pmc logo imageJournal ListSearchpmc logo image
Logo of brjpharmJournal URL: redirect3.cgi?&&auth=0dITyhdtTrRuaqYxvUUFzRDzCiIe8S_8woqRWVAsz&reftype=publisher&artid=1575421&article-id=1575421&iid=134961&issue-id=134961&jid=282&journal-id=282&FROM=Article|Banner&TO=Publisher|Other|N%2FA&rendering-type=normal&&http://www.nature.com/bjp
Br J Pharmacol. 2004 October; 143(4): 495–507.
Published online 2004 October 5. doi: 10.1038/sj.bjp.0705986.
PMCID: PMC1575421
Allosteric modulation of semicarbazide-sensitive amine oxidase activities in vitro by imidazoline receptor ligands
Andrew Holt,1,2* Barbara Wieland,3 and Glen B Baker2
1Department of Pharmacology, 9-70 Medical Sciences Building, University of Alberta, Edmonton, Alberta, Canada T6G 2H7
2Neurochemical Research Unit, Department of Psychiatry, WC Mackenzie Health Sciences Centre, University of Alberta, Edmonton, Alberta, Canada T6G 2R7
3Department of Chemistry, University of Alberta, Edmonton, Alberta, Canada T6G 2G2
*Author for correspondence: Email: Andy.Holt/at/ualberta.ca
Received June 16, 2004; Revised August 12, 2004; Accepted August 25, 2004.
Abstract
  • Evidence indicates that imidazoline I2 binding sites (I2BSs) are present on monoamine oxidase (MAO) and on soluble (plasma) semicarbazide-sensitive amine oxidase enzymes. The binding site on MAO has been described as a modulatory site, although no effects on activity are thought to have been observed as a result of ligands binding to these sites.
  • We examined the effects in vitro of several imidazoline binding site ligands on activities of bovine plasma amine oxidase (BPAO) and porcine kidney diamine oxidase (PKDAO) in a spectrophotometric protocol.
  • While both enzymes were inhibited at high concentrations of all ligands, clonidine, cirazoline and oxymetazoline were seen, at lower concentrations, to increase activity of BPAO versus benzylamine, but not of PKDAO versus putrescine. This effect was substrate dependent, with mixed or biphasic inhibition of spermidine, methylamine, p-tyramine and β-phenylethylamine oxidation observed at cirazoline concentrations that increased benzylamine oxidation.
  • With benzylamine as substrate, clonidine decreased KM (EC50 8.82 μM, Emax 75.1% of control) and increased Vmax (EC50 164.6 μM, Emax 154.1% of control). Cirazoline decreased Vmax (EC50 2.15 μM, Emax 91.4% of control), then decreased KM (EC50 5.63 μM, Emax 42.6% of control) and increased Vmax (EC50 49.0 μM, Emax 114.4% of decreased Vmax value).
  • Data for clonidine fitted a mathematical model for two-site nonessential activation plus linear intersecting noncompetitive inhibition. Data for cirazoline were consistent with involvement of a fourth site.
  • These results reveal an ability of imidazoline ligands to modulate BPAO kinetics allosterically. The derived mechanism may have functional significance with respect to modulation of MAO by I2BS ligands.
Keywords: MAO, semicarbazide-sensitive amine oxidase (SSAO), bovine plasma amine oxidase (BPAO), imidazoline, modulation, activation, cirazoline, clonidine, benzylamine, in vitro
Introduction

Following the discovery that clonidine exerts its hypotensive effect at least in part through an action at novel nonadrenergic receptors in the ventrolateral medulla (Bousquet et al., 1984), interest in these imidazoline receptors, as well as in other binding sites for clonidine and related drugs, has blossomed (see Eglen et al., 1998; Head, 2003 for reviews). At least three subclasses of binding sites have been claimed to exist, with only the imidazoline type-1 receptor (I1R), which is thought to play a role in central control of blood pressure (see Bousquet & Feldman, 1999), having been shown thus far to be associated with a signal transduction system (Takada et al., 1997; Edwards et al., 2001). Imidazoline-3 binding sites (I3BSs) encompass a nebulous group of loci associated predominantly with various ion channel subunits, and while ligands may modulate processes such as insulin secretion and catecholamine release (see Molderings, 1997; Morgan & Chan, 2001 for reviews), some of these effects may be due to drugs acting at two or more different binding sites in target tissues (Morgan & Chan, 2001).

While a variety of physiological effects can be ascribed to ligands acting at I1Rs or I3BSs, the same may not be said for those binding to imidazoline-2 binding sites (I2BSs).

The observation that I2BSs were located on the outer membranes of mitochondria (Tesson et al., 1991) quickly led to the realisation that many such sites were actually located on the monoamine oxidase (MAO) protein (Olmos et al., 1993; Alemany et al., 1995; Carpéné et al., 1995; Tesson et al., 1995). Imidazoline ligands bound with high affinity to these sites and were able to inhibit MAO and other amine oxidases in vitro (Carpéné et al., 1995), while chronic treatment of rats with amine oxidase inhibitors caused either a downregulation (Alemany et al., 1995) or an upregulation (Wiest & Steinberg, 1997) of I2BSs in the brain or liver. Some functional interplay between the I2BS on MAO and enzyme activity was thus presumed to occur, and the I2BS became increasingly referred to as a modulatory or regulatory site on MAO (Tesson et al., 1995).

In support of such a designation, numerous studies have shown clearly that the I2BS is distinct from the MAO active site (Raddatz et al., 1997; 1999) and that, in rats at least, MAO-B appears to be the predominant enzyme isoform upon which these sites might be found (Raddatz et al., 2000; Remaury et al., 2000). However, while imidazolines certainly bind with high affinity to the I2BS on MAO, inhibition of enzyme activity requires rather higher ligand concentrations than are necessary to saturate the I2BS (Ozaita et al., 1997).

A few other enzymes also show affinity for I2BS ligands, including several amine oxidases of the EC 1.4.3.6 subclass (Carpéné et al., 1995; Holt & Baker, 1995; Holt et al., 2003). This family of copper-containing amine oxidases includes diamine oxidase, plasma amine oxidases and tissue-bound semicarbazide-sensitive amine oxidase (SSAO) (Callingham et al., 1991). It is likely that the active sites of these enzymes are responsible, at least in part, for binding imidazoline ligands, since several endogenous ligands for I1Rs, I2BSs and I3BSs, including agmatine, tryptamine, histamine and polyamines, are substrates for one or more of these enzymes (Holt & Baker, 1995; Lyles, 1996). However, the possibility remains that an I2BS separate from the active site may exist on semicarbazide-sensitive amine oxidase proteins.

If the high-affinity I2BS on MAO, and perhaps on semicarbazide-sensitive amine oxidases, corresponds to a modulatory site and not simply to an artefactual binding site, then modulation of substrate turnover would necessarily take place on ligand binding to the enzyme. Some precedent does exist for rapid activation of amine oxidase activities. Human platelet MAO and rat brain MAO can be activated in vitro in a substrate-selective manner by a component of human plasma, as a result of a reduction in substrate KM values (Yu & Boulton, 1979; Wahlund et al., 1984). Human plasma also activates human lung SSAO, once again in a substrate-selective fashion, by increasing substrate affinity (Dalfó et al., 2003), and captopril, although not possessing an imidazoline structure, activates rat MAO in vitro through increasing Vmax (Raasch et al., 2002).

An extensive study by Carpéné et al. (1995) showed concentration-dependent inhibition of MAO by high concentrations of several imidazoline ligands; at low concentrations, MAO activity was similar to control values. However, in the same study, tranylcypromine and cirazoline, an imidazoline I2BS/I1R ligand, appeared to induce a small, statistically insignificant potentiation of bovine plasma amine oxidase (BPAO), an EC 1.4.3.6 semicarbazide-sensitive amine oxidase enzyme, at low concentrations of ligand. This effect was not commented upon by the authors. A related amine oxidase from porcine plasma was stimulated in vitro, in a pH-dependent manner, by including ammonia or imidazole in the enzyme reaction (Kelly et al., 1981; Yadav & Knowles, 1981).

