ArticlePDF Available

Genetic Analysis Reveals Domain Interactions of Arabidopsis Hsp100/ClpB and Cooperation with the Small Heat Shock Protein Chaperone System

Authors:

Abstract and Figures

We have defined amino acids important for function of the Arabidopsis thaliana Hsp100/ClpB chaperone (AtHsp101) in acquired thermotolerance by isolating recessive, loss-of-function mutations and a novel semidominant, gain-of-function allele [hot1-4 (A499T)]. The hot1-4 allele is unusual in that it not only fails to develop thermotolerance to 45 degrees C after acclimation at 38 degrees C, but also is sensitive to 38 degrees C, which is a permissive temperature for wild-type and loss-of-function mutants. hot1-4 lies between nucleotide binding domain 1 (NBD1) and NBD2 in a coiled-coil domain that is characteristic of the Hsp100/ClpB proteins. We then isolated two classes of intragenic suppressor mutations of hot1-4: loss-of-function mutations (Class 1) that eliminated the 38 degrees C sensitivity, but did not restore thermotolerance function to hot1-4, and Class 2 suppressors that restored acquired thermotolerance function to hot1-4. Location of the hot1-4 Class 2 suppressors supports a functional link between the coiled-coil domain and both NBD1 and the axial channel of the Hsp100/ClpB hexamer. In addition, the strongest Class 2 suppressors restored solubility of aggregated small heat shock proteins (sHsps) after heat stress, revealing genetic interaction of the Hsp100/ClpB and sHsp chaperone systems. These results also demonstrate that quantitative phenotypes can be used for in vivo genetic dissection of protein mechanism in Arabidopsis.
Content may be subject to copyright.
Genetic Analysis Reveals Domain Interactions of Arabidopsis
Hsp100/ClpB and Cooperation with the Small Heat Shock
Protein Chaperone System
W
Ung Lee,
a
Chris Wie,
a
Mindy Escobar,
a
Ben Williams,
a
Suk-Whan Hong,
b
and Elizabeth Vierling
a
,
1
a
Department of Biochemistry and Molecular Bioph ysics, Uni versit y of Arizona, Tucson, Ari zona 85721
b
Department of Appli ed Plant Sc ienc es, Agricultural Plant Stress Research Center,
Chonnam National Universi ty, Kwang Ju, 500-757, South Korea
We have defined amino acids important for function of the Arabidopsis thaliana Hsp100/ClpB chaperone (AtHsp101) in
acquired thermotolerance by isolating recessive, loss-of-function mutations and a novel semidominant, gain-of-function
allele [hot1-4 (A499T)]. The hot1-4 allele is unusual in that it not only fails to develop thermotolerance to 458C after
acclimation at 388C, but also is sensitive to 388C, which is a permissive temperature for wild-type and loss-of-function
mutants. hot1-4 lies between nucleotide binding domain 1 (NBD1) and NBD2 in a coiled-coil domain that is characteristic of
the Hsp100/ClpB proteins. We then isolated two classes of intragenic suppressor mutations of hot1-4: loss-of-function
mutations (Class 1) that eliminated the 388C sensitivity, but did not restore thermotolerance function to hot1-4, and Class 2
suppressors that restored acquired thermotolerance function to hot1-4. Location of the hot1-4 Class 2 suppressors
supports a functional link between the coiled-coil domain and both NBD1 and the axial channel of the Hsp100/ClpB
hexamer. In addition, the strongest Class 2 suppressors restored solubility of aggregated small heat shock proteins (sHsps)
after heat stress, revealing genetic interaction of the Hsp100/ClpB and sHsp chaperone systems. These results also
demonstrate that quantitative phenotypes can be used for in vivo genetic dissection of protein mechanism in Arabidopsis.
INTRODUCTION
Hsp100/ClpB chaperones are hexameric members of the AAAþ
family of proteins, which are ATPases that couple ATP binding
and hydrolysis to a variety of protein-remodeling activities
(Neuwald et al., 1999; Vale, 2000; Ogura and Wilkinson, 2001;
Lupas and Martin, 2002). Hsp100/ClpB proteins have been
shown to be essential for the development of thermotolerance
to high temperature in bacteria (Goloubinoff et al., 1999), yeast
(Glover and Lindquist, 1998), some parasitic protozoa, and
higher plants (Hong and Vierling, 2000, 2001), which are the
only higher eukaryotes known to express the Hsp100/ClpB
subgroup of chaperones. The demonstration that the highly
heat-induced Hsp100/ClpB protein in Arabidopsis thaliana,
AtHsp101, was essential for acquisition of tolerance to high
temperature also represented a direct link of a heat shock protein
to heat tolerance in plants. Both in vivo and in vitro evidence
indicates that the protective function of these chaperones is
a result of their ability to resolubilize protein aggregates in co-
operation with the Hsp70/DnaK chaperone system (Goloubinoff
et al., 1999; Zolkiewiski, 1999). It remains unknown, however,
whether general removal of aggregated proteins or rescue of
certain critical substrates defines the essential activity of
Hsp100/ClpB in thermotol erance.
This unique subgroup of ATPases contains two AAA modules,
each comprised of a nucleotide binding domain (NBD1 or 2) with
conserved Walker A, Walker B, and sensor motifs and a smaller
C-terminal domain. The AAA modules are flanked by additional
N- and C-terminal domains and are separated by a large coiled-
coil domain, which is actually an insertion into the C-terminal
small domain of AAA module 1 (see Figure 1A). The length of the
coiled-coil domain is a major feature that distinguishes Hsp100/
ClpB proteins from other Clp proteins with two AAA modules
(e.g., ClpA) (Schirmer et al., 1996; Celerin et al., 1998; Nieto-
Sotelo et al., 1999). The recent 3.0-A
˚
structure of a complete
Thermus thermophilus ClpB subunit, and a cryoelectron micros-
copy reconstruction of the hexamer (21-A
˚
resolution), has de-
fined the structure of the different ClpB domains and how they
are oriented relative to each other (Lee et al., 2003). In each AAA
module, NBD1 and 2 have a RecA-like mononucleotide binding
fold, and the C-terminal smaller domain has a mixed a/b
structure. The two AAA modules are stacked on each other
within a subunit such that the hexamer appears as two rings of
six ATPase sites, surrounding a pore or axial channel on the order
of
;16 A
˚
in diameter. The coiled-coil domain, which is inserted
into the smaller domain of NBD1, is 85-A
˚
long and resembles
a two-bladed propeller ringing the outside of the hexamer. Based
on the observation of three different conformations in the crystal
structure, the N-terminal domain and coiled-coil domain are
1
To whom correspondence should be addressed. E-mail vierling@
email.arizona.edu; fax 520-621-3709.
The author responsible for distribution of materials integral to the
findings presented in this article in accordance with the policy described
in the Instructions for Authors (www.plantcell.org) is: Elizabeth Vierling
(vierling@email.arizona.edu).
W
Online version contains Web-only data.
Article, publication date, and citation information can be found at
www.plantcell.org/cgi/doi/10.1105/tpc.104.027540.
The Plant Cell, Vol. 17, 559–571, February 2005, www.plantcell.org ª 2005 Ameri can Society of Plant Biologists
proposed to be mobile. This is consistent with the fact that
deletion of either domain does not disrupt the hexamer (Mogk
et al., 2003b).
Biochemical studies using both in vitro assays for ATP hydro-
lysis and protein disaggregation (chaperone activity), along with
in vivo assays for thermotolerance, have also begun to dissect
functional domains of Hsp100/ClpB (Smith et al., 1999; Barnett
et al., 2000; Clarke and Eriksson, 2000; Schirmer et al., 2001;
Beinker et al., 2002; Cashikar et al., 2002; Mogk et al., 2003b;
Strub et al., 2003). However, details of the Hsp100/ClpB reaction
cycle are unknown. The ATPase activity of both NBDs, as well as
an intact coiled-coil domain, are essential for chaperone activity
and in vivo thermotolerance (Kim et al., 2000a; Schirmer et al.,
2001; Hattendorf and Lindquist, 2002b; Mogk et al., 2003b). A
majority of data indicate that deletion of the N-terminal domain
does not impair measured activities, and its function remains
unknown (Clarke and Eriksson, 2000; Beinker et al., 2002). The
C-terminal domain is required for activity, but because deletion
of this domain prevents hexamerization, which is required for
ATPase activity, its direct role is not revealed by such experi-
ments. Studies with model substrates and effectors of the
ATPase activity have suggested that the C-terminal domain is
a substrate binding site (Smith et al., 1999), but others conclude
this domain modulates ATPase activity only (Strub et al., 2003).
The N-terminal domain, NBD1, and the coiled-coil domain have
all also been proposed to be involved in substrate interactions
(Dougan et al., 2002; Liu et al., 2002; Weibezahn et al., 2003).
However, the only data on binding of a natural substrate (TrfA to
Escherichia coli ClpB) reveals an interaction only with the surface
of NBD1 near the axial pore (Schlieker et al., 2004). Lindquist and
coworkers (Cashikar et al., 2002; Hattendorf and Lindquist,
2002a) proposed that in yeast Hsp104, stimulation of NBD2
activity drives a conformational change in the coiled-coil domain,
which in turn stimulates ATPase activity in NBD1. Cross-linking
studies with ClpB further suggest that movement of the coiled-
coil domain is necessary for chaperone activity in vitro (Lee et al.,
2003). How these motions are connected to protein disaggre-
gation remains to be defined. It has been suggested that ClpB
has a crowbar action in protein disaggregation, perhaps medi-
ated by the coiled-coil domain (Vale, 2000; Maurizi and Xia, 2004;
Weibezahn et al., 2004). Alternatively, or in addition, substrate
unfolding may be coupled to threading through the axial channel
of the ClpB hexamer in a mechanism related to that used by ClpA
(Ishikawa et al., 2001) and ClpX (Kim et al., 2000b) for protein
unfolding and transfer to the ClpP protease. Recent experiments
indicate mutations that affect protein loops in the axial channel
eliminate yeast Hsp104 function in protein disaggregation,
perhaps because of disruption of protein threading through the
channel (Lum et al., 2004).
We have taken a genetic approach to investigating the function
and mechanism of Hsp100/ClpB in Arabidopsis. In previous
work we showed that AtHsp101, encoded by the HOT1 gene, is
a major component controlling acquired thermotolerance, based
on the phenotypes of missense and protein null alleles, hot1-1
(E637K; in NBD2) and hot1-3 (a T-DNA insertion in exon 1),
respectively (Hong and Vierling, 2000, 2001). Loss of AtHsp101
function, however, does not appear to have major effects on
plant growth and development under optimal temperature
Figure 1. Location and Phenotype of Mutations in AtHsp101 (See Also
Supplemental Figure 1 Online).
(A) Location of hot1 mutations with a thermotolerance phenotype on
a schematic diagram of the AtHsp101 protein. The conserved AAA
modules consist of two domains, a nucleotide binding domain (NBD1 or
2), containing conserved motifs [Walker A, Walker B, Sensor 1(S1), and
Arg finger (R)], and a C-terminal small domain (gray boxes), containing
Sensor 2 (S2) in AAA2. The coiled-coil domain contains two signature
motifs (Lee et al., 2003), with hot1-4 located in signature motif II.
(B) Location of missense mutations that are wild type for thermotoler-
ance, which were obtained by Tilling analysis (see Methods).
(C) Quantitative assessment of thermotolerance in hot1 mutant seedlings
compared with their corresponding wild type (Columbia for hot1-1 and
hot1-4, Columbia erecta for all others). After growth for 2.5 d in the dark
at 228C, seedlings were pretreated at 388C for 90 min and returned to
228C for 2.5 d (388C samples), or pretreatment was followed by 2 h at
228C then 2 or 3 h at 458C before 2.5 d of recovery (388C>458C samples).
hot1-4 seedlings are more sensitive to 388C treatment than wild-type or
null alleles. Asterisk s indicate values equal to zero. Mean and standard
deviations were derived by measurement from three independent ex-
periments performed with 60 or more seedlings per mutant and values
plotted as a percentage of the 228C value.
(D) Protein gel blot analysis of AtHsp101 protein levels in wild-type and
hot1 seedlings after treatment at 388C.
560 The Plant Cell
conditions. To dissect further the mechanism of Hsp101 action in
thermotolerance, we sought additional mutant alleles either by
direct screening for mutants in a seedling thermotolerance assay
(Hong and Vierling, 2000) or by analysis of point mutations in
AtHsp101 obtained through the Arabidopsis Tilling Resource
(http://tilling.fhcrc.org:9366). We describe four new mutant al-
leles of Hsp101 that provide insight into structure and function of
this class of chaperone ATPase. One mutant allele, hot1-4, which
is a missense mutation in the coiled-coil domain (A499T) pro-
vides direct in vivo evidence for an essential function of this
unique domain. Using this allele, we then screened for suppres-
sor mutations and defined 13 intragenic suppressors. The
location of these suppressor mutations defines features essen-
tial for in vivo activity of Hsp101 and provides in vivo evidence for
a novel mechanistic link between the coiled-coil domain and the
axial channel of Hsp100/ClpB proteins. Furthermore, certain
suppressors restore the solubility of small heat shock proteins
(sHsps) after heat stress in the hot1-4 mutant, providing genetic
evidence for interaction of the Hsp100/ClpB and sHsp chaper-
one systems in higher plants.
