of substrate into the 20S
cavity. The lid is composed of nine non-ATPase subunits:
Rpn3, Rpn5–Rpn9, Rpn11, Rpn12, and Sem1. The major
activity of the lid is proposed to be deubiquitination [5–7].
Rpn10 and additional ubiquitin receptor proteins that
reversibly associate with the proteasome, such as Rad23 and
Dsk2, deliver the ubiquitinated protein to the proteasome [14].
The proteasome lid subunits exhibit high homology to the
COP9 signalosome complex (CSN) [11], an essential regulator
of diverse cellular and developmental processes, and has been
found in most eukaryotic organisms ranging from yeast to
human (for reviews see [15–17]). The CSN contains eight core
subunits that assemble into a 450-kDa particle. The subunits
show a remarkable one-to-one sequence correspondence with
those of the 19S lid, suggesting a common ancestry [11]. Of
the lid components, Rpn11 is the most highly conserved
subunit, with 65% identity between yeast and human proteins
[6]. Recently it was identified as a Zn
2þ
-dependent metal-
loprotease responsible for substrate deubiquitination during
proteasomal degradation [5,6,18,19].
Although in the last few years functional characterization of
the 19S lid has progressed considerably, the assembly poses a
considerable challenge to structural biologists. High-resolu-
tion structural analyses are difficult since the complex is
assembled in vivo and different subunits are expected to be
dynamic and heterogeneous [20]. Only a low-resolution 3-D
image of the 26S has been observed by electron microscopy
[3], and the structural features of the lid complex were
difficult to discern. Pairwise interactions among the lid
subunits have been studied using the yeast two-hybrid system
[20–26]. Insight into the organization of the lid has also been
gleaned from analysis of lid subcomplexes in cells that express
mutant forms of specific lid subunits [25,26]. However, a
comprehensive model of subunit interactions within the lid
has not b een proposed, and no distinct ive t opolo gical
features, such as rings or well-defined channels, have been
identified.
In this study, we aim to obtain novel information regarding
Academic Editor: Michael Glickman, Biology Department, Technion-Israel Institute
of Technology, Israel
Received March 31, 2006; Accepted June 9, 2006; Published August 1, 2006
DOI: 10.1371/jou rnal.pbio.0040267
Copyright: Ó 2006 Sharon et al. This is an open-access article distributed under the
terms of the Creative Commons Attribution License, which permits unrestricted
use, distribution, and reproduction in any medium, provided the original author
and source are credited.
Abbreviations: CSN, COP9 signalosome; MALDI, matrix-assisted laser desorption/
ionization; MS, mass spectrometry; MS/MS, tandem mass spectrometry; TOF, time
of flight
* To whom correspondence should be addressed. E-mail: cvr24@cam.ac.uk
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671314
P
L
o
S
BIOLOGY
the structural organization of the 19S lid from Saccharomyces
cerevisiae using both tandem mass spectrometry (MS/MS) of
the intact nine-component complex together with chemical
cross-linking and MS analysis. It is established that MS and
MS/MS can be applied to the study of intact noncovalent
complexes, but to date, the majority of examples were of
various complexes reconstituted in vitro [27–29]. While some
studies have been carried out on complexes isolated directly
from cells, including bacterial RNA polymerase [30], ribo-
somes [31,32], and RNA-processing complexes from yeast
[33], existing structural information aided the interpretation
of the resulting MS and MS/MS spectra. Here we show that the
end ogenous wild-type complex can be studied directly,
without the need to generate mutants, and the results
incorporated with existing yeast two-hybrid data to yield a
comprehensive model. This represents a significant advance
in MS methodology given the absence of existing high-
resolution structural data. Moreover, the fact that all directly
or indirectly interacting subunits can be observed within a
single spectrum highlights one of the fundamental advantages
of our methodology.
Using this approach the mass spectrum revealed not only
the intact complex but also a substoichiometric complex. In
addition, we identified the presence of an independent
subcomplex containing Rpn5, Rpn6, Rpn8, and Rpn9, where-
in Rpn8 occupies a central p osition. We could also
demonstrate that Rpn6, Rpn9, and Rpn12 are found on the
periphery of the complex. Complementary information from
cross-linking approaches enabled us to show that Rpn5 binds
to Rpn3. In addition, we could determine that the recently
identified lid subunit Sem1 [9] binds to both Rpn7 and Rpn3.
Taken together, our data allow us to propose a comprehen-
sive model for the subunit organization of the 19S lid that
incorporates our findings from MS with those of prior
analyses of pairwise interactions.
Results
Electrospray Ionizatio n MS and MS/MS of the Intact Lid
The electrospray mass spectrum of the intact 19S lid
isolated from S. cerevisiae is shown in Figure 1. Despite the fact
that the sample analyzed was the endogenous complex and
was not overexpressed, the spectrum is well resolved and has
one major series of peaks centered at 8,000 m/z. This major
charge state series corresponds to the nine protein compo-
nents of the lid, demonstrating clearly that all subunits must
be interacting, either directly or indirectly, and present at
unit stoichiometry. The measured mass of the intact complex
is 376,151 6 369 Da (Table 1). Interestingly, an overlapping
less-intense charge state series discerned at 7,750 m/z,
indicates the existence of a substoichiometric complex that
results from the absence of a subunit from the complex.
Calculation of the mass difference of these series indicates a
mass that is 50,270 Da less than that of the intact lid. This
corresponds to the absence of Rpn6 from the intact lid. The
similarity in the number of positive charges between the two
complexes suggests that both complexes are present in
solution (see below). An additional minor series of peaks is
centered at 5,500 m/z. The calculated mass for this lower m/z
charge series is 185,556 6 137 Da, indicating that a smaller
subcomplex of the lid is also present in solution.
To probe the composition and subunit organization of the
lid we used an MS/MS approach [31,34]. In such experiments,
a specific well-defined m/z range that encompasses the parent
ion is selected and accelerated through a collision cell at
increased argon pressures. This process gives rise to multiple
collisions in which the internal energy of the ions is
accumulated. When this energy reaches a threshold value,
dissociation occurs, yielding product ions. It is established
that for noncovalent assemblies, individual highly charged
subunits are dissociated from the parent ion, producing
‘‘ stripped’’ complexes with lower charge states than the
original ion [35–38]. As a consequence, we can distinguish
complexes formed in solution from those generated in the
gas phase since the former will have charge states commen-
surate with their mass [39], while those formed in the gas
phase will have anomalously low charge states. If we consider
the propensity for dissociation of different protein subunits
in the gas phase, it is known that under vacuum ionic
interactions are enhanced, while hydrophobic interactions
are weakened [40,41]. However, we cannot rule out the effect
of other factors that govern protein–protein interactions
such as contact area, planarity, shape complementarity, etc.
There is, however, a growing number of studies demonstrat-
ing that individual proteins that are expelled during tandem
MS are those that are known from high-resolution structures
to be more exposed and peripheral, while those at the core of
the assembly are retained in the ‘‘ stripped’’ complex
[30,32,42,43].
Applying this tandem MS strategy to the 376-kDa lid
complex, using a collision cell acceleration voltage of 100 V,
five series of ions were formed (Figure 2A). At m/z values
below 2,400, two highly charged series are observed. Their
measured masses correspond to 45,782 6 4 and 31,796 6 10
Da, in very close agreement with the mass calculated for Rpn9
and Rpn12, respectively. At high m/z values between 10,000
and 18,500 three further charge series are observed. Their
masses are consistent with the lid stripped of Rpn9 and/or
Rpn12. These results indicate that both Rpn9 and Rpn12 have
Figure 1. Electrospray Mass Spectrum of the 19S Lid Complex Isolated
from S. cerevisiae
The intact lid complex is observed between 7,250 and 9,000 m/z, with
the most intense peak at 8,000 m/z. Well-resolved charge states are
observed, and the measured mass confirms the presence of all nine
subunits. An overlapping charge series is detected and corresponds to a
substoichiometric complex in which the Rpn6 subunit is absent (labeled
with asterisks). A subcomplex of the lid gives rise to a signal in the range
of 5,000–6,250 m/z.
DOI: 10.1371/journal.pbio.0040267.g001
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671315
MS of the Proteasome Lid
weaker interactions with other subunits and are peripheral,
and hence are first to dissociate. Increasing the collision cell
voltage, and hence the internal energy of the lid complex,
induces the dissociation of an additional subunit (Figure 2B).
The new charge series observed at 2,000–2,400 m/z is assigned
to Rpn6, and the corresponding stripped complex appears
between 13,000 to 17,000 m/z. These results enable us to
deduce therefore that Rpn6, Rpn9, and Rpn12 interact
weakly and are presumably at the periphery of the lid
subcomplex.
Tandem Mass Spectrum of the Lid Subcomplex
In order to examine the composition of the 185-kDa
subcomplex observed in Figure 1 we again applied MS/MS. A
single charge state isolated from this complex and subjected
to collision with argon at an accelerating potential of 130 V
yielded a well-resolved tandem mass spectrum, despite the
low intensity of the subcomplex relative to the intact complex
(Figure 3A). Three series of ions are observed at the low m/z
region (Figure 3B). On the basis of their measured masses we
assign these species to Rpn5, Rpn6, and Rpn9. Between 8,000
to 16,000 m/z, ions arising from the removal of a single
subunit are observed (Figure 3C). Although Rpn8 and the
modified Rpn11 have similar masses, by simulating the charge
states based on the known sequence of the subunits we can
clearly distinguish between the two possibilities. We could
then conclude that Rpn8 and not Rpn11 is a component of
the subcomplex. These ions are therefore attributed to three
different heterotrimers; Rpn5:6:8, Rpn5:8:9, and Rpn6:8:9.