Given the clear precedent for rapid modulation of several amine oxidase enzymes and the ability of imidazole, and perhaps cirazoline and tranylcypromine, to stimulate plasma amine oxidases in vitro, it was decided to examine the abilities of several imidazoline binding site (IBS) ligands to influence the activities both of BPAO and of diamine oxidase from pig kidney (PKDAO). Both of these enzymes are semicarbazide-sensitive amine oxidases of the EC 1.4.3.6 subclass, and they share a common quinone cofactor and similar reaction mechanisms (Klinman, 2003).

Methods

Enzyme preparation
Commercially available BPAO and PKDAO were dissolved in potassium phosphate buffer (0.2 M, pH 7.4), usually at 1 or 3 mg ml−1 (BPAO) or 7.5 mg ml−1 (PKDAO), without further purification. Final assay concentrations were 0.167 or 0.5 mg ml−1, and 1.25 mg ml−1, respectively. Enzyme solutions were prepared freshly on the day of assay and were retained on ice until required.

Amine oxidase activity measurements
Enzyme activities were measured by a modification of the colorimetric protocol of Holt et al. (1997). Assays were performed in 96-well polystyrene plates in a total volume of 300 μl and generally contained 50 μl enzyme, 50 μl chromogenic solution, 50 μl amine substrate or water (blanks) and 150 μl water (controls) or test compound. The chromogenic solution contained vanillic acid, 4-aminoantipyrine and horseradish peroxidase in potassium phosphate buffer (Holt et al., 1997), with the peroxide generated during amine turnover initiating the peroxidase-dependent formation of a red quinoneimine dye. Amine substrates were benzylamine (final concentration 740 μM) for BPAO, and putrescine (final concentration 288 μM) for PKDAO, concentrations similar to their respective KM values. In some kinetic experiments (see below), a range of benzylamine concentrations was used, between 10 μM and 3 mM. In examinations of ligand effects upon turnover of different substrates, p-tyramine, β-phenylethylamine, methylamine and spermidine were also used as BPAO substrates, at 135, 275, 420 and 50 μM, respectively. These final concentrations are equal to one-half of their KM values.

Assays were performed in duplicate or triplicate wells, at 37°C, with reactions usually initiated by adding amine substrates that had been prewarmed to 37°C. The increase in absorbance at 498 nm was monitored continuously in a Molecular Devices SpectraMax Plus 384 plate reader for an appropriate period of time, and initial rates of enzyme activities were determined by linear regression of data (SOFTmax PRO, version 3.1.2).

BPAO and drug purity assessments
The BPAO used was purchased from Sigma and was included in assays without further purification. However, since more than one plasma amine oxidase activity versus benzylamine has been detected in other ruminants, such as the sheep (Elliott et al., 1992), it was necessary to clarify how many activities in the bovine plasma extract might contribute to amine metabolism.

Proteins in the solid obtained from Sigma (86% protein, w w−1) were separated by SDS–polyacrylamide gel electrophoresis. Samples were separated under denaturing conditions and lanes were either stained with Coomassie blue or were electroblotted to a nitrocellulose membrane. The membrane was then incubated in 2 M potassium glycinate buffer (pH 10) containing nitroblue tetrazolium (0.24 mM) for 45 min in the dark. This procedure results in deposition of purple formazan on the nitrocellulose paper over protein bands, such as that for BPAO, containing a quinone or quinoid cofactor (Paz et al., 1991; Holt et al., 1998). Gel electrophoresis (not shown) revealed that the commercial enzyme preparation contained several major and numerous minor protein components. Quinone staining of an electroblot identified one major band, at 82 kDa, consistent with the mass of a single BPAO subunit (Mu et al., 1994). Very faint formazan staining was also observed at 101 kDa, similar to a minor band at 102 kDa seen with partially purified porcine aorta SSAO (Holt et al., 1998). Thus, although the commercial BPAO preparation contains only one major quinone amine oxidase activity, it is possible that the 101 kDa component represents a subunit from vascular smooth muscle SSAO. Nevertheless, kinetic evidence in this study did not support participation of a second enzyme activity in benzylamine turnover.

Since benzylamine is also a substrate for MAO-B, the possibility was examined that some platelet MAO-B activity might be present in the bovine plasma extract and thus contribute to reaction velocity. Colorimetric assays were performed as described above, but with benzylamine substrate present at a final concentration of 2 mM, and either in the presence or absence of the MAO inhibitor pargyline (10 μM), added 10 min before substrate. While electrophoresis indicated that one or more proteins were present with masses close to 60 kDa, not inconsistent with the potential presence of MAO-B, the failure of pargyline to have any effect on benzylamine turnover (not shown) indicated that no active MAO-B enzyme was present in the preparation.

In order to assess the purity of drug compounds, aqueous solutions of clonidine and cirazoline were diluted in HPLC-grade methanol and were injected directly onto a Waters Micromass ZQ quadrupole mass spectrometer, with positive electrospray ionisation. Mass spectra (not shown) revealed the presence of parent ions with m/z ratios of 229.93, 231.94 and 233.95 (clonidine, containing 35Cl and 37Cl) and 217.03 (cirazoline), consistent with calculated M+H+ values for the drugs. No other charged species could be detected above baseline noise in either sample, suggesting that drugs were devoid of contaminants that might demonstrate efficacy in subsequent assays.

Kinetic analyses of BPAO activity
In kinetic analyses, BPAO was incubated in the presence of a range of concentrations of benzylamine substrate (10 μM–3 mM), in the absence (controls) or presence of imidazoline test drugs. In order to obtain kinetic constants, data were fitted to a Michaelis–Menten hyperbola with the nonlinear regression facility of GraphPad Prism version 4.00 (GraphPad Software, San Diego, CA, U.S.A.). Sigmoidal curves were fitted to data, where appropriate, with the nonlinear regression facility, and Dixon plots and kinetic replots with the linear regression facility of GraphPad Prism.

Based on results obtained from kinetic studies, a reaction mechanism was proposed and a novel equation formulated to describe the observed behaviour. Details of the derivation of this mechanism will be published elsewhere.

Assessment of IBS characteristics
I2BSs are differentiated from I1Rs partly on the basis of the former group's high affinity for idazoxan, while subclassification of I2 sites is dependent on a high or low affinity for the diuretic amiloride (Michel & Ernsberger, 1992). Although data from Figure 1a (see Results) suggested that the affinities of moxonidine and idazoxan for BPAO were rather low, it was possible that the modest inhibition observed resulted from interactions of these drugs with the active site, and the possibility remained that moxonidine or idazoxan may bind to the ‘high-affinity' cirazoline/clonidine site, but lack the intrinsic efficacy to alter substrate turnover. Therefore, in order to determine whether the high-affinity sites through which cirazoline and clonidine act resembled typical I1, I2A or I2B sites, the abilities of cirazoline (10 μM) or clonidine (100 μM) to potentiate benzylamine turnover were assessed in the absence or presence of a range of concentrations of moxonidine, idazoxan or amiloride. Results would reveal whether or not the effects of cirazoline or clonidine could be antagonised in a concentration-dependent manner at lower concentrations of the selective ligands than those required to inhibit enzyme activity directly.
Figure 1Figure 1
Preliminary screening results for effects of imidazoline ligands on benzylamine oxidation by BPAO (a) or putrescine oxidation by PKDAO (b). Benzylamine (740 μM) was incubated with BPAO (1 mg ml−1), or putrescine (288 μM) with PKDAO (more ...)

Materials
BPAO (marketed incorrectly as EC 1.4.3.4; bovine plasma MAO) and PKDAO were obtained from Sigma (Oakville, Ontario, Canada). BPAO is no longer available from Sigma.

