Open access peer-reviewed chapter

New Discoveries on the Roles of “Other” HECT E3 Ubiquitin Ligases in Disease Development

Written By

Emma I. Kane and Donald E. Spratt

Submitted: 12 February 2020 Reviewed: 14 February 2020 Published: 21 March 2020

DOI: 10.5772/intechopen.91770

From the Edited Volume

Ubiquitin - Proteasome Pathway

Edited by Xianquan Zhan

Chapter metrics overview

925 Chapter Downloads

View Full Metrics

Abstract

HECT E3 ubiquitin ligases selectively recognize, bind, and ubiquitylate their substrate proteins to target them for 26S proteasomal degradation. There is increasing evidence that HECT E3 ubiquitin ligase dysfunction due to misfolding and/or the gene encoding the protein being mutated is responsible for the development of different diseases. Apart from the more prominent and well-characterized E6AP and members of the NEDD4 family, new studies have begun to reveal how other members of the HECT E3 ubiquitin ligase family function as well as their links to disease and developmental disorders. This chapter provides a comprehensive discussion on the more mysterious members of the HECT E3 ubiquitin ligase family and how they control intracellular processes. Specifically, AREL1, HACE1, HECTD1, HECTD4, G2E3, and TRIP12 will be examined as these enzymes have recently been identified as contributors to disease development.

Keywords

  • apoptosis
  • AREL1
  • cancer
  • HECT E3 ubiquitin ligase
  • G2E3
  • HACE1
  • HECT
  • HECTD1
  • HECTD4
  • neurodevelopment
  • proteasomal degradation
  • TRIP12
  • ubiquitin
  • ubiquitylation

1. Introduction

1.1 HECT E3 ubiquitin ligase-dependent ubiquitylation

Ubiquitylation is an essential post-translational modification that regulates numerous intracellular processes including protein localization and trafficking, DNA damage response, immune system and viral response, apoptosis and proteolysis [1, 2]. E3 ubiquitin ligases play an important role in recognizing, binding, and covalently attaching ubiquitin to their various substrates to elicit a specific cellular response [3]. The homologous to E6AP C-terminus (HECT) E3 ubiquitin ligases are a unique subfamily that use a multistep pathway to selectively target substrate proteins for ubiquitylation [4]. HECT-dependent ubiquitylation requires the recruitment of an E2 ubiquitin conjugating enzyme charged with ubiquitin to the N-terminal lobe of the HECT domain on the E3 ligase [5, 6]. The ubiquitin cargo is then transferred from the E2 enzyme to the conserved catalytic cysteine within the C-terminal lobe of the HECT domain via a transthiolation reaction to form a thioester bond. The HECT E3~ubiquitin complex will then bind to a substrate and covalently attach ubiquitin on to a lysine residue of the substrate protein forming a stable isopeptide bond between the C-terminus of ubiquitin and the ε-amine of the substrate lysine [3, 5, 6]. This process can be repeated numerous times to form different polyubiquitin chain linkages with the specific HECT E3 ubiquitin ligase dictating the type(s) of ubiquitin linkages that are built [2, 7].

1.2 Ubiquitin attachment site(s) and chain type linkages determine the fate of a substrate protein

The destiny of a ubiquitin-tagged protein is dependent on (i) the site(s) of ubiquitin attachment on the substrate, (ii) the number of ubiquitin moieties attached to the substrate (i.e. mono-, multi-mono-, or polyubiquitin), and (iii) the specific type(s) of linkages between the different ubiquitin molecules in a polyubiquitin chain (i.e. K48, K63, branched, etc.) [1, 2, 7]. Potential fates of a ubiquitin-tagged substrate include changes in cellular localization/trafficking, enhanced/inhibited protein activity, changes in protein–protein affinity/interactions, and proteolysis [1, 2, 6, 7, 8].

Differences in ubiquitin lysine linkage specificity determine the destination and/or fate of the targeted protein in the cell (Table 1). For example, the well-established K48-polyubiquitylation chain, heterotypic K11/K48-polyubiquitin, K29/K48-polyubiquitin monoubiquitin tagged peptides and multiple monoubiquitin tagged proteins have also been found to signal for 26S proteasomal degradation [7, 8]. K63-polyubiquitin chains signal for protein degradation through the initiation of K48/K63 polyubiquitin branch formation [5] but cannot be recognized by the 26S proteosome [26]. To date, many different varieties of ubiquitin chain types have been identified, but their distinct biological functions remain unclear.

Chain types Linker Proposed function
Monoubiquitylation
Monoubiquitylation/multi-monoubiquitylation Endocytosis [9]
DNA damage repair [10, 11, 12, 13, 14, 15]
Histone regulation [10, 11, 12, 13, 14, 15]
Mitophagy [10, 11, 12, 13, 14, 15]
Protein localization [10, 11, 12, 13, 14, 15]
Protein interactions [10, 11, 12, 13, 14, 15]
Protein transportation [10, 11, 12, 13, 14, 15]
Transcription activation [10, 11, 12, 13, 14, 15]
Polyubiquitylation
Chain (homotypic) M1 Innate immunity [2, 9, 16]
Linear chain formation [9]
NF-κB activation [9, 16]
Signaling cascades [9, 16]
K6 DNA damage response [14]
NF-κB regulation [14]
Mitophagy [14]
K11 Cell cycle regulation [17]
DNA damage response [18]
Mitophagy [17]
NF-κB activation [16]
Protein degradation [17]
K27 DNA damage response [18]
Kinase activation [19]
Protein degradation [20]
Protein scaffolding [21]
Protein trafficking [22]
K29 DNA damage response [18]
Kinase activation [19]
Protein degradation [9]
K33 DNA damage response [10, 11, 12, 13, 14, 15, 18]
Kinase activation [23]
Post-golgi trafficking [24]
T-cell signaling [23]
K48 Protein degradation [1, 2, 25]
K63 DNA damage response [9, 18]
NF-κB activation [9, 16]
Protein trafficking [9]
Chain (heterotypic; branched) M1/K63 NF-κB activation [16]
K11/K48 Protein degradation [26, 27]
K29/K48 Protein degradation [26]
K48/K63 Protein degradation [26]
K11/K63 Endocytosis [28]

Table 1.

Ubiquitin conjugation determines the intracellular fate of a substrate protein.

