Next Article in Journal
An Improved Dual-Sorting NSGA-II Method for Optimal Radiation Shielding Design
Previous Article in Journal
Synthesizing Time-Series Gene Expression Data to Enhance Network Inference Performance Using Autoencoder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Ocean to Market: Technical Applications of Fish Protein Hydrolysates in Human Functional Food, Pet Wellness, Aquaculture and Agricultural Bio-Stimulant Product Sectors

1
Food BioSciences Department, Teagasc Food Research Centre, Ashtown, Dublin 15, D15 DY05 Dublin, Ireland
2
Department of Food Science and Environmental Health, Technological University Dublin (TU Dublin), Grangegorman, D7, D07 H6K8 Dublin, Ireland
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(10), 5769; https://doi.org/10.3390/app15105769
Submission received: 4 March 2025 / Revised: 25 April 2025 / Accepted: 2 May 2025 / Published: 21 May 2025

Abstract

:
Sustainability in food production is a pressing priority due to environmental and political crises, the need for long-term food security, and feeding the populace. Food producers need to increasingly adopt sustainable practices to reduce negative environmental impacts and food waste. The ocean is a source for sustainable food systems; deforestation, water scarcity, and greenhouse gas emissions burden traditional, terrestrial resources. Our oceans contain the largest unexploited resource in the world in the form of mesopelagic fish species, with an estimated biomass of 10 billion metric tons. This resource is largely untapped due in part to the difficulties in harvesting these species. To ensure sustainability of this resource, management of fish stocks and fish processing practices must be optimised. Generation of fish protein hydrolysates from by-catch/underutilised species creates high-value, functional ingredients while also reducing waste. Marine hydrolysates offer a renewable source of nutrition and align with the principles of the circular economy, where waste is minimised and resources are reused efficiently. Ocean-derived solutions demand fewer inputs, generate less pollution, and have a smaller carbon footprint compared to traditional agriculture. This review collates clearly and succinctly the current and potential uses of FPHs for different market sectors and highlights the advantages of their use in terms of the scientifically validated health benefits for humans and animals and fish, and the protection and crop yield benefits that are documented to date from scientific studies.

1. Introduction

Protein hydrolysates are manufactured primarily through enzymatic or chemical hydrolysis using proteolytic enzymes. While techniques such as high-pressure processing (HPP), ultrasound, and heat-treatment are used to aid in protein unfolding and enhance the exposure of specific amino acid residues, thereby improving enzyme accessibility and hydrolysis efficiency [1,2,3]. Protein hydrolysates are known to exhibit a wide spectrum of bioactivities—ranging from antimicrobial, anti-inflammatory, and antioxidant effects to anti-hypertensive, anti-diabetic, appetite-regulating, and weight management benefits. However, these functional properties are intrinsically linked to the release of specific bioactive peptides and are highly influenced by the amino acid composition and the positional arrangement of these residues within the peptide chains [4,5]. Bioactive peptides are defined as short chains of amino acids—typically comprising two to thirty residues—that, once liberated from their parent protein, can exert physiological effects and confer health benefits [6]. These peptides are incorporated into foods, nutraceuticals, and pharmaceutical formulations, and can be delivered through various routes including oral, nasal, and rectal administration, or applied topically, as in the case of cosmetics and dermatological treatments [7].
Fish protein hydrolysate (FPH) generation has gained popularity as a strategy to add value to fish processing co-products of primary production and to reduce food waste and develop new ingredients and new markets, adding value to the harvested fish resource or fish processed from aquaculture production. The health benefits of FPHs are well documented and several FPH products are sold globally for human health benefits [8,9]. However, regulations with regard to making health claims in Europe and elsewhere including the USA, Japan, and China are strict and require significant financial and scientific investments to enable products to be placed on the market [9]. Development of functional foods/nutraceuticals is similar in terms of the strategy required for the development of pharmaceutical bioactives and requires careful characterisation of bioactives, pre-clinical, and clinical trials. Different markets can be developed for FPHs, intended for the human market, during characterisation of the hydrolysates [10]. Herein, we detail the use of FPHs and bioactive peptide products in pet care ingredients and palatants as ingredients for use in aquaculture and as biostimulants agents in agriculture. Existing details of products in these sectors are collated and studies concerning the use of FPHs in pet feed and aquaculture for animal health benefit are described. Work concerning the regulations of FPHs as human, pet, and aquaculture feed ingredients and as biostimulants are cited within.

