Next Article in Journal
Sulforaphane-Induced Klf9/Prdx6 Axis Acts as a Molecular Switch to Control Redox Signaling and Determines Fate of Cells
Previous Article in Journal
Spontaneous Development of Dental Dysplasia in Aged Parp-1 Knockout Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoparticles Targeting STATs in Cancer Therapy

1
Department of Basic Science, Faculty of Veterinary Medicine, University of Tabriz, Tabriz 5166616471, Iran
2
Department of Basic Science, Shoushtar Branch, Islamic Azad University, Shoushtar 6451741117, Iran
3
Centre for Research in Medical Devices (CÚRAM), National University of Ireland Galway, Newcastle, Galway H91 W2TY, Ireland
4
Neuroscience Research Center, Institute of Neuropharmacology, Kerman University of Medical Sciences, Kerman 7619813159, Iran
5
Department of Basic Medical Sciences, Neyshabur University of Medical Sciences, Neyshabur 9318614139, Iran
6
Pharmaceutics Research Center, Institute of Neuropharmacology, Kerman University of Medical Sciences, Kerman 7616911319, Iran
7
Department of Pharmacology, Yong Loo Lin School of Medicine, National University of Singapore, Singapore 117600, Singapore
*
Authors to whom correspondence should be addressed.
Cells 2019, 8(10), 1158; https://doi.org/10.3390/cells8101158
Submission received: 16 August 2019 / Revised: 20 September 2019 / Accepted: 25 September 2019 / Published: 27 September 2019

Abstract

:
Over the past decades, an increase in the incidence rate of cancer has been witnessed. Although many efforts have been made to manage and treat this life threatening condition, it is still one of the leading causes of death worldwide. Therefore, scientists have attempted to target molecular signaling pathways involved in cancer initiation and metastasis. It has been shown that signal transducers and activator of transcription (STAT) contributes to the progression of cancer cells. This important signaling pathway is associated with a number of biological processes including cell cycle, differentiation, proliferation and apoptosis. It appears that dysregulation of the STAT signaling pathway promotes the migration, viability and malignancy of various tumor cells. Hence, there have been many attempts to target the STAT signaling pathway. However, it seems that currently applied therapeutics may not be able to effectively modulate the STAT signaling pathway and suffer from a variety of drawbacks such as low bioavailability and lack of specific tumor targeting. In the present review, we demonstrate how nanocarriers can be successfully applied for encapsulation of STAT modulators in cancer therapy.

1. Introduction

As a multidisciplinary field, nanotechnology can be extensively applied in medicine, chemistry and engineering [1,2]. Nanotechnology aims to the development of materials and structures with low size (1–1000 nm) [3]. Over the past decades, significant attention has been directed towards nanotechnology for diagnosis and management of cancer [4]. Clinically, application of a number of strategies such as chemotherapy, radiotherapy and surgery seems to be beneficial in the inhibition of tumorigenesis. However, metastasis and subsequent recurrence are the most challenging problems in cancer therapy [5,6]. Accumulating data demonstrates that there are few major drawbacks associated with conventional cancer therapeutic strategies including the resistance of cancer cells to chemotherapy and radiotherapy, the invasive feature of surgery, unexpected side effects and poor tumor targeting as well as low bioavailability of anti-tumor drugs [7], thereby demanding novel strategies for cancer therapy.
Nanocarriers can be considered as potential candidates in cancer therapy. The low particle size of nanocarriers enables them to effectively penetrate into the blood–brain barrier (BBB) [7]. It appears that application of nanocarriers is associated with enhanced bioavailability of the drug. In fact, nanocarriers can encapsulate the drug to protect it against degradation thus resulting in its enhanced bioavailability for therapeutic application [8]. It is noteworthy that nanoparticles (NPs) provide a minimally invasive-cancer therapy [9] and simultaneously, significantly diminish the chance of resistance and adverse impacts by using a low amount of anti-tumor drug, while the anti-tumor activity is at its highest level [10]. It is possible that mild pH of the tumor microenvironment degrades the drug and more importantly, conventional cancer therapeutic strategies suffer from a lack of specific targeting of cancer cells leading to their toxicity against normal cells. A variety of receptors undergo upregulation in tumor cells and receptor-targeted NPs are of importance in enhancing the delivery of drug into cancer cells [11]. Therefore, based on the high incidence rate of cancer [12], using nanotechnology seems to be a promising approach against this life threatening condition due to its capability in enhancing the anti-tumoral actions of drugs. Currently, various NPs are applied for the delivery of anti-tumor drugs such as solid lipid nanoparticles (SLNs) [13], liposomes [14], niosomes [15], micelles [16], polymeric NPs [17,18,19], carbon nanostructures [20], viral NPs [21], mesoporous silica NPs [22] and gold NPs [23]. Besides, different methods can be used for drug loading. It has been established that various drugs can be predominantly loaded on nanocarriers by encapsulation, as well as covalent or electrostatic binding [24,25,26,27,28].
Cancer is considered as a malignant condition and deregulation of various oncogenic signaling pathways are generally involved in its progression [29]. For example, Wnt signaling pathway is one of the major signaling cascades that can enhance the proliferation and metastasis of cancer cells [30,31,32]. On the contrary, nuclear factor erythroid 2-related factor 2 (Nrf2) can also be targeted to overcome resistance of cancer cells to chemotherapy [33]. These studies demonstrate that diverse oncogenic signaling pathways can be effectively modulated to develop novel strategies for cancer therapy [34,35,36,37]. In the present review, we describe the various ongoing efforts for delivery of anti-tumor drugs primarily targeting oncogenic STAT3 signaling pathway.

2. STATs Family: Members and Signaling Pathways

The discovery of signal transducers and activator of transcription (STAT) signaling pathway returns back to 1997, when the scientists have found that STATs are involved in mediation of interferon signaling [38]. A variety of hormones, cytokines and growth factors function as upstream modulators of Janus kinase (JAK)/STAT signaling pathway resulting in regulation of important biological mechanisms such as cell cycle, cell differentiation, cell proliferation and apoptosis [39,40,41,42,43]. Besides, the JAK/STAT signaling pathway is involved in complicated mechanisms such as immune regulation and cancer [44,45]. In mammals, there are four genes encoding JAK1, JAK2, JAK3 and TYK2, and seven genes encoding STAT1, STAT2, STAT3, STAT4, STAT 5A and 5B, and STAT6 [46,47,48]. The expression of JAK3 occurs primarily in hematopoietic cells, while JAK1, JAK2 and TYK2 are ubiquitously expressed [49]. Four major domains are associated with JAKs including N-terminal FERM-domain, SH2-like domain, pseudokinase domain and JH1 domain. It has been demonstrated that FERM and SH2-like domains can contribute to the interaction of JAKs with their receptors [50,51]. On the contrary, STATs effectively affect the transcription of target genes by interaction with DNA regulatory elements (DREs) [52].
Hormones, cytokines and growth factors bind to the receptor leading to the phosphorylation of receptor-associated JAKs. This phosphorylation occurs on the tyrosine (Tyr) residue of JAK that is necessary for stimulation of kinase activity [53]. Importantly, attachment of a ligand to the cell membrane receptor promotes the interaction of receptor-JAK complex to facilitate the phosphorylation of tyrosine residues of cytoplasmic domains of receptors [54], which can than form docking sites for SH2 domain-containing STAT proteins. Then, phosphorylation of Tyr residues within the C terminal domain of receptor-bound STATs occurs resulting in detachment of STATs from receptors and generation of homo- and heterodimers. The STAT proteins accumulate in the cytoplasm and then, translocate into the nucleus where they bind to the members of gamma-activated sites (GASs) and interferon-stimulated response elements (ISREs) [55,56,57,58,59,60]. ISREs are limited to interferon (IFN) signaling, while GASs are present at the promoter of genes including acute-phase proteins [61]. It is noteworthy that STAT3 is capable of transferring from the cytoplasm to nucleus and vice versa, regardless of its phosphorylation status [62].
A number of proteins play a significant role in regulation of the JAK/STAT signaling pathway. These characteristic proteins include the suppressor of cytokine signaling (SCOS), protein tyrosine phosphatases (PTP) and protein inhibitors of activated STATs (PIAS) [63]. SCOS proteins suppress JAK/STAT signaling pathway via A) inhibition of JAK phosphorylation, and B) blocking STAT recruitment [64,65,66]. PIAS proteins prevent the interaction of STAT proteins with DNA. PTP are involved in suppressing JAK proteins [63].

3. Role of STATs in Cancer Hallmarks

Importantly, it has been shown that dysregulation of the STAT signaling pathway is associated with development of a number of pathological conditions, particularly cancer. Notably, it seems that STAT1 is considered as a pro-tumorigenic pathway, so that several studies have revealed that the STAT1 signaling pathway significantly enhances the proliferation and malignancy of cancer cells [9,67,68]. However, there are a variety of studies that demonstrate that down-regulation of STAT1 is related to the enhanced invasion and metastasis of tumor cells [69]. Taking these reports into account, dysregulation of STAT1 (upregulation and down-regulation) occurs in tumor cells. It has been shown that interleukin-6 (IL-6) stimulates the malignancy and proliferation of tumor cells. It appears that STAT2 enhances the proliferation of cancer cells by elevating the level of IL-6/STAT3 [70]. A similar story occurs for STAT3, so that various research studies have confirmed that the STAT3 signaling pathway incredibly increases tumor migration, tumor size and tumor malignancy [71,72,73,74,75,76,77]. However, targeting the STAT4 signaling pathway can be considered as a promising strategy in cancer therapy. For example, an upregulation of STAT4 protein can enhance the survival time of patients [78]. Notably, STAT proteins may also act as prognostic signatures in gastric cancer. Moreover, it has been demonstrated that among STAT proteins, STAT4 can determine the prognosis of gastric cancer due to its association with high levels of dendritic cells and CD8+ T cells, whereas STAT3 and STAT6 have minimal prognostic value [79]. Furthermore, miRNA-141-3p inhibits the viability and metastasis of gastric cancer cells through the upregulation of STAT4 [80]. STAT5 and STAT6 contribute in the progression of cancer [41,81,82]. The various members of STAT proteins may also function as upstream modulators of other STAT proteins. STAT5 is an example of this case and it is capable of regulating the expression of STAT3 in tumor cells [83]. Besides, the interaction between STAT proteins may be vital in regulating gene transcription [84]. The STAT signaling pathway can also be involved in the resistance of cancer cells to chemotherapy [85]. For example, accumulating data shows that the RAS signaling pathway may be a key to the malignancy of colorectal cancer (CRC) cells [86,87]. It was found that the interaction between RAS and IFN/STAT signaling pathways [88] can be vital for the induction of the resistance of tumor cells to chemotherapy with trametinib. RAS triggers IFN/STAT signaling pathway by stimulation of STAT1 phosphorylation. Although administration of trametinib is associated with MEK inhibition, the phosphorylation of STAT1 was not found to be affected [89]. IFN/STAT signaling pathway can induce drug resistance in colorectal cancer (CRC) cells via interaction with RAS [89]. Cancer stem cells develop resistance to chemotherapeutic agents by stimulation of the JAK-STAT signaling pathway. Disruption of the JAK-STAT pathway reduces the proliferation and viability capabilities of cancer stem cells [90]. In respect to the potential role of STAT proteins in cancer invasion and metastasis, a number of studies have been performed to elucidate the upstream modulators of STAT signaling pathway. Long non-coding RNA (lncRNA) PART1 is suggested to be involved in enhancing the malignancy of lung cancer cells via induction of JAK-STAT signaling pathway [43]. MicroRNAs (miRs) are short non-coding RNA molecules, which can affect the invasion of cancer cells due to their role in regulation of important biological processes such as cell differentiation, cell proliferation, cell growth and apoptosis [91,92,93]. It appears that miR-15a-3p effectively diminishes the malignancy of liver cancer cells by down-regulation of STAT3 [94]. SOCS plays a significant role in induction of immune system [95]. Moreover, in lung cancer, a reduction in SOCS3 enhances the expression of STAT3 thus causing the progression of cancer cells. MiR-410 down-regulation increases the expression of SOCS3 leading to the decreased level of STAT3 protein and minimized progression of lung cancer cells [96]. Notably, application of STAT3 inhibitor is suggested to be beneficial in the treatment of head and neck cancers [97]. These findings highlight this notion that STAT signaling pathway perturbation is involved in various cancers and targeting this pathway using synthetic or naturally occurring drugs is of importance in cancer therapy. Besides, detecting the mediators of the STAT signaling pathway such as lncRNAs and miRs can be beneficial in genetic manipulation. Based on the complexity and dynamic feature of the STAT signaling pathway, providing an effective modulation of the STAT pathway depends on targeting various signaling molecules involved in regulating this multifunctional pathway.

4. STATs Inhibitors

Contemporary therapy is based on targeting the pathways and mechanisms that diseases use. To accomplish this, we should first identify these mechanisms and then create individual molecular drugs that specifically target these pathways. From the theoretical standpoint, targeting one pathway seems very beneficial, but in practice this single therapy is not completely effective and we have not witnessed substantial progress in the eradication of sophisticated pathological disorders, particularly cancer. Besides, using one drug enhances the chance of resistance, so the application of several drugs that affect various molecular pathways diminishes the risk of resistance developing. The targeted therapy of STATs has been advanced due to identification of the unique roles of STATs in various cellular processes. However, over the recent decades, natural and synthetic inhibitors have been developed that can target STAT signaling pathway in various disorders, specifically cancer [98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113]. Among the STAT proteins, there have been many efforts to detect the inhibitors of STAT3, leading to development of more synthetic and naturally occurring inhibitors of STAT3 compared to other STAT proteins. This may be due to this fact that STAT3 and STAT1 proteins are involved in the progression of several tumor cells [114,115]. It can be concluded that STAT3 inhibitors may negatively affect STAT3 signaling pathway via four major actions [116]: i) Inhibition of SH2 domain or dimerization, ii) influencing upstream mediators of STAT3 such as JAK, iii) suppressing STAT3-DNA domain binding, and iv) endogenous modulators of STAT3. However, there are a variety of difficulties that restrict targeting STAT signaling pathway. For instance, it has been demonstrated that there is a similarity among the structures of STAT proteins, leading to reduced specificity in targeting. Moreover, there is a need for more studies to confirm the safety of these inhibitors in clinical trials.
Furthermore, there have been some attempts to interfere with the transcription of genes. However, these strategies suffer from low specificity and a lack of knowledge about appropriate therapeutic doses [117].
Curcumin is a naturally occurring nutraceutical compound with diverse pharmacological impacts such as antioxidant, anti-inflammatory, anti-diabetic and anti-tumor [118,119,120,121]. It appears that curcumin is capable of targeting different signaling pathways in stimulation of its anti-tumor activity and JAK-STAT pathway is one of them [122,123,124,125]. The induction of apoptotic cell death in H-Ras human mammary epithelial cells is a consequence of direct interaction of curcumin with cysteine (Cys) 259 residue of STAT3. This interaction can lead to the inactivation of STAT3 and subsequently, sensitize tumor cells into apoptotic cell death [126]. Pravastatin is one of the key members of statins with the capability of reducing cholesterol and improving cardiovascular parameters [127]. The administration of pravastatin has been found to be associated with down-regulation of IFN-γ levels and amelioration of atherosclerosis via reducing the expression of STAT1 phosphorylation [128]. It has been demonstrated that pimozide as a neuroleptic drug is capable of targeting STAT proteins [129]. Pimozide can remarkably diminish the phosphorylation level of STAT5 resulting in high cytotoxicity against K562 cells [130]. As an immunosuppressive compound, leflunomide effectively inhibits IgG1 generation by suppressing tyrosine phosphorylation of JAK3 and STAT6 [131]. Niflumic acid has demonstrated great potential in treatment of asthma by modulation of STAT signaling pathway. It seems that IL-13 is vital in induction of asthma through stimulation of chronic inflammation, eosinophilic infiltration, reversible airway narrowing and airway hyperresponsiveness (AHR) [132,133,134,135]. Niflumic acid prevents IL-13-mediated asthma by down-regulation of JAK2 and STAT6 [136]. Cinnamon has a long story in traditional medicine and is extensively used in amelioration of pathological conditions, particularly cancer [137]. The immunomodulatory impact of cinnamon can be attributed to the modulation of STAT proteins, as it suppresses the expression of STAT4 to inhibit the production of IFN-γ [138]. Taking these reports into account, it appears that inhibiting the phosphorylation may be an important strategy for STAT suppression. However, some of them directly bind to the target STAT and suppress its activity. Table 1 and Table 2 summarize the selected pharmacological inhibitors of STAT proteins.

