Next Article in Journal
Wolfram Syndrome 1: A Pediatrician’s and Pediatric Endocrinologist’s Perspective
Next Article in Special Issue
DNA Nanomachine (DNM) Biplex Assay for Differentiating Bacillus cereus Species
Previous Article in Journal
Rose Bengal Induced Photothrombosis in CAM Integrated Human Split Skin Grafts—A Feasibility Study
Previous Article in Special Issue
Integrating Analysis to Identify Differential circRNAs Involved in Goat Endometrial Receptivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modulation of AKT Pathway-Targeting miRNAs for Cancer Cell Treatment with Natural Products

1
Division of Breast Oncology and Surgery, Department of Surgery, Kaohsiung Medical University Hospital, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
2
Graduate Institute of Medicine, College of Medicine, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
3
School of Dentistry, Taipei Medical University, Taipei 11031, Taiwan
4
Department of Oral and Maxillofacial Surgery, Chi-Mei Medical Center, Tainan 71004, Taiwan
5
Graduate Institute of Natural Products, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
6
Department of Biomedical Science and Environmental Biology, College of Life Science, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
7
School of Post-Baccalaureate Medicine, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
8
Department of Radiation Oncology, Kaohsiung Medical University Hospital, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
9
Institute of Medical Science and Technology, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
10
Center for Cancer Research, Kaohsiung Medical University, Kaohsiung 80708, Taiwan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(4), 3688; https://doi.org/10.3390/ijms24043688
Submission received: 21 November 2022 / Revised: 6 February 2023 / Accepted: 10 February 2023 / Published: 12 February 2023
(This article belongs to the Special Issue Molecular Biology of RNA: Recent Progress)

Abstract

:
Many miRNAs are known to target the AKT serine-threonine kinase (AKT) pathway, which is critical for the regulation of several cell functions in cancer cell development. Many natural products exhibiting anticancer effects have been reported, but their connections to the AKT pathway (AKT and its effectors) and miRNAs have rarely been investigated. This review aimed to demarcate the relationship between miRNAs and the AKT pathway during the regulation of cancer cell functions by natural products. Identifying the connections between miRNAs and the AKT pathway and between miRNAs and natural products made it possible to establish an miRNA/AKT/natural product axis to facilitate a better understanding of their anticancer mechanisms. Moreover, the miRNA database (miRDB) was used to retrieve more AKT pathway-related target candidates for miRNAs. By evaluating the reported facts, the cell functions of these database-generated candidates were connected to natural products. Therefore, this review provides a comprehensive overview of the natural product/miRNA/AKT pathway in the modulation of cancer cell development.

1. Introduction

MicroRNAs (miRNAs) are part of a group of noncoding RNAs that exist in eukaryotic cells. They are approximately 21~22 nucleotides in length. miRNAs are responsible for post-transcriptional gene regulation and are classified according to their oncogenic and tumor suppression functions in the regulation of diverse cell functions [1]. miRNAs are known to have several functions in the regulation of proliferation, apoptosis [2], autophagy [3,4,5], endoplasmic reticulum (ER) stress [1,6], ferroptosis [7,8,9,10], necroptosis [11,12,13,14], DNA damage response (DDR) [15,16,17], senescence [18,19], and migration [20,21,22,23] relating to cancer cells [23,24,25]. In the following, we briefly introduce the background for these cell functions and their relationships to miRNA.
Apoptosis is a form of programmed cell death involving a series of activations for caspase signaling. Different miRNAs may exhibit functions opposite to apoptosis. Oncogenic miRNAs suppress cancer cell apoptosis, while tumor-suppressor miRNAs promote such apoptosis [2]. Autophagy refers to self-eating biomolecule recycling for energy restoration and abnormal organelle digestion of cells in response to starvation and other stressors. Autophagy-regulating miRNAs exhibit inhibitory or promoting effects in chemotherapy, radiotherapy, endocrine therapy, and target therapy for breast cancer [3]. Additionally, some autophagy-associated miRNAs contribute to modulating drug resistance to cancers [4,5].
ER stress occurs when protein misfolding exceeds the ER unfolding capacity. Unfolded protein response (UPR) is triggered by ER stress. When the ER stress is tolerable, cells adapt and can survive, but massive ER stress exceeding the unfolding capacity causes cell death. Oncogenic miRNAs promote early UPR and improve cancer cell survival but suppress late cell death response. In contrast, the effects of ER stress caused by tumor-suppressor miRNAs show the opposite function to that of oncogenic miRNAs [6]. Some miRNAs upregulate and downregulate ER stress in digestive cancer [1].
Ferroptosis refers to nonapoptosis-programmed cell death involving the upregulation of iron uptake, lipid peroxidation, and glutathione peroxidase 4 (GPX4) downregulation [26]. Ferroptosis-suppressing and -promoting miRNAs have been widely reviewed [7]. These miRNAs exhibit distinct functions targeting different components of ferroptosis signaling [8] and regulate ferroptosis among cancer cells [9,10]. Necroptosis refers to the nonapoptosis-programmed, necrosis-based cell death involving receptor-interacting serine/threonine-protein kinase 3 (RIPK3) and the mixed lineage kinase domain-like (MLKL) protein. Some miRNAs upregulate and downregulate necroptosis in the liver [11], kidneys [12], and lungs [13], as well as in colon cancer [14].
Canonical DDR includes the DNA damage sensors (MRE11/RAD50/NBS1 (MRN)), the Ataxia telangiectasia mutated (ATM) gene, the Ataxia telangiectasia and Rad3d-related (ATR) serine/threonine kinase, cell cycle checkpoints (checkpoint kinase 1/2 (CHK1/2) and p53), and downstream proapoptosis signaling [15]. Several miRNAs have been found to target DDR signaling, such as DNA damage sensing [15] and DNA repair [16,17]. Cellular senescence is characterized by permanent quiescent proliferation while maintaining metabolic activity. Senescent cells show a senescence-associated secretory phenotype (SASP) [18]. Modulating the SASP may control tumor growth. Several senescence-targeting miRNAs have been reviewed [18,19]. Moreover, migration and invasion are critical for metastases in cancer progression. Several miRNAs have been reported to regulate the detachment, epithelial-to-mesenchymal transition (EMT), migration, invasion, angiogenesis, and metastasis of cancer cells [20,21,22].
Notably, various miRNAs cooperate to control several cell functions. For example, apoptosis, necroptosis, and ferroptosis, which have close interactions, are mediated by reactive oxygen species (ROS) and lipid peroxidation [27]. Consequently, miRNAs regulating these cell functions may show complex interactions. However, the functions of miRNAs are commonly studied separately. Hence, detailed investigations into the combined interactions of miRNAs are warranted. This holds, for example, for the combined transfections of miR-424-5p and miR-142-3p, which cause separately greater antiproliferation, apoptosis, and autophagy in breast cancer cells than independent transfections [28]. Additionally, miR-34a-5p (miR-34a), miR-449a, and miR-16 are regarded as tumor-suppressor miRNAs. miR-34a and miR-16 are responsible for regulating senescence, autophagy, and apoptosis. miR-449a can control senescence and apoptosis but not autophagy. In cooperation with miR-16 and/or miR-34a-5p, miR-449a induces synergistic autophagy in cervical cancer cells [29]. Consequently, cell functions may be coordinately regulated by a network of miRNAs.
As mentioned, several miRNAs modulating cell functions have been reviewed in detail. However, the involvement of AKT serine-threonine kinase (AKT; protein kinase B (PKB)) and its effectors is less studied and the research lacks systemic organization. We introduce here the AKT (Section 2) and AKT effectors (Section 3) to connect the miRNA-mediated regulation of cell functions. The impacts of miRNAs (Section 4) and AKT- and AKT effector-targeting miRNAs (Section 5) on the treatment of cancer cells with natural products are particularly emphasized.

2. Relationship between miRNA, AKT, and Cell Functions

Many types of cancer cells show high expression of AKT [30], which controls sophisticated cell functions (apoptosis, autophagy, ER stress, ferroptosis, necroptosis, DDR, senescence, and migration) [31,32,33,34] by cooperating with its downstream AKT effectors, such as forkhead box transcription factors (FOXO); c-Myc, the mechanistic target of rapamycin complex 1/2 (mTORC1/2); the mechanistic target of the rapamycin kinase (mTOR) substrate S6 kinase 1/2 (S6K1/2; PRS6KB1/2); sterol regulatory element-binding protein 1 (SREBP1); eukaryotic translation initiation factor 4E-binding protein 1 (4EBP1); and hypoxia-inducible factor (HIF) [33,34]. Furthermore, mTORC1 is composed of mTOR, RPTOR (regulatory-associated protein of mTOR complex 1; RAPTOR), mTOR-associated protein, LST8 homolog (MLST8), AKT1S1 (AKT1 substrate 1; PRAS40), and DEP domain-containing mTOR-interacting protein (DEPTOR).
The various noncoding RNAs (ncRNAs), such as lncRNAs and circular RNAs, targeting AKT and its effectors in the regulation of a range of cell functions are well-reviewed [25,34]. However, the connections involving other ncRNAs, such as between miRNAs and AKT/AKT signaling, lack comprehensive and systemic integration. Recently, the impact of miRNAs in the regulation of AKT signaling has been reported in several cancer studies, such as for breast [35], glioblastoma [36], and nasopharyngeal [37] cancer cells. Moreover, miRNA and AKT signaling exhibit a reciprocal interaction in cancer cells [38]. Consequently, this review explored the relationship between miRNAs, AKT, and AKT effectors in connection to cell functions (Figure 1).
In the following (Section 2.1, Section 2.2, Section 2.3, Section 2.4, Section 2.5, Section 2.6, Section 2.7 and Section 2.8), AKT is connected to miRNA-mediated cell functions (Table 1).

2.1. miRNAs Targeting AKT-Mediated Pathways Regulate Apoptosis in Cancer Cells

Several miRNAs show apoptosis-modulating effects on cancer cells (Table 1). miR-330-3p (miR-330) shows lower expression in prostate cancer (PC-3) cells than normal cells. miR-330-3p can target several genes, such as E2F transcription factor 1 (E2F1) [39] and coatomer protein complex subunit beta2 (COPB2) [40], and regulate cancer cell proliferation. For example, miR-330-3p targets the 3′-untranslated region (3′-UTR) of E2F1. miR-330-3p triggers apoptosis of prostate cancer cells by downregulating E2F1 and AKT phosphorylation [39]. Furthermore, lysine demethylase 5A (KDM5A), a histone demethylase, is overexpressed in prostate cancer [40]. KDM5A silencing inhibits the proliferation and migration of prostate cancer cells. miR-330-3p targets the COPB2 gene. KDM5A enhances prostate cancer (PRAD) cell proliferation by downregulating miR-330-3p and activating COPB2 and AKT expression [40].
Table 1. Connecting AKT to miRNA-regulated cell functions.
Table 1. Connecting AKT to miRNA-regulated cell functions.
ApoptosisAutophagyER StressFerroptosisNecroptosisDNA Damage ResponseSenescenceMigration
upregulationmiR-330-3p [40] ( AKT, prostate ca)miR-21-5p [41] ( AKT, glioma) miR-374b [42] ( AKT, colon ca)miR-22-3p [43] ( AKT3, endothelial cells)miR-136 [44] ( AKT, gastric ca)
miR-641 [45] ( AKT, breast ca)miR-29a-3p [46] ( AKT, osteosarcoma) miR-22-3p [47] ( AKT)miR-181a-5p and miR-182-5p [48] ( AKT3, uterine ca)
miR-107 [49] ( AKT, hypopharyngeal ca) miR-34a-5p [50] ( AKT, chondrosarcoma)
miR-373-3p [51] ( AKT, liver ca) miR-145-5p [52] (vascular smooth muscle)
miR-1306-5p [53] ( AKT, colon ca)
miR-181c [54] ( AKT, resistant ovarian ca)
downregulationmiR-14669 [55] ( AKT, colon ca) miR-495-3p [56] ( AKT, breast ca)miR-7-5p [57] (oral ca)miR-26a-5p [58] ( AKT, chicken kidney) miR-495-3p [56] ( AKT, breast ca)
miR-126-3p [59] ( AKT, lung ca) miRNA-124-3p [60] (renal ca) miR-200c-3p [61] ( AKT, endometrial ca)
miR-381-3p [62] (renal ca) miR-148a [63] ( AKT, liver ca)
miR-29b-3p [64] (colon ca) miR-107 [49] ( AKT, hypopharyngeal ca)
miR-373-3p [51] ( AKT, liver ca)
miR-1306-5p [53] ( AKT, liver ca)
ca, cancer; , enhance or activate; , inhibit or inactivate. AKT or tissue information is not shown for some miRNAs because they were not provided in the literature collected by searching PubMed.
Several miRNAs are expressed at low levels in some cancer tissues. Overexpression of nonabundant miRNAs triggers apoptosis in cancer cells (Table 1). miR-641 is expressed at a low level and NUCKS1 shows a high level in breast cancer cells (Hs-578T, MCF7, HCC1937, and MAD-MB-231) [45]. Nuclear casein kinase and cyclin-dependent kinase substrate 1 (NUCKS1), a phosphoinositide 3-kinase (PI3K)/AKT enhancer, targets miR-641. miR-641 inhibits breast cancer cell migration and enhances apoptosis, which is reversed by PI3K inhibitor, suggesting miR-641-induced apoptosis through the targeting of the NUCKS1/PI3K/AKT axis [45]. miR-107 is downregulated in hypopharyngeal cancer compared to normal control tissues. miR-107 overexpression causes antiproliferative effects and apoptosis in hypopharyngeal cancer cells through the inactivation of PI3K/AKT [49]. miR-373-3p is downregulated [51] and transcription factor-activating enhancer-binding protein 4 (TFAP4) is upregulated [65] in liver cancer tissues. TFAP4 activates PI3K/AKT expression [65]. Upregulating miR-373-3p triggers apoptosis in liver cancer cells by downregulating AKT and TFAP4, which is reversed by TFAP4 overexpression [51]. miR-1306-5p is downregulated in colon cancer cells and tissues [53]. miR-1306-5p targets solute carrier organic anion transporter family member 2A1 (SLCO2A1). Overexpressing miR-1306-5p suppresses AKT signaling. Overexpressing miR-1306-5p inhibits proliferation and enhances apoptosis in colon cancer cells by inactivating AKT and targeting SLCO2A1 [53].
Several miRNAs reduce drug resistance and induce apoptosis in cancer cells (Table 1). Paclitaxel is a first-line drug for ovarian cancer therapy, but its resistance problem limits its treatment effects [66]. miR-181c binds to the 3′-UTR of glucose-regulated protein 78 (GRP78) and downregulates its expression, suppressing the resistance of ovarian cancer (SKOV3-PTX) cells toward paclitaxel by inducing apoptosis [54]. miR-181c is downregulated in paclitaxel-resistant ovarian cancer cells. In contrast, upregulating miR-181c alleviates this paclitaxel resistance by inactivating AKT [54]. Therefore, miR-181c reduces the paclitaxel resistance involving AKT.
In contrast, some miRNAs inhibit apoptosis in cancer cells (Table 1). miR-126-3p (miR-126) silencing upregulates apoptosis gene expression and inhibits the proliferation of lung cancer cells by upregulating p-PI3K and p-mTOR expression [59]. Ginsenoside Rg1 shows antiproliferative and apoptotic effects in lung cancer (A540) cells by downregulating miR-126-3p [59]. This warrants a detailed assessment of the apoptotic role of miR-126-3p and AKT signaling in ginsenoside Rg1-treated lung cancer cells.
Other apoptosis-inhibitory miRNAs, such as miR-14669, enhance drug resistance (Table 1). Drug resistance in breast cancer can be increased due to the activation of the PI3K/AKT/mTOR axis [67]. miRNA-mediated apoptosis in colon cancer cells and PI3K/AKT signaling may partly contribute to drug resistance [55]. Vincristine, which inhibits miR-14669, induces apoptosis in colon cancer cells and alleviates drug resistance by inactivating PI3K/AKT signaling, which is reversed by the upregulation of miRNA-14669 [55]. As mentioned above, some miRNAs regulate apoptosis with the involvement of AKT.

2.2. miRNAs Targeting AKT Regulate Autophagy in Cancer Cells

Several miRNAs show autophagy-modulating effects in cancer cells (Table 1). miR-21-5p (miR-21) is an oncogenic miRNA in malignant glioma. miR-21-5p silencing by antisense oligonucleotide reduces radioresistance and AKT phosphorylation in γ-ray irradiated glioma cells, which is reversed by miR-21-5p overexpression [41]. Moreover, miR-21-5p silencing also induces autophagosome formation for autophagy. Hence, miR-21-5p stimulates AKT activation and suppresses autophagy, conferring radioresistance in glioma cells.
Some miRNAs show tumor-suppressive effects in cancer cells (Table 1). miR-29a-3p, a tumor-suppressive miRNA, exhibits low levels in osteosarcoma cells compared to normal cells [46]. Insulin-like growth factor 1 (IGF1) is a target of miR-29a-3p. IGF1 exerts oncogenic effects in osteosarcoma by inducing IGF-1R/PI3K/AKT activation, which is reversed by upregulating miR-29a-3p, triggering autophagy and antiproliferative effects [46]. Accordingly, some miRNAs regulate autophagy with the involvement of AKT.

2.3. miRNAs Targeting AKT Regulate ER Stress in Cancer Cells

Several miRNAs show ER stress-modulating effects in cancer cells (Table 1). miR-495-3p mimics suppress the expression of ER chaperone GRP78 and inhibit the proliferation and migration of breast cancer (MDA-MB-231) cells, reducing pirarubicin resistance by inactivating AKT expression, which is reversed by miR-495-3p inhibition [56]. Although studies on the targeting of ER stress-associated AKT by miRNAs are rare (Table 1), the literature on ER stress-associated AKT effectors targeted by various miRNAs is discussed later (Section 3.3).

2.4. miRNAs Targeting AKT Regulate Ferroptosis in Cancer Cells

Several miRNAs show ferroptosis-modulating effects in cancer cells (Table 1). Clinically relevant radioresistant oral cancer cells [68] exhibit miR-7-5p (miR-7) overexpression, which is reversed by miR-7-5p inhibition [57]. miR-7-5p silencing enhances ROS and intracellular Fe2+ content and upregulates ferroptosis gene (arachidonate 12-lipoxygenase, 12S type; ALOX12) expression and lipid peroxidation [57]. Although the reported study did not assess the involvement of AKT, miR-7-5p was predicted to target AKT3 according to the miRDB database (retrieval date: 12 October 2022) [69]. This warrants a detailed evaluation of the role of miR-7-5p in regulating ferroptosis through the targeting of AKT in the future. Although studies on the targeting of ferroptosis-associated AKT by miRNA are rare (Table 1), the literature on ferroptosis-associated AKT effectors targeted by various miRNAs is discussed later (Section 3.4).