RESULTS
AtHsp101 Mutants with Defects in Thermotoleran ce
Arabidopsis seedlings will continue to grow after a 458C, 2-h heat
treatment if first preconditioned at 388C for 90 min followed by 1
to 2 h at 228C. This ability to acclimate to 458C is lost in the
previously characterized missense (hot1-1; E637K) and protein
null alleles of AtHsp101 (hot1-3; T-DNA insertion in exon 1) (Hong
and Vierling, 2000, 2001). We used this assay to screen for
additional AtHsp101 mutants in a population of mutagenized
Arabidopsis seedlings and also to assay identified point muta-
tions in AtHsp101 from the Arabidopsis Tilling Resource (http://
tilling.fhcrc.org:9366) for loss of thermotolerance. From our
screen for additional thermotolerance mutants (Hong et al.,
2003), only one other mutant was recovered that failed to
complement hot1-1 in allelism tests and that mapped to the
Hsp101 chromosomal region (data not shown). Sequencing of
the AtHsp101 gene from this mutant identified a single missense
mutation resulting in an exchange of Thr for Ala-499 in the coiled-
coil domain (Figure 1A; see Supplemental Figure 1 online).
Analysis of the AtHsp101 tilling mutants uncovered three addi-
tional hot1 alleles that were defective in thermotolerance, R706K
(hot1-5), E509K (hot1-6), and R815D (hot1-7), all of which lie in
AAA module 2 (Figure 1A). The locations of an additional 35
missense mutations obtained by tilling showed no defect in
thermotolerance and are indicated in Figure 1B.
The relative severity of each of the newly identified defective
alleles (hot1-4 to hot1-7) was compared with the wild type and
the functionally null hot1-1 allele (Hong and Vierling, 2001) in
a quantitative assay for seedling thermotolerance (Figure 1C).
The hot1-4 and hot1-5 alleles are the most severe, showing no
ability to acclimate to treatment at 458C for 2 h, whereas the hot1-6
and hot1-7 alleles have a less severe phenotype, with hot1-6
showing the mildest phenotype. Differences in severity are not
because of differences in protein levels as all the alleles accu-
mulate AtHsp101 protein to wild-type levels (Figure 1D). There-
fore, we conclude that relative thermotolerance of the alleles
reflects the degree to which each mutation disrupts AtHsp101
function in vivo and defines critical AtHsp101 residues in the
coiled-coil domain and NBD2.
hot1-4 Is a Semidominant, Gain-of-Function
AtHsp101 Allele
We noted that in contrast with all the other hot1 alleles, hot1-4
was not only defective in acquired thermotolerance to 458C
treatment, but also showed sensitivity to the 388C acclimation
pretreatment (Figure 1C), which is a permissive treatment for the
other hot1 alleles. To determine the extent of this sensitivity,
elongation of hot1-4 seedlings was compared with the wild type,
protein null (hot1-3), and the missense hot1-1 alleles after in-
creasing time at 388C. The hot1-4 mutant displays dramatic
sensitivity to 388C treatment compared with the wild type and the
other loss-of-function alleles (Figure 2). The fact that the pheno-
type of hot1-4 at 388C is worse than the absence of the protein
(hot1-3 phenotype, Figure 1D) defines hot1-4 as a gain-of-
function mutation and indicates that the hot1-4 protein interferes
with essential cell functions at 38 8C. Furthermore, whereas the
hot1-5, hot1-6, and hot1-7 as well as the previously isolated
AtHsp101 alleles are recessive, loss-of function mutants, the
hot1-4 phenotype proved to be semidominant (Table 1). Com-
pared with wild-type and hot1-4 homozygous seedlings, hetero-
zygous hot1-4 plants showed an intermediate hypocotyl length
after a 458C heat treatment with preconditioning. Despite the
sensitivity of hot1-4 to mild heat treatment, the hot1-4 mutation
had no obvious effect on growth and development in the
absence of stress (data not shown).
To confirm that the hot1-4 mutant protein causes the observed
heat-sensitivity phenotypes, we cloned a hot1-4 mutant genomic
fragment, including its own 1.5-kb promoter region, and trans-
formed wild-type Arabidopsis plants with the mutated gene. As
Figure 2. Unusual Sensitivity of hot1-4 to 388C.
Hypocotyl elongation of hot1 mutant seedlings treated for different times
at 388C, as measured after an additional 2.5 d of growth in the dark at
228C. Replicated measurements performed as in Figure 1C.
AtHsp101 Structure and Thermotolerance 561
shown in Figure 3, the phenotype of hot1-4 can be recapitulated
by expression of the hot1-4 mutant protein in wild-type plants.
Protein gel blot analysis showed that levels of HSP accumulation
were unaffected in the transgenic plants. Thus, the hot1-4 mutant
protein functions in a dominant-negative fashion. In total, these
results provide evidence that the coiled-coil domain is required
for AtHsp101 function during an early step in the development of
thermotolerance. In addition, the semidominant, gain-of-function
phenotype of hot1-4 suggests that this mutation results in a
nonproductive interaction with a limiting substrate or cofactor.
Isolation of Suppressors of ho t1-4
The dominant-negative, 388C-sensitive phenotype of hot1-4
provided an excellent opportunity to screen for suppressor
mutations that could restore growth. Suppressor mutations
should provide additional insight into structure-function relation-
ships of AtHsp101 and potentially identify important AtHsp101
substrates or cofactors. Approximately 7500 homozygous
hot1-4 seeds were mutagenized with ethyl methanesulfonate,
and nearly 110,000 dark-grown, 2.5-d-old seedlings from the
M2 generation were screened for continued growth after a 2-h
treatment at 388C. In total, 43 lines were isolated that showed
a significant reversion of hot1-4 sensitivity to 388C treatment and
that exhibited the same phenotype in the subsequent M3
generation (data not shown).
To separate intragenic from extragenic suppressors, we
sequenced the AtHsp101 gene from all 43 suppressors. Seven
of the suppressors did not carry a second mutation in the
AtHsp101 gene and are unlikely to be linked to the original
hot1-4 mutation. Presumably, they represent extragenic sup-
pressors of hot1-4. These suppressors could define factors
specific to plant stress tolerance and/or identify genes that
encode AtHsp101 cofactors or critical substrates. Further ge-
netic and molecular analysis of these mutants is in progress. The
other 34 suppressors had a second mutation in the hot1-4 gene,
resulting in 13 different amino acid substitutions: eight in the first
nucleotide binding domain (NBD1) (R223K, A270T/A305T,
A297T, G313D, E319K, A329V, G384S, and P388S), four within
NBD2 (G611D, G649E, G653E, and A723V), and one in the
C-terminal region (V813M) (Figure 4).
To confirm that the loss of sensitivity to a 388C treatment in
mutants carrying a second site mutation in the Hsp101 gene was
because of this mutation and not to another extragenic mutation,
a plant carrying each of the individual mutations was crossed
to either the wild type or hot1-4. Segregation of the wild type,
hot1-4, and suppressed phenotype was then scored in the F1
and F2 generations for each cross. As shown in Table 2 and Sup-
plemental Tables 1 and 2 online, suppression of the 388C pheno-
type was recessive and linked to the hot1-4 mutation. These data
confirm that the second site mutation in Hsp101 is the cause of
the recovery of wild-type growth at 388C for each of these
suppressors.
The Suppresso r Mutations Include Both Loss-of-Fun ction
and Restoration-of-Function Alleles
The phenotypes of intragenic suppressor mutants, representing
each amino acid substitution, were retested at 388C to quantify
the degree of phenotypic reversion. All the intragenic suppressors
Table 1. Genetic Analysis of hot1 Mutants
Cross Generation Total WT
Semi-
dominant hot x
2
P
hot1-1 F1 23 23 0 0
3 WT F2
a
159 127 0 32 1.58 >0.1
hot1-3 F1 31 31 0 0
3 WT F2 252 187 0 65 0.08 >0.5
hot1-4 F1 30 0 30 0
3 WT F2 268 64 134 70 0.26 >0.5
hot1-5 F1 7 7 0 0
3 WT F2 42 33 0 9 0.28 >0.5
hot1-6 F1 8 8 0 0
3 WT F2 94 72 0 22 0.12 >0.5
hot1-7 F1 4 4 0 0
3 WT F2 48 38 0 10 0.44 >0.5
The phenotypic scoring was based on the hypocotyl elongation phe-
notypes of the parental lines after 458C for 2 h with pretreatment (388C
for 90 min). WT, wild type.
a
The hypocotyl length ranges of the phenotypic classes of F2 progenies
were 6.5 to 8 mm for the wild-type phenotype, 2.0 to 4.0 mm for the
semidominant phenotype, and 0 mm for hot1-4 mutant phenotype after
heat stress.
Figure 3. Expression of Mutant HOT1-4 Protein in Wild-Type Plants
Recapitulates the 388C Sensitivity Phenotype of hot1-4.
The hot1-4 genomic DNA under its own promoter was expressed in
transgenic wild-type plants. Two independent transgenic lines (T3:1 and
T3:2) are shown compared with the wild type and hot1-4 (three seedlings
each). The hypocotyl elongation assay was performed as in Figure 1C.
Equal protein samples from heat-stressed seedlings (388C for 90 min
followed by 2 h at 228C) were analyzed with the indicated antisera against
AtHsp101, sHsps, or glyceraldehyde-3-phosphate dehydrogenase
(GAPC).
562 The Plant Cell
showed a strong restoring phenotype of hypocotyl elongation
after the same 388C treatment that was used for their initial
identification (Table 3). In addition, when tested for sensitivity to
388C as 10-d-old seedlings, the suppressors also showed re-
covery of the wild-type phenotype (Figure 5A). Thus, all of the
suppressors clearly eliminated the dominant-negative function
of the hot1-4 allele at two distinct growth stages.
The intragenic suppressor mutants were then tested to de-
termine if the function of AtHsp101 in acquired thermotolerance
had been restored by measurement of hypocotyl elongation after
2hat458C of seedlings that had been preconditioned at 388C
(Table 3). The suppressors could clearly be separated into two
classes based on their response to this assay. Suppressors in
Class 1, which includes A270T/A305T, E319K, G384S, P388S,
G611D, A723V, and V813M, were completely arrested in hypo-
cotyl elongation after 2 h at 458C with preconditioning. Genetic
analysis confirmed that this phenotype also segregated with the
hot1-4 mutation, again confirming that it arises as a result of the
second site mutation in AtHsp101 (see Supplemental Tables 1
and 2 online). This acquired thermotolerance test indicates that
the Class 1 suppressors are most likely loss-of-function alleles
because they eliminate the dominant-negative phenotype of
hot1-4 but do not restore function in acquired thermotolerance.
By contrast, the second class of suppressors (Class 2) (R223K,
A297T, G313D, A329V, G649E, and G653E) partially restored
acquired thermotolerance of hypocotyl elongation, with the best
suppression shown by A329V, which elongated to
;75% of
wild-type levels (Table 3).
The same phenotypic behavior was observed when 10-d-old
suppressor seedlings were tested for acquired thermotolerance.
Class 1 suppressors failed to acquire thermotolerance; a 458C
treatment for 2 h (with preconditioning) completely blocked
production of additional leaves, and existing leaves and cotyle-
dons turned white (Figure 5B). Of the Class 2 suppressors, those
which restored hypocotyl elongation after 458C stress to <35% of
the wild type (G313D, G649E, and G653E; Table 3) did not exhibit
acquired thermotolerance at the 10-d-old seedling stage (Figure
5B). However, suppressors R223K, A297T, and A329V, which
showed stronger reversion of the hypocotyl phenotype (>43% of
the wild type), also showed acquired thermotolerance at the
Figure 4. Location of hot1-4 Suppressor Mutations in AtHsp101.
Class 1 loss-of-function suppressors are labeled in smaller gray font and
indicated with closed circles. Restoration-of-function Class 2 suppres-
sors are indicated in larger black font with open circles. The strongest
Class 2 suppressors, as discussed in the text, are underlined (see also
Supplemental Figure 1 online).
Table 2. Genetic Segregation Analysis of Intragenic Suppressors
Genotypes
No. of plants
x
2
P
Total WT hot Sup.
