Skip to main content

A review on marine collagen: sources, extraction methods, colloids properties, and food applications

Abstract

The growing interest in valorizing industrial by-products has led researchers to focus on exploring different sources and optimizing collagen extraction conditions over the past decade. While bovine hide, cattle bones, pork, and pig skins remain the most abundant collagen sources, there is a growing trend in the industrial utilization of collagen from non-mammalian species. This review explores alternative marine collagen sources and summarizes emerging trends in collagen recovery from marine sources, with a particular focus on environmentally friendly methods. Additionally, this review covers the colloidal structure-forming properties of marine collagens, including foam, film, gel, and emulsion formation. It also highlights the potential and important applications of marine collagen in various food products. Based on the currently reported marine sources, collagens extracted from fish, jellyfish, and sea cucumbers were found to have the highest yield and mostly comprised type-I collagen, while crustaceans and mollusks yielded lower percentages of collagen. Traditional extraction techniques isolate collagen based on acetic acid and pepsin treatment, but they come with drawbacks such as being time-consuming, causing sample destruction, and using solvents. Conversely, marine collagen extracted using conventional methods assisted with ultrasonication resulted in higher yields and strengthened the triple-stranded helical structures. Recently, an increasing number of new applications have been found in the food industry for marine collagens, such as biodegradable film-forming materials, colloid stabilizers, foaming agents, and micro-encapsulating agents. Furthermore, collagen is a modern foodstuff and is extensively used in the beverage, dairy, and meat industries to increase the stability, consistency, and elasticity of products.

Graphical abstract

1 Introduction

Collagen, a group of structural proteins in the extracellular matrix with a fibrillar arrangement, contributes to the integrity and mechanics of bio-tissues [1]. Found ubiquitously in living organisms, collagen maintains conserved forms in gene and amino acid sequences, notably in the triple-helix structure. Collagen can also be very abundant, especially in mammals, constituting up to 30% of total proteins and serving as vital components in tissues such as skin, bone, and cartilage [2]. Currently, 29 types of collagen have been identified, each differing in bio-physiological properties, morphological structure, amino acid sequence, and distribution [1]. Among the 29 identified collagen types, type I is the most common, characterized by a triple-helix structure [2]. The collagen molecule, called tropocollagen, is structurally organized in a lengthy amino acid sequence, consisting of three domains, i.e., –COOH non-triple-helix terminal (C-terminal), triple-helix, and –NH2 non-triple-helix terminal (N-terminal), primarily stabilized by hydrogen bonds [3]. These ~ 300 nm long to ~ 1.5 nm diameter collagen molecules self-assemble into collagen fibrils (diameter ~ 100 nm), and thousands (~ 4000) of these fibrils can intertwine to create collagen fibers, which further pack and crosslink to generate diverse hierarchical and robust bulk collagen materials (Fig. 1) [3]. The hydrogen and covalent bonds stabilizing the collagen triple helix are broken during the extraction process, resulting in a polypeptide mixture called gelatin, i.e., collagen partially degrades to form gelatin. Gelatin, produced via partial hydrolysis of collagen, has molecular weights ranging from 15 to 250 kDa [4]. Two types of gelatin, type A and B, can be obtained based on pre-treatment methods like alkaline or acid conditions [5]. Enzyme pre-treatment targeting specific labile peptide bonds is also involved in gelatin production. The final product comprises polypeptides of varied conformations and sizes due to diverse pre-treatment and extraction procedures. It may include higher molecular weight fractions, such as β-chains (covalently linked α-chains dimers), γ-chains (covalently linked α-chain trimmers), and microgels with higher orders [5]. Despite its denatured state, gelatin retains the same amino acid composition as collagen [4].

Fig. 1
figure 1

Multi-hierarchical structure of type I collagen

Collagen derived from land-based animals, such as pigs, cattles, and chickens, finds extensive applications in tissue repair, medical devices, pharmaceuticals, and the food industry. Despite its widespread use, challenges arise, including potential immune reactions, limited availability, high costs, and substantial land requirements for land-based animals [6]. Concerns about infectious diseases, like transmissible spongiform encephalopathy, bovine spongiform encephalopathy, foot and mouth disease, have limited the use of collagen from these sources. Religious prohibitions also impact a substantial global population. Conversely, marine collagen exhibits physicochemical characteristics similar to mammalian collagen, with distinct advantages such as a lower denaturation temperature, lower denaturation weight, reduced risk of disease transmission, simpler extraction methods, and minimal inflammatory responses [7]. Marine collagen molecules consist of three α-chains (α1-α1-α2) of about 1000 amino acid residues or multi-peptide α-chains, forming a stable triple helix structure arranged vertically and bilaterally in a periodic fiber structure. The high stability of the characteristic triple helix region in marine collagen mirrors its similarity in quaternary structure and essential amino acid composition to terrestrial mammalian collagen [8]. Despite this similarity, the marine environment's higher complexity and diversity result in varied marine collagen structures influenced by species, origin, growth cycle, season, environment, and other factors. As a consequence, there are subtle differences in the structure and composition of marine collagen compared to collagen from terrestrial animals.

The global demand for functional proteins is projected to reach $7.98 billion by 2026, experiencing a 6.93% annual growth rate from 2019 to 2026 [7]. Collagen, a crucial macromolecule derived from marine vertebrates and invertebrates, is widely utilized in the pharmaceuticals, food, and healthcare industries. The increased demand for collagen and its hydrolyzed gelatin is evident in various sectors, including food packaging, beverages, dairy products, meat-based items, drug delivery systems, and food supplements. Recent data indicates a 9.4% growth in the global collagen market from 2015 to 2023, with a projected market value of $9.37 billion by 2023, up from $4.13 billion in 2014, primarily sourced from marine origins [1, 9].

According to current research reports, collagen extraction and separation methods from various marine sources involve acidic, enzymatic, ultrasound-assisted, and physical-aided techniques [10,11,12]. Researchers have also introduced environmentally friendly processes, such as extrusion-hydro-extraction [13], supercritical fluid extraction [14] and deep eutectic solvent extraction [15] to enhance yield and reduce pollution. The distribution of collagen fibers, binding tightness, and cross-linking degree with other components impact separation difficulty, extraction rate, purity, and structural integrity [16, 17].

To date, no review article has comprehensively summarized and discussed the diverse marine collagen sources, including sea cucumbers, mollusks, marine sponges, fish, jellyfish, and crustaceans. Addressing this gap, the current review provides a comprehensive exploration of prevalent novel marine collagen sources, established and developing methods employed in their extraction, and the associated challenges. Additionally, this review aims to provide an updated and detailed overview of the colloidal structure-forming properties of marine collagens, such as gels, films, foams, and emulsions. Subsequently, the paper delves into various applications of marine collagen, encompassing food packaging materials, beverages, dairy products, supplements, and meat-based items.

2 Marine collagen sources

Collagens from various marine sources, including, sea cucumbers, mollusks, sponges, crustaceans, jellyfish, and particularly fish, have been extracted and characterized to varying extents. These sources exhibit significant variations in their physicochemical properties (Table 1).

Table 1 Physicochemical properties of collagens derived from marine sources

2.1 Sea cucumbers

Sea cucumbers, classified as invertebrates in the Holothuroidea class within the phylum Echinodermata, are traditional food sources in Korea, Japan, China, and parts of Southeast Asia [47]. Often referred to as gamat or bêche-de-mer, these marine invertebrates comprise over 1250 species globally, many of which are not only edible but also possess significant nutritional and bioactive functions [48]. The nutritional content and chemical composition of sea cucumbers may vary based on the species and growing environment. For example, Holothuria arenicola sea cucumbers from the Bndar-e-lengeh coast in southern Iran showed protein, moisture, ash, crude fiber, and fat percentages of 24.4%, 69.5%, 10.9%, 2.29%, and 2.9%, respectively, while Holothuria parva sea cucumbers had percentages of 17.6%, 67.8%, 32.7%, 1.97%, and 2.4% [49].

The primary edible part of the sea cucumber is its body wall, where collagen constitutes approximately 70% of total protein, plays a crucial role [47]. The sea cucumber body wall is a unique mutable collagenous tissue, fabricated from basic structural components such as collagen, proteoglycan and glycoprotein. These elements assemble into collagen fibrils, microfibrils, and collagen fibers, with insoluble collagen fibrils dominating the majority of the total body wall proteins [48]. Collagen fibers, encircled by a microfibrillar network, contribute to organizational stability and provide a long-range restoring force. Senadheera, Dave [3] investigated the structures and biomechanics of collagen's microfibrillar network derived from the sea cucumber (Cucumaria frondosa), observing morphological characteristics similar to vertebrate fibrillin microfibrils.

The collagen in sea cucumbers is characterized by symmetrically spindle-shaped and short collagen fibrils in echinoderms. Molecularly, these fibrils are bipolar and associate with surface-bound proteoglycans. Internal covalent crosslinks, which are similar to mammalian collagen, contribute to collagen stabilization. Solubilized collagen from the body wall of the sea cucumber exhibits a distinctive triple helix structure built by three homologous α1 chains as (α1)3 and is notably rich in glutamic acid [50]. Tian, Xue [47] employed bioinformatic methods and proteomic techniques to study the constituents within sea cucumber collagen. Phylogenetic analyses of collagen sequences indicated that these sequences did not align with typical collagen branches. The authors established that the complex and heterogeneous nature of sea cucumber collagen warrants thorough investigations.

Sea cucumbers predominantly contain type I collagen, playing a crucial role in the food quality, particularly the textural properties, of sea cucumbers and their processed products (e.g., microwave-dried, freeze-dried, salt-dried, boiled-dried, and ready-to-eat products). Previous studies, including Acaudina leucoprocta [32], Stichopus monotuberculatus [51], Stichopus japonicus [50], Holothuria parva [35], Stichopus vastus [33], and Parastichopus californicus [31], focusing on sea cucumber autolysis and processing activities such as soaking, boiling, drying, and rehydration attribute changes in food properties to sea cucumber collagen. In addition, rheological properties of sea cucumber collagen, indicative of its molecular structure and chain conformation, significantly impact the physical characteristics of sea cucumber products. Liu, Oliveira [31] found that the reduced rheological properties of collagen from the giant red sea cucumber (Parastichopus californicus) compared to calf skin collagen, was ascribed to a lower imino acid residue content in both connective tissue and skin (15.3% and 14.2%, respectively). These values were even lower than those observed in cod or walleye pollock. Additionally, sea cucumber collagen exhibited lower thermal stability than terrestrial animal collagen. For example, the calculated Td values for collagen from the connective tissue and skin of the giant red sea cucumber were 17.9 and 18.5 °C [31], respectively, markedly lower by about 19.1 and 18.5 °C than the Td value of porcine skin collagen (Td = 37.0 °C) [47].

2.2 Crustaceans

Crustaceans, which belong to the Phylum Arthropoda, encompass a varied range of marine invertebrates, such as shrimp, mantis shrimp, lobsters, crabs, prawns, krills, crayfish, ostracods, and copepods. Crustaceans are globally produced in large quantities, with an annual value of over $57 billion, equating to 23% of the total global aquaculture market. Notably, according to the 2021 China Fishery Yearbook, China alone recorded a global shrimp production of approximately 2.69 million tons in 2021 [52]. Several large crustaceans, such as lobsters, shrimp, crabs, and prawns, are consumed as food due to their excellent nutritional characteristics, including high unsaturated fatty acids, protein, and essential micro-elements (e.g., Mg2+, Ca2+, Zn2+) [52]. Additionally, the processing of crustacean seafood (crabs, lobsters, shrimp, and prawns) generates substantial waste in the form of carapace and head [40]. Approximately 60–85% of lobster, 60–70% of crab, 65–70% of crayfish, 70–75% of krill, and 60–80% of shrimp are discarded during processing [53]. These by-products can be separated into their chemical components: 15–40% chitin, 20–40% protein, and 20–50% calcium carbonate [54].

Crustacean muscles primarily consist of myofibrillar proteins, with sarcoplasmic proteins also present, while collagen constitutes only a small fraction of the total protein content [55]. For instance, Hiransuchalert, Oonwiset [41] isolated collagen from the muscles of four different mantis shrimp species: Odontodactylus cultrifer, Erugosquilla woodmasoni, Harpiosquilla harpax, and Miyakella nepa. The collagen exhibited type I characteristics similar to vertebrate muscles, despite low content percentages ranging from 0.015% to 0.4878%. Noteworthy collagen content variations may exist among different crustaceans along with differences in collagen types. Li, Han [56] found that the collagen concentration in the skeletal muscles of three crustacean species (spiny lobster, fleshy prawn, and giant river prawn) ranged from 2.4 to 2.6% of the total protein. In muscle tissues, collagen content showed variations, such as 5.9–6.2% in squilla (quilla Oratosquilla oratoria), 3.4–3.6% in crayfish (Procambarus clarkia), 2.6–2.9% in prawn (Penaeus japonicas), 2.5–2.7% in lobster (Panulirus Iongipes), and 1.1–2.2% in the shrimp (Pandalus borealis). For thoracic and pereiopod muscles of four crab species (Chionoecetes opilio, Erimacrus isenbeckii, Portunus trituberculatus, and Charybdis japonica), the collagen content varied from 0.2 to 0.8% [57].