5. STATs Gene Silencing by RNA Interference

The introduction of the RNA interference (RNAi) mechanism returns back to two decades ago [163]. This phenomenon has resulted in a great advancement in the investigation of the function of RNAs [12]. At this mechanism, small RNAs containing 18–30 nucleotides are designed to act on long RNAs. This action involves stimulation or inhibition of cleavage at the post-transcriptional level [164]. In respect to the modulatory effect of RNAs on STATs, it seems that regulation of RNAs using RNAi is beneficial in the treatment of pathological disorders associated with dysregulation of STAT proteins [165]. Modulation of STAT3 using RNAi is advantageous in treatment of a laryngeal tumor. An animal model was induced to examine the anti-tumor activity. This animal model included immunocompromised mice in that HepG2 cells were transplanted. Suppressing STAT3 protein remarkably diminished the growth rate of tumors. It appears that STAT3 down-regulation is associated with reduced expression of Bcl-2, cyclin D1 and survivin genes leading to the stimulation of apoptotic cell death [166]. A similar observation was noted in pancreatic cancer cells [167], where after suppressing STAT3 expression using STAT3 short hairpin RNA (shRNA) expression vectors, the malignancy and metastasis of pancreatic cancer cells remarkably reduced. Besides, the mRNA expression of matrix metalloproteinase-2 (MMP-2) and the vascular endothelial growth factor (VEGF) underwent down-regulation after STAT3 knockdown, demonstrating the pivotal role of STAT proteins in progression of cancer cells. In spite of much progress in cancer therapy and developing novel drugs targeting various signaling pathways, scientists are not yet able to effectively remedy this life threatening condition. Another study puts emphasis on the potential role of STAT3, STAT5A and STAT5B in the malignancy and invasion of leukemia. In this study, K-562 cells were transfected by anti-STAT3, anti-STAT5A and anti-STAT5B small interfering RNAs (siRNAs). Importantly, the expression of mentioned STAT proteins significantly reduced. It was found that preventing the expression of STAT3, STAT5A and STAT5B is related to the enhanced apoptosis in cancer cells [168]. Finding a new way in treatment of astrocytoma attracts much attention due to the high incident rate of this primary central nervous system tumor. Based on the vital role of STAT3 in the malignancy of tumor cells, inhibition of STAT3 in astrocytoma cells can diminish the mortality resulted from this disorder [169]. STAT3 knockdown promotes the sensitivity of astrocytoma cells into apoptosis.
Furthermore, in respect to the role of STAT3 in inducing the expression of anti-apoptotic factors such as Bcl-xL and survivin, down-regulation of STAT3 is related to the decreased viability and proliferation of cancer cells. However, scientists have faced challenges in the treatment of other brain tumors, particularly glioblastoma. In spite of much effort in the treatment of glioblastoma, it still remains one of the most malignant brain tumors [170]. The capabilities of cells to initiate, progress and recur have led to the high malignancy of these tumor cells [171,172,173,174,175]. Gene manipulation is of importance in reducing the malignancy of glioblastoma cells. Interestingly, inhibition of STAT3 using RNAi can stimulate apoptotic cell death in glioblastoma cells by upregulation of caspase-3 and BAX, and down-regulation of Bcl-2 and cyclin-D. Besides, STAT3 inhibition decreases the CD133+ cell proportion and subsequently, sensitizes cancer cells to apoptosis [176]. On the other hand, one of the difficulties in radio- and chemo-therapy is the resistance of cancer cells. Investigation of molecular signaling pathways and subsequently, regulation of them can be beneficial in enhancing the efficacy of radio- and chemo-therapy. It seems that STAT3 knockdown remarkably elevates the efficacy of radio-therapy in laryngeal carcinoma by reducing the expression of Bcl-2 and VEGF, and enhancing the number of apoptotic cell death [177]. These studies obviously highlight this fact that STAT proteins have vital roles in migration, proliferation and malignancy of cancer cells and modulation of their expression using RNAi interference is a great strategy in combating cancer cells.

6. Nano-Technological Approaches for Targeting STATs

6.1. Nanoparticles

6.1.1. In Vitro

Based on the statistics reported by American Cancer Society, the efforts for management of cancer should be continued to prevent the high mortality and morbidity associated with this life threatening condition [178]. Cancer cells apply various signaling pathways to ensure their progression. These dynamic and flexible molecular pathways provide a challenge in the treatment of cancer [9,179,180]. On the other hand, although anti-tumor drugs targeting signaling pathways have been introduced in cancer therapy, low bioavailability and lack of targetability diminish the anti-tumor activity of these drugs. To date, NPs have been used for the treatment of various pathological disorders [180] and this capability has been applied in cancer therapy. Hydroxyapatite (HAP) is an important biomaterial with extensive applications in tissue engineering and bone repair [181,182]. HAP has demonstrated great potential in the delivery of DNA and proteins due to its excellent properties such as biocompatibility and porosity [183]. HAP-based NPs can be considered as a promising strategy in the delivery of anti-STAT3 shRNA. HAP NPs effectively deliver anti-STAT3 shRNA to prostate cancer cells leading to the induction of apoptosis and decreased viability of cancer cells. During this transfection, STAT3 down-regulation significantly diminished the expression of Bcl-2, VEGF and cyclin D1. Furthermore, the expression of caspase-3 and BAX underwent upregulation [184]. SLNs are another option in the delivery of small molecule drugs and genetic materials. High biocompatibility and great stability have resulted in application of SLNs for gene delivery [185]. Loading a STAT3 inhibitor on SLNs is of importance in combating lung cancer cells. SLNs protected genetic materials against DNasel and serum-mediated degradation. Encapsulation of DNA by SLNs preserved its supercoiled and circular formation. STAT3 inhibitor-loaded SLNs significantly sensitized lung cancer cells to cisplatin-mediated apoptosis (Table 3) [186]. In respect to the potential role of STAT3 in enhancing the malignancy of cancer cells [187], this signaling pathway has obtained much attention in triple negative breast cancer (TNBC) therapy and a number of drugs approved by the Food and Drug Administration (FDA) such as niclosamide have been used in treatment of TNBC as inhibitors of STAT3 [188]. In accordance to the efficacy of SLNs in the delivery of STAT inhibitors, loading a STAT3 inhibitor on SLNs remarkably decreases the viability of cancer cells by stimulation of apoptosis via down-regulation of STAT3 phosphorylation [13]. SLNs have been applied in treatment of ovarian cancer due to their potential in delivery of STAT3 siRNA and consequently, stimulation of apoptotic cell death through down-regulation of Bcl-2 and survivin [189].
Accumulating data demonstrates that SHP-1 may be able to modulate stemness and the epithelial-to-mesenchymal transition (EMT) of tumor cells by targeting the JAK2/STAT3 signaling pathway [190,191,192]. Therefore, NP-mediated SHP-1 regulation is of interest in cancer therapy. ZnAs@SiO2 NPs use the same strategy in reducing hepatocellular carcinoma malignancy. It seems that application of ZnAs@SiO2 NPs significantly diminishes the expression of stemness markers such as CD133, Sox-2 and Oct-4. Besides, these NPs are capable of induction of apoptotic cell death and reducing the metastasis and migration of hepatocellular carcinoma cells by EMT inhibition. These anti-tumor activities arise as a result of disruption in the SHP-1/JAK2/STAT3 signaling pathway [193]. Receptor-targeted delivery enhances the capability of NPs in decreasing the viability of cancer cells. It has been demonstrated that CD38 has a minimal expression in normal cells, while its overexpression occurs in multiple myeloma (MM) cells [194]. There have been efforts to target CD38 at MM cells and daratumumab has been used for this purpose [195,196,197]. Moreover, anti-CD38-decorated NPs carrying the STAT3 inhibitor have been reported to have high cellular uptake with great anti-tumor activity [198].
Poly(lactic-co-glycolic acid) (PLGA) has a variety of excellent properties such as biocompatibility and biodegradability. FDA has approved the application of PLGA for human uses. PLGA NPs have a size similar to pathogens leading to their phagocytosis by dendritic cells (DCs). This feature has resulted in application of PLGA NPs for delivery of drugs into DCs [199,200,201,202,203,204]. It appears that PLGA provides a suitable platform for conjugation of JSI-124, as a STAT3 inhibitor. JSI-124 PLGA NPs have great anti-tumor activity against B16 melanoma cells by reducing the expression of STAT3 in DCs and enhancing the function of DCs in terms of promoting the production of T cells leading to the cancer immunotherapy [205]. The capability of PLGA NPs in releasing drugs in a sustained-released behavior is of importance in co-delivery of paclitaxel, a chemotherapeutic agent and STAT siRNA to sensitize lung cancer cells to apoptotic cell death [206]. Taking everything into account, in respect to the ability of PLGA NPs in delivery, anti-STAT3-loaded PLGA NPs can be considered as promising agents in cancer immunotherapy by targeting DCs [207]. It has been demonstrated that STAT3-siRNA-loaded NPs have high cellular uptake by tumor cells leading to their high efficacy in reducing the malignancy of cancer cells. It appears that clathrin-mediated endocytosis participates in cellular uptake of STAT3-siRNA-loaded NPs by melanoma cells [208].

6.1.2. In Vivo

Melanoma is one of the malignant skin cancers that proliferation of pigment producing melanocytes occurs in the epidermis. Surgery and chemotherapy are considered as current strategies in melanoma therapy [209,210,211]. However, one of the problems associated with chemotherapy is the resistance of tumor cells [29]. Using gene therapy enhances the anti-tumor activity of chemotherapeutic agents. Co-delivery of imatinib and anti-STAT3 siRNA (non-invasive topical iontophoretic administration) using gold NPs is related to a remarkable decrease in tumor volume and tumor weight in melanoma tumor bearing mice, showing the efficacy of gold NPs in treatment of melanoma by inhibition of STAT3 [212]. Erlotinib (ELTN) is extensively used in chemotherapy with the capability of targeting epidermal growth factor receptor (EGFR) gene. However, resistance of cancer cells challenges the potential of this agent in chemotherapy [213,214]. Fedratinib (FDTN) is a small molecule known as the JAK2 inhibitor and is applied in the treatment of myelofibrosis [215]. Co-administration of ELTN and FDTN using biodegradable NPs leads to the satisfactory results in ELTN-resistance non-small cell-lung cancer (NSCLC) cells. The biodegradable NPs had great stability and effectively released drug at mild acidic pH of the tumor microenvironment. Loading a combination of ELTN and FDTN on NPs not only enhances the anti-tumor activity by inhibition of the JAK2/STAT3 signaling pathway, but also diminishes the systemic adverse effects [216]. As it was mentioned, HAP has human applications due to its high biocompatibility. Besides, it seems that HAP has anti-tumor activity making its appropriate for cancer therapy [217,218,219,220,221,222,223]. HAP NPs are capable of inhibiting the progression and invasion of prostate tumor cells in mouse model by reducing the expression of STAT3 resulting in down-regulation of Bcl-2, VEGF and cyclin D1 [224]. A newly developed nanocarrier for the delivery of siRNA should be capable of protection of siRNA against degradation, promoting siRNA potency and simultaneously, improving the biodistribution and pharmacokinetics [224]. Polymeric NPs have demonstrated great potential in this field and polyethyleneimine is among them [225,226,227,228]. Loading STAT3 siRNA on lipid-substituted PEI is associated with decreased viability and proliferation of tumor cells by upregulation of caspase-3 and IL-6, and down-regulation of STAT3 and VEGF [229].
5,2/-4/-trihydroxy-6,7,5/-trimethoxyflavone (TTF1) is a naturally occurring compound exclusively found in Sorbaria sorbifolia (SS) [230,231]. TTF1 has great pharmacological effects such as anti-tumor activity. However, low bioavailability and biodegradation restrict the therapeutic activities of TTF1 [232]. TTF1-loaded NPs are able to remarkably suppress angiogenesis and metastasis of human hepatoma cancer cells by down-regulation of STAT3. It appears that decreased invasion of cancer cells is a consequence of MMP-2 and MMP-9 down-regulation. Besides, anti-angiogenic effect of TTF1-NPs is mediated by reducing the expression of VEGF [233]. BBB is considered as one of the most challenging problems in penetration of drugs into brain. PEI-PLGA NPs solve this problem by enhancing the crossing of STAT3 siRNA through BBB [234].

6.2. Liposomes

6.2.1. In Vivo

It seems that liposomes are potential candidates in the treatment of skin cancer. This notion emanates from the capability in crossing over the stratum corneum layer of skin [235]. It has been demonstrated that edge activators (transferosomes)- or ethanol (ethosomes)-based liposomes are able to deeply penetrate into the skin [236]. Besides, physical techniques such as iontophoresis have enhanced the penetration potential of liposomes into the skin [237,238,239,240]. Therefore, liposomes can serve as promising candidates for delivery of STAT proteins in the treatment of skin cancers [241]. It appears that curcumin- and STAT3 siRNA-loaded liposomes significantly down-regulate the expression of STAT3 protein leading to the inhibition of tumor invasion and a remarkable reduction in tumor weight and tumor volume [242].

6.2.2. In Vitro

Targeting tumor-associated macrophages (TAMs) is of importance in cancer therapy due to the potential role of TAMs in the tumor microenvironment and enhancing the malignancy, invasion, angiogenesis and resistance of cancer cells [243,244]. It is held that enhanced TAM-infiltration is associated with a decrease in survival time of patients with cancer [245,246,247]. Notably, disruption in the STAT3 signaling pathway effectively promotes anti-tumor immunity by enhancing the production of TNF-α and stimulation of M1-like reprogramming of macrophages [248,249,250,251,252]. Hence, providing STAT3 modulation in macrophages is of interest in improving anti-tumor immunity. CD163-targeted crosolic acid-containing liposomes prevent the expression of STAT3 in macrophages, resulting in enhanced anti-tumor immunity by increasing TNF-α, IFN-γ, IL-12 and IL-2 levels, and decreasing the IL-10 level [245]. Similar to in vivo findings, co-delivery of curcumin and STAT3 by deformable cationic liposomes is associated with cell growth inhibition and apoptosis induction. It is held that clathrin-induced endocytosis mediates the penetration of liposomes into skin [241].

6.3. Micelles

In Vitro and In Vivo

Micelles were first introduced in 1984 for the delivery of drugs [253,254]. Micelles are able to remarkably improve the bioavailability and anti-tumor activity of drugs [255,256]. It seems that polymeric micelles have higher permeability and retention effect compared to the conventional micellar nanocarriers [257,258] making them appropriate for drug delivery. There are two studies that have investigated the efficiency of micellar NPs in the delivery of STAT inhibitors in melanoma cells both in vitro and in vivo. It was found that administration of STAT3 inhibitor-loaded polymeric micelles results in apoptotic cell death in melanoma cells and down-regulates VEGF expression. Besides, these nanocarriers have greater biocompatibility and improve anti-tumor immunity by enhancing DC-mediated IL-12 production [259,260]. The potential application of nanoparticles in targeting STATs is summarized in Figure 1.

7. Conclusion and Future Trends

In respect to the vital role of STAT proteins in various important biological processes including cell cycle, differentiation, apoptosis and cell proliferation, any impairment in the STAT signaling pathway is associated with the development of pathological conditions, particularly cancer. As a consequence, targeting the STAT signaling pathway has demonstrated a great potential in cancer therapy. On the other hand, there have been some difficulties in the delivery of drugs that target the STAT signaling pathway. Therefore, it seems that application of nanocarriers for loading STAT modulators may be important in terms of releasing drug into the tumor site and inhibition of resistance of cancer cells by loading the optimum amount of drug. Until now, various nanoparticles have been designed for targeting the STAT signaling pathway, especially STAT3, which include gold nanoparticles, hydroapatite nanocarriers, PLGA nanoparticles, micelles, solid lipid nanoparticles, liposomes and microbubbles. These nanocarriers have been applied in various cancers both in vitro and in vivo, and exhibited high potentiality in reducing the viability, migration and malignancy of tumor cells by regulating the expression of STAT proteins. However, more studies are needed to elucidate the efficacy of nanoparticles in targeting the STAT signaling pathway for cancer therapy. Although huge emphasis has been put on the capabilities and benefits of using NPs in delivery of STAT to cancer cells, it has been reported that only 0.7% of administered NPs are found to be delivered to the tumor site, thereby challenging the potential role of NPs in drug delivery [271]. This issue needs to be carefully addressed in future studies.

