eISSN: 2353-9461
ISSN: 0860-7796
BioTechnologia
Current issue Archive About the journal Editorial board Abstracting and indexing Subscription Contact Instructions for authors Publication charge Ethical standards and procedures
Editorial System
Submit your Manuscript
3/2023
vol. 104
 
Share:
Share:
REVIEW PAPER

Microbial canthaxanthin: an orange-red keto carotenoid with potential pharmaceutical applications

Vinita Gaur
1
,
Surojit Bera
1

1.
Department of Microbiology, School of Bioengineering and Biosciences, Lovely Professional University, Punjab, India
BioTechnologia vol. 104(3) ∙ 315 – 328 ∙ 2023
Online publish date: 2023/09/25
Article file
Get citation
 
PlumX metrics:
 

Introduction

Carotenoids are pigments naturally found in subcellular structures called chloroplasts of photosynthetic organisms, as well as certain bacteria and fungi. Several carotenoids have been characterized to date (Rebelo et al., 2020). These compounds are considered major pigment categories and are produced as secondary metabolites in various fruits, vegetables, and a few microbes (Bera, 2019). Carotenes are a class of hydrocarbons that are not combined with other elements, constituting less than 10% of the total carotenoid species. On the other hand, xanthophylls are characterized by the presence of oxygen atoms, and their side-chain derivatives may include hydroxy-, keto-, methoxy-, epoxy-, or carboxyl groups (Mordi et al., 2020). Canthaxanthin, a prominent keto-carotenoid, is found in various plants and is responsible for the coloring of fruits, vegetables, and flowers. It is also present in green algae, bacteria, crustaceans, fish, and vertebrates (Bera, 2020). Canthaxanthin exhibits potent oxygen-quenching and free radical deactivating properties (Dewanjee et al., 2021). The biological synthesis of canthaxanthin occurs through the activity of β-C-4 oxygenase (β-carotene ketolase), an enzyme involved in catalyzing the conversion of β-carotene into canthaxanthin (Rebelo et al., 2020). Canthaxanthin has the molecular formula C40H52O2, a molar mass of 564.84 g/mol, and the CAS number 514-78-3. This symmetrical ketocarotenoid contains two conjugated carbonyl groups, and its spectrum exhibits a broad symmetrical peak with a maximum at 466 nm in petroleum ether (%III/II = 0) (Rodriguez-Amaya, 2001). While canthaxanthin can be synthesized artificially, there is a growing demand for naturally derived canthaxanthin and other commercially valuable carotenoids obtained from plants and various other microbes. The global market for carotenoid pigments has been steadily increasing due to consumer preference for naturally sourced active components over synthetic products (Torregrosa-Crespo et al., 2018).

In 2018, the global carotenoid industry was valued at approximately USD 1.40 billion. It is projected to reach USD 1.85 billion by the end of 2026, with a compound yearly growth rate of 3.57%. Between 2007 and 2017, there was a significant 1500-fold increase in the utilization of carotenoids worldwide. In the commercial food market, carotenoids accounted for 1609.8 metric tons, while nonfood-based products accounted for 63.0 metric tons (Yaqoob et al., 2021). A recent study suggests that there will be a notable surge in global canthaxanthin trade, with an expected growth from USD 64220 million in 2019 to USD 69100 million by 2025, representing a compound annual growth rate of 1.85% (Fig. 1) (https://www.wboc.com).

Fig. 1

Prospective growth of the global canthaxanthin market between the years 2019 and 2025 (https://www.wboc.com)

/f/fulltexts/BTA/51306/BTA-104-3-51306-g001_min.jpg

The canthaxanthin market is experiencing continuous growth primarily driven by the increasing demand for a rational and regular source of coloring components.

The market expansion is attributed to factors such as the widespread availability of canthaxanthin, its various medical benefits, and its established role as a standard coloring agent.

Canthaxanthin holds a significant share of the carotenoid market, thanks to its extensive use as a coloring agent in feed supplements, food additives, personal care products, and pharmaceutical applications (https://www.globenewswire.com/news). Currently, there are several commercially available canthaxanthin products on the market (Table 2).

Table 1
SourcesApplicationsMarket localization
Natural sources
 microbes (bacteria, algae, fungi)
 animals (crustaceans, few fishes & birds)
 genetically modified organisms
Synthetic sources
 by chemical synthesis
As supplement, feed, food & in cosmetics
Expected role in:
  1. inherited blood disorder erythropoietic protoporphyria (EPP)

  2. an autoimmune disorder cutaneous lupus erythematosus (CLE)

  3. skin rashes, skin redness and irritation

  4. a skin discoloration disorder (vitiligo)

  5. sun sensitivity as a result of the use of specific medications, itching resulting from excessive sunlight, artificial tanning, etc.

North America
Europe (largest marcet allocation)
Asia Pacific
Rest of the world
Table 2

Companies associated with the commercial production of canthaxanthin

CompanyCommercialized productCountry of originURL
BASFLucantin® redGermanyhttps://nutrition.basf.com
Allied BiotechCanthatene® CanthaxanthinTaiwanhttps://www.altratene.com
Dynadis SARLMIXDYN®Francehttps://www.dynadis.com
Ardap Care GmbHQuiko Intensive RedGermanyhttps://ardapcare.com
Ardap Care GmbHQuikocarophyllGermanyhttps://ardapcare.com
The Nekton GmbHNekton-RGermanyhttps://allbirdproducts.com
Taiyo GroupPetslife Intensive RedIndiahttps://www.taiyogroup.in
DSMCarophyll® Red 10%USAhttps://www.dsm.com
ZMCCanthaxanthin Beadlet 10%USAhttps://www.ulprospector.com
NovephaSell CanthaxanthinChinahttps://detail.en.china.cn
ParchemCanthaxanthinUSAhttps://www.parchem.com

This article provides an extensive review of recently discovered sources of canthaxanthin and their vital applications in pharmacology and as feed additives. Special attention is given to the utilization of canthaxanthin and other carotenoids as feed ingredients for fish and poultry, considering their growing market importance. The article offers a comprehensive explanation of the past, present, and future perspectives of canthaxanthin.

Structure and isomerism of canthaxanthin

Carotenoids, from a structural perspective, exhibit a geometric pattern where two residual parts are connected by a double bond, either in the same direction (forming the E form) or in the opposite direction (forming the Z form) with respect to the plane (Moldenhauer et al., 2023). Isomers, aside from having differences in melting points, stability, and solubility, also display significant differences in absorption affinity, color, and color intensity (Venugopal, 2008).

A canthaxanthin molecule consists of a polyene chain comprising 40 carbon atoms, with double bonds and two unreacted hydroxyls (OH) groups located at each terminal ring. Notably, there is an extended double-bond system centrally positioned in the molecule. The electron density is higher towards the terminal end of the chain, while β-electrons are delocalized throughout the entire polyene chain (Bart and MacGillavry, 1968). By modifying the polyene skeleton of canthaxanthin through processes such as cyclization, atom rearrangement, chain shortening, hydrogenation, oxygen incorporation, or dehydrogenation, a wide range of structures can be generated. The β-rings of canthaxanthin form dihedral angles of 43º at their ends. The two keto groups in a canthaxanthin molecule are approximately 2.7 nm apart (Bart and MacGillavry, 1968).

Canthaxanthin is readily soluble in small volumes of water, lipids, or a mixture of organic solvent and water, and it can form molecular aggregates. In nature, all-trans-canthaxanthin, along with 9-cis- and 13-cis-canthaxanthin, has been reported to occur. Typically, HPLC is used for the separation and quantification of canthaxanthin isomers, while the technique of High-performance liquid chromatography-Atmospheric pressure chemical ionization-Mass spectrometry (HPLC–APCI–MS) is a powerful tool for confirming the presence of canthaxanthin (Schlatterer and Breithaupt, 2006). The isomerization process of canthaxanthin has gained increasing importance, and various factors such as temperature, acid, and ions have been found to influence it (Qiu et al., 2014). High temperature and light are known to influence the isomerization of carotenoids, transforming them into cis forms, which can also alter their biological functions (Bera and Dutta, 2017). Successful separation of 13 canthaxanthin isomers has been achieved using a column consisting of calcium hydroxide, and the identification of seven isomers has been accomplished using the proton nuclear magnetic resonance (1H-NMR) technique (Hashimoto et al., 1988). Among the isomers, 9-cis-canthaxanthin exhibits higher proapoptotic activity compared to all-trans-canthaxanthin (Venugopalan et al., 2013). Modern methods for the separation of canthaxanthin isomers or impurities involve the use of C30 and Si60-HPLC–APCI–MS combined with a diode array detector (DAD), which yields satisfactory results (Qiu et al., 2014). Thus far, some of the identified isomer compositions using these approaches include 9,13'-di-cis-, 13,15-di-cis-, 9,15-di-cis-, 15-cis-, 13-cis-, 9,13-di-cis-, all-trans-, 11-cis-, 7-cis-, and 9-cis-canthaxanthin (Qiu et al., 2014).