R223K F1 5 5 0 0
3 WT F2 36 36 0 0
R223K F1 5 0 0 5
3 hot1-4 F2 36 0 10 26 0.14 >0.9
A270/A305 F1 3 3 0 0
3 WT F2 38 38 0 0
A270/A305 F1 7 0 0 7
3 hot1-4 F2 40 0 9 31 0.13 >0.9
A297T F1 7 7 0 0
3 WT F2 36 36 0 0
A297T F1 6 0 0 6
3 hot1-4 F2 29 0 6 23 0.73 >0.5
G313D F1 5 5 0 0
3 WT F2 36 36 0 0
G313D F1 6 0 0 6
3 hot1-4 F2 32 0 7 25 0.16 >0.9
E319K F1 7 7 0 0
3 WT F2 32 32 0 0
E319K F1 2 0 0 2
3 hot1-4 F2 28 0 5 23 0.76 >0.5
A329V F1 5 5 0 0
3 WT F2 46 46 0 0
A329V F1 5 0 0 5
3 hot1-4 F2 39 0 8 31 0.14 >0.9
G384S F1 6 6 0 0
3 WT F2 36 36 0 0
G384S F1 5 0 0 5
3 hot1-4 F2 32 0 7 25 0.16 >0.9
P388S F1 10 10 0 0
3 WT F2 40 40 0 0
P388S F1 3 0 0 3
3 hot1-4 F2 40 0 9 31 0.13 >0.9
G611D F1 4 4 0 0
3 WT F2 46 46 0 0
G611D F1 5 0 0 5
3 hot1-4 F2 43 0 10 33 0.07 >0.9
G649E F1 5 5 0 0
3 WT F2 45 45 0 0
G649E F1 6 0 0 6
3 hot1-4 F2 44 0 8 36 0.36 >0.5
G653E F1 5 5 0 0
3 WT F2 37 37 0 0
G653E F1 5 0 0 5
3 hot1-4 F2 34 0 7 27 0.35 >0.5
A723V F1 5 5 0 0
3 WT F2 27 27 0 0
A723V F1 8 0 0 8
3 hot1-4 F2 35 0 8 27 0.08 >0.9
V813M F1 5 5 0 0
3 WT F2 46 46 0 0
V813M F1 5 0 0 5
3 hot1-4 F2 48 0 9 39 1 >0.5
Phenotype scoring was done by comparing the hypocotyl length
distribution patterns of the F2 seedlings with those of the parental lines.
The hypocotyl length ranges of the phenotypic classes of F2 progenies
were 9.5 to 10.5 mm for the wild-type phenotype, 9.0 to 11.0 mm for the
suppressor phenotype (Sup.), and 2.0 to 4.0 mm for hot1-4 mutant
phenotype after 388C heat treatment for 2 h. WT, wild type.
AtHsp101 Structure and Thermotolerance 563
10-d-old seedling stage (Figure 5B). Furthermore, the A329V
suppressor recovered as much as the wild type, whereas the
R223K and A297T suppressors had pale green leaves and were
smaller than the A329V suppressor, also consistent with the
relative growth of these suppressors in the hypocotyl assay.
Thus, the strong suppressors, R223K, A297T, and A329V, clearly
compensate for the primary functional defect caused by hot1-4,
although with somewhat different effectiveness. These three
mutations are all located in NBD1 (Figure 4), supporting a func-
tional interaction of this domain with the coiled-coil domain
containing the primary hot1-4 mutation.
The V813M and G384S Suppressors Accumulate
Truncated AtHsp101
To determine whether the restored growth phenotype of the
suppressors was because of a dose effect of the mutant protein,
protein gel blot analysis was performed on protein samples from
heat-treated, 2.5-d-old dark-grown seedlings. Except for V813M
and G384S, all the mutant proteins accumulated to similar levels
when compared with the wild type and hot1-4 (Figure 6). Although
the Class 1 mutants V813M and G384S are obviously missense
mutations,
;90- and 41-kD bands, respectively, were detected
with AtHsp101 antibody in these samples (Figure 6). Because the
antibody is directed specifically to the N-terminal domain of
AtHsp101 (Hong and Vierling, 2001), these bands most likely
represent C-terminally truncated proteolytic fragments gener-
ated in vivo as a result of protease sensitivity near the mutation
site. Thus, these two suppressor mutations clearly produce
a nonfunctional protein. This result supports the above conclu-
sion that the defective phenotype in acquired thermotolerance of
the Class 1 suppressors is caused by loss of AtHsp101 function.
Strong Suppressors Resto re the Function of AtHsp101 in
Protein Resol ubiliz atio n
The current model for Hsp100/ClpB function in thermotolerance
proposes that these proteins are required for reactivation of
aggregated proteins after heat stress (Glover and Lindquist,
1998; Goloubinoff et al., 1999; Motohashi et al., 1999). Therefore,
it can be assumed that during recovery from heat stress, heat-
aggregated proteins requiring AtHsp101 function should remain
in the insoluble cell fraction in hot1 mutants, whereas these
proteins should return to the soluble fraction in the wild type. It
also follows that if resolubilization is an essential function of
AtHsp101 for acquired thermotolerance, the strong hot1-4
suppressor mutations should restore protein solubility after
heat stress. As a first step to test this hypothesis, we sought
conditions for comparing the insoluble protein fractions from
wild-type and mutant plants during recovery from heat stress,
along with a suitable protein to assay for solubility. We de-
termined that in wild-type plants, the cytosolic Class I and II
Figure 5. Thermotolerance Phenotype of 10-d-Old Suppressor Mutants.
After growth for 10 d in the light at 228C, seedlings were treated either at
388Cfor3h(A) or at 388C for 90 min followed by 2 h at 228C and then 2 h
at 458C (B). Seedlings were photographed 7 d after treatment.
Table 3. Hypocotyl Elongation Phenotype of Intragenic Suppressors
Genotype
No. of
Isolates
Hypocotyl Length (%)
a
388C388C/458C(%)
WT 91.8 6 9.6 68.0 6 9.2 (100)
hot1-4 (A499T) 26.2 6 4.5 0 6 0
hot1-3 (null) 90.8 6 9.6 0 6 0
Class 1 suppressors
A270T/A305T 6 89.9 6 6.4 0 6 0
E319K 1 84.7 6 6.6 0 6 0
G384S 3 89.7 6 8.3 0 6 0
P388S 3 87.5 6 6.8 0 6 0
G611D 5 90.1 6 6.1 0 6 0
A723V 3 79.4 6 9.8 0 6 0
V813M 2 91.4 6 8.3 0 6 0
Class 2 suppressors
R223K 2 83.6 6 5.1 28.7 6 5.8 (42)
A297T 2 82.1 6 9.8 29.3 6 6.3 (43)
G313D 3 75.3 6 6.7 9.7 6 4.1 (14)
A329V 1 94.1 6 6.2 51.3 6 6.7 (75)
G649E 1 86.7 6 8.0 20.9 6 6.8 (30)
G653E 2 85.2 6 7.3 24.1 6 7.3 (35)
a
Dark-grown, 2.5-d-old seedlings were total hypocotyl elongation after
the heat stress was measured as a percentage of the 228C value. Mean
and standard deviations were derived by measurement from three
independent experiments performed with 60 or more seedlings per F3
seedling. The numbers in parentheses indicate percentage of elongation
compared with that of wild-type plants.
564 The Plant Cell
sHsps, partitioned during 458C heat stress into the insoluble cell
fraction and almost completely transitioned back to the soluble
fraction after 3 h of recovery from the heat stress (data not
shown). We therefore compared the solubility of either Class I
AtHsp17.6C-I or Class II AtHsp17.6C-II in wild-type, hot1-4, and
the strong suppressors, R223K, A297T, and A329V, after 3 h of
recovery from a 388C, 3-h heat stress or from a 458C, 45-min heat
stress with preconditioning. In contrast with the wild type,
a majority of the sHsps remained in the pellet fraction during
recovery from 458C stress in the hot1-4, hot1-1, and hot1-3
mutants (Figure 7). The mutants also showed detectably more
pelleted sHsps after the 388C stress. We then examined the
solubility of these sHsps in the strong suppressors. sHsp solu-
bility was directly correlated with the strength of the suppressor
(Figure 7), with the most effective suppressor, A329V, being
almost identical to the wild type and the weaker suppressors,
R223K and A297T, showing slightly more sHsp retained in the
insoluble fraction. The same samples were also used to examine
the solubility of AtHsp101. In contrast with the sHsps, neither the
wild type nor the hot1-1 allele of AtHsp101 becomes insoluble
during heat stress; similar behavior is reported for E. coli ClpB
(Mogk et al., 2003a, 2003c). A fraction of hot1-4 becomes
insoluble during heat stress, but insolubility does not correspond
to the phenotype, as at 388C the hot1-4 phenotype is already
severe, whereas little protein is insoluble. Only the strongest
suppressor reverses the AtHsp101 insolubility. Altogether, these
results support the model that one essential thermotolerance
function of AtHsp101 involves protein disaggregation and ge-
netically link Hsp101 to the sHsp chaperone system.
DISCUSSION
We have been able to identify features critical to Hsp100/ClpB
function in vivo using a sensitive, quantitative bioassay for
thermotolerance. We identified four missense mutations in the
Hsp100/ClpB protein, Arabidopsis AtHsp101, that compromise
the function of AtHsp101 in thermotolerance, including an un-
usual allele, hot1-4 (A499T), which shows a semidominant, gain-
of-function sensitivity to a mild, usually nonlethal heat stress.
This mutation, which lies in the coiled-coil domain of AtHsp101,
supports an essential role for this structural domain in Hsp100/
ClpB action in vivo. We also suggest that the gain-of-function
phenotype of hot1-4 is consistent with the interpretation that the
mutation disrupts AtHsp101 activity such that a critical substrate
or cofactor becomes limiting for the cell. Locations of the primary
missense mutations hot1-5, -6, and -7 and of suppressors of
hot1-4 define other features essential for in vivo function of
AtHsp101 and point to protein domain interactions required for
Hsp100/ClpB action. Finally, appearance of sHsps in the soluble
cell fraction during recovery from heat stress does not occur in
hot1-4 but is restored by strong suppressor mutants, in line with
the model that disruption of protein aggregates is an essential
AtHsp101 function in thermotolerance and supporting the model
that sHsps and Hsp100/ClpB proteins interact in a chaperone
network in the cell.
The coiled-coil domain of Hsp100/ClpB is unique to this class
of AAAþ proteins (Schirmer et al., 1996; Celerin et al., 1998).
Although referred to alternatively as the linker, middle region, or
spacer between AAA modules 1 and 2, it is actually an insertion
into the C-terminal small domain of AAA module 1 (Figures 1 and
8; see Supplemental Figure 1 online). Deletion of the coiled-coil
domain from E. coli ClpB has shown that it is not essential for
hexamerization or ATPase activity but is required for thermotol-
erance in vivo and for disaggregation of firefly luciferase in vitro
(Mogk et al., 2003b). The structure of T. thermophilus ClpB
reveals that this domain is formed by four helices (L1 to L4) that
Figure 6. Accumulation of AtHsp101 Protein in the Suppressor Mutants.
Protein samples from heat-stressed seedlings (388C for 90 min followed
by 2 h at 228C) were analyzed with AtHsp101 antiserum.
(A) Class 1 suppressor mutants.
(B) Class 2 suppressor mutants.
Figure 7. Reversal of sHsp Insolubility in the Suppressor Mutants.
Ten-day-old seedlings were treated either at 388C for 3 h (38) or at 388C
for 90 min followed by 2 h at 228C and then 60 min at 458C (45), and total
protein (T) was isolated after further recovery for 3 h at 228C. The
insoluble protein fraction (P) was separated from the soluble fraction (S)
by centrifugation and analyzed with AtHsp101, AtHsp17.6C-I (sHsp I),
and AtHsp17.6C-II (sHsp II) antiserum.
AtHsp101 Structure and Thermotolerance 565
interact in two coiled-coil motifs comprised of a-helices L1 and
L2, and L2 with L3/4 (Figures 8 and 9). Both coiled-coils have
features typical of Leu zippers. The hot1-4 mutation (A499T) lies
in helix L3, within the conserved sequence segment 494YDLAR-
AADL502 (484YDLNRAAEL492, in T. thermophilus) (Figures 8A
and 8B). This sequence motif was previously identified by
Schirmer et al. (1996) as characteristic of Hsp100/ClpB chaper-
ones. The weak mutation hot1-6 (E509K) lies in the coiled-coil
domain directly outside this conserved motif (Figure 8B). Neither
the hot1-4 nor hot1-6 residues are predicted to contact residues
in neighboring subunits, which along with the normal accumula-
tion of the proteins and the similar solubility of hot1-4 to the wild
type, suggest that these mutations do not significantly disrupt the
hexamer structure. In vitro cross-linking experiments also pro-
duce the same large, presumably oligomer forms of AtHsp101 in
the wild type and the hot1-4 mutants (U. Lee, unpublished data).
Eight other mutations in this domain have no effect on AtHsp101
function (Figure 1). Altogether, the hot1-4 and hot1-6 mutations
provide direct genetic evidence for the importance of helix L3 and
the coiled-coil domain in AtHsp101 function.
