U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Probe Reports from the NIH Molecular Libraries Program [Internet]. Bethesda (MD): National Center for Biotechnology Information (US); 2010-.

Cover of Probe Reports from the NIH Molecular Libraries Program

Probe Reports from the NIH Molecular Libraries Program [Internet].

Show details

ML345, A Small-Molecule Inhibitor of the Insulin-Degrading Enzyme (IDE)

, , , , , , , , , , , and .

Author Information and Affiliations

Received: ; Last Update: May 13, 2014.

Insulin-degrading enzyme (IDE) is a thiol-sensitive zinc-metallopeptidase that is strongly implicated in the pathogenesis of multiple highly prevalent diseases, including type 2 diabetes and Alzheimer’s disease (AD). Because IDE is the principal insulin-degrading protease in vivo, IDE inhibitors should enhance insulin signaling and thus have efficacy in relevant animal models of diabetes and also in therapy. Despite decades of study, a strong need yet exists for the identification of potent, selective, in vivo stable small molecule experimental probes that inhibit IDE with significantly high potency and target selectivity. We herein describe an IDE inhibitor molecular probe, ML345, which targets a specific cysteine residue (Cys819) in IDE. ML345 arose from an ultra high-throughput screening (uHTS) campaign that was supplemented with medicinal chemistry SAR optimization and with biochemical mechanistic profiling efforts. The probe is distinguished from prior art inhibitors, such as the potent and selective (but peptide-derived) hydroxamic acids and from weaker and far less selective small molecules that have been described in the literature. Probe ML345 is well-suited for use as a pharmacophore for drug development in diabetes research and as an experimental probe to understand the array of effects displayed by IDE in biological systems.

Assigned Assay Grant #: 1 R03 DA024888-01 (MLSCN cycle 6)

Screening Center Name & PI: The Scripps Research Institute Molecular Screening Center (SRIMSC), H Rosen

Chemistry Center Name & PI: SRIMSC, H Rosen

Assay Submitter & Institution: M Leissring, Mayo Clinic Jacksonville

PubChem Summary Bioassay Identifier (AID): 434984

Probe Structure & Characteristics

ML345.

ML345

CID57390068

SID144241486

SR-03000002959-2

MW 479.5

cLogP 2.0

tPSA 113

H-bond donors: 0

H-bond acceptors: 7

Ro5 compliant

CID/ML#Target NameEC50 [SID, AID]Anti-target(s)IC50 [SID, AID]Fold SelectiveSecondary Assay(s): IC50 (nM) [SID, AID]
CID 57390068
SID 136920229 synthesized (small scale)

SID 144241486 synthesized (larger scale)
ML345
IDEIC50 = 188 nM

[SID 136920229, AID 624065]
HEK cytotoxicityEC50 >10 μM (inactive)

[SID136920229, AID 588709]
>50IC50 >100 μM, cysteine null mutant IDE enzyme [SID 136920229, AID 624338]

Activity-based protein profiling analysis: minimal effects vs. the cysteine-reactive proteome @ 2 μM [SID 144241486, AID pending]

Batch data, IDE Inhibitor Probe

CIDSIDsourcePurity
57390068136920229Scripps synthesis>98%, (HPLC, MS, NMR)
57390068144241486Scripps synthesis>98% (HPLC, MS, NMR)
Probe ML345 chemical descriptors and highlighted properties.

Probe ML345 chemical descriptors and highlighted properties

Physical Properties (measured at the SRIMSC):

Stability (t1/2) in PBS: >>48 hours

t1/2, liver microsome stability:

>120 min (human), 24 min (mouse), 50 min (rat)

CYP450 inhib. 81%, 75%, 57%, 90%

@10 μM (1A2, 2D6, 3A4, 2C9)

Solubility in PBS = 0.4 μM

Solubility in simulated general assay buffer = 5.1 μM

Other attributes of ML345:

The probe is readily synthesized,

rule-of-five compliant,

target-selective (as determined by activity-based protein profiling),

has a defined mechanism of action,

has IDE activity that greatly exceeds the small molecule prior art.

Recommendations for scientific use of the probe: The IDE inhibitor probe ML345 will be used to understand how IDE processes its natural substrates, including insulin and beta-amyloid, defining the effects of disrupting IDE’s actions. The probe and derived follow-up compounds may show efficacy in models of insulin homeostasis, both in cell culture and in animal models. It is anticipated that IDE inhibitors can enhance insulin signaling, perhaps in synergy with agents such as GLP-1 agonists and DPPIV inhibitors, thus providing an opportunity for a significant advance in anti-diabetes therapy.

1. Introduction

Overview. Insulin-degrading enzyme (IDE) is a thiol-sensitive zinc-metallopeptidase that is strongly implicated in the pathogenesis of multiple highly prevalent diseases, including type 2 diabetes[1,2], Alzheimer’s disease (AD)[2,3], and Varicella-Zoster virus (VZV) infection[4]. IDE is the primary protease responsible for insulin degradation in vivo[1,2], mediating the critical, final step of insulin action, the termination of the insulin response[5]. IDE inhibitors thus should enhance insulin signaling and find use in treating diabetes[5]. They may also aid wound healing[6], among other medicinal benefits.

The IDE superfamily of zinc-metalloproteases evolved separately from the widely-studied “conventional” zinc-metalloproteases, a fascinating example of convergent evolution[7]. Members of this superfamily (clan M16) are often referred to as “inverzincins” because they feature a characteristic inverted zinc-binding motif (HxxEH) relative to that found within most conventional zinc-metalloproteases (HExxH)[8]. This structural change alters function and may further explain why conventional zinc-metalloproteases have been targeted for small molecule inhibition often and frequently with success, while no drug-like inhibitors of IDE (or any other inverzincin family member, for that matter) are known.

IDE was discovered in 1949 by I. Arthur Mirsky, who recognized that blocking IDE-mediated insulin catabolism was desirable for promoting endogenous insulin signaling and, thereby, was a promising target for the treatment of type 2 diabetes[9]. Mirsky showed that a purified, nonproteinaceous endogenous (but unknown) inhibitor of IDE could potentiate the hypoglycemic action of insulin in rabbits[10]. These and other results prompted strong interest in the development of orally bioavailable inhibitors of IDE. Indeed, several antidiabetic drugs emerged in the late 1950s that were initially believed to act by selective inhibition of IDE, though they were later found to act by other mechanisms[1114]. Truly specific and drug-like IDE inhibitors have remained elusive.

IDE is subject to inhibition by chelation of the key active site zinc atom, though an inherent concern of this approach is the difficult task of achieving selectivity for IDE over the wide array of other zinc-dependent metalloproteases. Another concern is that animal toxicity is often associated with the incorporation of potent zinc-binding motifs, such as hydroxamic acids.

The Leissring group (the assay provider for this effort) has shown that IDE’s actions are dependent upon two specific cysteine residues, C812 and C819[15], suggesting that thiol-modifying agents may also be IDE inhibitors. Such an approach raises the concern, however, about the potentially difficult task of attaining selectivity for IDE over the many other enzymes bearing reactive cysteine residues. IDE inhibitors that bind to allosteric sites of IDE may, in principle, also be used, though no allosteric inhibition sites of IDE have yet been identified.

IDE’s unique clamshell-like macromolecular structure[15,16] and its sensitivity both to zinc chelation and to thiol modification may help explain many of its unusual structural dynamics, extraordinary pharmacological properties, and the failure of previous efforts to find small molecule modulators of its activity. The Leissring group has aimed to study and to exploit these critical features for therapeutic means, pursuing the development of IDE inhibitors (and also activators) for more than a decade[5,1719]. Using a rational design approach, the Leissring group succeeded in developing the first potent (Ki =1.7 nM) and selective (10,000-fold selectivity vs. other zinc-metalloproteases) IDE inhibitor, a peptide-derived hydroxamic acid known as “Ii1” (IDE inhibitor 1)[5]. Despite its merits, Ii1 is not ideal as a chemical probe, especially for in vivo studies, owing to its high molecular weight (~750 Daltons) and its peptidic nature that renders it susceptible to rapid metabolism (t1/2 ~9 minutes in mice).

Relevance to Type 2 Diabetes. A natural substrate for IDE, insulin, possesses a distinctive tertiary structure, with two peptide chains linked by intra- and inter-chain disulfide bonds. Insulin is quite inert to processing by most common multifunctional proteases. IDE, however, efficiently degrades insulin and other peptide substrates such as glucagon, TGF alpha, and β-endorphin[8]. IDE knockout (IDE-KO) mice confirm the in vivo relevance of IDE in insulin catabolism, exhibiting fasting serum insulin levels ~3-fold higher than wild-type (WT) littermates[1,2]. Moreover, insulin degradation in IDE-KO tissue extracts was shown to be decreased by as much as 97%[2]. Because IDE is essentially the only protease that efficiently degrades insulin, inhibition of IDE is a straightforward way to augment insulin’s actions experimentally and, potentially, therapeutically.