Funding

R.M. acknowledges financial supports of Kerman University of Medical Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

BBBblood-brain barrier
NPsnanoparticles
STATsignal transducers and activator of transcription
DREDNA regulatory elements
Tyrtyrosine
GASgamma-activated sites
ISREsinterferon-stimulated response elements
IFNinterferon
SCOSsuppressor of cytokine signaling
PTPprotein tyrosine phosphatase
PIASprotein inhibitors of activated STATs
ILinterleukin
CRCcolorectal cancer
lncRNAlong non-coding RNA
miRmicroRNA
Cyscysteine
AHRairway hyper responsiveness
RNAiRNA interference
MMPmatrix metalloproteinase
VEGFvascular endothelial growth factor
SLNssolid lipid NPs
HAPhydroxyapatite
TNBCtriple negative breast cancer
Nicloniclosamide
FDAFood and Drug Administration
EMTepithelial-to-mesenchymal transition
MMmultiple myeloma
DCsdendritic cells
ELTNerlotinib
EGFRepidermal growth factor receptor
FDTNfedratinib
NSCLCnon-small cell-lung cancer
PEIpolyethylenimine
TTF1trimethoxyflavone
SSSorbaria sorbifolia
TAMtumor-associated macrophage
TADCstumor-associated CDs
JAKJanus kinase

