Next Article in Journal
A Comparison between Patient- and Physician-Reported Late Radiation Toxicity in Long-Term Prostate Cancer Survivors
Next Article in Special Issue
KIT Mutations Correlate with Higher Galectin Levels and Brain Metastasis in Breast and Non-Small Cell Lung Cancer
Previous Article in Journal
Harnessing Liquid Biopsies to Guide Immune Checkpoint Inhibitor Therapy
Previous Article in Special Issue
BRCA 1/2 Germline Mutation Predicts the Treatment Response of FOLFIRINOX with Pancreatic Ductal Adenocarcinoma in Korean Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Current Impact of Inhibitors of Apoptosis Proteins and Their Repression in Cancer

by
Pierina Cetraro
1,
Julio Plaza-Diaz
2,3,4,*,
Alex MacKenzie
3,5 and
Francisco Abadía-Molina
6,7,*
1
Research and Advances in Molecular and Cellular Immunology, Center of Biomedical Research, University of Granada, Armilla, 18016 Granada, Spain
2
Department of Biochemistry and Molecular Biology II, School of Pharmacy, University of Granada, 18071 Granada, Spain
3
Children’s Hospital of Eastern Ontario Research Institute, Ottawa, ON K1H 8L1, Canada
4
Instituto de Investigación Biosanitaria IBS.GRANADA, Complejo Hospitalario Universitario de Granada, 18014 Granada, Spain
5
Department of Cellular and Molecular Medicine, Faculty of Medicine, University of Ottawa, Ottawa, ON K1H 8M5, Canada
6
Institute of Nutrition and Food Technology “José Mataix”, Biomedical Research Center, University of Granada, Armilla, 18016 Granada, Spain
7
Department of Cell Biology, School of Sciences, University of Granada, 18071 Granada, Spain
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(7), 1671; https://doi.org/10.3390/cancers14071671
Submission received: 31 January 2022 / Revised: 19 March 2022 / Accepted: 23 March 2022 / Published: 25 March 2022
(This article belongs to the Special Issue Inside Cancer Genomics: From Structure to Therapy)

Abstract

:

Simple Summary

The Inhibitor of Apoptosis (IAP) family of proteins has emerged as a potential pharmacological target in cancer. Abnormal expression of IAPs can lead to dysregulated cell suicide, promoting the development of different pathologies. In several cancer types, members of this protein family are overexpressed while their natural antagonist (Smac) appears to be downregulated, contributing to the acquisition of resistance to traditional therapy. The development of compounds that mimic the action of Smac showed promise in the re-sensitization of the cell suicide defense mechanism in cancer cells, particularly in combination with other treatments. Interaction with other molecules, such as tumor necrosis factor-α, in the tumor microenvironment reveals a complex interplay between IAPs and cancer.

Abstract

The Inhibitor of Apoptosis (IAP) family possesses the ability to inhibit programmed cell death through different mechanisms; additionally, some of its members have emerged as important regulators of the immune response. Both direct and indirect activity on caspases or the modulation of survival pathways, such as nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), have been implicated in mediating its effects. As a result, abnormal expression of inhibitor apoptosis proteins (IAPs) can lead to dysregulated apoptosis promoting the development of different pathologies. In several cancer types IAPs are overexpressed, while their natural antagonist, the second mitochondrial-derived activator of caspases (Smac), appears to be downregulated, potentially contributing to the acquisition of resistance to traditional therapy. Recently developed Smac mimetics counteract IAP activity and show promise in the re-sensitization to apoptosis in cancer cells. Given the modest impact of Smac mimetics when used as a monotherapy, pairing of these compounds with other treatment modalities is increasingly being explored. Modulation of molecules such as tumor necrosis factor-α (TNF-α) present in the tumor microenvironment have been suggested to contribute to putative therapeutic efficacy of IAP inhibition, although published results do not show this consistently underlining the complex interaction between IAPs and cancer.

1. Introduction

Many cancers demonstrate evasion of apoptosis enhancing cell survival [1]. Alteration in structure, expression, or even epigenetic regulation of genes, such as TP53 or BCL2 and the caspase-family of genes (or their products) leading to a dysregulated apoptotic process, are classically involved in this hallmark of cancer [2]. Dysregulated apoptosis is often observed in other pathological conditions, such as neurodegenerative disorders (Alzheimer’s Disease, motor system disorders, stroke and many others) or autoimmune diseases (type I diabetes, multiple sclerosis, rheumatoid arthritis and more) [3,4]. Many molecular pathways participating in this dysregulation have been associated with resistance to anticancer therapy that are based on the induction of endogenous cell death mechanisms [5]. Therefore, new therapeutic approaches aiming at the re-engagement of the apoptotic process by resistant cells are promising strategies to combat cancer.
The Inhibitor of Apoptosis Protein (IAP) family is a group of apoptosis negative regulators, characterized by the presence of at least one copy of the Baculovirus IAP Repeat (BIR) domain in their N-terminal portion, rather than by sharing a common function [6]. The human IAP family is composed of eight members: NAIP, cIAP1, cIAP2, XIAP, Survivin, Bruce/Apollon, ML-IAP/Livin and ILP-2, although XIAP, cIAP1 and cIAP2 are the most extensively studied [7,8,9,10]. First identified in viruses, they received the name of inhibitor of apoptosis proteins due to the function they exert on host cells [11]. In this regard, the process can be regulated both directly on caspases, through the blockage of substrate entry and ubiquitin-dependent mechanisms, or in an indirect manner through the regulation of nuclear factor kappa-light-chain-enhancer of activated B cell (NF-κB) canonical and non-canonical signaling cascades [12]. In addition to BIR domains, this function as an inhibitor of apoptosis is supported by the presence of various other domains which are not conserved among the different family members. For that reason, it is not surprising that IAPs have additional functions, participating in other processes, such as cell-cycle regulation or cell migration [13,14,15]. In this regard, the multiplicity of pathways in which IAPs participate and, most importantly, their abnormal function, due to dysregulation in IAP expression, may add to the pathogenesis of diseases, such as cancer. Importantly, in these pathologies, including breast, colorectal and hematological cancers, the eight identified members are often overexpressed. As outlined above, cancers are not the only pathologies affected by IAP dysregulation. In fact, different IAPs seem to play important roles in the correct function of the immune system. In this sense, aberrant expression of IAPs have been observed to be present in both primary and secondary immunodeficiencies and inflammasome-related pathologies [16].
On the other hand, second mitochondrial-derived activator of caspases (Smac), the second mitochondrial-derived activator of caspases, also known as DIABLO, is a mitochondrial protein released to the cytosol upon apoptotic stimuli that antagonizes IAPs endogenously, leading to the activation of caspases and the promotion of programmed cell death in normal conditions [17]. Given that in some cancers, Smac is downregulated along with its inherent IAP inhibitory structural properties, peptides mimicking its activity have been developed. These compounds are commonly known as Smac mimetics and have been systematically tested during the past decade. In spite of promising pre-clinical results, further clinical trials evaluating the anticancer activity of Smac mimetics/IAP antagonists as a monotherapy failed to provide a homogeneous pattern of response. Instead, these studies showed a limited scope of action in only a subset of cancer types and patients. Additionally, Smac mimetics’ killing efficacy was suggested to be dependent on other molecules, such as interferons (IFNs) or tumor necrosis factor-α (TNF-α) [18]. This, again, pointed to the well-established complex nature of IAPs’ biology, emphasized by the fact that the inhibition of apoptosis seems to be only one of this class of protein’s many functions. Despite all that, when tested in combination with different anticancer agents, Smac mimetic-based therapy showed synergizing effects and led to remission in many cases [19,20].
Recently, some investigations have reinforced the fact that these cell death pathways (apoptosis, pyroptosis, and necroptosis) could interact with each other [21]. Pyroptosis is characterized by diminution of plasma membrane integrity and induced by activation of so-called inflammasome sensors, such as the Nod-like receptor (NLR) family, the Pyrin receptor, and DNA receptor Absent in Melanoma 2 [21,22], and necroptosis, which is a complementary type of regulated cell death, mimicking features of necrosis and apoptosis [23]. Indeed, the simultaneous engagement of three processes has been termed PANoptosis [22].
In this work, IAP structure, function, and role in cancer will be addressed, as well as anticancer therapies based on their inhibition to promote apoptosis. In addition, possible reasons for their modest effects, shall be presented incorporating new findings which may contribute to understanding of this phenomenon. Furthermore, the current state of combination therapy involving Smac mimetics with traditional agents as well as with immunotherapy, will be reviewed.

2. IAPs

The first human IAP was described in spinal muscular atrophy (SMA), when it was found to be partially deleted or mutated in type I SMA patients. Together with the causal SMN gene, NAIP was proposed to cause or contribute to SMA pathogenesis, which included dysregulated motor neuron apoptosis [24]. The following year, the following three human IAPs were identified: XIAP, cIAP1 and cIAP2 [25]. Soon after, survivin was identified in cancer and lymphoma, Bruce/Apollon in brain cancer, livin/ML-IAP as part of a homology search and ILP-2 in testis.
BIR domains are essential zinc-binding folds which mediate protein-protein interaction [26]. Consecutive BIR domains are suggested to increase affinity to their ligand and their binding properties are conserved in those molecules that harbor more than one [27]. XIAP, cIAP1 and cIAP2 possess three BIR domains. In these cases, BIR2 and BIR3 bind to IAP binding motif (IBM)-containing proteins, such as caspases, while BIR1 interacts with other signaling intermediates, such as TAB1 or TRAF1.
Additional domains are also observed in some IAPs (Figure 1). XIAP, cIAP-1 and cIAP-2 have RING (Really Interesting New Gene) domains. These contribute to ubiquitin-protein ligase (E3) activity which promotes the addition of ubiquitin residues to lysines or methionines on proteins [28]. CIAP-1 and c-IAP2 also possess a caspase-recruitment domain (CARD) which seems to participate in the regulation of E3 activity by preventing RING-mediated dimerization as well as E2 binding [14]. NAIP does not present CARD or RING domains. Instead, NAIP displays a nucleotide-binding oligomerization domain (NOD), which, together with the BIR domains, allows its participation in the inflammasome assembly in cooperation with NLRC4. NAIP also possesses a C-terminal leucine-rich (LRR) domain. Furthermore, other domains in IAPs are the evolutionarily conserved UBA (ubiquitin-associated) domains, which enable IAPs to bind to Lys-linked polyubiquitination [29].

2.1. Anti-Apoptotic Mechanisms

2.1.1. Caspase Inhibition

IAPs participate in the inhibition of apoptosis through several mechanisms allowing their involvement in both extrinsic and intrinsic apoptotic programs. The first mechanism discovered was the inhibition of caspase activity [11]. However, under physiological conditions, IAPs are not enough to evade apoptosis, since the accumulation of apoptotic signals rapidly decides the cell’s fate. In turn, IAPs set an inhibitory threshold that caspases must overcome to result in cell death [30].
As caspase inhibitors, IAPs can reportedly either block substrate entry or add ubiquitin (Ub) chains on caspases targeting them for proteasomal degradation. According to Eckelman et al. [31], XIAP is the only direct caspase inhibitor in a strict biochemical sense since it inhibits their action through physical contact. Caspase-3 and -7 are blocked, as the linker region between the BIR1 and BIR2 domains of XIAP bind to the catalytic site of the cysteine proteases, preventing substrate entry. In the meantime, BIR2 can have contact with the IAP-binding motif (IBM) of the protein. Binding to both the catalytic pocket and the IBM results in an effective dock and inhibition of the effector caspases. Caspase 9 inhibition takes place when the BIR3 domain binds to the homodimerization surface of the cysteine protease, interfering with the production of a catalytic pocket [31].
CIAPs, on the other hand, inhibit caspases-3 and -7 by targeting them for proteasomal degradation through the addition of poly-Ub chains, by the action of their RING finger domains [32]. Knockout experiments in mice suggested overlapping activities of both cIAP1 and cIAP2, with cIAP1 having more biological impact. In this regard, later experiments showed viability in XIAP and cIAP1 null mice, with a marked upregulation of cIAP2 observed in double knockouts [33].
In addition to E3 ligase activity targeting other proteins for degradation, the three IAPs also share autodegradative capacity [34]. This function, observed in healthy cells, where RING bearing IAPs are downregulated, has been proposed to restrict levels of accidentally released IBM proteins by inducing co-degradation. This way explain normal phenotypes in mice deficient for these proteins [20]. However, in humans, defective expression of XIAP, due to deleterious mutations, is found to cause X-linked lymphoproliferative syndrome (XLP) type 2 [35,36].
Bruce/Apollon can modulate caspase-3, caspase-7 and caspase-9. For instance, it can polyubiquitylate caspase-9, when ectopically expressed in vitro [37,38,39].
NAIP is capable of inhibiting effector caspases-3 and -7. Interestingly, it can also process caspase-9 once it is in the apoptosome complex, through the action of the BIR3 domain in a unique way, different from the other IAPs’ modes of action [40].
Both survivin and livin are normally expressed during fetal development and not in adult tissues [41]. Consistent with its cell cycle-dependent expression, survivin is translocated from the cytoplasm to the nucleus during the G2/M phase. It is believed that the E3 activity of both survivin’s and livin’s RING domain might be responsible for their subcellular localization [42]. As for their anti-apoptotic activities, survivin and livin contain one BIR domain that interacts with both caspase-3 and -7, regarding the former, and with caspase-3, -7 and -9, respectively [15].
In addition to direct function as caspase inhibition, the IAP family modulates apoptotic events modulating the signaling cascade of NF-κB pathways leading to cell survival. These pathways are associated with innate and adaptive immune responses, cell survival, proliferation, growth, or motility, depending on the cell type.