2.3 Marine sponges

Marine sponges or poriferans, belonging to the phylum Porifera, constitute a diverse group of filter-feeding benthic invertebrates, predominantly inhabiting saltwater but occasionally found in freshwater (around 220 species) [58]. Unlike other animal groups, sponges exhibit a simple organizational structure devoid of real organs and tissues. They are composed of gelatinous internal tissue (mesohyl) surrounded by a layer of epithelial cells, including pinacocytes and choanocytes. These cells are integrated into a complex 3-D matrix network rich in collagen, commonly referred to as spongin or spongin-like collagen [45]. Spongin, a collagenous protein found in the exoskeleton of certain sponges, forms a complex fibrous network that provides flexural rigidity to the sponge. However, its precise and complete molecular composition remains unclear due to the extensive diversity within this sponge family, posing a significant challenge for future elucidation [58]. For example, Araújo, de Souza [44] reported a fibrillar structure in spongin-like collagen from Chondrilla caribensis, contrasting with the nodular/particulate aggregate structure observed in the spongin-like collagen from Aplysina fulva. Tziveleka, Ioannou [46] suggested that collagen extracted from the demosponges Suberites carnosus and Axinella cannabina can be differentiated based on various characteristics, including thermal behavior, solubility, amino acid composition, isoelectric point, and microscopic observations. The study's findings indicated that Axinella cannabina had intercellular collagen, insoluble collagen, and spongin-like collagen content of 3.0%, 12.6%, and 42.8%, respectively, while for Suberites carnosus, they were 1.9%, 5.0%, and 21.9%.

Despite the identification of around 15,000 sponge species, only a few dozen have been explored for collagen extraction and characterization, highlighting the largely untapped biotechnological potential of marine sponges [58]. While sponge collagen holds promise due to its unique physicochemical properties, its industrial/large-scale production faces two key obstacles: (1) sponge-derived collagen typically exhibits lower thermal stability compared to that of homeothermic mammals and birds, given the poikilothermic nature of sponges living in lower temperatures than the human body [46]. (2) Despite the growing demand for sponge-derived natural products, there is a scarcity of sustainable and economically viable culture methods to generate substantial and stable sponge biomasses [58].

2.4 Jellyfish

Jellyfish, also known as 'medusae,' have been consumed as a traditional food source in numerous Asian countries, notably in Japan and China, for over 1700 years due to their nutritional and pharmacological value [59]. Global jellyfish productions now surpass 800,000 tons/year, exceeding catches of popular seafood such as lobsters, clams, and mussels. Despite the existence of over 1400 jellyfish species worldwide, only 23 species have been extensively explored as a sustainable, collagen-rich food source [59]. With a richness in collagenous protein and minerals, jellyfish's low-fat and low-calorie content makes them as an ideal, healthy seafood commodity. De Rinaldis, Leone [60] highlighted jellyfish as an untapped marine collagen resource. Jellyfish collagen exhibits potential advantages over terrestrial animal collagens, including lower inflammatory and immunogenic responses, as well as fewer biological toxins and contaminants [59]. León-Campos, Claudio-Rizo [61] reported the high biocompatibility, low potential for transmitting zoonotic diseases to humans, and low allergic response (requiring caution for consumers with fish allergies) associated with jellyfish collagen. The invertebrate nature of jellyfish suggests the production of unique and commercially appealing products with novel physicochemical and functional properties.

Fresh jellyfish exhibit a proximate composition that differs slightly among species, primarily comprising minerals (1–2%), moisture (95–97%), and crude protein (1–3%) with a significant portion of the protein up to 75% being collagen [62]. The types of jellyfish collagen, akin to other marine and mammalian collagens, vary across species, and no two species share identical collagen makeup. Commonly, types I and II collagen are identified, with occasional examples of types III, IV, and V. Contrary to conventional collagen-type nomenclature, Smith, Domingos [63] suggested that jellyfish collagen, particularly from invertebrates, deviates due to negligible calcified tissues, a high collagen-to-insoluble extracts ratio, and similarities in function and morphology to mammalian collagen. Hoyer et al. (2014) reported that the jellyfish collagen in S. meleagris shares similarities with vertebrate collagen type II, as indicated by solubility properties, salting-out concentration, molecular mobility, a high hydroxylysine content, a highly hygroscopic nature, and the absence of disulfide bonds. Despite low sequence similarities to vertebrate collagen, invertebrate collagen from different jellyfish species can exhibit unique physical and functional properties [63].

2.5 Mollusca

Mollusca, a diverse phylum containing over 120,000 species such as the squids (Doryteuthis singhalensis), mussels (Mytilus chilensis), scallops (Patinopecten yessoensis), oysters (Crassostrea gigas), and clams (Meretrix meretrix), displays significant morphological, ecological, and chemical diversity [10]. Distributed across tropical seas, temperate waters, and polar regions, mollusk species can be differentiated based on the size and shape of their bodies. Mollusks are protein-rich, with 80% of their fleshy material suitable for human consumption [64]. The edible portions of mollusks are characterized by high water content, while their dry mass consists of proteins and micro/macro minerals, but is low in fats. For example, Wu, Guo [42] found that the surf clam shell (Coelomactra antiquata) had moisture, protein, carbohydrate, and ash contents of 82.46%, 11.56%, 3.05%, and 2.38%, respectively, with a fat content of only 0.55%. Certain parts of mollusks have received special attention due to their successful use in collagen extraction. For example, the muscle and mantle of the bivalve mollusk Anadara broughtonii contain 6.8% and 7.2% collagen, respectively. Similarly, Mactra chinensis showed 7.1% and 7.9% collagen content, as highlighted by [64].

Studies have shown that the bio-physiological properties, morphological structure, and molecular composition of mollusk collagen have not been fully elucidated owing to the great diversity within this family of mollusks. Vallejos, González [10] provided the first comprehensive proteomic description under industrial conditions of the Chilean mussel (Mytilus chilensis), a species recognized for its favorable characteristics in collagen production. They observed high concentrations of proline, hydroxyproline, and glycine in Chilean mussels collagen, characterized by β and γ bands, indicating significant cross-linking in the telopeptide region. Multiple studies have reported the potential availability of type I collagen in cephalopods [65], squid [65], clam Shell [42], and byssus [10], indicating their suitability as potential raw material sources for food applications. Veeruraj, Arumugam [65] identified type I collagen in squid (Doryteuthis singhalensis) with α1 and α2 chains, showing thermal denaturation temperatures of 34.80–35.70 °C and a glycine residue range of 32.8–33.2%. This highlights the potential of squid skin collagen as a thermostable alternative for commercial use.

2.6 Fish

Fish is globally recognized as a nutritious and sought-after food. Currently, over 33,600 fish species have been identified and classified into four groups based on their habitats: ice-water, cold-water, warm-water, and hot-water. During the production of fish meat, various byproducts like scales, bones, cartilages, and skin are often discarded as waste (about 7.3 million tons/year) by fish processing industries, contributing to environmental pollution [66]. Approximately 30% of fish bone comprises organic collagen, with 60–70% constituting inorganic substances, mainly calcium phosphate and hydroxyapatite [66]. Fish collagen has specific amino acid compositions, including low concentrations of hydroxyproline, proline, and glycine, resulting in a lower denaturing temperature compared to mammalian collagen. According to Caruso [7], fish collagens from skin, bone, cartilage, and scales surpass porcine or bovine collagen in bioavailability, demonstrating greater absorption efficiency (up to 1.5 times) and quicker bloodstream circulation due to their low molecular weight and small particle size.

Fish collagen's structural characteristics have been delineated via amino acid analysis and physicochemical characterization. Research on the imino acids of collagen from various fish species indicates substantial differences in terms of proline and hydroxyproline imino rings, imparting conformational stability and imposing rigid constraints on rotational movement along the N-Cα bond in the backbone. Cruz-López, Rodríguez-Morales [6] reported that the imino acid content in the skin of gulf corvina (Cynoscion othonopterus) (16.8%) was lower than the skin of giant croaker (19.1%) [67], but similar to that from pacific cod (Gadus macrocephalus) (15.9%) [68], bighead carp (16.5%) [24], and grass carp (16.6%) [69]. Differences between imino acid profiles and structures are probably due to factors such as fish species, tissues, origin, habitats, growth cycle, extraction methods, and other factors [69].

The thermal stability of fish collagen can vary based on the relative abundance of imino acids (proline and hydroxyproline), which closely tied to the species' living conditions and body temperature [68]. Fish collagen, influenced by colder habitats, typically displays lower imino acid levels and thermal stability compared to collagen from warm-blooded land animals. For instance, chicken sternal cartilage collagen exhibited a higher imino acid content (23.2%) and a greater denaturation temperature (32.8 °C) than shark cartilage collagen 15.6% and 16.8 °C, respectively [70, 71]. Despite similar imino acid levels to land animals, sturgeon collagen displays a lower denaturation temperature. However, the challenges lie in the low denaturation temperature (25–30 °C for most fish species) and diverse composition of fish collagen, hindering its applications. While strict comparisons are challenging due to methodological differences across studies, the denaturation temperatures of collagens in deep-sea redfish, inhabiting an ocean temperature of 3–8 °C, resembled those of cold-water fish like chum salmon (19.4 °C), Argentine hake (10 °C), Baltic cod (15 °C), and Alaska Pollack (16.8 °C) [24, 68, 69, 71]. Notably, these temperatures were substantially lower than those observed in temperate and tropical fish species, including common Nile perch (36.5 °C), ayu (29.7 °C), Japanese seabass (26.5 °C), skipjack tuna (29.7 °C), eel (29.3 °C), and mackerel (26.1 °C) [72].

3 Extraction methods

Due to collagen's robust structure and its low solubility in water, two key steps are involved for extraction: (a) pretreatment of raw materials; and (b) collagen extraction. Pretreatment phase involves cleaning, dehydration, degreasing, and decalcification of the raw material. Four commonly employed degreasing methods include combined salt solution, enzymatic digestion and lipase degreasing, degreasing solution usage, and organic solution extraction [23]. Non-collagenous proteins, fats, and pigments may be present in some samples, and they are typically removed at this stage using NaOH, alcohols (ethanol or butyl-alcohol), and oxygen peroxide. Additionally, demineralization of the raw material with HCl or EDTA before extraction enhances the efficiency of collagen extraction from mineral-rich body parts, such as cartilage [3]. Different methods designed for collagen extraction from marine sources are discussed below:

3.1 Acidic extraction

Collagen proteins, predominantly types I, II, III, and V, are fibrous and exhibit lower solubility in aqueous mediums than in acidic mediums, and therefore, are typically extracted through acidic treatments following the removal or degradation of noncollagenous molecules [23]. The mechanism involves enhanced repulsive forces between tropocollagen molecules under acidic conditions, leading to the solubilisation of less cross-linked collagens, resulting in acid-soluble collagen. This process breaks down various amino acids within the sample, achieving the dissolution of non-covalent intramolecular and intermolecular bonds through non-selective chemical hydrolysis of collagen chains. Typically, 0.5 M acetic acid is commonly employed because it attains the highest yield (90%), while diluted organic (chloroacetic, citric and lactic) and inorganic (hydrochloric) acids result in a relatively low yield (≤ 20%) [22]. These acidic solutions offer a notable advantage by effectively disrupting Schiff bases and ionic bonds between molecules at low acidic concentrations [22, 23]. The extractability of collagens depends on marine age and tissue type, with extraction factors including temperature, treatment duration, acid concentration, and the ratio of acidic solution to raw material [73]. The key drawback of this extraction method is that it is time-consuming (Fig. 2), requiring 2–4 days and has a relatively low yield [23].

Fig. 2
figure 2

The advantages and disadvantages of the different extraction methods

The acidic extraction solution, ranging from 0.5 to 1 M, facilitates cleavage of inter- and intramolecular cross-links without compromising the structure of collagen chains. Arumugam, Sharma [74] investigated the impact of acetic acid concentration (0.2–1 M) on the extraction of collagen from sole fish skin, keeping other variables constant. The collagen yield exhibited a gradual increase with acetic acid concentration, reaching a peak at 0.54 M with a maximum yield of 19.27 mg/g. However, concentrations beyond 0.6 M resulted in a decrease in collagen yield. In a study on Baltic cod (Gadus morhua) skin, Skierka and Sadowska [16] reported a higher recovery of collagen using acetic acid (90%), followed by citric acid (60%) and hydrochloric acid (18%). Conversely, Tan and Chang [73] found that HCl-assisted extraction produced the highest collagen yield (42.36%), followed by acetic acid extraction (39.45%). The discrepancy may result from varying pH values in the two studies; the former at pH 2.4 and the latter at pH 0.87. Under lower pH conditions, amine groups bond with anions (Cl), decreasing electrostatic repulsion between charged groups. This leads to tightened collagen fibers, diminishing their water bonding ability and reducing collagen solubility. Additionally, despite equal total H concentration, the ionized H+ in the solutions differed.

3.2 Enzymatic extraction

Traditional collagen extraction involves acidic solutions, but collagens from different sources may not fully dissolve in these mediums, resulting in the retention of intermolecular crosslinks [75]. To enhance collagen yield, enzymatic extraction methods prefer the use of enzymes like pepsin, papain, trypsin, and various collagenases under specific environmental conditions and pH levels [67]. Pepsin, widely used for seafood collagen extraction, is employed alone or with varying acetic acid concentrations, resulting in pepsin-soluble collagens [71]. Enzymatic extraction is preferred for increased efficiency, reduced collagen antigenicity, and preservation of the collagen triple helical structures. Maintaining a low temperature (4–10 °C) during pepsin use is crucial due to its sensitivity to higher temperatures (above 60 °C), risking self-digestion and deactivation [24].