Benzylamine hydrochloride, putrescine dihydrochloride, methylamine hydrochloride, spermidine trihydrochloride, vanillic acid, 4-aminoantipyrine, peroxidase (type II, from horseradish), amiloride hydrochloride, clonidine hydrochloride, efaroxan hydrochloride, guanabenz, idazoxan hydrochloride, moxonidine hydrochloride and oxymetazoline hydrochloride were from Sigma. 2-(2-Benzofuranyl)-2-imidazoline hydrochloride (2-BFI) was purchased from Tocris (Ballwin, MO, U.S.A.). Cirazoline was a generous gift from Sanofi-Synthélabo (Bagneux, France), and guanfacine a gift from Sandoz (Basel, Switzerland; now Novartis Pharma). Sodium phosphate and potassium phosphate buffer salts were of molecular biology grade, and were obtained from Sigma. HPLC-grade methanol was purchased from Fisher Scientific (Nepean, Ontario, Canada). All other chemicals were of reagent grade.

Results

Preliminary screen for modulation of enzyme activities
Figure 1a shows the effects of several imidazoline ligands on turnover of benzylamine by BPAO, with the substrate present at 740 μM, close to its KM concentration. Guanfacine and guanabenz inhibited activity completely, while 2-BFI and efaroxan appeared to be able to inhibit only around 50% of turnover. However, cirazoline and clonidine both caused a substantial increase in turnover rate, while oxymetazoline had a modest effect in this regard. The most efficacious activator was cirazoline, which increased turnover to 182% of control at 300 μM. Modest activation by oxymetazoline appeared to be lost as its concentration was increased, although no inhibition per se was evident. Moxonidine and idazoxan began to inhibit activity only at the highest ligand concentrations used. Similar results were obtained when the same compounds were included in an assay of benzylamine turnover by sheep plasma amine oxidase (not shown).

In contrast, PKDAO was inhibited by all compounds tested (Figure 1b), with no indication of activation at lower ligand concentrations. For those compounds that inhibited both enzymes, the rank orders of inhibitory potency were similar (guanfacine=guanabenz>2-BFI), and in the absence of activation, weak inhibition of PKDAO by cirazoline, clonidine and oxymetazoline was evident. While 2-BFI appeared to be able to inhibit BPAO only by around 50%, inhibition of PKDAO by 2-BFI appeared uniphasic and largely complete at 1 mM. Efaroxan, a structural analogue of 2-BFI, which inhibited BPAO to a degree and with a potency similar to that of 2-BFI was virtually without effect on PKDAO.

Activation of BPAO-mediated benzylamine oxidation by cirazoline and clonidine was only observed above pH 6.00 and below pH 8.33 (data not shown). Although activity was highest between pH 7.33 and 7.67, both in control wells and in the presence of cirazoline or clonidine, the greatest degree of potentiation was seen at pH 6.33 for 30 μM cirazoline (152% of control activity at pH 6.33) and at pH 6.67 for 300 μM clonidine (164% of control activity at pH 6.67) in 167 μM sodium phosphate buffer.

The activation of BPAO in the absence of similar effects upon PKDAO suggests that the effects observed with BPAO are not due simply to nonspecific interactions of some imidazoline ligands with all copper-containing quinone cofactor amine oxidases. Furthermore, neither cirazoline nor clonidine interacted directly with components of the chromogenic solution, including peroxidase, or acted as BPAO substrates from which peroxide would be generated (results not shown).

Assessment of IBS characteristics
Figure 2 illustrates that idazoxan, amiloride and moxonidine did not compete with cirazoline and prevent the stimulation of enzyme activity that occurred in the presence of cirazoline at the low (nanomolar) concentrations at which these ligands have been shown to bind to I1 or I2 sites. Similar results were obtained when BPAO was instead stimulated with clonidine (not shown). These observations suggest that the sites through which cirazoline and clonidine act to stimulate enzyme activity do not possess characteristics typical of I1, I2A or I2B sites (but see Discussion).
Figure 2Figure 2
Effects of the I1R ligand moxonidine (MOX), the I2BS ligand idazoxan (IDZ) and the I2ABS ligand amiloride (AM) on the ability of cirazoline (CIR; 60 μM) to potentiate oxidation of benzylamine (300 μM) by BPAO. Moxonidine, idazoxan or amiloride (more ...)

Although inhibition of BPAO by moxonidine, idazoxan and amiloride could only be observed at high concentrations of these drugs, with the consequence that IC50 values could not be determined with any certainty, it was nevertheless evident that inhibitory potencies of these compounds were increased in the presence of cirazoline (Figure 2). The effects of clonidine were less substantial in this regard, with the potency of amiloride in particular remaining entirely unaffected by the presence of clonidine (not shown). These data suggest that binding of cirazoline, and to a lesser extent clonidine, to BPAO increases the affinity of the enzyme for at least some imidazoline ligands.

Effects of imidazoline ligands on turnover of other substrates
In addition to benzylamine, BPAO can oxidise other aromatic amines such as p-tyramine and β-phenylethylamine, as well as aliphatic amines such as methylamine, and polyamines such as spermidine. Following preliminary experiments to determine kinetic constants for these substrates (not shown), BPAO was incubated with benzylamine, p-tyramine, β-phenylethylamine, methylamine or spermidine, at concentrations equal to one-half of their KM values, in the presence of a range of concentrations of cirazoline (100 nM–3 mM). Figure 3 shows that cirazoline once again caused a concentration-dependent increase in benzylamine turnover, which was followed, at higher cirazoline concentrations, by enzyme inhibition. However, turnover of other aromatic and aliphatic amines was not potentiated. Rather, enzyme activity versus these substrates was inhibited; inhibition of oxidation of p-tyramine, β-phenylethylamine and methylamine was clearly biphasic, while inhibition of spermidine turnover appeared to be inhibited by cirazoline in a uniphasic manner. These observations suggest that cirazoline may have at least two binding sites on the BPAO enzyme, with binding of the drug to either site resulting in a distinct, substrate-dependent change in amine oxidation kinetics.
Figure 3Figure 3
Effects of cirazoline (0.1 μM–3 mM) on oxidation by BPAO of benzylamine (BZ; 300 μM), p-tyramine (TYR; 135 μM), β-phenylethylamine (PEA; 275 μM), methylamine (MA; 420 μM) or spermidine (SPD; 50 (more ...)

Modulation of kinetic constants by imidazoline ligands
To determine whether or not kinetic constants (KM and Vmax) were altered in a concentration-dependent manner by imidazoline ligands, a range of concentrations of benzylamine (10 μM–3 mM) was incubated with BPAO in the presence of a range of concentrations of cirazoline (1 μM–10 mM) or clonidine (1 μM–50 mM). Data were fitted initially to simple Michaelis–Menten hyperbolae and kinetic constants determined by nonlinear regression. Excellent fits for all data sets were generally obtained, with r2 typically higher than 0.995.

Figure 4a is a representative plot illustrating that cirazoline caused a concentration-dependent reduction in the KM of BPAO for benzylamine, with an EC50 value in this example of 5.9 μM and an Emax reducing KM to 44.8% of control. At higher concentrations, the KM value was seen to increase, due to inhibition of activity by cirazoline. Vmax was increased by cirazoline to 110% of control (Figure 4b), with an EC50 of 44.0 μM. However, it was observed that at very low micromolar concentrations, cirazoline appeared already to have reduced Vmax by around 6%; two separate experiments suggested that this minor effect was reproducible, and occurred with an EC50 of approximately 2 μM. Thus, the effect of cirazoline in increasing Vmax was, in this representative experiment, to activate the enzyme to 117% of baseline activity.

Figure 4Figure 4
Effects of cirazoline on the kinetic constants KM (a) and Vmax (b) for benzylamine turnover by BPAO. Benzylamine (10 μM–3 mM) was incubated at 37°C with BPAO (0.167 mg ml−1) in the presence of water (controls) or cirazoline (more ...)

Figure 5a illustrates results from a representative experiment, showing that clonidine reduced KM in a concentration-dependent manner to 78.2% of the control value. The EC50 in this example was 7.0 μM. This stimulatory effect was rapidly obscured by the apparent increase in KM as a result of enzyme inhibition by clonidine. The major effect of clonidine was to increase Vmax (Figure 5b). The EC50 in this example was 133.6 μM, and the Emax of clonidine was apparent as an increase in substrate turnover to 156% of the control Vmax value, consistent with the effect on v shown in Figures 1a and 3.