2.5. miRNAs Targeting AKT Regulate Necroptosis in Cancer Cells

Several miRNAs show necroptosis-modulating effects in noncancer cells (Table 1). Selenium yeast (Se-Y), an organic selenium source, suppresses cadmium-induced necroptosis in chicken kidney by downregulating RIP1, RIP3, and MLKL [58]. Moreover, Se-Y upregulates miR-26a-5p expression, causing PTEN downregulation and leading to AKT upregulation. Accordingly, miR-26a-5p may regulate necroptosis with the involvement of AKT.
Several miRNAs also show necroptosis-modulating effects in cancer cells (Table 1). miRNAs such as miRNA-124-3p (miRNA-124), miR-381-3p, and miR-29b-3p (miR-29b) are reported to show the ability to modulate necroptosis [62,64,69]. However, the role of AKT in this has not been investigated as yet. Cisplatin upregulates miRNA-124-3p in renal cancer (Caki-1) cells [60]. Calpain small subunit 1 (CAPNS1; CAPN4) is a target of miRNA-124-3p. Upregulating miRNA-124-3p alleviates cisplatin sensitivity, and cisplatin-induced necroptosis of renal cancer cells is reversed by CAPN4 overexpression [60]. miR-381-3p overexpression enhances cell proliferation and suppresses TNF-induced apoptosis and necroptosis of renal cancer cells [62]. Papillary renal cancer patients show high miR-381-3p expression and exhibit poor survival. Hence, miR-381-3p plays an oncogenic role in suppressing apoptosis and necroptosis. Additionally, miR-29b-3p is overexpressed, and TRAF5 is downregulated in colon cancer [64]. TNF receptor-associated factor 5 (TRAF5), a regulator of necroptosis, is a target of miR-29b-3p. Consequently, miR-29b-3p suppresses 5-fluorouracil-promoted necroptosis of colon cancer cells and enhances its resistance by downregulating TRAF5 [64].
Notably, after data mining using the miRDB database (retrieval date: 12 October 2022) [69], miRNA-124-3p and miR-29b-3p were predicted to target AKT2 and AKT3, while miR-381-3p was predicted to target AKT3. Consequently, this warrants a careful examination of the role of AKT in miRNA-modulating (miRNA-124-3p, miR-381-3p, and miR-29b-3p) necroptosis in the future.

2.6. miRNAs Targeting AKT Regulate DDR in Cancer Cells

Several miRNAs show DDR-modulating effects, such as DNA damage in cancer cells (Table 1). Bleomycin causes DNA damage in colon cancer (HCT116 and HT29) cells accompanied by downregulation of AKT1 protein expression, which is reversed by p53 and miR-374b knockdown [42]. Overexpressing p53 promotes expression of the bleomycin-induced AKT1 regulator miR-374b, which is reversed by p53 knockdown. Hence, the p53/miR-374b/AKT1 axis regulates bleomycin-induced DNA damage in colon cancer cells.
In osteosarcoma cells, AKT1 and senescence suppress the DDR function, such as homologous recombination for DNA repair, of the mediator of DNA damage checkpoint 1 (MDC1) by overexpressing miR-22-3p (miR-22) [47]. Therefore, senescence downregulates MDC1 by upregulating miR-22-3p and downregulating DNA repair involving AKT activation. Accordingly, some miRNAs inhibit DDR with the involvement of AKT.

2.7. miRNAs Targeting AKT Regulate Senescence in Cancer Cells

Several miRNAs show senescence-promoting effects in cancer cells (Table 1). miR-22-3p is overexpressed in aged endothelial progenitor cells [43]. AKT3 is a target of miR-22-3p. Upregulating miR-22-3p in young endothelial progenitor cells causes senescence and suppresses proliferation, migration, and angiogenesis, which is reversed by miR-22-3p silencing and AKT3 upregulation [43]. Hence, miR-22-3p regulates senescence with the involvement of AKT3. In addition to miR-22-3p, AKT3 is also regulated by other miRNAs. For example, the overexpression of miR-181a-5p (miR-181a) and miR-182-5p (miR-182) causes senescence in uterine leiomyoma cells by suppressing AKT3 and CCND2, respectively [48].
miR-34a-5p upregulation commonly occurs in the cartilage of osteoarthritis patients [50]. Delta-like protein 1 (DLL1) mRNA is a target of miR-34a-5p. DLL1 overexpression upregulates AKT phosphorylation (activation). Overexpressing miR-34a-5p has antiproliferative effects on and causes the senescence of chondrosarcoma cells by downregulating DLL1 and phosphorylated AKT, which is reversed by an miR-34a-5p inhibitor (Table 1) [50]. Hence, miR-34a-5p regulates senescence with the involvement of AKT. miR-145-5p (miR-145) overexpression is associated with DNA damage and senescence in vascular smooth muscle cells [52]. Moreover, miR-145-5p was predicted to target AKT3 in accordance with the miRDB database (retrieval date: 12 October 2022) [69]. This warrants an advanced investigation of the AKT response in miR-145-5p-modulating senescence in the future. Although studies on the targeting of senescence-associated AKT by miRNA are rare (Table 1), the literature on senescence-associated AKT effectors targeted by several miRNAs is discussed later (Section 3.7).

2.8. miRNA Targeting AKT Regulates Migration in Cancer Cells

Several miRNAs function as tumor-suppressor miRNAs inhibiting cancer cell migration (Table 1). miR-200c-3p (miR-200c) suppresses EMT and dephosphorylates AKT in endometrial cancer cells by targeting the BMI1 proto-oncogene polycomb ring finger (BMI-1) gene, which is reversed by miR-200c inhibition [61]. miR-148a shows low expression in liver cancer tissues [63]. Death receptor-5 (DR-5) is a target of miR-148a. Overexpressing miR-148a suppresses proliferation, migration, and invasion and enhances apoptosis in liver cancer cells by downregulating EMT and AKT through the targeting of DR-5 [63]. Upregulating oncogenic miR-107 inhibits the migration and invasion of hypopharyngeal cancer (FaDu) cells by inactivating AKT [49]. miR-373-3p functions as a tumor-suppressor miRNA and is generally downregulated in liver cancer cells. Overexpressing miR-373-3p inhibits liver cancer (Huh7, HLE, HCCLM6, and HCCLM3) cell migration, metastasis, and EMT by downregulating AKT activation [51]. miR-1306-5p, a tumor-suppressor miRNA, is expressed at a low level in colon cancer. Upregulating miR-1306-5p suppresses migration and invasion of colon cancer cells by inhibiting AKT activation [53].
In contrast, some miRNAs function as oncogenic miRNAs that inhibit cancer cell migration (Table 1). miR-136 is overexpressed in gastric cancer tissues and cells [44]. Downregulating miR-136 promotes antiproliferative effects and suppresses invasion in gastric cancer (MGC-803 and SGC-7901) cells. PTEN, a target of miR-136 and an inhibitor of AKT, shows low expression in gastric cancer tissues. miR-136 inhibition downregulates AKT phosphorylation [44]. Accordingly, some miRNAs regulate migration with the involvement of AKT.

3. Relationship between miRNA, AKT Effectors, and Cell Functions

In the following (Section 3.1, Section 3.2, Section 3.3, Section 3.4, Section 3.5, Section 3.6, Section 3.7 and Section 3.8), AKT effectors are connected to miRNA-mediated cell functions (Table 2).

3.1. miRNAs Targeting AKT Effectors Regulate Apoptosis in Cancer Cells

3.1.1. FOXO-Targeting miRNAs and Apoptosis

FOXO plays a crucial role in promoting apoptosis by inducing proapoptotic protein expressions [70,71]. Several miRNAs regulate apoptosis in cancer cells with the involvement of FOXO (Table 2). miR-181a-5p is highly expressed in cervical cancer cells and is an oncogenic miRNA. Downregulating miR-181a-5p inhibits proliferation and invasion and triggers apoptosis in cervical cancer (HeLa) cells by upregulating PTEN and downregulating AKT and FOXO1 [72]. miR-335-5p (miR-335) shows a lower level in liver cancer cells than normal. miR-335-5p binds FOXO3a and targets FOXO3a 3′-UTR promoter activity. The miR-335-5p mimic enhances antiproliferative effects and apoptosis in liver cancer cells, which is reversed by the miR-335-5p inhibitor. Consequently, miR-335-5p triggers apoptosis in liver cancer (SMMC7721) cells by targeting FOXO3a [73].
Table 2. Connecting AKT effectors to miRNA-regulated cell functions.
Table 2. Connecting AKT effectors to miRNA-regulated cell functions.
AKT EffectorsApoptosisAutophagyER StressFerroptosisNecroptosisDNA Damage ResponseSenescenceMigration
FOXOmiR-181a-5p [72] ( \ ), cervical camiR-223-3p [74] ( \ ), liver camiR-132-3p [1] ( \ ), colon camiR-670-3p [75] ( \ ), liver camiR-6852-5p [76] ( \ ), cervical camiR-223-3p [77] ( \ damage), breast camiR-34a-5p [78] ( \ ), endothelial cellsmiR-135a [79] ( \ ), liver ca
miR-335-5p [73] ( \ ), liver ca miR-494-3p [80] ( FOXO3\ ) miR-96 [81,82] \ repair), breast ca
c-MycmiR-196b-5p [83] ( \ ), endometriotic stromal cellsmiR-27b-3p [84] ( \ ), resistant colon camiR-1291 [85] ( \ ), prostate camiR-25-3p [86] ( \ ), prostate camiR-494-3p [87] ( \ ), ovarian camiR-1245 [88] ( \ repair), breast camiR-34a-5p [89] ( \ ), liver camiR-33b-5p [90] ( \ ), osteosarcoma
miR-150-5p [91] ( \ ), lung ca miR-449a [92] ( \+ damage), prostate ca miR-34a-5p [93] ( \ ), bladder ca
mTORC1mi-21-5p [94] ( mTOR\ ), renal cancer cellsmiR-126-3p [95] ( mTOR\ ), colon camiR-99b-5p and miR-100-5p [96] mTOR\ )miR-137 [97] ( mTOR\ ), melanomamiR-32 [98,99] ( mTOR\ ), epithelialmiR-18a-5p and miR-421-3p [100] ( mTOR\ )miR-100-5p and miR-101-3p ( mTOR\ ), endothelial cellsmiR-520a-3p [101] ( mTOR\ ), lung ca
miR-1908 [102] AKT1S1\ ), lung camiR-56 [103] ( AKT1S1\ ), glioblastoma miR-107 [104] ( mTOR\ ), endothelial cellsmiR-155-5p [105] ( DEPTOR\ ), lymphoma
miR-155-5p [105] ( DEPTOR\ ), B-cell lymphomamiR-125a-5p [106] ( AKT1S1\ ), breast ca
miR-182-5p [107] ( DEPTOR\ ), intestine I/R
S6K1/2XmiR-495-3p [108] ( \ ), gastric caXXXXXX
SREBP1miR-185-5p and miR-342 [109] ( \ ), prostate camiR-34a-5p [110] ( \ ), fat liver
miR-132-3p [111] ( \ ), glioma cellsmiR-33 [112] ( \ ), macrophagesmiR-34a-5p [110] ( \ ), fat livermiR-670-3p [75] ( \ ), liver caXXmiR-21-5p [113] ( \ ), prostate camiR-18a-5p [114] ( \ ), breast ca
4EBP1miR-149-3p [115] ( \ ), T-ALLmiR-495-3p [108] ( \ ), gastric caXmiR-1911-3p [116] ( \ ), lung caXXmiR-125a-5p and miR-125b-5p [117] ( \ ), ovarian camiR-125a-5p and miR-125b-5p ( \ ), ovarian ca
miR-101-3p [118] ( \ ), endometrial ca
HIFmiR-199a-5p [119] ( \ ), hemangioma-derived endothelial cellsmiR-210 [120] ( \ ), colon camiR-205-5p [121] ( \ ), renal tubule cellsmiR-147a [122] ( \ ), glioblastomamiR-210 ( \ and miR-383 ( \ ), [123] macrophagemiR-210 and miR-373 [124] ( \ ), breast camiR-126-3p [125] ( \ ), endothelial cellsmiR-200c [126] ( \ ), lung ca
ca, cancer; , enhance or activate; , inhibit or inactivate; left\right (e.g., \ or \ ), AKT effectors\cell function. mTORC1 is composed of mTOR, RPTOR, MLST8, AKT1S1, and DEPTOR. X indicates that miRNAs were rarely reported for these cell functions. X, data not available.

3.1.2. c-Myc-Targeting miRNAs and Apoptosis

Several miRNAs regulate apoptosis in cancer cells with the involvement of c-Myc (Table 2). miR-196b-5p (miR-196b) overexpression suppresses proliferation and induces apoptosis in endometriotic stromal cells by targeting and downregulating c-Myc and Bcl-2 mRNA expression [83].

3.1.3. mTOR-, AKT1S1-, and DEPTOR-Targeting miRNAs and Apoptosis

Several miRNAs regulate apoptosis in cancer cells with the involvement of MTORC1 (Table 2), which composed of mTOR, RPTOR, MLST8, and AKT1S1. miR-21-5p is overexpressed in renal cancer (ACHN) cells [94]. miR-21-5p silencing causes the apoptotic activation of caspase 3 and the dephosphorylation of signal transduction and activators of transcription 3 (STAT3). miR-21-5p enhances proliferation and suppresses apoptosis in renal cancer cells, which is reversed by mTOR inhibitor (XL388) [94].
The ribosomal protein-p53 pathway, a critical function for regulating apoptosis and senescence in cancer cells, is activated by the miR-1908 mimic and inhibited by AKT1S1 [127]. AKT1S1 downregulation promotes antiproliferative effects and apoptosis in melanoma (UACC 903) cells (Table 2) [102]. miR-1908 is downregulated in lung cancer cells. miR-1908 can target AKT1S1 and downregulate ribosomal protein L11 (RPL11)-p53-p21 signaling [127]. Consequently, miR-1908 may induce apoptosis in lung cancer cells by targeting AKT1S1.
DEPTOR is a potential target of miR-155-5p (miR-155) [105], which regulates B-cell development, lymphomagenesis, and differentiation [128]. Inhibiting miR-155-5p enhances the sensitivity of ibrutinib-triggered apoptosis in diffuse large B-cell lymphoma cells to ibrutinib (Table 2) [105].

3.1.4. SREBP1-Targeting miRNAs and Apoptosis

Several miRNAs regulate apoptosis in cancer cells with the involvement of SREBP1 (Table 2). SREBP1 is upregulated and miR-185-5p (miR-185) and miR-342-3p (miR-342) are downregulated in prostate cancer (LNCaP) cells compared to normal cells [109]. miR-185-5p and miR-342 suppress SREBP-1 and SREBP-2 expression, causing the inhibition of cell proliferation and migration in prostate cancer cells. Overexpression of miR-185-5p and miR-342-3p causes apoptosis in prostate cancer cells [109]. Additionally, miR-132-3p (miR-132) downregulates SREBP-1c expression; inhibits proliferation, migration, and invasion; and triggers apoptosis in glioma (U251) cells [111]. Hence, several miRNAs modulate SREBP1 expression in the regulation of apoptosis.

3.1.5. EBP1-Targeting miRNAs and Apoptosis

Several miRNAs inhibit apoptosis in cancer cells with the involvement of 4EBP1 (Table 2). miR-149-3p (miR-149*) is overexpressed in T-cell acute lymphoblastic leukemia (T-ALL) [115]. miR-149-3p mimics enhance T-ALL cell proliferation and suppress apoptosis in T-ALL cells by upregulating 4EBP1 and S6K (PRS6KB1), which is reversed by miR-149-3p inhibitors [115]. In contrast, several miRNAs promote apoptosis in cancer cells with the involvement of SREBP1. miR-101-3p (miR-101) is downregulated and mTOR and 4EBP1 are upregulated in endometrial cancer cells, promoting proliferation and invasion and suppressing apoptosis, which is reversed by the miR-101-3p mimic and/or si-mTOR [118]. Hence, several miRNAs modulate 4EBP1 expression in the regulation of apoptosis.

3.1.6. HIF1A-Targeting miRNAs and Apoptosis

Several miRNAs inhibit apoptosis in cancer cells with the involvement of HIF1A (Table 2). Proliferating hemangioma (HDEC) cells downregulate miR-199a-5p (miR-199a) and upregulate HIF1A, a target of miR-199a-5p [119]. The miR-199a-5p mimic suppresses the proliferation of hemangioma cells, which is reversed by HIF1A overexpression. The overexpression of miR-199a-5p causes antiproliferative effects and apoptosis in hemangioma-derived endothelial cells [119]. Hence, miRNA may regulate apoptosis with the involvement of HIF1A.

3.2. miRNAs Targeting AKT Effectors Regulate Autophagy in Cancer Cells

3.2.1. FOXO- and c-Myc-Targeting miRNAs and Autophagy

Several miRNAs regulate apoptosis in cancer cells with the involvement of FOXO (Table 2). Doxorubicin downregulates miR-223-3p (miR-223) in liver cancer cells. FOXO3a is a direct target of miR-223-3p. miR-223-3p overexpression downregulates FOXO3a and suppresses doxorubicin-triggered autophagy, leading to chemoresistance [74].
Several miRNAs regulate apoptosis in cancer cells with the involvement of c-Myc (Table 2). After oxaliplatin treatment, resistant colon cancer cells show lower miR-27b-3p expression and higher autophagy than parental cells [84]. c-Myc downregulates miR-27b-3p expression. Hence, miR-27b-3p inhibits autophagy by downregulating c-Myc. Accordingly, c-Myc/miR-27b-3p signaling confers oxaliplatin resistance to colon cancer cells [84]. Furthermore, c-Myc also modulates other miRNAs. c-Myc and miR-150-5p (miR-150) are highly expressed in lung cancer (A549) cells [91]. c-Myc functions as a miR-150-5p transcriptional factor for upregulating miR-150-5p expression. miR-150-5p inhibits autophagy and enhances the proliferation of lung cancer cells. Consequently, the inhibition of c-Myc suppresses the proliferation of lung cancer cells.

