Next Article in Journal
Development and Evaluation of Duplex MIRA-qPCR Assay for Simultaneous Detection of Staphylococcus aureus and non-aureus Staphylococci
Previous Article in Journal
Ferulic Acid Esterase Producing Lactobacillus johnsonii from Goat Feces as Corn Silage Inoculants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Do Bacteria Provide an Alternative to Cancer Treatment and What Role Does Lactic Acid Bacteria Play?

Department of Microbiology, Stellenbosch University, Private Bab X1, Matieland, Stellenbosch 7602, South Africa
*
Author to whom correspondence should be addressed.
Microorganisms 2022, 10(9), 1733; https://doi.org/10.3390/microorganisms10091733
Submission received: 28 July 2022 / Revised: 17 August 2022 / Accepted: 26 August 2022 / Published: 27 August 2022
(This article belongs to the Section Food Microbiology)

Abstract

:
Cancer is one of the leading causes of mortality and morbidity worldwide. According to 2022 statistics from the World Health Organization (WHO), close to 10 million deaths have been reported in 2020 and it is estimated that the number of cancer cases world-wide could increase to 21.6 million by 2030. Breast, lung, thyroid, pancreatic, liver, prostate, bladder, kidney, pelvis, colon, and rectum cancers are the most prevalent. Each year, approximately 400,000 children develop cancer. Treatment between countries vary, but usually includes either surgery, radiotherapy, or chemotherapy. Modern treatments such as hormone-, immuno- and antibody-based therapies are becoming increasingly popular. Several recent reports have been published on toxins, antibiotics, bacteriocins, non-ribosomal peptides, polyketides, phenylpropanoids, phenylflavonoids, purine nucleosides, short chain fatty acids (SCFAs) and enzymes with anticancer properties. Most of these molecules target cancer cells in a selective manner, either directly or indirectly through specific pathways. This review discusses the role of bacteria, including lactic acid bacteria, and their metabolites in the treatment of cancer.

1. Introduction

Cancer is one of the leading causes of death, according to 2022 statistics published by the American Cancer Society [1]. At the 17th World Health Assembly [2], the WHO passed the resolution “cancer prevention and control in the context of an integrated approach” and predicted that the number of cancer cases world-wide could increase to 21.6 million by 2030. The American Cancer Society predicted 1.9 million new cancer cases and 609 360 deaths in the United States of America in 2022, rating cancer as the second leading cause of death [1]. According to the latest statistics from the South African National Cancer Registry (NCR), published by the National Institute for Communicable Diseases (NICD), 85 302 new cancer cases have been reported in 2019 [3]. Of all cancers, breast, lung, prostate, colon, rectum, bladder, kidney, renal, pelvis, pancreatic, thyroid and liver cancer are most common [4].
Cancer is normally treated with radiotherapy or chemotherapy and tumours are surgically removed. Less severe cancers are treated with hormone-, immune- and antibody-based therapies [5]. Most of these treatments, however, do not discriminate between normal cells and cancer cells, and vary concerning the ability to infiltrate tumours. Patients that receive radiotherapy or chemotherapy complain of adverse side effects such as flu-like symptoms, heart problems, diarrhoea, nausea, lockjaw (trismus) and chronic bladder spams [6,7,8,9]. Radiotherapy of the neck and head may elicit difficulty in swallowing (dysphagia), dry mouth feel (xerostomia), necrosis, inflammation of the spinal cord, and even permanent trismus or neurological damage [6,10]. Approximately 90% of patients with advanced cancer experience severe pain after surgery [11] and need to take opioids. This may lead to drug abuse [12]. In some cases, cancer cells have developed resistance to conventional treatments [13], which emphasizes the need to search for alternative treatments and anticancer drugs. This led to the search for anticancer compounds produced by plants, marine organisms, fungi, algae, and bacteria [14]. Several toxins, antibiotics, bacteriocins, non-ribosomal peptides, polyketides, phenylpropanoids, phenylflavonoids, purine nucleosides, short chain fatty acids (SCFAs) and enzymes with anticancer properties, mostly produced by bacteria, have been described [15]. The challenge is to identify compounds that only target cancer cells and not normal cells.
Cancer cells differ from normal cells by having more fluidic cell membranes [16], a net negative charge (due to elevated levels of phosphatidylserine, O-glycosylated mucins, sialylated gangliosides and heparin sulfates), and more microvilli, thus a larger surface area [17,18,19,20]. Because of these characteristics, cationic peptides such as bacteriocins may find it easier to adhere to cancer cells than normal cells [21,22].
The aim of this review is not to discuss various cancers and symptoms, nor the advantages and disadvantages of conventional therapies, but to provide the reader with the latest developments in the use of bacterial toxins, antibiotics, bacteriocins, non-ribosomal peptides, polyketides, phenylpropanoids, phenylflavonoids, purine nucleosides, SCFAs and enzymes in cancer treatment. The possibility of using lactic acid bacteria or their bacteriocins as anticancer agents is investigated.

2. Bacterial-Mediated Cancer Therapy

Experiments with viable or attenuated microorganisms to treat cancer dates back more than a century [7,23,24]. The first cancer vaccine, composed of viable Streptococcus pyogenes cells, was developed by Dr Coley in 1891 [23,24,25]. The cells activated macrophages and lymphocytes, and stimulated production of tumour necrosis factor α (TNFα) that regulates inflammatory responses required to attack malignant neoplasm [26,27]. The “Coley-toxin”, that consisted of heat-treated culture supernatants of S. pyogenes and Serratia marcescens, was rejected by the USA Food and Drug Administration (FDA) in 1962 due to lack of unequivocal scientific support and reports of organ damage [23,28]. Subsequent reports of Clostridium, Corynebacterium, Bacillus Calmette-Guérin, Salmonella, Escherichia coli, Bifidobacterium and Listeria associated with cancer cells [29,30,31,32,33] led to renewed interest in the search for bacterial cells with anticancer properties, especially species autochthonous to the human gut. Several papers have been published advocating the use of Salmonella in cancer treatment.
Salmonella typhimurium VNP20009 was made less toxic by deleting msbB encoding lipopolysaccharide (LPS) production [34]. By deleting purI the strain became auxotrophic for adenine [35]. In a later study, King et al. [36] engineered strain VNP20009 to express cytosine deaminase (CD) that converts 5-fluorocytosine (5-FC) to 5-fluorouracil (5-FU). Fluorouracil, commercially known as Adrucil, is a cytotoxin used in treatment of colorectal, oesophageal, stomach, pancreatic, breast, and cervical cancers [37]. In murine models 5-FU formed at tumour sites, resulting in a dramatic repression of cell growth [36,38]. A CD-expressing strain of Salmonella enterica yielded similar results when tested in mice [39], raising the hope that engineered strains of Salmonella may be used to activate cytotoxic drugs within tumour cells. Another strain of S. typhimurium (TAPET-CD, also referred to as VNP20029) expressed genes encoding CD and colonized tumour cells for at least 15 days. The strain converted 5-FC to 5-FU in 2 of the 3 patients treated [40]. No reports of follow-up clinical trials with any of these strains have been reported.
Inducible promoters responding to specific conditions in tumour cells (e.g., hypoxia) were used to design delivery systems for anti-cancer treatment [41]. The hypoxia-inducible promoter (HIP) and a fluorescent marker were cloned into S. typhimurium VNP20009 [42]. A 15-fold increase in fluorescence was recorded in HCT116 human colorectal carcinoma cells. Non-cancerous cells did not fluoresce [42], which suggested the Salmonalla delivery vector is very specific. Other HIPs experimented with included those regulating the expression of pflE and ansB [43]. Swofford et al. [44] transformed the luxI/luxR quorum-sensing system of Vibrio fischeri into S. typhimurium VPN200010. These reporter genes were expressed in 4T1 mammary tumours, but only when cell densities were above a certain threshold.
Other strains of S. typhimurium tested for anti-cancer therapy included strain AR-1, an arginine and leucine auxotroph defective in ppGpp synthesis (strain ΔppGpp) [45], and strain SF200 with mutations in lipid A and flagella synthesis [46]. Clinical trials on metastatic cancer patients intravenously injected with maximum tolerable cell numbers of strain VNP20009 (3 × 108 cfu/m2) proved ineffective [47]. Higher cell numbers led to anaemia, low platelet counts, bacteraemia, high blood bilirubin, nausea, vomiting, diarrhoea, hypophosphatemia, and an increase in alkaline phosphatase [47]. A genetically engineered strain of VNP20009 that produced TNF-related apoptosis-inducing ligand (TRAIL) under control of a γ-irradiation-inducible RecA promoter, stimulated caspase-3-mediated apoptosis in 4T1 mammary carcinoma cells [48]. S. typhimurium ΔppGpp that expressed tissue inhibitor of metalloproteinases 2 (TIMP-2) reduced the size of glioma brain tumours in BALB/c mice and increased survival by 60% [49]. Modification of S. typhimurium ΔppGpp to express the mitochondrial targeting domain of Noxa (MTD), fused with the cell-penetrating peptide DS4.3 (DS4.3-MTD), led to the complete destruction of colon carcinoma tumours [50]. An engineered strain of Salmonella choleraesuis that expressed the angiogenesis inhibitor thrombospondin-1 inhibited the growth of B16F10 melanoma cells in mice [51]. No clinical trial data have been published on any of these genetically engineered strains.
A strain of S. typhimurium genetically engineered to produce truncated human interleukin-2 (SalpIL2), reduced adenocarcinoma metastases of the liver [52,53,54,55]. Sorensen et al. [56] reported that metastatic osteosarcoma in a mouse model could be treated with a single oral dose of attenuated SalpIL2-producing S. typhimurium. Barnett et al. [57] have shown that SalpIL2 reduced the volume and mass of retroperitoneal neuroblastoma tumours in a murine model. In vitro experiments have shown that SalpIL2-producing strains invade and divide within K7M2 osteosarcoma cells [58], suggesting that SalpIL2 may persist for long periods in malignant tissue.
Significant lysis of tumor cells in mice were recorded with Clostridium histolyticum treatment [58]. Shrinking of tumor cells were observed when these cells were exposed to Clostridium tetani [59]. Clostridium novyi, made non-pathogenic by deleting the gene encoding α-toxin NT, destroyed tumours and secreted liposomase [60]. The latter has been experimented with in enhancing the release of liposome-encapsulated drugs within tumours [61]. Endospores of C. novyi-NT colonized the hypoxic regions of tumours, elicited cell lysis and an immune inflammatory response that resulted in immunogenic cell death [60,62]. In a phase I study, injection of C. novyi-NT into tumour cells led to a decrease in the size of the cells, resulted in a systemic cytokine response and enhanced systemic tumor-specific T-cell responses [63]. Bettegowda et al. [64] reported a long-term remediation effect when tumours in a mice model were treated with a combination of C novyi-NT endospores and radiation therapy. The authors suggested that a combination of radioactive iodine with C novyi-NT might enable patients to be treated with lower doses of radiolabeled antibodies, which limits injury to normal tissue such as bone marrow.
Nuyts et al. [65] has shown that the recA and recN genes in Clostridium acetobutylicum DSM792 could be activated with a radiation dose of 2 Gy. The authors argued that the activation of the recA promoter could increase TNFα production in recombinant clostridia. In a later study, Jiang et al. [66] showed that E coli K12, harbouring plasmid pAClyA, produced higher levels of cytolysin A, which enhanced the therapeutic effects of radiation.
Cloning the genes encoding nitroreductase from E. coli to Clostridium beijerinckii resulted in the activation of CB1954 (nontoxic) into a drug with anticancer properties. Intravenous injection of activated CB1954 into mice destroyed tumours [67]. Clostridium spp. engineered to deliver CD to tumours [68] have also been modified to secrete TNFα [69] and may, in future, be used in cancer therapy. Li et al. [70] suppressed the growth of Heps mouse liver cancer cells in vivo by using a genetically engineered strain of Bifidobacterium adolescentis. Bifodobacterium spp. have also been used to deliver active CD enzymes to hypoxic regions of solid tumours in mice [71,72,73,74].
The first report using Listeria monocytogenes to direct an immune response to tumours was published by Pan et al. [75]. The authors have genetically engineered L. monocytogenes to secrete influenza virus nucleoprotein and have shown that the protein could repress tumours in colon and renal cancer models. Several other model studies were published using recombinant strains of Listeria to repress cervical, head and neck, breast, skin, and renal cancers; reviewed by Guirnalda et al. [76] and Cory and Chu [77]. An immunotherapy-based treatment for cervical cancer was developed based on live attenuated L. monocytogenes that secretes the fusion protein Lm-LLO-E7 [78]. The protein, referred to as ADXS11-001 (ADXS-HPV), targets human papillomavirus (HPV)-associated tumours. The immune response elicited by the ADXS11-001 vaccine against HPV oncoprotein E7 led to the reduction in tumour cells in animal models. Phase I and II clinical studies were later conducted [79,80,81]. Clinical trials are being conducted to evaluate another genetically engineered vaccine, protein ADXS-504, for treatment of biochemically recurrent (early) prostate cancer (https://www.globenewswire.com/en/news-release/2020/09/24/2098634/0/en/Advaxis-Announces-FDA-Clearance-of-New-IND-for-ADXS-504-for-Treatment-of-Prostate-Cancer.html#:~:text=filingsmedia%20partners-,Advaxis%20Announces%20FDA%20Clearance%20of%20New%20IND%20for%20ADXS%2D504,%7C%20Source%3A%20Advaxis%2C%20Inc; https://www.globenewswire.com/news-release/2022/06/29/2471231/0/en/Advaxis-Updates-on-the-Phase-1-Clinical-Trial-of-ADXS-504-for-the-Treatment-of-Early-Prostate-Cancer.html, assessed on 15 August 2022).
E. coli Nissle 1917 was genetically engineered to convert NH3 to L-arginine (L-arg). The recombinant strain, referred to as SYNB1020, reduced systemic hyperammonemia in mouse models [82]. A phase 1 clinical study showed that SYNB1020 was well tolerated at daily doses of up to 1.5 × 1012 cfu (colony-forming units) administered for up to 14 days. An increase in urinary nitrate, plasma 15N-nitrate and urinary 15N-nitrate was reported, suggesting that SYNB1020 could be used to treat hyperammonemia, including urea cycle disorders and hepatic encephalopathy [82]. Another genetically engineered strain of E. coli Nissle 1917, strain SYNB1618, yielded promising results when tested for the ability to alleviate phenylketonuria (PKU), a disorder caused by defective phenylalanine hydrolase, thus the inability to convert phenylalanine (Phe) to tyrosine [83]. Dose-responsive increases were observed in plasma (trans-cinnamic acid) and urine (hippuric acid) levels of Phe metabolites, suggesting that genetically engineered E. coli may be used in the treatment of rare metabolic disorders [83].
Intestinal bacteria influence various inflammatory and immune processes, many of which are implicated in tumour etiology, such as in colorectal cancer (CRC) [84]. Bacteroides fragilis and Fusobacterium spp. are directly associated with tumours, including CRC [85,86]. Fusobacterium nucleatum suppresses the immune response that leads to the induction of chronic inflammation [85]. Bacteroides fragilis alters (damages) the DNA of host cells, increases cell proliferation, and induces pro-inflammatory processes through the production of toxins [86]. A gene encoding B. fragilis toxin detected in colonic mucosa is associated with late-stage CRC [86]. Further research is required to determine if these species could be genetically modified to prevent the proliferation of cancer cells.
Although gut microbiota may prevent CRC, they also pose a risk of inducing CRC. This is mostly diet related. High levels of secondary bile acids (BAs) are produced from a high fat content diet [87,88]. Abnormal high levels of BAs in the colon induces inflammation [89,90] and forms reactive oxygen species that disrupts cell membranes and mitochondria [88]. Species primarily responsible for production of BAs are Clostridium scindens, Clostridium hiranonis, Clostridium hylemonae and Clostridium sordellii [91]. A diet rich in proteins and low in carbohydrates my also cause CRC, as reported by Russel et al. [92]. Fermentation of proteins in the distal colon leads to the production of toxic ammonia, amines, phenols and sulfides [93]. Lithocholic acid (LCA), a derivative of cholic acid, is an exception to the rule, as it inhibits the growth of human prostate cancer cells LNCaP and PC-3 by induction of caspase-3, 8 and 9 mediated apoptosis [94]. LCA not only induces endoplasmic reticulum (ER) stress that triggers the unfolded protein response (UPR) activating cell death [95], but transforms growth factor-β in HepG2 liver cancer cells, and suppresses the growth of breast cancer cells [96]. In addition, LCA induces oxidative phosphorylation, inhibits epithelial-mesenchymal transition and expression of vascular endothelial growth factor A, and stimulates antitumor immunity [96].
The anticarcinogenic properties of lactic acid bacteria (LAB) is addressed in far fewer publications and focuses mainly on exopolysaccharides (EPS), peptidoglycan, nucleic acid, bacteriocins, and S-layer proteins [97]. Viable cells of Lactobacillus casei BL23, intranasally administered using the human pappilomavirus (HPV)-induced model, reduced tumour growth [98]. Lactobacillus reuteri BCRC14652, tested in vitro, damaged the cell membranes of colon carcinoma HT29 cells [99], suppressed tumor necrosis factor (TNF)-induced NF-κB activation, and repressed the growth of cancer cells by apoptosis [100]. EPS produced by Lactobacillus acidophilus and Lactobacillus rhamnosus repressed the growth of HT-29 cells by inducing the activity of Beclin-1 (an autophagy protein) and GRP78 (an endoplasmic reticulum chaperone) directly, and indirectly by regulating apoptosis through stimulation of Bcl-2 (B-cell lymphoma 2) and Bak, a pro-apoptotic gene of the Bcl-2 family [101]. A combination of L. acidophilus and L. casei, used with 5-FU, induced apoptosis of LS513 cancer cells [102], suggesting that these species may be used as adjuvants in anticancer chemotherapy.
Anti-tumor activities were also reported for cell-free supernatants of LAB, and irradiation-inactivated and heat-killed cells of LAB [99,102,103,104,105]. Exopolysaccharides (EPS) produced by L. casei 01 reduced the cytotoxicity of 4-nitroquinoline N-oxide (4-NQO), a pro-mutagen [106]. EPS isolated from L. acidophilus 606 repressed the growth of cancerous cells [103] and EPS from Lactobacillus plantarum and L. acidophilus significantly reduced tumour growth [107,108,109]. An interesting observation is that the repression of tumour growth exerted by an EPS-producing strain of L. acidophilus (strain LA1) may be associated with the suppression of lactate dehydrogenase (LDH) and alkaline phosphatase (ALP). Inhibition of LDH in the glycolytic pathway of cancer cells results in lower ATP production, hence slower growth [110,111]. EPS116, produced by L. plantarum NCU116, binds to TLR2 and activates the TLR2/MyD88/TRAF6/MKK7 pathway, which, in turn, activates JNK/c-Jun that upregulates the transcription and translation of Fas and Fasl. The Fas/Fasl signaling pathway activates FADD of caspase-8 and caspase-3. Activated Caspase-3 facilitates apoptosis by upregulating the expression of cellular target proteins PARPs and Rock1, followed by cleavage of PARP1 and inhibition of CT26 growth [112]. Many probiotic LAB produce EPS and may, in future, be used as alternative or complementary treatment of cancer. The anticancer properties of EPS are reviewed by Wu et al. [113].
Several reports highlighted the importance of LAB in the prevention of CRC, reviewed by Zhong [114]. The health and quality of life of patients that underwent surgical resection of CRC were significantly improved when administered L. acidophilus and Bacillus natto [115]. Despite this, there is no consensus on the role LAB play in CRC treatment. It is, however, certain that a select few LAB activate mechanisms involved in the killing or repression of cancer cells, and that they regulate immune response [116]. This includes neutralizing free radicals [117] and inactivation of reactive oxygen species (ROS) by NADH oxidase/peroxidase and catalase [118,119]. Strains of Bifidobacterium longum and L. acidophilus displayed antioxidative activity by inhibiting linoleic acid peroxidation [120]. Heat-killed cells of L. acidophilus 606 and EPS produced by the strain has potent antioxidative activity [103]. According to Kumar et al. [121] and Annuk et al. [122], obligate homofermentative lactobacilli display high antioxidant activity, but is highly strain-dependent among facultative and obligate heterofermentative lactobacilli.
LAB play an important role in stimulating the immune system, especially anti-inflammatory cytokine IL-10 [123]. Lipoteichoic acid, present in the cell walls of all LAB, stimulate DCs through Toll-like receptor 2, resulting in the release of cytokines [124,125]. Some lactobacilli stimulate DCs to produce IL-12 and IL-10 [126,127]. Disruption of LTA in L. acidophilus resulted in the production of IL-10 by DCs and the downregulation of IL-12, which led to T cell-mediated colitis in mice [128]. It thus seems possible to treat CRC by altering the cell surface components of L. acidophilus. A strain of L. acidophilus deficient in LTA (strain NCK2025) repressed the growth of colonic polyps by downregulating IL-12, TNF-α and IL-10. This activated CD4+ T-cells, as observed in a mice model [129]. In vivo studies have shown an increase in cytoplasmic levels of TNF-α, interferon-γ (IFN-γ) and IL-10 in animals administered L. casei and B. longum [130]. This correlated with an increase in T cells, NK cells and MHC class II+ cells, and CD4-CD8+ T cells in a murine model [131]. L. casei Shirota (LcS) suppressed chemically induced carcinogenesis [131], supported by an increase in IFN-γ, interleukin-β (IL-1 β) and TNF-α levels [132]. A butanol extract prepared from the cell-free supernatant of B. adolescentis significantly increased the production of TNF-α and NO, which regulated immune modulation and repressed tumor growth [133].
Messenger RNA (mRNA) and interferon gamma (IFN-γ) levels of leukemia KHYG-1 cells increased when treated with a combination of Lactococcus lactis subsp. lactis, Lactococcus lactis subsp. cremoris, L. Lactococcus lactis subsp. lactis biovar diacetylactis, L. plantarum, Leuconostoc mesenteroides subsp. cremoris, and L. casei [134]. The six LAB enhanced the cytotoxicity to human chronic myelogenous leukemia K562 cells and colorectal tumor HCT116 cells [134]. L. reuteri ATCC-PTA-6475 reduced the growth of mammary tumors in Swiss mice by blocking NFκ-B-p65 nuclear translocation and the expression of c-jun, an oncogenic transcription factor [135]. Phenyllactic acid (PLA), hydroxyphenyllactic acid (OH-PLA), lactic acid, and indole lactic acid (ILA) produced by L. plantarum UM55 inhibited the growth of Aspergillus flavus and thus production of carcinogenic aflatoxins in food products [136].
Peptidoglycan isolated from Bifidobacterium infantis ATCC 15,697 repressed the growth of Meth A fibrosarcoma in BALB/c mice [137]. A cell wall-derived polysaccharide-peptidoglycan complex (PSPG) from L. casei Shirota prevented the activation of IL-6/STAT3 signalling and repressed ileal cancer [138]. Peptidoglycan from Lactococcus and Bifidobacterium inhibited the growth of bladder cancer HT-1376, colon cancer DLD-1 and SNUC2A cells, and kidney cancer A498 cells [139]. S-layer proteins from L. acidophilus CICC 6074 up-regulating the expression of p53, p21, and p16 and down-regulated the expression of CDK1 (cyclin-dependent kinase) and cyclin B in colon cancer HT-29 cells [140].
RNA extracted from the cell-free supernatant of Lactobacillus DM9811 inhibited the growth of colon cancer HT-29 cells and mouse ascites hepatoma cells [141]. The authors ascribed the anticarcinogenic activity to increased activity of NK and CD4+ T cells and the upregulation of cellular immunity. DNA fragments of Lactobacillus bulgaricus and Streptococcus thermophilus promoted mitosis in mouse spleen B and Pierre spot cells, resulting in enhanced immune functions [141]. These studies may be the first towards nucleic acid-based vaccines in cancer therapy.
Probiotic LAB and bifidobacteria inhibit signaling of epidermal growth factor receptor (EGFR) pathways [142], leading to an increase in phosphorylation of cytoplasmic tyrosine kinase domains. This causes the activation of cell proliferation, apoptosis, migration and differentiation [143].
Jackson Laboratory (JAX) mice, known to be easily colonized by commensal microbes, had higher cell numbers of Bifidobacterium in their colons. These mice showed reduced growth of skin cancer cells and had higher levels of antitumor cytotoxic T lymphocytes (CTL). Species linked to antitumor immune responses were Bifidobacterium breve, B. longum and B. adolescentis [144]. Mice devoid of these species recovered from melanoma and showed an increased in tumour specific CTLs when administered B. breve or B. longum [144]. Once administered, bifidobacteria proliferate in the nutrient-rich environment created by cell death and necrosis [144]. The specific mechanism by which bifidobacteria or other intestinal bacteria stimulate antitumor immune responses is unknown. They may stimulate the maturation of dendritic cells (DCs) that, similar to antigen-presenting cells (APC), play a role in T-cell activation. B. longum BB536 stimulate the development and maturation of interferon γ (IFN-γ) secreting cells. Newborn infants showed an increase in the ratio of IFN-γ/IL-4 secreting T helper (Th) cells (Th1/Th2) when they received B. longum BB536 [145].
Enterococcus faecalis downregulates the expression of the FIAF (angiopoietin-like protein 4) gene associated with the development of some cancer types [146]. In a mouse model of ulcerative colitis, E. faecalis inhibited inflammation by suppressing T helper (Th)-1 and Th17 responses [147]. Heat-killed cells of E. faecalis YM-73 enhanced immune modulation by increasing Th1 and reducing Th2-associated cytokines [148].
Gut bacteria could also be used to reduce normal tissue damage during or after radiotherapy. Several studies have shown that probiotic strains of Lactobacillus and Bifidobacterium reduce radiotherapy side effects [149,150,151,152,153].
LAB are of paramount importance in the prevention of CRC. Insights into the cellular and molecular mechanisms which include apoptosis, antioxidant, immune responses, and epigenetics opened the door for the development of novel therapeutic approaches. Although a wide range of studies has shown the remarkable potential of LAB strains in interfering with colorectal carcinogenesis, conclusive clinical evidence supporting the role of probiotics in CRC treatment is still lacking. More epigenetic studies on LAB are required to demonstrate their effects in cancer prevention. Although several mechanisms of action of LAB in carcinogenesis have been described in in vitro and animal model studies, we are still far from pinpointing the exact cellular signals.

3. Bacterial Toxins

Bacterial toxins with medicinal applications have been well studied, especially those produced by Bacillus thuringiensis [154], Clostridium spp. and Bacillus spp. [155]. Botulinum neurotoxin (BoNT), produced by Clostridium botulinum, is widely used in ophthalmology, dermatology, and neurology [156]. To date, 108 clinical trials have been registered to evaluate the anticancer properties of bacterial toxins and an additional 98 trials have been registered to study the anticancer properties of immunotoxins (https://www.clinicaltrials.gov/, accessed on 1 June 2022).

