Next Article in Journal
Effects of Defective Unloading and Recycling of PCNA Revealed by the Analysis of ELG1 Mutants
Next Article in Special Issue
Effects of Usnic Acid to Prevent Infections by Creating a Protective Barrier in an In Vitro Study
Previous Article in Journal
Cyto-Genotoxic and Behavioral Effects of Flubendiamide in Allium cepa Root Cells, Drosophila melanogaster and Molecular Docking Studies
Previous Article in Special Issue
Involvement of Hypoxia-Inducible Factor 1-α in Experimental Testicular Ischemia and Reperfusion: Effects of Polydeoxyribonucleotide and Selenium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Endoplasmic Reticulum Stress and Cancer: Could Unfolded Protein Response Be a Druggable Target for Cancer Therapy?

by
Gregorio Bonsignore
,
Simona Martinotti
and
Elia Ranzato
*,†
DiSIT-Dipartimento di Scienze e Innovazione Tecnologica, University of Piemonte Orientale, Viale Teresa Michel 11, 15121 Alessandria, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(2), 1566; https://doi.org/10.3390/ijms24021566
Submission received: 16 December 2022 / Revised: 4 January 2023 / Accepted: 10 January 2023 / Published: 13 January 2023
(This article belongs to the Collection State-of-the-Art Bioactives and Nutraceuticals in Italy)

Abstract

:
Unfolded protein response (UPR) is an adaptive response which is used for re-establishing protein homeostasis, and it is triggered by endoplasmic reticulum (ER) stress. Specific ER proteins mediate UPR activation, after dissociation from chaperone Glucose-Regulated Protein 78 (GRP78). UPR can decrease ER stress, producing an ER adaptive response, block UPR if ER homeostasis is restored, or regulate apoptosis. Some tumour types are linked to ER protein folding machinery disturbance, highlighting how UPR plays a pivotal role in cancer cells to keep malignancy and drug resistance. In this review, we focus on some molecules that have been revealed to target ER stress demonstrating as UPR could be a new target in cancer treatment.

1. ER Stress and the UPR

The ER plays an important role in eukaryotic cells, as organelle which influence proteins in their synthesis, maturation and folding [1]. ER executes and regulates many of protein post-translational modifications guarantying a correct protein working [2,3].
An optimal protein folding is related to the disulfate-bond formation, that are influenced by several factors, for example ATP, calcium, and oxidizing environment [4]. To guarantee a correct protein folding preventing the accumulation of misfolded or unfolded proteins, ER possesses the ERQC mechanism (ER Quality Control Compartment) [5,6].
ER can be distinguished into two categories: rough and smooth [7]. RER (rough endoplasmic reticulum) is dotted with ribosomes and it is crucial for production, correct folding, quality check and delivery of protein [7].
SER (smooth endoplasmic reticulum) is not dotted with ribosomes, and it is paired with smooth slippery fats [6]. Into the SER take place both the production and the metabolism of steroid hormones and fats [8].
The protein quality control is a critical step for the cell survival. ER controls proteins and, in case of misfolding or unfolding, it mediates the ERAD (Endoplasmic-reticulum-associated protein degradation) [9,10,11]. If proteins are unable to pass ERQC, ERAD identifies and destructs them by a proteolytic system [9,12]. Unfortunately, sometimes ERQC functions can be impaired, and this condition causes the development of different serious protein folding diseases, such as neurodegenerative diseases, cardiac diseases, and cancer [13,14].
The first cell component which is involved into signal transduction and homeostatic changes sensing is the ER, giving feedback to other organelles [15]. Typically, proteins are folded (tertiary or quaternary structure) in the ER [1]. However, alterations of ATP level, calcium concentration or redox conditions limit the protein-folding ability of the ER, and as a consequence, there is an aggregation and accumulation of unfolded proteins, determining ER stress [16].
UPR is an adaptive response which is used for re-establishing protein homeostasis, and it is triggered by ER stress [17,18]. Three ER proteins mediates UPR activation, i.e., IRE1, PERK, and ATF6. These proteins have a luminal domain binding an ER chaperone called GRP78 which is inactive in normal conditions. In case of ER stress, the luminal domains dissociate from GRP78, causing their activation [19]. UPR mainly plays three roles:
-
reducing ER stress and re-establishing ER homeostasis → adaptive response;
-
blocking UPR when ER homeostasis is restored → feedback control;
-
regulating apoptosis [20,21].

2. UPR Pathways

2.1. IRE1

IRE1 (Inositol-requiring enzyme 1) is a transmembrane protein of ER (type I), with serine/threonine kinase activity, detecting ER stress through its N-terminal luminal domain, starting also the most common UPR signalling pathway [22]. IRE1 has two isoforms: IRE1α and IRE1β.
IRE1α is the most studied and is present in all eukaryotic cells. In case of unfolded proteins accumulation, IRE1 oligomerizes in the ER lumen and it starts the autophosphorylation [23].
After the activation, IRE1 cuts XBP1 mRNA, causing a shift in its codon reading frame, this condition triggers the formation of a new C-terminal domain which includes an active domain of transactivation, sXBP1 [4,5,24,25].
sXBP1, in turn, provokes the upregulation of UPR-related genes implicated in protein folding, and in translocation to the ER and ERAD [26,27].
IRE1 enlists TRAF2, and triggers ASK1 [28]. ASK1 causes the activation of JNK and p38, a MAPK [28,29].
Then, JNK molecules move to the membrane of mitochondria leading to Bim activation and the inhibition of Bcl-2. On the other hand, the phosphorylation of p38 MAPK provokes the activation of a transcriptional factor CHOP, causing the increasing of Bim and DR5 expression, at the same time decreasing Bcl-2 expression, this condition leads to the apoptosis initiation [30,31].
Bax and Bak can bind to IRE1 and trigger it, interact with IP3R inducing the release of Ca2+ from the ER [32].

2.2. PERK

The attenuation of mRNA translation is induced by an ER-resident transmembrane protein called PERK (protein kinase R-like endoplasmic reticulum kinase), which works as sensor of ER stress and contains a luminal domain similar to IRE1 [33,34].
Usually, PERK is linked to GRP78, and, after its activation, it blocks the entrance of newly synthesized proteins into the ER (which is already stressed). After the estrangement from GRP78, PERK forms a dimer and induces its autophosphorylation and activation [35].
This condition is possible for the inactivation of elF2 by the phosphorylation of serine 51 [13]. The inhibition of elF2α is caused by a guanine nucleotide exchange factor complex that leads elF2 to its active GTP-bound form [36], this last reduces the excess of misfolded proteins and mitigates ER stress [11].
Moreover, elF2 phosphorylation plays another role; in fact, it allows the translation of UPR-dependent genes, for example ATF4, codifying for different upstream open reading frames [37,38]. ATF4 triggers the expression of ER stress target genes (like CHOP, growth arrest, GADD34 and ATF3) [39,40].

2.3. ATF6

ATF6 (activating transcription factor 6) is a transmembrane protein of ER (type II), in case of ER stress conditions, dissociates from GRP78 (the release from GRP78 enables ATF6) and moves to the Golgi apparatus for additional proteolytic processing [41,42]. Into the Golgi, two enzymes called site proteases-1 and 2 (S1P-S2P) make the proteolytic cleavage of the full length ATF6 (90 kDa) [43,44].
Then, the cleaved N-terminal cytosolic domain of 50-kDa bZIP (cytosolic basic leucine zipper) moves to the nucleus binding to CRE (the ATF/cAMP response elements) and ERSE-1 to induce the transcription of some target protein like GRP78, XBP-1, and CHOP [45,46]. Thereby, during a prolonged ER stress, CHOP can be activated by IRE1, PERK, and ATF6, and then it leads to apoptosis [47,48].
The cleaved ATF6 translocate into the nucleus and works as an active transcription factor to upregulate proteins that improve ER folding capacity (for example two chaperons like GRP78 and GR94), and folding enzymes (for example PDI) [45,49,50,51].

2.4. GRP78

GRP78, also known as BiP, is a member of the HSP70 family. It is localized on the membrane of ER of all eukaryotes [52]. GRP78 is composed of 654 amino acids. It edits the folding and assembly, and also avoids the transport of protein (or subunits of them) that are misfolded [53,54,55]. The expression of GRP78 is increased in case of ER stress. For example, GRP78 is upregulated during
-
sugar abrogation;
-
inhibition of protein glycosylation induced by particular reagents;
-
intercellular calcium storage disturbance [56].
GRP78 is soluble in water, with only small hydrophobic parts; however, these last components are essential for its function, for example to recognize the unfolded proteins addressed either to the degradation or refolding mechanisms [57]. GRP78 possesses two domains called ABD (or NBD) locates at the N-terminal and SBD at the C-terminal [58].
There is 60% homology between GRP78 and the HSP70 family, and in particular, the most conserved domains are ABD and SBD. The most conserved sequence of the HSP70 family belongs to ABD domain [52,58,59]. However, GRP78 differs in protein expression regulation, nevertheless it is an HSP70 family A member, and owns the properties of abnormal protein binding in case of stress. In addition, the protein synthesis inhibitor cycloheximide is effective on GRP78 [60,61] (cycloheximide can inhibit protein synthesis in eukaryotic cells) [62].
The state of ER protein folding is checked by three ER-localized transmembrane UPR signal sensors, such as PERK, IRE1, and the transcription factor ATF6.
Under homeostatic conditions, these UPR sensors are maintained inactive by the interaction of their luminal domains with GRP78.
In response to ER stress, accumulated unfolded proteins sequester GRP78 from the UPR sensors, which promotes the activation of IRE1, PERK, and ATF6 and the induction of GRP78 [25,63].

2.5. UPR and Apoptosis

UPR acts through the transitional (at transcriptional and transcriptional level) attenuation of the enzymes folding, the induction of ER chaperones, and ERAD involved proteins to relieve protein aggregation in the ER as an adaptive response.
In case of important and extended ER stress, the UPR triggers some apoptotic pathways [50]. ER stress induces conformational change in ER membrane through the proapoptotic BH3 protein; moreover, it allows Ca2+ transmigration to the cytosol, which provokes the activation of m-calpain and, subsequently, the cleavage and the activation of procaspase 12 and caspase cascade [64,65,66,67].
CHOP, one of the principal UPR downstream effectors, inhibits Bcl-2, provoking growth arrest and the activation of GADD34 and ERO1, and all these elements promote apoptosis [68,69].
The upregulation of GADD34 by CHOP causes a feedback inhibition of elF2α phosphorylation; as a consequence, in cell death and survival, the role of CHOP may be context dependent. This phenomenon could allow the restoration of translation, which could be positive; however, if translation continues also in ER-stress conditions, the accumulation of anomalous proteins may compromise the ER folding ability, causing the death of cells.
After IRE1 activation, JNK is bound by IRE1 which recruits TRAF2; this condition causes both the release of the procaspase 12 from the ER and the activation of apoptosis signal-regulating kinase 1 and JNK [28,70]. Additionally, PUMA and NOXA are activated by ER stress, and this condition provokes BAX and BAK activation and apoptosis [71].
Activation and maintenance of representative UPR pathways in cells treated with low concentrations of chemical ER stress inducers, showed that survival, in case of stress, is reached through intrinsic mRNA instabilities and proteins that induce apoptosis [72].
Autophagy is a catabolic process, leading to several effects, in particular, the degradation and recycling of cytosolic, ageing, or misfolded proteins and excess or faulty organelles. ER stress provokes autophagy and promotes cell survival; in fact, in case of starving conditions, it enables the intracellular resources utilization [73].
In cells where GRP78 is reduced by small interfering RNA (siRNA), the ER structure is compromised, and autophagosome formation under ER-stress or starvation is suppressed [74].

2.6. ER Stress/Calcium-Mediated Apoptosis

The decrease of ER Ca2+ causes protein misfolding and chronic mitochondrial Ca2+ overload, which lead to apoptosis through Bcl-2-dependent pathway [75]. The localization and oligomerization of pro-apoptotic Bcl-2 proteins, Bax and Bak, are triggered by ER stress, promoting Ca2+ release from the ER into the cytosol [50], through IP3Rs and RyRs [67,76], which are linked to the apoptotic signal transduction mechanism [77,78,79].
When [Ca2+]cyt increases, it leads to the activation of Ca2+-dependent cysteine protease m-calpain, which is involved in several intracellular processes, for example apoptosis, cell cycle progression, differentiation, and signal transduction [80,81].
m-calpain is well-known to cleave and trigger the ER-resident procaspase-12 [64,82], involved in the ER stress-induced cell death pathway in differentiated PC12 cells [83]. The activation of caspase-12 leads to the activation of procaspase-9 and then to the activation of caspase-3 apoptotic mechanism [65].
The increasing of cytosolic Ca2+ causes its uptake into the mitochondrial matrix, inducing a depolarization of the inner mitochondrial membrane and a perturbation of the outer membrane permeability [35]. This leads to the cytochrome c release and to the activation of the apoptosome by Apaf-1, causing apoptosis [84].
The most important actor in the regulation of the ER stress-induced apoptosis is CHOP [85]. CHOP is a basic leucine zipper-containing transcription factor, suppressing the expression of Bcl-2 and triggering the transcription of many genes favouring apoptosis [69,86]. The release of procaspase-12 from TRAF2 (and then its activation) is triggered by the association of IRE1/TRAF2 and ER stress [87,88]. The activated caspase-12 results in apoptosis activation [70].