2.1.2. Regulation of Survival Pathways

The family of transcription factors known as NF-κB comprises five different proteins belonging to the Rel Homology Domain (RHD)-containing family: RelA (p65), RelB, c-Rel, p105/50 (NF-kB1) and p100/52 (NF-kB1). Only three of the members (RelA, RelB and c-Rel) are produced in their mature forms, while the other two (NF-kB1 and NF-kB2) require a proteolytic cleavage of the precursor’s C-terminal portion. These later molecules do not contain transactivation domains (TAD), which are required for the activation of NF-κB genes; so, they rely on interaction with the other members. In order to become active, different members combine to form dimers. Different signals leading to NF-κB formation combined with different dimer combinations account for the modulation of the different NF-κB transcriptional programs. The two main signaling cascades that activate NF-κB can be classified either as classical (canonical) or alternative (non-canonical) pathways. Importantly, both processes are governed by the addition or deletion of Ub residues on complexes that lead to the activation of the kinases controlling NF-κB activation [43].
The NF-κB canonical pathway is classically activated upon stimulation of TNFR1, IL-1E1, TLRs, BCR or TCR. It depends on the assembly of complex I and relies on the phosphorylation of IkBα (NF-κB inhibitor) by IKK. Complex I, composed of TNFR1, TRADD, RIPK1, TRAF2 and cIAP1 is formed upon TNF-α binding to TNFR. CIAP1 ubiquitylates many components of the complex amongst which RIPK1 is essential. Once ubiquitylated, these components recruit LUBAC, TAK1-TAB2-TAB3 and IKKγ-IKKα-IKKβ. LUBAC, in turn, conjugates M1-linked Ub chains onto IKK which stabilizes the complex. IKK then phosphorylates NF-kB1β, targeting it for proteasomal degradation. NF-κB (p50p65) translocates to the nucleus and transcribes survival and inflammation-related genes. In addition, it promotes the expression of cIAP1, cIAP2, and XIAP. In the absence of IAPs, complex II is formed. This new complex is derived from complex I when it detaches from the receptor. The formation of complex II is dependent on RIPK1 and leads to the recruitment and activation of caspase-8, leading to apoptosis [44].
The NF-κB alternative pathway can be initiated by the stimulation of BAFFR, CD40, HTLV or EBV and activates non-canonical NF-κB complexes. In resting conditions, cIAPs target NIK for ubiquitylation and proteasomal degradation. This is possible through the action of TRAF2 and TRAF3. However, upon substrate binding after recognition of CD40L by CD40, the ligand-receptor complex recruits TRAF3-TRAF2-cIAP for ubiquitylation of TRAF3. Ubiquitylated TRAF3 undergoes proteasomal degradation impeding NIK elimination. Accumulation of NIK activates IKKα which, in turn, phosphorylates NF-κB 2 (p100), transforming it into p52 and allowing nuclear translocation of the p52-RelB heterodimer [45,46].
In such a way, cIAPs exert both positive and negative regulation on NF-κB activation (Figure 2). Therefore, in cells are deficient for cIAPs, the classical NF-κB pathway is proapoptotic, halting the expression of target genes, including IAPs or inflammatory cytokines. In this scenario, the positive feedback loop, in which the production of IAPs and the NF-κB pathway participate, stops. However, in these same conditions, the alternative pathway is activated, due to lack of destruction of NIK, which allows activation of non-canonical NF-κB complexes. Since constitutive NF-κB activation and promoted inflammation have critical roles in tumor development, and have been observed in several cancer types, the regulation of said pathways has been suggested as a therapeutical approach. Along these lines, realization that IAPs play critical roles in the modulation of said pathways opens new doors in terms of new access points to their regulation.

3. IAPs in Cancer

Traditional cancer chemotherapies are frequently based on the induction of apoptosis in cancerous cells. Many molecular mechanisms can be involved in resistance to chemotherapeutics, amongst which the evasion of apoptosis is frequently observed. Indeed, the avoidance of programmed cell death is a characteristic hallmark in cancers. Dysregulation in pathways that modulate apoptosis-like JNK, which affects p53 expression, or Notch-1, is found to promote this effect. Mutations on genes coding proteins involved in the apoptotic process, or overexpression of pro-apoptotic proteins, are also common [47]. Consistently, observation that IAPs are overexpressed in many hematological and solid cancers suggests their direct or indirect participation in their refractory nature. Also, different levels of IAPs’ expression in distinct patient groups have hinted at a possible involvement in treatment response and have been suggested to possess great potential as prognostic markers. As a result, new therapies targeting IAPs’ expression have been designed to re-sensitize cancerous cells to apoptosis. The latest evidence supporting the role of IAPs in some death-resistant cancers and cancer progression are reviewed next.

3.1. Acute Myeloid Leukemia (AML)

Studies evaluating the role of different IAPs in AM have revealed heterogeneous patient profiles while analyses of IAPs’ expression have failed to provide strong evidence that establishes IAPs as potent individual prognostic markers. In the early 2000s, two studies found a correlation between lower levels of XIAP protein with longer survival [48,49]. However, a third study showed no correlation between the levels of XIAP protein expression and survival was found [50]. In 2007, Hess et al. [51], analyzed pro- and anti-apoptotic gene expression in AML patients, documenting a three-gene, (including BIRC3 (c-IAP2), expression signature associated with poor overall survival (OS) [51]. High levels of ML-IAP/Livin protein have also been correlated with poor OS [52]. A later study, by Pluta et al. [53], evaluated correlations both between XIAP, c-IAP1, c-IAP2 and survivin with each other and with Smac/Diablo in newly diagnosed AML patients. Although XIAP showed strong correlations with c-IAP1 and c-IAP2 and the latter two with each other, no correlation with survivin was observed. Low levels of IAPs were associated with 100% complete remission (CR), while high levels of either one, two or three IAPs significantly reduced the percentage of patients attaining CR. On the other hand, a short OS was influenced by older age (>50), poor risk karyotype, low levels of Smac/DIABLO and high levels of survivin [53]. Along these lines, survivin expression was observed in 98% of the samples and protein and mRNA levels were shown to be higher than any other IAP measured in these patients. Taken together, results from Tamm et al. [48], Carter et al. [50], and Pluta et al. [53], appear to suggest an influence of survivin in OS that is stronger than that of XIAP, although further studies assessing how the presence of individual IAPs affect treatment and CR achievement are clearly required [48,50,53].
Importantly, different authors failed to find similar results when studying younger AML patients. Indeed, despite the classical assumption that childhood cancers are similar to adult ones, evidence against this model is becoming more obvious. Childhood AML was recently defined as a significantly different disease from adult AML at the genetic level [54]. However, genes encoding IAPs were not included in the genomic analysis performed on the infant cohorts as part of the Therapeutically Applicable Research to Generate Effective Treatments (TARGET) initiative. Therefore, over time, more and more studies evaluating childhood malignancies as independent diseases have started to appear. Along these lines, some studies have assessed the role of IAPs’ expression on survival and therapeutic response in children. Over the past decade, high levels of XIAP were found to positively correlate with risk groups, worse response to chemotherapy and shorter overall and relapse-free survival [49]. Livin and survivin levels were recently analyzed in two AML subtypes, acute promyelocytic (APL) and non-promyelocytic (non-APL). Survivin-negative patients were observed to exhibit longer survival independently of the AML group. On the other hand, livin negative patients presented longer survivals but only in the APL group. Furthermore, risk of relapse was associated with higher expression of survivin but not livin, although both correlated with higher levels of primary WBC. Taken together, none of them were identified as independent prognostic factors, although associations with other AML poor prognostic factors were evident [55]. While these findings agree with those of Makhlouf et al. [56] and El-Mesallamy et al. [52], they stand in strong contrast to the results presented by Sung et al., Carter et al, Moore et al. or Ibrahim et al. [57,58,59,60].
The overexpression of apollon was correlated with an unfavorable prognosis in pediatric patients, a finding first observed by Sung et al. [60] and later confirmed by Ismail et al. [61]. In both studies, apollon was identified as a promising prognostic factor due to its association with a higher risk of relapse, worse response to chemotherapy and decreased survival. Indeed, cutoff values for the expression of the BIRC6 gene were observed to be potent prognostic metrics for both poor clinical outcomes and worse response to therapy [60,61].
All in all, although the cited studies point towards associations between IAPs and AML, larger cohorts, stratification by relevant factors such as disease subtype, cytogenetic characteristics or age, and unified methodology must be analyzed in future studies. Additionally, data obtained from mRNA levels must be assessed and compared to those of protein in both adult and child populations in order to validate these IAPs as prognostic factors.

3.2. Chronic Lymphocytic Leukemia (CLL)

Chronic lymphocytic leukemia is the most common leukemia in adults. CLL slowly develops for extended periods of time before showing any symptoms, allowing patients to have normal lives during the early stages. However, in contrast to acute leukemias, which develop much faster, it is harder to cure. CLL is a heterogeneous disease characterized by the accumulation of monoclonal mature CD5+ B cells in lymphoid organs, bone marrow and peripheral blood. The accumulation of CLL cells has classically been suggested to be caused by the inhibition of spontaneous apoptosis, rather than by cell proliferation. In turn, given CLL cells fail to engage apoptotic programs in vivo although not in vitro, as observed by Collin et al. [62], it becomes apparent that the tumor microenvironment might have a role in these differences [62]. In fact, spontaneous apoptosis has been observed to be prohibited in patients with progressive disease compared to those with stable disease, which also correlated with higher expression of cIAP1, cIAP2, XIAP and survivin [63]. Lower Smac/DIABLO protein levels also correlated with progressive disease. Additionally, Grzybowska et al. [63] found the co-expression of survivin and cIAP1to be a negative prognostic factor, since patients presenting it showed shorter OS [63].
Despite the classical definition of CLL as a non-proliferative disease, some studies have paid attention to the proliferative nature of CLL cells found in ‘pseudo-follicles’ formed in lymphoid organs [64]. These cells are known to receive support from accessory cells, such as monocyte-derived nurse-like cells, follicular dendritic cells, BMSCs or T lymphocytes, which provide pro-survival and proliferative signals. Indeed, several pro-survival signaling pathways, including NF-kB, are known to be constitutively activated in CLL [65]. It is believed that such a microenvironment is accountable for resistance to chemotherapeutics. Growing evidence indicates that CLL proliferative cells might be responsible for disease relapse after treatment, so any approach targeting said microenvironment could mean a complete clear out of malignant cells [66]. Purroy et al. [67] sensitized resistant CLL cells to cytotoxic agents by inhibiting survivin expression and inducing apoptosis in an ex vivo model of this microenvironment [67]. Likewise, Zhu et al. [68] achieved sensitization of CLL cells to chemotherapy by modulating the expression of XIAP, among other antiapoptotic proteins, with microRNAs shown to be downregulated in ‘p53 wild-type’ patients from a Chinese cohort [68].

3.3. Colorectal Carcinoma

Colorectal cancer (CRC), also known as bowel cancer, was the third most diagnosed cancer, and the second most fatal, worldwide in 2018, according to “Las cifras del cancer” report from the SEOM (Sociedad Española de Oncología Médica). Although it is mostly a curable disease, when detected at early stages, when symptoms arise, cancer may already have advanced. Lack of activity, obesity, high-fat diet as well as alcohol and smoking are all risk factors for sporadic CRC. Altered molecular pathways documented in CRC have delineated a large number of potential biomarkers. However, given that current screening methods present specificity, sensitivity and cost limitations, make it has proven difficult to rapidly advance towards personalized and targeted therapies. Additionally, studies aiming for validation of these markers show small sample sizes, have data interpretation limitations and lack standard methodologies causing reproducibility challenges [69]. Despite these barrier, promising advances are being made. EGFR-targeted therapies use the KRAS gene as a predictive marker of response, and anti-VEGF is sometimes used in metastatic disease in combination with chemotherapy [69].
Survivin and cIAP-2 expression were distinctive in elderly groups with CRC compared to younger patients, which suggested age-related differences in CRC. Interestingly, XIAP expression was comparable in normal and cancerous tissues in both young and older patients [70]. The location of both cIAP-1 and cIAP-2 was assessed by Ponnelle et al. [71] who found that, although the proteins were present in both nucleus and cytoplasm, cIAP-1 was most frequently expressed in the nucleus (85%) and cIAP-2 in the cytoplasm (82%) in CRC samples [71]. However, nuclear expression of cIAP-2 was associated with lymphoid infiltrate in the stroma, suggesting a nuclear localization role of both proteins in the pathogenesis of CRC. Similarly, Karasawa et al. [72] observed higher cIAP-2 expression in stage II CRC with lymph node metastasis [72]. These patients also presented early recurrence after fluorouracil-based chemotherapy. Interestingly, lymph node metastases with low levels of cIAP-2 and TUCAN (CARD8) showed a larger percentage of five-year survival in a 2005 study of IAPs’ expression [73].
Correlation between XIAP expression and clinical features or prognosis in CRC have yielded conflicting results with some studies reporting no differential expression between cancerous and surrounding normal tissues, while others observed both higher protein and mRNA expression in the former [70,74]. Guoan et al. [75] showed XIAP staining with immunocytochemistry in the cytoplasm of CRC cells [75]. Also, a significant correlation between higher levels of XIAP and tumor differentiation, venous invasion and Duke’s staging was found. Furthermore, XIAP expression levels varied inversely with different survival indexes, such as PFS or OS, in most studies [74]. Takeuchi et al. [76] described the Akt signaling pathway as an important cascade through which XIAP appeared to be upregulated in C-Met overexpressing CRC cells [76].
Interestingly, livin expression significantly correlated with longer survival in a study by Lee et al. [74] in contrast to what was observed for XIAP, while Takeuchi et al. [76] did not find mRNA or protein overexpression in malignant cells [74,76].
BRUCE/Apollon mRNA was also found to be overexpressed in CRC tissues compared with non-neoplastic samples through cDNA microarrays [77] correlating with unfavorable clinical features. Smac/DIABLO levels have also been evaluated with respect to their prognostic potential in CRC and one study reported its decreased expression as an independent factor for poor prognosis [78].