3.1. Diphtheria Toxin

In 1884, Loeffler injected a pure culture of Clostridium diphtheriae into rabbits and pigeons and described the formation of lesions in several organs [157]. Follow-up studies have shown that these lesions were caused by diphtheria toxin (DT), encoded by the tox gene located on the genome of corynebacteriophage β [157,158]. The toxin targets elongation factor II (aminoacyl transferase II) and inhibits peptide synthesis [159,160]. The mode of action of DT is summarised in Figure 1.
In many human cancer cells, including hepatocarcinoma, melanoma, and colon, breast, myeloma, prostate, bladder and oral tumours, the HBEGF gene, encoding the membrane-anchored precursor of proHB-EGF, is significantly upregulated [161]. Increased levels of proHB-EGF and HB-EGF have been implicated in resistance to chemotherapeutic agents [162]. Cross-reactive material 197 (CRM197), a non-toxic variation of DT, binds to pro-HB-EGF and HB-EGF and inhibits the mitogenic action of HB-EGF by preventing its binding to ErbB (epidermal growth factor) receptors [163]. CRM197 has been used in the treatment of oral squamous cell carcinoma [161]. Tumour formation was completely inhibited in vivo when CRM197 was used in combination with cisplatin, a chemotherapeutic drug [164]. In humans, CRM197 led to an increase in neutrophil and TNF α levels, and a decrease in lymphocyte numbers. The drug, also referred to as BK-UM, proved effective in the treatment of resistant cancers [162,164]. Nam et al. [164] tested the efficacy of intravenously administered BK-UM against the triple-negative breast cancer cell line MDA-MB-231. The size of tumours treated with BK-UM decreased after two weeks. Furthermore, no adverse side effects such as weight loss were reported. In phase I clinical studies, BK-UM was tested against recurrent ovarian and peritoneal cancer cells [162]. Patients were administered four dose levels (1.0, 2.0, 3.3 and 5.0 mg/m2), of which 2.0 mg/m2 was the most effective. Those treated with 3.3 mg/m2 complained of nausea, hypotension, fever, and irritation of the peritoneum [162]. A truncated version of DT, in which the cell receptor-binding domain was replaced by proteins that selectively binds to the surface of cancer cells (referred to as DT385), inhibited angiogenesis and decreased tumour growth in chick chorioallantoic membranes [165]. DT385 also inhibited the growth of Lewis lung carcinoma (LLC) tumours in mice models [165]. Although the results were promising, DT385 entered cells without its R-domain, which triggered non-specific toxicity. To circumvent the problem, immunotoxins that target specific receptors on cancer cells and biochemical processes were developed and approved by the FDA in 1978 [166].
First-generation immunotoxins developed were antibodies randomly linked to non-binding toxins or derivatives thereof [167]. These immunotoxins were non-specific and targeted several receptors on cell surfaces. Second-generation immunotoxins were without cell surface-receptor domains and were much larger, which made penetration of solid tumours more difficult. Both generations of immunotoxins caused severe side effects, such as vascular leak syndrome, haemolytic uremic syndrome and pleuritis [166]. Progress in recombinant DNA technology led to the development of third-generation immunotoxins with less side-effects and improved tumour-penetrating properties. In these immunotoxins the non-specific receptor-binding domain is replaced by the Fv domain of an antibody, either genetically or chemically. Despite being more specific in the targeting of cancer cells, third-generation immunotoxins were not effective in patients vaccinated against diphtheria [168]. Attempts to suppress immune responses with anti-monoclonal antibodies [169] and binding to polyethylene glycol (PEGylation) were unsuccessful [170]. This led to the developing of humanized immunotoxins, i.e., immunotoxins containing a human protein with anti-cancer properties [168]. The first FDA approved humanized immunotoxin, Denileukin Diftitox (Ontak, DAB389 IL-2, Eisai Medical Research, Inc., Tokyo, Japan), was constructed by fusing human interleukin-2 (IL-2) with fragment A of DT [171]. IL-2 was genetically fused to the first 388 amino acids of DT, thereby replacing the R-domain [167]. Success rates with Ontak ranged from 30% to 50% [172]. Side effects reported were nausea, diarrhoea and vascular leak syndrome [173]. Despite difficulties encountered in purification of the first recombinant protein expressed by E. coli [173], several humanized immunotoxins were developed, all based on DT (Table 1).

3.2. Clostridium perfringens Enterotoxin

Clostridium perfringens enterotoxin (CPE) is a polypeptide of 319 amino acids [183], arranged in three domains. Domain I represents the binding domain (residues 162 to 309) and domains II and III the cytotoxic domains [184,185]. CPE is produced intracellularly and is only released during endospore germination [183]. Mutations in the TM1 region of CPE (amino acids 81 to 106) resulted in loss of membrane insertion, indicating that it plays a role in pore formation. The mode of action of CPE is illustrated in Figure 2. The amphipathic TM1 region forms a β-hairpin and a β-barrel pore once inserted into the membrane [186]. Amino acids at positions 45 to 53, located upstream of the TM1 region, are responsible for CPE oligomerization [187]. Domains II and III consist of eight β-sheets (two of which span the entire length of the module), two α-helices, and two 310 helical segments [185]. Domain I is a nine-stranded β-sandwich [188]. Briggs et al. [189] reported only two domains. Despite the discrepancy in the number of domains, both studies agreed on the structure of CPE and that the N-terminal domain is divided into two sections.
CPE, encoded by genes α, β, ε and ι located on the genome or on a plasmid, are transcriptionally regulated [190]. The spo0A gene, encoding the master sporulation regulator Spo0A, in conjunction with NanR, a transcriptional regulator, and three sporulation-associated sigma factors (SigE, SigK and SigF), are responsible for CPE production and gene regulation [191]. Transcriptional regulation of CPE expression has not been fully elucidated. However, a proposed mechanism of regulation has been compiled from various studies [188,192,193]. An Agr-like quorum-sensing (QS) system phosphorylates Spo0A, which activates the transcription of SigF, SigG, SigK and SigE [190]. SigK and SigE are required for CPE production, as they bind promotors to the cpe open reading frame (ORF) [190]. Three promoters (P1-P3) are located upstream of the ORF, with P1 being SigK-dependent, and P2 and P3 SigE-dependent [194].
The β-hairpin loops inserted in the membrane of the target cell create a pore through which cations are channelled. CPE binds to claudins (CLDNs) 3, 4, 5, 6, 7, 8, 9, 14 and 19 [187,188,195,196]. CLDN 4 is localized at tight junctions in normal human prostate epithelial cells (PrECs) but are distributed along the entire surface of cancer PrECs [197]. The size of cancerous tumours treated with CPE was reduced by 59%, suggesting that CLDN 4-targetted CPE treatment may be used to treat prostate cancer. Abedi et al. [198] constructed a recombinant plasmid containing the CPE and prostate stem cell antigen (PSCA) and named it pBudCE4.1-CPE-PSCA. The expression of transgenes introduced into cancer cells, referred to as suicide gene therapy, may be the first step towards developing a vaccine against prostate cancer. The authors [198] reported a 62.6% death rate of PC3 prostate cancer cells. Genes encoding apoptosis were overexpressed, whereas genes encoding cell cycling were repressed. The influx of Ca2+ ions activates Ca2-dependent proteases and causes cell lysis. Cytochrome C is released from the mitochondrion and caspase 3/7 is activated, leading to apoptosis [13,187,188,199,200,201].
Pahle et al. [202] used suicide gene therapy to treat mice with colorectal cancer. The authors amplified cDNA of CPE by PCR from plasmid pCpG-optCPE to construct a translation-optimized CPE vector (optCPE) and fused the amplicon to genes encoding green fluorescent protein (GFP), resulting in pcDNA3-optCPE-GFP (optCPE-GFP). This construct, and recombinant CPR (recCPE) were used to transfect different cell lines, including CaCo-2 and HT-29, and isogenic Sk-Mel5 and Sk-Mel5 Cldn-3-YFP melanoma cell lines. Colon carcinoma cell lines that overexpressed CLDN 3 and 4 were highly sensitive to recCPE and optCPE, but cells transfected with optCPE displayed rapid cytotoxic effects such as membrane disruption and necrosis. This suggested that suicide gene therapy may be used to suppress colon cancer in cells overexpressing CLDN 3 and 4. Gabig et al. [203] compared the cytotoxicity of CPE against the chemotherapeutics Dasatinib (Das) and Mitomycin C (MMC) used in the treatment of bladder cancer. The cells were killed within one hour when exposed to CPE, compared to 24 h when treated with Das or MMC. Furthermore, after one hour of treatment, 75% of primary bladder cancer cells died (in a 3D culture). Normal cells and cells derived from highly aggressive tumours survived all treatments.
Despite the success of CPE with experimental models, its use in treatment of cancer is limited due to the abundance of CLDNs in normal cells [204]. Shim et al. [200] tested DOX-C-SNP (doxorubicin-loaded C-CPE-polysialic acid) nanoparticles against pancreatic tumour cells in vitro and in vivo and have shown that DOX-C-SNPs accumulated only in tumour cells, without displaying significant cytotoxicity towards non-target cells [200]. A 5.9-fold increase in apoptosis was recorded in orthotopic murine models. Gao et al. [204] reported a decrease in CLDN 4 expression when epithelial ovarian cells (EOCs) were treated with C-CPE and ascribed this to the disruption of TJ proteins. CLDN 4+ EOC cell lines were also more sensitive to chemotherapeutic agents, as shown with 59% suppression of tumour growth when cells were treated with a combination of C-CPE and Taxol [204]. Treatment with C-CPE also resulted in the upregulation of genes in the ubiquitin-proteasome pathway that regulates apoptosis and angiogenesis, and downregulated genes involved in metabolic pathways. Becker et al. [205] linked gold nanoparticles (AuNPs) to C-CPE to form a C-CPE-AuNP complex that targets CLDN-overexpressing cancer cells. The Strep-Tag Strep-Tactin fusion system developed by Becker et al. [205] could also be used to conjugate C-CPE to chromophores, thereby allowing imaging and detection of cancer cells. Photonic activation of the AuNPs, referred to as AuNP-mediated laser perforation (GNOME-LP), used in combination with C-CPE is highly specific and targets only CLDN+ cells. A 30% and 40% reduction in cell viability was recorded for MCF-7 and OE-33 cells, respectively [205]. C-CPE-targeted GNOME-LP had no significant effect on the survival of cells in the control group. Gabig et al. [203] have shown that C-CPE treatment of RT4 (non-invasive superficial) cancer cells enhanced the toxicity of Das and MMC. Moreover, a drastic decrease in CLDN 4 expression was recorded, without affecting normal cells [203]. Nanoparticles loaded with fluorescent rhodamine dye and superparamagnetic iron oxide, linked to C-CPE (CPE290–319) were used to target CLDN 3 and CLDN 4 in cancer cells [206]. This technique may be used to determine the aggressiveness of cancer tumours.

3.3. Botulinum Toxins

Botulinum toxins (BoNT), produced by C. botulinum, Clostridium butyrricum, Clostridium barati, and Clostridium argentinensis are used in the treatment of muscle disorders [207], anismus [208], tremors [209], dystonia [210], cancer [211], and severe pain [212]. The inactive single-chain polypeptide is nicked by a protease to form a di-peptide of 100 kDa and 50 kDa [207]. The light chain (LC) is located at the N-terminal and contains the catalytic domain (C-domain), whereas the heavy chain (HC) is divided into the central translocation (T) domain and the C-terminal receptor-binding (R) domain [213]. The toxin associates with non-toxic neurotoxin-associated proteins (NAPs) to form a 300 to 900 kDa protoxin resistant to stomach acid and improved ability to be translocated across the intestinal epithelial barrier [207]. Eight types of botulinum toxins (A, B, C1, C2, D, E, F, and G) and a novel serotype (BoNT/H), isolated from an infant with botulism, have been described [207,214]. The mode of action of BoNT is illustrated in Figure 3.
BoNT binds to ecto-acceptors (polysialogangliosides) on the presynaptic cell surface of cholinergic neurons and is internalized via synaptic vesicles (SVs) or membrane-trafficking proteins (synaptotagmins). Acidification of the SVs leads to the activation of acetylcholine (Ach) transporter proteins that import Ach, the LC and HC domains of BoNT. Heat shock protein 90 (Hsp90) and thioredoxin reductase-thioredoxin (TrxR-Trx) cleaves the LC and liberates it into the cytosol. The C-domain is a Zn2+-dependent endopeptidase that cleaves proteins in the soluble SNARE protein complex. This complex is responsible for exocytosis and the fusing of Ach-containing vesicles with the plasma membrane, allowing the release of Ach. For more information BoNT mode of action, the reader is referred Nigam and Nigam [156], Gul et al. [213], Dolly et al. [215], Choudhury et al. [216] and Huang et al. [217].
Huang et al. [217] performed one of the earliest studies using BoNT/A to treat HIT-T15 insulinoma cells. Although the authors did not determine whether BoNT/A can be used to kill cancer cells, they showed that a transient transfection with the toxin can inhibit insulin expression. This paved the way for more in vivo and in vitro studies using BoNT. Treatment with BoNT/A render cancer cells radiosensitive [218] and may be used in the treatment breast and prostate cancer [211,219]. BoNT may also be used as an immunotoxin [220]. Toxin A induces apoptosis, inhibits the proliferation of LNCaP (infraclinical prostate cancer) cells, as shown with in vitro and in vivo studies [219]. In a separate study [221], toxin A showed cytotoxicity against cell lines LNCaP and PC-3 (prostate cancer), most probably due to the phosphorylation of phospholipase A2. Cell death of T47D breast cancer cells was attributed to the induction of caspase 3- and 7-dependent apoptosis [222]. Toxin C induced apoptosis and cell death in differentiated human neuroblastoma cells (SH-SY5Y and SiMa) [223]. Other cell-line anticancer studies performed with BoNT are listed in Table 2.
BoNT inhibits the release of neurotransmitters [207] and may be used as a painkiller in cancer treatment. Van Daele et al. [6] were the first to report on the analgesic effect of BoNT/A. The authors treated patients with painful spams of the sternocleidomastoid muscle. Injection with BoNT/A relieved the pain in four of six patients. Wang et al. [228] administered BoNT/A to a lung cancer patient with Raynaud phenomenon and previously treated with chemotherapy. Conventional treatment of neoplasms is generally ineffective. However, after BoNT/A treatment patients reported relief in symptoms, with no adverse side effects. Incobotulinumtoxin A (INCO), a BoNT/A preparation used in clinical settings, has also been used in the treatment of cancer-related pain [12]. Twelve patients with head, neck and breast cancer were enrolled in the study. Two patients passed away due to advanced cancer, one developed a skin rash, and another did not return due to poor general health. Three of the remaining eight patients reported an improvement in their quality of life. Pain amelioration was assessed using the Visual Analog Scale (VAS). All eight patients reported a significant improvement. A significant satisfaction of treatment was reported by seven of the eight patients by self-assessment using the Patients’ Global Impression of Change Scale. Dana et al. [10] used Botox and Dysport to test the efficacy of BoNT/A in pain relief of neck and head cancer patients suffering from radiotherapy-induced trismus and masticator spasms. One month after BoNT/A injection, a significant improvement in pain and spams were recorded, but no improvement in trismus. No adverse side effects were observed and the authors concluded that BoNT/A may be prescribed to patients with radiotherapy-induced pain or muscle spasms. De Groef et al. [229] studied the effect of a single BoNT/A injection in conjunction with physical therapy on breast cancer patients that underwent a mastectomy. Of the 50 patients, 25 received the injection (intervention) and 25 a placebo. After three months of treatment, a significant change in pain was observed in the upper limb of patients from the intervention group. Other studies using BoNT/A in the treatment of cancer-related or cancer therapy-related pain that were successful included post-radiosurgical neck contracture [230], frontotemporal glioblastoma related pain [231], and postoperative pain in patients that underwent a mastectomy and tissue expander reconstruction [232].

3.4. Pseudomonas aeruginosa Toxin

Pseudomonas aeruginosa produces a potent 66-kDa A-B toxin (PE, also known as exotoxin A or ETA) that inhibits protein translation [233,234]. The A domain has enzymatic activity, and the B domain acts as a cell-binding moiety [235,236]. The first 25 amino acids form a highly hydrophobic signal sequence that is removed during secretion [237]. The remaining 613 amino acids is divided into three domains. Domain Ia is the receptor binding domain (first 252 amino acids) that attaches to target cells, domain II (amino acids 253–364) facilitates the translocation of PE across the cell membrane, and domain Ib (amino acids aa 365–404) together with domain III (amino acids 405–613), represents the catalytic part of toxin PE [238]. When secreted, the terminal amino acid residue of PE (lysine 613) is cleaved by a host plasma carboxypeptidase, which converts the REDLK (amino acids 609–613) motif into REDL (amino acids 609–612) [239]. PE interacts with the low-density lipoprotein receptor-related protein 1 (LRP-1) via the Ia domain and is then internalized by endocytosis. In the early endosome, which is acidic, PE dissociates from the receptor and changes conformation, exposing the furin-cleavable motif to be cleaved by furin into two fragments of approximately 28 kDa (279 amino acids) and 37 kDa (333 amino acids) [240]. The 28 kDa fragment consists of domain Ia and parts of domain II. The 37 kDa fragment contains parts of domain II, domains Ib, and domain III and has enzymatic activity. The 37-kDa fragment migrates from the late endosome to the trans-Golgi network (TGN) and from there to the ER via the retrograde pathway. The C-terminal REDL motif interact with KDEL receptors on the TGN [241]. The mode of action of PE is illustrated in Figure 4.
Moxetumomab pasudotox [241] is a recombinant anti-CD22 immunotoxin consisting of a single chain antibody fragment of a mouse anti-CD22 monoclonal antibody (scFv) fused to a Pseudomonas endotoxin A (PE38 domain). The immunotoxin binds to CD22 antigen expressed on B cells in various hematological malignancies. Upon internalization and intracellular proteolysis, the cytotoxic fragment (PE38) is released, which then induces cell death by apoptosis [241,242]. In September 2018, Moxetumomab pasudotox (LUMOXITI™; AstraZeneca, Cambridge, UK) was approved by the U.S. Food and Drug Administration (FDA) for the treatment of relapsed or refractory hairy cell leukemia [243]. Immunotoxins containing PE are highly immunogenic in patients with normal immune systems, but less so in patients with hematologic malignancies, whose immune systems are often compromised. SS1P, a first-generation, mesothelin-targeted immunotoxin, demonstrated little activity as a single agent [244]. In patients with solid tumors, neutralizing antidrug antibodies (ADAs) directed against PE developed after only three infusions of SS1P. To delay the development of high-titer ADAs, SS1P was combined with a preconditioning regimen of lymphocyte-depleting chemotherapy [245]. LMB-100, a second-generation recombinant immunotoxin that targets the glycoprotein mesothelin on the surface of cancer cells is composed of a humanized antimesothelin antibody fragment fused to a truncated PE A [246]. The maximum tolerated dose (MTD) of LMB-100 was 140 µg/kg, administered every other day over 3 weeks [246]. Although LMB-100 was less immunogenic than SS1P, most patients developed antidrug antibodies after 2 cycles. Phase 2 clinical studies with LMB-100 plus pembrolizumab is conducted on patients with mesothelioma and lung cancer [246]. Several studies are devoted towards the developing of PE-based recombinant immunotoxins (RITs), especially against mesothelin and other proteins on solid tumors. For more information, the reader is referred to the review by Mazor and Pastan [247].

4. Antibiotics

Although antibiotics are mainly used as bactericidal or bacteriostatic agents, some display anticancer properties and are classified as anthracyclines, of which Actinomycin D (Dactinomycin), Bleomycin, Doxorubicin (adriamycin and doxil), Epirubicin (ellence), Cerubine and Daunorubicin (DaunoXome), Novantrone (mitoxantrone), Mitomycin C, Spergualin and Epothilone are the best known.

4.1. Actinomycin D

Actinomycin D, produced by Streptomyces antibioticus and Streptomyces parvulus has two pentapeptide lactone rings and a 2-aminophenoxazine-based chromophore [15]. Intercalation of the chromophore into DNA inhibits transcription and prevents the growth of tumour cells [248]. Interaction of actinomycin D with DNA is facilitated by a GpC (guanosine-cytosine) base pair. Increased binding was obtained with the formation of hydrogen bonds between L-threonine residues in the pentapeptide rings and the amino-terminal of guanosine [249]. Other amino acids of the two lactone rings, including proline, N-methylglycine and methylvaline, facilitates the binding of actinomycin D to the minor groove of DNA, thereby improving the stability of the interaction [15,249]. Despite the toxic effects of actinomycin D, such as tissue necrosis, dermatoxicity and gastrointestinal enterotoxicity, the drug has been approved for treatment of Wilms’ tumour, gestational choriocarcinoma, neuroblastoma, germ cell cancers, trophoblastic tumours, Ewing sarcoma and rhabdomyosarcoma [250,251,252,253]. Actinomycin D has also been used in combination with antitumor agents to treat high-risk malignancies [249]. Actinomycin D increased the therapeutic efficacy of antitumor agents such as RG7787, a mesothelin-targeting immunotoxin. RG7787 (also referred to as LMB-100) is a recombinant immunotoxin formed through conjugation between exotoxin A of P. aeruginosa and anti-mesothelin Fab [246]. Synergistic cytotoxicity of actinomycin D and LMB-100 towards mouse xenografts of pancreatic and stomach cancer cells were illustrated with Phase I clinical trials [254]. This combination led to apoptosis via an extrinsic pathway and resulted in noteworthy regression of the tumours. Anticancer drugs containing Actinomycin D are available in the market under the trade names Cosmegen and Lyovac [8].

4.2. Bleomycin

Bleomycin is produced by Streptomyces verticillus [255,256]. A combination of Bleomycin A2 (C55H84N17O21S3, Mw. 1415.56 Da) and Bleomycin B2 (C55H84N20O21S2, Mw. 1425.52 Da) was approved by the FDA in July 1973 [207]. The mode of action is described as a two-step process. In the first step Bleomycin chelates metal ions (primarily iron) and produces a pseudoenzyme. The second step is the enzymatic conversion of oxygen to superoxide and hydroxyl free radicals damaging DNA [257,258]. Bleomycin is used to treat Hodgkin’s lymphoma, non-Hodgkin’s lymphoma, testicular cancer, ovarian cancer, and cervical cancer [259,260], and is commercially available as Bleomycin USP and Blenoxane [8].

4.3. Doxorubicin, Epirubicin, Daunorubicin and Novantrone

Doxorubicin (adriamycin and doxil; C27H29NO11), produced by Streptomyces peucetius var. caesius [15,27,261], has amphipathic properties owing to a water-insoluble aglycone (adriamycinone, C21H18O9) and water-soluble amino sugar group (daunosamine, C6H13NO3). Its anti-cancer properties was first reported in 1969 [8] and has been approved by the FDA for treatment of malignant lymphoma, soft tissue sarcoma, and breast-, liver-, ovary-, neck-, head-, gastric- and childhood cancers [15,213]. Doxorubicin binds to DNA and RNA polymerases, which prevents DNA replication and transcription [261], intercalates with DNA and removes histones from chromatin during transcription [8]. The formation of covalent complexes between topoisomerase-II and DNA [15] leads to single- and double-strand DNA breaks and apoptosis [261]. Doxorubicin also binds to cardiolipin and mitochondrial creatine kinase [8]. Patients treated with doxorubicin displayed adverse side effects such as fatal cardiotoxicity and nonspecific cytotoxicity [15]. A nano-drug delivery system, based on liposome-encapsulation of doxorubicin (e.g., Doxil®), has been approved for treatment of ovarian, breast- and AIDS-related Kaposi’s sarcoma. Doxorubicin and mitomycin C encapsulated in polymer-lipid hybrid nanoparticles (PLNs) were 20 to 30 times more active than the drugs in separate form and killed multidrug-resistant human breast cells MDA435/LCC6 more effectively [262]. Examples of drugs containing doxorubicin are Myocet, Doxorubicin-Ebewe, Adriblastine PFS, Caelyx, Doxorubicin medac, and Doxorubicinum Accord [8].
Epirubicin (ellence) is a semisynthetic derivative of doxorubicin, with a hydroxyl group in the 4′ position of the daunosamine ring [263]. It is mainly used in the treatment of early or metastatic breast cancer, but also other tumours such as lung, bladder, gastric, ovarian and hepatocellular carcinoma, and lymphatic cancers [264]. The mode of activity of Epirubicin is similar to that of Doxorubicin and also inhibits topoisomerase II activity [265].
Daunorubicin or Daunomycin, also referred to as Cerubine, is produced by Streptomyces peucetius and differs from doxorubicin by lacking a hydroxyl group at the the 14th position [266]. The name Daunomycin is derived from the pigment aglycone daunomycinone and the amino sugar daunosamine [267]. The liposome-encapsulated form of daunorubicin, referred to as DaunoXome, is more stable in an aqueous solution and is more toxic towards certain types of solid tumours [266].
Novantrone (mitoxantrone) is chemically related to doxorubicin and acts as a potent immunosuppressive agent for treatment of multiple sclerosis [268]. Novantrone inhibits the proliferation of B and T lymphocytes as well as macrophages, kills antigen-presenting cells, and suppresses the migration of activated leukocytes [269]. Other modes of action for mitoxantrone include lowering the secretion of IFN-γ, TNF-α, and IL-2 [270].

4.4. Mitomycin C, Duramycin and Epothilones

Mitomycin C (C15H18N4O5, Mw. 334 Da), produced by Streptomyces caespitosus [271], is an aziridine [(CH₂)₂NH]-containing antibiotic that cross-links DNA and inhibits alkylation [272]. The drug is used in the treatment of bladder, colorectal and pancreatic cancers, head and neck sarcoma, and lung-, hepatic and esophageal carcinoma [273]. Spergualin (C17H37N7O4, Mw. 403.53 Da), produced by Brevibacillus laterosporus BMG162-aF2, repressed fibrosarcoma cells (M5076) and cell lines of rat hepatomas (AH66, AH66F), as well as leukemia in mice models [274,275]. Duramycin induces apoptosis and reduces the proliferation in tumour cells [276,277], a phenomenon that may be ascribed to its high affinity for phosphatidylethanolamine [278]. The cytotoxicity of duramycin was reduced by fusion to IgG [279,280]. This did not influence binding to phosphatidylethanolamine. Fusion to IgG guide host immune cells to apoptotic cells, resulting in enhanced phagocytosis. Tumour growth in mice was inhibited after treatment with duramycin-IgG [280]. Since duramycin binds to PE and the Fc region (fused) IgG antibodies, it interacts with phagocytic cells to enhance phagocytosis. Duramycin is cleared from the site effectively soon after inducing apoptosis in cancer cells, via phagocytosis, which would explain its lower cytotoxicity to surrounding normal cells. Epothilone A and B, produced by Sorangium cellulosum, is classified as a macrolide polyketide [15]. Both variants inhibit mitosis and induce the formation of α/β-tubulin polypeptide heterodimers [186]. Low dosages of epothilone inhibits cell growth without blocking mitosis [186]. Ixabepilone (IXEMPRA), a synthetic analogue of epothilone B, represses the growth of a variety cancer cells. In 2007, the FDA approved IXEMPRA for treatment of breast cancer cells resistant to treatment with taxanes such as paclitaxel and docetaxel, anthracycline and capecitabine. IXEMPRA combined with capecitabine was more effective than capecitabine in the treatment of breast cancer cells resistant to taxane and anthracycline.

5. Bacteriocins

Bacteriocins of Gram-negative bacteria are divided into four main classes: colicins, colicin-like, phage-tail-like bacteriocins, and microcins [281]. Colicins are protease-sensitive, heat-sensitive and has a high-molecular weight (30–80 kDa). Most E. coli strains have genes encoding colicins. These proteins are expressed when cells experience stress and usually leads to self-destruction due to co-expression with lysis protein [282]. Depending on the mechanism of interaction with the target cell, colicins are divided into three main groups, i.e., pore-forming, nuclease, and peptidoglycan degrading. Bacteriocins of Gram-positive bacteria are ribosomally-synthesized peptides with bacteriostatic or bacteriolytic activity and are usually below 10 kDa in size. These antimicrobial peptides are grouped into four classes. Class I are linear peptides, produced as pre-peptides, and are converted to active, mature, peptides after post-translational modification. They contain several disulphide bridges and unusual polycyclic thioether amino acids such as dehydrobutyrine, dehydroalanine and lanthionine. Class II bacteriocins do not contain lanthionine and do not undergo post-translation modification, except for the removal of the leader peptide [16]. Bacteriocins from both classes are thermostable. Class III bacteriocins are thermolabile and have a molecular mass exceeding 10 kDa. Class IV bacteriocins are circular [9,283].
The first report of a bacteriocin displaying anticancer properties was published in the late 1970s, but this was a crude extract [284]. Since then, several reports of bacteriocins with anticancer properties (also from our own group, unpublished), have been reported [285]. To the best of our knowledge only three bacteriocins have been patented for their anticancer properties, i.e., plantaricin A produced by L. plantarum C11 [8,286,287], microcin E492 produced by Klebsiella pneumoniae [288,289,290,291] and Pep27anal2, a 27-amino acid peptide produced by Streptococcus pneumoniae [291,292]. Anticancer properties have also been reported for bovicin HC5 produced by Streptococcus bovus HC5 [293,294], colicins A, E1, E3, and U isolated from E. coli [295,296,297,298,299,300], pyocins from P. aeruginosa [301], nisin from Lactococus lactis [302], and pediocins from Pediococcus and other lactic acid bacteria [303,304].
Pyocin S2, produced by P. aeruginosa M47 (PAO 3047), inhibited the growth of cancer cells XC, TSV-5, mKS-A TU-7, HeLa-S3, and AS-II, but did not affect normal cells such as mice cells BALB/3T3, rat kidney cells and human lung cells [305]. Abdi-Ali et al. [306] reported cytotoxic activity of pyocin S2 towards HepG2 cells and human immunoglobulin secreting (Im9) cells isolated from multiple myeloma [306]. Pep27anal2, an analog of signal peptide Pep27, produced by S. pneumoniae, causes caspase-dependent and cytochrome C-independent apoptosis in cancer cells, as shown with studies on OCI-AML-2 and HL60 (leukemia) cells, Jurkat cells, and MCF-7 and SNU-601 (adenocarcinoma) cell lines [292]. Bovicin (2.4 kDa), produced by Streptococcus bovis, showed cytotoxicity towards human breast cancer (MCF-7) and human liver hepatocellular carcinoma (HepG2) cell lines in a concentration-dependent manner [294]. Colicins, produced by E. coli, are larger than the average bacteriocins (>20 kDa). Colicin E3 showed cytotoxic and cytocidal effects against HeLa (human cervical cancer) cells and cleaves rRNA [297]. Colicin produced by E. coli HSC10 degraded DNA and is cytotoxic towards mammalian cells [295]. Colicin from the same strain, referred to as verotoxin 1, acted anticarcinogenic against human ovarian cell lines but protected cells in a murine metastatic fibrosarcoma model [296]. Colicins E1-E5 displayed cytotoxicity towards hamster V79 fibroblast cells [297]. Colicin A, E1, E3 and U caused cell cycle arrest in human fibroblast cell lines MRC5, MCF-7, MDA-MB-231 (mammary tumor), HOS (bone osteosarcoma), and HS913T (fibrosarcoma) [298]. In general, colicins depolarize the cytoplasmic membrane, prevents the synthesis of peptidoglycan, degrade DNA and RNA, seize cell cycles, and causes necrosis [299].
The interaction of bacteriocins with cancer cells is ill researched. Kaur and Kaur [16] hypothesized that cancer cells increase the expression of negatively charged cell-surface molecules when exposed to bacteriocins and, by doing so, promote cytotoxicity and apoptotic cell death. Azurin, an anticancer copper-containing bacteriocin of 14 kDa (Figure 4) produced by P. aeruginosa [307], provides some indication as to how bacteriocins may enter cancer cells. The peptide enters human cancer cells through receptor-mediated endocytosis [308] or, as reported with studies on breast cancer cells lines MCF-7, ZR-75-1 and T47D, via caveolin-mediated pathways [309]. The p28 domain of azurin (Figure 5) facilitates cell crossing and promotes apoptosis [310]. Sections within p28 have different adhesion and cell penetration properties, as summarized in the legend of Figure 5. Once inside a cancer cell, azurin attaches to the DNA-binding domain (DBD) of the tumor-suppressor protein p53 (Figure 6) and increases the intracellular level of the protein by inhibiting the binding of the E3 ubiquitin ligase COP1 to p53 [311]. This leads to repression of cell growth and apoptosis [312]. Amino acids 88 to 113 in the C-terminal of azurin repressed the growth of MCF-7 breast- and DU145 prostate cancer cells [313]. More recent studies have shown that azurin also interferes with non-receptor tyrosine kinase (NRTK) signalling pathways [314]. Increased intracellular levels of p53 and Bax protein (a central cell death regulator) were detected in cells exposed to azurin [314]. This led to the release of cytochrome C in the cytoplasm, and activation of caspases 9 and 7 [8]. Azurin also reduces VEGFR-2 tyrosine kinase activity, thereby preventing the formation of new blood vessels and the expression of P-cadherin, a glycoprotein maintaining the structural integrity in epithelial tissue [314]. Phase I clinical trials were performed with p28 on recurrent, difficult to treat, and progressive solid tumours in adults, and on tumours from the central nervous system (CNS) of paediatric patients. No significant immune response or adverse side effects were reported [315]. Protein p21, a short amino acid fragment of azurin overexpressed in MCF-7 cells, inhibited cyclin-dependent kinases and prevented cell growth [316].
A few bacteriocins produced by LAB are worth pointing out and are summarized here. Plantaricin A (2.98 kDa), a peptide pheromone produced by L. plantarum C11, binds to negatively charged phospholipids and glycosylated proteins in cell membranes of cancerous and normal cells, leading to the destabilization of cytoplasmic membranes [318]. The peptide also induces apoptosis and necrosis in Jurkat cells, accompanied by an increase in caspase 3 levels [286,287]. Microcin E492 (7.8 kDa) causes cell shrinkage, DNA fragmentation, release of phosphatidylserine and calcium ions, and apoptosis of cancerous cells such as HeLa, Jurkat. RJ2.25 and colorectal carcinoma cells [289]. Normal bone marrow cells, splenocytes, KG-1, human tonsil cells, and nontumor macrophage-derived cells were not affected [289].
An in vitro study showed that lacticin A164 and BH5, produced by L. lactis subsp. lactis, inhibited the growth of Helicobacter pylori [319], which is responsible for a sizeable number of gastrointestinal cancers. Probiotic preparations with these strains may thus be used to control of H. pylori infection and related intestinal cancers. Nisin, also produced by L. lactis subspecies lactis, has been used in the food industry for over 50 years [16,27]. Joo et al. [320] reported a decrease in the proliferation of head and neck squamous cell carcinoma (HNSCC) cells when exposed to nisin and illustrated decreased tumour growth in an oral cancer floor-of-mouth mouse model [320]. Athymic nude mice (NCr-nu/nu strain) were used. Gene array analyses performed on HSC-3 oral squamous cell carcinoma (SCC) cells treated with nisin revealed an increase in genes regulating calcium transport, membrane lipid functioning and apoptosis. The gene most upregulated encodes the cation transport regulator CHAC1 (ChaC Glutathione Specific Gamma-Glutamylcyclotransferase 1), which is known to be activated by oxidized phospholipids [320]. Since nisin interacts with phospholipids in cell membranes, especially phosphatidylcholine, it is possible that CHAC1 may be a “downstream” nisin target. In in vitro experiments conducted with nisin ZP on HNSCC cells and with NCr-nu/nu mice much higher levels of apoptosis were reported compared to treatment with nisin [321]. Apoptosis of HNSCC cells increased with higher concentrations of nisin ZP [321]. This coincided with a decrease in cell proliferation, clonogenic capacity, and sphere formation [321]. Nisin ZP induced apoptosis through a calpain-dependent pathway in HNSCC cells but not in human oral keratinocytes [321]. Apoptosis of human umbilical vein endothelial cells (HUVEC) induced by nisin ZP coincided with a decrease in vascular sprout formation (in vitro) and lowering of intratumoral microvessel density (in vivo) [321]. No abnormalities in organ tissue, inflammation, fibrosis or signs of necrosis were observed in mice treated [321], suggesting that nisin ZP may be a promising alternative in cancer therapy. Reports of nisin showing potential in the treatment of colorectal cancers [321,322,323] and growth inhibition of blood, breast, brain, colon, gastrointestinal, liver and skin cancer cells [324] have been published.
Preet et al. [325] studied the synergistic effect of nisin and doxorubicin on dimethylbenzanthracene-induced skin cancer in murine models. Doxorubicin-alone-treatment and nisin-alone-treatment resulted in a mean tumour burden reduction of 51.3% and 14.18%, respectively. Mice treated with a combination of nisin and doxorubicin displayed a reduction of 66.82% in the mean tumour burden. Nisin combined with 5-FU killed skin cancer cells, induced with 7,12-dimethylbenz(a)anthracene, in vivo [326]. Anticarcinogenic properties was ascribed to modulation of apoptotic, angiogenic and cell proliferative pathways.
Pediocin PA-1, produced by Pediococcus acidilactici K2a2-3, is cytotoxic towards human colon adenocarcinoma (HT29) and HeLa cells [303]. Pediocin CP-2, produced by P. acidilactici MTCC 5101 and rec-pediocin CP-2 (recombinant) cells are cytotoxic towards MCF-7, HepG2, HeLa and mouse spleen lymphoblast (Sp2/O-Ag14) cells and induces apoptosis [304].
Bacteriocins may thus play a key role in prevention of intestinal and skin cancers. Only a few of such studies have been published and most were performed in vitro. Should these results be confirmed by in vivo studies, methods will have to be developed to protect these peptides from degradation by gastrointestinal enzymes and enhance their activity. Bacteriocin-producing strains with anticarcinogenic properties may be included in probiotic supplements. We are, however, still a long way from understanding the efficacy of bacteriocins in anticancer therapy and whether these peptides should be used alone or in combination with chemotherapeutic agents.

6. Non-Ribosomal Peptides and Polyketides

Non-ribosomal peptide synthetases (NRPS) were discovered in 1968 by Gevers and co-workers when they studied gramicidin S production by Bacillus brevis [327]. By adding RNAse and puromycin (a ribosome inhibitor) to B. brevis extracts containing gramicidin S, Gevers et al. [328] observed that aminoacyl-tRNA synthetases and tRNA were not used in the production of gramicidin S. Subsequent studies have shown that non-ribosomal peptides are produced by enzyme complexes in a nucleic acid-independent manner [329]. Further research is needed to determine if non-ribosomal peptides can be used in cancer therapy [8]. Polyketides are produced non-ribosomally by type 1 polyketide synthetases (PKSs) [330]. Almost one-third of all pharmaceuticals are polyketides which can be attributed to their structural and biological diversity. Type 1 PKSs are modular enzyme assemblies and act successively to elongate the polyketide chain. Different domains can be found in PKSs such as ketoreductase, dehydratase, and enoylreductase with conserved domains such as acyl carrier protein, ketosynthase and acyltransferase. Readers are referred to Baindara and Mandal [8] for further information non-ribosomal peptides and polyketides. Noteworthy is that no non-ribosomal peptides and polyketides with anticancer properties have been reported for lactic acid bacteria.

7. Phenylpropanoids

Phenylpropanoid-derived metabolites from plants inhibits the growth of several cancer cell types [331]. However, Bacteroides thetaiotaomicron, Bacteroides eggerthii, Bacteroides ovatus, Bacteroides fragilis, Parabacteroides distasonis, Eubacterium hallii and Clostridium bartlettii ferment phenylalanine, tyrosine and tryptophan to phenylacetic acid (PAA) and 4-hydroxylphenylacetic acid (4-hydroxyPAA) in the colon [332]. An increase in cell numbers of Bacteroides fragilis was observed in patients with advanced stages III and IV) CRC [333].
The phenylpropanoid verbascoside protects the GIT from oxidative stress and represses the growth of MKN45 gastric epithelial cancer cells, but also stimulates appetite [334,335]. Further research on the stability of phenylpropanoids is required. Acteoside, a verbascoside, has anti-inflammatory properties [336] and prevents the onset of chronic diseases, including cancer [337,338], but is degraded by gut microbiota [339]. Further research is required to understand the degradation of phenylpropanoid by gut microbiota, and exactly how these compounds modulate inflammatory and microbial processes.

8. Phenylflavonoids

Xanthohumol (XN), a prenylated flavonoid found in hops, has promising anticancer properties [340]. Gut microbiota metabolites XN to 8-prenylnaringenin (8-PN), a very potent phytoestrogen with anticancer activity, as demonstrated with SK-MEL-28 and BLM human metastatic melanoma cells [341] and MCF7 breast cancer cells [342]. 8-PN also inhibited cell proliferation of the HT-115 and HT-116 colon cancer cells [342,343].

9. Natural Purine Nucleosides

Adenosine and inosine (formed by the catabolism of adenosine) bind to adenosine receptors, and initiates cAMP production and the phosphorylation of kinase-1 and -2 [344]. Inosine also enhances T cell antitumour activity in colorectal, bladder, and melanoma cancer cells [345]. Studies on bladder cancer cells indicated that inosine enhanced the function of anti-CTLA-4, causing to increase infiltration of IFN- γ + CD4+ and IFN- γ + CD8+ T-cells into the tumour, as well as reducing overall tumour weight [345]. CTLA-4, also known as CD152, is a protein receptor that functions as an immune checkpoint downregulating immune responses, a process referred to as immune checkpoint blockade (ICB). Inosine produced by Bifidobacterium pseudolongum increased the activation of a cDC-dependent TH 1 cell circuit and led to the enhancing of ICB in murine models with intestinal and epithelial tumours [345].

10. Short Chain Fatty Acids

Diet plays an important role in CRC and may promote the formation of tumours [346]. SCFAs such as butyrate, acetate, propionate and lactate are largely produced in the colon by Bifodobacterium, Lactobacillus, Lachnospiraceae, Blautia, Coprococcus, Roseburia, Faecalibacterium, Clostridium, Eubacterium, and species converting lactate and acetate, e.g., Anaerostipes spp. [347,348]. These SCFAs adhere to free fatty acid receptors (FFARs), e.g., GPR43 (FFAR2) and GPR41 (FFAR3) located on the surface of IECs [349]. Patients diagnosed with CRC had less Bifidobacterium spp. and lower levels of SCFAs [350]. Butyrate also protects intestinal barrier function by up-regulating the tight junction protein claudin-1 [351,352,353]. Other functions of butyrate include maintaining a balanced state of oxygen [301] and suppressing inflammatory responses [354,355]. The latter is achieved by downregulating histone deacetylase (also referred to as lysine deacetylase) inhibitors (HDACi) [354,355]. An increase in de-acetylated histones (due to the inhibition of HDACi), together with a decline in gene transcriptions, leads to autophagic cell death, the activation of extrinsic and/or intrinsic apoptotic pathways, an increase in production of reactive oxygen species (ROS), and a decrease in the expression of pattern recognition receptors, kinases, transcription regulators, cytokines, chemokines, and growth factors [356,357]. Molecules released from dying cells are recognised by nucleotide-binding oligomerization domain (NOD)-like receptors (NLRs) and form specific protein-protein interactions. These interactions, also prevalent in lymphocytes, macrophages and dendritic cells, play a key role in the regulation of cytokines, chemokines and the expression of genes coding for the production of antimicrobial compounds, collectively referred to as the innate immune response [358,359,360]. Downregulation of NLRs prevents the formation of multi-protein inflammasomes, the signaling of caspase-independent nuclear factor kappa B (NF-κB) and mitogen-activated protein kinase (MAPK). These cascades of events counteract autoimmune and inflammatory disorders and as recently shown, represses the growth of cancerous cells [361].
Butyrate and propionate are more effective than acetate in inhibiting the growth of HT29 cells [362]. Butyrate significantly increased apoptosis in cancer cells [363] and activates ornithine decarboxylase, resulting in the inhibition of polyamine metabolism and an increase in alkaline phosphatase activity [364]. Butyrate also activates GPR41/GPR43 receptor signaling pathways [364], thus preventing the proliferation of cancer cells [365]. GPR43 is predominantly present in the large intestine of healthy cells, but less so in colon cancer cells [366]. Manipulation of HCT8 human colonic adenocarcinoma cells to express the GPR43 receptor led to an increase in apoptosis and G0/G1 cell cycle arrest [366]. Concluded from these studies, the GPR43 receptor serves as a tumour suppressor. SCFAs were ineffective against HCT-116 colon cancer cells with a deletion in the p21 gene, suggesting that p21, a cell cycle inhibitor and anti-proliferative effector, is important in repressing tumour growth. Another study has shown that p21 is regulated by the p53 transcription factor [367]. Diets rich in SCFAs suppresses T cell-mediated autoimmune responses, most probably through regulation of cytokine expression [364]. The ability of butyrate to de-repress epigenetically silenced genes in cancer cells, such as cell cycle inhibitor p21 and the pro-apoptotic protein Bcl-2, has important implications for cancer prevention and therapy. Lightfoot et al. [368] tested possible epigenetic modifications induced by LTA-deficient L. acidophilus and found that oral NCK 2025 enhances the expression of tumor suppressor genes [368,369]. This indicates that differential epigenetic regulation of CRC-related genes by NCK2025 represents a potential therapy against CRC.
Overall, SCFAs are promising specifically in the context of colon cancer. Future studies should evaluate the effect of SCFAs on other cancer types, e.g., pancreatic and gastric cancer. Studies should also explore the impact of SCFAs on the efficacy and safety of standard chemotherapy and the prognosis of cancer.

11. Enzymes

Four enzymes with anticancer properties have been reported, i.e., arginine deiminase (ADI), produced by Mycoplasma hominis and Mycoplasma arginine, asparaginase (ASNase), produced by E. coli and Erwinia chrysanthemi, glutaminase [370] and methionase [371]. ADI converts arginine to citrulline and ammonia in vivo [372]. Pegilation of ADI (ADI-PEG20) significantly increased the half-life of ADI in serum and decreased its antigenicity [373], rendering ADI much more effective against cancerous cells. The mode of activity of ADI-PEG20 is ascribed to loss of argininosuccinate synthetase (ASS) activity and, thus, the inability to synthesize arginine from citrulline [374]. Due to this, growth of hepatocellular carcinoma cells (HCC), auxotrophic to arginine, were repressed, with evidence of apoptosis [375]. Treatment of metastatic hepatocellular carcinoma cells with ADI-PEG20 entered phase II clinical trials [376]. Autophagy was induced in prostate cancer cells (CWR22Rv1) treated with ADI-PEG20 [374]. ASNase degrades asparagine, which results in a severe reduction in protein synthesis and thus growth inhibition of cancer cells [377,378]. Studies performed with human cells lines (pediatric medulloblastoma, DAOY, and glioblastomas GBM-ES and U87) indicated that inhibition of cell growth by ASNase is dose-dependent [379]. ASNase has been used in the treatment of acute lymphoblastic leukemia (ALL), acute myeloid leukemia, ovarian carcinoma, myelosarcoma, Hodgkin lymphomas, and extranodal NK/T cell lymphoma [380,381,382,383]. Glutaminase, specifically glutaminase 1 (GLS1) expressed in mitochondria, converts glutamine to glutamate [384], but also stimulates the growth of tumour cells [385] and is involved in autophagy [386], signal transduction [387], and radioresistance [388]. Recent evidence emerged showing that GLS1 might be involved in tumorigenesis and progression of human cancers [370]. GLS1 is overexpressed in metastatic lymph nodes and colorectal cancer cells [370]. Methionase, also known as L-Methionine-γ-lyase (EC 4.4.1.11; MGL), methioninase, L-methionine-γ-demethiolase, and L-methionine methanethiol-lyase (deaminating), is produced by Pseudomonas putida, Pseudomonas ovalis, Micrococcus luteus, Arthrobacter spp., Corynebacterium glutamicum, Staphylococcus equorum, Citrobacter spp., Clostridium sporogenes, Trichomonas vaginalis and Entamoeba histolytica, but has also been isolated from protozoans, fungal, archaeon, and plants [371]. Melignant tumours are highly dependent on methionine. Depletion of methionine through methionase-based therapy may be an important strategy to control the growth of cancer cells. One approach experimented with was linking L-methionase to human annexin-V to generate a fusion protein. The fusion protein catalyzed the conversion of nontoxic prodrug selenomethionine into toxic methylselenol and restricted the growth of tumour cells by depriving the cells from access to methionine [389,390,391]. The advantage of using the fusion protein is that it does not require to be delivered directly to the tumour cells but only to the bloodstream.

12. Conclusions and Future Directions

Cancer is a major health concern and treatment remains a challenge due to cells developing resistance. Despite numerous efforts to use viable bacteria in the treatment of cancer, the idea is still viewed with scepticism, as many strains experimented with are considered obsolete or opportunistic pathogens. To date, bacterial anticancer compounds studied were either toxins, antibiotics, bacteriocins, non-ribosomal peptides, polyketides, phenylpropanoids, phenylflavonoids, purine nucleosides, short chain fatty acids (SCFAs) or enzymes. Other natural antitumor compounds discovered are spliceostatins, such as spliceostatin B purified from cell-free extracts of Pseudomonas sp. 2663 and pladienolide B, a macrolide produced by Streptomyces platensis Mer-11107. More research on spliceostatins is required. It is interesting to note that many of the bacteria that produce anticancer compounds are either obligate or facultative anaerobes. Many anticancer compounds naturally produced by bacteria are specific in their mode of action, and some are selective in attacking only cancerous cells. Lactic acid bacteria form a major part of the gut microbiome, yet the role they play in cancer treatment is ill researched. The combined use of natural anticancer compounds with conventional anticancer therapy warrants further research. Progress in metagenomics, proteomics, heterologous gene expressions and nanotechnology, combined with the use of artificial intelligence software such as AlphaFold 2 (https://alphafold.ebi.ac.uk/) and Chemistry42 (https://arxiv.org/abs/2101.09050, accessed on 10 May 2022), may lead to the discovery and design of novel anticancer molecules.

Author Contributions

Conceptualization, L.M.T.D.; writing—original draft preparation, both authors; supervision, L.M.T.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. Am. Cancer Soc. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  2. World Health Organization. World Health Assembly, 70 Cancer Prevention and Control in the Context of an Integrated Approach. Available online: https://apps.who.int/iris/handle/10665/275676 (accessed on 21 June 2022).
  3. National Institute for Communicable Diseases National Cancer Registry 2019, Full Report. Available online: https://www.nicd.ac.za/centres/national-cancer-registry/cancer-statistics (accessed on 21 June 2022).
  4. American Cancer Society, Cancer Facts and Figures. 2018. Available online: https://doi.org/10.1182/blood-2015-12-687814 (accessed on 20 June 2022).
  5. Soleimanpour, S.; Hasanian, S.M.; Avan, A.; Yaghoubi, A.; Khazaei, M. Bacteriotherapy in gastrointestinal cancer. Life Sci. 2020, 254, 117754. [Google Scholar] [CrossRef] [PubMed]
  6. Van Daele, D.J.; Finnegan, E.M.; Rodnitzky, R.L.; Zhen, W.; McCulloch, T.M.; Hoffman, H.T. Head and neck muscle spasm after radiotherapy. Otolaryngol. Head Neck Surg. 2002, 128, 956. [Google Scholar] [CrossRef] [PubMed]
  7. Sharma, P.C.; Sharma, D.; Sharma, A.; Bhagat, M.; Ola, M.; Thakur, V.K.; Bhardwaj, J.K.; Goyal, R.K. Recent advances in microbial toxin-related strategies to combat cancer. Semin. Cancer Biol. 2021. [Google Scholar] [CrossRef]
  8. Baindara, P.; Mandal, S.M. Bacteria and bacterial anticancer agents as a promising alternative for cancer therapeutics. Biochimie 2020, 177, 164–189. [Google Scholar] [CrossRef] [PubMed]
  9. Diwan, D.; Cheng, L.; Usmani, Z.; Sharma, M.; Holden, N.; Willoughby, N.; Sangwan, N.; Baadhe, R.R.; Liu, C.; Gupta, V.K. Microbial cancer therapeutics: A promising approach. Semin. Cancer Biol. 2021. [Google Scholar] [CrossRef] [PubMed]
  10. Dana, M.H.; Cohen, M.; Morbize, J.; Marandas, P.; Janot, F.; Bourhis, J. Botulinum toxin for radiation-induced facial pain and trismus. Otolaryngol. Head Neck Surg. 2008, 138, 459–463. [Google Scholar] [CrossRef]
  11. Namazi, H.A. Novel use of botulinum toxin to ameliorate bone cancer pain. Ann. Surg. Oncol. 2007, 15, 1259–1260. [Google Scholar] [CrossRef]
  12. Rostami, R.; Mittal, S.; Radmand, R.; Jabbari, B. Incobotulinum Toxin-A improves post-surgical and post-radiation pain in cancer patients. Toxins 2016, 8, 22. [Google Scholar] [CrossRef]
  13. Gao, Z.; McClane, B.A. Use of Clostridium perfringens enterotoxin and the enterotoxin receptor-binding domain (C-CPE) for cancer treatment: Opportunities and challenges. J. Toxicol. 2012, 2012, 981626. [Google Scholar] [CrossRef] [Green Version]
  14. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef] [PubMed]
  15. Mohan, C.D.; Rangappa, S.; Nayak, S.C.; Jadimurthy, R.; Wang, L.; Sethi, G.; Garg, M.; Rangappa, K.S. Bacteria as a treasure house of secondary metabolites with anticancer potential. Semin. Cancer Biol. 2021. [Google Scholar] [CrossRef] [PubMed]
  16. Kaur, S.; Kaur, S. Bacteriocins as potential anticancer agents. Front. Pharmacol. 2015, 6, 272. [Google Scholar] [CrossRef]
  17. Utsugi, T.; Schroit, A.J.; Connor, J.; Bucana, C.D.; Fidler, I.J. Elevated expression of phosphatidylserine in the outer membrane leaflet of human tumor cells and recognition by activated human blood monocytes. Cancer Res. 1991, 51, 3062–3066. [Google Scholar] [PubMed]
  18. Schweizer, F. Cationic amphiphilic peptides with cancer-selective toxicity. Eur. J. Pharmacol. 2009, 625, 190–194. [Google Scholar] [CrossRef]
  19. Riedl, S.; Rinner, B.; Asslaber, M.; Schaider, H.; Walzer, S.; Novak, A.; Lohner, K.; Zweytick, D. In search of a novel target—phosphatidylserine exposed by non-apoptotic tumor cells and metastases of malignancies with poor treatment efficacy. Biochim. Biophys. Acta 2011, 1808, 2638–2645. [Google Scholar] [CrossRef]
  20. Sok, M.; Šentjurc, M.; Schara, M. Membrane fluidity characteristics of human lung cancer. Cancer Lett. 1999, 139, 215–220. [Google Scholar] [CrossRef]
  21. Hoskin, D.W.; Ramamoorthy, A. Studies on anticancer activities of antimicrobial peptides. Biochim. Biophys. Acta 2008, 1778, 357–375. [Google Scholar] [CrossRef]
  22. Dobrzynska, I.; Szachowicz-Petelska, B.; Sulkowski, S.; Figaszewski, Z. Changes in electric charge and phospholipids composition in human colorectal cancer cells. Mol. Cell. Biochem. 2005, 276, 113–119. [Google Scholar] [CrossRef]
  23. Kramer, G.M.; Masner, M.; Ferreira, F.A.; Hoffman, R.M. Bacterial therapy of cancer: Promises, limitations, and insights for future directions. Front. Microbiol. 2018, 9, 16. [Google Scholar] [CrossRef] [Green Version]
  24. Łukasiewicz, K.; Fol, M. Microorganisms in the treatment of cancer: Advantages and limitations. J. Immunol. Res. 2018, 2018, 2397808. [Google Scholar] [CrossRef] [PubMed]
  25. Guo, C.; Manjili, M.H.; Subjeck, J.R.; Sarkar, D.; Fisher, P.B.; Wang, X.-Y. Therapeutic cancer vaccines: Past, present and Future. Adv. Cancer Res. 2013, 119, 421–475. [Google Scholar] [CrossRef] [PubMed]
  26. McCarthy, E.F. The toxins of William B. Coley and the treatment of bone and soft-tissue sarcomas. Iowa Orthop. J. 2006, 26, 154–158. [Google Scholar]
  27. Karpiński, T.; Adamczak, A. Anticancer activity of bacterial proteins and peptides. Pharmaceutics 2018, 10, 54. [Google Scholar] [CrossRef] [PubMed]
  28. Forbes, N.S.; Coffin, R.S.; Deng, L.; Evgin, L.; Fiering, S.; Giacalone, M.; Gravekamp, C.; Gulley, J.L.; Gunn, H.; Hoffman, R.M.; et al. White paper on microbial anti-cancer therapy and prevention. J. Immunother. Cancer 2018, 6, 78. [Google Scholar] [CrossRef]
  29. Yoshimura, K.; Jain, A.; Allen, H.E.; Laird, L.S.; Chia, C.Y.; Ravi, S.; Brockstedt, D.G.; Giedlin, M.A.; Bahjat, K.S.; Leong, M.L.; et al. Selective targeting of antitumor immune responses with engineered live-attenuated Listeria monocytogenes. Cancer Res. 2006, 66, 1096–1104. [Google Scholar] [CrossRef]
  30. Yazawa, K.; Fujimori, M.; Nakamura, T.; Sasaki, T.; Amano, J.; Kano, Y.; Taniguchi, S. Bifidobacterium longum as a delivery system for gene therapy of chemically induced rat mammary tumors. Breast Cancer Res. Treat. 2001, 66, 165–170. [Google Scholar] [CrossRef]
  31. Min, J.-J.; Kim, H.-J.; Park, J.H.; Moon, S.; Jeong, J.H.; Hong, Y.-J.; Cho, K.-O.; Nam, J.H.; Kim, N.; Park, Y.-K.; et al. Noninvasive real-time imaging of tumors and metastases using tumor-targeting light-emitting Escherichia coli. Mol. Imaging Biol. 2008, 10, 54–61. [Google Scholar] [CrossRef]
  32. Nguyen, V.H.; Kim, H.-S.; Ha, J.-M.; Hong, Y.; Choy, H.E.; Min, J.-J. Genetically engineered Salmonella typhimurium as an imageable therapeutic probe for cancer. Cancer Res. 2010, 70, 18–23. [Google Scholar] [CrossRef]
  33. Nuyts, S.; van Mellaert, L.; Theys, J.; Landuyt, W.; Lambin, P.; Anné, J. Clostridium spores for tumor-specific drug delivery. Anticancer Drugs 2002, 13, 115–125. [Google Scholar] [CrossRef]
  34. Khan, S.A.; Everest, P.; Servos, S.; Foxwell, N.; Zähringer, U.; Brade, H.; Rietschel, E.T.; Dougan, G.; Charles, I.G.; Maskell, D.J. A lethal role for lipid A in Salmonella infections. Mol. Microbiol. 1998, 29, 571–579. [Google Scholar] [CrossRef] [PubMed]
  35. Pawelek, J.M.; Low, K.B.; Bermudes, D. Tumor-targeted Salmonella as a novel anticancer vector. Cancer Res. 1997, 57, 4537–4544. [Google Scholar] [PubMed]
  36. King, I.; Bermudes, D.; Lin, S.; Belcourt, M.; Pike, J.; Troy, K.; Le, T.; Ittensohn, M.; Mao, J.; Lang, W.; et al. Tumor-targeted Salmonella expressing cytosine deaminase as an anticancer agent. Hum. Gene Ther. 2002, 13, 1225–1233. [Google Scholar] [CrossRef] [PubMed]
  37. American Society of Health-System Pharmacists: Fluorouracil. 2016. Available online: https://www.drugs.com/monograph/fluorouracil-systemic.html (accessed on 14 July 2022).
  38. Mei, S.; Theys, J.; Landuyt, W.; Anne, J.; Lambin, P. Optimization of tumor-targeted gene delivery by engineered attenuated Salmonella typhimurium. Anticancer Res. 2002, 22, 3261–3266. [Google Scholar] [PubMed]
  39. Royo, J.L.; Becker, P.D.; Camacho, E.M.; Cebolla, A.; Link, C.; Santero, E.; Guzmán, C.A. In vivo gene regulation in Salmonella spp. by a salicylate-dependent control circuit. Nat. Methods 2007, 4, 937–942. [Google Scholar] [CrossRef]
  40. Nemunaitis, J.; Cunningham, C.; Senzer, N.; Kuhn, J.; Cramm, J.; Litz, C.; Cavagnolo, R.; Cahill, A.; Clairmont, C.; Sznol, M. Pilot trial of genetically modified, attenuated Salmonella expressing the E. coli cytosine deaminase gene in refractory cancer patients. Cancer Gene Ther. 2003, 10, 737–744. [Google Scholar] [CrossRef]
  41. Leschner, S.; Deyneko, I.V.; Lienenklaus, S.; Wolf, K.; Bloecker, H.; Bumann, D.; Loessner, H.; Weiss, S. Identification of tumor-specific Salmonella typhimurium promoters and their regulatory logic. Nucleic Acids Res. 2012, 40, 2984–2994. [Google Scholar] [CrossRef]
  42. Mengesha, A.; Dubois, L.; Lambin, P.; Landuyt, W.; Chiu, R.K.; Wouters, B.G.; Theys, J. Development of a flexible and potent hypoxia-inducible promoter for tumor-targeted gene expression in attenuated Salmonella. Cancer Biol. Ther. 2006, 5, 1120–1128. [Google Scholar] [CrossRef]
  43. Arrach, N.; Zhao, M.; Porwollik, S.; Hoffman, R.M.; McClelland, M. Salmonella promoters preferentially activated inside tumors. Cancer Res. 2008, 68, 4827–4832. [Google Scholar] [CrossRef]
  44. Swofford, C.A.; van Dessel, N.; Forbes, N.S. Quorum-sensing Salmonella selectively trigger protein expression within tumors. Proc. Natl. Acad. Sci. USA 2015, 112, 3457–3462. [Google Scholar] [CrossRef]
  45. Mi, Z.; Feng, Z.-C.; Li, C.; Yang, X.; Ma, M.-T.; Rong, P.-F. Salmonella-mediated cancer therapy: An innovative therapeutic strategy. J. Cancer 2019, 10, 4765–4776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Felgner, S.; Kocijancic, D.; Frahm, M.; Heise, U.; Rohde, M.; Zimmermann, K.; Falk, C.; Erhardt, M.; Weiss, S. Engineered Salmonella enterica serovar Typhimurium overcomes limitations of anti-bacterial immunity in bacteria-mediated tumor therapy. OncoImmunology 2018, 7, e1382791. [Google Scholar] [CrossRef] [PubMed]
  47. Toso, J.F.; Gill, V.J.; Hwu, P.; Marincola, F.M.; Restifo, N.P.; Schwartzentruber, D.J.; Sherry, R.M.; Topalian, S.L.; Yang, J.C.; Stock, F.; et al. Phase I study of the intravenous administration of attenuated Salmonella typhimurium to patients with metastatic melanoma. Am. J. Clin. Oncol. 2002, 20, 142–152. [Google Scholar] [CrossRef]
  48. Ganai, S.; Arenas, R.B.; Forbes, N.S. Tumour-targeted delivery of TRAIL using Salmonella typhimurium enhances breast cancer survival in mice. Br. J. Cancer 2009, 101, 1683–1691. [Google Scholar] [CrossRef]
  49. Wen, M.; Zheng, J.H.; Choi, J.M.; Pei, J.; Li, C.-H.; Li, S.-Y.; Kim, I.-Y.; Lim, S.-H.; Jung, T.-Y.; Moon, K.-S.; et al. Genetically-engineered Salmonella typhimurium expressing TIMP-2 as a therapeutic intervention in an orthotopic glioma mouse model. Cancer Lett. 2018, 433, 140–146. [Google Scholar] [CrossRef]
  50. Jeong, J.-H.; Kim, K.; Lim, D.; Jeong, K.; Hong, Y.; Nguyen, V.H.; Kim, T.-H.; Ryu, S.; Lim, J.-A.; Kim, J.I.; et al. Anti-tumoral effect of the mitochondrial target domain of noxa delivered by an engineered Salmonella typhimurium. PLoS ONE 2014, 9, e80050. [Google Scholar] [CrossRef]
  51. Lee, C.-H.; Wu, C.-L.; Shiau, A.-L. Systemic administration of attenuated Salmonella choleraesuis carrying thrombospondin-1 gene leads to tumor-specific transgene expression, delayed tumor growth and prolonged survival in the murine melanoma model. Cancer Gene Ther. 2005, 12, 175–184. [Google Scholar] [CrossRef]
  52. Feltis, B.A.; Miller, J.S.; Sahar, D.A.; Kim, A.S.; Saltzman, D.A.; Leonard, A.S.; Wells, C.L.; Sielaff, T.D. Liver and circulating NK1.1(+)CD3(-) cells are increased in infection with attenuated Salmonella typhimurium and are associated with reduced tumor in murine liver cancer. J. Surg. Res. 2002, 107, 101–107. [Google Scholar] [CrossRef]
  53. Saltzman, D.A.; Heise, C.P.; Hasz, D.E.; Zebede, M.; Kelly, S.M.; Curtiss, R., 3rd; Leonard, A.S.; Anderson, P.M. Attenuated Salmonella typhimurium containing interleukin-2 decreases MC-38 hepatic metastases: A novel anti-tumor agent. Cancer Biother. Radiopharm. 1996, 11, 145–153. [Google Scholar] [CrossRef]
  54. Saltzman, D.A.; Katsanis, E.; Heise, C.P.; Hasz, D.E.; Vigdorovich, V.; Kelly, S.M.; Curtiss, R., 3rd; Leonard, A.S.; Anderson, P.M. Antitumor mechanisms of attenuated Salmonella typhimurium containing the gene for human interleukin-2: A novel antitumor agent? J. Pediatr. Surg. 1997, 32, 301–306. [Google Scholar] [CrossRef]
  55. Smith, K.A. Lowest dose interleukin-2 immunotherapy. Blood 1993, 81, 1414–1423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Sorenson, B.S.; Banton, K.L.; Frykman, N.L.; Leonard, A.S.; Saltzman, D.A. Attenuated Salmonella typhimurium with IL-2 gene reduces pulmonary metastases in murine osteosarcoma. Clin. Orthop. Relat. Res. 2008, 466, 1285–1291. [Google Scholar] [CrossRef] [PubMed]
  57. Barnett, S.J.; Soto, L.J., 3rd; Sorenson, B.S.; Nelson, B.W.; Leonard, A.S.; Saltzman, D.A. Attenuated Salmonella typhimurium invades and decreases tumor burden in neuroblastoma. J. Pediatr. Surg. 2005, 40, 993–997. [Google Scholar] [CrossRef] [PubMed]
  58. Parker, R.C.; Plummer, H.C.; Siebenmann, C.O.; Chapman, M.G. Effect of histolyticus infection and toxin on transplantable mouse tumors. Exp. Biol. Med. 1947, 66, 461–467. [Google Scholar] [CrossRef]
  59. Malmgren, R.A.; Flanigan, C.C. Localization of the vegetative form of Clostridium tetani in mouse tumors following intravenous spore administration. Cancer Res. 1955, 15, 473–478. [Google Scholar]
  60. Dang, L.H.; Bettegowda, C.; Huso, D.L.; Kinzler, K.W.; Vogelstein, B. Combination bacteriolytic therapy for the treatment of experimental tumors. Proc. Natl. Acad. Sci. USA 2001, 98, 15155–15160. [Google Scholar] [CrossRef]
  61. Cheong, I.; Huang, X.; Bettegowda, C.; Diaz, L.A.; Kinzler, K.W.; Zhou, S.; Vogelstein, B. A bacterial protein enhances the release and efficacy of liposomal cancer drugs. Science 2006, 314, 1308–1311. [Google Scholar] [CrossRef]
  62. Roberts, N.J.; Zhang, L.; Janku, F.; Collins, A.; Bai, R.-Y.; Staedtke, V.; Rusk, A.W.; Tung, D.; Miller, M.; Roix, J.; et al. Intratumoral injection of Clostridium novyi-NT spores induces antitumor responses. Sci. Translat. Med. 2014, 6, 249ra111. [Google Scholar] [CrossRef]
  63. Janku, F.; Zhang, H.H.; Pezeshki, A.M.; Goel, S.; Murthy, R.; Wang-Gillam, A.; Shepard, D.R.; Helgason, T.; Masters, T.; Hong, D.S.; et al. Intratumoral injection of Clostridium novyi-NT spores in patients with treatment-refractory advanced solid tumors. Clin. Cancer Res. 2020, 27, 96–106. [Google Scholar] [CrossRef]
  64. Bettegowda, C.; Dang, L.H.; Abrams, R.; Huso, D.L.; Dillehay, L.; Cheong, I.; Agrawal, N.; Borzillary, S.; McCaffery, J.M.; Watson, E.L.; et al. Overcoming the hypoxic barrier to radiation therapy with anaerobic bacteria. Proc. Natl. Acad. Sci. USA 2003, 100, 15083–15088. [Google Scholar] [CrossRef] [PubMed]
  65. Nuyts, S.; Van Mellaert, L.; Theys, J.; Landuyt, W.; Lambin, P.; Anné, J. The use of radiation-induced bacterial promoters in anaerobic conditions: A means to control gene expression in clostridium-mediated therapy for cancer. Radiat Res. 2001, 155, 716–723. [Google Scholar] [CrossRef]
  66. Jiang, S.-N.; Phan, T.X.; Nam, T.-K.; Nguyen, V.H.; Kim, H.-S.; Bom, H.-S.; Choy, H.E.; Hong, Y.; Min, J.-J. Inhibition of tumor growth and metastasis by a combination of Escherichia coli–mediated cytolytic therapy and radiotherapy. Mol Ther. 2010, 18, 635–642. [Google Scholar] [CrossRef] [PubMed]
  67. Lemmon, M.; van Zijl, P.; Fox, M.; Mauchline, M.; Giaccia, A.; Minton, N.; Brown, J. Anaerobic bacteria as a gene delivery system that is controlled by the tumor microenvironment. Gene Ther. 1997, 4, 791–796. [Google Scholar] [CrossRef]
  68. Liu, S.-C.; Minton, N.; Giaccia, A.; Brown, J. Anticancer efficacy of systemically delivered anaerobic bacteria as gene therapy vectors targeting tumor hypoxia/necrosis. Gene Ther. 2002, 9, 291–296. [Google Scholar] [CrossRef] [PubMed]
  69. Nuyts, S.; van Mellaert, L.; Theys, J.; Landuyt, W.; Bosmans, E.; Anné, J.; Lambin, P. Radio-responsive RecA promoter significantly increases TNFα production in recombinant Clostridia after 2 Gy irradiation. Gene Ther. 2001, 8, 1197–1201. [Google Scholar] [CrossRef]
  70. Li, X.; Fu, G.-F.; Fan, Y.-R.; Liu, W.-H.; Liu, X.-J.; Wang, J.-J.; Xu, G.-X. Bifidobacterium adolescentis as a delivery system of endostatin for cancer gene therapy: Selective inhibitor of angiogenesis and hypoxic tumor growth. Cancer Gene Ther. 2003, 10, 105–111. [Google Scholar] [CrossRef] [PubMed]
  71. Nakamura, T.; Sasaki, T.; Fujimori, M.; Yazawa, K.; Kano, Y.; Amano, J.; Taniguchi, S. Cloned cytosine deaminase gene expression of Bifidobacterium longum and application to enzyme/pro-drug therapy of hypoxic solid tumors. Biosci. Biotechnol. Biochem. 2002, 66, 2362–2366. [Google Scholar] [CrossRef]
  72. Hidaka, A.; Hamaji, Y.; Sasaki, T.; Taniguchi, S.; Fujimori, M. Exogeneous cytosine deaminase gene expression in Bifidobacterium breve I-53-8w for tumor-targeting enzyme/prodrug therapy. Biosci. Biotechnol. Biochem. 2007, 71, 2921–2926. [Google Scholar] [CrossRef]
  73. Sasaki, T.; Fujimori, M.; Hamaji, Y.; Hama, Y.; Ito, K.; Amano, J.; Taniguchi, S. Genetically engineered Bifidobacterium longum for tumor-targeting enzyme-prodrug therapy of autochthonous mammary tumors in rats. Cancer Sci. 2006, 97, 649–657. [Google Scholar] [CrossRef]
  74. Yi, C.; Huang, Y.; Guo, Z.; Wang, S. Antitumor effect of cytosine deaminase/5-fluorocytosine suicide gene therapy system mediated by Bifidobacterium infantis on melanoma. Acta Pharmacol. Sin. 2005, 26, 629–634. [Google Scholar] [CrossRef]
  75. Pan, Z.K.; Ikonomidis, G.; Lazenby, A.; Pardoll, D.; Paterson, Y. A recombinant Listeria monocytogenes vaccine expressing a model tumour antigen protects mice against lethal tumour cell challenge and causes regression of established tumours. Nat. Med. 1995, 1, 471–477. [Google Scholar] [CrossRef] [PubMed]
  76. Guirnalda, P.; Wood, L.; Paterson, Y. Listeria monocytogenes and its products as agents for cancer immunotherapy. Adv. Immunol. 2012, 113, 81–118. [Google Scholar] [CrossRef] [PubMed]
  77. Cory, L.; Chu, C. ADXS-HPV: A therapeutic Listeria vaccination targeting cervical cancers expressing the HPV E7 antigen. Hum Vaccin Immunother. 2014, 10, 3190–3195. [Google Scholar] [CrossRef] [PubMed]
  78. Wallecha, A.; French, C.; Petit, R.; Singh, R.; Amin, A.; Rothman, J. Lm-LLO-based immunotherapies and HPV-associated disease. J. Oncol. 2012, 2012, 542851. [Google Scholar] [CrossRef] [PubMed]
  79. Maciag, P.C.; Radulovic, S.; Rothman, J. The first clinical use of a live-attenuated Listeria monocytogenes vaccine: A phase I safety study of lm-LLO-E7 in patients with advanced carcinoma of the cervix. Vaccine 2009, 27, 3975–3983. [Google Scholar] [CrossRef]
  80. ClinicalTrials.gov. Safety Study of Recombinant Listeria Monocytogenes(lm)Based Vaccine Virus Vaccine to Treat Oropharyngeal Cancer (REALISTIC:). ClinicalTrials.gov Identifier: NCT01598792. Available online: https://clinicaltrials.gov/ct2/show/NCT01598792 (accessed on 14 June 2022).
  81. ClinicalTrials.gov. A Phase I/II Evaluation of ADXS11-001, Mitomycin, 5-Fluorouracil (5-FU) and IMRT for Anal Cancer (276). ClinicalTrials.gov Identifier: NCT01671488. Available online: https://clinicaltrials.gov/ct2/show/NCT01671488 (accessed on 12 June 2022).
  82. Kurtz, C.B.; Millet, Y.A.; Puurunen, M.K.; Perreault, M.; Charbonneau, M.R.; Isabella, V.M.; Kotula, J.W.; Antipov, E.; Dagon, Y.; Denney, W.S.; et al. An engineered E. coli Nissle improves hyperammonemia and survival in mice and shows dose-dependent exposure in healthy humans. Sci. Translat. Med. 2019, 11, eaau7975. [Google Scholar] [CrossRef]
  83. Puurunen, M.K.; Vockley, J.; Searle, S.L.; Sacharow, S.J.; Phillips, J.A.; Denney, W.S.; Goodlett, B.D.; Wagner, D.A.; Blankstein, L.; Castillo, M.J.; et al. Safety and pharmacodynamics of an engineered E. coli Nissle for the treatment of phenylketonuria: A first-in-human phase 1/2a study. Nat. Metab. 2021, 3, 1125–1132. [Google Scholar] [CrossRef]
  84. Longhi, G.; van Sinderen, D.; Ventura, M.; Turroni, F. Microbiota and cancer: The emerging beneficial role of bifidobacteria in cancer immunotherapy. Front. Microbiol. 2020, 11, 575072. [Google Scholar] [CrossRef]
  85. Kostic, A.D.; Gevers, D.; Pedamallu, C.S.; Michaud, M.; Duke, F.; Earl, A.M.; Ojesina, A.I.; Jung, J.; Bass, A.J.; Tabernero, J.; et al. Genomic analysis identifies association of Fusobacterium with colorectal carcinoma. Genome Res. 2012, 22, 292–298. [Google Scholar] [CrossRef]
  86. Boleij, A.; Hechenbleikner, E.M.; Goodwin, A.C.; Badani, R.; Stein, E.M.; Lazarev, M.G.; Ellis, B.; Carroll, K.C.; Albesiano, E.; Wick, E.C.; et al. The Bacteroides fragilis toxin gene is prevalent in the colon mucosa of colorectal cancer patients. Clin. Infect. Dis. 2015, 60, 208–215. [Google Scholar] [CrossRef]
  87. Bernstein, C.; Holubec, H.; Bhattacharyya, A.K.; Nguyen, H.; Payne, C.M.; Zaitlin, B.; Bernstein, H. Carcinogenicity of deoxycholate, a secondary bile acid. Arch. Toxicol. 2011, 85, 863–871. [Google Scholar] [CrossRef] [PubMed]
  88. Ajouz, H.; Mukherji, D.; Shamseddine, A. Secondary bile acids: An underrecognized cause of colon cancer. World J. Surg. Oncol. 2014, 12, 164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Zeng, H.; Umar, S.; Rust, B.; Lazarova, D.; Bordonaro, M. Secondary bile acids and short chain fatty acids in the colon: A focus on colonic microbiome, cell proliferation, inflammation, and cancer. Int. J. Mol. Sci. 2019, 20, 1214. [Google Scholar] [CrossRef]
  90. Wong, J.M.; de Souza, R.; Kendall, C.W.C.; Emam, A.; Jenkins, D.J.A. Colonic health: Fermentation and short chain fatty acids. J. Clin. Gastroenterol. 2006, 40, 235–243. [Google Scholar] [CrossRef]
  91. Ridlon, J.M.; Kang, D.J.; Hylemon, P.B.; Bajaj, J.S. Bile acids and the gut microbiome. Curr. Opin. Gastroenterol. 2014, 30, 332–338. [Google Scholar] [CrossRef] [PubMed]
  92. Russell, W.R.; Gratz, S.W.; Duncan, S.H.; Holtrop, G.; Ince, J.; Scobbie, L.; Duncan, G.; Johnstone, A.M.; Lobley, G.E.; Wallace, R.J.; et al. High-protein, reduced-carbohydrate weight-loss diets promote metabolite profiles likely to be detrimental to colonic health. Am. J. Clin. Nutr. 2011, 93, 1062–1072. [Google Scholar] [CrossRef]
  93. Windey, K.; de Preter, V.; Verbeke, K. Relevance of protein fermentation to gut health. Mol. Nutr. Food Res. 2012, 56, 184–196. [Google Scholar] [CrossRef]
  94. Goldberg, A.A.; Titorenko, V.I.; Beach, A.; Sanderson, J.T. Bile acids induce apoptosis selectively in androgen-dependent and -independent prostate cancer cells. PeerJ 2013, 1, e122. [Google Scholar] [CrossRef]
  95. Yadav, R.K.; Chae, S.-W.; Kim, H.-R.; Chae, H.J. Endoplasmic reticulum stress and cancer. J. Cancer Prev. 2014, 19, 75–88. [Google Scholar] [CrossRef]
  96. Adachi, T.; Kaminaga, T.; Yasuda, H.; Kamiya, T.; Hara, H. The involvement of endoplasmic reticulum stress in bile acid-induced hepatocellular injury. J. Clin. Biochem. Nutr. 2014, 54, 129–135. [Google Scholar] [CrossRef]
  97. Liu, C.; Zheng, J.; Ou, X.; Han, Y. Anti-cancer substances and safety of lactic acid bacteria in clinical treatment. Front. Microbiol. 2021, 12, 2893. [Google Scholar] [CrossRef] [PubMed]
  98. Jacouton, E.; Michel, M.-L.; Torres-Maravilla, E.; Chain, F.; Langella, P.; Bermúdez-Humarán, L.G. Elucidating the immune-related mechanisms by which probiotic strain Lactobacillus casei BL23 displays anti-tumoral properties. Front. Microbiol. 2019, 9, 3281. [Google Scholar] [CrossRef] [PubMed]
  99. Chen, Z.-Y.; Hsieh, Y.-M.; Huang, C.-C.; Tsai, C.-C. Inhibitory effects of probiotic Lactobacillus on the growth of human colonic carcinoma cell line HT-29. Molecules 2017, 22, 107. [Google Scholar] [CrossRef] [PubMed]
  100. Iyer, C.; Kosters, A.; Sethi, G.; Kunnumakkara, A.B.; Aggarwal, B.B.; Versalovic, J. Probiotic Lactobacillus reuteri promotes TNF-induced apoptosis in human myeloid leukemia-derived cells by modulation of NF-ΚB and MAPK signalling. Cell. Microbiol. 2008, 10, 1442–1452. [Google Scholar] [CrossRef]
  101. Kim, Y.; Oh, S.; Yun, H.S.; Oh, S.; Kim, S.H. Cell-bound exopolysaccharide from probiotic bacteria induces autophagic cell death of tumour cells. Lett. Appl. Microbiol. 2010, 51, 123–130. [Google Scholar] [CrossRef]
  102. Baldwin, C.; Millette, M.; Oth, D.; Ruiz, M.T.; Luquet, F.-M.; Lacroix, M. Probiotic Lactobacillus acidophilus and L. casei mix sensitize colorectal tumoral cells to 5-fluorouracil-induced apoptosis. Nutr. Cancer 2010, 62, 371–378. [Google Scholar] [CrossRef]
  103. Choi, S.S.; Kim, Y.; Han, K.S.; You, S.; Oh, S.; Kim, S.H. Effects of Lactobacillus strains on cancer cell proliferation and oxidative stress in vitro. Lett. Appl. Microbiol. 2006, 42, 452–458. [Google Scholar] [CrossRef]
  104. Hsieh, S.-C.; Liu, J.-M.; Pua, X.-H.; Ting, Y.; Hsu, R.-J.; Cheng, K.-C. Optimization of Lactobacillus acidophilus cultivation using taro waste and evaluation of its biological activity. Appl. Microbiol. Biotechnol. 2016, 100, 2629–2639. [Google Scholar] [CrossRef]
  105. Tuo, Y.; Jiang, S.; Qian, F.; Mu, G.; Liu, P.; Guo, Y.; Ma, C. Short communication: Antiproliferative effect of 8 different Lactobacillus strains on K562 cells. Int. J. Dairy Sci. 2015, 98, 106–110. [Google Scholar] [CrossRef]
  106. Liu, C.-T.; Chu, F.-J.; Chou, C.-C.; Yu, R.-C. Antiproliferative and anticytotoxic effects of cell fractions and exopolysaccharides from Lactobacillus casei 01. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2011, 721, 157–162. [Google Scholar] [CrossRef]
  107. Haroun, B.M.; Refaat, B.M.; El-Menoufy, H.A.; Amin, H.A.; El-Waseif, A.A. Structure analysis and antitumor activity of the exopolysaccharide from probiotic Lactobacillus plantarum NRRL B- 4496 in vitro and in vivo. J. Appl. Sci. 2013, 9, 425–434. [Google Scholar]
  108. Abd El Ghany, K.; Hamouda, R.; Abd Elhafez, E.; Mahrous, H.; Salem-Bekhit, M.; Hamza, H.A. A potential role of Lactobacillus acidophilus LA1 and its exopolysaccharides on cancer cells in male albino mice. Biotechnol. Biotechnol. Equip. 2015, 29, 977–983. [Google Scholar] [CrossRef]
  109. Abd El Ghany, K.; Elhafez, E.A.; Hamouda, R.A.; Mahrous, H.; Ahmed, F.A.H.; Hamza, H.A. Evaluation of antioxidant and antitumor activities of Lactobacillus acidophilus bacteria isolated from Egyptian infants. Int. J. Pharmacol. 2014, 10, 282–288. [Google Scholar] [CrossRef] [Green Version]
  110. Ji, F.; Fu, S.-J.; Guo, Z.-Y.; Pang, H.; Ju, W.-Q.; Wang, D.-P.; Hua, Y.-P.; He, X.-S. Prognostic value of combined preoperative lactate dehydrogenase and alkaline phosphatase levels in patients with resectable pancreatic ductal adenocarcinoma. Medicine 2016, 95, e4065. [Google Scholar] [CrossRef] [PubMed]
  111. Varma, G.; Seth, P.; Coutinho de Souza, P.; Callahan, C.; Pinto, J.; Vaidya, M.; Sonzogni, O.; Sukhatme, V.; Wulf, G.M.; Grant, A.K. Visualizing the effects of lactate dehydrogenase (LDH) inhibition and LDH-A genetic ablation in breast and lung cancer with hyperpolarized pyruvate NMR. NMR Biomed. 2021, 34, e4560. [Google Scholar] [CrossRef]
  112. Zhou, B.; Deng, J.; Chen, L.; Zheng, S. Preoperative neutrophil-to-lymphocyte ratio and tumor-related factors to predict lymph node metastasis in nonfunctioning pancreatic neuroendocrine tumors. Sci. Rep. 2017, 7, 17506. [Google Scholar] [CrossRef]
  113. Wu, J.; Zhang, Y.; Ye, L.; Wang, C. The anti-cancer effects and mechanisms of lactic acid bacteria exopolysaccharides in vitro: A review. Carbohydr. Polym. 2021, 253, 117308. [Google Scholar] [CrossRef]
  114. Zhong, L. Emerging roles of lactic acid bacteria in protection against colorectal cancer. World J. Gastroenterol. 2014, 20, 7878. [Google Scholar] [CrossRef]
  115. Ohigashi, S.; Hoshino, Y.; Ohde, S.; Onodera, H. Functional outcome, quality of life, and efficacy of probiotics in postoperative patients with colorectal cancer. Surg. Today 2011, 41, 1200–1206. [Google Scholar] [CrossRef]
  116. Hirayama, K.; Rafter, J. The role of lactic acid bacteria in colon cancer prevention: Mechanistic considerations. J. Clin. Exp. Oncol. 1999, 76, 391–394. [Google Scholar] [CrossRef]
  117. Amaretti, A.; di Nunzio, M.; Pompei, A.; Raimondi, S.; Rossi, M.; Bordoni, A. Antioxidant properties of potentially probiotic bacteria: In vitro and in vivo activities. Appl. Microbiol. Biotechnol. 2013, 97, 809–817. [Google Scholar] [CrossRef]
  118. Bruno-Bárcena, J.M.; Andrus, J.M.; Libby, S.L.; Klaenhammer, T.R.; Hassan, H.M. Expression of a heterologous manganese superoxide dismutase gene in intestinal lactobacilli provides protection against hydrogen peroxide toxicity. Appl. Environ. Microbiol. 2004, 70, 4702–4710. [Google Scholar] [CrossRef] [PubMed]
  119. Lee, J.; Hwang, K.-T.; Heo, M.-S.; Lee, J.-H.; Park, K.-Y. Resistance of Lactobacillus plantarum KCTC 3099 from Kimchi to oxidative stress. J. Med. Food 2005, 8, 299–304. [Google Scholar] [CrossRef] [PubMed]
  120. Lin, M.Y.; Chang, F.J. Antioxidative effect of intestinal bacteria Bifidobacterium longum ATCC 15708 and Lactobacillus acidophilus ATCC 4356. Dig. Dis. Sci. 2000, 45, 1617–1622. [Google Scholar] [CrossRef]
  121. Kumar, M.; Kumar, A.; Nagpal, R.; Mohania, D.; Behare, P.; Verma, V.; Kumar, P.; Poddar, D.; Aggarwal, P.K.; Henry, C.J.K.; et al. Cancer-preventing attributes of probiotics: An update. Int. J. Food Sci. Nutr. 2010, 61, 473–496. [Google Scholar] [CrossRef]
  122. Annuk, H.; Shchepetova, J.; Kullisaar, T.; Songisepp, E.; Zilmer, M.; Mikelsaar, M. Characterization of intestinal lactobacilli as putative probiotic candidates. J. Appl. Microbiol. 2003, 94, 403–412. [Google Scholar] [CrossRef]
  123. di Giacinto, C.; Marinaro, M.; Sanchez, M.; Strober, W.; Boirivant, M. Probiotics ameliorate recurrent Th1-mediated murine colitis by inducing IL-10 and IL-10-dependent TGF-beta-bearing regulatory cells. J. Immunol. 2005, 174, 3237–3246. [Google Scholar] [CrossRef] [PubMed]
  124. Mayer, M.L.; Phillips, C.M.; Townsend, R.A.; Halperin, S.A.; Lee, S.F. Differential activation of dendritic cells by toll-like receptor agonists isolated from the Gram-positive vaccine vector Streptococcus gordonii. Scand. J. Immunol. 2009, 69, 351–356. [Google Scholar] [CrossRef] [PubMed]
  125. Mayer, M.L.; Phillips, C.M.; Stadnyk, A.W.; Halperin, S.A.; Lee, S.F. Synergistic BM-DC activation and immune induction by the oral vaccine vector Streptococcus gordonii and exogenous tumor necrosis factor. Mol. Immunol. 2009, 46, 1883–1891. [Google Scholar] [CrossRef]
  126. Mohamadzadeh, M.; Olson, S.; Kalina, W.V.; Ruthel, G.; Demmin, G.L.; Warfield, K.L.; Bavari, S.; Klaenhammer, T.R. Lactobacilli activate human dendritic cells that skew T cells toward T helper 1 polarization. Proc. Natl. Acad. Sci. USA 2005, 102, 2880–2885. [Google Scholar] [CrossRef]
  127. Konstantinov, S.R.; Smidt, H.; de Vos, W.M.; Bruijns, S.C.M.; Singh, S.K.; Valence, F.; Molle, D.; Lortal, S.; Altermann, E.; Klaenhammer, T.R.; et al. S Layer protein A of Lactobacillus acidophilus NCFM regulates immature dendritic cell and T cell functions. Proc. Natl. Acad. Sci. USA 2008, 105, 19474–19479. [Google Scholar] [CrossRef] [PubMed]
  128. Mohamadzadeh, M.; Pfeiler, E.A.; Brown, J.B.; Zadeh, M.; Gramarossa, M.; Managlia, E.; Bere, P.; Sarraj, B.; Khan, M.W.; Pakanati, K.C.; et al. Regulation of induced colonic inflammation by Lactobacillus acidophilus deficient in lipoteichoic acid. Proc. Natl. Acad. Sci. USA 2011, 108 (Suppl. 1), 4623–4630. [Google Scholar] [CrossRef] [PubMed]
  129. Khazaie, K.; Zadeh, M.; Khan, M.W.; Bere, P.; Gounari, F.; Dennis, K.; Blatner, N.R.; Owen, J.L.; Klaenhammer, T.R.; Mohamadzadeh, M. Abating colon cancer polyposis by Lactobacillus acidophilus deficient in lipoteichoic acid. Proc. Natl. Acad. Sci. USA 2012, 109, 10462–10467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Lee, J.-W.; Shin, J.-G.; Kim, E.H.; Kang, H.E.; Yim, I.B.; Kim, J.Y.; Joo, H.-G.; Woo, H.J. Immunomodulatory and antitumor effects in vivo by the cytoplasmic fraction of Lactobacillus casei and Bifidobacterium longum. J. Vet. Sci. 2004, 5, 41–48. [Google Scholar] [PubMed]
  131. Takagi, A.; Matsuzaki, T.; Sato, M.; Nomoto, K.; Morotomi, M.; Yokokura, T. Enhancement of natural killer cytotoxicity delayed murine carcinogenesis by a probiotic microorganism. J. Carcinog. 2001, 22, 599–605. [Google Scholar] [CrossRef]
  132. Matsuzaki, T.; Yokokura, T.; Mutai, M. Antitumor effect of intrapleural administration of Lactobacillus casei in mice. Cancer Immunol. Immunother. 1988, 26, 209–214. [Google Scholar] [CrossRef]
  133. Lee, D.K.; Jang, S.; Kim, M.J.; Kim, J.H.; Chung, M.J.; Kim, K.J.; Ha, N.J. Anti-proliferative effects of Bifidobacterium adolescentis SPM0212 extract on human colon cancer cell lines. BMC Cancer 2008, 8, 310. [Google Scholar] [CrossRef]
  134. Yamane, T.; Sakamoto, T.; Nakagaki, T.; Nakano, Y. Lactic acid bacteria from kefir increase cytotoxicity of natural killer cells to tumor cells. Foods 2018, 7, 48. [Google Scholar] [CrossRef]
  135. Lakritz, J.R.; Poutahidis, T.; Levkovich, T.; Varian, B.J.; Ibrahim, Y.M.; Chatzigiagkos, A.; Mirabal, S.; Alm, E.J.; Erdman, S.E. Beneficial bacteria stimulate host immune cells to counteract dietary and genetic predisposition to mammary cancer in mice. Int. J. Cancer Res. 2014, 135, 529–540. [Google Scholar] [CrossRef]
  136. Guimarães, A.; Santiago, A.; Teixeira, J.A.; Venâncio, A.; Abrunhosa, L. Anti-aflatoxigenic effect of organic acids produced by Lactobacillus plantarum. Int. J. Food Microbiol. 2018, 264, 31–38. [Google Scholar] [CrossRef]
  137. Sekine, K.; Toida, T.; Saito, M.; Kuboyama, M.; Kawashima, T.; Hashimoto, Y. A New morphologically characterized cell-wall preparation (whole peptidoglycan) from Bifidobacterium infantis with a higher efficacy on the regression of an established tumor in mice. Cancer Res. 1985, 45, 1300–1307. [Google Scholar]
  138. Matsumoto, S.; Hara, T.; Nagaoka, M.; Mike, A.; Mitsuyama, K.; Sako, T.; Yamamoto, M.; Kado, S.; Takada, T. A component of polysaccharide peptidoglycan complex on Lactobacillus induced an improvement of murine model of inflammatory bowel disease and colitis-associated cancer. Immunology 2009, 128, e170–e180. [Google Scholar] [CrossRef]
  139. Kim, J.Y.; Woo, H.J.; Kim, Y.-S.; Lee, H.J. Screening for antiproliferative effects of cellular components from lactic acid bacteria against human cancer cell lines. Biotechnol. Lett. 2002, 24, 1431–1436. [Google Scholar] [CrossRef]
  140. Zhang, T.; Pan, D.; Yang, Y.; Jiang, X.; Zhang, J.; Zeng, X.; Wu, Z.; Sun, Y.; Guo, Y. Effect of Lactobacillus acidophilus CICC 6074 S-layer protein on colon cancer HT-29 cell proliferation and apoptosis. J. Agric. Food Chem. 2020, 68, 2639–2647. [Google Scholar] [CrossRef]
  141. Zang, C.; Wen, S.; Tang, L. The effects of nucleic acids from Lactobacillus fermented filtrate on anti-tumor and immunological function in tumor-bearing mice. Chin. J. Microecol. 2011, 23, 577–581. [Google Scholar]
  142. Asadollahi, P.; Ghanavati, R.; Rohani, M.; Razavi, S.; Esghaei, M.; Talebi, M. Anti-cancer effects of Bifidobacterium species in colon cancer cells and a mouse model of carcinogenesis. PLoS ONE. 2020, 15, e0232930. [Google Scholar] [CrossRef]
  143. Krasinskas, A.M. EGFR signaling in colorectal carcinoma. Pathol. Res. Int. 2011, 2011, 932932. [Google Scholar] [CrossRef]
  144. Sivan, A.; Corrales, L.; Hubert, N.; Williams, J.B.; Aquino-Michaels, K.; Earley, Z.M. Commensal Bifidobacterium promotes antitumor immunity and facilitates anti-PD-L1 efficacy. Science. 2015, 350, 1084–1089. [Google Scholar] [CrossRef] [PubMed]
  145. Wu, B.B.; Yang, Y.; Xu, X.; Wang, W.P. Effects of Bifidobacterium supplementation intestinal microbiota composition and the immune response in healthy infants. World J. Pediatr. 2016, 12, 177–182. [Google Scholar] [CrossRef]
  146. Grootaert, C.; Van De Wiele, T.; Van Roosbroeck, I.; Possemiers, S.; Vercoutter-Edouart, A.-S.; Verstraete, W.; Bracke, M.; Vanhoecke, B. Bacterial monocultures, propionate, butyrate and H2O2 modulate the expression, secretion and structure of the fasting-induced adipose factor in gut epithelial cell lines. Environ. Microbiol. 2011, 13, 1778–1789. [Google Scholar] [CrossRef]
  147. Shen, F.; Feng, J.; Wang, X. Vinegar treatment prevents the development of murine experimental colitis via inhibition of inflammation and apoptosis. J. Agric. Food Chem. 2016, 64, 1111–1121. [Google Scholar] [CrossRef] [PubMed]
  148. Ou, C.C.; Lin, S.L.; Tsai, J.-J.; Lin, M.-Y. Heat-killed lactic acid bacteria enhance immunomodulatory potential by skewing the immune response toward Th1 polarization. J. Food Sci. 2011, 76, M260–M267. [Google Scholar] [CrossRef] [PubMed]
  149. Kouhsari, E.; Ghadimi-Daresajini, A.; Abdollahi, H.; Amirmozafari, N.; Mahdavi, S.R.; Abbasian, S.; Mousavie, S.H.; Yaseri, H.F.; Moghaderi, M. The potential roles of bacteria to improve radiation treatment outcome. Clin. Transl. Oncol. 2017, 19, 1–13. [Google Scholar] [CrossRef]
  150. Abdollahi, H. Probiotic-based protection of normal tissues during radiotherapy. Nutrition 2014, 30, 495–496. [Google Scholar] [CrossRef]
  151. Khademi, S.; Abdollahi, H. Application of hydrogen producing microorganisms in radiotherapy: An idea. Iran. J. Public Health 2014, 43, 1018–1019. [Google Scholar]
  152. Abdollahi, H.; Atashzar, M.; Amini, M. The potential use of biogas producing microorganisms in radiation protection. J. Med. Hypotheses Idea. 2015, 9, 67–71. [Google Scholar] [CrossRef] [Green Version]
  153. Sedighi, M.; Bialvaei, A.Z.; Hamblin, M.R.; Ohadi, E.; Asadi, A.; Halajzadeh, M.; Lohrasbi, V.; Mohammadzadeh, N.; Amiriani, T.; Krutova, M.; et al. Therapeutic bacteria to combat cancer; current advances, challenges, and opportunities. Cancer Med. 2019, 8, 3167–3181. [Google Scholar] [CrossRef]
  154. Palma, L.; Muñoz, D.; Berry, C.; Murillo, J.; Caballero, P. Bacillus thuringiensis toxins: An overview of their biocidal activity. Toxins 2014, 6, 3296–3325. [Google Scholar] [CrossRef]
  155. Ghasemi-Dehkordi, P.; Doosti, A.; Jami, M.-S. The functions of azurin of Pseudomonas aeruginosa and human mammaglobin-A on proapoptotic and cell cycle regulatory genes expression in the MCF-7 breast cancer cell line. Saudi J. Biol. Sci. 2020, 27, 2308–2317. [Google Scholar] [CrossRef]
  156. Nigam, P.K.; Nigam, A. Botulinum toxin. Indian J. Dermatol. 2010, 55, 8–14. [Google Scholar] [CrossRef]
  157. Holmes, R.K. Biology and molecular epidemiology of diphtheria toxin and the tox gene. J. Infect. Dis. 2000, 181, S156–S167. [Google Scholar] [CrossRef] [PubMed]
  158. Bennett, M.J.; Eisenberg, D. Refined structure of monomeric diphtheria toxin at 2.3A resolution. Protein Sci. 1994, 3, 1464–1475. [Google Scholar] [CrossRef]
  159. Honjo, T.; Nishizuka, Y.; Hayaishi, O.; Kato, I. Diphtheria toxin-dependent adenosine diphosphate ribosylation of aminoacyl transferase II and inhibition of protein synthesis. J. Biol. Chem. 1968, 243, 3553–3555. [Google Scholar] [CrossRef]
  160. Collier, R.J.; Pappenheimer, A.M. Studies on the mode of action of diphtheria toxin. J. Exp. Med. 1964, 120, 1019–1039. [Google Scholar] [CrossRef]
  161. Dateoka, S.; Ohnishi, Y.; Kakudo, K. Effects of CRM197, a specific inhibitor of HB-EGF, in oral cancer. Med. Mol. Morphol. 2012, 45, 91–97. [Google Scholar] [CrossRef]
  162. Miyamoto, S.; Yotsumoto, F.; Ueda, T.; Fukami, T.; Sanui, A.; Miyata, K.; Nam, S.O.; Fukagawa, S.; Katsuta, T.; Maehara, M.; et al. BK-UM in patients with recurrent ovarian cancer or peritoneal cancer: A first-in-human phase-I study. BMC Cancer 2017, 17, 89. [Google Scholar] [CrossRef] [Green Version]
  163. Mitamura, T.; Higashiyama, S.; Taniguchi, N.; Klagsbrun, M.; Mekada, E. Diphtheria toxin binds to the epidermal growth factor (EGF)-like domain of human heparin-binding EGF-like growth factor/diphtheria toxin receptor and inhibits specifically its mitogenic activity. J. Biol. Chem. 1995, 270, 1015–1019. [Google Scholar] [CrossRef]
  164. Nam, S.O.; Yotsumoto, F.; Miyata, K.; Fukagawa, S.; Odawara, T.; Manabe, S.; Ishikawa, T.; Kuroki, M.; Yasunaga, S.; Miyamoto, S. Anti-tumor effect of intravenous administration of CRM197 for triple-negative breast cancer therapy. Anticancer Res. 2016, 36, 3651–3657. [Google Scholar]
  165. Zhang, Y.; Schulte, W.; Pink, D.; Phipps, K.; Zijlstra, A.; Lewis, J.D.; Waisman, D.M. Sensitivity of cancer cells to truncated diphtheria toxin. PLoS ONE 2010, 5, e10498. [Google Scholar] [CrossRef]
  166. Allahyari, H.; Heidari, S.; Ghamgosha, M.; Saffarian, P.; Amani, J. Immunotoxin: A new tool for cancer therapy. Tumor Biol. 2017, 39, 101042831769222. [Google Scholar] [CrossRef]
  167. Shapira, A.; Benhar, I. Toxin-based therapeutic approaches. Toxins 2010, 2, 2519–2583. [Google Scholar] [CrossRef]
  168. Mathew, M.; Verma, R.S. Humanized immunotoxins: A new generation of immunotoxins for targeted cancer therapy. Cancer Sci. 2009, 100, 1359–1365. [Google Scholar] [CrossRef] [PubMed]
  169. Jin, F.S.; Youle, R.J.; Johnson, V.G.; Shiloach, J.; Fass, R.; Longo, D.L.; Bridges, S.H. Suppression of the immune response to immunotoxins with anti-CD4 monoclonal antibodies. J. Immunol. Res. 1991, 146, 1806–1811. [Google Scholar]
  170. Tsutsumi, Y.; Onda, M.; Nagata, S.; Lee, B.; Kreitman, R.J.; Pastan, I. Site-specific chemical modification with polyethylene glycol of recombinant immunotoxin anti-tac(Fv)-PE38 (LMB-2) improves antitumor activity and reduces animal toxicity and immunogenicity. Proc. Natl. Acad. Sci. USA 2000, 97, 8548–8553. [Google Scholar] [CrossRef]
  171. Kiyokawa, T.; Shirono, K.; Hattori, T.; Nishimura, H.; Yamaguchi, K.; Nichols, J.C.; Strom, T.B.; Murphy, J.R.; Takatsuki, K. Cytotoxicity of interleukin 2-toxin toward lymphocytes from patients with adult T-cell leukemia. Cancer Res. 1989, 49, 4042–4046. [Google Scholar]
  172. Wang, Z.; Zheng, Q.; Zhang, H.; Bronson, R.T.; Madsen, J.C.; Sachs, D.H.; Huang, C.A.; Wang, Z. Ontak-like Human IL-2 fusion toxin. J. Immunol. Methods 2017, 448, 51–58. [Google Scholar] [CrossRef]
  173. Zahaf, N.-I.; Schmidt, G. Bacterial toxins for cancer therapy. Toxins 2017, 9, 236. [Google Scholar] [CrossRef] [PubMed]
  174. Hosseininejad-Chafi, M.; Alirahimi, E.; Ramezani, B.; Oghalaie, A.; Sotoudeh, N.; Ghaderi, H.; Kazemi-Lomedasht, F.; Habibi-Anbouhi, M.; Moazzami, R.; Behdani, M. In vivo solid tumor targeting with recombinant VEGF-diphtheria immunotoxin. Iran. J. Basic Med. Sci. 2022, 25, 27–31. [Google Scholar] [CrossRef]
  175. Vallera, D.A.; Li, C.; Jin, N.; Panoskaltsis-Mortari, A.; Hall, W.A. Targeting urokinase-type plasminogen activator receptor on human glioblastoma tumors with diphtheria toxin fusion protein DTAT. J. Natl. Cancer Inst. 2002, 94, 597–606. [Google Scholar] [CrossRef] [Green Version]
  176. Shafiee, F.; Rabbani, M.; Jahanian-Najafabadi, A. Evaluation of cytotoxic and apoptotic effects of DT386–BR2: A promising anticancer fusion protein. J. Rep. Pharm. Sci. 2020, 9, 68. [Google Scholar] [CrossRef]
  177. Shafiee, F.; Rabbani, M.; Jahanian-Najafabadi, A. Production and evaluation of cytotoxic effects of DT386-BR2 fusion protein as a novel anti-cancer agent. J. Microbiol. Methods 2016, 130, 100–105. [Google Scholar] [CrossRef]
  178. Syed, Y.Y. Tagraxofusp: First global approval. Drugs 2019, 79, 579–583. [Google Scholar] [CrossRef] [PubMed]
  179. Ghodrati Siahmazgi, M.; Nasiri Khalili, M.A.; Zeinoddini, M.; Ahmadpour, F.; Khodadadi, S. Purification and characterization of DT389GCSF fusion protein: A unique immunotoxin against the human granulocyte-colony stimulating factor receptor. Int. J. Pept. Res. Ther. 2020, 26, 767–774. [Google Scholar] [CrossRef]
  180. Xie, G.; Shan, L.; Liu, Y.; Wu, T.-C.; Gu, X. Antitumor efficacy of EGFR-targeted recombinant immunotoxin in human head and neck squamous cell carcinoma. Biology 2022, 11, 486. [Google Scholar] [CrossRef] [PubMed]
  181. Mousavi, A.; Sabouri, A.; Eskafi, A.H.; Alirahimi, E.; Kazemi-Lomedasht, F.; Ghaderi, H.; Behdani, M. In Vivo tumor therapy with novel immunotoxin containing programmed cell death protein-1 and diphtheria toxin. Monoclon. Antibodies Immunodiagn. Immunother. 2021, 40, 113–117. [Google Scholar] [CrossRef]
  182. Hashemi Yeganeh, H.; Heiat, M.; Kieliszek, M.; Alavian, S.M.; Rezaie, E. DT389-YP7, a recombinant immunotoxin against glypican-3 that inhibits hepatocellular cancer cells: An in vitro study. Toxins 2021, 13, 749. [Google Scholar] [CrossRef]
  183. Kokai-Kun, J.F.; McClane, B.A. Deletion analysis of the Clostridium perfringens enterotoxin. Infect. Immun. 1997, 65, 1014–1022. [Google Scholar] [CrossRef]
  184. Hanna, P.C.; Wnek, A.P.; McClane, B.A. Molecular Cloning of the 3′ Half of the Clostridium perfringens enterotoxin gene and demonstration that this region encodes receptor-binding activity. J. Bacteriol. 1989, 171, 6815–6820. [Google Scholar] [CrossRef]
  185. Kitadokoro, K.; Nishimura, K.; Kamitani, S.; Fukui-Miyazaki, A.; Toshima, H.; Abe, H.; Kamata, Y.; Sugita-Konishi, Y.; Yamamoto, S.; Karatani, H.; et al. Crystal structure of Clostridium perfringens enterotoxin displays features of β-pore-forming toxins. J. Biol. Chem. 2011, 286, 19549–19555. [Google Scholar] [CrossRef] [Green Version]
  186. Chen, J.-G.; Horwitz, S.B. Differential mitotic responses to microtubule-stabilizing and -destabilizing drugs. Cancer Res. 2002, 62, 1935–1938. [Google Scholar]
  187. Hashimoto, Y.; Yagi, K.; Kondoh, M. Roles of the first-generation claudin binder, Clostridium perfringens enterotoxin, in the diagnosis and claudin-targeted treatment of epithelium-derived cancers. Pflug. Arch. Eur. J. Physiol. 2017, 469, 45–53. [Google Scholar] [CrossRef] [PubMed]
  188. Freedman, J.; Shrestha, A.; McClane, B. Clostridium perfringens enterotoxin: Action, genetics, and translational applications. Toxins 2016, 8, 73. [Google Scholar] [CrossRef] [PubMed]
  189. Briggs, D.C.; Naylor, C.E.; Smedley, J.G.; Lukoyanova, N.; Robertson, S.; Moss, D.S.; McClane, B.A.; Basak, A.K. Structure of the food-poisoning Clostridium perfringens enterotoxin reveals similarity to the aerolysin-like pore-forming toxins. J. Mol. Biol. 2011, 413, 138–149. [Google Scholar] [CrossRef]
  190. Harry, K.H.; Zhou, R.; Kroos, L.; Melville, S.B. Sporulation and enterotoxin (CPE) synthesis are controlled by the sporulation-specific sigma factors SigE and SigK in Clostridium perfringens. J. Bacteriol. 2009, 191, 2728–2742. [Google Scholar] [CrossRef] [PubMed]
  191. Hu, W.S.; Woo, D.U.; Kang, Y.J.; Koo, O.K. Biofilm and spore formation of Clostridium perfringens and its resistance to disinfectant and oxidative stress. Antibiotics 2021, 10, 396. [Google Scholar] [CrossRef] [PubMed]
  192. Li, J.; Paredes-Sabja, D.; Sarker, M.R.; McClane, B.A. Clostridium perfringens sporulation and sporulation-associated toxin production. Microbiol. Spectr. 2016, 4. [Google Scholar] [CrossRef]
  193. Mi, E.; Li, J.; McClane, B.A. NanR regulates sporulation and enterotoxin production by Clostridium perfringens type F strain F4969. Infect. Immun. 2018, 86, e00416-18. [Google Scholar] [CrossRef]
  194. Zhao, Y.; Melville, S.B. Identification and characterization of Sporulation-dependent promoters upstream of the enterotoxin gene (Cpe) of Clostridium perfringens. J. Bacteriol. 1998, 180, 136–142. [Google Scholar] [CrossRef]
  195. Neuhaus, W.; Piontek, A.; Protze, J.; Eichner, M.; Mahringer, A.; Subileau, E.-A.; Lee, I.-F.M.; Schulzke, J.D.; Krause, G.; Piontek, J. Reversible opening of the blood-brain barrier by claudin-5-binding variants of Clostridium perfringens enterotoxin’s claudin-binding domain. Biomaterials 2018, 161, 129–143. [Google Scholar] [CrossRef]
  196. Vecchio, A.J.; Stroud, R.M. Claudin-9 structures reveal mechanism for toxin-induced gut barrier breakdown. Proc. Natl. Acad. Sci. USA 2019, 116, 17817–17824. [Google Scholar] [CrossRef] [Green Version]
  197. Maeda, T.; Murata, M.; Chiba, H.; Takasawa, A.; Tanaka, S.; Kojima, T.; Masumori, N.; Tsukamoto, T.; Sawada, N. Claudin-4-targeted therapy using Clostridium perfringens enterotoxin for prostate cancer. Prostate 2012, 72, 351–360. [Google Scholar] [CrossRef] [PubMed]
  198. Abedi, S.; Doosti, A.; Jami, M.-S. Impacts of the prostate stem cell antigen (PSCA) and Clostridium perfringens enterotoxin (CPE) on the apoptosis and cell cycle regulatory genes in PC3. Prep. Biochem. Biotechnol. 2020, 50, 47–55. [Google Scholar] [CrossRef] [PubMed]
  199. Navarro, M.; McClane, B.; Uzal, F. Mechanisms of action and cell death associated with Clostridium perfringens toxins. Toxins 2018, 10, 212. [Google Scholar] [CrossRef] [PubMed]
  200. Shim, M.K.; Na, J.; Cho, I.K.; Jang, E.H.; Park, J.; Lee, S.; Kim, J.-H. Targeting of claudin-4 by Clostridium perfringens enterotoxin-conjugated polysialic acid nanoparticles for pancreatic cancer therapy. J. Control. Release 2021, 331, 434–442. [Google Scholar] [CrossRef]
  201. Tsukita, S.; Tanaka, H.; Tamura, A. The Claudins: From tight junctions to biological systems. Trends Biochem. Sci. 2019, 44, 141–152. [Google Scholar] [CrossRef]
  202. Pahle, J.; Menzel, L.; Niesler, N.; Kobelt, D.; Aumann, J.; Rivera, M.; Walther, W. Rapid eradication of colon carcinoma by Clostridium perfringens enterotoxin suicidal gene therapy. BMC Cancer 2017, 17, 129. [Google Scholar] [CrossRef]
  203. Gabig, T.G.; Waltzer, W.C.; Whyard, T.; Romanov, V. Clostridium perfringens enterotoxin as a potential drug for intravesical treatment of bladder cancer. Biochem. Biophys. Res. Commun. 2016, 478, 887–892. [Google Scholar] [CrossRef]
  204. Gao, Z.; Xu, X.; McClane, B.; Zeng, Q.; Litkouhi, B.; Welch, W.R.; Berkowitz, R.S.; Mok, S.C.; Garner, E.I.O. C terminus of Clostridium perfringens enterotoxin downregulates CLDN4 and sensitizes ovarian cancer cells to taxol and carboplatin. Clin. Cancer Res. 2011, 17, 1065–1074. [Google Scholar] [CrossRef]
  205. Becker, A.; Leskau, M.; Schlingmann-Molina, B.L.; Hohmeier, S.C.; Alnajjar, S.; Murua Escobar, H.; Ngezahayo, A. Functionalization of gold-nanoparticles by the Clostridium perfringens enterotoxin C-terminus for tumor cell ablation using the gold nanoparticle-mediated laser perforation technique. Sci. Rep. 2018, 8, 14963. [Google Scholar] [CrossRef]
  206. Martin, D.T.; Lee, J.S.; Liu, Q.; Galiana, G.; Sprenkle, P.C.; Humphrey, P.A.; Petrylak, D.P.; Weinreb, J.C.; Schulam, P.G.; Weiss, R.M.; et al. Targeting prostate cancer with Clostridium perfringens enterotoxin functionalized nanoparticles co-encapsulating imaging cargo enhances magnetic resonance imaging specificity. Nanomed. Nanotechnol. Biol. Med. 2022, 40, 102477. [Google Scholar] [CrossRef]
  207. Dolly, J.O.; Aoki, K.R. The structure and mode of action of different botulinum toxins. Eur. J. Neurol. 2006, 13, 1–9. [Google Scholar] [CrossRef] [PubMed]
  208. Ron, Y.; Avni, Y.; Lukovetski, A.; Wardi, J.; Geva, D.; Birkenfeld, S.; Halpern, Z. Botulinum toxin type-a in therapy of patients with anismus. Dis. Colon Rectum 2001, 44, 1821–1826. [Google Scholar] [CrossRef] [PubMed]
  209. Jankovic, J.; Schwartz, K. Botulinum toxin treatment of tremors. Neurology 1991, 41, 1185. [Google Scholar] [CrossRef] [PubMed]
  210. Jankovic, J.; Schwartz, K.; Donovan, D.T. Botulinum toxin treatment of cranial-cervical dystonia, spasmodic dysphonia, other focal dystonias and hemifacial spasm. J. Neurol. Neurosurg. Psychiatry 1990, 53, 633–639. [Google Scholar] [CrossRef] [PubMed]
  211. Bandala, C.; Terán-Melo, J.L.; Anaya-Ruiz, M.; Mejía-Barradas, C.M.; Domínguez-Rubio, R.; de la Garza-Montano, P.; Alfaro-Rodríguez, A.; Lara-Padilla, E. Effect of botulinum neurotoxin type A (BoNTA) on the morphology and viability of 3T3 murine fibroblasts. Int. J. Clin. Exp. Pathol. 2015, 8, 9458–9462. [Google Scholar]
  212. Aoki, K.R. Evidence for antinociceptive activity of botulinum toxin type A in pain management. Headache 2003, 43, 9–15. [Google Scholar] [CrossRef]
  213. Gul, N.; Smith, L.A.; Ahmed, S.A. Light chain separated from the rest of the type A botulinum neurotoxin molecule is the most catalytically active form. PLoS ONE 2010, 5, e12872. [Google Scholar] [CrossRef]
  214. Dover, N.; Barash, J.R.; Hill, K.K.; Xie, G.; Arnon, S.S. Molecular characterization of a novel botulinum neurotoxin type H gene. J. Infect. Dis. 2014, 209, 192–202. [Google Scholar] [CrossRef]
  215. Dolly, J.O.; Black, J.; Williams, R.S.; Melling, J. Acceptors for botulinum neurotoxin reside on motor nerve terminals and mediate its internalization. Nature 1984, 307, 457–460. [Google Scholar] [CrossRef] [PubMed]
  216. Choudhury, S.; Baker, M.R.; Chatterjee, S.; Kumar, H. Botulinum toxin: An update on pharmacology and newer products in development. Toxins 2021, 13, 58. [Google Scholar] [CrossRef]
  217. Huang, X.; Wheeler, M.B.; Kang, Y.-H.; Sheu, L.; Lukacs, G.L.; Trimble, W.S.; Gaisano, H.Y. Truncated SNAP-25 (1-197), like botulinum neurotoxin A, can inhibit insulin secretion from HIT-T15 insulinoma cells. Mol. Endocrinol. 1998, 12, 1060–1070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Ansiaux, R.; Baudelet, C.; Cron, G.O.; Segers, J.; Dessy, C.; Martinive, P.; de Wever, J.; Verrax, J.; Wauthier, V.; Beghein, N.; et al. Botulinum toxin potentiates cancer radiotherapy and chemotherapy. Clin. Cancer Res. 2006, 12, 1276–1283. [Google Scholar] [CrossRef] [PubMed]
  219. Karsenty, G.; Rocha, J.; Chevalier, S.; Scarlata, E.; Andrieu, C.; Zouanat, F.Z.; Rocchi, P.; Giusiano, S.; Elzayat, E.A.; Corcos, J. Botulinum toxin type A inhibits the growth of LNCaP human prostate cancer cells in Vitro and in Vivo. Prostate 2009, 69, 1143–1150. [Google Scholar] [CrossRef]
  220. Hajighasemlou, S.; Alebouyeh, M.; Rastegar, H.; Manzari, M.T.; Mirmoghtadaei, M.; Moayedi, B.; Ahmadzadeh, M.; Parvizpour, F.; Johari, B.; Naeini, M.M.; et al. Preparation of immunotoxin herceptin-botulinum and killing effects on two breast cancer cell lines. Asian Pac. J. Cancer Prev. 2015, 16, 5977–5981. [Google Scholar] [CrossRef] [PubMed]
  221. Proietti, S.; Nardicchi, V.; Porena, M.; Giannantoni, A. Botulinum toxin type-A toxinactivity in prostate cancer cell lines. Urol. J. 2012, 79, 135–141. [Google Scholar] [CrossRef] [PubMed]
  222. Bandala, C.; Perez-Santos, J.L.M.; Lara-Padilla, E.; Lopez, M.G.D.; Anaya-Ruiz, M. Effect of botulinum toxin A on proliferation and apoptosis in the T47D breast cancer cell line. Asian Pac. J. Cancer Prev. 2013, 14, 891–894. [Google Scholar] [CrossRef]
  223. Rust, A.; Leese, C.; Binz, T.; Davletov, B. Botulinum neurotoxin type C protease induces apoptosis in differentiated human neuroblastoma cells. Oncotarget 2016, 7, 33220–33228. [Google Scholar] [CrossRef]
  224. Coarfa, C.; Florentin, D.; Putluri, N.; Ding, Y.; Au, J.; He, D.; Ragheb, A.; Frolov, A.; Michailidis, G.; Lee, M.; et al. Influence of the neural microenvironment on prostate cancer. Prostate 2018, 78, 128–139. [Google Scholar] [CrossRef]
  225. He, D.; Manzoni, A.; Florentin, D.; Fisher, W.; Ding, Y.; Lee, M.; Ayala, G. Biologic effect of neurogenesis in pancreatic cancer. Hum. Pathol. 2016, 52, 182–189. [Google Scholar] [CrossRef]
  226. Rodrigues, A.N.; Duarte, A. Study of the neutralization of cell multiplication, in vitro, with botulinum neurotoxin type A. Int. J. Adv. Eng. Res. Sci. 2021, 8, 390–393. [Google Scholar] [CrossRef]
  227. Akyuva, Y. Clostridium botulinum neurotoxin A inhibits DBTRG glioblastoma cell proliferation and TRPV1 channel signaling pathways. J. Cell. Neurosci. Oxidative Stress 2020, 12, 903–913. [Google Scholar] [CrossRef]
  228. Wang, L.; Lei, Q.; Liu, Y.; Song, G.; Song, C. A case report of the beneficial effects of botulinum toxin type A on Raynaud phenomenon in a patient with lung cancer. Medicine 2016, 95, e5092. [Google Scholar] [CrossRef] [PubMed]
  229. de Groef, A.; Devoogdt, N.; van Kampen, M.; Nevelsteen, I.; Smeets, A.; Neven, P.; Geraerts, I.; Dams, L.; van der Gucht, E.; Debeer, P. Effectiveness of botulinum toxin A for persistent upper limb pain after breast cancer treatment: A double-blinded randomized controlled trial. Arch. Phys. Med. Rehabil. 2018, 99, 1342–1351. [Google Scholar] [CrossRef] [PubMed]
  230. Bach, C.-A.; Wagner, I.; Lachiver, X.; Baujat, B.; Chabolle, F. Botulinum toxin in the treatment of post-radiosurgical neck contracture in head and neck cancer: A novel approach. Eur. Ann. Otorhinolaryngol. Head Neck Dis. 2012, 129, 6–10. [Google Scholar] [CrossRef]
  231. Fu, J.; Ngo, A.; Shin, K.; Bruera, E. Botulinum toxin injection and phenol nerve block for reduction of end-of-life pain. J. Palliat. Med. 2013, 16, 1637–1640. [Google Scholar] [CrossRef]
  232. Layeeque, R.; Hochberg, J.; Siegel, E.; Kunkel, K.; Kepple, J.; Henry-Tillman, R.S.; Dunlap, M.; Seibert, J.; Klimberg, V.S. Botulinum toxin infiltration for pain control after mastectomy and expander reconstruction. Ann. Surg. 2004, 240, 608–614. [Google Scholar] [CrossRef]
  233. Iglewski, B.H.; Kabat, D. NAD-dependent inhibition of protein synthesis by Pseudomonas aeruginosa toxin. Proc Natl Acad Sci USA. 1977, 72, 2284–2288. [Google Scholar] [CrossRef]
  234. Iglewski, B.H.; Liu, P.V.; Kabat, D. Mechanism of action of Pseudomonas aeruginosa exotoxin A: Adenosine diphosphate-ribosylation of mammalian elongation factor 2 in vitro and in vivo. Infect Immun. 1977, 15, 138–144. [Google Scholar] [CrossRef]
  235. Wedekind, J.E.; Trame, C.B.; Dorywalska, M.; Koehl, P.; Raschke, T.M.; McKee, M.; FitzGerald, D.; Collier, R.J.; McKay, D.B. Refined crystallographic structure of Pseudomonas aeruginosa exotoxin A and its implications for the molecular mechanism of toxicity. J. Mol. Biol. 2001, 314, 823–837. [Google Scholar] [CrossRef]
  236. Odumosu, O.; Nicholas, D.; Yano, H.; Langridge, W. AB toxins: A paradigm switch from deadly to desirable. Toxins 2010, 2, 1612–1645. [Google Scholar] [CrossRef]
  237. Wolf, P.; Elsässer-Beile, U. Pseudomonas exotoxin A: From virulence factor to anti-cancer agent. Int. J. Med. Microbiol. 2009, 299, 161–176. [Google Scholar] [CrossRef]
  238. Weldon, J.E.; Pastan, I.A. Guide to taming a toxin-recombinant immunotoxins constructed from Pseudomonas exotoxin A for the treatment of cancer. FEBS J. 2011, 278, 4683–4700. [Google Scholar] [CrossRef] [PubMed]
  239. Michalska, M.; Wolf, P. Pseudomonas exotoxin A: Optimized by evolution for effective killing. Front. Microbiol. 2015, 6, 963. [Google Scholar] [CrossRef]
  240. Kreitman, R.J.; Pastan, I. Importance of the glutamate residue of KDEL in increasing the cytotoxicity of Pseudomonas exotoxin derivatives and for increased binding to the KDEL receptor. Biochem. J. 1995, 307, 29–37. [Google Scholar] [CrossRef]
  241. Kreitman, R.J.; Pastan, I. Antibody fusion proteins: Anti-CD22 recombinant immunotoxin moxetumomab pasudotox. Clin. Cancer Res. 2011, 17, 6398–6405. [Google Scholar] [CrossRef]
  242. Schneider, A.K.; Vainshtein, I.; Roskos, L.K.; Chavez, C.; Sun, B.; Liang, M. An immunoinhibition approach to overcome the impact of pre-existing antibodies on cut point establishment for immunogenicity assessment of moxetumomab pasudotox. J. Immunol. Meth. 2016, 435, 68–76. [Google Scholar] [CrossRef] [PubMed]
  243. Anon. AstraZeneca Pharmaceuticals LP-Wilmington, DE. 2018. Available online: https://lawyers.findlaw.com/profile/view/3358972_1 (accessed on 10 June 2022).
  244. Hassan, R.; Bullock, S.; Premkumar, A.; Kreitman, R.J.; Kindler, H.; Willingham, M.C.; Pastan, I. Phase I study of SS1P, a recombinant anti-mesothelin immunotoxin given as a bolus I.V. infusion to patients with mesothelin-expressing mesothelioma, ovarian, and pancreatic cancers. Clin Cancer Res. 2007, 13, 5144–5149. [Google Scholar] [CrossRef] [PubMed]
  245. Hassan, R.; Miller, A.C.; Sharon, E.; Thomas, A.; Reynolds, J.C.; Ling, A.; Kreitman, R.J.; Miettinen, M.M.; Steinberg, S.M.; Fowler, D.H.; et al. Major cancer regressions in mesothelioma after treatment with an anti-mesothelin immunotoxin and immune suppression. Sci. Transl. Med. 2013, 5, 208ra147. [Google Scholar] [CrossRef]
  246. Hassan, R.; Alewine, C.; Mian, I.; Spreafico, A.; Siu, L.L.; Gomez-Roca, C.; Delord, J.P.; Italiano, A.; Lassen, U.; Soria, J.C.; et al. Phase 1 study of the immunotoxin LMB-100 in patients with mesothelioma and other solid tumors expressing mesothelin. Cancer 2020, 126, 4936–4947. [Google Scholar] [CrossRef]
  247. Mazor, R.; Pastan, I. Immunogenicity of immunotoxins containing Pseudomonas exotoxin A: Causes, consequences, and mitigation. Front. Immunol. 2020, 11, 1261. [Google Scholar] [CrossRef]
  248. Sobell, H.M. Actinomycin and DNA transcription. Proc. Natl. Acad. Sci. USA 1985, 82, 5328–5331. [Google Scholar] [CrossRef] [Green Version]
  249. Sha, F.; Chen, F.M. Actinomycin D binds strongly to d(CGACGACG) and d(CGTCGTCG). Biophys. J. 2000, 79, 2095–2104. [Google Scholar] [CrossRef]
  250. D’Angio, G.J.; Evans, A.; Breslow, N.; Beckwith, B.; Bishop, H.; Farewell, V.; Goodwin, W.; Leape, L.; Palmer, N.; Sinks, L.; et al. The treatment of Wilms’ tumor: Results of the second national Wilms’ tumor study. Cancer 1981, 47, 2302–2311. [Google Scholar] [CrossRef]
  251. Khatua, S.; Nair, C.N.; Ghosh, K. Immune-mediated thrombocytopenia following dactinomycin therapy in a child with alveolar rhabdomyosarcoma. J. Pediatr. Hematol. Oncol. 2004, 26, 777–779. [Google Scholar] [CrossRef] [PubMed]
  252. Cortes, C.L.; Veiga, S.R.; Almacellas, E.; Hernández-Losa, J.; Ferreres, J.C.; Kozma, S.C.; Ambrosio, S.; Thomas, G.; Tauler, A. Effect of low doses of actinomycin D on neuroblastoma cell lines. Mol. Cancer 2016, 15, 1. [Google Scholar] [CrossRef]
  253. Subbiah, V.; Anderson, P.; Lazar, A.J.; Burdett, E.; Raymond, K.; Ludwig, J.A. Ewing’s sarcoma: Standard and experimental treatment options. Curr. Treat. Options Oncol. 2009, 10, 126–140. [Google Scholar] [CrossRef]
  254. Liu, X.F.; Xiang, L.; Zhou, Q.; Carralot, J.-P.; Prunotto, M.; Niederfellner, G.; Pastan, I. Actinomycin D enhances killing of cancer cells by immunotoxin RG7787 through activation of the extrinsic pathway of apoptosis. Proc. Natl. Acad. Sci. USA 2016, 113, 10666–10671. [Google Scholar] [CrossRef]
  255. Egger, C.; Cannet, C.; Gérard, C.; Jarman, E.; Jarai, G.; Feige, A.; Suply, T.; Micard, A.; Dunbar, A.; Tigani, B.; et al. Administration of bleomycin via the oropharyngeal aspiration route leads to sustained lung fibrosis in mice and rats as quantified by UTE-MRI and histology. PLoS ONE 2013, 8, e63432. [Google Scholar] [CrossRef]
  256. Umezawa, H.; Maeda, K.; Takeuchi, T.; Okami, Y. New antibiotics, bleomycin A and B. J. Antibiot. 1966, 19, 200–209. [Google Scholar]
  257. Dorr, R.T. Bleomycin Pharmacology: Mechanism of action and resistance, and clinical pharmacokinetics. Semin. Oncol. 1992, 19, 3–8. [Google Scholar]
  258. Chen, J.; Ghorai, M.K.; Kenney, G.; Stubbe, J. Mechanistic studies on bleomycin-mediated DNA damage: Multiple binding modes can result in double-stranded DNA cleavage. Nucleic Acids Res. 2008, 36, 3781–3790. [Google Scholar] [CrossRef] [PubMed]
  259. da Rocha, A. Natural products in anticancer therapy. Curr. Opin. Pharmacol. 2001, 1, 364–369. [Google Scholar] [CrossRef]
  260. Mir, L.M.; Tounekti, O.; Orlowski, S. Bleomycin: Revival of an old drug. Vascul. Pharmacol. 1996, 27, 745–748. [Google Scholar] [CrossRef]
  261. Gao, Y.; Shang, Q.; Li, W.; Guo, W.; Stojadinovic, A.; Mannion, C.; Man, Y.; Chen, T. Antibiotics for cancer treatment: A double-edged sword. J. Cancer 2020, 11, 5135–5149. [Google Scholar] [CrossRef] [PubMed]
  262. Shuhendler, A.J.; Cheung, R.Y.; Manias, J.; Connor, A.; Rauth, A.M.; Wu, X.Y. A Novel doxorubicin-mitomycin C co-encapsulated nanoparticle formulation exhibits anti-cancer synergy in multidrug resistant human breast cancer cells. Breast Cancer Res. Treat. 2010, 119, 255–269. [Google Scholar] [CrossRef]
  263. Khasraw, M.; Bell, R.; Dang, C. Epirubicin: Is it like doxorubicin in breast cancer? A clinical review. Breast 2012, 21, 142–149. [Google Scholar] [CrossRef]
  264. Plosker, G.L.; Faulds, D. Epirubicin. Drugs 1993, 45, 788–856. [Google Scholar] [CrossRef]
  265. Bonadonna, G.; Gianni, L.; Santoro, A.; Bonfante, V.; Bidoli, P.; Casali, P.; Demicheli, R.; Valagussa, P. Drugs ten years later: Epirubicin. Ann. Oncol. 1993, 4, 359–369. [Google Scholar] [CrossRef]
  266. Rohiwal, S.S.; Tiwari, A.P. Consensus Protocols for Animal Experimentation and Nanomedicine Trials at Clinical Stage in Breast Cancer. In Nanomedicines for Breast Cancer Theranostics; Elsevier: Amsterdam, The Netherlands, 2020; pp. 331–349. [Google Scholar]
  267. Tan, C.; Tasaka, H.; Yu, K.-P.; Murphy, M.L.; Karnofsky, D.A. Daunomycin, an antitumor antibiotic, in the treatment of neoplastic disease. Clinical evaluation with special reference to childhood leukemia. Cancer 1967, 20, 333–353. [Google Scholar] [CrossRef]
  268. Maghzi, A.-H.; Borazanci, A.; McGee, J.; Steven Alexander, J.; Gonzalez-Toledo, E.; Minagar, A. Multiple sclerosis: Pathophysiology, Clinical Features, Diagnosis, and Management. In Neuroinflammation; Elsevier: Amsterdam, The Netherlands, 2011; pp. 1–23. [Google Scholar]
  269. Bellosillo, B.; Piqué, M.; Barragán, M.; Castaño, E.; Villamor, N.; Colomer, D.; Montserrat, E.; Pons, G.; Gil, J. Aspirin and salicylate induce apoptosis and activation of caspases in B-cell chronic lymphocytic leukemia cells. Blood 1998, 92, 1406–1414. [Google Scholar] [CrossRef]
  270. Fidler, J.M.; DeJoy, S.Q.; Gibbons, J.J. Selective immunomodulation by the antineoplastic agent mitoxantrone. I. Suppression of B lymphocyte function. J. Immunol. Res. 1986, 137, 727–732. [Google Scholar]
  271. Tomasz, M. Mitomycin C: Small, fast and deadly (but very selective). Chem. Biol. 1995, 2, 575–579. [Google Scholar] [CrossRef]
  272. Verweij, J.; Pinedo, H.M. Mitomycin C: Mechanism of action, usefulness and limitations. Anticancer Drugs 1990, 1, 5–13. [Google Scholar] [CrossRef] [PubMed]
  273. Bradner, W.T. Mitomycin C: A clinical update. Cancer Treat. Rev. 2001, 27, 35–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Takeuchi, T.; Iinuma, H.; Kunimoto, S.; Masuda, T.; Ishizuka, M.; Takeuchi, M.; Hamada, M.; Naganawa, H.; Kondo, S.; Umezawa, H. A new antitumor antibiotic, spergualin: Isolation and antitumor activity. J. Antibiot. 1981, 34, 1619–1621. [Google Scholar] [CrossRef]
  275. Nishikawa, K.; Shibasaki, C.; Takahashi, K.; Nakamura, T.; Takeuchi, T.; Umezawa, H. Antitumor Activity of spergualin, a novel antitumor antibiotic. J. Antibiot. 1986, 39, 1461–1466. [Google Scholar] [CrossRef]
  276. Broughton, L.J.; Crow, C.; Maraveyas, A.; Madden, L.A. Duramycin-induced calcium release in cancer cells. Anticancer Drugs 2016, 27, 173–182. [Google Scholar] [CrossRef]