2.7. MAM and ER Stress

The ER is closely associated with mitochondria through mitochondria associated membranes (MAM), leading to a strong functional interaction between these two organelles [89]. For example, the acute ER stress impairs mitochondrial function in rat and mouse hearts, while inhibition of mitochondrial respiration by using rotenone or antimycin A also increases the ER stress [90].
ER is a key site of intracellular calcium storage, and ER stress leads to intracellular calcium overload by disrupting calcium homeostasis [91]. ER stress-mediated calcium overload contributes to mitochondrial damage. Furthermore, in response to Ca2+ overload, the ER stress also rises ROS production directly through NADPH oxidase 4, and then through impairment of mitochondrial electron transport [92].
Taken together, these data point out the existence of a positive feedback loop between mitochondrial dysfunction and ER stress.
Moreover, the activation of initial phase of UPR, i.e., the pathways to solve the ER stress by expanding the ER, upregulating chaperones and by causing a (temporary) translation stop, is accompanied by ER morphology changes and the consolidation of ER and mitochondria contact sites.
In fact, considering that the folding of newly proteins is one of most energy-requiring processes, this strengthening of the contact sites makes sense. In ER stress-induced by tunicamycin, an important number of mitochondria relocate towards the perinuclear ER. These mitochondria exhibit a transmembrane potential increase, Ca2+ uptake rise, higher ATP production, and increased oxygen consumption [93]. Therefore, tightening the ER-mitochondrial contacts could favour the increase in intracellular ATP level necessary to sustain the pro-survival ER machinery.
Conversely, if ER stress is too severe and cannot be resolved by the UPR, the signalling pathways initiated by the ER stress sensors will turn on a lethal signal, ultimately causing cell death, typically in an apoptotic way [94]. Several studies have also validated a direct relationship between changes in MAM components and deregulated Ca2+ transfer and apoptotic sensitivity during ER stress [95,96].
Likewise, new data suggested that the MAM could be involved not only in Ca2+ signals, but also in some toxic lipid oxidation products to the mitochondria [97]. This phenomenon could be very important as lipid peroxides may be responsible for spreading ROS signalling, to control mitochondrial Ca2+ uptake and cell death choice after ER stress [98].
Additionally, to increase importance of MAM involvement in ER stress, recent studies displayed that ER-mitochondria contacts may be crucial in the formation of autophagosomes [99].

3. UPR as Therapeutic Target

One of tumorigenesis hallmark is the unrestrained growth of the transformed cells; cancers are continuously challenged by a limited oxygen and nutrients supply due to inadequate vascularisation. Moreover, some hematopoietic cancers often show that secretory proteins have increased production, such as multiple myeloma cells producing immunoglobulins [100]. All these circumstances induce an ER stress and UPR activation.
Different cancer types are linked to ER protein folding machinery disturbance, demonstrating as the correct folding process is a key for signalling pathway proteins [101]. A huge amount of evidence demonstrates how the UPR is a crucial process which is very important for cancer cells to keep malignancy and drug resistance.
UPR is a very active research area because different components are triggered or repressed in malignant cancers [102]. Recent indications suggest that UPR molecular components could be useful as prognostic and diagnostic markers for cancer progression and response to chemotherapeutics [94].
In a great variety of cancers, UPR pathways are activated, and they are essential for tumour microenvironment creation and maintenance. In fact, in cancer cells a great number of events can take place, such as:
-
over-expression of XBP1;
-
activation of ATF6;
-
phosphorylation of eIF2α;
-
induction of ATF4 and CHOP
-
upregulation of GRP78;
-
upregulation of glucose-regulated protein 94 (GRP94, also known as gp96 or HSP90b1);
-
upregulation of GRP170 [103,104].
From animal studies, it was discovered that XBP1 is important for tumour growth in vivo. In fact, in Xbp1-/- and Xbp1-knockdown mice cells, cancers cannot develop [105]. In addition, ER stress can cause anti-apoptotic reactions. After XBP1 activation, glycogen GSK3b [106] induces p53 phosphorylation, which causes an increasing of its degradation and avoids p53 dependent apoptosis for cancer cells. Furthermore, during ER stress, NFκB is activated and induces anti-apoptotic responses [107].
An important role is played by heat shock proteins; in fact, they help cancer cell adaptation against stress associated with oncogenesis both to repair damaged proteins (protein refolding) and to degrade them. Moreover, heat shock proteins are involved in cell proliferation and in resistance to different anti-tumoral drugs that cause apoptosis. For example, HSP90 interfaces with different key proteins in inducing prostate cancer progression, comprised both wild-type and mutated AR, HER2, ErbB2, Src, Abl, Raf, and Akt [108,109].
In a great variety of tumours, GRP78 is largely expressed at high levels conferring resistance against therapies in both proliferating and dormant cancer cells. Animal models with reduced GRP78 level shown significant impediments in tumour growth.
GRP78 mediated-cancer progression has been explained though three major mechanisms:
-
activation of cancer proliferation;
-
suppression of apoptosis;
-
stimulation of angiogenesis [110,111].
ER stress has been involved in different levels of tumour development. The current idea is that, during early tumorigenesis and before angiogenesis occurs, activation of the UPR causes the cell cycle arrest in G1 phase and the activation of p38, both of which induce a dormant state.
ER stress also promotes anti-apoptotic NF-κB and silences p53-dependent apoptotic signals. If the balance of early cancer development takes off against cell death, ER stress can further induce the aggressive growth of cancer cells by enhancing their angiogenic ability. For example, the induction of GRP170 (which is a BiP-like protein that acts as a chaperone for VEGF [112]) causes an increasing of VEGF secretion.
Emerging data indicate that agents disturbing UPR pathway may be utilized as promising anticancer drug. However, due to the dual role of the UPR in cell survival/death, and depending on the type of cancer, molecules that either provoke severe ER stress and cell death or compounds blocking the pro-survival role of disturbed UPR of tumour cells could be used either alone or in combination with conventional anticancer treatments [113].

3.1. GRP78 as Target

In different kinds of cancer, the expression of GRP78 is often elevated if compared to healthy tissues. This has been noticed in different diseases such as hepatocellular carcinoma [114], gliomas [115], prostate [116], and gastric cancers [117]. However, a study on lung cancer showed a link between the overexpression of GRP78 and better prognosis [118]. Despite this, GRP78 results overexpressed in more cases, and, for this reason, the consensus is that its expression is linked to poor prognosis and strong tumour aggressiveness.
For example, it is well-known that an important expression of GRP78 is correlated with poor prognosis and lymph node metastasis in gastric tumours [117]; moreover, these data are reinforced by preclinical studies which show how the silencing of GRP78 reduces invasion in vitro and tumour growth and metastasis in vivo [117].
In prostate tumours, high GRP78 activity is linked to a reduction of patient survival [116], while, in breast tumours [119,120], a shorter time to recidivism is associated with high GRP78 expression [120].
In addition, a GRP78 overexpression has been described both in tumour with acquired anti-oestrogen resistance and in oestrogen-receptor positive breast cancer cells [119]. However, a study reported a correlation between oestrogen receptor positivity and GRP78 or XBP1 expression, showing that the oestrogen upgrade can promote the expression of GRP78 and XBP1 [121].
Moreover, GRP78 might contribute significantly to therapy resistance. In glioma cells and breast cancer cells, the inhibition of GRP78 expression improves sensitivity to chemotherapy [115,122]; in particular, in resistant breast cancer cells, it re-establishes sensitivity to anti-oestrogens [119]. On the other hand, GRP78 overexpression induces resistance to chemotherapeutics [114,115,121,123] and anti-oestrogens [114,119].
GRP78-mediated therapy resistance is not a general rule; in renal carcinoma cells, GR78 knockdown causes resistance to chemotherapy [124].
Nevertheless, GRP78 is a possible target for cancer therapies. Chen et al. described the use of the GRP78-promoter to induce the expression of an apoptotic pathway [125,126]. Another strategy is the direct attack against GRP78 expressing cells. For example, EGCG ((−)-epigallocatechin-3-gallate), a green tea polyphenol, can block GRP78 binding on its ATP-binding domain. A treatment with EGCG can damage glioma cells resistance to temozolomide [115]. EGCG is also effective to reduce the proliferation of mesothelioma cell lines, overexpressing GRP78 [127].
An alternative is versipelostatin, a transcriptional inhibitor of GRP78, which causes selective death of glucose-deprived cells and stops cancer development in vivo [128]. XPB1 and ATF4 expression are repressed by the treatment with this compound, but only during glucose deprivation, not after treatment with tunicamycin or A23187. Combining versipelostatin with cisplatin induces growth inhibition [129].
The evidence indicates that GRP78 plays different roles in cancer cells, not only in the ER [129]. In fact, some isoforms of GRP78 have been detected in the nucleus, mitochondria, cytosol, and also in the ER membrane [123,130,131,132,133].
Expression of GRP78 on the cell surface could be a great therapeutic target using specific therapeutic antibodies. An example is PAT-SM6, an autoantibody isolated from a gastric cancer patient [133]. PAT-SM6 is directed against a particular isoform of GRP78, showing the ability to inhibit the development of gastric carcinoma in vivo [134].
Moreover, this antibody can induce specific apoptosis in multiple myeloma cells, while it is safe against non-malignant cells [135]. A phase I clinical trial shown that PAT-SM6 could also be effective against melanoma [136].

3.2. PERK and IRE1 as Targets

A comparison between malignant tissue and normal tissue shows different expression of ATF4, the downstream factor of PERK [137]. In neoplastic tissues, the expression of ATF4 is higher than in normal tissues; so, this factor plays a crucial role in the response against chemotherapy. In addition, expression of ATF4 correlates with cisplatin resistance in lung cancer cell lines [138]. In the case of an ATF4 overexpression, it could induce multidrug resistance against cisplatin, doxorubicin, etoposide, irinotecan, and vincristine, but not to 5-fluorouracil [138,139]. The resistance to cisplatin causes an intracellular glutathione level increase [139]. On the contrary, ATF4 knockout cells show a decreased glutathione biosynthesis and higher sensitivity to anti-cancer treatments.
However, the UPR is helpful not only to improve chemotherapy; in fact, it can also have a significant impact on the radiotherapy efficacy. In breast cancer, radiotherapy triggers the PERK-pathway of the UPR, and this is induced by increased PERK protein levels, phosphorylated eIF2α, ATF4, and LAMP3 [140]. Moreover, in vivo studies on rat intestinal epithelial cells show how irradiation promotes three important consequences:
-
GRP78 expression;
-
eIF2α phosphorylation;
-
XBP1 splicing [141].
According to experimental data, this stress response is not linked to ATF6 [141]. Only during knockdown of PERK, ATF4 or LAMP3 can enhance the sensitivity of breast cancer cells to radiotherapy [140]. Furthermore, a knockdown of GADD34, an essential element to prolong eIF2α phosphorylation, causes radio-resistance; on the contrary, a treatment with a pharmacological PERK inhibitor (GSK2606414) provokes radio-sensitization [140].
Another PERK inhibitor (GSK2656157) shows the ability to suppress the development of xenografted tumours through the reduction of the vascular density [142,143]. Tamoxifen is a well-known anti-oestrogen drug which can induce UPR components, such as GRP78 and LAMP3 [121,140].
In addition to radio-sensitization of breast cancer cells, knockdown of LAMP3 sensitizes cells to tamoxifen [140]. In breast tamoxifen-tolerant cancer, cells have been noted an increased expression of LAMP3 which, if silenced, provokes a decreasing of tamoxifen-resistance [140].
Another arm of UPR is upregulated in breast cancer cells: XBP1-pathway [144]. In fact, in this kind of tumour, XBP1 is co-expressed with the oestrogen receptor. XBP1 overexpression causes the independent growth of oestrogen receptor positive cells regardless of the presence of the hormones. Furthermore, this condition makes cells more resistant to the anti-oestrogen drugs such as tamoxifen and faslodex [144]. According to data, higher levels of spliced XBP1 are linked to more aggressive breast tumours and poor prognosis [145].

3.3. ER Stress-Inducing Agents as Anti-Cancer Therapies

There are different molecules which can induce ER stress and some of them have the potential to be used as anti-cancer therapies. For example, eeyarestatin 1 is a small molecule which induces ER stress by preventing ER-associated degradation [146]. Eeyarestatin 1 shows a synergistic behaviour with bortezomib and is selective against cancer cells. Furthermore, tunicamycin-induced ER stress can increase sensitivity of some cancers to therapies, for example: breast cancer cells sensitivity to radiotherapy [147] and ovarian cancer cells sensitivity to cisplatin and carboplatin [148].
Other data have shown different situations where ER-stress can induce resistance to chemotherapy [149,150]. Tunicamycin, through induction of GRP78, can significantly reduce the apoptosis induced by chemotherapy [151] and, as a consequence, GRP78 silencing can reduce the sensitivity to tunicamycin effects.
Thus, depending on the treatment given, pharmacological induction of ER stress can be effective both for promoting sensitivity and inducing resistance to anti-tumoral therapies.
Another important study led by Ledoux et al. shows that via the UPR, glucose withdrawal induces the expression of P-glycoprotein in hepatic cancer cells [152], enhancing the efflux of chemotherapeutic drugs.