Interestingly, in a screen for mutants in Saccharomyces
cerevisiae Hsp104, Schirmer et al. (2004) recently reported
isolation of a mutation in the same motif, at the adjacent Ala
residue (A503V; 497YDTATA
ADL505 in S. cerevisiae). The
Hsp104 A503V mutant was an active ATPase in vitro (Cashikar
et al., 2002), exhibiting actually an approximately twofold
Figure 8. Location of the hot1 Mutations on the T. thermophilus ClpB Structure.
(A) The available structure of the T. thermophilus ClpB monomer (PDB file 1QVR; chain in [A]) is shown with the coiled-coil domain and AAA modules
indicated. Within each AAA module, the NBD domain is colored blue, with the Walker A and B motifs in orange, sensor 1 in pink, and the Arg finger in
yellow (see also Supplemental Figure 1 online). The small, a-helical domain of each AAA module is shown in gray. Nucleotide (AMP-PNP) is space-filled
in CPK coloring (hydrogen, white; oxygen, red; phosphorous, orange; sodium, blue). Motif 1 as defined by Schirmer et al. (1996) is colored purple in the
coiled-coil domain. The residue corresponding to each of the hot1 mutation sites is space-filled and colored green.
(B) Alternative view of the motif II region of the coiled-coil domain indicating the L2, L3, and L4 helices and the residue positions of the hot1-4 and hot1-6
mutations (in CPK coloring).
(C) View of interactions of the hot1-5 position (in CPK coloring) with residues within NDB2 and NBD1 as discussed in the text.
(D) Position of hot1-7 (CPK coloring) relative to the NBD2 ATP binding site. The sensor 2 Arg residue, which is part of the GAR motif, as well as sensor 1
are shown in pink. Residue numbering corresponds to AtHsp101. Figure was prepared with Swiss PDB viewer.
566 The Plant Cell
increase in basal ATPase activity (i.e., activity in the absence of
substrate). Therefore, the A503V mutation cannot disrupt the
hexameric structure, which is required for ATPase activity. This
observation further supports our interpretation that the hot1-4
mutant protein retains its hexameric structure. Furthermore, like
hot1-4, the Hsp104 A503V mutation also exhibits a gain-of-
function phenotype, leading to cell lethality when induced at the
normally permissive temperature of 378C (Schirmer et al., 2004).
The similarity of phenotype of these AtHsp101 and yeast Hsp104
mutations, along with the ability of AtHsp101 to support thermo-
tolerance in yeast (Schirmer et al., 1994), argue that these
proteins are true functional homologs.
The other mutations studied also provide new information
about interactions within the Hsp100/ClpB proteins. The strong
hot1-5 allele alters a residue (R706K) that is conserved in E. coli
(R705) and T. thermophilus (R695), and which is in a loop
between b-strands 4 and 5 in NBD2, positioned away from the
active site (Figures 8A and 8C; see Supplemental Figure 1 online).
Normal accumulation and solubility of the hot1-5 protein argues
that the protein is not significantly compromised structurally.
Based on the T. thermophilus structure, this Arg residue makes
intramolecular contact with three other residues conserved in
Arabidopsis and E. coli, including a Leu residue directly adjace nt
to sensor 1 in NBD1, and has side chain interactions with
conserved Glu and Arg residues in NBD2 helix D7 (Figure 8C).
Notably, a Lys residue is present at the hot1-5 position in yeast
Hsp104, and correspondingly the predicted interacting residues
are different. The unique importance of hot1-5 interactions, as
opposed to variation in the loop itself, is also indirectly supported
by the absence of a phenotype for another mutation in the loop
connecting b-strands 4 and 5, G703E (Figures 1 and 8C).
Hattendorf and Lindquist (2002a) have proposed that ATP hydro-
lysis at NBD1 depends on the nucleotide bound at NBD2 in yeast
Hsp104. We speculate that hot1-5 is in a position involved in
communicating nucleotide status between NBD2 and NBD1 and
note that the distance between the hot1-5 Arg residue and the
conserved Leu in NBD1 varies in the three ClpB molecules in the
crystal structure, consistent with a dynamic interaction.
Figure 9. Location of the hot1-4 Suppressor Mutations on the T. thermophilus ClpB Monomer Structure.
The major domains and motifs of HCP100/ClpB are colored as indicated and described in Figure 8. The sites of Class 1 suppressor mutations are
space-filled dark gray, and sites of the Class 2 suppressor mutations are space-filled in red. The three strongest Class 2 suppressors are indicated in
bold and are underlined. For A297, G653, and G649, which are in unresolved segments of the structure, dotted lines are used to indicate the relative
positions of those segments. The arrow indicates the general position of the axial channel of the hexameric form of the protein. Residue numbering
corresponds to AtHsp101. Figure was prepared with Swiss PDB viewer.
AtHsp101 Structure and Thermotolerance 567
It is difficult to define the possible effects of some of the other
AtHsp101 mutations. The hot1-7 mutation represents a dramatic
change (G815D) in a conserved motif (GAR) that includes the
sensor 2 Arg residue already recognized as important for
function (Figure 8D). The residues in the sensor 2 motif are in
contact with nucleotides, and the Arg residue in yeast Hsp104 is
proposed to contribute binding energy, but not discrimination
between ATP and ADP (Hattendorf and Lindquist, 2002a).
Mutations in sensor 2, or of all three residues of the GAR motif,
do not severely disrupt hexamerization or ATPase activity, but
prevent ability to dissociate protein aggregates (Hattendorf and
Lindquist, 2002b; Mogk et al., 2003b). In vivo, the moderate
phenotype of hot1-7 is consistent with a moderate defect in
thermotolerance observed for an R to M mutation of sensor 2 in
yeast (Hattendorf and Lindquist, 2002b). The absence of phe-
notype for 35 other missense mutations analyzed is not surpris-
ing, given that many lie in nonconserved residues or represent
conservative changes (Figure 1; see Supplemental Figure 1
online). Perhaps the most surprising of the aphenotypic muta-
tions is P671L, which represents a Pro residue conserved not
only in NBD2 of all ClpB proteins, but also found in ClpA.
The unusual sensitivity of the hot1-4 allele to 388C, which is
otherwise permissive for wild-type and AtHsp101 null plants, led
us to focus further on analysis of this mutation. We performed
a screen that identified both loss-of-function (Class 1) and
restoration-of-function (Class 2) intragenic suppressor muta-
tions in AtHsp101. Genetic dissection of Hsp100/ClpB function
by intragenic suppressor analysis has not been performed in any
organism, and we are also unaware of any previous studies using
intragenic suppressors for dissection of protein mechanism in
higher plants. We argue that the seven Class 1 mutations that
eliminated the 388Cheatsensitivityofhot1-4, but did not restore
the acquired thermotolerance function, represent loss-of-function
AtHsp101 mutants. Consistent with this interpretation, one of
these suppressors is in the Walker A motif of NBD 2 (G611D),
and three others are predicted to affect the nucleotide binding
pocket (E319K, P388S, and A723V) (Figure 9), which should
result in defective hexamerization or ATP hydrolysis as demon-
strated in vitro for other Hsp100/ClpB proteins (Kim et al., 2000a;
Watanabe et al., 2002; Mogk et al., 2003b). These findings
suggest that in hot1-4, AtHsp101 retains normal ATP hydrolysis
activity and hexamerization and indicates that these properties
are essential for the dominant-negative effect of this allele. Two
other Class 1 suppressors, G384S and V813M, are both located
between the second and third helices in the C-terminal sub-
domain of NBD1 and NBD2, respectively (Figure 9; see Sup-
plemental Figure 1 online). Both of these mutations lead to
accumulation of a truncated AtHsp101 protein with sizes that
predict that in vivo cleavage occurs in close proximity to the
mutation. Proteolytic removal of the C terminus in V813M
supports the requirement of this domain for function, which as
stated before, may involve not only hexamerization, but also
effector binding (Smith et al., 1999; Strub et al., 2003). Although
our screen was not performed to saturation, it is notable that no
loss-of-function suppressors were obtained in the N-terminal
domain, which has also not been associated with an essential
function in other organisms (Clarke and Eriksson, 2000; Beinker
et al., 2002; Mogk et al., 2003b).
Whereas the coiled-coil domain is clearly essential for Hsp100/
ClpB function, its role in the catalytic cycle is not known. The
position of hot1-4 (A499T) (and A503V in S. cerevisiae Hsp104)
suggests that this mutation could disrupt interactions of L3 with
L2 within a subunit. Lee et al. (2003) have proposed that motion of
the coiled-coil domain is required for Hsp100/ClpB chaperone
activity. They introduced Cys residues into T. thermophilus ClpB
to effect cross-linking of the L2/L3-4 coiled-coil (in which hot1-4
is located) to NBD1 and observed in vitro that ATP hydrolysis
continued, but the protein disaggregation activity of ClpB was
lost. The three hot1-4 suppressors that restore >43% of acquired
thermotolerance of hypocotyl growth and also restore acquired
thermotolerance of 10-d-old seedlings are all located in NBD1
(R223K, A297T, and A329V). The strongest of these, A329V, has
>70% wild-type activity in the hypocotyl assay and restores
solubility of sHsps during recovery from heat stress. Interest-
ingly, A329V lies in close proximity to the NBD1 Arg finger, which
is believed to interact with the nucleotide in an adjacent subunit
(Figure 9). Another weak suppressor, G313D, is also in NBD1
(Figure 9). Taken together, the location of the Class 2 restoration-
of-function suppressors of hot1-4 is consistent with the inter-
pretation that dynamic interaction of the coiled-coil domain with
NBD1 is indeed essential in vivo.
How Hsp100/ClpB proteins act to disaggregate proteins has
been proposed to involve a crowbar action of the coiled-coil
domain and/or threading of substrate through the axial channel
of the hexamer, analogous to the mechanism of other hexameric
ATPases (Lee et al., 2003; Lum et al., 2004; Schlieker et al., 2004).
Location of the Class 2 suppressors provides potential support
for the latter mechanism. The strong suppressor A297V is
located in a presumably flexible, disordered segment that was
not resolved in the TtClpB structure, but which is in the axial
channel region. A329V, another strong suppressor, is positioned
to potentially alter interactions with another unresolved axial
channel loop (TtClpB residues 234 to 246; corresponding to 243
to 255 in AtHsp101), which has recently been proposed to
contain residues that bind the E. coli ClpB substrate TrfA
(Schlieker et al., 2004). The only restoration-of-function sup-
pressors outside NBD1 are the two weak suppressors, G649E
and G653E in NBD 2, which also lie in an unresolved axial channel
loop (TtClpB residues 638 to 650; corresponding to AtHsp101
648 to 660) (Figure 9; see Supplemental Figure 1 online). This
loop contains the conserved motif GYVG found in other AAAþ
proteins (with the first G corresponding to G653 in AtHsp101).
Mutation of the Tyr residue in this loop impairs function of E. coli
ClpX (Siddiqui et al., 2004) and HslU (Wang et al., 2001b).
Structural studies also indicate that this loop gates the axial
channel of HslU during the ATPase cycle (Wang et al., 2001a).
Recent experiments with yeast Hsp104 also support an essential
role for this loop in the protein disaggregation reaction (Lum et al.,
2004). In total, we suggest that these intragenic suppressors
point to a model in which motions of the coiled-coil domain
modulate the ATPase activity of NBD1 and the position of axial
channel loops in AtHsp101, which may be involved in threading
of substrate through the axial channel.
Analysis of the solubility of plant cytosolic sHsps in the hot1-4
mutant and the strong suppressors also provides data to
link these two chaperone systems in a eukaryote. Genetic
568 The Plant Cell
interactions of sHsps and ClpB have been reported in both E. coli
and Synechocystis (Giese and Vierling, 2002; Mogk et al., 2003a)
and are supported by in vitro analysis of E. coli ClpB activity in
combination with E. coli, Synechocystis, and plant cytosolic
class I sHsps (Mogk et al., 2003c). The current model for sHsp
function proposes that sHsps bind and maintain denaturing
proteins in a form that is accessible to ATP-dependent refolding
chaperones. Furthermore, their presence in large protein aggre-
gates makes these aggregates better substrates for the action of
Hsp100/ClpB (Mogk et al., 2003c). However, although Hsp100/
ClpB disaggregates sHsp containing complexes, there is no
evidence for direct physical interaction of Hsp100/ClpB with
sHsps (Mogk et al., 2003a, 2003c), and we did not observe
association of AtHsp101 with sHsps in the insoluble protein
fraction of the cell. In total, the data here indicate that the plant
cytosolic Class II, as well as cytosolic Class I sHsps, which
diverged on the order of 400 million years ago (Waters and
Vierling, 1999), may both facilitate protein disaggregation by
Hsp100/ClpB in plants.
