ArticlePDF Available

Identification and Characterization of the First Small Molecule Inhibitor of MDMX

Authors:

Abstract and Figures

The p53 pathway is disrupted in virtually every human tumor. In approximately 50% of human cancers, the p53 gene is mutated, and in the remaining cancers, the pathway is dysregulated by genetic lesions in other genes that modulate the p53 pathway. One common mechanism for inactivation of the p53 pathway in tumors that express wild-type p53 is increased expression of MDM2 or MDMX. MDM2 and MDMX bind p53 and inhibit its function by distinct nonredundant mechanisms. Small molecule inhibitors and small peptides have been developed that bind MDM2 in the p53-binding pocket and displace the p53 protein, leading to p53-mediated cell cycle exit and apoptosis. To date, peptide inhibitors of MDMX have been developed, but no small molecule inhibitors have been reported. We have developed biochemical and cell-based assays for high throughput screening of chemical libraries to identify MDMX inhibitors and identified the first MDMX inhibitor SJ-172550. This compound binds reversibly to MDMX and effectively kills retinoblastoma cells in which the expression of MDMX is amplified. The effect of SJ-172550 is additive when combined with an MDM2 inhibitor. Results from a series of biochemical and structural modeling studies suggest that SJ-172550 binds the p53-binding pocket of MDMX, thereby displacing p53. This lead compound is a useful chemical scaffold for further optimization of MDMX inhibitors that may eventually be used to treat pediatric cancers and various adult tumors that overexpress MDMX or have similar genetic lesions. When combined with selective MDM2 inhibitors, SJ-172550 may also be useful for treating tumors that express wild-type p53.
Identification of diverse chemotypes with candidate MDMX inhibitors. A, work flow schematic of the primary HTS, secondary analysis, and dose-response and cell-based assays. Numbers of compounds that were selected for each round of analysis are indicated. B, histogram of the distribution of MDMX activities for the 1,152 compounds in this study. Gray bars represent the number of compounds within the indicated range of activity from the FITC-FP primary screen; black bars represent the number of those compounds that were confirmed as true positives via dose response. Similarly, the light and dark blue bars represent the distributions from the Texas Red FP retest screen. The Texas Red FP assay is better than the FITC FP assay at discriminating true-positives from false-positives. C, distribution of EC 50 values for MdmX, MDMX, and MDM2 for all 1,152 compounds calculated from the dose response in triplicate using the Texas Red FP assay. D–J, visual representation of seven chemotype clusters. The black triangles are Murcko scaffolds, and screened compounds are represented as nodes that are connected to their parent scaffold by gray lines. Nodes are colored by potency against MDMX (blue high and gray low) and sized according to selective cytotoxicity for retinoblastoma cells versus BJ cells (large selective for retinoblastoma). Large dark blue circles have low binding constants for MDMX and selective cytotoxicity for retinoblastoma cells. Eleven compounds were selected from these chemotype clusters for further characterization.
… 
A and B, space-filling model of the overlaid MDM2-nutlin-3a (teal) with MDMX-p53 (pink) showing SJ-172550 bound to the p53-binding pocket of MDM2/MDMX in the secondary docking pose. C, position of p53 peptide based on the MDM2/peptide crystal structure. D, primary docking pose of SJ-172550 to MDMX superimposed on the crystal structure of p53 peptide bound to MDM2. C and D, the gray is the solvent-excluded surface of MDM2 (Protein Data Bank code 2Z5T), and the blue is the surface for MDMX. Green residues represent single mutations, and residues shown in red formed a quadruple mutant. Mutations designed to displace SJ-172550 based on the models of the most energetically favorable docking poses (Fig. 5, B and D) are Q58D, M61F, Y66I, and Q71D. A second series of residues were changed to make the MDMX-binding pocket more like MDM2 to determine whether SJ-172550 was binding in the p53-binding pocket. These mutations include M53L, H54F, and a quadruple mutant (QUAD) with P95H/S96R/P97K/R103Y. From top to bottom, the green residues are Met-53, His-54, Gln-58, Met-61, Tyr-66, and Gln-71 with the red residues being Arg-103, Pro-97, Ser-96, and Pro-95. E, plot of direct binding of each MDMX protein to Texas Red-labeled p53 peptide exactly as described for the HTS. Each point is the mean S.D. of triplicate assays. The EC 50 values for each protein are indicated. F, competition experiments with each MDMX mutant and increasing concentrations of SJ-172550. Each data point is the mean S.D. of triplicate assays. The proteins that did not show direct binding of p53 peptide could not be analyzed in this competition experiment. G, summary of the EC 50 values for direct binding of p53 peptide and competition by nutlin-3a and SJ-172550. The shaded column (M53L) is one of the mutants that was predicted to make MDMX more like MDM2 and increase nutlin-3a binding while inhibiting SJ-172550 binding without affecting peptide binding. n.d., not determined; WT, wild type.
… 
Content may be subject to copyright.
Michael A. Dyer
Aart G. Jochemsen, R. Kiplin Guy and
Samantha A. Cicero, Brenda A. Schulman,
Catherine A. Regni, Donald Bashford,
Zhu, Nicholas Mills, David C. Smithson,
FangyiLeggy A. Arnold, Antonio M. Ferreira,
Damon Reed, Ying Shen, Anang A. Shelat,
First Small Molecule Inhibitor of MDMX
Identification and Characterization of the
Cell Biology:
doi: 10.1074/jbc.M109.056747 originally published online January 15, 2010
2010, 285:10786-10796.J. Biol. Chem.
10.1074/jbc.M109.056747Access the most updated version of this article at doi:
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
Supplemental material:
http://www.jbc.org/content/suppl/2010/01/15/M109.056747.DC1.html
http://www.jbc.org/content/285/14/10786.full.html#ref-list-1
This article cites 39 references, 18 of which can be accessed free at
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
Identification and Characterization of the First Small
Molecule Inhibitor of MDMX
*
S
Received for publication, August 17, 2009, and in revised form, January 8, 2010 Published, JBC Papers in Press, January 15, 2010, DOI 10.1074/jbc.M109.056747
Damon Reed
‡1
, Ying Shen
‡1
, Anang A. Shelat
§
, Leggy A. Arnold
§
, Antonio M. Ferreira
, Fangyi Zhu
§
, Nicholas Mills
§
,
David C. Smithson
§
, Catherine A. Regni
, Donald Bashford
, Samantha A. Cicero
, Brenda A. Schulman
,
Aart G. Jochemsen**, R. Kiplin Guy
§
, and Michael A. Dyer
2
From the Departments of
Developmental Neurobiology,
§
Chemical Biology and Therapeutics, and
Structural Biology and
Howard Hughes Medical Institute, St. Jude Children’s Research Hospital, Memphis, Tennessee 38105 and the **Department of
Molecular and Cellular Biology, Leiden University Medical Center, 2333 ZA Leiden, Netherlands
The p53 pathway is disrupted in virtually every human tumor.
In 50% of human cancers, the p53 gene is mutated, and in the
remaining cancers, the pathway is dysregulated by genetic
lesions in other genes that modulate the p53 pathway. One com-
mon mechanism for inactivation of the p53 pathway in tumors
that express wild-type p53 is increased expression of MDM2 or
MDMX. MDM2 and MDMX bind p53 and inhibit its function by
distinct nonredundant mechanisms. Small molecule inhibitors
and small peptides have been developed that bind MDM2 in the
p53-binding pocket and displace the p53 protein, leading to
p53-mediated cell cycle exit and apoptosis. To date, peptide
inhibitors of MDMX have been developed, but no small mole-
cule inhibitors have been reported. We have developed bio-
chemical and cell-based assays for high throughput screening of
chemical libraries to identify MDMX inhibitors and identified
the first MDMX inhibitor SJ-172550. This compound binds
reversibly to MDMX and effectively kills retinoblastoma cells in
which the expression of MDMX is amplified. The effect of
SJ-172550 is additive when combined with an MDM2 inhibitor.
Results from a series of biochemical and structural modeling
studies suggest that SJ-172550 binds the p53-binding pocket of
MDMX, thereby displacing p53. This lead compound is a useful
chemical scaffold for further optimization of MDMX inhibitors
that may eventually be used to treat pediatric cancers and vari-
ous adult tumors that overexpress MDMX or have similar
genetic lesions. When combined with selective MDM2 inhibi-
tors, SJ-172550 may also be useful for treating tumors that
express wild-type p53.
Tumorigenesis is a multistep process that involves dysregu-
lation of several pathways that are crucial for cell growth and
survival (1). The p53 pathway regulates cell survival in response
to cellular stress (e.g. DNA damage) or oncogenic stress (e.g. Rb
pathway dysregulation) (2, 3) and is suppressed in virtually
every human cancer by genetic lesions in the p53 gene or other
components of the pathway (4). Approximately half of all can-
cers express wild-type p53, and considerable research over the
past decade has focused on inducing p53-mediated cell death in
these tumors (4, 5). Most efforts to date have focused on inhib-
iting MDM2, a negative regulator of p53 (6–14).
Another key regulator of the p53 pathway is a protein related
to MDM2 called MDMX (15–17). MDM2 and MDMX share
homology in their p53-binding domains, but MDMX is
believed to regulate p53 through distinct mechanisms. Specifi-
cally, MDM2 primarily regulates p53 stability and subcellular
localization, whereas MDMX may directly regulate p53 tran-
scription (17–21). MDMX is genetically amplified in 19% of
breast carcinomas, 19% of colon carcinomas, 18% of lung car-
cinomas, and a smaller percentage of gliomas (17). One of the
best characterized tumors with an MDMX amplification is ret-
inoblastoma. Approximately 65% of human retinoblastomas
have increased MDMX copy number, which correlates with
increased MDMX mRNA and protein (22). Previous studies
have demonstrated that the MDMX amplification suppresses
p53-mediated cell death in Rb pathway-deficient retinoblasts
(22).