3.4. Breast Carcinoma

In 2020, breast cancer was the most commonly diagnosed cancer worldwide, surpassing lung cancer. It is the most common cancer among women and the second most common cause of mortality in this population. However, the establishment of prevention programs, although still challenging, has allowed an increase in early detection and survival of patients [79].
In the last few decades, the discovery of hormonal alterations in breast cancer patients and the application of targeted therapies in populations with specific molecular characteristics have brought therapeutic promise. In terms of IAPs, overexpression in cancerous tissues has been generally observed in comparison to normal tissues. A trend of increasing expression in more advanced diseases is also apparent. Pluta et al., [80] evaluated XIAP, cIAP1, cIAP2 and survivin expression in breast cancer samples at diagnosis. The expression of XIAP and survivin was higher in more advanced cancers. Expression of cIAP-2 was higher in node-positive breast cancer, although cIAP-1 levels did not correlate with clinicopathological features of this malignancy. Higher expression of XIAP was found in most patients compared to control. Its overexpression was also associated with advanced cancer [80]. Yang et al. [81], observed an increase in XIAP expression in higher grades of ductal invasive breast carcinoma and ductal breast carcinoma in situ [81]. Similarly, immunohistochemical assays confirmed its expression in 84% of breast invasive ductal carcinomas with high immunoscores [82]. The distinction between cytoplasmic and nuclear localization has also been made. Zhang et al. [82], reported differential subcellular expressions in breast cancer samples when cytoplasmic XIAP was observed in all the studied cases but nuclear staining in only 43% [82]. Likewise, Xu et al. [83], found higher cytoplasmic expression compared to normal tissues, while nuclear expression showed no difference [83]. Cytoplasmic localization correlated to HER-2 expression status, and mutant-type (P53) status, and was considered to be a prognostic biomarker for basal-like breast cancer. Patients with higher tumor levels of XIAP showed an increased risk of relapse, as observed in database analysis. Zhang et al. [82], reported a significant relationship between nuclear staining of XIAP protein and shorter OS while Pluta et al. [80], observed associations between shorter PFS and XIAP protein expression [80,82]. However, Xu et al. [83], did not find a correlation with DFS or OS [83]. Tumor size, extranodal extension, triple-negative status and poorly differentiated subtypes showed direct associations with XIAP expression in the Middle Eastern population. This study showed a significant relationship with the PI3K pathway, p-AKT, Ki-67 and PARP. XIAP was proven to be an independent prognostic marker [84].
Similar to XIAP expression, survivin expression has been correlated with worse clinicopathological features, such as metastasis, and advanced stage, or size, of the tumor. In addition, ER and progesterone receptor-negative hormonal status was observed in to correlate with survivin level [85]. However, previous studies had not found any correlation with said features [86]. In particular, Adamkov et al. [87], suggested that the disagreement observed between different studies could have an origin in the subcellular localization of survivin [87]. Their experiments led to the determination of survivin expression as a potential prognostic factor for ductal breast carcinoma. A meta-analysis encompassing 15 studies, with a total of 2202 breast cancer patients, confirmed significant associations between positive expression of survivin and worse OS [88].
High expression of livin has been associated with highly-invasive breast cancer cells in opposition to a lower expression in non-invasive cells [13]. Li et al. [13], described the role of livin in cancer cell migration and invasion by means of the activation of Akt signaling and the induction of EMT in vitro and in vivo [13]. Furthermore, as in other malignancies, Smac/DIABLO protein was observed to be lowered in breast cancer samples compared to control [89].
Finally, despite the clear tendency of IAP expression to correlate with poor prognosis, Pluta et al. [80], who observed XIAP, cIAP1, cIAP2 and survivin expression in breast cancer samples at diagnosis, its effect on survival could not be confirmed [80]. IAPs’ patterns of expression in other malignancies, such as bladder, lung or renal cancers, are similar to those presented above and have been further developed by Fulda and Vucic [90] and Che et al. [91]. In glioblastomas, livin is correlated with worse PFS and OS rates. C-IAP1 and cIAP2 have been found to be overexpressed in osteosarcoma, and XIAP’s expression levels were high in esophageal, ovarian and pancreatic cancers [92,93]. Table 1 summarizes the IAPs and cancer.

4. Smac Mimetics

4.1. Smac/DIABLO

Smac, also known as DIABLO (Direct IAP Binding protein with Low pI), is a mitochondrial protein encoded in the nuclear genome. Although it is synthesized as a much longer precursor, in its mature form, it is composed of 184 amino acids. Upon induction of the intrinsic apoptotic pathway, mitochondria release Smac from their intermembrane space into the cytosol along with cytochrome C [17,94,95]. Apoptosis is a tightly regulated process in which the balance between different pro- and anti-apoptotic proteins decides the cell’s fate. Programmed cell death occurring through the intrinsic pathway is modulated by the Bcl-2 family of proteins, which govern mitochondrial membrane permeability [96]. Further down the pathway, IAPs have critical roles on caspase inhibition, setting an inhibitory threshold for these enzymes. Both IAPs’ direct caspase inhibition and poly-Ub-dependent caspase degradation can be easily overcome after massive release of IBM-containing proteins from mitochondria, including Smac/DIABLO. Dimeric Smac/DIABLO interacts with BIR2 and BIR3 IAP domains, ultimately leading to proteolytic cleavage and activation of procaspase-3 and procaspase-9, and the induction of caspase-3 enzymatic activity [97].

4.2. Smac Mimetics

Smac/DIABLO interaction with IAPs requires its first four N-terminal residues, also known as AVPI (Ala-Val-Pro-Ile). As previously outlined, Smac/DIABLO and IAPs regulate each other. IAPs auto-ubiquitylate when Smac/DIABLO induces their RING domain’s E3 activity. However, when overexpressed, IAPs can inhibit apoptosis as observed in cancers, contributing to the acquisition of resistance to first-line treatment traditionally based on inducing DNA damage. Inhibition of IAPs in experimental models with IAP overexpression achieved sensitization to apoptosis suggesting a potential therapeutic application [20,98,99,100,101]. Similarly, the use of exogenous Smac/DIABLO protein, both in vitro and in vivo, in several resistant cancer types, such as melanoma, neuroblastoma or glioblastoma, promoted the engagement of extrinsic and intrinsic apoptotic pathways [102]. However, Smac/DIABLO is a large protein that can have significant off target effects when delivered to patients at clinical levels. Therefore, compounds mimicking Smac/DIABLO’s antagonistic activity over IAPs have been developed in an effort to induce apoptosis in IAP-dependent apoptotic-resistant cells. The AVPI fragment, the minimal fragment conserving Smac’s inhibitory activity, has been systematically modified to optimize its pharmacological properties [103]. A total of four generations of these compounds, known as Smac mimetics, have been developed. GDC-0512, LCL16, SM-406/AT-406, Birinapant or ASTX660 stand out as promising compounds in current clinical evaluation.

4.2.1. Smac Mimetics as Single Agents

Here, the development and effectiveness of the most clinically relevant Smac mimetics during pre-clinical experimentation will be described, as well as the results that preliminary investigation in the clinical contexts.
GDC-0152, is described as a pan-IAP antagonist because it has high affinities for cIAP-1, c-IAP2, XIAP-BIR3 and livin [104]. GDC-0152’s in vitro effectiveness was demonstrated on A2058 melanoma and different glioblastoma (GL261, U87MG, GBM6 and GBM9) cell lines, in which it successfully induced c-IAP degradation, activation of caspase-3 and -7 and apoptosis [93]. MDA-MB231 breast cancer xenograft mouse models benefited from antitumor activity in vivo when GDC-0152 was administered orally as a single agent. Also, U87MG orthotopic xenografts demonstrated delayed tumor growth upon administration of GDC-0152. The only trial assessing its safety and pharmacokinetics was terminated in 2017 due to patient or efficacy-unrelated matters (NCT00977067).
LCL161 is a peptidomimetic being evaluated in several clinical trials. It is orally available and capable of binding XIAP, c-IAP1 and c-IAP2 with IC50 of 52.7, 10.4 and 12.9 nM, respectively [105]. It was assessed in leukemia and hepatocellular carcinoma cell lines inducing cell death. Interestingly, in vitro studies with multiple myeloma cell lines observed resistance to LCL161, while in vivo they were sensitive to it [106,107,108]. Response to LCL161 was limited in 23 cell lines and 46 xenograft models of pediatric cancers [109]. A phase I study on solid tumors revealed good tolerability but poor efficacy [110]. Although the stable response was observed in 19% of patients, no objective response was observed. Additionally, the maximum tolerated dose (MTD) was determined at 1800 mg when higher doses provoked symptoms of cytokine release syndrome (CRS) in 6% of the patients: vomiting, fatigue, nausea and anorexia. However, cytokine levels, such as TNF-α augmented in circulation and cIAP1 degradation, were observed. Lack of tumoral TNF-α production and sensitivity was identified as a possible cause for the low efficacy obtained. On the other hand, an interesting thing was found in triple-negative breast cancer with a specific gene signature characterized by positive expression of TNF-α [111]. Promising results were observed when evaluated in combination with paclitaxel; although toxicity was highlighted as an important and limiting factor in the therapeutical scope.
Evidence of Birinapant’s efficacy has been modest but promising. According to Benetatos et al. [112], only 16% of malignancies were shown to be sensitive to Birinapant on a large-scale cellular screen. Human melanoma cell lines 451 Lu and 1025 Lu were resistant to Birinapant in vitro. Surprisingly, in vivo assessment of these same cell lines as xenotransplants showed their decreased growth. Similarly, one-third of the 50 patient-derived xenotransplants of ovarian, colorectal and melanoma cell lines studied acquired growth inhibition. Again, such differences between in vivo and in vitro results were attributed to higher expression of TNF in the tumor’s microenvironment [112]. Clinical evaluation of this compound was initially on advanced solid tumors or lymphoma (NCT00993239), where MTD was established at 47 mg/m2. Adverse effects were evident for higher doses and comprised headaches, nausea, vomiting and Bell’s palsy in 2 of the 3 patients receiving the highest dose. Similar to LCL161, no complete or partial response was observed in patients, although stable disease was achieved in 27%. In a different study for relapsed AML or MDS, one case of Bell’s palsy was also observed but, at the same time, there was a reduction of blasts in the bone marrow. Moreover, in a relapsed platinum-resistant, or refractory epithelial ovarian cancer phase II trial (NCT01681368) no clinical benefit was observed and therefore, the trial was terminated, despite observation of potent on-target inhibition of IAPs. Moreover, a trial assessing its efficacy on advanced ovarian, fallopian tube and peritoneal cancer had to be canceled early due to lack of clinical benefit (NCT01681368). Similarly, NCT01486784, evaluating Birinapant in AML, ALL and Myelodysplasic Syndrome never initiated its phase II portion, due to early termination. All in all, Birinapant showed moderate antitumoral activity, but efficacy was far from ideal.
Debio 1143/AT-406 was firstly described in 2011 shown to mimic AVPI’s binding to BIR3. It induces cIAP1 degradation and binds to XIAP and cIAP-2. Researchers assessed this compound effectiveness as a single agent in more than 100 human cancer cell lines, although only 15% of the cell lines were responsive to the treatment, MDA-MB-231 breast and SK-OV-3 ovarian cancer lines demonstrated high levels of cell death. Furthermore, xenograft models of MDA-MB-231 cell line had efficient induction of caspase-8 dependent apoptosis with a reduction in cIAP1 levels [113]. In the first-in-man study evaluating Debbio1143 in advanced cancer, results were very modest and were not able to shed light on a recommendable dose to work with in further studies despite its good orally bioavailability. Very limited antitumor activity was hinted at, rather than observed, through the degradation of cIAP-1 [114]. Therefore, its use in combination with other therapies was proposed as a follow-up.
ASTX660 antagonizes cIAP-1, ciap-2 and XIAP through binding to BIR3. In vitro evaluation of its efficacy has demonstrated direct XIAP antagonism and depletion of cIAP-1 and cIAP-2 activating the alternative NF-κB pathway and leading to TNF-α dependent apoptosis in MDA-MB-231, HEK293, A375 and SK-MEL-28 cell lines, among others [115]. In this case, the first-in-human phase I study in advanced cancer or lymphoma succeeded in establishing an MTD of 210 mg/day. Severe adverse events included anemia, and increased lipase and lymphopenia, which were observed, but only in seven patients. As for its efficacy against IAPs, it successfully inhibited cIAP-1 in peripheral blood mononuclear cells and reduced skin lesions in patients with cutaneous T-cell lymphoma [116]. Phase II studies are still ongoing, but results obtained in the previous phase I are promising and support further evaluation (NCT02503423). Additionally, a recent clinical trial is currently recruiting participants to evaluate ASTX660′s safety and efficacy in patients with refractory or relapsing T-cell lymphoma (NCT04362007).

4.2.2. The Role of TNF-α in the Efficacy of Smac Mimetics

The binding of TNF-α to TNFR can initiate the extrinsic apoptotic pathway. As outlined above, cIAPs have important roles in the regulation of the signaling cascade following the binding of TNF-α to its receptor. Several studies have revealed the importance of TNF-α in the effectiveness of Smac mimetics treatment. To sensitize cells to apoptosis, some cancers require production and response to TNF-α. For instance, LCL161 has been suggested to require the TNF-α pathway and TNF-α production to promote apoptosis [117]. Cancers showing no response to, or no expression of, TNF-α, have been observed to be resistant to Smac mimetics treatment in some circumstances. Therefore, some therapeutic approaches are based on the induction of TNF-α expression by tumor cells. One hypothesis argues that if tumors have access to TNF-α, the suppression of IAPs will halt the NF-κB classical pathway, re-routing it towards caspase-8 dependent apoptosis. At the same time, the lack of said cIAPs would allow the NF-κB alternative pathway to produce further TNF-α, which would feed an autocrine cycle leading to cell death (Figure 3).
Smac mimetics EAG40730 and ASTX660 have proven to induce cIAP autoubiquitination, allowing NIK accumulation and alternative NF-κB pathway activation. As a consequence, TNF-α provoked apoptosis [118]. Similar results have been obtained when assessing BV-6’s and Debbio 1143’s efficacies [119,120]. Hence, the administration of exogenous TNF-α has been evaluated. For safety reasons, due to the high toxicity of systemic delivery of TNF-α, targeted delivery methods had to be developed. The use of viral vectors and the induction of parallel pathways have been considered. In addition, radiation therapy appears to induce TNF-α production, and, in turn, TNF-α enhances radiation-mediated killing. Along these lines, the results obtained by Krepler et al. [121] are consistent with this role of TNF-α. They observed synergism between Birinapant and exogenously administered TNF-α on antitumoral activity over several melanoma cell lines [121]. Overall, Smac mimetics’ effects are generally understood to be enhanced by pro-inflammatory cytokines, such as TNF-α or TRAIL [46,122,123,124]. This phenomenon links IAPs to immunity, although it was not the first evidence of said relationship. This interesting association with immune activity will be further explored below.
Cancer showing resistance to chemotherapy, which is the first line of treatment for many cancer types, has also shown dysregulation in IAP expression. Therefore, inhibition of IAPs is being trialed in combination with chemotherapy. However, some cancers can be both responsive and non-responsive to Smac mimetics, irrespective of their TNF-α status. For instance, NSCLC cell lines were sensitized to chemotherapy by the SM JP1201 even when TNF-α expression was not high [125]. Later, Langemann et al. [126], showed that the effect of LCL161 on vincristine-induced neuroblastoma apoptosis was independent of NF-κB and TNF-α [126]. Similarly, Birinapant efficacy was observed to be TNF-α independent in SUM149 and SUM190 inflammatory breast cancer cell lines [127].
In addition to apoptosis, necrosis can also be promoted by IAPs. It can start from TNFR1 signaling, which can induce both apoptosis and necroptosis. In some cells, alterations in the caspase-8-dependent apoptotic pathway, provoked by caspase inhibitors or genetic defects, can switch the cell death to necrosis, depending on the levels of RIP3. BV6 and Birinapant are both capable of inducing TNF-induced cell death. Before being used as clinical agents, these Smac mimetics were used in the study of necrosis [83]. Indeed, according to Yabal et al. [128], XIAP protein might be involved in RIP1 and RIP3 ubiquitylation in the necrosome but, instead of promoting it, they regulate it, since deletion of XIAP renders elevated RIP1 ubiquitylation [128]. Apoptotic/necrotic death induced by Smac mimetics and TNF-α crucially depends on RIPK1 because RIPK1 deficient MEFs, and cancer cells in which RIPK1 is depleted, are resistant to TNFR1 killing following treatment with Smac mimetics [46,123]. Finally, other pathways leading to apoptosis, such as that derived from endoplasmic reticulum (ER) stress induction, appear to be, in part, modulated by the presence of IAPs too. Thus, Smac mimetics treatment might exert its function in a wider spectrum of signaling pathways than previously anticipated [129].