Figure 5Figure 5
Effects of clonidine on the kinetic constants KM (a) and Vmax (b) for benzylamine turnover by BPAO. Benzylamine (10 μM–3 mM) was incubated at 37°C with BPAO (0.167 mg ml−1) in the presence of water (controls) or clonidine (more ...)

Table 1 lists mean effects of cirazoline and clonidine on kinetic constants for benzylamine turnover by BPAO.

Table 1Table 1
Effects of imidazoline ligands on kinetic constants for benzylamine turnover by BPAO

The inhibition by cirazoline of spermidine oxidation (see Figure 3) was examined in greater detail (results not shown). Kinetic plots revealed a mixed pattern of inhibition, with an increase in KM accompanied by a decrease in Vmax. These observations are inconsistent with simple competitive inhibition at the BPAO active site, and instead indicate that cirazoline inhibits activity through binding to at least one site on the enzyme that is distinct from the active site.

Mechanism of inhibition of benzylamine oxidation at high concentrations of imidazolines
At concentrations of imidazoline compounds above approximately 1 mM, the rate of benzylamine oxidation was seen to decrease, as a combined result of a reduced Vmax value and an apparent increase in KM. Lineweaver–Burk plots of velocity data obtained in the presence of ligand concentrations (cirazoline, 300 μM–10 mM; clonidine, 3–50 mM) at which activating effects were already maximal, and at which inhibition was clearly present, showed common intercepts to the left of the 1/v axis, and above the 1/[S] axis (not shown). These observations are indicative of linear mixed-type inhibition (Segel, 1993). Abscissal intercepts from replots of KM/Vmax versus [cirazoline] (Figure 6a) or [clonidine] (Figure 6c) indicated that inhibitor dissociation constant (Ki) values for these compounds in the absence of bound substrate were approximately 2.64 and 8.8 mM, respectively. However, as revealed by abscissal intercepts from replots of 1/Vmax versus [cirazoline] (Figure 6b) or [clonidine] (Figure 6d), constants increased by a factor γ if substrate was also bound, such that Ki values increased to approximately 14.55 mM (cirazoline) and 270 mM (clonidine). Thus, values for γ, which also reflect the increase in KM that occurs on binding of the inhibitor to the enzyme–substrate (ES) complex, could be estimated as 5.51 (for E-benzylamine-cirazoline) and 30.7 (for E-benzylamine-clonidine) (Segel, 1993).
Figure 6Figure 6
Replots of kinetic constants for oxidation of benzylamine by BPAO in the presence of high (inhibitory) concentrations of cirazoline or clonidine. KM and Vmax values for benzylamine oxidation were obtained by nonlinear regression analysis of velocity data, (more ...)

Dixon plots (1/v versus [I]) at a range of substrate concentrations for a representative data set were linear for high (inhibitory) concentrations of both cirazoline and clonidine (not shown), suggesting that the complexes formed between enzyme, benzylamine and either cirazoline or clonidine were unable to yield product (Segel, 1993). However, while common intercepts of Dixon plots confirmed the estimated Ki value for cirazoline of 2.64 mM, they suggested that the Ki for clonidine may be two- to three-fold higher than the value of 8.8 mM, determined from Figure 6c, perhaps a consequence of the outlying point for 50 mM clonidine on that plot. Dixon slope replots (Figure 7) were thus used to confirm estimates for inhibitor constants, and yielded respective values for Ki, γKi and γ of 2.85 mM, 15.95 mM and 5.6 (cirazoline) and 19.1 mM, 379 mM and 19.8 (clonidine).

Figure 7Figure 7
Replots of slopes obtained from Dixon plots (1/v versus [I]) at several substrate concentrations and at high (inhibitory) concentrations of cirazoline and clonidine. Mean velocity data from triplicate determinations, representative (more ...)

The present data are consistent with intersecting linear noncompetitive inhibition (Segel, 1993), a simple mixed inhibition system in which the EI complex has a lower affinity for substrate than does free enzyme, and complexes containing both substrate and inhibitor are nonproductive.

Proposed reaction scheme
The U-shaped (hormetic) changes observed in kinetic constants in the presence of clonidine and cirazoline could occur as a result of binding, at higher ligand concentrations, of more than one drug molecule within a single binding site. Alternatively, the results obtained might also suggest that clonidine interacts with the enzyme at three distinct sites, while cirazoline might also interact with a fourth site. BPAO is a dimer composed of two identical subunits, each of which contains an active site. While it is highly improbable that each monomer also contains two or more distinct allosteric sites, it is far more likely that a single allosteric site exists on each monomer, and that a degree of communication takes place between the active site and the allosteric site within each monomer, and also between these sites on opposing monomers. Although we found no evidence for cooperativity with respect to substrate turnover in the present study, an extreme form of negative cooperativity, half-site reactivity, has been found to occur when phenylhydrazine binds to BPAO (Morpurgo et al., 1992) and when p-nitrophenylhydrazine binds to a copper amine oxidase from Aspergillus niger (Frébort et al., 1995), and we would have been unable to detect such an effect under the present experimental conditions. Such a mechanism would allow for oxidation kinetics at the active site on one monomer to be influenced allosterically by binding of an imidazoline ligand to the opposite active site, to an allosteric site on the same monomer and to an allosteric site on the opposite monomer. A reaction scheme modelling this mechanism has been devised, which reflects observations made with clonidine. Details of this mechanistic scheme and the equation derived therefrom will be published elsewhere. In this model, enzyme can bind with substrate, and/or clonidine at one, two, or all three of three drug binding sites. Binding of clonidine to site ‘H' (in this case, a high-affinity site) with dissociation constant KH alters the affinity for benzylamine substrate by a factor, α. Binding to site ‘L' (in this case, a low-affinity site) with dissociation constant KL alters the rate of release of product (Vmax) by a factor, β. Binding to site ‘I' with dissociation constant Ki reduces the affinity of the enzyme for benzylamine substrate by a factor, γ, and any complex of ES in which clonidine is also bound at site ‘I' is unable to release product. The model makes no assumptions about the location of the drug binding sites on or within the enzyme protein, relative to each other or to the active site, other than that the dissociation constant for loss of clonidine from any one site remains unaffected by binding of clonidine to any combination of the other two clonidine binding sites.

A velocity equation (not shown) was devised based on this scheme by an approach that assumed the existence of rapid equilibrium conditions (Segel, 1993). The equation was entered into the user-defined nonlinear regression equations facility of GraphPad Prism, and was then applied to the representative v versus [S] data sets from which Figures 4 and 5 were generated. The variables α, β, KH and KL were constrained initially to the estimated values listed in parentheses in Table 2, which were deduced from the results listed in Table 1, while estimates for γ and Ki were obtained from Figure 7. Thereafter, β, KL, α and KH were allowed to vary in turn, in order to assess goodness of fit. However, the values for γ and Ki were held constant at their initial values, since the very low affinities of the imidazoline drugs for the inhibitory site meant that at most drug concentrations, large variations in Ki or γ would have a negligible effect on calculated values for v at any given substrate concentration.

Table 2Table 2
Variables determined for clonidine and cirazoline acting to induce two-site nonessential activation and intersecting linear noncompetitive (mixed) inhibition of BPAO, with benzylamine as substrate, by applying a novel equation (see text) to a representative (more ...)

Figure 8a shows that velocity curves at all concentrations of clonidine could be described successfully by the equation. Table 2 lists the best-fit variables obtained by nonlinear regression. With most data sets, it was subsequently found that when three of four variables were constrained to their best-fit values, the fourth could be re-obtained by regression even when the initial estimate was quite different from the calculated best-fit value, indicating that the calculated best-fit values were probably not localised minima or maxima.

Figure 8Figure 8
Representative velocity versus [S] data for clonidine (a) and cirazoline (b) obtained as described in the legend to Figure 4, and fitted to an equation describing enzyme activity in the presence of a drug that is a nonessential activator (more ...)