A general consensus is emerging that to efficiently induce a
p53 response in tumor cells that express wild-type p53, it may
be necessary to inactivate both MDM2 and MDMX (18, 23, 24).
To date, no screens to identify small molecule inhibitors of
MDMX have been reported, and MDM2 inhibitors probably do
not bind as efficiently to MDMX because of structural differ-
ences in the p53-binding pockets of the two proteins (25–27).
Consistent with this theory, nutlin-3a binds MDMX with at
least a 40-fold weaker equilibrium binding constant than for
MDM2 (22).
Therefore, to identify small molecules that bind MDMX and
prevent its interaction with p53, we developed biochemical and
cell-based assays suitable for high throughput screening (HTS)
3
of chemical libraries. Using this approach, we have identified
the first MDMX inhibitor, SJ-172550, and demonstrated that it
*This work was supported, in whole or in part, by National Institutes of Health
grants from NCI, Cancer Center Support. This work was also supported by
Howard Hughes Medical Institute, American Cancer Society, Research to
Prevent Blindness, Pearle Vision Foundation, International Retinal
Research Foundation, Pew Charitable Trust, and American Lebanese Syr-
ian Associated Charities (to M. A. D.).
S
The on-line version of this article (available at http://www.jbc.org) contains
supplemental “Materials and Methods,” Figs. 1–9, Table 1, and additional
references.
1
Both authors contributed equally to this work.
2
To whom correspondence should be addressed: MS323, St. Jude Children’s
Research Hospital, 262 Danny Thomas Place, Memphis, TN 38105. Tel.: 901-
595-2257; Fax: 901-595-3143; E-mail: michael.dyer@stjude.org.
3
The abbreviations used are: HTS, high throughput screening; FP, fluores-
cence polarization; FITC, fluorescein isothiocyanate; MALDI, matrix-as-
sisted laser desorption ionization; IR, ionizing radiation; GST, glutathione
S-transferase; shRNA, small hairpin RNA.
THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 285, NO. 14, pp. 10786 –10796, April 2, 2010
© 2010 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.
10786 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 14• APRIL 2, 2010
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
can efficiently kill MDMX-amplified retinoblastoma cells.
SJ-172550 functions in an additive manner with the MDM2
inhibitor nutlin-3a, thereby confirming the importance of tar-
geting both of these negative regulators of p53 in cancer cells.
This validated MDMX inhibitor provides a valuable lead com-
pound and chemical scaffold for further chemical modification
to develop a high affinity MDMX inhibitor with good bioavail-
ability, pharmacokinetics, and pharmacodynamics.
EXPERIMENTAL PROCEDURES
Plasmid Constructs and Protein Production—The p53-bind-
ing domain of mouse and human MDMX (amino acids 1–185)
and human MDM2 (amino acids 1–188) were amplified by PCR
and cloned into the pGEX-4T1 plasmid. Recombinant GST
fusion proteins were prepared in BL21 (DE3) Escherichia coli
cells. The lysates were cleared by spinning at 100,000 g,
and the supernatant was loaded onto a 5-ml GSTrap Fast-
Flow column (GE Healthcare). Subsequent purification
included a Mono Q column and an S200 gel filtration col-
umn. Peak fractions were combined and dialyzed against
phosphate-buffered saline (pH 7.6) containing 2 mMphenyl-
methylsulfonyl fluoride.
Fluorescence Polarization Assays—Fluorescence polarization
(FP) assays were conducted in assay buffer containing 10 mM
Tris (pH 8.0), 42.5 mMNaCl, and 0.0125% Tween 20. The wild-
type p53 peptide (amino acids 15–29) was GSGSSQETF-
SDLWKLLPEN, and the mutant AAA-p53 peptide was
GSGSSQETFADLAKLAPEN. The FP assays were carried out
using 2.5 nMFITC peptide (or 15 nMTexas Red) and 1
M
GST-MDMX or GST-MDM2. For MDM2-p53 or MDMX-p53
inhibitor assays, small molecules were preincubated with the
recombinant protein for 30 min. The labeled peptide was then
added and incubated for 45 min. FP assays were conducted in
384-well black microplates (Corning Glass). The FP FITC
assays were analyzed using an EnVision multilabel plate reader
with a 480-nm excitation filter, a 535-nm static and polarized
filter, and an FP FITC dichroic mirror. The unlabeled compet-
itor peptide and nutlin-3 were used as positive controls, and the
alanine-substituted p53 peptide (AAA-p53) was used as a neg-
ative control. To minimize the possibility of false-positives
caused by endogenous fluorescence from the compounds in the
library, we also developed an FP assay with the Texas Red fluo-
rophore. This assay was conducted as described above, except it
required a 555-nm excitation filter, a 632-nm static and polar-
ized filter, and a Texas Red FP dichroic mirror.
Chemical Library and High Throughput Screening—The
screening library consisted of 295,848 unique compounds from
commercial sources (ChemDiv, ChemBridge, and Life Chemi-
cals) arrayed individually at 10 mMin DMSO in 384-well
polypropylene plates. The purity of compounds was reported
by the vendor as 90%. HTS was carried out on a system devel-
oped by high resolution engineering with integrated plate incu-
bators (Liconic). Plates were transferred from instrument to
instrument by a Stau¨bli T60 robot arm. Assay materials were
dispensed in bulk by using Matrix Wellmates (Matrix Technol-
ogies). Compound plates were centrifuged in a Vspin plate cen-
trifuge (Velocity11). All compound transfers were accom-
plished by using a 384-well pin tool with 10-nl hydrophobic
surface-coated pins (V & P Scientific). These pins allowed for
the delivery of 25 nl to achieve a final compound concentration
of 10
M. The fluorescent signal was measured using an EnVision
multilabel plate reader.
RESULTS
Characterization of an MDMX-p53 Binding Assay for High
Throughput Screening—To identify MDMX inhibitors by HTS
of chemical libraries, we developed an FP assay (28) to detect
the binding of the p53 peptide to GST-MDMX-(1–185) in 384-
well plates. This assay is based on the retention of polarization
during fluorescence spectroscopy of the p53 peptide conju-
gated to a fluorophore such as FITC (Fig. 1A). FP is inversely
proportional to the rotational diffusion of the fluorophore, and
for our experiments, it was measured using an FP spectrometer.
The polarization of free p53-FITC peptide (2.5 nM) was 100 mP,
and with increasing concentrations of purified GST-MDMX-
(1–185) protein (supplemental Fig. 1, AD), the polarization
peaked (100%) at 280 mP (Fig. 1B). Similar data were obtained
using GST-MDM2-(1–188) (Fig. 1Band supplemental
Fig. 1, AD). From these curves, the EC
50
for p53 peptide-
MDMX was 0.36
Mand that for p53 peptide-MDM2 was 0.23
M(Fig. 1B). Biacore experiments provided similar binding
constants for GST-MDMX-(1–185) (K
d
1.05
M) and GST-
MDM2-(1–188) (K
d
1.03
M)(supplemental Fig. 1E). Iso-
thermal titration calorimetry measurements using a purified
minimal p53-binding domain of MDMX-(23–111) confirmed
these binding constants (supplemental Fig. 1, FH).
To test the specificity of our FP assay for MDMX/MDM2
binding to p53, we performed a competition experiment with
unlabeled p53 peptide and a version of the AAA-p53 peptide
that is defective for binding to MDM2/MDMX (29). The pro-
tein concentration in this assay was 1
M, and the p53-FITC
peptide concentration was 2.5 nM. The EC
50
values from the
competition experiments were 0.42 and 0.30
Mfor GST-
MDMX-(1–185) and GST-MDM2-(1–188), respectively (Fig.
1C), whereas AAA-p53 showed no evidence of competition at
any concentration tested (0.5–200
M) (Fig. 1C).
To test whether our FP assay was suitable for identifying
small molecule inhibitors of MDMX-p53 binding, we per-
formed a dose-response experiment using nutlin-3a at con-
centrations ranging from 0.5 nMto 300
M(Fig. 1D). Nut-
lin-3a was originally identified as an MDM2 inhibitor (6), but
it also binds to MDMX, albeit with a much weaker K
d
value
(22). The protein concentration was held constant at 1
M,
and the peptide concentration was 2.5 nMfor each concen-
tration of nutlin-3a tested. The EC
50
value for binding of
nutlin-3a to MDM2 was 0.28
Mand that to MDMX was 20.1
M(Fig. 1D).
High Throughput Screening of a Chemical Library to Identify
Novel MDMX Inhibitors—To identify novel MDMX inhibitors,
we performed an HTS of the St. Jude chemical library (295,848
unique compounds) by using the GST-MDMX-(1–185)/p53-
FITC peptide FP assay (Fig. 1E). A total of 356,352 wells (com-
pounds and controls) were screened over the course of 13 days
by using 928 plates (chemical structures and screening data are
available for download at the following url: www.stjuderesearch.
org/guy/data/mdmx). We selected the mouse MDMX (MdmX)
First MDMX Inhibitor
APRIL 2, 2010VOLUME 285• NUMBER 14 JOURNAL OF BIOLOGICAL CHEMISTRY 10787
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
protein for the HTS, because expression of the recombinant
protein in E. coli was more efficient than human MDMX.