There is also convincing evidence that IDE can regulate insulin action via degradation occurring downstream of insulin receptor (IR) binding[20]. For example, the Leissring group recently showed that that the activity of insulin pre-bound to its receptor is potentiated in the presence of IDE inhibitors[5], and Roth and colleagues found that the converse is true in cells over-expressing IDE21]. These and other results strongly suggest that IDE acts to regulate the termination of insulin action within cells, likely by modulating the off rate of insulin from its receptor[5]. Effective IDE inhibitors that are sufficiently stable for long-term use in cultured cells and in vivo will constitute essential tools for clarifying how—and where—IDE regulates the termination of insulin action.

Besides the expected elevated fasting serum insulin levels and accelerated glucose clearance of IDE-KO mice, these animals also show pronounced compensatory changes that are perhaps not linked to effects that would be expected from pharmacological IDE inhibition, such as chronic hyperinsulinemia[1,2]. The complete systemic ablation of IDE in the knockout state, levels far lower than what might ideally be achieved pharmacologically, may lead to hyperinsulinemia and to other compensatory effects such as pronounced glucose and insulin intolerance[1]. Such effects complicate the elucidation of IDE’s normal role in insulin signaling and in the pathogenesis of diabetes (as well as in other disease pathologies). Without potent and selective small molecule modulators of IDE activity available, it has been difficult to determine if such effects are artifacts that are not entirely relevant to possible IDE inhibitor therapy. We propose that such effects need not arise via short-term pharmacological inhibition of IDE, using compounds having a rapid onset of action, in adult individuals, with an ability to inhibit IDE partially and/or transiently. IDE inhibitors may also affect different pools of IDE (e.g.; intracellular vs. extracellular). Suitable tool compounds to cleanly assess the physiological and pathophysiological role of IDE proteolytic activity in vivo are needed. Importantly, as we will show, suitable molecular probes to answer these important questions have arisen from our probe development effort.

Other roles of IDE. A recent study[4] identified IDE as the cellular receptor for Varicella-Zoster virus (VZV) infection and invasion. Recent analysis suggests that infectivity is linked to IDE activity[15], thus IDE inhibitors might be useful against VZV infection.

IDE also plays an important role in Alzheimer’s disease (AD) pathogenesis, as the principal extracellular protease responsible for the degradation of Aß, an amyloidogenic peptide that accumulates in the brains of AD patients[22]. IDE-KO mice show a ~95% reduction in the rate of degradation of physiological levels of Aß applied exogenously[2]. IDE-KO mice also harbor significant increases in steady-state levels of endogenous cerebral Aß[2]. Using a transgenic approach, the Leissring group showed that a doubling of IDE levels expressed in neurons dramatically reduced steady-state brain Aß levels, amyloid plaque burden, and downstream cytopathology in a mouse model of AD[3]. Literature evidence strongly supports a link between IDE genetic variation and the incidence/onset of AD in man[23].

Because of expected Aß effects, IDE inhibitors used in diabetes therapy should not be highly brain-exposed. IDE inhibitors administered to affected brain tissue in animal models may, however, manipulate IDE activity and permit a more clearly assessment of its role in AD progression. Evidence of IDE inhibitory effects would support efforts to thwart the progression of AD by the use of IDE activators, which are the subject of another ongoing MLPCN probe development effort in our labs.

Structural Aspects. IDE contains an unusually large internal cavity, uniquely capable of accommodating insulin. Crystal structures of human IDE in co-complex with four substrates were reported in 2006[16]. The Leissring group has also solved IDE-inhibitor co-crystal structures (see PDB Code 3E4A) and found that IDE possesses a number of unusual structural features, including that it must undergo large conformational changes in its catalytic cycle. IDE’s unique structure is the key to understanding why IDE inhibitors have remained elusive and also in designing effective new classes of IDE inhibitors.

The overall structure of IDE resembles a clamshell, with two bowl-shaped halves connected by a flexible linker (see Figure 1 for a schematic diagram highlighting the Zn atom in the active site, and relevant cysteine residues). IDE encapsulates an unusually large (~13,000 Å3) internal chamber. The co-crystal structure of insulin in complex with IDE[24] revealed that the insulin molecule fits neatly within this cavity, explaining IDE’s unique ability to accommodate bulky substrates that make extensive contacts not just with the active site but also with multiple contact points on the surface of IDE’s inner chamber[16]. This unusual feature explains IDE’s exquisite substrate specificity, wherein it is the tertiary rather than the primary protein structure of substrates that best determines fit, and also rationalizes IDE’s low primary cleavage-site specificity, since interactions distal to the active site will determine the peptide bond that is accessible to the catalytic zinc atom for coordination and eventual cleavage.

slide1

Figure 1

IDE schematic structure.

Comparison of Tang et al.’s X-Ray crystal structures of IDE[16] to the structure for pitrilysin, a bacterial IDE homolog, suggests that IDE can switch between its “open” and “closed” states[8] (Figure 2). Tang et al. also showed that the activity of IDE can be controlled by engineering a disulfide bridge connecting the two halves, preventing the protease from opening[16]. Enzymatic analysis of these mutants showed them to be inactive under oxidizing conditions yet active under reducing conditions, providing direct evidence that the protease must switch between open and closed states to complete a catalytic cycle[16].

Fig-2-flat

Figure 2

IDE closed state, and open form of Pitrilysin.

The finding that IDE is a thiol-sensitive protease distinguishes it from most conventional zinc-metalloproteases. Indeed, IDE was initially classified as a cysteine protease by the Enzyme Commission. The Leissring group elucidated the precise molecular basis underlying IDE’s thiol sensitivity[25] through a comprehensive analysis of the 13 cysteine residues in IDE. Cys819 and, to a lesser extent, Cys812, are the principal mediators of this effect. Somewhat counter-intuitively, the cysteine residues that mediate IDE’s thiol-sensitivity reside within the C-terminal half of the protease (IDE-C), ~700 amino acids distal to the zinc-binding motif within its N-terminal half (IDE-N) (His108-His112). These results illustrate that IDE’s active site is actually bipartite.

Zn site-acting IDE inhibitors act as latches, holding IDE in the closed conformation. While no suitable small molecules are effective IDE inhibitors, larger peptide hydroxamic acid inhibitors act at the zinc site, as shown by a co-crystal structure of IDE and inhibitor Ii1[5]. This peptide hydroxamic acid binds not only to residues within IDE-N, which contains the active-site Zn atom, but also to numerous residues within IDE-C that comprise the second half of IDE’s bipartite active site[5]. Ii1 thus spans the active site of the protease, latching it in the closed conformation. It is reasonable to conclude that early failures to identify effective IDE inhibitors through HTS efforts owe to the fact that the screening libraries used had only small molecules (molecular weight <500) incapable of spanning the active site in this manner, i.e., unable to form interactions with both halves of IDE’s bipartite active site simultaneously.

Our uHTS campaign was designed to identify IDE inhibitors that target any part of the protein, including the zinc site. Given that the MLSMR library is mostly comprised of small molecules, however, we anticipated that compounds binding by other modes were more likely to arise. Compounds showing IDE inhibition by acting at allosteric sites (though such sites are not yet known to exist for IDE) would be particularly interesting. Also of interest are small molecules that bind near to, or that even covalently modify, one of the crucial cysteine residues, particularly if they are specific for one cysteine residue of IDE and if, moreover, they are also IDE-selective. Allosteric and cysteine-binding modes for IDE inhibition might not require a large inhibitor to be present in order to achieve a high level of inhibition.

Our uHTS effort was complemented by a focused follow-up and SAR strategy and biochemical strategy to define mode of action, a coordinated effort that has successfully delivered the first potent and selective IDE inhibitor molecular probe, ML345.

2. Materials and Methods

Chemistry: All chemical reagents and solvents were acquired from commercial vendors. Reactions were monitored by LC/MS (Thermo/Finnegan LCQ Duo system with MS/MS capability). An Agilent 1200 analytical HPLC was used for quantitative purity assessment. Teledyne-Isco “combiflash” automated silica gel MPLC instruments were used for chromatographic purifications. A 400 Brüker MHz NMR instrument was used for NMR analysis.

Biology: All protocols are reported in the relevant PubChem AIDs, provided below.

Compound Properties: Solubility, stability, and glutathione reactivity analyses were conducted in accordance with NIH guidelines. CYP450 inhibition and microsome stability analyses were performed as previously described[26].

2.1. Assays

Table 1IDE Inhibitor PubChem AIDs (click on hyperlink for details)

Assay descriptions follow.

