Open access peer-reviewed chapter - ONLINE FIRST

Phycocyanin, The Microalgae Bio-Treasure

Written By

Joana Campos, Raquel Fernandes and Ana Novo Barros

Submitted: 08 April 2024 Reviewed: 20 May 2024 Published: 05 June 2024

DOI: 10.5772/intechopen.115108

Functional Food - Upgrading Natural and Synthetic Sources IntechOpen
Functional Food - Upgrading Natural and Synthetic Sources Edited by Ana Novo Barros

From the Edited Volume

Functional Food - Upgrading Natural and Synthetic Sources [Working Title]

Dr. Ana Novo Barros, Ph.D. Joana Campos and Dr. Alice Vilela

Chapter metrics overview

32 Chapter Downloads

View Full Metrics

Abstract

The growing demand for natural alternatives to synthetic compounds has propelled the large-scale production of microalgae and their bioactive constituents. Among these, phycocyanin, a prominent pigment abundant in blue-green algae, has emerged as a subject of intense research interest due to its multifaceted biological activities, which include antioxidative, anti-inflammatory, anticancer, and neuroprotective properties. Its versatility has led to widespread use across various industries, from food and cosmetics to pharmaceuticals, underscoring its economic significance. As a result, efforts have been intensified to refine production processes, enhance purity, and ensure stability to increase its market value. Furthermore, the exploration of secondary metabolites derived from microalgae production holds promise for cross-industry applications, fostering industrial symbiosis and a circular economy. This chapter aims to elucidate the antioxidant capacity of phycocyanin derived from microalgae and delve into its potential for therapeutic approaches.

Keywords

  • phycocyanin
  • microalgae
  • biological functions
  • antioxidant activity
  • therapeutic application

1. Introduction

Over the years, there has been an increase in demand for natural compounds as alternatives to chemical and synthetic molecules. This new era brings forth promising opportunities, including the exploration of emerging novel targets and the utilization of cutting-edge technologies tailored for drug discovery from natural sources [1]. These natural compounds have demonstrated the potential to provide reduced side effects and lower toxicity when compared to certain synthetic molecules used in cancer chemotherapy drugs [2, 3]. While synthesized compounds serve as valuable resources in current drug discovery programs, natural products remain significant due to their unique chemical structures and biological activities [4]. Furthermore, the progression of a natural compound from a “screening hit” to a “drug lead” and eventually to a “marketed drug” requires larger quantities of the compound, often surpassing what can be obtained through re-isolation from plant sources [5]. Natural products have also played a pivotal role as lead compounds in health promotion and in the discovery and development of new drugs [6].

Marine resources, particularly microalgae, have gained global recognition for their remarkable nutritional profile and their potential as alternative sources of essential biomolecules, including proteins, carbohydrates, and lipids. These bioactive compounds derived from microalgae exhibit significant biological functions and find applications across diverse industries [7, 8, 9]. Notably, microalgae have emerged as promising candidates for producing metabolites with pharmaceutical applications, offering natural alternatives for the development of analog compounds to therapeutic drugs [10]. As a result, there is a rising demand for algae products, underscoring the increasing recognition of the significant contributions that microalgae can make across several industrial sectors [7, 8, 9].

Over the years, extensive research has been conducted on highly valuable proteins produced by certain species of microalgae, namely phycocyanin, which has been reported to be associated with relevant and beneficial properties for human health [11]. Phycocyanin is highly recognized for its pharmacological properties, including cardiovascular protection, anti-inflammatory effects, anticancer, and anti-neurodegenerative properties, most of them related to its antioxidant activity. This therapeutic potential underscore phycocyanin’s significance as a molecule of great interest in biomedical research [11, 12].

Hence, the aim of this chapter is to review the antioxidative activity of phycocyanin derived from several microalgae species and explore its potential for the treatment of different diseases.

Advertisement

2. Phycocyanin: origin and structure

Phycocyanin (Figure 1) is a water-soluble, non-toxic, blue pigment, and biologically active compound that exists in various species of microalgae [12, 13].

Figure 1.

Phycocyanin structure. Adapted from Fernandes et al. [11].

Based on their spectral properties, phycocyanin can be classified into three main groups: allophycocyanin, which is obtained from both red and blue-green algae, cyano-phycocyanin (C-PC), which is derived from blue-green algae, and R-phycocyanin (R-PC), which is formed from red algae [12, 13]. The protein consists of an α polypeptide (10–19 kDa) and a β polypeptide (14–21 kDa) forming a monomer, which contributes to its stability [11]. While the structure and function of phycocyanin are largely similar across different microalgae species, differences in the composition and properties of the molecule may exist. Some possible distinctions between phycocyanin molecules from different microalgae species include the amino acid sequence composition, which may affect its physical and chemical properties; the absorption spectra, which can impact its light-harvesting efficiency and coloration; the stability of phycocyanin, which can be influenced by factors such as pH, temperature, and salt concentration; oligomeric structure; and its localization within the cell, such as the thylakoid membrane or the cytoplasm [13, 14].

