Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Determinants of the Differential Antizyme-Binding Affinity of Ornithine Decarboxylase

  • Yen-Chin Liu,

    Affiliation Department of Life Sciences and Institute of Genomics and Bioinformatics, National Chung-Hsing University, Taichung, Taiwan

  • Den-Hua Hsu,

    Affiliation Department of Life Sciences and Institute of Genomics and Bioinformatics, National Chung-Hsing University, Taichung, Taiwan

  • Chi-Liang Huang,

    Affiliation Department of Life Sciences and Institute of Genomics and Bioinformatics, National Chung-Hsing University, Taichung, Taiwan

  • Yi-Liang Liu,

    Affiliations Institute of Microbiology and Immunology, Chung Shan Medical University, Taichung, Taiwan, Division of Allergy, Immunology, and Rheumatology, Chung Shan Medical University Hospital, Taichung, Taiwan

  • Guang-Yaw Liu ,

    hchung@dragon.nchu.edu.tw (HCH); liugy@csmu.edu.tw (GYL)

    Affiliations Institute of Microbiology and Immunology, Chung Shan Medical University, Taichung, Taiwan, Division of Allergy, Immunology, and Rheumatology, Chung Shan Medical University Hospital, Taichung, Taiwan

  • Hui-Chih Hung

    hchung@dragon.nchu.edu.tw (HCH); liugy@csmu.edu.tw (GYL)

    Affiliations Department of Life Sciences and Institute of Genomics and Bioinformatics, National Chung-Hsing University, Taichung, Taiwan, Agricultural Biotechnology Center, National Chung Hsing University, Taichung, Taiwan

Abstract

Ornithine decarboxylase (ODC) is a ubiquitous enzyme that is conserved in all species from bacteria to humans. Mammalian ODC is degraded by the proteasome in a ubiquitin-independent manner by direct binding to the antizyme (AZ). In contrast, Trypanosoma brucei ODC has a low binding affinity toward AZ. In this study, we identified key amino acid residues that govern the differential AZ binding affinity of human and Trypanosoma brucei ODC. Multiple sequence alignments of the ODC putative AZ-binding site highlights several key amino acid residues that are different between the human and Trypanosoma brucei ODC protein sequences, including residue 119, 124,125, 129, 136, 137 and 140 (the numbers is for human ODC). We generated a septuple human ODC mutant protein where these seven bases were mutated to match the Trypanosoma brucei ODC protein sequence. The septuple mutant protein was much less sensitive to AZ inhibition compared to the WT protein, suggesting that these amino acid residues play a role in human ODC-AZ binding. Additional experiments with sextuple mutants suggest that residue 137 plays a direct role in AZ binding, and residues 119 and 140 play secondary roles in AZ binding. The dissociation constants were also calculated to quantify the affinity of the ODC-AZ binding interaction. The Kd value for the wild type ODC protein-AZ heterodimer ([ODC_WT]-AZ) is approximately 0.22 μM, while the Kd value for the septuple mutant-AZ heterodimer ([ODC_7M]-AZ) is approximately 12.4 μM. The greater than 50-fold increase in [ODC_7M]-AZ binding affinity shows that the ODC-7M enzyme has a much lower binding affinity toward AZ. For the mutant proteins ODC_7M(-Q119H) and ODC_7M(-V137D), the Kd was 1.4 and 1.2 μM, respectively. These affinities are 6-fold higher than the WT_ODC Kd, which suggests that residues 119 and 137 play a role in AZ binding.

Introduction

Ornithine decarboxylase (ODC, EC 4.1.1.17) is a pyridoxal 5′phosphate-dependent enzyme that catalyzes the decarboxylation of ornithine to putrescine [1], [2]. This reaction is the first and rate-limiting reaction in polyamine biosynthesis [3], [4], which is essential for eukaryotic cell growth and differentiation [5][7]. Polyamines and ODC play important roles in many biological functions, including the cell cycle, cellular proliferation, differentiation, apoptosis and embryonic development [8][17]. High levels of polyamines and ODC have also been associated with human disease and a variety of cancers [3], [18][23]. Because the concentration of ODC and polyamine is critical for cell proliferation [11], as well as during the development of neoplastic disease [24][28], ODC is considered to be an oncogenic enzyme. Regulation of ODC and polyamine levels is a current target for therapeutic studies involving numerous types of cancer [13], [29][33].