3.4. UPR-Induced Autophagy Helps to Survive ER Stress

Some conditions (hypoxia and detachment from the extracellular matrix) and compounds (A23187, tunicamycin, thapsigargin and brefeldin A) which cause ER-stress can induce both the UPR and the autophagy [153,154].
Autophagy is an important way to reduce ER stress and to allow survival of cells [73,153,155] but this effect was discovered only in cancer cells, not in normal cells [153].
As shown by Kouroku et al., polyglutamine aggregates induce ER stress leading to autophagy as a degradation mechanism [156]. In this case, the activation of autophagy is provoked by the PERK-arm of the UPR. Dominant-negative PERK and mutated non-phosphorylatable eIF2α avoid the conversion of LC3-I to LC3-II, while the phosphorylation of eIF2α induces ATG12 expression [156].
In addition, in neoplastic conditions, a resistance against therapies is provoked by ER stress-induced autophagy.
Pharmacological-induced ER stress, before the cisplatin therapy, can induce autophagy and confer resistance against cisplatin-induced apoptosis [157], whereas in breast cancer, PERK-dependent autophagy is induced by radiotherapy [147,158]. It is possible to sensitize cells to radiotherapy though pharmacological inhibition of autophagy or PERK-pathway silencing [140,158].
Moreover, autophagy can be induced also by tamoxifen treatment, and this process is mediated by ATF4-induced LAMP3 [159,160] or by GRP78-dependent inhibition of mTOR [119]. Another possible treatment of breast cancer is bortezomib, that leads to an ATF4-dependent increase in LC3B and autophagy, favouring bortezomib-resistance [161].
Other reports have indicated that not the PERK-arm, but the IRE1-arm is responsible for UPR-mediated autophagy. In neuroblastoma, ER stress caused by amino acid starvation, thapsigargin, or tunicamycin promotes autophagy [73]. However, this condition can be blocked by IRE1 silencing or a pharmacological treatment with a JNK inhibitor.
Interestingly, cells lacking PERK or ATF6 induce autophagy similar to wild-type cells, suggesting how autophagy is independently triggered.
UPR-independent pathway can also induce autophagy in response to ER stress. ER stress caused by thapsigargin or tunicamycin triggers autophagy via protein kinase Cθ (PKCθ) [162]. PKCθ activation takes place independently from UPR sensors. ER stress-activated autophagy can be blocked by PKCθ silencing or its pharmacological inhibition but, this condition can be prevented by chelating intracellular Ca2+, while the kinase is not reactive to amino acid starvation.
Furthermore, when cytosolic Ca2+ levels increase, AMPK is triggered by CAMKK-β, this condition leads to ER-stress which causes the inhibition of mTOR [162], inducing autophagy.

4. Compounds Targeting the UPR

As stated, ER is an important organelle which controls protein folding, calcium reserve, and lipid production. New proteins move to the ER for other adjustments, such as glycosylating, folding, and disulphide bond formation [163]. To guarantee the correct maturation and folding, ER uses the ERQC mechanism.
Usually, proteins are inspected by ERQC, or they are degraded through by ERAD [164]. When cells are affected by bad stimuli, such as calcium disruption, glucose deprivation and redox imbalance, ER is full of unfolded or misfolded proteins, triggering ER stress.
UPR is linked to a correction of folding and also to a decreasing of the proteins amount into ER and gene translation [165], but, if the stress is out of control, cells could trigger the apoptotic mechanism [166]. In traditional medicine, natural compounds played an important role in the treatment of different diseases. In recent times, many of these methods were rediscovered. In particular, some natural molecules have been revealed to target ER stress, which has a central role in the pathophysiological progress of diseases (see Table 1). For example, thapsigargin, produced by Thapsia garganica, is deemed an ER stress inducer which uses competitive inhibition of SERCA.

4.1. Natural Compounds and ER Stress-Related Apoptosis

4.1.1. PERK-eIF2α-CHOP

Baicalein (5,6,7-trihydroxyflavone), a flavone originally isolated from the roots of Scutellaria baicalensis and Scutellaria lateriflora may trigger ER stress, regulating CHOP, JNK (upregulation) and Bcl-2 family (downregulation), thus, it can induce apoptosis and autophagy in hepatocellular carcinoma cells (HCC) [168].
Berberine, a quaternary ammonium salt from plants in the genus Berberis, has been employed against several diseases, such as dyslipidaemia, hyperglycaemia, and obesity. It is an antidiabetic agent which can reduce the phosphorylation levels of PERK, eIF2α, and IRS-1-ser307, inducing the reduction of ER stress to increase insulin sensitization in Hep G2 cells [169]. Moreover, berberine enhanced GRP78 by suppression of ubiquitination/proteasomal degradation of GRP78 and activation of ATF6, suggesting as berberine can induces autophagic death via enhancing GRP78 levels [170].
A natural triterpenoid called celastrol, contained into Tripterygium wilfordii Hook, can increase the effect of BH3 mimetic drug ABT-737. This combination activates the eIF2α-ATF4 signalling pathway, then upregulates Noxa expression, which can induce apoptosis in HCC cells [171].
Another important natural compound is curcumin. In acute promyelocytic leukaemia (APL), it can inhibit ERAD and protease-regulated degradation and then induce the amassing of phosphorylated misfolded N-CoR. This condition causes the sensitization of APL to UPR-induced apoptosis [172].
Honokiol, a compound separated from Magnolia officinalis, may attenuate ER stress caused by apoptosis in torsion/detorsion testicular injury, it downregulates the expression of p-eIF2α and CHOP [173]. Honokiol treatment can also induce apoptotic pathway in human chondrosarcoma cell lines but not primary cells, triggering ER stress, as shown by changes in Ca2+ levels [174].
From Astragalus membranaceus a natural compound (astragaloside IV) has been separated, showing protective activity in diabetic nephropathy because of ER stress inhibition [175]. Specifically, it can downregulate PERK-ATF4-CHOP signalling reducing podocyte apoptosis in diabetic rats [176]. Moreover, astragaloside IV sensitized non-small cell lung cancer cells to cisplatin through suppressing ER stress and autophagy [177]
Another natural compound, resveratrol, a polyphenol contained into different fruits and in wine, induces the reduction of fat and body weight trough the regulation of lipid and glucose metabolisms. Resveratrol can downregulate AMPK signalling pathway and activate ER stress through the expression of eIF2α phosphorylation and CHOP [178].

4.1.2. GRP78 and IRE1-XBP1

Metformin was originally developed from natural compounds found in the plant Galega officinalis (also known as goat’s rue), a traditional herbal medicine in Europe. The mechanism of action of metformin involves an activation of AMP-activated protein kinase (AMPK). The activation of AMPK decreases cell injury during oxidative stress in part through the inhibition of mitochondrial permeability transition pore (MPTP) opening. Moreover, in addition to its anti-diabetic effect, a number of studies suggest that AMPK activators might exert an anti-cancer effect through the modulation of the UPR in ER stress conditions. Interestingly, the effects of metformin on the UPR have recently been described in different cancer cell lines [179,180]. Metformin co-treatment with bortezomib suppresses the induction of GRP78, impairing the autophagosome formation in myeloma cells and enhancing apoptosis of cancer cells [181].
EGCG, a compound in green tea, is an inhibitor of GRP78. It can act both as an antioxidant and as a pro-oxidant. ECGC, due to GRP78 inhibition, may greatly improve the therapeutic effect of temozolomide to cause glioblastoma apoptosis [182] from in vivo studies, for its effect on GRP78, EGCG could protect against cisplatin-induced nephrotoxicity by suppression of ER stress-mediated apoptosis in mouse renal tubular epithelial cells [183]. In malignant mesothelioma, EGCG can induce synergistic effects with gemcitabine and ascorbic acid [184].
Another molecule is nicotine, the main component of tobacco. Nicotine can diminish apoptosis induced by tunicamycin-mediated ER stress in PC12 cells, provoking a reduction of expression for ATF6, GRP78, and IRE1-XBP1 [185].
Furthermore, resveratrol possesses two isomers (trans and cis). Both isomers show antioxidant, anti-inflammatory, antitumor, and immunomodulatory properties. In particular, cis-resveratrol may suppress the expression of GRP78 and reduce the production of ROS in human macrophages [186]; in addition, it may abrogate the pro-survival IRE1-XBP1 signalling and activate the pro-apoptotic responses activating ER stress-induced apoptosis [187].

4.2. Natural Compounds and Calcium-Mediated ER Stress

In many Curcuma species, there are three principal constituents for curcuminoids: bisdemethoxycurcumin (BDMC), demethoxycurcumin (DMC), and curcumin. BDMC has shown, in different studies, the ability to induce arrest at S phase and subsequent apoptosis in human lung cancer NCI H460 cells, acting through elevation of Ca2+ and ROS and activation of ER stress; in fact, these events are preceded by GRP78, IRE1α, IRE1β, CHOP, ATF6α, ATF6β, and caspase-4 upregulation [188]. In acute myeloid leukaemia, curcumin can cooperate with carnosic acid to provoke apoptosis, the combination is cancer-selective cytotoxic inducing the disruption of Ca2+ homeostasis.
Baicalein promotes apoptosis in retina ganglion cells (N18). It causes MDA-MB-231 cell apoptosis through stimulation of ER stress, reduction of Bcl-2 expression, elevation of Ca2+, downregulation of mitochondrial membrane potential and upregulation of Bax expression [168].
Another interesting natural compound is camphene, a molecule isolated from the essential oil of Piper cernuum. It can turn on apoptosis in melanoma cells, thus inducing ER stress, which could be linked to calcium perturbation and mitochondria disorder [189].
A triterpenoid saponins isolated from Gynostemma pentaphyllum, called gypenosides shown its ability to activate apoptosis in human hepatoma cells inducing calcium-modulated ER stress and mitochondrial disorder [190].

4.3. Natural Compounds and Inflammation-Mediated ER Stress

The anti-inflammatory activity of natural compounds is an important subject of study. In fact, there is a link between ER stress and the inflammatory response. New data have related the UPR induction with various pro-inflammatory factors release such as IL-6, IL-8, and TNF-α [191]. In fact, some UPR signals can facilitate processes which lead to different inflammatory phenomena related to cancer progression [94]. In particular, the maintenance of an inflammatory microenvironment is an essential component of all tumours [94].
For example, curcumin exhibit anti-inflammatory activity against bacterial invasion Caco-2 cells and T84 cells (intestinal epithelial cells), downregulating ER stress and decreasing the expression of GRP78 and IRE1α-XBP1 signalling [192].
Another example is the sterol ergosta-7,22-dien-3-ol separated from the echinoderm Marthasterias glacialis. It shows anti-inflammatory functions, as confirmed by the upregulation of COX-2, iNOS, IL-6, and NF-κB. It is well-known that CHOP-modulated ER stress upregulates the inflammatory responses, however it can be attenuated by sterol ergosta-7,22-dien-3-ol [193].
The TXNIP/NLRP3 inflammasome results ER stress-associated and causes flogosis and cell death in the endothelial disorder. Some natural molecules such as EGCG, quercetin, and luteolin shown the ability to turn on AMPK signalling, this condition leads to the suppression of ER stress and TXNIP/NLRP3 inflammasome [194]. Additionally, curcumin, ilexgenin A, and astragaloside IV can suppress TXNIP/NLRP3 inflammasome through their interaction with AMPK signalling [195,196,197].

4.4. Metal (Ruthenium and Iridium) Based Compounds

BOLD-100 (ruthenium complex sodium trans-[tetrachlorido-bis(1Hindazole) ruthenate(III)] (BOLD-100/KP1339) is an inhibitor of stress-induced GRP78 upregulation, disrupting endoplasmic reticulum (ER) homeostasis and inducing ER stress and UPR.
BOLD-100 is linked to UPR response and is the most studied non-platinum metal-based anticancer drugs [198]. It displayed promising anticancer effects in several tumour models; in addition, clinical trials prove its safety [198].
Several cancer cells lines as well as tumour models highlighted that BOLD-100 induces a down-modulation of the GRP78 and an induction of ER stress [198]. Moreover, ruthenium possesses an important redox activity which allows it to interfere with the cellular redox balance via direct as well as indirect mechanisms [199]. Ruthenium compounds can play a role in Fenton-like reactions, leading to generation of reactive oxygen species (ROS), and can also induce depletion of the intracellular GSH pools, which induce an increased susceptibility in cells to endogenous and exogenous oxidative stress. As a consequence of this (redox) stress, treatment with ruthenium induces apoptosis of cancer cells via the mitochondrial pathway [200,201].
Recently, BOLD-100 obtained encouraging results in the treatment of some non-responsive tumours like breast, gastric, colorectal, and pancreatic cancers.
In malignant mesothelioma cell lines, treatment with BOLD-100 induce an increase in ROS production and Ca2+ release from the ER, producing an activation leading to ER stress and, ultimately, to cell death [202].
Some authors have also described as iridium complexes may show effects on ER status. In particular, it has been suggested [203,204] that Ir complex–peptide hybrids (IPHs) induce ER stress and decrease the mitochondrial membrane potential, thus triggering intracellular signalling pathways and resulting in cytoplasmic vacuolization in Jurkat cells.

5. Conclusions

The role of UPR in cancer development and progression has been demonstrated; however, we do not recognize exactly the finest regulations and implications of UPR signalling in tumour cells.
UPR was initially considered to be pro-survival adaptive machinery designed to reduce levels of unfolded proteins and to restore ER homeostasis. This UPR implication may be true for cancer cells at least in the first stages of developments, while UPR could also be utilized in other ways to benefit cancer in advancement and invasion.
Some approaches, as well as some molecules, have demonstrated that UPR could be a druggable target to treat tumour, but the next challenge will be to correctly decipher the UPR signalling details and intricacies to specifically target cancer cells lines and select patients to benefit from this therapeutic option.