Interestingly, while Rpn8 is not detected as an individual
subunit, it is present in all three heterotrimers. By extrap-
olation, therefore, we can conclude that prior to the
collisional activation, the subcomplex is a tetramer composed
of Rpn5, Rpn6, Rpn8, and Rpn9. From the dissociation
pattern generated by the MS/MS experiment we can
determine that Rpn8 occupies a central position within this
subcomplex (Figure 3D). It is noteworthy that the two charge
series assigned to monomeric Rpn9 and its stripped hetero-
trimer, namely Rpn5:Rpn6:Rpn8, are the most intense series
in the spectra. This is further evidence for the ease of
dissociation of Rpn9, implying a relatively weak interaction
with neighboring subunits.
Chemical Cross-Linking Analysis
The cross-linking agent BS
3
is a bifunctional reagent
linking amines, and is selective for the N-terminal amino
group and lysine side chains, with an eight-carbon spacer arm
of 11.4 A
˚
. Denaturing gel electrophoresis before and after
chemical cross-linking [44] shows that the reaction resulted in
the production of covalently linked species labeled A to F
according to their electrophoretic mobility (Figure 4A). The
bands were assigned as described in Materials and Methods
(Table 2). As an example, Figure 4B shows part of the matrix-
assisted laser desorption/ionization (MALDI) mass spectrum
of the peptide mixture from band D containing Rpn3, Rpn5,
and Rpn8. The inset within the figure illustrates the MS/MS
analysis of one of the peptides corresponding to Rpn5.
Several conclusions can be drawn from the information
extracted from the six cross-linked bands. As Sem1 is present
in two bands cross-linked with Rpn7 and with Rpn3, we can
assume that Sem1 interacts with both subunits. Moreover,
due to the small molecular mass of Sem1, it is reasonable to
assume that Rpn3 and Rpn7 are in close proximity to one
another. In band C, Rpn3 is observed accompanied by Rpn5,
and we can therefore conclude that Rpn3 binds to Rpn5. In
addition, the subunits identified in bands D and E further
confirm the formation of the subcomplex comprising Rpn5,
Rpn8, and Rpn9 identified from our intact complex data. For
Table 1. Theoretical and Measured Masses of Proteins and Complexes
Protein/Complex Theoretical Mass (Da) Experimental Mass (Da) Figure Number
Individual components Rpn3
a,b
60,303 58,815 6 14
Rpn5
a,b
51,679 51,695 6 8—
Rpn6
a,b
49,685 49,688 6 11
Rpn7
a
48,827 48,829 6 11
Rpn8
a,b
38,223 38,224 6 8—
Rpn9 45,782 45,782 6 10
Rpn11
b,c
37,903 37,921 6 14
Rpn12
a
31,788 31,789 6 9—
Sem1
d
10,386 10,298 6 1—
Complex
e
Intact lid 373,040 376,151 6 369 1
Lid-Rpn6 323,352 325,881 6 160 1
Lid-Rpn6 323,352 323,836 6 294 2B
Lid-Rpn9 327,258 327,175 6 118 2A
Lid-Rpn12 341,252 341,082 6 98 2A
Lid-Rpn9:12 295,470 295,347 6 255 2A
Rpn5:6:8:9 185,389 185,556 6 137 1
Rpn5:6:8 139,607 139,574 6 42 3
Rpn5:8:9 135,701 135,575 6 261 3
Rpn6:8:9 133,694 133,435 6 146 3
a
The initiator Met is removed [54].
b
The N-terminal amino group is acetylated [54].
c
The sequence of Rpn11 was modified for purification purposes.
d
According to the experimental mass the initiator Met is removed and the N-terminal amino group is acetylated.
e
The masses were calculated from experimental masses.
DOI: 10.1371/journal.pbio.0040267.t001
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671316
MS of the Proteasome Lid
Rpn11 (band F) we can deduce that it binds either directly to
Rpn5:9 or is anchored by Rpn3:7 to Rpn5:9. In summary,
cross-linking analysis allows determination of four additional
interactions within the lid.
Discussion
We acquired remarkably well-resolved mass spectra of the
intact lid from S. cerevisiae despite the fact that we anticipated
that the nine-component endogenous complex would be
heterogeneous. From our data we were able to determine the
subunit composition of the intact complex, the substoichio-
metric complexes present in solution, and those generated as
a result of our MS/MS approach. Together with chemical
cross-linking, our MS approach allowed us to obtain precise
additional information regarding the subunit organization of
the lid complex. Based on our data we constructed an
interaction map of the lid subunits (Figure 5). We identified a
heterotetrameric core structure formed by Rpn5, Rpn6,
Rpn8, and Rpn9, in which Rpn8 is centered. The absence of
Rpn11 from this complex shows that it forms after
purification and indicates that it is a stable subcomplex in
solution. In addition, we revealed that Sem1 binds to both
Rpn7 and Rpn3. Considering the small size of Sem1, we
assume that Rpn3 and Rpn7 are also interacting. We could
also determine that Rpn3 binds Rpn5. Interestingly, we
determined a substoichiometric complex of the lid in which
Rpn6 is absent. Close inspection of the spectrum suggests that
20%–30% of the purified lid complexes lack Rpn6. The MS/
MS results clearly show that both Rpn9 and Rpn12 and, to a
lesser extent, Rpn6, readily dissociate from the lid complex. It
is established that the extent of interaction between protein
subunits as well as their exposure underlies the release of
subunits in tandem MS of heteroligomeric complexes [45].
Figure 2. MS/MS of the 47þ Charge State of the Intact 19S Lid
(A) Applying a voltage of 100 V to the collision cell generates product ions from the loss of only Rpn9 and/or Rpn12. Peaks between 12,500–17,000 m/z
correspond to the loss of Rpn9 (blue squares), while the series at 10,000–12,000 m/z correspond to the loss of Rpn12 (red squares). The low intensity
charge series at 17,000–18,000 m/z indicate the loss of both Rpn9 and Rpn12 (magenta squares). At low m/z 1,000–2,400 series of peaks are assigned to
individual Rpn9 (blue circles) and Rpn12 (red circles). The data indicate the Rpn9 and Rpn12 have a relatively small contact surface with the other
subunits, which implies that they are exposed.
(B) Increasing the voltage to 120 V induces the dissociation of Rpn6 in addition to Rpn9 and Rpn12. The highly charged Rpn6 series is centered at ;
2,200 m/z (green circles), while the remaining stripped complex is observed between 13,000 to 18,000 m/z (green squares). In summary, Rpn9 and
Rpn12 are easily dissociated from the lid complex (blue). An increase in the collision cell voltage also induces dissociation of Rpn6 (purple).
DOI: 10.1371/journal.pbio.0040267.g002
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671317
MS of the Proteasome Lid
This indicates that these subunits are therefore located at the
periphery of the complex. Consistent with this, Rpn9 is
known to be a nonessential subunit of the lid [10].
Interestingly, it has been shown previously that Rpn10, which
stabilizes the association of the lid and base, interacts with
both Rpn9 and Rpn12 [20]. This suggests that Rpn9 and
Rpn12 are located near the contact surface between the base
and the lid.
A very recent study has shown that an rpn6 temperature-
sensitive mutant grown at restrictive temperatures contained
a subcomplex comprising four out of the nine lid components,
Rpn5, Rpn8, Rpn9, and Rpn11 [26]. A similar study with an
rpn7 temperature-sensitive mutant demonstrated the pres-
ence of an Rpn5:6:8:9:11 subcomplex [25]. From our cross-
linking data, we could not conclude whether Rpn3:7 or Rpn5:9
were anchored to Rpn11. However the results from the rpn6
Figure 3. MS/MS of the Lid Subcomplex
(A) Spectrum showing the dissociation of the 34þ charge state.
(B) Expansion of the spectrum over the m/z range of 1,600–2,800. Three individual subunits are assigned Rpn5, Rpn6, and Rpn9 (orange, green, and blue
circles, respectively).
(C) Expansion of the 8,000–16,000 m/z region. Three different hetrotrimers are assigned: Rpn5:Rpn6:Rpn8, Rpn5:Rpn8:Rpn9, and Rpn6:Rpn8:Rpn9 (blue,
green, and orange stars, respectively). By extrapolation we can conclude that the subcomplex is composed of four subunits, Rpn5, Rpn6, Rpn8, and
Rpn9. Rpn8 is not observed in its monomeric form; however, because it exists in all three different hetrotrimers, we can conclude it adopts a central
position within the subcomplex.
(D) A schematic summarizes the structural data obtained from MS and MS/MS results. Red indicates core subunits, while purple and to a greater extent
blue specifies peripheral subunits.
DOI: 10.1371/journal.pbio.0040267.g003
Figure 4. Chemical Cross-Linking of the Lid Complex
(A) Bis-Tris 4%–12% denaturing gel, stained with Colloidal Blue, enables separation of proteins before () and after (þ) cross-linking revealed six new
bands, A-F. The absence of Sem1 in this analysis is attributed to its low molecular mass and is consistent with other analyses [10,25,56].
(B) A MALDI-TOF/TOF spectrum of trypsin digestion of band D. Three proteins were identified within this band: Rpn3 (yellow), Rpn5 (cyan), and Rpn8
(pink). Peaks labeled by asterisks were sequenced in order to identify the peptide unambiguously. The inset shows a low-energy collision-induced
dissociation of an Rpn5 peptide producing mainly C-terminal (y) sequence ions.