Synthesis of canthaxanthin

Chemical synthesis

Chemical synthesis of carotenoids, including canthaxanthin, involves several important factors such as procedure complexity, availability of new chemical routes, potential immune-boosting effects, and undesired side effects upon consumption (Ausich, 1997). A commonly used petrochemical-based material called “ketoisophorone” serves as a precursor for synthesizing xanthophylls. Among natural carotenoids, canthaxanthin and eight other carotenoids are utilized for large-scale industrial production (Ernst, 2002).

Canthaxanthin was historically produced commercially from β-carotene, following the method initially conceived by Karrer. The process involves allylic bromination of β-carotene, followed by solvolysis to form 4,4!-bis acetate. Subsequent hydrolysis and oxidation reactions lead to the formation of canthaxanthin (Rosenberg et al., 1979). In the late 1980s and early 1990s, Wittig olefination became an important technique for synthesizing polyenes, including carotenoids, despite the formation of unwanted by-products. Though carbon–carbon double bond formation remains a harsh but probing challenge in carotenoid synthesis, new methods were established like the aldehyde–sulfone route, to meet market demand as well as to avoid complexity (Bernhard and Mayer, 1991).

A novel method for canthaxanthin synthesis was described, involving the oxidation of 4,4′-dihydroxyl-β-carotene, which was synthesized from the condensation of C15-phosphonate 2 with C10-trienedial 3. The critical intermediate, C15-phosphonate 2, was synthesized in four steps using α-ionone as the starting compound (Pi et al., 2020). In another study, canthaxanthin was synthesized from β-carotene, and its properties were evaluated by synthesizing three different diodes using various combinations of carotene, canthaxanthin, silicon, and aluminum (Uzun et al., 2022).

Biological synthesis

In order to address growing environmental concerns and increase consumer awareness, the process of chemical synthesis, which often involves creating new isomers, is being scrutinized (Scaife et al., 2012). The health concerns associated with synthetic carotenoids have led to rising demand for naturally produced alternatives (Saini et al., 2019). One significant alternative method for canthaxanthin production is the utilization of various microbial species, offering a more sustainable approach compared to chemical synthesis (Chen et al., 2021). Exploring the secondary metabolic pathways of microorganisms can unveil new structural metabolites that can be produced at a minimal cost. Chlorella zofingiensis, a microalga, is widely used for canthaxanthin production. Additionally, the bacterium Gordonia jacobea is employed for industrial-scale canthaxanthin production (https://www.bccresearch.com/title). Moreover, noncarotenogenic microbes such as Blakeslea trispora, Saccharomyces cerevisiae, and Escherichia coli are being genetically engineered for the commercial production of carotenoids (Saini et al., 2019).

Canthaxanthin, an isoprenoid compound, is synthesized through the conversion of two basic isoprene compounds, namely isopentenyl diphosphate and dimethylallyl diphosphate. These compounds undergo condensation with geranyl diphosphate to form a C10 compound. Subsequently, this C10 compound reacts to produce geranylgeranyl pyrophosphate, a type of pyrophosphate (Wang et al., 2007). Cyclic carotenoids such as zeta-carotene, neurosporene, and lycopene are formed through the action of various desaturase enzymes. Enzymes located in the cellular membrane, such as phytoene desaturase and lycopene β-cyclase, catalyze successive steps in the synthesis of colored carotenoids, leading to the abundant formation of β-carotene (Tian and Hua, 2010).

The biosynthetic pathway for canthaxanthin production involves the enzymatic conversions of precursor molecules. The enzymes β-carotene ketolase (BKT) and β-carotene hydroxylase play crucial roles in these conversions.

In the case of canthaxanthin, β-carotene is converted to canthaxanthin through a series of steps. Initially, β-carotene is converted to echinenone by the action of β-carotene ketolase. β-Carotene ketolase introduces keto groups at C4 with or without hydroxylation at the C3 position. Echinenone is then further modified to form canthaxanthin through additional oxidations.

Another pathway involving β-carotene hydroxylase leads to the production of zeaxanthin. β-Carotene is converted to zeaxanthin via an intermediate called β-cryptoxanthin through the action of β-carotene hydroxylase.

High-yielding microalgae such as Haematococcus pluvialis (for astaxanthin) and Chlorella zofingiensis (for canthaxanthin) have been extensively studied for their commercial production of carotenoids. These microalgae possess well-characterized biosynthetic pathways that allow for the efficient production of specific carotenoids (Saini et al., 2019). The biosynthetic pathway for canthaxanthin production is depicted in Figure 2.

Fig. 2

Enzymes of carotenoid metabolic pathway: A) GGPP synthase; B) Phytoene synthase; C–F) Desaturases; G) Lycopene cyclase; H) β-Carotene oxygenase; I) β-Carotene ketolase; J, K) β-Carotene hydroxylase

/f/fulltexts/BTA/51306/BTA-104-3-51306-g002_min.jpg

Sources of canthaxanthin

Microbial sources

Isolation of canthaxanthin was first reported in the mid-19th century from the edible mushroom Cantharellus cinnabarinus (Haxo, 1950). Studies have shown that canthaxanthin is produced as a secondary carotenoid in several green algae, blue-green algae, and bacteria. It is synthesized either at the end of the growth phase instead of or in addition to primary carotenoids (Hertzberg et al., 1966; Saperstein et al., 1954).

In recent years, researchers have been increasingly interested in microbial sources of canthaxanthin due to their advantages over other natural sources. Microbial fermentation processes can be easily controlled to achieve higher growth rates and greater cell density, offering the potential for efficient canthaxanthin production (Das et al., 2007; Krupa et al., 2010). Moreover, microbial production of canthaxanthin does not pose significant limitations in terms of space and time, making it a promising approach for commercial production (Bhosale and Bernstein, 2005).

While there is limited information available regarding the commercial-scale production of canthaxanthin by various microbes, there is a wide range of microbial sources that have been identified as potential producers of canthaxanthin are listed in Table 3.

Table 3

Various microbial sources of canthaxanthin

Organism nameTypeReference
BacteriaDietzia marisbacteria(Venil et al., 2021)
Bradyrhizobium sp. Strain ORS278bacteria(George et al., 2020)
Bradyrhizobium sp.bacteria(Agarwal et al., 2023)
Escherichia colibacteria(Madhavan et al., 2022)
Gordonia jacobaeabacteria(de Miguel et al., 2001)
Dietzia natronolimnaea HS1bacteria(Nasri et al., 2010a)
Micrococcus roseusbacteria(Karbalaei-Heidari et al., 2020)
Brevibacterium sp.bacteria(Mitra et al., 2021)
Rhodococcusmarisbacteria(Nelis et al., 1991)
Corynebacterium michiganense (mutant)bacteria(Watcharawipas et al., 2022)
Dietzia sp. K44bacteria(Venugopalan et al., 2013)
AlgaeChlorella emersoniialgae(Cezare-Gomes et al., 2019)
Chlorella zofingiensisalgae(Pelah et al., 2004)
Dictyococcus cinnabarinusalgae(Rather et al., 2022)
Coelastrella striolata var. multistriatamicroalga(Abe et al., 2007)
Haematococcus lacustrisalgae(Chattopadhyay et al., 2008)
Scenedesmus sp.microalgae(Rajput et al., 2022)
Nostoc punctiforme PCC 73102algae(Llewellyn et al., 2020)
OthersCantharellus cinnabarinusfungus(Martinez-Camara et al., 2021)
Haloferax alexandrines Strain-TM Tarchaeon(Asker et al., 2002)
Mucor circinelloides (genetically modified)fungi(Naz et al., 2021)
Monascus roseusyeast(Dufosse et al., 2009)

Higher animal sources

Canthaxanthin is naturally produced in many mushrooms, and it can also be found in the eggs of fishes and crustaceans (https://www.cfs.gov.hk). The presence of canthaxanthin has been established in various fish species, including carp Cyprinus carpio [family: cyprinidae] (Katayama et al., 1973), golden mullet (Mugil auratus) [family: mugilidae], Diplodus annularis [family: sparidae], and Crenila brustinca [family: labridae] (Czeczuga, 1973). However, it is worth noting that canthaxanthin is not located in wild Atlantic salmon, although a minor portion of carotenoid was found in wild Pacific salmon (Caballo et al., 2012). Canthaxanthin has also been found in the wild trout Salmo trutta (Briers et al., 2013) and in the bird golden-crowned kinglet (Chui et al., 2011).