ODC activity in vivo is highly regulated through several pathways (reviewed in [3], [6], [18], [19]). For instance, the ODC protein has a short half-life and turns over very rapidly [34], [35]. A majority of proteins are degraded through ubiquitination, but ODC is degraded by the proteasome in a ubiquitin-independent manner via direct binding to the antizyme protein (AZ), which is regulated by polyamines [36][39]. AZ binds to ODC and promotes the dissociation of ODC homodimers and then forms the AZ-ODC heterodimer which is ultimately degraded by the 26S proteasome [3], [19], [40][42]. Several studies have shown that 37 residues in the C-terminus of ODC are important for degradation [41], [43], [44], and deletion of this region stabilizes ODC, even in the presence of AZ [45]. Moreover, additional studies have shown that residues 117–140 of ODC may play a role in AZ binding, which induces a conformational change in ODC that exposes the C-terminus and leads to recognition/degradation by the 26S proteasome [35], [36], [46].

The ODC protein circulates as a homodimer, and dimer formation is essential for enzyme activity [3], [47], [48]. The active site is located at the dimer interface, which is formed by the N-terminal domain of one subunit and the C-terminal domain of the second subunit [47]. Disruption of the dimer interface causes a loss of enzyme activity [42]. ODC can bind to AZ to form a heterodimer, and the AZ residues Glu-161, Glu-164 and Glu-165 seem to allow, through electrostatic interactions, the binding between of ODC to AZ to form a heterodimer [49]. The binding of AZ with ODC causes the ODC dimer to dissociate and thus inhibits ODC enzyme activity [3], [42].

AZ expression is regulated by cellular levels of polyamine [38]. The AZ mRNA transcript contains two overlapping open reading frames (ORFs). As the cellular concentration of polyamines increases, it induces a translational frame-shift of the AZ mRNA, which produces a longer functional AZ protein [19], [50][52]. Moreover, the cellular polyamines and polyamine transporter are regulated by AZ. AZ not only inhibits ODC activity to suppress polyamine production but also restrains polyamine uptake and stimulates polyamine excretion, thus controlling polyamine levels 38,5355. Through this mechanism, ODC activity is down-regulated by AZ if polyamines are excessively generated by ODC.

ODC is a ubiquitous enzyme that is conserved in all species from bacteria to humans. The fatal human disease African sleeping sickness is caused by the protozoan Trypanosoma brucei. The disease is currently treated with an irreversible inhibitor, DFMO (DL-α-difluoromethylornithine), which inhibits the Trypanosoma brucei ODC enzyme (tODC) [56][58]. tODC is more stable than hODC, and tODC lacks the C-terminus that appear to be important for human ODC protein degradation [34], [59]. Furthermore, tODC has a low binding affinity toward AZ and thus has a long half-life in the Trypanosoma brucei [34]. Mutation of mouse ODC residues 117–140 to match the tODC sequence disrupts both AZ binding and in vivo regulation, suggesting that this sequence within mouse ODC is important for AZ binding [34]. In this paper, we identified several amino acid residues that influence human ODC (hODC) binding to AZ. Sequence alignments of residues 117–140 of the hODC and tODC proteins show that there are seven non-conserved amino acid residues within this region (Table 1). We therefore mutated these seven amino acid residues in hODC to match the tODC sequence and subsequently examined the binding affinity of the mutant human ODC toward AZ.

thumbnail
Table 1. Amino acid residues at the putative AZ-binding site of human ODC (hODC), Trypanosoma brucei ODC (tODC) and human AZI (hAZI).

https://doi.org/10.1371/journal.pone.0026835.t001

Results and Discussion

Previous structural studies of human ODC suggest that residues 117–140 may be the putative AZ-binding site [34], [60]. Mutation of residues 117–140 in mouse ODC to match the tODC sequence resulted in decreased binding affinity toward AZ, implying that this region is important for AZ binding [34]. In this AZ-binding region of ODC, some amino acid residues are diverse and they may be the factors determining the differential AZ-binding affinity among these organisms. We aligned and compared the protein sequences of hODC and tODC and chose seven candidate residues to further study based on the charge dissimilarities and hydrophobicity (Table 1). The mutations that were made to the human ODC protein are Q119H, A124R, N125D, Q129D, E136V, V137D and M140E.

Analysis of AZ inhibition of the hODC septuple mutant

We initially simultaneously mutated all seven of these residues to generate a septuple mutant hODC protein, which was named ODC_7M (ODC_Q119H/A124R/N125D/Q129D/E136V/V137D/M140E). For ODC_WT, the enzyme activity decreased with increasing concentrations of AZ. At a molar ratio of 2∶1 monomeric AZ to dimeric ODC, the ODC enzyme activity was approximately 30% (Figure 1A, closed circles). In contrast, the septuple ODC mutant was much less sensitive to AZ inhibition. At a molar ratio of 1∶1 AZ:ODC, the septuple ODC enzyme activity was approximately 90% (Figure 1A, open circles), indicating that these seven residues play a role in AZ binding and regulation of ODC enzyme activity.

thumbnail
Figure 1. AZ inhibition of the wild type and ODC mutant proteins.