Compounds were screened at a final concentration of 10
M.
The scatterplot of activities demonstrates clear separation
between the positive and negative controls (Fig. 1E). The
average z-prime for the assays was positive but low (0.4)
because of variance in the control distributions (supple-
mental Fig. 2D). However, the reference peptide EC
50
value was
First MDMX Inhibitor
10788 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 14• APRIL 2, 2010
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
stable and within experiment error, and no major plate artifacts
were detected (supplemental Fig. 2).
To reduce the probability of selecting false-positive com-
pounds from the primary screen, we performed a receiver oper-
ating characteristic analysis using different activity thresholds
for compound selection. This analysis led us to select the 70%
activity threshold to obtain 3,596 compounds for subsequent
validation. To eliminate compounds with intrinsic fluorescence
emission spectra that overlapped with FITC, we performed a
secondary FP assay using a Texas Red-conjugated p53 peptide.
This analysis was performed on all 3,596 compounds that met
the 70% activity cutoff (10
M) in triplicate. The top 1,000 com-
pounds were then selected from the Texas Red FP assay based
on their percentage inhibition. An additional 152 structural
analogs of these 1000 compounds were included (total 1,152) to
provide sufficient coverage of chemical scaffolds to begin to
FIGURE 1. Biochemical assays for high throughput screening to identify MDMX inhibitors. A, schematic of the FP assay used to identify MDMX inhibitors.
The protein used for the screen consisted of residues 1–185 of MDMX fused to GST. The peptide (orange) was conjugated to FITC (green) for the primary screen
and to Texas Red for secondary assays. B, plot of the percentage of bound p53-FITC peptide (at a fixed concentration) with increasing concentrations of MDMX
(squares) or MDM2 (triangles). C, plot of the percentage of p53-FITC peptide associated with indicated proteins in the presence of increasing concentrations of
unlabeled wild-type p53 peptide or unlabeled alanine-substituted p53 peptide (AAA-p53) as a negative control. D, plot of the percentage of p53-FITC peptide
associated with the indicated proteins in the presence of increasing concentrations of nutlin-3a. B–D, each data point is the mean S.D. of triplicate
experiments. E, scatterplot of HTS of a chemical library for MdmX inhibitors. The blue data points indicate compounds that were selected for further analysis,
and the black data points are compounds that did not exhibit activity in the HTS. DMSO was used as a negative control (red), and the unlabeled p53 peptide
(green) was used as a positive control. The density plot illustrates the clear separation of the positive and negative control samples across the entire screen. Each
day of screening is separated by a yellow line.
FIGURE 2. Identification of diverse chemotypes with candidate MDMX inhibitors. A, work flow schematic of the primary HTS, secondary analysis, and
dose-response and cell-based assays. Numbers of compounds that were selected for each round of analysis are indicated. B, histogram of the distribution of
MDMX activities for the 1,152 compounds in this study. Gray bars represent the number of compounds within the indicated range of activity from the FITC-FP
primary screen; black bars represent the number of those compounds that were confirmed as true positives via dose response. Similarly, the light and dark blue
bars represent the distributions from the Texas Red FP retest screen. The Texas Red FP assay is better than the FITC FP assay at discriminating true-positives from
false-positives. C, distribution of EC
50
values for MdmX, MDMX, and MDM2 for all 1,152 compounds calculated from the dose response in triplicate using the
Texas Red FP assay. D–J, visual representation of seven chemotype clusters. The black triangles are Murcko scaffolds, and screened compounds are represented
as nodes that are connected to their parent scaffold by gray lines. Nodes are colored by potency against MDMX (blue high and gray low) and sized
according to selective cytotoxicity for retinoblastoma cells versus BJ cells (large selective for retinoblastoma). Large dark blue circles have low binding
constants for MDMX and selective cytotoxicity for retinoblastoma cells. Eleven compounds were selected from these chemotype clusters for further
characterization.
First MDMX Inhibitor
APRIL 2, 2010VOLUME 285• NUMBER 14 JOURNAL OF BIOLOGICAL CHEMISTRY 10789
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
First MDMX Inhibitor
10790 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 14• APRIL 2, 2010
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
establish structure-activity relationships for our candidate
MDMX inhibitors (Fig. 2A).
Analysis of Active Compounds from MDMX Inhibitor High
Throughput Screening—To further characterize the 1,152
active compounds, we measured the binding constant of each
compound to MdmX (mouse), MDMX (human), and MDM2
(human) by performing a dose-response assay in triplicate (Fig.
2, AC). Compounds that showed activity 70% in the primary
FITC MdmX assay displayed a wider range of activities in the
Texas Red retest assay (Fig. 2B, compare gray bars to light blue
bars). Moreover, the median retest activity was a better predic-
tor of validated compounds (compounds with a well behaved
dose response) than the primary screen single point activity.
Therefore, the to further characterize the 1,152 active com-
pounds, we measured the binding constant of each compound
to MdmX (mouse), MDMX (human), and MDM2 (human) by
performing a dose-response assay in triplicate (Fig. 2, AC). FP
assay demonstrated better discriminatory power than the FITC
FP assay.
To complement these biochemical studies, we carried out a
cell-based assay to further characterize the activity of the 1,152
compounds on retinoblastoma cells that have an MDMX
amplification (Weri1) or a cell line that is p53-deficient
(SJmRbl-8) (22). CellTiter-Glo (Promega) assay was used to
measure intracellular ATP levels as an indicator of viability.
SJmRbl-8 cells and Weri1 cells showed a linear relationship
between luminescence and cell number (supplemental
Fig. 3, Aand B). As a positive control for cytotoxicity, we used
vincristine, which is a microtubule inhibitor that disrupts chro-
mosome segregation during mitosis and kills both cell lines
with similar LC
50
values (supplemental Fig. 3C). As a positive
control for p53-selective cytotoxicity, we used nutlin-3a, which
selectively kills Weri1 cells with an MDMX amplification and is
less cytotoxic against p53-deficient SJmRbl-8 cells (supple-
mental Fig. 3D). We also used BJ cells, an hTERT-immortalized
human foreskin fibroblast cell line, as an additional control to
estimate the general cytotoxicity of compounds in our lead
compound collection. We carried out a dose-response cytotox-
icity assay for each cell line on all 1,152 compounds in triplicate.
Overall, the assay performed well in this HTS format, and we
identified a subset of compounds from our active compounds
set with significant selective cytotoxicity against the retinoblas-
toma cells (supplemental Fig. 3, Eand F).
To integrate and visualize the dose-response data from the
biochemical assays of MDMX and MDM2 and the cell-based
data and chemical scaffolds represented in the 1,152 com-
pounds, we overlaid biochemical and cell-based data onto a
network graph constructed to represent related families of che-
motypes within the lead compound set (supplemental Fig. 4).
This approach allowed us to quickly identify scaffolds with the
desired profiles showing strong binding to MDMX and cyto-
toxicity against an MDMX-amplified retinoblastoma cell line.
On the basis of these data, we selected 11 representative com-
pounds from clusters 1, 4, 5, 7, 8, 11, and 54 for further analysis
(Fig. 2, DJ,supplemental Fig. 4, and supplemental Table 1).
Several clusters that looked promising based on the aforemen-
tioned criteria were eliminated from further analysis because of
the presence of potentially problematic chemical functional-
ities such as unsubstituted quinones, maleimides, Michael
acceptors, and thioesters; potential redox activity; or other
medicinal chemistry liabilities (supplemental Fig. 4).
SJ-134433 and SJ-044557 Covalently Modify the MDMX
Protein—Among the 11 compounds selected for further analy-
sis, SJ-134433 and SJ-044557 had excellent profiles (sup-
plemental Fig. 6) with good binding constants for MDMX,
some selectivity for MDMX over MDM2, and efficient killing of
retinoblastoma cells with selectivity for the Weri1 line
(supplemental Table 1). To begin characterizing these com-
pounds, we performed an isothermal denaturation assay and a
redox assay on SJ-134433, SJ-044557, and the other nine com-
pounds (supplemental Fig. 5 and supplemental Table 1). The
rationale underlying the isothermal denaturation assay is that
binding of a small molecule to the p53-binding pocket of
MDMX may stabilize the protein, and in the presence of the
SYPRO orange hydrophobic dye, the temperature for denatur-
ation and dye binding would shift (30, 31). Indeed, GST-
MDMX-(1–185) showed a melting point shift in the presence
of p53 peptide from 46.9 0.6 °C for native GST-MDMX-(1–
185) protein to 50.8 0.6 °C for GST-MDMX-(1–185) protein
bound to the p53 peptide (supplemental Fig. 5, Aand B). How-
ever, neither SJ-044557 nor SJ-134433 exhibited a thermal
shift; in fact, they appeared to destabilize the protein
(supplemental Fig. 5).