Monoubiquitylation can occur at one site or at multiple sites (multi-monoubiquitylation) on a substrate. Polyubiquitylation can build off of a mono-ubiquitin attachment site with a specific lysine linkage (homotypic) or have multiple chains with different lysine linkages (branch) at the end of a growing ubiquitin chain (heterotypic). These modifications can also influence signaling pathways, whether it is through enhancing or inhibiting participating proteins and processes.

Advertisement

2. The “other” HECT E3 ubiquitin ligases: important players in disease, yet poorly understood

The HECT E3 ubiquitin ligases can be categorized into three subfamilies –NEDD4, HERC, and “other” – based on their sequence/structure similarity and domain architecture [4, 5]. Of the 28 HECT E3 ubiquitin ligases identified in humans, there are 12 “other” HECT E3 ubiquitin ligases that do not fall under the well-studied NEDD4 or HERC subfamilies. Each member of the “other” HECTs have variable N-terminal domains that are thought to be involved in protein–protein interactions and/or intracellular localization [4, 5]. Having prominent responsibilities in cellular homeostasis would leave the impression there is ample research on the HECT E3 ubiquitin ligase family as a whole, however, there remain many unanswered questions about the biological functions and mechanisms of this important E3 ligase family, particularly for members of the more mysterious “other” subfamily. With new research and discoveries becoming available, there is mounting evidence that the lesser known HECT E3 ubiquitin ligases play critical roles in regulating intracellular processes and their dysfunction have been suggested to contribute to the onset of many diseases and disorders [4, 29, 30, 31, 32]. Here we discuss the latest discoveries on these lesser known members of the HECT E3 ubiquitin ligases and on their emerging roles in developmental and neurological abnormalities, cancers, and embryogenesis.

2.1 AREL1, a key regulator of apoptosis and potential oncogenic drug target

Apoptosis resistant E3 ubiquitin protein ligase 1(AREL1; 823 residues) is a cytosolic enzyme responsible for regulating apoptosis through the inhibition of proapoptotic proteins via ubiquitylation [33]. AREL1 contains two immunoglobin-like folds (IGF) near its N-terminus that potentially mediate substrate binding and recognition (Figure 1) [35, 36]. IGF domains assemble into a sandwich-like form consisting of antiparallel β-strands that allow for protein–protein interactions [36, 37].

Figure 1.

AREL1 domain architecture. AREL1 contains a putative immunoglobulin fold domain (IGF, residues 52-158) and the canonical HECT domain (436-823) as annotated on UniProt and InterPro. Representative crystal structures of an IGF fold (human IGF FAB in yellow/orange; PDB 7FAB [34]), which is suggested to mediate AREL1 substrate binding and recognition, while the AREL1 HECT domain (HECTN−lobe in green, HECTC−lobe in blue, catalytic cysteine C790 in red; PDB 6JX5 [35]) is required for ubiquitylation activity. Structures were visualized using PyMol.

Apoptosis (aka programmed cell death) is an important and highly regulated biological process that occurs during early embryonic development and the immune response [38, 39]. Once apoptosis is initiated the cell is committed to die, which is mediated by a caspase cascade [40]. Intrinsic apoptosis is turned on by the release of intermembrane mitochondrial proteins when cells are under oxidative stress [38, 39, 41]. In contrast, the extrinsic apoptotic pathway is activated by extracellular signaling at the cell membrane leading to the formation of the death-inducing signaling complex (DISC) [42, 43, 44].

AREL1 was first identified in 2013 and was immediately recognized as an oncogenic target due to its inhibitory role in apoptosis [33]. Identified substrates for AREL1 include second mitochondrial activator of caspase (SMAC), HtrA serine peptidase 2 (HtrA2) and septin 4 (ARTS), which are known antagonists of inhibitor of apoptosis proteins (IAPs) [45]. Studies have shown that AREL1 can build K48 and K63 polyubiquitin chains to target substrates for proteolysis, as well as atypical K33 polyubiquitin chains whose biological function is still being clarified [35, 46]. Various IAPs, including SMAC, HtrA1, and ARTS, are released from the mitochondrial intermembrane into the cytosol when the cell is triggered or stressed. AREL1 inhibits apoptosis by ubiquitylating these IAP antagonists with K48-linked polyubiquitin chains targeting the IAPs for proteasomal degradation [33].

The induction of apoptosis is thought to require the release of numerous IAPs in the cytosol to allow different signaling pathways to initiate apoptosis depending on the cell’s specific stress. For example, the release of SMAC into the cytosol allows it to bind cellular inhibitor of apoptosis protein 1/2 (cIAP1/2), which then targets cIAP1/2 for proteasomal degradation to initiate apoptosis. However, when AREL1 is present, SMAC is ubiquitylated by AREL1 and degraded, thus blocking SMAC-cIAP1/2 complex formation enabling cell survival [33]. Many cancer therapies are interested in specifically turning on apoptosis through IAPs in cancer cells [47, 48, 49], thus AREL1 could prove to be a novel enzyme in drug development. Continued studies on AREL1’s mechanisms in controlling cell death are warranted.

2.2 HACE1, a prominent tumor suppressor with dual function

First identified in 2004, HECT domain and ankyrin repeat containing E3 ubiquitin protein ligase 1 (HACE1; 909 residues) has been shown to take part in various cellular processes. For example, HACE1 is best known as a tumor suppressor as altered HACE1 expression levels have been observed in various cancers including colorectal, breast, liver, kidney, osteosarcoma, lymphoma and gastric cancer [50, 51, 52, 53, 54]. HACE1 contains six ankyrin repeats near its N-terminus that likely take part in HACE1-substrate recognition and protein–protein interactions (Figure 2). While it is not yet fully understood how the ankyrin repeats support HACE1 function, ankyrin repeats in other proteins have been shown to instigate the development of a wide array of diseases including cancer [56]. HACE1 also supports Golgi membrane biogenesis during cell division by ubiquitylating members of the Ras-related superfamily of small G proteins [57].

Figure 2.

Predicted domain architecture of HACE1. HACE1 contains six putative ankyrin-repeats (ANK; residues 64-258) and a HECT domain (574-909) as annotated on UniProt and InterPro. Representative crystal structures of an ANK repeat fold (in shades of pink, PDB 4O60 [55]), which is likely involved in HACE1 substrate binding and recognition, and a HECT domain (HECTN-lobe in green, HECTC-lobe in blue; PDB 6JX5 [35]) found at HACE1’s C-terminus that is required of ubiquitylation activity. Structures were visualized using PyMol.