A: inhibition plots for the ODC_WT and ODC_7M septuple enzymes; B: inhibition plots for the ODC_7M(-A124R) and ODC_7M(-E136V) sextuple mutant enzymes; C: inhibition plots for the ODC_7M(-N125D) and ODC_7M(-Q129D) sextuple mutant enzymes; D: inhibition plots for the ODC_7M(-Q119H), ODC_7M(-V137D) and ODC_7M(-M140E) sextuple mutant enzymes; E: inhibition plots for the ODC_Q119H/V137D, ODC_Q119H/M140E, ODC_V137D/M140E and ODC_Q119H/V137D/M140E mutant enzymes. The enzyme concentration was kept constant at 20 μg/mL, while the AZ concentration ranged from 0 to 28 μg/mL.

https://doi.org/10.1371/journal.pone.0026835.g001

Analysis of AZ inhibition of hODC sextuple mutants

By using ODC_7M as the template, we generated ODC mutant enzymes that had each possible combination of six point mutations. We created 7 sextuple mutants: ODC_7M(-Q119H), ODC_7M(-A124R), ODC_7M(-N125D), ODC_7M(-Q129D), ODC_7M(-E136V), ODC_7M(-V137D) and ODC_7M(-M140E). For ODC_7M(-Q119H), all of the residues were mutated except for Q119H (ODC_A124R/N125D/Q129D/E136V/V137D/M140E). The purpose of generating these sextuple mutants was to identify the essential amino acid residue(s) that govern ODC-AZ binding.

The inhibitory plots for these sextuple mutants are also shown in Figures 1. ODC_7M(-A124 R) and ODC_7M(-E136V) had a pattern of inhibition that was similar to ODC_7M (Figure 1B). These mutants were not significantly inhibited by AZ, which suggests that AZ is no longer binding and that residues 124 and 136 alone do not play a major role in AZ inhibition/binding. The ODC_7M(-N125D) and ODC_7M(-Q129D) enzymes were moderately resistant to AZ inhibition (Figure 1C), indicating that residues 125 and 129 may play a small role in AZ inhibition. The ODC_7M(-Q119H), ODC_7M(-V137D) and ODC_7M(-M140E) mutants had very little resistance to AZ inhibition (Figure 1D). This is especially obvious for ODC_7M(-V137D), which had an inhibition plot that is similar to ODC_WT, implying that residue 137 may be the most important amino acid residue involved in AZ binding and inhibition. The inhibitory plots of ODC_7M(-Q119H) and ODC_7M(-M140E) suggest that residues 119 and 140 may play secondary roles in human ODC-AZ binding.

Analysis of AZ inhibition of hODC double and triple mutants

To further confirm the significance of residues 119, 137 and 140 in AZ binding, double and triple hODC mutants were created. The triple mutant ODC_Q119H/V137D/M140E had an inhibitory plot that is very similar to ODC_7M (Figure 1E, closed squares), again suggesting these three amino acid residues play a role in ODC binding to AZ. However, the ODC_Q119H/V137D double mutant was also resistant to AZ inhibition, similar to the triple mutant (Figure 1E, open circles), which suggests that residues 119 and 137 mediate AZ binding. Compared to the triple mutant ODC_Q119H/V137D/M140E, ODC_V137D/M140E displayed moderate AZ inhibition (Figure 1E, open triangles); absence of the Q119H mutation decreased the AZ inhibitory resistance. ODC_Q119H/M140E displayed minor AZ-inhibition resistance (Figure 1E, closed triangles), which again suggests that residue 137 plays a key role in hODC binding to AZ.

Dissociation constant of the hODC-AZ heterodimer

To quantify the effect of hODC mutations on the AZ binding affinity, dissociation constants (Kd) were determined for the human WT and mutant ODC proteins (Table 2). Sedimentation velocity (SV) experiments with increasing AZ concentrations were performed, and the data were globally fitted to determine the dissociation constant of the ODC-AZ heterodimer (Table 2). Figure 2 shows the distribution plots of the WT and mutant ODC proteins. In the absence of AZ, ODC formed a stable dimer with an S-value of approximately 6; when AZ was present, ODC was dissociated. The ODC dimer peak shifted left and an ODC-AZ complex was formed, which had an S-value of approximately 4.1 (Figure 2A). The Kd value of the [ODC_WT]-AZ heterodimer is approximately 0.22 μM, while the [ODC_7M]-AZ heterodimer Kd is approximately 12.4 μM. The greater than 50-fold increase in the [ODC_WT]-AZ complex Kd indicates that the ODC-7M enzyme really has an extremely low binding affinity toward AZ. The triple mutant ODC_Q119H/V137D/M140E had an AZ-binding affinity that is similar to ODC_7M. The Kd for the [ODC_Q119H/V137D/M140E]-AZ heterodimer was about 9.9 μM, which is approximately 45-fold greater than the WT_ODC one. The Kd values for ODC_7M(-Q119H) and ODC_7M(-V137D) were 1.4 and 1.2 μM, respectively, and are 6-fold higher than the WT_ODC. These data again suggest that residues 119 and 137 play a key role in AZ binding.

thumbnail
Figure 2. The continuous sedimentation coefficient distributions of human ODC mutant enzymes in the presence of AZ.

