Review article

J. KUDELOVA1, J. FLEISCHMANNOVA2,3, E. ADAMOVA2,4, E. MATALOVA2,5

PHARMACOLOGICAL CASPASE INHIBITORS:
RESEARCH TOWARDS THERAPEUTIC PERSPECTIVES

1Hospital Na Bulovce, Prague, Czech Republic; 2Institute of Animal Physiology and Genetics, CAS, v.v.i., Brno, Czech Republic; 3Regional Centre of Applied Molecular Oncology, Masaryk Memorial Cancer Institute, Brno, Czech Republic; 4University of Veterinary and Pharmaceutical Sciences, Brno, Czech Republic; 5Institute of Analytical Chemistry, CAS, v.v.i., Brno, Czech Republic
Caspases are key molecules of apoptosis and the inflammatory response. Up-regulation of the caspase cascade contributes to human pathologies such as neurodegenerative and immune disorders. Thus, blocking the excessive apoptosis by pharmacological inhibitors seems promising for therapeutic interventions in such diseases. Caspase inhibitors, both natural and artificial, have been used as research tools and have helped to define the role of the individual caspases in apoptosis and in non-apoptotic processes. Moreover, some caspase inhibitors have demonstrated their therapeutic efficiency in the reduction of cell death and inflammation in animal models of human diseases. However, no drug based on caspase inhibition has been approved on the market until now. Thus, the development of therapeutic approaches that specifically target caspases remains a great challenge and is now the focus of intense biological and clinical interest. Here, we provide a brief review of recent knowledge about pharmacological caspase inhibitors with special focus on their proposed clinical applications.
Key words:
caspase, caspase inhibitor, apoptosis, ischaemia-reperfusion injury, central nervous system, inflammation

INTRODUCTION

Caspases are key molecules executing cell death during ontogenesis and homeostasis of multicellular organisms (1, 2). They are activated in both main apoptotic pathways: extrinsic, mediated by death receptors, and intrinsic, where mitochondria play a central role. Moreover, caspases are essential for cytokine maturation during inflammation and contribute to biological processes such as proliferation, differentiation and migration.

According to their structure and function, caspases are classified into three groups. Caspases with long prodomains are mainly involved in inflammation (pro-inflammatory caspases) and apical phases of apoptosis (initiators); short prodomain caspases (executioners) act downstream of initiator caspases. Initiator caspases are synthesized as inactive monomers and typically require dimerization. Executioner caspases are synthesized as inactive dimers, they require cleavage to achieve their active conformation (3).

Increased apoptosis and caspase activity are associated with a number of disorders including hepatic and transplant injuries, myocardial infarction and neurodegenerative diseases, while pro-inflammatory caspases are implicated in rheumatoid arthritis or psoriasis (4, 5). For such diseases, pharmacological inhibition of caspases represents a possible therapeutic approach, as proved in animal models and initial clinical trials (6, 7). No caspase inhibitor-based drug has been introduced to the praxis until now and development of pharmaceuticals specifically targeting caspases remains a great challenge and is now the focus of intense biological and clinical interest.

CASPASE INHIBITORS

Hundreds of synthetic caspase inhibitors have been designed with the aim of bringing more insight into the complex caspase network and also to find new pharmaceuticals. Most of the inhibitors are pseudosubstrates (peptide based inhibitors, peptidomimetics and non-peptidic compounds), which block the active site of the enzyme. Allosteric inhibitors seem to be an innovative way to improve target selectivity (8, 9) (Fig. 1).

Figure 1
Fig. 1. Scheme of caspase inhibition. Covalent caspase inhibitors act as pseudosubstrates that block their catalytic site. A llosteric inhibitors form disulphide bonds with an allosteric site of the enzyme in its inactive state thereby preventing the formation of the active conformation of the enzyme and substrate binding to the active site.

The first synthetic caspase inhibitors were peptides based on aspartic acid modified with a reactive electrophilic group (so-called warhead or cysteine trap), forming reversible/irreversible covalent linkage with the nucleophilic active thiol site of the enzyme. A modified aspartate residue per se may serve as an inhibitor (Boc-Asp-FMK), but it is usually a part of a tri/tetrapeptide (P4)-P3-P2, corresponding with the substrate specificity of caspases (Fig. 1).

Inhibitors with different characteristics have been designed by using different electrophilic warheads. Peptide aldehydes, ketones and nitriles are reversible inhibitors with easily hydrolysable linkage (10). In contrast, a-substituted ketones, such as fluormethylketone (FMK) and chlormethylketone (CMK) irreversibly inactivate caspases trough the formation of thiomethylketone with the active site cysteine (10). Moreover, structure-activity relationship (SAR) revealed that 2, 6-dichlorobenzoyl (DCB) and trifluoperazine (TFP) warheads are successful irreversible moieties (11).

FMK inhibitors have a wide range of specificity and efficiency to block caspase activity. Specific tetrapeptide inhibitors Z/AC-DEVD-FMK, AC-YVAD-FMK/CMK, as well as the broad spectrum tripeptide inhibitor Z-VAD-FMK, are the most widely used caspase inhibitors for research purposes (12) and have substantially contributed to the understanding of apoptosis pathologies in animal models. FMK inhibitors show sufficient permeability compared to aldehyde warheads; however, they also affect other cysteine proteases and are metabolized to toxic fluoracetate (13, 14). Thus, next generation broad spectrum caspase inhibitors have been synthesized with an O-phenoxy group added at the C terminus together with a quinolyl group at the N terminus (Q-VD-OPH), which show higher effectiveness, reduced toxicity and enhanced cellular permeability compared to FMK inhibitors and are able to cross the blood-brain barrier (14).

The use of peptide inhibitors is limited for drug development. Despite substantial improvement, their pharmacokinetic properties (poor bioavailability, cell permeability, metabolic stability) are still not sufficient. Therefore, the peptidic nature of caspase inhibitors was reduced by the use of compounds mimicking the peptide backbone, active warheads were modified and chemical libraries were searched for small molecule caspase inhibitors (15). Application of these strategies resulted in the development of several peptidomimetic and non-peptidic caspase inhibitors based on heterocyclic scaffolds (isatin sulphonamides, aza-peptide epoxides, benzyl- and cyclohexyl-amines, pyrazinone mono-amides) with enhanced selectivity, stability and in vivo activity (16-18); some of them even entered clinical testing, such as IDN-6556; VX-765 (Table 1).

Table 1. Prospective clinical applications of caspase inhibitors.
Table 1

The design of specific caspase inhibitors via active site targeting is very difficult due to a stringent requirement for active site Asp for all caspases. Therefore, researchers focused on the dimerisation interface of the enzymes. This is very variable among caspases and thus offers new unique targets, e.g. BIR3 allosteric inhibition mechanism typical for caspase-9 (19). The newly designed allosteric inhibitors form disulphide bonds with cysteine allosteric sites in the dimerization interface and alter their productive conformation towards one that strongly resembles caspases in their zymogenic form (20, 21) (Fig. 1).