The reaction scheme devised to describe reactions between BPAO, benzylamine and clonidine provides for one allosteric site, not including the low-affinity inhibitory site, through which modulation of Vmax might be effected. However, cirazoline was observed to reduce Vmax slightly at concentrations lower than those that subsequently decreased KM and increased Vmax, and a reaction scheme that would account for this further interaction would be rather more complex than that devised for clonidine. As such, the scheme for clonidine may be a little too simplistic with respect to providing an indication of all pathways that might alter reaction kinetics. Nevertheless, since the effect of cirazoline in reducing Vmax is rather small, and since it occurs at cirazoline concentrations lower than those required to decrease KM and increase Vmax, it might be expected that the equation would provide an adequate fit for data obtained at higher drug concentrations. Representative data were subsequently analysed as described above for clonidine, and it was found that good fits could be obtained for velocity data in the presence of cirazoline at 100 μM or higher (Figure 8b). At 10 and 30 μM, fits were poor and could be improved only by altering α, KH, β or KL to values that were unrealistic. Fits could not be obtained for data obtained in the presence of 1 or 3 μM cirazoline. Table 2 lists the best-fit variables obtained from data sets for 100 μM cirazoline and higher.

The fits obtained support the contention that allosteric modulation of BPAO by cirazoline and clonidine occurs through interactions of these drugs with at least three distinct sites, and thus that alternative explanations, such as binding of two or more ligands within a single site leading to a loss of efficacy, are unlikely to offer a satisfactory explanation for the present observations. Accordingly, these data also provide indirect support for communication between copper amine oxidase subunits.

Discussion

The present data provide the first conclusive evidence for allosteric modulation of amine oxidase enzymes by imidazoline receptor/binding site ligands. The effects observed with cirazoline and clonidine in this study occurred at micromolar concentrations or higher, and it is doubtful if, for these particular drugs, such effects would be of much consequence in vivo. Rather, these drugs have provided the first indication that imidazoline ligands are able to modulate SSAO enzymes through acting at novel binding sites on these enzymes. Consequently, it is anticipated that other ligands that are rather more potent in this regard will be identified or synthesised.

The present observations are of particular interest in view of the abilities of a number of imidazoline ligands to bind with high affinity to a site on MAO that corresponds, for a subset of human MAO-B enzymes, to a region between Lys 149 and Met 222 (Raddatz et al., 1997; 2000), and it has been proposed that MAO activity may be regulated in a tissue-specific manner by imidazoline ligands interacting with this site (Raddatz et al., 2000). However, with the exception of MAO inhibition through an action claimed to be at the active site (Carpéné et al., 1995; Raasch et al., 1999), effects of these ligands on MAO activity have not been demonstrated at low concentrations around the ligand KD values for these sites. The enzyme examined in the present study is not a membrane-bound EC 1.4.3.4 flavin enzyme from the MAO family, but rather a soluble copper-containing EC 1.4.3.6 quinoprotein from the SSAO family (Janes et al., 1990). Nevertheless, despite numerous physicochemical differences, these enzyme classes share many substrates and inhibitors, and perhaps some functions (Lyles, 1984; 1996; Callingham et al., 1991). For example, chronic inhibition of both activities in vivo results in a downregulation of rat brain I1R, presumably as a result of a prolonged increase in the levels of an endogenous I1R agonist that is also a substrate for both MAO and SSAO (Holt et al., 2003).

Enzyme activation is usually essential in nature, and is most often evident as a requirement for a metal ion (activator) to form a complex with substrate before turnover can occur (Segel, 1993). Nonessential activation, where activity is simply enhanced following binding of the activator to the enzyme protein, is a common phenomenon in cooperative enzyme systems, and activation by low concentrations of a competitive inhibitor at low substrate concentrations is indicative of cooperative binding (MacRae et al., 2000). SSAO enzymes are homodimeric, with one active site per subunit. Crystal structures for EC 1.4.3.6 enzymes from Escherichia coli (Parsons et al., 1995), pea seedling (Kumar et al., 1996) and Arthrobacter globiformis (Wilce et al., 1997) reveal β-ribbon arms that extend from each subunit into the other, with residues from these arms partially lining a channel within the opposite subunit that extends from bulk solvent to the active site and which allows substrates to access the active site (Wilce et al., 1997). Thus, it is conceivable that binding of substrate (or inhibitor) to one active site might influence access of substrate (or inhibitor) to the second through an effect on the second substrate channel mediated through the β-ribbon arm. Since access of substrate to the active site is not thought to be rate limiting in catalysis by these enzymes (Mure et al., 2002), only by limiting access of substrate might a cooperative effect be induced; however, no evidence has been uncovered for negative (or, indeed, positive) cooperativity with respect to substrate turnover in SSAO dimers. This is also true for MAO-B, which is thought to exist as a dimer on the mitochodrial membrane (Binda et al., 2001). Nevertheless, half-site reactivity, an extreme form of negative cooperativity, has been shown to occur with hydrazine and hydrazide inhibitors of SSAO enzymes (Morpurgo et al., 1992; Frébort et al., 1995). Proof of the occurrence of half-site reactivity constitutes proof of crosstalk between subunits. Consequently, the possibility that cooperative substrate binding could occur in BPAO, and thus that low concentrations of an imidazoline ligand binding at one active site might alter substrate binding kinetics at the other active site, is very real.

While the existence of such crosstalk would provide an adequate explanation for the effects of a modulator that acts at a single site, the multisite effects observed with clonidine and cirazoline clearly implicate sites other than the dimer active sites, and the mechanism(s) responsible may thus include, but are not limited to, classical cooperativity. The present data may therefore describe a highly unusual and perhaps unique mechanism.

The model that has been proposed to explain the effects of imidazolines on the BPAO dimer contains a substrate binding site, two sites through which activity is modulated allosterically (either positively or negatively) and a site through which activity is inhibited. The model makes few assumptions, although, for simplicity, it was presumed that binding of an imidazoline drug to any of its sites was without effect on subsequent binding of the drug to the other vacant sites. Experimental data for clonidine fitted well to the derived equation. However, the possibility that cirazoline may bind to one more site than does clonidine indicates that the model may be too simplistic. While clonidine may have bound to an extra site that was not revealed experimentally, the drug failed to influence enzyme activity through binding to that site, and, for clonidine, the model therefore represents an acceptable simplification of the true system.

Cirazoline appears to bind to four sites, causing both a reduction and an increase in Vmax, a reduction in KM, and mixed inhibition at high concentrations. It is possible, but difficult to envisage, that a monomer could contain three or more different allosteric sites with affinities for one drug. Rather, the results can be explained adequately through a homodimeric model with one active site and one modulatory site on each monomer. In order that an explanation based on such a model might fit the present observations with benzylamine and BPAO, the existence of cooperativity between subunits would be necessary. With substrate bound at one active site, cirazoline might then bind to the active site on the second subunit, influencing turnover at the first active site, or at a regulatory site on the first subunit, influencing turnover at the first active site, or at a regulatory site on the second subunit, influencing either turnover of benzylamine at the first active site or binding or efficacy of the cirazoline at the second active site. The latter point is supported by the observation that the inhibitory potencies of cirazoline and clonidine appeared to be increased in the presence of other imidazoline ligands (Figure 2). This suggests that, while it is possible that inhibition is mediated through binding within the active site pocket at which substrate is oxidised, albeit not directly at the substrate binding site itself, it is equally likely that the mixed inhibition instead takes the form of negative cooperativity, with drugs binding to the active site at which substrate is not bound. In deriving an equation to describe this model, no allowance was made for the possibility that a ligand might influence its own affinity or efficacy at one site through binding to an adjacent allosteric site; the hypercubic model including such an interaction would be comprised of 32 enzyme species, 88 equilibrium constants and eight rate constants, and would be both highly complex and of little predictive usefulness.