StageAssay Type# Tested# ActivePubChem AID
HTS
Scripps: MLSMR Liquids
Primary Assay (1X%INH)324,8581,316434962
Confirmation Assay (3X%INH)1,179598435028
HEK CYTOX Counterscreen (3X%INH)1,179334449730
Dose Response (3XIC50)12744463220
HEK CYTOX Dose Response Counterscreen (3XIC50)1271463221
SAR
Scripps: powders (Rounds 0 + 1)
Dose Response (3XIC50)327588712
HEK CYTOX Dose Response Counterscreen (3XCC50)322588709
Cell-free Wildtype IDE Assay (FP) (3XIC50)329588711
Artifact FL Profiling Counterscreen (3XIC50)321588718
SAR
Leissring Lab: powders (Rounds 1 + 2)
Insulin Degradation Assay (3XIC50) Round 1173624065
QFRET FRET1 Artifact Assay Round 15510624066
WT IDE Cell-free Assay (FP FABB) (3XIC50) Round 166624067
Mutant IDE Cell-free Assay (FRET1) (3XIC50) Round 1132624306
Mutant IDE Cell-free Assay (FRET1) (3XIC50) Round 2260624338
QFRET FRET1 Artifact Assay Round 2261624340
WT IDE Cell-free Assay (FP FABB) (3XIC50) Round 2263624353

Primary Assay, IDE Inhibition (AID 434962, AID 435028, AID 463220, AID 588712)

The purpose of this assay is to determine the ability of test compounds to inhibit endogenous cellular IDE activity by measuring shifts in the proportion of intact and cleaved forms of the IDE substrate Aβ. This fluorescence polarization (FP)-based assay employs a derivatized Abeta- peptide [Fluorescein-Abeta-(1–40)-Lys-Biotin (FAbetaB)]. Cleavage of FAbetaB by IDE (provided by live HEK cells in the reaction) separates the Fluorescein moiety from the biotinylated moiety. Subsequent addition of avidin to the reaction increases the effective molecular weight of intact, biotinylated FAbetaB substrate, slowing their rotation rate and reducing the degree of depolarization of plane polarized light. In contrast, cleaved FAbetaB substrate species have their Fluorescein moiety separated from the rest of the molecule, which has a low molecular weight, hence rotating rapidly and causing strong depolarization. Thus, the relative amounts of cleaved and intact forms of the FAbetaB substrate can be measured. As designed, compounds that act as IDE inhibitors will inhibit FAbeta-B cleavage, resulting in a population of large avidin-bound, slowly rotating FAbeta-B molecules. Compounds are tested in triplicate at a nominal concentration of 5.59 micromolar.

Secondary Assay #1, HEK Counterscreen (AID 449730, AID 463221, AID 588709)

The purpose of this assay is to determine whether compounds identified as active in the high throughput primary assay to identify inhibitors of insulin-degrading enzyme are nonselective due to HEK cytotoxicity. In this assay, HEK cells are incubated with test compounds, followed by determination of cell viability. The assay utilizes the CellTiter-Glo luminescent reagent to measure intracellular ATP in viable cells. Luciferase present in the reagent catalyzes the oxidation of beetle luciferin to oxyluciferin and light in the presence of cellular ATP. Well luminescence is directly proportional to ATP levels and cell viability. As designed, compounds that reduce cell viability will reduce ATP levels, luciferin oxidation and light production, resulting in decreased well luminescence. Compounds are tested in triplicate using a 10-point 1:3 dilution series starting at a nominal test concentration of 56 micromolar.

Secondary Assay #2, Biochemical Assay for IDE Inhibition (AID 688711, AID 624067)

The purpose of this biochemical assay is to determine whether compounds act directly on IDE. This assay uses the FP-format as used in the Primary Assay, but is cell-free and employs recombinant IDE. Briefly, a range of concentrations of candidate inhibitors were incubated with 2 nM recombinant IDE and 100 nM FabB, and hydrolysis assessed by FP. Compounds that effectively inhibit recombinant IDE will thereby be shown to have a definitive mechanism of action. Compounds that do not inhibit in this assay act by another mechanism. If such compounds are of interest, downstream experiments are to be conducted in the laboratory of the assay provider to elucidate mechanism of action.

Secondary Assay #3, Biochemical Assay for Inhibition of a cysteine-free IDE mutant (AID 624306, AID 624338)

The purpose of this assay is to identify compounds that inhibit IDE activity via thiol-alkylation. This assay also uses an FP-format, but is cell-free and employs a cysteine-free mutant of recombinant IDE. In this assay, compounds that inhibit IDE via thiol alkylation in the wild-type enzyme will be ineffective at inhibiting the cysteine free mutant, thereby identifying them as thiol-alkylating agents. Non-covalent probes are likely to be active in inhibiting cysteine-free IDE. Covalent-acting thiol-modifying compounds should be inactive in inhibiting cysteine-free IDE. Inhibitors of either type are of interest. Identification of mode of action will guide studies intended to define probe selectivity.

Secondary Assay #4, Fluorescence Artifact Assay (AID 588718, AID 624353)

The purpose of this assay is to determine whether powder samples of compounds identified as possible IDE inhibitor probe candidates are nonselective due to their optical activity or fluorescence properties. This fluorescence polarization (FP)-based assay employs a derivatized Abeta- peptide [Fluorescein-Abeta-(1–40)-Lys-Biotin (FAbetaB)]. Cleavage of FAbetaB by IDE separates the Fluorescein moiety from the biotinylated moiety. Subsequent addition of avidin to the reaction increases the effective molecular weight of intact, biotinylated FAbetaB substrate, slowing their rotation rate and reducing the degree of depolarization of plane polarized light. In contrast, cleaved FAbetaB substrate species have their Fluorescein moiety separated from the rest of the molecule, which has a low molecular weight, hence rotating rapidly and causing strong depolarization. Thus, the relative amounts of cleaved and intact forms of the FAbetaB substrate can be measured. As designed, compounds that act as IDE inhibitors will inhibit FAbetaB cleavage, resulting in a population of large avidin-bound, slowly rotating FAbetaB molecules. Quenchers and fluorescent compounds may interfere with the fluorescence polarization readout and appear as inhibiting IDE. In order to identify such assay artifact, tests compounds are added at the end of the reaction, just before measuring fluorescence polarization, so the calculated %inhibition is a result of the optical interference by the compound rather than its impact on IDE activity. Compounds are tested in triplicate using a 10-point 1:3 dilution series starting at a nominal test concentration of 55 μM. The probe was not active in this assay (IC50 value > 55 μM).

Secondary Assay #5, QFRET Fluorescence Artifact Assay (AID 624066, AID 624340)

The purpose of this biochemical assay is to determine whether powder samples of compounds identified as possible IDE inhibitor probe candidates are artifacts due to their interference with the FP assay format. This IDE activity assay differs from the primary FP-based HTS assay (AID 434962). In this assay, a fluorogenic IDE substrate (FRET1) is incubated with wild-type (WT) recombinant IDE in the presence of compounds in reaction buffer. The reaction is monitored by quantification of fluorescence intensity as a function of time. Percent activity is calculated relative to controls containing no experimental compound (100%) or no enzyme (0%). Compounds are tested in triplicate using a 10-point, semilog dilution series starting at a nominal test concentration of 100 μM.

Protocol Summary: Ten microliters of wildtype (WT) recombinant IDE enzyme (1 nM final, i.e., EC80) were dispensed into each well of 384-well microtiter plates, together with control wells loaded with buffer only (Low_Control). Next, 10 μL of test compounds in DMSO (1% final concentration) or DMSO only (High_Control) were added to the appropriate wells. The assay was started by dispensing 10 μL FRET1 substrate (1 μM final) in Buffer A [100 mM NaCl, 10 mM MgCl2, 50 mM HEPES, pH 7.4 supplemented with 0.1% bovine serum albumin (BSA)]. Activity was read continuously at 2 min intervals for 2 h at room temperature (22 C) using a Molecular Devices SpectraMAX 5e multilabel plate reader (excitation = 335 nm, emission = 385 nm).

For each test compound activity was determined from the rate of increase of fluorescence as a function of time relative to controls, using the linear portion of the progress curves (i.e., 0–30% of complete hydrolysis). For each test compound, percent activity relative to High_Control (100%) was plotted against compound concentration using Microsoft Excel, according to the following formula: %_Activity = 100 * ((Test_Compound − Median_Low_Control)/( Median_High_Control − Median_Low_Control)). Where:

  • Test_Compound indicates wells containing test compound
  • Low_Control indicates wells containing Buffer only
  • High_Control indicates wells containing DMSO only

The IC50 values were determined by fitting a sigmoidal curve to the data set using Prism 5.0 (GraphPad Software, Inc.) Technologies Inc), solving for the concentration corresponding to 50% activity. In cases where the highest concentration tested (i.e. 100 μM) did not result in greater than 50% inhibition, the IC50 value was determined manually as greater than 100 μM.

Compounds with an IC50 value greater than 10 μM were considered inactive. Compounds with an IC50 value equal to or less than 10 μM were considered active. Probe candidates should not exhibit activity in this artifact assay.