3.2.2. mTOR-, AKT1S1-, and DEPTOR-Targeting miRNAs and Autophagy

Several miRNAs induce the autophagy of cancer cells with the involvement of mTORC1 (Table 2). miR-126-3p shows a lower level in colon cancer tissues and cells than normal. miR-126-3p binds to 3′-UTR to downregulate mTOR expression. The overexpression of miR-126-3p inhibits proliferation and mTOR expression and induces autophagy and apoptosis in colon cancer cells, which is reversed by autophagy inhibitor bafilomycin A1, suggesting that miR-126-3p-induced apoptosis depends on autophagy through the modulation of mTOR expression [95].
In contrast, several miRNAs inhibit the autophagy of cancer cells with the involvement of mTORC1 (Table 2). miR-56 can directly target the 3′-UTR of AKT1S1. Overexpressing miR-56 enhances proliferation and suppresses autophagy in glioblastoma cells [103]. Hypermethylation of the CpG island located in the miR-125a-5p (miR-125a) promoter occurs in breast cancer tissues and cells, leading to miR-125a-5p downregulation, which is reversed by 5-Aza-dC, a methylation inhibitor. miR-125a-5p overexpression inhibits autophagy and enhances proliferation in breast cancer cells [106]. miR-125a-5p was predicted to target AKT effectors, such as AKT1S1, in accordance with the miRDB database [69] (retrieval date: 12 October 2022).
Intestinal ischemia/reperfusion (I/R) inhibits miR-182-5p expression. Overexpression of miR-182-5p inhibits autophagy and DEPTOR expression, alleviating intestinal damage under I/R (Table 2) [107]. This warrants an examination of DEPTOR in cancer cells.

3.2.3. SREBP1-, 4EBP1-, S6K-, and HIF1A-Targeting miRNAs and Autophagy

Several miRNAs regulate the autophagy of cancer cells with the involvement of SREBP1 (Table 2). miR-33 has two isoforms, miR-33a-5p and miR-33b-5p (miR-33b), derived from intron 16 of SREBP2 and intron 17 of SREBP1 genes, respectively [129]. miR-33 can modulate the autophagy of macrophages by inhibiting several autophagy-related gene expressions [112]. In a high-fat-diet rat model, autophagy was induced in liver tissues by upregulating miR-34a-5p and SREBP1c [110].
Several miRNAs inhibit the autophagy of cancer cells with the involvement of 4EBP1 and S6K (Table 2). mTOR is an upstream regulator of autophagy that inhibits autophagosome formation. Multiple-drug-resistance (MDR) cells exhibit greater autophagy than parent cells. miR-495-3p upregulation suppresses autophagy in MDR gastric cancer (SGC7901) cells by phosphorylating mTOR substrates, such as 4EBP1 and PRS6KB1 (S6K1) [108]. Accordingly, miR-495-3p alleviates MDR by activating mTOR and inhibiting autophagy.
Under hypoxia, miR-210 is overexpressed and induces autophagy and radioresistance in colon cancer cells by upregulating HIF1A (Table 2) [120]. Hence, miRNA may regulate autophagy with the involvement of HIF1A. This warrants further identification of HIF1A-targeting miRNAs in the regulation of autophagy.

3.3. miRNAs Targeting AKT Effectors Regulate ER Stress in Cancer Cells

3.3.1. FOXO-Targeting miRNAs and ER Stress

Several miRNAs control ER stress in digestive cancer with the involvement of FOXO (Table 2) [1]. miR-132-3p is overexpressed in gastric cancers [130]. Upregulated miR-132-3p promotes gastric cancer (AGS) cell growth by binding to the 3′-UTR of FOXO1 mRNA and, consequently, blocking FOXO1 function [130]. Folate depletion upregulates miR-132-3p and downregulates ER stress-associated gene expression in colon cancer cells [1]. Accordingly, FOXO1 is involved in miR-132-3p-regulated ER stress in cancer development.
Several studies have shown interactions between ER stress and miRNAs involving FOXO (Table 2). miR-494-3p is upregulated by ER stress inducers, such as tunicamycin and thapsigargin. miR-494-3p pretreatment suppresses tunicamycin-induced ER stress and promotes cell proliferation, which is reversed by miR-494-3p inhibition [80]. Hence, miR-494-3p shows reciprocal regulation with ER stress. Although this study did not assess the involvement of AKT effectors, miR-494-3p was predicted to target FOXO3 in accordance with the miRDB database (retrieval date: 12 October 2022) [69]. This warrants a detailed evaluation of miR-494-3p in the regulation of ferroptosis through the targeting of AKT effectors in the future.

3.3.2. c-Myc-Targeting miRNAs and ER Stress

Several miRNAs modulate ER stress in cancer cells with the involvement of c-Myc (Table 2). IRE1α, an ER stress sensor, is directly targeted by miR-1291 [85]. IRE1α enhances prostate tumor growth by activating c-Myc [131]. The inhibition of IRE1α downregulates c-Myc expression in natural killer cells. Accordingly, c-Myc is involved in miRNA-regulated ER stress.

3.3.3. mTOR- and DEPTOR-Targeting miRNAs and ER Stress

Several miRNAs regulate ER stress in cancer cells with the involvement of mTORC1 (Table 2). mTOR, one of the targets of miR-99b-5p and miR-100-5p (miR-100), can trigger amyloid β-induced apoptosis by causing ER stress [96]. DEPTOR silencing causes multiple myeloma death without activating the UPR [132]. The role of DEPTOR in regulating ER stress in conjunction with miRNAs needs further investigation.

3.3.4. SREBP1- and HIF1A-Targeting miRNAs and ER Stress

Several miRNAs regulate ER stress with the involvement of SREBP1 and HIF1A (Table 2). In a high-fat-diet rat model, liver miR-34a-5p was overexpressed and induced ER stress by downregulating sirtuin 1 (SIRT1), upregulating SREBP1c and GRP78 [110]. ER stress downregulates miR-205-5p (miR-205) and inhibits HIF1 expression in renal tubule cells [121]. This warrants a detailed investigation of SREBP1- and HIF1A-targeting miRNAs and their regulation of cancer cell autophagy.

3.4. miRNAs Targeting AKT Effectors Regulate Ferroptosis in Cancer Cells

3.4.1. FOXO-, c-Myc-, and SREBP1-Targeting miRNAs and Ferroptosis

Several miRNAs modulate ferroptosis in cancer cells with the involvement of FOXO, c-Myc, and SREBP1 (Table 2). miR-670-3p, overexpressed in human glioblastoma [133], blocks ferroptosis by targeting acyl-CoA synthase long-chain family member 4 (ACSL4) [133]. Moreover, ACSL4 downregulates FOXO1 through ubiquitination and acetylation in β-cell-specific Rictor-knockout islets [75]. SREBP1 induces ACSL4 and lipogenesis-associated genes to improve the metastasis of liver cancer cells with c-Myc [134]. Additionally, the overexpression of lncRNA PCAT1 promotes docetaxel resistance and suppresses ferroptosis in prostate cancer (PC-3) cells by upregulating c-Myc and solute carrier family 7 member 11 (SLC7A11) [86]. SLC7A11 can compete with miR-25-3p and bind to c-Myc to enhance protein stability [86]. Accordingly, FOXO, c-Myc, and SREBP1 are involved in miRNA-regulated ferroptosis.

3.4.2. mTOR-, 4EBP1-, and HIF1A-Targeting miRNAs and Ferroptosis

Several miRNAs modulate ferroptosis in cancer cells with the involvement of mTOR, 4EBP1, and HIF1A (Table 2). miR-137 inhibits ferroptosis by downregulating glutamine transporter SLC1A5 in melanoma (A375) cells [97]. SLC1A5 induces the uptake of glutamine and activates mTORC1 signaling [135]. Additionally, miR-1911-3p mimics downregulate mTORC1 enhancers, such as p-4EBP1, in lung cancer (H1299) cells [116]. Activated mTORC1 suppresses ferroptosis by upregulating 4EBP1 [136]. miR-147a, induced by HIF1A [137], binds to the 3′-UTR of solute carrier family 40 member 1 (SLC40A1) to trigger ferroptosis in glioblastoma (U87MG) cells [122]. Accordingly, mTOR, 4EBP1, and HIF1A are involved in miRNA-regulated ferroptosis.

3.5. miRNAs Targeting AKT Effectors Regulate Necroptosis in Cancer Cells

3.5.1. FOXO- and c-Myc-Targeting miRNAs and Necroptosis

Several miRNAs modulate necroptosis in cancer cells with the involvement of FOXO and c-Myc (Table 2). Interleukin-27 upregulates miR-6852-5p and causes necrosis in cervical cancer cells by inhibiting Forkhead box protein M1 (FOXM1) expression [76], with the FOXM1 being suppressed by FOXO3a [138]. Upregulated miR-494-3p inhibits the expression of receptor-interacting serine/threonine-protein kinase 1 (RIPK1) in epilepsy rats [139], which is a necroptosis regulator [140]. Moreover, miR-494-3p targets the c-Myc and SIRT1 genes in ovarian cancer (OVCAR3) cells [87]. Accordingly, FOXO and c-Myc are involved in miRNA-regulated necroptosis.

3.5.2. mTOR- and HIF1A-Targeting miRNAs and Necroptosis

Several miRNAs modulate necroptosis is cancer cells with the involvement of mTOR and HIF1A (Table 2). Antibiotics decrease microbiota levels and downregulate the expression of receptor-interacting serine/threonine-protein kinase 3 (RIPK3), reducing mTOR hyperactivation-enhanced epithelial necroptosis [99]. Moreover, mTOR is downregulated by miR-32 [98]. Hence, the role of miR-32 in mTOR-regulated necroptosis warrants detailed assessments. Furthermore, miR-383 upregulates ATP contents and suppresses necroptosis. HIF1A triggers necroptosis and ATP depletion in inflammatory macrophages by upregulating miR-210 and downregulating miR-383 [123]. Accordingly, mTOR and HIF1A are involved in miRNA-regulated necroptosis.

3.6. miRNAs Targeting AKT Effectors Regulate DDR in Cancer Cells

3.6.1. FOXO-Targeting miRNAs and DDR

Several miRNAs modulate DDR in cancer cells with the involvement of FOXO (Table 2). DNA damage agents, such as N-methyl-N’-nitro-N-nitrosoguanidine (MNNG), upregulate FOXO1 expression in lung cancer cells [141]. miR-223-3p downregulates FOXO1 to promote the proliferation of breast cancer, which is reversed by the miR-223-3p inhibitor [77]. Similarly, miR-96 inhibits FOXO3a expression in breast cancer cells through direct targeting [81]. FOXO3a exhibits a DNA repair function [82]. This warrants detailed assessments to examine the roles of miR-223-3p and miR-96 in FOXO-regulated DDR. Accordingly, FOXO is potentially involved in regulating miRNA-mediated DDR.

3.6.2. c-Myc-Targeting miRNAs and DDR

Several miRNAs modulate DDR in cancer cells with the involvement of c-Myc (Table 2). c-Myc directly targets the miR-1245 promoter and enhances its expression, as well as causing downregulation of BRCA2 DNA repair-associated (BRCA2) gene expression and suppressing the ability for homologous recombination in breast cancer cells [88]. X-rays induce DNA damage, and they upregulate miR-449a but downregulate c-Myc in prostate cancer cells [92]. Accordingly, c-Myc is involved in regulating miRNA-mediated DDR.

3.6.3. mTOR- and HIF1A-Targeting miRNAs and DDR

Several miRNAs modulate DDR in cancer cells with the involvement of mTOR and HIF1A (Table 2). mTORC1, a complex consisting of mTOR, MLST8, AKT1S1, and DEPTOR, activates S6K1 to inhibit ATM, a DNA damage sensor, by upregulating miR-18a-5p and miR-421-3p to target ATM [100]. In hypoxia, miR-210 and miR-373 are upregulated in HIF1A-positive breast cancer cells but not in HIF1A knockdown cells [124]. miR-210 overexpression downregulates the homologous recombination repair protein RAD52. miR-373 overexpression downregulates nucleotide excision repair (NER) proteins, such as RAD23B and RAD52 [124]. Consequently, miR-210 and miR-373 may function as oncogenic miRNAs by suppressing DNA repair.

3.7. miRNAs Targeting AKT Effectors Regulate Senescence in Cancer Cells

3.7.1. FOXO-, c-Myc-, and mTOR-Targeting miRNAs and Senescence

Several miRNAs modulate senescence with the involvement of FOXO, c-Myc, and mTOR (Table 2). The overexpression of miR-34a-5p, a tumor-suppressor miRNA, inhibits SIRT1 and upregulates SIRT1 effector-acetylated FOXO1, suppressing angiogenesis by inducing the senescence of endothelial progenitor cells [78]. Furthermore, oncogene-induced senescence upregulates miR-34a-5p and downregulates c-Myc expression [89]. miR-34a-5p can bind to the 3′-UTR of c-Myc and FOXM1 and downregulate telomerase reverse transcriptase (hTERT) activity, causing senescence in liver cancer cells [89].
mTOR plays a vital role in miRNA-regulated senescence in endothelial cells [142]. miR-100-5p and miR-101-3p function as mTOR suppressors and modulate senescence [142]. Moreover, miR-107 can target PTEN, an MTORC1 suppressor, and upregulate its downstream factor MTORC1, leading to endothelial cell senescence (Table 2) [104].

3.7.2. SREBP1-, 4EBP1-, and HIF1A-Targeting miRNAs and Senescence

Several miRNAs modulate senescence with the involvement of SREBP1, 4EBP1, and HIF1A (Table 2). miR-21-5p inhibition downregulates SREBP1 expression by suppressing insulin receptor substrate 1 (IRS1)-mediated transcription and inducing senescence in prostate cancer cells, which is reversed by miR-21-5p overexpression [113]. Moreover, 4EBP1 silencing triggers p53-dependent senescence [143]. The downregulation of 4EBP1 expression in ovarian cancer (SKOV3) cells causes the upregulation of miR-125a-5p and miR-125b-5p (miR-125b) [117]. HIF1A is a target of miR-126-3p [125]. Both miR-126-3p and HIF1A are downregulated in senescent endothelial cells. Consequently, SREBP1, 4EBP1, and HIF1A potentially regulate miRNA-mediated senescence.

3.8. miRNAs Targeting AKT Effectors Regulate Migration in Cancer Cells

3.8.1. FOXO- and c-Myc-Targeting miRNAs and Migration

Several miRNAs modulate migration with the involvement of FOXO (Table 2). FOXO1 is the direct target of miR-135a [79], which is overexpressed in liver cancer tissues and cells. Overexpression of miR-135a promotes the migration and invasion of liver cancer cells by upregulating MMP2 and downregulating FOXO3a, which is reversed by miR-135a inhibition [79]. Moreover, osteosarcoma tissues and cells exhibit low miR-33b-5p expression. miR-33b-5p inhibits osteosarcoma cell migration and invasion by downregulating c-Myc expression [90]. The CD44 and c-Myc genes are the targets of miR-34a-5p. Overexpression of miR-34a-5p suppresses the invasion of bladder cancer (UMUC3) cells by downregulating CD44 and c-Myc [93]. Accordingly, FOXO and c-Myc are involved in regulating miRNA-mediated cancer cell migration.

3.8.2. mTOR- and DEPTOR-Targeting miRNAs and Migration

Several miRNAs modulate migration with the involvement of mTOR and DEPTOR (Table 2). Overexpressing miR-520a-3p suppresses proliferation, migration, and invasion and triggers apoptosis in lung cancer (NCI-H1975) cells by inactivating AKT and downregulating mTOR, MMP2, and MMP9 [101]. miR-155-5p can directly bind to the 3′-UTR of DEPTOR to inhibit DEPTOR-mediated antimigration in diffuse large B-cell lymphoma cells [105]. Accordingly, MTORC1 is involved in regulating the miRNA-mediated migration of cancer cells.

3.8.3. SREBP1-, 4EBP1-, and HIF1A-Targeting miRNAs and Migration

Several miRNAs modulate migration with the involvement of SREBP1 and 4EBP1 (Table 2). SREBP1, highly expressed in breast cancer, enhances proliferation and metastasis, contributing to the poor survival of breast cancer patients. SREBP1 is modulated by miRNAs, such as miR-18a-5p, which target SREBP1 to suppress EMT status and invasion in breast cancer (MDA-MB-231) [114]. Additionally, miR-125a-5p and miR-125b-5p, potential tumor-suppressor miRNAs, are suppressed in ovarian cancer tissue and cells. The overexpression of miR-125a-5p and miR-125b-5p decreases the invasion and migration of ovarian cancer cells by downregulating 4EBP1 expression [117].
HIF1A can modulate several hypoxia-induced miRNAs (Table 2). miR-200c inhibits HIF1A expression and blocks the migration of lung cancer cells [126]. Therefore, overexpression of miR-200c may exhibit anticancer effects in tumors associated with hypoxia. Accordingly, SREBP1, 4EBP1, and HIF1A regulate the miRNA-mediated migration of cancer cells.