The IBS on MAO fits the profile for an I2BS, based partly on an affinity for idazoxan in the nanomolar range (Eglen et al., 1998). While the low micromolar affinity of at least one of the sites on BPAO for clonidine is consistent with binding to an I2BS (Eglen et al., 1998), the low potencies of idazoxan and amiloride as antagonists of the potentiating effects of cirazoline and clonidine in the present study would seem to suggest that the sites to which cirazoline and clonidine bind are not I2BSs. Although several of the ligands demonstrated here to have modulatory actions versus BPAO activity are I1R ligands, binding profiles are also not entirely consistent with an involvement of I1Rs, which in any case, like I3BSs, are not associated with amine oxidase enzyme proteins. However, dissociation constants for the interaction of these ligands with BPAO, measured as Ki values versus amine turnover, were determined in the presence of amine substrates, and would not thus be expected to be similar to KD values determined from radioligand binding assays. Consequently, the possibility that these allosteric sites on the BPAO protein possess characteristics typical of the I2 sites already identified on MAO might not be precluded at the present time.

The very different effects of cirazoline on five different substrates indicate that this drug is unlikely to influence only the oxidative (second) half-reaction, since this copper- and oxygen-dependent step involves oxidation of an aminoquinol cofactor to regenerate topa quinone cofactor at the enzyme active site (Klinman, 2003); accordingly, this step should be similar with all substrates used. Rather, the reductive (first) half-reaction, which results in the release of aldehyde following binding and oxidation of amine substrate, is more likely to be the process affected by binding of low concentrations of imidazoline drugs, although a concomitant effect on the oxidative half-reaction may not be ruled out at this time. In support of this hypothesis, the rate-limiting step in benzylamine oxidation by BPAO is thought to be C–H bond cleavage during aldehyde production (Mure et al., 2002), an important step in the reductive half-reaction, and thus the rate of C–H bond cleavage must necessarily be increased in any activation process. This would not occur if the mechanism of action of imidazoline ligands was simply to control access of substrates to the hydrophobic channel, which, as has been shown in a bacterial SSAO enzyme, leads from bulk solvent to the active site (Wilce et al., 1997). The unusual temperature dependence of BPAO activity and the large kinetic isotope effect for benzylamine with this enzyme indicate that hydrogen tunnelling, a process by which activation energy requirements are anomalously reduced, occurs during this C–H bond cleavage step (Grant & Klinman, 1989) in a copper-dependent manner. The distance between the active site aspartate base (proton acceptor) and the substrate (proton donor) has a major influence on the degree to which hydrogen tunnelling may occur (Kohen & Klinman, 1999), and very small changes in the distances between donor, acceptor and active site copper following binding of an imidazoline drug to a binding site in the vicinity of the active site could be reflected in very large changes in rates of catalysis. This might explain the differential effects of cirazoline on oxidation of benzylamine, β-phenylethylamine, p-tyramine, methylamine and spermidine, since a minor spatial rearrangement of residues bordering the active site might facilitate hydrogen tunnelling when benzylamine is bound as substrate, but reduce the extent of tunnelling with the other amines.

Activation of SSAO enzymes in vitro is not entirely without precedent. Elegant kinetic work with porcine plasma amine oxidase by Kelly et al. (1981) revealed that this enzyme could be activated and inhibited by ammonia, a product of oxidation of primary amines by SSAO, and by imidazole, a common constituent in biological buffers. Furthermore, the authors concluded that the simplest model that might explain their observations would possess two sites, in addition to the active site, at which modifiers might bind (Kelly et al., 1981). More recently, a component of human plasma was found to activate SSAO from human lung through increasing affinity for benzylamine substrate (Dalfó et al., 2003). The Emax in that instance was to reduce KM to 36.5% of control, comparable with the Emax of 42.6% determined for cirazoline in this study. While lysophosphatidylcholine was shown also to activate lung SSAO, the efficacy of this plasma phospholipid was rather lower than that of plasma itself, and it was suggested that other components in the plasma may be of greater importance in this regard (Dalfó et al., 2003). Nevertheless, the demonstration that a tissue-bound SSAO enzyme can be activated in vitro through effects of an endogenous species on substrate affinity in a manner analogous to the present observations with soluble SSAO enzymes, and to those of others (Kelly et al., 1981), suggests that allosteric sites similar to those identified in the present study may be present on several different amine oxidase enzymes, and may be physiologically relevant. Furthermore, the findings here and by others of SSAO activation in enzymes from several species and tissue sources, and in both impure and purified enzyme preparations, argue against the likelihood that the imidazoline drugs bind to a second contaminating protein and that this protein–drug complex then interacts with the SSAO dimer to activate or inhibit substrate turnover.

Active site-directed inhibitors of SSAO are being developed as novel anti-inflammatory agents (Koskinen et al., 2004). Such drugs may also reduce atherosclerotic plaque deposition (Yu et al., 2002) and reduce formation of advanced glycation end products and thus vascular damage in diabetics (Yu et al., 2003). In addition, inhibition of adipocyte SSAO results in a reduction in glucose transport into fat cells, and substantial weight loss in obese mice (Yu et al., 2004). However, while selective SSAO inhibition would seem, in some patients, to be beneficial, there is a dearth of inhibitors that show good selectivity for SSAO versus MAO enzymes in vivo, while retaining high potency. The data presented here provide novel information regarding kinetic mechanisms of amine turnover by SSAO, and indicate that novel inhibitors might target sites on the protein separate from the active site, perhaps offering a means by which selectivity could be improved. That such drugs could act in a substrate-dependent manner is an intriguing thought. Furthermore, the observation that benzylamine oxidation by BPAO could be inhibited only by 50% by 2-BFI and efaroxan suggests that these compounds will prove to be useful tools in future examinations of possible cooperative behaviour in SSAO dimers.

In summary, the present study has revealed an ability of some imidazoline ligands to alter reaction kinetics of the EC 1.4.3.6 amine oxidase from bovine plasma in a manner dependent on the substrate used. Similar observations were made in a preliminary experiment with a related enzyme from sheep plasma. Effects, both activatory and inhibitory, appeared to be mediated through drugs binding to three or four sites, and the present results are consistent with some degree of crosstalk occurring between monomers in the SSAO homodimer. A review of the literature indicates that a number of drugs and chemicals can alter reaction kinetics in vitro and potentiate activity of SSAO, and also MAO enzymes, which are known to express a high-affinity I2BS. An ability of imidazoline ligands to alter MAO reaction kinetics and influence substrate specificity in a tissue-specific manner through an action at the I2BS on MAO would reveal a novel target for future antidepressant and neuroprotective compounds. Work continues in our laboratory to clarify the nature of the IBSs on SSAO enzymes, and to identify compounds that alter MAO reaction kinetics through binding to the imidazoline site on that enzyme.

Acknowledgments

We are indebted to Irwin Segel (University of California, Davis) for his extensive advice on appropriate kinetic analytical procedures and to Keith Tipton (Trinity College, Dublin) for assistance with interpretation of data. We also thank Brian Callingham (Cambridge University), David Nutt and Alan Hudson (Bristol University) and Peter Knowles (University of Leeds) for helpful discussions. We are grateful to Monica Palcic (University of Alberta) for assistance with electrophoresis and electroblotting procedures, and to Gail Rauw (University of Alberta) for mass spectrometric analyses. Funding has been provided by the Canadian Institutes of Health Research (CIHR), the Canada Research Chair programme and the George and Dorothy Davey Endowment.