Over the years, several microalgae species, mostly cyanobacteria, have been reported to be natural producers of phycocyanin (Table 1), namely Arthrospira platensis, Arthrospira fusiformis, Arthrospira maxima, Arthronema africanum, Anabaena sp., Aphanizomenon flos-aquae, Calothrix sp., Cyanidioschyzon merolae, Cyanidium caldarium, Cyanothece sp., Galdieria sulphuraria, Geitlerinema sp., Limnothrix sp., Lyngbya sp., Microcystis aeruginosa, Nostoc commune, Oscillatoria minima, Oscillatoria quadripunctulata, Phormidium sp., Plectonema sp., Pseudanabaena sp., Synechocystis sp., Synechococcus sp., Synechococcus vulcanus, and Thermosynechcoccus vulcanus. Interestingly, red macroalgae were also reported as producers of phycocyanin, Gracilaria chilensis, and Polysiphonia urceolata (Table 1) [13, 15, 16, 17, 18, 19, 20, 21, 22].

PhylumClassOrderFamilySpecies
CyanobacteriotaCyanophyceaeOscillatorialesMicrocoleaceaeArthrospira fusiformis, Arthrospira maxima, Arthrospira platensis
OscillatoriaceaeLyngbya sp., Oscillatoria minima, Oscillatoria quadripunctulata, Phormidium sp., Plectonema sp.,
SynechococcalesSynechococcaceaeSynechococcus sp., Thermosynechcoccus vulcanus
MerismopediaceaeSynechocystis sp.
PseudanabaenalesPseudanabaenaceaeLimnothrix sp., Pseudanabaena sp.
NostocalesNostocaceaeNostoc sp., Anabaena sp.
AphanizomenonaceaeAphanizomenon flos-aquae
CalotrichaceaeCalothrix sp.
GomontiellalesCyanothecaceaeCyanothece sp.
LeptolyngbyalesLeptolyngbyaceaeArthronema africanum
GeitlerinematalesGeitlerinemataceaeGeitlerinema sp.
ChroococcalesMicrocystaceaeMicrocystis aeruginosa
RhodophytaBangiophyceaePorphyridialesPorphyridiaceaePorphyridium cruentum
GaldierialesGaldieriaceaeGaldieria sulphuraria
CyanidialesCyanidiaceaeCyanidioschyzon merolae, Cyanidium caldarium
FlorideophyceaeGracilarialesGracilariaceaeGracilaria chilensis
CeramialesRhodomelaceaePolysiphonia urceolata

Table 1.

Taxonomy of reported algae producers of phycocyanin.

Data from NCBI taxonomy database (Available from: https://www.ncbi.nlm.nih.gov/taxonomy).

Nevertheless, A. platensis, commonly known as Spirulina, stands out as the principal natural source of phycocyanin (14–25%) and is the most used among various microalgae species [11, 15, 23]. In fact, one of the reasons for the current large-scale cultivation of Spirulina is the increasing demand for natural compounds, particularly for the production of high-value proteins such as phycocyanin. Depending on its purity, phycocyanin finds applications in several industries, including food, cosmetics, and pharmaceuticals [11].

The purity of phycocyanin is assessed using the A620/A280 ratio, where A620 represents the absorbance of phycocyanin at 620 nm and A280 corresponds to the absorbance of other proteins at 280 nm. This purity index will determine its market value and, consequently, its applicability. Phycocyanin is considered food, cosmetic, reagent, or analytic grade when A620/A280 is, respectively, greater than 0.7, 1.5, 3.9, or 4.0 [11]. Furthermore, enhancing methods or processes of extraction (e.g., freeze/thaw, mixing/homogenization, bead milling, ultrasonication, electric field, high-pressure homogenization, and enzymatic extraction) and purification (e.g., ammonium sulfate precipitation, membrane filtration, chromatography), as well as developing more stable formulations (e.g., light, temperature, pH, preservatives, biomass/solvent ratio), can significantly increase the commercial value of phycocyanin [11, 15, 24]. In fact, it is estimated that the market of phycocyanin will grow at a Compound Annual Growth Rate (CAGR) of 28.1% from 2023 to 2030, reaching $279.6 million. Also, the global phycocyanin market is anticipated to reach 3587.2 tonnes by 2030, expanding at a CAGR of 33.8% between 2023 and 2030 [15].

Nevertheless, several factors will determine phycocyanin production, including environmental conditions, cultivation methods, and the genetic diversity among strains or populations of the same microalgae species [11, 13, 16]. The phycocyanin molecular structure, composition, and purity will establish its potential biological functions including antioxidant, anti-cancer, anti-inflammatory, and neuroprotective activities [11, 12].

Advertisement

3. Antioxidant activity

Phycocyanin derived from different microalgae species has demonstrated the capacity to mitigate damage induced by reactive oxygen species (ROS) and free radicals, thus endowing it with antioxidant properties [25]. This antioxidative potential of phycocyanin has received significant interest in biomedical research due to its potential therapeutic applications. Moreover, studies suggest that phycocyanin can modulate antioxidant enzymes, enhancing cellular defenses against oxidative stress. Additionally, phycocyanin has been shown to reduce lipid peroxidation and prevent DNA damage induced by oxidative stress.

3.1 Free radical scavenging

Free radicals are highly reactive molecules characterized by unpaired electrons, rendering them unstable and capable of damaging cells, proteins, and DNA through oxidative stress. Elevated levels of free radicals are implicated in the onset of various diseases, such as cancer, cardiovascular, metabolic, and neurodegenerative diseases [26]. The mechanism through which these reactive molecules are neutralized and eliminated is known as free radical scavenging. Antioxidant compounds play a pivotal role in this process by facilitating the donation of electrons or hydrogen groups, thereby neutralizing free radicals and avoiding their propagation within cells [26]. In phycocyanin, its chromophore named phycocyanobilin has a special structure that allows it to directly neutralize free radicals by donating an electron, thus preventing cell damage [27].