The concentration of ODC was fixed at 0.3 mg/mL with an AZ concentration of 0.03, 0.06 or 0.09 mg/mL (the molar ratio of AZ/ODC was 0.24, 0.47 and 0.71, respectively). The sedimentation velocity data were globally fitted with the SEDPHAT program to calculate the Kd values of the ODC-AZ heterodimer (Table 2). A: ODC_WT; B: ODC_Q119H/V137D/M140E; C: ODC_7M; D: ODC_7M(-Q119H); E: ODC_7M(-V137D).

https://doi.org/10.1371/journal.pone.0026835.g002

thumbnail
Table 2. Dissociation constants of the human ODC-AZ complex.

https://doi.org/10.1371/journal.pone.0026835.t002

Kinetic properties of human WT and mutant ODC enzymes

The kinetic parameters of the WT and mutant ODC enzymes were determined (Table S1). There were no obvious differences in Km for ornithine substrate and PLP cofactor; additionally, the kcat values for the WT and mutant enzymes were similar. These data suggest that the putative AZ-binding site mutations do not affect ODC enzymatic activity and may not induce a significant change in enzyme conformation.

Charge effect in the putative AZ-binding site for differential AZ-binding affinity

Mutation of residues 117–140 in mouse ODC to mimic the Trypanosoma brucei ODC protein sequence abolished AZ binding [34], which suggests that several or all of these amino acid residues play a role in AZ binding. Our data clearly show that residues 137 and 119 in human ODC play key roles in AZ binding and influence the differential AZ-binding affinity of hODC and tODC. Mutation of Val137 to Asp and Gln119 to His (ODC_Q119H/V137D) may introduce a new charge to the AZ-binding element that may repel AZ and prevent binding. Whether Val137 and Gln119 in hODC directly contact with AZ or simply stabilize the conformation in the AZ-binding site of hODC cannot be clearly elucidated at this time. Crystal structural analysis of hODC-AZ interaction may show the binding of these residues to their counterparts in AZ, however, this complex structure is not available.

Residues 125 and 129 are both aspartic acids, and residue 140 is glutamic acid in tODC; however, in hODC these positions are amino acid residues with neutral side chains (Table 1). Introduction of these negatively charged amino acid residues with the mutant hODC decreases the binding affinity between ODC and AZ. The structural superimposition of hODC and tODC demonstrates a perfect overlapping with a RMSD value of 0.81, and the Cα positioning highlights the conserved AZ-binding element. The inability of tODC to bind AZ and the weak binding affinity of the [ODC_7M]-AZ complex may result from the mutated charged amino acid residues at positions 119, 125, 129, 137 and 140 of the ODC enzyme.

The putative AZ-binding residues of ODC

Sequence alignments at the putative AZ-binding site of hODC, tODC and human antizyme inhibitor (hAZI) is shown in Table 1. The AZI protein structure is homologous to ODC, and AZI has a higher binding affinity toward AZ [42]. Gln119 is highly conserved in all of the ODC and AZI enzymes, except in trypanosomes, suggesting that residue 119 may be crucial for the differential AZ-binding affinity between hODC and tODC but not between ODC and AZI. In contrast, residues 125 and 140 are not strictly conserved; in a majority of the ODC sequences, residue 125 is a neutral Asn or Ser and 140 is Met or Ser; however, in tODC these residues are negatively charged Asp and Glu, respectively, while in AZI, they are a positively charged Lys. Our previous report has suggested that the differences in residues 125 and 140 between human ODC and AZI are responsible for the differential AZ-binding affinities [61]. Here we suggest that electrostatic effects are responsible for the differential AZ-binding affinities among hODC, hAZI and tODC. Furthermore, residue 129 is Asn or Gln and residue 137 is Val or Ile in a majority of the ODC and AZI sequences, while these residues are Asp in tODC. Based on the sequence comparisons and the mutagenesis analyses, we hypothesize that repulsive effect may occur in tODC-AZ binding. The ODC_7M mutant, which has the tODC sequence with Asp125, Asp129, Asp137 and Glu140, has a 50-fold smaller AZ-binding affinity than hODC, suggesting the differences in charge properties of these residues play a key role in AZ-binding affinity of ODC.