Studies have shown that HACE1 expression levels are altered when comparing normal and cancerous tissue. Specifically, in the Wilms’ tumor cell line, HACE1 expression was essentially nonexistent, whereas in other cancer cell lines expression levels were lower than average [50]. This study concluded that HACE1 was essential in repressing cancer development as the lowered expression levels of HACE1 were not mutation dependent. Low expression levels of HACE1 have also been observed in other cancer cell lines. For instance, it was found that the methylation of the HACE1 promoter resulted in decreased HACE1 expression in liver cancer cells, which in turn decreased HACE1’s ability to ubiquitylate its identified substrates optineurin (OPTN) and microtubule-associated proteins 1A/1B light chain 3B protein [53]. Many different substrates of HACE1 have been identified to date (summarized in [4]), including β2-adrenergic receptor (ß2AR) [58], OPTN [59], retinoic acid receptor beta (RAR-β) [57], tumor necrosis factor receptor-2 (TNFR2) [60], and various Ras-related proteins [57, 58, 61, 62, 63, 64, 65]. Expanded studies on how HACE1 binds and recognizes its substrates are needed to further clarify the role of HACE1 in cancer development.

HACE1 also plays an essential role in neurodevelopment as it was recently shown to be involved in the spastic paraplegia and psychomotor retardation with or without seizures (SPPRS) phenotype [66]. HACE1 also has cardiac protection function during hemodynamic stress when it was shown in mice with HACE1 deficiency, their susceptibility to accelerated heart failure greatly increased [67]. This suggests that HACE1 has a critical role in protecting the heart from various stresses, thus making it a potential cardiac drug target.

2.3 HECTD1, an important regulator in neurodevelopment

HECT domain containing E3 ubiquitin protein ligase 1 (HECTD1; 2610 residues) was discovered in 2007 as a novel and important regulator of neurodevelopment [68]. HECTD1 has similar domain architecture to HACE1 with four ankyrin repeats near its N-terminus and a C-terminal HECT domain (Figure 3). HECTD1 plays an important role in pulmonary fibrosis during endothelial-mesenchymal transition (EndMT) with reports of increased circular RNA HECTD1 (circHECTD1) transcription, which causes decreased HECTD1 protein expression in lung tissue [72, 73]. Elevated circHECTD1 gene expression has also been found in patients with acute ischemic stroke (AIS) [74]. HECTD1 also contains a Smad4 activation SAD1/UNC (SUN) domain and a mind bomb (MIB) domain, with each having unique roles in intracellular signaling due to Smad-DNA complex formation and cellular interactions through the Notch pathway, respectively [75, 76].

Figure 3.

HECTD1 domain architecture. HECTD1 contains putative protein-protein interaction domains including two armadillo-repeat containing domains (ARM1, residues 8-254; ARM2 residues 892-925 in purple; PDB 4DB8 [69]), four ankyrin-repeats (residues 395-612 in shades of pink; PDB 4O60 [56]), a SAD1/UNC domain (SUN, residues 1115-1244 in yellow; PDB 3UNP [70]), a mind bomb domain (MIB, residues 1266–1338 in red; PDB 2DK3), a helix-bundle domain (H, residues 1896–1968 in magenta; PDB 2KZS [71]) and a HECT domain (HECTN−lobe in green, HECTC−lobe in blue; PDB 6JX5 [35]). Domain boundaries are denoted according to UniProt and InterPro. Structures were visualized using PyMol.

HECTD1 supports fetal growth and proper placenta development. Specifically, HECTD1 aids in the development of the labyrinthine and junctional zones of the placenta, regions where the fetus acquires nutrients and disposes of waste, as well as a bilayer between the labyrinthine and decidual cells, respectively [77, 78]. HECTD1 ensures the proper size development of the labyrinthine, yet the mechanisms to ensure this are still not fully understood. Mutations within HECTD1 lead to the onset of irregular labyrinthine development, which in turn depletes nutrients for the fetus. Fetal fatality can occur without proper maintenance of these placental regions, suggesting that HECTD1 expression is essential for the proper development and survival of fetuses in utero. HECTD1 also plays a role in proper neural tube closure. Anencephaly occurs when the neural tube does not close properly, which has been linked to HECTD1 control of heat shock protein 90 (Hsp90) levels [79]. When Hsp90 secretion levels are not properly regulated by HECTD1 ubiquitylation, abnormal neural tube development can occur. The continued examination of this important enzyme will hopefully clarify the molecular basis for HECTD1’s role in neurodevelopment.

2.4 HECTD4, a genetically linked precursor to cancer and cardiovascular disease

HECT domain containing E3 ubiquitin protein ligase 4 (HECTD4; 3996 residues) was recently discovered in 2014. A pleiotropic gene screen showed that there were links between metabolic syndromes and inflammation, specifically with single nucleotide polymorphisms (SNPs) in the HECTD4 gene [80]. Since 2014, HECTD4 has also been found to be associated with diabetes, hypertension and cardiovascular disease, lung adenocarcinoma, urothelial carcinoma and ovarian endometriosis [81, 82, 83, 84].

Being one of the larger enzymes within the “other” HECT E3 ubiquitin ligase subfamily, it is intriguing that there have been no putative domains annotated for HECTD4 except the C-terminal HECT domain (residues 3627–3996). There are likely different domains located in the N-terminal region of the HECTD4 protein that need to be identified and functionally examined.

Genetic screening has identified various mutations in HECTD4 in cancer cells. For example, HECTD4 was recently identified as one of nine genes that correlated with the onset of lung adenocarcinoma [83]. Tumor genetic screens have revealed in patients with urothelial carcinoma in the bladder (UCB) that mutations in HECTD4, Fibrillin-3 Precursor (FBN3) and Citron Rho-Interacting Kinase (CIT) were correlated to UCB disease progression [81]. HECTD4 may also be linked to the development of ovarian endometriosis (OEM) [82]. However, future studies are still needed to verify this relationship.

Already having various genetic links to cancers and cardiovascular disease, HECTD4 deserves more attention by the research community to further clarify its biological and functional roles in the cell. Currently very little is known about HECTD4, therefore it will be imperative to first identify potential similarities in protein sequence and/or domain architecture. To better clarify HECTD4’s role in ubiquitin biology, it will also be important to discover HECTD4 substrates and annotate the sites of HECTD4-dependent ubiquitylation to answer how this mysterious HECT E3 ubiquitin ligase contributes to disease development.