2. FPHs for Human Functional Benefits

Functional foods impart health benefits to the consumer that go beyond basic, human nutrition [11]. They provide effects such as anti-hypertension, anti-inflammatory, and anti-microbial actions, bone and skin health benefits, inhibition of enzymes that play roles in diseases associated with metabolic syndrome like type-2 diabetes (T2D), obesity and others [12,13,14,15,16]. FPHs, rich in bioactive peptides generated through enzymatic protein degradation, have been shown to modulate key enzymes involved in maintaining human health. These include alpha-amylase and dipeptidyl peptidase-IV, which contribute to type 2 diabetes (T2D) management and satiety regulation, as well as angiotensin-converting enzymes (ACE-1 and ACE-2) and renin, whose inhibition is associated with blood pressure reduction [17]. Hydrolysates find applications as human functional foods in different food categories, including Speciality Foods that encompass different product formats including snack products, frozen refrigerated goods, cheese, and plant-based products, breads/baked goods, entrées, coffee, chocolate/confectionary ingredients, desserts, and water. They are also found within the health and wellness category of food products, where they are promoted for their positive effects on inflammation, heart health, mood, anxiety prevention, and the promotion of sleep, gut health benefits, as well as several other health impacts [18]. A range of FPHs, including those derived from pelagic species, are commercially available and marketed with various health-related claims, as summarised in Table 1. These products are predominantly sold as ingredients for use in functional foods and dietary supplements. However, only a limited number have secured approval from the European Food Safety Authority (EFSA) for novel food status or authorised health claims within the European Union (EU) [19]. An exception to this is the thermolysin Bonito fish hydrolysate with active ingredient LKPNM. This hydrolysate has antihypertensive activities and has held Foods of Specified Health Use (FOSHU) status since 1997 in Japan [20]. This product does not have a cause and effect EFSA health claim for antihypertensive activity. EFSA found that a cause-and-effect relationship was not possible for the maintenance of normal blood pressure claims due to the active ingredient LKPNM, but the Bonito hydrolysate did receive Novel Food Status from EFSA [21]. There is a need to carry out human dietary intervention and clinical trials for FPHs if health claims are sought. This is just one, largely economic barrier towards the use of FPHs on a wide-scale basis in preventative medicine.

2.1. FPH with Anti-Inflammatory Activities

FPH may be rich in anti-inflammatory peptides with the potential to inhibit/impact various enzymes involved in the process of inflammation. These peptides are usually positively charged, containing amino acids like lysine, arginine and histidine, short peptides (usually 2–10 amino acids in length) with low molecular weights, and contain hydrophobic amino acid residues including Alanine, Valine, Proline, Isoleucine, Leucine, and others. Molecular weight cut off filtration may be used to generate extracts rich in these peptides (usually <3000 Da in size) [22]. Anti-inflammatory activity may be due to the amino acid composition or the folding of the peptide and sidechain interactions with the targets. Targets include cyclooxygenase (COX-1, COX-2) enzymes IL-6, IL-8 and TNF-α [23]. Examples of fish derived anti-inflammatory hydrolysates containing characterised peptides include a blue mussel hydrolysate >5 kDa in size where the anti-inflammatory activity is likely due to the secondary and tertiary structure of larger peptides found in the hydrolysate [24]. An Alcalase hydrolysate of the Green-lipped mussel containing the peptide sequence EGLLGDVG could downregulate COX-2 protein expression and iNOS in RAW 264.7 mice macrophage cells [25]. The peptides SNKGGGRPN, PGVATAPTH, LLGLGLPPA derived using papain applied to salmon bones also was found to inhibit COX-2, NO, IL-6, iNOS, and TNF-α mRNA in RAW 264.7 mice macrophage cells [26]. Other examples of FPH with anti-inflammatory activity are available in the excellent review of Kemp and colleagues [24]. Antihypertensive hydrolysates were also developed from Blue whiting and mackerel co-products of processing recently for use in companion animals [27].