Using pepsin has been found to increase extraction yields and decrease the duration of collagen extraction from giant croaker (Nibea japonica) [67]. The results showed that the extraction yield increased from 66.35 to 79.93% by increasing the pepsin level from 800 to 1200 U/g, and triple-helical structure of collagen was largely intact through hydrogen bonding, suggesting insignificant occurrence of protein breakdown. Most recently, attention has been focused on the use of protease extracted from proteolytic bacteria. For instance, Ahmed, Getachew [76] reported that proteases isolated from Bacillus cereus (FRCY9-2) and Bacillus cereus (FORC005) were employed in extracting collagen from bigeye tuna (Thunnus obesus) skin, resulting in yields of 177.2 and 188 g/kg, respectively, compared to using acid alone (134.5 g/kg).

3.3 Deep eutectic solvent extraction

Deep eutectic solvents, defined as mixtures of two or more components showing a significant melting point depression at a specific composition, transform into liquids at room temperature. These solvents, formed through the interaction between a hydrogen-bond acceptor and a hydrogen-bond donor, have gained prominence as green and sustainable alternatives to conventional industrial processes [15]. The complexing agent, usually a hydrogen-bond donor, interacts with the halide anion, increasing its effective size and reducing the anion interaction with the cation, resulting in mixture melting points far below those of individual components. Particular interest arises when these deep eutectic solvents are composed of natural components (urea, oxalic acid, ethylene glycol, and choline chloride), making the method particularly suitable for extracting volatile aromatic and phenolic compounds, metals, and collagen proteins [77]. Notably, they offer advantages such as cost-effective, numerous combinations, biodegradability and low toxicity [77].

Batista, Fernández [15] introduced a green and sustainable collagen extraction methodology from blue shark by means of a deep eutectic solvent comprising xylitol:citric acid:water at a 1:1:10 molar ratio. This approach resulted in isolating over 21% of the protein content from blue shark skin, surpassing conventional methods by 2.5 times, without the need for raw material pre-treatment and decreasing the procedure time from 96 to 1 h. Among the six deep eutectic solvents studied by Bai, Wei [78], the choline chloride and oxalic acid combination demonstrated remarkable efficacy, achieving an extraction efficiency 90% for cod skins. Furthermore, elevating the oxalic acid quantity resulted in a corresponding increase in free protons within the solution, which, in turn, enhanced interactions with collagen helices, thereby facilitating the leaching of peptides.

3.4 Supercritical fluid extraction

Supercritical fluid extraction has emerged an alternative to traditional solvent extraction by utilizing supercritical fluids, generally carbon dioxide, to extract bioactive compounds from various marine sources, such as microalgae, brown algae, and green seaweed. These fluids, existing above critical pressure and temperature, demonstrate gas-like compressibility, lower density, and potent solvating capabilities similar to liquids. Their high diffusivity facilitates easy penetration into solid materials, enabling dissolution and accessibility [79]. Supercritical fluid extraction offers various advantages over traditional extraction methods, including increased extraction yields, enhanced selectivity, improved fractionation capabilities, and a reduced environmental impact.

Silva, Barros [14] explored the application of supercritical fluid extraction at pressures of 10, 30, and 50 bar for 3 h, using CO2-acidified water for collagen extraction from demosponge (Chondrosia reniformis). The study demonstrated a 30% increase in collagen recovery at 10 bar pressure compared to traditional enzymatic/acidic methods. Similarly, a study by Sousa, Martins [79] utilized supercritical fluid extraction for collagen extraction from Atlantic cod (Gadus morhua) skin. The extraction, conducted with CO2-acidified water at 37 °C and 50 bar for 3 h, resulted in a 13.8% collagen yield. Notably, this method proved significantly more efficient than the conventional approach, which requires nearly 200 h to extract only 10.9% collagen from the same fish species.

3.5 Extrusion-hydro-extraction

Collagen in fish scales is tightly bound to hydroxyapatite, posing challenges in separation. The extrusion–hydro extraction process utilizes extrusion to break the strong linkage between collagen and hydroxyapatite, easing collagen release through water extraction from fish scales [13]. This method provides benefits such as continuous production, minimal labor, reduced waste, high yield, straightforward operation, and versatility in product outcomes. Huang, Kuo [13] introduced a novel extrusion–hydro-extraction method for collagen extraction from tilapia fish scales. Their findings demonstrated that the high pressures reached during extrusion resulted in collagen yields ranging from 7.5 to 12.3%, which was 2–3 times higher compared to non-extruded scale samples. Moreover, the amino acid profiles, and moisture absorption and retention properties of collagens obtained through this process closely resembled those of collagen isolated by conventional methods.

3.6 Physical-aided extraction

Physical-aided collagen extraction methods contribute to enhanced solubilization and tissue homogeneity, significantly improving the efficiency and yield of collagen extraction within a limited timeframe compared to traditional acidic and enzymatic methods [79]. Khong et al. [9] reported a new technique to isolate collagen from marine sources, combining acidic treatment with a sequence of mechanical and physical processes, including adjusting pH, mixing, sonication as well as homogenization. Khong, Yusoff [11] employed these techniques to extract collagen from jellyfish (Acromitus hardenbergi) oral arms and bells. The physical-aided processes led to a remarkable increase in extraction efficiency, with nearly a five times improvement from oral arms and a seven times enhancement from bells compared to acid-assisted extraction. Additionally, there was a two times increase compared to pepsin-assisted extraction. Kuwahara [80] suggested that the combination of acetic acid and ultrafine bubbles of various gases (ozone, carbon dioxide, and oxygen) enhances both the yield and quality of collagen extraction from tilapia scales. The most effective method was using CO2 ultrafine bubbles in a 0.1 M acetic acid solution, resulting in a noteworthy 1.58% collagen yield after 5 h of aeration. Araújo, de Souza [44] introduced a specific procedure for extracting spongin-like collagen from marine sponges (Chondrosia reniformis), involving a mildly basic solution with a chaotropic agent. The procedure included treating sponges with a 100 mM Tris–HCl buffer (8 M urea, 100 mM 2-mercaptoethanol, 10 mM EDTA, and pH 9) for 24 h at room temperature. Subsequently, sponge collagens were solubilized, separated through centrifugation, and further precipitated by decreasing the pH to 4 with the addition of acetic acid, then freeze-dried for preservation of the collagen.

Ultrasonic extraction, utilizing high-intensity sound waves beyond the human hearing limit (20 kHz–10 MHz), has gained significant interest in the food industry owing to its non-toxic, eco-friendly, and cost-effective nature [81]. This method relies on high-frequency sound waves creating low- and high-pressure regions, disrupting cell walls, and boosting acidic and enzymatic treatment, thereby substantially decreasing extraction time compared to conventional methods [81]. The technique enhances extraction efficiency at lower temperatures, aiding in extracting temperature-sensitive proteins with higher yield and minimal damage. Researchers optimized ultrasound effectiveness by adjusting its amplitude, frequency, propagation cycle (discontinuous or continuous), device nominal power, and system geometry, including probe dimensions. Ultrasonic treatment at 200–750 W, amplitude 20–100%, 20–35 kHz, and pulsation 2/2–20/20 s takes about 10–30 min or even 0–24 h. For instance, 300 W ultrasound power with 25 min exposure improved the collagen recovery rate from yellowfin tuna skin to 57.06% [82].

Ali, Kishimura [83] investigated the efficiency of ultrasonication in extracting pepsin- and acid-soluble collagens from golden carp (Probarbus Jullieni). The study's findings revealed that applying ultrasonication at 80% amplitude significantly improved the extraction efficiency, resulting in yields of 81.53% for pepsin-soluble collagen and 94.88% for acid-soluble collagen from the golden carp skin, compared to conventional method yields of 51.90% and 79.27%, respectively. The study suggested that ultrasonication strengthened the triple-stranded helical structures through interchain hydrogen bonding. Kim, Kim [84] reported the improved collagen yield through ultrasonication from sea bass skin (Lateolabrax japonicus) compared to traditional extraction, with no alteration in collagen components.

Despite its promise for yield improvement, high-intensity ultrasound might be destructive. After applying ultrasonication with an intensity of 11.35 kW × cm−2 for 36 min, Kim et al. [77] found that the secondary structure of the collagen was disordered. Additionally, long-term ultrasonication apparently led to destroy α-chain in collagen from sea bass skin. Petcharat, Benjakul [85] studied the impact of ultrasonication on collagen extraction from clown featherback (Chitala ornata) skin, varying amplitudes (20–80%) and durations (10–30 min), observing an enhanced collagen yield from 27.18 to 57.35%. However, ultrasound-induced protein degradation led to reduced hydroxyproline level and compromised collagen purity, especially evident with higher amplitudes and longer durations [84]. Thus, maintaining optimal amplitude and extraction time is crucial for harnessing ultrasound's potential to enhance both quantity and quality of collagen.

3.7 Salt solubilization extraction

Collagen extraction from marine sources using saline solutions (e.g., citrates, phosphates, Tris–HCl, or sodium chloride) is known as salt-soluble collagen. Despite its potential, this technique is rarely employed due to collagen's limited solubility in saline solutions. Liang, Wang [75] reported the successful extraction of salt-soluble collagen from the cartilage of Amur sturgeon (Acipenser schrenckii) using NaCl (0.45 M at a ratio of 1:1000 w/v) with constant stirring for 24 h. Apart from differences in molecular structure, amino acid profile, and thermal stability, there were remarkable differences in yields using salt-soluble collagen (2.18%), acid-solubilized collagen (27.04%), and pepsin-solubilized collagen (55.92%).

4 Collagen-based colloidal structures

Marine collagen is ideal for thickening, texturizing, and gel production due to its high water absorption capacity. Additionally, it exhibits interesting surface characteristics, such as emulsification, foaming, and film-forming properties. These attributes may be attributed to charged groups and hydrophobic/hydrophilic elements in the protein side chains. The source and extraction method significantly impact these functional attributes of collagen.

4.1 Emulsions

An emulsion is a dispersion of small liquid droplets in another immiscible liquid, stabilized by surface-active emulsifiers [86]. These emulsifiers, such as marine collagen-based ones, fall into two categories: soluble molecules and insoluble aggregates. Soluble collagen, which is rich in hydrophobic and hydrophilic amino acids, enhances emulsion stability by lowering interfacial tension and inducing electrostatic repulsion between droplets. This effect is measured by emulsifying activity and stability indices [81]. For instance, the emulsifying activity and stability indices of type II collagen from soft-shelled turtle calipash were lower than those of collagen from chicken sternal cartilage at pH 7, while the results at pH 4 or 10 are opposite [81, 87]. Kulkarni, Maniyar [88] observed that collagen from fish (Cirrhinus mrigala) scales exhibited good emulsifying potential with an emulsion activity index and emulsion stability index of 21.49 mg−1 and 15.67 min, respectively. Shaik, Asrul Effendi [89] revealed that collagen extracted from the sharpnose stingray's skin using acetic acid (15.01 min) exhibited significantly higher emulsifying stability compared to that extracted with hydrochloric acid (12.91 min). This difference is likely attributed to changes in the collagen's triple helix structure during emulsion formation, thereby modifying protein surface hydrophobicity and partially influencing its emulsifying properties. As an insoluble protein, collagen is generally considered an ineffective emulsifier, but it can embed in the oil–water interfaces, forming Pickering emulsions stabilized by solid particles [90]. For that purpose, processing treatments can convert insoluble collagen into soluble forms like gelatin, which acts as a surface-active agent, reducing interfacial tension or improving droplet zeta-potential.

Dey, Kadharbasha [91] isolated collagen type I (110–120 kDa) from seven fish by-products, yielding 9.15% to 92.38%. It was found that collagen hydrolysate from Tilapia bones and Pacu skin showed over 80% solubility across a wide pH range, 92–96% moisture retention, zeta potential > 50 mV, and emulsification activity 53–70 mg−1 with stability lasting 62–85 min. This collagen improved drug emulsion stability by 14 times, revealing a polyproline-II conformation forming a quasifibrillar network. Surface activity relied on small size, varied hydrophilic/lipophilic ratios, hydroxyproline abundance, and peptide assembly at the emulsion interface, forming a mimic-helix-based quasifibrillar network for optimal orientation and interaction with multiple phases. Razali, Zainol [92] utilized fish scale collagen to stabilize a water-in-virgin coconut oil emulsion, demonstrating excellent physical stability without separation layers, making it suitable for skin application. They found that the addition of collagen slightly reduced emulsion droplet size and resulted in shear-thinning behavior, aligning well with the requirements for topical application.

4.2 Foams

Foam, characterized as a two-phase system containing air bubbles dispersed in a continuous liquid/solid phase, is a fundamental component of various food products, including smoothies, desserts, whipped cream, ice cream, carbonated beverages, meringues, marshmallows, and mousses [1]. Foam structures are typically stabilized by alcohols, fats, proteins, and surfactants. Their half-life stability, lasting 50 min or more, makes them especially valuable in the food industry for preserving the desired appearance, color, and texture of aerated products. Marine collagen, with a hydrophilic-hydrophobic amphipathic structure, can quickly migrate and adsorb to the air–water interface through diffusion and resetting, improving interfacial rheological viscoelasticity and forming cohesive three-dimensional gels for gas packing [81]. These proteins effectively reduce interfacial tension, creating protective films around air bubbles [20].