Secondary Assay #6, TR-FRET-based IDE activity assay (AID 624065)

The purpose of this biochemical assay is to determine whether powder samples of compounds identified as possible IDE inhibitor probe candidates can inhibit the ability of recombinant IDE to degrade insulin in vitro. This assay is a proximity-based immunoassay that uses two monoclonal antibodies, one labeled with Eu3+-Cryptate and one labeled with XL665 that recognize distinct epitopes on the insulin molecule. Compounds are tested in triplicate using a 10-point semi-log dilution series starting at a nominal test concentration of 100 μM (highest dose). Details of this assay are available at the manufacturer’s website: http://www.htrf.com/products/cns/insulin/

Ten microliters of WT recombinant IDE enzyme (1 nM final, i.e. EC80) were dispensed into each well of 384-well microtiter plates. Next, 10 μL of test compound in DMSO, low control (1% DMSO final concentration), or high control (1 μM IDE inhibitor Ii1) were added to the appropriate wells. To define background fluorescence (background control) a subset of wells were filled with 10 μL buffer only (no insulin). The assay was started by dispensing 10 μL of insulin (100 nM final) in Buffer A [100 mM NaCl, 10 mM MgCl2, 50 mM HEPES, pH 7.4 supplemented with 0.1% bovine serum albumin (BSA)]. Plates were then incubated for various lengths of time at room temp (22 C), then the reactions were terminated by addition of 5 μL of the broad-spectrum zinc-metalloprotease inhibitor 1,10 phenanthroline (2 mM final). TR-FRET was measured using a Molecular Devices SpectraMAX 5e multi-label plate reader (excitation = 337 nm, emission A = 665 nm and emission B: 620 nm, with a 400 ms delay time). Raw TR-FRET values (TRFraw) were determined from the following formula: TRFraw = Em665/Em620

Next, background-subtracted TR-FRET values (TRFsub) were calculated according to the following formula: TRFsub = TRFraw − Mean_Background_Control_TRFraw

Where: Mean_Background_Control_TRFraw is defined as the raw TR-FRET values obtained in wells containing no insulin. For each test compound, percent activity relative to Low_Control (100%) was plotted against compound concentration using Microsoft Excel, according to the following formula:

100*((Test_Compound − Median_Low_Control)/(Median_High_Control − Median_Low_Control))

Where:

  • Low_Control is defined as DMSO-treated wells only.
  • Test_Compound is defined as wells containing test compound.
  • High_Control is defined as wells containing reference IDE inhibitor Ii1.

Secondary Assay #7, Proteome-wide profiling of cysteine reactivity

The purpose of this assay is to determine if covalent, cysteine-reactive probes measurably affect other cysteine-reactive members of the proteome at doses ranging from 2 nM to 20 μM. The assay is an activity-based protein profiling assay for late-stage analysis of a potential IDE inhibitor probes. Briefly, test compounds are incubated for 30 min at 37 °C, then labeled for 30 minutes with 5 μM chloroacetamide-Rh at 25 °C giving results as shown in Section 3.6.

2.2. Probe Chemical Characterization

The chemical structure of the probe was verified by analysis of its 400 MHz 1H NMR spectra (Figure 3) obtained on a Brüker 400 MHz instrument and supported by analysis of a 19F NMR spectrum obtain on the same instrument. The chemical structure was also corroborated by its LC/MS molecular ion (calc for M+1: 480.1, found 479.9, using a Thermo/Finnegan LCQ Duo system). Purity was measured at >98% (LC/MS analysis, confirmed by analytical HLPC analysis; HPLC purity data is shown in Figure 4). HPLC data was obtained using an Agilent 1200 analytical HPLC with an Agilent Eclipse XDB-C18 column, 4.6×150mm. The HPLC solvents used were acetonitrile and water with 0.1% formic acid added to each mobile phase as the pH modifier.

Figure 3. 1H NMR spectrum.

Figure 3

1H NMR spectrum.

Figure 4. HPLC Spectrum of ML345 at various wavelengths; purity >98%.

Figure 4

HPLC Spectrum of ML345 at various wavelengths; purity >98%.

The solubility of the probe in PBS at pH 7.4 was determined to be moderately low (0.4 μM), however, the solubility in simulated general assay buffer (DMEM + FBS) was substantially higher (5.1 μM). Its solubility is fully adequate to provide the high potency seen in cell-based assays. The measured solubility suggests that, when appropriately formulated, the solubility of ML345 is adequate for broad use as a biological probe. The FBS effect could indicate high protein binding, which we did not directly measure.

The probe has a half-life of >>48 hours in PBS at room temperature when tested at 10 μM (Figure 5).

Figure 5. 48 hour stability study of ML345 in PBS.

Figure 5

48 hour stability study of ML345 in PBS.

Disappearance of the LC peak for the probe is altered slowly but significantly by the addition of excess (50 μM) glutathione, with ~25% erosion of the parent peak area after 3 hours and ~41% erosion after 6 hours, suggesting a half-life of ~7.5 hours with 10-fold excess glutathione present. This agrees with the expectation that, while ML345 can be made to react with free thiols, it has significant selectivity for IDE, and specifically for reacting with Cys819 of IDE. The probe is stable in DMSO solution at 23°C (no erosion of peak intensity after 7 days) and is also stable as a free base dry powder. It is also stable under assay conditions, as indicated by potency in cell-based assays that is independent of incubation time.

The following compounds have been submitted to the MLSMR (MLS numbers pending):

Table 2Probe ML345 and Analogs

DesignationStructureCIDSIDSR IDAmount submitted (mg)Date Submitted
Probe ML345
Image ml345fu3.jpg
57390068144241486SR-03000002959-22012/17/2012
Analog 1
Image ml345fu4.jpg
57390067144241485SR-03000002958-21512/17/2012
Analog 2
Image ml345fu5.jpg
5295225144241488SR-03000003056-11512/17/2012
Analog 3
Image ml345fu6.jpg
4317404144241489SR-03000003057-11512/17/2012
Analog 4
Image ml345fu7.jpg
5023882144241490SR-03000003058-11512/17/2012
Analog 5
Image ml345fu8.jpg
1510390144241491SR-03000003059-11512/17/2012

2.3. Probe Preparation

Overview. The probe ML345 was synthesized in a straightforward fashion in a convergent manner, in six steps overall, with the longest linear sequence being 4 steps, as summarized in Figure 6. The overall yield of the process (after preparative HLPC purification of the final product) is 26%. Analogs for SAR evaluation were prepared by similar methods.

Figure 6. Synthesis of probe ML345.

Figure 6

Synthesis of probe ML345.

Step 1

Image ml345fu9

A mixture of acid chloride 1 (2.36 g, 10 mmol) and triethylamine (5.06 g, 50.0 mmol) in CH2Cl2 (50 mL) was treated with dropwise with morpholine (2.61g, 30 mmol) at room temperature. After addition was complete the reaction was stirred at room temperature for 14 h, quenched with saturated NH4Cl, and extracted with ethyl acetate. The combined organic extracts were washed with brine and dried over Na2SO4. The solvent was removed and the residue was purified by flash column chromatography (Hexanes: ethyl acetate = 1:1, Rf = 0.15) to afford 3.261 g (91%) of compound 2 as a yellow solid. Calc’d for C14H19N3O6S: 357.1; found [M+H]+: 358.0; 1H NMR (400 MHz, CDCl3) δ (ppm) 3.03–3.06 (m, 4H), 3.21–3.23 (m, 4H), 3.76–3.79 (m, 4H), 3.87–3.89 (m, 4H), 7.18 (d, J=8.8 Hz, 1H), 7.79 (dd, J=2.0, 8.8 Hz, 1H), 8.17 (d, J=2.0 Hz, 1H).

Step 2

Image ml345fu10

To a solution of compound 2 (488 mg, 1.37 mmol) in THF (10 mL) and MeOH (10 mL) was added Pd/C (10%, 50 mg). The reaction mixture was then stirred under atmosphere of H2 for 5 h, filtered and concentrated to afford 378 mg (85%) of compound 3 as white solid. Calc’d for C14H21N3O4S: 327.1; found [M+H]+: 328.1.

Step 3

Image ml345fu11

A solution of Acid 4 (344 mg, 2.0 mmol) in MeOH (12 mL) was treated with NaOH (160 mg, 4.0 mmol). The reaction mixture was stirred at room temperature for 10 min. The solvent was removed to afford a solid which was dissolved in acetone (18 mL). Benzyl bromide (376 mg, 2.2 mmol) was added. The reaction was sonicated for 5 min and then stirred at 0 °C for 1 h. The precipitate was collected by vacuum filtration. The solid was dissolved in H2O and then treated with 1 N HCl. The precipitate was collected by vacuum filtration and dried in air, affording 462 mg (88%) of compound 5 as white solid. 1H NMR (400 MHz, acetone-d6) δ (ppm) 4.25 (s, 2H), 7.27–7.36 (m, 4H), 7.45–7.47 (m, 2H), 7.54–7.57 (m, 1H), 7.70–7.73 (m, 1H).

Steps 4 and 5

Image ml345fu12

A suspension of acid 5 (570 mg, 2.17 mmol) in CH2Cl2 (24 mL) was treated with (COCl)2 (441 mg, 3.48 mmol) and a drop of DMF at 0 °C under atmosphere of N2. The reaction was stirred at 0 °C for 30 min, and then room temperature for 3 h. The solvent was removed. The residue was dissolved in CH2Cl2 (24 mL). Compound 3 (781 mg, 2.39 mmol) was added and cooled to 0 °C. Diisopropylethylamine (841 mg, 6.51 mmol) was added dropwise. The reaction mixture was stirred at room temperature overnight, quenched with H2O, extracted with CH2Cl2. The combined organic extracts were washed with brine, dried over Na2SO4. The solvent was removed and the residue was purified by flash column (hexanes: ethyl acetate = 1:1, Rf = 0.20) to afford 1.04 g (84%) of compound 6 as a yellow solid. Calc’d for C28H30FN3O5S2: 571.1; found [M+H]+: 572.0;. 1H NMR (400 MHz, CDCl3) δ (ppm) 2.93–2.95 (m, 4H), 3.13–3.15 (m, 4H), 3.78–3.83 (m, 8H), 4.05 (s, 2H), 7.07–7.09 (m, 3H), 7.19–7.20 (m, 3H), 7.30–7.33 (m, 2H), 7.42–7.45 (m, 1H), 7.58–7.60 (m, 2H), 8.92 (br s, 1H); 19F NMR (400 MHz, CDCl3) δ (ppm) −111.6.