Phycocyanin (200 μg/mL) extracted from the cyanobacteria Geitlerinema sp. revealed a scavenging activity of 78.75% and a hydrogen-peroxide-radical-scavenging activity of 95.27%, suggesting it as a promising pharmaceutical and nutraceutical compound [28]. Another study on 150 μg/mL phycocyanin extracted from this species demonstrated a reducing power efficiency of 85.15%, recognizing the correlation between the reducing power and antioxidant activity of phycocyanin [29].

Purified phycocyanin from Synechococcus sp. displayed 89% antioxidant activity of 2,2-diphenyl-1-picryhydrazyl (DPPH) and 11% of superoxide free radical scavenging activity [30]. Another study demonstrated its proton-donating ability and a 70% scavenging capacity at a dose of 80 μg [31].

Phycocyanin obtained from Pseudanabaena sp. and Limnothrix sp. cyanobacteria revealed antioxidant properties assessed by the 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid (ABTS) method using concentrations ranging from 0.25 to 1.00 mg/mL [32].

The antioxidant activity of phycocyanin isolated from three cyanobacterial species Lyngbya, Phormidium, and Spirulina was also studied in vitro. The results demonstrated that phycocyanin from Lyngbya, Phormidium, and Spirulina were able to scavenge peroxyl radicals with a relative rate constant ratios of, respectively, 3.13, 1.89, and 1.80. Interestingly, phycocyanin extracted from Lyngbya was found to be a more effective inhibitor of peroxyl radicals (IC50 6.63 μM), compared to Spirulina (IC50 12.15 μM) and Phormidium (IC50 12.74 μM) phycocyanin. In this study, the electron spin resonance spectra of phycocyanin indicated the presence of free radical active sites, which may play an important role in its radical scavenging property [33].

Moreover, phycocyanin at 1 mg/mL isolated from Oscillatoria minima showed 44% of DPPH radical scavenging activity and 95% of ABTS radical scavenging activity [34].

Purified phycocyanin at 0.15 mg/mL isolated from Limnothrix sp. revealed 100% antioxidant activity measured by DPPH [35].

A study using phycocyanin extract (10–150 nM) obtained from Aphanizomenon flos-aquae demonstrated a protective role against oxidative damage induced by a pro-oxidant agent in human erythrocytes [20]. In this study, the authors demonstrated a decreased production of free radicals after phycocyanin extract treatment [20].

Interestingly, a study has shown that the antioxidant activity of phycocyanobilin from Spirulina is comparable to that of phycocyanin when adjusting the concentrations of phycocyanobilin yield. This finding suggests that phycocyanobilin likely accounts for the majority of the antioxidant activity associated with phycocyanin [36]. Other studies demonstrated the higher capacity of allophycocyanin and phycocyanin to scavenge, respectively, peroxyl and hydroxyl radicals [37].

3.2 Boosting antioxidant enzymes

Antioxidant enzymes play a pivotal in shielding cells against oxidative harm by counteracting detrimental free radicals and ROS. Superoxide dismutase (SOD) catalyzes the conversion of superoxide radicals (O2) into hydrogen peroxide (H2O2) and molecular oxygen (O2), preventing oxidative damage to cells and tissues [38]. Similarly, catalase facilitates the breakdown of H2O2 into water (H2O) and O2, thereby diminishing the accumulation of H2O2 [39]. Hydrogen peroxide, a reactive oxygen species, can induce oxidative stress and harm cellular components. Catalase aids in neutralizing hydrogen peroxide, safeguarding cells from oxidative injury. Glutathione peroxidase (GPx) reduces H2O2 to H2O, limiting its harmful effects, while also mediating growth factor-mediated signal transduction and mitochondrial function, and preserving normal thiol redox-balance [40]. Glutathione reductase (GR) generates reduced glutathione (GSH), an antioxidant molecule protecting cells from oxidative damage, from its oxidized form (GSSG) [41].

A study demonstrated that a 5 mg/mL phycocyanin dosage from A. platensis mitigated cisplatin-induced mortality and inflammation by increasing GPx, γ-glutamyl transferase, and glutathione levels in the liver, along with catalase and SOD levels in the kidney in mice [42]. Supplementation of fish diet with phycocyanin from A. platensis also exhibited an increase of SOD, catalase, and GPx [43]. Moreover, the administration of 0.45 g of phycocyanin from A. platensis per day resulted in an increase in the liver’s reduced-to-oxidized glutathione ratio and a decrease in liver gene expression of SOD and GPx in male mice born from supplemented mothers in a model of atherosclerosis [44].

Phycocyanin from A. maxima at doses of 250 μg/mL in vitro and 400 mg/kg in in vivo models, prevented SOD and GPx activity, indicating a protective role against free radical-induced damage caused by carbon tetrachloride-induced hepatocyte damage [45].

Furthermore, a study on Microcystis aeruginosa exhibited a positive correlation between phycocyanin, GSH content, and SOD enzymes, underscoring its antioxidant potential and the regulation of redox-sensitive signaling [19].

3.3 Protecting cells from lipid peroxidation

Lipid peroxidation is a process where ROS attack lipids, particularly polyunsaturated fatty acids (PUFAs), within cell membranes, leading to the formation of lipid peroxides and other toxic compounds [46].

Phycocyanobilin from A. platensis, ranging from 100 to 1000 μM, effectively inhibited the peroxidation of methyl linoleate and the oxidation of phosphatidylcholine liposomes [36]. Additionally, phycocyanin administered at doses between 25 and 50 mg/kg prevented ethanol-induced lipid peroxidation, thereby reducing oxidative stress, ethanol-induced anxiety, and ethanol-induced testicular damage in rats [47, 48].