Genetically modified organisms as canthaxanthin sources

Importance of canthaxanthin

Canthaxanthin as a natural antioxidant

Canthaxanthin has been found to possess antiaging and antioxidant capabilities, making it effective in scavenging free radicals, combating oxidative stress, and enhancing endogenous antioxidant defenses (Mathimaran et al., 2021). In vitro studies have shown that keto-carotenoids like canthaxanthin exhibit higher antioxidant and free radical scavenging capacities compared to carotene carotenoids like lycopene or β-carotene. This is attributed to the stability provided by the conjugation of the keto group with the polyene backbone, which enhances the ability to stabilize carbon-centered radicals (Esatbeyoglu et al., 2017).

Canthaxanthin acts as a strong antioxidant [reduction potential of dietary canthaxanthin in triton X micelles is 1041 {E0 (CAR•+/CAR)/mV}] (Bohm et al., 2012) owing to the location of keto groups at the 4 and 4′ positions in the β-ionone ring. The rate constant for singlet oxygen quenching for canthaxanthin is near 1.45 × 1010 mol/s (Choe et al., 2009). Canthaxanthin acts as a strong antioxidant by reacting with phenoxyl radicals generated by the one-electron metabolism of phenolic compounds, the reaction being catalyzed by the peroxidase enzyme (Tyurin et al., 1997). The antioxidant potential of canthaxanthin against photo-oxidation by energy dissipation is well known (Albrecht et al., 2001).

Comparative studies have shown that lycopene exhibits the highest radical scavenging ability among carotenoids is lycopene > β-cryptoxanthin ≈ β-carotene > lutein ≈ zeaxanthin > α-carotene > echinenone > canthaxanthin = astaxanthin (Muller et al., 2011). Similarly, another experiment ranked carotenoids in terms of stability, with α-tocopherol > lycopene ~ β-tocopherol ~ γ-tocopherol > β-carotene > zeaxanthin ~ 8-tocopherol > lutein > echinenone » canthaxanthin ~ P-apo-8′-carotenal > astaxanthin (Mortensen et al., 1997).

Storage time is critical for canthaxanthin pellets, as approximately 20% of the content can be lost over 2 months under ambient temperature. To mitigate this, natural antioxidants can be used during storage in combination with other pigments (Choubert et al., 1991). Canthaxanthin supplementation has been shown to alter α and γ tocopherol concentrations in murine tissues in a dose-dependent and time-dependent manner, with gender and tissue specificity (Nys et al., 2000). These findings indicate the involvement of canthaxanthin in cellular oxidative stress and its potential for in vivo protection against chronic diseases (Palozza et al., 1998).

Studies have indicated that the 9-cis isomers of canthaxanthin exhibit higher antioxidant activity compared to the all-trans form (Venugopalan et al., 2013). Additionally, microbial canthaxanthin produced through continuous culture systems has been found to possess greater in vitro antioxidant activity compared to canthaxanthin obtained through batch and fed-batch fermentation systems using Dietzia natronolimnaea (Gharibzahedi et al., 2013a). Supplementation of the diet with carotenoids (xanthophyll/canthaxanthin) in Chinese soft-shelled turtles has been shown to enhance antioxidant ability, reduce lipid peroxidation in serum and liver, and upregulate the expression of enzymes β-carotene oxygenase 2 (BCO2), catalase (CAT), and the superoxide dismutase 2 (SOD2) involved in antioxidant defense (Wang et al., 2021).

Canthaxanthin and lipid metabolism

Carotenoid pigments, including canthaxanthin, can influence the structural and dynamic properties of lipid membranes. Differential scanning calorimetry studies have shown that canthaxanthin can alter the thermotropic phase transition of lipid bilayers, particularly in the gel phase, but has minimal effects on phosphatidylethanolamines (Rengel et al., 2000). Canthaxanthin can restrict molecular motion in both the head groups and hydrophobic core of lipid molecules within the membrane bilayer, leading to the elongation of alkyl lipid chains. It also modifies the surface of lipid membranes, particularly in the gel state, and promotes the assembly of lipid vesicles (Sujak et al., 2005, 2009, 2012).

In terms of its impact on atherosclerosis, an independent study found that incubation of monocyte-macrophages with canthaxanthin-incorporated low-density lipoprotein (LDL) demonstrated more effective prevention of atherosclerosis compared to β-carotene and zeaxanthin (Carpenter et al., 1997). Dietary carotenoids, including canthaxanthin, have been shown to reduce the progression of atherosclerosis (Carpenter et al., 1997).

The absorption efficiency of canthaxanthin is lower in liposomes, liver microsomes, and retinal epithelial cells, although its efficacy remains unaltered (Shafaa et al., 2007). A study by Shih et al. in 2008 suggested that canthaxanthin modulates the balance between pro-oxidation and antioxidation, suppresses cholesterol-induced oxidative stress, and affects the antioxidant system and cholesterol metabolism. Canthaxanthin significantly reduces LDL-cholesterol levels in the blood by increasing the concentration of glutathione transferase and elevating the concentration of catalase in the serum and liver (Kumar et al., 2011). Canthaxanthin has also been shown to significantly increase high-density lipoprotein (HDL) levels in rats (Balasubramaniam et al., 2012).

Excessive doses of canthaxanthin in the feed of monkeys have been associated with the accumulation of canthaxanthin–lipoprotein complexes, leading to retinopathy-related diseases (Sujak, 2009). However, controlled and consistent feeding of canthaxanthin over a prolonged period can result in its accumulation in the liver, fat, lungs, and small intestine without exhibiting any toxicity (Tang et al., 1995). Excessive and prolonged consumption of canthaxanthin as a superficial skin shading agent and antitanning agent has led to the accumulation of canthaxanthin crystals inside the eye. However, these crystals were reabsorbed after discontinuing canthaxanthin consumption without causing any clinical manifestations (Mussagy et al., 2019).

Canthaxanthin and cancer prevention

Canthaxanthin exhibits significant antimutagenic and anticarcinogenic effects (Azuine et al., 2012). Silver nanoparticles synthesized using bacterial canthaxanthin pigment have shown cytotoxic effects in HaCaT cell lines (immortalized keratinocyte line) and demonstrated no side effects when tested for their wound healing capacity (Venil et al., 2021). In mammalian cells, canthaxanthin is not transformed into active retinoids, suggesting that its lipid antioxidant effects may be responsible for its inhibitory activities on conversion (Pung et al., 1988).

The antiproliferative effect of methanolic extract of Moringa oleifera leaves (MOLME) was studied in Swiss albino mice using Ehrlich ascites carcinoma cell lines. Canthaxanthin, as an important bioactive component of MOLME, has been found to inhibit cancer initiation, showing a 65% decrease in mammary cancer incidence in rats induced by dimethylbenzanthracene (Das et al., 2021). Oral administration of canthaxanthin has led to a significant reduction in tumor size and number in hamster buccal pouch carcinogenesis (Schwartz et al., 1988). Huang et al. (1992) observed a significant inhibitory action of canthaxanthin on B16F10 melanomas and PYB6 fibrosarcoma tumor cells. In Human WiDr colon adenocarcinoma and SK-MEL-2 melanoma cells, canthaxanthin induced apoptosis (Palozza et al., 1992). Canthaxanthin has also shown a decrease in the number and size of liver preneoplastic foci (induced by aflatoxin B1) in male weanling rats through the variation of aflatoxin B1 metabolism towards aflatoxin M1 and thus dissuading aflatoxin B1 to commence genotoxic action (Gradelet et al., 1998).