  277. Yates, K.R.; Welsh, J.; Udegbunam, N.O.; Greenman, J.; Maraveyas, A.; Madden, L.A. Duramycin exhibits antiproliferative properties and induces apoptosis in tumour cells. Blood Coagul. Fibrinolysis 2012, 23, 396–401. [Google Scholar] [CrossRef]
  278. Van Staden, A.D.P.; van Zyl, W.F.; Trindade, M.; Dicks, L.M.T.; Smith, C. Therapeutic application of lantibiotics and other lanthipeptides: Old and new findings. Appl. Environ. Microbiol. 2021, 87, e00186-21. [Google Scholar] [CrossRef]
  279. Thorpe, P.; Soares, M.; He, J. Selected Antibodies and Duramycin Peptides Binding to Anionic Phospholipids and Aminophospholipids and Their Use in the Treatment of Viral Infections and Cancer. European Patent EP2283868B1, 30 March 2016. [Google Scholar]
  280. He, J.; Thorpe, P. Anti-tumor effects of duramycin-IgG conjugate. Cancer Res. 2004, 64, 1053. [Google Scholar]
  281. Ge, J.; Kang, J.; Ping, W. Effect of acetic acid on bacteriocin production by gram-positive bacteria. J. Microbiol. Biotechnol. 2019, 29, 1341–1348. [Google Scholar] [CrossRef] [PubMed]
  282. Alvarez-Sieiro, P.; Montalbán-López, M.; Mu, D.; Kuipers, O.P. Bacteriocins of lactic acid bacteria: Extending the family. Appl. Microbiol. Biotechnol. 2016, 100, 2939–2951. [Google Scholar] [CrossRef] [PubMed]
  283. García, P.; Rodríguez, L.; Rodríguez, A.; Martínez, B. Food biopreservation: Promising strategies using bacteriocins, bacteriophages and endolysins. Trends Food Sci. Technol. 2010, 21, 373–382. [Google Scholar] [CrossRef]
  284. Cornut, G.; Fortin, C.; Soulières, D. Antineoplastic properties of bacteriocins. Am. J. Clin. Oncol. 2008, 31, 399–404. [Google Scholar] [CrossRef] [PubMed]
  285. Cotter, P.D.; Ross, R.P.; Hill, C. Bacteriocins—a viable alternative to antibiotics? Nat. Rev. Microbiol. 2013, 11, 95–105. [Google Scholar] [CrossRef] [PubMed]
  286. Sand, S.L.; Nissen-Meyer, J.; Sand, O.; Haug, T.M. Plantaricin A, a cationic peptide produced by Lactobacillus plantarum, permeabilizes eukaryotic cell membranes by a mechanism dependent on negative surface charge linked to glycosylated membrane proteins. Biochim. Biophys. Acta. Biomembr. 2013, 1828, 249–259. [Google Scholar] [CrossRef]
  287. Zhao, H.; Sood, R.; Jutila, A.; Bose, S.; Fimland, G.; Nissen-Meyer, J.; Kinnunen, P.K.J. Interaction of the antimicrobial peptide pheromone plantaricin A with model membranes: Implications for a novel mechanism of action. Biochim. Biophys. Acta. Biomembr. 2006, 1758, 1461–1474. [Google Scholar] [CrossRef]
  288. Baquero, F.; Lanza, V.F.; Baquero, M.-R.; del Campo, R.; Bravo-Vázquez, D.A. Microcins in Enterobacteriaceae: Peptide antimicrobials in the eco-active intestinal chemosphere. Front. Microbiol. 2019, 10, 2261. [Google Scholar] [CrossRef]
  289. Hetz, C.; Bono, M.R.; Barros, L.F.; Lagos, R. Microcin E492, a channel-forming bacteriocin from Klebsiella pneumoniae, induces apoptosis in some human cell lines. Proc. Natl. Acad. Sci. USA 2002, 99, 2696–2701. [Google Scholar] [CrossRef]
  290. Lagos, R.; Tello, M.; Mercado, G.; Garcia, V.; Monasterio, O. Antibacterial and antitumorigenic properties of microcin E492, a pore-forming bacteriocin. Curr. Pharm. Biotechnol. 2009, 10, 74–85. [Google Scholar] [CrossRef]
  291. Kang, H.K.; Park, J.; Seo, C.H.; Park, Y. PEP27-2, a potent antimicrobial cell-penetrating peptide, reduces skin abscess formation during Staphylococcus aureus infections in mouse when used in combination with antibiotics. ACS Infect. Dis. 2021, 7, 2620–2636. [Google Scholar] [CrossRef] [PubMed]
  292. Lee, D.; Hahm, K.-S.; Park, Y.; Kim, H.-Y.; Lee, W.; Lim, S.-C.; Seo, Y.-K.; Choi, C.-H. Functional and structural characteristics of anticancer peptide Pep27 analogues. Cancer Cell Int. 2005, 5, 21. [Google Scholar] [CrossRef] [PubMed]
  293. Mantovani, H.C.; Hu, H.; Worobo, R.W.; Russell, J.B. Bovicin HC5, a bacteriocin from Streptococcus bovis HC5. Microbiology 2002, 148, 3347–3352. [Google Scholar] [CrossRef]
  294. Paiva, A.D.; de Oliveira, M.D.; de Paula, S.O.; Baracat-Pereira, M.C.; Breukink, E.; Mantovani, H.C. Toxicity of bovicin HC5 against mammalian cell lines and the role of cholesterol in bacteriocin activity. Microbiology 2012, 158, 2851–2858. [Google Scholar] [CrossRef]
  295. Farkas-Himsley, H.; Yu, H. Purified colicin as cytotoxic agent of neoplasia: Comparative study with crude colicin. Cytobios 1985, 42, 193–207. [Google Scholar] [PubMed]
  296. Farkas-Himsley, H.; Hill, R.; Rosen, B.; Arab, S.; Lingwood, C.A. The bacterial colicin active against tumor cells in Vitro and in Vivo is verotoxin 1. Proc. Natl. Acad. Sci. USA 1995, 92, 6996–7000. [Google Scholar] [CrossRef]
  297. Šmarda, J.; Keprtová, J. Cytotoxic effects of colicins E1–E5 and K on hamster fibroblasts. Folia Microbiol. 1987, 32, 133–136. [Google Scholar] [CrossRef]
  298. Chumchalová, J.; Šmarda, J. Human tumor cells are selectively inhibited by colicins. Folia Microbiol. 2003, 48, 111–115. [Google Scholar] [CrossRef]
  299. Smarda, J.; Fialová, M.; Smarda, J. Cytotoxic effects of colicins E1 and E3 on V-Myb-transformed chicken monoblasts. Folia Biol. 2001, 47, 11–13. [Google Scholar]
  300. Lancaster, L.E.; Wintermeyer, W.; Rodnina, M.V. Colicins and their potential in cancer treatment. Blood Cells Mol. Dis. 2007, 38, 15–18. [Google Scholar] [CrossRef]
  301. Michel-Briand, Y.; Baysse, C. The pyocins of Pseudomonas aeruginosa. Biochimie 2002, 84, 499–510. [Google Scholar] [CrossRef]
  302. Baindara, P.; Korpole, S.; Grover, V. Bacteriocins: Perspective for the development of novel anticancer drugs. Appl. Microbiol. Biotechnol. 2018, 102, 10393–10408. [Google Scholar] [CrossRef] [PubMed]
  303. Villarante, K.I.; Elegado, F.B.; Iwatani, S.; Zendo, T.; Sonomoto, K.; de Guzman, E.E. Purification, Characterization and in vitro cytotoxicity of the bacteriocin from Pediococcus acidilactici K2a2-3 against human colon adenocarcinoma (HT29) and human cervical carcinoma (HeLa) cells. World J. Microbiol. Biotechnol. 2011, 27, 975–980. [Google Scholar] [CrossRef]
  304. Kumar, B. In vitro Cytotoxicity of native and rec-pediocin CP2 against cancer cell lines: A comparative study. Pharm. Anal. Acta 2012, 3, 183. [Google Scholar] [CrossRef]
  305. Watanabe, T.; Saito, H. Cytotoxicity of pyocin S2 to tumor and normal cells and its interaction with cell surfaces. Biochim. Biophys. Acta. Gen. Subj. 1980, 633, 77–86. [Google Scholar] [CrossRef]
  306. Abdi-Ali, A.; Worobec, E.A.; Deezagi, A.; Malekzadeh, F. Cytotoxic effects of pyocin S2 produced by Pseudomonas aeruginosa on the growth of three human cell lines. Can. J. Microbiol. 2004, 50, 375–381. [Google Scholar] [CrossRef] [PubMed]
  307. Nguyen, C.; Nguyen, V.D. Discovery of azurin-like anticancer bacteriocins from human gut microbiome through homology modeling and molecular docking against the tumor suppressor P53. Biomed Res. Int. 2016, 2016, 1–12. [Google Scholar] [CrossRef]
  308. Yamada, T.; Fialho, A.M.; Punj, V.; Bratescu, L.; das Gupta, T.K.; Chakrabarty, A.M. Internalization of bacterial redox protein azurin in mammalian cells: Entry domain and specificity. Cell. Microbiol. 2005, 7, 1418–1431. [Google Scholar] [CrossRef]
  309. Yamada, T.; Mehta, R.R.; Lekmine, F.; Christov, K.; King, M.L.; Majumdar, D.; Shilkaitis, A.; Green, A.; Bratescu, L.; Beattie, C.W.; et al. A peptide fragment of azurin induces a P53-mediated cell cycle arrest in human breast cancer cells. Mol. Cancer Ther. 2009, 8, 2947–2958. [Google Scholar] [CrossRef]
  310. Taylor, B.N.; Mehta, R.R.; Yamada, T.; Lekmine, F.; Christov, K.; Chakrabarty, A.M.; Green, A.; Bratescu, L.; Shilkaitis, A.; Beattie, C.W.; et al. Noncationic peptides obtained from azurin preferentially enter cancer cells. Cancer Res. 2009, 69, 537–546. [Google Scholar] [CrossRef]
  311. Yamada, T.; Christov, K.; Shilkaitis, A.; Bratescu, L.; Green, A.; Santini, S.; Bizzarri, A.R.; Cannistraro, S.; Gupta, T.K.D.; Beattie, C.W. P28, a first in class peptide inhibitor of Cop1 binding to P53. Br. J. Cancer 2013, 108, 2495–2504. [Google Scholar] [CrossRef] [PubMed]
  312. Mehta, R.R.; Yamada, T.; Taylor, B.N.; Christov, K.; King, M.L.; Majumdar, D.; Lekmine, F.; Tiruppathi, C.; Shilkaitis, A.; Bratescu, L.; et al. A cell penetrating peptide derived from azurin inhibits angiogenesis and tumor growth by inhibiting phosphorylation of VEGFR-2, FAK and Akt. Angiogenesis 2011, 14, 355–369. [Google Scholar] [CrossRef]
  313. Chaudhari, A.; Mahfouz, M.; Fialho, A.M.; Yamada, T.; Granja, A.T.; Zhu, Y.; Hashimoto, W.; Schlarb-Ridley, B.; Cho, W.; Gupta, T.K.; et al. Cupredoxin−cancer interrelationship: Azurin binding with EphB2, interference in EphB2 tyrosine phosphorylation, and inhibition of cancer growth. Biochemistry 2007, 46, 1799–1810. [Google Scholar] [CrossRef] [PubMed]
  314. Gao, M.; Zhou, J.; Su, Z.; Huang, Y. Bacterial cupredoxin azurin hijacks cellular signaling networks: Protein-protein interactions and cancer therapy. Protein Sci. 2017, 26, 2334–2341. [Google Scholar] [CrossRef] [PubMed]
  315. Lulla, R.R.; Goldman, S.; Yamada, T.; Beattie, C.W.; Bressler, L.; Pacini, M.; Pollack, I.F.; Fisher, P.G.; Packer, R.J.; Dunkel, I.J.; et al. Phase I trial of P28 (NSC745104), a non-HDM2-mediated peptide inhibitor of P53 ubiquitination in pediatric patients with recurrent or progressive central nervous system tumors: A pediatric brain tumor consortium study. Neuro-oncology 2016, 18, 1319–1325. [Google Scholar] [CrossRef] [PubMed]
  316. Hasegawa, H.; Yamada, Y.; Iha, H.; Tsukasaki, K.; Nagai, K.; Atogami, S.; Sugahara, K.; Tsuruda, K.; Ishizaki, A.; Kamihira, S. Activation of P53 by nutlin-3a, an antagonist of MDM2, induces apoptosis and cellular senescence in adult T-Cell leukemia cells. Leukemia 2009, 23, 2090–2101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Yaghoubi, A.; Khazaei, M.; Avan, A.; Hasanian, S.M.; Cho, W.C.; Soleimanpour, S. P28 bacterial peptide, as an anticancer agent. Front. Oncol. 2020, 10, 1303. [Google Scholar] [CrossRef]
  318. Sand, S.L.; Oppegård, C.; Ohara, S.; Iijima, T.; Naderi, S.; Blomhoff, H.K.; Nissen-Meyer, J.; Sand, O. Plantaricin A, a peptide pheromone produced by Lactobacillus plantarum, permeabilizes the cell membrane of both normal and cancerous lymphocytes and neuronal cells. Peptides 2010, 31, 1237–1244. [Google Scholar] [CrossRef]
  319. Kim, T.-S.; Hur, J.-W.; Yu, M.-A.; Cheigh, C.-I.; Kim, K.-N.; Hwang, J.-K.; Pyun, Y.-R. Antagonism of Helicobacter pylori by bacteriocins of lactic acid bacteria. J. Food Prot. 2003, 66, 3–12. [Google Scholar] [CrossRef]
  320. Joo, N.E.; Ritchie, K.; Kamarajan, P.; Miao, D.; Kapila, Y.L. Nisin, an apoptogenic bacteriocin and food preservative, attenuates HNSCC tumorigenesis via CHAC1. Cancer Med. 2012, 1, 295–305. [Google Scholar] [CrossRef]
  321. Kamarajan, P.; Hayami, T.; Matte, B.; Liu, Y.; Danciu, T.; Ramamoorthy, A.; Worden, F.; Kapila, S.; Kapila, Y. Nisin ZP, a bacteriocin and food Preservative, inhibits head and neck cancer tumorigenesis and prolongs survival. PLoS ONE 2015, 10, e0131008. [Google Scholar] [CrossRef] [PubMed]
  322. Ahmadi, S.; Ghollasi, M.; Hosseini, H.M. The apoptotic impact of nisin as a potent bacteriocin on the colon cancer cells. Microb. Pathog. 2017, 111, 193–197. [Google Scholar] [CrossRef] [PubMed]
  323. Norouzi, Z.; Salimi, A.; Halabian, R.; Fahimi, H. Nisin, a potent bacteriocin and anti-bacterial peptide, attenuates expression of metastatic genes in colorectal cancer cell lines. Microb. Pathog. 2018, 123, 183–189. [Google Scholar] [CrossRef] [PubMed]
  324. Jaye, K.; Li, C.G.; Chang, D.; Bhuyan, D.J. The role of key gut microbial metabolites in the development and treatment of cancer. Gut Microbes 2022, 14, 2038865. [Google Scholar] [CrossRef]
  325. Preet, S.; Bharati, S.; Panjeta, A.; Tewari, R.; Rishi, P. Effect of nisin and doxorubicin on DMBA-induced skin carcinogenesis—a possible adjunct therapy. Tumor Biol. 2015, 36, 8301–8308. [Google Scholar] [CrossRef] [PubMed]
  326. Rana, K.; Sharma, R.; Preet, S. Augmented therapeutic efficacy of 5-fluorouracil in conjunction with lantibiotic nisin against skin cancer. Biochem. Biophys. Res. Commun. 2019, 520, 551–559. [Google Scholar] [CrossRef]
  327. Nikolouli, K.; Mossialos, D. Bioactive compounds synthesized by non-ribosomal peptide synthetases and type-I polyketide synthases discovered through genome-mining and metagenomics. Biotechnol. Lett. 2012, 34, 1393–1403. [Google Scholar] [CrossRef]
  328. Gevers, W.; Kleinkauf, H.; Lipmann, F. The activation of amino acids for biosynthesis of gramicidin S. Proc. Natl. Acad. Sci. USA 1968, 60, 269–276. [Google Scholar] [CrossRef]
  329. Sieber, S.A.; Marahiel, M.A. Learning from nature’s drug factories: Nonribosomal synthesis of macrocyclic peptides. J. Bacteriol. 2003, 185, 7036–7043. [Google Scholar] [CrossRef]
  330. Dutta, S.; Whicher, J.R.; Hansen, D.A.; Hale, W.A.; Chemler, J.A.; Congdon, G.R.; Narayan, A.R.H.; Håkansson, K.; Sherman, D.H.; Smith, J.L.; et al. Structure of a modular polyketide synthase. Nature 2014, 510, 512–517. [Google Scholar] [CrossRef]
  331. Abotaleb, M.; Liskova, A.; Kubatka, P.; Büsselberg, D. Therapeutic potential of plant phenolic acids in the treatment of cancer. Biomolecules 2020, 10, 221. [Google Scholar] [CrossRef] [PubMed]
  332. Russell, W.R.; Duncan, S.H.; Scobbie, L.; Duncan, G.; Cantlay, L.; Calder, A.G.; Anderson, S.E.; Flint, H.J. Major phenylpropanoid-derived metabolites in the human gut can arise from microbial fermentation of protein. Mol. Nutr. Food Res. 2013, 57, 523–535. [Google Scholar] [CrossRef] [PubMed]
  333. Viljoen, K.S.; Dakshinamurthy, A.; Goldberg, P.; Blackburn, J.M. Quantitative profiling of colorectal cancer-associated bacteria reveals associations between Fusobacterium spp., enterotoxigenic Bacteroides fragilis (ETBF) and clinicopathological features of colorectal cancer. PLoS ONE 2015, 10, e0119462. [Google Scholar] [CrossRef] [PubMed]
  334. Rossi, R.; Corino, C.; Modina, S.; di Giancamillo, A. Dietary Verbascoside influences gut morphology and the expression of α-transducin and α-gustducin in the small intestine of weaned piglets exposed to n-6 polyunsaturated fatty acids-induced oxidative stress. Animals 2019, 9, 20. [Google Scholar] [CrossRef] [PubMed]
  335. Zhang, F.; Jia, Z.; Deng, Z.; Wei, Y.; Zheng, R.; Yu, L. In vitro modulation of telomerase activity, telomere length and cell cycle in MKN45 cells by verbascoside. Planta Med. 2002, 68, 115–118. [Google Scholar] [CrossRef]
  336. Díaz, A.M.; Abad, M.J.; Fernández, L.; Silván, A.M.; de Santos, J.; Bermejo, P. Phenylpropanoid glycosides from Scrophularia scorodonia: In vitro anti-inflammatory activity. Life Sci. 2004, 74, 2515–2526. [Google Scholar] [CrossRef]
  337. Yahfoufi, N.; Alsadi, N.; Jambi, M.; Matar, C. The immunomodulatory and anti-inflammatory role of polyphenols. Nutrients 2018, 10, 1618. [Google Scholar] [CrossRef]
  338. Cui, Q.; Pan, Y.; Xu, X.; Zhang, W.; Wu, X.; Qu, S.; Liu, X. The metabolic profile of acteoside produced by human or rat intestinal bacteria or intestinal enzyme in vitro employed UPLC-Q-TOF–MS. Fitoterapia 2016, 109, 67–74. [Google Scholar] [CrossRef]
  339. Logan, I.; Miranda, C.; Lowry, M.; Maier, C.; Stevens, J.; Gombart, A. Antiproliferative and cytotoxic activity of xanthohumol and its non-estrogenic derivatives in colon and hepatocellular carcinoma cell lines. Int. J. Mol. Sci. 2019, 20, 1203. [Google Scholar] [CrossRef]
  340. Venturelli, S.; Niessner, H.; Sinnberg, T.; Berger, A.; Burkard, M.; Urmann, C.; Donaubauer, K.; Böcker, A.; Leischner, C.; Riepl, H.; et al. 6- and 8-Prenylnaringenin, novel natural histone deacetylase inhibitors found in hops, exert antitumor activity on melanoma cells. Cell. Physiol. Biochem. 2018, 51, 543–556. [Google Scholar] [CrossRef]
  341. Brunelli, E.; Minassi, A.; Appendino, G.; Moro, L. 8-Prenylnaringenin, inhibits estrogen receptor-α mediated cell growth and induces apoptosis in MCF-7 breast cancer cells. J. Steroid Biochem. Mol. Biol. 2007, 107, 140–148. [Google Scholar] [CrossRef] [PubMed]
  342. Allsopp, P.; Possemiers, S.; Campbell, D.; Gill, C.; Rowland, I. A comparison of the anticancer properties of isoxanthohumol and 8-prenylnaringenin using in vitro models of colon cancer. Biofactors 2013, 39, 441–447. [Google Scholar] [CrossRef] [PubMed]
  343. Koosha, S.; Mohamed, Z.; Sinniah, A.; Ibrahim, Z.A.; Seyedan, A.; Alshawsh, M.A. Antiproliferative and apoptotic activities of 8-prenylnaringenin against human colon cancer cells. Life Sci. 2019, 232, 116633. [Google Scholar] [CrossRef] [PubMed]
  344. Welihinda, A.A.; Kaur, M.; Greene, K.; Zhai, Y.; Amento, E.P. The adenosine metabolite inosine is a functional agonist of the adenosine A2A receptor with a unique signaling bias. Cell. Signal. 2016, 28, 552–560. [Google Scholar] [CrossRef]
  345. Mager, L.F.; Burkhard, R.; Pett, N.; Cooke, N.C.A.; Brown, K.; Ramay, H.; Paik, S.; Stagg, J.; Groves, R.A.; Gallo, M.; et al. Microbiome-derived inosine modulates response to checkpoint inhibitor immunotherapy. Science 2020, 369, 1481–1489. [Google Scholar] [CrossRef]
  346. McNabney, S.; Henagan, T. Short chain fatty acids in the colon and peripheral tissues: A focus on butyrate, colon cancer, obesity and insulin resistance. Nutrients 2017, 9, 1348. [Google Scholar] [CrossRef] [Green Version]
  347. Parada Venegas, D.; de la Fuente, M.K.; Landskron, G.; González, M.J.; Quera, R.; Dijkstra, G.; Harmsen, H.J.M.; Faber, K.N.; Hermoso, M.A. Short chain fatty acids (SCFAs)-mediated gut epithelial and immune regulation and its relevance for inflammatory bowel diseases. Front. Immunol. 2019, 10, 277. [Google Scholar] [CrossRef]
  348. Cushing, K.; Alvarado, D.M.; Ciorba, M.A. Butyrate and mucosal inflammation: New scientific evidence supports clinical observation. Clin. Transl. Gastroenterol. 2015, 6, e108. [Google Scholar] [CrossRef]
  349. Furusawa, Y.; Obata, Y.; Fukuda, S.; Endo, T.A.; Nakato, G.; Takahashi, D.; Nakanishi, Y.; Uetake, C.; Kato, K.; Kato, T.; et al. Commensal microbe-derived butyrate induces the differentiation of colonic regulatory T cells. Nature 2013, 504, 446–450. [Google Scholar] [CrossRef]
  350. Yusuf, F.; Adewiah, S.; Syam, A.F.; Fatchiyah, F. Altered profile of gut microbiota and the level short chain fatty acids in colorectal cancer patients. J. Phys.: Conf. Ser. 2019, 1146, 012037. [Google Scholar] [CrossRef]
  351. Wang, H.-B.; Wang, P.-Y.; Wang, X.; Wan, Y.-L.; Liu, Y.-C. Butyrate enhances intestinal epithelial barrier function via up-regulation of tight junction protein claudin-1 transcription. Dig. Dis. Sci. 2012, 57, 3126–3135. [Google Scholar] [CrossRef]
  352. De Vadder, F.; Kovatcheva-Datchary, P.; Goncalves, D.; Vinera, J.; Zitoun, C.; Duchampt, A.; Bäckhed, F.; Mithieux, G. Microbiota-generated metabolites promote metabolic benefits via gut-brain neural circuits. Cell 2014, 156, 84–96. [Google Scholar] [CrossRef] [PubMed]
  353. Liu, J.; Fu, Y.; Zhang, H.; Wang, J.; Zhu, J.; Wang, Y.; Guo, Y.; Wang, G.; Xu, T.; Chu, M.; et al. The hepatoprotective effect of the probiotic Clostridium butyricum against carbon tetrachloride-induced acute liver damage in mice. Food Funct. 2017, 8, 4042–4052. [Google Scholar] [CrossRef] [PubMed]
  354. Jin, U.-H.; Cheng, Y.; Park, H.; Davidson, L.A.; Callaway, E.S.; Chapkin, R.S.; Jayaraman, A.; Asante, A.; Allred, C.; Weaver, E.A.; et al. Short chain fatty acids enhance aryl hydrocarbon (Ah) responsiveness in mouse colonocytes and caco-2 human colon cancer cells. Sci. Rep. 2017, 7, 10163. [Google Scholar] [CrossRef] [PubMed]
  355. Zhang, Q.; Yang, F.; Li, X.; Zhang, H.-Y.; Chu, X.-G.; Zhang, H.; Wang, L.-W.; Gong, Z.-J. Trichostatin A protects against intestinal injury in rats with acute liver failure. J. Surg. Res. 2016, 205, 1–10. [Google Scholar] [CrossRef]
  356. Xu, W.S.; Parmigiani, R.B.; Marks, P.A. Histone deacetylase inhibitors: Molecular mechanisms of action. Oncogene 2007, 26, 5541–5552. [Google Scholar] [CrossRef] [Green Version]
  357. Roger, T.; Lugrin, J.; le Roy, D.; Goy, G.; Mombelli, M.; Koessler, T.; Ding, X.C.; Chanson, A.-L.; Reymond, M.K.; Miconnet, I.; et al. Histone deacetylase inhibitors impair innate immune responses to toll-like receptor agonists and to infection. Blood 2011, 117, 1205–1217. [Google Scholar] [CrossRef]
  358. Nøhr, M.K.; Pedersen, M.H.; Gille, A.; Egerod, K.L.; Engelstoft, M.S.; Husted, A.S.; Sichlau, R.M.; Grunddal, K.V.; Seier Poulsen, S.; Han, S.; et al. GPR41/FFAR3 and GPR43/FFAR2 as cosensors for short-chain fatty acids in enteroendocrine cells vs FFAR3 in enteric neurons and FFAR2 in enteric leukocytes. Endocrinology 2013, 154, 3552–3564. [Google Scholar] [CrossRef]
  359. Feng, Y.; Wang, Y.; Wang, P.; Huang, Y.; Wang, F. Short-chain fatty acids manifest stimulative and protective effects on intestinal barrier function through the inhibition of NLRP3 inflammasome and autophagy. Cell. Physiol. Biochem. 2018, 49, 190–205. [Google Scholar] [CrossRef]
  360. Priyadarshini, M.; Kotlo, K.U.; Dudeja, P.K.; Layden, B.T. Role of short chain fatty acid receptors in intestinal physiology and pathophysiology. Compr. Physiol. 2018, 8, 1091–1115. [Google Scholar]
  361. Saxena, M.; Yeretssian, G. NOD-Like Receptors: Master regulators of inflammation and cancer. Front. Immunol. 2014, 5, 327. [Google Scholar] [CrossRef] [PubMed]
  362. Gamet, L.; Daviaud, D.; Denis-Pouxviel, C.; Remesy, C.; Murat, J.-C. Effects of short-chain fatty acids on growth and differentiation of the human colon-cancer cell Line HT29. Int. J. Cancer Res. 1992, 52, 286–289. [Google Scholar] [CrossRef] [PubMed]
  363. Hinnebusch, B.F.; Meng, S.; Wu, J.T.; Archer, S.Y.; Hodin, R.A. The effects of short-chain fatty acids on human colon cancer cell phenotype are associated with histone hyperacetylation. J. Nutr. 2002, 132, 1012–1017. [Google Scholar] [CrossRef] [PubMed]
  364. Luu, M.; Visekruna, A. Short-chain Fatty Acids: Bacterial messengers modulating the immunometabolism of T cells. Eur. J. Immunol. 2019, 49, 842–848. [Google Scholar] [CrossRef] [PubMed]
  365. Davie, J.R. Inhibition of histone deacetylase activity by butyrate. J. Nutr. 2003, 133, 2485S–2493S. [Google Scholar] [CrossRef] [PubMed]
  366. Tang, Y.; Chen, Y.; Jiang, H.; Robbins, G.T.; Nie, D. G-protein-coupled receptor for short-chain fatty acids suppresses colon cancer. Int. J. Cancer Res. 2011, 128, 847–856. [Google Scholar] [CrossRef] [PubMed]
  367. Shamloo, B.; Usluer, S. P21 in Cancer Research. Cancers 2019, 11, 1178. [Google Scholar] [CrossRef] [Green Version]