Step 6

Image ml345fu13

A mixture of [bis(trifluoroacetoxy)iodo]benzene (PIFA, 1.02 g, 2.36 mmol) and TFA (0.415 g, 3.64 mmol) in CH2Cl2 (18 mL) was cooled to 0 °C under atmosphere of N2. A solution of compound 6 (1.04 g, 1.82 mmol) in CH2Cl2 (35 mL) was added dropwise. The reaction mixture was stirred at 0 °C for 30 min, room temperature 30 min, and then refluxed for 40 h. The crude product was concentrated and the residue was purified by flash column chromatography (hexanes : ethyl acetate = a gradient of 1:1 to 1:4) and then by preparative HPLC to afford 343 mg (40%) of ML345 as a white solid. Calc’d for M+H = C21H23FN3O5S2: 480.1; found [M+H]+: 479.9; 1H NMR (400 MHz, CDCl3) δ (ppm) 2.98–3.01 (m, 8H), 3.67–3.71 (m, 8H), 7.11 (d, J=8.8 Hz, 1H), 7.36–7.41 (m, 1H), 7.48–7.52 (m, 1H), 7.63 (dd, J=2.4, 8.4 Hz, 1H), 7.70 (dd, J=2.4, 8.0 Hz, 1H), 7.84 (d, J=2.0 Hz, 1H); 19F NMR (400 MHz, CDCl3) δ (ppm) −115.4.

3. Results

3.1. Summary of Screening Results

An overall summary of the screening campaign in shown in Figure 7. Briefly, uHTS primary screening of ~325,000 compounds gave 1,316 primary IDE INH hits. Counterscreening for toxicity and for (artifact) fluorescence interference gave 139 hits of highest interest. 127 available compounds were profiled in dose-response format for cell-based IDE activity and a total of 43 non-toxic compounds were then considered for follow-up testing of independent, freshly prepared powder samples by the SRIMSC. All results are available in PubChem using the indicated AIDs.

Figure 7. Screening campaign overview.

Figure 7

Screening campaign overview.

The series of efforts used to select probe candidate scaffolds and ultimately to identify molecular probes is summarized in the general hit-to-probe optimization flow chart (Figure 8). Chemistry and cheminformatics-based analyses of the results of the μHTS campaign allow for reactive, non-selective, assay-promiscuous, and otherwise non-tractable hits to be removed from further consideration.

Figure 8. Probe development strategy, post-uHTS.

Figure 8

Probe development strategy, post-uHTS.

Following these analyses, the most tractable compounds are ordered through for confirmation screening of a second batch. Any non-repeating samples are ignored for further probe development, as are compounds lacking targeted activity in secondary assays. Confirmed hits then allow for selection of scaffolds for probe development, aided by computational analysis of active and inactive analogs from the uHTS effort, to gauge potential of scaffolds to have meaningful non-flat SAR. Iterative rounds of medicinal chemistry and SAR development, complemented by assessment of compound properties, aids in selection of a high-quality probe.

The most interesting IDE screening hits are shown in Table 3 below. The screening hit selection criteria include structural desirability, high potency in IDE Inhibition dose-response (cell-based EC50 <10 μM), high maximal degree of inhibition (>75%), and lack of activity in the HEK toxicity and fluorescence artifact counterscreens (toxic and non-selective agents were excluded from probe development consideration).

Table 3. Hit summary, promiscuity analysis (reflects PubChem data available at time of uHTS).

Table 3

Hit summary, promiscuity analysis (reflects PubChem data available at time of uHTS).

Each of the compounds shown in Table 3 also displayed IDE activity when an independent batch was assayed (these “powder confirmation” EC50 values were within 2-fold of the uHTS result). All new batches of compounds were also inactive in the HEK toxicity counterscreen and also in the fluorescence artifact counterscreens (EC50 or IC50 in each case >>10 μM).

Our initial focus was upon three scaffolds shown in the middle of Table 3:

  • triazoles (CIDs 667188 and 664941),
  • the thiazole amide CID 1483690, and
  • the diacylpiperazine CID 17583130.

Analogs in each series are synthetically readily accessible. Further, literature analysis suggested that each series was tractable. We also preferred hits that were expected to be non-covalent inhibitors. Nine close structural analogs of the triazole CID667188 were prepared. However, seven of these analogs were found to be completely inactive vs. IDE, while the remaining two compounds were less potent than were the original hits. Even minor structural changes tended to eliminate all IDE activity. Poor water solubility was also a concern in the triazole series (e.g.; the cLogP for CID667188 is ~5.0). For these reasons, non-triazole probe development scaffolds were instead pursued. Seven structural analogs of the thiazole amide CID 1483690 were prepared and all were found to be completely inactive vs. IDE, prompting us to also discontinue efforts in this chemical series.

We then focused our attention upon the diacylpiperazine CID 17583130, initially encouraged by the simple tractable structure, IDE inhibitory activity of this compound, and by similar potency observed for several selected analogs that were purchased from a commercial vendor. However, when a sample of CID 17583130 was prepared internally (SID 124756577), this analytically pure sample was completely inactive vs. IDE (EC50 >100 μM). HPLC traces showed that the internally-generated sample was essentially identical in composition, at least by UV absorption, to both purchased samples of the “same” compound (SID 49825176 and SID 117689475). 21 structural analogs of CID 17583130, prepared internally in high purity, were also found to be inactive, though several purchased analogs were at least weakly active. When the purchased (previously active!) sample of CID 17583130 was rigorously purified by preparative HPLC (given SID 134420263 after purification), all IDE inhibitory activity was lost.

The above series of observations indicates that a non-organic and non-UV-active impurity is responsible for the IDE inhibition observed for batches of diacylpiperazines tested as purchased, without purification, samples that nevertheless had passed LCMS purity criteria owing to the absence of secondary peaks in the UV spectrum. We feel that this disappointing series of false positive results is likely due to the high sensitivity of the zinc or thiol moieties in IDE toward trace impurities that were present in the commercial samples as obtained from the vendor. Efforts to identify the active impurity or impurities have been unsuccessful. Thus the diacylpiperazine series was also eliminated from further consideration as a probe class.

The two hits shown at the bottom of Table 3, CID 4869155 and CID 2574532, were also eliminated from consideration for follow-up due to undesirable properties. Specifically, the high molecular weight and high LogP of CID 4869155, as well as the lack of active analogs in the uHTS collection, were concerns. The rhodanine-like hit CID 2574532 was also considered undesirable for chemical tractability reasons.

We expected that the four compounds shown at the top of Table 3, in a structural class termed the benzoisothiazolidones, were acting as covalent reversible modifiers of a cysteine residue at the active site of IDE through the formation of a disulfide bond. This feature distinguishes the class from the other entries in Table 3, which were expected to be non-covalent inhibitors, but that had, in every instance, proved ill-suited for probe development. This proposed mechanism of action for inhibition by the benzoisothiazolidones is illustrated in Figure 9.

Figure 9. Benzoisothiazolones target a cysteine residue of IDE.

Figure 9

Benzoisothiazolones target a cysteine residue of IDE.

Notably, the most potent HTS hit, CID 2325815 (top entry in Table 3) has a morpholine group present at the R2 position (see the highlighted R2 group depicted in Figure 9). Were these benzoisothiazolidone inhibitors merely acting as non-selective and chemically reactive “traps” for a cysteine residue, without significant cooperative binding to the enzyme, such a bulky and electron-donating R2 group would be expected to instead decrease the ability of the ligand in interact with IDE. Further, we found non-flat SAR found in this series, first as determined by cheminformatics analysis of uHTS results (percent IDE inhibition results for structural analogs of CID 2325815 that were present in the screening collection), and then verified by dose-response studies of several of these analogs (SAR data is shown later in section 3.4). These two lines of evidence suggested that there would be a high potential for finding a potent and selective IDE inhibitor probe in the benzoisothiazolidone scaffold.

When a freshly prepared second sample of CID 2325815 was obtained (SID 104144357) this analytically pure powder confirmation sample nearly duplicated the uHTS cell-based results, as shown in Figure 10, with EC50 ~2 μM (green curve, comparable with EC50 = 1.4 μM found for SID 24840227 in the uHTS campaign). Moreover, we saw promisingly high potency in an IDE biochemical assay (IC50 ~600 nM, black curve). Further, no HEK toxicity was seen at any dose (red curve) and no artifact fluorescence interference was seen below 10 μM (blue curve). This favorable profile supported the selection of the benzoisothiazolidones as a probe series.

Figure 10. Profile of most potent uHTS hit, CID 2325815.

Figure 10

Profile of most potent uHTS hit, CID 2325815. Data is for SID 104144357. Black = cell-free IDE assay

The use of covalent modifiers as molecular probes, and even as drugs, has been an area of considerable interest and substantial ongoing debate[27]. The stable covalent complex can confer favorable properties associated with long target residence time. Selectivity for the target in question (in this instance IDE) over other members of the proteome having reactive cysteine residues is of higher concern, however. A second concern is possible immunogenic effects of a covalently modified protein target, though this concern is lessened when the covalent modification is reversible, as in the case of benzoisothiazolidones. Our initial probe development plan expressed a preference for development a non-covalent probe, for the reasons listed above. In this case, however, we were thwarted by repeated difficulty in finding non-covalent hits that were not assay artifacts.