2.5 G2E3, a unique multifunctional HECT E3 ubiquitin ligase with RING-like features

G2/M-phase specific E3 ubiquitin protein ligase (G2E3; 706 residues) was first identified in 2006 and named for its role during the G2/M phase of cell division and for having a conserved C-terminal HECT domain [85]. Knockout studies of G2E3 in mice demonstrated that this enzyme is essential in preventing apoptotic cell death during early embryonic development [86]. Expression levels in G2E3 were also observed to increase during early embryogenesis, specifically during central nervous system development. This enzyme is also implicated in cell cycle progression and DNA damage response [85, 87, 88].

G2E3 contains three plant homeodomain (PHD)-type zinc finger repeats, a domain typically known to bind to modified histones and act as epigenetic readers [89], making it the only known HECT E3 ubiquitin ligase to possess “RING”-like characteristics (Figure 4). G2E3 is primarily found within the nucleus due to its N-terminal nucleolar localization signal sequence, while the PHD domains have been suggested to cause the translocation of G2E3 to the cytoplasm [85, 86]. Previous biochemical studies showed that the ubiquitin ligase activity of G2E3 was exclusively found in two of the putative N-terminal PHD domains while the C-terminal HECT domain of G2E3 had apparently lost its ubiquitylation activity [86]. Since the PHD domains of G2E3 appear to be capable of recruiting E2 enzymes to build K48-linked polyubiquitin chains [86], this suggests that the HECT domain of G2E3 may have become vestigial through evolutionary pressure. Intriguingly, G2E3 has aspects of the RING and HECT E3 ubiquitin ligase families, analogous to members of the RING-between-RING (aka RING-BRcat-Rcat) E3 ubiquitin ligases that includes parkin and HOIL-interacting protein (HOIP) of the LUBAC complex [91, 92].

Figure 4.

Domain architecture of G2E3. G2E3 contains three putative plant homeodomain (PHD)-type domains (residues 78-128) and a unique HECT-like domain (371-678) as annotated on UniProt and InterPro. Representative crystal structures of a PHD domain (human PHD finger protein 13 in cyan with Zn-coordinating residues in yellow; PDB 3O70 [90]), which is suggested to interact with DNA and/or proteins in the nucleus, while the G2E3 HECT domain (HECTN−lobe in green, HECTC−lobe in blue; PDB 6JX5 [35]) may have lost its ability to catalyze the transfer of ubiquitin. Structures were visualized using PyMol.

G2E3 was recently identified as a potential drug target to increase the efficacy of chemotherapy drugs, specifically with Cisplatin [88]. Since Cisplatin is the most common chemotherapy drug, much research has been dedicated to increasing Cisplatin’s ability to specifically trigger the DNA damage response in cancer cells to initiate apoptosis while limiting its exposure time and prescribed duration for patients [93]. Clearly, G2E3 is an important nuclear protein whose mechanism is currently unresolved. Further studies are needed to clarify how this divergent HECT-domain containing E3 ubiquitin ligase works in the cell.

2.6 TRIP12, the multifunctioning E3 ubiquitin ligase essential for embryogenesis and DNA damage repair

Thyroid hormone receptor interactor 12 (TRIP12; 2040 residues), was first identified in 2001 for containing a unique tryptophan-tryptophan-glutamate (WWE) domain that is predicted to be involved in ubiquitylation and ADP-ribosylation [94] (Figure 5). It also contains two N-terminal ARM domains, similar to HECTD1, and a conserved HECT domain at its C-terminus. TRIP12 is a novel HECT E3 ubiquitin ligase that has been shown to take part in various cellular pathways and processes including embryogenesis, DNA damage response and the neddylation pathway [95, 96, 97]. It has been reported that TRIP12 preferentially builds mono- as well as K48 and K63 polyubiquitin chains to tag its substrates for degradation and for DNA damage site recruitment, respectively [96].

Figure 5.

TRIP12 domain architecture. TRIP12 contains two putative armadillo-repeat containing domains (ARM1, residues 437-713; ARM2, residues 826-938), a tryptophan-tryptophan-glutamate (WWE)-domain (residues 749-836) and a conserved HECT domain (1885-1992) as annotated on UniProt and InterPro. Representative crystal structures of an ARM domain (in purple; PDB 4DB8 [69]) and WWE domain (in orange; PDB 6MIW), both of which are suggested to be involved in protein-protein and substrate interactions, as well as a HECT domain (HECTN−lobe in green, HECTC−lobe in blue; PDB 6JX5 [35]) required for ubiquitylation activity. Structures were visualized using PyMol.

TRIP12 has been shown to be directly and/or indirectly involved in cancer progression. For example, TRIP12 may serve as an oncogenic drug target for patients with acute myeloid leukemia (AML) by blocking a TRIP12 alternative splicing event, specifically excising exon3 from the mature TRIP12 mRNA [98]. TRIP12 also targets pancreas transcription factor 1a (PTF1a) for proteasomal degradation, a protein essential for pancreatic cancer development [99]. TRIP12 forms a ternary complex with deubiquitylase ubiquitin-specific protease 7 (USP7) that aids in hepatocellular carcinoma (HCC) proliferation; when USP7 expression levels are heightened, TRIP12 cannot tag ARF tumor suppressor (p14ARF) for ubiquitylation [100]. Furthermore, TRIP12 is associated with human papillomavirus (HPV)-positive head and neck squamous cell carcinoma (HNSCC) due to its mediation of p16-related radiation efficacy [101].

Members of the HECT E3 ubiquitin ligase family play important roles in neurodevelopment and their malfunction may be causative in different neurological diseases and disorders (reviewed in [4]). Recent genetic screens have been looking to identify genetic markers for autism spectrum disorder (ASD) and intellectual disability (ID). Interestingly, a de novo mutation in TRIP12 was found in patients with or without ASD and displaying characteristics of ID [102]. Further studies to clarify the specific mechanism(s) for how mutations in the TRIP12 gene contribute to ASD and ID phenotypes are needed.

Advertisement

3. Conclusion

Although much research has and continues to be performed for E6AP and members of the NEDD4 family, greater attention on the mysterious “other” HECT E3 ubiquitin ligases is warranted due to their emerging involvement in various diseases and neurological disorders. Combining genetic, cell biology, biochemical, and biophysical approaches to study these unique HECT E3 ubiquitin ligases will help to decipher their specific roles and/or functions in the cell as well as potentially aid in novel therapy development to treat rare conditions caused by the dysfunction of these lesser known members of the HECT E3 ubiquitin ligase family.