DOI: 10.1371/journal.pbio.0040267.g004
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671318
MS of the Proteasome Lid
and rpn7 mutants support the scenario in which Rpn5:9
recruits Rpn11 into the Rpn5:6:8:9:11 subcomplex. In addi-
tion, the rpn7 mutation prevented the incorporation of Rpn3,
Rpn7, and Rpn12 into the lid. These results, together with the
evidence that Rpn5:6:8:9 exists as a stable independent
subcomplex in solution (Figure 1), are consistent with the
Rpn5:6:8:9 subcomplex forming the scaffolding core of the lid.
Taking together our data with existing data from analysis
of mutants [25,26] and results of previous two-hybrid analysis
[20,22,23,46] (Table 3), we propose a detailed interaction map
(Figure 6). The fact that all the data could be integrated into a
single model increases our confidence in the MS analysis. In
the model, two clusters become apparent. The first structural
cluster includes Rpn5, Rpn6, Rpn8, Rpn9, and Rpn11, in
agreement with the cluster suggested previously [20]. Within
the second cluster Rpn3, Rpn7, Rpn12, and Sem1 are
included, in accord with recent observations [25,26]. The
link between the two subcomplexes is between Rpn5:Rpn3,
and not Rpn7 as suggested by the analysis of an rpn7 mutant
[25]. These results explain previous observations in Schizo-
saccharomyces pombe that in cells in which rpn5 is deleted the
26S proteasomes misassemble [47]. Moreover, our structural
organization indicates that Rpn11 forms extensive interac-
tions with all three subunits, Rpn5, Rpn8, and Rpn9.
Furthermore, in our model Rpn6 is opposite Rpn9 and
Rpn12. As these may interact with Rpn10 at the lid–base
junction, it follows that Rpn6 is exposed to the cellular
environment.
Given the significant genetic similarity shared between the
lid and the COP9 signalosome, we investigated their
structural homology. Table 4 summarizes the protein–protein
interactions identified within the CSN subunits. Out of the 21
interactions detected, nine were observed among the 19S lid
complex (Figure 6), and eight of the interactions, although
not physically recognized, are feasible with our current
model, while 4 interactions are not in correspondence with
the lid structure. We speculate that the dense web of protein–
protein interactions detected within the CSN complex could
also be due to the tendency of the two-hybrid system to
produce false positives [22]. However, it is interesting to note
that six of the nine similar interactions and two of the eight
feasible interactions are found within the Rpn5, Rpn6, Rpn8,
Rpn9, and Rpn11 cluster. It is also worth noting that a 2-D
electron microscopy study [48] did not deduce a common
architecture for the two complexes. However, since both
particles have similar sizes, show asymmetric arrangements of
their subunits, possess a central channel, and share a large
number of similar interactions, the two particles are likely to
possess some common structural features.
The major function of the lid is determined by Rpn11, a
specialized isopeptidase that tightly couples the deubiquiti-
nation and degradation of substrates [5–7]. So far, no
catalytic activity has been described for any other subunit.
A recent study suggests that the role of these subunits is of a
scaffold for the other binding partners [49], namely Rpn11,
and the base subunits as well as the proteolyic substrates and
soluble cofactors. Interestingly, Rpn11 is unable to fold as an
Figure 5. An Interaction Map Summarizing the Data Obtained from Both
Native MS and Cross-Linking MS Analysis
Rpn5, Rpn6, Rpn8, and Rpn9 form a tetrameric structure, in which Rpn8
occupies a central position. The dashed line circling Rpn6 indicates that it
is present in substoichiometric amounts. Both Rpn7 and Rpn3 bind to
Sem1. Considering the small size of Sem1, it is reasonable to suggest that
Rpn3 and Rpn7 are in close proximity. In addition, we could determine that
Rpn3 binds to Rpn5. Rpn11 interacts with Rpn3:7 or Rpn5:9; these possible
interactions are labeled with a dashed line. No Interactions were
determined for Rpn12. Subunits are colored either red, purple, or blue,
indicating their increasing exposure within the complex.
DOI: 10.1371/journal.pbio.0040267.g005
Table 3. Subunit–Subunit Interactions within the 19S Lid from
Saccharomyces cerevisiae
Pair Species Method References
Rpn3-Sem1 Sc, MSc This study
Rpn7-Sem1 Sc, MSc This study
Rpn3-Rpn7 Sc 2h, MSc [20,23,25,46], and this study
Rpn3-Rpn5 Sc MSc This study
Rpn3-Rpn12 Sc 2h [20]
Rpn5-Rpn6 Sc 2h [20,26]
Rpn5-Rpn8 Sc 2h, MSn [20], this study
Rpn5-Rpn9 Sc 2h [20]
Rpn5-Rpn11 Sc 2h [20]
Rpn6-Rpn8 Sc MSn This study
Rpn8-Rpn9 Sc, Ce 2h, MSn [20,46], this study
Rpn8-Rpn11 Sc, Ce 2h [20,22,46]
Rpn9-Rpn11 Sc, Ce 2h [20,46]
Species: Ce, C. elegans; Sc, S. cerevisiae; methods: 2h, yeast two-hybrid; MSn, native mass
spectrometry; MSc, cross-linking mass spectrometry.
DOI: 10.1371/journal.pbio.0040267.t003
Table 2. Cross-Linked Subunits Obtained from Bands A–F
Band Cross-Linked Proteins
A Rpn7-Sem1
B Rpn3-Sem1
C Rpn3-Rpn5
D Rpn3-Rpn5-Rpn8
E Rpn3-Rpn5-Rpn8-Rpn9
F Rpn3-Rpn5-Rpn7-Rpn9-Rpn11
DOI: 10.1371/journal.pbio.0040267.t002
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671319
MS of the Proteasome Lid
isolated subunit into a native, active conformation [6].
Neighboring subunit interactions are probably required to
position conjugates for cleavage by Rpn11. The multiple
protein–protein interactions observed in this study for S.
cerevisiae Rpn11 are in accordance with this assumption.
Given that the only interaction observed linking the two
protein clusters in the lid is between Rpn3 and Rpn5, this
implies that there is a flexible hinge region which allows
movement necessary for productive interaction of substrates
with Rpn11. Taking into consideration the fact that Rpn6 is
present at substoichiometric amounts, and that Rpn11 forms
extensive protein interactions, it is interesting to speculate
that after substrate binding, Rpn6 dissociates to expose active
Rpn11. Support for our hypothesis comes from the finding
that Rpn6 is essential for assembly of the lid; however, rpn6
mutants do not affect the activity of 26S proteasome after
assembly [26]. This may imply that Rpn6 acts as a fail-safe
mechanism, at least in S. cerevisiae, ensuring that active sites
are exposed only after assembly is completed. Given the high
natural abundance of proteasome particles in the cell and the
need to prevent indiscriminate proteolysis of proteins, the
masking of active sites before full assembly could be an
essential regulatory mechanism. We anticipate that further
examination of the levels of Rpn6 within the 19S lid complex
will validate this scenario.
In summary, the results presented here extend significantly
previous models based on pairwise interactions and allow
interesting comparison with the structure of the signalosome.
Our model is consistent with a core assembly of four subunits
that provides a scaffold for recruitment of additional
subunits. Also apparent from our analysis is the fact that
subunits are assembled into two protein clusters, possibly
providing a cleft in which a polyubiquitinated substrate can
be accommodated. Based on our finding of substoichiometric
quantities of Rnp6, and its position within the complex, we
speculate that a regulatory mechanism involving dissociation
of Rpn6 is responsible for the exposure of the active site of
Rpn11. Overall, therefore, the subunit architecture that we
have described will not only benefit further structural analysis
but has also provided insight into the function and regulation
of the 19S lid.
More generally, the results of this study highlight major
advantages over existing approaches for generating interac-
tion maps. For example, it is established that the yeast two-
hybrid approach is complicated by the fact that protein
interactions are queried pairwise. Interactions of subunits
within protein complexes, such as the anaphase-promoting
complex, can be underrepresented in two-hybrid datasets
because assembly of such complexes may require higher-order
interactions between multiple subunits. A second potential
problem is that endogenous proteins may facilitate two-
hybrid interactions that are not direct, and the presence of
endogenous copies of proteins may confuse the analysis [22].