Abbreviations

αfactor by which KM is altered by imidazoline drug
AMamiloride
βfactor by which Vmax is altered by imidazoline drug
2-BFI2-(2-benzofuranyl)-2-imidazoline
BPAObovine plasma amine oxidase
BZbenzylamine
CIRcirazoline
Dimidazoline drug binding to any site on BPAO
Eenzyme
Emaxmaximum degree to which measured parameter can be changed by drug of interest
γfactor by which affinity of enzyme for substrate is reduced by a mixed inhibitor
Himidazoline drug binding to high-affinity site on BPAO
Iimidazoline drug binding to (very low-affinity) inhibitory site on BPAO
IBSimidazoline binding site
IDZidazoxan
I1Rimidazoline type-1 receptor
kprate constant for enzymatic formation of product
KHdissociation constant for ligand H from high-affinity site on BPAO
KLdissociation constant for ligand L from low-affinity site on BPAO
KSdissociation constant for substrate S from active site of BPAO
Limidazoline drug binding to low-affinity site on BPAO
MAmethylamine
MOXmoxonidine
Pproduct (of an enzyme reaction)
PEAβ-phenylethylamine
PKDAOporcine kidney diamine oxidase
Ssubstrate
SPDspermidine
SSAOsemicarbazide-sensitive amine oxidase
TYRp-tyramine
venzyme reaction velocity
Vmaxenzyme reaction velocity at saturating substrate concentration