Advertisement

Acknowledgments

This work was supported by the National Institutes of Health (R15GM126432 to D.E.S.) and start-up funds from Clark University (to D.E.S.).

Advertisement

Conflict of interest

The authors declare that there is no conflict of interest regarding the publication of this chapter.

Advertisement

Abbreviations

AISacute ischemic stroke
AMLacute myeloid leukemia
ANKankyrin repeat
AREL1apoptosis resistant E3 ubiquitin protein ligase 1
ARMarmadillo-repeat domain
ARTSseptin 4
ASDautism spectrum disorder
β2ARadrenergic receptor β2AR
cIAP1/2cellular inhibitor of apoptosis protein 1/2
circHECTD1circular RNA HECTD1
CITcitron rho-interacting kinase
EndMTendothelial-mesenchymal transition
FBN3fibrillin-3 precursor
G2E3G2/M-phase specific E3 ubiquitin protein ligase
Hhelix-bundle domain
HACE1HECT domain and ankyrin repeat containing E3 ubiquitin protein ligase 1
HCChepatocellular carcinoma
HECThmologous to E6AP C-terminus
HECTD1HECT domain containing E3 ubiquitin protein ligase 1
HECTD4HECT domain containing E3 ubiquitin protein ligase 4
HERCHECT and RLD domain-containing
HNSCChead and neck squamous cell carcinoma
HOIPHOIL-interacting protein
HPVhuman papillomavirus
Hsp90heat shock protein 90
HtrA2HtrA serine peptidase 2
IDintellectual disability
IGFimmunoglobulin-like fold
MIBmind bomb domain
NEDD4neuronal precursor cell-expressed developmentally down-regulated 4
SMACsecond mitochondrial activator of caspase
OEMovarian endometriosis
OPTNoptineurin
p14ARFARF tumor suppressor
PHDplant homeodomain-type zinc finger
PTF1apancreas transcription factor 1a
RAR-βretinoic acid receptor beta
RINGreally interesting new gene
TNFR2tumor necrosis factor receptor-2
SNPssingle nucleotide polymorphisms
SPPRSspastic paraplegia and psychomotor retardation with or without seizures
SUNSmad4 activation SAD1/UNC domain
TRIP12thyroid hormone receptor interactor 12
UCBurothelial carcinoma in the bladder
USP7deubiquitylase ubiquitin-specific protease 7
WWEtryptophan-tryptophan-glutamate domain