Furthermore, foaming capacity and stability of collagen proteins depend on the physical properties of the formed film [81]. For instance, conformational aggregation and flexibility of the protein can create higher interfacial activity and a thicker interfacial layer, increasing viscosity and facilitating a multilayer cohesive protein film at the interface [93]. Moreover, macromolecular collagen proteins consist of covalently bonded amino acid residues, resulting in a rare thermodynamic rupture of the bubble membrane compared to that caused by small molecules. For instance, Chen, Li [94] reported that the foam stability values of acid- and pepsin-soluble collagens, extracted from red stingray skin, ranged from 12.50 to 72.0%, and 5.32 to 61.85%, respectively, which were higher than those of basmati rice protein concentrate (0.65–2.50%) and casein (0.17–0.54%). In addition, foaming capacity of acid- and pepsin-soluble collagens ranged from 47.62 to 146.67%, and 71.43 to 151.67%, respectively, which exceeded those of rice bran protein concentrate (5.2–10.03%), casein (3.95–10.15%), and scallop gonad protein isolates (25–90%) [91]. However, collagen proteins, while widely used, may lack the necessary foam capacity or stability to meet the stringent demands of the evolving food industry. For instance, adjusting the pH to the isoelectric point reduces foaming capacity for acid- and pepsin-soluble collagens extracted from rainbow trout (Oncorhynchus mykiss), rendering "soluble" collagen insoluble and hindering the necessary protein-water interaction for foaming [20]. Zou, Wang [81] reported that soft-shelled turtle (Pelodiscus sinensis) collagen had the highest foam capacity and foam stability values at pH 10, followed by pH 4 and 7. Changes in protein flexibility and exposure of hydrophobic groups in alkaline conditions could induce foam formation and stability changes.

4.3 Films

Marine collagen is employed to fabricate edible and biodegradable films, providing both economic advantages and environmental protection [93]. These collagen-based films are typically fabricated through casting or extrusion of colloids aqueous dispersions containing collagen fibers [95]. Pure collagen-based films generally exhibit a uniform and continuous structure characterized by smooth surfaces without holes or cracks, which can be attributed to the good compatibility between biopolymers. The distinctive triple-helical structure of collagen is noted for enhancing film density, resulting in superior mechanical properties compared to other biopolymer films. For instance, 5 mg/mL collagen showed the tensile strength of 88.46 MPa, while 70 mg/mL soy protein isolate, 50 mg/mL gelatin, 25 mg/mL chitosan, and 10 mg/mL κ-carrageenan exhibited 13.6 MPa, 22.42 MPa, 1.76 MPa, and 22.59 MPa, respectively [96,97,98,99]. Moreover, various pretreatments may be applied to collagen, influencing its structure and the characteristics of the resulting film. Xu, Wei [95] reported that films cast from collagen fiber dispersion pretreated at temperatures above 39 °C exhibit reduced tensile strength, but increased water resistance properties. Acid swelling within the pH range of 1.5–4.0, with pH 3 being optimal, enhances the mechanical and thermodynamic properties of the films by maximizing the swelling ratio of collagen fibers. Ma, Teng [100] employed high-pressure homogenization to fabricate micro/nano collagen fibers, suggesting that smaller fiber sizes lead to improved mechanical strength and water resistance in collagen films, with minimal impact on thermal stability. Ahmad, Nirmal [101] prepared biodegradable films from acid-solubilized collagen extracted from starry triggerfish, which showed the highest contact angle, elastic modulus, and tensile strength, but the lowest water vapor permeability and elongation at break compared with the pepsin-solubilized collagen film. The former film exhibited a heat-stable mass residue of 30.9% (w/w) in the temperature range from 50 to 600 °C, while the latter one showed 14.3% (w/w), indicating robust protein–protein interactions within the film network.

The film-forming properties of collagen can also be enhanced by combination with proteins/polysaccharides since protein-polysaccharide complexation effectively improves the mechanical and functional properties (Table 2). Liang, Feng [102] designed composite films from fish scale collagen and polyvinyl alcohol containing varying amounts of potassium sorbate, and found the improved elongation at break, tensile strength, and antimicrobial properties. According to Nuerjiang, Bai [9], compared to a pure fish collagen films, the Young's modulus, tensile strength, and elongation at break of collagen-gelatin composite films enhanced by 79.2%, 37.1%, and 34.4%, respectively, while the water vapor permeability reduced by 51.5%. Ghosh, Grosvenor [17] found that spongia collagens exhibited hydrophilic properties, and films made solely from these collagens were mechanically fragile, disintegrating instantly when wet. However, blending chitosan with Spongia collagens at a ratio of 30:70 (dry mass basis) not only enhanced mechanical properties but also improved structural integrity. This improvement was further intensified by the addition of organic cross-linkers, such as glyceraldehyde and genipin, as evidenced by stress–strain spectra, microscopic images, and mass swelling in water. In a study by Sionkowska, Kozłowska [103], UV-irradiation of fish scale (Esox lucius) collagen film altered the surface free energy and contact angle due to the decrease in amino acids present. Also, the addition of Ag+, Ca2+, and Fe3+ proved to be an effective in improving the functional and mechanical properties of collagen-based films [100, 104].

Table 2 Mechanical and functional properties of marine collagen-based films

In recent years, the incorporation of Pickering emulsion into collagen-based films has emerged as an innovative hydrophobic modification technique, suggesting great potential for introducing novel functionalities and enhancing packaging material characteristics [112, 113]. Pickering emulsion excels in transporting hydrophobic bioactive ingredients or essential oils, improving their compatibility with hydrophilic matrices and consequently enhancing film physical properties [112]. The schematic diagram in Fig. 3 illustrates the fabrication process of Pickering emulsion-loaded collagen-based film. In a study by Ran, Zheng [113], cinnamon essential oil-loaded Pickering emulsions were fabricated using solid particles soy protein isolate/chitosan. These particles irreversibly adsorbed onto the oil through hydrophobic interactions, electrostatic interactions, van der Waals forces, and hydrogen bonding. The resulting Pickering emulsions were incorporated into the collagen film-forming solution, which not only increased the water barrier properties, but also improved UV-blocking properties of the collagen film. The incorporation of Pickering emulsion into collagen films increased water contact angle, water vapor permeability, and elongation at break from 81.87° to 105.37°, 1.34 to 4.60 g × m−1 × s−1 × Pa−1, and 9.22 to 36.85%, respectively, while decreased the tensile strength from 93.43 to 41.77 MPa. Furthermore, this integration enhanced the film's antimicrobial and antioxidant activities while improving thermal stability.

Fig. 3
figure 3

Schematic diagram of the formation process of Pickering emulsion-loaded collagen-based film

4.4 Gels

The gel formation of marine collagen referred to a continuous process, which includes the self-assembly of collagen through self-aggregation and intermolecular forces, micro-fibrils produced via fibrillogenesis by collagen (4–8 molecules), and collagen fibrils formed from these micro-fibrils [1]. Generally, during the nucleation phase, the size and number of microfibrils formed play a crucial role in shaping the final structure of self-assembled collagen fibrils. These factors considerably effect the features of the collagen gel networks, such as uniformity of pore distribution, size, and fiber density [114]. For instance, Jiang, Wang [115] established that ultrasonic treatment reduced the diameter and uniformity of grass carp skin collagen fibrils by enhancing microfiber number and homogeneity during nucleation. This results in collagen gels with more heterogeneous pore structures and larger pore sizes. In addition, Xu, Wei [95] employed ultraviolet radiation at low temperatures to induce collagen gel formation. They observed molecular degradation and cross-linking, which produced collagen fibrils with increased "branches," promoting fibril intertwining. Notably, the amino acid content and characteristics of marine collagen differ from mammalian collagen, giving rise to unique features in the self-assembly of marine collagen [116]. For instance, Bao, Sun [117] observed that higher amounts of hydroxyproline and cysteine induced more cross-linking, resulting in a more homogeneous structure, for example, tilapia collagen gel showed higher mechanical properties and elasticity than porcine collagen gel [118].

Marine collagen gels can be induced by altering the temperature, pH, or ionic conditions. When collagen solutions are vigorously heated, the stabilizing bonds of collagen's triple helical structures break, resulting in a disordered conformation. Typically, collagen proteins denature in the temperature range of 53–63 °C, likely involving the initial breakage of hydrogen bonds, followed by the loss of collagen fibril and molecule contraction [116]. Shi, Tian [119] demonstrated that varying pH (5.0–8.0) affects the gel formation of golden pompano collagen, leading to an increased collagenous fibril number and diameter with rising pH. Moreover, collagen gel networks exhibit reduced fibril diameters and pore spaces as temperature increases from 4 to 37 °C. Tian, Ren [118] reported that tilapia skin collagen gels develop a denser fibril network with higher NaCl concentrations in simulated body fluid. This is attributed to chloride ions neutralizing collagen molecule surface charges, reducing inter-molecular repulsion, and promoting the formation of compactly packed collagen fibril aggregate gels with D-periodicity structures.

Generally, collagen gels produced from native marine sources are not strong enough to meet the requirements of different food applications [120, 121]. One possible way to manipulate the properties of a low-gelling collagen to achieve greater viscoelasticity and mechanical strength than mammalian collagen is to induce cross-linking through physical, chemical, or enzymatic actions [120, 122, 123]. However, the use of chemical cross-linking agents (e.g., carbodiimide, glutaraldehyde, and formaldehyde) has been reported as a means of reinforcing collagen structures, but they are potentially toxic [120]. On the other hand, the formation of collagen-based composite gels, such as combinations of collagen/protein or collagen/polyphenol, not only enhances the oxidative stability of gels but also maintains their structural stability and provides mechanical properties. The formation of networks in collagen-myosin composite gels involves various interactions, including ionic bonds, disulfide bonds, hydrogen bonds, hydrophobic interactions, and electrostatic interactions (Fig. 4) [123,124,125]. Zhao, Lu [125] reported that collagen type I, with increased water solubility and a higher concentration of charged amino acids compared to type II, efficiently integrates with surimi myofibrillar proteins. This integration promotes heightened exposure of protein functional domains, induces significant conformational changes in myosin, and fosters stronger chemical forces among proteins. Consequently, these improvements accelerate the gelation rate, resulting in enhanced water holding capacity and superior textural profiles in the collagen-myosin composite gels. Recently, An, Duan [124] introduced a cross-linking technique to boost tilapia collagen fibril density using chlorogenic acid and procyanidin, which resulted in a roughly threefold increase in enzymatic resistance, mechanical properties, and water absorption and retention capacity of the collagen gel, while thermal stability decreased. In addition, polyphenol cross-linking granted better antioxidant activity to the gel, especially procyanidin, resulting in higher DPPH radical scavenging and antibacterial activity, while chlorogenic acid showed a higher Fe (II) chelation ratio.

Fig. 4
figure 4

Proposed collagen-myosin interaction model to illustrate different chemical forces in collagen-based gel

5 Applications of marine collagen in food systems

In this section, we explored the potential applications of marine collagen in the food industry, considering its enhanced texture, mechanical properties, as well as antibacterial and antioxidant activities. The schematic diagram illustrating the simulated application of marine collagen is presented in Fig. 5.

Fig. 5
figure 5

An overview of food applications of marine collagens

5.1 Meat products

Many processed meat products, such as sausages, patties, and burgers, include ingredients to enhance their functional properties. Collagen proteins, especially those derived from marine sources, are frequently added to comminuted meat recipes to boost stability through improved gel structure and texture. While collagen has limited nutritive value concerning essential amino acids, several authors suggested that substituting 10–20% of the total meat protein content with collagen could enhance emulsifying stability, resilience, hardness, and water-binding capacity in processed meat products [126]. According to Zhao, Lu [125], adding 9% cod skin collagen to surimi gel enhances the formation of ionic bonds, disulfide, and hydrogen bonds. This results in the highest gel strength of 554 g × cm and the highest water-holding capacity of 74.66%, which could be attributed to more charged amino acids intertwining with surimi myofibrillar proteins. Ibrahim, Ismail‐Fitry [126] demonstrated that partially substituting buffalo fat with fish collagen hydrolysates (ranging from 2.5% to 10%) in buffalo patties affected cooking loss, water-holding capacity, and pH value of the cooked patties. Sousa, Fragoso [127] replaced pork backfat with hydrolyzed collagen in Frankfurt-type sausages, finding that 50% collagen addition resulted in improved aroma, protein profile, oxidative stability, texture properties, and water-holding capacity compared to sausages prepared from whole pork meat. Prestes, Carneiro [128] reported a higher water-holding capacity (92.8–98.8%) in turkey ham processed with added collagen, guar gum, and starch, attributing it to the synergistic effect of biopolymers for retaining water. Ham, Hwang [129] investigated the replacement of fat with combinations of collagen and dietary fibers in the fermented sausage products, and found the increased moisture, protein, and ash contents but decreased fat content.

5.2 Dairy products

Dairy products currently constitute approximately 40% of functional foods, aligning with consumer preferences that extend beyond essential nutrition to focus on well-being and disease prevention. Probiotics and prebiotics are notably key ingredients in these dairy products [130]. However, recently reported bioactive compounds, such as marine collagen, are gaining attention for their use as emulsifiers in various products like ice cream, chocolate, yogurt, and butter. Apart from providing functional properties, marine collagen enhances the microbiological, rheological, and physicochemical aspects of these products (Table 3). These foods can be classified into two emulsion-based categories: oil-in-water-based foods (e.g., chocolate sauce) and water-in-oil-based foods (e.g., butter) [91].