References

  1. Gao, A.; Hu, X.-L.; Saeed, M.; Chen, B.-F.; Li, Y.-P.; Yu, H.-J. Overview of recent advances in liposomal nanoparticle-based cancer immunotherapy. Acta Pharmacol. Sin. 2019, 40, 1129–1137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Liu, J.; Zhang, R.; Xu, Z.P.J.S. Nanoparticle-Based Nanomedicines to Promote Cancer Immunotherapy: Recent Advances and Future Directions. Small 2019, 15, 1900262. [Google Scholar] [CrossRef] [PubMed]
  3. Ferrari, M. Cancer nanotechnology: Opportunities and challenges. Nat. Rev. Cancer 2005, 5, 161. [Google Scholar] [CrossRef] [PubMed]
  4. Singh, R.P.; Sharma, G.; Kumari, L.; Koch, B.; Singh, S.; Bharti, S.; Rajinikanth, P.S.; Pandey, B.L.; Muthu, M.S. RGD-TPGS decorated theranostic liposomes for brain targeted delivery. Colloids Surf. B Biointerfaces 2016, 147, 129–141. [Google Scholar]
  5. Wan, L.; Pantel, K.; Kang, Y. Tumor metastasis: Moving new biological insights into the clinic. Nat. Med. 2013, 19, 1450. [Google Scholar] [CrossRef] [PubMed]
  6. Barker, H.E.; Paget, J.T.; Khan, A.A.; Harrington, K.J. The tumour microenvironment after radiotherapy: Mechanisms of resistance and recurrence. Nat. Rev. Cancer 2015, 15, 409. [Google Scholar] [CrossRef] [PubMed]
  7. Ahmadi, Z.; Mohammadinejad, R.; Ashrafizadeh, M. Drug delivery systems for resveratrol, a non-flavonoid polyphenol: Emerging evidence in last decades. J. Drug Deliv. Sci. Technol. 2019, 51, 591–604. [Google Scholar] [CrossRef]
  8. Machado, N.D.; Fernández, M.A.; Díaz, D.D. Recent Strategies in Resveratrol Delivery Systems. ChemPlusChem 2019, 84, 951–973. [Google Scholar] [CrossRef]
  9. Ashrafizadeh, M.; Mohammadinejad, R.; Tavakol, S.; Ahmadi, Z.; Roomiani, S.; Katebi, M. Autophagy, anoikis, ferroptosis, necroptosis, and endoplasmic reticulum stress: Potential applications in melanoma therapy. J. Cell. Physiol. 2019, 234, 19471–19479. [Google Scholar] [CrossRef]
  10. Rawal, S.; Patel, M.M. Threatening cancer with nanoparticle aided combination oncotherapy. J. Control. Release 2019, 301, 76–109. [Google Scholar] [CrossRef]
  11. Johnsen, K.B.; Moos, T. Revisiting nanoparticle technology for blood–brain barrier transport: Unfolding at the endothelial gate improves the fate of transferrin receptor-targeted liposomes. J. Control. Release 2016, 222, 32–46. [Google Scholar] [CrossRef] [PubMed]
  12. Muhammad, T.; Zhang, F.; Zhang, Y.; Liang, Y. RNA interference: A natural immune system of plants to counteract biotic stressors. Cells 2019, 8, 38. [Google Scholar] [CrossRef] [PubMed]
  13. Pindiprolu, S.K.S.; Chintamaneni, P.K.; Krishnamurthy, P.T.; Ratna Sree Ganapathineedi, K. Formulation-optimization of solid lipid nanocarrier system of STAT3 inhibitor to improve its activity in triple negative breast cancer cells. Drug Dev. Ind. Pharm. 2019, 45, 304–313. [Google Scholar] [CrossRef] [PubMed]
  14. Askarizadeh, A.; Butler, A.E.; Badiee, A.; Sahebkar, A. Liposomal nanocarriers for statins: A pharmacokinetic and pharmacodynamics appraisal. J. Cell. Physiol. 2019, 234, 1219–1229. [Google Scholar] [CrossRef] [PubMed]
  15. Rouholamini, S.E.Y.; Moghassemi, S.; Maharat, Z.; Hakamivala, A.; Kashanian, S.; Omidfar, K. Effect of silibinin-loaded nano-niosomal coated with trimethyl chitosan on miRNAs expression in 2D and 3D models of T47D breast cancer cell line. Artif. Cells Nanomed. Biotechnol. 2018, 46, 524–535. [Google Scholar] [CrossRef] [PubMed]
  16. Gao, Y.-E.; Bai, S.; Ma, X.; Zhang, X.; Hou, M.; Shi, X.; Huang, X.; Chen, J.; Wen, F.; Xue, P. Codelivery of Doxorubicin and Camptothecin by Dual-responsive Unimolecular Micelle-based β-cyclodextrin for Enhanced Chemotherapy. Colloids Surf. B Biointerfaces 2019, 183, 110428. [Google Scholar] [CrossRef] [PubMed]
  17. Nadimi, A.E.; Ebrahimipour, S.Y.; Afshar, E.G.; Falahati-Pour, S.K.; Ahmadi, Z.; Mohammadinejad, R.; Mohamadi, M. Nano-scale drug delivery systems for antiarrhythmic agents. Eur. J. Med. Chem. 2018, 157, 1153–1163. [Google Scholar] [CrossRef] [PubMed]
  18. Jafari, R.; Zolbanin, N.M.; Majidi, J.; Atyabi, F.; Yousefi, M.; Jadidi-Niaragh, F.; Aghebati-Maleki, L.; Shanehbandi, D.; Zangbar, M.-S.S.; Rafatpanah, H. Anti-Mucin1 Aptamer-Conjugated Chitosan Nanoparticles for Targeted Co-Delivery of Docetaxel and IGF-1R siRNA to SKBR3 Metastatic Breast Cancer Cells. Iran. Biomed. J. 2019, 23, 21–33. [Google Scholar] [CrossRef] [Green Version]
  19. Misra, S.K.; De, A.; Pan, D. Targeted delivery of STAT-3 modulator to breast cancer stem-like cells downregulates a series of stemness genes. Mol. Cancer Ther. 2018, 17, 119–129. [Google Scholar] [CrossRef]
  20. Mohammadinejad, R.; Dadashzadeh, A.; Moghassemi, S.; Ashrafizadeh, M.; Dehshahri, A.; Pardakhty, A.; Sassan, H.A.; Sohrevardi, S.M.; Mandegary, A. Shedding light on gene therapy: Carbon dots for the minimally invasive image-guided delivery of plasmids and noncoding RNAs. J. Adv. Res. 2019, 18, 81–93. [Google Scholar] [CrossRef]
  21. Le, D.H.; Commandeur, U.; Steinmetz, N.F. Presentation and Delivery of Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand via Elongated Plant Viral Nanoparticle Enhances Antitumor Efficacy. ACS Nano 2019, 13, 2501–2510. [Google Scholar] [CrossRef] [PubMed]
  22. Sinha, A.; Chakraborty, A.; Jana, N.R. Dextran-gated, multifunctional mesoporous nanoparticle for glucose-responsive and targeted drug delivery. ACS Appl. Mater. Interfaces 2014, 6, 22183–22191. [Google Scholar] [CrossRef] [PubMed]
  23. Peng, C.; Xu, J.; Yu, M.; Ning, X.; Huang, Y.; Du, B.; Hernandez, E.; Kapur, P.; Hsieh, J.T.; Zheng, J. Tuning the In Vivo Transport of Anticancer Drugs Using Renal-Clearable Gold Nanoparticles. Angew. Chem. 2019, 131, 8567–8571. [Google Scholar]
  24. Mu, Q.; Kievit, F.M.; Kant, R.J.; Lin, G.; Jeon, M.; Zhang, M. Anti-HER2/neu peptide-conjugated iron oxide nanoparticles for targeted delivery of paclitaxel to breast cancer cells. Nanoscale 2015, 7, 18010–18014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Davarani, F.-H.; Ashrafizadeh, M.; Riseh, R.S.; Afshar, E.G.; Mohammadi, H.; Razavi, S.H.; Mandegary, A.; Mohammadinejad, R. Antifungal nanoparticles reduce aflatoxin contamination in pistachio. PHJ 2018, 1, 26–33. [Google Scholar]
  26. Saulite, L.; Pleiko, K.; Popena, I.; Dapkute, D.; Rotomskis, R.; Riekstina, U. Nanoparticle delivery to metastatic breast cancer cells by nanoengineered mesenchymal stem cells. Beilstein J. Nanotechnol. 2018, 9, 321–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Feng, T.; Ai, X.; Ong, H.; Zhao, Y. Dual-responsive carbon dots for tumor extracellular microenvironment triggered targeting and enhanced anticancer drug delivery. ACS Appl. Mater. Interfaces 2016, 8, 18732–18740. [Google Scholar] [CrossRef] [PubMed]
  28. Truffi, M.; Colombo, M.; Sorrentino, L.; Pandolfi, L.; Mazzucchelli, S.; Pappalardo, F.; Pacini, C.; Allevi, R.; Bonizzi, A.; Corsi, F. Multivalent exposure of trastuzumab on iron oxide nanoparticles improves antitumor potential and reduces resistance in HER2-positive breast cancer cells. Sci. Rep. 2018, 8, 6563. [Google Scholar] [CrossRef] [PubMed]
  29. Ahmadi, Z.; Roomiani, S.; Bemani, N.; Ashrafizadeh, M. The Targeting of Autophagy and Endoplasmic Reticulum Stress Mechanisms by Honokiol Therapy. Rev. Clin. Med. 2019, 6, 66–73. [Google Scholar]
  30. Daulat, A.M.; Wagner, M.S.; Walton, A.; Baudelet, E.; Audebert, S.; Camoin, L.; Borg, J.P. The Tumor Suppressor SCRIB is a Negative Modulator of the Wnt/beta-Catenin Signaling Pathway. Proteomics 2019, e1800487. [Google Scholar] [CrossRef]
  31. Liu, S.; Wang, Q.; Liu, Y.; Xia, Z.Y. miR-425-5p suppresses tumorigenesis and DDP resistance in human-prostate cancer by targeting GSK3beta and inactivating the Wnt/beta-catenin signaling pathway. J. Biosci. 2019, 44, 102. [Google Scholar] [CrossRef] [PubMed]
  32. Xie, W.; Zhang, Y.; Zhang, S.; Wang, F.; Zhang, K.; Huang, Y.; Zhou, Z.; Huang, G.; Wang, J. Oxymatrine enhanced anti-tumor effects of Bevacizumab against triple-negative breast cancer via abating Wnt/beta-Catenin signaling pathway. Am. J. Cancer Res. 2019, 9, 1796–1814. [Google Scholar] [PubMed]
  33. Ahmadi, Z.; Ashrafizadeh, M. Melatonin as a potential modulator of Nrf2. Fundam. Clin. Pharmacol. 2019. [Google Scholar] [CrossRef] [PubMed]
  34. Woo, C.C.; Hsu, A.; Kumar, A.P.; Sethi, G.; Tan, K.H.B. Thymoquinone inhibits tumor growth and induces apoptosis in a breast cancer xenograft mouse model: The role of p38 MAPK and ROS. PLoS ONE 2013, 8, e75356. [Google Scholar] [CrossRef] [PubMed]
  35. Nair, A.S.; Shishodia, S.; Ahn, K.S.; Kunnumakkara, A.B.; Sethi, G.; Aggarwal, B.B. Deguelin, an Akt inhibitor, suppresses IκBα kinase activation leading to suppression of NF-κB-regulated gene expression, potentiation of apoptosis, and inhibition of cellular invasion. J. Immunol. 2006, 177, 5612–5622. [Google Scholar] [CrossRef] [PubMed]
  36. Chua, A.W.L.; Hay, H.S.; Rajendran, P.; Shanmugam, M.K.; Li, F.; Bist, P.; Koay, E.S.; Lim, L.H.; Kumar, A.P.; Sethi, G. Butein downregulates chemokine receptor CXCR4 expression and function through suppression of NF-κB activation in breast and pancreatic tumor cells. Biochem. Pharmacol. 2010, 80, 1553–1562. [Google Scholar] [CrossRef]
  37. Singh, S.S.; Yap, W.N.; Arfuso, F.; Kar, S.; Wang, C.; Cai, W.; Dharmarajan, A.M.; Sethi, G.; Kumar, A.P. Targeting the PI3K/Akt signaling pathway in gastric carcinoma: A reality for personalized medicine? World J. Gastroenterol. 2015, 21, 12261. [Google Scholar] [CrossRef]
  38. Darnell, J.E. STATs and gene regulation. Science 1997, 277, 1630–1635. [Google Scholar] [CrossRef]
  39. Cui, C.; Cheng, X.; Yan, L.; Ding, H.; Guan, X.; Zhang, W.; Tian, X.; Hao, C. Downregulation of TfR1 promotes progression of colorectal cancer via the JAK/STAT pathway. Cancer Manag. Res. 2019, 11, 6323. [Google Scholar] [CrossRef]
  40. de Haas, N.; de Koning, C.; di Blasio, S.; Flórez-Grau, G.; de Vries, I.J.M.; Hato, S.V. STAT Family Protein Expression and Phosphorylation State during moDC Development Is Altered by Platinum-Based Chemotherapeutics. J. Immunol. Res. 2019, 2019, 7458238. [Google Scholar] [CrossRef]
  41. Jung, Y.; Shanmugam, M.; Narula, A.; Kim, C.; Lee, J.; Namjoshi, O.; Blough, B.; Sethi, G.; Ahn, K. Oxymatrine attenuates tumor growth and deactivates STAT5 signaling in a lung cancer xenograft model. Cancers 2019, 11, 49. [Google Scholar] [CrossRef] [PubMed]
  42. Lau, C.P.; Fung, C.S.; Wong, K.C.; Wang, Y.H.; Huang, L.; Tsui, S.K.; Lee, O.K.; Kumta, S.M. Simvastatin Possesses Antitumor and Differentiation-Promoting Properties that Affect Stromal Cells in Giant Cell Tumor of Bone. J. Orthop. Res. 2019. [Google Scholar] [CrossRef] [PubMed]
  43. Zhu, D.; Yu, Y.; Wang, W.; Wu, K.; Liu, D.; Yang, Y.; Zhang, C.; Qi, Y.; Zhao, S. Long noncoding RNA PART1 promotes progression of non-small cell lung cancer cells via JAK-STAT signaling pathway. Cancer Med. 2019. [Google Scholar] [CrossRef] [PubMed]
  44. Mohrherr, J.; Haber, M.; Breitenecker, K.; Aigner, P.; Moritsch, S.; Voronin, V.; Eferl, R.; Moriggl, R.; Stoiber, D.; Győrffy, B. JAK-STAT inhibition impairs K-RAS-driven lung adenocarcinoma progression. Int. J. Cancer 2019. [Google Scholar] [CrossRef] [PubMed]
  45. Park, A.; Yang, Y.; Lee, Y.; Kim, M.S.; Park, Y.-J.; Jung, H.; Kim, T.-D.; Lee, H.G.; Choi, I.; Yoon, S.R. Indoleamine-2, 3-Dioxygenase in Thyroid Cancer Cells Suppresses Natural Killer Cell Function by Inhibiting NKG2D and NKp46 Expression via STAT Signaling Pathways. J. Clin. Med. 2019, 8, 842. [Google Scholar] [CrossRef] [PubMed]
  46. Herrera, S.C.; Bach, E.A. JAK/STAT signaling in stem cells and regeneration: From Drosophila to vertebrates. Development 2019, 146, dev167643. [Google Scholar] [CrossRef] [PubMed]
  47. Alunno, A.; Padjen, I.; Fanouriakis, A.; Boumpas, D.T. Pathogenic and Therapeutic Relevance of JAK/STAT Signaling in Systemic Lupus Erythematosus: Integration of Distinct Inflammatory Pathways and the Prospect of Their Inhibition with an Oral Agent. Cells 2019, 8, 898. [Google Scholar] [CrossRef] [PubMed]
  48. Moresi, V.; Adamo, S.; Berghella, L. The JAK/STAT Pathway in Skeletal Muscle Pathophysiology. Front. Physiol. 2019, 10, 500. [Google Scholar] [CrossRef] [Green Version]
  49. Yamaoka, K.; Saharinen, P.; Pesu, M.; Holt, V.E.; Silvennoinen, O.; O’Shea, J.J. The janus kinases (jaks). Genome Biol. 2004, 5, 253. [Google Scholar] [CrossRef]
  50. Radtke, S.; Haan, S.; Jörissen, A.; Hermanns, H.M.; Diefenbach, S.; Smyczek, T.; Schmitz-VandeLeur, H.; Heinrich, P.C.; Behrmann, I.; Haan, C. The Jak1 SH2 domain does not fulfill a classical SH2 function in Jak/STAT signaling but plays a structural role for receptor interaction and up-regulation of receptor surface expression. J. Biol. Chem. 2005, 280, 25760–25768. [Google Scholar] [CrossRef]
  51. Zhao, L.; Ma, Y.; Seemann, J.; Huang, L.J.-S. A regulating role of the JAK2 FERM domain in hyperactivation of JAK2 (V617F). Biochem. J. 2010, 426, 91–98. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.; Levy, D.E. Comparative evolutionary genomics of the STAT family of transcription factors. Jak-Stat 2012, 1, 23–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Durham, G.A.; Williams, J.J.; Nasim, M.T.; Palmer, T.M. Targeting SOCS Proteins to Control JAK-STAT Signalling in Disease. Trends Pharmacol. Sci. 2019, 40, 298–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Babon, J.J.; Lucet, I.S.; Murphy, J.M.; Nicola, N.A.; Varghese, L.N. The molecular regulation of Janus kinase (JAK) activation. Biochem. J. 2014, 462, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Vargas-Hernandez, A.; Forbes, L.R. JAK/STAT proteins and their biological impact on NK cell development and function. Mol. Immunol. 2019. [Google Scholar] [CrossRef]
  56. Moshapa, F.T.; Riches-Suman, K.; Palmer, T.M. Therapeutic Targeting of the Proinflammatory IL-6-JAK/STAT Signalling Pathways Responsible for Vascular Restenosis in Type 2 Diabetes Mellitus. Cardiol. Res. Pract. 2019, 2019, 9846312. [Google Scholar] [CrossRef]
  57. Mogensen, T.H. IRF and STAT Transcription Factors—From Basic Biology to Roles in Infection, Protective Immunity, and Primary Immunodeficiencies. Front. Immunol. 2019, 9, 3047. [Google Scholar] [CrossRef] [PubMed]
  58. Trivedi, S.; Starz-Gaiano, M. Drosophila Jak/STAT Signaling: Regulation and Relevance in Human Cancer and Metastasis. Int. J. Mol. Sci. 2018, 19, 4056. [Google Scholar] [CrossRef]
  59. Ko, Y.S.; Rugira, T.; Jin, H.; Park, S.W.; Kim, H.J. Oleandrin and Its Derivative Odoroside A, Both Cardiac Glycosides, Exhibit Anticancer Effects by Inhibiting Invasion via Suppressing the STAT-3 Signaling Pathway. Int. J. Mol. Sci. 2018, 19, 3350. [Google Scholar] [CrossRef] [PubMed]
  60. Stabile, H.; Scarno, G.; Fionda, C.; Gismondi, A.; Santoni, A.; Gadina, M.; Sciume, G. JAK/STAT signaling in regulation of innate lymphoid cells: The gods before the guardians. Immunol. Rev. 2018, 286, 148–159. [Google Scholar] [CrossRef]
  61. Chen, X.; Vinkemeier, U.; Zhao, Y.; Jeruzalmi, D.; Darnell, J.E., Jr.; Kuriyan, J. Crystal structure of a tyrosine phosphorylated STAT-1 dimer bound to DNA. Cell 1998, 93, 827–839. [Google Scholar] [CrossRef]
  62. Pranada, A.L.; Metz, S.; Herrmann, A.; Heinrich, P.C.; Muller-Newen, G. Real time analysis of STAT3 nucleocytoplasmic shuttling. J. Biol. Chem. 2004, 279, 15114–15123. [Google Scholar] [CrossRef] [PubMed]