2.2. FPH with Anti-Hypertensive Activities

Examples of FPH with anti-hypertensive activities that are commercially available include Valytron® with the active, bioactive peptide VY derived from Sardine muscle as well as Vasotensin Peptide 90 derived from Katsuwonus pelamis containing the active peptide LPK and LKPNM (Table 1) [28,29]. Bioinformatics tools are used to predict in vitro ACE-1 and Renin inhibitory effects. Indeed, Abdelhedi and colleagues demonstrated that ACE-1 inhibitory peptides are usually derived from actin and collagen (alpha-1 and alpha-2) proteins and enzymes including papain and Alcalase [30]. ACE inhibition is either competitive or non-competitive inhibition of the ACE-1 enzyme. Competitive enzyme involves interaction of the inhibitor with the active enzyme sites to prevent substrate binding and non-competitive inhibition involves binding of the molecule to the free enzyme and the enzyme–substrate complex [31]. Naik, Mora, and Hayes identified several ACE-1 inhibitory peptides from hydrolysates generated from Blue mussels [32]. Peptides identified as ACE-1 inhibitory in these hydrolysates include those with sequences FNAEKGFGF, KPEAPKVP, and SSDVPGV [32]. Renin inhibitory FPH were developed from cod previously and the most active peptides were assessed for their ability to reduce systolic blood pressure (SBP) in spontaneously hypertensive rats (SHRs) [33]. ACE-1 and Renin inhibition IC50 values of 0.13 mg/mL and 0.16 mg/mL, respectively, were reported [33].

2.3. FPH with Anti-Diabetes Type 2 (T2D) and Satiety Activities

FPH may contain peptides that can affect the progression or prevent T2D. Peptides from FPH can stimulate glucagon-like peptide-1 (GLP-1) secretions and enhance insulin release from pancreatic cells and may also inhibit enzymes like dipeptidyl peptidase-IV activity (DPP-IV), resulting in an increase in glucose uptake and tolerance, a reduction in blood glucose concentration, and the up-regulation of GLUT4 [34]. Zhou and colleagues [35] recently collated a review on fish and milk derived peptides from hydrolysates and their impact on T2D development [35]. A review of the literature concerning FPH and in vivo studies came to the conclusion that even though there is methodological variation between several studies performed to date, there is significant potential for the application of FPH to control hyperglycaemia and weight [35]. A study by Cudennec and colleagues demonstrated both in vitro and in vivo evidence for a satiating effect of FPH obtained from blue whiting (Micromesistius poutassou) muscle [36]. Several in vivo studies with FPH looking at the anti-hyperglycaemic effect of FPH have shown a reduction in blood glucose with doses of FPH as low as 50 mg kg−1 bodyweight [37].
Furthermore, all in vivo studies investigating the satiating effect of FPH have shown a decrease in bodyweight or a reduction in weight gain [38]. Antidiabetic peptides mainly inhibit dipeptidyl peptidase-IV (DPP-IV) activity. Inhibitors of DPP-IV can cleave incretins like glucagon-like peptide 1 (GLP-1) and glucose-dependent insulinotropic peptide (GDP). Li-Chan and colleagues generated the peptide GPAE from salmon gelatine previously and this peptide was shown to have considerable inhibitory effects on DPP-IV [39]. DPP-IV inhibitory peptides were also identified from discarded Sardine pilchardus protein previously [39,40].