Table 3 Food applications of marine collagen-based packaging materials

Dey, Kadharbasha [91] reported that collagen hydrolysates preserved the emulsion properties of butter and chocolate sauce, leading to a significant extension of their shelf life. Moreover, no cytotoxic effects were observed on leukocytes and Vero cells. Shori, Tin [141] found that the presence of fish collagen in Codonopsis pilosula-yogurt improved O-phthaldialdehyde peptide concentrations and scored positively in sensory evaluations. Additionally, Shori, Baba [142] investigated the effect of fish collagen on cheddar cheese (Allium sativum) in terms of acidity and proteolysis of milk protein during a 60-day ripening period. The results revealed that fish collagen increased the free amino acid content in cheddar cheese from 1.66 to 2.47 mM/L (Leucine equivalent), suggesting a significant increase in the proteolysis of milk protein in the presence of fish collagen. Shori, Yong [143] also observed the highest peptide content in plant extract-based cheeses (Illicium verum-, Psidium guajava-, or Curcuma longa) by adding fish collagen. In another development, Bhagwat and Dandge [130] created a milk-based food product, paneer, by incorporating collagen extracted from fish scales. The resulting paneer contained a higher concentration of moisture and protein, contributing to favorable sensorial and textural attributes.

5.3 Drinks

The integration of collagen into beverages has been found to provide various benefits, driving the growing demand for products such as soy collagen, cocoa collagen, and cappuccino collagen. Collagen inclusion in these drinks supports the body's natural ability to generate fatty tissues and can stimulate bodily mechanisms, contributing to the strengthening of tissues and reducing skin wrinkles. Hydrolyzed collagen, due to its low viscosity and high water solubility, is commonly added to beverages to enhance their functional and nutritional properties without causing technological issues in production. Bilek and Bayram [144] prepared new functional drinks from white grape, apple, and orange juice blends containing hydrolyzed fish collagen. Results indicated that fruit juice drinks formulated with a substitution of up to 2.5% collagen increased the protein content of the product from 0.56 to 2.22–2.48 g/100 mL. In vitro bioavailability results showed that apple (90.71%) and orange (95.37%) drinks exhibited higher bioavailability (5–14%) compared to white grape juice blends (1.43%). The drinks also varied widely in terms of total phenolic content (86.93–117.43 mg GAE/100 mL), ascorbic acid content (81.39–113.5 mg/100 mL), and antioxidative results (104–127 µmol TEAC/100 mL). Sae-leaw, Aluko [145] investigated the impact of salmon scale ossein-derived collagen on the sensory, antioxidant, and physicochemical aspects of chrysanthemum tea over a 6-month storage period. The research revealed that incorporating collagen into the tea led to elevated α-amino group content, pH, turbidity, and browning index. Additionally, it yielded a beverage with antioxidant properties, extending its shelf life to 6 months at 30 °C.

5.4 Food packaging materials

Marine collagen, renowned for its native triple helices and fibril networks, finds extensive application in the packaging of meat, poultry, and seafood products, particularly as artificial casing due to its sustainable, biodegradable, and biocompatible nature (Table 4) [146]. Collagen-based packaging materials serve as excellent barriers to moisture and oxygen due to their tightly packed, ordered covalent-bonded network structure. Additionally, collagen packaging materials exhibit enhanced mechanical properties, including adsorption, adhesion, solubility, transparency, and favorable sensory and organoleptic properties [146].

Table 4 Applications of marine collagen in dairy products

Bhuimbar, Bhagwat [132] utilized thick black ruff skin collagen to develop collagen-chitosan-based antibacterial food packaging films with pomegranate peel extract, proving effective against water vapor permeability and foodborne pathogens. In another study, Song, Liu [131] characterized tilapia fish skin collagen for developing active food packaging materials to preserve tilapia muscle. It was found that collagen/zein electrospun film encapsulating different concentrations of gallic acid inhibited tilapia muscle quality deterioration, prolonging the shelf life of fish muscle for at least two days. Wang, Hu [133] reported that collagen-lysozyme coatings (4% collagen + 0.5% lysozyme, w/v) on fresh-salmon fillets (Salmo salar) inhibited bacterial growth, reduced weight loss, and improved texture. Ben Slimane and Sadok [2] highlighted the effectiveness of a biofilm prepared from cartilaginous fish (Mustelus mustelus) collagen and chitosan, demonstrating superior UV barrier properties and antioxidant activity compared to chitosan alone. This biofilm shows promise as a green bioactive film for preserving nutraceutical products. Nicklas, Schatton [155] formulated an aqueous gastro-resistant coating using freeze-dried sponge collagen at a concentration of 15% (w/w) as the film-forming agent. The findings indicated that tablets coated with this collagen successfully resisted 0.1 M HCl for over 2 h while disintegrating within 10 min in a phosphate buffer solution with a pH of 6.8. These coated tablets exhibited favorable mechanical properties and maintained their enteric characteristics for a minimum of 6 months during storage.

When employed in sausage casings, marine collagen provides several advantages over natural casings, including uniformity, sanitation, and flexibility, making it recognized as one of the most promising casings. Its unique ability to shrink and stretch allows it to mimic the expansion and contraction of flesh batter during continuous processing [8]. This characteristic not only improves shelf life but also reduces cooking shrinkage, enhances juiciness, and minimizes elastic stretch after heat treatment. Additionally, marine collagen finds extensive application in producing edible casings for preserving meat products, such as salami and snack sticks. Regardless its numerous benefits, collagen casings face certain challenges, including bending, folding, and stretching during shirring, resulting in increased friction. Their low mechanical strength may lead to ruptures during stuffing, and the casings may easily rupture during sausage filling, often due to excessive local resistance [8]. Nevertheless, researchers have explored solutions to these challenges. The addition of aldehydes [156], metal ions [104], keratin [157], and carbodiimide [158] to collagen casings has been shown to increase their mechanical properties, addressing issues related to deformation and rupture during processing. For instance, Chen, Liu [156] reported that glutaraldehyde improved the tensile strength and Young’s modulus of collagen casings, resulting in enhanced deformation resistance and mechanical strength.

5.5 Food supplements

Marine collagen serves as a food supplement, contributing to lean muscle growth, joint repair, accelerated recovery, and improved cardiovascular performance. The demand for collagen supplements, driven by their myriad benefits, has soared in sports nutrition [159]. Notably, collagen has demonstrated effectiveness in treating rheumatoid arthritis, with studies revealing significant reductions in pain, morning stiffness, and joint swelling [160]. Collagen is recommended as a dietary supplement for the treatment of hand, knee, and hip osteoarthritis in the United States [161]. As a food supplement, the daily collagen recommended dosage ranges from 800 to 1200 mg [161]. In sports nutrition products for athletes, sardine scales collagen, available in various forms such as powdered and liquid supplements, multicomponent bars, and chewing marmalade, is valued for its rich amino acid composition [162]. Shark cartilage supplements have also gained attention, claimed to be natural anti-inflammatory and anti-carcinogenic agents that relieve arthritis and reduce joint pain, thanks to their chondroitin and glycosaminoglycan contents [163]. Furthermore, marine collagen peptide derived from fish waste has been utilized as a fortifier in biscuits, contributing to elevated protein and antioxidant content [161].

6 Conclusions and future perspectives

Marine collagen is rapidly gaining global acceptance for its interesting functional and physicochemical properties, with vast potential in various food applications, including packaging materials, beverages, dairy products, food supplements, and meat-based items. This review has discussed the current state of knowledge on diverse marine collagen sources, including sea cucumbers, mollusks, marine sponges, fish, jellyfish, and crustaceans, emphasizing low-cost collagen extraction based on various studies. Fish and jellyfish notably yield predominantly type I collagen, while crustacean muscle collagen constitutes a minor fraction (0.5–1%) of total protein content. The prevalent method involves solvent extraction with acetic acid, yet acidic treatments are impractical for marine sponge collagens due to their insolubility. Some studies have explored enzymatic extraction, showing significant potential for high collagen yield. While emerging technologies like extrusion-hydro-extraction, supercritical fluid extraction, and deep eutectic solvent extraction hold promise, further research is needed for their practical application in efficiently and cost-effectively extracting collagen proteins from marine sources. Therefore, it is recommended that research efforts should be focused on the development of environmentally friendly and cost-effective methods for extracting collagen from diverse marine sources.

Marine collagen, when utilized in its native form, has demonstrated the capability to enhance both biodegradability and the antioxidative/antimicrobial properties of films, especially those incorporating Pickering emulsions. However, challenges such as brittleness, low mechanical strength, high water solubility, and increased water vapor permeability are associated with marine collagen films. A major drawback, particularly with collagen sourced from cold-water organisms, is its lower denaturation temperature, affecting both processing conditions and film properties. While techniques like crosslinking, blending, or chemical modifications can improve thermal stability, exploring collagens from warm water marine species with superior thermal stability offers a more sustainable approach. This could facilitate collagen processing under milder conditions, reducing the necessity for extensive modifications to achieve thermal stability in food packaging films. Consequently, ongoing research and development are essential to establish packaging alternatives that can compete with currently employed petroleum-based polymers.

Marine collagen-based novel functional food ingredients contain nutritional benefits, including non-essential and essential amino acids, enhancing the quality of various food products. Studies suggest that incorporating collagenous proteins can improve technological and sensory attributes in comminuted meat products undergoing fat replacement with dietary fiber. Additionally, they serve as texturing agents and natural antioxidants, potentially reducing reliance on chemical preservatives and meeting consumer demands for safer and environmentally friendly food products.

Availability of data and materials

The authors declare that all the data supporting the findings of this study are available within the article.

References

  1. Gómez-Guillén MC, Giménez B, López-Caballero ME, Montero MP. Functional and bioactive properties of collagen and gelatin from alternative sources: a review. Food Hydrocolloids. 2011;25:1813–27.

    Article  Google Scholar 

  2. Ben Slimane E, Sadok S. Collagen from cartilaginous fish by-products for a potential application in bioactive film composite. Mar Drugs. 2018;16:211.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Senadheera TRL, Dave D, Shahidi F. Sea cucumber derived type i collagen: a comprehensive review. Mar Drugs. 2020;18:471.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Ahmad MI, Li Y, Pan J, Liu F, Dai H, Fu Y, et al. Collagen and gelatin: structure, properties, and applications in food industry. Int J Biol Macromol. 2024;254: 128037.

    Article  CAS  PubMed  Google Scholar 

  5. Duconseille A, Astruc T, Quintana N, Meersman F, Sante-Lhoutellier V. Gelatin structure and composition linked to hard capsule dissolution: a review. Food Hydrocolloids. 2015;43:360–76.

    Article  CAS  Google Scholar 

  6. Cruz-López H, Rodríguez-Morales S, Enríquez-Paredes LM, Villarreal-Gómez LJ, Olivera-Castillo L, Cortes-Santiago Y, et al. Comparison of collagen characteristic from the skin and swim bladder of Gulf corvina (Cynoscion othonopterus). Tissue Cell. 2021;72: 101593.

    Article  PubMed  Google Scholar 

  7. Caruso GJ. Fishery wastes and by-products: a resource to be valorised. J Fish Sci. 2016;10:12–5.

    Google Scholar 

  8. Suurs P, Barbut S. Collagen use for co-extruded sausage casings: a review. Trends Food Sci Technol. 2020;102:91–101.

    Article  CAS  Google Scholar 

  9. Nuerjiang M, Bai X, Sun L, Wang Q, Xia X, Li F. Size effect of fish gelatin nanoparticles on the mechanical properties of collagen film based on its hierarchical structure. Food Hydrocolloids. 2023;144: 108931.

    Article  CAS  Google Scholar 

  10. Vallejos N, González G, Troncoso E, Zúñiga RN. Acid and enzyme-aided collagen extraction from the byssus of chilean mussels (Mytilus chilensis): effect of process parameters on extraction performance. Food Biophys. 2014;9:322–31.

    Article  Google Scholar 

  11. Khong NMH, Yusoff FM, Jamilah B, Basri M, Maznah I, Chan KW, et al. Improved collagen extraction from jellyfish (Acromitus hardenbergi) with increased physical-induced solubilization processes. Food Chem. 2018;251:41–50.

    Article  CAS  PubMed  Google Scholar 

  12. Farooq S, Ahmad MI, Ali U, Li Y, Shixiu C, Zhang H. A review of advanced techniques for detecting the authenticity and adulteration of camellia oil.n/a.

  13. Huang C-Y, Kuo J-M, Wu S-J, Tsai H-T. Isolation and characterization of fish scale collagen from tilapia (Oreochromis sp.) by a novel extrusion–hydro-extraction process. Food Chem. 2016;190:997–1006.

    Article  CAS  PubMed  Google Scholar 

  14. Silva JC, Barros AA, Aroso IM, Fassini D, Silva TH, Reis RL, et al. Extraction of collagen/gelatin from the marine demosponge chondrosia reniformis (nardo, 1847) using water acidified with carbon dioxide: process optimization. Ind Eng Chem Res. 2016;55:6922–30.

    Article  CAS  Google Scholar 

  15. Batista MP, Fernández N, Gaspar FB, Bronze MDR, Duarte ARC. Extraction of biocompatible collagen from blue shark skins through the conventional extraction process intensification using natural deep eutectic solvents. Front Chem. 2022;10: 937036.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Skierka E, Sadowska M. The influence of different acids and pepsin on the extractability of collagen from the skin of Baltic cod (Gadus morhua). Food Chem. 2007;105:1302–6.

    Article  CAS  Google Scholar 

  17. Ghosh A, Grosvenor AJ, Dyer JM. Marine Spongia collagens: protein characterization and evaluation of hydrogel films. J Appl Polym Sci. 2019;136:47996.

    Article  Google Scholar 

  18. Nurilmala M, Pertiwi RM, Nurhayati T, Fauzi S, Batubara I, Ochiai Y. Characterization of collagen and its hydrolysate from yellowfin tuna Thunnus albacares skin and their potencies as antioxidant and antiglycation agents. Fish Sci. 2019;85:591–9.

    Article  CAS  Google Scholar 

  19. Ahmed R, Haq M, Chun B-S. Characterization of marine derived collagen extracted from the by-products of bigeye tuna (Thunnus obesus). Int J Biol Macromol. 2019;135:668–76.