  63. Rawlings, J.S.; Rosler, K.M.; Harrison, D.A. The JAK/STAT signaling pathway. J. Cell Sci. 2004, 117, 1281–1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zhang, J.-G.; Metcalf, D.; Rakar, S.; Asimakis, M.; Greenhalgh, C.J.; Willson, T.A.; Starr, R.; Nicholson, S.E.; Carter, W.; Alexander, W.S. The SOCS box of suppressor of cytokine signaling-1 is important for inhibition of cytokine action in vivo. Proc. Natl. Acad. Sci. USA 2001, 98, 13261–13265. [Google Scholar] [CrossRef] [PubMed]
  65. Cohney, S.J.; Sanden, D.; Cacalano, N.A.; Yoshimura, A.; Mui, A.; Migone, T.S.; Johnston, J.A. SOCS-3 is tyrosine phosphorylated in response to interleukin-2 and suppresses STAT5 phosphorylation and lymphocyte proliferation. Mol. Cell. Biol. 1999, 19, 4980–4988. [Google Scholar] [CrossRef] [PubMed]
  66. Krebs, D.L.; Hilton, D.J. SOCS proteins: Negative regulators of cytokine signaling. Stem Cells 2001, 19, 378–387. [Google Scholar] [CrossRef] [PubMed]
  67. Buttarelli, M.; Babini, G.; Raspaglio, G.; Filippetti, F.; Battaglia, A.; Ciucci, A.; Ferrandina, G.; Petrillo, M.; Marino, C.; Mancuso, M. A combined ANXA2-NDRG1-STAT1 gene signature predicts response to chemoradiotherapy in cervical cancer. J. Exp. Clin. Cancer Res. 2019, 38, 279. [Google Scholar] [CrossRef] [PubMed]
  68. Yu, L.; Ye, F.; Li, Y.-Y.; Zhan, Y.-Z.; Liu, Y.; Yan, H.-M.; Fang, Y.; Xie, Y.-W.; Zhang, F.-J.; Chen, L.-H. Histone methyltransferase SETDB1 promotes colorectal cancer proliferation through the STAT1-CCND1/CDK6 axis. Carcinogenesis 2019. [Google Scholar] [CrossRef]
  69. Jiang, L.; Liu, J.-Y.; Shi, Y.; Tang, B.; He, T.; Liu, J.-J.; Fan, J.-Y.; Wu, B.; Xu, X.-H.; Zhao, Y.-L.; et al. MTMR2 promotes invasion and metastasis of gastric cancer via inactivating IFNγ/STAT1 signaling. J. Exp. Clin. Cancer Res. 2019, 38, 206. [Google Scholar] [CrossRef]
  70. Gamero, A.M.; Young, M.R.; Mentor-Marcel, R.; Bobe, G.; Scarzello, A.J.; Wise, J.; Colburn, N.H. STAT2 contributes to promotion of colorectal and skin carcinogenesis. Cancer Prev. Res. 2010, 3, 495–504. [Google Scholar] [CrossRef]
  71. Dhir, R.; Ni, Z.; Lou, W.; DeMiguel, F.; Grandis, J.R.; Gao, A.C. Stat3 activation in prostatic carcinomas. Prostate 2002, 51, 241–246. [Google Scholar] [CrossRef] [PubMed]
  72. Hwang, S.T.; Kim, C.; Lee, J.H.; Chinnathambi, A.; Alharbi, S.A.; Shair, O.H.; Sethi, G.; Ahn, K.S. Cycloastragenol can negate constitutive STAT3 activation and promote paclitaxel-induced apoptosis in human gastric cancer cells. Phytomedicine 2019, 59, 152907. [Google Scholar] [CrossRef] [PubMed]
  73. Lee, J.; Kim, C.; Lee, S.-G.; Sethi, G.; Ahn, K. Ophiopogonin d, a steroidal glycoside abrogates STAT3 signaling cascade and exhibits anti-cancer activity by causing GSH/GSSG imbalance in lung carcinoma. Cancers 2018, 10, 427. [Google Scholar] [CrossRef] [PubMed]
  74. Lee, J.H.; Kim, C.; Lee, J.; Um, J.-Y.; Sethi, G.; Ahn, K.S. Arctiin is a pharmacological inhibitor of STAT3 phosphorylation at tyrosine 705 residue and potentiates bortezomib-induced apoptotic and angiogenic effects in human multiple myeloma cells. Phytomedicine 2019, 55, 282–292. [Google Scholar] [CrossRef] [PubMed]
  75. Loh, C.-Y.; Arya, A.; Naema, A.F.; Wong, W.F.; Sethi, G.; Looi, C.Y. Signal transducer and activator of transcription (stats) proteins in cancer and inflammation: Functions and therapeutic implication. Front. Oncol. 2019, 9, 48. [Google Scholar] [CrossRef]
  76. Pan, S.; Deng, Y.; Fu, J.; Zhang, Y.; Zhang, Z.; Ru, X.; Qin, X. TRIM52 promotes colorectal cancer cell proliferation through the STAT3 signaling. Cancer Cell Int. 2019, 19, 57. [Google Scholar] [CrossRef]
  77. Yu, Z.Y.; Huang, R.; Xiao, H.; Sun, W.F.; Shan, Y.J.; Wang, B.; Zhao, T.T.; Dong, B.; Zhao, Z.H.; Liu, X.L. Fluacrypyrim, a novel STAT3 activation inhibitor, induces cell cycle arrest and apoptosis in cancer cells harboring constitutively-active STAT3. Int. J. Cancer 2010, 127, 1259–1270. [Google Scholar] [CrossRef] [PubMed]
  78. Nishi, M.; Batsaikhan, B.-E.; Yoshikawa, K.; Higashijima, J.; Tokunaga, T.; Takasu, C.; Kashihara, H.; Ishikawa, D.; Shimada, M. High STAT4 Expression Indicates Better Disease-free Survival in Patients with Gastric Cancer. Anticancer Res. 2017, 37, 6723–6729. [Google Scholar]
  79. Zhang, Y.; Yu, C. Prognostic values of signal transducers activators of transcription in gastric cancer. Biosci. Rep. 2019, 39, BSR20181695. [Google Scholar] [CrossRef]
  80. Zhou, Y.; Zhong, J.-H.; Gong, F.-S.; Xiao, J. MiR-141-3p suppresses gastric cancer induced transition of normal fibroblast and BMSC to cancer-associated fibroblasts via targeting STAT4. Exp. Mol. Pathol. 2019, 107, 85–94. [Google Scholar] [CrossRef]
  81. Chen, J.; Gong, C.; Mao, H.; Li, Z.; Fang, Z.; Chen, Q.; Lin, M.; Jiang, X.; Hu, Y.; Wang, W. E2F1/SP3/STAT6 axis is required for IL-4-induced epithelial-mesenchymal transition of colorectal cancer cells. Int. J. Oncol. 2018, 53, 567–578. [Google Scholar] [CrossRef] [PubMed]
  82. Miklossy, G.; Hilliard, T.S.; Turkson, J. Therapeutic modulators of STAT signalling for human diseases. Nat. Rev. Drug Discov. 2013, 12, 611. [Google Scholar] [CrossRef] [PubMed]
  83. Walker, S.; Xiang, M.; Frank, D. STAT3 Activity and Function in Cancer: Modulation by STAT5 and miR-146b. Cancers 2014, 6, 958–968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Wingelhofer, B.; Neubauer, H.A.; Valent, P.; Han, X.; Constantinescu, S.N.; Gunning, P.T.; Müller, M.; Moriggl, R. Implications of STAT3 and STAT5 signaling on gene regulation and chromatin remodeling in hematopoietic cancer. Leukemia 2018, 32, 1713. [Google Scholar] [CrossRef] [PubMed]
  85. Tabassum, S.; Abbasi, R.; Ahmad, N.; Farooqi, A.A. Targeting of JAK-STAT Signaling in Breast Cancer: Therapeutic Strategies to Overcome Drug Resistance. Adv. Exp. Med. Biol. 2019, 1152, 271–281. [Google Scholar] [CrossRef] [PubMed]
  86. Drost, J.; Van Jaarsveld, R.H.; Ponsioen, B.; Zimberlin, C.; Van Boxtel, R.; Buijs, A.; Sachs, N.; Overmeer, R.M.; Offerhaus, G.J.; Begthel, H. Sequential cancer mutations in cultured human intestinal stem cells. Nature 2015, 521, 43. [Google Scholar] [CrossRef] [PubMed]
  87. Matano, M.; Date, S.; Shimokawa, M.; Takano, A.; Fujii, M.; Ohta, Y.; Watanabe, T.; Kanai, T.; Sato, T. Modeling colorectal cancer using CRISPR-Cas9–mediated engineering of human intestinal organoids. Nat. Med. 2015, 21, 256. [Google Scholar] [CrossRef] [PubMed]
  88. Goel, S.; Huang, J.; Klampfer, L. K-Ras intestinal homeostasis and colon cancer. Curr. Clin. Pharmacol. 2015, 10, 73–81. [Google Scholar] [CrossRef] [PubMed]
  89. Sakahara, M.; Okamoto, T.; Oyanagi, J.; Takano, H.; Natsume, Y.; Yamanaka, H.; Kusama, D.; Fusejima, M.; Tanaka, N.; Mori, S.; et al. IFN/STAT signaling controls tumorigenesis and the drug response in colorectal cancer. Cancer Sci. 2019, 110, 1293. [Google Scholar] [CrossRef]
  90. Dolatabadi, S.; Jonasson, E.; Lindén, M.; Fereydouni, B.; Bäcksten, K.; Nilsson, M.; Martner, A.; Forootan, A.; Fagman, H.; Landberg, G. JAK–STAT signalling controls cancer stem cell properties including chemotherapy resistance in myxoid liposarcoma. Int. J. Cancer 2019, 145, 435–449. [Google Scholar] [CrossRef]
  91. Soleimani, A.; Khazaei, M.; Ferns, G.A.; Ryzhikov, M.; Avan, A.; Hassanian, S.M. Role of TGF-β signaling regulatory microRNAs in the pathogenesis of colorectal cancer. J. Cell. Physiol. 2019, 234, 14574–14580. [Google Scholar] [CrossRef] [PubMed]
  92. Paseban, M.; Marjaneh, R.M.; Banach, M.; Riahi, M.M.; Bo, S.; Sahebkar, A. Modulation of microRNAs by aspirin in cardiovascular disease. Trends Cardiovasc. Med. 2019. [Google Scholar] [CrossRef] [PubMed]
  93. Tajbakhsh, A.; Bianconi, V.; Pirro, M.; Hayat, S.M.G.; Johnston, T.P.; Sahebkar, A. Efferocytosis and Atherosclerosis: Regulation of Phagocyte Function by MicroRNAs. Trends Endocrinol. Metab. 2019, 30, 672–683. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, S.; Zhang, S.; He, Y.; Huang, X.; Hui, Y.; Tang, Y. HOXA11-AS regulates JAK-STAT pathway by miR-15a-3p/STAT3 axis to promote the growth and metastasis in liver cancer. J. Cell. Biochem. 2019, 120, 15941–15951. [Google Scholar] [CrossRef] [PubMed]
  95. Alston, C.I.; Dix, R.D. SOCS and Herpesviruses, With Emphasis on Cytomegalovirus Retinitis. Front. Immunol. 2019, 10, 732. [Google Scholar] [CrossRef] [PubMed]
  96. Li, M.; Zheng, R.; Yuan, F.L. MiR-410 affects the proliferation and apoptosis of lung cancer A549 cells through regulation of SOCS3/JAK-STAT signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 5987–5993. [Google Scholar] [CrossRef] [PubMed]
  97. Sen, M.; Thomas, S.M.; Kim, S.; Yeh, J.I.; Ferris, R.L.; Johnson, J.T.; Duvvuri, U.; Lee, J.; Sahu, N.; Joyce, S. First-in-human trial of a STAT3 decoy oligonucleotide in head and neck tumors: Implications for cancer therapy. Cancer Discov. 2012, 2, 694–705. [Google Scholar] [CrossRef]
  98. Aggarwal, B.B.; Sethi, G.; Baladandayuthapani, V.; Krishnan, S.; Shishodia, S. Targeting cell signaling pathways for drug discovery: An old lock needs a new key. J. Cell. Biochem. 2007, 102, 580–592. [Google Scholar] [CrossRef]
  99. Aggarwal, V.; Kashyap, D.; Sak, K.; Tuli, H.S.; Jain, A.; Chaudhary, A.; Garg, V.K.; Sethi, G.; Yerer, M.B. Molecular mechanisms of action of tocotrienols in cancer: Recent trends and advancements. Int. J. Mol. Sci. 2019, 20, 656. [Google Scholar] [CrossRef]
  100. Bishayee, A.; Sethi, G. Bioactive natural products in cancer prevention and therapy: Progress and promise. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2016; pp. 1–3. [Google Scholar]
  101. Chai, E.Z.P.; Shanmugam, M.K.; Arfuso, F.; Dharmarajan, A.; Wang, C.; Kumar, A.P.; Samy, R.P.; Lim, L.H.; Wang, L.; Goh, B.C. Targeting transcription factor STAT3 for cancer prevention and therapy. Pharmacol. Ther. 2016, 162, 86–97. [Google Scholar] [CrossRef]
  102. Jung, Y.Y.; Lee, J.H.; Nam, D.; Narula, A.S.; Namjoshi, O.A.; Blough, B.E.; Um, J.-Y.; Sethi, G.; Ahn, K.S. Anti-myeloma effects of icariin are mediated through the attenuation of jak/stat3-dependent signaling cascade. Front. Pharmacol. 2018, 9, 531. [Google Scholar] [CrossRef] [PubMed]
  103. Kim, C.; Lee, S.-G.; Yang, W.M.; Arfuso, F.; Um, J.-Y.; Kumar, A.P.; Bian, J.; Sethi, G.; Ahn, K.S. Formononetin-induced oxidative stress abrogates the activation of STAT3/5 signaling axis and suppresses the tumor growth in multiple myeloma preclinical model. Cancer Lett. 2018, 431, 123–141. [Google Scholar] [CrossRef] [PubMed]
  104. Baek, S.; Lee, J.; Kim, C.; Ko, J.-H.; Ryu, S.-H.; Lee, S.-G.; Yang, W.; Um, J.-Y.; Chinnathambi, A.; Alharbi, S. Ginkgolic acid C 17: 1, derived from Ginkgo biloba leaves, suppresses constitutive and inducible STAT3 activation through induction of PTEN and SHP-1 tyrosine phosphatase. Molecules 2017, 22, 276. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, J.; Ahn, K.S.; Kim, C.; Shanmugam, M.K.; Siveen, K.S.; Arfuso, F.; Samym, R.P.; Deivasigamanim, A.; Lim, L.H.K.; Wang, L. Nimbolide-induced oxidative stress abrogates STAT3 signaling cascade and inhibits tumor growth in transgenic adenocarcinoma of mouse prostate model. Antioxid. Redox Signal. 2016, 24, 575–589. [Google Scholar] [CrossRef] [PubMed]
  106. Baek, S.H.; Ko, J.-H.; Lee, H.; Jung, J.; Kong, M.; Lee, J.-W.; Lee, J.; Chinnathambi, A.; Zayed, M.; Alharbi, S.A. Resveratrol inhibits STAT3 signaling pathway through the induction of SOCS-1: Role in apoptosis induction and radiosensitization in head and neck tumor cells. Phytomedicine 2016, 23, 566–577. [Google Scholar] [CrossRef] [PubMed]
  107. Deorukhkar, A.; Krishnan, S.; Sethi, G.; Aggarwal, B.B. Back to basics: How natural products can provide the basis for new therapeutics. Expert Opin. Investig. Drugs 2007, 16, 1753–1773. [Google Scholar] [CrossRef] [PubMed]
  108. Sethi, G.; Shanmugam, M.; Warrier, S.; Merarchi, M.; Arfuso, F.; Kumar, A.; Bishayee, A. Pro-apoptotic and anti-cancer properties of diosgenin: A comprehensive and critical review. Nutrients 2018, 10, 645. [Google Scholar] [CrossRef] [PubMed]
  109. Shanmugam, M.K.; Kannaiyan, R.; Sethi, G. Targeting cell signaling and apoptotic pathways by dietary agents: Role in the prevention and treatment of cancer. Nutr. Cancer 2011, 63, 161–173. [Google Scholar] [CrossRef] [PubMed]
  110. Shanmugam, M.K.; Lee, J.H.; Chai, E.Z.P.; Kanchi, M.M.; Kar, S.; Arfuso, F.; Dharmarajan, A.; Kumar, A.P.; Ramar, P.S.; Looi, C.Y. Cancer prevention and therapy through the modulation of transcription factors by bioactive natural compounds. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA; pp. 35–47.
  111. Shanmugam, M.K.; Nguyen, A.H.; Kumar, A.P.; Tan, B.K.; Sethi, G. Targeted inhibition of tumor proliferation, survival, and metastasis by pentacyclic triterpenoids: Potential role in prevention and therapy of cancer. Cancer Lett. 2012, 320, 158–170. [Google Scholar] [CrossRef] [Green Version]
  112. Shrimali, D.; Shanmugam, M.K.; Kumar, A.P.; Zhang, J.; Tan, B.K.; Ahn, K.S.; Sethi, G. Targeted abrogation of diverse signal transduction cascades by emodin for the treatment of inflammatory disorders and cancer. Cancer Lett. 2013, 341, 139–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Zhang, J.; Sikka, S.; Siveen, K.S.; Lee, J.H.; Um, J.-Y.; Kumar, A.P.; Chinnathambi, A.; Alharbi, S.A.; Rangappa, K.S.; Sethi, G. Cardamonin represses proliferation, invasion, and causes apoptosis through the modulation of signal transducer and activator of transcription 3 pathway in prostate cancer. Apoptosis 2017, 22, 158–168. [Google Scholar] [CrossRef] [PubMed]
  114. Liu, J.; Qu, L.; Meng, L.; Shou, C. Topoisomerase inhibitors promote cancer cell motility via ROS-mediated activation of JAK2-STAT1-CXCL1 pathway. J. Exp. Clin. Cancer Res. CR 2019, 38, 370. [Google Scholar] [CrossRef] [PubMed]
  115. Park, S.Y.; Lee, C.J.; Choi, J.H.; Kim, J.H.; Kim, J.W.; Kim, J.Y.; Nam, J.S. The JAK2/STAT3/CCND2 Axis promotes colorectal Cancer stem cell persistence and radioresistance. J. Exp. Clin. Cancer Res. CR 2019, 38, 399. [Google Scholar] [CrossRef] [PubMed]
  116. Wong, A.L.; Hirpara, J.L.; Pervaiz, S.; Eu, J.-Q.; Sethi, G.; Goh, B.-C. Do STAT3 inhibitors have potential in the future for cancer therapy? Expert Opin. Investig. Drugs 2017, 26, 883–887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Furqan, M.; Akinleye, A.; Mukhi, N.; Mittal, V.; Chen, Y.; Liu, D. STAT inhibitors for cancer therapy. J. Hematol. Oncol. 2013, 6, 90. [Google Scholar] [CrossRef] [PubMed]
  118. Barati, N.; Momtazi-Borojeni, A.A.; Majeed, M.; Sahebkar, A. Potential therapeutic effects of curcumin in gastric cancer. J. Cell. Physiol. 2019, 234, 2317–2328. [Google Scholar] [CrossRef] [PubMed]
  119. Keihanian, F.; Saeidinia, A.; Bagheri, R.K.; Johnston, T.P.; Sahebkar, A. Curcumin, hemostasis, thrombosis, and coagulation. J. Cell. Physiol. 2018, 233, 4497–4511. [Google Scholar] [CrossRef]
  120. Tabeshpour, J.; Hashemzaei, M.; Sahebkar, A. The regulatory role of curcumin on platelet functions. J. Cell. Biochem. 2018, 119, 8713–8722. [Google Scholar] [CrossRef]
  121. Zendedel, E.; Butler, A.E.; Atkin, S.L.; Sahebkar, A. Impact of curcumin on sirtuins: A review. J. Cell. Biochem. 2018, 119, 10291–10300. [Google Scholar] [CrossRef]
  122. Fossey, S.L.; Bear, M.D.; Lin, J.; Li, C.; Schwartz, E.B.; Li, P.-K.; Fuchs, J.R.; Fenger, J.; Kisseberth, W.C.; London, C.A. The novel curcumin analog FLLL32 decreases STAT3 DNA binding activity and expression, and induces apoptosis in osteosarcoma cell lines. BMC Cancer 2011, 11, 112. [Google Scholar] [CrossRef]