The other existing approach that has been widely applied is to
generate mutants for functional analysis of this complex
[25,26]. This method can also be limited by the availability of
appropriate mutants, lackofknowledgeabouthowto
generate appropriate mutants, and interpreting the effects
of the mutations. The fundamental advantage of our approach
is that it is carried out with the wild-type endoge nous
complex, and heterogeneity and subcomplex formation are
immediately apparent from the spectra. Moreover, when
combined with gas-phase dissociation we are able to probe the
composition of protein subcomplexes and reveal peripheral
subunits leading to definition of the structural organization of
Table 4. Subunit–Subunit Interactions within the COP9
Pair Species Interaction
Type
References Similar Interaction
Observed in Lid
Csn1-Csn2 Sp, Hs 2h [24,48] (Rpn7-Rpn6)
Csn1-Csn3 Hs 2h, fb [24,48] Rpn7-Rpn3
Csn1-Csn4 At, Hs 2h [24] (Rpn7-Rpn5)
Csn1-Csn5 At, Hs 2h, fb [24,48] Rpn7-Rpn11
Csn1-Csn7 At 2h [20,24] Rpn7-Rpn9
Csn2-Csn3 Hs fb [48] Rpn6-Rpn3
Csn2-Csn5 Dm, Hs 2h, fb [24,48] (Rpn6-Rpn11)
Csn2-Csn6 Hs fb [48] Rpn6-Rpn8
Csn2-Csn7 Dm, Hs 2h, fb [24,48] (Rpn6-Rpn9)
Csn3-Csn4 At 2h [20] Rpn3-Rpn5
Csn3-Csn5 At 2h [20] Rpn3-Rpn11
Csn3-Csn7 At 2h [20] (Rpn3-Rpn9)
Csn3-Csn8 At 2h [20] Rpn3-Rpn12
Csn4-Csn5 At 2h [20,24] Rpn5-Rpn11
Csn4-Csn7 At, Dm 2h [20,24] Rpn5-Rpn9
Csn4-Csn8 At 2h [24] (Rpn5-Rpn12)
Csn5-Csn6 At 2h [20] Rpn11-Rpn8
Csn5-Csn7 Hs 2h, fb [20,48] Rpn11-Rpn9
Csn5-Csn8 At 2h [20] (Rpn11-Rpn12)
Csn6-Csn7 Hs fb [20,48] Rpn8-Rpn9
Csn7-Csn8 At, Hs 2h, fb [20,24,48] (Rpn9-Rpn12)
A bold font indicates that a similar interaction was identified in the 19S lid (Fig 6B); an
italic font corresponds to an interaction that was not found in the 19S lid complex;
parenthesis designates a possible interaction that could exist in the lid complex.
At, Arabidopsis thaliana; Dm, Drosophila melanogaster; Hm, Homo sapiens; Sp, S. pombe;
2h, yeast two-hybrid; fb, filter binding.
DOI: 10.1371/journal.pbio.0040267.t004
Figure 6. Structural Organization of the 19S Lid
A plot summarizing the interactions within the lid, based on interactions
identified here in combination with those determined previously. The
dashed line circling Rpn6 indicates that it is present in substoiciometric
amounts. Rpn10 (in gray) is added to the model to emphasize the
orientation of the complex relative to the base and the cellular
environment. The two apparent clusters are labeled. Subunits are
colored either red, purple, or blue, indicating their increasing exposure.
DOI: 10.1371/journal.pbio.0040267.g006
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671320
MS of the Proteasome Lid
the complex. The fa ct th at the assembly state of this
asymmetric complex, with multiple distinct subunits, can be
determined and results incorporated into a comprehensive
model highlights the tremend ous potential of MS. We
anticipate that this approach will be used for determining
the organization of many other important cellular complexes
for which very little structural data exists.
Materials and Methods
Materials. Bis(sulfosuccinimidyl) (BS
3
) suberimidate was purchased
from Pierce Biotechnolo gy (Rock ford, Illinois, United Stat es).
Modified trypsin and RNasin were obtained from Promega (Madison,
Wisconsin, United States). Centricon YM10 centrifugal filters were
purchased from Millipore (Billerica, Massachusetts, United States)
and Nanosep centrifugal devices were purchased from Pall (East Hills,
New York, United States). Other reagents and solvents used were
analytical reagents or high-performance liquid chromatography
grade where appropriate, and were purchased from Sigma-Aldrich
(Saint Louis, Missouri, United States).
Protein expression and purification. Plasmid pJS-TM53H [50] was
modified by overlap PCR to replace the Myc epitopes with 12
histidine codons, resulting in plasmid pT2H12. The S. cerevisiae
RJD415 strain (MATa can1–100 leu2–3,112 his3 trp1–1 ura3–1 ade2–1
pep4::TRP1 bar1::LEU2) was transformed with a PCR product from a
reaction with pT2H12 as the template using a forward primer
homologous to the 42 nucleotides preceding the stop codon of
RPN11, and a reverse primer with 41 nucleotides that are
homologous to the region following the stop codon. The last 21
nucleotides of the forward and reverse primers are respectively
homologous to the first and last regions of the tagging cassette of
pT2H12. Transformants were selected on synthetic dropout–HIS
agar plates. Integrants were screened for by nickel precipitation using
Ni-NTA Agarose (Qiagen, Valencia, California, United States)
followed by Western blot analysis of the eluates with anti-His and
anti-Pad1 (S. pombe Rpn11) antibodies. A positive transformant was
selected and designated as RJD2909.
The RJD2909 strain was grown in yeast/peptone/dextrose (YPD)
medium, and levels of Rpn11
H12
were monitored through late
stationary phase by nickel precipitation and detection by anti-His
antibodies. Expression levels of Rpn11
H12
remained constant through
all phases of growth. A large (170 l) growth of RJD2909 in YPD was
performed at the UCLA Fermentor Facility. The cells were harvested at
an OD
600
of 18.4 yielding a 1.7 kg cell pellet and were frozen at 80 8C.
A cell pellet of 70 g was thawed in 140 ml of chilled lysis buffer (25
mM Tris [pH 7.5], 200 mM NaCl, 20 mM imidazole, 10 mM b-
mercaptoethanol, 0.3% Triton X-100, 50 mM NaF, 1 mM PMSF) and 3
complete EDTA-free protease inhibitor cocktail tab lets (Roche,
Indianapolis, Indiana, United States), and added to 200 ml of chilled
500-lm glass beads inside a stainless steel chamber of a BeadBeater
(BioSpec Products, Bartlesville, O klahoma, United States). The
chamber jacket was filled with ice water, and the BeadBeater was
run for 1 min and allowed to cool for 1 min for six cycles at 4 8C. The
contents of the chamber were applied with vacuum to a filter unit
(Nalgene, Rochester, New York, United States) equipped with a 180-
lm nylon net filter (Millipore) to separate the glass beads from the
supernatant. The filtrate was subsequently spun at 5,000 rpm in an
Eppendorf 5804 centrifuge (Eppendorf, Westbury, New York, United
States) for 20 min to remove intact cells and precipitated material.
The cell lysis solution was applied to a column of Ni-NTA
Superflow resin (Qiagen) at a flow rate of 2 ml/min. The resin was
subsequently washed in wash buffer (25 mM Tris [pH 7.5], 500 mM
NaCl, 0.3% Triton X-100, 50 mM imidazole, 10 mM b-mercaptoe-
thanol) and starting buffer (25 mM Tris [pH 7.5], 200 mM NaCl, 20
mM imidazole, 10 mM b-mercaptoethanol), and the bound protein
was eluted with elution buffer (25 mM Tris [pH 7.5], 100 mM NaCl,
300 mM imidazole, 10 mM b-mercaptoethanol). Peak fractions were
collected and concentrated to a final volume of 2 ml by ultra-
centrifu gation in a Centric on concentrator (Millipore), with a
molecular weight cutoff of 100 kDa.
The sample was loaded and run on a 26:60 Sephacryl S-400 gel
filtration column (Amersham Biosciences, Little Chalfont, United
Kingdom) at a flow rate of 0.5 ml/min with gel filtration buffer (50
mM Tris [pH 7.5], 200 mM NaCl, 5% glycerol, 1 mM DTT). The lid
particle elutes in three peaks at 156 ml (peak 1), 164 ml (peak 2), and
196 ml (peak 3). Based on calculations derived from calibration
standards, the theoretical molecular masses of the three peaks are
1,800 kDa for peak 1, 200 kDa for peak 2, and 500 kDa for peak 3. In
peaks 1 and 2, the lid particle elutes as part of the 19S proteasome
regulatory particle. In peak 3, the lid particle elutes in a relatively
pure form. Fractions corresponding to peaks 2 and 3 were
concentrated to ;200 ll by ultracentrifugation in an Amicon
Ultra-15 concentrator (Millipore) with a molecular weight cutoff of
100 kDa. The protein concentrations of peak 2 and 3 were 4 mg/ml
and 10 mg/ml, respectively, as determined by a DC protein assay (Bio-
Rad, Hercules, California, United States).
MS conditions. To confirm the protein composition of the lid
complex the sample was analysed by established proteomics methods
using a high-resolution MALDI time of flight (TOF)/TOF (Applied
Biosystems, Foster City, California, United States), and a C18 75 lm 3
15 cm column on an UltiMate capillary system (Dionex, Sunnyvale,
California, United States) connected to a Probot fraction collector
(Dionex). All 9 subunits were identified by automated MALDI-TOF/
TOF peptide mapping and sequence database searching of S. cerevisiae
proteins. Electrospray ionization MS and MS/MS experiments were
conducted on a high mass quadrupole TOF–type instrument [42,51]
adapted for a QSTAR XL platform, [42,51]. In this instrument a flow-
restricting sleeve is installed in the front part of the Q
0
to increase
the pressure locally and allow collisional cooling of heavy ions. In
addition, the low-frequency extended mass range of Q
1
permits the
isolation of ions up to 35,000 m/z.
For analysis of the intact complex, prior to mass spectrometry 25
ll of the 25 lM sample of the lid complex was buffer exchanged using
a Nanosep centrifugal device with a molecular weight cutoff of 10
kDa into 1 M ammonium acetate. For accurate mass determination of
the individual protein subunits, the sample was washed and eluted
from a C
4
ZipTip column (Millipore) under denaturing conditions
(50% acetonitrle and 0.1% formic acid). All other mass spectra were
recorded from aqueous solution conditions in 1 M ammonium
acetate. Typically, 1.5 ll of solution was electrosprayed from gold-
coated borosilicate capillaries prepared in-house as described [52].