To explore the possibility that these two compounds exhib-
ited redox activity, we performed an assay to detect compounds
capable of reducing resazurin to resorufin, a redox couple rele-
vant to oxygen tension in mammalian cells (32). DMSO was
used as a negative control, and a 1,6-dimethylpyrimido[5,4-
e][1,2,4]triazine-5,7(1H,6H)-dione-containing compound was
used as the positive control. SJ-044557 and SJ-134433 showed
some redox activity (supplemental Table 1). Next, we explored
the stability of the compounds in our FP buffer to determine
whether they were unstable and if any of the degradation prod-
ucts were reactive species that covalently modified the MDMX
protein and blocked p53 binding. Both compounds were unstable
after 24 h, and SJ-134433 degraded after2hinFPbuffer
(supplemental Fig. 6, Band E). To directly determine whether the
purified MDMX-(23–111) protein was covalently modified by
FIGURE 3. SJ-172550 reversibly binds MDMX. A, heat map of normalized activity for the 11 compounds selected for follow-up characterization. Dark blue is
more favorable for each measurement, and the compounds are listed in order of binding constant for MDMX from best (top) to worst (bottom). These data
suggested that with subsequent biochemical analyses, SJ-134433 and SJ-044557 were less suitable for follow up, and SJ-172550 was preferred. High perform-
ance liquid chromatography of compound SJ-172550 showed that it is stable for 24 h in FP buffer (Band C), and MALDI mass spectrometry showed that it does
not covalently modify the MDMX-(23–111) protein (D). The MDMX protein was incubated with SJ-172550 for2hinFPbuffer and then dialyzed away in a large
excess of dialysis buffer. The ability of the dialyzed protein to bind the p53-FITC peptide was then measured using the FP assay (red line). E, EC
50
value for
binding to p53 peptide after removal of SJ-172550 was indistinguishable from that of the untreated protein. A similar experiment with nutlin-3a also
demonstrated that it binds reversibly to MDMX. F, SJ-044557 and SJ-134433, which covalently modified MDMX, did not reversibly bind MDMX (red lines). A.U.,
absorbance units.
First MDMX Inhibitor
APRIL 2, 2010VOLUME 285• NUMBER 14 JOURNAL OF BIOLOGICAL CHEMISTRY 10791
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
either compound, we performed high resolution mass spectrom-
etry following incubation with SJ-044557 or SJ-134433. Both
showed a shift in their mass consistent with covalent modifications
(supplemental Fig. 6, Hand I). Although these inhibitors might
prove useful as tools for interrogating MDMX function, they are
not suitable for further development and were thus abandoned.
First MDMX Inhibitor
10792 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 14• APRIL 2, 2010
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
SJ-172550 Is Stable and Reversibly Binds MDMX to Inhibit
p53 Binding—A more detailed analysis of the 11 selected com-
pounds revealed that SJ-172550 had an excellent chemical pro-
file, with respect to chemical stability, thermal stability, redox
potential, and solubility. Unlike SJ-134433 and SJ-044557,
SJ-172550 was stable in solution (Fig. 3, Band C) and did not mod-
ify the MDMX protein by covalent binding in our FP assay buffer
(Fig. 3D). Moreover, the compound exhibited strong thermal sta-
bilization (supplemental Fig. 5) and had undetectable redox activ-
ity (Fig. 3Aand supplemental Table 1). To confirm that SJ-172550
binds MDMX reversibly, we incubated the compound with
MDMX for2hinFPbuffer and then removed the compound by
dialysis. As a positive control, we used the p53 peptide and nutlin-
3a. Our results showed that SJ-172550 bound MDMX reversibly
(Fig. 3E), unlike SJ-044557 or SJ-172550 (Fig. 3F).
SJ-172550 Inhibits MDMX-p53 Binding in Cultured Cells
To test whether SJ-172550 and nutlin-3a have additive or
synergistic effects on retinoblastoma cells, we performed an
isobologram experiment with these two compounds. The data
suggested that nutlin-3a and SJ-172550 act in an additive man-
ner to kill MDMX-amplified human retinoblastoma cells (Fig.
4A). Next, we exposed Weri1 and RB355 retinoblastoma cells
and ML-1 leukemia cells (with wild-type p53) to SJ-172550 (20
M) for 20 h and analyzed the p53 and activated caspase-3 levels
by immunofluorescence to study the mechanism of cell death.
As positive controls, we exposed Weri1 cells to nutlin-3a (5
M)
for 20 h or 5 gray ionizing radiation (IR). DMSO was used as the
negative control. As expected, the Weri1 cells exposed to nut-
lin-3a or IR showed a robust accumulation of p53 (Fig. 4, Band
D). In contrast, the cells exposed to SJ-172550 did not exhibit
the same level of accumulation of p53 (Fig. 4, Band D). This is
consistent with the role of MDMX in regulating transcriptional
activation of p53-responsive promoters but not p53 protein levels.
Apoptosis was robustly induced after exposure to SJ-172550
(Fig. 4, Cand E), and cells exited the cell cycle (Fig. 4F). Real
time RT-PCR and immunoblotting analysis of these cells
revealed that there was also an induction of p53 target genes,
but it was not as robust as that observed with nutlin-3a or IR
(Fig. 4, Gand H). More importantly, the cell death mediated by
SJ-172550 was p53-dependent (Fig. 4, IK). In addition,
HCT116 cells were sensitive to SJ-172550, but p53-deficient
HCT-116 cells were not (supplemental Fig. 8). SJSA-X cells
expressing high levels of MDMX were also sensitive to
SJ-172550 (supplemental Fig. 8).
To determine whether SJ-172550 disrupts the MDMX-p53
interaction of cells in culture, we performed co-immunoprecipita-
tion experiments in the presence of the compound. Reciprocal
co-immunoprecipitation experiments with antibodies against
MDMX and p53 in C33A (human cervical carcinoma) cells dem-
onstrated partial inhibition of MDMX-p53 binding in cells
(supplemental Fig. 7). Similar data were obtained using human
embryonic retina cells and Weri1 retinoblastoma cells (supple-
mental Fig. 7). Together, these results suggest that the MDMX-
p53 interaction was at least partially inhibited by SJ-172550,
despite its relatively low cell permeability (see supple-
mental Table 1).
Computational Model of SJ-172550 Binding to MDMX—X-
ray crystallographic studies have provided high resolution
structures of MDM2 and MDMX bound to p53 (Fig. 5 and
supplemental Fig. 9) (26, 27, 33, 34). We overlaid these two
structures and determined that the C
root mean square
deviation was 3.9 Å, suggesting that although the overall fold
was well conserved, the tertiary structures of MDM2 or
MDMX bound to p53 are significantly different. This can be
more readily visualized using a space-filling representation
of the overlaid structures. In particular, the structure of the
p53-binding pocket of MDMX was smaller than that of
MDM2 (supplemental Fig. 9B). When nutlin-3a was bound
to MDM2, the tertiary structure of the pocket underwent a
small change (C
root mean square deviation 0.82 Å)
(supplemental Fig. 9C). When the structure of MDM2 bound
to nutlin-3a was overlaid with that of MDMX bound to p53,
the smaller binding pocket of MDMX may explain the lower
affinity binding of nutlin-3a to MDMX, as compared with
that of MDM2 (supplemental Fig. 9, Dand E).
We used both AutoDock 4.2 (35) and the fastdock algorithm in
Scigress Explorer version 7.7 to model the binding of nutlin-3a to
MDM2 and found excellent agreement between the computa-
tional model and the native conformation reported in previous
co-crystallization studies (supplemental Fig. 9F) (6). Using the
same computational approach, the binding of SJ-172550 to the
p53-binding pocket of MDMX was modeled, yielding two impor-
tant structures (Fig. 5, Band D, and supplemental Fig. 9G).
Together, these results provide a plausible mechanism of action
for SJ-172550. SJ-172550 may occlude the p53-binding pocket of
MDMX, thereby inhibiting p53 binding. To test this directly, we
generated a series of seven mutants in the MDMX-binding pocket
(Fig. 5, Cand D,supplemental Fig. 9, Fand G, and supplemen-
tal material) and purified the protein for binding studies (Fig. 5,
EG,supplemental Fig. 9, Hand Iand supplemental informa-
tion). Some of the mutants (e.g. H54F) were predicted to displace
SJ-172550 without affecting peptide binding, and other mutants
(e.g. M53L) were predicted to make the binding pocket of MDMX
more like MDM2 and thereby reduce binding of SJ-172550 and
FIGURE 4. SJ-172550 disrupts the MDMX-p53 interaction in cells maintained in culture. A, isobologram shows the additive inhibition of Weri1 cell growth when
the combination of SJ-172550 and nutlin-3a was used to treat cells. Multiple ratios were tested, and the plot shows that the LC
50
value (dashed line) was additive but
not synergistic (shaded area) or adverse (area above the dashed line). The error bars represent standard deviation from two independent experiments. Band C,
immunostaining of compound-treated Weri1 cells. The top panels are merged images of differential interference contrast (DIC) and 4,6-diamidino-2-phenylindole
(DAPI) staining, and the lower panels show p53 levels (B) or caspase-3 activation (C). Nutlin-3a and IR were used as positive controls, and DMSO was used as a negative
control. D–F, quantification of the percentage of immunopositive cells shown in Band Cand bromodeoxyuridine (BrdU). G, real time PCR quantification of p53 target
genes p21,MDM2, and E2F1 activated by drug or IR treatments. MDMX levels were also tested. H, immunoblot and quantification for p21 protein levels activated by
drug or IR treatment. Gy, gray. Iand J, histogram of the proportion of p53 or activated caspase-3 immunopositive cells following treatment with 5 gray IR, nutlin-3a or
SJ-172550. In parallel samples, p53 was knocked down using a p53 shRNA, and MDMX was knocked down using an MDMX shRNA (22), or the samples were treated
with a control (scrambled) shRNA. Each bar represents the mean S.D. scoring in triplicate of 100 cells for each condition and each shRNA. K, representative images
of Weri1 retinoblastoma cells following treatment with SJ-172550 when they lacked p53 (p53 shRNA) or MDMX (MDMX shRNA) as compared with a control shRNA.
Arrows indicate activated caspase-3 immunopositive cells in the red (Cy3) channel. shRNA, small hairpin RNA. Scale bars, 10
M.