At which step the hot1-4 mutation affects catalysis will require
additional biochemical studies. It does not appear that the
dominant-negative phenotype of hot1-4 is attributable to un-
regulated ATPase activity of the mutant AtHsp101 protein,
leading to depletion of cellular ATP. Measurement of ATP levels
in the wild type, hot1-3, and hot1-4 showed no significant
differences in ATP levels between the genotypes during heat
stress or recovery (U. Lee and E. Vierling, unpublished data). One
possibility is that the mutation traps AtHsp101 in a nonfunctional
protein complex such that another critical cofactor or substrate
becomes limiting for the cell. Such a complex must still be
soluble, as our analysis does not indicate that a significant
fraction of AtHsp101 becomes insoluble at 388C where the
hot1-4 phenotype is still severe. The Class 1 suppressors that
eliminate the dominant-negative phenotype are predicted to
disrupt the basic hexameric structure and ATPase activity,
consistent with the requirement of these activities to form
a trapped complex. By altering structural interactions of the
coiled-coil domain with NBD1, the strong Class 2 suppressors
may eliminate this trapped intermediate. Weaker Class 2 sup-
pressors may affect the protein’s conformation, but not enough
to fully disrupt the formation of the nonfunctional complex
between hot1-4 and its bound protein(s). Another possibility is
that the suppressors directly or indirectly alter a binding site for
substrate or functional partner protein, such that a nonfunctional
complex no longer forms, but sufficient substrate or cofactor
interactions are retained for partial function in vivo. Attempts to
coimmunoprecipitate hot1-4 in complex with other proteins has
not been successful. Biochemical characterization of hot1-4
along with identity of extragenic suppressors should provide
further insight into the nature of the hot1-4 defect and the
mechanism of Hsp100/ClpB action.
METHODS
Plant Material and Growth Conditions
Arabidopsis thaliana (Columbia accession) plants were grown on plates in
the dark or under long-day conditions (16 h light/8 h dark) in a controlled-
temperature growth chamber (228C/188C). Thermotolerance tests of
2.5-d-old dark-grown or 10-d-old light-grown seedlings were performed
basically according to Hong and Vierling (2000).
Isolation and Genetic Analysis of hot1 Muta nt Alle les
The hot1-4 mutation was isolated as a thermotolerance-defective mutant
in a screen based on hypocotyl elongation of 2.5-d-old dark-grown
seedlings (Hong and Vierling, 2000). Allelism tests showed that hot1-4
was tightly linked to the previously described AtHsp101 loss-of-function
missense allele, hot1-1 (Hong and Vierling, 2000). For most of the
experiments described in this work, the third backcrossed line of hot1-4
to Columbia wild type was used. Tilling analysis (in the Columbia ecotype,
carrying the erecta mutation) was performed on three segments of the
AtHsp101gene, encompassing approximately amino acid residues 1 to
230, 355 to 595, and 660 to 911 (Arabidopsis Tilling Resource, http://
tilling.fhcrc.org:9366) (see Supplemental Figure 1 online). The hot1-5,
hot1-6, and hot1-7 mutants were recovered from 37 missense mutations
in AtHsp101 obtained from this analysis. Three stop codon mutations
(Q409, Q422, and Q704) were also obtained, confirmed to be defective in
thermotolerance, and not studied further. Each line was assayed for
thermotolerance defects in the homozygous or heterozygous state, and
homozygous lines were isolated for lines showing a phenotype (hot1-5,
hot1-6,andhot1-7 ). One homozygous M4 plant from each mutant line
was then backcrossed to Columbia erecta wild-type plants, and one
homozygous F3 line for each mutation was used for quantitative analysis.
Isolation and Genetic Analysis of hot1-4 Suppressor Mutations
Approximately 7500 homozygous seeds of hot1-4 were mutagenized
with ethyl methanesulfonate, and
;110,000 M2 seed were screened for
suppressor mutants as follows. M2 seeds were surface-sterilized and
plated on minimal medium containing 0.5% sucrose. Plates were in-
cubated at 48C for 3 d and then placed in a vertical position at 228Cfor
2.5 d in the dark. The dark-grown seedlings were treated at 388Cfor2h,
and then the plates were returned to the dark at 228C for 1.5 d. Seedlings
that showed increased hypocotyl elongation compared with hot1-4 were
rescued by growth under light for approximately 1 week before being
transplanted to soil. The selected seedlings were retested for tolerance to
388C in the next generation (M3) as described above.
To distinguish if the suppressor mutations were intragenic or extra-
genic, the entire AtHsp101 gene from the candidate suppressors was
amplified by PCR using specific primers. The amplified DNA products
were sequenced on both strands. Thirty-four lines were found to contain
the original hot1-4 mutation together with an additional missense muta-
tion within the AtHsp101 coding region. The same suppressor mutation
was found in more than two independent lines (R223K, A270/A305T,
A297T, G313D, G384S, P388S, G611D, G653E, A723D, and V813M),
whereas E319K, A329V, and G649E mutations, which were isolated only
once, were confirmed by sequencing from two individual lines (Table 3).
To test genetic linkage between hot1-4 and the suppressors, the
homozygous M3 or M4 intragenic suppressors with second site muta-
tions in AtHSP101 were backcrossed to hot1-4. The F1 plants all showed
the suppressing phenotype after 388C heat treatment for 2 h. The
subsequent F2 progenies completely segregated in a 3:1 ratio of the
suppressor mutant to hot1-4, indicating that the suppressor mutations
are tightly linked to the hot1-4 mutation (Table 2; see Supplemental
Tables 1 and 2 online).
To obtain F3 homozygous intragenic suppressor lines for phenotypic
analysis, the homozygous M3 or M4 intragenic suppressors were out-
crossed to wild-type plants. All F1 plants showed the wild-type pheno-
type, and F2 progenies segregated either the hot1-4 mutant phenotype
(Class 1 suppressor) or the suppressor phenotype (Class 2 suppressor) at
AtHsp101 Structure and Thermotolerance 569
a ratio of 1:3 to the wild-type phenotype when tested by heating at 458C
for 2 h with pretreatment (388C for 90 min). Scoring of the hypocotyl length
phenotype in F2 progenies was based on the hypocotyl length distribu-
tion of the parental lines.
Plant Transformation
A 4.7-kb XhoI/XbaI AtHsp101 genomic region containing the hot1-4
mutation, including 1.5 kb of the promoter region, was cloned into pBin19
and transformed into Columbia wild-type plants by the floral dipping
method (Clough and Bent, 1998). A total of 20 lines (T1 generation) were
selected on minimal plates with kanamycin (30 mg/mL). The number of
T-DNA insertion loci was determined in the T2 generation based on the
segregation ratio of both kanamycin resistance and the hot1-4 pheno-
type. Two independent T3 homozygous lines were used for phenotypic
studies.
Fractionation of Heat-Denatured Proteins
Ten-day-old seedlings were pretreated at 388C followed by 2 h at 228C,
and then heat shocked at 458C for 60 min (denaturatio n phase), followed
by 3 h of recovery at 228C (recovery phase). Total protein was extracted
before the denaturation phase and after the recovery phase in non-
denaturing buffer (25 mM Hepes, pH 7.5, 0.5% Triton X-100, 200 mM
NaCl, 0.5 mM EDTA, 5 mM e-amino-N-caproic acid, and 1.0 mM
benzamidine). Protein concentration was determined using a Coomassie
Brilliant Blue dye binding assay (Ghosh et al., 1988) with BSA as
a standard. Total protein solutions were immediately centrifuged at
19,000g for 15 min at 48C. Supernatants were mixed with the same
volume of SDS sample buffer (60 mM Tris-HCl, pH 7.5, 60 mM DTT, 2.0%
[w/v] SDS, 15% [w/v] sucrose, 5 mM e-amino-N-caproic acid, and 1.0 mM
benzamidine). The pellet was suspended in nondenaturing buffer, washed
six times, and then resuspended to the original volume in SDS sample
buffer. Proteins were separated by SDS-PAGE on 15% acrylamide gels
and processed for protein gel blot analysis (Hong and Vierling, 2001).
SDS-PAGE and Protein Gel B lot Analysis
The 2.5-d-old dark-grown seedlings were treated at 388C for 90 min, and
then total protein was extracted in SDS sample buffer and separated by
SDS-PAGE on 7.5% or 15% acrylamide gels and processed for protein
gel blot analysis (Hong and Vierling, 2001). Protein blots were probed with
rabbit antiserum against AtHsp101 or against AtHsp17.6C-I or -II (Hong
and Vierling, 2001). As a loading control, blots were probed for cytosolic
glyceraldehyde-3-phosphate dehydrogenase using a GAPC antibody
(gift of Ming-Che Shih, University of Iowa) as described (Chan et al., 2002).
Blots were incubated with goat anti-rabbit horseradish peroxidase and
bands visualized by enhanced chemiluminescence (Amersham Interna-
tional, Piscataway, NJ).
ACKNOWLEDGMENTS
We thank Ming-Che Shih for the glyceraldehyde-3-phosphate dehydro-
genase antibodies, K. Giese, M. Mishkind, and C. Dieckmann for
critique of the manuscript, and Joseph T. Carroll for preparation of the
structure figures. Supported by Department of Energy Grant DE-FG03-
99ER20338 to E.V. S.W.H. was in part supported by the Agricultural
Plant Stress Research Center (Kwang Ju, South Korea).
Received September 3, 2004; accepted November 11, 2004.
REFERENCES
Barnett, M.E., Zolkiewska, A., and Zolkiewiski, M. (2000). Structure
and activity of ClpB from Escherichia coli. Role of the amino- and
carboxyl-terminal domains. J. Biol. Chem. 275, 37565–37571.
Beinker, P., Schlee, S., Groemping, Y., Seidel, R., and Reinstein, J.
(2002). The N terminus of ClpB from Thermus thermophilus is not
essential for the chaperone activity. J. Biol. Chem. 277, 47160–47166.
Cashikar, A.G., Schirmer, E.C., Hattendorf, D.A., Glover, J.R.,
Ramakrishnan, M.S., Ware, D.M., and Lindquist, S.L. (2002). De-
fining a pathway of communication from the C-terminal peptide
binding domain to the N-terminal ATPase domain in a AAA protein.
Mol. Cell 9, 751–760.
Celerin, M., Gilpin, A.A., Schisler, N.J., Ivanov, A.G., Miskiewicz, E.,
Krol, M., and Laudenbach, D.E. (1998). ClpB in a Cyanobacterium:
Predicted structure, phylogenic relationships, and regulation by light
and temperature. J. Bacteriol. 180, 5137–5182.
Chan, C.S., Peng, H.P., and Shih, M.C. (2002). Mutations affecting light
regulation of nuclear genes encoding chloroplast glyceraldehyde-3-
phosphate dehydrogenase in Arabidopsis. Plant Physiol. 130, 1476–
1486.
Clarke, A.K., and Eriksson, M.J. (2000). The truncated form of the
bacterial heat shock protein ClpB/HSP100 contributes to develop-
ment of thermotolerance in the cyanobacterium Synechococcus sp.
strain PCC 7942. J. Bacteriol. 182, 7092– 7096.
Clough, S.J., and Bent, A.F. (1998). Floral dip: A simple method for
Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant
J. 16, 735–743.
Dougan, D.A., Mogk, A., Zeth, K., Turgay, K., and Bukau, B. (2002).
AAAþ proteins and substrate recognition, it all depends on their
partner in crime. FEBS Lett. 529, 6–10.
Ghosh, S., Hepstein, S., Heikkila, J., and Dumbroff, E. (1988). Use of
scanning densitometer or an ELISA plate reader for measurement of
nanogram amounts of protein in crude extracts from biological tissue.
Anal. Biochem. 169, 227–233.
Giese, K.C., and Vierling, E. (2002). Changes in oligomerization are
essential for the chaperone activity of a small heat shock protein in
vivo and in vitro. J. Biol. Chem. 277, 46310–46318.
Glover, J.R., and Lindquist, S.L. (1998). Hsp104, Hsp70, and Hsp40: A
novel chaperone system that rescues previously aggregated proteins.
Cell 94, 73–82.
Goloubinoff, P., Mogk, A., Zvi, A.P., Tomoyasu, T., and Bukau, B.
(1999). Sequential mechanism of solubilization and refolding of stable
protein aggregates by a bichapher one network. Proc. Natl. Acad. Sci.
USA 96, 13732–13737.
Hattendorf, D.A., and Lindquist, S.L. (2002a). Cooperative kinetics of
both Hsp104 ATPase domains and interdomain communication
revealed by AAA sensor-1 mutants. EMBO J. 21, 12–21.
Hattendorf, D.A., and Lindquist, S.L. (2002b). Analysis of the AAA
sensor-2 motif in the C-terminal ATPase domain of Hsp104 with a site-
specific fluorescent probe of nucleotide binding. Proc. Natl. Acad.
Sci. USA 99, 2732–2737.
Hong, S.W., Lee, U., and Vierling, E. (2003). Arabidopsis hot mutants
define multiple functions required for acclimation to high temper-
atures. Plant Physiol. 132, 757–767.
Hong, S.W., and Vierling, E. (2000). Mutants of Arabidopsis thaliana
defective in the acquisition of tolerance to high temperature stress.
Proc. Natl. Acad. Sci. USA 97, 4392–4397.