CASPASE INHIBITORS IN ANIMAL MODELS OF VARIOUS DISEASES

Some current applications of caspase inhibitors are in the research field, where they have confirmed their irreplaceable role in understanding mechanisms of cell death in cell cultures and organ explants, as well as in vivo in physiological and mainly pathophysiological situations. In animal models, caspase inhibitors revealed beneficial effects on ischaemia-reperfusion injury, sepsis, hepatitis, fibrosis and neurodegenerative disorders (Table 2).

Table 2. Diseases associated with inappropriate caspase activation and caspase inhibitors used to investigate related disorders.
Table 2

Liver injury

Beneficial effects of various caspase inhibitors have been shown in several models of liver diseases: the application of caspase inhibitors reduced transaminase elevation and apoptosis and decreased mortality in rodent models of acute liver failure, reduced ischaemia-reperfusion injury in rodent models and reduced hepatic fibrosis in bile duct ligated and NASH (non-alcoholic steatohepatitis) mouse models (22, 23). One of the first reports of the use of caspase inhibitors in vivo was in mouse models of liver injury (24). Since that time, several caspase inhibitors, either broad spectrum (Z-VAD-FMK, MX1122) (25) or caspase-1 specific (YVAD-CMK) (26), have been shown to inhibit hepatic apoptosis and prevent signs of liver damage induced by anti-CD95 antibodies or tumor necrosis factor (TNF) administration and improved recovery from the treatment. VX-166, a pan-caspase inhibitor decreased caspase-3 and fibrosis in liver (23). Based on preclinical data from animal models, several clinical trials have been performed to treat human liver pathologies (Table 1).

Autoimmune disorders

Caspases play also an important role in autoimmune disorders and infections. The inhibition of caspase-1 was effective in models of rheumatoid arthritis, osteoarthritis and psoriasis (27, 28). Z-VAD-FMK was shown to be protective in the asthma model, decreasing the infiltration of airways, mucus hypersecretion, oedema, and Th2 cytokine release (29). Moreover, Z-VAD-FMK effectively blocked pathologic fibrosis, which might result from autoimmune disorders and infections, as shown in the bleomycin-induced lung fibrosis model (30).

Infections and sepsis

Regarding bacterial infections, Z-VAD-FMK accelerated peritoneal bacterial clearance in peritonitis (31) and decreased hippocampal neuronal cell death in pneumococcal meningitis (32) as well as prevented lymphocyte apoptosis and improved survival in the caecal ligation puncture and pneumonia model of sepsis (33-35). Bacterial sepsis is a growing health problem with high mortality rate associated with a substantial loss of lymphocytes. This may be due to dysregulated apoptosis affecting both intrinsic and extrinsic cell death pathways, as shown by the pharmacological as well as genetic ablation of particular apoptotic cascade players (36). Several broad spectrum caspase inhibitors along with caspase-1, caspase-9 and caspase-8 specific inhibitors have been shown to effectively improve sepsis (37-40). Inflammatory caspase-11 and -12 seem to be promising targets for selective anti-caspase therapy, since caspase-11 deficient mice are resistant to LPS-induced shock (41). Caspase-12 polymorphism in humans is related with the vulnerability to sepsis and caspase-12 deficient mice do better in the case of sepsis (42). Previously mentioned pan-caspase inhibitor VX-166 has also potential in treatment for sepsis (43).

Ischaemia-reperfusion injury

Ischaemia-reperfusion injury is a complex phenomenon that comprises oxidative stress, inflammation and apoptosis activation. Several studies in various animal models have revealed the positive effect of caspase inhibitors on the ischaemia-reperfusion injury in the liver (44), lung (45), heart (46), kidney (47), and brain (48, 49).

MMPSI (isatin sulphonamide, reversible caspase-3 and -7 inhibitor) reduced apoptosis in vitro in a hypoxia model of cardiomyocytes (50). The inhibitor Z-VAD-FMK attenuated cardiomyocyte apoptosis in myocardial injury in rats subjected to a coronary occlusion followed by reperfusion (51). Z-LEHD-FMK (a specific caspase-9 inhibitor) was able to protect the rat myocardium during coronary microembolization by inhibiting apoptosis and improving cardiac function (52).

In the case of acute neurologic ischaemia, a neuroprotective effect was achieved using broad spectrum (Z-VAD-FMK, YVAD-CMK, Boc-Asp-FMK, IDN-5370; Q-VD-OPH), caspase-3-specific (DEVD-FMK, M826), caspase-1-specific (VRT-018858) and caspase-9-specific inhibitors (53-57). Moreover, caspase-2 seems to be a promising target for the future treatment of perinatal brain hypoxia-ischaemia. Caspase-2 is highly expressed in the neonatal brain (58) and its inhibition (TRP601) protects the newborn rodent brain against excytotoxicity, hypoxia-ischaemia, and perinatal arterial stroke, with no adverse effects on physiological parameters (59).

Central nervous system traumatic injury

Acute traumatic injury of the brain and spinal cord includes the activation of multiple caspases in neurons and astrocytes, but also in oligodendrocytes and microglia (60, 61). Their general blockade by Z-VAD-FMK was shown to improve the neurological outcome (62). Particularly, caspase-3 and caspase-1 seem to be crucial for neurological dysfunction in a rat model of traumatic brain injury (63-65). Namely broad spectrum caspase inhibitors (Z-VAD-FMK, Q-VD-OPh) (66, 67) and caspase-3 inhibitors (Z-DEVD-FMK, Ac-DMQD-CHO) (68, 69) were effective in the reduction of trauma-induced apoptosis in rodent models of traumatic spinal cord injury as well as the caspase-9 inhibitor, Z-LEHD-FMK (70), which reduced post-traumatic lesion size and improved motor performance. The combination of Z-LEDH-FMK with MgSO4 (N-methyl-D aspartate receptor antagonist) was demonstrated to minimize the effects of secondary injury in spinal cord trauma (71). Additionally, caspases-6 and -8 seem to have a prominent role in the apoptosis of neurons in the central nervous system, as shown using selective caspase inhibitors in the model of injured adult retinal ganglion cells (72). Besides the activation of multiple caspases, neurologic trauma also involves caspase independent, AIF (apoptosis inducing factor) mediated, cell death; thus, superior neuroprotection may be achieved with combined therapeutic strategies directed at multiple pathways of programmed cell death (73).

Neurodegeneration

Caspases are also potential therapeutic targets in neurodegenerative disorders and thus caspases may be involved in retinal degeneration in diseases such as glaucoma, age-related macular degeneration, and diabetic retinopathy (74, 75).

Z-VAD-FMK blocked apoptosis of the hair cells treated with aminoglycoside and streptomycin and thus protected against vestibular function deficits in vivo (76). Furthermore, Z-VAD-FMK could prevent apoptosis resulting from gunshot noise trauma (77). Activation of multiple caspases has been demonstrated in Alzheimer disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD) and amyotrophic lateral sclerosis (ALS) (78). Caspase-3 and caspase-6 have been shown to cleave tau and APP and actively contribute to senile plaque formation in AD models (79). Similarly, caspase-6 cleaves the HD protein into toxic fragments which enter the nucleus of neurons and interfere with gene transcription (80).