First MDMX Inhibitor
APRIL 2, 2010VOLUME 285• NUMBER 14 JOURNAL OF BIOLOGICAL CHEMISTRY 10793
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
increase binding of nutlin-3a without affecting peptide binding.
These data provide additional validation for our proposed mech-
anism of SJ-172550 inhibition of the MDMX-p53 interaction.
DISCUSSION
Although several small molecule inhibitors of MDM2 have
been identified (6, 14), this is the first report to identify a small
molecule inhibitor of MDMX. There is growing evidence that
MDM2 and MDMX inhibit p53 through distinct mechanisms
and that simultaneous inhibition of these two proteins in tumor
cells that express wild-type p53 may be more effective at killing
the cells than the inhibition of MDM2 alone. Our results com-
plement important previous studies on high affinity peptide
inhibitors of MDMX and MDM2 (9, 34). MDMX inhibitors
FIGURE 5. Aand B, space-filling model of the overlaid MDM2-nutlin-3a (teal) with MDMX-p53 (pink) showing SJ-172550 bound to the p53-binding pocket
of MDM2/MDMX in the secondary docking pose. C, position of p53 peptide based on the MDM2/peptide crystal structure. D, primary docking pose of
SJ-172550 to MDMX superimposed on the crystal structure of p53 peptide bound to MDM2. Cand D, the gray is the solvent-excluded surface of MDM2
(Protein Data Bank code 2Z5T), and the blue is the surface for MDMX. Green residues represent single mutations, and residues shown in red formed a
quadruple mutant. Mutations designed to displace SJ-172550 based on the models of the most energetically favorable docking poses (Fig. 5, Band D)
are Q58D, M61F, Y66I, and Q71D. A second series of residues were changed to make the MDMX-binding pocket more like MDM2 to determine whether
SJ-172550 was binding in the p53-binding pocket. These mutations include M53L, H54F, and a quadruple mutant (QUAD) with P95H/S96R/P97K/R103Y.
From top to bottom, the green residues are Met-53, His-54, Gln-58, Met-61, Tyr-66, and Gln-71 with the red residues being Arg-103, Pro-97, Ser-96, and
Pro-95. E, plot of direct binding of each MDMX protein to Texas Red-labeled p53 peptide exactly as described for the HTS. Each point is the mean S.D.
of triplicate assays. The EC
50
values for each protein are indicated. F, competition experiments with each MDMX mutant and increasing concentrations
of SJ-172550. Each data point is the mean S.D. of triplicate assays. The proteins that did not show direct binding of p53 peptide could not be analyzed
in this competition experiment. G, summary of the EC
50
values for direct binding of p53 peptide and competition by nutlin-3a and SJ-172550. The shaded
column (M53L) is one of the mutants that was predicted to make MDMX more like MDM2 and increase nutlin-3a binding while inhibiting SJ-172550
binding without affecting peptide binding. n.d., not determined; WT, wild type.
First MDMX Inhibitor
10794 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 14• APRIL 2, 2010
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
alone (peptide or small molecule) may be useful for treating
tumors such as retinoblastoma that show increased MDMX
expression (22). In addition, they may be effective when com-
bined with MDM2 inhibitors to induce a robust p53 response in
cancer cells that express wild-type p53 (36).
We have developed and optimized a biochemical assay for
HTS of MDMX inhibitors. We screened a diverse chemical
library and identified compound SJ-172550 as the first small
molecule inhibitor of MDMX with a low micromolar binding
constant. SJ-172550 reduced p53 binding in vitro and had little
or no redox activity. It did not covalently modify MDMX but
rather thermostabilized the protein and reversibly bound it,
which was consistent with our modeling of SJ-172550 binding
to the p53-binding pocket of MDMX. When retinoblastoma
cells expressing wild-type p53 and high levels of MDMX were
exposed to SJ-172550 in vitro, they showed evidence of p53-
mediated cytotoxicity. More importantly, the death was p53-
dependent because an shRNA to p53 prevented SJ-172550-me-
diated cell death (Fig. 4, Jand K). Although these data point to a
p53-dependent cell death mechanism, they do not rule out the
possibility of off-target binding of SJ-172550. Indeed, many lead
compounds and drugs show off-target effects.
In combination with the MDM2 inhibitor nutlin-3a,
SJ-172550 showed additive cytotoxicity in cells that expressed
wild-type p53. Thus, we propose that SJ-172550 binds the p53-
binding pocket of MDMX, thereby freeing p53 to induce apo-
ptosis. This compound represents a bona fide lead molecule
and, together with the other promising lead scaffolds from this
study, can now be used for further medicinal chemical analyses,
including optimization of affinity, specificity, and cell perme-
ability and assessment of pharmacokinetics and toxicity.
It has been shown that MDM2 and MDMX can form a het-
erodimer through their Ring domains, and this may regulate
MDM2-mediated degradation of p53 (37–39). We found that
the overall p53 protein levels are not dramatically altered fol-
lowing exposure of cultured cells to SJ-172550. The identifica-
tion of the first small molecule inhibitor will allow researchers
to probe this mechanism further by comparing the effect of
MDMX protein loss to inhibition of MDMX-p53 binding on
p53 stability.
One of our key findings from the analysis of biochemical
and cell-based assay data was that the compounds that had
the best binding constants for MDMX were not necessarily
the ones that were most suitable for follow-up. For example,
SJ-134433 had a very good binding constant for MDMX, but
further analysis showed it was unstable in FP buffer, did not
thermostabilize MDMX, had significant redox activity, and
covalently modified the protein. These data clearly empha-
size the importance of performing comprehensive charac-
terization of active compounds to rule out nonspecific
mechanisms of action that make compounds unsuitable for
further development. They also emphasize the need to select
candidates for further work based on a balance of chemical
and biological properties, rather than purely on potency or
biochemical mechanism of action.
We have not yet fully validated the mechanism of action of
SJ-172550, but it seems probable based upon our modeling and
data that it binds the p53-binding pocket of MDMX and frees
p53 to activate its target genes leading to cell cycle exit and
apoptosis. Consistent with this model, we observed moderate
p53 pathway activation in MDMX-amplified retinoblastoma
cells and partial disruption of the MDMX-p53 interaction in
cell lysates from retinoblastoma cells and other cell lines. We do
not believe that the mechanism of action is through p53 protein
stabilization based on immunoblotting and single cell immuno-
staining. A structural alignment of the binding pockets of
MDMX and MDM2 was produced using the backbone atoms
and provides some clues about where SJ-172550 may bind
MDMX to induce p53 pathway activation. Mutations gener-
ated in the p53-binding pocket of MDMX provided additional
support for this proposed mechanism. Additional x-ray crystal-
lography and other structural studies are required to defini-
tively show that SJ-172550 binds to the p53-binding pocket of
MDMX. Nonetheless, this is the first small molecule MDMX
inhibitor that has been identified with a low micromolar bind-
ing constant. It is important to note that SJ-172550 also binds
MDM2, although less effectively. Our compound is clearly less
effective against MDM2 than nutlin-3a (6) or other MDM2
inhibitors (14), and it is difficult to predict whether further
refinement of MDMX binding will similarly improve MDM2
binding or lead to a high affinity MDMX-selective inhibitor.
Acknowledgments—We thank Taosheng Chen, Fu-Yue Zeng, Wenwei
Lin, Jimmy Cui, and David Bouck for assistance with HTS at the St.
Jude High Throughput Screening Center. We thank Aaron Kosinski
and David Miller for assistance with protein preparation. We thank
Erin H. Seeley, Jamie L. Allen, and Richard M. Caprioli (all of the
Mass Spectrometry Research Center, Vanderbilt Medical Center,
Nashville, TN) for assistance with high resolution MALDI mass spec-
trometry analyses. We also thank Brett Waddell for assistance with
Biacore analysis and Angie McArthur for editing the manuscript.
REFERENCES
1. Hanahan, D., and Weinberg, R. A. (2000) Cell 100, 57–70
2. Dang, J., Kuo, M. L., Eischen, C. M., Stepanova, L., Sherr, C. J., and Roussel,
M. F. (2002) Cancer Res. 62, 1222–1230
3. Vogelstein, B., Lane, D., and Levine, A. J. (2000) Nature 408, 307–310
4. Vousden, K. H., and Lu, X. (2002) Nat. Rev. Cancer 2, 594 604
5. Levine, E. M., Passini, M., Hitchcock, P. F., Glasgow, E., and Schechter, N.
(1997) J. Comp. Neurol. 387, 439–448
6. Vassilev, L. T., Vu, B. T., Graves, B., Carvajal, D., Podlaski, F., Filipovic, Z.,
Kong, N., Kammlott, U., Lukacs, C., Klein, C., Fotouhi, N., and Liu, E. A.