4.2.3. The Role of Endoplasmic Reticulum (ER) Stress in IAP Inhibition

In cancer, different factors, including hypoxia or genomic instability, increase the demand for protein production, resulting in accumulation of proteins, which end up promoting ER stress [130]. Likewise, ER-stress-derived-NF-κB activation, and consequent inflammation, have been linked to numerous cancer types.
During the last decades, the Unfolded Protein Response (UPR) to ER stress has been systematically studied and associated with cancer pathogenesis [131]. It is composed of three effector mechanisms that can lead to adaptation, survival or apoptosis, respectively, and depends on stimulus and cell type. Furthermore, although it is normally started as a response to ER stress to alleviate potential proteotoxicities, its pro-survival capacity can contribute to tumorigenesis and cancer progression [132,133]. For instance, ER stress has been described as participating in the acquisition of drug resistance in AML, where ER stress appears to be transmitted through extracellular vesicles among leukemic and bone marrow resident cells. This phenomenon forces the remodeling of the bone marrow niche contributing to AML expansion and is potentially involved in drug resistance [129]. In addition, the engagement of the UPR adaptive response to ER stress is known to provoke drug resistance in other malignancies, such as pancreatic and breast cancers.
Interestingly, among the different pathways elicited as part of the pro-survival UPR, some have been observed to somewhat regulate IAPs’ expression [134]. Nevertheless, while adaptive UPR has been shown to upregulate cIAP1, cIAP2, and XIAP through PERK (one of the sensors of the UPR), chronic ER stress was observed to downregulate XIAP protein levels, promoting a switch towards the apoptotic mechanism [134]. This downregulation is also mediated by PERK, which halts global protein translation after activation, and the activation of ATF4. Additionally, Hu et al. [135], demonstrated that ER stress might upregulate IAPs in different cancer cells through the PI3K/Akt pathway, which could be elicited by Ca2+ imbalances normally found in such conditions [135]. The ablation of IAPs in Hu and colleagues’ experiments did not completely decrease cell survival, indicating the contribution of other pathways in IAP upregulation. However, there is limited information regarding the contribution of both mechanisms to actual increases in IAP expression in specific cancer types. Therefore, further research is required to elucidate the contribution of each mechanism to both apoptosis and therapy resistance. Altogether, available data suggest a strong link between IAP-mediated apoptosis inhibition and ER stress induction, although it remains unclear whether the UPR apoptosis program contributes or not (and to what extent) to cell death promoted by Smac mimetics treatment [135]. In this regard, it should be noted that the evaluation of the Smac mimetic JP1202 on NSCLC cell line described TNF-α independent pathways as being responsible for cell death induced by the treatment. The researchers highlighted the participation of ER stress as observed by the activation of caspase-4 [125].

5. Combination Therapy

Cells with dysregulated apoptosis cannot induce cell death; they continue dividing. Upon IAPs’ downregulation by antisense oligonucleotides both in vitro and in vivo, pre-clinical resistant cell models are able to engage in the apoptotic programs [36]. As already stated, IAPs’ inhibition with Smac mimetics has demonstrated some clinical relevance, although their effectiveness has been limited when assessed as individual agents. In parallel, some chemotherapeutic approaches, despite being the main, or even the only, therapy for some cancers, have modest efficacy. For example, this is the case with NSCLC, where paclitaxel only shows modest efficacy and data is limited [136]. Consequently, combination approaches have been evaluated. The effect of LCL161 on NSCLC cell lines was studied in combination with paclitaxel, where increased expression of TNF-α accompanied by degradation of cIAPs, were observed to precede caspase-8 dependent cell death in vitro [136]. Xenograft models indicated similar results, where the combination therapy worked better than any of the two as single agents. Taking into account TNF-α expression as an important factor in antitumoral response, a clinical study on triple-negative breast cancer (TNBC) used TNF-α expression profiles as a biomarker to evaluate patients’ response to LCL161 and paclitaxel. Promising results showed a possible correlation between clinical efficacy and TNF-α positive TNBC patients with both LCL161 and paclitaxel. However, due to higher discontinuation provoked by AEs in the TNF-α negative TNBC patient group, lower efficacy observed could be attributed to either the consequent lower paclitaxel exposure or their TNF-α status [111]. The authors were therefore prevented from giving a more concrete opinion about the role of TNF-α in TNBC. Birinapant has been assayed with several chemotherapeutics due to its good tolerability. In vitro studies have revealed higher efficacy in combination with chemotherapy than that of any of the agents alone [137]. Combination with docetaxel achieved tumor volume reduction in HNSCC xenograft models, although survival was not affected. Carboplatin and Birinapant sensitized cells to caspase-8 dependent killing in High-Grade Serous Ovarian Cancer in both in vitro and in vivo models [138]. Interestingly, combination with gemcitabine showed no involvement of TNF-dependent death mechanisms. Debio 1143 is capable of synergizing with taxanes, among other agents, against lung adenocarcinoma cells, even reducing the tumor volume in mice xenografts [139]. A phase I/II trial of Debio in combination with cisplatin in patients with locally advanced squamous cell carcinoma of the head and neck provided encouraging results including high OS, CR and PFS rates [140]. Furthermore, it demonstrated efficacy in mice models of carboplatin-resistant ovarian cancer in combination with carboplatin, which is currently the first line of treatment, despite eventual relapse. Survival was improved and correlated with XIAP downregulation [141].
Other therapeutical approaches have been assessed in combination with Smac mimetics including TRAIL, radiation and monoclonal antibodies have been assayed together with different Smac mimetics. Importantly, and given the recent evidence of IAPs’ role in the immune system, different immunotherapeutic approaches have been suggested to potentially benefit from Smac mimetics’ co-treatment. Immunotherapy-based treatment harnesses the immune system to fight cancerous cells. On their own, immunotherapy approaches can show impressive efficacy. Despite limited effects on large patient cohorts and adverse effects, as severe as induction of autoimmunity, cancers meeting very specific criteria seem highly sensitive to this approach. For example, immune checkpoint inhibitors (ICIs) have already succeeded in stimulating T cell responses against tumors. Nevertheless, an important observation was made by Kearney et al. [142], when PD-L1 expression seemed to inhibit TNF-α production in cytotoxic lymphocytes, avoiding a possible paracrine interaction with tumoral cells and the subsequent TNF-mediated cell death [142]. Interestingly, Birinapant was not able to increase said TNF-α levels, although it did sensitize tumor cells to TNF-mediated death induced by cytotoxic lymphocytes. These results pointed out the potential benefits of using ICIs together with Smac mimetics as a way of increasing TNF-αproduction, while forcing the TNFR pathway towards a caspase-8 dependent death. Accordingly a number of studies, proving and revising the role of IAPs on immune modulation through the activation of NF-κB pathways have been published [143].
Combination therapy of Smac mimetics with ICIs have provided good results in pre-clinical models. PD-1 and CLTA-4 extended survival rates of glioblastoma mouse models when administered in combination with LCL161 or Birinapant, despite not showing any activity on their own. Combined blockade with anti-PD1 and anti-CLTA-4 was assessed as in previous studies [144]. The addition of Smac mimetics to said combination led to a 100% durable cure rate. Interestingly, effective penetration of blood-brain barrier was achieved. Likewise, models of multiple myeloma and mammary tumors obtained very high percentages for durable care rates [145]. This effect was ascribed to the presence of cytotoxic T cells, and decrease in immunosuppressive CD4+ T cells, and seemed to involve IFNs and TNF-α pathways. Indeed, Bv6 was found to synergize with IFN-γ in the promotion of necroptosis in a TNF-α independent manner [18].
The combination of Birinapant with several other therapies was being evaluated but different trials had to be terminated early due to Birinapant’s lack of clinical benefit (NCT02147873), (NCT01681368). Similarly, a trial assessing its combination with pembrolizumab in solid tumors (colorectal, ovarian and cervical carcinomas, among others) had to be canceled, based on futility analysis (NCT02587962). LCL161 evaluation with an anti-PD1 monoclonal antibody is currently active in a phase I trial against colorectal cancer, TNBC, NSCLC and renal carcinoma (NCT02890069), and completed for multiple myeloma, although results are not yet available (NCT03111992).
Furthermore, the realization that Smac mimetics may have different roles on the immune system, apart from cIAPs’ involvement in NF-κB activation, has catalyzed different uses of these agents. INKT cells can be stimulated to produce cytokines upon Smac mimetics’ stimulation [146]. M2 macrophages can transform into M1 inflammatory macrophages, as observed in mice models of ovarian carcinoma and sarcoma cells. Inflammation-induced assayed Smac mimetics promoted the production of TNF-α, IL-1β and IFNγ, which supposedly provoked macrophage phenotypic change. The immune response, started as a result of SM83 administration (a bivalent Smac mimetics), terminated in necrotic death of cancerous cells [147]. According to Rigaud et al. [36], patients with XLP resulting from XIAP mutations were observed to have low numbers of natural killer T-cells (NKTs), which suggested a role for this IAP in lymphocyte homeostasis in vivo [36]. Also, CARD domains present in cIAPs appear to cooperate with c-Myc in proliferation/migration of cancers [14]. And, importantly, Smac mimetics have demonstrated effects on T cell cytokine secretion, enhanced in vitro priming of naïve T cells by peripheral blood mononuclear cells and dendritic cell maturation, suggesting possible roles in modulation of immune responses [148].
Lastly, given that both the enhancement and the inhibition of stress responses in cancer can achieve reduction in tumor growth, efforts in both directions have been made. The results of some of these approaches are collected in a review from Pakos-Zebrucka et. al. [149]. Despite the above, a direct link between ER stress and Smac mimetics’ resistance has not been established yet and, therefore, there is no information about the study of ER stress modulators and Smac mimetics combined therapy. However, there is evidence pointing towards a possible relationship between ER stress and Smac mimetics’ effectivity that cannot be ignored by future evaluations. Figure 4 summarizes that association.

6. Conclusions

IAPs play a crucial role in setting a threshold for cells to undergo programmed cell death upon apoptotic stimuli. The mechanisms through which these molecules exert their functions are varied, ranging from direct physical interaction with caspases to the regulation of cell survival pathways. In this sense, overexpression of IAPs has been observed to directly correlate with cancer progression and decreased survival. Therefore, therapeutic approaches based on mode action of IAPs’ endogenous antagonist, Smac/DIABLO, have been proposed. Smac mimetics are thus developed in an effort to decrease IAPs’ levels observed in cancers and other diseases in which their expression is dysregulated. In cancer, cells exhibiting upregulated IAPs are sometimes observed to show resistance to chemotherapy and other therapies based on the induction of apoptosis. Along these lines, reduction of IAP levels can potentially re-sensitize cells to apoptosis. This phenomenon is usually described as being enhanced by the presence of TNF-α, as TNF-α combined with the inhibition of IAPs achieves a switch in the TNFR-dependent NF-κB pathway towards caspase-8 dependent cell death. At the same time, NIK accumulation in NF-κB alternative pathway feeds a positive loop, allowing inflammation and antitumoral effect in some cases. However, and despite supporting evidence, in other cases, death promoted by administration of Smac mimetics is not dependent on TNF-α or NF-kB. The cause of such discrepancy is still unknown but further studies are needed to determine in which cancers the TNF-α status can or cannot be a reliable prognostic/response marker. In parallel, different studies have demonstrated that ER stress can both induce and reduce the expression of IAPs. In addition, in some cancers, ER stress has been identified as responsible for the acquisition of resistance to chemotherapeutic agents. Together, a constitutive ER stress that promotes upregulation of IAPs could, at least in part, explain the loss of apoptotic capability. However, the total scope of the ER stress contribution on IAPs’ upregulation in cancer is not known, and neither are its connections with other routes also participating in IAPs’ overexpression. It might be worth studying the nature of Smac mimetics’ resistance. It would not be surprising to find that Smac mimetics’ resistance is not, at least in part, promoted by ER stress induction of IAPs’ overexpression, such that previously established doses of Smac mimetics might not be sufficient to overcome inhibition of apoptosis. This might be of particular interest in the context of TNF-α non-responsive cancers. Ultimately, a picture of multifactorial IAP overexpression and acquisition of resistance emerges. Therefore, it might be interesting to further investigate the pathways interacting to produce IAPs’ upregulation. Overall, this supports the evaluation of combination therapy in cancers, in which no therapeutic approach has provided positive results on its own.
Finally, not only apoptosis but also necrosis is promoted as part of Smac mimetics treatment, unpacking new layers of IAPs’ relationship with tumoral development/progression.
Furthermore, recent evidence demonstrating new immune-modulatory functions of IAPs, in both innate and adaptive immunity, opens new doors in terms of therapeutic opportunities. It is not only the inhibition of apoptosis, but also how these proteins interact with immune cells, that create a complex and fairly opaque mode of action and therefore make it difficult to predict Smac mimetics therapy outcomes. In this sense, further investigations should undertake, in terms of deciphering the complex web of interactions between Smac mimetics and inflammation, cell maturation and, in general, immune modulation for specific cancers. The elucidation of these converging routes would signify a huge advance in prognostic evaluation, which could be incorporated in pre-treatment screening to improve targeted therapies.