Blocking of caspases by Z-VAD-FMK delayed the disease onset and also mortality in ALS, and in mice expressing mutant human SOD1, and it also restored glutamatergic signalling and behavioural dysfunctions in transgenic animals (81). Q-VD-OPH revealed its neuroprotective role in mouse models of AD and HD (82). More recently, peptidomimetic or non-peptide inhibitors demonstrated great potential in neuroprotection. A specific reversible inhibitor of caspase-3, M826, prevented cell death of striatal neurons in a mouse model of HD (83).

Transplantation

Besides apoptosis and inflammation-related disorders, the application of caspase inhibitors might contribute to a better outcome of therapeutic applications, including transplantation and cryopreservation. Caspase inhibitors have been shown to inhibit ischaemia/reperfusion injury and thus improved graft survival after liver, lung, and pancreatic islet transplantation (45, 84-87). Namely, IDN-6556 improved glucose tolerance and pancreatic islet graft survival in mice (87) and even progressed to clinical testing.

Anticancer treatment

Cell death induction is a common approach in treatment based e. g. on cytostatics (88-90). Notably, caspase inhibitors have been shown to potentiate the effect of anticancer treatment. Z-VAD-FMK and M867 sensitized solid tumour cells to irradiation in vitro and in vivo in mouse hind limb tumour model (91, 92). Caspase inhibitors may be a component of differentiation therapy of leukaemia, perhaps in combination with derivatives of vitamin D (93).

Other applications

Recently, caspase inhibitors showed their potential for stem cell-based therapeutic applications. Cultivation of HSPCs in media with Z-VAD-FMK/Z-LLY-FMK and cytokines could be useful to enhance their homing potential (94).

Caspase inhibitors (Z-VAD-FMK) have been reported to improve cryopreservation efficiency when added to cryopreservation and post-thaw culture media. Positive results have been achieved in case of canine spermatozoa, porcine hepatocytes, human embryonic stem cells and amniotic fluid-derived stem cells (95-99). Similarly, the treatment of platelets with caspase-3 inhibitor could increase their functionality and survival in platelet concentrates (100).

CASPASE INHIBITORS IN CLINICAL TRIALS

The first inhibitor which entered clinical trial was VX-740 (Pralnacasan). Pralnacasan is an orally bioavailable pro-drug of the reversible specific inhibitor of caspase-1. The structure is derived from the tetrapeptide sequences YVAD, the preferred recognition sequence of caspase-1. Preclinical studies demonstrated that the drug is able to inhibit type II collagen-induced arthritis in mice and to reduce forepaw inflammation by decreasing disease severity by 70% (27). In addition to these effects Pralnacasan also suppressed the synthesis of IFN-γ in splenocytes (101). In initial Phase I/IIa studies, Pralnacasan was well tolerated in healthy volunteers and patients with rheumatoid arthritis and osteoarthritis. Moreover, the drug displayed good oral bioavailability and safety and significantly reduced joint symptoms and inflammation in patients with rheumatoid arthritis (27). Pralnacasan was withdrawn from clinical trials because of liver toxicity observed after long-term administration in animal studies, despite no adverse side effects occurring in the trial participants, even after long-term exposure (102).

VX-765 (Belnacasan), another specific caspase-1 inhibitor, also progressed into Phase IIa clinical trial in patients with psoriasis. In preclinical studies Belnacasan inhibited cytokine production in animal models of inflammatory, autoimmune joint and skin disease (103, 104). Phase I clinical trials demonstrated a dose-dependent reduction of cytokine levels in plasma and a Phase II trial conducted in patients proved the safety, tolerability and clinical activity. In 2011, a Phase IIb study to evaluate the efficacy and safety of VX-765 was started in subjects with treatment-resistant partial epilepsy (105). The drug was safe but not effective for epilepsy treatment (106).

IDN-6556 (Emricasan) is a broad spectrum irreversible caspase inhibitor that showed a strong effect in the mouse model of liver injury in vivo. In humans, it has been investigated for the treatment of a number of hepatic diseases and diabetes. To date, Emricasan has been administered to six Phase I and four Phase II clinical trials, and has generally been well-tolerated in both healthy volunteers and patients with liver disease. Clinical trials have demonstrated that Emricasan does not inhibit normal levels of caspase activity in healthy individuals (107). Emricasan was well tolerated and significantly improved markers of liver damage including reductions in serum AST and ALT in patients with hepatitis C (108, 109). The use of IDN-6556 has also been expanded to other important indications such as fatty liver disease (NASH), hepatitis B virus, alcoholic hepatitis and chronic liver failure (108). In 2013, dosing of Emricasan was initiated in the Phase II clinical trial in patients with severe alcoholic hepatitis (110), and last year, Phase II clinical trial started to evaluate the pharmacokinetics and pharmacodynamics of Emricasan in patients with acute or chronic liver failure (111). Besides hepatic diseases, IDN-6556 entered clinical trial in islet transplantation as treatment for diabetes (112) in order to minimize early post-transplant apoptosis (7, 87).

GS-9450 (Nivocasan), an irreversible inhibitor of caspases-1, -8, and -9, was applied in the case of hepatic injury. A Phase I trial dosing GS-9450 for 14 days in healthy volunteers proved the compound to be safe and well tolerated. Subsequently, Phase II trials were performed to examine the safety, tolerability, pharmacokinetics and initial activity of GS-9450 in preventing liver damage due to scarring or fibrosis caused by hepatitis C virus infection and non-alcoholic steatohepatitis. In both studies, treatment with GS-9450 resulted in lower ALT-values. Despite the four-week application of GS-9450 showing no adverse side effects in patients with NASH, a larger, 6-month study in hepatitis C subjects, reported episodes of drug-induced liver injury and the trial was terminated (6, 113) (Table 1).

NCX-1000 (2(acetyloxy) benzoic acid-3(nitrooxymethyl) phenyl ester), an antiapoptotic compund (inhibitor caspase-3, -8, -9), entered Phase IIa of clinical trial to prevent portal hypertension. Oral doses were administered for the period of 16 days. NCX-1000 was well tolerated but did not reduce portal hypertension due to the low selectivity (114, 115).

FUTURE OF ANTI-CASPASE TREATMENT

The ongoing clinical trials demonstrate that caspase inhibitors could have significant therapeutic potential; nonetheless, some clinical testing has been discontinued and none of the caspase inhibitors have yet reached Phase III. Thus, there are still key questions and obstacles which prevent drugs based on caspase inhibitors from their routine clinical practice as anti-caspase treatment.

(I) Efficient and specific caspase inhibitors that would minimize adverse effects are still lacking.

(II) Type of cell death, specific caspases activated, possible redundancies and non-apoptotic/non-inflammatory functions within the caspase family that may interfere with the treatment are not entirely clear in numerous clinical states (116-118).

(III) Drug dosing, timing of drug administration, its duration and drug delivery to the site of action are not yet well understood.

Adverse effects interrupted the clinical testing of VX-740 and GS-9450. These may be caused by interference with non-apoptotic or non-inflammatory roles of caspases as well as poor specificity and the selectivity of caspase inhibitors. In particular, medication with irreversible caspase inhibitors may lead to the modulation of non-target caspases or unrelated proteins (119) such as cysteine cathepsins and legumain. Since irreversible inhibitors show better effectivity in vivo, these could be beneficial, despite of their disadvantages, in short-term therapy of acute life threatening diseases. For the long-term therapy of chronic diseases like rheumatoid arthritis or hepatitis C, reversible non-covalent allosteric inhibitors would be the best choice (120). Such inhibitors are much more difficult to synthesize compared to irreversible inhibitors.