The following experimental parameters were used: capillary voltage
up to 1.2 kV; declustering potential, 150 V; focusing potential, 250V;
second declustering potential, 55V; and collision energy up to 200 V,
microchannel plate 2350 V. In MS/MS the relevant m/z value was
selected in the quadrupole and argon was used as a collision gas at
maximum pressure. All spectra were calibrated externally by using a
solution of cesium iodide (100 mg/ml). Spectra are shown here with
minimal smoothing and without background subtraction.
Assignment of peaks in mass spectra. The masses of the individual
subunits and the complexes were calculated from the spectra
according to a method in which the charge is iterated over the
measured mass value to determine the best fit to the experimental
data [53]. The accurate masses of the individual subunits were
determined from a denatured solution of the complex. The molecular
masses of Rpn5, Rpn6, Rpn7, Rpn8, Rpn9, Rpn11, and Rpn12 (Table
1) are consistent with those reported in the database and consistent
with the N-terminal modifications previously reported [54]. We could
also clearly identify that the initiator methionine is removed and the
N-terminal amino group is acetylated in the Sem1 subunit. In
addition, from the mass measured for Rpn3 we could determine that
the first 16 residues of this subunit are truncated.
Cross-linking reactions and tryptic digestion. Solutions of the
cross-linking reagent BS
3
were prepared immediately before use at a
concentration of 10 mM in buffer (200 mM NaCl, HEPES 50 mM [pH
7.5]). The cross-linking reaction was performed with a lid concen-
tration of 4.8 lM and BS
3
ratio of 1:50 for 2 h on ice. The reaction was
terminated by addition of 2 M Tris-HCl (pH 7.4) to give a final
concentration of 50 mM Tris-HCl. SDS-PAGE gel analysis was
performed on Bis-Tris 4%–12% gels using the Novex Colloidal Blue
kit (Invitrogen, Carlsbad, California, United States). For analysis of
gel bands by in-gel digest, proteins were digested with 2% trypsin wt/
wt at 37 8C for 12 h in 50 mM ammonium bicarbonate [55].
MS/MS analysis of cross-linked peptides. The digestion mixture was
subjected to a Nano LC column coupled to a Probot MALDI spotter
followed by MS and MS/MS analysis using air as collision gas on a
MALDI-TOF/TOF instrument (Applied Biosystems). Resolution on
this instrument was within 30 ppm and MS data were acquired using
software from Applied Biosystems. Typically, 5,000 spectra were
acquired for each sample in MS or each MS/MS experiment.
Throughout the assignment a set of four conditions were applied:
(1) the peak intensity had to be 10%; (2) only peptides identical to
those observed in the absences of cross-linking reagent were
considered; (3) each protein had to be represented by at least four
independent peptides; and (4) peptides for which there was more
than one amino acid composition possible had to be sequenced. The
assignment was validated further by analysis of the MALDI-MS/MS
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671321
MS of the Proteasome Lid
peptide sequencing spectra. Each protein had at least one peptide
confirmed by the MALDI-TOF/TOF sequence analysis. It was not
practical to apply these conditions to Sem1 due to its low molecular
weight (10 kDa). The presence of Sem1 was confirmed by identifying
two Sem1 peptides and sequencing by MALDI-MS/MS spectra.
Using our set of criteria for assignment, proteins identified in bands
A–F were considered cross-linked. In the case of band A three proteins,
Rpn3, Rpn7, and Sem1, were identified. Considering the apparent
migration mass for this band, the only combination of proteins
consistent with this mass was Rpn7-Sem1. Therefore, the presence of
Rpn3 is likely to arise from the two neighboring bands of native Rpn3
and band B. In all the other 5 bands the apparent masses deduced from
migration on the gel were within error, taking into account changes in
migration anticipated from cross-linking [44], and were therefore
entirely consistent with the cross-linked protein components.
Acknowledgments
We thank Rati Verma and Sarah Maslen for valuable assistance in the
preparation of this manuscript. CVR acknowledges the Walters-
Kundert trust.
Author contributions. MS, RJD, and CVR conceived and designed
the experiments. MS and TT performed the experiments. MS and TT
analyzed the data. XIA and RJD contributed reagents/materials/
analysis tools. MS, RJD, and CVR wrote the paper.
Funding. M.S. is grateful for funding from the European Molecular
Biology Organization and from the Wingate Scholarship.
Competing interests. The authors have declared that no competing
interests exist.
References
1. Hershko A, Ciechanover A (1998) The ubiquitin system. Annu Rev Biochem
67: 425–479.
2. Voges D, Zwickl P, Baumeister W (1999) The 26S proteasome: A molecular
machine designed for controlled proteolysis. Annu Rev Biochem 68: 1015–1068.
3. Walz J, Erdmann A, Kania M, Typke D, Koster AJ, et al. (1998) 26S
proteasome structure revealed by three-dimensional electron microscopy. J
Struct Biol 121: 19–29.
4. Baumeister W, Walz J, Zuhl F, Seemuller E (1998) The proteasome:
Paradigm of a self-compartmentalizing protease. Cell 92: 367–380.
5. Verma R, Aravind L, Oania R, McDonald WH, Yates JR 3rd, et al. (2002)
Role of Rpn11 metalloprotease in deubiquitination and degradation by the
26S proteasome. Science 298: 611–615.
6. Yao T, Cohen RE (2002) A cryptic protease couples deubiquitination and
degradation by the proteasome. Nature 419: 403–407.
7. Guterman A, Glickman MH (2004) Complementary roles for Rpn11 and
Ubp6 in deubiquitination and proteolysis by the proteasome. J Biol Chem
279: 1729–1738.
8. Verma R, Chen S, Feldman R, Schieltz D, Yates J, et al. (2000) Proteasomal
proteomics: Identification of nucleotide-sensitive proteasome-interacting
proteins by mass spectrometric analysis of affinity-purified proteasomes.
Mol Biol Cell 11: 3425–3439.
9. Sone T, Saeki Y, Toh-e A, Yokosawa H (2004) Sem1p is a novel subunit of
the 26 S proteasome from Saccharomyces cerevisiae. J Biol Chem 279: 28807–
28816.
10. Glickman MH, Rubin DM, Fried VA, Finley D (1998) The regulatory
particle of the Saccharomyces cerevisiae proteasome. Mol Cell Biol 18: 3149–
3162.
11. Glickman MH, Rubin DM, Coux O, Wefes I, Pfeifer G, et al. (1998) A
subcomplex of the proteasome regulatory particle required for ubiquitin-
conjugate degradation and related to the COP9-signalosome and eIF3. Cell
94: 615–623.
12. Braun BC, Glickman M, Kraft R, Dahlmann B, Kloetzel PM, et al. (1999) The
base of the proteasome regulatory particle exhibits chaperone-like activity.
Nat Cell Biol 1: 221–226.
13. Rubin DM, Glickman MH, Larsen CN, Dhruvakumar S, Finley D (1998)
Active site mutants in the six regulatory particle ATPases reveal multiple
roles for ATP in the proteasome. EMBO J 17: 4909–4919.
14. Elsasser S, Finley D (2005) Delivery of ubiquitinated substrates to protein-
unfolding machines. Nat Cell Biol 7: 742–749.
15. Cope GA, Deshaies RJ (2003) COP9 signalosome: A multifunctional
regulator of SCF and other cullin-based ubiquitin ligases. Cell 114: 663–
671.
16. Schwechheimer C (2004) The COP9 signalosome (CSN): An evolutionary
conserved proteolysis regulator in eukaryotic develop ment. Biochim
Biophys Acta 1695: 45–54.
17. Wei N, Deng XW (2003) The COP9 signalosome. Annu Rev Cell Dev Biol 19:
261–286.
18. Ambroggio XI, Rees DC, Deshaies RJ (2004) JAMM: A metalloprotease-like
zinc site in the proteasome and signalosome. PLoS Biol 2: e2. DOI: 10.1371/
journal.pbio.0020002
19. Tran HJ, Allen MD, Lowe J, Bycroft M (2003) Structure of the Jab1/MPN
domain and its implications for proteasome function. Biochemistry 42:
11460–11465.
20. Fu H, Reis N, Lee Y, Glickman MH, Vierstra RD (2001) Subunit interaction
maps for the regulatory particle of the 26S proteasome and the COP9
signalosome. EMBO J 20: 7096–7107.
21. Ferrell K, Wilkinson CR, Dubiel W, Gordon C (2000) Regulatory subunit
interactions of the 26S proteasome, a complex problem. Trends Biochem
Sci 25: 83–88.
22. Uetz P, Giot L, Cagney G, Mansfield TA, Judson RS, et al. (2000) A
comprehensive analysis of protein-protein interactions in Saccharomyces
cerevisiae. Nature 403: 623–627.
23. Ito T, C hiba T, Ozawa R, Yoshida M, Hattori M, et al. (2001) A
comprehensive two-hybrid analysis to explore the yeast protein inter-
actome. Proc Natl Acad Sci U S A 98: 4569–4574.
24. Hartmann-Petersen R, Tanaka K, Hendil KB (2001) Quaternary structure
of the ATPase complex of human 26S proteasomes determined by chemical
cross-linking. Arch Biochem Biophys 386: 89–94.
25. Isono E, Saeki Y, Yokosawa H, Toh-e A (2004) Rpn7 Is required for the
structural integrity of the 26 S proteasome of Saccharomyces cerevisiae. J Biol
Chem 279: 27168–27176.
26. Isono E, Saito N, Kamata N, Saeki Y, Toh EA (2005) Functional analysis of
Rpn6p, a lid component of the 26 S proteasome, using temperature-
sensitive rpn6 mutants of the yeast Saccharomyces cerevisiae. J Biol Chem 280:
6537–6547.