Author Contributions

P.C., J.P.-D., A.M. and F.A.-M. participated in the bibliographic search, discussion, and writing of the manuscript. P.C. and F.A.-M. designed the work. J.P.-D. and A.M. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

J.P.-D. is part of the “UGR Plan Propio de Investigación 2016” and the “Excellence actions: Unit of Excellence on Exercise and Health (UCEES), University of Granada”. J.P.-D. is supported by a fellowship to postdoctoral researchers at foreign universities and research centers from the “Fundación Ramón Areces”, Madrid, Spain.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Zhang, J.Y. Apoptosis-based anticancer drugs. Nat. Rev. Drug Discov. 2002, 1, 101. [Google Scholar] [CrossRef] [PubMed]
  3. Hashemi, M.; Kroczak, T.J. Apoptosis and autoimmune disease. Curr. Med. Chem. Anti-Inflamm. Anti-Allergy Agents 2005, 4, 429–437. [Google Scholar] [CrossRef]
  4. Mattson, M.P. Apoptosis in neurodegenerative disorders. Nat. Rev. Mol. Cell Biol. 2000, 1, 120–130. [Google Scholar] [CrossRef]
  5. Rebucci, M.; Michiels, C. Molecular aspects of cancer cell resistance to chemotherapy. Biochem. Pharmacol. 2013, 85, 1219–1226. [Google Scholar] [CrossRef]
  6. Silke, J.; Vaux, D.L. Two kinds of BIR-containing protein-inhibitors of apoptosis, or required for mitosis. J. Cell Sci. 2001, 114, 1821–1827. [Google Scholar] [CrossRef]
  7. Ambrosini, G.; Adida, C.; Altieri, D.C. A novel anti-apoptosis gene, survivin, expressed in cancer and lymphoma. Nat. Med. 1997, 3, 917–921. [Google Scholar] [CrossRef]
  8. Ashhab, Y.; Alian, A.; Polliack, A.; Panet, A.; Yehuda, D.B. Two splicing variants of a new inhibitor of apoptosis gene with different biological properties and tissue distribution pattern. FEBS Lett. 2001, 495, 56–60. [Google Scholar] [CrossRef] [Green Version]
  9. Chen, Z.; Naito, M.; Hori, S.; Mashima, T.; Yamori, T.; Tsuruo, T. A human IAP-family gene, apollon, expressed in human brain cancer cells. Biochem. Biophys. Res. Commun. 1999, 264, 847–854. [Google Scholar] [CrossRef]
  10. Richter, B.W.; Mir, S.S.; Eiben, L.J.; Lewis, J.; Reffey, S.B.; Frattini, A.; Tian, L.; Frank, S.; Youle, R.J.; Nelson, D.L. Molecular cloning of ILP-2, a novel member of the inhibitor of apoptosis protein family. Mol. Cell. Biol. 2001, 21, 4292–4301. [Google Scholar] [CrossRef] [Green Version]
  11. Crook, N.E.; Clem, R.J.; Miller, L.K. An apoptosis-inhibiting baculovirus gene with a zinc finger-like motif. J. Virol. 1993, 67, 2168–2174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Gyrd-Hansen, M.; Meier, P. IAPs: From caspase inhibitors to modulators of NF-κB, inflammation and cancer. Nat. Rev. Cancer 2010, 10, 561–574. [Google Scholar] [CrossRef] [PubMed]
  13. Li, F.; Yin, X.; Luo, X.; Li, H.-Y.; Su, X.; Wang, X.-Y.; Chen, L.; Zheng, K.; Ren, G.-S. Livin promotes progression of breast cancer through induction of epithelial–mesenchymal transition and activation of AKT signaling. Cell. Signal. 2013, 25, 1413–1422. [Google Scholar] [CrossRef] [PubMed]
  14. Lopez, J.; John, S.W.; Tenev, T.; Rautureau, G.J.; Hinds, M.G.; Francalanci, F.; Wilson, R.; Broemer, M.; Santoro, M.M.; Day, C.L. CARD-mediated autoinhibition of cIAP1′s E3 ligase activity suppresses cell proliferation and migration. Mol. Cell 2011, 42, 569–583. [Google Scholar] [CrossRef] [Green Version]
  15. Shin, S.; Sung, B.-J.; Cho, Y.-S.; Kim, H.-J.; Ha, N.-C.; Hwang, J.-I.; Chung, C.-W.; Jung, Y.-K.; Oh, B.-H. An anti-apoptotic protein human survivin is a direct inhibitor of caspase-3 and-7. Biochemistry 2001, 40, 1117–1123. [Google Scholar] [CrossRef]
  16. Hrdinka, M.; Yabal, M. Inhibitor of apoptosis proteins in human health and disease. Genes Immun. 2019, 20, 641–650. [Google Scholar] [CrossRef]
  17. Zou, H.; Henzel, W.J.; Liu, X.; Lutschg, A.; Wang, X. Apaf-1, a human protein homologous to C. elegans CED-4, participates in cytochrome c–dependent activation of caspase-3. Cell 1997, 90, 405–413. [Google Scholar] [CrossRef] [Green Version]
  18. Cekay, M.J.; Roesler, S.; Frank, T.; Knuth, A.-K.; Eckhardt, I.; Fulda, S. Smac mimetics and type II interferon synergistically induce necroptosis in various cancer cell lines. Cancer Lett. 2017, 410, 228–237. [Google Scholar] [CrossRef]
  19. Rathore, R.; McCallum, J.E.; Varghese, E.; Florea, A.M.; Busselberg, D. Overcoming chemotherapy drug resistance by targeting inhibitors of apoptosis proteins (IAPs). Apoptosis 2017, 22, 898–919. [Google Scholar] [CrossRef]
  20. Srinivasula, S.M.; Ashwell, J.D. IAPs: What′s in a name? Mol. Cell 2008, 30, 123–135. [Google Scholar] [CrossRef]
  21. Bertheloot, D.; Latz, E.; Franklin, B.S. Necroptosis, pyroptosis and apoptosis: An intricate game of cell death. Cell. Mol. Immunol. 2021, 18, 1106–1121. [Google Scholar] [CrossRef] [PubMed]
  22. Christgen, S.; Zheng, M.; Kesavardhana, S.; Karki, R.; Malireddi, R.K.S.; Banoth, B.; Place, D.E.; Briard, B.; Sharma, B.R.; Tuladhar, S.; et al. Identification of the PANoptosome: A Molecular Platform Triggering Pyroptosis, Apoptosis, and Necroptosis (PANoptosis). Front. Cell Infect. Microbiol. 2020, 10, 237. [Google Scholar] [CrossRef] [PubMed]
  23. Dhuriya, Y.K.; Sharma, D. Necroptosis: A regulated inflammatory mode of cell death. J. Neuroinflamm. 2018, 15, 199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Roy, N.; Mahadevan, M.S.; McLean, M.; Shutter, G.; Yaraghi, Z.; Farahani, R.; Baird, S.; Besner-Johnston, A.; Lefebvre, C.; Kang, X. The gene for neuronal apoptosis inhibitory protein is partially deleted in individuals with spinal muscular atrophy. Cell 1995, 80, 167–178. [Google Scholar] [CrossRef] [Green Version]
  25. Listen, P.; Roy, N.; Tamai, K.; Lefebvre, C.; Baird, S.; Cherton-Horvat, G.; Farahani, R.; McLean, M.; Mackenzie, A.; Korneluk, R.G. Suppression of apoptosis in mammalian cells by NAIP and a related family of IAP genes. Nature 1996, 379, 349–353. [Google Scholar] [CrossRef]
  26. Wei, Y.; Fan, T.; Yu, M. Inhibitor of apoptosis proteins and apoptosis. Acta Biochim. Et Biophys. Sin. 2008, 40, 278–288. [Google Scholar] [CrossRef] [Green Version]
  27. Huang, Y.; Rich, R.L.; Myszka, D.G.; Wu, H. Requirement of both the second and third BIR domains for the relief of X-linked inhibitor of apoptosis protein (XIAP)-mediated caspase inhibition by Smac. J. Biol. Chem. 2003, 278, 49517–49522. [Google Scholar] [CrossRef] [Green Version]
  28. Weissman, A.M. Themes and variations on ubiquitylation. Nat. Rev. Mol. Cell Biol. 2001, 2, 169–178. [Google Scholar] [CrossRef]
  29. Guarda, G.; So, A. Regulation of inflammasome activity. Immunology 2010, 130, 329–336. [Google Scholar] [CrossRef]
  30. Alberts, B.; Bray, D.; Hopkin, K.; Johnson, A.D.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Essential Cell Biology; Garland Science (Taylor & Francis Group): New York, NY, USA, 2014. [Google Scholar]
  31. Eckelman, B.P.; Salvesen, G.S.; Scott, F.L. Human inhibitor of apoptosis proteins: Why XIAP is the black sheep of the family. EMBO Rep. 2006, 7, 988–994. [Google Scholar] [CrossRef] [Green Version]
  32. Zheng, C.; Kabaleeswaran, V.; Wang, Y.; Cheng, G.; Wu, H. Crystal structures of the TRAF2: cIAP2 and the TRAF1: TRAF2: cIAP2 complexes: Affinity, specificity, and regulation. Mol. Cell 2010, 38, 101–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Heard, K.N.; Bertrand, M.J.; Barker, P.A. cIAP2 supports viability of mice lacking cIAP1 and XIAP. EMBO J. 2015, 34, 2393–2395. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, Y.; Fang, S.; Jensen, J.P.; Weissman, A.M.; Ashwell, J.D. Ubiquitin protein ligase activity of IAPs and their degradation in proteasomes in response to apoptotic stimuli. Science 2000, 288, 874–877. [Google Scholar] [CrossRef] [PubMed]
  35. Aguilar, C.; Latour, S. X-linked inhibitor of apoptosis protein deficiency: More than an X-linked lymphoproliferative syndrome. J. Clin. Immunol. 2015, 35, 331–338. [Google Scholar] [CrossRef]
  36. Rigaud, S.; Fondaneche, M.-C.; Lambert, N.; Pasquier, B.; Mateo, V.; Soulas, P.; Galicier, L.; Le Deist, F.; Rieux-Laucat, F.; Revy, P. XIAP deficiency in humans causes an X-linked lymphoproliferative syndrome. Nature 2006, 444, 110–114. [Google Scholar] [CrossRef]
  37. Bartke, T.; Pohl, C.; Pyrowolakis, G.; Jentsch, S. Dual role of BRUCE as an antiapoptotic IAP and a chimeric E2/E3 ubiquitin ligase. Mol. Cell 2004, 14, 801–811. [Google Scholar] [CrossRef]
  38. Hao, Y.; Sekine, K.; Kawabata, A.; Nakamura, H.; Ishioka, T.; Ohata, H.; Katayama, R.; Hashimoto, C.; Zhang, X.; Noda, T. Apollon ubiquitinates SMAC and caspase-9, and has an essential cytoprotection function. Nat. Cell Biol. 2004, 6, 849–860. [Google Scholar] [CrossRef]
  39. Qiu, X.-B.; Goldberg, A.L. The membrane-associated inhibitor of apoptosis protein, BRUCE/Apollon, antagonizes both the precursor and mature forms of Smac and caspase-9. J. Biol. Chem. 2005, 280, 174–182. [Google Scholar] [CrossRef] [Green Version]
  40. Davoodi, J.; Ghahremani, M.-H.; Es-Haghi, A.; Mohammad-Gholi, A.; MacKenzie, A. Neuronal apoptosis inhibitory protein, NAIP, is an inhibitor of procaspase-9. Int. J. Biochem. Cell Biol. 2010, 42, 958–964. [Google Scholar] [CrossRef]
  41. Chandele, A.; Prasad, V.; Jagtap, J.C.; Shukla, R.; Shastry, P.R. Upregulation of survivin in G2/M cells and inhibition of caspase 9 activity enhances resistance in staurosporine-induced apoptosis. Neoplasia 2004, 6, 29–40. [Google Scholar] [CrossRef] [Green Version]
  42. Kasof, G.M.; Gomes, B.C. Livin, a novel inhibitor of apoptosis protein family member. J. Biol. Chem. 2001, 276, 3238–3246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhang, Q.; Lenardo, M.J.; Baltimore, D. 30 years of NF-κB: A blossoming of relevance to human pathobiology. Cell 2017, 168, 37–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Dondelinger, Y.; Darding, M.; Bertrand, M.J.; Walczak, H. Poly-ubiquitination in TNFR1-mediated necroptosis. Cell. Mol. Life Sci. 2016, 73, 2165–2176. [Google Scholar] [CrossRef] [Green Version]
  45. Sun, S.C. The noncanonical NF-κB pathway. Immunol. Rev. 2012, 246, 125–140. [Google Scholar] [CrossRef] [Green Version]
  46. Vince, J.E.; Wong, W.W.-L.; Khan, N.; Feltham, R.; Chau, D.; Ahmed, A.U.; Benetatos, C.A.; Chunduru, S.K.; Condon, S.M.; McKinlay, M. IAP antagonists target cIAP1 to induce TNFα-dependent apoptosis. Cell 2007, 131, 682–693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pan, S.T.; Li, Z.L.; He, Z.X.; Qiu, J.X.; Zhou, S.F. Molecular mechanisms for tumour resistance to chemotherapy. Clin. Exp. Pharmacol. Physiol. 2016, 43, 723–737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Tamm, I.; Kornblau, S.M.; Segall, H.; Krajewski, S.; Welsh, K.; Kitada, S.; Scudiero, D.A.; Tudor, G.; Qui, Y.H.; Monks, A. Expression and prognostic significance of IAP-family genes in human cancers and myeloid leukemias. Clin. Cancer Res. 2000, 6, 1796–1803. [Google Scholar]
  49. Tamm, I.; Richter, S.; Oltersdorf, D.; Creutzig, U.; Harbott, J.; Scholz, F.; Karawajew, L.; Ludwig, W.-D.; Wuchter, C. High expression levels of x-linked inhibitor of apoptosis protein and survivin correlate with poor overall survival in childhood de novo acute myeloid leukemia. Clin. Cancer Res. 2004, 10, 3737–3744. [Google Scholar] [CrossRef] [Green Version]
  50. Carter, B.Z.; Kornblau, S.M.; Tsao, T.; Wang, R.Y.; Schober, W.D.; Milella, M.; Sung, H.G.; Reed, J.C.; Andreeff, M. Caspase-independent cell death in AML: Caspase inhibition in vitro with pan-caspase inhibitors or in vivo by XIAP or Survivin does not affect cell survival or prognosis. Blood 2003, 102, 4179–4186. [Google Scholar] [CrossRef] [Green Version]
  51. Hess, C.J.; Berkhof, J.; Denkers, F.; Ossenkoppele, G.J.; Schouten, J.P.; Oudejans, J.J.; Waisfisz, Q.; Schuurhuis, G.J. Activated intrinsic apoptosis pathway is a key related prognostic parameter in acute myeloid leukemia. J. Clin. Oncol. 2007, 25, 1209–1215. [Google Scholar] [CrossRef]
  52. El-Mesallamy, H.O.; Hegab, H.M.; Kamal, A.M. Expression of inhibitor of apoptosis protein (IAP) livin/BIRC7 in acute leukemia in adults: Correlation with prognostic factors and outcome. Leuk. Res. 2011, 35, 1616–1622. [Google Scholar] [CrossRef] [PubMed]
  53. Pluta, A.; Wierzbowska, A.; Cebula-Obrzut, B.; Pluta, P.; Stępka, K.; Szmigielska-Kapłon, A.; Grzybowska-Izydorczyk, O.; Czemerska, M.; Smolewski, P.; Wrzesien-Kus, A. Prognostic value of inhibitor of apoptosis protein family expression in patients with acute myeloid leukemia. Leuk. Lymphoma 2015, 56, 2529–2535. [Google Scholar] [CrossRef] [PubMed]