Hong, S.W., and Vierling, E. (2001). Hsp101 is necessary for heat
tolerance but dispensable for development and germination in the
absence of stress. Plant J. 27, 25–35.
Ishikawa, T., Beuron, F., Kessel, M., Wickner, S., Maurizi, M.R., and
570 The Plant Cell
Steven, A.C. (2001). Translocation pathway of protein substrates in
ClpAP protease. Proc. Natl. Acad. Sci. USA 98, 4328–4333.
Kim, K.I., Cheong, G.W., Park, S.C., Ha, J.S., Woo, K.M., Choi, S.J.,
and Chung, C.H. (2000a). Heptameric ring structure of the heat-
shock protein ClpB, a prot ein-activated ATPase in Escherichia coli.
J. Mol. Biol. 303, 655–666.
Kim, Y.I., Burton, R.E., Burton, B.M., Sauer, R.T., and Baker, T.A.
(2000b). Dynamics of substrate denaturation and translocation by the
ClpXP degradation machine. Mol. Cell 5, 639–648.
Lee, S., Sowa, M.E., Watanabe, Y., Sigler, P.B., Chiu, W., Yoshida,
M., and Tsai, F.T.F. (2003). The structure of ClpB: A molecular
chaperone that rescues protein from an aggregated state. Cell 115,
229–240.
Liu, Z., Tek, V., Akoev, V., and Zolkiewski, M. (2002). Conserved
amino acid residues within the amino-terminal domain of ClpB are
essential for the chaperone activity. J. Mol. Biol. 32, 111–120.
Lum, R., Tkach, J.M., Vier ling, E., and Glover, J.R. (2004). Evidence
for an unfolding/threading mechanism for protein disaggregation by
Saccharomyces cerevisiae Hsp104. J. Biol. Chem. 279, 29139–29146.
Lupas, A.N., and Martin, J. (2002). AAA proteins. Curr. Opin. Struct.
Biol. 12, 746–747.
Maurizi, M.R., and Xia, D. (2004). Protein binding and disruption by
Clp/Hsp100 chaperones. Structure 12, 175–183.
Mogk, A., Deuerling, E., Vorderwulbecke, S., Vierling, E., and Bukau,
B. (2003a). Small heat shock proteins, ClpB and the DnaK system
form a functional triade in reversing protein aggregation. Mol. Micro-
biol. 50, 585–595.
Mogk, A., Schlieker, C., Friedrich, K.L., Schonfeld, H.J., Vierling, E.,
and Bukau, B. (2003c). Refolding of substrates bound to small Hsps
relies on a disaggregation reaction mediated most efficiently by ClpB/
DnaK. J. Biol. Chem. 278, 31033–31042.
Mogk, A., Schlieker, C., Strub, C., Rist, W., Weiberzahn, J., and
Bukau, B. (2003b). Roles of individual domains and conserved motifs
of the AAAþ chaperone ClpB in oligomerization, ATP hydrosis, and
chaperone activity. J. Biol. Chem. 278, 17615–17624.
Motohashi, K., Watanabe, Y., Yohda, M., and Yoshida, M. (1999).
Heat-inactivated proteins are rescued by the DnaK.J-GrpE set and
ClpB chaperones. Proc. Natl. Acad. Sci. USA 96, 7184–7189.
Neuwald, A.F., Aravind, L., Spouge, J.L., and Koonin, E.V. (1999).
AAAþ: A class of chaperone-like ATPases associated with assembly,
operation, and disassembly of protein complexes. Genome Res. 9,
27–43.
Nieto-Sotelo, J., Kannan, K.B., Martı
´
nez, L.M., and Segal, C. (1999).
Characterization of a maize heat-shock protein 101 gene, HSP101,
encoding a ClpB/Hsp100 protein homologue. Gene 230, 187–195.
Ogura, T., and Wilkinson, A.J. (2001). AAAþ superfamily ATPases:
Common structure—diverse function. Genes Cells 6, 575–597.
Schirmer, E.C., Glover, J.R., Singer, M.A., and Lindquist, S.L. (1996).
Hsp100/Clp proteins: A common mechanism explains diverse func-
tions. Trends Biochem. Sci. 21, 289–296.
Schirmer, E.C., Homann, O.R., Kowal, A.S., and Lindquist, S.L.
(2004). Dominant gain-of-function mutations in Hsp104p reveal critical
roles for the middle region. Mol. Biol. Cell 15, 2061–2072 .
Schirmer, E.C., Lindquist, S., and Vierling, E. (1994). An Arabidopsis
heat shock protein complements a thermotolerance defect in yeast.
Plant Cell 6, 1899–1909.
Schirmer, E.C., Ware, D.M., Queitsch, C., Kowal, A.S., and Lindquist,
S.L. (2001). Subunit interactions influence the biochemical and bio-
logical properties of Hsp104. Proc. Natl. Acad. Sci. USA 98, 914–919.
Schlieker, C., Weiberzahn, J., Patzelt, H., Tessarz, P., Strub, C.,
Zeth, K., Erbse, A., Schneider-Mergener, J., Chin, J.W., Schiltz,
P.G., Bukau, B., and Mogk, A. (2004). Substrate recognition by the
AAAþ chaperone ClpB. Nat. Struct. Mol. Biol. 1, 607–615.
Siddiqui, S.M., Sauer, R.T., and Baker, T.A. (2004). Role of the
processing pore of the ClpX AAAþ ATPase in the recognition and
engagement of specific protein substrates. Genes Dev. 18, 369–374.
Smith, C.K., Baker, T.A., and Sauer, R.T. (1999). Lon and Clp family
proteases and chaperones share homologous substrate-recognition
domains. Proc. Natl. Acad. Sci. USA 96, 6678–6682.
Strub, C., Schlieker, C., Bukau, B., and Mogk, A. (2003). Poly-L-lysine
enhances the protein disaggregation activity of ClpB. FEBS Lett. 553,
125–130.
Vale, R.D. (2000). AAA proteins: Lords of the ring. J. Cell Biol. 150,
F13–F19.
Wang, J., Song, J.J., Franklin, M.C., Kamtekar, S., Im, Y.J., Rho,
S.H., Seong, I.S., Lee, C.S., Chung, C.H., and Eom, S.H. (2001b).
Crystal structures of the HslVU peptidase-ATPase complex reveal an
ATP-dependent proteolysis mechanism. Structure 9, 177–1 84.
Wang, J., Song, J.J., Seong, I.S., Franklin, M.C., Kamtekar, S., Eom,
S.H., and Chung, C.H. (2001a). Nucleotide-dependent conforma-
tional changes in a protease-associated ATPase HsIU. Structure 9,
1107–1116.
Watanabe, Y.H., Motohashi, K., and Yoshida, M. (2002). Roles of the
two ATP binding sites of ClpB from Thermus thermophilu s. J. Biol.
Chem. 277, 5804–5809.
Waters, E.R., and Vierling, E. (1999). The diversification of plant
cytosolic small heat shock proteins preceded the divergence of
mosses. Mol. Biol. Evol. 16, 127–139.
Weibezahn, J., Bukau, B., and Mogk, A. (2004). Unscrambling an egg:
Protein disaggregation by AAAþ proteins. Microb. Cell Fact. 3, 1–12.
Weibezahn, J., Schlieker, C., Bukau, B., and Mogk, A. (2003).
Characterization of a trap mutant of the AAAþ chaperone ClpB.
J. Biol. Chem. 278, 32608–32617.
Zolkiewiski, M. (1999). ClpB cooperates with DnaK, DnaJ, and GrpE in
suppressing protein aggregation. J. Biol. Chem. 274, 28083–28086.
NOTE ADDED IN PROOF
Additional published data support threading of the substrate through the
axial channel.
Weibezahn, J., Tessarz, P., Schlieker, C., Zahn, R., Maglica, Z.,
Lee, S., Zentgraf, H., Weber-Ban, E.U., Dougan, D.A., Tsai, F.T.F.,
Mogk, A., and Bukau, B. (2004). Thermotolerance requires refolding
of aggregated proteins by substrate translocation through the central
pore of ClpB. Cell 119, 653–665.
AtHsp101 Structure and Thermotolerance 571
... Notably, some MD mutations cause constitutive derepression of protein activity leading to toxicity in vivo (Schirmer et al., 2004;Oguchi et al., 2012;Lipińska et al., 2013). A gain-of-function mutation in the MD of A. thaliana HSP101 (hot1-4; A499T) results in plants that cannot survive acclimation temperatures that induce HSP101 expressiontemperatures that are not otherwise lethal to HSP101 null plants (hot1-3 mutants) and that are necessary to acclimate wild-type plants to survive severe heat stress (Lee et al., 2005). Sensitivity of hot1-4 plants to heat stress acclimation temperatures indicates that the hot1-4 mutant protein is toxic, as is seen for MD mutations in other organisms. ...
... To understand further the function of HSP101 in mechanisms of plant thermotolerance, we performed suppressor screening of ethyl methanesulfonate-mutagenized seeds of A. thaliana carrying the hot1-4 (A499T) missense mutation in HSP101. In addition to multiple intragenic suppressors (Lee et al., 2005), we also identified extragenic suppressors. We previously reported identification of an extragenic suppressor of hot1-4 1 (shot1), which is a bypass suppressor mutant with defects in mitochondrial function (Kim et al., 2012(Kim et al., , 2021. ...
... The degree of heat sensitivity of dark-grown seedlings can be quantitatively measured by hypocotyl growth inhibition after heat stress, which provides an assay to identify mutants with altered heat stress sensitivity . When A. thaliana seedlings are treated at 38°C for 1.5 h followed by 2 or more hours at optimal growth temperatures, the plants accumulate HSPs and can become acclimated and survive treatments at a more severe, 45°C temperature Lee et al., 2005). However, because 38°C treatment of seedlings carrying the hot1-4 allele produces a toxic, mutant HSP101 protein, hot1-4 seedlings do not grow after a 38°C treatment, which is an otherwise permissive treatment for wild-type and HSP101 null plants (Lee et al., 2005). ...
Article
Heat Shock Protein 101 (HSP101) in plants, and bacterial and yeast orthologs, is essential for thermotolerance. To investigate thermotolerance mechanisms involving HSP101, we performed a suppressor screen in Arabidopsis thaliana of a missense HSP101 allele (hot1-4). hot1-4 plants are sensitive to acclimation heat treatments that are otherwise permissive for HSP101 null mutants, indicating that the hot1-4 protein is toxic. We report one suppressor (shot2, suppressor of hot1-4 2) has a missense mutation of a conserved residue in CLEAVAGE STIMULATION FACTOR77 (CstF77), a subunit of the polyadenylation complex critical for mRNA 3' end maturation. We performed ribosomal RNA depletion RNA-Seq and captured transcriptional readthrough with a custom bioinformatics pipeline. Acclimation heat treatment caused transcriptional readthrough in hot1-4 shot2, with more readthrough in heat-induced genes, reducing the levels of toxic hot1-4 protein and suppressing hot1-4 heat sensitivity. Although shot2 mutants develop like the wild type in the absence of stress and survive mild heat stress, reduction of heat-induced genes and decreased HSP accumulation makes shot2 in HSP101 null and wild-type backgrounds sensitive to severe heat stress. Our study reveals the critical function of CstF77 for 3' end formation of mRNA and the dominant role of HSP101 in dictating the outcome of severe heat stress.
... The genes encoding the components of calcium transporting ATPase, a calcium-dependent protein kinase CPK1 adapter protein, calnexin precursor, annexin, calcium/calmodulin mediated signal pathway, calcineurin B, mitogen-activated protein kinases, cell wall-associated protein kinase were induced under heat treatment, in the present study. [10,36,37]. Heat stress also induces small proteins called sHSPs (Small heat shock proteins), having monomeric masses between 12 and 43 kDa. ...
... A strong correlation between sHSPs accumulation and plant tolerance to heat stress has been proved before [37]. In this study, most HSP genes, including HSP90, HSP70, and sHSPs, were up-regulated in response to heat stress, which aligns with other studies as mentioned above [15, [35][36][37]. Hence, the up-regulated HSP genes must play essential roles in heat tolerance in ragi. ...
Article
Full-text available
Background Eleusine coracana (L.) Gaertn is a crucial C4 species renowned for its stress robustness and nutritional significance. Because of its adaptability traits, finger millet (ragi) is a storehouse of critical genomic resources for crop improvement. However, more knowledge about this crop’s molecular responses to heat stress needs to be gained. Methods and results In the present study, a comparative RNA sequencing analysis was done in the leaf tissue of the finger millet, between the heat-sensitive (KJNS-46) and heat-tolerant (PES-110) cultivars of Ragi, in response to high temperatures. On average, each sample generated about 24 million reads. Interestingly, a comparison of transcriptomic profiling identified 684 transcripts which were significantly differentially expressed genes (DEGs) examined between the heat-stressed samples of both genotypes. The heat-induced change in the transcriptome was confirmed by qRT-PCR using a set of randomly selected genes. Pathway analysis and functional annotation analysis revealed the activation of various genes involved in response to stress specifically heat, oxidation-reduction process, water deprivation, and changes in heat shock protein (HSP) and transcription factors, calcium signaling, and kinase signaling. The basal regulatory genes, such as bZIP, were involved in response to heat stress, indicating that heat stress activates genes involved in housekeeping or related to basal regulatory processes. A substantial percentage of the DEGs belonged to proteins of unknown functions (PUFs), i.e., not yet characterized. Conclusion These findings highlight the importance of candidate genes, such as HSPs and pathways that can confer tolerance towards heat stress in ragi. These results will provide valuable information to improve the heat tolerance in heat-susceptible agronomically important varieties of ragi and other crops.