Materials and Methods

Site-directed mutagenesis

Site-directed mutagenesis was performed with a QuikChangeTM kit (Stratagene, USA) to generate plasmids with the mutated human ODC (hODC). Purified hODC DNA was used as the template, which was PCR amplified with high-fidelity Pfu DNA polymerase and specific primers with the appropriate codon. The primers with the desired mutations were between 25 to 45 bases in length, which is necessary for specific binding to the template DNA. The sequence was amplified for 16–18 cycles. The PCR products were then treated with DpnI to cleave the wild-type hODC template. The mutant PCR amplicons were cloned into a plasmid vector and transformed into XL-1 E. coli. The mutant plasmid DNA sequence was confirmed by autosequencing.

Recombinant hODC expression and purification

Human ODC or AZ was sub-cloned into the pQE30 vector (Qiagen, Hilden, Germany), which carries a N-terminal His6·Tag sequence for purification. This ampicillin-resistant vector was transformed into the JM109 strain of Escherichia coli. Recombinant protein expression was induced with 1.0 mM isopropyl-1-thio-β-D-galactoside (IPTG), and the cells were harvested at 25°C overnight. The recombinant protein was purified with a Ni-NTA Sepharose column (Sigma). The lysate-Ni-NTA mixture was washed with buffer that contained 10 mM imidazole, 500 mM NaCl and 30 mM Tris-HCl (pH 7.6). Recombinant ODC or AZ was eluted with buffer comprised of 250 mM imidazole, 500 mM NaCl, 2 mM β-mercaptoethanol and 30 mM Tris-HCl, pH 7.6. The purified ODC enzyme was buffer-exchanged and concentrated with 30 mM Tris-HCl (pH 7.6) and 2 mM β-mercaptoethanol with a 30 kDa molecular weight cutoff centrifugal filter device (Amicon Ultra-15, Millipore). The purified AZ protein was buffer-exchanged and concentrated with 250 mM NaCl, 30 mM Tris-HCl (pH 7.6) and 2 mM β-mercaptoethanol with a 10 kDa cutoff centrifugal filter device. The protein purity was verified by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), and the protein concentration was determined with the Bradford method [62].

ODC enzyme assays

The decarboxylation of ornithine by ODC was measured with a CO2-L3K assay kit (DCL, Charlottetown, Canada) at 37°C. A continuous measurement of ODC enzyme activity was coupled with the decarboxylation of ornithine to the carboxylation of phosphoenolpyruvate (PEP) to form oxaloacetate (OAA), which becomes malate following NADH oxidation according to a previously published protocol [42]. The assay reaction mixture for spectrophotometry contained 30 mM Tris-HCl (pH 7.8), 10 mM ornithine, 0.2 mM PLP and 0.4 mL of CO2-L3K assay buffer containing 12.5 mM phosphoenolpyruvate, 0.4 unit/mL microbial phosphoenolpyruvate carboxylase, 4.1 units/mL mammalian malate dehydrogenase and 0.6 mM NADH analog in a final volume of 0.5 mL. ODC enzyme was added to initiate the reaction and the decrease of absorbance at 405 nm was continuously recorded with a Perkin-Elmer Lamba-25 spectrophotometer. In this coupled assay method, 1 mol of CO2 was formed and 1 mol of NADH analog was oxidized under the assay conditions. An absorption coefficient of 2,410 m−1 was used for the NADH analog in the calculations. All of the calculations were performed with the Sigma Plot 10.0 software program (Jandel, San Rafael, CA).

Size-distribution analysis by analytical ultracentrifugation

Sedimentation velocity experiments were performed using a Beckman Optima XL-A analytical ultracentrifuge. Buffer (400 μl) and sample solutions (380 μl) were loaded into the double sector centerpiece individually and built up in a Beckman An-50 Ti rotor. The sedimentation velocity experiments were performed at 20°C with a rotor speed of 42,000 rpm. The protein samples were followed by continually monitoring UV absorbance at 280 nm with a time interval of 420 s and a step size of 0.002 cm. Multiple scans at different time points were fitted to a continuous size distribution model with the SEDFIT software [63], [64]. All of the size distributions were worked out at a confidence level of p = 0.95, a best fit average anhydrous frictional ratio (f/f0), and a resolution N of 250 sedimentation coefficients between 0.1 and 20.0 S.

To determine the dissociation constant (Kd) for human ODC binding toward AZ, sedimentation velocity experiments were carried out at three different concentrations of AZ (0.03, 0.06 or 0.09 mg/mL) with a constant concentration of human ODC (0.3 mg/mL). All of the sedimentation data were globally fitted into the AB hetero-association model using the SEDPHAT program [65], [66] to calculate the Kd value for the ODC-AZ heterodimer. The partial specific volumes of the proteins, the solvent densities and the viscosity were calculated by the SEDNTERP program [67].

Supporting Information

Table S1.

Kinetic parameters of the wild-type and mutant ODC enzymes.

https://doi.org/10.1371/journal.pone.0026835.s001

(DOC)

Author Contributions

Conceived and designed the experiments: HCH GYL. Performed the experiments: YCL DHH CLH YLL. Analyzed the data: YCL DHH HCH YLL. Contributed reagents/materials/analysis tools: HCH GYL. Wrote the paper: HCH.