2.4. FPH with Bone and Skin Health Implications

FPHs have emerged as powerful multifunctional ingredients for skin health, with growing evidence supporting their role in combating photoaging and pigmentation disorders [41,42]. Sourced from a variety of marine species, FPHs are rich in bioactive peptides that exert a wide range of beneficial effects, including antioxidant, anti-inflammatory, anti-apoptotic, and anti-melanogenic actions. These peptides not only help preserve extracellular matrix integrity and enhance skin hydration but also protect against UV-induced oxidative stress and DNA damage, making them ideal candidates for functional foods, nutraceuticals, and cosmeceuticals [43].
Recent studies have illuminated the molecular mechanisms behind these effects. Chen et al. [44] demonstrated that gelatine peptides from Pacific cod skin inhibited MMP-1, MMP-3, and MMP-9 by suppressing the MAPK pathway, thereby preserving collagen and enhancing dermal structure in UV-exposed skin. Similarly, Li et al. [45] found that silver carp scale-derived peptides (SCPs1) inhibited melanogenesis in B16 melanoma cells via downregulation of the cAMP-CREB-MITF axis. These peptides also exhibited strong antioxidant activity by reducing reactive oxygen species (ROS) and boosting intracellular glutathione levels, positioning them as dual-function agents for both skin brightening and oxidative protection.
Furthering this line of research, Losageanu et al. [46] investigated peptides from silver carp bones (FBBPs), reporting cytoprotective effects in UVB-irradiated skin cells, suppression of TNF-α in inflamed macrophages, and significant reductions in melanin production and tyrosinase activity. Complementing these findings, Kong et al. [47] identified three peptides—TCP3, TCP6, and TCP9—from skipjack tuna, which protected keratinocytes from UVB damage by stabilising mitochondrial membrane potential and regulating apoptosis-related markers through activation of the Nrf2–Keap1 pathway.