References

  1. 1. Hershko A, Ciechanover A. The ubiquitin system. Annual Review of Biochemistry. 1998;67:425-479
  2. 2. Komander D, Rape M. The ubiquitin code. Annual Review of Biochemistry. 2012;81:203-229
  3. 3. Morreale FE, Walden H. Types of ubiquitin ligases. Cell. 2016;165(1):248-e1
  4. 4. Wang Y, Argiles-Castillo D, Kane EI, Zhou A, Spratt DE. HECT E3 ubiquitin ligases: Emerging insights into their biological roles and disease relevance. Journal of Cell Science. 2020. in press
  5. 5. Lorenz S. Structural mechanisms of HECT-type ubiquitin ligases. Biological Chemistry. 2018;399(2):127-145
  6. 6. Metzger MB, Hristova VA, Weissman AM. HECT and RING finger families of E3 ubiquitin ligases at a glance. Journal of Cell Science. 2012;125(Pt 3):531-537
  7. 7. Swatek KN, Komander D. Ubiquitin modifications. Cell Research. 2016;26(4):399-422
  8. 8. Grice GL, Nathan JA. The recognition of ubiquitinated proteins by the proteasome. Cellular and Molecular Life Sciences. 2016;73(18):3497-3506
  9. 9. Pickart CM, Fushman D. Polyubiquitin chains: Polymeric protein signals. Current Opinion in Chemical Biology. 2004;8(6):610-616
  10. 10. Bergink S, Jentsch S. Principles of ubiquitin and SUMO modifications in DNA repair. Nature. 2009;458(7237):461-467
  11. 11. Durcan TM, Tang MY, Perusse JR, Dashti EA, Aguileta MA, McLelland GL, et al. USP8 regulates mitophagy by removing K6-linked ubiquitin conjugates from parkin. The EMBO Journal. 2014;33(21):2473-2491
  12. 12. Haglund K, Di Fiore PP, Dikic I. Distinct monoubiquitin signals in receptor endocytosis. Trends in Biochemical Sciences. 2003;28(11):598-603
  13. 13. Hicke L. Protein regulation by monoubiquitin. Nature Reviews. Molecular Cell Biology. 2001;2(3):195-201
  14. 14. Passmore LA, Barford D. Getting into position: The catalytic mechanisms of protein ubiquitylation. The Biochemical Journal. 2004;379(Pt 3):513-525
  15. 15. Paul A, Wang B. RNF8- and Ube2S-dependent ubiquitin lysine 11-linkage modification in response to DNA damage. Molecular Cell. 2017;66(4):458-472.e5
  16. 16. Iwai K. Diverse ubiquitin signaling in NF-kappaB activation. Trends in Cell Biology. 2012;22(7):355-364
  17. 17. Wickliffe KE, Williamson A, Meyer HJ, Kelly A, Rape M. K11-linked ubiquitin chains as novel regulators of cell division. Trends in Cell Biology. 2011;21(11):656-663
  18. 18. Liu P, Gan W, Su S, Hauenstein AV, Fu TM, Brasher B, et al. K63-linked polyubiquitin chains bind to DNA to facilitate DNA damage repair. Science Signaling. 2018;11(533)
  19. 19. Singh R, Karri D, Shen H, Shao J, Dasgupta S, Huang S, et al. TRAF4-mediated ubiquitination of NGF receptor TrkA regulates prostate cancer metastasis. The Journal of Clinical Investigation. 2018;128(7):3129-3143
  20. 20. Birsa N, Norkett R, Wauer T, Mevissen TE, Wu HC, Foltynie T, et al. Lysine 27 ubiquitination of the mitochondrial transport protein Miro is dependent on serine 65 of the Parkin ubiquitin ligase. The Journal of Biological Chemistry. 2014;289(21):14569-14582
  21. 21. Wang Q , Liu X, Cui Y, Tang Y, Chen W, Li S, et al. The E3 ubiquitin ligase AMFR and INSIG1 bridge the activation of TBK1 kinase by modifying the adaptor STING. Immunity. 2014;41(6):919-933
  22. 22. Castaneda CA, Dixon EK, Walker O, Chaturvedi A, Nakasone MA, Curtis JE, et al. Linkage via K27 bestows ubiquitin chains with unique properties among polyubiquitins. Structure. 2016;24(3):423-436
  23. 23. Huang H, Jeon MS, Liao L, Yang C, Elly C, Yates JR 3rd, et al. K33-linked polyubiquitination of T cell receptor-zeta regulates proteolysis-independent T cell signaling. Immunity. 2010;33(1):60-70
  24. 24. Yuan WC, Lee YR, Lin SY, Chang LY, Tan YP, Hung CC, et al. K33-linked polyubiquitination of coronin 7 by Cul3-KLHL20 ubiquitin E3 ligase regulates protein trafficking. Molecular Cell. 2014;54(4):586-600
  25. 25. Thrower JS, Hoffman L, Rechsteiner M, Pickart CM. Recognition of the polyubiquitin proteolytic signal. The EMBO Journal. 2000;19(1):94-102
  26. 26. Ohtake F, Tsuchiya H, Saeki Y, Tanaka K. K63 ubiquitylation triggers proteasomal degradation by seeding branched ubiquitin chains. Proceedings of the National Academy of Sciences of the United States of America. 2018;115(7):E1401-E1408
  27. 27. Meyer HJ, Rape M. Enhanced protein degradation by branched ubiquitin chains. Cell. 2014;157(4):910-921
  28. 28. Boname JM, Thomas M, Stagg HR, Xu P, Peng J, Lehner PJ. Efficient internalization of MHC I requires lysine-11 and lysine-63 mixed linkage polyubiquitin chains. Traffic. 2010;11(2):210-220
  29. 29. Bernassola F, Chillemi G, Melino G. HECT-type E3 ubiquitin ligases in cancer. Trends in Biochemical Sciences. 2019;44(12):1057-1075
  30. 30. Bielskiene K, Bagdoniene L, Mozuraitiene J, Kazbariene B, Janulionis E. E3 ubiquitin ligases as drug targets and prognostic biomarkers in melanoma. Medicina (Kaunas, Lithuania). 2015;51(1):1-9
  31. 31. Fajner V, Maspero E, Polo S. Targeting HECT-type E3 ligases - insights from catalysis, regulation and inhibitors. FEBS Letters. 2017;591(17):2636-2647
  32. 32. Weber J, Polo S, Maspero E. HECT E3 ligases: A tale with multiple facets. Frontiers in Physiology. 2019;10:370
  33. 33. Kim JB, Kim SY, Kim BM, Lee H, Kim I, Yun J, et al. Identification of a novel anti-apoptotic E3 ubiquitin ligase that ubiquitinates antagonists of inhibitor of apoptosis proteins SMAC, HtrA2, and ARTS. The Journal of Biological Chemistry. 2013;288(17):12014-12021
  34. 34. Saul FA, Poljak RJ. Crystal structure of human immunoglobulin fragment fab new refined at 2.0 a resolution. Proteins. 1992;14(3):363-371
  35. 35. Singh S, Ng J, Nayak D, Sivaraman J. Structural insights into a HECT-type E3 ligase AREL1 and its ubiquitination activities in vitro. The Journal of Biological Chemistry. 2019;294(52):19934-19949
  36. 36. Chen J, Wang B, Wu Y. Structural characterization and function prediction of immunoglobulin-like fold in cell adhesion and cell signaling. Journal of Chemical Information and Modeling. 2018;58(2):532-542
  37. 37. Halaby DM, Poupon A, Mornon J. The immunoglobulin fold family: Sequence analysis and 3D structure comparisons. Protein Engineering. 1999;12(7):563-571
  38. 38. Elmore S. Apoptosis: A review of programmed cell death. Toxicologic Pathology. 2007;35(4):495-516