The reported antitumor and anticancer activities of canthaxanthin may be attributed to its radical trapping or chain-breaking processes (Bendich et al., 1989; Tanaka et al., 2012). This keto-carotenoid caused a reduction of oral carcinogenesis via suppression of cell proliferation and decreasing polyamine levels in oral mucosal tissues of F344 rats (Tanaka et al., 1995). However, lower plasma canthaxanthin concentration has been observed in females with cervical cancer (Palan et al., 1996). Canthaxanthin has been effective in preventing the initiation of dimethylbenzanthracene-induced mammary cancer in female Sprague-Dawley rats but has shown no significant effect on methyl nitrosourea-induced carcinogenesis of the same type (Grubbs et al., 1991). In MCF-7 and MDA-MB-231 human breast cancer cells, canthaxanthin has not demonstrated inhibitory effects (Chew et al., 2004). Similarly, canthaxanthin has shown no inhibitory effect on n-butyl-n-(4-hydroxybutyl) nitrosamine-induced bladder cancer (Mathews-Roth et al., 1991). A recent study showed the induction of apoptosis in DU145 prostate cancer cells by partially saturated canthaxanthin obtained from mutant fungi (Kumaresan et al., 2008).

Neuroprotective activities of canthaxanthin

Canthaxanthin has shown partial neuroprotective activity in chemically induced rat adrenal medulla cell deaths at lower concentrations (Chang et al., 2013). Among carotenoids, β-carotene and canthaxanthin have been identified as potent stimulators in the gap junction system. Gap junctions function as water-filled pores that allow the exchange of low molecular weight compounds, connecting the cytosol of neighboring cells (Stahl et al., 2005). Zhang et al. (1992) predicted this activity of canthaxanthin earlier and demonstrated that canthaxanthin upregulates the connexin43 gene, which encodes a major gap junction protein, in a dose-dependent manner.

In PC12 cells differentiated by nerve growth factor, treatment with canthaxanthin inhibited the release of TNF-α, IL-1, and IL-6, and also inhibited the activity of caspase-3 (Chan et al., 2009). However, the exact mechanisms by which canthaxanthin exerts its neuroprotective effects have not been fully elucidated (Guest et al., 2012).

Immunomodulatory activities of canthaxanthin

Canthaxanthin has been shown to exhibit immuno-modulating activities by enhancing the proliferation and activity of murine immunocompetent cells (Okai et al., 1996). In rat spleen, supplementation with canthaxanthin increased the proliferation of T and B lymphocytes (Bendich et al., 1989). Additionally, canthaxanthin treatment has been found to increase mitogen-induced lymphocyte proliferation without exhibiting provitamin activity. In human peripheral blood mononuclear cells, canthaxanthin treatment enhanced the expression of activation markers for natural killer cells and T helper cells (Chew et al., 2004).

Canthaxanthin has been established for the treatment of photosensitive disorders and tanning (Pangestuti et al., 2020). It has also been observed that canthaxanthin specifically affects the expression of heme oxygenase-1 upon UVA-related injury to human dermal fibroblasts (Camera et al., 2009).

A recent study demonstrated that canthaxanthin weakly inhibits the activities of cytochrome P450 enzymes, particularly CYP2C19 and CYP3A4/5. This suggests that dietary supplements of canthaxanthin may have minimal effects on drugs metabolized by these enzymes (Zheng et al., 2013). Furthermore, canthaxanthin has been found to increase the synthesis of cytochrome oxidases and peroxidases in murine macrophages (Pechinskii et al., 2014).

Some independent studies have shown that canthaxanthin acts as a significant inducer of liver enzymes, such as P4501A1 and 1A2, in rats. It co-induces 4NP-UGT and QR, leading to the production of xenobiotic-metabolizing enzymes (Astorg et al., 1994; Gradelet et al., 1996). However, this kind of enzyme induction is absent in lower animals (Page et al., 2002).

Effect of canthaxanthin on growth performances and digestion in marine animals

Studies have shown that the addition of carotenoids, including canthaxanthin, as nutritional supplements, can enhance the growth of marine fishes by increasing the utilization of nutrients. Niu et al. (2014) reported that the addition of carotenoids as supplements resulted in increased growth of marine fishes. In a recent study by Kalidoss et al. (2020), the addition of astaxanthin–canthaxanthin as a nutritional complement has resulted in increased growth and survival of Atlantic salmon fry and juveniles.

Furthermore, the addition of canthaxanthin and astaxanthin as dietary complements has been shown to increase the apparent digestibility coefficients (ADC) values, indicating improved utilization of protein and lipid components of food. For example, rainbow trout fed with canthaxanthin-supplemented feed (100 mg/kg) showed an ADC value of 61.4% (Kalidoss et al., 2020).

In addition to growth promotion, the effects of canthaxanthin-loaded liposomes, along with α-tocopherol, on various growth aspects and canthaxanthin accumulation in rainbow trout (Oncorhynchus mykiss) fillets were studied by Toan et al. (2021). The study demonstrated that fish administered with a diet supplemented with 1 g/kg of canthaxanthin and α-tocopherol-loaded liposomes (IC = 0.5%) showed a notable increase in final weight compared to those fed a nonsupplemented diet. This supplementation also resulted in a noticeable change in the color of salmon muscle (Toan et al., 2021).

Conclusion and future perspectives

The production of naturally occurring canthaxanthin from various living organisms such as bacteria, fungi, and algae is currently in the developmental stages, showing promising potential for commercial-scale production. Recent studies have identified several high canthaxanthin-yielding strains, including Gordonia jacobaea MV-26 (Scaife et al., 2012b), Dietzia maris NIT-D (Goswami et al., 2012), Dietzia natronolimnaea HS-1 (Nasri et al., 2010), Yarrowia lipolytica (Grewal et al., 2022), Rhodotorula mucilaginosa (Aksu et al., 2005), Coelastrella sp. (Corato et al., 2022), Dietzia maris AURCCBT01 (Abuthahir et al., 2021), and Paracoccus carotinifaciens VTP20181 (Duy et al., 2021).

However, due to the limited options available for the biological production of canthaxanthin, the market is predominantly monopolized by chemical synthesis methods (Scaife et al., 2012a, b). Nevertheless, there are promising developments in utilizing inexpensive carbon sources and external nutritional supplements to enhance the production of keto-carotenoids from biological sources (Domínguez-Bocanegra et al., 2004; Nasri Nasrabadi et al., 2010a; Nasrabadi et al., 2010b; Gharibzahedi et al., 2013b). Despite these advancements, the current yields of biopigments are not sufficient from an industrial standpoint (Nasrabadi et al., 2010b). Consequently, new approaches combining molecular biology and genetic engineering methods have been explored to enhance and improve canthaxanthin production, leading to a revolution and a new era in the medicinal, nutraceutical, feed additive, and biocolorant sectors (Gharibzahedi et al., 2013b).

Canthaxanthin plays a vital role in animals by promoting maturity and protecting tissues, both in plants and animals, against the harmful effects of oxidizing free radicals. Many countries regularly use canthaxanthin as a supplement in animal and fish feed, which has shown promising results. Additionally, there are synthetic forms of canthaxanthin available in the market, particularly useful in the food industry (Munasinghe et al., 2021).

Canthaxanthin is extensively utilized as a dietary supplement and colorant in the poultry and aquaculture industries. Beyond its pigmenting properties, canthaxanthin also holds potential health benefits, acting as a scavenger of free radicals, an antioxidant, and playing a role in gene regulation (Esatbeyoglu et al., 2017).