(2004) Science 303, 844–848
7. Marine, J. E. (2006) Heart Rhythm 3, 342–344
8. Anderson, J. J., Challen, C., Atkins, H., Suaeyun, R., Crosier, S., and Lunec,
J. (2007) Int. J. Oncol. 31, 545–555
9. Hu, B., Gilkes, D. M., and Chen, J. (2007) Cancer Res. 67, 8810 8817
10. Udayakumar, T. S., Hachem, P., Ahmed, M. M., Agrawal, S., and Pollack,
A. (2008) Mol. Cancer Res. 6, 1742–1754
11. Tovar, C., Rosinski, J., Filipovic, Z., Higgins, B., Kolinsky, K., Hilton, H.,
Zhao, X., Vu, B. T., Qing, W., Packman, K., Myklebost, O., Heimbrook,
D. C., and Vassilev, L. T. (2006) Proc. Natl. Acad. Sci. U.S.A. 103,
1888–1893
12. Coll-Mulet, L., Iglesias-Serret, D., Santidria´n, A. F., Cosialls, A. M., de
Frias, M., Castan˜o, E., Campa`s, C., Barraga´n, M., de Sevilla, A. F., Do-
mingo, A., Vassilev, L. T., Pons, G., and Gil, J. (2006) Blood 107,
4109–4114
13. Kojima, K., Konopleva, M., Samudio, I. J., Shikami, M., Cabreira-Hansen,
M., McQueen, T., Ruvolo, V., Tsao, T., Zeng, Z., Vassilev, L. T., and An-
dreeff, M. (2005) Blood 106, 3150–3159
First MDMX Inhibitor
APRIL 2, 2010VOLUME 285• NUMBER 14 JOURNAL OF BIOLOGICAL CHEMISTRY 10795
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
14. Shangary, S., Qin, D., McEachern, D., Liu, M., Miller, R. S., Qiu, S., Ni-
kolovska-Coleska, Z., Ding, K., Wang, G., Chen, J., Bernard, D., Zhang, J.,
Lu, Y., Gu, Q., Shah, R. B., Pienta, K. J., Ling, X., Kang, S., Guo, M., Sun, Y.,
Yang, D., and Wang, S. (2008) Proc. Natl. Acad. Sci. U.S.A. 105, 3933–3938
15. Bartel, F., Schulz, J., Bo¨hnke, A., Blu¨ mke, K., Kappler, M., Bache, M.,
Schmidt, H., Wu¨rl, P., Taubert, H., and Hauptmann, S. (2005) Int. J. Can-
cer 117, 469–475
16. Shvarts, A., Bazuine, M., Dekker, P., Ramos, Y. F., Steegenga, W. T., Mer-
ckx, G., van Ham, R. C., van der Houven van Oordt, W., van der Eb, A. J.,
and Jochemsen, A. G. (1997) Genomics 43, 34–42
17. Danovi, D., Meulmeester, E., Pasini, D., Migliorini, D., Capra, M., Frenk,
R., de Graaf, P., Francoz, S., Gasparini, P., Gobbi, A., Helin, K., Pelicci,
P. G., Jochemsen, A. G., and Marine, J. C. (2004) Mol. Cell. Biol. 24,
5835–5843
18. Toledo, F., Krummel, K. A., Lee, C. J., Liu, C. W., Rodewald, L. W., Tang,
M., and Wahl, G. M. (2006) Cancer Cell 9, 273–285
19. Marine, J. C., Dyer, M. A., and Jochemsen, A. G. (2007) J. Cell Sci. 120,
371–378
20. Marine, J. C., and Jochemsen, A. G. (2004) Cell Cycle 3, 900 –904
21. Migliorini, D., Lazzerini Denchi, E., Danovi, D., Jochemsen, A., Capillo,
M., Gobbi, A., Helin, K., Pelicci, P. G., and Marine, J. C. (2002) Mol. Cell.
Biol. 22, 5527–5538
22. Laurie, N. A., Donovan, S. L., Shih, C. S., Zhang, J., Mills, N., Fuller, C.,
Teunisse, A., Lam, S., Ramos, Y., Mohan, A., Johnson, D., Wilson, M.,
Rodriguez-Galindo, C., Quarto, M., Francoz, S., Mendrysa, S. M., Guy,
R. K., Marine, J. C., Jochemsen, A. G., and Dyer, M. A. (2006) Nature 444,
61–66
23. Hu, B., Gilkes, D. M., Farooqi, B., Sebti, S. M., and Chen, J. (2006) J. Biol.
Chem. 281, 33030–33035
24. Gu, J., Kawai, H., Nie, L., Kitao, H., Wiederschain, D., Jochemsen, A. G.,
Parant, J., Lozano, G., and Yuan, Z. M. (2002) J. Biol. Chem. 277,
19251–19254
25. Bo¨ttger, V., Bo¨ ttger, A., Garcia-Echeverria, C., Ramos, Y. F., van der Eb,
A. J., Jochemsen, A. G., and Lane, D. P. (1999) Oncogene 18, 189–199
26. Popowicz, G. M., Czarna, A., Rothweiler, U., Szwagierczak, A., Krajewski,
M., Weber, L., and Holak, T. A. (2007) Cell Cycle 6, 2386–2392
27. Popowicz, G. M., Czarna, A., and Holak, T. A. (2008) Cell Cycle 7,
2441–2443
28. Stricher, F., Martin, L., Barthe, P., Pogenberg, V., Mechulam, A., Menez,
A., Roumestand, C., Veas, F., Royer, C., and Vita, C. (2005) Biochem. J. 390,
29–39
29. Lu, F., Chi, S. W., Kim, D. H., Han, K. H., Kuntz, I. D., and Guy, R. K. (2006)
J. Comb. Chem. 8, 315–325
30. Senisterra, G. A., and Finerty, P. J., Jr. (2009) Mol. Biosyst. 5, 217–223
31. Senisterra, G. A., Soo Hong, B., Park, H. W., and Vedadi, M. (2008) J. Bi-
omol. Screen. 13, 337–342
32. Lor, L. A., Schneck, J., McNulty, D. E., Diaz, E., Brandt, M., Thrall, S. H.,
and Schwartz, B. (2007) J. Biomol. Screen. 12, 881–890
33. Kallen, J., Goepfert, A., Blechschmidt, A., Izaac, A., Geiser, M., Tavares, G.,
Ramage, P., Furet, P., Masuya, K., and Lisztwan, J. (2009) J. Biol. Chem.
284, 8812–8821
34. Pazgier, M., Liu, M., Zou, G., Yuan, W., Li, C., Li, C., Li, J., Monbo, J., Zella,
D., Tarasov, S. G., and Lu, W. (2009) Proc. Natl. Acad. Sci. U.S.A. 106,
4665–4670
35. Morris, G. M., Huey, R., Lindstrom, W., Sanner, M. F., Belew, R. K., Good-
sell, D. S., and Olson, A. J. (2009) J. Comput. Chem. 30, 2785–2791
36. Laurie, N. A., Shih, C. S., Schin-Shih, C., and Dyer, M. A. (2007) Curr.
Cancer Drug Targets 7, 689–695
37. Stad, R., Little, N. A., Xirodimas, D. P., Frenk, R., van der Eb, A. J., Lane,
D. P., Saville, M. K., and Jochemsen, A. G. (2001) EMBO Rep. 2,
1029–1034
38. Linke, K., Mace, P. D., Smith, C. A., Vaux, D. L., Silke, J., and Day, C. L.
(2008) Cell Death Differ. 15, 841–848
39. Jackson, M. W., and Berberich, S. J. (2000) Mol. Cell. Biol. 20, 1001–1007
First MDMX Inhibitor
10796 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 14• APRIL 2, 2010
at St. Jude Children's Research Hospital on July 25, 2013http://www.jbc.org/Downloaded from
... Small molecule antagonists targeting the p53 binding site of MDMX, such as SJ-172550 and CTX1, typically lead to accumulation of the p53 protein and expression of p53-regulated genes resulting in similar effects to MDM2 antagonists: growth arrest and apoptosis ( Figure 2B) [148]. Though reducing toxicity associated with MDM2 inhibition was the primary driving force behind the development of MDMX antagonists, these compounds display potent synergism when combined [149,150]. One of the most exciting MDMX inhibitors is an antisense oligonucleotide (ASO) directed against full-length MDMX. ...
... SAHA and ganetespib facilitate degradation of mup53 protein that abrogates tumorigenic potential in animal models of cancer but failed to perform in the clinic [164][165][166][167]. When combined with MDM2 and MDMX inhibitors, the efficacy of treatment was enhanced [149,150], indicating that combining modalities may offer some treatment benefits as long as there is no overlapping toxicity (Table 1). Therefore, it may be possible to improve treatment responses by combining these Hsp90 inhibitors with therapeutics such as Ad-p53 [113][114][115], which would activate WT p53 signaling while simultaneously blocking the effects of mup53-preventing a dominant negative effect from occurring [122]-while also thwarting GOF mup53 signaling pathways. ...
Article
Full-text available
While chemotherapy is a key treatment strategy for many solid tumors, it is rarely curative, and most tumor cells eventually become resistant. Because of this, there is an unmet need to develop systemic treatments that capitalize on the unique mutational landscape of each patient’s tumor. The most frequently mutated protein in cancer, p53, has a role in nearly all cancer subtypes and tumorigenesis stages and therefore is one of the most promising molecular targets for cancer treatment. Unfortunately, drugs targeting p53 have seen little clinical success despite promising preclinical data. Most of these drug compounds target specific aspects of p53 inactivation, such as through inhibiting negative regulation by the mouse double minute (MDM) family of proteins. These treatment strategies fail to address cancer cells’ adaptation mechanisms and ignore the impact that p53 loss has on the entire p53 network. However, recent gene therapy successes show that targeting the p53 network and cellular dysfunction caused by p53 inactivation is now possible and may soon translate into successful clinical responses. In this review, we discuss p53 signaling complexities in cancer that have hindered the development and use of p53-targeted drugs. We also describe several current therapeutics reporting promising preclinical and clinical results.
... For example, MDM2 antagonists, such as Nutlins could inhibit MDM2-P53 interactions, resulting in the apoptosis of malignant cells (Demicco, 2019). Additionally, efforts have been made in searching for MDM4 antagonists, for example, compounds such as SJ-172550 and XI-006 could inhibit MDM4 effectively (Reed et al., 2010;Pishas et al., 2015). Moreover, several dual MDM2-MDM4 ...