Moreover, a daily dosage of 50 mg/kg of phycocyanin from A. maxima resulted in a 57% reduction in lipid peroxidation, thus preventing oxidative stress induced by acute myocardial infarction in rats [49].

Phycocyanin extracted from Geitlerinema sp. at 200 μg/mL revealed an anti-lipid peroxidation activity of 53.65% [28].

Treatment of rats with phycocyanin from Plectonema species, ranging from 100 to 200 mg/kg, protected against type 2 diabetes mellitus by maintaining the redox state through the reduction of lipid peroxidation [18].

A phycocyanin extract (10–100 nM) from A. flos-aquae inhibited the extent of lipid peroxidation induced by cupric chloride in human plasma samples [20].

Finally, individual and released phycocyanobilin from phycocyanin in vivo demonstrated its potential to scavenge reactive oxygen and nitrogen species by counteracting lipid peroxidation and inhibiting oxidase enzymes [50].

3.4 DNA damage control

Free radicals can induce DNA damage through oxidative stress. When free radicals interact with DNA molecules, they can trigger chemical modifications to the DNA bases, such as oxidation, thereby interfering with the accurate replication and transcription of DNA [51]. Antioxidants play a crucial role in protecting DNA from damage by neutralizing free radicals and reducing oxidative stress, maintaining genomic integrity.

Pretreatment and treatment of mice subjected to irradiation with 200 mg/kg phycocyanin from Spirulina via oral gavage for 7 days attenuated the expression of H2AX, an indicator of DMA damage in mice [52]. Another study showcased the protective effect of phycocyanin and phycocyanobilin in scavenging peroxynitrite, a molecule known to mediate oxidative damage to DNA [53]. Furthermore, Niu et al. [54] demonstrated the protective effect of phycocyanin on DNA damage and blastocyst autophagy in porcine embryo development. In a colon cancer rat model, a combination of 200 mg/kg of phycocyanin from Spirulina and piroxicam prevented genomic DNA instability and fragmentation [55].

Advertisement

4. Phycocyanin’s antioxidant activity: therapeutic insights

Elevated levels of free radicals, oxidative stress, and dysfunctional immunity are involved in many human diseases. Consequently, the therapeutic potential of phycocyanin, mainly extracted from Spirulina, owing to its potent antioxidant activity, has been extensively explored in biomedical research including inflammatory diseases, cancer, cardiovascular, and neurodegenerative conditions [56]. Spirulina exhibits remarkable adaptability to extreme climatic conditions, guaranteeing reliable and consistent production of phycocyanin. Additionally, these microalgae yield larger quantities of phycocyanin compared to other microalgae species, further enhancing its potential as a research model for exploring the therapeutic potential of phycocyanin’s antioxidant activity [11].

Several studies have reported the effect of phycocyanin on inflammation, resulting in a reduction in the production of tumor necrosis factor-α (TNF-α) and subsequently attenuating neutrophil infiltration [57, 58]. Moreover, phycocyanin has been shown to inhibit the activation of nuclear factor-kappa B (NF-κB) and decrease the production of nitric oxide (NO), inducible nitric oxide synthase (iNOS), and pro-inflammatory cytokines [58]. Research has demonstrated the protective role of phycocyanin in various inflammatory conditions, including acute myocardial infarction, atherosclerosis, liver, and lung injury [11]. Additionally, studies have reported the activity of phycocyanin in attenuating cardiovascular disease. Administration of 50 mg/kg of phycocyanin in rats with acute myocardial infarction reduced levels of ROS, Bax expression, and caspase-9 release [49].

The role of phycocyanin in cancer was demonstrated in liver cancer [59], breast cancer [60], colon cancer [61], leukemia [62], and lung cancer [63]. The potential mechanisms underlying phycocyanin’s anticancer activity against cancer cells include suppressing the cell cycle at specific phases, modifying the redox state of cells, and promoting the expression of various genes and receptors responsible for necrosis and apoptosis [12]. Phycocyanin initiates the activation of genes such as caspase-9 and -3, responsible for DNA fragmentation and cell shrinkage, indicating its significant role in apoptotic pathways. Additionally, it induces cleavage of poly [ADP-ribose] polymerase 1 (PARP-1) and alters the ratio of Bcl-2/Bax [64].

Phycocyanin has also been implicated in neurodegenerative disorders, including multiple sclerosis, acting as a scavenger of peroxynitrite species, an inhibitor of lipid peroxidation and DNA damage, and a promoter of remyelination of damaged brain tissue [65, 66]. Additionally, phycocyanin attenuates the production of enzymes involved in Alzheimer’s disease, such as Aβ production [67]. Further studies demonstrated that phycocyanin binds to the Aβ40/42 peptide, disrupting a key mechanism of Alzheimer’s disease [68]. Other reports showed that daily intraperitoneal injection of phycocyanin attenuated the production of pro-apoptotic (Bax, Bcl-2, caspase-9) and pro-inflammatory mediators (TNF-α, NF-κB, interleukin (IL)-6, IL-1β) [69, 70]. The potential of phycocyanin as a drug for neurodegenerative disorders is an area of interest for future development and utilization.