Recent advances in combinatorial chemistry, including innovative screening techniques, structural biology, and bioinformatics greatly accelerated the development of new caspase inhibitors. Studies addressing structure and function of particular caspases helped to identify differences between closely related caspases applicable for development of highly specific inhibitors (21, 121). Available bioinformatics tools allow searching for structures effectively inhibiting caspases as well as prediction of their properties such as oral bioavailability, metabolic stability and minimal toxicity. Pharmacophore modelling, docking studies and searches in libraries of potential drug compounds have been performed to find functional caspase inhibitors structures (11, 21, 122-123). These approaches have great potential to identify compounds that could progress into clinically applicable drugs.

Inhibition of the caspase pathway may be detrimental by blocking non-inflammatory apoptotic cell clearance and thereby promoting inflammation. The broad-spectrum caspase inhibitor ZVAD-FMK modulates the three major types of cell death; it blocks apoptotic cell death, sensitizes cells to necrotic cell death, and induces autophagic cell death. Inhibition of caspases by Z-VAD-FMK might induce hyper-acute TNF-induced shock in certain situations; this means exacerbated TNF-α toxicity, and enhanced oxidative stress and mitochondrial damage, resulting in hyper-acute haemodynamic collapse, kidney failure, and death of the patient (124).

Adverse effects of necrosis and inflammation may be minimized by proper timing and dosing of the drugs for every individual pathological state. For ischaemia-reperfusion injury, the therapeutic window of caspase inhibitors can change according to the type of damage, the duration of ischaemia, and also to the other medicines in combination. It has been demonstrated that inhibition of caspases leads to a decrease in apoptosis during early reperfusion, but with prolonged ischaemia it seems to be ineffective or rather detrimental at least in the case of kidney and myocardial ischaemia-reperfusion injury, respectively (117, 124, 125).

In order to lower the risk of toxicity in the long-term use of caspase inhibitors, drugs based on caspase inhibitors must only target the offending cells and not to disturb the natural turnover of normal cells and hypothetically start or extend tumour growth. Therefore, inhibition of specific caspases in selective cell types is the major challenge. For this purpose, the development of specific nontoxic nanocarriers able to cross different body barriers and to safety release caspase inhibitors is of major interest. Especially for neurological disorders, the drug must be able to cross the blood brain barrier; therefore, novel nanoparticulate drug delivery systems like liposomes are hot candidates (126, 127).

We have focused on apoptotic roles of caspases but there are also non-apoptotic functions of caspases (128-131). Caspase inhibitors could therefore potentially block more functions. Despite numerous attempts non-apoptotic functions of caspases are not completely understood yet.

In summary, caspase inhibitors are essential research tools that are useful in various experimental systems to monitor caspase activity, to understand functions of individual caspases in development and everyday tissue homeostasis and in pathological processes. Nevertheless, the path to “anti-caspase” therapy is still very difficult; several studies have been suspended and many important questions are still open regarding side effects or complications such as tumour growth during or after treatment. Despite the mentioned obstacles, there is no doubt about the prospective of caspase inhibitors for clinical praxis (Tables 1 and 2), and intense research in this field with running clinical trials may yield new safe therapeutic strategies towards personalized therapy.

Acknowledgements: Related research is supported by the Grant Agency of the Czech Republic, project P502/12/1285 at the UVPS, 14-37368G at the IAPG ASCR, v.v.i., and 14-28254s at the IAC ASCR, v.v.i. European Regional Development Fund and the State Budget of the Czech Republic (RECAMO, CZ.1.05/2.1.00/03.0101) supports research at the MMCI. E. Matalova is a holder of the L´Oreal UNESCO Fellowship 2015.

Conflict of interests: None declared.