Author Contributions

Conceptualization, G.B., S.M. and E.R.; Writing—original draft preparation, G.B. and E.R.; writing—review and editing, S.M. and E.R.; supervision, E.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Braakman, I.; Bulleid, N.J. Protein folding and modification in the mammalian endoplasmic reticulum. Annu. Rev. Biochem. 2011, 80, 71–99. [Google Scholar] [CrossRef] [PubMed]
  2. Ellgaard, L.; Helenius, A. Quality control in the endoplasmic reticulum. Nat. Rev. Mol. Cell Biol. 2003, 4, 181–191. [Google Scholar] [CrossRef] [PubMed]
  3. Ferri, K.F.; Kroemer, G. Organelle-specific initiation of cell death pathways. Nat. Cell Biol. 2001, 3, E255–E263. [Google Scholar] [CrossRef] [PubMed]
  4. Hetz, C.; Martinon, F.; Rodriguez, D.; Glimcher, L.H. The unfolded protein response: Integrating stress signals through the stress sensor IRE1α. Physiol. Rev. 2011, 91, 1219–1243. [Google Scholar] [CrossRef] [PubMed]
  5. Shen, X.; Ellis, R.E.; Lee, K.; Liu, C.Y.; Yang, K.; Solomon, A.; Yoshida, H.; Morimoto, R.; Kurnit, D.M.; Mori, K.; et al. Complementary signaling pathways regulate the unfolded protein response and are required for C. elegans development. Cell 2001, 107, 893–903. [Google Scholar] [CrossRef] [Green Version]
  6. Gaut, J.R.; Hendershot, L.M. The modification and assembly of proteins in the endoplasmic reticulum. Curr. Opin. Cell Biol. 1993, 5, 589–595. [Google Scholar] [CrossRef]
  7. English, A.R.; Zurek, N.; Voeltz, G.K. Peripheral ER structure and function. Curr. Opin. Cell Biol. 2009, 21, 596–602. [Google Scholar] [CrossRef] [Green Version]
  8. Voeltz, G.K.; Rolls, M.M.; Rapoport, T.A. Structural organization of the endoplasmic reticulum. EMBO Rep. 2002, 3, 944–950. [Google Scholar] [CrossRef]
  9. Benyair, R.; Ron, E.; Lederkremer, G.Z. Protein quality control, retention, and degradation at the endoplasmic reticulum. Int. Rev. Cell Mol. Biol. 2011, 292, 197–280. [Google Scholar] [CrossRef]
  10. Walter, P.; Ron, D. The unfolded protein response: From stress pathway to homeostatic regulation. Science 2011, 334, 1081–1086. [Google Scholar] [CrossRef]
  11. Yan, M.M.; Ni, J.D.; Song, D.; Ding, M.; Huang, J. Interplay between unfolded protein response and autophagy promotes tumor drug resistance. Oncol. Lett. 2015, 10, 1959–1969. [Google Scholar] [CrossRef] [Green Version]
  12. Brodsky, J.L.; Skach, W.R. Protein folding and quality control in the endoplasmic reticulum: Recent lessons from yeast and mammalian cell systems. Curr. Opin. Cell Biol. 2011, 23, 464–475. [Google Scholar] [CrossRef] [Green Version]
  13. Söti, C.; Csermely, P. Chaperones and aging: Role in neurodegeneration and in other civilizational diseases. Neurochem. Int. 2002, 41, 383–389. [Google Scholar] [CrossRef]
  14. Dissemond, J.; Busch, M.; Kothen, T.; Mörs, J.; Weimann, T.K.; Lindeke, A.; Goos, M.; Wagner, S.N. Differential downregulation of endoplasmic reticulum-residing chaperones calnexin and calreticulin in human metastatic melanoma. Cancer Lett. 2004, 203, 225–231. [Google Scholar] [CrossRef]
  15. Lawless, M.W.; Greene, C.M. Toll-like receptor signalling in liver disease: ER stress the missing link? Cytokine 2012, 59, 195–202. [Google Scholar] [CrossRef]
  16. Bravo, R.; Parra, V.; Gatica, D.; Rodriguez, A.E.; Torrealba, N.; Paredes, F.; Wang, Z.V.; Zorzano, A.; Hill, J.A.; Jaimovich, E.; et al. Endoplasmic reticulum and the unfolded protein response: Dynamics and metabolic integration. Int. Rev. Cell Mol. Biol. 2013, 301, 215–290. [Google Scholar] [CrossRef] [Green Version]
  17. Kozutsumi, Y.; Segal, M.; Normington, K.; Gething, M.J.; Sambrook, J. The presence of malfolded proteins in the endoplasmic reticulum signals the induction of glucose-regulated proteins. Nature 1988, 332, 462–464. [Google Scholar] [CrossRef]
  18. Adams, G.A.; Rose, J.K. Incorporation of a charged amino acid into the membrane-spanning domain blocks cell surface transport but not membrane anchoring of a viral glycoprotein. Mol. Cell. Biol. 1985, 5, 1442–1448. [Google Scholar]
  19. Yadav, R.K.; Chae, S.W.; Kim, H.R.; Chae, H.J. Endoplasmic reticulum stress and cancer. J. Cancer Prev. 2014, 19, 75–88. [Google Scholar] [CrossRef]
  20. Oslowski, C.M.; Urano, F. Measuring ER stress and the unfolded protein response using mammalian tissue culture system. Methods Enzymol. 2011, 490, 71–92. [Google Scholar] [CrossRef] [Green Version]
  21. Oslowski, C.M.; Urano, F. The binary switch between life and death of endoplasmic reticulum-stressed beta cells. Curr. Opin. Endocrinol. Diabetes Obes. 2010, 17, 107–112. [Google Scholar] [CrossRef] [PubMed]
  22. Urano, F.; Bertolotti, A.; Ron, D. IRE1 and efferent signaling from the endoplasmic reticulum. J. Cell Sci. 2000, 113 Pt 21, 3697–3702. [Google Scholar] [CrossRef] [PubMed]
  23. Shamu, C.E.; Walter, P. Oligomerization and phosphorylation of the Ire1p kinase during intracellular signaling from the endoplasmic reticulum to the nucleus. EMBO J. 1996, 15, 3028–3039. [Google Scholar] [CrossRef] [PubMed]
  24. Yoshida, H.; Matsui, T.; Yamamoto, A.; Okada, T.; Mori, K. XBP1 mRNA is induced by ATF6 and spliced by IRE1 in response to ER stress to produce a highly active transcription factor. Cell 2001, 107, 881–891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Calfon, M.; Zeng, H.; Urano, F.; Till, J.H.; Hubbard, S.R.; Harding, H.P.; Clark, S.G.; Ron, D. IRE1 couples endoplasmic reticulum load to secretory capacity by processing the XBP-1 mRNA. Nature 2002, 415, 92–96. [Google Scholar] [CrossRef]
  26. Lee, K.; Tirasophon, W.; Shen, X.; Michalak, M.; Prywes, R.; Okada, T.; Yoshida, H.; Mori, K.; Kaufman, R.J. IRE1-mediated unconventional mRNA splicing and S2P-mediated ATF6 cleavage merge to regulate XBP1 in signaling the unfolded protein response. Genes Dev. 2002, 16, 452–466. [Google Scholar] [CrossRef] [Green Version]
  27. Scheuner, D.; Vander Mierde, D.; Song, B.; Flamez, D.; Creemers, J.W.; Tsukamoto, K.; Ribick, M.; Schuit, F.C.; Kaufman, R.J. Control of mRNA translation preserves endoplasmic reticulum function in beta cells and maintains glucose homeostasis. Nat. Med. 2005, 11, 757–764. [Google Scholar] [CrossRef]
  28. Nishitoh, H.; Matsuzawa, A.; Tobiume, K.; Saegusa, K.; Takeda, K.; Inoue, K.; Hori, S.; Kakizuka, A.; Ichijo, H. ASK1 is essential for endoplasmic reticulum stress-induced neuronal cell death triggered by expanded polyglutamine repeats. Genes Dev. 2002, 16, 1345–1355. [Google Scholar] [CrossRef] [Green Version]
  29. Nishitoh, H.; Saitoh, M.; Mochida, Y.; Takeda, K.; Nakano, H.; Rothe, M.; Miyazono, K.; Ichijo, H. ASK1 is essential for JNK/SAPK activation by TRAF2. Mol. Cell 1998, 2, 389–395. [Google Scholar] [CrossRef]
  30. Puthalakath, H.; O’Reilly, L.A.; Gunn, P.; Lee, L.; Kelly, P.N.; Huntington, N.D.; Hughes, P.D.; Michalak, E.M.; McKimm-Breschkin, J.; Motoyama, N.; et al. ER stress triggers apoptosis by activating BH3-only protein Bim. Cell 2007, 129, 1337–1349. [Google Scholar] [CrossRef] [Green Version]
  31. Deng, X.; Xiao, L.; Lang, W.; Gao, F.; Ruvolo, P.; May, W.S. Novel role for JNK as a stress-activated Bcl2 kinase. J. Biol. Chem. 2001, 276, 23681–23688. [Google Scholar] [CrossRef]
  32. Sano, R.; Reed, J.C. ER stress-induced cell death mechanisms. Biochim. Biophys. Acta 2013, 1833, 3460–3470. [Google Scholar] [CrossRef] [Green Version]
  33. Bertolotti, A.; Zhang, Y.; Hendershot, L.M.; Harding, H.P.; Ron, D. Dynamic interaction of BiP and ER stress transducers in the unfolded-protein response. Nat. Cell Biol. 2000, 2, 326–332. [Google Scholar] [CrossRef]
  34. Shi, Y.; Vattem, K.M.; Sood, R.; An, J.; Liang, J.; Stramm, L.; Wek, R.C. Identification and characterization of pancreatic eukaryotic initiation factor 2 alpha-subunit kinase, PEK, involved in translational control. Mol. Cell. Biol. 1998, 18, 7499–7509. [Google Scholar] [CrossRef] [Green Version]
  35. Rutkowski, D.T.; Kaufman, R.J. A trip to the ER: Coping with stress. Trends Cell Biol. 2004, 14, 20–28. [Google Scholar] [CrossRef]
  36. Mearini, G.; Schlossarek, S.; Willis, M.S.; Carrier, L. The ubiquitin-proteasome system in cardiac dysfunction. Biochim. Biophys. Acta 2008, 1782, 749–763. [Google Scholar] [CrossRef] [Green Version]
  37. Harding, H.P.; Novoa, I.; Zhang, Y.; Zeng, H.; Wek, R.; Schapira, M.; Ron, D. Regulated translation initiation controls stress-induced gene expression in mammalian cells. Mol. Cell 2000, 6, 1099–1108. [Google Scholar] [CrossRef]
  38. Lu, P.D.; Harding, H.P.; Ron, D. Translation reinitiation at alternative open reading frames regulates gene expression in an integrated stress response. J. Cell Biol. 2004, 167, 27–33. [Google Scholar] [CrossRef]
  39. Novoa, I.; Zeng, H.; Harding, H.P.; Ron, D. Feedback inhibition of the unfolded protein response by GADD34-mediated dephosphorylation of eIF2alpha. J. Cell Biol. 2001, 153, 1011–1022. [Google Scholar] [CrossRef] [Green Version]
  40. Ron, D. Translational control in the endoplasmic reticulum stress response. J. Clin. Investig. 2002, 110, 1383–1388. [Google Scholar] [CrossRef]
  41. Johnson, A.J.; Hsu, A.L.; Lin, H.P.; Song, X.; Chen, C.S. The cyclo-oxygenase-2 inhibitor celecoxib perturbs intracellular calcium by inhibiting endoplasmic reticulum Ca2+-ATPases: A plausible link with its anti-tumour effect and cardiovascular risks. Biochem. J. 2002, 366, 831–837. [Google Scholar] [CrossRef] [PubMed]
  42. Shen, J.; Chen, X.; Hendershot, L.; Prywes, R. ER stress regulation of ATF6 localization by dissociation of BiP/GRP78 binding and unmasking of Golgi localization signals. Dev. Cell 2002, 3, 99–111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Shen, J.; Prywes, R. Dependence of site-2 protease cleavage of ATF6 on prior site-1 protease digestion is determined by the size of the luminal domain of ATF6. J. Biol. Chem. 2004, 279, 43046–43051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ye, J.; Rawson, R.B.; Komuro, R.; Chen, X.; Davé, U.P.; Prywes, R.; Brown, M.S.; Goldstein, J.L. ER stress induces cleavage of membrane-bound ATF6 by the same proteases that process SREBPs. Mol. Cell 2000, 6, 1355–1364. [Google Scholar] [CrossRef] [PubMed]
  45. Haze, K.; Yoshida, H.; Yanagi, H.; Yura, T.; Kazutoshi, M. Mammalian transcription factor ATF6 is synthesized as a transmembrane protein and activated by proteolysis in response MULTIFUNCTIONAL ROLE OF GRP78/BIP 2313 to endoplasmic reticulum stress. Mol. Biol. Cell 1999, 10, 3787–3799. [Google Scholar] [CrossRef] [Green Version]
  46. Yoshida, H.; Haze, K.; Yanagi, H.; Yura, T.; Mori, K. Identification of the cis-acting endoplasmic reticulum stress response element responsible for transcriptional induction of mammalian glucose-regulated proteins. Involvement of basic leucine zipper transcription factors. J. Biol. Chem. 1998, 273, 33741–33749. [Google Scholar] [CrossRef] [Green Version]
  47. Zinszner, H.; Kuroda, M.; Wang, X.; Batchvarova, N.; Lightfoot, R.T.; Remotti, H.; Stevens, J.L.; Ron, D. CHOP is implicated in programmed cell death in response to impaired function of the endoplasmic reticulum. Genes Dev. 1998, 12, 982–995. [Google Scholar] [CrossRef]