4. Modulation of miRNAs by Natural Products

Several natural products have been reported to modulate oncogenic and tumor-suppressor miRNAs and to control their targeting genes in the regulation of cancer cell proliferation [144,145,146]. Up-to-date, detailed information on the miRNAs, cancer cells, and functions related to these natural products is summarized in Table 3.
As shown in Table 3, ellagitannin, a fruit- and nut-derived polyphenol, has antiproliferative effects in liver cancer cells associated with miRNA modulation [147]. Ellagitannin upregulates 14 miRNAs (miR-526b-5p (miR-526b), miR-452-5p (miR-452), miR-194-5p (miR-194), miR-373-5p (miR-373*), miR-518-5p (miR-518f*), miR-302a-5p (miR-302a*), miR-424-5p (miR-424), let-7e-5p (let-7e), miR-525-5p (miR-525), miR-519e-5p (miR-519e*), miR-513a-3p (miR-513), miR-518c-5p (miR-518c*), miR-512-5p, and miR-346). It also downregulates seven miRNAs (miR-542-3p, let-7d-5p (let-7d), miR-299-3p, miR-200a-5p (miR-200a*), let-7f-5p (let-7f), let-7i-5p (let-7i), and let-7a-5p (let-7a)) [147].
The combined treatment including luteolin, ellagic acid, and punicic acid, which are pomegranate juice components, inhibits the metastasis of prostate cancer cells by enhancing the expression of tumor-suppressor miRNAs (miR-144-3p (miR-144), miR-133b, miR-1-3p (miR-1), miR-122-5p (miR-122), miR-34c-5p (miR-34c), miR-200c-3p, miR-127-3p (miR-127), miR-335-5p, miR-124-3p, miR-181a-5p, miR-7-5p, miR-15a-5p (miR-15a), and let-7d-5p). Moreover, this combined treatment also suppresses the expression of oncogenic miRNAs (miR-20a-5p (miR-20a), miR-21-5p, miR-9, miR-29b-3p, and miR-181b) [148] (Table 3).
A novel Berberis amurensis plant-derived berbamine analog, BBMD3, shows antiproliferative and apoptotic effects in glioblastoma cancer stem cells through the upregulation of miR-4284 [149]. Oridonin, a Rabdosia rubescens-derived diterpenoid, induces apoptosis in leukemia cells by suppressing miR-17-5p (miR-17) and miR-20a-5p [150]. Bufalin, a Chinese toad venom-derived toxin, suppresses differentiation and proliferation and triggers apoptosis in osteosarcoma cancer stem cells by upregulating miR-148a-3p [151]. Betulinic acid, a pentacyclic triterpenoid, suppresses liver cancer cell proliferation and enhances apoptosis by upregulating miR-21-5p [152] and miR-22-3p [153]. Piperlongumine, a Piper longum-derived alkaloid, causes ROS generation and apoptosis and downregulates c-Myc and its downstream miRNAs, such as miR-27a-3p (miR-27a), miR-20a-5p, and miR-17-5p, in pancreas, lung, and breast cancer cells [154] (Table 3).
Fucoidan has antiproliferative, apoptotic, and antimigration effects in liver cancer cells, accompanied by the downregulation of miR-522-3p [155]. Steviol, a Stevia rebaudiana Bertoni leaf-derived natural sweetener, shows anticancer effects, such as antiproliferative effects, G1 arrest, and apoptotic effects, in gastrointestinal cancer cells, accompanied by upregulation of miR-203a-3p and downregulation of miR-6088 in colon cancer (Caco-2) cells and upregulation of miR-1268b and downregulation of miR-23c in gastric cancer (HGC-27) cells [156]. The miRNAs involved in the anticancer effects of apigenin are well-reviewed [157]. For example, miR-101-3p (miR-101) was downregulated in doxorubicin-treated liver cancer cells, which was reversed by apigenin for apoptosis-inducible effects [158]. The Kanglaite Injection (KLT), made from the TCM herb yiyiren (Coicis Semen) through supercritical extraction [159], shows anticancer effects in lung cancer patients through the downregulation of miR-21-5p [160] (Table 3).
The natural product butylcycloheptyl prodiginine binds to precursor miR-21-5p to block the Dicer-mediated maturation of oncogenic miR-21-5p, inhibiting the growth of colon cancer cells [161]. Cucurbitacin B, a Cucurbitaceae plant-derived natural product, causes G2/M arrest and inhibits the proliferation of pancreatic cancer cells and xenograft tumors by upregulating miR-146b-5p [162]. Sanguinarine, an argemone oil-derived alkaloid, induces G1 arrest and oxidative stress-dependent apoptosis in liver cancer cells by upregulating miR-16-2 expression [163]. A methoxylated quercetin glycoside (quercetin-3′-methoxy-3-O-(4”-acetylrhamnoside)-7-O-α-rhamnoside) suppresses the proliferation and migration of liver cancer cells by upregulating TP53 and its downstream miRNAs miR-15a-5p (miR-15a)/miR-16-5p (miR-16), which can be reversed by inhibitors of miR-15a-5p/miR-16-5p (Table 3) [164].
Honokiol, a Magnolia officinalis-derived polyphenol, has antiproliferative and apoptotic effects in osteosarcoma cells through the downregulation of miR-21-5p (miR-21) [165]. miR-21-5p downregulates the phosphatase and tensin homolog (PTEN) [165], an AKT inhibitor. Consequently, honokiol inhibits miR-21-5p expression, causing AKT inactivation, which is reversed by miR-21-5p overexpression. Moreover, doxorubicin resistance develops from miR-188-5p overexpression, and honokiol can downregulate miR-188-5p to improve the doxorubicin sensitivity in breast cancer (MDA-MB-231) cells [166]. Shikonin, a Lithospermum erythrorhizon root-derived naphthoquinine, induces cell autophagy by upregulating miR-545-3p in colon cancer cells [167] (Table 3).
Sulforaphane, a cruciferous vegetable-derived sulfur-rich natural product, triggers cervical cancer cell apoptosis by activating MAPK through the downregulation of miR-1247-3p [168]. Sulforaphane also exhibits antiproliferative and apoptotic effects in nasopharyngeal cancer cells by increasing miR-124-3p expression [169]. Luteolin (3′,4′,5,7-tetrahydroxy flavone), a naturally occurring flavonoid modulating several miRNAs and inducing apoptosis in various types of cancer cells, is well-reviewed [170]. For example, miR-7-1-3p, miR-124-3p, miR-8080, miR-34a-5p, miR-384, miR-6809-5p, and miR-630 are upregulated by luteolin and exhibit tumor-suppressor functions in glioblastoma, brain, prostate, colon, gastric, liver, and prostate cancer cells, respectively [170] (Table 3).
Camptothecin, a Camptotheca acuminate-derived alkaloid, suppresses the migration, invasion, and proliferation of liver cancer Huh7 cells by upregulating miR-16-5p [171]. Piceatannol, a resveratrol analog, shows antiproliferative and apoptotic effects in osteosarcoma cells through the downregulation of miR-21-5p [172]. Matrine, a Sophora flavescens Ait-derived bioactive compound, triggers apoptosis in thyroid cancer (PTC) cells and inhibits tumor growth by decreasing miR-182-5p [173]. Maytenin and 22-β-hydroxymaytenin, which are Maytenus ilicifolia root-derived quinone-methide triterpenes, trigger ROS generation and apoptosis in head and neck cancer cells by downregulating miR-27a-3p, miR-20a-5p, and miR-17-5p [174]. Dihydromyricetin, an Ampelopsis plant-derived flavonoid, suppresses proliferation and migration, triggering cholangiocarcinoma cell apoptosis by downregulating miR-21-5p [175] (Table 3).
Table 3. Connecting natural products to miRNA-regulated cell functions.
Table 3. Connecting natural products to miRNA-regulated cell functions.
Natural ProductsmiRNAsCancerCell Functions
Ellagitannin   miR-526b-5p, miR-452-5p, miR-194-5p, miR-373-5p, miR-518-5p, miR-302a-5p, miR-424-5p, let-7e-5p, miR-525-5p, miR-519e-5p, miR-513a-3p, miR-518c-5p, miR-512-5p, miR-346 [147]Liver caAntiproliferation
  miR-542-3p, let-7d-5p, miR-299-3p, miR-200a-5p, let-7f-5p, let-7i-5p, let-7a-5p [147]
Luteolin, ellagic acid, punicic acid (combined)   miR-144-3p, miR-133b, miR-1-3p, miR-122-5p, miR-34c-5p, miR-200c-3p, miR-127-3p, miR-335-5p, miR-124-3p, miR-181a-5p, miR-7-5p, miR-15a-5p, let-7d-5p [148]Prostate caAntimetastasis
  miR-20a-5p, miR-21-5p, miR-9, miR-29b-3p, miR-181b [148]
Berbamine analog BBMD3   miR-4284 [149]GlioblastomaAntiproliferation, apoptosis
Oridonin   miR-17-5p, miR-20a-5p [150]LeukemiaApoptosis
Bufalin   miR-148a-3p [151]OsteosarcomaAntiproliferation, apoptosis
Betulinic acid   miR-21-5p [152], miR-22-3p [153]Liver caAntiproliferation, apoptosis
Piperlongumine   miR-27a-3p, miR-20a-5p, miR-17-5p [154]Pancreas, lung, breast caApoptosis
Fucoidan   miR-29b-3p [176]Liver caAntimigration
  miR-522-3p [155]Liver ca Antiproliferation,
 apoptosis,
 antimigration
Steviol   miR-203a-3p;   miR-6088 [156]Colon ca Antiproliferation, apoptosis
  miR-1268b;   miR-23c [156]Gastric caAntiproliferation, apoptosis
Baicalein   miR-3127-5p [177]Liver caAnti-apoptosis
Apigenin   miR-101-3p [158]Liver caApoptosis
Kanglaite injection   miR-21-5p [160]Lung ca patientsAnticancer
 Butylcycloheptyl prodiginine   miR-21-5p [161]Colon ca Antiproliferation
Cucurbitacin B   miR-146b-5p [162]Pancreatic caAntiproliferation
Sanguinarine   miR-16-2 expression [163]Liver caApoptosis
 Quercetin-3′-methoxy-3-O-(4″-acetylrhamnoside)-7-O-α-rhamnoside   miR-15a-5p, miR-16-5p [164]Liver caAntimigration
Honokiol   miR-21-5p (miR-21) [165]OsteosarcomaApoptosis
  miR-188-5p [166]Breast caDoxorubicin sensitization
Shikonin   miR-545-3p [167]Colon caAutophagy
Sulforaphane   miR-1247-3p [168]Cervical caApoptosis
  miR-124-3p expression [169] Nasopharyngeal caApoptosis
Luteolin   miR-7-1-3p, miR-124-3p, miR-8080, miR-34a-5p, miR-384, miR-6809-5p, miR-630 [170] Glioblastoma, brain, prostate, colon, gastric, liver, prostate caApoptosis
Camptothecin   miR-16-5p [171]Liver caAntimigration
Piceatannol   miR-21-5p [172]OsteosarcomaApoptosis
Panax ginseng CA Meyer   miR-21-5p [178] Gastric precancerous lesionApoptosis
Oleacein   miR-193a-3p, miR-193a-5p, miR-34a-5p, miR-16-5p [179]MelanomaApoptosis
  miR-214-3p [179] MelanomaApoptosis
Matrine   miR-182-5p [173]Thyroid caApoptosis
 Maytenin, 22-β-hydroxymaytenin   miR-27a-3p, miR-20a-5p, miR-17-5p [174]Head/neck caApoptosis
Dihydromyricetin   miR-21-5p [175]CholangiocarcinomaAntiproliferation, apoptosis
Toosendanin   miR-608 [180]GliomaApoptosis
( )-Sativan   miR-200c-3p [181]Breast caAntimigration, apoptosis
Thymoquinone   miR-34a-5p [182]Breast caAntimigration
Phytic acid   miR-224-5p, miR-200a-3p [183]Colon caApoptosis
 Polyphenol-enriched blueberry preparation   miR-200b-3p [184]MelanomaAntimigration
Formononetin   miR-542-5p [185]Gastric caAntimigration
Rutin   miR-590-5p [186]Liver caAnti-autophagy
  miR-129-1-3p [187]Breast caAntimetastasis
Asparanin A   miR-551a, miR-1303 [188]Endometrial caAntimigration
CoB1   miR-125b-5p [189]Lung caAutophagy
Curcumenol   miR-19b-3p [190]Lung caFerroptosis
Hesperidin, luteolin   miR-16-5p, miR-34a-5p [191]Breast caApoptosis
  miR-21-5p [191]Breast caApoptosis
 Sulforaphane/peptide nucleic acid (combined)   miR-15b-5p [192]GlioblastomaApoptosis
Solamargine   miR-192-5p [193]Liver caApoptosis, autophagy
Asiaticoside   miR-635 [194]Gastric caER stress
Icariside II   miR-324-3p [195]Renal caFerroptosis
, enhance or activate; , inhibit or inactivate.
Toosendanin, a Melia toosendan Sieb et Zucc-derived triterpenoid, promotes apoptosis in glioma cells by upregulating miR-608 [180]. ( )-Sativan, a Spatholobus suberectus-derived natural product, inhibits proliferation and migration and induces apoptosis in breast cancer cells by upregulating miR-200c-3p [181]. Thymoquinone, the main bioactive compound derived from Nigella sativa, inhibits the migration of breast cancer cells by upregulating miR-34a-5p [182]. Thymoquinone induces apoptosis by downregulating the activation (phosphorylation) of STAT3 [196]. Moreover, thymoquinone has chemosensitization effects. The combined treatment including thymoquinone and doxorubicin shows synergistic antiproliferative effects against murine solid Ehrlich carcinoma through the upregulation of miR-125a-5p [196]. Phytic acid, a plant seed-derived natural product, improved the apoptotic effects of oxaliplatin in a 1,2-dimethylhydrazine-induced colorectal cancer model by downregulating miR-224-5p (miR-224) and miR-200a-3p (miR-200a) [183] (Table 3).
The polyphenol-enriched blueberry preparation inhibits the proliferation and migration of melanoma cells by upregulating miR-200b-3p (miR-200b) [184]. Formononetin, an isoflavone derived from Astragalus membranaceus root, suppresses the migration and invasion of gastric cancer cells by downregulating miR-542-5p [185]. Rutin, a vegetable- and fruit-derived flavonoid, inhibits autophagy in liver cancer cells by downregulating the lncRNA BRAF-activated ncRNA (BANCR), a molecular sponge of miR-590-5p [186]. Accordingly, rutin upregulates miR-590-5p in liver cancer cells. Moreover, rutin also inhibits breast cancer cell proliferation and metastasis by upregulating miR-129-1-3p-dependent calcium signaling [187]. Asparanin A, an Asparagus officinalis-derived natural product, inhibits the migration and invasion of endometrial cancer cells by downregulating miR-551a and miR-1303 [188] (Table 3).
Curcumenol is a Curcuma zedoaria-derived sesquiterpene causing ferroptosis in lung cancer (H1299) cells by downregulating lncRNA H19 and upregulating miR-19b-3p [190]. Hesperidin and luteolin, both citrus fruit-derived flavanone glycosides and Reseda luteola-derived flavonoids, induce antiproliferative and apoptotic effects in breast cancer (MCF7) cells by downregulating miR-21-5p and upregulating miR-16-5p and miR-34a-5p [191]. A combined treatment including sulforaphane and peptide nucleic acid triggers the synergistic apoptosis of glioblastoma (U251) cells by downregulating miR-15b-5p [192]. Asiaticoside, a glycosylated triterpene of Centella asiatica, has antiproliferative effects and causes ER stress in gastric cancer cells by upregulating miR-635 [194]. Icariside II, the Epimedium brevicornum-derived natural product, induces ferroptosis in renal cancer cells by upregulating miR-324-3p to target GPX4 [195]. Fucoidan, a sulfated brown algal polysaccharide, suppresses EMT in liver cancer (HepG2) cells by upregulating miR-29b-3p to inhibit DNMT3B, a metastasis suppressor 1 (MTSS1) suppressor [176] (Table 3).
All chemical structures mentioned in Table 3 are provided in Figure 2.

5. Modulation of AKT- and AKT Effector-Targeting miRNAs by Natural Products

Many natural products show AKT- and AKT effector-modulating functions [197] with anticancer effects. However, a systemic understanding of the potential roles of miRNAs in regulating cell functions in coordination with natural products that modulate AKT and AKT effectors has not yet been achieved.
Several natural products that control various cancer cell functions by modulating miRNAs are summarized in Table 3. Some anticancer studies on natural products have emphasized the regulation of AKT and several AKT effectors in the miRNA-mediated cell functions exerted by natural products (Table 3). For example, solamargine, a Solanum incanum herb-derived natural product, promotes antiproliferative effects, apoptosis, and autophagy in liver cancer cells by downregulating leukemia inhibitory factor (LIF), cysteine-rich angiogenic inducer 61 (CYR61), and AKT and upregulating miR-192-5p (Table 3) [193]. Moreover, the downregulation of miR-192-5p enhances AKT activation. miR-3127-5p is downregulated to promote cell proliferation in several liver cancer cells. Baicalein can upregulate miR-3127-5p to suppress liver cancer (Bel-7402) cell proliferation, S phase arrest, and apoptosis by inactivating AKT (Table 3) [177]. Panax ginseng CA Meyer (Rg3) suppresses proliferation and promotes apoptosis in gastric precancerous lesion cells by upregulating miR-21-5p and downregulating PI3K/AKT (Table 3) [178]. Oleacein, an extra-virgin olive oil polyphenol, induces antiproliferative and apoptotic effects in melanoma cells by upregulating miR-193a-3p, miR-193a-5p (targeting mTOR), miR-34a-5p, and miR-16-5p (targeting mTOR) and downregulating miR-214-3p (targeting BAX) (Table 3) [179]. The cochlioquinone derivative CoB1 triggers autophagy in lung cancer (A549) cells by upregulating miR-125b-5p expression and downregulating AKT and FOXO3 expressions [189] (Table 3). Although these studies provide information on AKT and several AKT effectors, the potential functions of other AKT effectors in natural product-induced miRNA modulation remain unclear.
Notably, most studies listed in Table 3 did not investigate the involvement of AKT and its effectors. Utilizing bioinformatics (the miRDB database [69]), the miRNAs listed in Table 3 were input into the miRDB database to check the predicted target genes related to AKT and AKT effectors. The retrieval results for the natural product-modulating miRNAs from Table 3 are summarized in Table 4. For example, miR-7-1-3p was predicted to target AKT1, while miR-103a-3p, miR-107, miR-124-3p, miR-148a-3p, miR-29b-3p, and miR-29c-3p were predicted to target AKT2 (Table 4). Many miRNAs were predicted to target AKT3, as shown in Table 4.
Many AKT effectors were predicted to be targeted by several natural product-modulated miRNAs. For example, FOXO1 was predicted to be targeted by let-ff-1-3p, miR-143-3p, miR-144-3p, miR-145-5p, miR-182-5p, miR-183-5p, miR-223-3p, miR-27a-3p, miR-27b-3p, miR-324-5p, miR-486-5p, miR-513a-3p, miR-5195-3p, miR-7-1-3p, and miR-96-5p (Table 4). Similarly, FOXO3 and FOXO4 were predicted to be targeted by many of the natural product-modulated miRNAs from Table 3. Myc was predicted to be targeted by miR-16-2, miR-203a-3p, and miR-629-5p. mTOR was predicted to be targeted by miR-101-3p, miR-144-3p, miR-421, miR-545-3p, miR-616-3p, miR-629-5p, miR-96-5p, and miR-99a-5p. AKT1S1 was predicted to be targeted by miR-124-3p, miR-125a-5p, miR-125b-5p, and miR-129-1-3p.
Moreover, DEPTOR and HIF1A were also predicted to be targeted by several natural product-modulating miRNAs that were not reported to be associated with AKT and AKT effectors. Therefore, several miRNAs connected to natural product studies have bioinformatic predictions related to AKT and AKT effectors that should be verified in experiments. The contributions of AKT and AKT effectors to the modulation of miRNAs by natural products can thus be explored.

6. Conclusions

Many miRNAs regulate various cancer cell functions by targeting several genes. AKT and AKT effectors are highly expressed in various cancer cells and accompanied by the regulation of cell functions (apoptosis, autophagy, ER stress, ferroptosis, necroptosis, DDR, senescence, and migration). However, the connection between miRNAs and AKT and AKT effectors in controlling cell function remains unclear. Although several natural products may modulate miRNAs and the AKT pathway, their relationships lack systematic organization.
This review provided comprehensive information concerning the relationships between miRNAs and cancer cell functions (Figure 3). The roles of AKT and AKT effectors in miRNA-regulated cancer cell functions were clarified. Moreover, there is abundant information on the modulation of many miRNAs by natural products, but the involvement of AKT and AKT effectors has rarely been reported. Utilizing bioinformatics, the miRDB database was chosen for the prediction of AKT and AKT effectors related to miRNAs regulated by natural products. Consequently, the gaps between AKT, AKT effectors, miRNAs, and natural products were filled (Figure 3).
However, the predictions of the targets for miRNAs obtained by searching miRDB must be carefully examined via detailed experiments because these predictions may be derived from specific types of cancer cells that may have different responses to other types of cancer cells. This review sheds light on the connections between natural products, miRNAs, and AKT pathways and can provide a future direction for exploring the regulation of cancer cell functions by natural products.