REFERENCES

  1. Vaux DL, Korsmeyer SJ. Cell death in development. Cell 1999; 96: 245-254.
  2. Krammer PH. CD95’s deadly mission in the immune system. Nature 2000; 407: 789-795.
  3. Pop C, Salvesen GS. Human caspases: activation, specificity, and regulation. J Biol Chem 2009; 284: 21777-81.
  4. Fadeel B, Orrenius S. Apoptosis: a basic biological phenomenon with wide-ranging implications in human disease. J Intern Med 2005; 258: 479-517.
  5. Favaloro B, Allocati N, Graziano V, Di Ilio C, De Laurenzi V. Role of apoptosis in disease. Aging (Albany NY) 2012; 4: 330-349.
  6. Ratziu V, Sheikh MY, Sanyal AJ, et al. A phase 2, randomized, double-blind, placebo-controlled study of GS-9450 in subjects with non-alcoholic steatohepatitis. Hepatology 2012; 55: 419-428.
  7. McCall MD, Maciver AM, Kin T, et al. Caspase inhibitor IDN6556 facilitates marginal mass islet engraftment in a porcine islet auto-transplant model. Transplantation 2012; 94: 30-35.
  8. Hacker HG, Sisay MT, Gutschow M. Allosteric modulation of caspases. Pharmacol Ther 2011; 132: 180-195.
  9. Gao J, Wells JA. Identification of specific tethered inhibitors for caspase-5. Chem Biol Drug Des 2012; 79: 209-215.
  10. Callus BA, Vaux DL. Caspase inhibitors: viral, cellular and chemical. Cell Death Differ 2007; 14: 73-78.
  11. Ullman BR, Aja T, Deckwerth TL, et al. Structure-activity relationships within a series of caspase inhibitors: effect of leaving group modifications. Bioorg Med Chem Lett 2003; 13: 3623-3626.
  12. Vandenabeele P, Vanden Berghe T, Festjens N. Caspase inhibitors promote alternative cell death pathways. Sci STKE 2006; 2006 (358): pe44.
  13. Van Noorden CJF. The history of Z-VAD-FMK, a tool for understanding the significance of caspase inhibition. Acta Histochem 2001; 103: 241-251.
  14. Caserta TM, Smith AN, Gultice AD, Reedy MA, Brown TL. Q-VD-OPh, a broad spectrum caspase inhibitor with potent anti-apoptotic properties. Apoptosis 2003; 8: 345-352.
  15. Fischer U, Schulze-Osthoff K. New approaches and therapeutics targeting apoptosis in disease. Pharmacol Rev 2005; 57: 187-215.
  16. Asgian JL, James KE, Li ZZ, et al. Aza-peptide epoxides: a new class of inhibitors selective for clan CD cysteine proteases. J Med Chem 2002; 45: 4958-4860.
  17. James KE, Asgian JL, Li ZZ, et al. Design, synthesis, and evaluation of aza-peptide epoxides as selective and potent inhibitors of caspases-1, -3, -6, and -8. J Med Chem 2004; 47: 1553-1574.
  18. Leyva MJ, Degiacomo F, Kaltenbach LS, et al. Identification and evaluation of small molecule pan-caspase inhibitors in Huntington’s disease models. Chem Biol 2010; 17: 1189-1200. Erratum in: Chem Biol 2013; 20: 742.
  19. Huber KL, Ghosh S, Hardy JA. Inhibition of caspase-9 by stabilized peptides targeting the dimerization interface. Biopolymers 2012; 98: 451-465.
  20. Hardy JA, Lam J, Nguyen JT, O’Brien T, Wells JA. Discovery of an allosteric site in the caspases. Proc Natl Acad Sci USA 2004; 101: 12461-12466.
  21. Feldman T, Kabaleeswaran V, Jang SB, et al. A class of allosteric caspase inhibitors identified by high-throughput screening. Mol Cell 2012; 47: 585-595.
  22. Masuoka HC, Guicciardi ME, Gores GJ. Caspase inhibitors for the treatment of hepatitis C. Clin Liver Dis 2009; 13: 467-475.
  23. Witek RP, Stone WC, Karaca FG, et al. Pan-caspase inhibitor VX-166 reduces fibrosis in an animal model of non-alcoholic steatohepatitis. Hepatology 2009; 50: 1421-1430.
  24. Rodriguez I, Matsuura K, Ody C, Nagata S, Vassalli P. Systemic injection of a tripeptide inhibits the intracellular activation of CPP32-like proteases in vivo and fully protects mice against Fas-mediated fulminant liver destruction and death. J Exp Med 1996; 184: 2067-2072.
  25. Cai SX, Guan LF, Jia SJ, et al. Dipeptidyl aspartyl fluoromethylketones as potent caspase inhibitors: SAR of the N-protecting group. Bioorg Med Chem Lett 2004; 14: 5295-5300.
  26. Rouquet N, Pages JC, Molina T, Briand P, Joulin V. ICE inhibitor YVADcmk is a potent therapeutic agent against in vivo liver apoptosis. Curr Biol 1996; 6: 1192-1195.
  27. Rudolphi K, Gerwin N, Verzijl N, van der Kraan P, van den Berg W. Pralnacasan, an inhibitor of interleukin-1beta converting enzyme, reduces joint damage in two murine models of osteoarthritis. Osteoarthritis Cartilage 2003; 11: 738-746.
  28. Le GT, Abbenante G. Inhibitors of TACE and caspase-1 as anti-inflammatory drugs. Curr Med Chem 2005; 12: 2963-2977.
  29. Iwata A, Nishio K, Winn RK, Chi EY, Henderson WR Jr, Harlan JM. A broad-spectrum caspase inhibitor attenuates allergic airway inflammation in murine asthma model. J Immunol 2003; 170: 3386-3391.
  30. Kuwano K, Hagimoto N, Hara N. Molecular mechanisms of pulmonary fibrosis and current treatment. Curr Mol Med. 2001; 1: 551-573.
  31. Catalan MP, Esteban J, Subira D, Egido J, Ortiz A. Inhibition of caspases improves bacterial clearance in experimental peritonitis. Perit Dial Int 2003; 23: 123-126.
  32. Braun JS, Novak R, Herzog KH, Bodner SM, Cleveland JL, Tuomanen EI. Neuroprotection by a caspase inhibitor in acute bacterial meningitis. Nat Med 1999; 5: 298-302.
  33. Hotchkiss RS, Tinsley KW, Swanson PE, et al. Prevention of lymphocyte cell death in sepsis improves survival in mice. Proc Natl Acad Sci USA 1999; 96: 14541-14546.
  34. Hotchkiss RS, Dunne WM, Swanson PE, et al. Role of apoptosis in Pseudomonas aeruginosa pneumonia. Science 2001; 294: 1783.
  35. Hotchkiss RS, Tinsley KW, Swanson PE, et al. Sepsis-induced apoptosis causes progressive profound depletion of B and CD4(+) T lymphocytes in humans. J Immunol 2001; 166: 6952-6963.
  36. Wesche-Soldato DE, Lomas-Neira JL, Perl M, Jones L, Chung CS, Ayala A. The role and regulation of apoptosis in sepsis. J Endotoxin Res 2005; 11: 375-382.
  37. Hotchkiss RS, Chang KC, Swanson PE, et al. Caspase inhibitors improve survival in sepsis: a critical role of the lymphocyte. Nat Immunol 2000; 1: 496-501.
  38. Methot N, Huang JQ, Coulombe N, et al. Differential efficacy of caspase inhibitors on apoptosis markers during sepsis in rats and implication for fractional inhibition requirements for therapeutics. J Exp Med 2004; 199: 199-207.
  39. Fauvel H, Marchetti P, Chopin C, Formstecher P, Neviere R. Differential effects of caspase inhibitors on endotoxin-induced myocardial dysfunction and heart apoptosis. Am J Physiol Heart Circul Physiol 2001; 280: H1608-H1614.
  40. Oberhotzer C, Tschoeke SK, Moldawer LL, Oberholzer A. Local thymic caspase-9 inhibition improves survival during polymicrobial sepsis in mice. J Mol Med 2006; 84: 389-395.