2.5. FPH with Antioxidant and Antimicrobial Activities

Oxidative stress, a key contributor to the degradation of essential biomolecules in both biological systems and food matrices, is closely linked to cellular impairment, tissue damage, and the decline of food quality. Antioxidants serve as critical defenders against this process by neutralising free radicals and interrupting oxidative chain reactions [48]. In recent years, marine-derived proteins have gained significant attention as a rich source of natural antioxidant peptides, owing to their promising bioactivity and diverse structural profiles [49]. A growing body of research has focused on the enzymatic hydrolysis of proteins from various fish species to isolate and characterise peptides with strong antioxidative potential [50,51,52,53]. These peptides, typically under 6000 Da in molecular weight, display activity influenced by several key factors, including peptide length, amino acid sequence, hydrophobicity, and overall composition [54].
Halim et al. [50], in their review of FPHs, noted that short-chain peptides composed of 2 to 10 amino acid residues tend to exhibit superior antioxidant efficacy. This heightened activity is largely attributed to the presence of specific amino acids known for their redox potential. Hasani et al. [51], further supported this observation in a study utilising Indian mackerel waste hydrolysed with Alcalase and Flavourzyme, revealing that residues such as valine, leucine, proline, histidine, and particularly tyrosine play pivotal roles. Tyrosine, for instance, acts as an effective electron donor, helping to stabilise reactive species and prevent oxidative damage.
Peptides exert their antioxidant effects through various mechanisms. Firstly, they can donate hydrogen atoms or electrons to free radicals, effectively neutralising and stabilising these reactive species and thereby preventing oxidative damage to biological molecules [50,55]. Secondly, peptides possess the ability to chelate transition metal ions such as iron and copper, which are key catalysts in the generation of reactive oxygen species; by binding these metals, peptides inhibit the redox reactions that fuel oxidative stress [56,57]. A third mechanism involves the direct interaction of peptides with reactive oxygen species, through which they interrupt the lipid peroxidation chain reaction and halt the propagation of oxidative damage across cellular membranes or food lipids [58]. Lastly, peptides may function as physical barriers by forming protective layers around lipid droplets, thereby preventing the penetration of pro-oxidative initiators and safeguarding lipid molecules from degradation [59].
Bi and colleagues [60] identified three potent antimicrobial peptides from the enzymatic breakdown of Turbot viscera. These peptides, generally comprising between 1 and 50 amino acids, had molecular weights under 10 kDa. Structurally, they exhibited amphipathic properties, were notably enriched in cysteine, and carried a net cationic charge in their active conformation—features that enabled them to effectively target and disrupt bacterial cell structures.
In their study, Pezeshk and colleagues [52] enzymatically hydrolysed the viscera of yellowfin tuna (Thunnus albacares) and fractionated the hydrolysate into four distinct molecular weight fractions via membrane ultrafiltration (<3 kDa, 3–10 kDa, 10–30 kDa, and >30 kDa). They evaluated the antioxidant and antibacterial activities of both the unfractionated hydrolysate and its individual fractions. The findings revealed that the smallest molecular weight fraction (<3 kDa) exhibited the most potent antibacterial effects, significantly inhibiting both Gram-positive bacteria and Gram-negative strains. These results underscore the exceptional structural and functional traits of marine-derived antimicrobial peptides, which tend to show diverse structural motifs, broad-spectrum activity, limited sedimentation, and high bacterial target specificity.
Perez Espitia et al. [61] proposed that antimicrobial peptides exert their effects by first attaching to the pathogen cytoplasmic membrane. This initial interaction is followed by the formation of pores within the membrane structure. These pores compromise the membrane’s structural and functional integrity, which in turn interferes with vital processes such as cellular respiration and the maintenance of electrochemical gradients. As a result, uncontrolled movement of water and ions into the cell occurs, causing cellular swelling and, ultimately, lysis.
Two primary mechanisms have been proposed to describe how antimicrobial peptides form pores in bacterial membranes: the carpet model and the barrel-stave model [61]. The carpet model involves peptides spreading across the membrane surface like a “carpet”. Once a critical concentration is reached, these peptides destabilise the membrane through a combination of hydrophilic and hydrophobic interactions. This destabilisation leads to the formation of transient pores or micelle-like structures, disrupting membrane integrity and resulting in its complete disintegration. Conversely, in the barrel-stave model, peptides adopt amphipathic alpha-helical structures that insert themselves perpendicularly into the lipid bilayer. Their hydrophobic regions interact with the lipid tails of the membrane, while the hydrophilic sides face the interior of the formed pore. As more peptides accumulate, these channels enlarge, leading to membrane permeabilization, leakage of intracellular components, and eventual cell death.
Da Rocha [62] highlighted that the antimicrobial effectiveness of protein hydrolysates derived from Umbrina canosai is influenced by a range of interrelated factors. These include the peptide’s hydrophobicity, net charge, amino acid profile, size, and secondary structure, as well as environmental parameters such as pH, temperature, and ionic strength. Additionally, the source of the peptide, its concentration, and the presence of surfactants play significant roles. Importantly, the structural characteristics of the bacterial membrane also contribute to determining how effectively these peptides can exert their antimicrobial action.
Table 1. Fish protein hydrolysate (FPH) products with associated health benefits.
Table 1. Fish protein hydrolysate (FPH) products with associated health benefits.
Product NameManufacturerFish Source and Active IngredientHealth Benefit ClaimedReferences
Jellice (hydrolysate) found in CollametaTMJellice Pioneer Europe PV, (7821 BG, Emmen, The Netherlands); supplied into product CollaMeta produced by Glanbia Nutritionals Ltd. (Global)Skin from Tilapia and Pangasius sp./Collagen tripeptideSkin health, possible prevention of arthritis in humans[63]
Collagen HMTMCopalis, France
(Le Portel, Hauts, France)
Different fish sources including white fish species and salmon/Collagen HMTM polypeptides mean molecular weight is 3600 Da, making it soluble in aqueous phase and fully digestibleSkin and joint health[64]
NutripeptinTMCopalis, France
(Le Portel, Hauts, France)
Bioactive marine peptideReduces the glycaemic index of foods and thus helps to reduce fat storage in weight-control formulas[65]
ProtizenTMCopalis, France
(Le Portel, Hauts, France)
Pollock and coalfish digest produced by enzymatic hydrolysis/anti-stress peptide that promotes well-being, found in Serenlider product—possess anxiolytic properties and is considered as beneficial for mental health/stressAnti-stress, mood food, promotes restful sleep[66]
Molval®Dielen Laboratoire, France
(Cherbourg-en-Cotentin, Normandy, France)
Extracted from cod and mackerel/Gabolysat PC60.Beneficial effects on dyslipidaemia and cardiovascular risk, contains omega-3 fatty acids and Gabolysat PC60/anti-stress effects[67]
Curcumega®Dielen Laboratoire, France
(Cherbourg-en-Cotentin, Normandy, France)
Fish oil rich in omega-3 fatty acids combined with turmericNutritional impact and antioxidative effects[67]
Seagest®Trimedica, USA
(Santa Monica, CA, USA)
Deep ocean within fish species none found
PeptACE®Natural factors, USA
(14224 167TH Ave SE Monroe, WA, USA)
Bonito fish hydrolysate/nine small peptides derived from Bonito fishMaintains normal blood pressure[68]
Stabilium® 200Nutricology, Canada
(Salt Lake City, UT, USA)
Derived hydrolysate from
Molva dypterygia
Improves resilience to stress, may reduce fatigue and supports normal psychological
functions, such as memory, concentration, and cognitive
abilities
[69]
Amizate®Zymtech Production AS, Norway
(Oppland, Lesja)
Enzymatic process extracts amino acids, short peptides, and micronutrients from Atlantic salmon (Salmo salar)Excellent nutritive value; mood improving agent[70]
WhiteCal®BioMarine Ingredients Ireland (BII), Ltd.
(Ballybay, Monaghan, Ireland)
Blue whiting fish/WhiteCal is a marine mineral complex, naturally rich in collagen peptides and high in calcium, phosphorus, and magnesium. It can be easily used in a wide range of human nutrition applicationsGrowth and repair of the body and maintenance of good health, bone, and joint health.[71]
ProAtlantic®BioMarine Ingredients Ireland (BII), Ltd.
(Ballybay, Monaghan, Ireland)
Premium grade hydrolysed Fish Protein Isolates (FPI) containing 95% protein and 0.5% fat for human productsGlycaemic control, diabetes prevention, satiety[72]