  39. 39. D'Arcy MS. Cell death: A review of the major forms of apoptosis, necrosis and autophagy. Cell Biology International. 2019;43(6):582-592
  40. 40. McIlwain DR, Berger T, Mak TW. Caspase functions in cell death and disease. Cold Spring Harbor Perspectives in Biology. 2013;5(4):a008656
  41. 41. Wu CC, Bratton SB. Regulation of the intrinsic apoptosis pathway by reactive oxygen species. Antioxidants & Redox Signaling. 2013;19(6):546-558
  42. 42. Falschlehner C, Emmerich CH, Gerlach B, Walczak H. TRAIL signalling: Decisions between life and death. The International Journal of Biochemistry & Cell Biology. 2007;39(7-8):1462-1475
  43. 43. Walczak H. Death receptor-ligand systems in cancer, cell death, and inflammation. Cold Spring Harbor Perspectives in Biology. 2013;5(5):a008698
  44. 44. Walczak H, Haas TL. Biochemical analysis of the native TRAIL death-inducing signaling complex. Methods in Molecular Biology. 2008;414:221-239
  45. 45. Lotan R, Rotem A, Gonen H, Finberg JP, Kemeny S, Steller H, et al. Regulation of the proapoptotic ARTS protein by ubiquitin-mediated degradation. The Journal of Biological Chemistry. 2005;280(27):25802-25810
  46. 46. Michel MA, Elliott PR, Swatek KN, Simicek M, Pruneda JN, Wagstaff JL, et al. Assembly and specific recognition of k29- and k33-linked polyubiquitin. Molecular Cell. 2015;58(1):95-109
  47. 47. Zhu H, Li Y, Liu Y, Han B. Bivalent SMAC mimetics for treating cancer by antagonizing inhibitor of apoptosis proteins. ChemMedChem. 2019;14(23):1951-1962
  48. 48. Ali R, Singh S, Haq W. IAP proteins antagonist: An introduction and chemistry of Smac mimetics under clinical development. Current Medicinal Chemistry. 2018;25(31):3768-3795
  49. 49. Fulda S. Smac mimetics to therapeutically target IAP proteins in cancer. International Review of Cell and Molecular Biology. 2017;330:157-169
  50. 50. Anglesio MS, Evdokimova V, Melnyk N, Zhang L, Fernandez CV, Grundy PE, et al. Differential expression of a novel ankyrin containing E3 ubiquitin-protein ligase, Hace1, in sporadic Wilms' tumor versus normal kidney. Human Molecular Genetics. 2004;13(18):2061-2074
  51. 51. El-Naggar AM, Clarkson PW, Negri GL, Turgu B, Zhang F, Anglesio MS, et al. HACE1 is a potential tumor suppressor in osteosarcoma. Cell Death & Disease. 2019;10(1):21
  52. 52. Li JC, Chang X, Chen Y, Li XZ, Zhang XL, Yang SM, et al. Loss of the tumor suppressor HACE1 contributes to cancer progression. Current Drug Targets. 2019;20(10):1018-1028
  53. 53. Yu Z, Li Y, Han T, Liu Z. Demethylation of the HACE1 gene promoter inhibits the proliferation of human liver cancer cells. Oncology Letters. 2019;17(5):4361-4368
  54. 54. Zhou Z, Zhang HS, Zhang ZG, Sun HL, Liu HY, Gou XM, et al. Loss of HACE1 promotes colorectal cancer cell migration via upregulation of YAP1. Journal of Cellular Physiology. 2019;234(6):9663-9672
  55. 55. Tikhonova EB, Ethayathulla AS, Su Y, Hariharan P, Xie S, Guan L. A transcription blocker isolated from a designed repeat protein combinatorial library by in vivo functional screen. Scientific Reports. 2015;5:8070
  56. 56. Li J, Mahajan A, Tsai MD. Ankyrin repeat: A unique motif mediating protein-protein interactions. Biochemistry. 2006;45(51):15168-15178
  57. 57. Tang D, Xiang Y, De Renzis S, Rink J, Zheng G, Zerial M, et al. The ubiquitin ligase HACE1 regulates Golgi membrane dynamics during the cell cycle. Nature Communications. 2011;2:501
  58. 58. Lachance V, Degrandmaison J, Marois S, Robitaille M, Genier S, Nadeau S, et al. Ubiquitylation and activation of a Rab GTPase is promoted by a beta(2)AR-HACE1 complex. Journal of Cell Science. 2014;127(Pt 1):111-123
  59. 59. Liu Z, Chen P, Gao H, Gu Y, Yang J, Peng H, et al. Ubiquitylation of autophagy receptor optineurin by HACE1 activates selective autophagy for tumor suppression. Cancer Cell. 2014;26(1):106-120
  60. 60. Tortola L, Nitsch R, Bertrand MJM, Kogler M, Redouane Y, Kozieradzki I, et al. The tumor suppressor Hace1 is a critical regulator of TNFR1-mediated cell fate. Cell Reports. 2016;16(12):3414
  61. 61. Acosta MI, Urbach S, Doye A, Ng YW, Boudeau J, Mettouchi A, et al. Group-I PAKs-mediated phosphorylation of HACE1 at serine 385 regulates its oligomerization state and Rac1 ubiquitination. Scientific Reports. 2018;8(1):1410
  62. 62. Andrio E, Lotte R, Hamaoui D, Cherfils J, Doye A, Daugaard M, et al. Identification of cancer-associated missense mutations in hace1 that impair cell growth control and Rac1 ubiquitylation. Scientific Reports. 2017;7:44779
  63. 63. Castillo-Lluva S, Tan CT, Daugaard M, Sorensen PH, Malliri A. The tumour suppressor HACE1 controls cell migration by regulating Rac1 degradation. Oncogene. 2013;32(13):1735-1742
  64. 64. Daugaard M, Nitsch R, Razaghi B, McDonald L, Jarrar A, Torrino S, et al. Hace1 controls ROS generation of vertebrate Rac1-dependent NADPH oxidase complexes. Nature Communications. 2013;4:2180
  65. 65. Torrino S, Visvikis O, Doye A, Boyer L, Stefani C, Munro P, et al. The E3 ubiquitin-ligase HACE1 catalyzes the ubiquitylation of active Rac1. Developmental Cell. 2011;21(5):959-965
  66. 66. Hollstein R, Parry DA, Nalbach L, Logan CV, Strom TM, Hartill VL, et al. HACE1 deficiency causes an autosomal recessive neurodevelopmental syndrome. Journal of Medical Genetics. 2015;52(12):797-803
  67. 67. Zhang L, Chen X, Sharma P, Moon M, Sheftel AD, Dawood F, et al. HACE1-dependent protein degradation provides cardiac protection in response to haemodynamic stress. Nature Communications. 2014;5:3430
  68. 68. Zohn IE, Anderson KV, Niswander L. The Hectd1 ubiquitin ligase is required for development of the head mesenchyme and neural tube closure. Developmental Biology. 2007;306(1):208-221
  69. 69. Madhurantakam C, Varadamsetty G, Grutter MG, Pluckthun A, Mittl PR. Structure-based optimization of designed armadillo-repeat proteins. Protein Science. 2012;21(7):1015-1028
  70. 70. Zhou Z, Du X, Cai Z, Song X, Zhang H, Mizuno T, et al. Structure of Sad1-UNC84 homology (SUN) domain defines features of molecular bridge in nuclear envelope. The Journal of Biological Chemistry. 2012;287(8):5317-5326