  54. Bolouri, H.; Farrar, J.E.; Triche, T., Jr.; Ries, R.E.; Lim, E.L.; Alonzo, T.A.; Ma, Y.; Moore, R.; Mungall, A.J.; Marra, M.A.; et al. The molecular landscape of pediatric acute myeloid leukemia reveals recurrent structural alterations and age-specific mutational interactions. Nat. Med. 2018, 24, 103–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Zareifar, S.; Ghorbani, S.; Monabbati, A.; Bordbar, M.R.; Zekavat, O.R.; Abdolkarimi, B.; Haghpanah, S. Expression of antiapoptotic proteins livin and survivin in pediatric AML patients, as prognostic markers. Pediatr. Hematol. Oncol. 2018, 35, 250–256. [Google Scholar] [CrossRef]
  56. Makhlouf, M.M. Survivin and cyclin E2 genes expression in a cohort of Egyptian acute leukemia patients: Clinical importance and future prospects. Cancer Biomark. 2016, 16, 181–189. [Google Scholar] [CrossRef]
  57. Carter, B.Z.; Qiu, Y.; Huang, X.; Diao, L.; Zhang, N.; Coombes, K.R.; Mak, D.H.; Konopleva, M.; Cortes, J.; Kantarjian, H.M. Survivin is highly expressed in CD34+ 38− leukemic stem/progenitor cells and predicts poor clinical outcomes in AML. Blood J. Am. Soc. Hematol. 2012, 120, 173–180. [Google Scholar] [CrossRef] [PubMed]
  58. Ibrahim, L.; Aladle, D.; Mansour, A.; Hammad, A.; Al Wakeel, A.A.; Abd El-Hameed, S.A. Expression and prognostic significance of livin/BIRC7 in childhood acute lymphoblastic leukemia. Med. Oncol. 2014, 31, 1–8. [Google Scholar] [CrossRef]
  59. Moore, A.S.; Alonzo, T.A.; Gerbing, R.B.; Lange, B.J.; Heerema, N.A.; Franklin, J.; Raimondi, S.C.; Hirsch, B.A.; Gamis, A.S.; Meshinchi, S. BIRC5 (survivin) splice variant expression correlates with refractory disease and poor outcome in pediatric acute myeloid leukemia: A report from the Children′s Oncology Group. Pediatr. Blood Cancer 2014, 61, 647–652. [Google Scholar] [CrossRef] [Green Version]
  60. Sung, K.W.; Choi, J.; Hwang, Y.K.; Lee, S.J.; Kim, H.-J.; Lee, S.H.; Yoo, K.H.; Jung, H.L.; Koo, H.H. Overexpression of Apollon, an antiapoptotic protein, is associated with poor prognosis in childhood de novo acute myeloid leukemia. Clin. Cancer Res. 2007, 13, 5109–5114. [Google Scholar] [CrossRef] [Green Version]
  61. Ismail, E.A.R.; Mahmoud, H.M.; Tawfik, L.M.; Habashy, D.M.; Adly, A.A.M.; El-Sherif, N.H.; Abdelwahab, M.A. BIRC6/Apollon gene expression in childhood acute leukemia: Impact on therapeutic response and prognosis. Eur. J. Haematol. 2012, 88, 118–127. [Google Scholar] [CrossRef]
  62. Collins, R.J.; Verschuer, L.A.; Harmon, B.V.; Prentice, R.L.; Pope, J.H.; Kerr, J.F. Spontaneous programmed death (apoptosis) of B-chronic lymphocytic leukaemia cells following their culture in vitro. Br. J. Haematol. 1989, 71, 343–350. [Google Scholar] [CrossRef] [PubMed]
  63. Grzybowska-Izydorczyk, O.; Cebula, B.; Robak, T.; Smolewski, P. Expression and prognostic significance of the inhibitor of apoptosis protein (IAP) family and its antagonists in chronic lymphocytic leukaemia. Eur. J. Cancer 2010, 46, 800–810. [Google Scholar] [CrossRef] [PubMed]
  64. Messmer, B.T.; Messmer, D.; Allen, S.L.; Kolitz, J.E.; Kudalkar, P.; Cesar, D.; Murphy, E.J.; Koduru, P.; Ferrarini, M.; Zupo, S. In vivo measurements document the dynamic cellular kinetics of chronic lymphocytic leukemia B cells. J. Clin. Investig. 2005, 115, 755–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Cuni, S.; Perez-Aciego, P.; Perez-Chacon, G.; Vargas, J.; Sanchez, A.; Martin-Saavedra, F.; Ballester, S.; Garcia-Marco, J.; Jorda, J.; Durantez, A. A sustained activation of PI3K/NF-κ B pathway is critical for the survival of chronic lymphocytic leukemia B cells. Leukemia 2004, 18, 1391–1400. [Google Scholar] [CrossRef]
  66. Herishanu, Y.; Pérez-Galán, P.; Liu, D.; Biancotto, A.; Pittaluga, S.; Vire, B.; Gibellini, F.; Njuguna, N.; Lee, E.; Stennett, L. The lymph node microenvironment promotes B-cell receptor signaling, NF-κB activation, and tumor proliferation in chronic lymphocytic leukemia. Blood J. Am. Soc. Hematol. 2011, 117, 563–574. [Google Scholar] [CrossRef]
  67. Purroy, N.; Abrisqueta, P.; Carabia, J.; Carpio, C.; Calpe, E.; Palacio, C.; Castellví, J.; Crespo, M.; Bosch, F. Targeting the proliferative and chemoresistant compartment in chronic lymphocytic leukemia by inhibiting survivin protein. Leukemia 2014, 28, 1993–2004. [Google Scholar] [CrossRef]
  68. Zhu, D.-X.; Zhu, W.; Fang, C.; Fan, L.; Zou, Z.-J.; Wang, Y.-H.; Liu, P.; Hong, M.; Miao, K.-R.; Liu, P. miR-181a/b significantly enhances drug sensitivity in chronic lymphocytic leukemia cells via targeting multiple anti-apoptosis genes. Carcinogenesis 2012, 33, 1294–1301. [Google Scholar] [CrossRef] [Green Version]
  69. Mármol, I.; Sánchez-de-Diego, C.; Pradilla Dieste, A.; Cerrada, E.; Rodriguez Yoldi, M.J. Colorectal carcinoma: A general overview and future perspectives in colorectal cancer. Int. J. Mol. Sci. 2017, 18, 197. [Google Scholar] [CrossRef] [Green Version]
  70. Endo, T.; Abe, S.; Seidlar, H.; Nagaoka, S.; Takemura, T.; Utsuyama, M.; Kitagawa, M.; Hirokawa, K. Expression of IAP family proteins in colon cancers from patients with different age groups. Cancer Immunol. Immunother. 2004, 53, 770–776. [Google Scholar] [CrossRef]
  71. Ponnelle, T.; Chapusot, C.; Martin, L.; Bonithon-Kopp, C.; Bouvier, A.M.; Plenchette, S.; Rageot, D.; Faivre, J.; Solary, E.; Piard, F. Subcellular expression of c-IAP1 and c-IAP2 in colorectal cancers: Relationships with clinicopathological features and prognosis. Pathol. Res. Pract. 2003, 199, 723–731. [Google Scholar] [CrossRef]
  72. Karasawa, H.; Miura, K.; Fujibuchi, W.; Ishida, K.; Kaneko, N.; Kinouchi, M.; Okabe, M.; Ando, T.; Murata, Y.; Sasaki, H. Down-regulation of cIAP2 enhances 5-FU sensitivity through the apoptotic pathway in human colon cancer cells. Cancer Sci. 2009, 100, 903–913. [Google Scholar] [CrossRef] [PubMed]
  73. Krajewska, M.; Kim, H.; Kim, C.; Kang, H.; Welsh, K.; Matsuzawa, S.-i.; Tsukamoto, M.; Thomas, R.G.; Assa-Munt, N.; Piao, Z. Analysis of apoptosis protein expression in early-stage colorectal cancer suggests opportunities for new prognostic biomarkers. Clin. Cancer Res. 2005, 11, 5451–5461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Lee, K.-H.; Lee, S.; Lee, H.-M.; Back, S.-C.; Cho, S.-B.; Lee, J.-H. Inhibitors of apoptosis proteins expression and their prognostic significance in colorectal carcinoma. Tic 2011, 9, 10. [Google Scholar] [CrossRef]
  75. Guoan, X.; Xiaomin, W.; Hanning, W.; Kaiyun, C.; Hao, L. Expression of X-linked inhibitor of apoptosis protein in human colorectal cancer and its correlation with prognosis. J. Surg. Oncol. 2009, 100, 708–712. [Google Scholar] [CrossRef]
  76. Takeuchi, H.; Kim, J.; Fujimoto, A.; Umetani, N.; Mori, T.; Bilchik, A.; Turner, R.; Tran, A.; Kuo, C.; Hoon, D.S. X-Linked inhibitor of apoptosis protein expression level in colorectal cancer is regulated by hepatocyte growth factor/C-met pathway via Akt signaling. Clin. Cancer Res. 2005, 11, 7621–7628. [Google Scholar] [CrossRef] [Green Version]
  77. Bianchini, M.; Levy, E.; Zucchini, C.; Pinski, V.; Macagno, C.; De Sanctis, P.; Valvassori, L.; Carinci, P.; Mordoh, J. Comparative study of gene expression by cDNA microarray in human colorectal cancer tissues and normal mucosa. Int. J. Oncol. 2006, 29, 83–94. [Google Scholar] [CrossRef]
  78. Endo, K.; Kohnoe, S.; Watanabe, A.; Tashiro, H.; Sakata, H.; Morita, M.; Kakeji, Y.; Maehara, Y. Clinical significance of Smac/DIABLO expression in colorectal cancer. Oncol. Rep. 2009, 21, 351–355. [Google Scholar] [CrossRef] [Green Version]
  79. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA A Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  80. Pluta, P.; Jeziorski, A.; Cebula-Obrzut, A.; Wierzbowska, A.; Piekarski, J.; Smolewski, P. Expression of IAP family proteins and its clinical importance in breast cancer patients. Neoplasma 2015, 62, 666–673. [Google Scholar] [CrossRef] [Green Version]
  81. Yang, L.; Cao, Z.; Yan, H.; Wood, W.C. Coexistence of high levels of apoptotic signaling and inhibitor of apoptosis proteins in human tumor cells: Implication for cancer specific therapy. Cancer Res. 2003, 63, 6815–6824. [Google Scholar]
  82. Zhang, Y.; Zhu, J.; Tang, Y.; Li, F.; Zhou, H.; Peng, B.; Zhou, C.; Fu, R. X-linked inhibitor of apoptosis positive nuclear labeling: A new independent prognostic biomarker of breast invasive ductal carcinoma. Diagn. Pathol. 2011, 6, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Xu, Y.-C.; Liu, Q.; Dai, J.-Q.; Yin, Z.-Q.; Tang, L.; Ma, Y.; Lin, X.-L.; Wang, H.-X. Tissue microarray analysis of X-linked inhibitor of apoptosis (XIAP) expression in breast cancer patients. Med. Oncol. 2014, 31, 1–9. [Google Scholar] [CrossRef] [PubMed]
  84. Hussain, A.R.; Siraj, A.K.; Ahmed, M.; Bu, R.; Pratheeshkumar, P.; Alrashed, A.M.; Qadri, Z.; Ajarim, D.; Al-Dayel, F.; Beg, S. XIAP over-expression is an independent poor prognostic marker in Middle Eastern breast cancer and can be targeted to induce efficient apoptosis. BMC Cancer 2017, 17, 640. [Google Scholar] [CrossRef] [PubMed]
  85. Youssef, N.S.; Hewedi, I.H.; Abd Raboh, N.M. Immunohistochemical expression of survivin in breast carcinoma: Relationship with clinicopathological parameters, proliferation and molecular classification. J. Egypt Natl. Canc. Inst. 2008, 20, 348–357. [Google Scholar]
  86. Tanaka, K.; Iwamoto, S.; Gon, G.; Nohara, T.; Iwamoto, M.; Tanigawa, N. Expression of survivin and its relationship to loss of apoptosis in breast carcinomas. Clin. Cancer Res. 2000, 6, 127–134. [Google Scholar]
  87. Adamkov, M.; Kajo, K.; Vybohova, D.; Krajcovic, J.; Stuller, F.; Rajcani, J. Correlations of survivin expression with clinicomorphological parameters and hormonal receptor status in breast ductal carcinoma. Neoplasma 2012, 59, 30–37. [Google Scholar] [CrossRef] [Green Version]
  88. Song, J.; Su, H.; Zhou, Y.Y.; Guo, L.L. Prognostic value of survivin expression in breast cancer patients: A meta-analysis. Tumour Biol. 2013, 34, 2053–2062. [Google Scholar] [CrossRef]
  89. Pluta, P.; Cebula-Obrzut, B.; Ehemann, V.; Pluta, A.; Wierzbowska, A.; Piekarski, J.; Bilski, A.; Nejc, D.; Kordek, R.; Robak, T. Correlation of Smac/DIABLO protein expression with the clinico-pathological features of breast cancer patients. Neoplasma 2011, 58, 430–435. [Google Scholar] [CrossRef] [Green Version]
  90. Fulda, S.; Vucic, D. Targeting IAP proteins for therapeutic intervention in cancer. Nat. Rev. Drug Discov. 2012, 11, 109–124. [Google Scholar] [CrossRef]
  91. Che, X.; Yang, D.; Zong, H.; Wang, J.; Li, X.; Chen, F.; Chen, X.; Song, X. Nuclear cIAP1 overexpression is a tumor stage-and grade-independent predictor of poor prognosis in human bladder cancer patients. Urol Oncol. 2012, 30, 450–456. [Google Scholar] [CrossRef]
  92. Ma, O.; Cai, W.-W.; Zender, L.; Dayaram, T.; Shen, J.; Herron, A.J.; Lowe, S.W.; Man, T.-K.; Lau, C.C.; Donehower, L.A. MMP13, Birc2 (cIAP1), and Birc3 (cIAP2), amplified on chromosome 9, collaborate with p53 deficiency in mouse osteosarcoma progression. Cancer Res. 2009, 69, 2559–2567. [Google Scholar] [CrossRef] [Green Version]
  93. Tchoghandjian, A.; Soubéran, A.; Tabouret, E.; Colin, C.; Denicolaï, E.; Jiguet-Jiglaire, C.; El-Battari, A.; Villard, C.; Baeza-Kallee, N.; Figarella-Branger, D. Inhibitor of apoptosis protein expression in glioblastomas and their in vitro and in vivo targeting by SMAC mimetic GDC-0152. Cell Death Dis. 2016, 7, e2325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Li, P.; Nijhawan, D.; Budihardjo, I.; Srinivasula, S.M.; Ahmad, M.; Alnemri, E.S.; Wang, X. Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade. Cell 1997, 91, 479–489. [Google Scholar] [CrossRef] [Green Version]
  95. Verhagen, A.M.; Ekert, P.G.; Pakusch, M.; Silke, J.; Connolly, L.M.; Reid, G.E.; Moritz, R.L.; Simpson, R.J.; Vaux, D.L. Identification of DIABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell 2000, 102, 43–53. [Google Scholar] [CrossRef] [Green Version]
  96. Elmore, S. Apoptosis: A review of programmed cell death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef] [PubMed]
  97. Chai, J.; Du, C.; Wu, J.-W.; Kyin, S.; Wang, X.; Shi, Y. Structural and biochemical basis of apoptotic activation by Smac/DIABLO. Nature 2000, 406, 855–862. [Google Scholar] [CrossRef] [PubMed]