References

  1. 1. Smithson DC, Lee J, Shelat AA, Phillips MA, Guy RK (2010) Discovery of potent and selective inhibitors of Trypanosoma brucei ornithine decarboxylase. J Biol Chem 285: 16771–16781.
  2. 2. Jackson LK, Brooks HB, Osterman AL, Goldsmith EJ, Phillips MA (2000) Altering the reaction specificity of eukaryotic ornithine decarboxylase. Biochemistry 39: 11247–11257.
  3. 3. Pegg AE (2006) Regulation of ornithine decarboxylase. J Biol Chem 281: 14529–14532.
  4. 4. Pegg AE, McCann PP (1982) Polyamine metabolism and function. Am J Physiol 243: C212–221.
  5. 5. Pendeville H, Carpino N, Marine JC, Takahashi Y, Muller M, et al. (2001) The ornithine decarboxylase gene is essential for cell survival during early murine development. Mol Cell Biol 21: 6549–6558.
  6. 6. Mangold U (2006) Antizyme inhibitor: mysterious modulator of cell proliferation. Cell Mol Life Sci 63: 2095–2101.
  7. 7. Tabor CW, Tabor H (1984) Polyamines. Annu Rev Biochem 53: 749–790.
  8. 8. Bello-Fernandez C, Packham G, Cleveland JL (1993) The ornithine decarboxylase gene is a transcriptional target of c-Myc. Proc Natl Acad Sci USA 90: 7804–7808.
  9. 9. Esko JD (2002) Tumor attenuation by combined heparan sulfate and polyamine depletion. Proc Natl Acad Sci USA 99: 371–376.
  10. 10. Childs AC, Mehta DJ, Gerner EW (2003) Polyamine-dependent gene expression. Cell Mol Life Sci 60: 1394–1406.
  11. 11. Auvinen M, Paasinen A, Andersson LC, Holtta E (1992) Ornithine decarboxylase activity is critical for cell transformation. Nature 360: 355–358.
  12. 12. O'Brien TG, Megosh LC, Gilliard G, Soler AP (1997) Ornithine decarboxylase overexpression is a sufficient condition for tumor promotion in mouse skin. Cancer Res 57: 2630–2637.
  13. 13. Liu GY, Hung YC, Hsu PC, Liao YF, Chang WH, et al. (2005) Ornithine decarboxylase prevents tumor necrosis factor alpha-induced apoptosis by decreasing intracellular reactive oxygen species. Apoptosis 10: 569–581.
  14. 14. Huang CC, Hsu PC, Hung YC, Liao YF, Liu CC, et al. (2005) Ornithine decarboxylase prevents methotrexate-induced apoptosis by reducing intracellular reactive oxygen species production. Apoptosis 10: 895–907.
  15. 15. Hsu PC, Hour TC, Liao YF, Hung YC, Liu CC, et al. (2006) Increasing ornithine decarboxylase activity is another way of prolactin preventing methotrexate-induced apoptosis: CROSSTALK BETWEEN ODC AND BCL-2. Apoptosis 11: 389–399.
  16. 16. Packham G, Cleveland JL (1994) Ornithine decarboxylase is a mediator of c-Myc-induced apoptosis. Mol Cell Biol 14: 5741–5747.
  17. 17. Gerner EW, Meyskens FL (2004) Polyamines and cancer: old molecules, new understanding. Nat Rev Cancer 4: 781–792.
  18. 18. Mangold U (2005) The antizyme family: polyamines and beyond. IUBMB Life 57: 671–676.
  19. 19. Medina MA, Quesada AR, Nunez de Castro I, Sanchez-Jimenez F (1999) Histamine, polyamines, and cancer. Biochem Pharmacol 57: 1341–1344.
  20. 20. Furumitsu Y, Yukioka K, Yukioka M, Ochi T, Morishima Y, et al. (2000) Interleukin-1beta induces elevation of spermidine/spermine N1-acetyltransferase activity and an increase in the amount of putrescine in synovial adherent cells from patients with rheumatoid arthritis. J Rheumatol 27: 1352–1357.
  21. 21. Morrison LD, Cao XC, Kish SJ (1998) Ornithine decarboxylase in human brain: influence of aging, regional distribution, and Alzheimer's disease. J Neurochem 71: 288–294.
  22. 22. Schluter KD, Frischkopf K, Flesch M, Rosenkranz S, Taimor G, et al. (2000) Central role for ornithine decarboxylase in beta-adrenoceptor mediated hypertrophy. Cardiovasc Res 45: 410–417.
  23. 23. Moshier JA, Dosescu J, Skunca M, Luk GD (1993) Transformation of NIH/3T3 cells by ornithine decarboxylase overexpression. Cancer Res 53: 2618–2622.
  24. 24. Murakami Y, Matsufuji S, Miyazaki Y, Hayashi S (1994) Forced expression of antizyme abolishes ornithine decarboxylase activity, suppresses cellular levels of polyamines and inhibits cell growth. Biochem J 304: 183–187.