Advertisement

5. Conclusion

Phycocyanin, a highly valuable protein, is widely used in the food, cosmetic, and pharmaceutical industries due to its biological activities, particularly antioxidative. Advances in microalgae cultivation techniques hold promise for enhancing phycocyanin production, while the development of efficient extraction and purification methods is crucial for obtaining high-quality products. Furthermore, strategies to stabilize phycocyanin during processing and storage are essential to maintain its antioxidant properties over time. Different studies highlight phycocyanin’s antioxidant activity as a valuable natural resource with diverse applications in biomedicine. However, further research efforts are needed to delve deeper into its antioxidative mechanisms and therapeutic potential against various diseases. In fact, phycocyanin holds significant promise as a therapeutic agent for conditions such as cancer, inflammation, neurodegenerative disorders, and metabolic syndromes. However, additional preclinical and clinical studies are necessary to thoroughly assess its efficacy and safety profiles. In summary, the antioxidant activity of phycocyanin represents a valuable asset with broad applications in biomedicine, emphasizing the importance of ongoing research efforts and technological advancements to fully explore its potential for enhancing human health and well-being.

Advertisement

Acknowledgments

This work was supported by National Funds from the FCT-Portuguese Foundation for Science and Technology, under the project 2023.03608.BD and funded by FCT, under the project UIDB/04033/2020 (https://doi.org/10.54499/UIDB/04033/2020, https://doi.org/10.54499/LA/P/0126/2020) [Accessed: April 6, 2024].