  98. Kulathila, R.; Vash, B.; Sage, D.; Cornell-Kennon, S.; Wright, K.; Koehn, J.; Stams, T.; Clark, K.; Price, A. The structure of the BIR3 domain of cIAP1 in complex with the N-terminal peptides of SMAC and caspase-9. Acta Crystallogr. Sect. D Biol. Crystallogr. 2009, 65, 58–66. [Google Scholar] [CrossRef] [PubMed]
  99. Song, Z.; Yao, X.; Wu, M. Direct interaction between survivin and Smac/DIABLO is essential for the anti-apoptotic activity of survivin during taxol-induced apoptosis. J. Biol. Chem. 2003, 278, 23130–23140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Vucic, D.; Deshayes, K.; Ackerly, H.; Pisabarro, M.T.; Kadkhodayan, S.; Fairbrother, W.J.; Dixit, V.M. SMAC negatively regulates the anti-apoptotic activity of melanoma inhibitor of apoptosis (ML-IAP). J. Biol. Chem. 2002, 277, 12275–12279. [Google Scholar] [CrossRef] [Green Version]
  101. Yang, Q.-H.; Du, C. Smac/DIABLO selectively reduces the levels of c-IAP1 and c-IAP2 but not that of XIAP and livin in HeLa cells. J. Biol. Chem. 2004, 279, 16963–16970. [Google Scholar] [CrossRef] [Green Version]
  102. Fulda, S.; Wick, W.; Weller, M.; Debatin, K.-M. Smac agonists sensitize for Apo2L/TRAIL-or anticancer drug-induced apoptosis and induce regression of malignant glioma in vivo. Nat. Med. 2002, 8, 808–815. [Google Scholar] [CrossRef] [PubMed]
  103. Cong, H.; Xu, L.; Wu, Y.; Qu, Z.; Bian, T.; Zhang, W.; Xing, C.; Zhuang, C. Inhibitor of apoptosis protein (IAP) antagonists in anticancer agent discovery: Current status and perspectives. J. Med. Chem. 2019, 62, 5750–5772. [Google Scholar] [CrossRef] [PubMed]
  104. Flygare, J.A.; Beresini, M.; Budha, N.; Chan, H.; Chan, I.T.; Cheeti, S.; Cohen, F.; Deshayes, K.; Doerner, K.; Eckhardt, S.G. Discovery of a potent small-molecule antagonist of inhibitor of apoptosis (IAP) proteins and clinical candidate for the treatment of cancer (GDC-0152). J. Med. Chem. 2012, 55, 4101–4113. [Google Scholar] [CrossRef] [PubMed]
  105. Baggio, C.; Gambini, L.; Udompholkul, P.; Salem, A.F.; Aronson, A.; Dona, A.; Troadec, E.; Pichiorri, F.; Pellecchia, M. Design of potent pan-IAP and Lys-covalent XIAP selective inhibitors using a thermodynamics driven approach. J. Med. Chem. 2018, 61, 6350–6363. [Google Scholar] [CrossRef]
  106. Chen, K.-F.; Lin, J.-P.; Shiau, C.-W.; Tai, W.-T.; Liu, C.-Y.; Yu, H.-C.; Chen, P.-J.; Cheng, A.-L. Inhibition of Bcl-2 improves effect of LCL161, a SMAC mimetic, in hepatocellular carcinoma cells. Biochem. Pharmacol. 2012, 84, 268–277. [Google Scholar] [CrossRef]
  107. Weisberg, E.; Ray, A.; Barrett, R.; Nelson, E.; Christie, A.; Porter, D.; Straub, C.; Zawel, L.; Daley, J.F.; Lazo-Kallanian, S. Smac mimetics: Implications for enhancement of targeted therapies in leukemia. Leukemia 2010, 24, 2100–2109. [Google Scholar] [CrossRef] [Green Version]
  108. Chesi, M.; Mirza, N.N.; Garbitt, V.M.; Sharik, M.E.; Dueck, A.C.; Asmann, Y.W.; Akhmetzyanova, I.; Kosiorek, H.E.; Calcinotto, A.; Riggs, D.L. IAP antagonists induce anti-tumor immunity in multiple myeloma. Nat. Med. 2016, 22, 1411–1420. [Google Scholar] [CrossRef] [Green Version]
  109. Houghton, P.J.; Kang, M.H.; Reynolds, C.P.; Morton, C.L.; Kolb, E.A.; Gorlick, R.; Keir, S.T.; Carol, H.; Lock, R.; Maris, J.M. Initial testing (stage 1) of LCL161, a SMAC mimetic, by the Pediatric Preclinical Testing Program. Pediatr. Blood Cancer 2012, 58, 636–639. [Google Scholar] [CrossRef] [Green Version]
  110. Infante, J.R.; Dees, E.C.; Olszanski, A.J.; Dhuria, S.V.; Sen, S.; Cameron, S.; Cohen, R.B. Phase I dose-escalation study of LCL161, an oral inhibitor of apoptosis proteins inhibitor, in patients with advanced solid tumors. J. Clin. Oncol. 2014, 32, 3103–3110. [Google Scholar] [CrossRef]
  111. Bardia, A.; Parton, M.; Kümmel, S.; Estévez, L.G.; Huang, C.-S.; Cortés, J.; Ruiz-Borrego, M.; Telli, M.L.; Martin-Martorell, P.; López, R. Paclitaxel with inhibitor of apoptosis antagonist, LCL161, for localized triple-negative breast cancer, prospectively stratified by gene signature in a biomarker-driven neoadjuvant trial. J. Clin. Oncol. 2018, 36, 3126–3133. [Google Scholar] [CrossRef]
  112. Benetatos, C.A.; Mitsuuchi, Y.; Burns, J.M.; Neiman, E.M.; Condon, S.M.; Yu, G.; Seipel, M.E.; Kapoor, G.S.; LaPorte, M.G.; Rippin, S.R. Birinapant (TL32711), a bivalent SMAC mimetic, targets TRAF2-associated cIAPs, abrogates TNF-induced NF-κB activation, and is active in patient-derived xenograft models. Mol. Cancer Ther. 2014, 13, 867–879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Cai, Q.; Sun, H.; Peng, Y.; Lu, J.; Nikolovska-Coleska, Z.; McEachern, D.; Liu, L.; Qiu, S.; Yang, C.-Y.; Miller, R. A potent and orally active antagonist (SM-406/AT-406) of multiple inhibitor of apoptosis proteins (IAPs) in clinical development for cancer treatment. J. Med. Chem. 2011, 54, 2714–2726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Hurwitz, H.I.; Smith, D.C.; Pitot, H.C.; Brill, J.M.; Chugh, R.; Rouits, E.; Rubin, J.; Strickler, J.; Vuagniaux, G.; Sorensen, J.M. Safety, pharmacokinetics, and pharmacodynamic properties of oral DEBIO1143 (AT-406) in patients with advanced cancer: Results of a first-in-man study. Cancer Chemother. Pharmacol. 2015, 75, 851–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Ward, G.A.; Lewis, E.J.; Ahn, J.S.; Johnson, C.N.; Lyons, J.F.; Martins, V.; Munck, J.M.; Rich, S.J.; Smyth, T.; Thompson, N.T. ASTX660, a novel non-peptidomimetic antagonist of cIAP1/2 and XIAP, potently induces TNFα-dependent apoptosis in cancer cell lines and inhibits tumor growth. Mol. Cancer Ther. 2018, 17, 1381–1391. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Mita, M.M.; LoRusso, P.M.; Papadopoulos, K.P.; Gordon, M.S.; Mita, A.C.; Ferraldeschi, R.; Keer, H.; Oganesian, A.; Su, X.Y.; Jueliger, S. A phase I study of ASTX660, an antagonist of inhibitors of apoptosis proteins, in adults with advanced cancers or lymphoma. Clin. Cancer Res. 2020, 26, 2819–2826. [Google Scholar] [CrossRef] [Green Version]
  117. Firestone, B.; Conway, C.; Yang, G.; Gao, H.; Porter, D.; Slisz, J.; He, D.; Mosher, R.; Monahan, J.; Straub, C. Abstract B27: Correlation between TNFα and LCL161 anti-tumor activity in patient derived xenograft models of human cancer. Mol. Cancer Ther. 2009, 8, B27. [Google Scholar]
  118. Cheung, H.; St Jean, M.; Beug, S.; Lejmi-Mrad, R.; LaCasse, E.; Baird, S.; Stojdl, D.; Screaton, R.; Korneluk, R. SMG1 and NIK regulate apoptosis induced by Smac mimetic compounds. Cell Death Dis. 2011, 2, e146. [Google Scholar] [CrossRef]
  119. Mak, D.H.; Manton, C.; Andreeff, M.; Carter, B.Z. SMAC-Mimetic BV-6 Sensitizes Therapeutic Agents-Induced Apoptosis in AML Cells. Blood 2010, 116, 2177. [Google Scholar] [CrossRef]
  120. Matzinger, O.; Viertl, D.; Tsoutsou, P.; Kadi, L.; Rigotti, S.; Zanna, C.; Wiedemann, N.; Vozenin, M.-C.; Vuagniaux, G.; Bourhis, J. The radiosensitizing activity of the SMAC-mimetic, Debio 1143, is TNFα-mediated in head and neck squamous cell carcinoma. Radiother. Oncol. 2015, 116, 495–503. [Google Scholar] [CrossRef]
  121. Krepler, C.; Chunduru, S.K.; Halloran, M.B.; He, X.; Xiao, M.; Vultur, A.; Villanueva, J.; Mitsuuchi, Y.; Neiman, E.M.; Benetatos, C. The novel SMAC mimetic birinapant exhibits potent activity against human melanoma cells. Clin. Cancer Res. 2013, 19, 1784–1794. [Google Scholar] [CrossRef] [Green Version]
  122. Beug, S.T.; Tang, V.A.; LaCasse, E.C.; Cheung, H.H.; Beauregard, C.E.; Brun, J.; Nuyens, J.P.; Earl, N.; St-Jean, M.; Holbrook, J. Smac mimetics and innate immune stimuli synergize to promote tumor death. Nat. Biotechnol. 2014, 32, 182–190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Gaither, A.; Porter, D.; Yao, Y.; Borawski, J.; Yang, G.; Donovan, J.; Sage, D.; Slisz, J.; Tran, M.; Straub, C. A Smac mimetic rescue screen reveals roles for inhibitor of apoptosis proteins in tumor necrosis factor-α signaling. Cancer Res. 2007, 67, 11493–11498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Li, L.; Thomas, R.M.; Suzuki, H.; De Brabander, J.K.; Wang, X.; Harran, P.G. A small molecule Smac mimic potentiates TRAIL-and TNFα-mediated cell death. Science 2004, 305, 1471–1474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Greer, R.M.; Peyton, M.; Larsen, J.E.; Girard, L.; Xie, Y.; Gazdar, A.F.; Harran, P.; Wang, L.; Brekken, R.A.; Wang, X. SMAC Mimetic (JP1201) Sensitizes Non–Small Cell Lung Cancers to Multiple Chemotherapy Agents in an IAP-Dependent but TNF-α–Independent Manner. Cancer Res. 2011, 71, 7640–7648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Langemann, D.; Trochimiuk, M.; Appl, B.; Hundsdoerfer, P.; Reinshagen, K.; Eschenburg, G. Sensitization of neuroblastoma for vincristine-induced apoptosis by Smac mimetic LCL161 is attended by G2 cell cycle arrest but is independent of NFκB, RIP1 and TNF-α. Oncotarget 2017, 8, 87763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Allensworth, J.L.; Aird, K.M.; Aldrich, A.J.; Batinic-Haberle, I.; Devi, G.R. XIAP inhibition and generation of reactive oxygen species enhances TRAIL sensitivity in inflammatory breast cancer cells. Mol. Cancer Ther. 2012, 11, 1518–1527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Yabal, M.; Müller, N.; Adler, H.; Knies, N.; Groß, C.J.; Damgaard, R.B.; Kanegane, H.; Ringelhan, M.; Kaufmann, T.; Heikenwälder, M. XIAP restricts TNF-and RIP3-dependent cell death and inflammasome activation. Cell Rep. 2014, 7, 1796–1808. [Google Scholar] [CrossRef] [Green Version]
  129. Doron, B.; Abdelhamed, S.; Butler, J.T.; Hashmi, S.K.; Horton, T.M.; Kurre, P. Transmissible ER stress reconfigures the AML bone marrow compartment. Leukemia 2019, 33, 918–930. [Google Scholar] [CrossRef]
  130. Ma, Y.; Hendershot, L.M. The role of the unfolded protein response in tumour development: Friend or foe? Nat. Rev. Cancer 2004, 4, 966–977. [Google Scholar] [CrossRef]
  131. Urra, H.; Dufey, E.; Avril, T.; Chevet, E.; Hetz, C. Endoplasmic reticulum stress and the hallmarks of cancer. Trends Cancer 2016, 2, 252–262. [Google Scholar] [CrossRef]
  132. Chen, X.; Iliopoulos, D.; Zhang, Q.; Tang, Q.; Greenblatt, M.B.; Hatziapostolou, M.; Lim, E.; Tam, W.L.; Ni, M.; Chen, Y. XBP1 promotes triple-negative breast cancer by controlling the HIF1α pathway. Nature 2014, 508, 103–107. [Google Scholar] [CrossRef] [PubMed]
  133. Masouleh, B.K.; Geng, H.; Hurtz, C.; Chan, L.N.; Logan, A.C.; Chang, M.S.; Huang, C.; Swaminathan, S.; Sun, H.; Paietta, E. Mechanistic rationale for targeting the unfolded protein response in pre-B acute lymphoblastic leukemia. Proc. Natl. Acad. Sci. USA 2014, 111, E2219–E2228. [Google Scholar]
  134. Hiramatsu, N.; Messah, C.; Han, J.; LaVail, M.M.; Kaufman, R.J.; Lin, J.H. Translational and posttranslational regulation of XIAP by eIF2α and ATF4 promotes ER stress–induced cell death during the unfolded protein response. Mol. Biol. Cell 2014, 25, 1411–1420. [Google Scholar] [CrossRef] [PubMed]
  135. Hu, P.; Han, Z.; Couvillon, A.D.; Exton, J.H. Critical role of endogenous Akt/IAPs and MEK1/ERK pathways in counteracting endoplasmic reticulum stress-induced cell death. J. Biol. Chem. 2004, 279, 49420–49429. [Google Scholar] [CrossRef] [Green Version]
  136. Chu, Q.; Vincent, M.; Logan, D.; Mackay, J.A.; Evans, W.K.; Care, L.C.D.S.G.o.C.C.O.s.P.i.E.-b. Taxanes as first-line therapy for advanced non-small cell lung cancer: A systematic review and practice guideline. Lung Cancer 2005, 50, 355–374. [Google Scholar] [CrossRef]
  137. Eytan, D.F.; Snow, G.E.; Carlson, S.G.; Schiltz, S.; Chen, Z.; Van Waes, C. Combination effects of SMAC mimetic birinapant with TNF α, TRAIL, and docetaxel in preclinical models of HNSCC. Laryngoscope 2015, 125, E118–E124. [Google Scholar] [CrossRef] [Green Version]
  138. Janzen, D.; Tiourin, E.; Salehi, J.; Paik, D.; Lu, J.; Pellegrini, M.; Memarzadeh, S. RETRACTED ARTICLE: An apoptosis-enhancing drug overcomes platinum resistance in a tumour-initiating subpopulation of ovarian cancer. Nat. Commun. 2015, 6, 1–18. [Google Scholar] [CrossRef] [Green Version]
  139. Langdon, C.G.; Wiedemann, N.; Held, M.A.; Mamillapalli, R.; Iyidogan, P.; Theodosakis, N.; Platt, J.T.; Levy, F.; Vuagniaux, G.; Wang, S. SMAC mimetic Debio 1143 synergizes with taxanes, topoisomerase inhibitors and bromodomain inhibitors to impede growth of lung adenocarcinoma cells. Oncotarget 2015, 6, 37410. [Google Scholar] [CrossRef] [Green Version]