However, further investigation is necessary to understand the molecular targets through which canthaxanthin provides health benefits and to explore its biological functions. It is crucial to conduct more studies to uncover its biological role beyond its colorant properties. Additionally, the development of low-cost downstream processes and ensuring biological safety with regard to genetically modified microorganisms should be carefully considered for future perspectives (Jin et al., 2003; Goswami et al., 2010).

To date, the advancement of new transgenic strains of microbes has contributed to increased canthaxanthin production with enhanced environmental tolerance (Gao et al., 2020). Therefore, further research is needed to enhance the biological production of canthaxanthin and make it competitive with synthetic methods.

Acknowledgments

The author wishes to thank Lovely Professional University for their help and necessary support to carry out the present work.

Conflict of interest

No conflicts, informed consent, or human or animal rights are applicable to this study.

References

1 

Abe K., Hattori H., Hirano M. (2007) Accumulation and antioxidant activity of secondary carotenoids in the aerial microalga Coelastrellastriolata var. multistriata. Food Chem. 100(2): 656–661.

2 

Abuthahir S.S.F., Venil C.K., Malathi M., Devi P.R. (2021) Optimization of submerged fermentation for enhanced production of canthaxanthin by Dietziamaris AURCCBT01. Mater. Today Proc. 47: 2132–2137.

3 

Agarwal H., Bajpai S., Mishra A., Kohli I., Varma A., Fouillaud, M., Joshi N.C. (2023) Bacterial pigments and their multifaceted roles in contemporary biotechnology and pharmacological Applications. Microorganisms 11(3): 614.

4 

Aksu Z., Eren A.T. (2005) Carotenoids production by the yeast Rhodotorula mucilaginosa:use of agricultural wastes as a carbon source. Process Biochem. 40(9): 2985–2991.

5 

Albrecht M., Steiger S., Sandmann G. (2001) Expression of a ketolase gene mediates the synthesis of canthaxanthin in Synechococcus leading to tolerance against photoinhibition, pigment degradation and UV-B sensitivity of photosynthesis. Photochem. Photobiol. 73(5): 551–555.

6 

Asker D., Awad T., Ohta Y. (2002) Lipids of Haloferaxalexandrinus strain TMT: an extremely halophilic canthaxanthin-producing archaeon. J. Biosci. Bioengin. 93(1): 37–43.

7 

Astorg P., Gradelet S., Leclerc J., Canivenc M.C., Siess M.H. (1994) Effects of $-carotene and canthaxanthin on liver xenobiotic-metabolizing enzymes in the rat. Food Chem. Toxicol. 32(8): 735–742.

8 

Ausich R.L. (1997) Commercial opportunities for carotenoid production by biotechnology. Pure Appl. Chem. 69(10): 2169–2174.

9 

Azuine M.A., Goswami U.C., Kayal J.J., Bhide S.V. (1992) Antimutagenic and anticarcinogenic effects of carotenoids and dietary palm oil. Nutrit. Cancer 17(3): 287–295.

10 

Bart J., MacGillavry C. (1968) The crystal and molecular structure of canthaxanthin. Acta Crystallogr. B: Struct. Crystallogr. Crystal Chem. 24(12): 1587–1606.

11 

Bendich A., Olson J.A. (1989) Biological actions of carotenoids 1. FASEB J. 3(8): 1927–1932.

12 

Bera S., Dutta D. (2017) Encapsulation and release of a bacterial carotenoid from hydrogel matrix: Characterization, kinetics and antioxidant study. Engin. Life Sci. 17(7): 739–748.

13 

Bera S. (2019) Carotenoids: updates on legal statutory and competence for nutraceutical properties. Curr. Res. Nutr. Food Sci. 7(2): 300.

14 

Bera S. (2020) Carotenoid accumulation in Dietzia maris NITD protects from macromolecular damage. BioTechnologia 101(3): 171–178.

15 

Bernhard K., Mayer H. (1991) Recent advances in the synthesis of achiral carotenoids. Pure Appl. Chem. 63(1): 35–44.

16 

Bhosale P., Bernstein P.S. (2005) Microbial xanthophylls. Appl. Microbiol. Biotech. 68(4): 445–455.

17 

Briers R.A., Waterman J.O., Galt K., Campbell R.N. (2013) Population differentiation and temporal changes of carotenoid pigments and stable isotope ratios in the offspring of anadromous and non-anadromous trout S almo trutta. Ecol. Freshwater Fish 22(1): 137–144.

18 

Britton G. (1998) Overview of carotenoid biosynthesis. Carotenoids 3: 13–147.

19 

Caballo C., Costi E.M., Sicilia M.D., Rubio S. (2012) Determination of supplemental feeding needs for astaxanthin and canthaxanthin in salmonids by supramolecular solvent-based microextraction and liquid chromatography–UV/VIS spectroscopy. Food Chem. 134(2): 1244–1249.

20 

Camera E., Mastrofrancesco A., Fabbri C., Daubrawa F., Picardo M., Sies H., Stahl W. (2009) Astaxanthin, canthaxanthin and $-carotene differently affect UVA-induced oxidative damage and expression of oxidative stress-responsive enzymes. Exp. Dermatol. 18(3): 222–231.

21 

Carpenter K.L., van der Veen C., Hird R., Dennis I.F., Ding T., Mitchinson M.J. (1997) The carotenoids $-carotene, canthaxanthin and zeaxanthin inhibit macrophage-mediated LDL oxidation. FEBS Lett. 401(2–3): 262–266.

22 

Castangia I., Manca M.L., Razavi S.H., Nácher A., Díez-Sales, O., Peris J.E., Manconi M. (2022) Canthaxanthin biofabrication, loading in green phospholipid vesicles and evaluation of in vitro protection of cells and promotion of their monolayer regeneration. Biomedicines 10(1): 157.

23 

Cezare-Gomes E.A., Mejia-da-Silva L.D.C., Pérez-Mora L.S., Matsudo M.C., Ferreira-Camargo L.S., Singh A.K., de Carvalho J.C.M. (2019) Potential of microalgae carotenoids for industrial application. Appl. Biochem. Biotechnol. 188: 602–634.

24 

Chan K.C., Mong M.C., Yin M.C. (2009) Antioxidative and anti-inflammatory neuroprotective effects of astaxanthin and canthaxanthin in nerve growth factor differentiated PC12 cells. J. Food Sci. 74(7): H225–H231.

25 

Chang C.S., Chang C.L., Lai G.H. (2013) Reactive oxygen species scavenging activities in a chemiluminescence model and neuroprotection in rat pheochromocytoma cells by astaxanthin, beta-carotene, and canthaxanthin. Kaohsiung J. Med. Sci. 29(8): 412–421.

26 

Chattopadhyay P., Chatterjee S., Sen S.K. (2008) Biotechnological potential of natural food grade biocolorants. Afr. J. Biotech. 7(17): 2972–2985.

27 

Chen M., Li M., Ye L., Yu H. (2021) Construction of canthaxanthin-producing yeast by combining spatiotemporal regulation and pleiotropic drug resistance engineering. ACS Synth. Biol. 11(1): 325–333.

28 

Chew B.P., Park J.S. (2004) Carotenoid action on the immune response. J. Nutr. 134(1): 257S–261S.

29 

Choe E., Min D.B. (2009) Mechanisms of antioxidants in the oxidation of foods. Compr. Rev. Food Sci. Food Safety 8(4): 345–358.

30 

Choubert G., de la Noüe J., Blanc J.M. (1991) Apparent digestibility of canthaxanthin in rainbow trout: effect of dietary fat level, antibiotics and number of pyloric caeca. Aquaculture 99(3–4): 323–329.

31 

Chui C.K., McGraw K.J., Doucet S.M. (2011) Carotenoid-based plumage coloration in golden-crowned kinglets Regulus satrapa: Pigment characterization and relationships with migratory timing and condition. J. Avian Biol. 42(4): 309–322.

32 

Cooney J.J., Berry R.A. (1981) Inhibition of carotenoid synthesis in Micrococcus roseus. Canad. J. Microbiol. 27(4): 421–425.

33 

Corato A., Le T.T., Baurain D., Jacques P., Remacle C., Franck F. (2022) A fast-growing oleaginous strain of coelastrella capable of astaxanthin and canthaxanthin accumulation in phototrophy and heterotrophy. Life 12(3): 334.