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Wainwright CL, Teixeira MM, Adelson DL, et al. Future directions for the discovery of natural product-derived immunomodulating drugs: An IUPHAR positional review. Pharmacological Research. 2022;177:106076. DOI: 10.1016/j.phrs.2022.106076
  2. 2. Siddiqui AJ, Jahan S, Singh R, et al. Plants in anticancer drug discovery: From molecular mechanism to chemoprevention. BioMed Research International. 2022;2022:5425485. DOI: 10.1155/2022/5425485
  3. 3. Nicolaou KC, Rigol S. Total synthesis in search of potent antibody-drug conjugate payloads. From the fundamentals to the translational. Accounts of Chemical Research. 2019;52:127-139. DOI: 10.1021/acs.accounts.8b00537
  4. 4. Futamura Y, Yamamoto K, Osada H. Phenotypic screening meets natural products in drug discovery. Bioscience, Biotechnology, and Biochemistry. 2017;81:28-31. DOI: 10.1080/09168451.2016.1248365
  5. 5. Atanasov AG, Waltenberger B, Pferschy-Wenzig EM, et al. Discovery and resupply of pharmacologically active plant-derived natural products: A review. Biotechnology Advances. 2015;33:1582-1614. DOI: 10.1016/j.biotechadv.2015.08.001
  6. 6. Bu M, Yang BB, Hu L. Natural endoperoxides as drug lead compounds. Current Medicinal Chemistry. 2016;23:383-405. DOI: 10.2174/0929867323666151127200949
  7. 7. Khan MI, Shin JH, Kim JD. The promising future of microalgae: current status, challenges, and optimization of a sustainable and renewable industry for biofuels, feed, and other products. Microbial Cell Factories. 2018;17:36. DOI: 10.1186/s12934-018-0879-x
  8. 8. Wu J, Gu X, Yang D, et al. Bioactive substances and potentiality of marine microalgae. Food Science & Nutrition. 2021;9:5279-5292. DOI: 10.1002/fsn3.2471
  9. 9. Susmita G, Tanmay S, Siddhartha P, Abdul KZ, Atan EH, Runu C. Novel bioactive compounds from marine sources as a tool for functional food development. Frontiers in Marine Science. 2022;9:2296-7745. DOI: 10.3389/fmars.2022.832957
  10. 10. Nowruzi B. Cyanobacteria natural products as sources for future directions in antibiotic drug discovery. In: Cyanobacteria – Recent Advances and New Perspectives. London, UK: IntechOpen; 2022. DOI: 10.5772/intechopen.106364
  11. 11. Fernandes R, Campos J, Serra M, et al. Exploring the benefits of phycocyanin: From spirulina cultivation to its widespread applications. Pharmaceuticals (Basel). 2023;16:592. DOI: 10.3390/ph16040592
  12. 12. Dranseikienė D, Balčiūnaitė-Murzienė G, Karosienė J, et al. Cyano-phycocyanin: Mechanisms of action on human skin and future perspectives in medicine. Plants (Basel). 2022;11:1249. DOI: 10.3390/plants11091249
  13. 13. Morya S, Kumar Chattu V, Khalid W, Zubair Khalid M, Siddeeg A. Potential protein phycocyanin: An overview on its properties, extraction, and utilization. International Journal of Food. 2023;26:3160-3176. DOI: 10.1080/10942912.2023.2271686
  14. 14. Carfagna S, Landi V, Coraggio F, Salbitani G, Vona V, Pinto G, et al. Different characteristics of C-phycocyanin (C-PC) in two strains of the extremophilic Galdieria phlegrea. Algal Research. 2018;31(31):406-412. DOI: 10.1016/j.algal.2018.02.030
  15. 15. Athiyappan KD, Routray W, Paramasivan B. Phycocyanin from spirulina: A comprehensive review on cultivation, extraction, purification, and its application in food and allied industries. Food and Humanity. 2024;2:100235. DOI: 10.1016/j.foohum.2024.100235
  16. 16. Chini Zittelli G, Lauceri R, Faraloni C, Silva Benavides AM, Torzillo G. Valuable pigments from microalgae: Phycobiliproteins, primary carotenoids, and fucoxanthin. Photochemical & Photobiological Sciences. 2023;22:1733-1789. DOI: 10.1007/s43630-023-00407-3
  17. 17. de Morais MG, da Fontoura PD, Moreira JB, Duarte JH, Alberto J. Phycocyanin from microalgae: Properties, extraction and purification, with some recent applications. Industrial Biotechnology. 2018;14:1550-9087. DOI: 10.1089/ind.2017.0009
  18. 18. Husain A, Alouffi S, Khanam A, Akasha R, Farooqui A, Ahmad S. Therapeutic efficacy of natural product 'C-Phycocyanin' in alleviating streptozotocin-induced diabetes via the inhibition of glycation reaction in rats. International Journal of Molecular Sciences. 2022;23:14235. DOI: 10.3390/ijms232214235
  19. 19. Liu S, Wu Z, Min X, et al. Synergism variation between intracellular Glutathione, phycocyanin and SOD in microalgae by carbon quantum dot fluorescence. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy. 2024;310:123833. DOI: 10.1016/j.saa.2023.123833
  20. 20. Benedetti S, Benvenuti F, Pagliarani S, Francogli S, Scoglio S, Canestrari F. Antioxidant properties of a novel phycocyanin extract from the blue-green alga Aphanizomenon flos-aquae. Life Sciences. 2004;75:2353-2362. DOI: 10.1016/j.lfs.2004.06.004
  21. 21. Hotos GN, Antoniadis TI. The effect of colored and white light on growth and phycobiliproteins, chlorophyll and carotenoids content of the marine cyanobacteria Phormidium sp. and Cyanothece sp. in batch cultures. Life (Basel). 2022;12:837. DOI: 10.3390/life12060837
  22. 22. Lauceri R, Chini Zittelli G, Torzillo G. A simple method for rapid purification of phycobiliproteins from Arthrospira platensis and Porphyridium cruentum biomass. Algal Research. 2019;44:101685. DOI: 10.1016/j.algal.2019.101685
  23. 23. Paniagua-Michel J. Chapter16 – Microalgal nutraceuticals. In: Kim S-K, editor. Handbook of Marine Microalgae. Boston: Academic Press; 2015. pp. 255-267. DOI: 10.1016/B978-0-12-800776-1.00016-9
  24. 24. Yuan B, Li Z, Shan H, et al. A review of recent strategies to improve the physical stability of phycocyanin. Current Research in Food Science. 2022;5:2329-2337. DOI: 10.1016/j.crfs.2022.11.019
  25. 25. Perera RMTD, Herath KHINM, Sanjeewa KKA, Jayawardena TU. Recent reports on bioactive compounds from marine cyanobacteria in relation to human health applications. Life (Basel). 2023;13:1411. DOI: 10.3390/life13061411
  26. 26. Haider K, Haider MR, Neha K, Yar MS. Free radical scavengers: An overview on heterocyclic advances and medicinal prospects. European Journal of Medicinal Chemistry. 2020;204:112607. DOI: 10.1016/j.ejmech.2020.112607
  27. 27. Sagara T, Nishibori N, Kishibuchi R, et al. Non-protein components of Arthrospira (Spirulina) platensis protect PC12 cells against iron-evoked neurotoxic injury. Journal of Applied Phycology. 2025;27:849-855. DOI: 10.1007/s10811-014-0388-1
  28. 28. Renugadevi K, Valli Nachiyar C, Sowmiya P, Sunkar S. Antioxidant activity of phycocyanin pigment extracted from marine filamentous cyanobacteria Geitlerinema sp TRV57. Biocatalysis and Agricultural Biotechnology. 2028;16:237-242. DOI: 10.1016/j.bcab.2018.08.009
  29. 29. Patel HM, Rastogi RP, Trivedi U, Madamwar D. Structural characterization and antioxidant potential of phycocyanin from the cyanobacterium Geitlerinema sp. H8DM. Algal Research. 2018;32:372-383. DOI: 10.1016/j.algal.2018.04.024
  30. 30. Lin JY, Tan SI, Yi YC, et al. High-level production and extraction of C-phycocyanin from cyanobacteria Synechococcus sp. PCC7002 for antioxidation, antibacterial and lead adsorption. Environmental Research. 2022;206:112283. DOI: 10.1016/j.envres.2021.112283
  31. 31. Sonani RR, Patel S, Bhastana B, et al. Purification and antioxidant activity of phycocyanin from synechococcus sp. R42DM isolated from industrially polluted site. Bioresource Technology. 2017;245(Pt A):325-331. DOI: 10.1016/j.biortech.2017.08.129
  32. 32. Aoki J, Sasaki D, Asayama M. Development of a method for phycocyanin recovery from filamentous cyanobacteria and evaluation of its stability and antioxidant capacity. BMC Biotechnology. 2021;21:40. DOI: 10.1186/s12896-021-00692-9
  33. 33. Patel A, Mishra S, Ghosh PK. Antioxidant potential of C-phycocyanin isolated from cyanobacterial species Lyngbya, Phormidium and Spirulina spp. Indian Journal of Biochemistry & Biophysics. 2006;43:25-31
  34. 34. Venugopal VC, Thakur A, Chennabasappa LK, et al. Phycocyanin extracted from Oscillatoria minima shows antimicrobial, algicidal, and antiradical activities: In silico and In vitro analysis. Anti-Inflammatory & Anti-Allergy Agents in Medicinal Chemistry. 2020;19:240-253. DOI: 10.2174/1871523018666190405114524
  35. 35. Gantar M, Simović D, Djilas S, Gonzalez WW, Miksovska J. Isolation, characterization and antioxidative activity of C-phycocyanin from Limnothrix sp. strain 37-2-1. Journal of Biotechnology. 2012;159:21-26. DOI: 10.1016/j.jbiotec.2012.02.004
  36. 36. Hirata T, Tanaka M, Ooike M, et al. Antioxidant activities of phycocyanobilin prepared from spirulina platensis. Journal of Applied Phycology. 2000;12:435-439. DOI: 10.1023/A:1008175217194
  37. 37. Cherdkiatikul T, Suwanwong Y. Production of the α and β subunits of spirulina allophycocyanin and C-Phycocyanin in Escherichia coli: A comparative study of their antioxidant activities. Journal of Biomolecular Screening. 2014;19:959-965. DOI: 10.1177/1087057113520565
  38. 38. Wang Y, Branicky R, Noë A, Hekimi S. Superoxide dismutases: Dual roles in controlling ROS damage and regulating ROS signaling. The Journal of Cell Biology. 2018;217:1915-1928. DOI: 10.1083/jcb.201708007
  39. 39. Lee SE, Park YS. The emerging roles of antioxidant enzymes by dietary phytochemicals in vascular diseases. Life (Basel). 2021;11:199. DOI: 10.3390/life11030199
  40. 40. Pei J, Pan X, Wei G, Hua Y. Research progress of glutathione peroxidase family (GPX) in redoxidation. Frontiers in Pharmacology. 2023;14:1147414. DOI: 10.3389/fphar.2023.1147414
  41. 41. Couto N, Wood J, Barber J. The role of glutathione reductase and related enzymes on cellular redox homoeostasis network. Free Radical Biology & Medicine. 2016;95:27-42. DOI: 10.1016/j.freeradbiomed.2016.02.028
  42. 42. Zhang Y, Li L, Qin S, et al. C-phycocyanin alleviated cisplatin-induced oxidative stress and inflammation via gut microbiota-metabolites axis in mice. Frontiers in Nutrition. 2022;9:996614. DOI: 10.3389/fnut.2022.996614
  43. 43. Hassaan MS, Mohammady EY, Soaudy MR, Sabae SA, Mahmoud AMA, El-Haroun ER. Comparative study on the effect of dietary β-carotene and phycocyanin extracted from Spirulina platensis on immune-oxidative stress biomarkers, genes expression and intestinal enzymes, serum biochemical in Nile tilapia, Oreochromis niloticus. Fish & Shellfish Immunology. 2021;108:63-72. DOI: 10.1016/j.fsi.2020.11.012
  44. 44. Coué M, Croyal M, Habib M, et al. Perinatal administration of C-phycocyanin protects against atherosclerosis in apoE-Deficient mice by modulating cholesterol and trimethylamine-N-oxide metabolisms. Arteriosclerosis, Thrombosis, and Vascular Biology. 2021;41:e512-e523. DOI: 10.1161/ATVBAHA.121.316848
  45. 45. Ou Y, Zheng S, Lin L, Jiang Q , Yang X. Protective effect of C-phycocyanin against carbon tetrachloride-induced hepatocyte damage in vitro and in vivo. Chemico-Biological Interactions. 2010;185:94-100. DOI: 10.1016/j.cbi.2010.03.013
  46. 46. Valgimigli L. Lipid peroxidation and antioxidant protection. Biomolecules. 2023;13:1291. DOI: 10.3390/biom13091291
  47. 47. Oumayma B, Wahid K, Soumaya G, et al. Phycocyanin improved alcohol-induced hepatorenal toxicity and behavior impairment in Wistar rats. Drug and Chemical Toxicology. 2023;46:1187-1192. DOI: 10.1080/01480545.2022.2139843