27. Loo JA, Berhane B, Kaddis CS, Wooding KM, Xie Y, et al. (2005)
Electrospray ionization mass spectrometry and ion mobility analysis of
the 20S proteasome complex. J Am Soc Mass Spectrom. 16: 998–1008.
28. McCammon MG, Hernandez H, Sobott F, Robinson CV (2004) Tandem
mass spectrometry defines the stoichiometry and quaternary structural
arrangement of tryptophan molecules in the multiprotein complex TRAP. J
Am Chem Soc 126: 5950–5951.
29. Rostom AA, Robinson CV (1999) Detection of the intact GroEL chaperonin
assembly by mass spectrometry. J Am Chem Soc 121: 4718–4719.
30. Ilag LL, Westblade LF, Deshayes C, Kolb A, Busby SJ, et al. (2004) Mass
spectrometry of Escherichia coli RNA polymerase: Interactions of the core
enzyme with sigma70 and Rsd protein. Structure (Camb) 12: 269–275.
31. Ilag LL, Videler H, McKay AR, Sobot t F, Fucini P, et al. (2005) Heptameric
(L12)6/L10 rather than canonical pentameric complexes are found by
tandem MS of intact ribosomes from thermophilic bacteria. Proc Natl Acad
Sci U S A 102: 8192–8197.
32. Hanson CL, Fucini P, Ilag LL , Nierhaus KH, Robinson CV (2003)
Dissociation of intact Escherichia coli ribosomes in a mass spectrometer.
Evidence for conformational change in a ribosome elongation factor G
complex. J Biol Chem 278: 1259–1267.
33. Herna
´
ndez H, Dziembows k A, Taverner T, Se´raphin B, Robinson CV (2006)
Subunit architecture of multimeric complexes isolated directly from cells.
EMBO Rep 7: 605–610.
34. Sharon M, Witt S, Felderer K, Rockel B, Baumeister W, et al. (2006) 20S
proteasomes have the potential to keep substrates in store for continual
degradation. J Biol Chem 281: 9569–9575.
35. Sobott F, McCammon MG, Robinso n CV (2003) Gas-phase dissociation
pathways of a tetrameric protein complex. Int J Mass Spectro 230: 193–200.
36. Jurchen JC, Williams ER (2003) Origin of asymmetric charge partitioning in
the dissociation of gas-phase protein homodimers. J Am Chem Soc 125:
2817–2826.
37. Smith RD, Lightwahl KJ, Winger BE, Loo JA (1992) Preservation of
noncovalent associations in electrospray ionization mass-spectrometry-
multiply charged polypeptide and protein dimers. Org Mass Spectrom 27:
811–821.
38. Versluis C, van der Staaij A, Stokvis E, Heck AJ, de Craene B (2001)
Metastable ion formation and disparate charge separation in the gas-phase
dissection of protein assemblies studied by orthogonal time-of-flight mass
spectrometry. J Am Soc Mass Spectrom 12: 329–336.
39. Benesch JL, Robinson CV (2006) Mass spectrometry of macromolecular
assemblies: Preservation and dissociation. Curr Opin Struct Biol 16: 245–251.
40. Loo JA (1997) Studying noncovalent protein complexes by electrospray
ionization mass spectrometry. Mass Spectrom Rev 16: 1–23.
41. Robinson CV, Chung EW, Kragelund BB, Knudsen J, Aplin RT, et al. (1996)
Probing the nature of noncovalent interactions by mass spectrometry. A
study of protein-CoA ligand binding and assembly. J Am Chem Soc 118:
8646–8653.
42. Chernushevich IV, Thomson BA (2004) Collisional cooling of large ions in
electrospray mass spectrometry. Anal Chem 76: 1754–1760.
43. Fandrich M, Tito MA, Leroux MR, Rostom AA, Hartl FU, et al. (2000)
Observation of the noncovalent assembly and disassembly pathways of the
chaperone complex MtGimC by mass spectrometry. Proc Natl Acad Sci U S
A 97: 14151–14155.
44. Rappsilber J, Siniossoglou S, Hurt EC, Mann M (2000) A generic strategy to
analyze the spatial organization of multi-protein complexes by cross-
linking and mass spectrometry. Anal Chem 72: 267–275.
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671322
MS of the Proteasome Lid
45. Benesch JLP, Aquilina JA, Ruotolo BT, Sobott F, Robinson CV (2006)
Tandem mass spectrometry reveals the quaternary organization of macro-
molecular assemblies. Chem Biol 13: 597–605.
46. Davy A, Bello P, Thierry-Mieg N, Vaglio P, Hitti J, et al. (2001) A protein-
protein interaction map of the Caenorhabditis elegans 26S proteasome.
EMBO Rep 2: 821–828.
47. Yen HC, Gordon C, Chang EC (2003) Schizosaccharomyces pombe Int6 and Ras
homologs regulate cell division and mitotic fidelity via the proteasome. Cell
112: 207–217.
48. Kapelari B, Bech-Otschir D, Hegerl R, Schade R, Dumdey R, et al. (2000)
Electron microscopy and subunit-subunit interact ion studies reveal a first
architecture of COP9 signalosome. J Mol Biol 300: 1169–1178.
49. Scheel H, Hofmann K (2005) Prediction of a common structural scaffold
for proteasome lid, COP9-signalosome and eIF3 complexes. BMC Bio-
informatics 6: 71.
50. Seol JH, Shevchenko A, Deshaies RJ (2001) Skp1 forms multiple protein
complexes, including RAVE, a regulator of V-ATPase assembly. Nat Cell
Biol 3: 384–391.
51. Sobott F, Hernandez H, McCammon MG, Tito MA, Robinson CV (2002) A
tandem mass spectrometer for improved transmission and analysis of large
macromolecular assemblies. Anal Chem 74: 1402–1407.
52. Nettleton EJ, Sunde M, Lai Z, Kelly JW, Dobson CM, et al. (1998) Protein
subunit interactions and structural integrity of amyloidogenic trans-
thyretins: Evidence from electrospray mass spectrometry. J Mol Biol 281:
553–564.
53. Tito MA, Tars K, Valegard K, Hajdu J, Robinson CV (2000) Electrospray
time-of-flight mass spectrometry of the intact MS2 virus capsid. J Am Chem
Soc 122: 3550–3551.
54. Kimura Y, Saeki Y, Yokosawa H, Polevoda B, Sherman F, et al. (2003) N-
Terminal modifications of the 19S regulatory particle subunits of the yeast
proteasome. Arch Biochem Biophys 409: 341–348.
55. Rosenfeld J, Capdevielle J, Guillemot JC, Ferrara P (1992) In-gel digestion of
proteins for internal sequenc e analysis after one- or two-dimensional gel
electrophoresis. Anal Biochem 203: 173–179.
56. Wendler P, Lehmann A, Janek K, Baumgart S, Enenkel C (2004) The
bipartite nuclear localization sequence of Rpn2 is required for nuclear
import of proteasomal base complexes via karyopherin alphabeta and
proteasome functions. J Biol Chem 279: 37751–37762.
PLoS Biology | www.plosbiology.org August 2006 | Volume 4 | Issue 8 | e2671323
MS of the Proteasome Lid
... Our knowledge of lid assembly comes almost entirely from studies in budding yeast [49][50][51][52][53][54][55][56], and in contrast to the CP and the base, the lid appears to assemble without the need for any dedicated assembly factors. In yeast, lid assembly begins with the formation of two complementary precursors via parallel paths ( Figure 3) [50,[52][53][54]. ...
... Our knowledge of lid assembly comes almost entirely from studies in budding yeast [49][50][51][52][53][54][55][56], and in contrast to the CP and the base, the lid appears to assemble without the need for any dedicated assembly factors. In yeast, lid assembly begins with the formation of two complementary precursors via parallel paths ( Figure 3) [50,[52][53][54]. These precursors then combine, followed by joining of one final subunit to form the mature subcomplex. ...
... One path is initiated by dimerization of Rpn8 and Rpn11, which then recruits Rpn6 [52]. This trimeric complex then recruits Rpn5 and Rpn9 to form an intermediate referred to as Module 1 [54]. In the second path, Rpn3 and Rpn7 are tethered together by Sem1 to form an intermediate called lid particle 3 (LP3) [50]. ...
Article
Full-text available
The 26S proteasome is the largest and most complicated protease known, and changes to proteasome assembly or function contribute to numerous human diseases. Assembly of the 26S proteasome from its ~66 individual polypeptide subunits is a highly orchestrated process requiring the concerted actions of both intrinsic elements of proteasome subunits, as well as assistance by extrinsic, dedicated proteasome assembly chaperones. With the advent of near-atomic resolution cryo-electron microscopy, it has become evident that the proteasome is a highly dynamic machine, undergoing numerous conformational changes in response to ligand binding and during the proteolytic cycle. In contrast, an appreciation of the role of conformational dynamics during the biogenesis of the proteasome has only recently begun to emerge. Herein, we review our current knowledge of proteasome assembly, with a particular focus on how conformational dynamics guide particular proteasome biogenesis events. Furthermore, we highlight key emerging questions in this rapidly expanding area.
... The obtained compounds may therefore have the potential to compensate for the unbalanced proteostasis found in aging and age-related diseases. non-ATPase subunits and a heterohexameric ring of ATPases (Rpt1-6), three of which are terminated with a conserved tripeptide sequence, the HbYX (hydrophobic-penultimate Tyrany amino acid) [9]. To allosterically stimulate gate-opening the 19S docks the C-terminal tails of Rpt subunits in the intersubunit pockets of the 20S α ring. ...
Article
Full-text available
Aging and age-related diseases are associated with a decline in the capacity of protein turnover. Intrinsically disordered proteins, as well as proteins misfolded and oxidatively damaged, prone to aggregation, are preferentially digested by the ubiquitin-independent proteasome system (UIPS), a major component of which is the 20S proteasome. Therefore, boosting 20S activity constitutes a promising strategy to counteract a decrease in total proteasome activity during aging. One way to enhance the proteolytic removal of unwanted proteins appears to be the use of peptide-based activators of the 20S. In this study, we synthesized a series of peptides and peptidomimetics based on the C-terminus of the Rpt5 subunit of the 19S regulatory particle. Some of them efficiently stimulated human 20S proteasome activity. The attachment of the cell-penetrating peptide TAT allowed them to penetrate the cell membrane and stimulate proteasome activity in HEK293T cells, which was demonstrated using a cell-permeable substrate of the proteasome, TAS3. Furthermore, the best activator enhanced the degradation of aggregation-prone α-synuclein and Tau-441. The obtained compounds may therefore have the potential to compensate for the unbalanced proteostasis found in aging and age-related diseases.
... For example, CID was used to determine the relative binding strength and topology of 19S proteasome lid. 26 However, CID might result in the release of monomers with high charge state from a protein complex, such as 14-mer GroEL, 27 failing to provide enough structural information of the protein complex compared to another activation method, SID, that will be discussed next. ...
... For example, CID was used to determine the relative binding strength and topology of 19S proteasome lid. 26 However, CID might result in the release of monomers with high charge state from a protein complex, such as 14-mer GroEL, 27 failing to provide enough structural information of the protein complex compared to another activation method, SID, that will be discussed next. ...
Thesis
Full-text available
Mass spectrometry (MS) has become a powerful tool in biological sciences research. This dissertation reports the progress made in uncovering the effect of Salmonella on the intestinal environment by MS-based quantification and characterization of small molecules, and in characterizing a Salmonella deglycase that is a potential drug target. The MS- inspired discovery of protein biomarkers in Aspergillus fumigatus that causes invasive aspergillosis (IA) is also documented. Chapters 2, 3, and 4 focus on Salmonella enterica serovar Typhimurium, one of the major causes of food-borne diseases. The hypothesis of Chapter 2 is that the use of metabolomics, metagenomics, and metatranscriptomics will uncover the effect of Salmonella enterica serovar Typhimurium on the gastrointestinal environment, specifically in terms of biomolecules, microbiota, and host cells and tissues. Untargeted metabolite analysis using liquid chromatography-mass spectrometry (LC-MS), shows that the Salmonella-infected intestine has a significantly different distribution of nutrients and metabolites compared with that for uninfected animals, which is consistent with the shift in the distribution of microbes. The observed decrease of short-chain fatty acids (SCFA) was also validated by using targeted analysis, and is consistent with the decrease in Clostridia, a well-known SCFA producer. This multi-omics approach can be applied to discovering prebiotics and probiotics to prevent Salmonella infection. An intermediate of the fructose-asparagine (F-Asn) pathway, unique in Salmonella and few other species, has been found to be toxic to a fraB mutant of Salmonella. Results shown in Chapters 3 and 4 lay the foundation for a new strategy to specifically inhibit Salmonella using the combination of a FraB inhibitor and F-Asn. Chapter 3 describes the quantification of F-Asn in various foods, using a hydrophilic interaction liquid chromatography (LC) coupled to a triple-quadrupole mass spectrometer. F-Asn is present in a wide range of human and animal foods, and is transported to the gastrointestinal tract, where it can be consumed by Salmonella and a few other microorganisms. These results pave the way for the study of dosing of F-Asn in a drug cocktail. Chapter 4 focuses on studying the native structure and catalytic mechanism of FraB. Native MS analysis was performed on purified recombinant FraB. The oligomeric state, collision-cross section, dimer interface, and ligand-binding stoichiometry derived from MS data support a homology model proposed by our collaborators. MS also supports that various site-directed FraB mutants preserve the native structure, suggesting that the loss of enzymatic activity in these mutants does not result from global structural alterations. The successful detection of the binding between FraB and its substrate or products can be transferred to study FraB-inhibitor interaction. The discovery of protein biomarkers for Aspergillus fumigatus infection in IA diagnosis is discussed in Chapter 5. A reversed-phase/reversed-phase two-dimensional LC-MS was used for the shotgun proteomics of bronchoalveolar lavage (BAL) fluids from both mice and humans with Aspergillus fumigatus infection. Several proteins from Aspergillus fumigatus were detected, including thioredoxin reductase GliT. The protein has low sequence homology to mouse and human proteins and was found in both mice and human BAL samples, showing its great potential as a biomarker.
... Experiments with Arabidopsis conditional mutant strains revealed that developmental phenotypes caused by defects in the CSN do not necessarily correlate with the deneddylation activity of the complex (30). Establishment of two main clusters as intermediates for the assembly of heptameric complexes could be a common theme for Zomes complex formation (16,21,22,(31)(32)(33)(34)(35)(36)(37). Albeit there is no evidence for transcriptional control of A. nidulans csn genes (38), our data disclosed various cellular mechanisms, which collaborate to secure appropriate protein levels and correct subcellular localization for complex assembly. ...
Article
Full-text available
The conserved eight-subunit COP9 signalosome (CSN) is required for multicellular fungal development. The CSN deneddylase cooperates with the Cand1 exchange factor to control replacements of E3 ubiquitin cullin RING ligase receptors, providing specificity to eukaryotic protein degradation. Aspergillus nidulans CSN assembles through a heptameric pre-CSN, which is activated by integration of the catalytic CsnE deneddylase. Combined genetic and biochemical approaches provided the assembly choreography within a eukaryotic cell for native fungal CSN. Interactomes of functional GFP-Csn subunit fusions in pre-CSN deficient fungal strains were compared by affinity purifications and mass spectrometry. Two distinct heterotrimeric CSN subcomplexes were identified as pre-CSN assembly intermediates. CsnA-C-H and CsnD-F-G form independently of CsnB, which connects the heterotrimers to a heptamer and enables subsequent integration of CsnE to form the enzymatically active CSN complex. Surveillance mechanisms control accurate Csn subunit amounts and correct cellular localization for sequential assembly since deprivation of Csn subunits changes the abundance and location of remaining Csn subunits.
... The fully assembled 26S proteasome is a macromolecular complex consis ng of a 20S core par cle and one or two 19S regulatory par cles. The 19S is composed of two subcomplexes, the base which directly binds to the 20S, and the lid [5]. The base is formed by six AAA-ATPase subunits (Rpt1-6), and four non-ATPase subunits (Rpn1, Rpn2, Rpn10, and Rpn13). ...
Preprint
The 26S proteasome is the main proteolytic machine involved in protein degradation, thus contributing to homeostasis or stress response of eukaryotic cells. This macromolecular complex, consisting of a 20S core particle assembled with one or two 19S regulatory particles, is highly regulated by phosphorylation. Here we describe the Plasmodium berghei proteasome AAA-ATPase regulatory subunit Rpt3 and show that it binds to protein phosphatase 1, the major parasite phosphatase. In addition, PbRpt3 regulates the activity of the phosphatase both in vitro and in a heterologous model of Xenopus oocytes. Using mutagenesis approaches, we observed that the RVXF motifs of PbRpt3 are involved in this binding and activity. Further use of Xenopus oocyte model and mutagenesis based on the 3D model that we established revealed that the binding capacity of PbRpt3 to ATP may also contribute to its phosphatase-regulating activity. In the parasite, reverse genetic studies suggested an essential role for PbRpt3 since no viable knock-out line could be obtained. Additionally, immunoprecipitation assays followed by mass spectrometry analyses using transgenic PbRpt3-tagged parasites not only confirmed that PbRpt3 belongs to the 19S regulatory particle of the proteasome, but also revealed potential interaction with proteins already shown to play a role in the phospholipid membrane dynamics.
Chapter
The ubiquitin–proteasome system is an essential protein degradation machinery responsible for maintaining cellular protein homeostasis. It is a multi-subunit complex comprising of the 20S core particle and the 19S regulatory particle. Recent studies have highlighted the conservation of several subunits of the proteasome in Plasmodium, and the Plasmodium proteasome is presented as an attractive target for novel antimalarials. Several classes of proteasome inhibitors were developed from chemical libraries and were shown to have selectively potent activities against the parasite. Proteasome inhibitors have the potential to inhibit all the stages of the parasite as well as different strains, including the drug-resistant parasites. Proteasome inhibitors have also been shown to act synergistically with artemisinin, which might be useful in combating drug resistance. This chapter highlights the different classes of proteasome inhibitors developed including β-lactones, peptide aldehydes, α′β′ epoxyketone, asparagine ethylene diamines, peptide sulphonyl fluorides, peptide boronates, and cyclic peptides. This chapter also highlights how the proteasome inhibitors are validated through in vitro evolution techniques. Rapid progress in the field together with available cryo-EM structure of the Plasmodium 20S proteasome, shows great promise for developing proteasome inhibitors as novel antimalarials.