  123. Saydmohammed, M.; Joseph, D.; Syed, V. Curcumin suppresses constitutive activation of STAT-3 by up-regulating protein inhibitor of activated STAT-3 (PIAS-3) in ovarian and endometrial cancer cells. J. Cell. Biochem. 2010, 110, 447–456. [Google Scholar] [CrossRef] [PubMed]
  124. Strimpakos, A.S.; Sharma, R.A. Curcumin: Preventive and therapeutic properties in laboratory studies and clinical trials. Antioxid. Redox Signal. 2008, 10, 511–546. [Google Scholar] [CrossRef] [PubMed]
  125. Weissenberger, J.; Priester, M.; Bernreuther, C.; Rakel, S.; Glatzel, M.; Seifert, V.; Kögel, D. Dietary curcumin attenuates glioma growth in a syngeneic mouse model by inhibition of the JAK1, 2/STAT3 signaling pathway. Clin. Cancer Res. 2010, 16, 5781–5795. [Google Scholar] [CrossRef] [PubMed]
  126. Hahn, Y.-I.; Kim, S.-J.; Choi, B.-Y.; Cho, K.-C.; Bandu, R.; Kim, K.P.; Kim, D.-H.; Kim, W.; Park, J.S.; Han, B.W. Curcumin interacts directly with the Cysteine 259 residue of STAT3 and induces apoptosis in H-Ras transformed human mammary epithelial cells. Sci. Rep. 2018, 8, 6409. [Google Scholar] [CrossRef] [PubMed]
  127. Ashrafizadeh, M.; Ahmadi, Z. Effects of Statins on Gut Microbiota (Microbiome). Rev. Clin. Med. 2019, 6, 55–59. [Google Scholar]
  128. Zhou, X.X.; Gao, P.J.; Sun, B.G. Pravastatin attenuates interferon-γ action via modulation of stat1 to prevent aortic atherosclerosis in apolipoprotein e-knockout mice. Clin. Exp. Pharmacol. Physiol. 2009, 36, 373–379. [Google Scholar] [CrossRef] [PubMed]
  129. Levy, D.E.; Darnell, J., Jr. Signalling: Stats: Transcriptional control and biological impact. Nat. Rev. Mol. Cell Biol. 2002, 3, 651. [Google Scholar] [CrossRef]
  130. Rondanin, R.; Simoni, D.; Maccesi, M.; Romagnoli, R.; Grimaudo, S.; Pipitone, R.M.; Meli, M.; Cascio, A.; Tolomeo, M. Effects of pimozide derivatives on pSTAT5 in K562 cells. ChemMedChem 2017, 12, 1183–1190. [Google Scholar] [CrossRef]
  131. Siemasko, K.; Chong, A.S.; Jäck, H.-M.; Gong, H.; Williams, J.W.; Finnegan, A. Inhibition of JAK3 and STAT6 tyrosine phosphorylation by the immunosuppressive drug leflunomide leads to a block in IgG1 production. J. Immunol. 1998, 160, 1581–1588. [Google Scholar]
  132. Bousquet, J.; Chanez, P.; Lacoste, J.Y.; Barneon, G.; Ghavanian, N.; Enander, I.; Venge, P.; Ahlstedt, S.; Simony-Lafontaine, J.; Godard, P.; et al. Eosinophilic inflammation in asthma. N. Engl. J. Med. 1990, 323, 1033–1039. [Google Scholar] [CrossRef]
  133. Grünig, G.; Warnock, M.; Wakil, A.E.; Venkayya, R.; Brombacher, F.; Rennick, D.M.; Sheppard, D.; Mohrs, M.; Donaldson, D.D.; Locksley, R.M. Requirement for IL-13 independently of IL-4 in experimental asthma. Science 1998, 282, 2261–2263. [Google Scholar] [CrossRef] [PubMed]
  134. Roche, W.; Williams, J.; Beasley, R.; Holgate, S. Subepithelial fibrosis in the bronchi of asthmatics. Lancet 1989, 333, 520–524. [Google Scholar] [CrossRef]
  135. Wills-Karp, M.; Luyimbazi, J.; Xu, X.; Schofield, B.; Neben, T.Y.; Karp, C.L.; Donaldson, D.D. Interleukin-13: Central mediator of allergic asthma. Science 1998, 282, 2258–2261. [Google Scholar] [CrossRef] [PubMed]
  136. Nakano, T.; Inoue, H.; Fukuyama, S.; Matsumoto, K.; Matsumura, M.; Tsuda, M.; Matsumoto, T.; Aizawa, H.; Nakanishi, Y. Niflumic acid suppresses interleukin-13–induced asthma phenotypes. Am. J. Respir. Crit. Care Med. 2006, 173, 1216–1221. [Google Scholar] [CrossRef] [PubMed]
  137. Sadeghi, S.; Davoodvandi, A.; Pourhanifeh, M.H.; Sharifi, N.; ArefNezhad, R.; Sahebnasagh, R.; Moghadam, S.A.; Sahebkar, A.; Mirzaei, H. Anti-cancer effects of cinnamon: Insights into its apoptosis effects. Eur. J. Med. Chem. 2019, 178, 131–140. [Google Scholar] [CrossRef]
  138. Lee, B.-J.; Kim, Y.-J.; Cho, D.-H.; Sohn, N.-W.; Kang, H. Immunomodulatory effect of water extract of cinnamon on anti-CD3-induced cytokine responses and p38, JNK, ERK1/2, and STAT4 activation. Immunopharmacol. Immunotoxicol. 2011, 33, 714–722. [Google Scholar] [CrossRef] [PubMed]
  139. Romagnoli, R.; Baraldi, P.G.; Prencipe, F.; Lopez-Cara, C.; Rondanin, R.; Simoni, D.; Hamel, E.; Grimaudo, S.; Pipitone, R.M.; Meli, M. Novel iodoacetamido benzoheterocyclic derivatives with potent antileukemic activity are inhibitors of STAT5 phosphorylation. Eur. J. Med. Chem. 2016, 108, 39–52. [Google Scholar] [CrossRef]
  140. Schafranek, L.; Nievergall, E.; Powell, J.; Hiwase, D.; Leclercq, T.; Hughes, T.; White, D. Sustained inhibition of STAT5, but not JAK2, is essential for TKI-induced cell death in chronic myeloid leukemia. Leukemia 2015, 29, 76–85. [Google Scholar] [CrossRef] [PubMed]
  141. Rzymski, T.; Mikula, M.; Żyłkiewicz, E.; Dreas, A.; Wiklik, K.; Gołas, A.; Wójcik, K.; Masiejczyk, M.; Wróbel, A.; Dolata, I. SEL120-34A is a novel CDK8 inhibitor active in AML cells with high levels of serine phosphorylation of STAT1 and STAT5 transactivation domains. Oncotarget 2017, 8, 33779–33795. [Google Scholar] [CrossRef]
  142. Peter, B.; Bibi, S.; Eisenwort, G.; Wingelhofer, B.; Berger, D.; Stefanzl, G.; Blatt, K.; Herrmann, H.; Hadzijusufovic, E.; Hoermann, G. Drug-induced inhibition of phosphorylation of STAT5 overrides drug resistance in neoplastic mast cells. Leukemia 2018, 32, 1016–1022. [Google Scholar] [CrossRef]
  143. Simpson, H.M.; Furusawa, A.; Sadashivaiah, K.; Civin, C.I.; Banerjee, A. STAT5 inhibition induces TRAIL/DR4 dependent apoptosis in peripheral T-cell lymphoma. Oncotarget 2018, 9, 16792–16806. [Google Scholar] [CrossRef] [PubMed]
  144. Faderl, S.; Ferrajoli, A.; Harris, D.; Van, Q.; Kantarjian, H.M.; Estrov, Z. Atiprimod blocks phosphorylation of JAK-STAT and inhibits proliferation of acute myeloid leukemia (AML) cells. Leuk. Res. 2007, 31, 91–95. [Google Scholar] [CrossRef] [PubMed]
  145. Shi, Z.; Zhou, Q.; Gao, S.; Li, W.; Li, X.; Liu, Z.; Jin, P.; Jiang, J. Silibinin inhibits endometrial carcinoma via blocking pathways of STAT3 activation and SREBP1-mediated lipid accumulation. Life Sci. 2019, 217, 70–80. [Google Scholar] [CrossRef] [PubMed]
  146. Granato, M.; Rizzello, C.; Montani, M.S.G.; Cuomo, L.; Vitillo, M.; Santarelli, R.; Gonnella, R.; D’Orazi, G.; Faggioni, A.; Cirone, M. Quercetin induces apoptosis and autophagy in primary effusion lymphoma cells by inhibiting PI3K/AKT/mTOR and STAT3 signaling pathways. J. Nutr. Biochem. 2017, 41, 124–136. [Google Scholar] [CrossRef] [PubMed]
  147. Sun, S.; Zhang, X.; Xu, M.; Zhang, F.; Tian, F.; Cui, J.; Xia, Y.; Liang, C.; Zhou, S.; Wei, H. Berberine downregulates CDC6 and inhibits proliferation via targeting JAK-STAT3 signaling in keratinocytes. Cell Death Dis. 2019, 10, 274. [Google Scholar] [CrossRef] [PubMed]
  148. Oz, B.; Yildirim, A.; Yolbas, S.; Celik, Z.B.; Etem, E.O.; Deniz, G.; Akin, M.; Akar, Z.A.; Karatas, A.; Koca, S.S. Resveratrol inhibits Src tyrosine kinase, STAT3, and Wnt signaling pathway in collagen induced arthritis model. BioFactors 2019, 45, 69–74. [Google Scholar] [CrossRef]
  149. Su, D.; Gao, Y.-Q.; Dai, W.-B.; Hu, Y.; Wu, Y.-F.; Mei, Q.-X. Helicteric acid, oleanic acid, and betulinic acid, three triterpenes from Helicteres angustifolia L., inhibit proliferation and induce apoptosis in HT-29 colorectal cancer cells via suppressing NF-κB and STAT3 Signaling. Evid. Based Complement. Altern. Med. 2017, 2017, 5180707. [Google Scholar] [CrossRef] [PubMed]
  150. Pandey, M.K.; Sung, B.; Ahn, K.S.; Aggarwal, B.B. Butein suppresses constitutive and inducible signal transducer and activator of transcription (STAT) 3 activation and STAT3-regulated gene products through the induction of a protein tyrosine phosphatase SHP-1. Mol. Pharmacol. 2009, 75, 525–533. [Google Scholar] [CrossRef] [PubMed]
  151. Agilan, B.; Rajendra Prasad, N.; Kanimozhi, G.; Karthikeyan, R.; Ganesan, M.; Mohana, S.; Velmurugan, D.; Ananthakrishnan, D. Caffeic Acid Inhibits Chronic UVB-Induced Cellular Proliferation Through JAK-STAT 3 Signaling in Mouse Skin. Photochem. Photobiol. 2016, 92, 467–474. [Google Scholar] [CrossRef]
  152. Jung, J.E.; Kim, H.S.; Lee, C.S.; Park, D.-H.; Kim, Y.-N.; Lee, M.-J.; Lee, J.W.; Park, J.-W.; Kim, M.-S.; Ye, S.K. Caffeic acid and its synthetic derivative CADPE suppress tumor angiogenesis by blocking STAT3-mediated VEGF expression in human renal carcinoma cells. Carcinogenesis 2007, 28, 1780–1787. [Google Scholar] [CrossRef]
  153. Bhutani, M.; Pathak, A.K.; Nair, A.S.; Kunnumakkara, A.B.; Guha, S.; Sethi, G.; Aggarwal, B.B. Capsaicin is a novel blocker of constitutive and interleukin-6–inducible STAT3 activation. Clin. Cancer Res. 2007, 13, 3024–3032. [Google Scholar] [CrossRef] [PubMed]
  154. Rajendran, P.; Li, F.; Shanmugam, M.K.; Kannaiyan, R.; Goh, J.N.; Wong, K.F.; Wang, W.; Khin, E.; Tergaonkar, V.; Kumar, A.P. Celastrol suppresses growth and induces apoptosis of human hepatocellular carcinoma through the modulation of STAT3/JAK2 signaling cascade in vitro and in vivo. Cancer Prev. Res. 2012, 5, 631–643. [Google Scholar] [CrossRef] [PubMed]
  155. Ma, W.; Xiang, Y.; Yang, R.; Zhang, T.; Xu, J.; Wu, Y.; Liu, X.; Xiang, K.; Zhao, H.; Liu, Y. Cucurbitacin B induces inhibitory effects via the CIP2A/PP2A/C-KIT signaling axis in t (8; 21) acute myeloid leukemia. J. Pharmacol. Sci. 2019, 139, 304–310. [Google Scholar] [CrossRef] [PubMed]
  156. Li, F.; Fernandez, P.P.; Rajendran, P.; Hui, K.M.; Sethi, G. Diosgenin, a steroidal saponin, inhibits STAT3 signaling pathway leading to suppression of proliferation and chemosensitization of human hepatocellular carcinoma cells. Cancer Lett. 2010, 292, 197–207. [Google Scholar] [CrossRef] [PubMed]
  157. Ahn, K.S.; Sethi, G.; Sung, B.; Goel, A.; Ralhan, R.; Aggarwal, B.B. Guggulsterone, a farnesoid X receptor antagonist, inhibits constitutive and inducible STAT3 activation through induction of a protein tyrosine phosphatase SHP-1. Cancer Res. 2008, 68, 4406–4415. [Google Scholar] [CrossRef] [PubMed]
  158. Sengupta, S.; Nagalingam, A.; Muniraj, N.; Bonner, M.; Mistriotis, P.; Afthinos, A.; Kuppusamy, P.; Lanoue, D.; Cho, S.; Korangath, P. Activation of tumor suppressor LKB1 by honokiol abrogates cancer stem-like phenotype in breast cancer via inhibition of oncogenic Stat3. Oncogene 2017, 36, 5709. [Google Scholar] [CrossRef] [PubMed]
  159. Haridas, V.; Nishimura, G.; Xu, Z.-X.; Connolly, F.; Hanausek, M.; Walaszek, Z.; Zoltaszek, R.; Gutterman, J.U. Avicin D: A protein reactive plant isoprenoid dephosphorylates Stat 3 by regulating both kinase and phosphatase activities. PLoS ONE 2009, 4, e5578. [Google Scholar] [CrossRef] [PubMed]
  160. Song, H.; Jung, J.I.; Cho, H.J.; Her, S.; Kwon, S.-H.; Yu, R.; Kang, Y.-H.; Lee, K.W.; Park, J.H.Y. Inhibition of tumor progression by oral piceatannol in mouse 4T1 mammary cancer is associated with decreased angiogenesis and macrophage infiltration. J. Nutr. Biochem. 2015, 26, 1368–1378. [Google Scholar] [CrossRef]
  161. Tahara, T.; Streit, U.; Pelish, H.E.; Shair, M.D. STAT3 inhibitory activity of structurally simplified withaferin A analogues. Org. Lett. 2017, 19, 1538–1541. [Google Scholar] [CrossRef]
  162. Subramaniam, A.; Shanmugam, M.K.; Ong, T.H.; Li, F.; Perumal, E.; Chen, L.; Vali, S.; Abbasi, T.; Kapoor, S.; Ahn, K.S. Emodin inhibits growth and induces apoptosis in an orthotopic hepatocellular carcinoma model by blocking activation of STAT3. Br. J. Pharmacol. 2013, 170, 807–821. [Google Scholar] [CrossRef] [Green Version]
  163. Vogel, E.; Santos, D.; Mingels, L.; Verdonckt, T.-W.; Broeck, J.V. RNA interference in insects: Protecting beneficials and controlling pests. Front. Physiol. 2018, 9, 1912. [Google Scholar] [CrossRef] [PubMed]
  164. Siomi, H.; Siomi, M.C. On the road to reading the RNA-interference code. Nature 2009, 457, 396. [Google Scholar] [CrossRef] [PubMed]
  165. Linder, B.; Weirauch, U.; Ewe, A.; Uhmann, A.; Seifert, V.; Mittelbronn, M.; Harter, P.N.; Aigner, A.; Kögel, D. Therapeutic Targeting of Stat3 Using Lipopolyplex Nanoparticle-Formulated siRNA in a Syngeneic Orthotopic Mouse Glioma Model. Cancers 2019, 11, 333. [Google Scholar] [CrossRef] [PubMed]
  166. Gao, L.-F.; Wen, L.-J.; Yu, H.; Zhang, L.; Meng, Y.; Shao, Y.-T.; Xu, D.-Q.; Zhao, X.-J. Knockdown of Stat3 expression using RNAi inhibits growth of laryngeal tumors in vivo. Acta Pharmacol. Sin. 2006, 27, 347. [Google Scholar] [CrossRef]
  167. Gao, Z.; Huang, C.; Qiu, Z.; Jiang, T.; Zhu, L.; Cao, J.; Zhang, F.; Huang, K. Effect of RNAi-mediated STAT3 gene inhibition on metastasis of human pancreatic cancer cells. Zhonghua Wai Ke Za Zhi 2008, 46, 1010–1013. [Google Scholar] [PubMed]
  168. Kaymaz, B.T.; Selvi, N.; Gündüz, C.; Aktan, Ç.; Dalmızrak, A.; Saydam, G.; Kosova, B. Repression of STAT3, STAT5A, and STAT5B expressions in chronic myelogenous leukemia cell line K–562 with unmodified or chemically modified siRNAs and induction of apoptosis. Anna. Hematol. 2013, 92, 151–162. [Google Scholar] [CrossRef] [PubMed]
  169. Konnikova, L.; Kotecki, M.; Kruger, M.M.; Cochran, B.H. Knockdown of STAT3 expression by RNAi induces apoptosis in astrocytoma cells. BMC Cancer 2003, 3, 23. [Google Scholar] [CrossRef]
  170. Furnari, F.B.; Fenton, T.; Bachoo, R.M.; Mukasa, A.; Stommel, J.M.; Stegh, A.; Hahn, W.C.; Ligon, K.L.; Louis, D.N.; Brennan, C. Malignant astrocytic glioma: Genetics, biology, and paths to treatment. Genes Dev. 2007, 21, 2683–2710. [Google Scholar] [CrossRef] [PubMed]
  171. Altaner, C. Glioblastoma and stem cells-Minireview. Neoplasma 2008, 55, 369. [Google Scholar]
  172. Das, S.; Srikanth, M.; Kessler, J.A. Cancer stem cells and glioma. Nat. Rev. Neurol. 2008, 4, 427. [Google Scholar] [CrossRef]
  173. Singh, S.K.; Clarke, I.D.; Terasaki, M.; Bonn, V.E.; Hawkins, C.; Squire, J.; Dirks, P.B. Identification of a cancer stem cell in human brain tumors. Cancer Res. 2003, 63, 5821–5828. [Google Scholar] [PubMed]
  174. Yu, S.-C.; Ping, Y.-F.; Yi, L.; Zhou, Z.-H.; Chen, J.-H.; Yao, X.-H.; Gao, L.; Wang, J.M.; Bian, X.-W. Isolation and characterization of cancer stem cells from a human glioblastoma cell line U87. Cancer Lett. 2008, 265, 124–134. [Google Scholar] [CrossRef] [PubMed]
  175. Yuan, X.; Curtin, J.; Xiong, Y.; Liu, G.; Waschsmann-Hogiu, S.; Farkas, D.L.; Black, K.L.; John, S.Y. Isolation of cancer stem cells from adult glioblastoma multiforme. Oncogene 2004, 23, 9392. [Google Scholar] [CrossRef] [PubMed]
  176. Li, G.-H.; Wei, H.; Lv, S.-Q.; Ji, H.; Wang, D.-L. Knockdown of STAT3 expression by RNAi suppresses growth and induces apoptosis and differentiation in glioblastoma stem cells. Int. J. Oncol. 2010, 37, 103–110. [Google Scholar] [PubMed]
  177. Wang, H.; Li, X.; Lu, X. Silencing of signal transducer and activator of transcription 3 gene expression using RNAi enhances the efficacy of radiotherapy for laryngeal carcinoma in vivo. Zhonghua Er Bi Yan Hou Tou Jing Wai Ke Za Zhi 2009, 44, 591–596. [Google Scholar] [PubMed]
  178. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2019. CA Cancer J. Clin. 2019, 69, 7–34. [Google Scholar] [CrossRef] [PubMed]
  179. Mohammadinejad, R.; Ahmadi, Z.; Tavakol, S.; Ashrafizadeh, M. Berberine as a potential autophagy modulator. J. Cell. Physiol. 2019, 234, 14914–14926. [Google Scholar] [CrossRef]
  180. Ashrafizadeh, M.; Ahmadi, Z.; Mohammadinejad, R.; Kaviyani, N.; Tavakol, S. Monoterpenes modulating autophagy: A review study. Basic Clin. Pharmacol. Toxicol. 2019. [Google Scholar] [CrossRef] [PubMed]