3. FPHs—As a Pet Health Food Ingredient

The use of FPHs in pet food exemplifies the intersection of advanced nutritional science and sustainable innovation. Valued at USD 1.4 billion in 2023, the market for fish-based pet health products is projected to surge to USD 16.7 billion by 2033, with an annual growth rate of 5% [73]. This growth is driven by rising pet ownership, the humanisation of pets, and a growing demand for premium, health-focused products [74]. Pet owners are increasingly willing to invest in diets offering superior taste, nutrition, and functional benefits [75]. FPH-based formulations are associated with enhanced digestibility, nutrient bioavailability, and multiple health advantages [76].
Age-related conditions in pets—such as inflammation, allergies, coat deterioration, and general health decline—also influence consumer preferences for fish-derived ingredients known to address these issues [77]. Moreover, sustainability concerns are shaping purchasing behaviour, with consumers seeking eco-friendlier alternatives to traditional meat-based pet foods [78]. FPH stand out for their hypoallergenic properties, high nutritional value, and alignment with sustainable resource utilisation [79]. FPHs are produced via enzymatic hydrolysis, a process that breaks down fish proteins into smaller peptides and amino acids, enhancing their digestibility and bioavailability—encompassing bioaccessibility, cellular uptake, and metabolic transformation [80,81]. This method also yields bioactive peptides that provide functional health benefits beyond basic nutrition [81]. They deliver a well-balanced amino acid profile, rich in essential amino acids such as lysine, methionine, and tryptophan—vital for muscle development, metabolism, and neurotransmitter synthesis in pets [82]. Their hypoallergenic nature makes them ideal for pets with food sensitivities, offering a superior alternative to conventional proteins like beef, chicken, or soy [83,84].
Moreover, FPH are highly palatable, enhancing food acceptance and consistent intake [85]. Certain peptides exhibit antimicrobial activity, promoting gut health by suppressing pathogenic bacteria [86,87,88,89,90], which is increasingly relevant due to the gut–brain axis’s influence on pet behaviour and neurological well-being [91]. Anti-inflammatory peptides present in FPH may also support the management of chronic conditions common in ageing pets [92,93,94,95,96]. Additionally, the collagen content in FPH supports joint health and mobility, with collagen peptides offering superior digestibility and bioavailability [97].
The production of FPHs exemplifies a sustainable approach rooted in circular bio economy principles, promoting the valorisation of fish processing by-products such as skins, bones, and viscera. This not only reduces the environmental impact but also enhances resource efficiency by converting waste into high-value ingredients for pet food [98,99].
FPH are produced through selective enzymatic hydrolysis and controlled fermentation, enabling precise modulation of peptide molecular weight and optimisation of bioactive properties [100]. In silico techniques further support this process by predicting the functional potential of peptides, particularly when integrated with in vitro hydrolysis methods [101]. These precision approaches enable the development of targeted pet food solutions for managing chronic conditions such as obesity, diabetes, and kidney disease. For example, low molecular weight peptides derived from fish have demonstrated antihypertensive effects, highlighting their role in preventing cardiovascular issues in pets [102,103].
With rising incidences of chronic ailments—such as cardiovascular disease, arthritis, digestive disorders, oxidative stress, renal dysfunction, and cognitive decline—often linked to ageing, poor diets, or genetic predisposition, FPH are gaining attention as a functional ingredient for health-oriented pet nutrition [8,104,105]. Rich in bioactive peptides and essential fatty acids like EPA and DHA, FPH exhibit anti-inflammatory, antioxidant, and immunomodulatory properties, offering a holistic strategy to enhance pet health, well-being, and longevity [106].