Author Contributions

Conceptualization, Y.-T.C., J.-Y.T. and H.-W.C.; Methodology, Y.-T.C., J.-P.S., C.-Y.Y., F.-R.C., K.-H.Y. and M.-F.H.; Supervision, J.-Y.T. and H.-W.C.; Writing—original draft, J.-Y.T., Y.-T.C. and H.-W.C.; Writing—review and editing, J.-Y.T. and H.-W.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partly supported by funds from the Ministry of Science and Technology (MOST 111-2320-B-037-015-MY3 and MOST 110-2314-B-037-074-MY3), the Kaohsiung Medical University (KMU-DK(A)111008), and the Kaohsiung Medical University Research Center (KMU-TC108A04).

Acknowledgments

The authors thank our colleague Hans-Uwe Dahms for editing the manuscript.

Conflicts of Interest

The authors declare that there are no conflicts of interest.

References

  1. Zhang, Y.; Huang, S.; Yang, G.; Zou, L.; Huang, X.; Liu, S. The role of miRNAs during endoplasmic reticulum stress induced apoptosis in digestive cancer. J. Cancer 2021, 12, 6787–6795. [Google Scholar] [CrossRef] [PubMed]
  2. Xu, L.; Huang, X.; Lou, Y.; Xie, W.; Zhao, H. Regulation of apoptosis, autophagy and ferroptosis by non-coding RNAs in metastatic non-small cell lung cancer (Review). Exp. Ther. Med. 2022, 23, 352. [Google Scholar] [CrossRef] [PubMed]
  3. Shekari, N.; Asadi, M.; Akbari, M.; Baradaran, B.; Zarredar, H.; Mohaddes-Gharamaleki, F.; Anvarnia, A.; Baghbanzadeh, A.; Shanehbandi, D. Autophagy-regulating microRNAs: Two-sided coin in the therapies of breast cancer. Eur. Rev. Med. Pharmacol. Sci. 2022, 26, 1268–1282. [Google Scholar] [PubMed]
  4. Shahverdi, M.; Hajiasgharzadeh, K.; Sorkhabi, A.D.; Jafarlou, M.; Shojaee, M.; Jalili Tabrizi, N.; Alizadeh, N.; Santarpia, M.; Brunetti, O.; Safarpour, H.; et al. The regulatory role of autophagy-related miRNAs in lung cancer drug resistance. Biomed. Pharmacother. 2022, 148, 112735. [Google Scholar] [CrossRef]
  5. Lei, Y.; Chen, L.; Liu, J.; Zhong, Y.; Deng, L. The microRNA-based strategies to combat cancer chemoresistance via regulating autophagy. Front. Oncol. 2022, 12, 841625. [Google Scholar] [CrossRef]
  6. Kim, T.; Croce, C.M. MicroRNA and ER stress in cancer. Semin. Cancer. Biol. 2021, 75, 3–14. [Google Scholar] [CrossRef]
  7. Fuhrmann, D.C.; Brune, B. A graphical journey through iron metabolism, microRNAs, and hypoxia in ferroptosis. Redox Biol. 2022, 54, 102365. [Google Scholar] [CrossRef]
  8. Xie, B.; Guo, Y. Molecular mechanism of cell ferroptosis and research progress in regulation of ferroptosis by noncoding RNAs in tumor cells. Cell Death Discov. 2021, 7, 101. [Google Scholar] [CrossRef]
  9. Zhi, Y.; Gao, L.; Wang, B.; Ren, W.; Liang, K.X.; Zhi, K. Ferroptosis holds novel promise in treatment of cancer mediated by non-coding RNAs. Front. Cell. Dev. Biol. 2021, 9, 686906. [Google Scholar] [CrossRef]
  10. Zuo, Y.B.; Zhang, Y.F.; Zhang, R.; Tian, J.W.; Lv, X.B.; Li, R.; Li, S.P.; Cheng, M.D.; Shan, J.; Zhao, Z.; et al. Ferroptosis in cancer progression: Role of noncoding RNAs. Int. J. Biol. Sci. 2022, 18, 1829–1843. [Google Scholar] [CrossRef]
  11. Visalli, M.; Bartolotta, M.; Polito, F.; Oteri, R.; Barbera, A.; Arrigo, R.; Di Giorgio, R.M.; Navarra, G.; Aguennouz, M. miRNA expression profiling regulates necroptotic cell death in hepatocellular carcinoma. Int. J. Oncol. 2018, 53, 771–780. [Google Scholar] [CrossRef] [PubMed]
  12. Bao, J.H.; Li, J.B.; Lin, H.S.; Zhang, W.J.; Guo, B.Y.; Li, J.J.; Fu, L.M.; Sun, Y.P. Deciphering a novel necroptosis-related miRNA signature for predicting the prognosis of clear cell renal carcinoma. Anal. Cell. Pathol. Amst. 2022, 2022, 2721005. [Google Scholar] [CrossRef]
  13. Hong, X.; Wang, G.; Pei, K.; Mo, C.; Rong, Z.; Xu, G. A novel prognostic model based on seven necroptosis-related miRNAs for predicting the overall survival of patients with lung adenocarcinoma. Biomed. Res. Int. 2022, 2022, 3198590. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, Z.; Lu, S.; Wang, Y.; Tang, H.; Wang, B.; Sun, X.; Qu, J.; Rao, B. A novel defined necroptosis-related miRNAs signature for predicting the prognosis of colon cancer. Int. J. Gen. Med. 2022, 15, 555–565. [Google Scholar] [CrossRef] [PubMed]
  15. Visser, H.; Thomas, A.D. MicroRNAs and the DNA damage response: How is cell fate determined? DNA Repair. 2021, 108, 103245. [Google Scholar] [CrossRef]
  16. Ketley, R.F.; Gullerova, M. Jack of all trades? The versatility of RNA in DNA double-strand break repair. Essays. Biochem. 2020, 64, 721–735. [Google Scholar] [PubMed]
  17. Peraza-Vega, R.I.; Valverde, M.; Rojas, E. Interactions between miRNAs and double-strand breaks DNA repair genes, pursuing a fine-tuning of repair. Int. J. Mol. Sci. 2022, 23, 3231. [Google Scholar] [CrossRef]
  18. Wang, Z.; Gao, J.; Xu, C. Tackling cellular senescence by targeting miRNAs. Biogerontology 2022, 23, 387–400. [Google Scholar] [CrossRef] [PubMed]
  19. Popov, A.; Mandys, V. Senescence-associated miRNAs and their role in pancreatic cancer. Pathol. Oncol. Res. 2022, 28, 1610156. [Google Scholar] [CrossRef] [PubMed]
  20. Jafri, M.A.; Al-Qahtani, M.H.; Shay, J.W. Role of miRNAs in human cancer metastasis: Implications for therapeutic intervention. Semin. Cancer Biol. 2017, 44, 117–131. [Google Scholar] [CrossRef]
  21. Petri, B.J.; Klinge, C.M. Regulation of breast cancer metastasis signaling by miRNAs. Cancer Metastasis Rev. 2020, 39, 837–886. [Google Scholar] [CrossRef] [PubMed]
  22. Teo, A.Y.T.; Xiang, X.; Le, M.T.; Wong, A.L.; Zeng, Q.; Wang, L.; Goh, B.C. Tiny miRNAs play a big role in the treatment of breast cancer metastasis. Cancers 2021, 13, 337. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, J.; Hu, S.; Liu, K.; Sun, G.; Zhang, Y. The role of microRNA in the regulation of tumor epithelial-mesenchymal transition. Cells 2022, 11, 1981. [Google Scholar] [CrossRef] [PubMed]
  24. Huang, W.; Wu, X.; Xiang, S.; Qiao, M.; Cen, X.; Pan, X.; Huang, X.; Zhao, Z. Regulatory mechanism of miR-20a-5p expression in Cancer. Cell Death Discov. 2022, 8, 262. [Google Scholar] [CrossRef]
  25. Frederick, M.I.; Siddika, T.; Zhang, P.; Balasuriya, N.; Turk, M.A.; O’Donoghue, P.; Heinemann, I.U. miRNA-dependent regulation of AKT1 phosphorylation. Cells. 2022, 11, 821. [Google Scholar] [CrossRef]
  26. Xie, Y.; Hou, W.; Song, X.; Yu, Y.; Huang, J.; Sun, X.; Kang, R.; Tang, D. Ferroptosis: Process and function. Cell Death Differ. 2016, 23, 369–379. [Google Scholar] [CrossRef]
  27. Su, L.J.; Zhang, J.H.; Gomez, H.; Murugan, R.; Hong, X.; Xu, D.; Jiang, F.; Peng, Z.Y. Reactive oxygen species-induced lipid peroxidation in apoptosis, autophagy, and ferroptosis. Oxid. Med. Cell. Longev. 2019, 2019, 5080843. [Google Scholar] [CrossRef]
  28. Dastmalchi, N.; Safaralizadeh, R.; Khojasteh, S.M.B.; Shadbad, M.A.; Hosseinpourfeizi, M.A.; Azarbarzin, S.; Rajabi, A.; Baradaran, B. The combined restoration of miR-424-5p and miR-142-3p effectively inhibits MCF-7 breast cancer cell line via modulating apoptosis, proliferation, colony formation, cell cycle and autophagy. Mol. Biol. Rep. 2022, 49, 8325–8335. [Google Scholar] [CrossRef]
  29. Gupta, S.; Panda, P.K.; Hashimoto, R.F.; Samal, S.K.; Mishra, S.; Verma, S.K.; Mishra, Y.K.; Ahuja, R. Dynamical modeling of miR-34a, miR-449a, and miR-16 reveals numerous DDR signaling pathways regulating senescence, autophagy, and apoptosis in HeLa cells. Sci. Rep. 2022, 12, 4911. [Google Scholar] [CrossRef]
  30. Song, M.; Bode, A.M.; Dong, Z.; Lee, M.H. AKT as a therapeutic target for cancer. Cancer Res. 2019, 79, 1019–1031. [Google Scholar] [CrossRef]
  31. Manning, B.D.; Toker, A. AKT/PKB signaling: Navigating the network. Cell 2017, 169, 381–405. [Google Scholar] [CrossRef] [PubMed]
  32. Tang, J.Y.; Cheng, Y.B.; Chuang, Y.T.; Yang, K.H.; Chang, F.R.; Liu, W.; Chang, H.W. Oxidative stress and AKT-associated angiogenesis in a zebrafish model and its potential application for withanolides. Cells 2022, 11, 961. [Google Scholar] [CrossRef] [PubMed]
  33. Shiau, J.P.; Chuang, Y.T.; Cheng, Y.B.; Tang, J.Y.; Hou, M.F.; Yen, C.Y.; Chang, H.W. Impacts of oxidative stress and PI3K/AKT/mTOR on metabolism and the future direction of investigating fucoidan-modulated metabolism. Antioxidants 2022, 11, 911. [Google Scholar] [CrossRef] [PubMed]
  34. Tang, J.Y.; Chuang, Y.T.; Shiau, J.P.; Yang, K.H.; Chang, F.R.; Hou, M.F.; Farooqi, A.A.; Chang, H.W. Long noncoding RNAs and circular RNAs regulate AKT and its effectors to control cell functions of cancer cells. Cells 2022, 11, 2940. [Google Scholar] [CrossRef]
  35. Taefehshokr, S.; Taefehshokr, N.; Derakhshani, A.; Baghbanzadeh, A.; Astamal, R.V.; Safaei, S.; Abbasi, S.; Hajazimian, S.; Maroufi, N.F.; Isazadeh, A.; et al. The regulatory role of pivotal microRNAs in the AKT signaling pathway in breast cancer. Curr. Mol. Med. 2022, 22, 263–273. [Google Scholar] [CrossRef]
  36. Ghaffarian Zirak, R.; Tajik, H.; Asadi, J.; Hashemian, P.; Javid, H. The Role of Micro RNAs in Regulating PI3K/AKT Signaling Pathways in Glioblastoma. Iran. J. Pathol. 2022, 17, 122–136. [Google Scholar] [CrossRef]
  37. Li, H.L.; Deng, N.H.; He, X.S.; Li, Y.H. Small biomarkers with massive impacts: PI3K/AKT/mTOR signalling and microRNA crosstalk regulate nasopharyngeal carcinoma. Biomark. Res. 2022, 10, 52. [Google Scholar] [CrossRef]
  38. Akbarzadeh, M.; Mihanfar, A.; Akbarzadeh, S.; Yousefi, B.; Majidinia, M. Crosstalk between miRNA and PI3K/AKT/mTOR signaling pathway in cancer. Life Sci. 2021, 285, 119984. [Google Scholar] [CrossRef]
  39. Lee, K.H.; Chen, Y.L.; Yeh, S.D.; Hsiao, M.; Lin, J.T.; Goan, Y.G.; Lu, P.J. MicroRNA-330 acts as tumor suppressor and induces apoptosis of prostate cancer cells through E2F1-mediated suppression of Akt phosphorylation. Oncogene 2009, 28, 3360–3370. [Google Scholar] [CrossRef]
  40. Mi, Y.; Zhang, L.; Sun, C.; Feng, Y.; Sun, J.; Wang, J.; Yang, D.; Qi, X.; Wan, H.; Xia, G.; et al. Lysine demethylase 5A promotes prostate adenocarcinoma progression by suppressing microRNA-330-3p expression and activating the COPB2/PI3K/AKT axis in an ETS1-dependent manner. J. Cell Commun. Signal. 2022, 16, 579–599. [Google Scholar] [CrossRef]
  41. Gwak, H.S.; Kim, T.H.; Jo, G.H.; Kim, Y.J.; Kwak, H.J.; Kim, J.H.; Yin, J.; Yoo, H.; Lee, S.H.; Park, J.B. Silencing of microRNA-21 confers radio-sensitivity through inhibition of the PI3K/AKT pathway and enhancing autophagy in malignant glioma cell lines. PLoS ONE 2012, 7, e47449. [Google Scholar] [CrossRef] [PubMed]
  42. Gong, H.; Cao, Y.; Han, G.; Zhang, Y.; You, Q.; Wang, Y.; Pan, Y. p53/microRNA-374b/AKT1 regulates colorectal cancer cell apoptosis in response to DNA damage. Int. J. Oncol. 2017, 50, 1785–1791. [Google Scholar] [CrossRef] [PubMed]
  43. Zheng, Y.; Xu, Z. MicroRNA-22 induces endothelial progenitor cell senescence by targeting AKT3. Cell. Physiol. Biochem. 2014, 34, 1547–1555. [Google Scholar] [CrossRef]
  44. Chen, X.; Huang, Z.; Chen, R. MicroRNA-136 promotes proliferation and invasion ingastric cancer cells through Pten/Akt/P-Akt signaling pathway. Oncol. Lett. 2018, 15, 4683–4689. [Google Scholar] [CrossRef]
  45. Li, L.; Wei, D.; Zhang, J.; Deng, R.; Tang, J.; Su, D. miR-641 inhibited cell proliferation and induced apoptosis by targeting NUCKS1/PI3K/AKT signaling pathway in breast cancer. Comput. Math. Methods Med. 2022, 2022, 5203839. [Google Scholar] [CrossRef]
  46. Qi, S.; Xu, L.; Han, Y.; Chen, H.; Cheng, A. miR-29a-3p mitigates the development of osteosarcoma through modulating IGF1 mediated PI3k/Akt/FOXO3 pathway by activating autophagy. Cell Cycle 2022, 21, 1980–1995. [Google Scholar] [CrossRef]
  47. Lee, J.H.; Park, S.J.; Jeong, S.Y.; Kim, M.J.; Jun, S.; Lee, H.S.; Chang, I.Y.; Lim, S.C.; Yoon, S.P.; Yong, J.; et al. MicroRNA-22 suppresses DNA repair and promotes genomic instability through targeting of MDC1. Cancer Res. 2015, 75, 1298–1310. [Google Scholar] [CrossRef]
  48. Xu, X.; Kim, J.J.; Li, Y.; Xie, J.; Shao, C.; Wei, J.J. Oxidative stress-induced miRNAs modulate AKT signaling and promote cellular senescence in uterine leiomyoma. J. Mol. Med. 2018, 96, 1095–1106. [Google Scholar] [CrossRef]
  49. Gao, X.; Fan, X.; Zeng, W.; Liang, J.; Guo, N.; Yang, X.; Zhao, Y. Overexpression of microRNA-107 suppressed proliferation, migration, invasion, and the PI3K/Akt signaling pathway and induced apoptosis by targeting Nin one binding (NOB1) protein in a hypopharyngeal squamous cell carcinoma cell line (FaDu). Bioengineered 2022, 13, 7881–7893. [Google Scholar] [CrossRef]
  50. Zhang, W.; Hsu, P.; Zhong, B.; Guo, S.; Zhang, C.; Wang, Y.; Luo, C.; Zhan, Y.; Zhang, C. MiR-34a enhances chondrocyte apoptosis, senescence and facilitates development of osteoarthritis by targeting DLL1 and regulating PI3K/AKT pathway. Cell. Physiol. Biochem. 2018, 48, 1304–1316. [Google Scholar] [CrossRef]
  51. Li, H.; Wang, N.; Xu, Y.; Chang, X.; Ke, J.; Yin, J. Upregulating microRNA-373-3p promotes apoptosis and inhibits metastasis of hepatocellular carcinoma cells. Bioengineered 2022, 13, 1304–1319. [Google Scholar] [CrossRef] [PubMed]
  52. Hemmings, K.E.; Riches-Suman, K.; Bailey, M.A.; O’Regan, D.J.; Turner, N.A.; Porter, K.E. Role of microRNA-145 in DNA damage signalling and senescence in vascular smooth muscle cells of type 2 diabetic patients. Cells 2021, 10, 919. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, W.; Zhang, J.; Fan, Y.; Zhang, L. MiR-1306-5p predicts favorable prognosis and inhibits proliferation, migration, and invasion of colorectal cancer cells via PI3K/AKT/mTOR pathway. Cell Cycle 2022, 21, 1491–1501. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, L.Y.; Yu, J.Y.; Leng, Y.L.; Zhu, R.R.; Liu, H.X.; Wang, X.Y.; Yang, T.T.; Guo, Y.N.; Tang, J.L.; Zhang, X.C. MiR-181c sensitizes ovarian cancer cells to paclitaxel by targeting GRP78 through the PI3K/Akt pathway. Cancer Gene Ther. 2022, 29, 770–783. [Google Scholar] [CrossRef]