34 

Czeczuga B. (1973) Carotenoids in fish II. Carotenoids and vitamin A in some fishes from the coastal region of the Black Sea. Hydrobiologia 41(1): 113–125.

35 

Das A., Yoon S.-H., Lee S.-H., Kim J.-Y., Oh D.-K., Kim S.-W. (2007) An update on microbial carotenoid production: application of recent metabolic engineering tools. Appl. Microbiol. Biotech. 77(3): 505–512.

36 

Das P.K., Asha S.Y., Siddika A., Siddika A., Tareq A.R.M., Islam F., Rakib M.A. (2021) Methanolic extract of Moringa oleifera leaves mediates anticancer activities through inhibiting NF-κB and enhancing ROS in Ehrlich ascites carcinoma cells in mice. J. Adv. Biotechnol. Exp. Ther. 4: 161–170.

37 

Domínguez-Bocanegra A.R., Torres-Muñoz J.A. (2004) Astaxanthin hyperproduction by Phaffia rhodozyma (now Xanthophyllomyces dendrorhous) with raw coconut milk as sole source of energy. Appl. Microbiol. Biotech. 66(3): 249–252.

38 

de Miguel T., Sieiro C., Poza M., Villa T.G. (2001) Analysis of canthaxanthin and related pigments from Gordonia jacobaea mutants. J. Agric. Food Chem. 49(3): 1200–1202.

39 

Dewanjee S., Bhattacharjee N., Chakraborty P., Bhattacharjee S. (2021) Carotenoids as antioxidants. carotenoids: structure and function in the human body. Springer: 447–473.

40 

Dufosse L. (2009) Pigments. Microbial: Encyclopedia of Microbiology (Third Edition). Academic Press 457–471.

41 

Duy L.X., Toan T.Q., Anh D.V., Hung N.P., Huong T.T.T., Long P.Q., Manh D.V. (2021) Optimization of canthaxanthin extraction from fermented biomass of Paracoccus carotinifacuens VTP20181 bacteria strain isolated in Vietnam. Foods Raw Mater. 9(1): 117–125.

42 

Entschel R., Karrer P. (1958) Carotinoidsynthesen XXII. Umsetzungsprodukte des β-Carotins mit Bromsuccinimid. (Einführung von Äther-und Hydroxyl-Gruppen in den Kohlenwasserstoff). Helvet. Chim. Acta 41(2): 402–413.

43 

Esatbeyoglu T., Rimbach G. (2017) Canthaxanthin: From molecule to function. Mol. Nutrit. Food Res. 61(6): 1600469.

44 

Ernst H. (2002) Recent advances in industrial carotenoid synthesis. Pure Appl. Chem. 74(11): 2213–2226.

45 

Gao X., Xu H., Zhu Z., She Y., Ye S. (2020) Improved production of echinenone and canthaxanthin in transgenic Nostoc sp. PCC 7120 overexpressing a heterologous crtO gene from Nostoc flagelliforme. Microbiol. Res. 236: 126455.

46 

George D.M., Vincent A.S., Mackey H.R. (2020) An overview of anoxygenic phototrophic bacteria and their applications in environmental biotechnology for sustainable Resource recovery. Biotech. Rep. 28: e00563.

47 

Gharibzahedi S.M.T., Razavi S.H., Mousavi S.M., Moayedi V. (2012) High efficiency canthaxanthin production by a novel mutant isolated from Dietzia natronolimnaea HS-1 using central composite design analysis. Industr. Crops Prod. 40: 345–354.

48 

Gharibzahedi S.M.T., Razavi S.H., Mousavi S.M. (2013a) Comparison of antioxidant and free radical scavenging activities of biocolorant synthesized by Dietzia natronolimnaea HS-1 cells grown in batch, fed-batch and continuous cultures. Industr. Crops Prod. 49: 10–16.

49 

Gharibzahedi S.M.T., Razavi S.H., Mousavi S.M. (2013b) Microbial canthaxanthin: perspectives on biochemistry and biotechnological production. Eng. Life Sci. 13(4): 408–417.

50 

Goswami G., Chaudhuri S., Dutta D. (2010) The present perspective of astaxanthin with reference to biosynthesis and pharmacological importance. World J. Microbiol. Biotech. 26(11): 1925–1939.

51 

Gradelet S., Astorg P., Leclerc J., Chevalier J., Vernevaut M.F., Siess M.H. (1996) Effects of canthaxanthin, astaxanthin, lycopene and lutein on liver xenobiotic-metabolizing enzymes in the rat. Xenobiotica 26(1): 49–63.

52 

Gradelet S., Le Bon A. M., Bergès R., Suschetet M., Astorg P. (1998) Dietary carotenoids inhibit aflatoxin B1-induced liver preneoplastic foci and DNA damage in the rat: role of the modulation of aflatoxin B1 metabolism. Carcinogenesis 19(3): 403–411.

53 

Grewal J., Wołacewicz M., Pyter W., Joshi N., Drewniak L., Pranaw K. (2022) Colorful treasure from agro-industrial wastes: a sustainable chassis for microbial pigment production. Front. Microbiol. 13: 89.

54 

Grubbs C.J., Eto I., Juliana M., Whitaker L.M. (1991) Effect of canthaxanthin on chemically induced mammary carcinogenesis. Oncology 48(3): 239–245.

55 

Guest J.A., Grant R.S. (2012) Effects of dietary derived antioxidants on the central nervous system. Int. J. Nutr. Pharmacol. Neurolog. Dis. 2(3): 185.

56 

Hashimoto H., Koyama Y., Shimamura T. (1988) Isolation of cis-trans isomers of canthaxanthin by high-performance liquid chromatography using a calcium hydroxide column and identification of their configurations by 1H NMR spectroscopy. J. Chromatogr. A. 448: 182–187.

57 

Haxo F. (1950) Carotenoids of the mushroom Cantharellus cinnabarinus. Bot. Gazet. 112(2): 228–232.

58 

Hertzberg S., Jensen S.L. (1966) The carotenoids of blue-green algae—II.: the carotenoids of Aphanizomenon flosaquae. Phytochemistry 5(4): 565–570.

59 

Huang D.S., Odeleye O.E., Watson R.R. (1992) Inhibitory effects of canthaxanthin on in vitro growth of murine tumor cells. Cancer Lett. 65(3): 209–213.

60 

Jin E.S., Polle J.E., Lee H.K., Hyun S.M., Chang M. (2003) Xanthophylls in microalgae: from biosynthesis to biotechnological mass production and application. J. Microbiol. Biotech. 13(2): 165–174.

61 

Kalidoss Manikandan N.F., Prabu E. (2020) A review on the application and effect of carotenoids with respect to canthaxanthin in the culture of fishes and crustaceans. Int. J. Fish. Aqua. Stud. 8(5): 128–133.

62 

Karbalaei-Heidari H.R., Partovifar M., Memarpoor-Yazdi M. (2020) Evaluation of the bioactive potential of secondary metabolites produced by a new marine micrococcus species isolated from the Persian Gulf. Avicenna J. Med. Biotech. 12(1): 61.

63 

Katayama T., Miyahara T., Tanaka Y., Sameshima M. (1973) Memoirs of the Faculty of Fisheries. Kagoshima Univ. J. 22(1): 39–45.

64 

Krupa D., Nakkeeran E., Kumaresan N., Vijayalakshmi G., Subramanian R. (2010) Extraction, purification and concentration of partially saturated canthaxanthin from Aspergillus carbonarius. Bioresour. Technol. 101(19): 7598–7604.

65 

Kumar A., Srikanta A.H., Muthukumar S.P., Sukumaran U.K., Govindaswamy V. (2011) Antioxidant and lipid peroxidation activities in rats fed with Aspergillus carbonarius carotenoid. Food Chem. Toxicol. 49(12): 3098–3103.

66 

Kumaresan N., Sanjay K.R., Venkatesh K.S., Kadeppagari, R.K., Vijayalakshmi G., Umesh-Kumar S. (2008) Partially saturated canthaxanthin purified from Aspergillus carbonarius induces apoptosis in prostrate cancer cell line. Appl. Microbiol. Biotech. 80(3): 467–473.