  41. Kang SJ, Wang S, Kuida K, Yuan J. Distinct downstream pathways of caspase-11 in regulating apoptosis and cytokine maturation during septic shock response. Cell Death Differ 2002; 9: 1115-1125.
  42. Saleh M, Mathison JC, Wolinski MK, et al. Enhanced bacterial clearance and sepsis resistance in caspase-12-deficient mice. Nature 2006; 440: 1064-1068.
  43. Weber P, Wang P, Maddens S, et al. VX-166: a novel potent small molecule caspase inhibitor as a potential therapy for sepsis. Crit Care 2009; 13: R136.
  44. Cursio R, Gugenheim J, Ricci JE, et al. A caspase inhibitor fully protects rats against lethal normothermic liver ischemia by inhibition of liver apoptosis. FASEB J 1999; 13: 253-261.
  45. Quadri SM, Segall L, de Perrot M, et al. Caspase inhibition improves ischemia-reperfusion injury after lung transplantation. Am J Transplant 2005; 5: 292-299.
  46. Okamura T, Miura T, Takemura, et al. Effect of caspase inhibitors on myocardial infarct size and myocyte DNA fragmentation in the ischaemia-reperfused rat heart. Cardiovasc Res 2000; 45: 642-650.
  47. Chatterjee PK, Todorovic Z, Sivarajah A, et al. Differential effects of caspase inhibitors on the renal dysfunction and injury caused by ischemia-reperfusion of the rat kidney. Eur J Pharmacol 2004; 503: 173-183.
  48. Cheng Y, Deshmukh M, D’Costa A, et al. Caspase inhibitor affords neuroprotection with delayed administration in a rat model of neonatal hypoxic-ischemic brain injury. J Clin Invest 1998; 101: 1992-1999.
  49. Wiessner C, Sauer D, Alaimo D, Allegrini PR. Protective effect of a caspase inhibitor in models for cerebral ischemia in vitro and in vivo. Cell Mol Biol (Noisy le grand) 2000; 46: 53-62.
  50. Chapman JG, Magee WP, Stukenbrok HA, Beckius GE, Milici AJ, Tracey WR. A Novel nonpeptidic caspase-3/7 inhibitor, (S)-(+)-5-[1-(2-methoxymethylpyrrolidinyl)sulfonyl]isatin reduces myocardial ischemic injury. Eur J Pharmacol 2002; 456: 59-68.
  51. Yaoita H, Ogawa K, Maehara K, Maruyama Y. Attenuation of ischemia/reperfusion injury in rats by a caspase inhibitor. Circulation 1998; 97: 276-281.
  52. Su Q, Li L, Liu YC, Zhou Y, Lu YG, Wen WM. Effect of metoprolol on myocardial apoptosis and caspase-9 activation after coronary microembolization in rats. Exp Clin Cardiol 2013; 18: 161-165.
  53. Feng YZ, Fratkin JD, LeBlanc MH. Inhibiting caspase-9 after injury reduces hypoxic ischemic neuronal injury in the cortex in the newborn rat. Neurosci Lett 2003; 344: 201-204.
  54. Han BH, Xu DG, Choi JJ, et al. Selective, reversible caspase-3 inhibitor is neuroprotective and reveals distinct pathways of cell death after neonatal hypoxic-ischemic brain injury. J Biol Chem 2002; 277: 30128-30136.
  55. Ross J, Brough D, Gibson RM, Loddick SA, Rothwell NJ. A selective, non-peptide caspase-1 inhibitor, VRT-018858, markedly reduces brain damage induced by transient ischemia in the rat. Neuropharmacology 2007; 53: 638-642.
  56. Braun JS, Prass K, Dirnagl U, Meisel A, Meisel C. Protection from brain damage and bacterial infection in murine stroke by the novel caspase-inhibitor Q-VD-OPH. Exp Neurol 2007; 206: 183-191.
  57. Han W, Sun YY, Wang XY, Zhu CL, Blomgren K. Delayed, long-term administration of the caspase inhibitor Q-VD-OPh reduced brain injury induced by neonatal hypoxia-ischemia. Dev Neurosci 2014; 36: 64-72.
  58. Carlsson Y, Schwendimann L, Vontell R, et al. Genetic inhibition of caspase-2 reduces hypoxic-ischemic and excitotoxic neonatal brain injury. Ann Neurol 2011; 70: 781-789.
  59. Chauvier D, Renolleau S, Holifanjaniaina S, et al. Targeting neonatal ischemic brain injury with a pentapeptide-based irreversible caspase inhibitor. Cell Death Dis 2011; 2: e203.
  60. Venero JL, Burguillos MA, Brundin P, Joseph B. The executioners sing a new song: killer caspases activate microglia. Cell Death Differ 2011; 18: 1679-1691.
  61. Fricker M, Vilalta A, Tolkovsky AM, Brown GC. Caspase inhibitors protect neurons by enabling selective necroptosis of inflamed microglia. J Biol Chem 2013; 288: 9145-9152.
  62. Knoblach SM, Nikolaeva M, Huang XL, et al. Multiple caspases are activated after traumatic brain injury: evidence for involvement in functional outcome. J Neurotrauma 2002; 19: 1155-1170.
  63. Yakovlev AG, Knoblach SM, Fan L, Fox GB, Goodnight R, Faden AI. Activation of CPP32-like caspases contributes to neuronal apoptosis and neurological dysfunction after traumatic brain injury. J Neurosci 1997; 17: 7415-7424.
  64. Clark RS, Kochanek PM, Watkins SC, et al. Caspase-3 mediated neuronal death after traumatic brain injury in rats. J Neurochem 2000; 74: 740-753.
  65. Mejia RO, Ona VO, Li M, Friedlander RM. Minocycline reduces traumatic brain injury-mediated caspase-1 activation, tissue damage, and neurological dysfunction. Neurosurgery 2001; 48: 1393-1399.
  66. Li M, Ona VO, Chen M, et al. Functional role and therapeutic implications of neuronal caspase-1 and-3 in a mouse model of traumatic spinal cord injury. Neuroscience 2000; 99: 333-342.
  67. Colak A, Antar V, Karaoglan A, et al. Q-VD-OPh, a pancaspase inhibitor, reduces trauma-induced apoptosis and improves the recovery of hind-limb function in rats after spinal cord injury. Neurocirugia (Astur) 2009; 20: 533-540.
  68. Barut S, Unlu YA, Karaoglan A, et al. The neuroprotective effects of z-DEVD.fmk, a caspase-3 inhibitor, on traumatic spinal cord injury in rats. Surg Neurol 2005; 64: 213-220.
  69. Akdemir O, Berksoy I, Karaoglan A, et al. Therapeutic efficacy of Ac-DMQD-CHO, a caspase 3 inhibitor, for rat spinal cord injury. J Clin Neurosci 2008; 15: 672-678.
  70. Colak A, Karaoglan A, Barut S, Kokturk S, Akyildiz AI, Tasyurekli M. Neuroprotection and functional recovery after application of the caspase-9 inhibitor z-LEHD-fmk in a rat model of traumatic spinal cord injury. J Neurosurg Spine 2005; 2: 327-334.
  71. Sencer A, Aydoseli A, Aras Y, et al. Effects of combined and individual use of N-methyl-D aspartate receptor antagonist magnesium sulphate and caspase-9 inhibitor z-LEDH-fmk in experimental spinal cord injury. Ulus Travma Acil Cerrahi Derg 2013; 19: 313-319.
  72. Monnier PP, D’Onofrio PM, Magharious M, et al. Involvement of caspase-6 and caspase-8 in neuronal apoptosis and the regenerative failure of injured retinal ganglion cells. J Neurosci 2011; 31: 10494-10505.
  73. Piao CS, Loane DJ, Stoica BA, et al. Combined inhibition of cell death induced by apoptosis inducing factor and caspases provides additive neuroprotection in experimental traumatic brain injury. Neurobiol Dis 2012; 46: 745-758.