Article
Full-text available
The term “undruggable” is to describe molecules that are not targetable or at least hard to target pharmacologically. Unfortunately, some targets with potent oncogenic activity fall into this category, and currently little is known about how to solve this problem, which largely hampered drug research on human cancers. Ras, as one of the most common oncogenes, was previously considered “undruggable”, but in recent years, a few small molecules like Sotorasib (AMG-510) have emerged and proved their targeted anti-cancer effects. Further, myc, as one of the most studied oncogenes, and tp53, being the most common tumor suppressor genes, are both considered “undruggable”. Many attempts have been made to target these “undruggable” targets, but little progress has been made yet. This article summarizes the current progress of direct and indirect targeting approaches for ras, myc, two oncogenes, and tp53, a tumor suppressor gene. These are potential therapeutic targets but are considered “undruggable”. We conclude with some emerging research approaches like proteolysis targeting chimeras (PROTACs), cancer vaccines, and artificial intelligence (AI)-based drug discovery, which might provide new cues for cancer intervention. Therefore, this review sets out to clarify the current status of targeted anti-cancer drug research, and the insights gained from this review may be of assistance to learn from experience and find new ideas in developing new chemicals that directly target such “undruggable” molecules.
... In this study, we sought to determine whether pharmacological inhibition of MDM4 could alleviate lung fibrosis. We first tested the effects of two small molecular inhibitors of MDM4, SJ-172550 [18,19] and NSC149109 (XI-011) [20], on MDM4 expression and p53 activity. Immunoblotting analysis showed that the MDM4 protein level was dose-dependently reduced after treatment with XI-011 for 6 -12 h in A549 cells ( Figure S2A). ...
Article
Full-text available
Idiopathic pulmonary fibrosis (IPF) is a progressive and fatal lung disease of unknown etiology with no cure. A better understanding of the disease processes and identification of druggable targets will benefit the development of effective therapies for IPF. We previously reported that MDM4 promoted lung fibrosis through the MDM4-p53-dependent pathway. However, it remained unclear whether targeting this pathway would have any therapeutic potential. In this study, we evaluated the efficacy of XI-011, a small molecular inhibitor of MDM4, for treating lung fibrosis. We found that XI-011 significantly reduced MDM4 expression and increased the expression of total and acetylated p53 in primary human myofibroblasts and a murine fibrotic model. XI-011 treatment resulted in the resolution of lung fibrosis in mice with no notable impact on normal fibroblast death or the morphology of healthy lungs. Based on these findings, we propose that XI-011 might be a promising therapeutic drug candidate for treating pulmonary fibrosis.
... Yet another strategy of overcoming MDMX inactivation is by using small molecule MDMX inhibitors such as SJ-172550, XI-011, and XI-006. These inhibitors yield additive effects in combination with nutlin-3a and induce apoptosis in the ER+ MCF-7 cell line with wild-type p53 [49,50]. ...
Article
Full-text available
The transcription factor p53 is an important regulator of a multitude of cellular processes. In the presence of genotoxic stress, p53 is activated to facilitate DNA repair, cell cycle arrest, and apoptosis. In breast cancer, the tumor suppressive activities of p53 are frequently inactivated by either the overexpression of its negative regulator MDM2, or mutation which is present in 30-35% of all breast cancer cases. Notably, the frequency of p53 mutation is highly subtype dependent in breast cancers, with majority of hormone receptor-positive or luminal subtypes retaining the wild-type p53 status while hormone receptor-negative patients predominantly carry p53 mutations with gain-of-function oncogenic activities that contribute to poorer prognosis. Thus, a two-pronged strategy of targeting wild-type and mutant p53 in different subtypes of breast cancer can have clinical relevance. The development of p53-based therapies has rapidly progressed in recent years, and include unique small molecule chemical inhibitors, stapled peptides, PROTACs, as well as several genetic-based approaches using vectors and engineered antibodies. In this review, we highlight the therapeutic strategies that are in pre-clinical and clinical development to overcome p53 inactivation in both wild-type and mutant p53-bearing breast tumors, and discuss their efficacies and limitations in pre-clinical and clinical settings.
... This class of compounds is ineffective against MDMX [144], and similar kind of results are obtained for other compounds such as AMG-232 (Figure 7), which binds to MDMX with a much smaller affinity as compared to MDM2. Small molecule MDMX inhibitors have also failed to work in cultured cells [145]. These molecules are designed to mimic the three hydrophobic residues found in the p53 binding epitope to reduce toxicity. ...
Article
Full-text available
Background: Protein-protein interactions (PPIs) are appealing targets for designing novel small-molecule inhibitors. The role of PPIs in various infectious and neurodegenerative disorders makes them potential targets with a broad therapeutic spectrum, though they were portrayed as un-druggable targets due to their flat surfaces, disordered conformations, and absence of grooves. However, recent progresses in computational biology have led researchers to reconsider PPIs as an important area in drug discovery. Areas covered: In this review, we introduce in-silico methods used to identify PPI interfaces and present an in-depth overview of various computational methodologies that are successfully applied to annotate the PPI interfaces. We also discuss several successful case studies that use computational tools to understand PPIs modulation and their key roles in various physiological processes. Expert opinion: Computational methods face challenges due to the inherent flexibility of proteins, which makes them expensive, and result in the use of rigid models. This problem becomes more significant in PPIs due to their flexible and flat interfaces. Computational methods like molecular dynamics (MD) simulation and machine learning can integrate the chemical structure data into biochemical and can be used for target identification and modulation. These computational methodologies have been crucial in understanding the structure of PPIs, designing PPI modulators, discovering new drug targets, and predicting treatment outcomes.
... Firstly, we conducted inhibition assay of the small-molecule PPI inhibitors against Mdmx-p53 and Mdm2-p53 interaction ( Fig. 2A,B). SJ-172550 46 was used for positive control as a well-studied Mdmx-p53 inhibitor. Compared to the control group, K-181 and its free thiol derivative K-181SH showed significant inhibitory effects on Mdmx-p53 interaction, as previously reported. ...
Article
Full-text available
Mdmx and Mdm2 are two major suppressor factors for the tumor suppressor gene p53. In central nervous system, Mdmx suppresses the transcriptional activity of p53 and enhances the binding of Mdm2 to p53 for degradation. But Mdmx dynamics in cerebral infarction remained obscure. Here we investigated the role of Mdmx under ischemic conditions and evaluated the effects of our developed small-molecule Protein–Protein Interaction (PPI) inhibitors, K-181, on Mdmx–p53 interactions in vivo and in vitro. We found ischemic stroke decreased Mdmx expression with increased phosphorylation of Mdmx Serine 367, while Mdmx overexpression by AAV-Mdmx showed a neuroprotective effect on neurons. The PPI inhibitor, K-181 attenuated the neurological deficits by increasing Mdmx expression in post-stroke mice brain. Additionally, K-181 selectively inhibited HDAC6 activity and enhanced tubulin acetylation. Our findings clarified the dynamics of Mdmx in cerebral ischemia and provide a clue for the future pharmaceutic development of ischemic stroke.
... However, p53-MDM2 inhibitors are inadequate when MDMX is also expressed at a high level in some cancer cells. A few p53-MDMX inhibitors have also been reported such as SJ-172550 [22], NSC207895 [23], and NSC146109 [24]. ...
Article
Full-text available
A series of novel indolone derivatives were synthesized and evaluated for their binding affinities toward MDM2 and MDMX. Some compounds showed potent MDM2 and moderate MDMX activities. Among them, compound A13 exhibited the most potent affinity toward MDM2 and MDMX, with a Ki of 0.031 and 7.24 μM, respectively. A13 was also the most potent agent against HCT116, MCF7, and A549, with IC50 values of 6.17, 11.21, and 12.49 μM, respectively. Western blot analysis confirmed that A13 upregulated the expression of MDM2, MDMX, and p53 by Western blot analysis. These results indicate that A13 is a potent dual p53-MDM2 and p53-MDMX inhibitor and deserves further investigation.
Article
The function of the p53 protein is impaired by the overexpression of its negative regulator murine double minute 2 protein (MDM2) and homologous protein MDMX. Disruption of the p53-MDM2/MDMX interaction to restore the transcriptional function of p53 is considered a promising strategy for cancer therapy. To design dual MDM2/MDMX inhibitors, the binding modes of MDM2 or MDMX with their inhibitors are elucidated. Several hot-spot residues of MDM2 or MDMX are identified by molecular dynamics simulations, alanine scanning and MM-GBSA calculations. Then, focusing on the interaction with hot-spot residues, two series of derivatives bearing 1,3-diketone and α-aminoketone scaffolds are designed and synthesized. Among these compounds, C16 is identified as the most potent compound with low micromolar binding affinities with MDM2 and MDMX. C16 also displays moderate antiproliferative activities against MDM2-overexpressing and MDMX-overexpressing cells, with IC50 values of 0.68 μM in HCT116 cells and 0.54 μM in SH-SY5Y cells. Furthermore, C16 inhibits cell migration and invasion, reactivates the function of p53, arrests the cell cycle and induces cellular apoptosis in HCT116 and SH-SY5Y cells. Collectively, C16 can be developed as a dual MDM2 and MDMX inhibitor for cancer therapy.
Article
Full-text available
1,3‐Diarylpropenes firstly serve as efficient alkenylated reagents for the functionalization of pyrazolinones. In the presence of 2,3‐Dichloro‐5,6‐dicyano‐1,4‐benzoquinone (DDQ), it undergoes oxidative coupling and final dehydrogenation to give the highly substituted E,E‐α,β,γ,δ‐dienpyrazolinones in moderate to excellent yields.