    Article  CAS  PubMed  Google Scholar 

  20. Heidari MG, Rezaei M. Extracted pepsin of trout waste and ultrasound-promoted method for green recovery of fish collagen. Sustain Chem Pharm. 2022;30: 100854.

    Article  CAS  Google Scholar 

  21. Chen J, Wang G, Li Y. Preparation and characterization of thermally stable collagens from the scales of lizardfish (Synodus macrops). Mar Drugs. 2021;19:597.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Truong TM, Nguyen VM, Tran TT, Le TM. Characterization of acid-soluble collagen from food processing by-products of snakehead fish (Channa striata). Processes. 2021;9:1188.

    Article  CAS  Google Scholar 

  23. Luo Q-B, Chi C-F, Yang F, Zhao Y-Q, Wang B. Physicochemical properties of acid- and pepsin-soluble collagens from the cartilage of Siberian sturgeon. Environ Sci Pollut Res. 2018;25:31427–38.

    Article  CAS  Google Scholar 

  24. Liu D, Liang L, Regenstein JM, Zhou P. Extraction and characterisation of pepsin-solubilised collagen from fins, scales, skins, bones and swim bladders of bighead carp (Hypophthalmichthys nobilis). Food Chem. 2012;133:1441–8.

    Article  CAS  Google Scholar 

  25. Calejo MT, Almeida AJ, Fernandes AI. Exploring a new jellyfish collagen in the production of microparticles for protein delivery. J Microencapsul. 2012;29:520–31.

    Article  CAS  PubMed  Google Scholar 

  26. Barzideh Z, Latiff AA, Gan C-Y, Benjakul S, Karim AA. Isolation and characterisation of collagen from the ribbon jellyfish (Chrysaora sp.). Int J Food Sci Technol. 2014;49:1490–9.

    Article  CAS  Google Scholar 

  27. Zhang J, Duan R, Huang L, Song Y, Regenstein JM. Characterisation of acid-soluble and pepsin-solubilised collagen from jellyfish (Cyanea nozakii Kishinouye). Food Chem. 2014;150:22–6.

    Article  CAS  PubMed  Google Scholar 

  28. Cheng X, Shao Z, Li C, Yu L, Raja MA, Liu C. Isolation, characterization and evaluation of collagen from jellyfish rhopilema esculentum kishinouye for use in hemostatic applications. PLoS ONE. 2017;12: e0169731.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Miki A, Inaba S, Baba T, Kihira K, Fukada H, Oda M. Structural and physical properties of collagen extracted from moon jellyfish under neutral pH conditions. Biosci Biotechnol Biochem. 2015;79:1603–7.

    Article  CAS  PubMed  Google Scholar 

  30. Rastian Z, Pütz S, Wang Y, Kumar S, Fleissner F, Weidner T, et al. Type I collagen from jellyfish catostylus mosaicus for biomaterial applications. ACS Biomater Sci Eng. 2018;4:2115–25.

    Article  CAS  PubMed  Google Scholar 

  31. Liu Z, Oliveira ACM, Su Y-C. Purification and characterization of pepsin-solubilized collagen from skin and connective tissue of giant red sea cucumber (Parastichopus californicus). J Agric Food Chem. 2010;58:1270–4.

    Article  CAS  PubMed  Google Scholar 

  32. Lin S, Xue Y-P, San E, Keong TC, Chen L, Zheng Y-G. Extraction and characterization of pepsin soluble collagen from the body wall of sea cucumber Acaudina leucoprocta. J Aquat Food Prod Technol. 2017;26:502–15.

    Article  CAS  Google Scholar 

  33. Abedin MZ, Karim AA, Ahmed F, Latiff AA, Gan C-Y, Che Ghazali F, et al. Isolation and characterization of pepsin-solubilized collagen from the integument of sea cucumber (Stichopus vastus). J Sci Food Agric. 2013;93:1083–8.

    Article  CAS  PubMed  Google Scholar 

  34. Zhong M, Chen T, Hu C, Ren C. Isolation and characterization of collagen from the body wall of sea cucumber Stichopus monotuberculatus. J Food Sci. 2015;80:C671–9.

    Article  CAS  PubMed  Google Scholar 

  35. Adibzadeh N, Aminzadeh S, Jamili S, Karkhane AA, Farrokhi N. Purification and characterization of pepsin-solubilized collagen from skin of sea cucumber Holothuria parva. Appl Biochem Biotechnol. 2014;173:143–54.

    Article  CAS  PubMed  Google Scholar 

  36. Zhu B, Dong X, Zhou D, Gao Y, Yang J, Li D, et al. Physicochemical properties and radical scavenging capacities of pepsin-solubilized collagen from sea cucumber Stichopus japonicus. Food Hydrocolloids. 2012;28:182–8.

    Article  ADS  CAS  Google Scholar 

  37. Shaik MI, Kadir ANA, Sarbon NM. Physicochemical and thermal properties of pepsin- and acid-soluble collagen isolated from the body wall of sea cucumbers (Stichopus hermanni). J Food Sci. 2023;89:320.

    Article  PubMed  Google Scholar 

  38. Li P-H, Lu W-C, Chan Y-J, Ko W-C, Jung C-C, Le Huynh DT, et al. Extraction and characterization of collagen from sea cucumber (Holothuria cinerascens) and its potential application in moisturizing cosmetics. Aquaculture. 2020;515: 734590.

    Article  CAS  Google Scholar 

  39. Saallah S, Roslan J, Julius FS, Saallah S, Mohamad Razali UH, Pindi W, et al. Comparative study of the yield and physicochemical properties of collagen from sea cucumber (holothuria scabra), obtained through dialysis and the ultrafiltration membrane. Molecules. 2021;26:2564.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Jose H, Murugesan P, Kumar K, Muthuvel A. Isolation and characterization of acid and pepsin: solubilised collagen from the muscle of mantis shrimp (Oratosquilla Nepa). Int J Pharm Pharm Sci. 2014;6:2014.

    Google Scholar 

  41. Hiransuchalert R, Oonwiset N, Imarom Y, Chindudsadeegul P, Laongmanee P, Arnupapboon SJF, et al. Extraction and characterization of pepsin-soluble collagen from different mantis shrimp species. Fish Aquatic Sci. 2021;24:406.

    Article  CAS  Google Scholar 

  42. Wu J, Guo X, Liu H, Chen L. Isolation and comparative study on the characterization of guanidine hydrochloride soluble collagen and pepsin soluble collagen from the body of surf clam shell (Coelomactra antiquata). Foods. 2019;8:11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Coelho RCG, Marques ALP, Oliveira SM, Diogo GS, Pirraco RP, Moreira-Silva J, et al. Extraction and characterization of collagen from Antarctic and Sub-Antarctic squid and its potential application in hybrid scaffolds for tissue engineering. Mater Sci Eng, C. 2017;78:787–95.

    Article  CAS  Google Scholar 

  44. Araújo TAT, de Souza A, Santana AF, Braga ARC, Custódio MR, Simões FR, et al. Comparison of different methods for spongin-like collagen extraction from marine sponges (Chondrilla caribensis and Aplysina fulva): physicochemical properties and in vitro biological analysis. Membranes. 2021;11:522.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Barros AA, Aroso IM, Silva TH, Mano JF, Duarte ARC, Reis RL. Water and carbon dioxide: green solvents for the extraction of collagen/gelatin from marine sponges. ACS Sustain Chem Eng. 2015;3:254–60.

    Article  CAS  Google Scholar 

  46. Tziveleka LA, Ioannou E, Tsiourvas D, Berillis P, Foufa E, Roussis V. Collagen from the marine sponges axinella cannabina and suberites carnosus: isolation and morphological, biochemical, and biophysical characterization. Mar Drugs. 2017;15:152.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Tian M, Xue C, Chang Y, Shen J, Zhang Y, Li Z, et al. Collagen fibrils of sea cucumber (Apostichopus japonicus) are heterotypic. Food Chem. 2020;316: 126272.

    Article  CAS  PubMed  Google Scholar 

  48. Bordbar S, Anwar F, Saari N. High-value components and bioactives from sea cucumbers for functional foods: a review. Mar Drugs. 2011;92011:1761–805.

    Article  Google Scholar 

  49. Salarzadeh A, Afkhami M, Bastami KD, Ehsanpour M, Khazaali A, Mokhleci AJ. Proximate composition of two sea cucumber species holothuria pavra and holothuria arenicola in Persian Gulf. Ann Biol Res. 2012;3:1305–11.

    CAS  Google Scholar 

  50. Yan L-J, Zhan C-L, Cai Q-F, Weng L, Du C-H, Liu G-M, et al. Purification, characterization, cDNA cloning and in vitro expression of a serine proteinase from the intestinal tract of sea cucumber (Stichopus japonicus) with collagen degradation activity. J Agric Food Chem. 2014;62:4769–77.

    Article  CAS  PubMed  Google Scholar 

  51. Wang J, Chang Y, Wu F, Xu X, Xue C. Fucosylated chondroitin sulfate is covalently associated with collagen fibrils in sea cucumber Apostichopus japonicus body wall. Carbohyd Polym. 2018;186:439–44.

    Article  CAS  Google Scholar 

  52. Peng S, Wei H, Zhan S, Yang W, Lou Q, Deng S, et al. Spoilage mechanism and preservation technologies on the quality of shrimp: an overview. Trends Food Sci Technol. 2022;129:233–43.

    Article  CAS  Google Scholar 

  53. Suresh PV, Kudre TG, Johny LC. Sustainable valorization of seafood processing by-product/discard. In: Singhania RR, Agarwal RA, Kumar RP, Sukumaran RK, editors. Waste to Wealth. Singapore: Springer; 2018. p. 111–39.

    Chapter  Google Scholar 

  54. Yan N, Chen X. Sustainability: don’t waste seafood waste. Nature. 2015;524:155–7.

    Article  ADS  CAS  PubMed  Google Scholar 

  55. Xiao X-C, Lin D, Cao K-Y, Sun L-C, Chen Y-L, Weng L, et al. Properties of Pacific white shrimp (Litopenaeus vannamei) collagen and its degradation by endogenous proteinases during cold storage. Food Chem. 2023;419: 136071.

    Article  CAS  PubMed  Google Scholar 

  56. Li X, Han T, Zheng S, Wu G. Nutrition and functions of amino acids in aquatic crustaceans. In: Wu G, editor. amino acids in nutrition and health: Amino acids in the nutrition of companion, zoo and farm animals. Cham: Springer; 2021. p. 169–98.

    Chapter  Google Scholar 

  57. Sriket P, Benjakul S, Visessanguan W, Kijroongrojana K. Comparative studies on chemical composition and thermal properties of black tiger shrimp (Penaeus monodon) and white shrimp (Penaeus vannamei) meats. Food Chem. 2007;103:1199–207.

    Article  CAS  Google Scholar 

  58. Pallela R, Bojja S, Janapala VR. Biochemical and biophysical characterization of collagens of marine sponge, Ircinia fusca (Porifera: Demospongiae: Irciniidae). Int J Biol Macromol. 2011;49:85–92.

    Article  CAS  PubMed  Google Scholar 

  59. Brotz L, Schiariti A, López-Martínez J, Álvarez-Tello J, Peggy Hsieh YH, Jones RP, et al. Jellyfish fisheries in the Americas: origin, state of the art, and perspectives on new fishing grounds. Rev Fish Biol Fish. 2017;27:1–29.

    Article  Google Scholar 

  60. De Rinaldis G, Leone A, De Domenico S, Bosch-Belmar M, Slizyte R, Milisenda G, et al. Biochemical characterization of cassiopea andromeda (Forsskål, 1775), another red sea jellyfish in the western Mediterranean sea. Mar Drugs. 2021;19:498.

    Article  PubMed  PubMed Central  Google Scholar 

  61. León-Campos MI, Claudio-Rizo JA, Rodriguez-Fuentes N, Cabrera-Munguía DA, Becerra-Rodriguez JJ, Herrera-Guerrero A, et al. Biocompatible interpenetrating polymeric networks in hydrogel state comprised from jellyfish collagen and polyurethane. J Polym Res. 2021;28:291.

    Article  Google Scholar 

  62. Khong NMH, Yusoff FM, Jamilah B, Basri M, Maznah I, Chan KW, et al. Nutritional composition and total collagen content of three commercially important edible jellyfish. Food Chem. 2016;196:953–60.

    Article  CAS  PubMed  Google Scholar 

  63. Smith IP, Domingos M, Richardson SM, Bella J. Characterization of the biophysical properties and cell adhesion interactions of marine invertebrate collagen from Rhizostoma pulmo. Mar Drugs. 2023;212023:59.

    Article  Google Scholar 

  64. Tabakaeva OV, Tabakaev AV, Piekoszewski W. Nutritional composition and total collagen content of two commercially important edible bivalve molluscs from the Sea of Japan coast. J Food Sci Technol. 2018;55:4877–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Veeruraj A, Arumugam M, Ajithkumar T, Balasubramanian T. Isolation and characterization of collagen from the outer skin of squid (Doryteuthis singhalensis). Food Hydrocolloids. 2015;43:708–16.

    Article  CAS  Google Scholar 

  66. Ozogul F, Cagalj M, Šimat V, Ozogul Y, Tkaczewska J, Hassoun A, et al. Recent developments in valorisation of bioactive ingredients in discard/seafood processing by-products. Trends Food Sci Technol. 2021;116:559–82.