  55. Dong, W.; Wang, F.; Liu, Q.; Wang, T.; Yang, Y.; Guo, P.; Li, X.; Wei, B. Downregulation of miRNA-14669 reverses vincristine resistance in colorectal cancer cells through PI3K/AKT signaling pathway. Recent Pat. Anticancer Drug. Discov. 2022, 17, 178–186. [Google Scholar]
  56. Liu, M.; Yang, J.; Lv, W.; Wang, S.; Du, T.; Zhang, K.; Wu, Y.; Feng, X. Down-regulating GRP78 reverses pirarubicin resistance of triple negative breast cancer by miR-495-3p mimics and involves the p-AKT/mTOR pathway. Biosci. Rep. 2022, 42, BSR20210245. [Google Scholar] [CrossRef]
  57. Tomita, K.; Nagasawa, T.; Kuwahara, Y.; Torii, S.; Igarashi, K.; Roudkenar, M.H.; Roushandeh, A.M.; Kurimasa, A.; Sato, T. MiR-7-5p is involved in ferroptosis signaling and radioresistance thru the generation of ROS in radioresistant HeLa and SAS cell lines. Int. J. Mol. Sci. 2021, 22, 8300. [Google Scholar] [CrossRef]
  58. Chen, H.; Li, P.; Shen, Z.; Wang, J.; Diao, L. Protective effects of selenium yeast against cadmium-induced necroptosis through miR-26a-5p/PTEN/PI3K/AKT signaling pathway in chicken kidney. Ecotoxicol. Environ. Saf. 2021, 220, 112387. [Google Scholar] [CrossRef]
  59. Chen, P.; Li, X.; Yu, X.; Yang, M. Ginsenoside Rg1 suppresses non-small-cell lung cancer via microRNA-126-PI3K-AKT-mTOR pathway. Evid.-Based Complement. Altern. Med. 2022, 2022, 1244836. [Google Scholar] [CrossRef]
  60. Mao, Q.; Zhuang, Q.; Shen, J.; Chen, Z.; Xue, D.; Ding, T.; He, X. MiRNA-124 regulates the sensitivity of renal cancer cells to cisplatin-induced necroptosis by targeting the CAPN4-CNOT3 axis. Transl. Androl. Urol. 2021, 10, 3669–3683. [Google Scholar] [CrossRef]
  61. Li, F.; Liang, A.; Lv, Y.; Liu, G.; Jiang, A.; Liu, P. MicroRNA-200c inhibits epithelial-mesenchymal transition by targeting the BMI-1 gene through the phospho-AKT pathway in endometrial carcinoma cells in vitro. Med. Sci. Monit. 2017, 23, 5139–5149. [Google Scholar] [CrossRef] [PubMed]
  62. Zhao, C.; Zhou, Y.; Ran, Q.; Yao, Y.; Zhang, H.; Ju, J.; Yang, T.; Zhang, W.; Yu, X.; He, S. MicroRNA-381-3p functions as a dual suppressor of apoptosis and necroptosis and promotes proliferation of renal cancer cells. Front. Cell Dev. Biol. 2020, 8, 290. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, N.; Zhou, J.; Zhou, Y.; Guan, F. MicroRNA-148a inhibits hepatocellular carcinoma cell growth via epithelial-to-mesenchymal transition and PI3K/AKT signaling pathways by targeting death receptor-5. Appl. Biochem. Biotechnol. 2022, 194, 2731–2746. [Google Scholar] [CrossRef] [PubMed]
  64. Wu, S.; Zhou, Y.; Liu, P.; Zhang, H.; Wang, W.; Fang, Y.; Shen, X. MicroRNA-29b-3p promotes 5-fluorouracil resistance via suppressing TRAF5-mediated necroptosis in human colorectal cancer. Eur. J. Histochem. 2021, 65, 3247. [Google Scholar] [CrossRef]
  65. Huang, T.; Chen, Q.F.; Chang, B.Y.; Shen, L.J.; Li, W.; Wu, P.H.; Fan, W.J. TFAP4 promotes hepatocellular carcinoma invasion and metastasis via activating the PI3K/AKT signaling pathway. Dis. Markers 2019, 2019, 7129214. [Google Scholar] [CrossRef]
  66. Ben-Hamo, R.; Zilberberg, A.; Cohen, H.; Bahar-Shany, K.; Wachtel, C.; Korach, J.; Aviel-Ronen, S.; Barshack, I.; Barash, D.; Levanon, K.; et al. Resistance to paclitaxel is associated with a variant of the gene BCL2 in multiple tumor types. NPJ Precis. Oncol. 2019, 3, 12. [Google Scholar] [CrossRef]
  67. Dong, C.; Wu, J.; Chen, Y.; Nie, J.; Chen, C. Activation of PI3K/AKT/mTOR pathway causes drug resistance in breast cancer. Front. Pharmacol. 2021, 12, 628690. [Google Scholar] [CrossRef]
  68. Kuwahara, Y.; Mori, M.; Oikawa, T.; Shimura, T.; Ohtake, Y.; Mori, S.; Ohkubo, Y.; Fukumoto, M. The modified high-density survival assay is the useful tool to predict the effectiveness of fractionated radiation exposure. J. Radiat. Res. 2010, 51, 297–302. [Google Scholar] [CrossRef]
  69. Chen, Y.; Wang, X. miRDB: An online database for prediction of functional microRNA targets. Nucleic Acids Res. 2020, 48, D127–D131. [Google Scholar] [CrossRef]
  70. Fu, Z.; Tindall, D.J. FOXOs, cancer and regulation of apoptosis. Oncogene 2008, 27, 2312–2319. [Google Scholar] [CrossRef]
  71. Zhang, X.; Tang, N.; Hadden, T.J.; Rishi, A.K. Akt, FoxO and regulation of apoptosis. Biochim. Biophys. Acta 2011, 1813, 1978–1986. [Google Scholar] [CrossRef]
  72. Xu, H.; Zhu, J.; Hu, C.; Song, H.; Li, Y. Inhibition of microRNA-181a may suppress proliferation and invasion and promote apoptosis of cervical cancer cells through the PTEN/Akt/FOXO1 pathway. J. Physiol. Biochem. 2016, 72, 721–732. [Google Scholar] [CrossRef]
  73. Yang, Z.; Zhang, L. MiRNA-335 modulates hepatoma cell lines apoptosis and proliferation by targeting forkhead box O3a (FOXO3a). J. Biomater. Tiss. Eng. 2022, 12, 417–421. [Google Scholar] [CrossRef]
  74. Zhou, Y.; Chen, E.; Tang, Y.; Mao, J.; Shen, J.; Zheng, X.; Xie, S.; Zhang, S.; Wu, Y.; Liu, H.; et al. miR-223 overexpression inhibits doxorubicin-induced autophagy by targeting FOXO3a and reverses chemoresistance in hepatocellular carcinoma cells. Cell Death Dis. 2019, 10, 843. [Google Scholar] [CrossRef] [PubMed]
  75. Cui, C.; Li, T.; Xie, Y.; Yang, J.; Fu, C.; Qiu, Y.; Shen, L.; Ni, Q.; Wang, Q.; Nie, A.; et al. Enhancing Acsl4 in absence of mTORC2/Rictor drove beta-cell dedifferentiation via inhibiting FoxO1 and promoting ROS production. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 166261. [Google Scholar] [CrossRef]
  76. Poudyal, D.; Herman, A.; Adelsberger, J.W.; Yang, J.; Hu, X.; Chen, Q.; Bosche, M.; Sherman, B.T.; Imamichi, T. A novel microRNA, hsa-miR-6852 differentially regulated by Interleukin-27 induces necrosis in cervical cancer cells by downregulating the FoxM1 expression. Sci. Rep. 2018, 8, 900. [Google Scholar] [CrossRef] [PubMed]
  77. Wei, Y.T.; Guo, D.W.; Hou, X.Z.; Jiang, D.Q. miRNA-223 suppresses FOXO1 and functions as a potential tumor marker in breast cancer. Cell. Mol. Biol. 2017, 63, 113–118. [Google Scholar] [CrossRef]
  78. Zhao, T.; Li, J.; Chen, A.F. MicroRNA-34a induces endothelial progenitor cell senescence and impedes its angiogenesis via suppressing silent information regulator 1. Am. J. Physiol. Endocrinol. Metab. 2010, 299, E110–E116. [Google Scholar] [CrossRef]
  79. Zeng, Y.B.; Liang, X.H.; Zhang, G.X.; Jiang, N.; Zhang, T.; Huang, J.Y.; Zhang, L.; Zeng, X.C. miRNA-135a promotes hepatocellular carcinoma cell migration and invasion by targeting forkhead box O1. Cancer Cell Int. 2016, 16, 63. [Google Scholar] [CrossRef]
  80. Chatterjee, N.; Fraile-Bethencourt, E.; Baris, A.; Espinosa-Diez, C.; Anand, S. MicroRNA-494 regulates endoplasmic reticulum stress in endothelial cells. Front. Cell Dev. Biol. 2021, 9, 671461. [Google Scholar] [CrossRef]
  81. Lin, H.; Dai, T.; Xiong, H.; Zhao, X.; Chen, X.; Yu, C.; Li, J.; Wang, X.; Song, L. Unregulated miR-96 induces cell proliferation in human breast cancer by downregulating transcriptional factor FOXO3a. PLoS ONE 2010, 5, e15797. [Google Scholar] [CrossRef] [PubMed]
  82. Tran, H.; Brunet, A.; Grenier, J.M.; Datta, S.R.; Fornace, A.J., Jr.; DiStefano, P.S.; Chiang, L.W.; Greenberg, M.E. DNA repair pathway stimulated by the forkhead transcription factor FOXO3a through the Gadd45 protein. Science 2002, 296, 530–534. [Google Scholar] [CrossRef] [PubMed]
  83. Abe, W.; Nasu, K.; Nakada, C.; Kawano, Y.; Moriyama, M.; Narahara, H. miR-196b targets c-myc and Bcl-2 expression, inhibits proliferation and induces apoptosis in endometriotic stromal cells. Hum. Reprod. 2013, 28, 750–761. [Google Scholar] [CrossRef]
  84. Sun, W.; Li, J.; Zhou, L.; Han, J.; Liu, R.; Zhang, H.; Ning, T.; Gao, Z.; Liu, B.; Chen, X.; et al. The c-Myc/miR-27b-3p/ATG10 regulatory axis regulates chemoresistance in colorectal cancer. Theranostics 2020, 10, 1981–1996. [Google Scholar] [CrossRef] [PubMed]
  85. Maurel, M.; Dejeans, N.; Taouji, S.; Chevet, E.; Grosset, C.F. MicroRNA-1291-mediated silencing of IRE1alpha enhances Glypican-3 expression. RNA 2013, 19, 778–788. [Google Scholar] [CrossRef]
  86. Jiang, X.; Guo, S.; Xu, M.; Ma, B.; Liu, R.; Xu, Y.; Zhang, Y. TFAP2C-mediated lncRNA PCAT1 inhibits ferroptosis in docetaxel-resistant prostate cancer through c-Myc/miR-25-3p/SLC7A11 signaling. Front. Oncol. 2022, 12, 862015. [Google Scholar] [CrossRef]
  87. Yang, A.; Wang, X.; Yu, C.; Jin, Z.; Wei, L.; Cao, J.; Wang, Q.; Zhang, M.; Zhang, L.; Zhang, L.; et al. microRNA-494 is a potential prognostic marker and inhibits cellular proliferation, migration and invasion by targeting SIRT1 in epithelial ovarian cancer. Oncol. Lett. 2017, 14, 3177–3184. [Google Scholar] [CrossRef]
  88. Song, L.; Dai, T.; Xie, Y.; Wang, C.; Lin, C.; Wu, Z.; Ying, Z.; Wu, J.; Li, M.; Li, J. Up-regulation of miR-1245 by c-myc targets BRCA2 and impairs DNA repair. J. Mol. Cell Biol. 2012, 4, 108–117. [Google Scholar] [CrossRef]
  89. Christoffersen, N.R.; Shalgi, R.; Frankel, L.B.; Leucci, E.; Lees, M.; Klausen, M.; Pilpel, Y.; Nielsen, F.C.; Oren, M.; Lund, A.H. p53-independent upregulation of miR-34a during oncogene-induced senescence represses MYC. Cell Death Differ. 2010, 17, 236–245. [Google Scholar] [CrossRef]
  90. Xu, N.; Li, Z.; Yu, Z.; Yan, F.; Liu, Y.; Lu, X.; Yang, W. MicroRNA-33b suppresses migration and invasion by targeting c-Myc in osteosarcoma cells. PLoS ONE 2014, 9, e115300. [Google Scholar] [CrossRef]
  91. Li, H.; Liu, J.; Cao, W.; Xiao, X.; Liang, L.; Liu-Smith, F.; Wang, W.; Liu, H.; Zhou, P.; Ouyang, R.; et al. C-myc/miR-150/EPG5 axis mediated dysfunction of autophagy promotes development of non-small cell lung cancer. Theranostics 2019, 9, 5134–5148. [Google Scholar] [CrossRef] [PubMed]
  92. Mao, A.; Zhao, Q.; Zhou, X.; Sun, C.; Si, J.; Zhou, R.; Gan, L.; Zhang, H. MicroRNA-449a enhances radiosensitivity by downregulation of c-Myc in prostate cancer cells. Sci. Rep. 2016, 6, 27346. [Google Scholar] [CrossRef] [PubMed]
  93. Chen, P.C.; Yu, C.C.; Huang, W.Y.; Huang, W.H.; Chuang, Y.M.; Lin, R.I.; Lin, J.M.J.; Lin, H.Y.; Jou, Y.C.; Shen, C.H.; et al. c-Myc acts as a competing endogenous RNA to sponge miR-34a, in the upregulation of CD44, in urothelial carcinoma. Cancers 2019, 11, 1457. [Google Scholar] [CrossRef] [PubMed]
  94. Liang, T.; Hu, X.Y.; Li, Y.H.; Tian, B.Q.; Li, Z.W.; Fu, Q. MicroRNA-21 regulates the proliferation, differentiation, and apoptosis of human renal cell carcinoma cells by the mTOR-STAT3 signaling pathway. Oncol. Res. 2016, 24, 371–380. [Google Scholar] [CrossRef] [PubMed]
  95. Wei, L.; Chen, Z.; Cheng, N.; Li, X.; Chen, J.; Wu, D.; Dong, M.; Wu, X. MicroRNA-126 inhibit viability of colorectal cancer cell by repressing mTOR induced apoptosis and autophagy. OncoTargets Ther. 2020, 13, 2459–2468. [Google Scholar] [CrossRef]
  96. Ye, X.; Luo, H.; Chen, Y.; Wu, Q.; Xiong, Y.; Zhu, J.; Diao, Y.; Wu, Z.; Miao, J.; Wan, J. MicroRNAs 99b-5p/100-5p regulated by endoplasmic reticulum stress are involved in abeta-induced pathologies. Front. Aging Neurosci. 2015, 7, 210. [Google Scholar] [CrossRef]
  97. Luo, M.; Wu, L.; Zhang, K.; Wang, H.; Zhang, T.; Gutierrez, L.; O’Connell, D.; Zhang, P.; Li, Y.; Gao, T.; et al. miR-137 regulates ferroptosis by targeting glutamine transporter SLC1A5 in melanoma. Cell Death Differ. 2018, 25, 1457–1472. [Google Scholar] [CrossRef]
  98. Shirjang, S.; Mansoori, B.; Asghari, S.; Duijf, P.H.G.; Mohammadi, A.; Gjerstorff, M.; Baradaran, B. MicroRNAs in cancer cell death pathways: Apoptosis and necroptosis. Free. Radic. Biol. Med. 2019, 139, 1–15. [Google Scholar] [CrossRef]
  99. Xie, Y.; Zhao, Y.; Shi, L.; Li, W.; Chen, K.; Li, M.; Chen, X.; Zhang, H.; Li, T.; Matsuzawa-Ishimoto, Y.; et al. Gut epithelial TSC1/mTOR controls RIPK3-dependent necroptosis in intestinal inflammation and cancer. J. Clin. Investig. 2020, 130, 2111–2128. [Google Scholar] [CrossRef]
  100. Ma, Y.; Vassetzky, Y.; Dokudovskaya, S. mTORC1 pathway in DNA damage response. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 1293–1311. [Google Scholar] [CrossRef]
  101. Lv, X.; Li, C.Y.; Han, P.; Xu, X.Y. MicroRNA-520a-3p inhibits cell growth and metastasis of non-small cell lung cancer through PI3K/AKT/mTOR signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 2321–2327. [Google Scholar] [PubMed]
  102. Madhunapantula, S.V.; Sharma, A.; Robertson, G.P. PRAS40 deregulates apoptosis in malignant melanoma. Cancer Res. 2007, 67, 3626–3636. [Google Scholar] [CrossRef] [PubMed]
  103. Xie, Y.; Chen, L.; Zhou, J.; Yang, C.; Xu, C.; Fan, X.; Tan, Y.; Wang, Y.; Kang, C.; Fang, C. TGFbeta signaling-induced miRNA participates in autophagic regulation by targeting PRAS40 in mesenchymal subtype of glioblastoma. Cancer Biol. Med. 2020, 17, 664–675. [Google Scholar] [CrossRef] [PubMed]
  104. Khor, E.S.; Wong, P.F. Endothelial replicative senescence delayed by the inhibition of MTORC1 signaling involves MicroRNA-107. Int. J. Biochem. Cell Biol. 2018, 101, 64–73. [Google Scholar] [CrossRef]
  105. Jablonska, E.; Bialopiotrowicz, E.; Szydlowski, M.; Prochorec-Sobieszek, M.; Juszczynski, P.; Szumera-Cieckiewicz, A. DEPTOR is a microRNA-155 target regulating migration and cytokine production in diffuse large B-cell lymphoma cells. Exp. Hematol. 2020, 88, 56–67.e52. [Google Scholar] [CrossRef]
  106. Ahmadpour, F.; Igder, S.; Babaahmadi-Rezaei, H.; Khalili, E.; Kanani, M.; Soleimani, V.; Mohammadzadeh, G. Methylation-mediated silencing of miR-125a-5p facilitates breast cancer progression by inducing autophagy. Mol. Biol. Rep. 2022, 49, 6325–6339. [Google Scholar] [CrossRef]
  107. Li, Y.; Luo, Y.; Li, B.; Niu, L.; Liu, J.; Duan, X. miRNA-182/Deptor/mTOR axis regulates autophagy to reduce intestinal ischaemia/reperfusion injury. J. Cell. Mol. Med. 2020, 24, 7873–7883. [Google Scholar] [CrossRef]