  48. Schröder, M.; Kaufman, R.J. The mammalian unfolded protein response. Annu. Rev. Biochem. 2005, 74, 739–789. [Google Scholar] [CrossRef]
  49. Li, M.; Baumeister, P.; Roy, B.; Phan, T.; Foti, D.; Luo, S.; Lee, A.S. ATF6 as a transcription activator of the endoplasmic reticulum stress element: Thapsigargin stress-induced changes and synergistic interactions with NF-Y and YY1. Mol. Cell Biol. 2000, 20, 5096–5106. [Google Scholar] [CrossRef] [Green Version]
  50. Wu, J.; Kaufman, R.J. From acute ER stress to physiological roles of the Unfolded Protein Response. Cell Death Differ. 2006, 13, 374–384. [Google Scholar] [CrossRef]
  51. Yoshida, H.; Okada, T.; Haze, K.; Yanagi, H.; Yura, T.; Negishi, M.; Mori, K. ATF6 activated by proteolysis binds in the presence of NF-Y (CBF) directly to the cis-acting element responsible for the mammalian unfolded protein response. Mol. Cell Biol. 2000, 20, 6755–6767. [Google Scholar] [CrossRef]
  52. Brocchieri, L.; de Macario, E.C.; Macario, A.J. hsp70 genes in the human genome: Conservation and differentiation patterns predict a wide array of overlapping and specialized functions. BMC Evol. Biol. 2008, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  53. Hendershot, L.M.; Valentine, V.A.; Lee, A.S.; Morris, S.W.; Shapiro, D.N. Localization of the gene encoding human BiP/GRP78, the endoplasmic reticulum cognate of the HSP70 family, to chromosome 9q34. Genomics 1994, 20, 281–284. [Google Scholar] [CrossRef]
  54. Haas, I.G. BiP—A heat shock protein involved in immunoglobulin chain assembly. Curr. Top Microbiol. Immunol. 1991, 167, 71–82. [Google Scholar] [CrossRef]
  55. Gething, M.J.; Sambrook, J. Protein folding in the cell. Nature 1992, 355, 33–45. [Google Scholar] [CrossRef]
  56. Lee, A.S. Coordinated regulation of a set of genes by glucose and calcium ionophores in mammalian cells. Trends Biochem. Sci. 1987, 12, 20–23. [Google Scholar] [CrossRef]
  57. Ting, J.; Lee, A.S. Human gene encoding the 78,000-dalton glucose-regulated protein and its pseudogene: Structure, conservation, and regulation. DNA 1988, 7, 275–286. [Google Scholar] [CrossRef]
  58. Lindquist, S.; Craig, E.A. The heat-shock proteins. Annu. Rev. Genet. 1988, 22, 631–677. [Google Scholar] [CrossRef]
  59. Pelham, H.R. Speculations on the functions of the major heat shock and glucose-regulated proteins. Cell 1986, 46, 959–961. [Google Scholar] [CrossRef]
  60. Resendez, E.; Ting, J.; Kim, K.S.; Wooden, S.K.; Lee, A.S. Calcium ionophore A23187 as a regulator of gene expression in mammalian cells. J. Cell Biol. 1986, 103, 2145–2152. [Google Scholar] [CrossRef] [Green Version]
  61. Kim, Y.K.; Kim, K.S.; Lee, A.S. Regulation of the glucose-regulated protein genes by beta-mercaptoethanol requires de novo protein synthesis and correlates with inhibition of protein glycosylation. J. Cell. Physiol. 1987, 133, 553–559. [Google Scholar] [CrossRef] [PubMed]
  62. Müller, F.; Ackermann, P.; Margot, P. Fungicides, Agricultural, 3. Toxicology. In Ullmann’s Encyclopedia of Industrial Chemistry; Verlag Chemie: Hoboken, NJ, USA, 2000. [Google Scholar]
  63. Lee, A.H.; Iwakoshi, N.N.; Glimcher, L.H. XBP-1 regulates a subset of endoplasmic reticulum resident chaperone genes in the unfolded protein response. Mol. Cell Biol. 2003, 23, 7448–7459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Nakagawa, T.; Yuan, J. Cross-talk between two cysteine protease families. Activation of caspase-12 by calpain in apoptosis. J. Cell Biol. 2000, 150, 887–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Rao, R.V.; Castro-Obregon, S.; Frankowski, H.; Schuler, M.; Stoka, V.; del Rio, G.; Bredesen, D.E.; Ellerby, H.M. Coupling endoplasmic reticulum stress to the cell death program. An Apaf-1-independent intrinsic pathway. J. Biol. Chem. 2002, 277, 21836–21842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Scorrano, L.; Oakes, S.; Opferman, J.; Cheng, E.; Sorcinelli, M.; Pozzan, T.; Korsmeyer, S. BAX and BAK regulation of endoplasmic reticulum Ca2þ: A control point for apoptosis. Science 2003, 300, 135–139. [Google Scholar] [CrossRef]
  67. Zong, W.X.; Li, C.; Hatzivassiliou, G.; Lindsten, T.; Yu, Q.C.; Yuan, J.; Thompson, C.B. Bax and Bak can localize to the endoplasmic reticulum to initiate apoptosis. J. Cell Biol. 2003, 162, 59–69. [Google Scholar] [CrossRef] [Green Version]
  68. Marciniak, S.J.; Yun, C.Y.; Oyadomari, S.; Novoa, I.; Zhang, Y.; Jungreis, R.; Nagata, K.; Harding, H.P.; Ron, D. CHOP induces death by promoting protein synthesis and oxidation in the stressed endoplasmic reticulum. Genes Dev. 2004, 18, 3066–3077. [Google Scholar] [CrossRef] [Green Version]
  69. McCullough, K.D.; Martindale, J.L.; Klotz, L.O.; Aw, T.Y.; Holbrook, N.J. Gadd153 sensitizes cells to endoplasmic reticulum stress by down-regulating Bcl2 and perturbing the cellular redox state. Mol. Cell Biol. 2001, 21, 1249–1259. [Google Scholar] [CrossRef] [Green Version]
  70. Yoneda, T.; Imaizumi, K.; Oono, K.; Yui, D.; Gomi, F.; Katayama, T.; Tohyama, M. Activation of caspase-12, an endoplastic reticulum (ER) resident caspase, through tumor necrosis factor receptor-associated factor 2-dependent mechanism in response to the ER stress. J. Biol. Chem. 2001, 276, 13935–13940. [Google Scholar] [CrossRef] [Green Version]
  71. Li, J.; Lee, B. Lee, AS Endoplasmic reticulum stressinduced apoptosis: Multiple pathways and activation of p53-up-regulated modulator of apoptosis (PUMA) and NOXA by p53. J. Biol. Chem. 2006, 281, 7260–7270. [Google Scholar] [CrossRef] [Green Version]
  72. Rutkowski, D.T.; Arnold, S.M.; Miller, C.N.; Wu, J.; Li, J.; Gunnison, K.M.; Mori, K.; Akha, A.A.S.; Raden, D.; Kaufman, R.J. Adaptation to ER stress is mediated by differential stabilities of pro-survival and pro-apoptotic mRNAs and proteins. PLoS Biol. 2006, 4, e374. [Google Scholar] [CrossRef]
  73. Ogata, M.; Hino, S.; Saito, A.; Morikawa, K.; Kondo, S.; Kanemoto, S.; Murakami, T.; Taniguchi, M.; Tanii, I.; Yoshinaga, K.; et al. Autophagy is activated for cell survival after endoplasmic reticulum stress. Mol. Cell Biol. 2006, 26, 9220–9231. [Google Scholar] [CrossRef] [Green Version]
  74. Li, J.; Ni, M.; Lee, B.; Barron, E.; Hinton, D.R.; Lee, A.S. The unfolded protein response regulator GRP78/BiP is required for endoplasmic reticulum integrity and stress-induced autophagy in mammalian cells. Cell Death Differ. 2008, 15, 1460–1471. [Google Scholar] [CrossRef]
  75. Grosskreutz, J.; Van Den Bosch, L.; Keller, B.U. Calcium dysregulation in amyotrophic lateral sclerosis. Cell Calcium 2010, 47, 165–174. [Google Scholar] [CrossRef]
  76. Zong, W.X.; Lindsten, T.; Ross, A.J.; MacGregor, G.R.; Thompson, C.B. BH3-only proteins that bind pro-survival Bcl-2 family members fail to induce apoptosis in the absence of Bax and Bak. Genes Dev. 2001, 15, 1481–1486. [Google Scholar] [CrossRef] [Green Version]
  77. Pan, Z.; Bhat, M.B.; Nieminen, A.L.; Ma, J. Synergistic movements of Ca2+ and Bax in cells undergoing apoptosis. J. Biol. Chem. 2001, 276, 32257–32263. [Google Scholar] [CrossRef] [Green Version]
  78. Jayaraman, T.; Marks, A.R. T cells deficient in inositol 1,4,5-trisphosphate receptor are resistant to apoptosis. Mol. Cell. Biol. 1997, 17, 3005–3012. [Google Scholar] [CrossRef] [Green Version]
  79. Andjelíc, S.; Khanna, A.; Suthanthiran, M.; Nikolić-Zugić, J. Intracellular Ca2+ elevation and cyclosporin A synergistically induce TGF-beta 1-mediated apoptosis in lymphocytes. J. Immunol. 1997, 158, 2527–2534. [Google Scholar] [CrossRef]
  80. Hata, S.; Sorimachi, H.; Nakagawa, K.; Maeda, T.; Abe, K.; Suzuki, K. Domain II of m-calpain is a Ca2+-dependent cysteine protease. FEBS Lett. 2001, 501, 111–114. [Google Scholar] [CrossRef] [Green Version]
  81. Sorimachi, H.; Ishiura, S.; Suzuki, K. Structure and physiological function of calpains. Biochem. J. 1997, 328 Pt 3, 721–732. [Google Scholar] [CrossRef] [Green Version]
  82. Rao, R.V.; Hermel, E.; Castro-Obregon, S.; del Rio, G.; Ellerby, L.M.; Ellerby, H.M.; Bredesen, D.E. Coupling endoplasmic reticulum stress to the cell death program. Mechanism of caspase activation. J. Biol. Chem. 2001, 276, 33869–33874. [Google Scholar] [CrossRef]
  83. Martinez, J.A.; Zhang, Z.; Svetlov, S.I.; Hayes, R.L.; Wang, K.K.; Larner, S.F. Calpain and caspase processing of caspase-12 contribute to the ER stress-induced cell death pathway in differentiated PC12 cells. Apoptosis 2010, 15, 1480–1493. [Google Scholar] [CrossRef] [PubMed]
  84. Crompton, M. The mitochondrial permeability transition pore and its role in cell death. Biochem. J. 1999, 341 Pt 2, 233–249. [Google Scholar] [CrossRef] [PubMed]
  85. Liu, Z.; Lv, Y.; Zhao, N.; Guan, G.; Wang, J. Protein kinase R-like ER kinase and its role in endoplasmic reticulum stress-decided cell fate. Cell Death Dis. 2015, 6, e1822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Ma, Y.; Brewer, J.W.; Diehl, J.A.; Hendershot, L.M. Two distinct stress signaling pathways converge upon the CHOP promoter during the mammalian unfolded protein response. J. Mol. Biol. 2002, 318, 1351–1365. [Google Scholar] [CrossRef]
  87. Saleh, M.; Mathison, J.C.; Wolinski, M.K.; Bensinger, S.J.; Fitzgerald, P.; Droin, N.; Ulevitch, R.J.; Green, D.R.; Nicholson, D.W. Enhanced bacterial clearance and sepsis resistance in caspase-12-deficient mice. Nature 2006, 440, 1064–1068. [Google Scholar] [CrossRef]
  88. Boya, P.; Cohen, I.; Zamzami, N.; Vieira, H.L.; Kroemer, G. Endoplasmic reticulum stress-induced cell death requires mitochondrial membrane permeabilization. Cell Death Differ. 2002, 9, 465–467. [Google Scholar] [CrossRef]
  89. Wieckowski, M.R.; Giorgi, C.; Lebiedzinska, M.; Duszynski, J.; Pinton, P. Isolation of mitochondria-associated membranes and mitochondria from animal tissues and cells. Nat. Protoc. 2009, 4, 1582–1590. [Google Scholar] [CrossRef]
  90. Mohsin, A.A.; Thompson, J.; Hu, Y.; Hollander, J.; Lesnefsky, E.J.; Chen, Q. Endoplasmic reticulum stress-induced complex I defect: Central role of calcium overload. Arch. Biochem. Biophys. 2020, 683, 108299. [Google Scholar] [CrossRef]
  91. Chami, M.; Oulès, B.; Szabadkai, G.; Tacine, R.; Rizzuto, R.; Paterlini-Bréchot, P. Role of SERCA1 truncated isoform in the proapoptotic calcium transfer from ER to mitochondria during ER stress. Mol. Cell 2008, 32, 641–651. [Google Scholar] [CrossRef] [Green Version]
  92. Zhang, Y.; Ren, J. Thapsigargin triggers cardiac contractile dysfunction via NADPH oxidase-mediated mitochondrial dysfunction: Role of Akt dephosphorylation. Free Radic. Biol. Med. 2011, 51, 2172–2184. [Google Scholar] [CrossRef]
  93. Bravo, R.; Vicencio, J.M.; Parra, V.; Troncoso, R.; Munoz, J.P.; Bui, M.; Quiroga, C.; Rodriguez, A.E.; Verdejo, H.E.; Ferreira, J.; et al. Increased ER-mitochondrial coupling promotes mitochondrial respiration and bioenergetics during early phases of ER stress. J. Cell Sci. 2011, 124, 2143–2152. [Google Scholar] [CrossRef] [Green Version]