Article
The immunoproteasome is a specialized form of proteasome equipped with modified catalytic subunits that was initially discovered to play a pivotal role in MHC class I antigen processing and immune system modulation. However, over the last years, this proteolytic complex has been uncovered to serve additional functions unrelated to antigen presentation. Accordingly, it has been proposed that immunoproteasome synergizes with canonical proteasome in different cell types of the nervous system, regulating neurotransmission, metabolic pathways and adaptation of the cells to redox or inflammatory insults. Hence, studying the alterations of immunoproteasome expression and activity is gaining research interest to define the dynamics of neuroinflammation as well as the early and late molecular events that are likely involved in the pathogenesis of a variety of neurological disorders. Furthermore, these novel functions foster the perspective of immunoproteasome as a potential therapeutic target for neurodegeneration. In this review, we provide a brain and retina-wide overview, trying to correlate present knowledge on structure-function relationships of immunoproteasome with the variety of observed neuro-modulatory functions.
Article
Field-free capillary vibrating sharp-edge spray ionization (cVSSI) is evaluated for its ability to conduct native mass spectrometry (MS) experiments. The charge state distributions for nine globular proteins are compared using field-free cVSSI, field-enabled cVSSI, and electrospray ionization (ESI). In general, for both positive and negative ion mode, the average charge state (qavg) increases for field-free cVSSI with increasing molecular weight similar to ESI. A clear difference is that the qavg is significantly lower for field-free conditions in both analyses. Two proteins, leptin and thioredoxin, exhibit bimodal charge state distributions (CSDs) upon the application of voltage in positive ion mode; only a monomodal distribution is observed for field-free conditions. In negative ion mode, thioredoxin exhibits a multimodal CSD upon the addition of voltage to cVSSI. Extensive molecular dynamics (MD) simulations of myoglobin and leptin in nanodroplets suggest that the multimodal CSD for leptin may originate from increased conformational "breathing" (decreased packing) and association with the droplet surface. These properties along with increased droplet charge appear to play critical roles in shifting ionization processes for some proteins. Further exploration and development of field-free cVSSI as a new ionization source for native MS especially as applied to more flexible biomolecular species is warranted.
Article
Full-text available
The E. coli RNA polymerase core enzyme is a multisubunit complex of 388,981 Da. To initiate transcription at promoters, the core enzyme associates with a σ subunit to form holo RNA polymerase. Here we have used nanoflow electrospray mass spectrometry, coupled with tandem mass spectrometry, to probe the interaction of the RNA polymerase core enzyme with the most abundant σ factor, σ70. The results show remarkably well-resolved spectra for both the core and holo RNA polymerases. The regulator of σ70, Rsd protein, has previously been identified as a protein that binds to free σ70. We show that Rsd also interacts with core enzyme. In addition, by adding increasing amounts of Rsd, we show that σ70 is displaced from holo RNA polymerase, resulting in complexes of Rsd with core and σ70. The results argue for a model in which Rsd not only sequesters σ70, but is also an effector of core RNA polymerase.
Article
Full-text available
The COP9 signalosome (CSN) is composed of eight distinct subunits and is highly homologous to the lid sub-complex of the 26S proteasome. CSN was initially defined as a repressor of photomorphogenesis in Arabidopsis, and it has now been found to participate in diverse cellular and developmental processes in various eukaryotic organisms. Recently, CSN was revealed to have a metalloprotease activity centered in the CSN5/Jab1 subunit, which removes the post-translational modification of a ubiquitin-like protein, Nedd8/Rub1, from the cullin component of SCF ubiquitin E3 ligase (i.e., de-neddylation). In addition, CSN is associated with de-ubiquitination activity and protein kinase activities capable of phosphorylating important signaling regulators. The involvement of CSN in a number of cellular and developmental processes has been attributed to its control over ubiquitin-proteasome-mediated protein degradation.
Article
Full-text available
We examined the different steps necessary for the enzymatic digestion of proteins in the polyacrylamide matrix after gel electrophoresis. As a result, we developed an improved method for obtaining peptides for internal sequence analysis from 1-2 micrograms of in-gel-digested proteins. The long washing-lyophilization-equilibration steps necessary to eliminate the dye, sodium dodecyl sulfate, and other gel-associated contaminants that perturb protein digestion in Coomassie blue-stained gels have been replaced by washing for 40 min with 50% acetonitrile, drying for 10 min at room temperature, and then rehydrating with a protease solution. The washing and drying steps result in a substantial reduction of the gel slice volume that, when next swollen in the protease solution, readily absorbs the enzyme, facilitating digestion. The Coomassie blue staining procedure has also been modified by reducing acetic acid and methanol concentrations in the staining solution and by eliminating acetic acid in the destaining solution. The peptides resulting from the in-gel digestion are easily recovered by passive elution, in excellent yields for structural characterization. This simple and rapid method has been successfully applied for the internal sequence analysis of membrane proteins from the rat mitochondria resolved in preparative two-dimensional gel electrophoresis.
Article
The gas-phase dissociation of the tetrameric complex transthyretin (TTR) has been investigated with tandem-mass spectrometry (tandem-MS) using a nanoflow-electrospray interface and a quadrupole time-of-flight (Q-TOF) mass spectrometer. The results show that highly charged monomeric product ions dissociate from the macromolecular complex to form trimeric products. Manipulating the pressure conditions within the mass spectrometer facilitates the formation of metastable ions. These were observed for the transitions from tetrameric to monomeric and trimeric product ions and additionally for losses of small molecules associated with the protein complex in the gas phase. These results are interpreted in the light of recent mechanisms for the electrospray process and provide insight into the composition and factors governing the stability of macromolecular ions in the gas phase.
Article
A series of noncovalent complexes formed between the 86 residue acyl CoA binding protein (ACBP) and a series of acyl CoA derivatives has been studied by electrospray ionization mass spectrometry. Conditions were found under which CoA ligands can be observed in the mass spectrometer bound to ACBP. Despite the very low dissociation constants (10-7 to 10-10 M) of the acyl CoA ligand complexes high ratios of ligand-to-protein concentration in the electrospray solution were found to increase the proportion of intact complex observed in the spectrum. Variation in the length of the hydrophobic acyl chain of the ligand (C16, C12, C8, C0) resulted in similar proportions of complex observed in the mass spectrum even though significant variation in solution dissociation constants has been measured. A substantially reduced proportion of complex was, however, found for the mutant proteins, Y28N, Y31N, and Y73F, lacking tyrosine residues involved in critical interactions with the CoA ligand. These results have been interpreted in terms of the different factors stabilizing complexes in the gas phase environment of the mass spectrometer. The complexed species were also investigated by hydrogen−deuterium exchange methods combined with mass spectrometric analysis and the results show that folding of ACBP occurs prior to complex formation in solution. The results also show increased hydrogen exchange protection in the complex when compared with the free protein. Furthermore, even after dissociation of the complex, under these nonequilibrium gas phase exchange conditions, increased protection from hydrogen exchange in the complex is maintained.
Article
The ‘softness’ of the electrospray ionization (ESI) method provides a direct link between solution chemistry and the inherent gas-phase environment of mass Spectrometry. Available results related to the preservation of non-covalent associations into the gas phase after ESI are reviewed. These associations include the possible retention of elements of higher order protein structure, non-covalent polypeptide–heme associations and enzyme complexes. Experimental results are presented showing that non-covalently bound polypeptide and protein dimer ions are relatively common as low level contributions to ESI mass spectra. It is argued that these dimers are reflective of multimeric species in solution since Coulombic barriers preclude dimerization after ESI although uncertainty remains regarding whether they exist prior to the formation of highly charged droplets. The dissociation of dimers is facile and for proteins can yield monomers having a broad distribution of charge states. The detection of non-covalently associated dimers requires gentle ESI mass spectrometer interface conditions, yielding relatively low levels of internal excitation. Under such conditions incomplete molecular ion desolvation can result in experimental artifacts for tandem mass spectrometric experiments. ESI mass Spectrometry may have broad potential for the study of noncovalent liquid phase associations.
Article
Electrospray ionization mass spectrometry has been used to study protein interactions driven by noncovalent forces. The gentleness of the electrospray ionization process allows intact protein complexes to be directly detected by mass spectrometry. Evidence from the growing body of literature suggests that the ESI‐MS observations for these weakly bound systems reflect, to some extent, the nature of the interaction found in the condensed phase. Stoichiometry of the complex can be easily obtained from the resulting mass spectrum because the molecular weight of the complex is directly measured. For the study of protein interactions, ESI‐MS is complementary to other biophysical methods, such as NMR and analytical ultracentrifugation. However, mass spectrometry offers advantages in speed and sensitivity. The experimental variables that play a role in the outcome of ESI‐MS studies of noncovalently bound complexes are reviewed. Several applications of ESI‐MS are discussed, including protein interactions with metal ions and nucleic acids and subunit protein structures (quaternary structure). © 1997 John Wiley & Sons, Inc., Mass Spectrom Rev 16(1), 25–49, 1997
Article
In 26S proteasomes, "19S cap complexes" associate with either one or both ends of the barrel-shaped 20S core complex. These regulatory complexes which comprise about 20 different subunits, including 6 ATPases of the AAA family, are thought to recognize ubiquitinated substrate proteins, to dissociate and unfold them before threading them into the 20S core where they are degraded. Here, we examine the structure of 26S proteasomes from Drosophila embryos and Xenopus oocytes by electron microscopy. Image analysis reveals a rather flexible linkage between the 19S caps and the 20S core, with a peculiar wagging-type movement of the caps relative to the core. At this stage of the analysis, it is not clear whether this movement is relevant in terms of function. Three-dimensional reconstructions, taking this into account, provide first insights into the remarkably complex structure of the 19S caps and allows us to put forward a composite model of the entire 26S complex.