  108. Chen, S.; Wu, J.; Jiao, K.; Wu, Q.; Ma, J.; Chen, D.; Kang, J.; Zhao, G.; Shi, Y.; Fan, D.; et al. MicroRNA-495-3p inhibits multidrug resistance by modulating autophagy through GRP78/mTOR axis in gastric cancer. Cell Death Dis. 2018, 9, 1070. [Google Scholar] [CrossRef]
  109. Li, X.; Chen, Y.T.; Josson, S.; Mukhopadhyay, N.K.; Kim, J.; Freeman, M.R.; Huang, W.C. MicroRNA-185 and 342 inhibit tumorigenicity and induce apoptosis through blockade of the SREBP metabolic pathway in prostate cancer cells. PLoS ONE 2013, 8, e70987. [Google Scholar] [CrossRef]
  110. Stacchiotti, A.; Grossi, I.; Garcia-Gomez, R.; Patel, G.A.; Salvi, A.; Lavazza, A.; De Petro, G.; Monsalve, M.; Rezzani, R. Melatonin effects on non-alcoholic fatty liver disease are related to microRNA-34a-5p/Sirt1 axis and autophagy. Cells 2019, 8, 1053. [Google Scholar] [CrossRef]
  111. Li, Y.; Zhang, J.; He, J.; Zhou, W.; Xiang, G.; Xu, R. MicroRNA-132 cause apoptosis of glioma cells through blockade of the SREBP-1c metabolic pathway related to SIRT1. Biomed. Pharmacother. 2016, 78, 177–184. [Google Scholar] [CrossRef] [PubMed]
  112. Ouimet, M.; Koster, S.; Sakowski, E.; Ramkhelawon, B.; van Solingen, C.; Oldebeken, S.; Karunakaran, D.; Portal-Celhay, C.; Sheedy, F.J.; Ray, T.D.; et al. Mycobacterium. tuberculosis. induces the miR-33 locus to reprogram autophagy and host lipid metabolism. Nat. Immunol. 2016, 17, 677–686. [Google Scholar] [CrossRef] [PubMed]
  113. Kanagasabai, T.; Li, G.; Shen, T.H.; Gladoun, N.; Castillo-Martin, M.; Celada, S.I.; Xie, Y.; Brown, L.K.; Mark, Z.A.; Ochieng, J.; et al. MicroRNA-21 deficiency suppresses prostate cancer progression through downregulation of the IRS1-SREBP-1 signaling pathway. Cancer Lett. 2022, 525, 46–54. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, N.; Zhang, H.; Liu, Y.; Su, P.; Zhang, J.; Wang, X.; Sun, M.; Chen, B.; Zhao, W.; Wang, L.; et al. SREBP1, targeted by miR-18a-5p, modulates epithelial-mesenchymal transition in breast cancer via forming a co-repressor complex with Snail and HDAC1/2. Cell Death Differ. 2019, 26, 843–859. [Google Scholar] [CrossRef]
  115. Fan, S.J.; Li, H.B.; Cui, G.; Kong, X.L.; Sun, L.L.; Zhao, Y.Q.; Li, Y.H.; Zhou, J. miRNA-149* promotes cell proliferation and suppresses apoptosis by mediating JunB in T-cell acute lymphoblastic leukemia. Leuk. Res. 2016, 41, 62–70. [Google Scholar] [CrossRef]
  116. Mendonca, D.B.; Nguyen, J.T.; Haidar, F.; Fox, A.L.; Ray, C.; Amatullah, H.; Liu, F.; Kim, J.K.; Krebsbach, P.H. MicroRNA-1911-3p targets mEAK-7 to suppress mTOR signaling in human lung cancer cells. Heliyon 2020, 6, e05734. [Google Scholar] [CrossRef]
  117. Lee, M.; Kim, E.J.; Jeon, M.J. MicroRNAs 125a and 125b inhibit ovarian cancer cells through post-transcriptional inactivation of EIF4EBP1. Oncotarget 2016, 7, 8726–8742. [Google Scholar] [CrossRef]
  118. Zhang, S.; Wang, M.; Li, Q.; Zhu, P. MiR-101 reduces cell proliferation and invasion and enhances apoptosis in endometrial cancer via regulating PI3K/Akt/mTOR. Cancer Biomark. 2017, 21, 179–186. [Google Scholar] [CrossRef]
  119. Wang, Y.; Dai, Y.X.; Wang, S.Q.; Qiu, M.K.; Quan, Z.W.; Liu, Y.B.; Ou, J.M. miR-199a-5p inhibits proliferation and induces apoptosis in hemangioma cells through targeting HIF1A. Int. J. Immunopathol. Pharmacol. 2018, 31, 749357. [Google Scholar] [CrossRef]
  120. Sun, Y.; Xing, X.; Liu, Q.; Wang, Z.; Xin, Y.; Zhang, P.; Hu, C.; Liu, Y. Hypoxia-induced autophagy reduces radiosensitivity by the HIF-1alpha/miR-210/Bcl-2 pathway in colon cancer cells. Int. J. Oncol. 2015, 46, 750–756. [Google Scholar] [CrossRef]
  121. Muratsu-Ikeda, S.; Nangaku, M.; Ikeda, Y.; Tanaka, T.; Wada, T.; Inagi, R. Downregulation of miR-205 modulates cell susceptibility to oxidative and endoplasmic reticulum stresses in renal tubular cells. PLoS ONE 2012, 7, e41462. [Google Scholar] [CrossRef]
  122. Xu, P.; Ge, F.H.; Li, W.X.; Xu, Z.; Wang, X.L.; Shen, J.L.; Xu, A.B.; Hao, R.R. MicroRNA-147a targets SLC40A1 to induce ferroptosis in human glioblastoma. Anal. Cell Pathol. 2022, 2022, 2843990. [Google Scholar] [CrossRef] [PubMed]
  123. Karshovska, E.; Wei, Y.; Subramanian, P.; Mohibullah, R.; Geissler, C.; Baatsch, I.; Popal, A.; Corbalan Campos, J.; Exner, N.; Schober, A. HIF-1alpha (hypoxia-inducible factor-1alpha) promotes macrophage necroptosis by regulating miR-210 and miR-383. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 583–596. [Google Scholar] [CrossRef] [PubMed]
  124. Crosby, M.E.; Kulshreshtha, R.; Ivan, M.; Glazer, P.M. MicroRNA regulation of DNA repair gene expression in hypoxic stress. Cancer Res. 2009, 69, 1221–1229. [Google Scholar] [CrossRef] [PubMed]
  125. Alique, M.; Bodega, G.; Giannarelli, C.; Carracedo, J.; Ramirez, R. MicroRNA-126 regulates Hypoxia-Inducible Factor-1alpha which inhibited migration, proliferation, and angiogenesis in replicative endothelial senescence. Sci. Rep. 2019, 9, 7381. [Google Scholar] [CrossRef] [PubMed]
  126. Byun, Y.; Choi, Y.C.; Jeong, Y.; Lee, G.; Yoon, S.; Jeong, Y.; Yoon, J.; Baek, K. MiR-200c downregulates HIF-1alpha and inhibits migration of lung cancer cells. Cell. Mol. Biol. Lett. 2019, 24, 28. [Google Scholar] [CrossRef] [PubMed]
  127. Ma, Y.; Feng, J.; Xing, X.; Zhou, B.; Li, S.; Zhang, W.; Jiang, J.; Zhang, J.; Qiao, Z.; Sun, L.; et al. miR-1908 overexpression inhibits proliferation, changing Akt activity and p53 expression in hypoxic NSCLC cells. Oncol. Res. 2016, 24, 9–15. [Google Scholar] [CrossRef]
  128. Musilova, K.; Mraz, M. MicroRNAs in B-cell lymphomas: How a complex biology gets more complex. Leukemia 2015, 29, 1004–1017. [Google Scholar] [CrossRef]
  129. Cora, D.; Bussolino, F.; Doronzo, G. TFEB signalling-related micrornas and autophagy. Biomolecules 2021, 11, 985. [Google Scholar] [CrossRef]
  130. Li, W.; Zhang, J.; Chen, T.; Yin, P.; Yang, J.; Cao, Y. miR-132 upregulation promotes gastric cancer cell growth through suppression of FoxO1 translation. Tumour Biol. 2015. [Google Scholar] [CrossRef]
  131. Sheng, X.; Nenseth, H.Z.; Qu, S.; Kuzu, O.F.; Frahnow, T.; Simon, L.; Greene, S.; Zeng, Q.; Fazli, L.; Rennie, P.S.; et al. IRE1alpha-XBP1s pathway promotes prostate cancer by activating c-MYC signaling. Nat. Commun. 2019, 10, 323. [Google Scholar] [CrossRef] [PubMed]
  132. Yang, Y.; Bardeleben, C.; Frost, P.; Hoang, B.; Shi, Y.; Finn, R.; Gera, J.; Lichtenstein, A. DEPTOR is linked to a TORC1-p21 survival proliferation pathway in multiple myeloma cells. Genes Cancer 2014, 5, 407–419. [Google Scholar] [CrossRef] [PubMed]
  133. Bao, C.; Zhang, J.; Xian, S.Y.; Chen, F. MicroRNA-670-3p suppresses ferroptosis of human glioblastoma cells through targeting ACSL4. Free Radic. Res. 2021, 55, 853–864. [Google Scholar] [CrossRef] [PubMed]
  134. Chen, J.; Ding, C.; Chen, Y.; Hu, W.; Yu, C.; Peng, C.; Feng, X.; Cheng, Q.; Wu, W.; Lu, Y.; et al. ACSL4 reprograms fatty acid metabolism in hepatocellular carcinoma via c-Myc/SREBP1 pathway. Cancer Lett. 2021, 502, 154–165. [Google Scholar] [CrossRef]
  135. Duran, R.V.; Oppliger, W.; Robitaille, A.M.; Heiserich, L.; Skendaj, R.; Gottlieb, E.; Hall, M.N. Glutaminolysis activates Rag-mTORC1 signaling. Mol. Cell 2012, 47, 349–358. [Google Scholar] [CrossRef]
  136. Cheng, Q.; Chen, M.; Liu, M.; Chen, X.; Zhu, L.; Xu, J.; Xue, J.; Wu, H.; Du, Y. Semaphorin 5A suppresses ferroptosis through activation of PI3K-AKT-mTOR signaling in rheumatoid arthritis. Cell Death Dis. 2022, 13, 608. [Google Scholar] [CrossRef]
  137. Wang, F.; Zhang, H.; Xu, N.; Huang, N.; Tian, C.; Ye, A.; Hu, G.; He, J.; Zhang, Y. A novel hypoxia-induced miR-147a regulates cell proliferation through a positive feedback loop of stabilizing HIF-1alpha. Cancer Biol. Ther. 2016, 17, 790–798. [Google Scholar] [CrossRef]
  138. Yao, S.; Fan, L.Y.; Lam, E.W. The FOXO3-FOXM1 axis: A key cancer drug target and a modulator of cancer drug resistance. Semin. Cancer Biol. 2018, 50, 77–89. [Google Scholar] [CrossRef]
  139. Qi, Y.; Qian, R.; Jia, L.; Fei, X.; Zhang, D.; Zhang, Y.; Jiang, S.; Fu, X. Overexpressed microRNA-494 represses RIPK1 to attenuate hippocampal neuron injury in epilepsy rats by inactivating the NF-kappaB signaling pathway. Cell Cycle 2020, 19, 1298–1313. [Google Scholar] [CrossRef]
  140. Xu, D.; Zou, C.; Yuan, J. Genetic Regulation of RIPK1 and Necroptosis. Annu. Rev. Genet. 2021, 55, 235–263. [Google Scholar] [CrossRef]
  141. Ju, Y.; Xu, T.; Zhang, H.; Yu, A. FOXO1-dependent DNA damage repair is regulated by JNK in lung cancer cells. Int. J. Oncol. 2014, 44, 1284–1292. [Google Scholar] [CrossRef] [PubMed]
  142. Khor, E.S.; Wong, P.F. The roles of MTOR and miRNAs in endothelial cell senescence. Biogerontology 2020, 21, 517–530. [Google Scholar] [CrossRef] [PubMed]
  143. Petroulakis, E.; Parsyan, A.; Dowling, R.J.; LeBacquer, O.; Martineau, Y.; Bidinosti, M.; Larsson, O.; Alain, T.; Rong, L.; Mamane, Y.; et al. p53-dependent translational control of senescence and transformation via 4E-BPs. Cancer Cell 2009, 16, 439–446. [Google Scholar] [CrossRef] [PubMed]
  144. Samec, M.; Liskova, A.; Kubatka, P.; Uramova, S.; Zubor, P.; Samuel, S.M.; Zulli, A.; Pec, M.; Bielik, T.; Biringer, K.; et al. The role of dietary phytochemicals in the carcinogenesis via the modulation of miRNA expression. J. Cancer Res. Clin. Oncol. 2019, 145, 1665–1679. [Google Scholar] [CrossRef]
  145. Zhang, B.; Tian, L.; Xie, J.; Chen, G.; Wang, F. Targeting miRNAs by natural products: A new way for cancer therapy. Biomed. Pharmacother. 2020, 130, 110546. [Google Scholar] [CrossRef]
  146. Alnuqaydan, A.M. Targeting micro-RNAs by natural products: A novel future therapeutic strategy to combat cancer. Am. J. Transl. Res. 2020, 12, 3531–3556. [Google Scholar]
  147. Wen, X.Y.; Wu, S.Y.; Li, Z.Q.; Liu, Z.Q.; Zhang, J.J.; Wang, G.F.; Jiang, Z.H.; Wu, S.G. Ellagitannin (BJA3121), an anti-proliferative natural polyphenol compound, can regulate the expression of miRNAs in HepG2 cancer cells. Phytother. Res. 2009, 23, 778–784. [Google Scholar] [CrossRef]
  148. Wang, L.; Ho, J.; Glackin, C.; Martins-Green, M. Specific pomegranate juice components as potential inhibitors of prostate cancer metastasis. Transl. Oncol. 2012, 5, 344–355. [Google Scholar] [CrossRef]
  149. Yang, F.; Nam, S.; Brown, C.E.; Zhao, R.; Starr, R.; Ma, Y.; Xie, J.; Horne, D.A.; Malkas, L.H.; Jove, R.; et al. A novel berbamine derivative inhibits cell viability and induces apoptosis in cancer stem-like cells of human glioblastoma, via up-regulation of miRNA-4284 and JNK/AP-1 signaling. PLoS ONE 2014, 9, e94443. [Google Scholar] [CrossRef]
  150. Weng, H.; Huang, H.; Dong, B.; Zhao, P.; Zhou, H.; Qu, L. Inhibition of miR-17 and miR-20a by oridonin triggers apoptosis and reverses chemoresistance by derepressing BIM-S. Cancer Res. 2014, 74, 4409–4419. [Google Scholar] [CrossRef]
  151. Chang, Y.; Zhao, Y.; Gu, W.; Cao, Y.; Wang, S.; Pang, J.; Shi, Y. Bufalin inhibits the differentiation and proliferation of cancer stem cells derived from primary osteosarcoma cells through Mir-148a. Cell Physiol. Biochem. 2015, 36, 1186–1196. [Google Scholar] [CrossRef]
  152. Yang, J.; Qiu, B.; Li, X.; Zhang, H.; Liu, W. p53-p66(shc)/miR-21-Sod2 signaling is critical for the inhibitory effect of betulinic acid on hepatocellular carcinoma. Toxicol. Lett. 2015, 238, 1–10. [Google Scholar] [CrossRef]
  153. Chen, F.; Zhong, Z.; Tan, H.Y.; Guo, W.; Zhang, C.; Cheng, C.S.; Wang, N.; Ren, J.; Feng, Y. Suppression of lncRNA MALAT1 by betulinic acid inhibits hepatocellular carcinoma progression by targeting IAPs via miR-22-3p. Clin. Transl. Med. 2020, 10, e190. [Google Scholar] [CrossRef]
  154. Karki, K.; Hedrick, E.; Kasiappan, R.; Jin, U.H.; Safe, S. Piperlongumine induces reactive oxygen species (ROS)-dependent downregulation of specificity protein transcription factors. Cancer Prev. Res. 2017, 10, 467–477. [Google Scholar] [CrossRef] [PubMed]
  155. Ma, D.; Wei, J.; Chen, S.; Wang, H.; Ning, L.; Luo, S.H.; Liu, C.L.; Song, G.; Yao, Q. Fucoidan Inhibits the progression of hepatocellular carcinoma via causing lncRNA LINC00261 overexpression. Front. Oncol. 2021, 11, 653902. [Google Scholar] [CrossRef] [PubMed]
  156. Chen, J.; Xia, Y.; Sui, X.; Peng, Q.; Zhang, T.; Li, J.; Zhang, J. Steviol, a natural product inhibits proliferation of the gastrointestinal cancer cells intensively. Oncotarget 2018, 9, 26299–26308. [Google Scholar] [CrossRef]
  157. Ozbey, U.; Attar, R.; Romero, M.A.; Alhewairini, S.S.; Afshar, B.; Sabitaliyevich, U.Y.; Hanna-Wakim, L.; Ozcelik, B.; Farooqi, A.A. Apigenin as an effective anticancer natural product: Spotlight on TRAIL, WNT/β-catenin, JAK-STAT pathways, and microRNAs. J. Cell. Biochem. 2019, 120, 1060–1067. [Google Scholar] [CrossRef]
  158. Gao, A.M.; Zhang, X.Y.; Ke, Z.P. Apigenin sensitizes BEL-7402/ADM cells to doxorubicin through inhibiting miR-101/Nrf2 pathway. Oncotarget 2017, 8, 82085–82091. [Google Scholar] [CrossRef]
  159. Yao, C.L.; Zhang, J.Q.; Li, J.Y.; Wei, W.L.; Wu, S.F.; Guo, D.A. Traditional Chinese medicine (TCM) as a source of new anticancer drugs. Nat. Prod. Rep. 2021, 38, 1618–1633. [Google Scholar] [CrossRef] [PubMed]
  160. Wu, Y.; Zhang, J.; Hong, Y.; Wang, X. Effects of kanglaite injection on serum miRNA-21 in patients with advanced lung cancer. Med. Sci. Monit. 2018, 24, 2901–2906. [Google Scholar] [CrossRef]
  161. Matarlo, J.S.; Krumpe, L.R.H.; Heinz, W.F.; Oh, D.; Shenoy, S.R.; Thomas, C.L.; Goncharova, E.I.; Lockett, S.J.; O’Keefe, B.R. The natural product butylcycloheptyl prodiginine binds pre-miR-21, inhibits Dicer-mediated processing of pre-miR-21, and blocks cellular proliferation. Cell. Chem. Biol. 2019, 26, 1133–1142.e1134. [Google Scholar] [CrossRef] [PubMed]