  140. Le Tourneau, C.; Tao, Y.; Gomez-Roca, C.; Cristina, V.; Borcoman, E.; Deutsch, E.; Bahleda, R.; Calugaru, V.; Modesto, A.; Rouits, E. Phase I Trial of Debio 1143, an Antagonist of Inhibitor of Apoptosis Proteins, Combined with Cisplatin Chemoradiotherapy in Patients with Locally Advanced Squamous Cell Carcinoma of the Head and Neck. Clin. Cancer Res. 2020, 26, 6429–6436. [Google Scholar] [CrossRef]
  141. Brunckhorst, M.K.; Lerner, D.; Wang, S.; Yu, Q. AT-406, an orally active antagonist of multiple inhibitor of apoptosis proteins, inhibits progression of human ovarian cancer. Cancer Biol. Ther. 2012, 13, 804–811. [Google Scholar] [CrossRef] [Green Version]
  142. Kearney, C.J.; Lalaoui, N.; Freeman, A.J.; Ramsbottom, K.M.; Silke, J.; Oliaro, J. PD-L1 and IAPs co-operate to protect tumors from cytotoxic lymphocyte-derived TNF. Cell Death Differ. 2017, 24, 1705–1716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Gyrd-Hansen, M.; Darding, M.; Miasari, M.; Santoro, M.M.; Zender, L.; Xue, W.; Tenev, T.; Da Fonseca, P.C.; Zvelebil, M.; Bujnicki, J.M. IAPs contain an evolutionarily conserved ubiquitin-binding domain that regulates NF-κB as well as cell survival and oncogenesis. Nat. Cell Biol. 2008, 10, 1309–1317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Reardon, D.A.; Gokhale, P.C.; Klein, S.R.; Ligon, K.L.; Rodig, S.J.; Ramkissoon, S.H.; Jones, K.L.; Conway, A.S.; Liao, X.; Zhou, J. Glioblastoma eradication following immune checkpoint blockade in an orthotopic, immunocompetent model. Cancer Immunol. Res. 2016, 4, 124–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Beug, S.T.; Beauregard, C.E.; Healy, C.; Sanda, T.; St-Jean, M.; Chabot, J.; Walker, D.E.; Mohan, A.; Earl, N.; Lun, X. Smac mimetics synergize with immune checkpoint inhibitors to promote tumour immunity against glioblastoma. Nat. Commun. 2017, 8, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Clancy-Thompson, E.; Ali, L.; Bruck, P.T.; Exley, M.A.; Blumberg, R.S.; Dranoff, G.; Dougan, M.; Dougan, S.K. IAP antagonists enhance cytokine production from mouse and human iNKT cells. Cancer Immunol. Res. 2018, 6, 25–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Lecis, D.; De Cesare, M.; Perego, P.; Conti, A.; Corna, E.; Drago, C.; Seneci, P.; Walczak, H.; Colombo, M.; Delia, D. Smac mimetics induce inflammation and necrotic tumour cell death by modulating macrophage activity. Cell Death Dis. 2013, 4, e920. [Google Scholar] [CrossRef] [Green Version]
  148. Knights, A.J.; Fucikova, J.; Pasam, A.; Koernig, S.; Cebon, J. Inhibitor of apoptosis protein (IAP) antagonists demonstrate divergent immunomodulatory properties in human immune subsets with implications for combination therapy. Cancer Immunol. Immunother. 2013, 62, 321–335. [Google Scholar] [CrossRef]
  149. Pakos-Zebrucka, K.; Koryga, I.; Mnich, K.; Ljujic, M.; Samali, A.; Gorman, A.M. The integrated stress response. EMBO Rep. 2016, 17, 1374–1395. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic representation of the IAP family members’ structure. The nomenclature, length, and domains of the different proteins are shown. Abbreviations: BIR (Baculoviral IAP Repeat), NOD (Nucleotide-binding and Oligomerization Domain), LRR (Leucine Rich Repeat), UBA (Ubiquitin-Associated), CARD (Caspase-Associated Recruitment Domain), RING (Really Interesting New Gene), UBC (Ubiquitin-conjugating).
Figure 1. Schematic representation of the IAP family members’ structure. The nomenclature, length, and domains of the different proteins are shown. Abbreviations: BIR (Baculoviral IAP Repeat), NOD (Nucleotide-binding and Oligomerization Domain), LRR (Leucine Rich Repeat), UBA (Ubiquitin-Associated), CARD (Caspase-Associated Recruitment Domain), RING (Really Interesting New Gene), UBC (Ubiquitin-conjugating).
Cancers 14 01671 g001
Figure 2. Participation of cIAPs in NF-κB classical and alternative pathways. E3 ligase activity from IAPs’ RING domain allow the addition of Ub residues onto proteins involved in these signaling cascades. In the classical pathway, cIAPs have a positive regulatory function while they exhibit negative regulation in the alternative pathway. Abbreviations: NF-κB (nuclear factor kappa-light-chain-enhancer of activated B cells), TNF-α (tumor necrosis factor-α), RIPK (Receptor-interacting serine/threonine-protein kinase), TRADD (TNF Receptor Associated Death Domain), TRAF (TNF Receptor Associated Factor), cIAP (cellular-Inhibitor of Apoptosis Protein), LUBAC (Linear Ubiquitination Assembly Complex), IKK (IκB Kinase), NIK (NF-κB-inducing kinase), Ub (ubiquitin).
Figure 2. Participation of cIAPs in NF-κB classical and alternative pathways. E3 ligase activity from IAPs’ RING domain allow the addition of Ub residues onto proteins involved in these signaling cascades. In the classical pathway, cIAPs have a positive regulatory function while they exhibit negative regulation in the alternative pathway. Abbreviations: NF-κB (nuclear factor kappa-light-chain-enhancer of activated B cells), TNF-α (tumor necrosis factor-α), RIPK (Receptor-interacting serine/threonine-protein kinase), TRADD (TNF Receptor Associated Death Domain), TRAF (TNF Receptor Associated Factor), cIAP (cellular-Inhibitor of Apoptosis Protein), LUBAC (Linear Ubiquitination Assembly Complex), IKK (IκB Kinase), NIK (NF-κB-inducing kinase), Ub (ubiquitin).
Cancers 14 01671 g002
Figure 3. The role of Smac mimetics in the regulation of TNF-α dynamics and the promotion of apoptosis. The inhibition of cIAPs alters NF-κB pathways. The classical NF-κB pathway is pushed towards caspase-8 dependent apoptosis through the formation of RIPK-FADD-Caspase-8 complex and it is fed by the production of TNF-α in the alternative pathway. Abbreviations: TNF-α (Tumor Necrosis Factor-α), TNFR (Tumor Necrosis Factor Receptor), cIAP (cellular- Inhibitor of Apoptosis Protein), RIPK (Receptor-interacting serine/threonine-protein kinase), FADD (Fas-associated protein with death domain), TRAF (TNF receptor associated factor), NIK (NF-κB-inducing kinase), NF-κB (nuclear factor kappa-light-chain-enhancer of activated B cells).
Figure 3. The role of Smac mimetics in the regulation of TNF-α dynamics and the promotion of apoptosis. The inhibition of cIAPs alters NF-κB pathways. The classical NF-κB pathway is pushed towards caspase-8 dependent apoptosis through the formation of RIPK-FADD-Caspase-8 complex and it is fed by the production of TNF-α in the alternative pathway. Abbreviations: TNF-α (Tumor Necrosis Factor-α), TNFR (Tumor Necrosis Factor Receptor), cIAP (cellular- Inhibitor of Apoptosis Protein), RIPK (Receptor-interacting serine/threonine-protein kinase), FADD (Fas-associated protein with death domain), TRAF (TNF receptor associated factor), NIK (NF-κB-inducing kinase), NF-κB (nuclear factor kappa-light-chain-enhancer of activated B cells).
Cancers 14 01671 g003
Figure 4. ER stress and Smac mimetics. Abbreviations, ER stress, endoplasmic reticulum stress; IAPs, inhibitor apoptosis proteins; UPR, unfolded protein response.
Figure 4. ER stress and Smac mimetics. Abbreviations, ER stress, endoplasmic reticulum stress; IAPs, inhibitor apoptosis proteins; UPR, unfolded protein response.
Cancers 14 01671 g004
Table 1. IAPs and cancer. Abbreviations: AML, acute myeloid leukemia; CLL, chronic lymphocytic leukemia; CRC, colorectal cancer; FU, fluorouracil; OS; overall survival.
Table 1. IAPs and cancer. Abbreviations: AML, acute myeloid leukemia; CLL, chronic lymphocytic leukemia; CRC, colorectal cancer; FU, fluorouracil; OS; overall survival.
Cancer TypeObservationsReferences
Acute Myeloid Leukemia (AML)
XIAPLower levels of XIAP correlate with longer survivalTamm et al. [48], Tamm et al. [49]
No correlation between XIAP protein levels and survivalCarter et al. [50]
XIAP expression strongly correlates with cIAP1 and cIAP2 but no correlation
with survivin was found. Low IAPs levels point to higher complete remission rate.
Pluta et al. [53]
SurvivinAppear to influence on the OS and it might be stronger than that from XIAP
expression.
Pluta et al. [53]
cIAP2High gene expression of cIAP2 is associated with poor OSEl-Mesallamy et al. [52]
LivinHigh level of Livin is correlated with poor OSEl-Mesallamy et al. [52]
Childhood AML
XIAPHigh levels of XIAP positively correlate with risk groups, worse response to
chemotherapy and shorter OS.
Tamm et al. [49]
LivinLivin negative patients showed longer OS in the acute promyelocytic AMLZareifar et al. [55]
SurvivinSurvivin negative patients showed longer OS Zareifar et al. [55]
ApollonOverexpression of apollon correlated with unfavorable prognosis. It was
identified as a prognostic factor.
Ismail et al. [61], Sung et al. [60]
Chronic Lymphocytic Leukemia
XIAPHigher expression of XIAP, cIAP1, cIAP2 and survivin was observed in
patients with progressive disease which also exhibited inhibition of
spontaneous apoptosis.
Grzybowska-Izydorczyk et al. [63]
Modulation of XIAP expression sensitized CLL cells to chemotherapy in
humans.
Zhu et al. [68]
SurvivinCo-expression of survivin and cIAP1 was related with shorter OS and
identified as a negative prognostic factor.
Grzybowska-Izydorczyk et al. [63]
Inhibition of survivin sensitized CLL cells to cytotoxic agents and induced
apoptosis in and ex vivo model.
Purroy et al. [67]
SmacLower levels of Smac protein correlated with progressive disease.Grzybowska-Izydorczyk et al. [63]
Colorectal Carcinoma (CRC)
SurvivinSurvivin and cIAP2 expression was characteristic of elderly groups Endo et al. [70]
XIAPXIAP expression was comparable in both normal and cancerous tissue of
old and young patients.
Endo et al. [70]
XIAP protein and mRNA levels are higher in cancerous tissue compared to
surrounding normal tissue.
Lee et al. [74], Guoan et al. [75]
cIAPscIAP1 is most frequently expressed in nucleus while cIAP2 in the cytoplasm.
A role of nuclear localization is suggested to be involved in the pathogenesis
of CRC.
Ponnelle et al. [71]
Higher cIAP2 levels are observed in stage II CRC either lymphoid metastasis
and higher rate of chemotherapy failure with FU.
Karasawa et al. [72]
Together with TUCAN, cIAP2 low levels presented positive correlation with
higher five-year survival.
Krajewska et al. [73]
LivinLivin expression correlated with longer survival Lee et al. [74]
ApollonApollon was observed overexpressed in CRC tissue and correlated with
unfavorable clinical features.
Bianchini et al. [77]
SmacDecreased expression of Smac was considered an independent factor for
poor prognosis.
Endo et al. [78]
Breast Carcinoma
XIAPXIAP and survivin expression was observed to be increased in advanced cancer.Pluta et al. [80]
Increased in XIAP expression was observed in higher grades of ductal
invasive breast carcinoma and ductal breast carcinoma in situ.
Yang et al. [81]
High percentage of breast invasive ductal carcinoma with high immunoscore
showed high XIAP expression.
Zhang et al. [82]
Higher cytoplasmic expression is observed compared to normal tissues and
correlated with other molecular abnormalities. It was considered a
prognostic biomarker.
Xu et al. [83]
High XIAP protein related with shorter OS
cIAP2Levels were higher in node positive breast cancerPluta et al. [80]
cIAP1cIAP1 levels did not correlate with clinicopathological featuresPluta et al. [80]
SurvivinCorrelated with metastasis, advanced stage and tumor size.Youssef et al. [85]
Positive expression of survivin have significant association with worse OSSong et al. [88]
LivinHigh expression is observed in high-invasive breast cancer cells compared to
non-invasive cells. Authors described potential role in migration of cancer cells.
Li et al. [13]
SmacLower expression was observed compared to healthy tissuePluta et al. [89]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cetraro, P.; Plaza-Diaz, J.; MacKenzie, A.; Abadía-Molina, F. A Review of the Current Impact of Inhibitors of Apoptosis Proteins and Their Repression in Cancer. Cancers 2022, 14, 1671. https://doi.org/10.3390/cancers14071671

AMA Style

Cetraro P, Plaza-Diaz J, MacKenzie A, Abadía-Molina F. A Review of the Current Impact of Inhibitors of Apoptosis Proteins and Their Repression in Cancer. Cancers. 2022; 14(7):1671. https://doi.org/10.3390/cancers14071671

Chicago/Turabian Style

Cetraro, Pierina, Julio Plaza-Diaz, Alex MacKenzie, and Francisco Abadía-Molina. 2022. "A Review of the Current Impact of Inhibitors of Apoptosis Proteins and Their Repression in Cancer" Cancers 14, no. 7: 1671. https://doi.org/10.3390/cancers14071671

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop