Skip to main content

Biotechnological Activities and Applications of Bacterial Pigments Violacein and Prodigiosin

Abstract

In this review, we discuss violacein and prodigiosin, two chromogenic bacterial secondary metabolites that have diverse biological activities. Although both compounds were “discovered” more than seven decades ago, interest into their biological applications has grown in the last two decades, particularly driven by their antimicrobial and anticancer properties. These topics will be discussed in the first half of this review. The latter half delves into the current efforts of groups to produce these two compounds. This includes in both their native bacterial hosts and heterogeneously in other bacterial hosts, including discussing some of the caveats related to the yields reported in the literature, and some of the synthetic biology techniques employed in this pursuit.

Introduction

Bacterial strains are capable of producing many different secondary metabolites, including anti-cancer and antibiotic drugs. Here, we discuss two such compounds that are gaining interest due to their diverse biological activities, namely violacein and prodigiosin. Both of these compounds are synthesized by Gram-negative hosts and have been shown in studies from a wide berth of groups to possess important biological activities, including as potent antibiotics against multidrug resistant pathogens. Although both compounds were “discovered” nearly a century ago in the mid-20th century [1,2,3], their biological activities are still being studied to this day. However, one critical factor limiting research with either compound is their cost, which range from $360 to $760 per milligram [4]. Within this review, therefore, discussion will be given primarily to the biological activities of these compounds, focusing on ecological and medical considerations of both violacein and prodigiosin, as well as current methods to over-produce these remarkable compounds.

Violacein and Prodigiosin – Hydrophobic Bacterial Chromogenic Pigments

Prodigiosin and violacein are both colorful secondary metabolites, a trait that makes isolating and identifying the bacterial strains that produce these compounds in sufficient quantities easier. As shown in Fig. 1, violacein is a purple-hued bacterial pigment. The fact that this compound is produced by a range of natural bacterial strains [5,6,7,8], including Chromobacterium [9] and Janthinobacterium [10], and in a wide-array of environmental locales, including the deep seas [11], rivers [9, 12], agricultural and forest soils [8, 13, 14], within polar and alpine glacial regions [7, 15, 16], and even on the leaves of white clover [17] and the skin of amphibians [18], all suggest the production of violacein should be relatively advantageous for the host. However, the octanol-water partitioning coefficient (Log POW) for violacein is 3.34 [19, 20], classifying this compound as highly hydrophobic and suggesting it is not readily secreted by the host into the surrounding environment.

Fig. 1
figure 1

Violacein and prodigiosin, showing the chemical structure and the colored phenotypes of the bacterial strains that produce these compounds

Similarly, prodigiosin is vibrant red in color (Fig. 1) and is produced by a number of different Gram-negative and Gram-positive bacterial strains, including Serratia marcescens [21] and Streptomyces. As a compound, prodigiosin is a member of the prodiginines, a group of chemicals with the same parent nucleus but differing side groups. For this review, emphasis will be given primarily to prodigiosin as this is the most extensively studied compound within this group. When compared with violacein, prodigiosin is even more hydrophobic, with a Log POW of 5.16 [22].

Violacein and Prodigiosin as Antimicrobials

The antimicrobial activities of these two compounds have been extensively studied (Tables 1 and 2), particularly for violacein. It is historically recognized that very few Gram-negative bacteria are susceptible to violacein, data that is supported by independent groups in many recent studies [3, 39,40,41, 58]. The fact that violacein has been produced in recombinant strains of E. coli, as well as in Salmonella typhimurium VNP20009, Enterobacter aerogenes IAM1183 and Citrobacter freundii ACCC 05411, with no clear detriment to the growth or viability of these strains [59,60,61,62] supports this further. However, individual studies from some groups recently claim violacein exhibits low MIC or growth inhibitory activities with Gram-negative strains [63,64,65]. Given the historicity and wide range of reports suggesting otherwise, the veracity of these studies needs to be demonstrated independently by other research groups.

Table 1 Prodigiosin’s antibiotic activity against microorganisms
Table 2 Violacein’s antibiotic activity against microorganisms

In contrast, the activity of violacein against many different Gram-positive bacterial strains (Table 1), including Staphylococcus, Bacillus and Streptococcus [3, 40], is well established. Despite this, its spectrum does not extend to all Gram-positive strains. For instance, Enterococcus faecalis ATCC 29212 was not affected by the addition of violacein [66], while Corynebacterium glutamicum ATCC 21850 was genetically engineered to produce violacein [67]. It also exhibits antibiotic activities against Mycobacterium tuberculosis and M. smegmatis, which are acid-fast microbes, and the Gram-variable Micrococcus luteus [7, 68].

Stemming from its recognized activities against Gram-positive strains, many recent studies have evaluated the use of violacein against antibiotic-resistant strains of S. aureus [8, 41, 58, 66]. For instance, the minimal inhibitory concentrations (MICs) for several S. aureus associated with Bovine Mastitis were between 6.25 and 25.00 μM violacein, even though these strains displayed penicillin, ampicillin and/or intermediary erythromycin resistance [58]. Moreover, violacein acted synergistically with penicillin [58], an idea that was expanded on in another study [64]. A separate study using methicillin-resistant S. aureus (MRSA) reported MICs in basically the same range, i.e., 7.5 to 30 μM [66], while research from our group found a multidrug-resistant S. aureus clinical isolate with resistance to seven different antibiotics was also susceptible to violacein [8]. In that study, the MICs for both the clinical isolate and the non-resistant type strain (S. aureus ATCC 25923) were identical (15 μM) while bactericidal effects against both were seen when 30 μM or more violacein was employed [8]. This proved the antibacterial mechanism used by violacein differs from that of the other antibiotics and also that cross-resistance was not present.

For both compounds, their antimicrobial activities stem in part due to their lipophilic natures. When introduced into a bacterial culture, prodigiosin and violacein rapidly insert into the membranes of the microbe and disrupt their integrity, leading to ATP and protein leakage [22, 69, 70]. Interactions between violacein and bacterial membranes were recently modeled [70], and suggested that this compound does not embed very deeply within the lipid bilayer. The same study looked at the release of carboxyfluorescein from large unilamellar vesicles (LUVs) prepared using the lipids from three different bacteria, i.e., E. coli ATCC 25922, B. subtilis PY79 and S. aureus ATCC 25923. They found, regardless of the strain, the LUVs were equally susceptible [70], implying E. coli cellular membranes are just as likely to be attacked by violacein and that its inherent resistance to violacein stems from the protective nature of the outer membrane, which absorbs this antibiotic and prevents its access to the cytoplasmic membrane. Recent work from our group studied this further, but from a different perspective, by asking how violacein acts as an antibiotic in nature if it is hydrophobic and remains embedded primarily within the membrane of the strain that produced it. It was found C. violaceum secretes violacein within membrane vesicles (MVs) [20]. These vesicles bud off of the bacterium as it grows and contained more violacein than proteins (mg/mg), increasing the apparent water solubility of violacein. Using S. aureus and a violacein-deficient vioA mutant, the violacein-carrying MVs were proven to be bactericidal, although a greater overall amount of violacein was required to achieve the same killing efficiencies as crude purified violacein. In contrast, MVs from the vioA mutant had no impact on S. aureus viabilities, proving violacein was the bactericidal factor responsible.

A recent study also performed molecular dynamic simulations with prodigiosin [71]. The authors found, in contrast to violacein, prodigiosin embedded itself much deeper within the membrane lipid bilayer, a finding that helps explain why this compound is effective against some Gram-negative strains as this would increase the chances for prodigiosin to penetrate the outer membrane and enter the cytoplasmic membrane. However, it still remains to be seen if MVs are also used by prodigiosin-producing strains to transport this antibiotic to susceptible microbes.

In addition to membrane disruption, prodigiosin apparently causes additional damage within the bacterium, including the generation of reactive oxygen species (ROS) [23, 72] and, based on the study by Darshan and Manonmani (2016) [23], interacting with the bacterial genomic DNA. This latter facet of its activities corroborates an earlier study where prodigiosin was shown to cleave double-stranded DNA in vitro [73], an activity that is mediated by oxidative radicals (i.e., ROS) and requires the presence of a redox-active transition metal since the addition of either catalase or EDTA inhibited cleavage. Taken together, both studies suggest the ROS production by prodigiosin and its interactions with redox-active transition metals may act in concert in vivo to cause DNA damage within the bacterial cell, although this would benefit from further verification.

Prodigiosin and Violacein as Antifungals

In addition to their application towards bacterial pathogens, violacein (and its deoxyviolacein derivative) and prodigiosin also work widely and effectively against many pathogenic fungi (Tables 1 and 2). For violacein, representative examples of fungi that are susceptible include the plant pathogen Rhizoctonia solani [42, 46] and Batrachochytrium dendrobatidis [43, 44], a fungus that is lethal to amphibians. In the latter case, the presence of a violacein-producing bacterium, J. lividum, on the skin of the black-backed salamander (Plethodon cinereus) [44] or frog (Rana muscosa) [43] provided protection against B. dendrobatidis. Under these conditions, this bacterium was clearly able to produce a significant amount of violacein as the skin-associated concentrations with the frogs averaged around 100 μM, which was much higher than the 18 μM MIC needed to prevent mortality and morbidity caused by B. dendrobatidis based on the salamander study.

Although not studied as extensively, several reports have also discussed prodigiosin and its activities against different fungal species [30, 74,75,76,77]. Much like the two studies mentioned above, one group even looked at the ability of S. marcescens to protect Acris blanchardi (Blanchard’s Cricket frog) from B. dendrobatidis infections, reporting a slight, yet significant, increase in survival rates when compared against a pig mutant that is unable to synthesize this compound [77]. Moreover, although the mechanism of action is not fully understood, detailed observation of S. marcescens invading into fungus was reported recently [31]. In that study, prodigiosin increased the membrane permeability of target cell, enabling S. marcescens to invade into F. oxysporum. Given prodigiosin’s ability to damage the target cell’s membrane was also suggested as a mechanism of action against other bacterial cells [22], it would appear this compound has similar properties against organisms spanning different kingdoms.

Violacein and Prodigiosin as Nematicidal and Anti-Protozoan Agents

A benefit of violacein and prodigiosin for the producing bacteria is that it confers a survival advantage against competitors and predators, providing selective advantages against neighboring bacteria and an effective defense and deterrent against bacterivores, such as protozoa and nematodes (Tables 1 and 2).

Nematodes have caused detrimental disease to both humans and agriculture worldwide. Pine wilt disease, a serious epidemic that has devastated pine forests globally, especially in East Asia, is caused by the nematode Bursaphelenchus xylophilus. This nematode, also called pine wilt nematode, attacks the water transport system of pine trees, causing them to wilt and die [78]. Expensive nematicides have commonly been used to combat pine wilt nematode with little success. Recently, a violacein5'-O-glucoside derivative was constructed by expressing the glycosyltransferase (YjiC) from a Bacillus sp. in E. coli along with the vioABCDE [49]. This novel violacein derivative had increased water solubility and was an effective treatment against the pine wood nematode [49], suggesting its potential use in the future as an anti-nematodal agent against pine wilt disease.

Violacein also negatively impacted the nematode genetic model organism C. elegans. When fed on violacein-producing Janthinobacterium, C. elegans displayed developmental arrest in early larval stages [50]. Similar developmental arrest and delay was seen when violacein was expressed in E. coli OP50 [79] (Fig. 2), the normal laboratory diet of C. elegans [50]. Consumption of this compound induced the expression of several detoxification genes regulated by the insulin-like signaling pathway [80]. Interestingly, supplementation of unsaturated fatty acids, especially oleate, alleviated the worm growth and survival in violacein, whereas saturated fatty acids had no effect [79]. In addition to highlighting the anti-nematodal potential of violacein, studies in C. elegans may help also elucidate if a conserved mechanism of violacein-induced toxicity in metazoans exists. With the extensive genetic and molecular tools available for C. elegans, exploring how unsaturated fatty acids are able to mitigate violacein’s toxicity may provide a window into this mechanism, and may also shed light on its activities within cancer cells.

Fig. 2
figure 2

Violacein stunts the growth and development of C. elegans. (A) Lawn of E. coli strain OP50 (left) and violacein-expressing OP50 (OP50-vio, right). (B) Body length of worms grown on OP50 and OP50-vio from L1 larvae stage for 4 days. (C) Development of worms grown on OP50 and OP50-vio. Day 1 image show L1 synchronized worms that has never been fed. Scale bar = 100 μm. Figure originally published in [79]

Anticancer Activities of Prodigiosin and Violacein

Another well-known characteristic of these two compounds is their anti-tumor activities. Cancer is the second leading cause of death globally [81], and although recent therapeutics have been developed for some cancers, still it remains as devastating as ever. In the laboratory, prodigiosin has been reported to kill human cancer cell lines by a process called programmed cell death or apoptosis. Prodigiosin can induce apoptosis in haematopoietic cancer cells [82], human lung cancer cells [83], B cells and T cells in chronic lymphocytic leukemia [84], gastric cancer cells [85], multidrug resistant breast cancer cells [86], colorectal cancer cells [87] and glioblastoma multiforme cancer cells [88] (Table 3).

Table 3 List of cell lines evaluated with prodigiosin

Despite the strong evidence that prodigiosin can work against multiple types of cancer cells, how this compound targets cancer cell death by apoptosis is not yet clear. Prodigiosin can interact with and cleave DNA [73, 92], supporting one possible mechanism of cell death. Prodigiosin also facilitates proton and chloride ion symport and can affect the acidification of cellular compartments [94, 95], providing support for an alternative mechanism of cancer cell apoptosis [90]. Finally, prodigiosins also inhibit protein phosphatase activity in vitro [96, 97], suggesting another possible mechanism of how this compound may inhibit cancer cell growth.

More recent studies have suggested that prodigiosin causes cell death by affecting a cellular process called autophagy. The process of autophagy causes an accumulation of specific vesicles in the cell called autophagosomes that can break down damaged organelles or proteins [98]. Autophagy has also been a target for cancer therapy [99], especially due to the fact that this cellular process also regulates apoptosis in cancers [100]. In a recent laboratory study, prodigiosin treatment induced the death of glioblastoma cancer cells and reduced neurosphere growth, a marker associated with increased death in glioblastoma patients [88]. The authors further showed that apoptotic death of the glioblastoma cells by prodigiosin treatment was due to increased autophagy in the cancer cells. In another recent study, colorectal cancer cells that were treated with the chemical 5-fluorouracil, a common chemotherapy treatment for colorectal cancer, showed increased apoptosis in the presence of prodigiosin [91]. Interestingly, prodigiosin impaired autophagic flux which actually promoted cell death in the cancer cells in response to 5-fluorouracil.

Combination therapy, which uses two or more therapeutic agents as a cancer treatment, has become a main strategy in cancer therapy in recent years [101]. The use of prodigiosin in combination with other cancer therapies is a promising strategy that is currently being explored. As mentioned previously, 5-fluorouracil in combination with prodigiosin effectively killed colorectal cancer cells by increasing apoptosis [91]. In addition, a recent study showed that the combination of prodigiosin and PU-H71, a candidate therapy for triple negative breast cancer, induced apoptosis in a metastatic breast cancer cell line killing many of the cancer cells [93]. These studies, as well as others, confirm that prodigiosin promotes the killing of cancer cells in the laboratory and demonstrate that it is an excellent candidate for cancer therapy either as a combination therapy or singular treatment. However, whether this activity can actually translate to a treatment for cancer patients remains unknown. Several phase I and phase II clinical studies with various cancer patients have occurred with a prodigiosin derivative called obatoclax [102,103,104,105], and the jury is still out on whether prodigiosin is an effective therapy for human cancer patients.

Similar with prodigiosin, violacein is also a promising anti-tumor bacterial metabolite (Table 4). As with prodigiosin, violacein leads to mitochondrial dysfunction, brought on by mitochondrial membrane hyperpolarization, in MRC-5 and HeLa cells [111]. It was also confirmed in RAS-mutated metastatic melanoma cell lines that the autophagy process employed to resolve mitochondrial damage is impaired due to inhibition of AKT and AXL [115]. Subsequent processes followed a general apoptotic pathway leading to p38 MAP kinase phosphorylation, NFκB pathway activation, and activation of caspases when treated with 1 μM of violacein in HL60 [113]. However, in TF1, which is known to have apoptosis resistance, the IC50 was still only 2 μM despite co-treatment with inhibitors of pro-apoptotic caspases, leading the authors to conclude that violacein induces cell death via the activation of a non-canonical mechanism of cell death [116]. Interestingly, an in vitro study showed that violacein inhibits PKA and PKC activity [117]. While the results do not exclude other possible targets, and whether this leads to cancer cell death in vivo awaits to be examined, it suggests PKA and PKC could be a direct target of violacein.

Table 4 List of cell lines evaluated with violacein

This sequence of cell death mechanisms resulting from mitochondrial damage brought on by violacein is due to the profound threat to the energy metabolism of cells. As a good indication of this, violacein has enhanced anti-cancer activities against some cell lines in hypoxia, such as HCT 116 (4.8-fold), HN5 (6.5-fold), HT29 (12.6-fold), and MCF7 (4-fold) [60]. Moreover, violacein treatment (1μM) led to the downregulated expression of chemokine/receptor CXCL12/CXCR4, which is important for angiogenesis [118]. Since carcinoma development without angiogenesis leads to hypoxic conditions, these results suggest violacein may actually induce the conditions within the tumor that increase its effectiveness as an anticancer agent, as was reported in one study [119].

Other studies have confirmed that oral administration of violacein contributes to NSAID-induced gastric damage healing. This led to a decrease in inflammatory cytokines, particularly TNF-α, and an increase in epidermal growth factor (EGF), vascular endothelial growth factor (VEGF) and hepatocyte growth factor (HGF) [120]. These appear to play an essential role in healing angiogenesis and mucin secretion. In other words, violacein administered orally plays a role in inhibiting inflammation, maintaining the balance of cytokines, while also inhibiting apoptosis, angiogenesis, and promoting healing.

Immunomodulatory Activities of Prodigiosin and Violacein

Prodigiosin is also known to have immunosuppressive effects. Specifically, this compound shows suppressive effects on T-cell proliferation, while having no effect in B-cells [121]. Its mechanism of action is to inhibit expression of the interleukin-2 receptor α(IL-2Rα) chain, an important contributor of T-cell activation [122]. In another study, the authors developed a prodigiosin-analogue molecule, PNU156804, which suppressed both T-cell and B-cell activation [123]. This compound also worked through inhibiting IL-2 dependent signaling, i.e., not by preventing IL-2Rα induction but rather by preventing activation of AP-1 and NF-κB. Prodigiosin was also synergistically active when administered with cyclosporine A, each working through different pathways to suppress T-cell activation [124], while another study found it inhibited macrophage and NK killer cell activities and splenocyte proliferation [125]..

Violacein was also shown to have immunomodulatory functions and inhibit inflammation. For instance, this compound had antipyretic, analgesic, and immunomodulatory reactions when orally administered to rats [126]. In ulcer rat models, violacein relieved inflammation of the gastrointestinal tract, possibly working through the COX-1 mediated pathways [120], while another study reported that, when injected directly into the intraperitoneal cavity, violacein can have immunomodulatory effects by regulating cytokine production: it down-regulated the expression of IL-6 and TNF-α but induced expression of IL-10 [127].

Some of the immunomodulatory mechanisms and findings associated with violacein seem contradictory with the cancer studies, however. Unlike the above study that reported violacein inhibits TNF-α expression [127], TNF-α expression was elevated in HL60, and TNF receptor 1 signaling was also activated when this cell line was exposed to violacein [113]. It is also known to increase the expression of TNF-α and upregulate the p53-dependent mitochondrial pathway in MCF-7 [128], while treatment with violacein also induced TNF-α expression in Raw 264.7 and ANA-1 cells [129]. These differences may be due to the experimental protocols, though, as the above studies were performed in vitro, i.e., violacein treatment directly into human or murine cell cultures [128, 129], rather than in vivo, i.e., the oral administration or injection of violacein into the digestive tract or intraperitoneal cavity [120, 126, 127]. In other words, vastly different results may result depending on the method of administration and the type of cells, but all of the above studies confirmed that violacein has immunomodulatory aspects.

Bioproduction - Measurement of Prodigiosin and Violacein – Spectrophotometry vs. HPLC

The classical method for prodigiosin extraction from the bacterial host and culture is to use acidified ethanol (4% 1M HCl v/v) to prevent the rapid decomposition of this molecule when above pH 5. The impurities present in the extracted prodigiosin are then removed using a solvent such as dichloromethane or n-hexane:chloroform and the final product purified through chromatography [130,131,132].

The simplest way to measure the extracted prodigiosin is to use a spectrophotometer using an absorption wavelength of 530-540 nm and convert this to the concentration using an extinction coefficient (ε) and the Beer–Lambert law. However, this is not without issue as the value of ε varies from study to study. Traditionally, the value of ε535 is 0.159 L/mg-cm [133]. The most detailed study on the extinction coefficient of prodigiosin is Domröse et al. (2015) [130], where ε535 was calculated to be 0.4322 L/mg-cm in acidified ethanol, a value that was confirmed through quantitative 1H-NMR. This value is near identical with that reported by another group, i.e., ε535 = 0.4311 L/mg-cm [134]. Consequently, due to the difference in the extinction coefficients, the prodigiosin concentration using the classical ε value will be over-estimated by 270%.

Similarly, violacein has often been quantified using a spectrophotometer and its absorbance peaks at 575-590nm [8, 135,136,137]. However, because of differences in reported ε between research groups, the yields claimed in the literature are inconsistent. For example, the ε values for violacein include, from lowest to highest, ε570 = 10.955 L/g-cm in ethanol [138], ε = 29.700 L/g-cm [137], ε565 = 31.3 L/g-cm in acetone-water [139], ε570 = 46 L/g-cm in ethanol [67], ε575 = 56.010 L/g-cm in ethanol [135] and ε575 = 74.3 L/g-cm in ethanol [140]. This disparity was raised in the study by Rodrigues et al. (2013) [140] and in previous reviews [141, 142], potentially inflating the violacein yields by as much as 670%.

To address this issue, Rodrigues et al. (2013) [140] elected to quantify violacein through HPLC [140], a protocol that has been successfully applied within several of our own studies [20, 143, 144]. At present, similar protocols have not been applied to quantify prodigiosin and HPLC may consolidate the yields in the literature, an idea that should be evaluated further. However, given the wide-spread problems raised by this issue, the concentrations of these two compounds reportedly produced in the literature will not be discussed, but rather the qualitative results of the studies.

Production by Natural Isolates

As discussed above, a wide-range of natural bacterial strains are capable of synthesizing violacein and prodigiosin. It should come as no surprise, therefore, that researchers have sought out a variety of strains for the lab scale production and application of these two compounds. For instance, S. marcescens FZSF02 was isolated from the soil in the region of Fuzhou, China, and is capable of producing prodigiosin in sufficient quantities that it reportedly pellets out of solution [145]. Another natural strain, S. marcescens MO-1 was isolated from a grasshopper [146] while S. marcescens UCP1459 and S. marcescens UTM1 were isolated from semi-arid soil in Brazil and an oxidation pond in Malaysia, respectively [147, 148]. A related bacterium, S. rubidaea, also produced prodigiosin and was initially isolated from a spoiled coconut, where it was discovered since it changed the color of the inside of the coconut, making it pink [149].

Similarly, violacein production has been studied in different natural strains. For instance, production of this compound in C. violaceum CCT 3496 was increased around 2.5-fold when tryptone and yeast extract were added, but the yields dropped with glucose [135]. In a separate study, optimization in Duganella sp. B2 found tryptophan, beef extract, and potassium nitrate were all major factors impacting violacein yields [138] while in Massilia sp. EP15224, an isolate known to be closely related to Duganella sp., the MM2 broth used to cultivate this strain was improved by adjusting the amount of phosphate, leading to faster production rates and slightly better final violacein yields [150].

Some violacein-producing bacteria are also psychrotrophic, such as strain RT102, which is related with J. lividum, reported by Nakmura et al (2003) [40]. The authors found that the conditions leading to optimum production levels were a slightly acid pH of 6, the growth temperature set to 20°C and with 1 mg/L of dissolved oxygen. Although not as psychrophilic as RT102, J. lividum was also successfully used to produce violacein, albeit at 25 °C and a pH of 7.0 [65]. Notably, in this study, the addition of 0.2 mg/mL of the antibiotic of silver ampicillin improved the yields by a factor of 1.3 while glycerol was used as a carbon source, a choice the authors claimed improves the violacein production relative to the cell mass.

The idea of using ampicillin and glycerol to increase violacein yields was actually reported more than a decade earlier in the study by Pantanella et al (2007), where glycerol enhanced violacein production levels by approximately 12-fold, while ampicillin led to an estimated 3-fold increase [136]. These factors, unfortunately, were not additive when used together – the maximum level with glycerol with or without ampicillin were basically identical.

The use of more natural feed stocks was also considered, as in the case with C. violaceum UTM5 where agricultural wastes were used [151], or in a separate study where liquid pineapple waste was used as the carbon source along with addition of L-tryptophan [41]. However, as noted above, since these papers do not provide the extinction coefficient and did not use HPLC techniques when quantifying their yields, it is difficult to directly compare their results with other studies.

Random Mutations to Enhance Prodigiosin Production

One strategy used by researchers to enhance production of prodigiosin is to generate random mutations within the genome of the natural host, typically with radiation. Since prodigiosin is a red pigment, screening is a simple and quick method for researchers to identify those colonies that overproduce this compound based on their color intensity. This was successfully used by one group with microwave irradiation to increase the prodigiosin yields from S. marcescens jx1 by just over two-fold [26], while a separate group used gamma irradiation [152]. In the latter study, the authors varied more than just the radiation dose and rate, including the pH and inoculum size, to identify conditions that optimize for prodigiosin production. However, as in the microwave radiation study, the yields were only improved by about 2-fold.

Heterogeneous Expression and Metabolic Pathway Engineering to Increase Prodigiosin and Violacein Yields

The above yields, although definitely improved, are not very significant and highlight potential limitations linked with random mutation studies, namely that improvements may not be very substantial, particularly when they involve complex metabolic pathways encoded in multiple genes such as those involved in prodigiosin biosynthesis. As such, researchers have often sought to clone and express the genes in other hosts where the metabolic and biosynthetic pathways can be engineered.

The prodigiosin gene cluster (pig) includes many genes, pigA to pigN, but may vary in gene order as well as include some auxiliary genes depending on the bacterial host [153, 154]. During the mid-20th century, studies in prodigiosin biosynthesis focused on related molecular components and constructing the pathway [21, 155,156,157], including the role of quorum sensing mechanisms [158,159,160,161], while recent studies have provided a more detailed understanding of the biosynthetic pathways involved [153, 154]. Violacein research has followed a similar path, with the biosynthetic pathway first mapped in 1991 [162] and the roles of the individual genes and enzymes characterized further in the early 2000’s [163, 164]. In addition, during the same period many articles, were published discussing the roles of quorum sensing in the production of this metabolite [165,166,167]. This led to the eventual development and application of C. violaceum CV026 as a quorum-sensing reporter strain, as it visually responded to the presence of acyl homoserine lactones (AHLs) with the production or inhibition of violacein synthesis [166, 168, 169]. Recently, this strain was reclassified as C. subtsugae [170].

As the last two decades have seen sequencing techniques and comparative genome analyses dramatically improve, a new era of prodigiosin and violacein production has opened. Gene clusters related with prodigiosin production were sequenced, analyzed and compared among different species and subspecies [171, 172], as have the genomes of numerous violacein-producing bacteria [173,174,175], particularly by Dr. Brooke Jude at Bard University who, in the last couple of years, has published several genomes [176,177,178,179]. Of particular note, one of the Janthinobacter sp. sequenced by her group actually lacked the genes for violacein but carried the pig gene cluster, allowing it to produce prodigiosin [180]. They concluded that, since this strain was isolated from the region where other violacein-producing strains were also located, including other Janthinobacter sp., the production of prodigiosin by this strain may represent a combined effort by the two groups to combat other bacterial species.

All of this information will aid researchers in further efforts to clone and express the genes required in other bacterial strains. This is not to say that this has not been done already, as a few groups reported the heterogeneous production of prodigiosin [130, 181], one as far back as 1984 [182]. However, only one study truly sought to use the new host, in this case Pseudomonas putida KT2440, as a platform for the production of this compound [130]. In their study, the authors introduced the pig cluster randomly into the genome of P. putida using a plasmid bearing a transposon and screened the resulting clones for prodigiosin production, looking for insertions where the cluster was expressed by a strong promoter. Using this method, they were able to increase prodigiosin production on agar plates by approximately five-fold over the original S. marcescens and as much as 94 mg/L, based on their quantification methods, in liquid cultures.

In contrast, the expression of violacein in other bacterial hosts is widespread, with the vioABCDE genes cloned and expressed within many plasmids and bacterial hosts. Some examples of this include Citrobacter freundii [61, 62], Klebsiella aerogenes (formerly Enterobacter aerogenes) [62] and E. coli [62, 140, 162, 183,184,185]. Other studies have sought to improve on the violacein yields through synthetic biology, often with E. coli as the host [54, 186], albeit not always for purification, as illustrated in two recent studies where its expression was used as a bioreporter [187, 188]. One prime example where synthetic biology was employed to improve violacein production is the study by Jeshek et al. (2016) where they introduced the Reduced Libraries algorithm [189]. These used this system to design smart combinatorial libraries for pathway optimization based on the ribosomal binding sites and, in this case, focused on increasing violacein production while minimizing that of deoxyviolacein. A second group used a different approach and elected to express each gene independently by their own promoter [59]. By controlling the strengths of each individual promoter, and using a combinatorial assembly of the genes, they were able to increase the violacein titers by more than 60-fold over the control, where each gene was expressed under the T7 promoter. In addition to E. coli, other hosts have been used for the heterogeneous production of violacein, including yeasts [190, 191]. One such study used Yarrowia lipolytica, an oleaginous yeast, as the host, where the vio genes were expressed using three different promoters and assembled using the Golden Gate assembly method to build combinatorial pathway libraries [191]. From this, three yeast strains, each producing a different chromogenic compound, i.e., violacein, deoxyviolacein and proviolacein, were constructed.

Conclusions

This review presented many biological traits of both prodigiosin and violacein reported in the recent and current literature. Fig. 3 is a plot showing the number of peer-reviewed articles listed in the National Center for Biotechnology Information’s PubMed website [192] for each year, providing visual evidence of the growing interest into these two compounds and their activities. Although the numbers may not be as great as some other hot-topics, the data makes it clear that many research groups continue to study and explore the biological activities of these two compounds and different methods for producing them in greater quantities. As this field continues to expand and mature, other derivatives of violacein and prodigiosin are expected to move towards clinical trials as antimicrobials and for the treatment of human diseases, including cancer, as was noted above for obatoclax. This will be supported in no small part by synthetic biologists and chemical engineers who are currently developing novel and more efficient protocols and strains to increase the productivity and yields of these two secondary metabolites, a trend that is also expected to reduce the costs of these compounds, which at present are too high for conventional medical research.

Fig. 3
figure 3

Number of research articles related with prodigiosin and violacein published each year according to the data available at the NCBI PubMed website [192]. (Accessed Jan 20th, 2021)

Availability of data and materials

None.

References

  1. Wasserman HH, Mc KJ, Santer UV. Studies related to the biosynthesis of prodigiosin in Serratia marcescens. Biochem Biophys Res Commun. 1960;3:146–9.

    Article  Google Scholar 

  2. Rapoport H, Holden KG. The Synthesis of Prodigiosin. J Am Chem Soc. 1962;84(4):635–42.

    Article  Google Scholar 

  3. Lichstein HC, Vandesand VF. Violacein, an Antibiotic Pigment Produced by Chromobacterium-Violaceum. J Infect Dis. 1945;76(1):47–51.

    Article  Google Scholar 

  4. Sigma-Aldrich U.S. website. https://www.sigmaaldrich.com/. Accessed 20 Jan 2021.

  5. Yang LH, Xiong H, Lee OO, Qi SH, Qian PY. Effect of agitation on violacein production in Pseudoalteromonas luteoviolacea isolated from a marine sponge. Lett Appl Microbiol. 2007;44(6):625–30.

  6. Matz C, Webb JS, Schupp PJ, Phang SY, Penesyan A, Egan S, Steinberg P, Kjelleberg S. Marine biofilm bacteria evade eukaryotic predation by targeted chemical defense. PLoS One. 2008;3(7):e2744.

  7. Hakvag S, Fjaervik E, Klinkenberg G, Borgos SEF, Josefsen KD, Ellingsen TE, Zotchev SB. Violacein-Producing Collimonas sp from the Sea Surface Microlayer of Costal Waters in Trondelag, Norway. Mar Drugs. 2009;7(4):576–88.

    Article  Google Scholar 

  8. Choi SY, Kim S, Lyuck S, Kim SB, Mitchell RJ. High-level production of violacein by the newly isolated Duganella violaceinigra str. NI28 and its impact on Staphylococcus aureus. Sci Rep. 2015;5:15598.

  9. Moss MO, Ryall C, Logan NA. Classification and Characterization of Chromobacteria from a Lowland River. J Gen Microbiol. 1978;105(Mar):11–21.

    Article  Google Scholar 

  10. Osullivan J, Mccullough J, Johnson JH, Bonner DP, Clark JC, Dean L, Trejo WH. Janthinocin-a, Janthinocin-B and Janthinocin-C, Novel Peptide Lactone Antibiotics Produced by Janthinobacterium-Lividum .1. Taxonomy, Fermentation, Isolation, Physicochemical and Biological Characterization. J Antibiot. 1990;43(8):913–9.

    Article  Google Scholar 

  11. Yada S, Wang Y, Zou Y, Nagasaki K, Hosokawa K, Osaka I, Arakawa R, Enomoto K. Isolation and characterization of two groups of novel marine bacteria producing violacein. Mar Biotechnol. 2008;10(2):128–32.

    Article  Google Scholar 

  12. Riveros R, Haun M, Duran N. Effect of Growth-Conditions on Production of Violacein by Chromobacterium-Violaceum (Bb-78 Strain). Braz J Med Biol Res. 1989;22(5):569–77.

    Google Scholar 

  13. Aranda S, Montes-Borrego M, Landa BB. Purple-Pigmented Violacein-Producing Duganella spp. Inhabit the Rhizosphere of Wild and Cultivated Olives in Southern Spain. Microb Ecol. 2011;62(2):446–59.

    Article  Google Scholar 

  14. Li WJ, Zhang YQ, Park DJ, Li CT, Xu LH, Kim CJ, Jiang CL. Duganella violaceinigra sp nov., a novel mesophilic bacterium isolated from forest soil. Int J Syst Evol Micr. 2004;54:1811–4.

    Article  Google Scholar 

  15. Baricz A, Teban A, Chiriac CM, Szekeres E, Farkas A, Nica M, Dascălu A, Oprișan C, Lavin P, Coman C. Investigating the potential use of an Antarctic variant of Janthinobacterium lividum for tackling antimicrobial resistance in a One Health approach. Sci Rep. 2018;8(1):15272.

  16. Kim SJ, Shin SC, Hong SG, Lee YM, Lee H, Lee J, Choi IG, Park H. Genome Sequence of Janthinobacterium sp Strain PAMC 25724, Isolated from Alpine Glacier Cryoconite. J Bacteriol. 2012;194(8):2096.

    Article  Google Scholar 

  17. Kampfer P, Wellner S, Lohse K, Martin K, Lodders N. Duganella phyllosphaerae sp. nov., isolated from the leaf surface of Trifolium repens and proposal to reclassify Duganella violaceinigra into a novel genus as Pseudoduganella violceinigra gen. nov., comb. nov. (vol 35, pg 19, 2012). Syst Appl Microbiol. 2012;35(4):278.

    Article  Google Scholar 

  18. Brucker RM, Harris RN, Schwantes CR, Gallaher TN, Flaherty DC, Lam BA, Minbiole KPC. Amphibian Chemical Defense: Antifungal Metabolites of the Microsymbiont Janthinobacterium lividum on the Salamander Plethodon cinereus. J Chem Ecol. 2008;34(11):1422–9.

    Article  Google Scholar 

  19. Suryawanshi RK, Patil CD, Borase HP, Narkhede CP, Stevenson A, Hallsworth JE, Patil SV. Towards an understanding of bacterial metabolites prodigiosin and violacein and their potential for use in commercial sunscreens. Int J Cosmetic Sci. 2015;37(1):98–107.

    Article  Google Scholar 

  20. Choi SY, Lim S, Cho G, Kwon J, Mun W, Im H, Mitchell RJ. Chromobacterium violaceum delivers violacein, a hydrophobic antibiotic, to other microbes in membrane vesicles. Environ Microbiol. 2020;22(2):705–13.

    Article  Google Scholar 

  21. Williams RP. Biosynthesis of prodigiosin, a secondary metabolite of Serratia marcescens. Appl Microbiol. 1973;25(3):396–402.

    Article  Google Scholar 

  22. Suryawanshi RK, Patil CD, Koli SH, Hallsworth JE, Patil SV. Antimicrobial activity of prodigiosin is attributable to plasma-membrane damage. Nat Product Res. 2016;31(5):572–7.

    Article  Google Scholar 

  23. Darshan N, Manonmani HK. Prodigiosin inhibits motility and activates bacterial cell death revealing molecular biomarkers of programmed cell death. AMB Express. 2016;6(1):50.

  24. Danevčič T, Borić Vezjak M, Tabor M, Zorec M, Stopar D. Prodigiosin Induces Autolysins in Actively Grown Bacillus subtilis Cells. Front Microbiol. 2016;7:27.

  25. Li D, Liu J, Wang X, Kong D, Du W, Li H, Hse CY, Shupe T, Zhou D, Zhao K. Biological Potential and Mechanism of Prodigiosin from Serratia marcescens Subsp. lawsoniana in Human Choriocarcinoma and Prostate Cancer Cell Lines. Int J Mol Sci. 2018;19(11):3465.

  26. Liu X, Wang Y, Sun S, Zhu C, Xu W, Park Y, Zhou H. MUTANT BREEDING OFSerratia marcescensSTRAIN FOR ENHANCING PRODIGIOSIN PRODUCTION AND APPLICATION TO TEXTILES. Prep Biochem Biotechnol. 2013;43(3):271–84.

    Article  Google Scholar 

  27. Lapenda JC, Silva PA, Vicalvi MC, Sena KXFR, Nascimento SC. Antimicrobial activity of prodigiosin isolated from Serratia marcescens UFPEDA 398. World J Microbiol Biotechnol. 2014;31(2):399–406.

    Article  Google Scholar 

  28. Woodhams DC, LaBumbard BC, Barnhart KL, Becker MH, Bletz MC, Escobar LA, Flechas SV, Forman ME, Iannetta AA, Joyce MD, et al. Prodigiosin, Violacein, and Volatile Organic Compounds Produced by Widespread Cutaneous Bacteria of Amphibians Can Inhibit Two Batrachochytrium Fungal Pathogens. Microb Ecol. 2018;75(4):1049–62.

    Article  Google Scholar 

  29. Someya N, Nakajima M, Hirayae K, Hibi T, Akutsu K. Synergistic Antifungal Activity of Chitinolytic Enzymes and Prodigiosin Produced by Biocontrol Bacterium, Serratia marcescens Strain B2 against Gray Mold Pathogen, Botrytis cinerea. J General Plant Pathol. 2001;67(4):312–7.

    Article  Google Scholar 

  30. John Jimtha C, Jishma P, Sreelekha S, Chithra S, Radhakrishnan EK. Antifungal properties of prodigiosin producing rhizospheric Serratia sp. Rhizosphere. 2017;3:105–8.

    Article  Google Scholar 

  31. Hazarika DJ, Gautom T, Parveen A, Goswami G, Barooah M, Modi MK, Boro RC. Mechanism of interaction of an endofungal bacterium Serratia marcescens D1 with its host and non-host fungi. PLoS One. 2020;15(4):e0224051.

  32. Gutiérrez-Román MI, Holguín-Meléndez F, Dunn MF, Guillén-Navarro K, Huerta-Palacios G. Antifungal activity of Serratia marcescens CFFSUR-B2 purified chitinolytic enzymes and prodigiosin against Mycosphaerella fijiensis, causal agent of black Sigatoka in banana (Musa spp.). BioControl. 2015;60(4):565–72.

    Article  Google Scholar 

  33. Someya N, Kataoka N, Komagata T, Hirayae K, Hibi T, Akutsu K. Biological Control of Cyclamen Soilborne Diseases by Serratia marcescens Strain B2. Plant Disease. 2000;84(3):334–40.

    Article  Google Scholar 

  34. Suryawanshi RK, Koujah L, Patil CD, Ames JM, Agelidis A, Yadavalli T, Patil SV, Shukla D. Bacterial Pigment Prodigiosin Demonstrates a Unique Antiherpesvirus Activity That Is Mediated through Inhibition of Prosurvival Signal Transducers. J Virol. 2020;94(13):e00251-20.

  35. Isaka M, Jaturapat A, Kramyu J, Tanticharoen M, Thebtaranonth Y. Potent In Vitro Antimalarial Activity of Metacycloprodigiosin Isolated from Streptomycesspectabilis BCC 4785. Antimicrobial Agents Chemother. 2002;46(4):1112–3.

    Article  Google Scholar 

  36. Papireddy K, Smilkstein M, Kelly JX, Shweta SSM, Alhamadsheh M, Haynes SW, Challis GL, Reynolds KA. Antimalarial Activity of Natural and Synthetic Prodiginines. J Med Chem. 2011;54(15):5296–306.

    Article  Google Scholar 

  37. Genes C, Baquero E, Echeverri F, Maya JD, Triana O. Mitochondrial dysfunction in Trypanosoma cruzi: the role of Serratia marcescens prodigiosin in the alternative treatment of Chagas disease. Parasit Vectors. 2011;4:66.

  38. Patil CD, Patil SV, Salunke BK, Salunkhe RB. Prodigiosin produced by Serratia marcescens NMCC46 as a mosquito larvicidal agent against Aedes aegypti and Anopheles stephensi. Parasitol Res. 2011;109(4):1179–87.

    Article  Google Scholar 

  39. Im H, Choi SY, Son S, Mitchell RJ. Combined Application of Bacterial Predation and Violacein to Kill Polymicrobial Pathogenic Communities. Sci Rep. 2017;7(1):14415.

  40. Nakamura Y, Asada C, Sawada T. Production of antibacterial violet pigment by psychrotropic bacterium RT102 strain. Biotechnol Bioproc E. 2003;8(1):37–40.

    Article  Google Scholar 

  41. Aruldass CA, Rubiyatno VCK, Ahmad WA. Violet pigment production from liquid pineapple waste by Chromobacterium violaceum UTM5 and evaluation of its bioactivity. Rsc Adv. 2015;5(64):51524–36.

    Article  Google Scholar 

  42. Sasidharan A, Sasidharan NK, Amma DBNS, Vasu RK, Nataraja AV, Bhaskaran K. Antifungal activity of violacein purified from a novel strain of Chromobacterium sp NIIST (MTCC 5522). J Microbiol. 2015;53(10):694–701.

    Article  Google Scholar 

  43. Harris RN, Brucker RM, Walke JB, Becker MH, Schwantes CR, Flaherty DC, Lam BA, Woodhams DC, Briggs CJ, Vredenburg VT, et al. Skin microbes on frogs prevent morbidity and mortality caused by a lethal skin fungus. ISME J. 2009;3(7):818–24.

    Article  Google Scholar 

  44. Becker MH, Brucker RM, Schwantes CR, Harris RN, Minbiole KPC. The Bacterially Produced Metabolite Violacein Is Associated with Survival of Amphibians Infected with a Lethal Fungus. Appl Environ Microb. 2009;75(21):6635–8.

    Article  Google Scholar 

  45. Shirata A, Tsukamoto T, Yasui H, Hata T, Hayasaka S, Kojima A, Kato H. Isolation of bacteria producing bluish-purple pigment and use for dyeing. Jarq-Jpn Agr Res Q. 2000;34(2):131–40.

    Google Scholar 

  46. Wang H, Wang F, Zhu X, Yan Y, Yu X, Jiang P, Xing X-H. Biosynthesis and characterization of violacein, deoxyviolacein and oxyviolacein in heterologous host, and their antimicrobial activities. Biochem Eng J. 2012;67:148–55.

    Article  Google Scholar 

  47. Lee Y-R, Mitchell RJ, Whang K-S. Isolation and characterization of antifungal violacein producing bacterium Collimonas sp. DEC-B5. Korean J Microbiol. 2016;52(2):212–9.

    Article  Google Scholar 

  48. Andrighetti-Frohner CR, Antonio RV, Creczynski-Pasa TB, Barardi CRM, Simoes CMO. Cytotoxicity and potential antiviral evaluation of violacein produced by Chromobacterium violaceum. Mem I Oswaldo Cruz. 2003;98(6):843–8.

    Article  Google Scholar 

  49. Lee YJ, Bashyal P, Pandey RP, Sohng JK. Enzymatic and Microbial Biosynthesis of Novel Violacein Glycosides with Enhanced Water Solubility and Improved Anti-nematode Activity. Biotechnol Bioproc E. 2019;24(2):366–74.

    Article  Google Scholar 

  50. Hornung C, Poehlein A, Haack FS, Schmidt M, Dierking K, Pohlen A, Schulenburg H, Blokesch M, Plener L, Jung K, Bonge A, Krohn-Molt I, Utpatel C, Timmermann G, Spieck E, Pommerening-Röser A, Bode E, Bode HB, Daniel R, Schmeisser C, Streit WR. The Janthinobacterium sp. HH01 genome encodes a homologue of the V. cholerae CqsA and L. pneumophila LqsA autoinducer synthases. PLoS One. 2013;8(2):e55045.

  51. Leon LL, Miranda CC, De Souza AO, Durán N. Antileishmanial activity of the violacein extracted from Chromobacterium violaceum. J Antimicrob Chemother. 2001;48(3):449–50.

  52. Leon LL, Miranda CC, De Souza AO, Duran N. Antileishmanial activity of the violacein extracted from Chromobacterium violaceum. J Antimicrob Chemoth. 2001;48(3):449–50.

    Article  Google Scholar 

  53. Lopes SCP, Blanco YC, Justo GZ, Nogueira PA, Rodrigues FLS, Goelnitz U, Wunderlich G, Facchini G, Brocchi M, Duran N, et al. Violacein Extracted from Chromobacterium violaceum Inhibits Plasmodium Growth In Vitro and In Vivo. Antimicrobial Agents Chemother. 2009;53(5):2149–52.

    Article  Google Scholar 

  54. Bilsland E, Tavella TA, Krogh R, Stokes JE, Roberts A, Ajioka J, Spring DR, Andricopulo AD, Costa FTM, Oliver SG. Antiplasmodial and trypanocidal activity of violacein and deoxyviolacein produced from synthetic operons. BMC Biotechnol. 2018;18(1):22.

  55. Rahul S, Chandrashekhar P, Hemant B, Bipinchandra S, Mouray E, Grellier P, Satish P. In vitro antiparasitic activity of microbial pigments and their combination with phytosynthesized metal nanoparticles. Parasitol Int. 2015;64(5):353–6.

  56. Lozano GL, Guan C, Cao Y, Borlee BR, Broderick NA, Stabb EV, Handelsman J. A Chemical Counterpunch: Chromobacterium violaceum ATCC 31532 Produces Violacein in Response to Translation-Inhibiting Antibiotics. mBio. 2020;11(3):e00948–20.

  57. Baskar K, Ignacimuthu S. Bioefficacy of violacein against Asian armyworm Spodoptera litura Fab. (Lepidoptera: Noctuidae). J Saudi Soc Agric Sci. 2012;11(1):73–7.

    Google Scholar 

  58. Cazoto LL, Martins D, Ribeiro MG, Duran N, Nakazato G. Antibacterial activity of violacein against Staphylococcus aureus isolated from Bovine Mastitis. J Antibiot. 2011;64(5):395–7.

    Article  Google Scholar 

  59. Jones JA, Vernacchio VR, Lachance DM, Lebovich M, Fu L, Shirke AN, Schultz VL, Cress B, Linhardt RJ, Koffas MA. ePathOptimize: A Combinatorial Approach for Transcriptional Balancing of Metabolic Pathways. Sci Rep. 2015;5:11301.

  60. Hashimi SM, Xu TF, Wei MQ. Violacein anticancer activity is enhanced under hypoxia. Oncol Rep. 2015;33(4):1731–6.

    Article  Google Scholar 

  61. Yang C, Jiang PX, Xiao S, Zhang C, Lou K, Xing XH. Fed-batch fermentation of recombinant Citrobacter freundii with expression of a violacein-synthesizing gene cluster for efficient violacein production from glycerol. Biochem Eng J. 2011;57:55–62.

    Article  Google Scholar 

  62. Jiang PX, Wang HS, Zhang C, Lou K, Xing XH. Reconstruction of the violacein biosynthetic pathway from Duganella sp B2 in different heterologous hosts. Appl Microbiol Biot. 2010;86(4):1077–88.

    Article  Google Scholar 

  63. Asencio G, Paris Lavin, Karen Alegrà a, et al. Antibacterial activity of the Antarctic bacterium Janthinobacterium sp. SMN 33.6 against multi-resistant Gram-negative bacteria Electronic Journal of Biotechnology. 2014;17:1-5.

  64. Subramaniam S, Ravi V, Sivasubramanian A. Synergistic antimicrobial profiling of violacein with commercial antibiotics against pathogenic micro-organisms. Pharm Biol. 2014;52(1):86–90.

    Article  Google Scholar 

  65. Kanelli M, Mandic M, Kalakona M, Vasilakos S, Kekos D, Nikodinovic-Runic J, Topakas E. Microbial Production of Violacein and Process Optimization for Dyeing Polyamide Fabrics With Acquired Antimicrobial Properties. Front Microbiol. 2018;9:1495.

  66. Martins D, Costa FTM, Brocchi M, Duran N. Evaluation of the antibacterial activity of poly-(D,L-lactide-co-glycolide) nanoparticles containing violacein. J Nanopart Res. 2011;13(1):355–63.

    Article  Google Scholar 

  67. Sun H, Zhao D, Xiong B, Zhang C, Bi C. Engineering Corynebacterium glutamicum for violacein hyper production. Microb Cell Fact. 2016;15(1):148.

  68. Mojib N, Philpott R, Huang JP, Niederweis M, Bej AK. Antimycobacterial activity in vitro of pigments isolated from Antarctic bacteria. Anton Leeuw Int J G. 2010;98(4):531–40.

    Article  Google Scholar 

  69. Aruldass CA, Masalamany SRL, Venil CK, Ahmad WA. Antibacterial mode of action of violacein from Chromobacterium violaceum UTM5 against Staphylococcus aureus and methicillin-resistant Staphylococcus aureus (MRSA). Environ Sci Pollut Res. 2017;25(6):5164–80.

    Article  Google Scholar 

  70. Cauz ACG, Carretero GPB, Saraiva GKV, Park P, Mortara L, Cuccovia IM, Brocchi M, Gueiros-Filho FJ. Violacein Targets the Cytoplasmic Membrane of Bacteria. ACS Infect Dis. 2019;5(4):539–49.

    Article  Google Scholar 

  71. Ravindran A, Anishetty S, Pennathur G. Molecular dynamics of the membrane interaction and localisation of prodigiosin. J Mol Graph Model. 2020;98:107614.

    Article  Google Scholar 

  72. Kimyon Ö, Das T, Ibugo AI, Kutty SK, Ho KK, Tebben J, Kumar N, Manefield M. Serratia Secondary Metabolite Prodigiosin Inhibits Pseudomonas aeruginosa Biofilm Development by Producing Reactive Oxygen Species that Damage Biological Molecules. Front Microbiol. 2016;7:972.

  73. Melvin MS, Tomlinson JT, Saluta GR, Kucera GL, Lindquist N, Manderville RA. Double-Strand DNA Cleavage by Copper·Prodigiosin. J Am Chem Soc. 2000;122(26):6333–4.

    Article  Google Scholar 

  74. Hazarika DJ, Gautom T, Parveen A, Goswami G, Barooah M, Modi MK, Boro RC. Mechanism of interaction of an endofungal bacterium Serratia marcescens D1 with its host and non-host fungi. Plos One. 2020;15(4):e0224051.

    Article  Google Scholar 

  75. Clements T, Ndlovu T, Khan W. Broad-spectrum antimicrobial activity of secondary metabolites produced by Serratia marcescens strains. Microbiol Res. 2019;229:126329.

    Article  Google Scholar 

  76. Dhar Purkayastha G, Mangar P, Saha A, Saha D. Evaluation of the biocontrol efficacy of a Serratia marcescens strain indigenous to tea rhizosphere for the management of root rot disease in tea. Plos One. 2018;13(2):e0191761.

    Article  Google Scholar 

  77. Madison JD, Ouellette SP, Schmidt EL, Kerby JL. Serratia marcescens shapes cutaneous bacterial communities and influences survival of an amphibian host. Proc Biol Sci. 2019;286(1914):20191833.

    Google Scholar 

  78. Mamiya Y. Pathology of the Pine Wilt Disease Caused by Bursaphelenchus xylophilus. Ann Rev Phytopathol. 1983;21(1):201–20.

    Article  Google Scholar 

  79. Yoon KH, Lee TY, Moon JH, Choi SY, Choi YJ, Mitchell RJ, Il Lee J. Consumption of Oleic Acid During Matriphagy in Free-Living Nematodes Alleviates the Toxic Effects of the Bacterial Metabolite Violacein. Sci Rep. 2020;10(1):8087.

    Article  Google Scholar 

  80. Ballestriero F, Daim M, Penesyan A, Nappi J, Schleheck D, Bazzicalupo P, Di Schiavi E, Egan S. Correction: Antinematode Activity of Violacein and the Role of the Insulin/IGF-1 Pathway in Controlling Violacein Sensitivity in Caenorhabditis elegans. PLoS One. 2018;13(12):e0210026.

  81. Forouzanfar MH, Afshin A, Alexander LT, Anderson HR, Bhutta ZA, Biryukov S, Brauer M, Burnett R, Cercy K, Charlson FJ, et al. Global, regional, and national comparative risk assessment of 79 behavioural, environmental and occupational, and metabolic risks or clusters of risks, 1990–2015: a systematic analysis for the Global Burden of Disease Study 2015. Lancet. 2016;388(10053):1659–724.

    Article  Google Scholar 

  82. Montaner B, Navarro S, Piqué M, Vilaseca M, Martinell M, Giralt E, Gil J, Pérez-Tomás R. Prodigiosin from the supernatant of Serratia marcescens induces apoptosis in haematopoietic cancer cell lines. Br J Pharmacol. 2000;131(3):585–93.

    Article  Google Scholar 

  83. Llagostera E, Soto-Cerrato V, Montaner B, PÉRez-TomÁS R: Prodigiosin Induces Apoptosis by Acting on Mitochondria in Human Lung Cancer Cells. Ann New York Acad Sci 2003, 1010(1):178-181.

  84. Campàs C, Dalmau M, Montaner B, Barragán M, Bellosillo B, Colomer D, Pons G, Pérez-Tomás R, Gil J. Prodigiosin induces apoptosis of B and T cells from B-cell chronic lymphocytic leukemia. Leukemia. 2003;17(4):746–50.

    Article  Google Scholar 

  85. Diaz-Ruiz C, Montaner B, Perez-Tomas R. Prodigiosin induces cell death and morphological changes indicative of apoptosis in gastric cancer cell line HGT-1. Histol Histopathol. 2001;16(2):415–21.

    Google Scholar 

  86. Soto-Cerrato V, Llagostera E, Montaner B, Scheffer GL, Perez-Tomas R. Mitochondria-mediated apoptosis operating irrespective of multidrug resistance in breast cancer cells by the anticancer agent prodigiosin. Biochem Pharmacol. 2004;68(7):1345–52.

    Article  Google Scholar 

  87. Prabhu VV, Hong B, Allen JE, Zhang S, Lulla AR, Dicker DT, El-Deiry WS. Small-Molecule Prodigiosin Restores p53 Tumor Suppressor Activity in Chemoresistant Colorectal Cancer Stem Cells via c-Jun-Mediated ΔNp73 Inhibition and p73 Activation. Cancer Res. 2016;76(7):1989–99.

    Article  Google Scholar 

  88. Cheng S-Y, Chen N-F, Kuo H-M, Yang S-N, Sung C-S, Sung P-J, Wen Z-H, Chen W-F. Prodigiosin stimulates endoplasmic reticulum stress and induces autophagic cell death in glioblastoma cells. Apoptosis. 2018;23(5-6):314–28.

    Article  Google Scholar 

  89. Zhang J, Shen Y, Liu J, Wei D. Antimetastatic effect of prodigiosin through inhibition of tumor invasion. Biochem Pharmacol. 2005;69(3):407–14.

    Article  Google Scholar 

  90. Sessler JL, Eller LR, Cho W-S, Nicolaou S, Aguilar A, Lee JT, Lynch VM, Magda DJ. Synthesis, Anion-Binding Properties, and In Vitro Anticancer Activity of Prodigiosin Analogues. Angewandte Chemie Int Edition. 2005;44(37):5989–92.

    Article  Google Scholar 

  91. Zhao C, Qiu SZ, He J, Peng Y, Xu HM, Feng ZQ, Huang HL, Du YL, Zhou YJ, Nie YQ. Prodigiosin impairs autophagosome-lysosome fusion that sensitizes colorectal cancer cells to 5-fluorouracil-induced cell death. Cancer Lett. 2020;481:15–23.

    Article  Google Scholar 

  92. Montaner B, Castillo-Ávila W, Martinell M, Öllinger R, Aymami J, Giralt E, Pérez-Tomás R. DNA Interaction and Dual Topoisomerase I and II Inhibition Properties of the Anti-Tumor Drug Prodigiosin. Toxicol Sci. 2005;85(2):870–9.

    Article  Google Scholar 

  93. Anwar MM, Shalaby M, Embaby AM, Saeed H, Agwa MM, Hussein A. Prodigiosin/PU-H71 as a novel potential combined therapy for triple negative breast cancer (TNBC): preclinical insights. Sci Rep. 2020;10(1):14706.

  94. Ohkuma S, Sato T, Okamoto M, Matsuya H, Arai K, Kataoka T, Nagai K, Wasserman HH. Prodigiosins uncouple lysosomal vacuolar-type ATPase through promotion of H+/Cl− symport. Biochem J. 1998;334(3):731–41.

    Article  Google Scholar 

  95. Sato T, Konno H, Tanaka Y, Kataoka T, Nagai K, Wasserman HH, Ohkuma S. Prodigiosins as a New Group of H+/Cl−Symporters That Uncouple Proton Translocators. J Biol Chem. 1998;273(34):21455–62.

    Article  Google Scholar 

  96. Fürstner A, Reinecke K, Prinz H, Waldmann H. The Core Structures of Roseophilin and the Prodigiosin Alkaloids Define a New Class of Protein Tyrosine Phosphatase Inhibitors. ChemBioChem. 2004;5(11):1575–9.

    Article  Google Scholar 

  97. Soliev AB, Hosokawa K, Enomoto K. Effects of prodigiosin family compounds fromPseudoalteromonassp. 1020R on the activities of protein phosphatases and protein kinases. J Enzyme Inhib Med Chem. 2014;30(4):533–8.

    Article  Google Scholar 

  98. Xie Z, Klionsky DJ. Autophagosome formation: core machinery and adaptations. Nat Cell Biol. 2007;9(10):1102–9.

    Article  Google Scholar 

  99. Levy JMM, Towers CG, Thorburn A. Targeting autophagy in cancer. Nat Rev Cancer. 2017;17(9):528–42.

    Article  Google Scholar 

  100. Razaghi A, Heimann K, Schaeffer PM, Gibson SB. Negative regulators of cell death pathways in cancer: perspective on biomarkers and targeted therapies. Apoptosis. 2018;23(2):93–112.

    Article  Google Scholar 

  101. Mokhtari RB, Homayouni TS, Baluch N, Morgatskaya E, Kumar S, Das B, Yeger H. Combination therapy in combating cancer. Oncotarget. 2017;8(23):38022–43.

    Article  Google Scholar 

  102. Schimmer AD, Raza A, Carter TH, Claxton D, Erba H, DeAngelo DJ, Tallman MS, Goard C, Borthakur G. A multicenter phase I/II study of obatoclax mesylate administered as a 3- or 24-hour infusion in older patients with previously untreated acute myeloid leukemia. PLoS One. 2014;9(10):e108694.

  103. Langer CJ, Albert I, Ross HJ, Kovacs P, Blakely LJ, Pajkos G, Somfay A, Zatloukal P, Kazarnowicz A, Moezi MM, et al. Randomized phase II study of carboplatin and etoposide with or without obatoclax mesylate in extensive-stage small cell lung cancer. Lung Cancer. 2014;85(3):420–8.

    Article  Google Scholar 

  104. Oki Y, Copeland A, Hagemeister F, Fayad LE, Fanale M, Romaguera J, Younes A. Experience with obatoclax mesylate (GX15-070), a small molecule pan–Bcl-2 family antagonist in patients with relapsed or refractory classical Hodgkin lymphoma. Blood. 2012;119(9):2171–2.

    Article  Google Scholar 

  105. Arellano ML, Borthakur G, Berger M, Luer J, Raza A. A Phase II, Multicenter, Open-Label Study of Obatoclax Mesylate in Patients With Previously Untreated Myelodysplastic Syndromes With Anemia or Thrombocytopenia. Clin Lymphoma Myeloma Leukemia. 2014;14(6):534–9.

    Article  Google Scholar 

  106. Saraiva VS, Marshall JC, Cools-Lartigue J, Burnier MN. Cytotoxic effects of violacein in human uveal melanoma cell lines. Melanoma Res. 2004;14(5):421–4.

    Article  Google Scholar 

  107. Mehta T, Vercruysse K, Johnson T, Ejiofor AO, Myles E, Quick QA. Violacein induces p44/42 mitogen-activated protein kinase-mediated solid tumor cell death and inhibits tumor cell migration. Mol Med Rep. 2015;12(1):1443–8.

    Article  Google Scholar 

  108. de Carvalho DD, Costa FTM, Duran N, Haun M. Cytotoxic activity of violacein in human colon cancer cells. Toxicol in Vitro. 2006;20(8):1514–21.

    Article  Google Scholar 

  109. Kodach LL, Bos CL, Duran N, Peppelenbosch MP, Ferreira CV, Hardwick JCH. Violacein synergistically increases 5-fluorouracil cytotoxicity, induces apoptosis and inhibits Akt-mediated signal transduction in human colorectal cancer cells. Carcinogenesis. 2006;27(3):508–16.

    Article  Google Scholar 

  110. Masuelli L, Pantanella F, La Regina G, Benvenuto M, Fantini M, Mattera R, Di Stefano E, Mattei M, Silvestri R, Schippa S, et al. Violacein, an indole-derived purple-colored natural pigment produced by Janthinobacterium lividum, inhibits the growth of head and neck carcinoma cell lines both in vitro and in vivo. Tumor Biol. 2016;37(3):3705–17.

    Article  Google Scholar 

  111. Leal AM, de Queiroz JD, de Medeiros SR, Lima TK, Agnez-Lima LF. Violacein induces cell death by triggering mitochondrial membrane hyperpolarization in vitro. BMC Microbiol. 2015;15:115.

  112. Bromberg N, Dreyfuss JL, Regatieri CV, Palladino MV, Duran N, Nader HB, Haun M, Justo GZ. Growth inhibition and pro-apoptotic activity of violacein in Ehrlich ascites tumor. Chem-Biol Interact. 2010;186(1):43–52.

    Article  Google Scholar 

  113. Ferreira CV, Bos CL, Versteeg HH, Justo GZ, Duran N, Peppelenbosch MP. Molecular mechanism of violacein-mediated human leukemia cell death. Blood. 2004;104(5):1459–64.

    Article  Google Scholar 

  114. Melo PD, Maria SS, Vidal BD, Haun M, Duran N. Violacein cytotoxicity and induction of apoptosis in V79 cells. In Vitro Cell Dev-An. 2000;36(8):539–43.

    Article  Google Scholar 

  115. Goncalves PR, Rocha-Brito KJP, Fernandes MRN, Abrantes JL, Duran N, Ferreira-Halder CV. Violacein induces death of RAS-mutated metastatic melanoma by impairing autophagy process. Tumor Biol. 2016;37(10):14049–58.

    Article  Google Scholar 

  116. Queiroz KC, Milani R, Ruela-de-Sousa RR, Fuhler GM, Justo GZ, Zambuzzi WF, Duran N, Diks SH, Spek CA, Ferreira CV, Peppelenbosch MP. Violacein induces death of resistant leukaemia cells via kinome reprogramming, endoplasmic reticulum stress and Golgi apparatus collapse. PLoS One. 2012;7(10):e45362.

  117. Balibar CJ, Walsh CT. In vitro biosynthesis of violacein from L-tryptophan by the enzymes VioA-E from Chromobacterium violaceum. Biochemistry. 2006;45(51):15444-57.

  118. Platt D, Amara S, Mehta T, Vercuyssee K, Myles EL, Johnson T, Tiriveedhi V. Violacein inhibits matrix metalloproteinase mediated CXCR4 expression: Potential anti-tumor effect in cancer invasion and metastasis. Biochem Bioph Res Co. 2014;455(1-2):107–12.

    Article  Google Scholar 

  119. Liekens S, Schols D, Hatse S. CXCL12-CXCR4 Axis in Angiogenesis, Metastasis and Stem Cell Mobilization. Curr Pharm Design. 2010;16(35):3903–20.

    Article  Google Scholar 

  120. Antonisamy P, Kannan P, Aravinthan A, Duraipandiyan V, Arasu MV, Ignacimuthu S, Al-Dhabi NA, Kim JH. Gastroprotective activity of violacein isolated from Chromobacterium violaceum on indomethacin-induced gastric lesions in rats: investigation of potential mechanisms of action. ScientificWorldJournal. 2014;2014:616432.

  121. Han SB, Kim HM, Kim YH, Lee CW, Jang E-S, Son KH, Kim SU, Kim YK. T-cell specific immunosuppression by prodigiosin isolated from Serratia marcescens. Int J Immunopharmacol. 1998;20(1-3):1–13.

    Article  Google Scholar 

  122. Han SB, Park SH, Jeon YJ, Kim YK, Kim HM, Yang KH. Prodigiosin blocks T cell activation by inhibiting interleukin-2Ralpha expression and delays progression of autoimmune diabetes and collagen-induced arthritis. J Pharmacol Exp Ther. 2001;299(2):415–25.

    Google Scholar 

  123. Mortellaro A, Songia S, Gnocchi P, Ferrari M, Fornasiero C, D'Alessio R, Isetta A, Colotta F, Golay J. New immunosuppressive drug PNU156804 blocks IL-2-dependent proliferation and NF-kappa B and AP-1 activation. J Immunol. 1999;162(12):7102–9.

    Article  Google Scholar 

  124. Han S-B, Lee CW, Yoon YD, Kang JS, Lee KH, Yoon WK, Kim YK, Lee K, Park S-K, Kim HM. Effective prevention of lethal acute graft-versus-host disease by combined immunosuppressive therapy with prodigiosin and cyclosporine A. Biochem Pharmacol. 2005;70(10):1518–26.

    Article  Google Scholar 

  125. Huh J-E, Koo H-J, Kim K-H, Yim J-H, Lee H-K, Sohn E-W, Pyo S-N. Immunosuppressive Effect of Prodigiosin on Murine Splenocyte and Macrophages. Biomol Ther. 2008;16(4):351–5.

    Article  Google Scholar 

  126. Antonisamy P, Ignacimuthu S. Immunomodulatory, analgesic and antipyretic effects of violacein isolated from Chromobacterium violaceum. Phytomed. 2010;17(3-4):300–4.

    Article  Google Scholar 

  127. Verinaud L, Lopes SC, Prado IC, Zanucoli F, Alves da Costa T, Di Gangi R, Issayama LK, Carvalho AC, Bonfanti AP, Niederauer GF, Duran N, Costa FT, Oliveira AL, Höfling MA, Machado DR, Thomé R. Violacein Treatment Modulates Acute and Chronic Inflammation through the Suppression of Cytokine Production and Induction of Regulatory T Cells. PLoS One. 2015;10(5):e0125409.

  128. Alshatwi AA, Subash-Babu P, Antonisamy P. Violacein induces apoptosis in human breast cancer cells through up regulation of BAX, p53 and down regulation of MDM2. Exp Toxicol Pathol. 2016;68(1):89–97.

    Article  Google Scholar 

  129. Venegas FA, Köllisch G, Mark K, Diederich WE, Kaufmann A, Bauer S, Chavarría M, Araya JJ, García-Piñeres AJ. The Bacterial Product Violacein Exerts an Immunostimulatory Effect Via TLR8. Sci Rep. 2019;9(1):13661.

  130. Domröse A, Klein AS, Hage-Hülsmann J, Thies S, Svensson V, Classen T, Pietruszka J, Jaeger KE, Drepper T, Loeschcke A. Efficient recombinant production of prodigiosin in Pseudomonas putida. Front Microbiol. 2015;6:972.

  131. Gallardo K, Candia JE, Remonsellez F, Escudero LV, Demergasso CS. The Ecological Coherence of Temperature and Salinity Tolerance Interaction and Pigmentation in a Non-marine Vibrio Isolated from Salar de Atacama. Front Microbiol. 2016;7:1943.

  132. Song MJ, Bae J, Lee DS, Kim CH, Kim JS, Kim SW, Hong SI. Purification and characterization of prodigiosin produced by integrated bioreactor from Serratia sp KH-95. J Biosci Bioeng. 2006;101(2):157–61.

    Article  Google Scholar 

  133. Williams RP, Gott CL, Green JA. Studies on Pigmentation of Serratia Marcescens V. J Bacteriol. 1961;81(3):376–9.

    Article  Google Scholar 

  134. Elahian F, Moghimi B, Dinmohammadi F, Ghamghami M, Hamidi M, Mirzaei SA. The Anticancer Agent Prodigiosin Is Not a Multidrug Resistance Protein Substrate. DNA and Cell Biology. 2013;32(3):90–7.

    Article  Google Scholar 

  135. Mendes AS, de Carvalho JE, Duarte MCT, Duran N, Bruns RE. Factorial design and response surface optimization of crude violacein for Chromobacterium violaceumi production. Biotechnol Lett. 2001;23(23):1963–9.

    Article  Google Scholar 

  136. Pantanella F, Berlutti F, Passariello C, Sarli S, Morea C, Schippa S. Violacein and biofilm production in Janthinobacterium lividum. J Appl Microbiol. 2007;102(4):992–9.

    Google Scholar 

  137. Rettori D, Duran N. Production, extraction and purification of violacein: an antibiotic pigment produced by Chromobacterium violaceum. World J Microb Biot. 1998;14(5):685–8.

    Article  Google Scholar 

  138. Wang HS, Jiang PX, Lu Y, Ruan ZY, Jiang RB, Xing XH, Lou K, Wei D. Optimization of culture conditions for violacein production by a new strain of Duganella sp B2. Biochem Eng J. 2009;44(2-3):119–24.

    Article  Google Scholar 

  139. DeMoss RD. Violacein. In: Biosynthesis; 1967. p. 77–81.

    Google Scholar 

  140. Rodrigues AL, Trachtmann N, Becker J, Lohanatha AF, Blotenberg J, Bolten CJ, Korneli C, Lima AOD, Porto LM, Sprenger GA, et al. Systems metabolic engineering of Escherichia coli for production of the antitumor drugs violacein and deoxyviolacein. Metab Eng. 2013;20:29–41.

    Article  Google Scholar 

  141. Choi SY, Yoon KH, Lee JI, Mitchell RJ. Violacein: Properties and Production of a Versatile Bacterial Pigment. Biomed Res Int. 2015;2015:465056.

  142. Duran N, Justo GZ, Duran M, Brocchi M, Cordi L, Tasic L, Castro GR, Nakazato G. Advances in Chromobacterium violaceum and properties of violacein-Its main secondary metabolite: A review. Biotechnol Adv. 2016;34(5):1030–45.

    Article  Google Scholar 

  143. Mun W, Kwon H, Im H, Choi SY, Monnappa AK, Mitchell RJ. Cyanide Production by Chromobacterium piscinae Shields It from Bdellovibrio bacteriovorus HD100 Predation. mBio. 2017;8(6):e01370-17.

  144. Cybulski O, Dygas M, Mikulak-Klucznik B, Siek M, Klucznik T, Choi SY, Mitchell RJ, Sobolev YI, Grzybowski BA. Concentric liquid reactors for chemical synthesis and separation. Nature. 2020;586(7827):57–63.

    Article  Google Scholar 

  145. Lin C, Jia X, Fang Y, Chen L, Zhang H, Lin R, Chen J. Enhanced production of prodigiosin by Serratia marcescens FZSF02 in the form of pigment pellets. Electron J Biotechn. 2019;40:58–64.

    Article  Google Scholar 

  146. Kurbanoglu EB, Ozdal M, Ozdal OG, OF A. Enhanced production of prodigiosin by Serratia marcescens MO-1 using ram horn peptone. Brazilian J Microbiol. 2015;46(2):631–7.

    Article  Google Scholar 

  147. Casullo de Araújo HW, Fukushima K, GMC T. Prodigiosin Production by Serratia marcescens UCP 1549 Using Renewable-Resources as a Low Cost Substrate. Molecules. 2010;15(10):6931–40.

    Article  Google Scholar 

  148. Aruldass CA, Venil CK, Zakaria ZA, Ahmad WA. Brown sugar as a low-cost medium for the production of prodigiosin by locally isolated Serratia marcescens UTM1. Int Biodeterioration Biodegradation. 2014;95:19–24.

    Article  Google Scholar 

  149. Siva R, Subha K, Bhakta D, Ghosh AR, Babu S. Characterization and Enhanced Production of Prodigiosin from the Spoiled Coconut. Appl Biochem Biotechnol. 2011;166(1):187–96.

    Article  Google Scholar 

  150. Koo B-S, Hahn B-S, Sim J-S, Lee C-M, Kwon S-W, Baek H-J, Yoon S-H. Production of Violacein by a Novel Bacterium, Massilia sp. EP15224 Strain. Korean J Microbiol Biotechnol. 2014;42(4):317–23.

    Article  Google Scholar 

  151. Ahmad WA, Yusof NZ, Nordin N, Zakaria ZA, Rezali MF. Production and Characterization of Violacein by Locally Isolated Chromobacterium violaceum Grown in Agricultural Wastes. Appl Biochem Biotechnol. 2012;167(5):1220–34.

    Article  Google Scholar 

  152. Elkenawy NM, Yassin AS, Elhifnawy HN, Amin MA. Optimization of prodigiosin production by Serratia marcescens using crude glycerol and enhancing production using gamma radiation. Biotechnol Rep. 2017;14:47–53.

    Article  Google Scholar 

  153. Harris AKP, Williamson NR, Slater H, Cox A, Abbasi S, Foulds I, Simonsen HT, Leeper FJ, Salmond GPC. The Serratia gene cluster encoding biosynthesis of the red antibiotic, prodigiosin, shows species- and strain-dependent genome context variation. Microbiology. 2004;150(11):3547–60.

    Article  Google Scholar 

  154. Williamson NR, Simonsen HT, Ahmed RAA, Goldet G, Slater H, Woodley L, Leeper FJ, Salmond GPC. Biosynthesis of the red antibiotic, prodigiosin, in Serratia: identification of a novel 2-methyl-3-n-amyl-pyrrole (MAP) assembly pathway, definition of the terminal condensing enzyme, and implications for undecylprodigiosin biosynthesis in Streptomyces. Mol Microbiol. 2005;56(4):971–89.

    Article  Google Scholar 

  155. Williams RP, Goldschmidt ME, Gott CL. Inhibition by temperature of the terminal step in biosynthesis of prodigiosin. Biochem Bioph Res Co. 1965;19(2):177–81.

    Article  Google Scholar 

  156. Morrison DA. Prodigiosin synthesis in mutants of Serratia marcesens. J Bacteriol. 1966;91(4):1599–604.

    Article  Google Scholar 

  157. Qadri SMH, Williams RP. Role of Methionine in Biosynthesis of Prodigiosin by Serratia marcescens. J Bacteriol. 1973;116(3):1191–8.

    Article  Google Scholar 

  158. Fineran PC, Slater H, Everson L, Hughes K, Salmond GPC. Biosynthesis of tripyrrole and β-lactam secondary metabolites inSerratia: integration of quorum sensing with multiple new regulatory components in the control of prodigiosin and carbapenem antibiotic production. Mol Microbiol. 2005;56(6):1495–517.

    Article  Google Scholar 

  159. Slater H, Crow M, Everson L, Salmond GPC. Phosphate availability regulates biosynthesis of two antibiotics, prodigiosin and carbapenem, in Serratia via both quorum-sensing-dependent and -independent pathways. Mol Microbiol. 2008;47(2):303–20.

    Article  Google Scholar 

  160. Thomson NR, Crow MA, McGowan SJ, Cox A, Salmond GPC. Biosynthesis of carbapenem antibiotic and prodigiosin pigment in Serratia is under quorum sensing control. Mol Microbiol. 2002;36(3):539–56.

    Article  Google Scholar 

  161. Coulthurst SJ, Kurz CL, GPC S. luxS mutants of Serratia defective in autoinducer-2-dependent ‘quorum sensing’ show strain-dependent impacts on virulence and production of carbapenem and prodigiosin. Microbiology. 2004;150(6):1901–10.

    Article  Google Scholar 

  162. Pemberton JM, Vincent KM, Penfold RJ. Cloning and Heterologous Expression of the Violacein Biosynthesis Gene-Cluster from Chromobacterium-Violaceum. Curr Microbiol. 1991;22(6):355–8.

    Article  Google Scholar 

  163. Sanchez C, Brana AF, Mendez C, Salas JA. Reevaluation of the violacein biosynthetic pathway and its relationship to indolocarbazole biosynthesis. Chembiochem. 2006;7(8):1231–40.

    Article  Google Scholar 

  164. Balibar CJ, Walsh CT. In vitro biosynthesis of violacein from L-tryptophan by the enzymes VioA-E from Chromobacterium violaceum. Biochemistry-Us. 2006;45(51):15444–57.

    Article  Google Scholar 

  165. August PR, Grossman TH, Minor C, Draper MP, MacNeil IA, Pemberton JM, Call KM, Holt D, Osburne MS. Sequence analysis and functional characterization of the violacein biosynthetic pathway from Chromobacterium violaceum. J Mol Microbiol Biotechnol. 2000;2(4):513–9.

    Google Scholar 

  166. McClean KH, Winson MK, Fish L, Taylor A, Chhabra SR, Camara M, Daykin M, Lamb JH, Swift S, Bycroft BW, et al. Quorum sensing and Chromobacterium violaceum: exploitation of violacein production and inhibition for the detection of N-acylhomoserine lactones. Microbiology. 1997;143(12):3703–11.

    Article  Google Scholar 

  167. Wang Y, Ikawa A, Okaue S, Taniguchi S, Osaka I, Yoshimoto A, Kishida Y, Arakawa R, Enomoto K. Quorum sensing signaling molecules involved in the production of violacein by Pseudoalteromonas. Biosci Biotechnol Biochem. 2008;72(7):1958–61.

    Article  Google Scholar 

  168. Fukumoto A, Murakami C, Anzai Y, Kato F. Maniwamycins: new quorum-sensing inhibitors against Chromobacterium violaceum CV026 were isolated from Streptomyces sp. TOHO-M025. J Antibiot (Tokyo). 2016;69(5):395–9.

    Article  Google Scholar 

  169. Ohta T, Fukumoto A, Iizaka Y, Kato F, Koyama Y, Anzai Y. Quorum Sensing Inhibitors against Chromobacterium violaceum CV026 Derived from an Actinomycete Metabolite Library. Biol Pharm Bull. 2020;43(1):179–83.

    Article  Google Scholar 

  170. Harrison AM, Soby SD. Reclassification of Chromobacterium violaceum ATCC 31532 and its quorum biosensor mutant CV026 to Chromobacterium subtsugae. AMB Express. 2020;10(1):202.

    Article  Google Scholar 

  171. Su C, Xiang Z, Liu Y, Zhao X, Sun Y, Li Z, Li L, Chang F, Chen T, Wen X, Zhou Y, Zhao F. Analysis of the genomic sequences and metabolites of Serratia surfactantfaciens sp. nov. YD25T that simultaneously produces prodigiosin and serrawettin W2. BMC Genomics. 2016;17(1):865.

  172. Li P, Kwok AH, Jiang J, Ran T, Xu D, Wang W, Leung FC. Comparative genome analyses of Serratia marcescens FS14 reveals its high antagonistic potential. PLoS One. 2015;10(4):e0123061.

  173. Smith HJ, Foreman CM, Akiyama T, Franklin MJ, Devitt NP, Ramaraj T. Genome Sequence of Janthinobacterium sp. CG23_2, a Violacein-Producing Isolate from an Antarctic Supraglacial Stream. Genome Announc. 2016;4(1):e01468-15.

  174. Valdes N, Soto P, Cottet L, Alarcon P, Gonzalez A, Castillo A, Corsini G, Tello M. Draft genome sequence of Janthinobacterium lividum strain MTR reveals its mechanism of capnophilic behavior. Stand Genomic Sci. 2015;10:110.

    Article  Google Scholar 

  175. Wu X, Deutschbauer AM, Kazakov AE, Wetmore KM, Cwick BA, Walker RM, Novichkov PS, Arkin AP, Chakraborty R. Draft Genome Sequences of Two Janthinobacteriumlividum Strains, Isolated from Pristine Groundwater Collected from the Oak Ridge Field Research Center. Genome Announc. 2017;5(26):e00582-17.

  176. Doing G, Perron GG, Jude BA. Draft Genome Sequence of a Violacein-Producing Iodobacter sp. from the Hudson Valley Watershed. Genome Announc. 2018;6(1):e01428-17.

  177. Jude BA. Draft Genome Sequence of a Chitinimonas Species from Hudson Valley Waterways That Expresses Violacein Pigment. Microbiol Resour Announc. 2019;8(35):e00683-19.

  178. Lamendella R, Jude BA. Draft Genome Sequences of Violacein-Producing Duganella sp. Isolates from a Waterway in Eastern Pennsylvania. Microbiol Resour Announc. 2018;7(12):e01196-18.

  179. Bettina AM, Doing G, O'Brien K, Perron GG, Jude BA. Draft Genome Sequences of Phenotypically Distinct Janthinobacterium sp. Isolates Cultured from the Hudson Valley Watershed. Genome Announc. 2018;6(3):e01426-17.

  180. O'Brien K, Perron GG, Jude BA. Draft Genome Sequence of a Red-Pigmented Janthinobacterium sp. Native to the Hudson Valley Watershed. Genome Announc. 2018;6(1):e01429-17.

  181. Kwon SK, Park YK, Kim JF. Genome-wide screening and identification of factors affecting the biosynthesis of prodigiosin by Hahella chejuensis, using Escherichia coli as a surrogate host. Appl Environ Microbiol. 2010;76(5):1661–8.

    Article  Google Scholar 

  182. Dauenhauer SA, Hull RA, Williams RP. Cloning and expression in Escherichia coli of Serratia marcescens genes encoding prodigiosin biosynthesis. J Bacteriol. 1984;158(3):1128–32.

    Article  Google Scholar 

  183. Fang MY, Zhang C, Yang S, Cui JY, Jiang PX, Lou K, Wachi M, Xing XH. High crude violacein production from glucose by Escherichia coli engineered with interactive control of tryptophan pathway and violacein biosynthetic pathway. Microb Cell Fact. 2015;14(1):8.

  184. Immanuel SRC, Banerjee D, Rajankar MP, Raghunathan A. Integrated constraints based analysis of an engineered violacein pathway in Escherichia coli. Biosystems. 2018;171:10–9.

    Article  Google Scholar 

  185. Rodrigues AL, Göcke Y, Bolten C, Brock NL, Dickschat JS, Wittmann C. Microbial production of the drugs violacein and deoxyviolacein: analytical development and strain comparison. Biotechnol Lett. 2011;34(4):717–20.

    Article  Google Scholar 

  186. Wilkinson MD, Lai HE, Freemont PS, Baum J. A Biosynthetic Platform for Antimalarial Drug Discovery. Antimicrob Agents Chemother. 2020;64(5):e02129-19.

  187. C-Y H, Guo Y, Liu L, Zhang N-X, Gao C-X, Yang X-Q, Yi J. Genetic control of violacein biosynthesis to enable a pigment-based whole-cell lead biosensor. Rsc Adv. 2020;10(47):28106–13.

    Article  Google Scholar 

  188. Jiang Y, Sigmund F, Reber J, Deán-Ben XL, Glasl S, Kneipp M, Estrada H, Razansky D, Ntziachristos V, Westmeyer GG. Violacein as a geneticallycontrolled, enzymatically amplified and photobleaching-resistant chromophore for optoacoustic bacterial imaging. Sci Rep. 2015;5:11048.

  189. Jeschek M, Gerngross D, Panke S. Rationally reduced libraries for combinatorial pathway optimization minimizing experimental effort. Nat Commun. 2016;7:11163.

  190. Chuang J, Boeke JD, Mitchell LA. Coupling Yeast Golden Gate and VEGAS for Efficient Assembly of the Violacein Pathway in Saccharomyces cerevisiae. In: Synthetic Metabolic Pathways; 2018. p. 211–25.

    Chapter  Google Scholar 

  191. Tong Y, Zhou J, Zhang L, Xu P. A Golden-Gate Based Cloning Toolkit to Build Violacein Pathway Libraries in Yarrowia lipolytica. ACS Synthetic Biol. 2021;10(1):115–24.

    Article  Google Scholar 

  192. NCBI PubMed https://pubmed.ncbi.nlm.nih.gov/. Accessed 20 Jan 2021.

Download references

Acknowledgements

Funding was provided through the National Research Foundation of Korea under the Mid-Career Project (Grant No. 2020R1A2C2012158), the General Research Project (Grant No. 2020R1F1A1073848), and the Early Career Research Project (Grant No. 2019R1C1C1008708). We appreciate the support.

Funding

Funding was provided through the National Research Foundation of Korea under the Mid-Career Project (Grant No. 2020R1A2C2012158), the General Research Project (Grant No. 2020R1F1A1073848), and the Early Career Research Project (Grant No. 2019R1C1C1008708). We appreciate the support.

Author information

Authors and Affiliations

Authors

Contributions

SYC, SL, KHY, JIL and RJM wrote the article. The author(s) read and approved the final manuscript.

Corresponding authors

Correspondence to Kyoung-hye Yoon, Jin I. Lee or Robert J. Mitchell.

Ethics declarations

Ethics approval and consent to participate

None.

Consent for publication

All authors agree with publishing this article within The Journal of Biological Engineering

Competing interests

The authors declare no competing interests with the publishing of this article

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Choi, S.Y., Lim, S., Yoon, Kh. et al. Biotechnological Activities and Applications of Bacterial Pigments Violacein and Prodigiosin. J Biol Eng 15, 10 (2021). https://0-doi-org.brum.beds.ac.uk/10.1186/s13036-021-00262-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s13036-021-00262-9

Keywords