  71. 71. Escobar-Cabrera E, Lau DK, Giovinazzi S, Ishov AM, McIntosh LP. Structural characterization of the DAXX N-terminal helical bundle domain and its complex with Rassf1C. Structure. 2010;18(12):1642-1653
  72. 72. Fang S, Guo H, Cheng Y, Zhou Z, Zhang W, Han B, et al. circHECTD1 promotes the silica-induced pulmonary endothelial-mesenchymal transition via HECTD1. Cell Death & Disease. 2018;9(3):396
  73. 73. Thiery JP, Acloque H, Huang RY, Nieto MA. Epithelial-mesenchymal transitions in development and disease. Cell. 2009;139(5):871-890
  74. 74. Peng X, Jing P, Chen J, Xu L. The role of circular RNA HECTD1 expression in disease risk, disease severity, inflammation, and recurrence of acute ischemic stroke. Journal of Clinical Laboratory Analysis. 2019;33(7):e22954
  75. 75. de Caestecker MP, Yahata T, Wang D, Parks WT, Huang S, Hill CS, et al. The Smad4 activation domain (SAD) is a proline-rich, p300-dependent transcriptional activation domain. The Journal of Biological Chemistry. 2000;275(3):2115-2122
  76. 76. Itoh M, Kim CH, Palardy G, Oda T, Jiang YJ, Maust D, et al. Mind bomb is a ubiquitin ligase that is essential for efficient activation of notch signaling by Delta. Developmental Cell. 2003;4(1):67-82
  77. 77. D'Alonzo D, Emch FH, Shen X, Bruder E, De Geyter C, Zhang H. Hectd1 is essential for embryogenesis in mice. Gene Expression Patterns. 2019;34:119064
  78. 78. Sarkar AA, Nuwayhid SJ, Maynard T, Ghandchi F, Hill JT, Lamantia AS, et al. Hectd1 is required for development of the junctional zone of the placenta. Developmental Biology. 2014;392(2):368-380
  79. 79. Sarkar AA, Zohn IE. Hectd1 regulates intracellular localization and secretion of Hsp90 to control cellular behavior of the cranial mesenchyme. The Journal of Cell Biology. 2012;196(6):789-800
  80. 80. Kraja AT, Chasman DI, North KE, Reiner AP, Yanek LR, Kilpelainen TO, et al. Pleiotropic genes for metabolic syndrome and inflammation. Molecular Genetics and Metabolism. 2014;112(4):317-338
  81. 81. Kittler R, Shiang C, Hutchinson R, Kollipara RK, Kapur P, Franto F, et al. Grade progression in urothelial carcinoma can occur with high or low mutational homology: A first-step toward tumor-specific care in initial low-grade bladder cancer. Oncotarget. 2018;9(10):9415-9424
  82. 82. Li L, Chen Q , Fan QB, Wang S, Shi HH, Zhu L, et al. Pathogenetic gene changes of eutopic endometrium in patients with ovarian endometriosis. Chinese Medical Journal (England). 2019;132(9):1107-1109
  83. 83. Wang X, Li J, Duan Y, Wu H, Xu Q , Zhang Y. Whole genome sequencing analysis of lung adenocarcinoma in Xuanwei, China. Thoracic Cancer. 2017;8(2):88-96
  84. 84. Zhang H, Mo XB, Xu T, Bu XQ , Lei SF, Zhang YH. Novel genes affecting blood pressure detected via gene-based association analysis. G3 (Bethesda). 2015;5(6):1035-1042
  85. 85. Brooks WS, Banerjee S, Crawford DF. G2E3 is a nucleo-cytoplasmic shuttling protein with DNA damage responsive localization. Experimental Cell Research. 2007;313(4):665-676
  86. 86. Brooks WS, Helton ES, Banerjee S, Venable M, Johnson L, Schoeb TR, et al. G2E3 is a dual function ubiquitin ligase required for early embryonic development. The Journal of Biological Chemistry. 2008;283(32):22304-22315
  87. 87. Schmidt F, Karnitz LM, Dobbelstein M. G2E3 attenuating replicative stress. Aging (Albany NY). 2015;7(8):527-528
  88. 88. Schmidt F, Kunze M, Loock AC, Dobbelstein M. Screening analysis of ubiquitin ligases reveals G2E3 as a potential target for chemosensitizing cancer cells. Oncotarget. 2015;6(2):617-632
  89. 89. Sanchez R, Zhou MM. The PHD finger: A versatile epigenome reader. Trends in Biochemical Sciences. 2011;36(7):364-372
  90. 90. Chung HR, Xu C, Fuchs A, Mund A, Lange M, Staege H, et al. PHF13 is a molecular reader and transcriptional co-regulator of H3K4me2/3. eLife. 2016;5:e10607
  91. 91. Dove KK, Stieglitz B, Duncan ED, Rittinger K, Klevit RE. Molecular insights into RBR E3 ligase ubiquitin transfer mechanisms. EMBO Reports. 2016;17(8):1221-1235
  92. 92. Spratt DE, Walden H, Shaw GS. RBR E3 ubiquitin ligases: New structures, new insights, new questions. The Biochemical Journal. 2014;458(3):421-437
  93. 93. Dasari S, Tchounwou PB. Cisplatin in cancer therapy: Molecular mechanisms of action. European Journal of Pharmacology. 2014;740:364-378
  94. 94. Aravind L. The WWE domain: A common interaction module in protein ubiquitination and ADP ribosylation. Trends in Biochemical Sciences. 2001;26(5):273-275
  95. 95. Kajiro M, Tsuchiya M, Kawabe Y, Furumai R, Iwasaki N, Hayashi Y, et al. The E3 ubiquitin ligase activity of Trip12 is essential for mouse embryogenesis. PLoS One. 2011;6(10):e25871
  96. 96. Liu X, Yang X, Li Y, Zhao S, Li C, Ma P, et al. Trip12 is an E3 ubiquitin ligase for USP7/HAUSP involved in the DNA damage response. FEBS Letters. 2016;590(23):4213-4222
  97. 97. Park Y, Yoon SK, Yoon JB. TRIP12 functions as an E3 ubiquitin ligase of APP-BP1. Biochemical and Biophysical Research Communications. 2008;374(2):294-298
  98. 98. Christen F, Hoyer K, Yoshida K, Hou HA, Waldhueter N, Heuser M, et al. Genomic landscape and clonal evolution of acute myeloid leukemia with t(8,21): An international study on 331 patients. Blood. 2019;133(10):1140-1151
  99. 99. Hanoun N, Fritsch S, Gayet O, Gigoux V, Cordelier P, Dusetti N, et al. The E3 ubiquitin ligase thyroid hormone receptor-interacting protein 12 targets pancreas transcription factor 1a for proteasomal degradation. The Journal of Biological Chemistry. 2014;289(51):35593-35604
  100. 100. Georges A, Marcon E, Greenblatt J, Frappier L. Identification and characterization of USP7 targets in cancer cells. Scientific Reports. 2018;8(1):15833
  101. 101. Wang L, Zhang P, Molkentine DP, Chen C, Molkentine JM, Piao H, et al. TRIP12 as a mediator of human papillomavirus/p16-related radiation enhancement effects. Oncogene. 2017;36(6):820-828
  102. 102. Bramswig NC, Ludecke HJ, Pettersson M, Albrecht B, Bernier RA, Cremer K, et al. Identification of new TRIP12 variants and detailed clinical evaluation of individuals with non-syndromic intellectual disability with or without autism. Human Genetics. 2017;136(2):179-192

Written By

Emma I. Kane and Donald E. Spratt

Submitted: 12 February 2020 Reviewed: 14 February 2020 Published: 21 March 2020