3.1. Cardiovascular and Hypertensive Health Issues

Cardiovascular disease (CVD) affects over 10% of dogs, with its incidence increasing by 1.5-fold annually [107]. Large, ageing dogs are particularly susceptible to dilated cardiomyopathy (DCM), while smaller breeds often develop degenerative mitral valve disease (DMVD), both of which contribute to significant veterinary costs and reduced quality of life [108]. Nutritional interventions—including coenzyme Q10, taurine, arginine, vitamin E, L-carnitine, and omega-3 fatty acids—have been shown to reduce inflammation, improve appetite, enhance cardiac function, and extend lifespan in affected dogs [109,110]. Fish-derived proteins and hydrolysates, rich in bioactive peptides and omega-3s, are increasingly recognised for their cardioprotective effects [111].
Pasławski et al. [112] reported that a fish-based diet enhanced cardiac and metabolic health in small-breed dogs with mitral valve disease. Similarly, Liu et al. [113] found that collagen hydrolysate from Atlantic salmon skin exhibited antioxidant, anti-inflammatory, endothelial-protective, and anti-platelet properties—comparable to aspirin—in atherosclerosis models. Key peptides, including FAGPPGGDGQPGAK and IAGPAGPRGPSGPA, reduced plaque buildup and arterial thickening in high-fat-fed mice. Maneesai et al. [114] further demonstrated that tuna protein hydrolysates (TPH) restored cardiovascular enzyme balance and reduced myocardial damage, oxidative stress, and inflammation in rats, yielding therapeutic effects comparable to metformin for managing obesity, hypertension, and hyperglycaemia.
Hypertension is another growing health concern in companion animals, particularly in ageing dogs and cats, often associated with chronic kidney disease, obesity, and diabetes [115]. Frequently asymptomatic in early stages, it can lead to severe complications such as heart failure, renal damage, and vision loss. The condition’s rising prevalence is closely tied to sedentary behaviour and poor diets, paralleling human trends [116]. Long-term treatment also imposes financial burdens on pet owners, prompting interest in natural, food-based interventions. Functional diets enriched with FPHs offer cardiovascular and antihypertensive benefits with minimal side effects [117].
Research supports the efficacy of FPHs in managing hypertension. In obesity-related renal insufficiency models, herring and salmon FPH significantly inhibited angiotensin-converting enzyme (ACE) activity, improved renal biomarkers, and demonstrated potent renoprotective effects. Herring and salmon hydrolysates yielded 81 and 49 ACE-inhibitory peptides, respectively. Obese Zucker rats on a 25% FPH diet exhibited decreased urinary protein, cystatin C, and glucose levels [118]. Aissaoui et al. [119] reported that red scorpionfish hydrolysates showed strong ACE-inhibitory and antioxidant activities. Likewise, Taheri and Bakhshizadeh [120] found that kawakawa FPH, especially peptides within the 1–3 kDa range rich in hydrophobic amino acids, possessed potent antioxidant and antihypertensive properties.