  94. Verfaillie, T.; Garg, A.D.; Agostinis, P. Targeting ER stress induced apoptosis and inflammation in cancer. Cancer Lett. 2013, 332, 249–264. [Google Scholar] [CrossRef]
  95. Muñoz, J.P.; Ivanova, S.; Sánchez-Wandelmer, J.; Martínez-Cristóbal, P.; Noguera, E.; Sancho, A.; Díaz-Ramos, A.; Hernández-Alvarez, M.I.; Sebastián, D.; Mauvezin, C.; et al. Mfn2 modulates the UPR and mitochondrial function via repression of PERK. EMBO J. 2013, 32, 2348–2361. [Google Scholar] [CrossRef] [Green Version]
  96. van Vliet, A.R.; Verfaillie, T.; Agostinis, P. New functions of mitochondria associated membranes in cellular signaling. Biochim. Biophys. Acta 2014, 1843, 2253–2262. [Google Scholar] [CrossRef] [Green Version]
  97. Rubio, N.; Coupienne, I.; Di Valentin, E.; Heirman, I.; Grooten, J.; Piette, J.; Agostinis, P. Spatiotemporal autophagic degradation of oxidatively damaged organelles after photodynamic stress is amplified by mitochondrial reactive oxygen species. Autophagy 2012, 8, 1312–1324. [Google Scholar] [CrossRef] [Green Version]
  98. Csordás, G.; Hajnóczky, G. SR/ER-mitochondrial local communication: Calcium and ROS. Biochim. Biophys. Acta 2009, 1787, 1352–1362. [Google Scholar] [CrossRef] [Green Version]
  99. Hamasaki, M.; Furuta, N.; Matsuda, A.; Nezu, A.; Yamamoto, A.; Fujita, N.; Oomori, H.; Noda, T.; Haraguchi, T.; Hiraoka, Y.; et al. Autophagosomes form at ER-mitochondria contact sites. Nature 2013, 495, 389–393. [Google Scholar] [CrossRef]
  100. Meister, S.; Schubert, U.; Neubert, K.; Herrmann, K.; Burger, R.; Gramatzki, M.; Hahn, S.; Schreiber, S.; Wilhelm, S.; Herrmann, M.; et al. Extensive immunoglobulin production sensitizes myeloma cells for proteasome inhibition. Cancer Res. 2007, 67, 1783–1792. [Google Scholar] [CrossRef] [Green Version]
  101. Davenport, E.L.; Morgan, G.J.; Davies, F.E. Untangling the unfolded protein response. Cell Cycle 2008, 7, 865–869. [Google Scholar] [CrossRef] [Green Version]
  102. Wang, G.; Yang, Z.Q.; Zhang, K. Endoplasmic reticulum stress response in cancer: Molecular mechanism and therapeutic potential. Am. J. Transl. Res. 2010, 2, 65–74. [Google Scholar] [PubMed]
  103. Fernandez, P.; Tabbara, S.; Jacobs, L.; Manning, F.; Tsangaris, T.; Schwartz, A.; Kennedy, K.; Patierno, S. Overexpression of the glucoseregulated stress gene GRP78 in malignant but not benign human breast lesions. Breast Cancer Res. Treat. 2000, 59, 15–26. [Google Scholar] [CrossRef] [PubMed]
  104. Ye, J.; Koumenis, C. ATF4, an ER stress and hypoxia-inducible transcription factor and its potential role in hypoxia tolerance and tumorigenesis. Curr. Mol. Med. 2009, 9, 411–416. [Google Scholar] [CrossRef] [PubMed]
  105. Romero-Ramirez, L.; Cao, H.; Nelson, D.; Hammond, E.; Lee, A.H.; Yoshida, H.; Mori, K.; Glimcher, L.H.; Denko, N.C.; Giaccia, A.J.; et al. XBP1 is essential for survival under hypoxic conditions and is required for tumor growth. Cancer Res. 2004, 64, 5943–5947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Qu, L.; Huang, S.; Baltzis, D.; Rivas-Estilla, A.M.; Pluquet, O.; Hatzoglou, M.; Koumenis, C.; Taya, Y.; Yoshimura, A.; Koromilas, A.E. Endoplasmic reticulum stress induces p53 cytoplasmic localization and prevents p53-dependent apoptosis by a pathway involving glycogen synthase kinase-3beta. Genes Dev. 2004, 18, 261–277. [Google Scholar] [CrossRef] [Green Version]
  107. Kitamura, M. Biphasic, bidirectional regulation of NF-kappaB by endoplasmic reticulum stress. Antioxid. Redox Signal. 2009, 11, 2353–2364. [Google Scholar] [CrossRef]
  108. Jolly, C.; Morimoto, R.I. Role of the heat shock response and molecular chaperones in oncogenesis and cell death. J. Natl. Cancer Inst. 2000, 92, 1564–1572. [Google Scholar] [CrossRef] [Green Version]
  109. So, A.; Hadaschik, B.; Sowery, R.; Gleave, M. The role of stress proteins in prostate cancer. Curr. Genom. 2007, 8, 252–261. [Google Scholar] [CrossRef]
  110. Dong, D.; Ni, M.; Li, J.; Xiong, S.; Ye, W.; Virrey, J.J.; Mao, C.; Ye, R.; Wang, M.; Pen, L.; et al. Critical role of the stress chaperone GRP78/BiP in tumor proliferation, survival, and tumor angiogenesis in transgene-induced mammary tumor development. Cancer Res. 2008, 68, 498–505. [Google Scholar] [CrossRef] [Green Version]
  111. Li, J.; Lee, A.S. Stress induction of GRP78/BiP and its role in cancer. Curr. Mol. Med. 2006, 6, 45–54. [Google Scholar] [CrossRef]
  112. 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]
  113. Saito, S.; Furuno, A.; Sakurai, J.; Sakamoto, A.; Park, H.R.; Shin-Ya, K.; Tsuruo, T.; Tomida, A. Chemical genomics identifies the unfolded protein response as a target for selective cancer cell killing during glucose deprivation. Cancer Res. 2009, 69, 4225–4234. [Google Scholar] [CrossRef] [Green Version]
  114. Al-Rawashdeh, F.Y.; Scriven, P.; Cameron, I.C.; Vergani, P.V.; Wyld, L. Unfolded protein response activation contributes to chemoresistance in hepatocellular carcinoma. Eur. J. Gastroenterol. Hepatol. 2010, 22, 1099–1105. [Google Scholar] [CrossRef]
  115. Pyrko, P.; Schönthal, A.H.; Hofman, F.M.; Chen, T.C.; Lee, A.S. The unfolded protein response regulator GRP78/BiP as a novel target for increasing chemosensitivity in malignant gliomas. Cancer Res. 2007, 67, 9809–9816. [Google Scholar] [CrossRef] [Green Version]
  116. Daneshmand, S.; Quek, M.L.; Lin, E.; Lee, C.; Cote, R.J.; Hawes, D.; Cai, J.; Groshen, S.; Lieskovsky, G.; Skinner, D.G.; et al. Glucose-regulated protein GRP78 is up-regulated in prostate cancer and correlates with recurrence and survival. Hum. Pathol. 2007, 38, 1547–1552. [Google Scholar] [CrossRef]
  117. Zhang, J.; Jiang, Y.; Jia, Z.; Li, Q.; Gong, W.; Wang, L.; Wei, D.; Yao, J.; Fang, S.; Xie, K. Association of elevated GRP78 expression with increased lymph node metastasis and poor prognosis in patients with gastric cancer. Clin. Exp. Metastasis 2006, 23, 401–410. [Google Scholar] [CrossRef]
  118. Uramoto, H.; Sugio, K.; Oyama, T.; Nakata, S.; Ono, K.; Yoshimastu, T.; Morita, M.; Yasumoto, K. Expression of endoplasmic reticulum molecular chaperone Grp78 in human lung cancer and its clinical significance. Lung Cancer 2005, 49, 55–62. [Google Scholar] [CrossRef]
  119. Cook, K.L.; Shajahan, A.N.; Warri, A.; Jin, L.; Hilakivi-Clarke, L.A.; Clarke, R. Glucoseregulated protein 78 controls cross-talk between apoptosis and autophagy to determine antiestrogen responsiveness. Cancer Res. 2012, 72, 3337–3349. [Google Scholar] [CrossRef] [Green Version]
  120. Lee, E.; Nichols, P.; Spicer, D.; Groshen, S.; Yu, M.C.; Lee, A.S. GRP78 as a novel predictor of responsiveness to chemotherapy in breast cancer. Cancer Res. 2006, 66, 7849–7853. [Google Scholar] [CrossRef] [Green Version]
  121. Scriven, P.; Coulson, S.; Haines, R.; Balasubramanian, S.; Cross, S.; Wyld, L. Activation and clinical significance of the unfolded protein response in breast cancer. Br. J. Cancer 2009, 101, 1692–1698. [Google Scholar] [CrossRef] [Green Version]
  122. Dong, D.; Ko, B.; Baumeister, P.; Swenson, S.; Costa, F.; Markland, F.; Stiles, C.; Patterson, J.B.; Bates, S.E.; Lee, A.S. Vascular targeting and antiangiogenesis agents induce drug resistance effector GRP78 within the tumor microenvironment. Cancer Res. 2005, 65, 5785–5791. [Google Scholar] [CrossRef] [PubMed]
  123. Reddy, R.K.; Mao, C.; Baumeister, P.; Austin, R.C.; Kaufman, R.J.; Lee, A.S. Endoplasmic reticulum chaperone protein GRP78 protects cells from apoptosis induced by topoisomerase inhibitors: Role of ATP binding site in suppression of caspase-7 activation. J. Biol. Chem. 2003, 278, 20915–20924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Lin, J.A.; Fang, S.U.; Su, C.L.; Hsiao, C.J.; Chang, C.C.; Lin, Y.F.; Cheng, C.W. Silencing glucose-regulated protein 78 induced renal cell carcinoma cell line G1 cell-cycle arrest and resistance to conventional chemotherapy. Urol. Oncol. 2014, 32, 29.e1–29.e11. [Google Scholar] [CrossRef] [PubMed]
  125. Chen, X.; Zhang, D.; Dennert, G.; Hung, G.; Lee, A.S. Eradication of murine mammary adenocarcinoma through HSVtk expression directed by the glucose-starvation inducible grp78 promoter. Breast Cancer Res. Treat. 2000, 59, 81–90. [Google Scholar] [CrossRef] [PubMed]
  126. Dong, D.; Dubeau, L.; Bading, J.; Nguyen, K.; Luna, M.; Yu, H.; Gazit-Bornstein, G.; Gordon, E.M.; Gomer, C.; Hall, F.L.; et al. Spontaneous and controllable activation of suicide gene expression driven by the stress-inducible grp78 promoter resulting in eradication of sizable human tumors. Hum. Gene Ther. 2004, 15, 553–561. [Google Scholar] [CrossRef] [PubMed]
  127. Martinotti, S.; Ranzato, E.; Burlando, B. (-)- Epigallocatechin-3-gallate induces GRP78 accumulation in the ER and shifts mesothelioma constitutive UPR into proapoptotic ER stress. J. Cell. Physiol. 2018, 233, 7082–7090. [Google Scholar] [CrossRef]
  128. Park, H.R.; Tomida, A.; Sato, S.; Tsukumo, Y.; Yun, J.; Yamori, T.; Hayakawa, Y.; Tsuruo, T.; Shin-ya, K. Effect on tumor cells of blocking survival response to glucose deprivation. J. Natl. Cancer Instig. 2004, 96, 1300–1310. [Google Scholar] [CrossRef] [Green Version]
  129. Ni, M.; Zhang, Y.; Lee, A.S. Beyond the endoplasmic reticulum: Atypical GRP78 in cell viability, signalling and therapeutic targeting. Biochem. J. 2011, 434, 181–188. [Google Scholar] [CrossRef]
  130. Ni, M.; Zhou, H.; Wey, S.; Baumeister, P.; Lee, A.S. Regulation of PERK signaling and leukemic cell survival by a novel cytosolic isoform of the UPR regulator GRP78/BiP. PLoS ONE 2009, 4, e6868. [Google Scholar] [CrossRef] [Green Version]
  131. Sun, F.C.; Wei, S.; Li, C.W.; Chang, Y.S.; Chao, C.C.; Lai, Y.K. Localization of GRP78 to mitochondria under the unfolded protein response. Biochem. J. 2006, 396, 31–39. [Google Scholar] [CrossRef] [Green Version]
  132. Zhang, Y.; Liu, R.; Ni, M.; Gill, P.; Lee, A.S. Cell surface relocalization of the endoplasmic reticulum chaperone and unfolded protein response regulator GRP78/BiP. J. Biol. Chem. 2010, 285, 15065–15075. [Google Scholar] [CrossRef]
  133. Rauschert, N.; Brandlein, S.; Holzinger, E.; Hensel, F.; Muller-Hermelink, H.K.; Vollmers, H.P. A new tumor-specific variant of GRP78 as target for antibody-based therapy. Lab. Investig. 2008, 88, 375–386. [Google Scholar] [CrossRef] [Green Version]
  134. Brandlein, S.; Rauschert, N.; Rasche, L.; Dreykluft, A.; Hensel, F.; Conzelmann, E.; Muller-Hermelink, H.K.; Vollmers, H.P. The human IgM antibody SAM-6 induces tumor-specific apoptosis with oxidized low-density lipoprotein. Mol. Cancer Ther. 2007, 6, 326–333. [Google Scholar] [CrossRef] [Green Version]
  135. Rasche, L.; Duell, J.; Morgner, C.; Chatterjee, M.; Hensel, F.; Rosenwald, A.; Einsele, H.M.; Topp, M.S.; Brandlein, S. The natural human IgMantibody PAT-SM6 induces apoptosis in primary human multiple myeloma cells by targeting heat shock protein GRP78. PLoS ONE 2013, 8, e63414. [Google Scholar] [CrossRef]