  368. Lightfoot, Y.L.; Mohamadzadeh, M. Tailoring gut immune responses with lipoteichoic acid-deficient Lactobacillus acidophilus. Front. Immunol. 2013, 4, 25. [Google Scholar] [CrossRef]
  369. Lightfoot, Y.L.; Yang, T.; Sahay, B.; Mohamadzadeh, M. Targeting aberrant colon cancer-specific DNA methylation with lipoteichoic acid-deficient Lactobacillus acidophilus. Gut Microbes 2013, 4, 84–88. [Google Scholar] [CrossRef]
  370. Xiang, L.; Mou, J.; Shao, B.; Wei, Y.; Liang, H.; Takano, N.; Semenza, G.L.; Xie, G. Glutaminase 1 expression in colorectal cancer cells is induced by hypoxia and required for tumor growth, invasion, and metastatic colonization. Cell Death Dis. 2019, 10, 40. [Google Scholar] [CrossRef]
  371. Sharma, B.; Singh, S.; Kanwar, S.S. L-methionase: A therapeutic enzyme to treat malignancies. BioMed. Res. Int. 2014, 2014, 506287. [Google Scholar] [CrossRef] [PubMed]
  372. Ni, Y.; Schwaneberg, U.; Sun, Z.-H. Arginine deiminase, a potential anti-tumor drug. Cancer Lett. 2008, 261, 1–11. [Google Scholar] [CrossRef] [PubMed]
  373. Holtsberg, F.W.; Ensor, C.M.; Steiner, M.R.; Bomalaski, J.S.; Clark, M.A. Poly(ethylene glycol) (PEG) conjugated arginine deiminase: Effects of PEG formulations on its pharmacological properties. J. Control. Release 2002, 80, 259–271. [Google Scholar] [CrossRef]
  374. Kim, R.H.; Coates, J.M.; Bowles, T.L.; McNerney, G.P.; Sutcliffe, J.; Jung, J.U.; Gandour-Edwards, R.; Chuang, F.Y.S.; Bold, R.J.; Kung, H.-J. Arginine deiminase as a novel therapy for prostate cancer Induces autophagy and caspase-independent apoptosis. Cancer Res. 2009, 69, 700–708. [Google Scholar] [CrossRef]
  375. Cheng, P.N.; Lo, T.W.H.; Leung, T.W.H.; Lam, A.T.L.; Lam, W.-M. Pegylated recombinant human arginase (RhArg-Peg 5000mw) has in vitro and in vivo anti-proliferative potential and apoptotic activities in human hepatocellular varcinoma (HCC). J. Clin. Oncol. 2005, 23, 3179. [Google Scholar] [CrossRef]
  376. Glazer, E.S.; Piccirillo, M.; Albino, V.; di Giacomo, R.; Palaia, R.; Mastro, A.A.; Beneduce, G.; Castello, G.; de Rosa, V.; Petrillo, A.; et al. Phase II study of pegylated arginine deiminase for nonresectable and metastatic hepatocellular carcinoma. J. Clin. Oncol. 2010, 28, 2220–2226. [Google Scholar] [CrossRef] [PubMed]
  377. Hak, L.J.; Relling, M.V.; Cheng, C.; Pei, D.; Wang, B.; Sandlund, J.T.; Rubnitz, J.; Pui, C.-H. Asparaginase pharmacodynamics differ by formulation among children with newly diagnosed acute lymphoblastic leukemia. Leukemia 2004, 18, 1072–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  378. Avramis, V.I.; Sencer, S.; Periclou, A.P.; Sather, H.; Bostrom, B.C.; Cohen, L.J.; Ettinger, A.G.; Ettinger, L.J.; Franklin, J.; Gaynon, P.S.; et al. A randomized comparison of native Escherichia coli asparaginase and polyethylene glycol conjugated asparaginase for treatment of children with newly diagnosed standard-risk acute lymphoblastic leukemia: A children’s cancer group study. Blood 2002, 99, 1986–1994. [Google Scholar] [CrossRef]
  379. Panosyan, E.H.; Wang, Y.; Xia, P.; Lee, W.-N.P.; Pak, Y.; Laks, D.R.; Lin, H.J.; Moore, T.B.; Cloughesy, T.F.; Kornblum, H.I.; et al. Asparagine depletion potentiates the cytotoxic effect of chemotherapy against brain tumors. Mol. Cancer Res. 2014, 12, 694–702. [Google Scholar] [CrossRef]
  380. Jaccard, A.; Gachard, N.; Marin, B.; Rogez, S.; Audrain, M.; Suarez, F.; Tilly, H.; Morschhauser, F.; Thieblemont, C.; Ysebaert, L.; et al. Efficacy of L-asparaginase with methotrexate and dexamethasone (AspaMetDex regimen) in patients with refractory or relapsing extranodal NK/T-cell lymphoma, a phase 2 study. Blood 2011, 117, 1834–1839. [Google Scholar] [CrossRef]
  381. Covini, D.; Tardito, S.; Bussolati, O.R.; Chiarelli, L.; Pasquetto, M.V.; Digilio, R.; Valentini, G.; Scotti, C. Expanding targets for a metabolic therapy of cancer: L-asparaginase. Recent Pat. Anticancer Drug Discov. 2012, 7, 4–13. [Google Scholar] [CrossRef] [PubMed]
  382. Yu, M.; Henning, R.; Walker, A.; Kim, G.; Perroy, A.; Alessandro, R.; Virador, V.; Kohn, E.C. L-Asparaginase inhibits invasive and angiogenic activity and induces autophagy in ovarian cancer. J. Cell. Mol. Med. 2012, 16, 2369–2378. [Google Scholar] [CrossRef] [PubMed]
  383. Shrivastava, A.; Khan, A.A.; Khurshid, M.; Kalam, M.A.; Jain, S.K.; Singhal, P.K. Recent developments in L-asparaginase discovery and its potential as anticancer agent. Crit. Rev. Oncol. Hematol. 2016, 100, 1–10. [Google Scholar] [CrossRef] [PubMed]
  384. Aledo, J.C.; Gomez-Fabre, P.M.; Olalla, L.; Marquez, J. Identification of two human glutaminase loci and tissue-specific expression of the two related genes. Mamm. Genome. 2000, 11, 1107–1110. [Google Scholar] [CrossRef]
  385. Wang, J.-B.; Erickson, J.W.; Fuji, R.; Ramachandran, S.; Gao, P.; Dinavahi, R.; Wilson, K.F.; Ambrosio, A.L.; Dias, S.M.; Dang, C.V.; et al. Targeting mitochondrial glutaminase activity inhibits oncogenic transformation. Cancer Cell. 2010, 18, 207–219. [Google Scholar] [CrossRef] [PubMed]
  386. Nicklin, P.; Bergman, P.; Zhang, B.; Triantafellow, E.; Wang, H.; Nyfeler, B.; Yang, H.; Hild, M.; Kung, C.; Wilson, C.; et al. Bidirectional transport of amino acids regulates mTOR and autophagy. Cell 2009, 136, 521–534. [Google Scholar] [CrossRef]
  387. Fuchs, B.C.; Finger, R.E.; Onan, M.C.; Bode, B.P. ASCT2 silencing regulates mammalian target-of-rapamycin growth and survival signaling in human hepatoma cells. Am. J. Physiol. Cell. Physiol. 2007, 293, C55–C63. [Google Scholar] [CrossRef]
  388. Xiang, L.; Xie, G.; Liu, C.; Zhou, J.; Chen, J.; Yu, S.; Li, J.; Pang, X.; Shi, H.; Liang, H. Knock-down of glutaminase 2 expression decreases glutathione, NADH, and sensitizes cervical cancer to ionizing radiation. Biochim. Biophys. Acta. 2013, 12, 2996–3005. [Google Scholar] [CrossRef]
  389. Harrison, R.G.; Norman, O.K. Enzyme Prodrug Cancer Therapy Selectively Targeted to Tumor Cells or Tumor Vasculature and Methods of Production and Use Thereof. US Patent 2011/0002978 A1, 6 January 2011. Available online: https://patentimages.storage.googleapis.com/4e/03/1d/bb6f1b96381c9a/US20110002978A1.pdf (accessed on 9 June 2022).
  390. Van Rite, B.D.; Krais, J.J.; Cherry, M.; Kakavetsis, V.I.; Kurkjian, C.; Harrison, R.G. Antitumor activity of an enzyme prodrug therapy targeted to the breast tumor vasculature. Cancer Investig. 2013, 31, 505–510. [Google Scholar] [CrossRef]
  391. Van Rite, B.D.; Lazrak, Y.A.; Pagnon, M.L.; Palwai, N.R.; Neves, L.F.; McFetridge, P.S.; Harrison, R.G. Enzyme prodrug therapy designed to target L-methioninase to the tumor vasculature. Cancer Lett. 2011, 301, 177–184. [Google Scholar] [CrossRef]
Figure 1. Mode of action of diphtheria toxin (DT). (A) The receptor (R domain) of DT, shown here as green speres, binds to the membrane-anchored precursor of heparin-binding epidermal-like growth factor (proHB-EGF). (B) DT is nicked and the DT-HB-EGF enters a clathrin-coated vesicle through endocytosis. Furin or furin-like proteases converts DT to mature form. (C) An early endosomal vesicle (EEV) is formed by replacing clathrin proteins with the GTPase Arf-1 and coat protein COPI (not shown). (D) EEV is acidified by the transport of protons (H+) across the membrane, instigated by vacuolar adenosine triphosphatase (vATPase). (E) The T-domain is translocated across the membrane, exposing the C-domain to the cytosol. (F) The disulphide bridge is reduced to liberate the catalytic C-domain. (G) The free C-domain catalyses the ADP-ribosylation of eukaryotic elongation factor 2 (eEF2) to ADPR-eEF2, which inhibits translation. This illustration was made using BioRender (https://biorender.com/, accessed on 12 May 2022).
Figure 1. Mode of action of diphtheria toxin (DT). (A) The receptor (R domain) of DT, shown here as green speres, binds to the membrane-anchored precursor of heparin-binding epidermal-like growth factor (proHB-EGF). (B) DT is nicked and the DT-HB-EGF enters a clathrin-coated vesicle through endocytosis. Furin or furin-like proteases converts DT to mature form. (C) An early endosomal vesicle (EEV) is formed by replacing clathrin proteins with the GTPase Arf-1 and coat protein COPI (not shown). (D) EEV is acidified by the transport of protons (H+) across the membrane, instigated by vacuolar adenosine triphosphatase (vATPase). (E) The T-domain is translocated across the membrane, exposing the C-domain to the cytosol. (F) The disulphide bridge is reduced to liberate the catalytic C-domain. (G) The free C-domain catalyses the ADP-ribosylation of eukaryotic elongation factor 2 (eEF2) to ADPR-eEF2, which inhibits translation. This illustration was made using BioRender (https://biorender.com/, accessed on 12 May 2022).
Microorganisms 10 01733 g001
Figure 2. Clostridium perfringens enterotoxin (CPE) mode of action. (A) Tyrosine residues in the C-terminal of CPE interact with the second extracellular loop (ECL-2) of a receptor claudin (CLDN) and form small complexes. (B) Approximately six small complexes oligomerize to form a larger CPE hexamer 1 (CH-1; prepore). (C) β-hairpin loops of CPE assemble to form a β-barrel that inserts into the cell membrane to create a cation-permeating pore. (D) Influx of Ca2+ ions disrupts the osmotic equilibrium and activates Ca2-dependent proteases to lyse the cell that forms calpain. (E) Release of cytochrome C from the mitochondrion. (F) Activation of caspase 3/7 and formation of a large new CH-2 complex (approximately 600 kDa, consisting of CLDNs, occludins and the CPE hexamer. (G and H) Apoptosis, leading to DNA fragmentation. This illustration was constructed using BioRender (https://biorender.com/, accessed on 12 May 2022).
Figure 2. Clostridium perfringens enterotoxin (CPE) mode of action. (A) Tyrosine residues in the C-terminal of CPE interact with the second extracellular loop (ECL-2) of a receptor claudin (CLDN) and form small complexes. (B) Approximately six small complexes oligomerize to form a larger CPE hexamer 1 (CH-1; prepore). (C) β-hairpin loops of CPE assemble to form a β-barrel that inserts into the cell membrane to create a cation-permeating pore. (D) Influx of Ca2+ ions disrupts the osmotic equilibrium and activates Ca2-dependent proteases to lyse the cell that forms calpain. (E) Release of cytochrome C from the mitochondrion. (F) Activation of caspase 3/7 and formation of a large new CH-2 complex (approximately 600 kDa, consisting of CLDNs, occludins and the CPE hexamer. (G and H) Apoptosis, leading to DNA fragmentation. This illustration was constructed using BioRender (https://biorender.com/, accessed on 12 May 2022).
Microorganisms 10 01733 g002
Figure 3. Botulinum neurotoxin (BoNT) mode of action. (A) BoNT binds to ecto-acceptors (polysialogangliosides) on the presynaptic cell surface of cholinergic neurons and is internalized via SV-2 or Syt-mediated endocytosis. (B) The synaptic vesicle is acidified with vesicular proton pumps, which in turn (C) activates Ach transporter proteins. The activated transporter proteins import acetylcholine (Ach) and the light chain (LC) of BoNT is translocated to the cytosol (D) with the heavy chain (HC) domain. (E) Heat shock protein 90 (Hsp90) and thioredoxin reductase-thioredoxin (TrxR-Trx) cleaves the LC and (F) liberates it into the cytosol. (G) The C-domain is a Zn2+-dependent endopeptidase that cleaves proteins in the soluble N-ethylmaleimide-sensitive adaptor receptor (SNARE) protein complex. This complex is responsible for exocytosis and the fusing of acetylcholine-containing vesicles with the plasma membrane, allowing the release of acetylcholine. The cleaved SNARE component is non-functional, thereby blocking the release of acetylcholine from the presynaptic membrane to muscles. Blocking exocytosis of acetylcholine leads to failed skeletal muscle contracture. This representation was constructed using BioRender (https://biorender.com/, accessed on 12 May 2022).
Figure 3. Botulinum neurotoxin (BoNT) mode of action. (A) BoNT binds to ecto-acceptors (polysialogangliosides) on the presynaptic cell surface of cholinergic neurons and is internalized via SV-2 or Syt-mediated endocytosis. (B) The synaptic vesicle is acidified with vesicular proton pumps, which in turn (C) activates Ach transporter proteins. The activated transporter proteins import acetylcholine (Ach) and the light chain (LC) of BoNT is translocated to the cytosol (D) with the heavy chain (HC) domain. (E) Heat shock protein 90 (Hsp90) and thioredoxin reductase-thioredoxin (TrxR-Trx) cleaves the LC and (F) liberates it into the cytosol. (G) The C-domain is a Zn2+-dependent endopeptidase that cleaves proteins in the soluble N-ethylmaleimide-sensitive adaptor receptor (SNARE) protein complex. This complex is responsible for exocytosis and the fusing of acetylcholine-containing vesicles with the plasma membrane, allowing the release of acetylcholine. The cleaved SNARE component is non-functional, thereby blocking the release of acetylcholine from the presynaptic membrane to muscles. Blocking exocytosis of acetylcholine leads to failed skeletal muscle contracture. This representation was constructed using BioRender (https://biorender.com/, accessed on 12 May 2022).
Microorganisms 10 01733 g003
Figure 4. Top of presentation: The four domains (Ia, II, Ib and III) of Pseudomonas exotoxin (PE), with the sizes of each domain indicated in amino acid numbers. Interaction of PE-IT (immunotoxin) with a cancer-specific antigen (CSA) or cancer-specific receptor (CSR) on the surface of cancer cells. Intracellular events leading to cell death is illustrated below the phospholipid membrane. Abbreviations: PE-L = Pseudomonas exotoxin A, fused to a cancer-specific ligand; Ab = antibody; EF2 = eukaryotic elongation factor-2 on ribosomes; Mcl1 = gene encoding anti-apoptotic protein; BAK = Bak protein involved in mitochondrion outer membrane (MOM) permeabilization; BCL-xL = B-cell lymphoma-extra large that inhibits the activation of Bak, thereby preventing a loss of MOM integrity. This illustration was constructed using BioRender (https://biorender.com/, accessed on 15 August 2022).
Figure 4. Top of presentation: The four domains (Ia, II, Ib and III) of Pseudomonas exotoxin (PE), with the sizes of each domain indicated in amino acid numbers. Interaction of PE-IT (immunotoxin) with a cancer-specific antigen (CSA) or cancer-specific receptor (CSR) on the surface of cancer cells. Intracellular events leading to cell death is illustrated below the phospholipid membrane. Abbreviations: PE-L = Pseudomonas exotoxin A, fused to a cancer-specific ligand; Ab = antibody; EF2 = eukaryotic elongation factor-2 on ribosomes; Mcl1 = gene encoding anti-apoptotic protein; BAK = Bak protein involved in mitochondrion outer membrane (MOM) permeabilization; BCL-xL = B-cell lymphoma-extra large that inhibits the activation of Bak, thereby preventing a loss of MOM integrity. This illustration was constructed using BioRender (https://biorender.com/, accessed on 15 August 2022).
Microorganisms 10 01733 g004
Figure 5. Azurin (128 amino acids), with an extended α-helix protein transduction domain, p28 (Leu50–Asp77), and four loop regions in its C-terminal (shown here in red, yellow, light blue and dark blue). The p28 peptide of 28 amino acids (amino acids 50 to 77) facilitates cell crossing and promotes apoptosis. Cell growth in tumours is repressed by 10 to 12 amino acids in the COOH terminal of p28. The α-helix, stretching over 18 amino acids (Leu50–Gly67), shown here as peptide p18, has a high affinity for cancer cells (less so for normal cells), excellent penetration abilities, and high binding to the tumour repressor protein p53. Peptide p18b also contains 18 amino acids (Val60–Asp77) but has a short α-helix and β-sheet and penetrates cancerous and normal cells. The p12 peptide of 12 amino acids (Gly66–Asp77) does not have an α-helical structure, binds poorly to p35, and penetrates cancer and normal cells. Adapted from Yaghoubi et al. [317].
Figure 5. Azurin (128 amino acids), with an extended α-helix protein transduction domain, p28 (Leu50–Asp77), and four loop regions in its C-terminal (shown here in red, yellow, light blue and dark blue). The p28 peptide of 28 amino acids (amino acids 50 to 77) facilitates cell crossing and promotes apoptosis. Cell growth in tumours is repressed by 10 to 12 amino acids in the COOH terminal of p28. The α-helix, stretching over 18 amino acids (Leu50–Gly67), shown here as peptide p18, has a high affinity for cancer cells (less so for normal cells), excellent penetration abilities, and high binding to the tumour repressor protein p53. Peptide p18b also contains 18 amino acids (Val60–Asp77) but has a short α-helix and β-sheet and penetrates cancerous and normal cells. The p12 peptide of 12 amino acids (Gly66–Asp77) does not have an α-helical structure, binds poorly to p35, and penetrates cancer and normal cells. Adapted from Yaghoubi et al. [317].
Microorganisms 10 01733 g005
Figure 6. Mode of action of Azurin. Once inside a cancer cell, azurin attaches to the DNA-binding domain of the tumour-suppressor protein p53 (middle of presentation) and prevents binding of the latter to E3 ubiquitin ligase COP1, resulting in an increase in cytoplasmic p53 levels. Cell growth is repressed, and cells are destroyed by apoptosis. Azurin also interferes with non-receptor tyrosine kinase (NRTK) signalling pathways. Bax protein (a central cell death regulator) and cytochrome C levels increase, and caspases 9 and 7 are activated. VEGFR-2 tyrosine kinase activity is reduced, preventing the formation of new blood vessels and the expression of P-cadherin. Prepared using BioRender (https://biorender.com/, accessed on 12 May 2022).
Figure 6. Mode of action of Azurin. Once inside a cancer cell, azurin attaches to the DNA-binding domain of the tumour-suppressor protein p53 (middle of presentation) and prevents binding of the latter to E3 ubiquitin ligase COP1, resulting in an increase in cytoplasmic p53 levels. Cell growth is repressed, and cells are destroyed by apoptosis. Azurin also interferes with non-receptor tyrosine kinase (NRTK) signalling pathways. Bax protein (a central cell death regulator) and cytochrome C levels increase, and caspases 9 and 7 are activated. VEGFR-2 tyrosine kinase activity is reduced, preventing the formation of new blood vessels and the expression of P-cadherin. Prepared using BioRender (https://biorender.com/, accessed on 12 May 2022).
Microorganisms 10 01733 g006
Table 1. Humanized immunotoxins developed from fusion to diphtheria toxin (DT).
Table 1. Humanized immunotoxins developed from fusion to diphtheria toxin (DT).
ImmunotoxinToxin/FragmentTargeting MoietyCancer or Cell LineResultReference
OntakDT389IL-2Adult T-cell leukaemia and CTCLSignificant activity. FDA approval for CTCL treatment[171]
mVEGF-DTDT386mVEGFTC1-induced solid tumoursTumour regression and an increase in survival rate[174]
DTATC- and T-domains of DTN-terminal of uPAGlioblastoma cellsSelective killing and regression in tumour growth[175]
DT386-BR2DT386Buforin II (BR2)MCF-7 and HeLa cells
K-562
Specific and significant reduction in survivability and apoptosis[176,177]
Tagraxofusp (ElzonrisTM)DT388IL-3BPDCN, AML, CMML, & MMFDA approval for BPDCN treatment[178]
DT389GCSFDT389GCSFHL-60Specific apoptotic death and nuclease activity[179]
hDT806DT390HuBiscFv8064 HNSCC cell lines Apoptosis, tumour size reduction and EGFR signalling disruption[180]
PD1-DTDT386PD1C57BL/6 tumorous mice67% decrease in tumour volume[181]
DT389-YP7DT389hYP7 scFvHepG2 HCCDecreased cell viability and specific toxicity[182]
CTCL: cutaneous T-cell lymphoma; mVEGF: mouse vascular endothelial growth factor; DTAT: DT fused to the amino (N)-terminal of uPA; uPA: urokinase-type plasminogen activator; BPDCN: blastic plasmacytoid dendritic cell neoplasm; AML: acute myeloid leukaemia; CMML: chronic myelomonocytic leukaemia; MM: multiple myeloma; GCSF: granulocyte colony-stimulating factor; HuBiscFv806: humanized bivalent single-chain variable fragment of monoclonal antibody 806; HNSCC: head and neck squamous cell carcinoma; EGFR: epidermal growth factor receptor; PD1: programmed cell death protein-1; hYP7 scFv: humanized YP7 single-chain variable fragment; HCC: hepatocellular carcinoma.
Table 2. Studies using BoNT as an anti-cancer agent.
Table 2. Studies using BoNT as an anti-cancer agent.
Cancer or Cell LineStudy TypeMethodologyResultsReference
VCap cells
Cancerous human prostate
In vivo
In vivo
OnaA injection into VCap cells transplanted into murine prostate
OnaA injection into prostate before prostatectomy
Inhibited cancer progression and increased apoptosis
Increased incidence of apoptosis
[224]
MIA PaCa-2 cells In vivoCo-injection of cancer cells and 20 U/kg BoNT, or BoNT injection followed by cancer cell injection (murine study)Increase in apoptosis and a decrease in tumour size[225]
SiMA and SH-SY5Y cell linesIn vitroBoNT/C injection into retinoic acid-treatedIncrease in apoptosis[223]
3T3 fibroblast cellsIn vitroBoNT/A treatmentCytoplasmic degradation and decreased cell viability[211]
SCC-25 and HUVEC cellsIn vitroCells grown in the presence of BoNTNo effect on cell growth[226]
DBTRG glioblastoma cell lineIn vitroBoNT/A and BoNT/A + AMGIncreased apoptosis and decreased cell proliferation[227]
Abbreviations: OnaA: OnabotulinumtoxinA (Botox); DBTRG: Denver Brain Tumour Research Group; AMG: transient receptor potential vanilloid 1 receptor antagonist.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dicks, L.M.T.; Vermeulen, W. Do Bacteria Provide an Alternative to Cancer Treatment and What Role Does Lactic Acid Bacteria Play? Microorganisms 2022, 10, 1733. https://doi.org/10.3390/microorganisms10091733

AMA Style

Dicks LMT, Vermeulen W. Do Bacteria Provide an Alternative to Cancer Treatment and What Role Does Lactic Acid Bacteria Play? Microorganisms. 2022; 10(9):1733. https://doi.org/10.3390/microorganisms10091733

Chicago/Turabian Style

Dicks, Leon M. T., and Wian Vermeulen. 2022. "Do Bacteria Provide an Alternative to Cancer Treatment and What Role Does Lactic Acid Bacteria Play?" Microorganisms 10, no. 9: 1733. https://doi.org/10.3390/microorganisms10091733

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