    Article  CAS  Google Scholar 

  67. Yu F, Zong C, Jin S, Zheng J, Chen N, Huang J, et al. Optimization of extraction conditions and characterization of pepsin-solubilised collagen from skin of giant croaker (Nibea japonica). Mar Drugs. 2018;16:29.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Sun L, Li B, Song W, Si L, Hou H. Characterization of Pacific cod (Gadus macrocephalus) skin collagen and fabrication of collagen sponge as a good biocompatible biomedical material. Process Biochem. 2017;63:229–35.

    Article  CAS  Google Scholar 

  69. Liu D, Zhang X, Li T, Yang H, Zhang H, Regenstein JM, et al. Extraction and characterization of acid- and pepsin-soluble collagens from the scales, skins and swim-bladders of grass carp (Ctenopharyngodon idella). Food Biosci. 2015;9:68–74.

    Article  Google Scholar 

  70. Cao H, Shi FX, Xu F, Yu JS. Molecular structure and physicochemical properties of pepsin-solubilized type II collagen from the chick sternal cartilage. Eur Rev Med Pharmacol Sci. 2013;17:1427–37.

    CAS  PubMed  Google Scholar 

  71. Jeevithan E, Bao B, Bu Y, Zhou Y, Zhao Q, Wu W. Type II collagen and gelatin from silvertip shark (Carcharhinus albimarginatus) cartilage: isolation, purification, physicochemical and antioxidant properties. Mar Drugs. 2014;12:3852–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Wang L, An X, Yang F, Xin Z, Zhao L, Hu Q. Isolation and characterisation of collagens from the skin, scale and bone of deep-sea redfish (Sebastes mentella). Food Chem. 2008;108:616–23.

    Article  CAS  PubMed  Google Scholar 

  73. Tan Y, Chang SKC. Isolation and characterization of collagen extracted from channel catfish (Ictalurus punctatus) skin. Food Chem. 2018;242:147–55.

    Article  ADS  CAS  PubMed  Google Scholar 

  74. Arumugam GKS, Sharma D, Balakrishnan RM, Ettiyappan JBP. Extraction, optimization and characterization of collagen from sole fish skin. Sustain Chem Pharm. 2018;9:19–26.

    Article  Google Scholar 

  75. Liang Q, Wang L, Sun W, Wang Z, Xu J, Ma H. Isolation and characterization of collagen from the cartilage of Amur sturgeon (Acipenser schrenckii). Process Biochem. 2014;49:318–23.

    Article  CAS  Google Scholar 

  76. Ahmed R, Getachew AT, Cho Y-J, Chun B-S. Application of bacterial collagenolytic proteases for the extraction of type I collagen from the skin of bigeye tuna (Thunnus obesus). LWT. 2018;89:44–51.

    Article  CAS  Google Scholar 

  77. Hansen BB, Spittle S, Chen B, Poe D, Zhang Y, Klein JM, et al. Deep eutectic solvents: a review of fundamentals and applications. Chem Rev. 2021;121:1232–85.

    Article  CAS  PubMed  Google Scholar 

  78. Bai C, Wei Q, Ren X. Selective extraction of collagen peptides with high purity from cod skins by deep eutectic solvents. ACS Sustain Chem Eng. 2017;5:7220–7.

    Article  CAS  Google Scholar 

  79. Sousa RO, Martins E, Carvalho DN, Alves AL, Oliveira C, Duarte ARC, et al. Collagen from Atlantic cod (Gadus morhua) skins extracted using CO2 acidified water with potential application in healthcare. J Polym Res. 2020;27:73.

    Article  CAS  Google Scholar 

  80. Kuwahara J. Extraction of type i collagen from tilapia scales using acetic acid and ultrafine bubbles. Processes. 2021;9:288.

    Article  CAS  Google Scholar 

  81. Zou Y, Wang L, Cai P, Li P, Zhang M, Sun Z, et al. Effect of ultrasound assisted extraction on the physicochemical and functional properties of collagen from soft-shelled turtle calipash. Int J Biol Macromol. 2017;105:1602–10.

    Article  CAS  PubMed  Google Scholar 

  82. Pezeshk S, Rezaei M, Abdollahi M. Impact of ultrasound on extractability of native collagen from tuna by-product and its ultrastructure and physicochemical attributes. Ultrason Sonochem. 2022;89: 106129.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Ali AMM, Kishimura H, Benjakul S. Extraction efficiency and characteristics of acid and pepsin soluble collagens from the skin of golden carp (Probarbus Jullieni) as affected by ultrasonication. Process Biochem. 2018;66:237–44.

    Article  CAS  Google Scholar 

  84. Kim HK, Kim YH, Kim YJ, Park HJ, Lee NH. Effects of ultrasonic treatment on collagen extraction from skins of the sea bass Lateolabrax japonicus. Fish Sci. 2012;78:485–90.

    Article  Google Scholar 

  85. Petcharat T, Benjakul S, Karnjanapratum S, Nalinanon S. Ultrasound-assisted extraction of collagen from clown featherback (Chitala ornata) skin: yield and molecular characteristics. J Sci Food Agric. 2021;101:648–58.

    Article  CAS  PubMed  Google Scholar 

  86. Farooq S, Zhang C, Xi Y, Zhang H. Physiochemical characteristics and rheological investigations of camellia oil body emulsions stabilized by gum tragacanth as a coating layer. Food Chem. 2022;377:131997.

    Article  CAS  PubMed  Google Scholar 

  87. Akram AN, Zhang C. Effect of ultrasonication on the yield, functional and physicochemical characteristics of collagen-II from chicken sternal cartilage. Food Chem. 2020;307: 125544.

    Article  CAS  PubMed  Google Scholar 

  88. Kulkarni P, Maniyar M, Nalawade M, Bhagwat P, Pillai S. Isolation, biochemical characterization, and development of a biodegradable antimicrobial film from Cirrhinus mrigala scale collagen. Environ Sci Pollut Res. 2022;29:18840–50.

    Article  CAS  Google Scholar 

  89. Shaik MI, Asrul Effendi NF, Sarbon NM. Functional properties of sharpnose stingray (Dasyatis zugei) skin collagen by ultrasonication extraction as influenced by organic and inorganic acids. Biocatal Agric Biotechnol. 2021;35: 102103.

    Article  CAS  Google Scholar 

  90. Zhu Q, Li Y, Li S, Wang W. Fabrication and characterization of acid soluble collagen stabilized Pickering emulsions. Food Hydrocolloids. 2020;106: 105875.

    Article  CAS  Google Scholar 

  91. Dey P, Kadharbasha S, Bajaj M, Das J, Chakraborty T, Bhat C, et al. Contribution of quasifibrillar properties of collagen hydrolysates towards lowering of interface tension in emulsion-based food leading to shelf-life enhancement. Food Bioprocess Technol. 2021;14:1566–86.

    Article  CAS  Google Scholar 

  92. Razali HC, Zainol I, Mahamod WRW, Zain H, Bakar NA, Alwi MA, et al. Effect of fish scale hydrolysed collagen on stability of water-in-virgin coconut oil emulsion for potential topical application. J Eng Appl Sci. 2019;14:7647.

    Article  CAS  Google Scholar 

  93. He L, Lan W, Cen L, Chen S, Liu S, Liu Y, et al. Improving catalase stability by its immobilization on grass carp (Ctenopharyngodon idella) scale collagen self-assembly films. Mater Sci Eng, C. 2019;105: 110024.

    Article  CAS  Google Scholar 

  94. Chen J, Li J, Li Z, Yi R, Shi S, Wu K, et al. Physicochemical and functional properties of type i collagens in red stingray (Dasyatis akajei) skin. Mar Drugs. 2019;17:558.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Xu C, Wei X, Shu F, Li X, Wang W, Li P, et al. Induction of fiber-like aggregation and gelation of collagen by ultraviolet irradiation at low temperature. Int J Biol Macromol. 2020;153:232–9.

    Article  CAS  PubMed  Google Scholar 

  96. Liu J, Song F, Chen R, Deng G, Chao Y, Yang Z, et al. Effect of cellulose nanocrystal-stabilized cinnamon essential oil Pickering emulsions on structure and properties of chitosan composite films. Carbohyd Polym. 2022;275: 118704.

    Article  CAS  Google Scholar 

  97. Ran R, Wang L, Su Y, He S, He B, Li C, et al. Preparation of pH-indicator films based on soy protein isolate/bromothymol blue and methyl red for monitoring fresh-cut apple freshness. J Food Sci. 2021;86:4594–610.

    Article  CAS  PubMed  Google Scholar 

  98. Dammak I, Lourenço RV, Sobral PJ. Active gelatin films incorporated with Pickering emulsions encapsulating hesperidin: Preparation and physicochemical characterization. J Food Eng. 2019;240:9–20.

    Article  CAS  Google Scholar 

  99. Liu Y, Qin Y, Bai R, Zhang X, Yuan L, Liu J. Preparation of pH-sensitive and antioxidant packaging films based on κ-carrageenan and mulberry polyphenolic extract. Int J Biol Macromol. 2019;134:993–1001.

    Article  CAS  PubMed  Google Scholar 

  100. Ma Y, Teng A, Zhao K, Zhang K, Zhao H, Duan S, et al. A top-down approach to improve collagen film’s performance: the comparisons of macro, micro and nano sized fibers. Food Chem. 2020;309: 125624.

    Article  CAS  PubMed  Google Scholar 

  101. Ahmad M, Nirmal NP, Chuprom J. Molecular characteristics of collagen extracted from the starry triggerfish skin and its potential in the development of biodegradable packaging film. RSC Adv. 2016;6:33868–79.

    Article  ADS  CAS  Google Scholar 

  102. Liang X, Feng S, Ahmed S, Qin W, Liu Y. Effect of potassium sorbate and ultrasonic treatment on the properties of fish scale collagen/polyvinyl alcohol composite film. Molecules. 2019;24:2363.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Sionkowska A, Kozłowska J, Lazare S. Modification by UV radiation of the surface of thin films based on collagen extracted from fish scales. Biointerphases. 2014;9: 029003.

    Article  CAS  PubMed  Google Scholar 

  104. Ma Y, Wang W, Wang Y, Guo Y, Duan S, Zhao K, et al. Metal ions increase mechanical strength and barrier properties of collagen-sodium polyacrylate composite films. Int J Biol Macromol. 2018;119:15–22.

    Article  CAS  PubMed  Google Scholar 

  105. Elango J, Robinson JS, Geevaretnam J, Rupia EJ, Arumugam V, Durairaj S, et al. Physicochemical and rheological properties of composite shark catfish (pangasius pangasius) skin collagen films integrated with chitosan and calcium salts. J Food Biochem. 2016;40:304–15.

    Article  CAS  Google Scholar 

  106. Zhuang Y, Ruan S, Yao H, Sun Y. Physical properties of composite films from tilapia skin collagen with pachyrhizus starch and rambutan peel phenolics. Mar Drugs. 2019;17:662.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Elango J, Bu Y, Bin B, Geevaretnam J, Robinson JS, Wu W. Effect of chemical and biological cross-linkers on mechanical and functional properties of shark catfish skin collagen films. Food Biosci. 2017;17:42–51.

    Article  CAS  Google Scholar 

  108. da Costa GF, Grisi CVB, de Albuquerque Meireles BRL, de Sousa S, de Magalhães Cordeiro AMT. Collagen films, cassava starch and their blends: physical–chemical, thermal and microstructure properties. Packag Technol Sci. 2022;35:229–40.

    Article  Google Scholar 

  109. Jiang Y, Lan W, Sameen DE, Ahmed S, Qin W, Zhang Q, et al. Preparation and characterization of grass carp collagen-chitosan-lemon essential oil composite films for application as food packaging. Int J Biol Macromol. 2020;160:340–51.

    Article  CAS  PubMed  Google Scholar 

  110. Adamiak K, Lewandowska K, Sionkowska A. The infuence of salicin on rheological and film-forming properties of collagen. Molecules. 2021;26:1661.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Ahmed M, Verma AK, Patel R. Physiochemical, antioxidant, and food simulant release properties of collagen-carboxymethyl cellulose films enriched with Berberis lyceum root extract for biodegradable active food packaging. J Food Process Preserv. 2022;46: e16485.

    Article  CAS  Google Scholar 

  112. Shi W-J, Tang C-H, Yin S-W, Yin Y, Yang X-Q, Wu L-Y, et al. Development and characterization of novel chitosan emulsion films via pickering emulsions incorporation approach. Food Hydrocolloids. 2016;52:253–64.

    Article  CAS  Google Scholar 

  113. Ran R, Zheng T, Tang P, Xiong Y, Yang C, Gu M, et al. Antioxidant and antimicrobial collagen films incorporating Pickering emulsions of cinnamon essential oil for pork preservation. Food Chem. 2023;420: 136108.

    Article  CAS  PubMed  Google Scholar 

  114. Farooq S, Ahmad MI, Ali U, Zhang H. Fabrication of curcumin-loaded oleogels using camellia oil bodies and gum arabic/chitosan coatings for controlled release applications. Int J Biol Macromol. 2024;254: 127758.

    Article  CAS  PubMed  Google Scholar 

  115. Jiang Y, Wang H, Deng M, Wang Z, Zhang J, Wang H, et al. Effect of ultrasonication on the fibril-formation and gel properties of collagen from grass carp skin. Mater Sci Eng, C. 2016;59:1038–46.

    Article  CAS  Google Scholar 

  116. Yan M, Li B, Zhao X, Qin S. Effect of concentration, pH and ionic strength on the kinetic self-assembly of acid-soluble collagen from walleye pollock (Theragra chalcogramma) skin. Food Hydrocolloids. 2012;29:199–204.