Article
Full-text available
MDMX, an MDM2-related protein, has emerged as yet another essential negative regulator of p53 tumor suppressor, since loss of MDMX expression results in p53-dependent embryonic lethality in mice. However, it remains unknown why neither homologue can compensate for the loss of the other. In addition, results of biochemical studies have suggested that MDMX inhibits MDM2-mediated p53 degradation, thus contradicting its role as defined in gene knockout experiments. Using cells deficient in either MDM2 or MDMX, we demonstrated that these two p53 inhibitors are in fact functionally dependent on each other. In the absence of MDMX, MDM2 is largely ineffective in down-regulating p53 because of its extremely short half-life. MDMX renders MDM2 protein sufficiently stable to function at its full potential for p53 degradation. On the other hand, MDMX, which is a cytoplasmic protein, depends on MDM2 to redistribute into the nucleus and be able to inactivate p53. We also showed that MDMX, when exceedingly overexpressed, inhibits MDM2-mediated p53 degradation by competing with MDM2 for p53 binding. Our findings therefore provide a molecular basis for the nonoverlapping activities of these two p53 inhibitors previously revealed in genetic studies.
Article
Full-text available
The oncoproteins MDM2 and MDMX negatively regulate the activity and stability of the tumor suppressor protein p53--a cellular process initiated by MDM2 and/or MDMX binding to the N-terminal transactivation domain of p53. MDM2 and MDMX in many tumors confer p53 inactivation and tumor survival, and are important molecular targets for anticancer therapy. We screened a duodecimal peptide phage library against site-specifically biotinylated p53-binding domains of human MDM2 and MDMX chemically synthesized via native chemical ligation, and identified several peptide inhibitors of the p53-MDM2/MDMX interactions. The most potent inhibitor (TSFAEYWNLLSP), termed PMI, bound to MDM2 and MDMX at low nanomolar affinities--approximately 2 orders of magnitude stronger than the wild-type p53 peptide of the same length (ETFSDLWKLLPE). We solved the crystal structures of synthetic MDM2 and MDMX, both in complex with PMI, at 1.6 A resolution. Comparative structural analysis identified an extensive, tightened intramolecular H-bonding network in bound PMI that contributed to its conformational stability, thus enhanced binding to the 2 oncogenic proteins. Importantly, the C-terminal residue Pro of PMI induced formation of a hydrophobic cleft in MDMX previously unseen in the structures of p53-bound MDM2 or MDMX. Our findings deciphered the structural basis for high-affinity peptide inhibition of p53 interactions with MDM2 and MDMX, shedding new light on structure-based rational design of different classes of p53 activators for potential therapeutic use.
Article
Full-text available
The significant increase in the demand for purified protein for crystallization and structural studies has made necessary the development of multi-sample methods for identifying solution conditions that affect protein stability and aggregation. Conditions that stabilize proteins can improve protein purification and crystallization. These methods can be used to identify small molecule compounds or inhibitors that interact with the purified proteins, and might serve as starting points for drug discovery. In this article three methods for measuring protein stability and aggregation are described and discussed: differential scanning fluorimetry (DSF), differential static light scattering (DSLS), and isothermal denaturation (ITD).
Article
Full-text available
p53 tumor suppressor activity is negatively regulated through binding to the oncogenic proteins Hdm2 and HdmX. The p53 residues Leu26, Trp23, and Phe19 are crucial to mediate these interactions. Inhibiting p53 binding to both Hdm2 and HdmX should be a promising clinical approach to reactivate p53 in the cancer setting, but previous studies have suggested that the discovery of dual Hdm2/HdmX inhibitors will be difficult. We have determined the crystal structures at 1.3 Å of the N-terminal domain of HdmX bound to two p53 peptidomimetics without and with a 6-chlorine substituent on the indole (which binds in the same subpocket as Trp23 of p53). The latter compound is the most potent peptide-based antagonist of the p53-Hdm2 interaction yet to be described. The x-ray structures revealed surprising conformational changes of the binding cleft of HdmX, including an “open conformation” of Tyr99 and unexpected “cross-talk” between the Trp and Leu pockets. Notably, the 6-chloro p53 peptidomimetic bound with high affinity to both HdmX and Hdm2 (Kd values of 36 and 7 nm, respectively). Our results suggest that the development of potent dual inhibitors for HdmX and Hdm2 should be feasible. They also reveal possible conformational states of HdmX, which should lead to a better prediction of its interactions with potential biological partners.
Article
Full-text available
Mdm2 and MDMX are two structurally related p53-binding proteins which show the highest level of sequence similarity in the N-terminal p53-binding domains. Apart from its ability to inhibit p53 mediated transcription, a feature it shares with mdm2, very little is known about the physiological functions of MDMX. It is clearly distinct from mdm2 since its expression appears not to be regulated by p53 and it cannot compensate for lack of mdm2 in early development. We present data on the structural similarity between the p53 binding pockets of mdm2 and MDMX using p53- and phage-selected peptides. From the results we conclude that our recently devised innovative approach to reverse the mdm2-mediated inhibition of p53's transactivation function in vivo would probably target MDMX as well. Strategies for selectively targeting mdm2 and MDMX are suggested and a possible mechanism for regulating the p53-mdm2/MDMX interactions by protein phosphorylation is discussed.
Article
We describe the testing and release of AutoDock4 and the accompanying graphical user interface AutoDockTools. AutoDock4 incorporates limited flexibility in the receptor. Several tests are reported here, including a redocking experiment with 188 diverse ligand-protein complexes and a cross-docking experiment using flexible sidechains in 87 HIV protease complexes. We also report its utility in analysis of covalently bound ligands, using both a grid-based docking method and a modification of the flexible sidechain technique.
Article
We have previously shown in separate studies that MDM2 knockdown via antisense MDM2 (AS-MDM2) and E2F1 overexpression via adenoviral-mediated E2F1 (Ad-E2F1) sensitized prostate cancer cells to radiation. Because E2F1 and MDM2 affect apoptosis through both common and independent pathways, we hypothesized that coupling these two treatments would result in increased killing of prostate cancer cells. In this study, the effect of Ad-E2F1 and AS-MDM2 in combination with radiation was investigated in three prostate cancer cell lines: LNCaP cells, LNCaP-Res cells [androgen insensitive with functional p53 and androgen receptor (AR)], and PC3 cells (androgen insensitive, p53(null), and AR(null)). A supra-additive radiosensitizing effect was observed in terms of clonogenic inhibition and induction of apoptosis (caspase-3 + caspase-7 activity) in response to Ad-E2F1 plus AS-MDM2 treatments in all three cell lines. In LNCaP and LNCaP-Res, these combination treatments elevated the levels of phospho-Ser(15) p53 with significant induction of p21(waf1/cip1), phospho-gammaH2AX, PUMA, and Bax levels and reduction of AR and bcl-2 expression. Similarly, AR(null) and p53(null) PC-3 cells showed elevated levels of Bax and phospho-gammaH2AX expression. These findings show that the combination of Ad-E2F1 and AS-MDM2 significantly increases cell death in prostate cancer cells exposed to radiation and that this effect occurs in the presence or absence of AR and p53.
Article
The Mdmx oncoprotein has only recently emerged as a critical-independent to Mdm2-regulator of p53 activation. We have determined the crystal structure of the N-terminal domain of human Mdmx bound to a 15-residue transactivation domain peptide of human p53. The structure shows why antagonists of the Mdm2 binding to p53 are ineffective in the Mdmx-p53 interaction.
Article
We recently reported the identification of a mouse cDNA encoding a new p53-associating protein that we called Mdmx because of its structural similarity to Mdm2, a well-known p53-binding protein. Here we report the isolation of a cDNA encoding the human homolog of Mdmx. The ORF of the cDNA encodes a protein of 490 amino acids, 90% similar to mouse Mdmx. The homology between Mdmx and Mdm2 is most prominent in the p53-binding domain and the putative metal-binding domains. The Mdmx protein, which, based on SDS-PAGE, has a MW of 80 kDa, can bind p53 in vitro. The human MDMX gene is transcribed in all tissues tested, with high levels in thymus. By fluorescence in situ hybridization analysis we mapped the mouse mdmx gene to chromosome 1 (region F-G) and the human MDMX gene to chromosome 1q32.
Article
The genetic linkages of the murine ocular retardation mutation with the Chx10 gene and the murine small eye mutation with the Pax-6 gene has demonstrated the importance of Paired class homeobox genes in the development of the mammalian retina. Previously, we identified a Paired-class homeobox gene, Vsx-1, whose expression in the adult goldfish retina is restricted to the inner nuclear layer (INL) and to postmitotic, differentiating progenitor cells in the growth zone at the retinal peripheral margin, where neurogenesis continues throughout life. Here, we report the molecular cloning and expression pattern of a new Paired class homeobox gene, Vsx-2, in the adult goldfish retina. Like Vsx-1, Vsx-2 expression is highly restricted to the retina in the adult goldfish and overlaps with Vsx-1 expression in the mature INL. At the peripheral margin, Vsx-2 is expressed in mitotically active neuronal progenitors and is downregulated as these cells become postmitotic and begin to differentiate. Comparison of the amino acid sequences of Vsx-2, Vsx-1, Chx10, and C. elegans ceh-10 reveal a conserved homeodomain and a unique domain termed the CVC domain. The similarities of the Vsx-2, Vsx-1, and Chx10 expression patterns suggest that genes containing the CVC domain have conserved functions during retinal development in vertebrates.