  74. Vincent JA, Mohr S. Inhibition of caspase-1/interleukin-1 beta signalling prevents degeneration of retinal capillaries in diabetes galactosaemia. Diabetes 2007; 56: 224-230.
  75. Vigneswara V, Berry M, Logan A, Ahmed Z. Pharmacological inhibition of caspase-2 protects axotomised retinal ganglion cells from apoptosis in adult rats. PLoS One 2012; 7: e53473.
  76. Matsui JI, Haque A, Huss D, et al. Caspase inhibitors promote vestibular hair cell survival and function after aminoglycoside treatment in vivo. J Neurosci 2003; 23: 6111-6122.
  77. Abaamrane L, Raffin F, Schmerber S, Sendowski I. Intracochlear perfusion of leupeptin and z-VAD-FMK: influence of anti-apoptotic agents on gunshot-induced hearing loss. Eur Arch Otorhinolaryngol 2011; 268: 987-993.
  78. Vila M, Przedborski S. Targeting programmed cell death in neurodegenerative diseases. Nat Rev Neurosci 2003; 4: 365-375.
  79. Zhao H, Zhao WJ, Lok K, Wang ZJ, Yin M. A synergic role of caspase-6 and caspase-3 in tau truncation at D421 induced by H2O2. Cell Mol Neurobiol 2014; 34: 369-378.
  80. Graham RK, Deng Y, Slow EJ, et al. Cleavage at the caspase-6 site is required for neuronal dysfunction and degeneration due to mutant huntingtin. Cell 2006; 125: 1179-1191.
  81. Li MW, Ona VO, Guegan C, et al. Functional role of caspase-1 and caspase-3 in an ALS transgenic mouse model. Science 2000; 288: 335-339.
  82. Rohn TT, Kokoulina P, Eaton CR, Poon WW. Caspase activation in transgenic mice with Alzheimer-like pathology: results from a pilot study utilizing the caspase inhibitor, Q-VD-OPh. Int J Clin Exp Med 2009; 2: 300-308.
  83. Toulmond S, Tang K, Bureau Y, et al. Neuroprotective effects of M826, a reversible caspase-3 inhibitor, in the rat malonate model of Huntington’s disease. Br J Pharmacol 2004; 141: 689-697.
  84. Emamaullee JA, Stanton L, Schur C, Shapiro AM. Caspase inhibitor therapy enhances marginal mass islet graft survival and preserves long-term function in islet transplantation. Diabetes 2007; 56: 1289-1298.
  85. Natori S, Higuchi H, Contreras P, Gores GJ. The caspase inhibitor IDN-6556 prevents caspase activation and apoptosis in sinusoidal endothelial cells during liver preservation injury. Liver Transpl 2003; 9: 278-284.
  86. Hoglen NC, Anselmo DM, Katori M, et al. A caspase inhibitor, IDN-6556, ameliorates early hepatic injury in an ex vivo rat model of warm and cold ischemia. Liver Transpl 2007; 13: 361-366.
  87. McCall M, Toso C, Emamaullee J, et al. The caspase inhibitor IDN-6556 (PF3491390) improves marginal mass engraftment after islet transplantation in mice. Surgery 2011; 150: 48-55.
  88. Gerl R, Vaux DL. Apoptosis in the development and treatment of cancer. Carcinogenesis 2005; 26: 263-70.
  89. Siedlecka-Kroplewska K, Jozwik A, Boguslawski W, et al. Pterostilbene induces accumulation of autophagic vacuoles followed by cell death in Hl60 human leukemia cells. J Physiol Pharmacol 2013; 64: 545-56.
  90. Toton E, Ignatowicz E, Bernard MK, Kujawski J, Rybczynska M. Evaluation of apoptotic activity of new condensed pyrazole derivatives. J Physiol Pharmacol 2013; 64: 115-123.
  91. Kim KW, Moretti L, Lu B. M867, a novel selective inhibitor of caspase-3 enhances cell death and extends tumor growth delay in irradiated lung cancer models. PLoS One 2008; 3: e2275.
  92. Moretti L, Kim KW, Jung DK, Willey CD, Lu B. Radiosensitization of solid tumors by Z-VAD, a pan-caspase inhibitor. Mol Cancer Ther 2009; 8: 1270-1279.
  93. Chen-Deutsch X, Kutner A, Harrison JS, Studzinski GP. The pan-caspase inhibitor Q-VD-OPh has anti-leukemia effects and can interact with vitamin D analogs to increase HPK1 signaling in AML cells. Leuk Res 2012; 36: 884-888.
  94. Sangeetha VM, Kadekar D, Kale VP, Limaye LS. Pharmacological inhibition of caspase and calpain proteases: A novel strategy to enhance the homing responses of cord blood HSPCs during expansion. PLoS One 2012; 7: e29383.
  95. Bissoyi A, Nayak B, Pramanik K, Sarangi SK. Targeting cryopreservation-induced cell death: A review. Biopreserv Biobank 2014; 12: 23-34.
  96. Yagi T, Hardin JA, Valenzuela YM, Miyoshi H, Gores GJ, Nyberg SL. Caspase inhibition reduces apoptotic death of cryopreserved porcine hepatocytes. Hepatology 2001; 33: 1432-1440.
  97. Peter AT, Linde-Forsberg C. Efficacy of the anti-caspase agent zVAD-fmk on post-thaw viability of canine spermatozoa. Theriogenology 2003; 59: 1525-1532.
  98. Heng BC, Clement MV, Cao T. Caspase inhibitor Z-VAD-FMK enhances the freeze-thaw survival rate of human embryonic stem cells. Biosci Rep 2007; 27: 257-264.
  99. Cho HJ, Lee SH, Yoo JJ, Shon YH. Evaluation of cell viability and apoptosis in human amniotic fluid-derived stem cells with natural cryoprotectants. Cryobiology 2014; 68: 244-250.
  100. Shiri R, Yari F, Ahmadinejad M, Vaeli S, Tabatabaei MR. The caspase-3 inhibitor (peptide Z-DEVD-FMK) affects the survival and function of platelets in platelet concentrate during storage. Blood Res 2014; 49: 49-53.
  101. Loher F, Bauer C, Landauer N, et al. The interleukin-1 beta-converting enzyme inhibitor pralnacasan reduces dextran sulfate sodium-induced murine colitis and T helper 1 T-cell activation. J Pharmacol Exp Ther 2004; 308: 583-590.
  102. URL 1: http://www.evaluategroup.com/Universal/View. aspx?type=Story&id=45867
  103. Wannamaker W, Davies R, Namchuk M, et al. (S)-1-((S)-2-{[1-(4-amino-3-chloro-phenyl)-methanoyl]-amino}-3,3-dimethy l-butanoyl)-pyrrolidine-2-carboxylic acid ((2R,3S)-2-ethoxy-5-oxo-tetrahydro-furan-3-yl)-amide (VX-765), an orally available selective interleukin (IL)-converting enzyme/caspase-1 inhibitor, exhibits potent anti-inflammatory activities by inhibiting the release of IL-1 beta and IL-18. J Pharmacol Exp Ther 2007; 321: 509-516.
  104. Stack JH, Beaumont K, Larsen PD, et al. IL-converting enzyme/caspase-I inhibitor VX-765 blocks the hypersensitive response to an inflammatory stimulus in monocytes from familial cold autoinflammatory syndrome patients. J Immunol 2005; 175: 2630-2634.
  105. URL 2: http://investors.vrtx.com/releasedetail.cfm? ReleaseID=555967
  106. Bialer M, Johannessen SI, Levy RH, Perucca E, Tomson T, White HS. Progress report on new antiepileptic drugs: a summary of the Eleventh Eilat Conference (EILAT XI). Epilepsy Res 2013; 103: 2-30.
  107. Valentino KL, Gutierrez M, Sanchez R, Winship MJ, Shapiro DA. First clinical trial of a novel caspase inhibitor: anti-apoptotic caspase inhibitor, IDN-6556, improves liver enzymes. Int J Clin Pharmacol Ther 2003; 41: 441-449.