Keeping the concerns over covalent modifiers in mind, while also factoring in the potency, counterscreen selectivity, and high maximal inhibition seen in the benzothiazolidone class (see Figure 10 above) we chose this series for probe optimization. We ultimately identified even more potent IDE inhibitors (shown later in section 3.4, SAR table) and found a probe with suitable target selectivity and potency, the declared probe ML345, a fluorinated analog of the screening hit CID 2325815 (Figure 11).

Figure 11. ML345 is a fluorinated analog of CID 2325815.

Figure 11

ML345 is a fluorinated analog of CID 2325815.

3.2. Dose-Response Curves for Probe

We found that other benzoisothiazolones that are related to the screening hit CID 2325815 maintain the same general profile shown above in Figure 10: highest potency in the biochemical assay, followed by the cell-based IDE assay, with minimal autofluoresence at inhibitory levels, and no inherent toxicity in the HEK counterscreen.

The IDE inhibition curves are shifted to the left when the benzoisothiazolone ring bears an appropriately-positioned electron withdrawing group, consistent with the covalent reactivity of the probe as discussed earlier and depicted in Figure 9. Figure 12 below shows the inhibition curve for ML347 in the biochemical (cell-free) assay, using DTT-free, wild-type, recombinant IDE. The IC50 measured in the Leissring lab was 188 nM (SID 136920229). When a larger, equally pure probe batch was tested at a later date (SID 144241486), its IC50 was measured at 1.1 μM. This slight discrepancy may be due to a variation in activity of different batches of IDE used in each assays (positive controls were also higher in the latter run). The IDE enzyme is quite sensitive to handling and purification methods, showing significant batch-to-batch variability. The lower-end values for the probe (~200 nM) are more typical and (we feel) more accurate, particularly considering the positive control was on outlier in the experiment giving the higher IC50 value. Nevertheless, the probe is a potent IDE inhibitor in biochemical assays, is only active vs. native IDE and not a mutant form lacking Cys819, and is a significant advance in the field given the absence of small molecule prior art.

Figure 12. Probe dose-response curve, IDE inhibition (AID 624065, SID 136920229).

Figure 12

Probe dose-response curve, IDE inhibition (AID 624065, SID 136920229).

3.3. Scaffold/Moiety Chemical Liabilities

We have calculated various chemical descriptors for the probe using the Accelrys Pipeline Pilot software and by applying standard “rule-of-five” and other lead-likeness or drug-likeness criteria[2830]. As shown in Figure 13 below, all calculated parameters meet these criteria. The probe has a moderately cLogP and modestly high polar surface area, with suitable molecular weight, H-bond donor/acceptor counts, and with no significant structural-based concerns from a stability, lead-likeness, drug-likeness, or general toxicity alert[31] perspective.

Figure 13. Lipiniski and other drug-likeness parameters for ML345.

Figure 13

Lipiniski and other drug-likeness parameters for ML345.

The calculated partition coefficients shown in Figure 13 for ML345 (e.g., LogD7.4 = 2.0) suggests the potential for significant aqueous solubility that could aid its utility as a probe. As was described earlier, the solubility of the probe in PBS was somewhat lower than anticipated, 0.4 μM. The solubility in simulated general assay buffer was more reasonable (5.1 μM) suggesting that formulation strategies, with solubilizing additives, should be considered by researchers using the probe. Also, as shown earlier (Figure 5), the probe is chemically stable in a PBS solution: neither a substantial erosion of parent LCMS peak intensity nor an emergence of new peaks was seen over a 48-hour study.

The potential for biological stability was assessed by incubation of the probe with hepatic microsome preparations (1 mg/mL) using known methods[26]. The probe is highly stable to liver microsomes, with a measured half-life of >120 minutes (human liver microsomes), 24 minutes (mouse microsomes), and 50 min (rat microsomes). High hepatic microsome stability is consistent with attributes of a compound intended to be used in the study of relevant in vivo effects.

Inhibition of cytochrome P450 enzymes (CYPs) is a marker for drug-drug interaction potential. An in vitro study used four relevant CYP isoforms (1A2, 2C9, 2D6, and 3A4). The appropriate positive controls, compounds known to inhibit individual CYP isoforms (furafylline, sulfaphenazole, quinidine, and ketoconazole), were used.[26] The analysis, shown in Table 4, determined that ML345 inhibits these CYP isoenzymes between 57–94% at 10 μM. This suggests that ML345 requires significant improvements with regard to CYP450 inhibition profile for safe use therapeutically; however, its profile should not hinder the utility of ML345 as a molecular probe and biological tool compound to study the importance of IDE, especially since it is a first-in-class small molecule inhibitor of IDE.

Table 4. CYP450 Inhibition of Probe ML345.

Table 4

CYP450 Inhibition of Probe ML345.

The selection of a covalently-reactive probe series, as discussed previously, demanded that we clearly show that the probe does not behave simply as a broadly reactive and non-selective cysteine trap. We have compiled several lines of evidence, including non-flat SAR (shown later SAR discussed later in Section 3.4) and profiling assays (discussed later in Section 3.6) to show that the probe has suitable target selectivity to be useful as a molecular probe.

The probe scaffold is termed a benzoisothiazolidone. We are not aware of any undesirable attributes of the scaffold, other than potential reactivity with cysteine-containing proteins (an issued addressed in Section 3.6, assay profiling) and the CYP450 inhibition we have observed, which is likely compound-specific rather than scaffold-specific. While in vivo studies are beyond the scope of the probe development effort, the electron-deficient aryl groups present in ML345 would tend to protect it from susceptibility to rapid oxidative metabolism. The sulfonamide group of ML345 is not normally prone to redox issues. The aniline moiety present is tri-substituted, sterically hindered, and electronically deactivated, thus this group is also unlikely to contribute to potential metabolic liabilities for the probe.

The relative ease of synthesis of the probe molecule and analogs by a convergent six step sequence (4 linear steps) in a reasonable yield (see Figure 6 in Section 2.3) is a strong asset for this scaffold and for this probe. Scale-up synthesis, if needed for extensive long-term studies, will be highly feasible.

An MLPCN probe with the benzoisothiazolidone scaffold has already been approved. This compound, ML089, is a potent, selective, and orally bioavailable phosphomannose isomerase inhibitor (IC50 = 1.3 μM)[32]. This same compound (entry 3 in Table 5, shown later in Section 3.4) is also an IDE inhibitor, though it is ~20-fold less potent vs. IDE than is ML345. Very importantly, however, the reported highly favorable DMPK properties of ML089, including high oral bioavailability[32], support the assertion that analogs tailored for potent and selective IDE inhibition may also have similarly favorable DMPK properties.

Table 5. Abbreviated SAR summary in the probe scaffold showing cell-free (biochemical) results.

Table 5

Abbreviated SAR summary in the probe scaffold showing cell-free (biochemical) results.

3.4. SAR Tables

After the team selected CID 2835215 as a probe development scaffold, we investigated structure-activity relationships (SAR) of the series that were apparent from uHTS results, since several related compounds were in the MLSMR library and had been screened. We also analyzed SAR of purchased analogs as well as related inhibitors synthesized internally. IDE inhibitory activity was measured in triplicate, using the cell-free IDE inhibition assay, and selected results are shown above in Table 5 (values for three other analogs present in the uHTS are shown in Table 3 and are not duplicated here). The screening hit and the eventual probe ML345 are shown at the top (entry 1) and bottom (entry 7) of Table 5, respectively, for comparison.

Briefly, entry 1 (the uHTS hit) has suitable activity for SAR development (The Scripps IC50 value was ~600 nM, as shown earlier in Figure 10, with EC50 in the cell-based assay = 1.4 μM; values shown in Table 5 are from the cell-free assay, measured in the Leissring lab).

Entry 2 shows that deletion of both the morpholine and morpholine sulfonamide groups present in the uHTS hit diminishes IDE activity more than 10-fold, suggesting that both of these groups play a role in IDE recognition.

Entries 3 and 4 show that electron withdrawing groups in either aryl ring enhances IDE inhibition (compare with entry 2).

Entry 5 shows that a pyrrolidine ring is less well-tolerated than is the morpholine ring present in the uHTS hit, again suggesting that morpholine group plays a important role in IDE recognition.

Entry 6 shows the very weak activity of a prototypical non-selective covalent “cysteine trap” N-ethyl maleimide.

Finally, the probe compound ML345 is shown in entry 7, which, like entry 3, bears a para-fluorine substituent to further augment IDE inhibitory activity (as suggested by entry 3). IDE inhibition is also aided by the presence of the morpholine ring and the sulfonamide group (as suggested by entries 3 and 4).

We draw another important conclusion from analysis of Table 5: the SAR in the benzoisothiazolidone series is not flat, i.e.; all compounds that are capable of interacting with a cysteine residue do not have roughly equal activity.