    Article  CAS  Google Scholar 

  117. Bao Z, Sun Y, Rai K, Peng X, Wang S, Nian R, et al. The promising indicators of the thermal and mechanical properties of collagen from bass and tilapia: synergistic effects of hydroxyproline and cysteine. Biomater Sci. 2018;6:3042–52.

    Article  CAS  PubMed  Google Scholar 

  118. Tian H, Ren Z, Shi L, Hao G, Chen J, Weng W. Self-assembly characterization of tilapia skin collagen in simulated body fluid with different salt concentrations. Process Biochem. 2021;108:153–60.

    Article  CAS  Google Scholar 

  119. Shi L, Tian H, Wang Y, Hao G, Chen J, Weng W. Effect of pH on properties of golden pompano skin collagen-based fibril gels by self-assembly in vitro. J Sci Food Agric. 2020;100:4801–7.

    Article  CAS  PubMed  Google Scholar 

  120. Andriakopoulou CE, Zadpoor AA, Grant MH, Riches PE. Development and mechanical characterisation of self-compressed collagen gels. Mater Sci Eng, C. 2018;84:243–7.

    Article  CAS  Google Scholar 

  121. Farooq S, Ahmad MI, Zhang Y, Zhang H. Impact of interfacial layer number and Schiff base cross-linking on the microstructure, rheological properties and digestive lipolysis of plant-derived oil bodies-based oleogels. Food Hydrocolloids. 2023;138: 108473.

    Article  CAS  Google Scholar 

  122. Farooq S, Ijaz Ahmad M, Zhang Y, Chen M, Zhang H. Fabrication, characterization and in vitro digestion of camellia oil body emulsion gels cross-linked by polyphenols. Food Chem. 2022;394: 133469.

    Article  CAS  PubMed  Google Scholar 

  123. Adamiak K, Sionkowska A. Current methods of collagen cross-linking: Review. Int J Biol Macromol. 2020;161:550–60.

    Article  CAS  PubMed  Google Scholar 

  124. An X, Duan S, Jiang Z, Chen S, Sun W, Liu X, et al. Role of chlorogenic acid and procyanidin in the modification of self-assembled fibrillar gel prepared from tilapia collagen. Polym Degrad Stab. 2022;206: 110177.

    Article  CAS  Google Scholar 

  125. Zhao Y, Lu K, Piao X, Song Y, Wang L, Zhou R, et al. Collagens for surimi gel fortification: type-dependent effects and the difference between type I and type II. Food Chem. 2023;407: 135157.

    Article  CAS  PubMed  Google Scholar 

  126. Ibrahim FN, Ismail‐Fitry MR, Yusoff MM, Shukri R. Effects of Fish Collagen Hydrolysate (FCH) as Fat Replacer in the Production of Buffalo Patties. In: Sciences AaF, (editor).2018.

  127. Sousa SC, Fragoso SP, Penna CRA, Arcanjo NMO, Silva FAP, Ferreira VCS, et al. Quality parameters of frankfurter-type sausages with partial replacement of fat by hydrolyzed collagen. LWT Food Sci Technol. 2017;76:320–5.

    Article  CAS  Google Scholar 

  128. Prestes RC, Carneiro EBB, Demiate IM. Hydrolyzed collagen, modified starch and guar gum addition in turkey ham. Agric Food Sci. 2012;42:1307–13.

    Google Scholar 

  129. Ham Y-K, Hwang K-E, Kim H-W, Song D-H, Kim Y-J, Choi Y-S, et al. Effects of fat replacement with a mixture of collagen and dietary fibre on small calibre fermented sausages. Int J Food Sci Technol. 2016;51:96–104.

    Article  CAS  Google Scholar 

  130. Bhagwat PK, Dandge PB. Isolation, characterization and valorizable applications of fish scale collagen in food and agriculture industries. Biocatal Agric Biotechnol. 2016;7:234–40.

    Article  Google Scholar 

  131. Song Z, Liu H, Huang A, Zhou C, Hong P, Deng C. Collagen/zein electrospun films incorporated with gallic acid for tilapia (Oreochromis niloticus) muscle preservation. J Food Eng. 2022;317: 110860.

    Article  CAS  Google Scholar 

  132. Bhuimbar MV, Bhagwat PK, Dandge PB. Extraction and characterization of acid soluble collagen from fish waste: development of collagen-chitosan blend as food packaging film. J Environ Chem Eng. 2019;7: 102983.

    Article  CAS  Google Scholar 

  133. Wang Z, Hu S, Gao Y, Ye C, Wang H. Effect of collagen-lysozyme coating on fresh-salmon fillets preservation. LWT. 2017;75:59–64.

    Article  Google Scholar 

  134. Liu J, Shibata M, Ma Q, Liu F, Lu Q, Shan Q, et al. Characterization of fish collagen from blue shark skin and its application for chitosan- collagen composite coating to preserve red porgy (Pagrus major) meat. J Food Biochem. 2020;44: e13265.

    Article  CAS  PubMed  Google Scholar 

  135. Qu W, Guo T, Zhang X, Jin Y, Wang B, Wahia H, et al. Preparation of tuna skin collagen-chitosan composite film improved by sweep frequency pulsed ultrasound technology. Ultrason Sonochem. 2022;82: 105880.

    Article  CAS  PubMed  Google Scholar 

  136. Juan W, Ya-ting S, Xin-xin Z, Jia-ni XUE, Shu-yan GUO, Hai-le MA. Preparation of collagen-chitosan film and its preservation of Pork. Mod Food Sci Technol. 2020;36:89–98.

    Google Scholar 

  137. Yu L, Chen S, Wang J, Wang Q. Preservation effect of collagen-chitosan blend film modified by tea polyphenols on grouper (Epinephelus coioides) Fillets Stored at 4 ℃. Food Sci. 2017;38:220–6.

    Google Scholar 

  138. Yin S, Zhang Y, Zhang X, Tao K, Li GJC. High-strength collagen/delphinidin film incorporated with Vaccinium oxycoccus pigment for active and intelligent food packaging. Collagen Leather. 2023;5:1–14.

    Article  Google Scholar 

  139. Guzman LE, Acevedo D, Romero L, Estrada J. Preparation of an edible film based on collagen incorporated as the antimicrobial agent nisin. Inf Tecnol. 2015;26:17–24.

    CAS  Google Scholar 

  140. Fadini AL, Rocha FS, Alvim ID, Sadahira MS, Queiroz MB, Alves RMV, et al. Mechanical properties and water vapour permeability of hydrolysed collagen–cocoa butter edible films plasticised with sucrose. Food Hydrocolloids. 2013;30:625–31.

    Article  CAS  Google Scholar 

  141. Shori AB, Tin YP, Baba AS. Codonopsis pilosula and fish collagen yogurt: Proteolytic, potential Angiotensin-I converting enzyme (ACE) inhibitory activity and sensory properties. LWT. 2022;165: 113729.

    Article  CAS  Google Scholar 

  142. Shori AB, Baba AS, Hoen Solear LS. Allium sativum and fish collagen enhanced proteolysis pattern of milk protein during cheddar cheese ripening. J Agric Food Res. 2020;2: 100059.

    Google Scholar 

  143. Shori AB, Yong YS, Baba AS. Effects of medicinal plants extract enriched cheese with fish collagen on proteolysis and in vitro angiotensin-I converting enzyme inhibitory activity. LWT. 2022;159: 113218.

    Article  CAS  Google Scholar 

  144. Bilek SE, Bayram SK. Fruit juice drink production containing hydrolyzed collagen. J Funct Foods. 2015;14:562–9.

    Article  CAS  Google Scholar 

  145. Sae-leaw T, Aluko RE, Chantakun K, Benjakul S. Physicochemical, antioxidant and sensory properties of ready-to-drink chrysanthemum tea fortified with hydrolyzed collagen from salmon scale ossein. J Aquat Food Prod Technol. 2021;30:1159–72.

    Article  CAS  Google Scholar 

  146. Antoniewski MN, Barringer SA. Meat shelf-life and extension using collagen/gelatin coatings: a review. Crit Rev Food Sci Nutr. 2010;50:644–53.

    Article  CAS  PubMed  Google Scholar 

  147. Shori AB, Hong YC, Baba AS. Proteolytic profile, angiotensin-I converting enzyme inhibitory activity and sensory evaluation of Codonopsis pilosula and fish collagen cheese. Food Res Int (Ottawa, Ont). 2021;143: 110238.

    Article  CAS  Google Scholar 

  148. Shori AB, Baba AS, Chuah PF. The effects of fish collagen on the proteolysis of milk proteins, ACE inhibitory activity and sensory evaluation of plain- and Allium sativum-yogurt. J Taiwan Inst Chem Eng. 2013;44:701–6.

    Article  CAS  Google Scholar 

  149. León-López A, Pérez-Marroquín XA, Campos-Lozada G, Campos-Montiel RG, Aguirre-Álvarez G. Characterization of whey-based fermented beverages supplemented with hydrolyzed collagen: antioxidant activity and bioavailability. Foods. 2020;9:1106.

    Article  PubMed  PubMed Central  Google Scholar 

  150. Gerhardt Â, Monteiro B, Gennari A, Lehn D, Souza C. Physicochemical and sensory characteristics of fermented dairy drink using ricotta cheese whey and hydrolyzed collagen. Revista do Instituto de Laticínios Cândido Tostes. 2013;68:41–50.

    Article  CAS  Google Scholar 

  151. Znamirowska A, Szajnar K, Pawlos M. Probiotic fermented milk with collagen. Dairy. 2020;1:126–34.

    Article  Google Scholar 

  152. Szopa K, Znamirowska-Piotrowska A, Szajnar K, Pawlos M. Effect of collagen types, bacterial strains and storage duration on the quality of probiotic fermented sheep Milk. Molecules. 2022;27:3028.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Shori AB, Baba AS, Keow JN. Effect of Allium sativum and fish collagen on the proteolytic and angiotensin-I converting enzyme-inhibitory activities in cheese and yogurt. Pak J Biol Sci. 2012;15:1160–7.

    Article  CAS  PubMed  Google Scholar 

  154. Shori AB, Ming KS, Baba AS. The effects of Lycium barbarum water extract and fish collagen on milk proteolysis and in vitro angiotensin I-converting enzyme inhibitory activity of yogurt. Biotechnol Appl Biochem. 2021;68:221–9.

    Article  CAS  PubMed  Google Scholar 

  155. Nicklas M, Schatton W, Heinemann S, Hanke T, Kreuter J. Enteric coating derived from marine sponge collagen. Drug Dev Ind Pharm. 2009;35:1384–8.

    Article  CAS  PubMed  Google Scholar 

  156. Chen C, Liu F, Yu Z, Ma Y, Goff HD, Zhong F. Improvement in physicochemical properties of collagen casings by glutaraldehyde cross-linking and drying temperature regulating. Food Chem. 2020;318: 126404.

    Article  CAS  PubMed  Google Scholar 

  157. Wu X, Liu Y, Liu A, Wang W. Improved thermal-stability and mechanical properties of type I collagen by crosslinking with casein, keratin and soy protein isolate using transglutaminase. Int J Biol Macromol. 2017;98:292–301.

    Article  CAS  PubMed  Google Scholar 

  158. Hu Y, Liu L, Gu Z, Dan W, Dan N, Yu X. Modification of collagen with a natural derived cross-linker, alginate dialdehyde. Carbohyd Polym. 2014;102:324–32.

    Article  CAS  Google Scholar 

  159. Praet SFE, Purdam CR, Welvaert M, Vlahovich N, Lovell G, Burke LM, et al. Oral supplementation of specific collagen peptides combined with calf-strengthening exercises enhances function and reduces pain in achilles tendinopathy patients. Nutrients. 2019;11:76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Zhang D, Wu X, Chen J, Lin K. The development of collagen based composite scaffolds for bone regeneration. Bioactive Mater. 2018;3:129–38.

    Article  Google Scholar 

  161. Restaino OF, Finamore R, Stellavato A, Diana P, Bedini E, Trifuoggi M, et al. European chondroitin sulfate and glucosamine food supplements: a systematic quality and quantity assessment compared to pharmaceuticals. Carbohydr Polym. 2019;222: 114984.

    Article  CAS  PubMed  Google Scholar 

  162. Pomanenko NY, Mezenova OY, Nekrasova YOJVM. Specialized sports nutrition products using protein hydrolysis compositions of collagen-containing fish raw materials. 2021.

  163. Isaacs RB, Hellberg RS. Shark cartilage supplement labeling practices and compliance with U.S. regulations. J Dietary Suppl. 2021;18:44–56.

    Article  Google Scholar 

Download references

Acknowledgements

This work was funded by the “Pioneer” and “Leading Goose” R & D Program of Zhejiang Province (2023C02044).

Author information

Authors and Affiliations

Authors

Contributions

SF: Investigation, Conceptualization, Writing—original draft. MIA and: Designed figures, UA: Data curation. SZ: Designed tables, YL and CS: Contributed to manuscript revision. HZ: Conceptualization, Project administration, Supervision, Writing—review & editing, Funding acquisition.

Corresponding author

Correspondence to Hui Zhang.

Ethics declarations

Competing interests

Hui Zhang serves as the Associate Editor-in-Chief of Collagen and Leather, and he was not involved in the editorial review, or the decision to publish this article. All authors declare that there are no competing interests.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Farooq, S., Ahmad, M.I., Zheng, S. et al. A review on marine collagen: sources, extraction methods, colloids properties, and food applications. Collagen & Leather 6, 11 (2024). https://doi.org/10.1186/s42825-024-00152-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s42825-024-00152-y

Keywords