  136. Hensel, F.; Eckstein, M.; Rosenwald, A.; Brändlein, S. Early development of PAT-SM6 for the treatment of melanoma. Melanoma Res. 2013, 23, 264–275. [Google Scholar] [CrossRef]
  137. Ameri, K.; Lewis, C.E.; Raida, M.; Sowter, H.; Hai, T.; Harris, A.L. Anoxic induction of ATF-4 through HIF-1-independent pathways of protein stabilization in human cancer cells. Blood 2004, 103, 1876–1882. [Google Scholar] [CrossRef] [Green Version]
  138. Tanabe, M.; Izumi, H.; Ise, T.; Higuchi, S.; Yamori, T.; Yasumoto, K.; Kohno, K. Activating transcription factor 4 increases the cisplatin resistance of human cancer cell lines. Cancer Res. 2003, 63, 8592–8595. [Google Scholar]
  139. Igarashi, T.; Izumi, H.; Uchiumi, T.; Nishio, K.; Arao, T.; Tanabe, M.; Uramoto, H.; Sugio, K.; Yasumoto, K.; Sasaguri, Y.; et al. Clock and ATF4 transcription system regulates drug resistance in human cancer cell lines. Oncogene 2007, 26, 4749–4760. [Google Scholar] [CrossRef] [Green Version]
  140. Nagelkerke, A.; Bussink, J.; van der Kogel, A.J.; Sweep, F.C.; Span, P.N. The PERK/ATF4/LAMP3-arm of the unfolded protein response affects radioresistance by interfering with the DNA damage response. Radiother. Oncol. 2013, 108, 415–421. [Google Scholar] [CrossRef]
  141. Zhang, B.; Wang, Y.; Pang, X.; Su, Y.; Ai, G.; Wang, T. ER stress induced by ionising radiation in IEC-6 cells. Int. J. Radiat. Biol. 2010, 86, 429–435. [Google Scholar] [CrossRef]
  142. Atkins, C.; Liu, Q.; Minthorn, E.; Zhang, S.Y.; Figueroa, D.J.; Moss, K.; Stanley, T.B.; Sanders, B.; Goetz, A.; Gaul, N.; et al. Characterization of a novel PERK kinase inhibitor with antitumor and antiangiogenic activity. Cancer Res. 2013, 73, 1993–2002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Axten, J.M.; Medina, J.R.; Feng, Y.; Shu, A.; Romeril, S.P.; Grant, S.W.; Li, W.H.; Heerding, D.A.; Minthorn, E.; Mencken, T.; et al. Discovery of 7-methyl-5-(1-{[3-(trifluoromethyl)phenyl]acetyl}-2,3-dihydro-1H-indol-5-yl)-7H-pyrrolo[2,3-d]pyrimidin-4-amine (GSK2606414), a potent and selective first-in-class inhibitor of protein kinase R (PKR)-like endoplasmic reticulum kinase (PERK). J. Med. Chem. 2012, 55, 7193–7207. [Google Scholar] [CrossRef] [PubMed]
  144. Gomez, B.P.; Riggins, R.B.; Shajahan, A.N.; Klimach, U.; Wang, A.; Crawford, A.C.; Zhu, Y.; Zwart, A.; Wang, M.; Clarke, R. Human X-box binding protein-1 confers both estrogen independence and antiestrogen resistance in breast cancer cell lines. FASEB J. 2007, 21, 4013–4027. [Google Scholar] [CrossRef] [PubMed]
  145. Davies, M.P.; Barraclough, D.L.; Stewart, C.; Joyce, K.A.; Eccles, R.M.; Barraclough, R.; Rudland, P.S.; Sibson, D.R. Expression and splicing of the unfolded protein response gene XBP-1 are significantly associated with clinical outcome of endocrine-treated breast cancer. Int. J. Cancer 2008, 123, 85–88. [Google Scholar] [CrossRef] [PubMed]
  146. Wang, Q.; Mora-Jensen, H.; Weniger, M.A.; Perez-Galan, P.; Wolford, C.; Hai, T.; Ron, D.; Chen, W.; Trenkle, W.; Wiestner, A.; et al. ERAD inhibitors integrate ER stress with an epigenetic mechanism to activate BH3-only protein NOXA in cancer cells. Proc. Natl. Acad. Sci. USA 2009, 106, 2200–2205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Kim, K.W.; Moretti, L.; Mitchell, L.R.; Jung, D.K.; Lu, B. Endoplasmic reticulum stress mediates radiation-induced autophagy by perk-eIF2alpha in caspase-3/7-deficient cells. Oncogene 2010, 29, 3241–3251. [Google Scholar] [CrossRef] [Green Version]
  148. Yamada, M.; Tomida, A.; Yun, J.; Cai, B.; Yoshikawa, Y.; Taketani, Y.; Tsuruo, T. Cellular sensitization to cisplatin and carboplatin with decreased removal of platinum-DNA adduct by glucose-regulated stress. Cancer Chemother. Pharmacol. 1999, 44, 59–64. [Google Scholar] [CrossRef]
  149. Hughes, C.S.; Shen, J.W.; Subjeck, J.R. Resistance to etoposide induced by three glucose-regulated stresses in Chinese hamster ovary cells. Cancer Res. 1989, 49, 4452–4454. [Google Scholar]
  150. Vichi, P.J.; Tritton, T.R. Protection from adriamycin cytotoxicity in L1210 cells by brefeldin A. Cancer Res. 1993, 53, 5237–5243. [Google Scholar]
  151. Hsu, J.L.; Chiang, P.C.; Guh, J.H. Tunicamycin induces resistance to camptothecin and etoposide in human hepatocellular carcinoma cells: Role of cell-cycle arrest and GRP78. Naunyn Schmiedebergs Arch Pharmacol. 2009, 380, 373–382. [Google Scholar] [CrossRef]
  152. Ledoux, S.; Yang, R.; Friedlander, G.; Laouari, D. Glucose depletion enhances P-glycoprotein expression in hepatoma cells: Role of endoplasmic reticulum stressresponse. Cancer Res. 2003, 63, 7284–7290. [Google Scholar]
  153. Ding, W.X.; Ni, H.M.; Gao, W.; Hou, Y.F.; Melan, M.A.; Chen, X.; Stolz, D.B.; Shao, Z.M.; Yin, X.M. Differential effects of endoplasmic reticulum stress-induced autophagy on cell survival. J. Biol. Chem. 2007, 282, 4702–4710. [Google Scholar] [CrossRef] [Green Version]
  154. Avivar-Valderas, A.; Salas, E.; Bobrovnikova-Marjon, E.; Diehl, J.A.; Nagi, C.; Debnath, J.; Aguirre-Ghiso, J.A. PERK integrates autophagy and oxidative stress responses to promote survival during extracellular matrix detachment. Mol. Cell. Biol. 2011, 31, 3616–3629. [Google Scholar] [CrossRef] [Green Version]
  155. Rouschop, K.M.; van den Beucken, T.; Dubois, L.; Niessen, H.; Bussink, J.; Savelkouls, K.; Keulers, T.; Mujcic, H.; Landuyt, W.; Voncken, J.W.; et al. The unfolded protein response protects human tumor cells during hypoxia through regulation of the autophagy genes MAP1LC3B and ATG5. J. Clin. Investig. 2010, 120, 127–141. [Google Scholar] [CrossRef]
  156. Kouroku, Y.; Fujita, E.; Tanida, I.; Ueno, T.; Isoai, A.; Kumagai, H.; Ogawa, S.; Kaufman, R.J.; Kominami, E.; Momoi, T. ER stress (PERK/eIF2alpha phosphorylation) mediates the polyglutamine-induced LC3 conversion, an essential step for autophagy formation. Cell Death Differ. 2007, 14, 230–239. [Google Scholar] [CrossRef]
  157. Chen, R.; Dai, R.Y.; Duan, C.Y.; Liu, Y.P.; Chen, S.K.; Yan, D.M.; Chen, C.N.; Wei, M.; Li, H. Unfolded protein response suppresses cisplatin-induced apoptosis via autophagy regulation in human hepatocellular carcinoma cells. Folia Biol. 2011, 57, 87–95. [Google Scholar]
  158. Chaachouay, H.; Ohneseit, P.; Toulany, M.; Kehlbach, R.; Multhoff, G.; Rodemann, H.P. Autophagy contributes to resistance of tumor cells to ionizing radiation. Radiother. Oncol. 2011, 99, 287–292. [Google Scholar] [CrossRef]
  159. Nagelkerke, A.; Sieuwerts, A.M.; Bussink, J.; Sweep, F.C.; Look, M.P.; Foekens, J.A.; Martens, J.W.; Span, P.N. LAMP3 is involved in tamoxifen resistance in breast cancer cells through the modulation of autophagy. Endocr. Relat. Cancer 2014, 21, 101–112. [Google Scholar] [CrossRef] [Green Version]
  160. Qadir, M.A.; Kwok, B.; Dragowska, W.H.; To, K.H.; Le, D.; Bally, M.B.; Gorski, S.M. Macroautophagy inhibition sensitizes tamoxifen-resistant breast cancer cells and enhances mitochondrial depolarization. Breast Cancer Res. Treat. 2008, 112, 389–403. [Google Scholar] [CrossRef]
  161. Milani, M.; Rzymski, T.; Mellor, H.R.; Pike, L.; Bottini, A.; Generali, D.; Harris, A.L. The role of ATF4 stabilization and autophagy in resistance of breast cancer cells treated with Bortezomib. Cancer Res. 2009, 69, 4415–4423. [Google Scholar] [CrossRef] [Green Version]
  162. Sakaki, K.; Wu, J.; Kaufman, R.J. Protein kinase Ctheta is required for autophagy in response to stress in the endoplasmic reticulum. J. Biol. Chem. 2008, 283, 15370–15380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Fewell, S.W.; Travers, K.J.; Weissman, J.S.; Brodsky, J.L. The action of molecular chaperones in the early secretory pathway. Annu. Rev. Genet. 2001, 35, 149–191. [Google Scholar] [CrossRef] [PubMed]
  164. Ruggiano, A.; Foresti, O.; Carvalho, P. Quality control: ERassociated degradation: Protein quality control and beyond. J. Cell Biol. 2014, 204, 869–879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Nishitoh, H. CHOP is amultifunctional transcription factor in the ER stress response. J. Biochem. 2012, 151, 217–219. [Google Scholar] [CrossRef] [Green Version]
  166. Kim, I.; Xu, W.; Reed, J.C. Cell death and endoplasmic reticulum stress: Disease relevance and therapeutic opportunities. Nat. Rev. Drug Discov. 2008, 7, 1013–1030. [Google Scholar] [CrossRef]
  167. Bonsignore, G.; Patrone, M.; Grosso, F.; Martinotti, S.; Ranzato, E. Cancer Therapy Challenge: It Is Time to Look in the “St. Patrick’s Well” of the Nature. Int. J. Mol. Sci. 2021, 22, 10380. [Google Scholar] [CrossRef]
  168. Lee, J.H.; Li, Y.C.; Ip, S.W.; Hsu, S.C.; Chang, N.W.; Tang, N.Y.; Yu, C.S.; Chou, S.T.; Lin, S.S.; Lino, C.C.; et al. The role of Ca2+ in baicalein-induced apoptosis in human breast MDA-MB-231 cancer cells through mitochondria- and caspase-3-dependent pathway. Anticancer. Res. 2008, 28, 1701–1711. [Google Scholar]
  169. Wang, Z.S.; Lu, F.E.; Xu, L.J.; Dong, H. Berberine reduces endoplasmic reticulumstress and improves insulin signal transduction inHepG2 cells. Acta Pharmacol. Sin. 2010, 31, 578–584. [Google Scholar] [CrossRef] [Green Version]
  170. La, X.; Zhang, L.; Li, Z.; Yang, P.; Wang, Y. Berberine-induced autophagic cell death by elevating GRP78 levels in cancer cells. Oncotarget 2017, 8, 20909–20924. [Google Scholar] [CrossRef] [Green Version]
  171. Zhu, H.; Yang, W.; He, L.J.; Ding, W.J.; Zheng, L.; Liao, S.D.; Huang, P.; Lu, W.; He, Q.J.; Yang, B. Upregulating Noxa by ER stress, celastrol exerts synergistic anti-cancer activity in combination with ABT-737 in human hepatocellular carcinoma cells. PLoS ONE 2012, 7, e52333. [Google Scholar] [CrossRef]
  172. Ng, A.P.P.; Chng, W.J.; Khan, M. Curcumin sensitizes acute promyelocytic leukemia cells to unfolded protein responseinduced apoptosis by blocking the loss of misfolded N-CoR protein. Mol. Cancer Res. 2011, 9, 878–888. [Google Scholar] [CrossRef] [Green Version]
  173. Huang, K.H.; Weng, T.I.; Huang, H.Y.; Huang, K.D.; Lin, W.C.; Chen, S.C.; Liu, S.H. Honokiol attenuates torsion/detorsion-induced testicular injury in rat testis by way of suppressing endoplasmic reticulum stress-related apoptosis. Urology 2012, 79, 967.e5–967.e11. [Google Scholar] [CrossRef]
  174. Chen, Y.J.; Wu, C.L.; Liu, J.F.; Fong, Y.C.; Hsu, S.F.; Li, T.M.; Su, Y.C.; Liu, S.H.; Tang, C.H. Honokiol induces cell apoptosis in human chondrosarcoma cells through mitochondrial dysfunction and endoplasmic reticulum stress. Cancer Lett. 2010, 291, 20–30. [Google Scholar] [CrossRef]
  175. Wang, Z.S.; Xiong, F.; Xie, X.H.; Chen, D.; Pan, J.H.; Cheng, L. Astragaloside IV attenuates proteinuria in streptozotocin induced diabetic nephropathy via the inhibition of endoplasmic reticulumstress. BMC Nephrol. 2015, 16, 44. [Google Scholar] [CrossRef] [Green Version]
  176. Chen, Y.; Gui, D.; Chen, J.; He, D.; Luo, Y.; Wang, N. Down-regulation of PERK-ATF4-CHOP pathway by Astragaloside IV is associated with the inhibition of endoplasmic reticulum stress-induced podocyte apoptosis in diabetic rats. Cell. Physiol. Biochem. 2014, 33, 1975–1987. [Google Scholar] [CrossRef]