... It is well known that HSP101 interacted with the sHSPs chaperone system to resolubilize protein aggregates after heat stress. Thereby improving tolerance to heat stress in plants such as Arabidopsis and maturing tomato pollen grains [10,[35][36]. Heat stress also induces small proteins called sHSPs (Small heat shock proteins), having monomeric masses between 12 to 43 kDa. ...
... A strong correlation between sHSPs accumulation and plant tolerance to heat stress has been proved before [36]. In this study, most HSP genes, including HSP90, HSP70, and sHSPs, were upregulated in response to heat stress, which is in correlation with other studies as mentioned above [15,[34][35][36]. Thus, the upregulated HSP genes play essential roles in heat tolerance in ragi. ...
Preprint
Full-text available
Background Eleusine coracana (L.) Gaertn is a crucial C4 species renowned for its stress robustness and nutritional significance. Because of its adaptability traits, finger millet (ragi) is a storehouse of critical genomic resources for crop improvement. However, more knowledge about this crop's molecular responses to heat stress must be gained. Methods and Results In the present study, a comparative RNA sequencing analysis was done in the leaf tissue of the finger millet between the heat-sensitive (KJNS-46) and heat-tolerant (PES-110) cultivars of Ragi in response to high temperatures. On average, each sample generated about 24 million reads. Interestingly, a comparison of transcriptomic profiling identified 684 transcripts which were significantly differentially expressed genes (DEGs) examined between the heat-stressed samples of both genotypes. The heat-induced change in the transcriptome was confirmed by qRT-PCR using a set of randomly selected genes. Pathway analysis and functional annotation analysis revealed the activation of various genes involved in response to stress, precisely heat, oxidation-reduction process, water deprivation, heat shock protein (HSP) and transcription factors, calcium, and kinase signaling. The basal regulatory genes, such as bZIP, were involved in response to heat stress, indicating that heat stress activates genes related to basal regulatory processes or housekeeping. A substantial percentage of the DEGs belonged to proteins of unknown functions (PUFs), i.e., uncharacterized. Conclusion The finding highlights the importance of HSPs, candidate genes, and pathways that can confer tolerance towards heat stress in ragi. These results will provide valuable information to improve heat tolerance in heat-susceptible agronomically important varieties of ragi and other crop plants.
... It is well known that HSP101 interacted with the sHSPs chaperone system to resolubilize protein aggregates after heat stress. Thereby improving tolerance to heat stress in plants such as Arabidopsis and maturing tomato pollen grains (Lee et al. 2005;Pressman et al. 2007;Frank et al. 2009). Heat stress also induces small proteins called sHSPs (Small heat shock proteins), having monomeric masses between 12 to 43 kDa. ...
... A strong correlation between sHSPs accumulation and plant tolerance to heat stress has been proved before (Pressman et al. 2007). In this study, most HSP genes, including HSP90, HSP70, and sHSPs, were upregulated in response to heat stress, which is in correlation with other studies as mentioned above (Myouga et al. 2006;Lee et al. 2005;Pressman et al. 2007;Wang et al. 2004). Thus, the upregulated HSP genes play essential roles in heat tolerance in ragi. ...
Preprint
Full-text available
Eleusine coracana (L.) Gaertn is a crucial C4 species renowned for its stress robustness and nutritional significance. Because of its adaptability traits, finger millet (ragi) is a storehouse of critical genomic resources for crop improvement. However, more knowledge about this crop's molecular responses to heat stress must be gained. Hence, in the present study, we generated RNA seq data from the leaf tissue of the finger millet to observe the physiological changes and gene expression study in heat-sensitive (KJNS-46) and heat-tolerant (PES-110) genotypes of Ragi in response to high temperatures. On average, each sample generated about 24 million reads. Nearly 684 transcripts were differentially expressed (DEGs) between the heat-stressed samples of both genotypes. Pathway analysis and functional annotation analysis revealed the activation of various genes involved in response to stress, precisely heat, oxidation-reduction process, water deprivation, heat shock protein and transcription factors, calcium, and kinase signaling. The basal regulatory genes, such as bZIP, were involved in response to heat stress, indicating that heat stress activates genes related to basal regulatory processes or housekeeping. A substantial percentage of the DEGs belonged to proteins of unknown functions (PUFs), i.e., uncharacterized. The expression pattern of a few selected DEGs genes was analyzed in both genotypes by quantitative RT-PCR. The present study found some candidate genes and pathways that may confer tolerance to heat stress in ragi. These results will provide valuable information to improve heat tolerance in heat-susceptible agronomically important varieties of ragi and other crop plants.
... All plant genomes sequenced contain ClpB1/Hsp100 (cytoplasmic form of ClpB) gene (Erdayani et al., 2020;Kumar et al., 2020;Grover, 2019, 2016). The transcript and protein amount of this gene are usually below detection limit under non-HS and are strongly up regulated upon HS Hong and Vierling, 2001;Lee et al., 2005;Singh et al., 2010). Arabidopsis Hsp101 protein interacts with proteasome to Abbreviations: AT-seedling, seedlings subjected to acquired heat stress treatment; BT-seed, seeds subjected to basal heat stress treatment; BT-seedling, seedlings subjected to basal heat stress treatment; C lines, lines developed with rice Hsp101 cDNA driven by CaMV 35 S promoter; CaMV35Sp, cauliflower mosaic virus 35S promoter; HS, heat stress; HT, heat tolerance; IN lines, lines developed with rice Hsp101 cDNA driven by Arabidopsis Hsp101 promoter; GF lines, lines developed with 4633 bp Hsp101 genomic fragment from Arabidopsis containing both its coding and the regulatory sequences; OX line, Hsp101 over-expressing GF lines; UX line, Hsp101 under-expressing GF lines. ...
... bring about clearing of the ubiquitinated protein aggregates (McLoughlin et al., 2019). AtHsp101 genetically interacts with sHsps (Lee et al., 2005): CI and CII sHsps together with Hsp101 protect protein translation factors under HS (McLoughlin et al., 2016). Hsp101 affects the release of ribosomal protein mRNAs from stress granules (SGs) during recovery after HS (Merret et al., 2017). ...
Article
Hsp101 chaperone is vital for survival of plants under heat stress. We generated transgenic Arabidopsis thaliana (Arabidopsis) lines with extra copies of Hsp101 gene using diverse approaches. Arabidopsis plants transformed with rice Hsp101 cDNA driven by Arabidopsis Hsp101 promoter (IN lines) showed high heat tolerance while the plants transformed with rice Hsp101 cDNA driven by CaMV35S promoter (C lines) were like wild type plants in heat stress response. Transformation of Col-0 plants with 4633bp Hsp101 genomic fragment (GF lines) from A. thaliana containing both its coding and the regulatory sequence resulted in mostly over-expressor (OX) lines and a few under-expressor (UX) lines of Hsp101. OX lines showed enhanced heat tolerance while the UX lines were overly heat sensitive. In UX lines, silencing of not only Hsp101 endo-gene was noted but also transcript of choline kinase (CK2) was silenced. Previous work established that in Arabidopsis, CK2 and Hsp101 are convergent gene pairs sharing a bidirectional promoter. The elevated AtHsp101 protein amount in most GF and IN lines was accompanied by lowered CK2 transcript levels under HS. We observed increased methylation of the promoter and gene sequence region in UX lines; however, methylation was lacking in OX lines.
... Plants express many different AAA+ proteins, which are found in all cellular compartments and membranes. These include the HSP100/ClpB chaperone family (which contains two AAA+ domains) involved in unfolding protein aggregates (Lee et al., 2005(Lee et al., , 2007McLoughlin et al., 2019), the Lon (Tsitsekian et al., 2019), FtsH (Yi et al., 2022) and Clp (Bouchnak & van Wijk, 2021) proteases, the chaperone p97/CDC48 (also with two AAA+ domains) (Bègue et al., 2017), the microtubule-severing protein katanin (Komis et al., 2017), subunits of the 19S regulatory particle of the proteasome , as well as others. All of these AAA+ proteins have protein substrates and are engaged in aspects of protein quality control or protein structure modulation. ...
Article
Full-text available
ATAD3 proteins (ATPase family AAA domain-containing protein 3) are unique mitochondrial proteins that arose deep in the eukaryotic lineage but that are surprisingly absent from the Fungi and Amoebozoa. These ~600 amino acid proteins are anchored in the inner mitochondrial membrane and are essential in metazoans and Arabidopsis thaliana. ATAD3s comprise a C-terminal AAA+ matrix domain and an ATAD3_N domain that is located primarily in the inner membrane space but potentially extends into cytosol to interact with the ER. Sequence and structural alignments indicate ATAD3 proteins are most similar to classic chaperone unfoldases in AAA+ family, suggesting that they operate in mitochondrial protein quality control. A. thaliana has four ATAD3 genes in two distinct clades that appear first in the seed plants, and both clades are essential for viability. The four genes are generally coordinately expressed, and transcripts are highest in growing apices and imbibed seeds. Plants with disrupted ATAD3 have reduced growth, aberrant mitochondrial morphology, diffuse nucleoids and reduced oxidative phosphorylation complex I. These and other pleiotropic phenotypes are also observed in ATAD3 mutants in metazoans. Here we discuss the distribution of ATAD3 proteins as they have evolved in the plant kingdom, their unique structure, what we know about their function in plants, and the challenges in determining their essential roles in mitochondria.
... This unique sequence feature of Hsp42 makes it difficult to predict whether other sHsp family members act as sequestrases as well. On the other hand, sHsps are frequently found associated with insoluble misfolded proteins in heat-stressed or aged cells and can even represent the most abundant single protein species of these aggregates (Coelho et al., 2014;Laskowska et al., 1996;Lee et al., 2005;Walther et al., 2015). This points to the critical role of sHsps in modulating protein aggregation, though direct evidence for a general sequestrase function is lacking. ...
Article
Full-text available
The chaperone-mediated sequestration of misfolded proteins into inclusions is a pivotal cellular strategy to maintain proteostasis in Saccharomyces cerevisiae, executed by small heat shock proteins (sHsps) Hsp42 and Btn2. Direct homologs of Hsp42 and Btn2 are absent in other organisms, questioning whether sequestration represents a conserved proteostasis strategy and, if so, which factors are involved. We examined sHsps from Escherchia coli, Caenorhabditis elegans, and humans for their ability to complement the defects of yeast sequestrase mutants. We show that sequestration of misfolded proteins is an original and widespread activity among sHsps executed by specific family members. Sequestrase positive C. elegans' sHsps harbor specific sequence features, including a high content of aromatic and methionine residues in disordered N-terminal extensions. Those sHsps buffer limitations in Hsp70 capacity in C. elegans WT animals and are upregulated in long-lived daf-2 mutants, contributing to lifespan extension. Cellular protection by sequestration of misfolded proteins is, therefore, an evolutionarily conserved activity of the sHsp family.
... Small heat shock proteins (sHSPs) are important molecular chaperones, which can prevent damaged protein aggregation caused by stress. sHSP help damaged protein refold to restore biological function by cooperating with other HSPs (HSP100 or HSP70) in the presence of ATP (Lee et al., 2005;Bernfur et al., 2017;Waters and Vierling, 2020). Previous studies have shown sHSP can enhance plant tolerance to external stress (Chauhan et al., 2012;Kuang et al., 2017;Guo et al., 2020). ...
Article
Full-text available
The F-box gene family is one of the largest gene families in plants. These genes regulate plant growth and development, as well as biotic and abiotic stress responses, and they have been extensively researched. Drought stress is one of the major factors limiting the yield and quality of soybean. In this study, bioinformatics analysis of the soybean F-box gene family was performed, and the role of soybean F-box-like gene GmFBL144 in drought stress adaptation was characterized. We identified 507 F-box genes in the soybean genome database, which were classified into 11 subfamilies. The expression profiles showed that GmFBL144 was highly expressed in plant roots. Overexpression of GmFBL144 increased the sensitivity of transgenic Arabidopsis to drought stress. Under drought stress, the hydrogen peroxide (H2O2) and malonaldehyde (MDA) contents of transgenic Arabidopsis were higher than those of the wild type (WT) and empty vector control, and the chlorophyll content was lower than that of the control. Y2H and bimolecular fluorescence complementation (BiFC) assays showed that GmFBL144 can interact with GmsHSP. Furthermore, our results showed that GmFBL144 can form SCFFBL144 (E3 ubiquitin ligase) with GmSkp1 and GmCullin1. Altogether, these results indicate that the soybean F-box-like protein GmFBL144 may negatively regulate plant drought stress tolerance by interacting with sHSP. These findings provide a basis for molecular genetics and breeding of soybean.