67 

Llewellyn C.A., Airs R.L., Farnham G., Greig C. (2020) Synthesis, regulation and degradation of carotenoids under low level UV-B radiation in the filamentous cyanobacterium Chlorogloeopsis fritschii PCC 6912. Front. Microbiol. 11: 163.

68 

Lotan T., Hirschberg J. (1995) Cloning and expression in Escherichia coli of the gene encoding $-C-4-oxygenase, that converts $-carotene to the ketocarotenoid canthaxanthin in Haematococcus pluvialis. FEBS Lett. 364(2): 125–128.

69 

Madhavan A., Arun K.B., Alex D., Anoopkumar A.N., Emmanual S., Chaturvedi P., Sindhu R. (2022) Microbial production of nutraceuticals: Metabolic engineering interventions in phenolic compounds, poly unsaturated fatty acids and carotenoids synthesis. J. Food Sci. Technol.: 1–13.

70 

Mathimaran A., Kumar A. (2020) Changes in morphogenesis and carotenogenesis to influence polygalacturonase secretion in Aspergillus carbonarius mutant. Arch. Microbiol. 202(6): 1285–1293.

71 

Mathimaran A., Kumar A., Prajapati G., Ampapathi R.S., Bora H.K., Guha R. (2021) Partially saturated canthaxanthin alleviates aging-associated oxidative stress in d-galactose administered male wistar rats. Biogerontology 22: 19–34.

72 

Martínez-Cámara S., Ibañez A., Rubio S., Barreiro C., Barredo J.L. (2021) Main carotenoids produced by microorganisms. Encyclopedia 1(4): 1223–1245.

73 

Mathews-Roth M.M., Lausen N., Drouin G., Richter A., Krinsky N.I. (1991) Effects of carotenoid administration on bladder cancer prevention. Oncology 48(3): 177–179.

74 

Moldenhauer M., Tseng H.W., Kraskov A., Tavraz N.N., Yaroshevich I.A., Hildebrandt P., Friedrich T. (2023) Parameterization of a single H-bond in Orange Carotenoid Protein by atomic mutation reveals principles of evolutionary design of complex chemical photosystems. Front. Mol. Biosci. 10: 43.

75 

Mitra R., Han J., Xiang H., Bera S. (2021) Importance of microbial secondary metabolites in health care applications. In Volatiles and Metabolites of Microbes. Academic Press: 349–383.

76 

Mordi R.C., Ademosun O.T., Ajanaku C.O., Olanrewaju I.O., Walton J.C. (2020) Free radical mediated oxidative degradation of carotenes and xanthophylls. Molecules 25(5): 1038.

77 

Mortensen A., Skibsted L.H. (1997) Relative stability of carotenoid radical cations and homologue tocopheroxyl radicals. A real time kinetic study of antioxidant hierarchy. FEBS Lett. 417(3): 261–266.

78 

Müller L., Fröhlich K., Böhm V. (2011) Comparative antioxidant activities of carotenoids measured by ferric reducing antioxidant power (FRAP), ABTS bleaching assay (αTEAC), DPPH assay and peroxyl radical scavenging assay. Food Chem. 129(1): 139–148.

79 

Munasinghe J., Bandara A., Sandakanthi P., Marcus A. (2021) Effect of natural and synthetic carotinoides on egg yolk color development. [in:] European Research Forum : 72–80.

80 

Mussagy C.U., Winterburn J., Santos-Ebinuma V.C., Pereira J.F.B. (2019) Production and extraction of carotenoids produced by microorganisms. Appl. Microbiol. Biotech. 103(3): 1095–1114.

81 

Nasri Nasrabadi M.R., Razavi S.H. (2010a) Enhancement of canthaxanthin production from Dietzia natronolimnaea HS-1 in a fed-batch process using trace elements and statistical methods. Brazil. J. Chem. Engin. 27: 517–529.

82 

Nasrabadi M.R.N., Razavi S.H. (2010b) Use of response surface methodology in a fed-batch process for optimization of tricarboxylic acid cycle intermediates to achieve high levels of canthaxanthin from Dietzia natronolimnaea HS-1. J. Biosci. Bioengin. 109(4): 361–368.

83 

Naz T., Yang J., Nosheen S., Sun C., Nazir Y., Mohamed H., Song Y. (2021) Genetic modification of Mucor circinelloides for canthaxanthin production by heterologous expression of $-carotene ketolase gene. Front. Nutrit. 8: 756218.

84 

Nelis H., De Leenheer A.P. (1991) Microbial sources of carotenoid pigments used in foods and feeds. J. Appl. Bacteriol. 70(3): 181–191.

85 

Nys Y. (2000) Dietary carotenoids and egg yolk coloration. Archiv Geflugelkun. 64: 45–54.

86 

Okai Y., Higashi-Okai K. (1996) Possible immunomodulating activities of carotenoids in in vitro cell culture experiments. Int. J. Immunopharmacol. 18(12): 753–758.

87 

Page G.I., Davies S.J. (2002) Astaxanthin and canthaxanthin do not induce liver or kidney xenobiotic-metabolizing enzymes in rainbow trout (Oncorhynchus mykiss Walbaum). Comparat. Biochem. Physiol. C: Toxicol. Pharmacol. 133(3): 443–451.

88 

Palan P.R., Mikhail M.S., Goldberg G.L., Basu J., Runowicz C.D., Romney S.L. (1996) Plasma levels of beta-carotene, lycopene, canthaxanthin, retinol, and alpha-and tau-tocopherol in cervical intraepithelial neoplasia and cancer. Clin. Cancer Res. 2(1): 181–185.

89 

Palozza P., Calviello G., Serini S., Moscato P., Bartoli G.M. (1998) Supplementation with canthaxanthin affects plasma and tissue distribution of α-and γ-tocopherols in mice. J. Nutrit. 128(11): 1989–1994.

90 

Palozza P., Maggiano N., Calviello G., Lanza P., Piccioni E., Ranelletti F.O., Bartoli G.M. (1998) Canthaxanthin induces apoptosis in human cancer cell lines. Carcinogenesis 19(2): 373–376.

91 

Pangestuti R., Suryaningtyas I.T., Siahaan E.A., Kim S.K. (2020) Cosmetics and cosmeceutical applications of microalgae pigments. [in:] Pigments from Microalgae Handbook 611–633.

92 

Pechinskii S.V., Kuregyan A.G. (2014) The impact of carotenoids on immunity. Pharmaceut. Chem. J. 47(10): 509–513.

93 

Pelah D., Sintov A., Cohen E. (2004) The effect of salt stress on the production of canthaxanthin and astaxanthin by Chlorella zofingiensis grown under limited light intensity. World J. Microbiol. Biotech. 20(5): 483–486.

94 

Pi S., Xi M., Deng L., Xu H., Feng C., Shen R., Wu C. (2020) Practical synthesis of canthaxanthin. J. Iranian Chem. Soc. 17: 493–497.

95 

Pung A.O., Rundhaug J.E., Yoshizawa C.N., Bertram J.S. (1988) β-Carotene and canthaxanthin inhibit chemically- and physically-induced neoplastic transformation in 10T1/2 cells. Carcinogenesis 9(9): 1533–1539.

96 

Qiu D., Zhu W.-L., Tang C.-K., Shi L.-F., Gao H.-Q. (2014) Identification of the composition of isomeric canthaxanthin sample by NMR, HPLC, and mass spectrometry. Food Analyt. Meth. 7(3): 597–605.

97 

Rajput A., Singh D.P., Khattar J.S., Swatch G.K., Singh Y. (2022) Evaluation of growth and carotenoid production by a green microalga Scenedesmus quadricauda PUMCC 4.1. 40. under optimized culture conditions. J. Basic Microbiol. 62(9): 1156–1166.

98 

Rather, L.J., Mir, S.S., Ganie, S.A., Islam, S.U., Li, Q. (2022) Research progress, challenges, and perspectives in microbial pigment production for industrial applications – a review. Dyes Pigments 110989.

99 

Rebelo B.A., Farrona S., Ventura M.R., Abranches R. (2020) Canthaxanthin, a red-hot carotenoid: applications, synthesis, and biosynthetic evolution. Plants 9(8): 1039.