  177. Lai, S.T.; Wang, Y.; Peng, F. Astragaloside IV sensitizes non-small cell lung cancer cells to cisplatin by suppressing endoplasmic reticulum stress and autophagy. J. Thorac. Dis. 2020, 12, 3715–3724. [Google Scholar] [CrossRef]
  178. Skrobuk, P.; von Kraemer, S.; Semenova, M.M.; Zitting, A.; Koistinen, H.A. Acute exposure to resveratrol inhibits AMPK activity in human skeletal muscle cells. Diabetologia 2012, 55, 3051–3060. [Google Scholar] [CrossRef]
  179. Conza, D.; Mirra, P.; Calì, G.; Insabato, L.; Fiory, F.; Beguinot, F.; Ulianich, L. Metformin Dysregulates the Unfolded Protein Response and the WNT/β-Catenin Pathway in Endometrial Cancer Cells through an AMPK-Independent Mechanism. Cells 2021, 10, 1067. [Google Scholar] [CrossRef]
  180. Ye, J.; Qi, L.; Chen, K.; Li, R.; Song, S.; Zhou, C.; Zhai, W. Metformin induces TPC-1 cell apoptosis through endoplasmic reticulum stress-associated pathways in vitro and in vivo. Int. J. Oncol. 2019, 55, 331–339. [Google Scholar] [CrossRef]
  181. Jagannathan, S.; Abdel-Malek, M.A.; Malek, E.; Vad, N.; Latif, T.; Anderson, K.C.; Driscoll, J.J. Pharmacologic screens reveal metformin that suppresses GRP78-dependent autophagy to enhance the anti-myeloma effect of bortezomib. Leukemia 2015, 29, 2184–2191. [Google Scholar] [CrossRef]
  182. Chen, T.C.; Wang, W.; Golden, E.B.; Thomas, S.; Sivakumar, W.; Hofman, F.M.; Louie, S.G.; Schönthal, A.H. Green tea epigallocatechin gallate enhances therapeutic efficacy of temozolomide in orthotopic mouse glioblastoma models. Cancer Lett. 2011, 302, 100–108. [Google Scholar] [CrossRef] [PubMed]
  183. Chen, B.; Liu, G.; Zou, P.; Li, X.; Hao, Q.; Jiang, B.; Yang, X.; Hu, Z. Epigallocatechin-3-gallate protects against cisplatin-induced nephrotoxicity by inhibiting endoplasmic reticulum stress-induced apoptosis. Exp. Biol. Med. 2015, 240, 1513–1519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Volta, V.; Ranzato, E.; Martinotti, S.; Gallo, S.; Russo, M.V.; Mutti, L.; Biffo, S.; Burlando, B. Preclinical demonstration of synergistic Active Nutrients/Drug (AND) combination as a potential treatment for malignant pleural mesothelioma. PLoS ONE 2013, 8, e58051. [Google Scholar] [CrossRef] [PubMed]
  185. Sasaya, H.; Utsumi, T.; Shimoke, K.; Nakayama, H.; Matsumura, Y.; Fukunaga, K.; Ikeuchi, T. Nicotine suppresses tunicamycin-induced, but not thapsigargin-induced, expression of GRP78 during ER stress-mediated apoptosis in PC12 cells. J. Biochem. 2008, 144, 251–257. [Google Scholar] [CrossRef] [PubMed]
  186. Huang, T.T.; Lai, H.C.; Chen, Y.B.; Chen, L.G.; Wu, Y.H.; Ko, Y.F.; Lu, C.C.; Chang, C.J.; Wu, C.Y.; Martel, J.; et al. cis-Resveratrol produces anti-inflammatory effects by inhibiting canonical and non-canonical inflammasomes in macrophages. Innate Immun. 2014, 20, 735–750. [Google Scholar] [CrossRef] [Green Version]
  187. Wang, F.M.; Galson, D.L.; Roodman, G.D.; Ouyang, H. Resveratrol triggers the pro-apoptotic endoplasmic reticulum stress response and represses pro-survival XBP1 signaling in human multiple myeloma cells. Exp. Hematol. 2011, 39, 999–1006. [Google Scholar] [CrossRef] [Green Version]
  188. Yang, S.-T.; Huang, A.-C.; Tang, N.-Y.; Liu, H.-C.; Liao, C.-L.; Ji, B.-C.; Chou, Y.-C.; Yang, M.-D.; Lu, H.-F.; Chung, J.-G. Bisdemethoxycurcumin-induced S phase arrest through the inhibition of cyclin A and E and induction of apoptosis via endoplasmic reticulum stress and mitochondria-dependent pathways in human lungcancer NCI H460 cells. Environ. Toxicol. 2015, 31, 1899–1908. [Google Scholar] [CrossRef]
  189. Girola, N.; Figueiredo, C.R.; Farias, C.F.; Azevedo, R.A.; Ferreira, A.K.; Teixeira, S.F.; Capello, T.M.; Martins, E.G.; Matsuo, A.L.; Travassos, L.R.; et al. Camphene isolated from essential oil of Piper cernuum (Piperaceae) induces intrinsic apoptosis in melanoma cells and displays antitumor activity in vivo. Biochem. Biophys. Res. Commun. 2015, 467, 928–934. [Google Scholar] [CrossRef]
  190. Sun, D.P.; Li, X.X.; Liu, X.L.; Zhao, D.; Qiu, F.Q.; Li, Y.; Ma, P. Gypenosides induce apoptosis by ca2+ overload mediated by endoplasmic-reticulum and store-operated ca2+ channels in human hepatoma cells. Cancer Biother. Radiopharm. 2013, 28, 320–326. [Google Scholar] [CrossRef] [Green Version]
  191. Li, Y.; Schwabe, R.F.; DeVries-Seimon, T.; Yao, P.M.; Gerbod-Giannone, M.C.; Tall, A.R.; Davis, R.J.; Flavell, R.; Brenner, D.A.; Tabas, I. Free cholesterol-loaded macrophages are an abundant source of tumor necrosis factor-alpha and interleukin-6: Model of NF-kappaB- and map kinase-dependent inflammation in advanced atherosclerosis. J. Biol. Chem. 2005, 280, 21763–21772. [Google Scholar] [CrossRef] [Green Version]
  192. Cho, J.A.; Park, E. Curcumin utilizes the anti-inflammatory response pathway to protect the intestine against bacterial invasion. Nutr. Res. Pract. 2015, 9, 117–122. [Google Scholar] [CrossRef] [Green Version]
  193. Pereira, D.M.; Correia-da-Silva, G.; Valentão, P.; Teixeira, N.; Andrade, P.B. Anti-inflammatory effect of unsaturated fatty acids and Ergosta-7,22-dien-3-ol from Marthasterias glacialis: Prevention of CHOP-mediated ER-stress and NF-κB activation. PLoS ONE 2014, 9, e88341. [Google Scholar] [CrossRef]
  194. Wu, J.; Xu, X.; Li, Y.; Kou, J.; Huang, F.; Liu, B.; Liu, K. Quercetin, luteolin and epigallocatechin gallate alleviate TXNIP and NLRP3-mediated inflammation and apoptosis with regulation of AMPK in endothelial cells. Eur. J. Pharmacol. 2015, 745, 59–68. [Google Scholar] [CrossRef]
  195. Li, Y.; Yang, J.; Chen, M.-H.; Wang, Q.; Qin, M.-J.; Zhang, T.; Chen, X.-Q.; Liu, B.-L.; Wen, X.-D. Ilexgenin A inhibits endoplasmic reticulum stress and ameliorates endothelial dysfunction via suppression of TXNIP/NLRP3 inflammasome activation in an AMPK dependent manner. Pharmacol. Res. 2015, 99, 101–115. [Google Scholar] [CrossRef]
  196. Song, J.; Li, J.; Hou, F.; Wang, X.; Liu, B. Mangiferin inhibits endoplasmic reticulum stress-associated thioredoxininteracting protein/NLRP3 inflammasome activation with regulation of AMPK in endothelial cells. Metabolism 2015, 64, 428–437. [Google Scholar] [CrossRef]
  197. Zhao, Y.; Li, Q.; Zhao, W.; Li, J.; Sun, Y.; Liu, K.; Liu, B.; Zhang, N. Astragaloside IV and cycloastragenol are equally effective in inhibition of endoplasmic reticulum stress-associated TXNIP/NLRP3 inflammasome activation in the endothelium. J. Ethnopharmacol. 2015, 169, 210–218. [Google Scholar] [CrossRef]
  198. Neuditschko, B.; Legin, A.A.; Baier, D.; Schintlmeister, A.; Reipert, S.; Wagner, M.; Keppler, B.K.; Berger, W.; Meier-Menches, S.M.; Gerner, C. Interaction with Ribosomal Proteins Accompanies Stress Induction of the Anticancer Metallodrug BOLD-100/KP1339 in the Endoplasmic Reticulum. Angew. Chem. Int. Ed. 2021, 60, 5063–5068. [Google Scholar] [CrossRef]
  199. Jungwirth, U.; Kowol, C.R.; Keppler, B.K.; Hartinger, C.G.; Berger, W.; Heffeter, P. Anticancer activity of metal complexes: Involvement of redox processes. Antioxid. Redox Signal. 2011, 15, 1085–1127. [Google Scholar] [CrossRef] [Green Version]
  200. Clarke, M.J.; Bitler, S.; Rennert, D.; Buchbinder, M.; Kelman, A.D. Reduction and subsequent binding of ruthenium ions catalyzed by subcellular components. J. Inorg. Biochem. 1980, 12, 79–87. [Google Scholar] [CrossRef]
  201. Smith, C.A.; Sutherland-Smith, A.J.; Keppler, B.K.; Kratz, F.; Baker, E.N. Binding of ruthenium(III) anti-tumor drugs to human lactoferrin probed by high resolution X-ray crystallographic structure analyses. JBIC J. Biol. Inorg. Chem. 1996, 1, 424–431. [Google Scholar] [CrossRef]
  202. Ranzato, E.; Bonsignore, G.; Martinotti, S. ER Stress Response and Induction of Apoptosis in Malignant Pleural Mesothelioma: The Achilles Heel Targeted by the Anticancer Ruthenium Drug BOLD-100. Cancers 2022, 14, 4126. [Google Scholar] [CrossRef] [PubMed]
  203. Balachandran, C.; Yokoi, K.; Naito, K.; Haribabu, J.; Tamura, Y.; Umezawa, M.; Tsuchiya, K.; Yoshihara, T.; Tobita, S.; Aoki, S. Cyclometalated Iridium(III) Complex-Cationic Peptide Hybrids Trigger Paraptosis in Cancer Cells via an Intracellular Ca. Molecules 2021, 26, 7028. [Google Scholar] [CrossRef] [PubMed]
  204. Haribabu, J.; Tamura, Y.; Yokoi, K.; Balachandran, C.; Umezawa, M.; Tsuchiya, K.; Yamada, Y.; Karvembu, R.; Aoki, S. Synthesis and anticancer properties of bis and mono(cationic peptide) hybrids of cyclometalated Ir(III) complexes: Effect of the number of peptide units on anticancer activity. Eur. J. Inorg. Chem. 2021, 2021, 1796–1814. [Google Scholar] [CrossRef]
Table 1. Natural products affecting ER stress and UPR pathways.
Table 1. Natural products affecting ER stress and UPR pathways.
Natural ProductCommon SourcesMechanisms of ActionChemical Structure Depiction
BaicaleinScutellaria baicalensis and Scutellaria laterifloraActivation of ER stress inducing apoptosis and autophagyIjms 24 01566 i001
BerberineRhizoma coptidisEffective against several diseases, like dyslipidaemia, hyperglycaemia, and obesity.Ijms 24 01566 i002
CelastrolTripterygium wilfordii HookIncreasing of BH3 mimetic drug ABT-737 effect.Ijms 24 01566 i003
CurcuminCurcuma speciesStimulation the UPR-induced apoptosis.Ijms 24 01566 i004
HonokiolMagnolia officinalisDecreasing of p-eIF2α and CHOP expression.Ijms 24 01566 i005
Astragaloside IVAstragalus membranaceusProtective activity in diabetic nephropathy because of ER stress inhibition.Ijms 24 01566 i006
ResveratrolAlmost 70 plant speciesReduction of fat and body weight trough the regulation of lipid and glucose metabolisms.Ijms 24 01566 i007
EGCGCamellia sinensisInhibitor of GRP78 [167].Ijms 24 01566 i008
NicotineNicotiana tabacumDownregulation of apoptosis induced by tunicamycin-mediated ER stress in cancer cells (causing the reduction of expression for ATF6, GRP78, and IRE1-XBP).Ijms 24 01566 i009
CamphenePiper cernuumturning on of apoptosis in melanoma cells inducing ER stress.Ijms 24 01566 i010
GypenosidesGynostemma pentaphyllumAbility to activate apoptosis in human hepatoma cells inducing calcium-modulated ER stress and mitochondrial disorder.Ijms 24 01566 i011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bonsignore, G.; Martinotti, S.; Ranzato, E. Endoplasmic Reticulum Stress and Cancer: Could Unfolded Protein Response Be a Druggable Target for Cancer Therapy? Int. J. Mol. Sci. 2023, 24, 1566. https://doi.org/10.3390/ijms24021566

AMA Style

Bonsignore G, Martinotti S, Ranzato E. Endoplasmic Reticulum Stress and Cancer: Could Unfolded Protein Response Be a Druggable Target for Cancer Therapy? International Journal of Molecular Sciences. 2023; 24(2):1566. https://doi.org/10.3390/ijms24021566

Chicago/Turabian Style

Bonsignore, Gregorio, Simona Martinotti, and Elia Ranzato. 2023. "Endoplasmic Reticulum Stress and Cancer: Could Unfolded Protein Response Be a Druggable Target for Cancer Therapy?" International Journal of Molecular Sciences 24, no. 2: 1566. https://doi.org/10.3390/ijms24021566

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 -