  25. 25. Iwata S, Sato Y, Asada M, Takagi M, Tsujimoto A, et al. (1999) Anti-tumor activity of antizyme which targets the ornithine decarboxylase (ODC) required for cell growth and transformation. Oncogene 18: 165–172.
  26. 26. Kim SW, Mangold U, Waghorne C, Mobascher A, Shantz LM, et al. (2006) Regulation of cell proliferation by the antizyme inhibitor: evidence for an antizyme independent mechanism. J Cell Sci 119: 2583–2591.
  27. 27. Feith DJ, Origanti S, Shoop PL, Sass-Kuhn S, Shantz LM (2006) Tumor suppressor activity of ODC antizyme in MEK-driven skin tumorigenesis. Carcinogenesis 27: 1090–1098.
  28. 28. Seiler N, Atanassov CL, Raul F (1998) Polyamine metabolism as target for cancer chemoprevention (Review). Int J Oncol 13: 993–1006.
  29. 29. Burns MR, Graminski GF, Weeks RS, Chen Y, O'Brien TG (2009) Lipophilic lysine-spermine conjugates are potent polyamine transport inhibitors for use in combination with a polyamine biosynthesis inhibitor. J Med Chem 52: 1983–1989.
  30. 30. Marton LJ, Pegg AE (1995) Polyamines as targets for therapeutic intervention. Annu Rev Pharmacol Toxicol 35: 55–91.
  31. 31. McCann PP, Pegg AE (1992) Ornithine decarboxylase as an enzyme target for therapy. Pharmacol Ther 54: 195–215.
  32. 32. Wallace HM, Fraser AV (2004) Inhibitors of polyamine metabolism: review article. Amino acid 26: 353–365.
  33. 33. Li X, Coffino P (1992) Regulated degradation of ornithine decarboxylase requires interaction with the polyamine-inducible protein antizyme. Mol Cell Biol 12: 3556–3562.
  34. 34. Li X, Coffino P (1993) Degradation of ornithine decarboxylase: exposure of the C-terminal target by a polyamine-inducible inhibitory protein. Mol Cell Biol 13: 2377–2383.
  35. 35. Murakami Y, Matsufuji S, Kameji T, Hayashi S, Igarashi K, et al. (1992) Ornithine decarboxylase is degradaded by the 26S proteasome without ubiquitination. Nature 360: 597–600.
  36. 36. Thomas T, Thomas TJ (2001) Polyamines in cell growth and cell death: Molecular mechanisms and therapeutic applications. Cell Mol Life Sci 58: 244–258.
  37. 37. Coffino P (1998) Degradation of ornithine decarboxylase, in Ubiquitin and the Biology of the Cell. In: Peters JM, Harris JR, Finley D, editors. New York: Plenum Press. pp. 411–427.
  38. 38. Coffino P (2001) Regulation of cellular polyamines by antizyme. Nat Rev Mol Cell Biol 2: 188–194.
  39. 39. Palanimurugan R, Scheel H, Hofmann K, Dohmen RJ (2004) Polyamines regulate their synthesis by inducing expression and blocking degradation of ODC antizyme. EMBO J 23: 4857–4867.
  40. 40. Zhang M, Pickart CM, Coffino P (2003) Determinants of proteasome recognition of ornithine decarboxylase, a ubiquitin independent substrate. EMBO J 22: 1488–1496.
  41. 41. Coffino P (2001) Antizyme, a mediator of ubiquitin-independent proteasomal degradation. Biochimie 83: 319–323.
  42. 42. Su KL, Liao YF, Hung HC, Liu GY (2009) Critical factors determining dimerization of human antizyme inhibitor. J Biol Chem 284: 26768–26777.
  43. 43. Zhang M, MacDonald AI, Hoyt MA, Coffino P (2004) Proteasomes begin ornithine decarboxylase digestion at the C-terminus. J Biol Chem 279: 20959–20965.
  44. 44. Ghoda L, van Daalen Wetters T, Macrae M, Ascherman D, Coffino P (1989) Prevention of rapid intracellular degradation of ODC by a carboxyl-terminal truncation. Science 243: 1493–1495.
  45. 45. Ghoda L, Sidney D, Macrae M, Coffino P (1992) Structural elements of ornithine decarboxylase required for intracellular degradation and polyamine-dependent regulation. Mol Cell Biol 12: 2178–85.
  46. 46. Mitchell JL, Chen HJ (1990) Conformational changes in ornithine decarboxylase enable recognition by antizyme Biochim Biophys Acta 1037: 115–121.