100 

Rengel D., Dlez-Navajas A., Serna-Rico A., Veiga P., Muga A., Milicua J.C.G. (2000) Exogenously incorporated ketocarotenoids in large unilamellar vesicles. Protective activity against peroxidation. Biochimica et Biophysica Acta (BBA)-Biomembranes 1463(1): 179–187.

101 

Rodriguez-Amaya D.B. (2001). A guide to carotenoid analysis in foods. Washington: ILSI Press, vol. 71.

102 

Rosenberger M., McDougal P., Saucy G., Bahr J. (1979) New approaches to the synthesis of canthaxanthin Carotenoids? 5. Elsevier: 871–886.

103 

Saini R.K., Keum Y.S. (2019) Microbial platforms to produce commercially vital carotenoids at industrial scale: an updated review of critical issues. J. Industr. Microbiol. Biotech. 46(5): 657–674.

104 

Scaife M.A., Ma C.A., Armenta R.E. (2012a) Efficient extraction of canthaxanthin from Escherichia coli by a 2-step process with organic solvents. Bioresour. Technol. 111: 276–281.

105 

Scaife M.A., Prince C.A., Norman A., Armenta R.E. (2012b) Progress toward an Escherichia coli canthaxanthin bioprocess. Process Biochem. 47(12): 2500–2509.

106 

Schlatterer J., Breithaupt D.E. (2006) Xanthophylls in commercial egg yolks: quantification and identification by HPLC and LC-(APCI) MS using a C30 phase. J. Agric. Food Chem. 54(6): 2267–2273.

107 

Schwartz J., Shklar G., Reid S., Trickier D. (1988) Prevention of experimental oral cancer by extracts of Spirulina-Dunaliella algae. Nutr. Cancer11(2): 127–134.

108 

Shafaa M.W.I., Diehl H.A., Socaciu C. (2007) The solubilisation pattern of lutein, zeaxanthin, canthaxanthin and β-carotene differ characteristically in liposomes, liver microsomes and retinal epithelial cells. Biophys. Chem. 129(2–3): 111–119.

109 

Shih C.K., Chang J.H., Yang S.H., Chou T.W., Cheng H.H. (2008) $-Carotene and canthaxanthin alter the pro-oxidation and antioxidation balance in rats fed a high-cholesterol and high-fat diet. Brit. J. Nutrit. 99(1): 59–66.

110 

Stahl W., Sies H. (2005) Bioactivity and protective effects of natural carotenoids. Biochim. Biophys. Acta Mol. Basis Dis. 1740(2): 101–107.

111 

Sujak A., Gabrielska J., Milanowska J., Mazurek P., Strzałka K., Gruszecki W.I. (2005) Studies on canthaxanthin in lipid membranes. Biochim. Biophys. Acta 1712(1): 17–28.

112 

Sujak A. (2009) Interactions between canthaxanthin and lipid membranes-possible mechanisms of canthaxanthin toxicity. Cell. Mol. Biol. Lett. 14(3): 395–410.

113 

Sujak A. (2012) Exceptional molecular organization of canthaxanthin in lipid membranes. Acta Biochim. Polon. 59(1): 31–33.

114 

Tanaka T., Shnimizu M., Moriwaki H. (2012) Cancer chemoprevention by carotenoids. Molecules 7(3): 3202–3242.

115 

Tanaka T., Makita H., Ohnishi M., Mori H., Satoh K., Hara A. (1995) Chemoprevention of rat oral carcinogenesis by naturally occurring xanthophylls, astaxanthin and canthaxanthin. Cancer Res. 55(18): 4059–4064.

116 

Tang G., M.C. Blanco J.G. Fox R.M. Russell (1995) Supplementing ferrets with canthaxanthin affects the tissue distributions of canthaxanthin, other carotenoids, vitamin A and vitamin E. J. Nutr. 125(7): 1945–1951.

117 

Tian B., Hua Y. (2010) Carotenoid biosynthesis in extremophilic Deinococcus–Thermus bacteria. Trends Microbiol. 18(11): 512–520.

118 

Toan T.Q., Dang V.A., Pham Q.L., Nguyen P.H., Trinh T.H., Tran T.H., Nguyen T.D. (2021) Effects of dietary inclusion of canthaxanthin-and α-tocopherol-loaded liposomes on growth and muscle pigmentation of rainbow trout (Oncorhynchus mykiss). J. Food Qual. 2021: 1–11.

119 

Torregrosa-Crespo J., Montero Z., Fuentes J.L., Reig García-Galbis M., Garbayo I., Vílchez C., Martínez-Espinosa R.M. (2018) Exploring the valuable carotenoids for the large-scale production by marine microorganisms. Marine Drugs 16(6): 203.

120 

Tyurin V.A., Carta G., Tyurina Y.Y., Banni S., Day B.W., Corongiu F.P., Kagan V.E. (1997) Peroxidase-catalyzed oxidation of β-carotene in HL-60 cells and in model systems: Involvement of phenoxyl radicals. Lipids 32(2): 131–142.

121 

Uzun İ., Orak İ., Sevgili Ö., Yiğit E., Karakaplan M. (2022) Synthesis of canthaxanthin from $-carotene and evaluation of both substances in diode construction. J. Mater. Sci. Mater. Electron. 2022: 1–11.

122 

Venil C.K., Malathi M., Velmurugan P., Renuka Devi P. (2021) Green synthesis of silver nanoparticles using canthaxanthin from Dietzia maris AURCCBT01 and their cytotoxic properties against human keratinocyte cell line. J. Appl. Microbiol. 130(5): 1730–1744.

123 

Venugopal V. (2008). Marine products for healthcare: functional and bioactive nutraceutical compounds from the ocean. CRC press: 1–528.

124 

Venugopalan V., Tripathi S.K., Nahar P., Saradhi P.P., Das R.H., Gautam H.K. (2013) Characterization of canthaxanthin isomers isolated from a new soil Dietzia sp. and their antioxidant activities. J. Microbiol. Biotech. 23(2): 237–245.

125 

Wang F., Jiang J.G., Chen Q. (2007) Progress on molecular breeding and metabolic engineering of biosynthesis pathways of C30, C35, C40, C45, C50 carotenoids. Biotech. Adv. 25(3): 211–222.

126 

Wang P., Zhou X., Xiong G., Zeng D., Ge L., Luo Z., Hu Y. (2021) Effect of dietary canthaxanthin and xanthophyll on growth, antioxidant capacity, body colour, and BCO2, CAT and SOD2 gene expression in Chinese soft-shelled turtle (Pelodiscus sinensis). Aquacult. Nutr. 27(6): 2365–2377.

127 

Watcharawipas A., Runguphan W. (2023) Red yeasts and their carotenogenic enzymes for microbial carotenoid production. FEMS Yeast Res. 23(2023): foac063. https://doi.org/10.1093/femsyr/foac063

128 

Yaqoob S., Riaz M., Shabbir A., Zia-Ul-Haq M., Alwakeel S S., Bin-Jumah M. (2021) Commercialization and marketing potential of carotenoids. Carotenoids: structure and function in the human body. Elsevier 2021: 799.

129 

Zhang L.X., Cooney R.V., Bertram J.S. (1992) Carotenoids upregulate connexin43 gene expression independent of their provitamin A or antioxidant properties. Cancer Res. 52(20): 5707–5712.

130 

Zheng Y.F., Bae S.H., Kwon M.J., Park J.B., Choi H.D., Shin W.G., Bae S.K. (2013) Inhibitory effects of astaxanthin, $-cryptoxanthin, canthaxanthin, lutein, and zeaxanthin on cytochrome P450 enzyme activities. Food Chem. Toxicol. 59: 78–85.

Copyright: © 2023 Institute of Bioorganic Chemistry, Polish Academy of Sciences This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND) (https://creativecommons.org/licenses/by-nc-nd/3.0/legalcode),.allowing third parties to download and share its works but not commercially purposes or to create derivative works.
 
Stosujemy się do standardu HONcode dla wiarygodnej informacji zdrowotnej This site complies with the HONcode standard for trustworthy health information: verify here