  108. Pockros PJ, Schiff ER, Shiffman ML, et al. Oral IDN-6556, an anti-apoptotic caspase inhibitor, may lower aminotransferase activity in patients with chronic hepatitis C. Hepatology 2007; 46: 324-329.
  109. Shiffman ML, Pockros P, McHutchison JG, Schiff ER, Morris M, Burgess G. Clinical trial: the efficacy and safety of oral PF-03491390, a pancaspase inhibitor - a randomized placebo-controlled study in patients with chronic hepatitis C. Aliment Pharmacol Ther 2010; 31: 969-978.
  110. URL 3: http://clinicaltrial.gov/ct2/show/NCT01912404 ?term=IDN-6556&rank=2
  111. URL 4: http://clinicaltrial.gov/ct2/show/NCT01937130 ?term=IDN-6556&rank=3
  112. URL 5: http://clinicaltrial.gov/ct2/show/NCT01653899 ?term=IDN-6556&rank=1
  113. Arends JE, Hoepelman AIM, Nanlohy NM, et al. Low doses of the novel caspase-inhibitor GS-9450 leads to lower caspase-3 and -8 expression on peripheral CD4+ and CD8+ T-cells. Apoptosis 2011; 16: 959-966.
  114. Fiorucci S, Mencarelli A, Palazzetti B, Del Soldato P, Morelli A, Ignarro LJ. An NO derivative of ursodeoxycholic acid protects against Fas-mediated liver injury by inhibiting caspase activity. Proc Natl Acad Sci 2001; 98: 2652-2627.
  115. Berzigotti A, Bellot P, De Gottardi A, et al. NCX-1000, a nitric oxide-releasing derivative of UDCA, does not decrease portal pressure in patients with cirrhosis: results of a randomized, double-blind, dose-escalating study. Am J Gastroenterol 2010; 105: 1094-1101.
  116. Van Molle W, Brouckaert P, Libert C. Caspase-1 is not involved in experimental hepatitis in mouse. FEBS Lett 1999; 445: 115-118.
  117. McCully JD, Wakiyama H, Hsieh YJ, Jones M, Levitsky S. Differential contribution of necrosis and apoptosis in myocardial ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol 2004; 286: 1923-1935.
  118. Mersmann J, Zacharowski PA, Schmitz I, Zacharowski K. Caspase inhibitor zVAD.fmk reduces infarct size after myocardial ischaemia and reperfusion in rats but not in mice. Resuscitation 2008; 79: 468-474.
  119. Berger AB, Sexton KB, Bogyo M. Commonly used caspase inhibitors designed based on substrate specificity profiles lack selectivity. Cell Res 2006; 16: 961-963.
  120. Heise CE, Murray J, Augustyn KE, et al. Mechanistic and structural understanding of uncompetitive inhibitors of caspase-6. PLoS One 2012; 7: e50864.
  121. Nakatsumi H, Yonehara S. Identification of functional regions defining different activity in caspase-3 and caspase-7 within cells. J Biol Chem 2010; 285: 25418-25425.
  122. Sharma S, Basu A, Agrawal RK. Pharmacophore modeling and docking studies on some nonpeptide-based caspase-3 inhibitors. Biomed Res Int 2013; 2013: 306081.
  123. Wu JH, Wang YR, Liang SG, Ma HC. Cytoprotective effect of selective small-molecule caspase inhibitors against staurosporine-induced apoptosis. Drug Des Devel Ther 2014; 8: 583-600.
  124. Cauwels A, Janssen B, Waeytens A, Cuvelier C, Brouckaert P. Caspase inhibition causes hyperacute tumor necrosis factor-induced shock via oxidative stress and phospholipase A2. Nat Immunol 2003; 4: 387-393.
  125. Daemen M, van’t Veer C, Denecker G, et al. Inhibition of apoptosis induced by ischemia-reperfusion prevents inflammation. J Clin Invest 1999; 104: 541-549.
  126. Spuch C, Navarro C. Liposomes for targeted delivery of active agents against neurodegenerative diseases (Alzheimer’s disease and Parkinson’s disease). J Drug Deliv 2011; 2011: 469679.
  127. Karatas H, Aktas Y, Gursoy-Ozdemir Y, et al. A nanomedicine transports a peptide caspase-3 inhibitor across the blood-brain barrier and provides neuroprotection. J Neurosci 2009; 29: 13761-13769.
  128. Svandova E, Lesot H, Vanden Berghe T, et al. Non-apoptotic functions of caspase-7 during osteogenesis. Cell Death Dis 2014; 5: e1366.
  129. Maelfait J, Beyaert R. Non-apoptotic functions of caspase-8. Biochem Pharmacol 2008; 76: 1365-1373.
  130. Gulyaeva NV. Non-apoptotic functions of caspase-3 in nervous tissue. Biochemistry (Mosc) 2003; 68: 1171-1180.
  131. Schwerk C, Schulze-Osthoff K. Non-apoptotic functions of caspases in cellular proliferation and differentiation. Biochem Pharmacol 2003; 66: 1453-1458.
  132. McDonald J, Wright MR, Wright CD, Kuehl PJ, Duncan M, Doyle-Eisele M. GS-9450-A novel caspase inhibitor provides protection against bleomycin induced pulmonary fibrosis by inhibition of apoptosis. Am J Respir Crit Care Med 2010; 181: A1053.
  133. Canbay A, Feldstein A, Baskin-Bey E, Bronk SF, Gores GJ. The caspase inhibitor IDN-6556 attenuates hepatic injury and fibrosis in the bile duct ligated mouse. J Pharmacol Exp Ther 2004; 308: 1191-1196.
  134. Maroso M, Balosso S, Ravizza T, et al. Interleukin 1b biosynthesis inhibition reduces acute seizures and drug resistant chronic epileptic activity in mice. Neurotherapeutics 2011; 8: 304-315.
  135. Hoglen NC, Chen LS, Fisher CD, Hirakawa BP, Groessl T, Contreras PC. Characterization of IDN-6556 (3-[2-(2-tert-butyl-phenylaminooxalyl)-amino]-propionylamino]-4-oxo-5-(2,3,5,6-tetrafluoro-phenoxy)-pentanoic acid): a liver-targeted caspase inhibitor. J Pharmacol Exp Ther 2004; 309: 634-640.
  136. Doitsh G, Galloway NL, Geng X, et al. Cell death by pyroptosis drives CD4 T-cell depletion in HIV-1 infection. Nature 2014; 505: 509-514.
  137. Schulz JB, Weller M, Matthews RT, et al. Extended therapeutic window for caspase inhibition and synergy with MK-801 in the treatment of cerebral histotoxic hypoxia. Cell Death Differ 1998; 5: 847-857.
  138. Yang L, Sugama S, Mischak RP, et al. A novel systemically active caspase inhibitor attenuates the toxicities of MPTP, malonate, and 3NP in vivo. Neurobiol Dis 2004; 17: 250-259.
  139. Deckwerth TL, Adams LM, Wiessner C, et al. Long-term protection of brain tissue from cerebral ischemia by peripherally administered peptidomimetic caspase inhibitors. Drug Develop Res 2001; 52: 579-586.
  140. Kuwano K, Kunitake R, Maeyama T, et al. Attenuation of bleomycin-induced pneumopathy in mice by a caspase inhibitor. Am J Physiol Lung Cell Mol Physiol 2001; 280: L316-L325.
  141. Montolio M, Tellez N, Biarnes M, Soler J, Montanya E. Short-term culture with the caspase inhibitor z-VAD.fmk reduces beta cell apoptosis in transplanted islets and improves the metabolic outcome of the graft. Cell Transplant 2005; 14: 59-65.
R e c e i v e d : December 15, 2014
A c c e p t e d : April 27, 2015
Author’s address: Dr. Eva Matalova, Institute of Animal Physiology and Genetics, CAS, v.v.i., 97 Veveri Street, 602 00 Brno, Czech Republic. e-mail: matalova@iach.cz