  162. Zhou, J.; Liu, M.; Chen, Y.; Xu, S.; Guo, Y.; Zhao, L. Cucurbitacin B suppresses proliferation of pancreatic cancer cells by ceRNA: Effect of miR-146b-5p and lncRNA-AFAP1-AS1. J. Cell. Physiol. 2019, 234, 4655–4667. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, B.; Wang, X.; Deng, J.; Zheng, H.; Liu, W.; Chen, S.; Tian, J.; Wang, F. p53-dependent upregulation of miR-16-2 by sanguinarine induces cell cycle arrest and apoptosis in hepatocellular carcinoma. Cancer Lett. 2019, 459, 50–58. [Google Scholar] [CrossRef]
  164. Ahmed Youness, R.; Amr Assal, R.; Mohamed Ezzat, S.; Zakaria Gad, M.; Abdel Motaal, A. A methoxylated quercetin glycoside harnesses HCC tumor progression in a TP53/miR-15/miR-16 dependent manner. Nat. Prod. Res. 2020, 34, 1475–1480. [Google Scholar] [CrossRef]
  165. Yang, J.; Zou, Y.; Jiang, D. Honokiol suppresses proliferation and induces apoptosis via regulation of the miR21/PTEN/PI3K/AKT signaling pathway in human osteosarcoma cells. Int. J. Mol. Med. 2018, 41, 1845–1854. [Google Scholar]
  166. Yi, X.; Lou, L.; Wang, J.; Xiong, J.; Zhou, S. Honokiol antagonizes doxorubicin resistance in human breast cancer via miR-188-5p/FBXW7/c-Myc pathway. Cancer Chemother. Pharmacol. 2021, 87, 647–656. [Google Scholar] [CrossRef]
  167. Hu, Z.; Zhou, X.; Zeng, D.; Lai, J. Shikonin induces cell autophagy via modulating the microRNA -545-3p/guanine nucleotide binding protein beta polypeptide 1 axis, thereby disrupting cellular carcinogenesis in colon cancer. Bioengineered 2022, 13, 5928–5941. [Google Scholar] [CrossRef]
  168. Meng, L.; Dian, F.; Yinan, X.; Hao, Z.; Dingyue, Z. Anticancer effect of natural product sulforaphane by targeting MAPK signal through miRNA-1247-3p in human cervical cancer cells. Biointerface Res. Appl. Chem. 2020, 11, 7943–7972. [Google Scholar]
  169. Li, X.; Zhao, Z.; Li, M.; Liu, M.; Bahena, A.; Zhang, Y.; Zhang, Y.; Nambiar, C.; Liu, G. Sulforaphane promotes apoptosis, and inhibits proliferation and self-renewal of nasopharyngeal cancer cells by targeting STAT signal through miRNA-124-3p. Biomed. Pharmacother. 2018, 103, 473–481. [Google Scholar] [CrossRef]
  170. Farooqi, A.A.; Butt, G.; El-Zahaby, S.A.; Attar, R.; Sabitaliyevich, U.Y.; Jovic, J.J.; Tang, K.F.; Naureen, H.; Xu, B. Luteolin mediated targeting of protein network and microRNAs in different cancers: Focus on JAK-STAT, NOTCH, mTOR and TRAIL-mediated signaling pathways. Pharmacol. Res. 2020, 160, 105188. [Google Scholar] [CrossRef]
  171. Wu, S.; Li, X.; Liu, Z. Camptothecin inhibits migration, invasion and clonogenic property of liver cancer cells by modulating microRNA expression. Acta Pol. Pharm.-Drug Res. 2020, 77, 295–304. [Google Scholar]
  172. Zheng, M.; Wu, Y. Piceatannol suppresses proliferation and induces apoptosis by regulation of the microRNA21/phosphatase and tensin homolog/protein kinase B signaling pathway in osteosarcoma cells. Mol. Med. Rep. 2020, 22, 3985–3993. [Google Scholar] [PubMed]
  173. Fu, S.; Zhao, N.; Jing, G.; Yang, X.; Liu, J.; Zhen, D.; Tang, X. Matrine induces papillary thyroid cancer cell apoptosis in vitro and suppresses tumor growth in vivo by downregulating miR-182-5p. Biomed. Pharmacother. 2020, 128, 110327. [Google Scholar] [CrossRef] [PubMed]
  174. Hernandes, C.; Miguita, L.; de Sales, R.O.; Silva, E.P.; Mendonca, P.O.R.; Lorencini da Silva, B.; Klingbeil, M.F.G.; Mathor, M.B.; Rangel, E.B.; Marti, L.C.; et al. Anticancer activities of the quinone-methide triterpenes maytenin and 22-beta-hydroxymaytenin obtained from cultivated Maytenus. ilicifolia. roots associated with down-regulation of miRNA-27a and miR-20a/miR-17-5p. Molecules 2020, 25, 760. [Google Scholar] [CrossRef] [PubMed]
  175. Chen, L.; Yang, Z.S.; Zhou, Y.Z.; Deng, Y.; Jiang, P.; Tan, S.L. Dihydromyricetin inhibits cell proliferation, migration, invasion and promotes apoptosis via regulating miR-21 in human cholangiocarcinoma cells. J. Cancer 2020, 11, 5689–5699. [Google Scholar] [CrossRef]
  176. Yan, M.D.; Yao, C.J.; Chow, J.M.; Chang, C.L.; Hwang, P.A.; Chuang, S.E.; Whang-Peng, J.; Lai, G.M. Fucoidan elevates microRNA-29b to regulate DNMT3B-MTSS1 axis and inhibit EMT in human hepatocellular carcinoma cells. Mar. Drugs 2015, 13, 6099–6116. [Google Scholar] [CrossRef]
  177. Bie, B.; Sun, J.; Li, J.; Guo, Y.; Jiang, W.; Huang, C.; Yang, J.; Li, Z. Baicalein, a natural anti-cancer compound, alters microRNA expression profiles in Bel-7402 human hepatocellular carcinoma cells. Cell. Physiol. Biochem. 2017, 41, 1519–1531. [Google Scholar] [CrossRef]
  178. Liu, W.; Pan, H.-F.; Yang, L.-J.; Zhao, Z.-M.; Yuan, D.-S.; Liu, Y.-L.; Lin, L.-Z. Panax. ginseng. CA Meyer (Rg3) ameliorates gastric precancerous lesions in Atp4a−/− mice via inhibition of glycolysis through PI3K/AKT/miRNA-21 pathway. Evid. Based. Complement. Altern. Med. 2020, 2020, 2672648. [Google Scholar]
  179. Carpi, S.; Polini, B.; Manera, C.; Digiacomo, M.; Salsano, J.E.; Macchia, M.; Scoditti, E.; Nieri, P. miRNA modulation and antitumor activity by the extra-virgin olive oil polyphenol oleacein in human melanoma cells. Front. Pharmacol. 2020, 11, 574317. [Google Scholar] [CrossRef]
  180. Wang, Q.; Wang, Z.; Hou, G.; Huang, P. Toosendanin suppresses glioma progression property and induces apoptosis by regulating miR-608/Notch axis. Cancer Manag. Res. 2020, 12, 3419–3431. [Google Scholar] [CrossRef]
  181. Peng, F.; Xiong, L.; Peng, C. (-)-Sativan inhibits tumor development and regulates miR-200c/PD-L1 in triple negative breast cancer cells. Front. Pharmacol. 2020, 11, 251. [Google Scholar] [CrossRef] [PubMed]
  182. Khan, M.A.; Tania, M.; Fu, J. Epigenetic role of thymoquinone: Impact on cellular mechanism and cancer therapeutics. Drug Discov. Today 2019, 24, 2315–2322. [Google Scholar] [CrossRef] [PubMed]
  183. Okda, T.M.; Katry, M.A.; Ragab, N.M.; Shalkami, A.S. Phytic acid potentiates oxaliplatin effects in colorectal cancer induced by 1,2-DMH: The role of miR-224 and miR-200a. Contemp. Oncol. 2021, 25, 118–124. [Google Scholar]
  184. Alsadi, N.; Mallet, J.F.; Matar, C. miRNA-200b signature in the prevention of skin cancer stem cells by polyphenol-enriched blueberry preparation. J. Cancer Prev. 2021, 26, 162–173. [Google Scholar] [CrossRef] [PubMed]
  185. Wang, W.S.; Zhao, C.S. Formononetin exhibits anticancer activity in gastric carcinoma cell and regulating miR-542-5p. Kaohsiung. J. Med. Sci. 2021, 37, 215–225. [Google Scholar] [CrossRef] [PubMed]
  186. Zhou, M.; Zhang, G.; Hu, J.; Zhu, Y.; Lan, H.; Shen, X.; Lv, Y.; Huang, L. Rutin attenuates sorafenib-induced chemoresistance and autophagy in hepatocellular carcinoma by regulating BANCR/miRNA-590-5P/OLR1 axis. Int. J. Biol. Sci. 2021, 17, 3595–3607. [Google Scholar] [CrossRef]
  187. Li, Q.; Xu, D.; Gu, Z.; Li, T.; Huang, P.; Ren, L. Rutin restrains the growth and metastasis of mouse breast cancer cells by regulating the microRNA-129-1-3p-mediated calcium signaling pathway. J. Biochem. Mol. Toxicol. 2021, 35, e22794. [Google Scholar] [CrossRef]
  188. Zhang, F.; Ni, Z.J.; Ye, L.; Zhang, Y.Y.; Thakur, K.; Cespedes-Acuna, C.L.; Han, J.; Zhang, J.G.; Wei, Z.J. Asparanin A inhibits cell migration and invasion in human endometrial cancer via Ras/ERK/MAPK pathway. Food Chem. Toxicol. 2021, 150, 112036. [Google Scholar] [CrossRef]
  189. Xu, N.; Zhao, Y.; Bu, H.; Tan, S.; Dong, G.; Liu, J.; Wang, M.; Jiang, J.; Yuan, B.; Li, R. Cochlioquinone derivative CoB1 induces cytostatic autophagy in lung cancer through miRNA-125b and Foxp3. Phytomedicine 2021, 93, 153742. [Google Scholar] [CrossRef]
  190. Zhang, R.; Pan, T.; Xiang, Y.; Zhang, M.; Xie, H.; Liang, Z.; Chen, B.; Xu, C.; Wang, J.; Huang, X.; et al. Curcumenol triggered ferroptosis in lung cancer cells via lncRNA H19/miR-19b-3p/FTH1 axis. Bioact. Mater. 2022, 13, 23–36. [Google Scholar] [CrossRef]
  191. Magura, J.; Moodley, R.; Mackraj, I. The effect of hesperidin and luteolin isolated from Eriocephalus. africanus. on apoptosis, cell cycle and miRNA expression in MCF-7. J. Biomol. Struct. Dyn. 2022, 40, 1791–1800. [Google Scholar] [CrossRef] [PubMed]
  192. Gasparello, J.; Papi, C.; Zurlo, M.; Gambari, L.; Rozzi, A.; Manicardi, A.; Corradini, R.; Gambari, R.; Finotti, A. Treatment of human glioblastoma U251 cells with sulforaphane and a Peptide Nucleic Acid (PNA) targeting miR-15b-5p: Synergistic effects on induction of apoptosis. Molecules 2022, 27, 1299. [Google Scholar] [CrossRef]
  193. Yin, S.; Jin, W.; Qiu, Y.; Fu, L.; Wang, T.; Yu, H. Solamargine induces hepatocellular carcinoma cell apoptosis and autophagy via inhibiting LIF/miR-192-5p/CYR61/Akt signaling pathways and eliciting immunostimulatory tumor microenvironment. J. Hematol. Oncol. 2022, 15, 32. [Google Scholar] [CrossRef] [PubMed]
  194. Zhang, C.; Ji, X.; Chen, Z.; Yao, Z. Asiaticoside suppresses gastric cancer progression and induces endoplasmic reticulum stress through the miR-635/HMGA1 axis. J. Immunol. Res. 2022, 2022, 1917585. [Google Scholar] [CrossRef] [PubMed]
  195. Yu, R.; Zhou, Y.; Shi, S.; Wang, X.; Huang, S.; Ren, Y. Icariside II induces ferroptosis in renal cell carcinoma cells by regulating the miR-324-3p/GPX4 axis. Phytomedicine 2022, 102, 154182. [Google Scholar] [CrossRef] [PubMed]
  196. Atteia, H.H.; Arafa, M.H.; Mohammad, N.S.; Amin, D.M.; Sakr, A.T. Thymoquinone upregulates miR-125a-5p, attenuates STAT3 activation, and potentiates doxorubicin antitumor activity in murine solid Ehrlich carcinoma. J. Biochem. Mol. Toxicol. 2021, 35, e22924. [Google Scholar] [CrossRef]
  197. Tewari, D.; Patni, P.; Bishayee, A.; Sah, A.N.; Bishayee, A. Natural products targeting the PI3K-Akt-mTOR signaling pathway in cancer: A novel therapeutic strategy. Semin. Cancer Biol. 2022, 80, 1–17. [Google Scholar] [CrossRef]
Figure 1. Overview of AKT, AKT effectors, and miRNAs and their regulation of multiple cell functions in cancer cells.
Figure 1. Overview of AKT, AKT effectors, and miRNAs and their regulation of multiple cell functions in cancer cells.
Ijms 24 03688 g001
Figure 2. Chemical structures of the natural products mentioned in Table 3.
Figure 2. Chemical structures of the natural products mentioned in Table 3.
Ijms 24 03688 g002
Figure 3. Schematic diagram of the natural products affecting the regulation of miRNAs that target AKT and AKT effectors to regulate multiple cell functions in cancer cells. More detailed information is provided in each of the tables above.
Figure 3. Schematic diagram of the natural products affecting the regulation of miRNAs that target AKT and AKT effectors to regulate multiple cell functions in cancer cells. More detailed information is provided in each of the tables above.
Ijms 24 03688 g003
Table 4. Connecting natural product-affected miRNAs to predict targeted AKT and AKT effectors.
Table 4. Connecting natural product-affected miRNAs to predict targeted AKT and AKT effectors.
AKT1/2/3AKT Effectors AKT1/2/3AKT Effectors
AKT1AKT2AKT3FOXO1FOXO3MYCMTORAKT1S1DEPTORHIF1A AKT1AKT2AKT3FOXO1FOXO3FOXO4MYCMTORDEPTORHIF1A
let-7f-1-3p FOXO1 miR-223-3p FOXO1FOXO3
miR-101-3p AKT3 mTOR miR-22-3p AKT3
miR-103a-3p AKT2 miR-23b-3p FOXO3FOXO4
miR-106a-5p AKT3 miR-23c FOXO3FOXO4
miR-106b-5p HIF1AmiR-24-3p FOXO4
miR-107 AKT2 miR-27a-3p FOXO1FOXO3
miR-122-5p AKT3 FOXO3 miR-27b-3p FOXO1FOXO3
miR-124-3p AKT2AKT3 AKT1S1DEPTOR miR-29b-3p AKT2AKT3 FOXO3FOXO4
miR-125a-5p AKT1S1 miR-29c-3p AKT2AKT3 FOXO3FOXO4
miR-125b-5p AKT1S1 miR-302a-5p HIF1A
miR-129-1-3p AKT1S1 miR-324-5p FOXO1
miR-1303 FOXO3 miR-381-3p AKT3 FOXO3 DEPTOR
miR-133b FOXO3 miR-384 FOXO3
miR-143-3p FOXO1 miR-421 FOXO3 MTOR
miR-144-3p FOXO1 MTOR miR-424-5p AKT3 DEPTOR
miR-145-5p AKT3FOXO1 miR-4262 AKT3 DEPTOR
miR-148a-3p AKT2 miR-4284 FOXO3
miR-151a-3p AKT3 HIF1AmiR-486-5p FOXO1
miR-15a-5p AKT3 miR-503-5p AKT3
miR-15b-5p AKT3 DEPTOR miR-506-3p AKT3 DEPTOR
miR-16-2 MYC HIF1AmiR-513a-3p FOXO1
miR-16-5p AKT3 DEPTOR miR-518c-5p AKT3
miR-17-5p AKT3 HIF1AmiR-518f-5p AKT3 FOXO3
miR-181a-5p AKT3 DEPTOR miR-5195-3p AKT3FOXO1
miR-181b-5p AKT3 DEPTOR miR-520h DEPTORHIF1A
miR-181c-5p AKT3 DEPTOR miR-545-3p AKT3 MTORDEPTOR
miR-182-5p FOXO1FOXO3 DEPTOR miR-590-5p FOXO3
miR-183-5p FOXO1 miR-608 FOXO4
miR-186-5p AKT3 HIF1AmiR-616-3p MTOR
miR-18a-5p HIF1AmiR-622 HIF1A
miR-18b-5p HIF1AmiR-629-5p AKT3 MYC
miR-195-5p AKT3 DEPTOR miR-7-1-3pAKT1 FOXO1 HIF1A
miR-199a-5p HIF1AmiR-7-5p AKT3
miR-203a-3p MYC HIF1AmiR-93-5p AKT3 HIF1A
miR-20a-5p AKT3 HIF1AmiR-9-5p FOXO3
miR-21-5p FOXO3 miR-96-5p FOXO1 MTORDEPTOR
miR-22-3P AKT3 miR-99a-5p MTOR
The above natural product-affected miRNAs were derived from Table 3. Some AKT effectors are not listed because they could not be retrieved using the miRDB database (date: 11 November 2022).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shiau, J.-P.; Chuang, Y.-T.; Yen, C.-Y.; Chang, F.-R.; Yang, K.-H.; Hou, M.-F.; Tang, J.-Y.; Chang, H.-W. Modulation of AKT Pathway-Targeting miRNAs for Cancer Cell Treatment with Natural Products. Int. J. Mol. Sci. 2023, 24, 3688. https://doi.org/10.3390/ijms24043688

AMA Style

Shiau J-P, Chuang Y-T, Yen C-Y, Chang F-R, Yang K-H, Hou M-F, Tang J-Y, Chang H-W. Modulation of AKT Pathway-Targeting miRNAs for Cancer Cell Treatment with Natural Products. International Journal of Molecular Sciences. 2023; 24(4):3688. https://doi.org/10.3390/ijms24043688

Chicago/Turabian Style

Shiau, Jun-Ping, Ya-Ting Chuang, Ching-Yu Yen, Fang-Rong Chang, Kun-Han Yang, Ming-Feng Hou, Jen-Yang Tang, and Hsueh-Wei Chang. 2023. "Modulation of AKT Pathway-Targeting miRNAs for Cancer Cell Treatment with Natural Products" International Journal of Molecular Sciences 24, no. 4: 3688. https://doi.org/10.3390/ijms24043688

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