  181. Do, T.N.T.; Lee, W.-H.; Loo, C.-Y.; Zavgorodniy, A.V.; Rohanizadeh, R. Hydroxyapatite nanoparticles as vectors for gene delivery. Ther. Deliv. 2012, 3, 623–632. [Google Scholar]
  182. Bose, S.; Tarafder, S. Calcium phosphate ceramic systems in growth factor and drug delivery for bone tissue engineering: A review. Acta Biomater. 2012, 8, 1401–1421. [Google Scholar] [CrossRef] [Green Version]
  183. Olton, D.; Li, J.; Wilson, M.E.; Rogers, T.; Close, J.; Huang, L.; Kumta, P.N.; Sfeir, C. Nanostructured calcium phosphates (NanoCaPs) for non-viral gene delivery: Influence of the synthesis parameters on transfection efficiency. Biomaterials 2007, 28, 1267–1279. [Google Scholar] [CrossRef] [PubMed]
  184. Liang, Z.; Wang, H.; Guo, B.; Li, F.; Liu, J.; Liu, Z.; Xu, L.; Yun, W.; Zhao, X.; Zhang, L. Inhibition of prostate cancer RM1 cell growth in vitro by hydroxyapatite nanoparticle-delivered short hairpin RNAs against Stat3. Mol. Med. Rep. 2017, 16, 459–465. [Google Scholar] [CrossRef] [PubMed]
  185. Jin, S.-E.; Kim, C.-K. Charge-mediated topical delivery of plasmid DNA with cationic lipid nanoparticles to the skin. Colloids Surf. B Biointerfaces 2014, 116, 582–590. [Google Scholar] [CrossRef] [PubMed]
  186. Kotmakçı, M.; Çetintaş, V.B.; Kantarcı, A.G. Preparation and characterization of lipid nanoparticle/pDNA complexes for STAT3 downregulation and overcoming chemotherapy resistance in lung cancer cells. Int. J. Pharm. 2017, 525, 101–111. [Google Scholar] [CrossRef] [PubMed]
  187. Pindiprolu, S.K.S.; Krishnamurthy, P.T.; Chintamaneni, P.K. Pharmacological targets of breast cancer stem cells: A review. Naunyn Schmiedeberg Arch. Pharmacol. 2018, 391, 463–479. [Google Scholar] [CrossRef] [PubMed]
  188. Li, R.; You, S.; Hu, Z.; Chen, Z.G.; Sica, G.L.; Khuri, F.R.; Curran, W.J.; Shin, D.M.; Deng, X. Inhibition of STAT3 by niclosamide synergizes with erlotinib against head and neck cancer. PLoS ONE 2013, 8, e74670. [Google Scholar] [CrossRef] [PubMed]
  189. Ma, Y.; Zhang, X.; Xu, X.; Shen, L.; Yao, Y.; Yang, Z.; Liu, P. STAT3 decoy oligodeoxynucleotides-loaded solid lipid nanoparticles induce cell death and inhibit invasion in ovarian cancer cells. PLoS ONE 2015, 10, e0124924. [Google Scholar] [CrossRef]
  190. Kim, S.H.; Yoo, H.S.; Joo, M.K.; Kim, T.; Park, J.-J.; Lee, B.J.; Chun, H.J.; Lee, S.W.; Bak, Y.-T. Arsenic trioxide attenuates STAT-3 activity and epithelial-mesenchymal transition through induction of SHP-1 in gastric cancer cells. BMC Cancer 2018, 18, 150. [Google Scholar] [CrossRef]
  191. Zhu, L.; Cheng, X.; Shi, J.; Lin, J.; Chen, G.; Jin, H.; Liu, A.B.; Pyo, H.; Ye, J.; Zhu, Y. Crosstalk between bone marrow-derived myofibroblasts and gastric cancer cells regulates cancer stemness and promotes tumorigenesis. Oncogene 2016, 35, 5388. [Google Scholar] [CrossRef]
  192. Chang, J.C. Cancer stem cells: Role in tumor growth, recurrence, metastasis, and treatment resistance. Medicine 2016, 95. [Google Scholar] [CrossRef]
  193. Huang, Y.; Zhou, B.; Luo, H.; Mao, J.; Huang, Y.; Zhang, K.; Mei, C.; Yan, Y.; Jin, H.; Gao, J. ZnAs@SiO2 nanoparticles as a potential anti-tumor drug for targeting stemness and epithelial-mesenchymal transition in hepatocellular carcinoma via SHP-1/JAK2/STAT3 signaling. Theranostics 2019, 9, 4391. [Google Scholar] [CrossRef]
  194. Malavasi, F.; Funaro, A.; Alessio, M.; DeMonte, L.B.; Ausiello, C.M.; Dianzani, U.; Lanza, F.; Magrini, E.; Momo, M.; Roggero, S. CD38: A multi-lineage cell activation molecule with a split personality. Int. J. Clin. Lab. Res. 1992, 22, 73–80. [Google Scholar] [CrossRef]
  195. Beum, P.V.; Lindorfer, M.A.; Peek, E.M.; Stukenberg, P.T.; de Weers, M.; Beurskens, F.J.; Parren, P.W.; van de Winkel, J.G.; Taylor, R.P. Penetration of antibody-opsonized cells by the membrane attack complex of complement promotes Ca2+ influx and induces streamers. Eur. J. Immunol. 2011, 41, 2436–2446. [Google Scholar] [CrossRef] [PubMed]
  196. de la Puente, P.; Luderer, M.J.; Federico, C.; Jin, A.; Gilson, R.C.; Egbulefu, C.; Alhallak, K.; Shah, S.; Muz, B.; Sun, J. Enhancing proteasome-inhibitory activity and specificity of bortezomib by CD38 targeted nanoparticles in multiple myeloma. J. Control. Release 2018, 270, 158–176. [Google Scholar] [CrossRef]
  197. De Weers, M.; Tai, Y.-T.; Van Der Veer, M.S.; Bakker, J.M.; Vink, T.; Jacobs, D.C.; Oomen, L.A.; Peipp, M.; Valerius, T.; Slootstra, J.W. Daratumumab, a novel therapeutic human CD38 monoclonal antibody, induces killing of multiple myeloma and other hematological tumors. J. Immunol. 2011, 186, 1840–1848. [Google Scholar] [CrossRef] [PubMed]
  198. Huang, Y.-H.; Vakili, M.R.; Molavi, O.; Morrissey, Y.; Wu, C.; Paiva, I.; Soleimani, A.H.; Sanaee, F.; Lavasanifar, A.; Lai, R. Decoration of Anti-CD38 on Nanoparticles Carrying a STAT3 Inhibitor Can Improve the Therapeutic Efficacy Against Myeloma. Cancers 2019, 11, 248. [Google Scholar] [CrossRef] [PubMed]
  199. Elamanchili, P.; Lutsiak, C.M.; Hamdy, S.; Diwan, M.; Samuel, J. “Pathogen-mimicking” nanoparticles for vaccine delivery to dendritic cells. J. Immunother. 2007, 30, 378–395. [Google Scholar] [CrossRef]
  200. Foged, C.; Sundblad, A.; Hovgaard, L. Targeting vaccines to dendritic cells. Pharm. Res. 2002, 19, 229–238. [Google Scholar] [CrossRef]
  201. Jiang, W.; Gupta, R.K.; Deshpande, M.C.; Schwendeman, S.P. Biodegradable poly (lactic-co-glycolic acid) microparticles for injectable delivery of vaccine antigens. Adv. Drug Deliv. Rev. 2005, 57, 391–410. [Google Scholar] [CrossRef]
  202. Schlosser, E.; Mueller, M.; Fischer, S.; Basta, S.; Busch, D.H.; Gander, B.; Groettrup, M. TLR ligands and antigen need to be coencapsulated into the same biodegradable microsphere for the generation of potent cytotoxic T lymphocyte responses. Vaccine 2008, 26, 1626–1637. [Google Scholar] [CrossRef] [Green Version]
  203. Waeckerle-Men, Y.; Gander, B.; Groettrup, M. Delivery of tumor antigens to dendritic cells using biodegradable microspheres. In Adoptive Immunotherapy: Methods and Protocols; Springer: Berlin, Germany, 2005; pp. 35–46. [Google Scholar]
  204. Waeckerle-Men, Y.; Groettrup, M. PLGA microspheres for improved antigen delivery to dendritic cells as cellular vaccines. Adv. Drug Deliv. Rev. 2005, 57, 475–482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Molavi, O.; Mahmud, A.; Hamdy, S.; Hung, R.W.; Lai, R.; Samuel, J.; Lavasanifar, A. Development of a poly (d, l-lactic-co-glycolic acid) nanoparticle formulation of STAT3 inhibitor JSI-124: Implication for cancer immunotherapy. Mol. Pharm. 2010, 7, 364–374. [Google Scholar] [CrossRef] [PubMed]
  206. Su, W.-P.; Cheng, F.-Y.; Shieh, D.-B.; Yeh, C.-S.; Su, W.-C. PLGA nanoparticles codeliver paclitaxel and Stat3 siRNA to overcome cellular resistance in lung cancer cells. Int. J. Nanomed. 2012, 7, 4269. [Google Scholar] [CrossRef]
  207. Alshamsan, A.; Haddadi, A.; Hamdy, S.; Samuel, J.; El-Kadi, A.O.; Uludag, H.; Lavasanifar, A. STAT3 silencing in dendritic cells by siRNA polyplexes encapsulated in PLGA nanoparticles for the modulation of anticancer immune response. Mol. Pharm. 2010, 7, 1643–1654. [Google Scholar] [CrossRef] [PubMed]
  208. Labala, S.; Jose, A.; Venuganti, V.V.K. Transcutaneous iontophoretic delivery of STAT3 siRNA using layer-by-layer chitosan coated gold nanoparticles to treat melanoma. Colloids Surf. B Biointerfaces 2016, 146, 188–197. [Google Scholar] [CrossRef]
  209. Ferlay, J.; Soerjomataram, I.; Dikshit, R.; Eser, S.; Mathers, C.; Rebelo, M.; Parkin, D.M.; Forman, D.; Bray, F. Cancer incidence and mortality worldwide: Sources, methods and major patterns in GLOBOCAN 2012. Int. J. Cancer 2015, 136, E359–E386. [Google Scholar] [CrossRef]
  210. Shain, A.H.; Bastian, B.C. From melanocytes to melanomas. Nat. Rev. Cancer 2016, 16, 345. [Google Scholar] [CrossRef] [PubMed]
  211. Maverakis, E.; Cornelius, L.A.; Bowen, G.M.; Phan, T.; Patel, F.B.; Fitzmaurice, S.; He, Y.; Burrall, B.; Duong, C.; Kloxin, A.M. Metastatic melanoma–a review of current and future treatment options. Acta Derm. Venereol. 2015, 95, 516–527. [Google Scholar] [CrossRef]
  212. Labala, S.; Jose, A.; Chawla, S.R.; Khan, M.S.; Bhatnagar, S.; Kulkarni, O.P.; Venuganti, V.V.K. Effective melanoma cancer suppression by iontophoretic co-delivery of STAT3 siRNA and imatinib using gold nanoparticles. Int. J. Pharm. 2017, 525, 407–417. [Google Scholar] [CrossRef] [PubMed]
  213. Song, J.; Shi, J.; Dong, D.; Fang, M.; Zhong, W.; Wang, K.; Wu, N.; Huang, Y.; Liu, Z.; Cheng, Y. A new approach to predict progression-free survival in stage IV EGFR-mutant NSCLC patients with EGFR-TKI therapy. Clin. Cancer Res. 2018, 24, 3583–3592. [Google Scholar] [CrossRef]
  214. Gazdar, A. Activating and resistance mutations of EGFR in non-small-cell lung cancer: Role in clinical response to EGFR tyrosine kinase inhibitors. Oncogene 2009, 28, S24. [Google Scholar] [CrossRef] [PubMed]
  215. Pardanani, A.; Harrison, C.; Cortes, J.E.; Cervantes, F.; Mesa, R.A.; Milligan, D.; Masszi, T.; Mishchenko, E.; Jourdan, E.; Vannucchi, A.M. Safety and efficacy of fedratinib in patients with primary or secondary myelofibrosis: A randomized clinical trial. JAMA Oncol. 2015, 1, 643–651. [Google Scholar] [CrossRef] [PubMed]
  216. Chen, D.; Zhang, F.; Wang, J.; He, H.; Duan, S.; Zhu, R.; Chen, C.; Chen, Y.; Yin, L. Biodegradable Nanoparticles Mediated Co-Delivery of Erlotinib (ELTN) and Fedratinib (FDTN) toward the Treatment of ELTN-Resistant Non-Small Cell Lung Cancer (NSCLC) via Suppression of the JAK2/STAT3 Signaling Pathway. Front. Pharmacol. 2018, 9, 1214. [Google Scholar] [CrossRef] [PubMed]
  217. Hornez, J.-C.; Chai, F.; Monchau, F.; Blanchemain, N.; Descamps, M.; Hildebrand, H. Biological and physico-chemical assessment of hydroxyapatite (HA) with different porosity. Biomol. Eng. 2007, 24, 505–509. [Google Scholar] [CrossRef] [PubMed]
  218. Kumar, G.S.; Girija, E.; Thamizhavel, A.; Yokogawa, Y.; Kalkura, S.N. Synthesis and characterization of bioactive hydroxyapatite–calcite nanocomposite for biomedical applications. J. Colloid Interface Sci. 2010, 349, 56–62. [Google Scholar] [CrossRef] [PubMed]
  219. Levy-Nissenbaum, E.; Radovic-Moreno, A.F.; Wang, A.Z.; Langer, R.; Farokhzad, O.C. Nanotechnology and aptamers: Applications in drug delivery. Trends Biotechnol. 2008, 26, 442–449. [Google Scholar] [CrossRef] [PubMed]
  220. McAllister, K.; Sazani, P.; Adam, M.; Cho, M.J.; Rubinstein, M.; Samulski, R.J.; DeSimone, J.M. Polymeric nanogels produced via inverse microemulsion polymerization as potential gene and antisense delivery agents. J. Am. Chem. Soc. 2002, 124, 15198–15207. [Google Scholar] [CrossRef] [PubMed]
  221. Shi, Z.; Huang, X.; Cai, Y.; Tang, R.; Yang, D. Size effect of hydroxyapatite nanoparticles on proliferation and apoptosis of osteoblast-like cells. Acta Biomater. 2009, 5, 338–345. [Google Scholar] [CrossRef]
  222. Sun, H.; Jiang, M.; Zhu, S. In vitro and in vivo studies on hydroxyapatite nanoparticles as a novel vector for inner ear gene therapy. Zhonghua Er Bi Yan Hou Tou Jing Wai Ke Za Zhi 2008, 43, 51–57. [Google Scholar]
  223. Ye, F.; Guo, H.; Zhang, H.; He, X. Polymeric micelle-templated synthesis of hydroxyapatite hollow nanoparticles for a drug delivery system. Acta Biomater. 2010, 6, 2212–2218. [Google Scholar] [CrossRef]
  224. Liang, Z.W.; Guo, B.F.; Li, Y.; Li, X.J.; Li, X.; Zhao, L.J.; Gao, L.F.; Yu, H.; Zhao, X.J.; Zhang, L.; et al. Plasmid-based Stat3 siRNA delivered by hydroxyapatite nanoparticles suppresses mouse prostate tumour growth in vivo. Asian J. Androl. 2011, 13, 481–486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Behlke, M.A. Progress towards in vivo use of siRNAs. Mol. Ther. 2006, 13, 644–670. [Google Scholar] [CrossRef]
  226. Demeneix, B.; Behr, J.P. Polyethylenimine (PEI). Adv. Genet. 2005, 53, 215–230. [Google Scholar]
  227. Doody, A.; Putnam, D. RNA-interference effectors and their delivery. Crit. Rev. Ther. Drug Carr. Syst. 2006, 23, 137–164. [Google Scholar]
  228. Uprichard, S.L. The therapeutic potential of RNA interference. FEBS Lett. 2005, 579, 5996–6007. [Google Scholar] [CrossRef] [Green Version]
  229. Alshamsan, A.; Hamdy, S.; Samuel, J.; El-Kadi, A.O.; Lavasanifar, A.; Uludağ, H. The induction of tumor apoptosis in B16 melanoma following STAT3 siRNA delivery with a lipid-substituted polyethylenimine. Biomaterials 2010, 31, 1420–1428. [Google Scholar] [CrossRef]
  230. Liu, C.; Li, X.-W.; Cui, L.-M.; Li, L.-C.; Chen, L.-Y.; Zhang, X.-W. Inhibition of tumor angiogenesis by TTF1 from extract of herbal medicine. World J. Gastroenterol. WJG 2011, 17, 4875. [Google Scholar] [CrossRef] [PubMed]
  231. Li, Y.; Bian, L.; Cui, F.; Li, L.; Zhang, X. TTF1-induced apoptosis of HepG-2 cells through a mitochondrial pathway. Oncol. Rep. 2011, 26, 651–657. [Google Scholar]
  232. Li, Y.; Cui, F.; Zhang, X. Preparation technology of Sorbaria sorbifolia solid lipid nanoparticles. Lishizhen Med. Mater. Med. Res. 2012, 23, 2549–2550. [Google Scholar]
  233. Xiao, B.; Lin, D.; Zhang, X.; Zhang, M.; Zhang, X. TTF1, in the form of nanoparticles, inhibits angiogenesis, cell migration and cell invasion in vitro and in vivo in human hepatoma through STAT3 regulation. Molecules 2016, 21, 1507. [Google Scholar] [CrossRef]
  234. Das, J.; Das, S.; Paul, A.; Samadder, A.; Bhattacharyya, S.S.; Khuda-Bukhsh, A.R. Assessment of drug delivery and anticancer potentials of nanoparticles-loaded siRNA targeting STAT3 in lung cancer, in vitro and in vivo. Toxicol. Lett. 2014, 225, 454–466. [Google Scholar] [CrossRef] [PubMed]
  235. Benson, H.A. Transdermal drug delivery: Penetration enhancement techniques. Curr. Drug Deliv. 2005, 2, 23–33. [Google Scholar] [CrossRef] [PubMed]
  236. Paudel, K.S.; Milewski, M.; Swadley, C.L.; Brogden, N.K.; Ghosh, P.; Stinchcomb, A.L. Challenges and opportunities in dermal/transdermal delivery. Ther. Deliv. 2010, 1, 109–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Souza, J.G.; Dias, K.; Pereira, T.A.; Bernardi, D.S.; Lopez, R.F. Topical delivery of ocular therapeutics: Carrier systems and physical methods. J. Pharm. Pharmacol. 2014, 66, 507–530. [Google Scholar] [CrossRef] [PubMed]
  238. Kajimoto, K.; Yamamoto, M.; Watanabe, M.; Kigasawa, K.; Kanamura, K.; Harashima, H.; Kogure, K. Noninvasive and persistent transfollicular drug delivery system using a combination of liposomes and iontophoresis. Int. J. Pharm. 2011, 403, 57–65. [Google Scholar] [CrossRef] [PubMed]
  239. Han, I.; Kim, M.; Kim, J. Enhanced transfollicular delivery of adriamycin with a liposome and iontophoresis. Exp. Dermatol. 2004, 13, 86–92. [Google Scholar] [CrossRef] [PubMed]
  240. Eljarrat-Binstock, E.; Orucov, F.; Aldouby, Y.; Frucht-Pery, J.; Domb, A.J. Charged nanoparticles delivery to the eye using hydrogel iontophoresis. J. Control. Release 2008, 126, 156–161. [Google Scholar] [CrossRef]
  241. Jose, A.; Labala, S.; Venuganti, V.V.K. Co-delivery of curcumin and STAT3 siRNA using deformable cationic liposomes to treat skin cancer. J. Drug Target. 2017, 25, 330–341. [Google Scholar] [CrossRef]
  242. Jose, A.; Labala, S.; Ninave, K.M.; Gade, S.K.; Venuganti, V.V.K. Effective skin cancer treatment by topical co-delivery of curcumin and STAT3 siRNA using cationic liposomes. AAPS PharmSciTech 2018, 19, 166–175. [Google Scholar] [CrossRef]