  47. 47. Jackson LK, Goldsmith EJ, Phillips MA (2003) X-ray structure determination of Trypanosoma brucei ornithine decarboxylase bound to D-ornithine and to G418: insights into substrate binding and ODC conformational flexibility. J Biol Chem 278: 22037–22043.
  48. 48. Coleman CS, Stanley BA, Viswanath R, Pegg AE (1994) Rapid exchange of subunits of mammalian ornithine decarboxylase. J Biol Chem 269: 3155–3158.
  49. 49. Hoffman DW, Carroll D, Martinez N, Hackert ML (2005) Solution structure of a conserved domain of antizyme: a protein regulator of polyamines. Biochemistry 44: 11777–11785.
  50. 50. Ivanov IP, Gesteland RF, Atkins JF (1998) A second mammalian antizyme: conservation of programmed ribosomal frameshifting. Genomics 52: 119–129.
  51. 51. Matsufuji S, Matsufuji T, Miyazaki Y, Murakami Y, Atkins JF, et al. (1995) Autoregulatory frameshifting in decoding mammalian ornithine decarboxylase antizyme. Cell 80: 51–60.
  52. 52. Rom E, Kahana C (1994) Polyamines regulate the expression of ornithine decarboxylase antizyme in vitro by inducing ribosomal frame-shifting. Proc Natl Acad Sci USA 91: 3959–3963.
  53. 53. Kahana C (2009) Antizyme and antizyme inhibitor, a regulatory tango. Cell Mol Life Sci 66: 2479–2488.
  54. 54. Sakata K, Kashiwagi K, Igarashi K (2000) Properties of a polyamine transporter regulated by antizyme. Biochem J 347: 297–303.
  55. 55. Zhu C, Lang DW, Coffino P (1999) Antizyme2 is a negative regulator of ornithine decarboxylase and polyamine transport. J Biol Chem 274: 26425–2.
  56. 56. Bacchi CJ, Nathan H, Hutner SH, McCann PP, Sjoerdsma A (1980) Polyamine metabolism: a potential therapeutic target in trypanosomes. Science 210: 332–4.
  57. 57. Van Nieuwenhove S, Schechter PJ, Declercq J, Bone G, Burke J, et al. (1985) Treatment of gambiense sleeping sickness in the Sudan with oral DFMO (DL-alpha-difluoromethylornithine), an inhibitor of ornithine decarboxylase; first field trial. Trans R Soc Trop Med Hyg 79: 692–8.
  58. 58. Coleman CS, Stanley BA, Pegg AE (1993) Effect of mutations at active site residues on the activity of ornithine decarboxylase and its inhibition by active site-directed irreversible inhibitors. J Biol Chem 268: 24572–24579.
  59. 59. Ghoda L, Phillips MA, Bass KE, Wang CC, Coffino P (1990) Trypanosome ornithine decarboxylase is stable because it lacks sequences found in the carboxyl terminus of the mouse enzyme which target the latter for intracellular degradation. J Biol Chem 265: 11823–11826.
  60. 60. Almrud JJ, Oliveira MA, Kern AD, Grishin NV, Phillips MA, et al. (2000) Crystal structure of human ornithine decarboxylase at 2.1 Å resolution: structural insights to antizyme binding. J Mol Biol 295: 7–16.
  61. 61. Liu YC, Liu YL, Su JY, Liu GY, Hung HC (2011) Critical Factors Governing the Difference in Antizyme-Binding Affinities between Human Ornithine Decarboxylase and Antizyme Inhibitor. PLoS ONE 6(4): e19253.
  62. 62. Bradford MM (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72: 248–254.
  63. 63. Schuck P, Perugini MA, Gonzales NR, Howlett GJ, Schubert D (2002) Size distribution analysis of proteins by analytical ultracentrifugation: strategies and application to model systems. Biophys J 82: 1096–1111.
  64. 64. Schuck P (2003) On the analysis of protein self-association by sedimentation velocity analytical ultracentrifugation. Anal Biochem 320: 104–124.
  65. 65. Brown PH, Balbo A, Schuck P (2008) Characterizing protein-protein interactions by sedimentation velocity analytical ultracentrifugation. Curr Protoc Immunol Chapter 18, Unit 18 15.
  66. 66. Dam J, Schuck P (2005) Sedimentation velocity analysis of heterogeneous protein-protein interactions: sedimentation coefficient distributions c(s) and asymptotic boundary profiles from Gilbert-Jenkins theory. Biophys J 89: 651–666.
  67. 67. Laue TM, Shah BD, Ridgeway TM, Pelleter SL (1992) Analytical Ultracentrifugation in Biochemistry and Polymer Science. Cambridge, UK: The Royal Society of Chemistry.