... AtClpB3 can function as a chaperone completely independent of the AtClp holocomplex. The action of AtClpB3 most likely consists in resolubilizing the interprotein aggregates formed during A. thaliana exposure to thermal stress [17] and light with increased intensity [18], and in stress-free environmental conditions, AtClpB3 participates in biogenesis chloroplasts from proplastids [19]. ...
Article
Degradation of chloroplast proteins within the organelle is supported by the observation that chloroplasts contain several proteases of the ClpP, FtsH, Deg, and Lon families. Clp proteases were among the first identified chloroplasts’ proteases and may play an important role during chloroplast biogenesis. Some members of the ClpP family (i.e., nclpP3 and nclpP5) are up-regulated during senescence, whereas the expression of other Clp proteases is constitutive, with no changes during leaf ontogeny. Interestingly, the mRNA levels of erd1, a Clp regulatory subunit are up-regulated during senescence of Arabidopsis thaliana leaves, but the levels of the corresponding ERD1 protein decline. Homologs of the bacterial FtsH protease are also found in plastids. At least 12 FtsH proteases have been identified in Arabidopsis thaliana, and some of them may play roles in thylakoid protein degradation. An FtsH protease is involved in the breakdown of the 23-kDa fragment of the D1 protein of the PSII reaction centre, which is formed upon photooxidative damage. Chloroplast DegP and FtsH proteases seem to cooperate in D1 degradation during photoinhibition, and it seems likely that they might also be responsible for D1 degradation during senescence. In vitro studies with thylakoids isolated from knock-out lines for FtsH6 have implicated the involvement of this protease in LHCII degradation during senescence. Other FtsH subunits may function in chloroplast biogenesis rather than senescence. In this article, we show which proteases are involved in the degradation of chloroplast proteins. We will focus on both: intrachloroplast and non-chloroplast proteases and their mechanism of the process.
Article
J-domain proteins (JDPs) are critical components of the cellular protein quality control machinery, playing crucial roles in preventing the formation and solubilizing cytotoxic protein aggregates. Bacteria, yeast and plants additionally have large, multimeric Hsp100-class disaggregases that resolubilize protein aggregates. JDPs interact with aggregated proteins and specify the aggregate-remodeling activities of Hsp70s and Hsp100s. However the aggregate-remodeling properties of plant JDPs are not well understood. Here we identify eight orthologs of Sis1 (an evolutionarily-conserved Class II JDP of budding yeast) in Arabidopsis thaliana with distinct aggregate-remodeling functionalities. Six of these JDPs associate with heat-induced protein aggregates in vivo and colocalize with Hsp101 at heat-induced protein aggregate centers. Consistent with a role in solubilizing cytotoxic protein aggregates, an atDjB3 mutant had defects in both solubilizing heat-induced aggregates and acquired thermotolerance as compared to wild-type seedlings. Next, we used yeast prions as protein aggregate models to show that the six JDPs have distinct aggregate-remodeling properties. Results presented in this study, as well as findings from phylogenetic analysis, demonstrate that plants harbour multiple, evolutionarily conserved JDPs with capacity to process a variety of protein aggregate conformers induced by heat and other stressors.
Article
Full-text available
Article
Full-text available
Point mutations in either of the two nucleotide-binding domains (NBD) of Hsp104 (NBD1 and NBD2) eliminate its thermotolerance function in vivo. In vitro, NBD1 mutations virtually eliminate ATP hydrolysis with little effect on hexamerization; analogous NBD2 mutations reduce ATPase activity and severely impair hexamerization. We report that high protein concentrations overcome the assembly defects of NBD2 mutants and increase ATP hydrolysis severalfold, changing Vmax with little effect on Km. In a complementary fashion, the detergent 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate inhibits hexamerization of wild-type (WT) Hsp104, lowering Vmax with little effect on Km. ATP hydrolysis exhibits a Hill coefficient between 1.5 and 2, indicating that it is influenced by cooperative subunit interactions. To further analyze the effects of subunit interactions on Hsp104, we assessed the effects of mutant Hsp104 proteins on WT Hsp104 activities. An NBD1 mutant that hexamerizes but does not hydrolyze ATP reduces the ATPase activity of WT Hsp104 in vitro. In vivo, this mutant is not toxic but specifically inhibits the thermotolerance function of WT Hsp104. Thus, interactions between subunits influence the ATPase activity of Hsp104, play a vital role in its biological functions, and provide a mechanism for conditionally inactivating Hsp104 function in vivo.
Article
Full-text available
The heat shock protein Hsp104 of the yeast Saccharomyces cerevisiae plays a key role in promoting survival at extreme temperatures. We found that when diverse higher plant species are exposed to high temperatures they accumulate proteins that are antigenically related to Hsp104. We isolated a cDNA corresponding to one of these proteins from Arabidopsis. The protein, AtHSP101, is 43% identical to yeast Hsp104. DNA gel blot analysis indicated that AtHSP101 is encoded by a single- or low-copy number gene. AtHsp101 mRNA was undetectable in the absence of stress but accumulated to high levels during exposure to high temperatures. When AtHSP101 was expressed in yeast, it complemented the thermotolerance defect caused by a deletion of the HSP104 gene. The ability of AtHSP101 to protect yeast from severe heat stress strongly suggests that this HSP plays an important role in thermotolerance in higher plants.
Article
Small heat shock proteins (sHsps) can efficiently prevent the aggregation of unfolded proteins in vitro. However, how this in vitro activity translates to function in vivo is poorly understood. We demonstrate that sHsps of Escherichia coli, IbpA and IbpB, co-operate with ClpB and the DnaK system in vitro and in vivo, forming a functional triade of chaperones. IbpA/IbpB and ClpB support independently and co-operatively the DnaK system in reversing protein aggregation. A delta ibpAB delta clpB double mutant exhibits strongly increased protein aggregation at 42 degrees C compared with the single mutants. sHsp and ClpB function become essential for cell viability at 37 degrees C if DnaK levels are reduced. The DnaK requirement for growth is increasingly higher for delta ibpAB, delta clpB, and the double delta ibpAB delta clpB mutant cells, establishing the positions of sHsps and ClpB in this chaperone triade.
Article
The bacterial heat shock locus ATPase HslU is an AAA(+) protein that has structures known in many nucleotide-free and -bound states. Nucleotide is required for the formation of the biologically active HslU hexameric assembly. The hexameric HslU ATPase binds the dodecameric HslV peptidase and forms an ATP-dependent HslVU protease. We have characterized four distinct HslU conformational states, going sequentially from open to closed: the empty, SO(4), ATP, and ADP states. The nucleotide binds at a cleft formed by an alpha/beta domain and an alpha-helical domain in HslU. The four HslU states differ by a rotation of the alpha-helical domain. This classification leads to a correction of nucleotide identity in one structure and reveals the ATP hydrolysis-dependent structural changes in the HslVU complex, including a ring rotation and a conformational change of the HslU C terminus. This leads to an amended protein unfolding-coupled translocation mechanism. The observed nucleotide-dependent conformational changes in HslU and their governing principles provide a framework for the mechanistic understanding of other AAA(+) proteins.
Article
Molecular chaperones assist protein folding by facilitating their “forward” folding and preventing aggregation. However, once aggregates have formed, these chaperones cannot facilitate protein disaggregation. Bacterial ClpB and its eukaryotic homolog Hsp104 are essential proteins of the heat-shock response, which have the remarkable capacity to rescue stress-damaged proteins from an aggregated state. We have determined the structure of Thermus thermophilus ClpB (TClpB) using a combination of X-ray crystallography and cryo-electron microscopy (cryo-EM). Our single-particle reconstruction shows that TClpB forms a two-tiered hexameric ring. The ClpB/Hsp104-linker consists of an 85 Å long and mobile coiled coil that is located on the outside of the hexamer. Our mutagenesis and biochemical data show that both the relative position and motion of this coiled coil are critical for chaperone function. Taken together, we propose a mechanism by which an ATP-driven conformational change is coupled to a large coiled-coil motion, which is indispensable for protein disaggregation.
Article
Functional chaperone cooperation between Hsp70 (DnaK) and Hsp104 (ClpB) was demonstrated in vitro. In a eubacterium Thermus thermophilus, DnaK and DnaJ exist as a stable trigonal ring complex (TDnaK\cdot J complex) and the dnaK gene cluster contains a clpB gene. When substrate proteins were heated at high temperature, none of the chaperones protected them from heat inactivation, but the TDnaK\cdot J complex could suppress the aggregation of proteins in an ATP- and TGrpE-dependent manner. Subsequent incubation of these heated preparations at moderate temperature after addition of TClpB resulted in the efficient reactivation of the proteins. Reactivation was also observed, even though the yield was low, if the substrate protein alone was heated and incubated at moderate temperature with the TDnaK\cdot J complex, TGrpE, TClpB, and ATP. Thus, all these components were necessary for the reactivation. Further, we found that TGroEL/ES could not substitute TClpB.
Article
Molecular chaperones assist protein folding by facilitating their "forward" folding and preventing aggregation. However, once aggregates have formed, these chaperones cannot facilitate protein disaggregation. Bacterial ClpB and its eukaryotic homolog Hsp104 are essential proteins of the heat-shock response, which have the remarkable capacity to rescue stress-damaged proteins from an aggregated state. We have determined the structure of Thermus thermophilus ClpB (TClpB) using a combination of X-ray crystallography and cryo-electron microscopy (cryo-EM). Our single-particle reconstruction shows that TClpB forms a two-tiered hexameric ring. The ClpB/Hsp104-linker consists of an 85 A long and mobile coiled coil that is located on the outside of the hexamer. Our mutagenesis and biochemical data show that both the relative position and motion of this coiled coil are critical for chaperone function. Taken together, we propose a mechanism by which an ATP-driven conformational change is coupled to a large coiled-coil motion, which is indispensable for protein disaggregation.
Article
Hsp101 is a molecular chaperone that is required for the development of thermotolerance in plants and other organisms. We report that Arabidopsis thaliana Hsp101 is also regulated during seed development in the absence of stress, in a pattern similar to that seen for LEA proteins and small Hsps; protein accumulates during mid-maturation and is stored in the dry seed. Two new alleles of the locus encoding Hsp101 (HOT1) were isolated from Arabidopsis T-DNA mutant populations. One allele, hot1-3, contains an insertion within the second exon and is null for Hsp101 protein expression. Despite the complete absence of Hsp101 protein, plant growth and development, as well as seed germination, are normal, demonstrating that Hsp101 chaperone activity is not essential in the absence of stress. In thermotolerance assays hot1-3 shows a similar, though somewhat more severe, phenotype to the previously described missense allele hot1-1, revealing that the hot1-1 mutation is also close to null for protein activity. The second new mutant allele, hot1-2, has an insertion in the promoter 101 bp 5′ to the putative TATA element. During heat stress the hot1-2 mutant produces normal levels of protein in hypocotyls and 10-day-old seedlings, and it is wild type for thermotolerance at these stages. Thus this mutation has not disrupted the minimal promoter sequence required for heat regulation of Hsp101. The hot1-2 mutant also expresses Hsp101 in seeds, but at a tenfold reduced level, resulting in reduced thermotolerance of germinating seeds and underscoring the importance of Hsp101 to seed stress tolerance.
Article
Protein contents of crude extracts from plant and animal tissues can be rapidly assayed using a Coomassie blue dye-binding procedure combined with scanning densitometry. Total protein is extracted from 100 mg of fresh-frozen or dried-ground tissue using 1 ml of extraction buffer. One-microliter aliquots of standard solutions or crude extracts are spotted in rows on a suitably sized sheet of Whatman 3MM chromatography paper. The dried samples are stained with Coomassie brilliant blue R-250 (0.2%, w/v, in acidified 50% MeOH) for 20 min and rinsed twice with acidified 20% MeOH. After drying, protein concentrations are read as reflectance using a scanning densitometer and peak heights or peak areas recorded using a digital integrator. In an alternative procedure, each spot is cut from the sample sheet and the dye-protein complex eluted in 1% sodium dodecyl sulfate (SDS) using an ultrasonic cleaner. Absorbance is subsequently read in a microwell sample holder at 590 nm with an enzyme-linked immunosorbent assay plate reader. Both procedures offer distinct advantages over previously reported methods. They are significantly faster when large numbers of samples are processed, they avoid interference by chlorophyll, dithiothreitol, SDS, 2-mercaptoethanol, Nonidet P-40, and phenylmethylsulfonyl fluoride (and other protease inhibitors) and they yield marked improvements in sensitivity, providing measurements of protein concentration below 100 and 200 ng.microliter-1, respectively.