  243. Noy, R.; Pollard, J.W. Tumor-associated macrophages: From mechanisms to therapy. Immunity 2014, 41, 49–61. [Google Scholar] [CrossRef]
  244. Komohara, Y.; Fujiwara, Y.; Ohnishi, K.; Takeya, M. Tumor-associated macrophages: Potential therapeutic targets for anti-cancer therapy. Adv. Drug Deliv. Rev. 2016, 99, 180–185. [Google Scholar] [CrossRef] [PubMed]
  245. Andersen, M.N.; Etzerodt, A.; Graversen, J.H.; Holthof, L.C.; Moestrup, S.K.; Hokland, M.; Møller, H.J. STAT3 inhibition specifically in human monocytes and macrophages by CD163-targeted corosolic acid-containing liposomes. Cancer Immunol. Immunother. 2019, 68, 489–502. [Google Scholar] [CrossRef] [PubMed]
  246. Yang, M.; McKay, D.; Pollard, J.W.; Lewis, C.E. Diverse functions of macrophages in different tumor microenvironments. Cancer Res. 2018, 78, 5492–5503. [Google Scholar] [CrossRef] [PubMed]
  247. Zhang, Q.-W.; Liu, L.; Gong, C.-Y.; Shi, H.-S.; Zeng, Y.-H.; Wang, X.-Z.; Zhao, Y.-W.; Wei, Y.-Q. Prognostic significance of tumor-associated macrophages in solid tumor: A meta-analysis of the literature. PLoS ONE 2012, 7, e50946. [Google Scholar] [CrossRef] [PubMed]
  248. Cheng, F.; Wang, H.-W.; Cuenca, A.; Huang, M.; Ghansah, T.; Brayer, J.; Kerr, W.G.; Takeda, K.; Akira, S.; Schoenberger, S.P. A critical role for Stat3 signaling in immune tolerance. Immunity 2003, 19, 425–436. [Google Scholar] [CrossRef]
  249. Zhang, L.; Alizadeh, D.; Van Handel, M.; Kortylewski, M.; Yu, H.; Badie, B. Stat3 inhibition activates tumor macrophages and abrogates glioma growth in mice. Glia 2009, 57, 1458–1467. [Google Scholar] [CrossRef] [PubMed]
  250. Kortylewski, M.; Kujawski, M.; Wang, T.; Wei, S.; Zhang, S.; Pilon-Thomas, S.; Niu, G.; Kay, H.; Mulé, J.; Kerr, W.G. Inhibiting Stat3 signaling in the hematopoietic system elicits multicomponent antitumor immunity. Nat. Med. 2005, 11, 1314. [Google Scholar] [CrossRef] [PubMed]
  251. Herrmann, A.; Kortylewski, M.; Kujawski, M.; Zhang, C.; Reckamp, K.; Armstrong, B.; Wang, L.; Kowolik, C.; Deng, J.; Figlin, R. Targeting Stat3 in the myeloid compartment drastically improves the in vivo antitumor functions of adoptively transferred T cells. Cancer Res. 2010, 70, 7455–7464. [Google Scholar] [CrossRef]
  252. Giurisato, E.; Xu, Q.; Lonardi, S.; Telfer, B.; Russo, I.; Pearson, A.; Finegan, K.G.; Wang, W.; Wang, J.; Gray, N.S. Myeloid ERK5 deficiency suppresses tumor growth by blocking protumor macrophage polarization via STAT3 inhibition. Proc. Natl. Acad. Sci. USA 2018, 115, E2801–E2810. [Google Scholar] [CrossRef] [Green Version]
  253. Bader, H.; Ringsdorf, H.; Schmidt, B. Water-soluble polymers in medi cine. Angew. Makromol. Chem. 1984, 123, 457–485. [Google Scholar] [CrossRef]
  254. Jones, M.-C.; Leroux, J.-C. Polymeric micelles—A new generation of colloidal drug carriers. Eur. J. Pharm. Biopharm. 1999, 48, 101–111. [Google Scholar] [CrossRef]
  255. Kwon, G.S.; Forrest, M.L. Amphiphilic block copolymer micelles for nanoscale drug delivery. Drug Dev. Res. 2006, 67, 15–22. [Google Scholar] [CrossRef]
  256. Lu, Y.; Park, K. Polymeric micelles and alternative nanonized delivery vehicles for poorly soluble drugs. Int. J. Pharm. 2013, 453, 198–214. [Google Scholar] [CrossRef] [PubMed]
  257. Matsumura, Y.; Maeda, H. A new concept for macromolecular therapeutics in cancer chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs. Cancer Res. 1986, 46, 6387–6392. [Google Scholar] [PubMed]
  258. Maeda, H.; Tsukigawa, K.; Fang, J. A Retrospective 30 Years After Discovery of the Enhanced Permeability and Retention Effect of Solid Tumors: Next-Generation Chemotherapeutics and Photodynamic Therapy—Problems, Solutions, and Prospects. Microcirculation 2016, 23, 173–182. [Google Scholar] [CrossRef] [PubMed]
  259. Soleimani, A.H.; Garg, S.M.; Paiva, I.M.; Vakili, M.R.; Alshareef, A.; Huang, Y.-H.; Molavi, O.; Lai, R.; Lavasanifar, A. Micellar nano-carriers for the delivery of STAT3 dimerization inhibitors to melanoma. Drug Deliv. Transl. Res. 2017, 7, 571–581. [Google Scholar] [CrossRef] [PubMed]
  260. Molavi, O.; Ma, Z.; Mahmud, A.; Alshamsan, A.; Samuel, J.; Lai, R.; Kwon, G.S.; Lavasanifar, A. Polymeric micelles for the solubilization and delivery of STAT3 inhibitor cucurbitacins in solid tumors. Int. J. Pharm. 2008, 347, 118–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Luo, Z.; Wang, C.; Yi, H.; Li, P.; Pan, H.; Liu, L.; Cai, L.; Ma, Y. Nanovaccine loaded with poly I: C and STAT3 siRNA robustly elicits anti-tumor immune responses through modulating tumor-associated dendritic cells in vivo. Biomaterials 2015, 38, 50–60. [Google Scholar] [CrossRef] [PubMed]
  262. Jiang, Q.; Dai, L.; Cheng, L.; Chen, X.; Li, Y.; Zhang, S.; Su, X.; Zhao, X.; Wei, Y.; Deng, H. Efficient inhibition of intraperitoneal ovarian cancer growth in nude mice by liposomal delivery of short hairpin RNA against STAT 3. J. Obstet. Gynaecol. Res. 2013, 39, 701–709. [Google Scholar] [CrossRef] [PubMed]
  263. Kopechek, J.A.; Carson, A.R.; McTiernan, C.F.; Chen, X.; Hasjim, B.; Lavery, L.; Sen, M.; Grandis, J.R.; Villanueva, F.S. Ultrasound targeted microbubble destruction-mediated delivery of a transcription factor decoy inhibits STAT3 signaling and tumor growth. Theranostics 2015, 5, 1378. [Google Scholar] [CrossRef]
  264. Shi, K.; Fang, Y.; Gao, S.; Yang, D.; Bi, H.; Xue, J.; Lu, A.; Li, Y.; Ke, L.; Lin, X. Inorganic kernel-Supported asymmetric hybrid vesicles for targeting delivery of STAT3-decoy oligonucleotides to overcome anti-HER2 therapeutic resistance of BT474R. J. Control. Release 2018, 279, 53–68. [Google Scholar] [CrossRef]
  265. Garg, S.M.; Vakili, M.R.; Molavi, O.; Lavasanifar, A. Self-Associating Poly (ethylene oxide)-block-poly (α-carboxyl-ε-caprolactone) Drug Conjugates for the Delivery of STAT3 Inhibitor JSI-124: Potential Application in Cancer Immunotherapy. Mol. Pharm. 2017, 14, 2570–2584. [Google Scholar] [CrossRef] [PubMed]
  266. Mai, J.; Huang, Y.; Mu, C.; Zhang, G.; Xu, R.; Guo, X.; Xia, X.; Volk, D.E.; Lokesh, G.L.; Thiviyanathan, V. Bone marrow endothelium-targeted therapeutics for metastatic breast cancer. J. Control. Release 2014, 187, 22–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Falamarzian, A.; Montazeri Aliabadi, H.; Molavi, O.; Seubert, J.M.; Lai, R.; Uludağ, H.; Lavasanifar, A. Effective down-regulation of signal transducer and activator of transcription 3 (STAT3) by polyplexes of siRNA and lipid-substituted polyethyleneimine for sensitization of breast tumor cells to conventional chemotherapy. J. Biomed. Mater. Res. Part A 2014, 102, 3216–3228. [Google Scholar] [CrossRef] [PubMed]
  268. Li, Z.; Guan, Y.-Q.; Liu, J.-M. The role of STAT-6 as a key transcription regulator in HeLa cell death induced by IFN-γ/TNF-α co-immobilized on nanoparticles. Biomaterials 2014, 35, 5016–5027. [Google Scholar] [CrossRef] [PubMed]
  269. Molavi, O.; Ma, Z.; Hamdy, S.; Lavasanifar, A.; Samuel, J. Immunomodulatory and anticancer effects of intra-tumoral co-delivery of synthetic lipid A adjuvant and STAT3 inhibitor, JSI-124. Immunopharmacol. Immunotoxicol. 2009, 31, 214–221. [Google Scholar] [CrossRef] [PubMed]
  270. Pan, J.; Ruan, W.; Qin, M.; Long, Y.; Wan, T.; Yu, K.; Zhai, Y.; Wu, C.; Xu, Y. Intradermal delivery of STAT3 siRNA to treat melanoma via dissolving microneedles. Sci. Rep. 2018, 8, 1117. [Google Scholar] [CrossRef] [PubMed]
  271. Wilhelm, S.; Tavares, A.J.; Dai, Q.; Ohta, S.; Audet, J.; Dvorak, H.F.; Chan, W.C. Analysis of nanoparticle delivery to tumours. Nat. Rev. Mater. 2016, 1, 16014. [Google Scholar] [CrossRef]
Figure 1. Application of nanoparticles in targeting STATs.
Figure 1. Application of nanoparticles in targeting STATs.
Cells 08 01158 g001
Table 1. Signal transducers and activator of transcription (STAT) inhibitors except STAT3 inhibitors.
Table 1. Signal transducers and activator of transcription (STAT) inhibitors except STAT3 inhibitors.
DrugMolecular FormulaTargetEffectAnimal Model/Cell LineRefs
---2-(3′,4′,5′-trimethoxybenzoyl)-3-iodoacetamido-6-methoxy benzo[b]furan derivative 1STAT5Inhibition of STAT5 phosphorylationK562 cells[139]
---N’-(4-Oxo-4 H-chromen-3-yl)methylene) nicotinohydrazideSTAT5Inhibition of STAT5 phosphorylationChronic myeloid leukemia (CML) cells[140]
SEL120-34AC15H19Br2ClN4STAT1, STAT5Inhibition of STAT1 S727 and STAT5 S726 phosphorylationAcute myeloid leukemia (AML) cells[141]
R763---STAT5Inhibition of STAT5 phosphorylationNeoplastic mast cell[142]
PravastatinC23H36O7STAT1Prevention of STAT1 expressionMice[128]
PimozideC28H29F2N3OSTAT5Inhibition of STAT5 phosphorylationK562 cells, peripheral T-cell lymphoma[130,143]
LeflunomideC12H9F3N2O2STAT6Inhibition of tyrosine phosphorylation of STAT6B cells[131]
Niflumic acidC13H9F3N2O2JAK2, STAT6Blockade of STAT6 phosphorylationMouse[136]
CinnamonC36H32O19STAT4Blockade of STAT4 phosphorylationMice[138]
AtiprimodC22H44N2STAT5STAT3Inhibition of phosphorylationAML cells[144]
Table 2. Natural STAT3 inhibitors.
Table 2. Natural STAT3 inhibitors.
DrugMolecular FormulaEffectAnimal Model/Cell LineRefs
SilibininC25H22O10Blocking pathways of STAT3 activationEndometrial carcinoma cells[145]
QuercetinC15H10O7Inhibiting STAT3 signaling pathwaysLymphoma cells[146]
BerberineC20H18NO4+Decrease of STAT3 phosphorylationKeratinocytes[147]
ResveratrolC14H12O3Inhibition of STAT3Rat[148]
Triterpenes from Helicteres angustifolia---Inhibition of STAT3 phosphorylationHT-29 colorectal cancer cells[149]
ButeinC15H12O5Inhibition of STAT3 expressionMultiple myeloma cells[150]
Caffeic acidC9H8O4Inhibition of activity of STAT3
Inhibition of JAK/STAT3 signaling pathway
Mouse, Human renal carcinoma cells[151,152]
CapsaicinC18H27NO3Inhibition of STAT3Human multiple myeloma cells[153]
CelastrolC29H38O4Inhibition of STAT3 phosphorylationHuman hepatocellular carcinoma[154]
CucurbitacinC32H48O8Inhibition of STAT3 activationAML cells[155]
DiosgeninC27H42O3Inhibition of STAT3 phosphorylationHuman hepatocellular carcinoma cells[156]
GuggulsteroneC21H28O2Inhibition of STAT3 phosphorylationTumor cells[157]
HonokiolC18H18O2Modulation of STAT3 activationBreast cancer cells[158]
Avicin DC98H155NO46Inhibition of STAT3 phosphorylationU266 cells, myeloma cell lines[159]
PiceatannolC14H12O4Reduction of P-STAT3 expressionMouse[160]
Withaferin A analogues---Inhibition of STAT3 phosphorylationBreast cancer cell line[161]
EmodinC15H10O5Inhibition of STAT3 phosphorylationHepatocellular carcinoma cell lines[162]
Table 3. Potential use of nanocarriers for delivery of STAT inhibitors.
Table 3. Potential use of nanocarriers for delivery of STAT inhibitors.
Nano-carriersAgentIn vitro/In vivoCell Line/Animal ModelMajor OutcomesRefs
Gold nanoparticleSTAT3 siRNA and imatinibIn vitro and in vivoB16F10 (melanoma cells) and tumor bearing C57BL/6 mice In vitro: Inhibition of tumor growth and decreased expression of STAT3
In vivo: decreased weight and volume of tumor, reduced expression of STAT3
[212]
Hydroxyapatite nanoparticlesPlasmid-based STAT3 siRNAIn vivoMouse prostate cancer cellsThe downregulation of STAT3 downstream genes such as Bcl-2, VEGF and cyclin D1, and consequently, increased level of apoptosis in cancer cells[231]
PLGA nanoparticlessiRNA polyplexesIn vitroDCsDownregulation of STAT3 expression and increased level of maturation and functionality in DCs[207]
MicelleSTAT3 siRNAIn vivoMice with tumor-associated DCs (TADCs)Downregulation of STAT3 and stimulation of maturation and activation in TADCs[261]
Solid lipid nanoparticleSTAT3 decoy oligodeoxynucleotidesIn vitroHuman ovarian cancer cell lines A2780 and SKOV3Inhibition of STAT3 pathway, stimulation of cell death via increased expression of Bax, Beclin-1, caspase-3 and LC3-II, and prevention of invasion via upregulation of E-cadherin and downregulation of Snail and MMP-9[189]
PEI-PLGA-FITC nanoparticlessiRNA targeting STAT3In vitro and in vivoA549 cells and Balb/c miceIn vitro: Reduced rate of viability in A549 cells.
In vivo: Upregulation of caspase-3 and downregulation of IL-6 in mice
[234]
LiposomeshRNA against STAT3In vitroOvarian cancer cell lines A2780CP and A2780ssIncreased level of apoptosis and inhibition of cell proliferation[262]
Poly (D,L-lactic-co-glycolic-acid) nanoparticleJSI-124 (STAT3 inhibitor)In vitroDCsImproved function of DCs and increased level of T cell proliferation[205]
Ultrasound-targeted microbubble destructionTranscription factor decoy of STAT3In vivoSquamous cell tumorsDownregulation of STAT3 and inhibition of tumor growth[263]
Deformable cationic liposomesCurcumin and STAT3 siRNAIn vitroHuman epidermoid (A431) cancer cellsInhibition of cancer cell growth and stimulation of apoptosis[241]
Lipid-substituted polyethylenimineSTAT3 siRNAIn vitroMurine B16.F10 melanoma cellsRemarkable inhibition of STAT3 expression and induction of apoptosis[229]
Inorganic kernel-supported asymmetric hybrid vesiclesSTAT3-decoy oligonucleotideIn vivoNude mice bearing BT474R breast cancer xenograftSignificant inhibition of tumor growth and prevention of trastuzumab resistance[264]
Self-Associating Poly(ethylene oxide)-block-poly(α-carboxyl-ε-caprolactone) Drug ConjugatesJSI-124 (STAT3 inhibitor)In vitroB16F10 melanoma cells and tumor exposed bone marrow derived dendritic cellsInhibition of STAT3 and great anti-tumor activity[265]
E-selectin thioaptamer-conjugated multistage vectorsiRNAIn vivoMice bearing metastatic breast cancer and murine xenograft models of human MDA-MB-231 breast tumorDownregulation of STAT3 as much as 48.7% in cancer cells inside bone marrow, and increased rate of survival in mice[266]
Lipid-substituted polyethyleniminesiRNA polyplexesIn vitroWild-type MDA-MB-435 breast cancer cellsDownregulation of STAT3 and decreasedviability of cells[267]
Polymeric nanoparticlesSTAT6In vitro and in vivoHeLa cells and tumor bearing miceIn vitro: knockdown of IFN-γR2 and stimulation of cell death in HeLa human epithelial cells
In vivo: decreased volume of tumor and increased rate of survival
[268]
Gold nanoparticlesSTAT3 siRNAIn vitroB16F10 murine melanoma cellsRemarkable inhibition of cancer cell growth[208]
Lipid nanoparticleRNAi-mediating plasmid DNAIn vitroChemoresistant Calu1 cellsDownregulation of STAT3 and resensitize Calu human lung cancer cells to chemotherapy (cisplatin)[186]
PLGA nanoparticlesJSI-124 (STAT3 inhibitor)In vivoC57BL/b male miceGreat anti-tumor impact[269]
Dissolving microneedlesSTAT3 siRNAIn vivoFemale C57BL/b miceGreat gene silencing and inhibition of tumor cell growth[270]

Share and Cite

MDPI and ACS Style

Ashrafizadeh, M.; Ahmadi, Z.; Kotla, N.G.; Afshar, E.G.; Samarghandian, S.; Mandegary, A.; Pardakhty, A.; Mohammadinejad, R.; Sethi, G. Nanoparticles Targeting STATs in Cancer Therapy. Cells 2019, 8, 1158. https://doi.org/10.3390/cells8101158

AMA Style

Ashrafizadeh M, Ahmadi Z, Kotla NG, Afshar EG, Samarghandian S, Mandegary A, Pardakhty A, Mohammadinejad R, Sethi G. Nanoparticles Targeting STATs in Cancer Therapy. Cells. 2019; 8(10):1158. https://doi.org/10.3390/cells8101158

Chicago/Turabian Style

Ashrafizadeh, Milad, Zahra Ahmadi, Niranjan G. Kotla, Elham Ghasemipour Afshar, Saeed Samarghandian, Ali Mandegary, Abbas Pardakhty, Reza Mohammadinejad, and Gautam Sethi. 2019. "Nanoparticles Targeting STATs in Cancer Therapy" Cells 8, no. 10: 1158. https://doi.org/10.3390/cells8101158

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