It is also noteworthy that the ortho morpholine group present in entries 1 and 7 enhances activity, since both the electron-donating effect of this group, as well as the steric hindrance it imparts, would deactivate the sulfur atom in the benzoisothiazolidone ring for nucleophilic attack. This again indicates that the morpholine ring present in the probe plays a positive role in IDE recognition that overrides its deleterious effects upon ligand reactivity.

3.5. Cellular Activity

The uHTS primary assay is cell-based, so all hits (and the probe) demonstrate cellular activity. As shown in Figure 10 for the uHTS hit, cell-based EC50 is typically within 2- to 5-fold of biochemical IC50.

3.6. Profiling Assays

In order to ascertain the selectivity of ML345 for IDE over other enzymes in the proteome with reactive cysteine residues, a proteome-wide activity-based protein profiling (ABPP) assay was conducted in collaboration with Ben Cravatt and Andrea Zuhl at TSRI-La Jolla. These assays were performed with the general cysteine-reactive probe chloroacetamide-rhodamine (CA-Rh)[3335]. Briefly, the probe compound was incubated for 30 minutes at 37 °C in HEK293T proteomes then labeled for 30 minutes with 5 μM CA-Rh at 25 °C. Analysis by SDS-PAGE and in-gel fluorescence scanning gave the results shown below in Figure 14.

Figure 14. Proteome-wide activity-based protein profiling for cysteine reactivity of ML345.

Figure 14

Proteome-wide activity-based protein profiling for cysteine reactivity of ML345.

When ML345 was added to either membrane (A) or soluble (B) proteomes at concentrations of 2.0 μM or below, it did not inhibit labeling of cysteine-reactive proteins as determined by the retention of fluorescence signal compared to a DMSO control lane. For both proteome fractions, ML345 was significantly reactive at 20 μM, as indicated by bands that do not appear in that lane using the highest concentration of test compound. Overall these results show that the probe is not a broadly reactive “cysteine trap” and that it likely has on the order of at least 10-fold selectivity for IDE in a whole cell environment.

Ideally the proteome-wide profiling assay would be complemented by profiling assays, especially including thio-reactive enzymes. Such assays much be thiol-free to have meaningful results, however, and a suitable thiol-free “ready to go” panel was not available from screening vendors. We hope to gather profiling data against related targets in the future, however.

4. Discussion

The identified probe ML345 will serve as a valuable tool for investigators interested in exploring the endogenous function of IDE, particularly aspects of its role in diabetes and Alzheimer’s disease.

4.1. Comparison to existing art and how the new probe is an improvement

Prior art compounds that act as inhibitors of IDE are known, but none are both potent (IDE IC50 or EC50 <1 μM) and truly selective for IDE (Figure 15). These compounds include the cyclic peptide bacitracin (which has very low potency), chloroquine, 4-chloromercuribenzenesulfonate (aka PCMBS; CID11136; a highly cytotoxic compound)[36], and related p-chloromercuribenzoic acid (PCMBA). Weak and non-selective thiol-alkylating agents such as N-ethyl maleimide (NEM) and 2-iodoactetamide are known, as are weak and non-selective general zinc chelators such as 1,10-phenanthroline and EDTA[37]. IDE can also be inhibited in a dose-dependent manner with excess insulin[37,38]. Long-chain fatty acids and related acyl-CoA thioesters, such as linoleoyl-CoA and palmitoyl-CoA, act as IDE inhibitors with in vitro IC50 values ~10 μM and above[39]. ATP also inhibits IDE activity in vitro[37]. The protease inhibitor nelfinavir has been reported to inhibit insulin degradation mediated by partially purified IDE in cell-based assays; however, the IC50 was found to be greater than 100 μM[39].

Figure 15. Prior art compounds.

Figure 15

Prior art compounds. Exceptionally weak, non-selective, and non-drug-like prior art IDE inhibitors are not suitable starting points for small molecule probe development.

Recently, and subsequent to the initiation of this project, a peptide hydroxamic acid inhibitor of IDE that is highly potent (Ki for insulin ~23 nM) was developed by the assay submitter, Dr. Leissring[5]. However, this compound suffers from several limitations. First, due to its peptidic nature its utility in vivo is limited, which also limits its prospects for development into a more drug-like probe. Second, it is not a small molecule, with a molecular weight ~750 Daltons. Finally, the hydroxamic acid group present in this IDE inhibitor is a promiscuous zinc-binding moiety which will likely cross-react with other zinc-metalloproteases, thereby limiting its specificity.

From the above analysis, the potent and selective small molecule IDE inhibitor ML345 is well-differentiated from existing art and will be a significant advance in the field.

4.2. Mechanism of Action Studies

We made significant efforts to show precisely how benzoisothiazolidones, including the probe ML345, interact with IDE. Much of this work was performed using the uHTS hit, CID 2325815. This compound and its analogs similarly studied display activity in the cell-based IDE inhibition assays (AIDs 434962, 435028, 463220, and 588712) and also consistently have higher activity in related cell-free biochemical assays (AIDs 688711 and 624067). Typically a two- to ten-fold cell-based shift is seen, typical of the barrier to cell penetration seen for most effective on-target small molecules.

We were interested in testing the proposed covalent mechanism of action of this series of compounds (see Figure 9, in section 3.1). Toward this end, mutant forms of IDE that have alanine replacements for either Cys-819 or Cys-812 were prepared[25] and used in IDE inhibition assays. The benzoisothiazolidone screening hit CID 2325815, an inhibitor of the native IDE enzyme with an IC50 = 1.4 μM (Table 5) does not inhibit mutant IDE lacking Cys-819 (IC50 >100 μM), though this compound is an inhibitor of the mutant enzyme lacking Cys-812 but with Cys-819 intact (IC50 ~3 μM). This shows that benzoisothiazolidones are not indiscriminately reactive, since they can distinguish between these two cysteine residues of IDE, preferentially reacting at Cys-819 to inhibit the enzyme.

The covalent mechanism of action for ML345 was further confirmed by using electrospray ionization-mass-spectrometry (ESI-MS) to show that ML345 bonds covalently to a thiol-containing cysteine analog present in large excesss (N-acetylcysteine), yielding a product of the predicted mass (625.2 amu).

Finally, the inhibition of IDE by ML345 was shown to be reversible by addition of reducing agents, such as ß-mercaptoethanol or dithiothreitol (data not shown).

These experiments, along with non-flat SAR (Table 5 in Section 3.4), and combined with the results of the activity-based protein profiling experiment (Figure 14 in section 3.6) show that potent benzoisothiazolidones are direct, covalent, and reversible inhibitors of wild-type IDE that act preferentially at Cys-819 in the internal cavity of IDE.

4.3. Planned Future Studies

Professors Leissring, Bannister, and collaborators intend to extend and expand upon this work through future grant proposals. Such efforts, if funded, will aim to develop a diverse set of potent, selective, IDE inhibitors, including zinc-targeting IDE inhibitors, more potent cysteine-targeting IDE inhibitors, and compounds in each class that are particularly suited for animal use, while also further studying mechanistic aspects of IDE inhibition by these compounds.

4.4. Acknowledgement

We would like to gratefully recognize the assistance of Dr. Andrea Zuhl and Professor Benjamin Cravatt in obtaining the activity-based protein profiling data, for helpful discussions, and for preparing the relevant figure ABBP data (Figure 14 in Section 3.6).