  48. 48. Boukari O, Ghoghbane S, Khemissi W, et al. Phycocyanin alleviates alcohol-induced testicular injury in male Wistar rats. Clinical and Experimental Reproductive Medicine. Jun 2024;51(2):102-111. DOI: 10.5653/cerm.2023.06422
  49. 49. Blas-Valdivia V, Moran-Dorantes DN, Rojas-Franco P, et al. C-Phycocyanin prevents acute myocardial infarction-induced oxidative stress, inflammation and cardiac damage. Pharmaceutical Biology. 2022;60:755-763. DOI: 10.1080/13880209.2022.2055089
  50. 50. Piniella B, Prida JM, Pentón-Rol G. Nutraceutical and therapeutic potential of phycocyanobilin for treating Alzheimer’s disease. Journal of Biosciences. 2021;46:42. DOI: 10.1007/s12038-021-00161-7
  51. 51. Juan CA, Pérez de la Lastra JM, Plou FJ, Pérez-Lebeña E. The chemistry of reactive oxygen species (ROS) revisited: Outlining their role in biological macromolecules (DNA, lipids and proteins) and induced pathologies. International Journal of Molecular Sciences. 2021;22:4642. DOI: 10.3390/ijms22094642
  52. 52. Liu Q , Li W, Qin S. Therapeutic effect of phycocyanin on acute liver oxidative damage caused by X-ray. Biomedicine & Pharmacotherapy. 2020;130:110553. DOI: 10.1016/j.biopha.2020.110553
  53. 53. Bhat VB, Madyastha KM. Scavenging of peroxynitrite by phycocyanin and phycocyanobilin from Spirulina platensis: protection against oxidative damage to DNA. Biochemical and Biophysical Research Communications. 2001;285:262-266. DOI: 10.1006/bbrc.2001.5195
  54. 54. Niu YJ, Zhou W, Guo J, et al. C-Phycocyanin protects against mitochondrial dysfunction and oxidative stress in parthenogenetic porcine embryos. Scientific Reports. 2017;7:16992. DOI: 10.1038/s41598-017-17287-0
  55. 55. Saini MK, Vaiphei K, Sanyal SN. Chemoprevention of DMH-induced rat colon carcinoma initiation by combination administration of piroxicam and C-phycocyanin. Molecular and Cellular Biochemistry. 2012;361:217-228. DOI: 10.1007/s11010-011-1106-9
  56. 56. Wu Q , Liu L, Miron A, Klímová B, Wan D, Kuča K. The antioxidant, immunomodulatory, and anti-inflammatory activities of Spirulina: An overview. Archives of Toxicology. 2016;90:1817-1840. DOI: 10.1007/s00204-016-1744-5
  57. 57. Shih CM, Cheng SN, Wong CS, Kuo YL, Chou TC. Antiinflammatory and antihyperalgesic activity of C-phycocyanin. Anesthesia and Analgesia. 2009;108:1303-1310. DOI: 10.1213/ane.0b013e318193e919
  58. 58. Cherng SC, Cheng SN, Tarn A, Chou TC. Anti-inflammatory activity of c-phycocyanin in lipopolysaccharide-stimulated RAW 264.7 macrophages. Life Sciences. 2007;81:1431-1435. DOI: 10.1016/j.lfs.2007.09.009
  59. 59. Jiang L, Wang Y, Yin Q , et al. Phycocyanin: A potential drug for cancer treatment. Journal of Cancer. 2017;8:3416-3429. Published 2017 Sep 20. DOI: 10.7150/jca.21058
  60. 60. Heisnam R, Keithellakpam OS, Kshetrimayum V, Mukherjee PK, Sharma N. Phycocyanin purified from Westiellopsis sp. induces caspase 3 mediated apoptosis in breast cancer cell line MDA-MB-231. Algal Research. 2022;68:102852. DOI: 10.1016/j.algal.2022.102852
  61. 61. Wen Y, Wen P, Hu T-G, Linhardt RJ, Zong M-H, Wu H, et al. Encapsulation of phycocyanin by prebiotics and polysaccharides-based electrospun fibers and improved colon cancer prevention effects. International Journal of Biological Macromolecules. 2020;149:672-681. DOI: 10.1016/j.ijbiomac.2020.01.189
  62. 62. Subhashini J, Mahipal SV, Reddy MC, Mallikarjuna Reddy M, Rachamallu A, Reddanna P. Molecular mechanisms in C-Phycocyanin induced apoptosis in human chronic myeloid leukemia cell line-K562. Biochemical Pharmacology. 2004;68:453-462. DOI: 10.1016/j.bcp.2004.02.025
  63. 63. Chaowen H, Dongxuan H, Dongsheng H, et al. C-Phycocyanin Suppresses Cell Proliferation and Promotes Apoptosis by Regulating the AMPK Pathway in NCL-H292 Non-Small Cell Lung Cancer Cells. Folia Biologica(Praha). 2022;68:16-24
  64. 64. Fernandes E, Silva E, Figueira FDS, Lettnin AP, et al. C-Phycocyanin: Cellular targets, mechanisms of action and multi drug resistance in cancer. Pharmacological Reports. 2018;70(1):75-80. DOI: 10.1016/j.pharep.2017.07.018
  65. 65. Pentón-Rol G, Martínez-Sánchez G, Cervantes-Llanos M, et al. C-Phycocyanin ameliorates experimental autoimmune encephalomyelitis and induces regulatory T cells. International Immunopharmacology. 2011;11:29-38. DOI: 10.1016/j.intimp.2010.10.001
  66. 66. Pentón-Rol G, Marín-Prida J, Falcón-Cama V. C-Phycocyanin and Phycocyanobilin as Remyelination Therapies for Enhancing Recovery in Multiple Sclerosis and Ischemic Stroke: A Preclinical Perspective. Behavioral Sciences (Basel). 2018;8:15. DOI: 10.3390/bs8010015
  67. 67. Koh EJ, Kim KJ, Choi J, Kang DH, Lee BY. Spirulina maxima extract prevents cell death through BDNF activation against amyloid beta 1-42 (Aβ1-42) induced neurotoxicity in PC12 cells. Neuroscience Letters. 2018;673:33-38. DOI: 10.1016/j.neulet.2018.02.057
  68. 68. Liu Y, Jovcevski B, Pukala TL. C-Phycocyanin from Spirulina Inhibits α-Synuclein and Amyloid-β Fibril formation but not amorphous aggregation. Journal of Natural Products. 2019;82:66-73. DOI: 10.1021/acs.jnatprod.8b00610
  69. 69. Agrawal M, Perumal Y, Bansal S, Arora S, Chopra K. Phycocyanin alleviates ICV-STZ induced cognitive and molecular deficits via PI3-Kinase dependent pathway. Food and Chemical Toxicology. 2020;145:111684. DOI: 10.1016/j.fct.2020.111684
  70. 70. Li Z, Gan L, Yan S, Yan Y, Huang W. Effect of C-phycocyanin on HDAC3 and miRNA-335 in Alzheimer's disease. Translational Neuroscience. 2020;11:161-172. DOI: 10.1515/tnsci-2020-0101

Written By

Joana Campos, Raquel Fernandes and Ana Novo Barros

Submitted: 08 April 2024 Reviewed: 20 May 2024 Published: 05 June 2024