References
  • ALEMANY, R.; OLMOS, G.; GARCIA-SEVILLA, J.A. The effects of phenelzine and other monoamine oxidase inhibitor antidepressants on brain and liver I2 imidazoline-preferring receptors. Br. J. Pharmacol. 1995;114:837–845. [PubMed]
  • BINDA, C.; NEWTON-VINSON, P.; HUBALEK, F.; EDMONDSON, D.E.; MATTEVI, A. Structure of human monoamine oxidase B, a drug target for the treatment of neurological disorders. Nat. Struct. Biol. 2001;26:22–26.
  • BOUSQUET, P.; FELDMAN, J. Drugs acting on imidazoline receptors: a review of their pharmacology, their use in blood pressure control and their potential interest in cardioprotection. Drugs. 1999;58:799–812. [PubMed]
  • BOUSQUET, P.; FELDMAN, J.; SCHWARTZ, J. Central cardiovascular effects of alpha adrenergic drugs: differences between catecholamines and imidazolines. J. Pharmacol. Exp. Ther. 1984;230:232–236. [PubMed]
  • CALLINGHAM, B.A.; HOLT, A.; ELLIOTT, J. Properties and functions of the semicarbazide-sensitive amine oxidases. Biochem. Soc. Trans. 1991;19:228–233. [PubMed]
  • CARPÉNÉ, C.; COLLON, P.; REMAURY, A.; CORDI, A.; HUDSON, A.; NUTT, D.; LAFONTAN, M. Inhibition of amine oxidase activity by derivatives that recognize imidazoline I2 sites. J. Pharmacol. Exp. Ther. 1995;272:681–688. [PubMed]
  • DALFÓ, E.; HERNANDEZ, M.; LIZCANO, J.M.; TIPTON, K.F.; UNZETA, M. Activation of human lung semicarbazide-sensitive amine oxidase by a low molecular weight component present in human plasma. Biochim. Biophys. Acta. 2003;1638:278–286. [PubMed]
  • EDWARDS, L.; FISHMAN, D.; HOROWITZ, P.; BOURBON, N.; KESTER, M.; ERNSBERGER, P. The I1-imidazoline receptor in PC12 pheochromocytoma cells activates protein kinases C, extracellular signal-regulated kinase (ERK) and c- jun N-terminal kinase (JNK). J. Neurochem. 2001;79:931–940. [PubMed]
  • EGLEN, R.M.; HUDSON, A.L.; KENDALL, D.A.; NUTT, D.J.; MORGAN, N.G.; WILSON, V.G.; DILLON, M.P. ‘Seeing through a glass darkly': casting light on imidazoline ‘I' sites. Trends Pharmacol. Sci. 1998;19:381–390. [PubMed]
  • ELLIOTT, J.; CALLINGHAM, B.A.; SHARMAN, D.F. Amine oxidase enzymes of sheep blood vessels and blood plasma: a comparison of their properties. Comp. Biochem. Physiol. 1992;102C:83–89.
  • FRÉBORT, I.; TOYAMA, H.; MATSUSHITA, K.; ADACHI, O. Half-site reactivity with p-nitrophenylhydrazine and subunit separation of the dimeric copper-containing amine oxidase from Aspergillus niger. Biochem. Mol. Biol. Int. 1995;36:1207–1216. [PubMed]
  • GRANT, K.L.; KLINMAN, J.P. Evidence that both protium and deuterium undergo significant tunneling in the reaction catalysed by bovine serum amine oxidase. Biochemistry. 1989;28:6597–6605. [PubMed]
  • HEAD, G.A. Agmatine and imidazoline systems – fourth international symposium: 9–11 April 2003, San Diego, CA, USA. IDrugs. 2003;6:544–547. [PubMed]
  • HOLT, A.; ALTON, G.; SCAMAN, C.; LOPPNOW, G.R.; SZPACENKO, A.; SVENDSEN, I.; PALCIC, M.M. Identification of the quinone cofactor in mammalian semicarbazide-sensitive amine oxidase. Biochemistry. 1998;37:4946–4957. [PubMed]
  • HOLT, A.; BAKER, G.B. Metabolism of agmatine (clonidine-displacing substance) by diamine oxidase and the possible implications for studies of imidazoline receptors. Prog. Brain Res. 1995;106:187–197. [PubMed]
  • HOLT, A.; SHARMAN, D.F.; BAKER, G.B.; PALCIC, M.M. A continuous spectrophotometric assay for monoamine oxidase and related enzymes in tissue homogenates. Anal. Biochem. 1997;244:384–392. [PubMed]
  • HOLT, A.; TODD, K.G.; BAKER, G.B. The effects of chronic administration of inhibitors of flavin and quinone amine oxidases on imidazoline I1 receptor density in rat whole brain. Ann. NY Acad. Sci. 2003;1009:309–322.
  • JANES, S.M.; MU, D.; WEMMER, D.; SMITH, A.J.; KAUR, S.; MALTBY, D.; BURLINGAME, A.L.; KLINMAN, J.P. A new redox cofactor in eukaryotic enzymes: 6-hydroxydopa at the active site of bovine serum amine oxidase. Science. 1990;248:981–987. [PubMed]
  • KELLY, I.D.; KNOWLES, P.F.; YADAV, K.D.; BARDSLEY, W.G.; LEFF, P.; WAIGHT, R.D. Steady-state kinetic studies on benzylamine oxidase from pig plasma. Eur. J. Biochem. 1981;114:133–138. [PubMed]
  • KLINMAN, J.P. The multi-functional topa-quinone copper amine oxidases. Biochim. Biophys. Acta. 2003;1647:131–137. [PubMed]
  • KOHEN, A.; KLINMAN, J.P. Hydrogen tunneling in biology. Chem. Biol. 1999;6:R191–R198. [PubMed]
  • KOSKINEN, K.; VAINIO, P.J.; SMITH, D.J.; PIHLAVISTO, M.; YLA-HERTTUALA, S.; JALKANEN, S.; SALMI, M. Granulocyte transmigration through endothelium is regulated by the oxidase activity of vascular adhesion protein-1 (VAP-1). Blood. 2004;103:3388–3395. [PubMed]
  • KUMAR, V.; DOOLEY, D.M.; FREEMAN, H.C.; GUSS, J.M.; HARVEY, I.; MCGUIRL, M.A.; WILCE, M.C.J.; ZUBAK, V.M. Crystal structure of a eukaryotic (pea seedling) copper-containing amine oxidase at 2.2 Å resolution. Structure. 1996;4:943–955. [PubMed]
  • LYLES, G.A. The interaction of semicarbazide-sensitive amine oxidase with MAO inhibitors Monoamine Oxidase and Disease 1984. London: Academic Press547–556.ed. Tipton, K.F. pp.
  • LYLES, G.A. Mammalian plasma and tissue-bound semicarbazide-sensitive amine oxidases: biological, pharmacological and toxicological aspects. Int. J. Biochem. Cell Biol. 1996;28:259–274. [PubMed]
  • MACRAE, I.J.; HANNA, E.; HO, J.D.; FISHER, A.J.; SEGEL, I.H. Induction of positive cooperativity by amino acid replacements within the C-terminal domain of Penicillium chrysogenum ATP sulfurylase. J. Biol. Chem. 2000;275:36303–36310. [PubMed]
  • MICHEL, M.C.; ERNSBERGER, P. Keeping an eye on the I site: imidazoline-preferring receptors. Trends Pharmacol. Sci. 1992;13:369–370. [PubMed]
  • MOLDERINGS, G.J. Imidazoline receptors; basic knowledge, recent advances and future prospects for therapy and diagnosis. Drugs Future. 1997;22:757–772.
  • MORGAN, N.G.; CHAN, S.L.F. Imidazoline binding sites in the endocrine pancreas: can they fulfil their potential as targets for the development of new insulin secretagogues. Curr. Pharm. Des. 2001;7:1413–1431. [PubMed]
  • MORPURGO, L.; AGOSTINELLI, E.; MONDOVI, B.; AVIGLIANO, L.; SILVESTRI, R.; STEFANCICH, G.; ARTICO, M. Bovine serum amine oxidase: half-site reactivity with phenylhydrazine, semicarbazide, and aromatic hydrazides. Biochemistry. 1992;31:2615–2621. [PubMed]
  • MU, D.; MEDZIHRADSZKY, K.F.; ADAMS, G.W.; MAYER, P.; HINES, W.M.; BURLINGAME, A.L.; SMITH, A.J.; CAI, D.; KLINMAN, J.P. Primary structures for a mammalian cellular and serum copper amine oxidase. J. Biol. Chem. 1994;269:9926–9932. [PubMed]
  • MURE, M.; MILLS, S.A.; KLINMAN, J.P. Catalytic mechanism of the topa quinone containing copper amine oxidases. Biochemistry. 2002;41:9269–9278. [PubMed]
  • OLMOS, G.; GABILONDO, A.M.; MIRALLES, A.; ESCRIBA, P.V.; GARCIA-SEVILLA, J.A. Chronic treatment with the monoamine oxidase inhibitors clorgyline and pargyline down-regulates non-adrenoceptor [3H]-idazoxan binding sites in the rat brain. Br. J. Pharmacol. 1993;108:597–603. [PubMed]
  • OZAITA, A.; OLMOS, G.; BORONAT, M.A.; LIZCANO, J.M.; UNZETA, M.; GARCIA-SEVILLA, J.A. Inhibition of monoamine oxidase A and B activities by imidazol(ine)/guanidine drugs, nature of the interaction and distinction from I2-imidazoline receptors in rat liver. Br. J. Pharmacol. 1997;121:901–912. [PubMed]
  • PARSONS, M.R.; CONVERY, M.A.; WILMOT, C.M.; YADAV, K.D.; BLAKELEY, V.; CORNER, A.S.; PHILLIPS, S.E.; MCPHERSON, M.J.; KNOWLES, P.F. Crystal structure of a quinoenzyme: copper amine oxidase of Escherichia coli at 2 Å resolution. Structure. 1995;3:1127–1129. [PubMed]
  • PAZ, M.A.; FLUCKIGER, R.; BOAK, A.; KAGAN, H.M.; GALLOP, P.M. Specific detection of quinoproteins by redox-cycling staining. J. Biol. Chem. 1991;266:689–692. [PubMed]
  • RAASCH, W.; BARTELS, T.; GIESELBERG, A.; DENDORFER, A.; DOMINIAK, P. Angiotensin 1-converting enzyme inhibition increases cardiac catecholamine content and reduces monoamine oxidase activity via an angiotensin type 1 receptor-mediated mechanism. J. Pharmacol. Exp. Ther. 2002;300:428–434. [PubMed]
  • RAASCH, W.; MUHLE, H.; DOMINIAK, P. Modulation of MAO activity by imidazoline and guanidine derivatives. Ann. NY Acad. Sci. 1999;881:313–331.
  • RADDATZ, R.; PARINI, A.; LANIER, S.M. Localization of the imidazoline binding domain on monoamine oxidase B. Mol. Pharmacol. 1997;52:549–553. [PubMed]
  • RADDATZ, R.; SAVIC, S.L.; BAKTHAVACHALAM, V.; LESNICK, J.; JASPER, J.R.; MCGRATH, C.R.; PARINI, A.; LANIER, S.M. Imidazoline-binding domains on monoamine oxidase B and subpopulations of enzyme. J. Pharmacol. Exp. Ther. 2000;292:1135–1145. [PubMed]
  • RADDATZ, R.; SAVIC, S.L.; LANIER, S.M. Imidazoline binding domains on MAO-B. Localization and accessibility. Ann. NY Acad. Sci. 1999;881:26–31.
  • REMAURY, A.; RADDATZ, R.; ORDENER, C.; SAVIC, S.; SHIH, J.C.; CHEN, K.; SEIF, I.; DE MAEYER, E.; LANIER, S.M.; PARINI, A. Analysis of the pharmacological and molecular heterogeneity of I(2)-imidazoline-binding proteins using monoamine oxidase-deficient mouse models. Mol. Pharmacol. 2000;58:1085–1090. [PubMed]
  • SEGEL, I.H. Enzyme Kinetics. Behavior and Analysis of Rapid Equilibrium and Steady-State Enzyme Systems. New York: John Wiley & Sons Inc; 1993.
  • TAKADA, K.; HAYASHI, Y.; KAMIBAYASHI, T.; MAMMOTO, T.; YAMATODANI, A.; KITAMURA, S.; YOSHIYA, I. The involvement of pertussis toxin-sensitive G proteins in the post receptor mechanism of central I1-imidazoline receptors. Br. J. Pharmacol. 1997;120:1575–1581. [PubMed]
  • TESSON, F.; LIMON-BOULEZ, I.; URBAN, P.; PUYPE, M.; VANDEKERCKHOVE, J.; COUPRY, I.; POMPON, D.; PARINI, A. Localization of I2-imidazoline binding sites on monoamine oxidases. J. Biol. Chem. 1995;270:9856–9861. [PubMed]
  • TESSON, F.; PRIP-BUUS, C.; LEMOINE, A.; PEGORIER, J.P.; PARINI, A. Subcellular distribution of imidazoline-guanidinium-receptive sites in human and rabbit liver. Major localization to the mitochondrial outer membrane. J. Biol. Chem. 1991;266:155–160. [PubMed]
  • WAHLUND, L.O.; SAAF, J.; ROSS, S.B.; WETTERBERG, L. Activation of monoamine oxidase by high molecular weight fractions of human plasma. Acta Physiol. Scand. 1984;120:337–341. [PubMed]
  • WIEST, S.A.; STEINBERG, M.I. Binding of [3H]2-(2-benzofuranyl)-2-imidazoline (BFI) to human brain: potentiation by tranylcypromine. Life Sci. 1997;60:605–615. [PubMed]
  • WILCE, M.C.J.; DOOLEY, D.M.; FREEMAN, H.C.; GUSS, J.M.; MATSUNAMI, H.; MCINTIRE, W.S.; RUGGIERO, C.E.; TANIZAWA, K.; YAMAGUCHI, K. Crystal structures of the copper-containing amine oxidase from Arthrobacter globiformis in the holo and apo forms: implications for the biogenesis of topaquinone. Biochemistry. 1997;36:16116–16133. [PubMed]
  • YADAV, K.D.; KNOWLES, P.F. A catalytic mechanism for benzylamine oxidase from pig plasma. Stopped-flow kinetic studies. Eur. J. Biochem. 1981;114:139–144. [PubMed]
  • YU, P.H.; BOULTON, A.A. Activation of platelet monoamine oxidase by plasma in the human. Life Sci. 1979;25:31–36. [PubMed]
  • YU, P.H.; WANG, M.; DENG, Y.L.; FAN, H.; SHIRA-BOCK, L. Involvement of semicarbazide-sensitive amine oxidase-mediated deamination in atherogenesis in KKAy diabetic mice fed with high cholesterol diet. Diabetologia. 2002;45:1255–1262. [PubMed]
  • YU, P.H.; WANG, M.; FAN, H.; DENG, Y.; GUBISNE-HABERLE, D. Involvement of SSAO-mediated deamination in adipose glucose transport and weight gain in obese diabetic KKAy mice. Am. J. Physiol. Endocrinol. Metab. 2004;286:E634–E641. [PubMed]
  • YU, P.H.; WRIGHT, S.; FAN, E.H.; LUN, Z.R.; GUBISNE-HARBERLE, D. Physiological and pathological implications of semicarbazide-sensitive amine oxidase. Biochim. Biophys. Acta. 2003;1647:193–199. [PubMed]