5. References

1.
Abdul-Hay SO, Kang D, McBride M, Li L, Zhao J, Leissring MA. Deletion of insulin-degrading enzyme elicits antipodal, age-dependent effects on glucose and insulin tolerance. PLoS ONE. 2011;6(6):e20818. [PMC free article: PMC3111440] [PubMed: 21695259]
2.
Farris W, Mansourian S, Chang Y, Lindsley L, Eckman EA, Frosch MP, Eckman CB, Tanzi RE, Selkoe DJ, Guenette S. Insulin-degrading enzyme regulates the levels of insulin, amyloid beta-protein, and the beta-amyloid precursor protein intracellular domain in vivo. Proc Natl Acad Sci U S A. 2003;100(7):4162–7. [PMC free article: PMC153065] [PubMed: 12634421]
3.
Leissring MA, Farris W, Chang AY, Walsh DM, Wu X, Sun X, Frosch MP, Selkoe DJ. Enhanced proteolysis of beta-amyloid in APP transgenic mice prevents plaque formation, secondary pathology, and premature death. Neuron. 2003;40(6):1087–93. [PubMed: 14687544]
4.
Li Q, Ali MA, Cohen JI. Insulin degrading enzyme is a cellular receptor mediating varicella-zoster virus infection and cell-to-cell spread. Cell. 2006;127(2):305–16. [PMC free article: PMC7125743] [PubMed: 17055432]
5.
Leissring MA, Malito E, Hedouin S, Reinstatler L, Abdul-Hay SO, Choudhry S, Fauq AH, Huzarska M, May PS, Choi S, Logan TP, Turk BE, Cantley LC, Manolopoulou M, Tang WJ, Stein RL, Cuny GD, Selkoe DJ. Designed inhibitors of insulin-degrading enzyme regulate the catabolism and activity of insulin. PLoS One. 2010;5(5):e10504. [PMC free article: PMC2866327] [PubMed: 20498699]
6.
Shearer JD, Coulter CF, Engeland WC, Roth RA, Caldwell MD. Insulin is degraded extracellularly in wounds by insulin-degrading enzyme (EC 3.4.24.56). Am J Physiol. 1997;273(4 Pt 1):E657–64. [PubMed: 9357792]
7.
Makarova KS, Grishin NV. Thermolysin and mitochondrial processing peptidase: how far structure-functional convergence goes. Protein Sci. 1999;8(11):2537–40. [PMC free article: PMC2144204] [PubMed: 10595562]
8.
Becker AB, Roth RA. Insulysin and pitrilysin: insulin-degrading enzymes of mammals and bacteria. Methods Enzymol. 1995;248:693–703. [PubMed: 7674956]
9.
Mirsky IA, Broth-Kahn RH. The inactivation of insulin by tissue extracts. I. The distribution and properties of insulin inactivating extracts (insulinase). Arch Biochem. 1949;20:1–9. [PubMed: 18104389]
10.
Mirsky IA, Perisutti G. Effect of insulinase-inhibitor on hypoglycemic action of insulin. Science. 1955;122(3169):559–60. [PubMed: 13255892]
11.
Mirsky IA, Perisutti G, Diengott D. The inhibition of insulinase by hypoglycemic sulfonamides. Metabolism. 1956;5(2):156–61. [PubMed: 13296877]
12.
Mirsky IA, Perisutti G, Gitelson S. The role of insulinase in the hypoglycemic response to sulfonylureas. Ann N Y Acad Sci. 1957;71(1):103–11. [PubMed: 13459205]
13.
Marigo S, Panelli G. Insulinase and its inhibition by hypoglycemic sulfonamides; data on insulin sensitivity during tolbutamide therapy. Arch Sci Med (Torino). 1958;105(6):587–609. [PubMed: 13560210]
14.
Leites SM, Smirnov NP. Significance of insulin-inactivating properties of the liver (insulinase) in the mechanism of action of antidiabetic sulfonamides. Bull Eksp Biol Med. 1959;47(6):58–62. [PubMed: 13671070]
15.
Leissring MA, Selkoe DJ. Structural biology: enzyme target to latch on to. Nature. 2006;443(7113):761–2. [PubMed: 17051198]
16.
Shen Y, Joachimiak A, Rosner MR, Tang WJ. Structures of human insulin-degrading enzyme reveal a new substrate mechanism. Nature. 2006;443(7113):870–4. [PMC free article: PMC3366509] [PubMed: 17051221]
17.
Cabrol C, Huzarska MA, Dinolfo C, Rodriguez MC, Reinstatler L, Ni J, Yeh L-A, Cuny GD, Stein RL, Selkoe DJ, Leissring MA. Small-molecule activators of insulin-degrading enzyme discovered through high-throughput compound screening. PLoS ONE. 2009;4(4):e5274. [PMC free article: PMC2668070] [PubMed: 19384407]
18.
Leissring MA, Farris W, Turk BE, Cantley LC, Yeh L, Caclin C, Xing X, Cuny GD, Stein RL, Selkoe DJ. Small-molecule inhibitors of insulin-degrading enzyme. Society for Neuroscience Abstract Viewer/Itinerary Planner. 2004. Program No. 488.11(online)
19.
Leissring MA, Logan TP, Choi S, Cuny GD, Turk B, Cantley LC, Farris W, Selkoe DJ. Rationally designed peptide hydroxamate inhibitors of insulin-degrading enzyme. Society for Neuroscience Abstract Viewer/Itinerary Planner. 2005. Program No. 134.11(Online)
20.
Yonezawa K, Yokono K, Shii K, Hari J, Yaso S, Amano K, Sakamoto T, Kawase Y, Akiyama H, Nagata M, et al. Insulin-degrading enzyme is capable of degrading receptor-bound insulin. Biochem Biophys Res Commun. 1988;150(2):605–14. [PubMed: 3277630]
21.
Seta KA, Roth RA. Overexpression of insulin degrading enzyme: cellular localization and effects on insulin signaling. Biochem Biophys Res Commun. 1997;231(1):167–71. [PubMed: 9070242]
22.
Leissring MA. The AßCs of Aß-cleaving proteases. J Biol Chem. 2008;283(44):29645–9. [PMC free article: PMC2662051] [PubMed: 18723506]
23.
Turner AJ, Nalivaeva NN. New insights into the roles of metalloproteinases in neurodegeneration and neuroprotection. Int Rev Neurobiol. 2007;82:113–35. [PMC free article: PMC7112344] [PubMed: 17678958]
24.
Manolopoulou M, Guo Q, Malito E, Schilling AB, Tang WJ. Molecular basis of catalytic chamber-assisted unfolding and cleavage of human insulin by human insulin degrading enzyme. J Biol Chem. 2009;284(21):14177–88. [PMC free article: PMC2682866] [PubMed: 19321446]
25.
Neant-Fery M, Garcia-Ordonez RD, Logan TP, Selkoe DJ, Li L, Reinstatler L, Leissring MA. Molecular basis for the thiol sensitivity of insulin-degrading enzyme. Proc Natl Acad Sci U S A. 2008;105(28):9582–7. [PMC free article: PMC2474492] [PubMed: 18621727]
26.
Brothers SP, Saldanha SA, Spicer TP, Cameron M, Mercer BA, Chase P, McDonald P, Wahlestedt C, Hodder PS. Selective and brain penetrant neuropeptide Y Y2 receptor antagonists discovered by whole-cell high-throughput screening. Mol Pharmacol. 2010;77(1):46–57. [PMC free article: PMC2802430] [PubMed: 19837904]
27.
Singh J, Petter RC, Baillie TA, Whitty A. The resurgence of covalent drugs. Nature Reviews Drug Discovery. 2011;10:307–317. [PubMed: 21455239]
28.
Lipinski CA, Lombardo F, Dominy BW, Feeney PJ. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Advanced Drug Delivery Reviews. 2001;46:3–26. [PubMed: 11259830]
29.
Walters WP. Going further than Lipinski’s rule in drug design. Expert Opinion on Drug Discovery. 2012;7(2):99–107. [PubMed: 22468912]
30.
Rishton GM. Nonleadlikeness and leadlikeness in biochemical screening. Drug Discovery Today. 2003;8(2):86–96. [PubMed: 12565011]
31.
Benigni R, Bossa C. Mechanisms of Chemical Carcinogenicity and Mutagenicity: A Review with Implications for Predictive Toxicology. Chem Rev. 2011;111(4):2507–2536. [PubMed: 21265518]
32.
Dahl R, Bravo Y, Sharma V, Ichikawa M, Dhanya R-P, Hedrick M, Brown B, Rascon J, Vicchiarelli M, Mangravita-Novo A, Yang L, Stonich D, Su Y, Smith LH, Sergienko E, Freeze HH, Cosford NDP. Potent, Selective, and Orally Available Benzoisothiazolone Phosphomannose Isomerase Inhibitors as Probes for Congenital Disorder of Glycosylation Ia. J Med Chem. 2011;54:3661–3668. [PMC free article: PMC3437750] [PubMed: 21539312]
33.
Weerapana E, Simon GM, Cravatt BF. Disparate proteome reactivity profiles of carbon electrophiles. Nature Chemical Biology. 2008;4(7):405–407. [PMC free article: PMC2440582] [PubMed: 18488014]
34.
Bachovchin DA, Zuhl AM, Speers AE, Wolfe MR, Weerapana E, Brown SJ, Rosen H, Cravatt BF. Discovery and Optimization of Sulfonyl Acrylonitriles as Selective, Covalent Inhibitors of Protein Phosphatase Methylesterase-1. Journal of Medicinal Chemistry. 2011;54(14):5229–5236. [PMC free article: PMC3144155] [PubMed: 21639134]
35.
Weerapana E, Wang C, Simon GM, Richter F, Khare S. Dillon; MB, Bachovchin DA, Mowen K, Baker D, Cravatt BF, Quantitative reactivity profiling predicts functional cysteines in proteomes. Nature. 2010;468:790–795. [PMC free article: PMC3058684] [PubMed: 21085121]
36.
Duckworth WC, Hamel FG, Peavy DE. Two pathways for insulin metabolism in adipocytes. Biochim Biophys Acta. 1997;1358(2):163–71. [PubMed: 9332452]
37.
Del Carmen Camberos M, Cresto JC. Insulin-degrading enzyme hydrolyzes ATP. Exp Biol Med (Maywood). 2007;232(2):281–92. [PubMed: 17259336]
38.
Farris W, Mansourian S, Chang Y, Lindsley L, Eckman EA, Frosch MP, Eckman CB, Tanzi RE, Selkoe DJ, Guenette S. Insulin-degrading enzyme regulates the levels of insulin, amyloid beta-protein, and the beta-amyloid precursor protein intracellular domain in vivo. Proc Natl Acad Sci U S A. 2003;100(7):4162–7. [PMC free article: PMC153065] [PubMed: 12634421]
39.
Hamel FG, Upward JL, Bennett RG. In vitro inhibition of insulin-degrading enzyme by long-chain fatty acids and their coenzyme A thioesters. Endocrinology. 2003;144(6):2404–8. [PubMed: 12746301]

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this page (1.4M)

Related information

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...