3.2. Arthritis

Arthritis is increasingly prevalent in ageing pets, impairing mobility, reducing quality of life, and leading to long-term veterinary costs [121]. In response, pet owners are turning to FPHs for their natural anti-inflammatory and joint-supporting properties, making them a promising dietary intervention for managing arthritis symptoms [122]. In a 16-week study on dogs with osteoarthritis (OA), deep sea fish oil significantly reduced malondialdehyde (MDA), a key marker of oxidative stress, and increased free carnitine levels—both indicators of improved metabolic health [122]. Compared to corn oil, fish oil also had more favourable effects on immune and metabolic markers, including reductions in specific white blood cells.
Manfredi et al. further demonstrated that a commercial fish-based diet enriched with fish oil fatty acids significantly reduced the incidence and severity of elbow dysplasia and OA in growing Labrador retrievers [123]. Similarly, studies on German shepherds with early-stage OA showed that collagen hydrolysates, sulphated glucosamine, and fatty acid-enriched diets (e.g., Hill’s JD) improved mobility and agility [124]. Researchers concluded that combining marine-derived nutraceuticals with vitamin- and lipid-rich diets offers the most effective joint health support in pets.

3.3. Hair and Coat Issues

Hair and coat disorders—such as shedding, dullness, and skin irritation—are common in pets and often stem from poor nutrition, allergies, or underlying conditions [125]. FPHs, rich in essential amino acids and omega-3 fatty acids, are increasingly used to promote skin and coat health. They help strengthen the skin barrier, reduce inflammation, and support hair growth, leading to shinier coats and healthier skin [126,127].
Fritsch et al. reported that Hill’s® Prescription Diet® Canine Skin Support, enriched with salmon protein and fish oil, significantly improved redness, skin thickness, and sores in 101 dogs with chronic pruritus, while owners noted reduced itching and better coat condition [128]. Noli et al. [84] reported that a hydrolysed fish and rice starch-based diet (Farmina Ultra Hypo—FUH) effectively alleviated skin allergy symptoms in dogs with pruritus while also modulating gut microbiota composition. The study highlighted that the FPHs reduced immune reactivity to dietary proteins, promoted beneficial gut microbes, and significantly eased allergic skin responses. Similarly, Szczepanik et al. demonstrated that a hypoallergenic diet with 18% hydrolysed salmon significantly reduced skin lesions and itching in dogs and cats with food-related allergic dermatitis, showing over 50% symptom reduction using clinical scoring systems [129]. These findings highlight the therapeutic role of FPH in managing dermatological issues in companion animals.

3.4. Digestive, Renal, and Metabolic Health Issues

Digestive disorders such as chronic enteropathy (CE) and inflammatory bowel disease (IBD) frequently affect pets, leading to symptoms like vomiting, diarrhoea, weight loss, and malnutrition. These conditions often result from abnormal immune responses to dietary or environmental triggers [130,131,132]. FPHs, composed of highly digestible and low-allergen peptides, have demonstrated efficacy in reducing gut inflammation and minimising immune activation [71,133,134]. CE, which includes food-responsive (FRE), antibiotic-responsive (ARE), and immunosuppressant-responsive (IRE) forms, is most commonly food-responsive. More severe forms, such as protein-losing enteropathy (PLE), pose greater risks due to protein loss and poor prognosis [134,135]. Diets based on hydrolysed or novel proteins have shown significant benefits in FRE and PLE cases [135,136].
Ontsouka et al. [137] found that a fish meal and potato protein diet enriched with omega-3 fatty acids reduced inflammation and improved fatty acid absorption in dogs with FRE and IBD. Simpson et al. [138] observed clinical recovery and sustained remission in PLE dogs fed hydrolysed fish-based diets, with notable gains in weight, serum albumin, and micronutrient levels. Zinn et al. [139] compared fish-based ingredients—including pink salmon hydrolysate, milt meal, and white fish meal—to poultry by-product diets in senior dogs, finding improved nutrient digestibility and reduced pro-inflammatory cytokines, underlining FPHs’ role in gut health and immune modulation.
Oxidative stress, resulting from an imbalance between reactive oxygen species (ROS) and antioxidant defences, contributes significantly to cellular damage, inflammation, and the progression of chronic kidney disease (CKD), particularly in ageing pets [140]. FPHs are rich in antioxidant peptides that combat oxidative stress, reduce inflammation, and support renal function, making them a natural dietary strategy for CKD management [141].
Nasri et al. [142] showed that goby FPH improved antioxidant status and metabolic markers in rats fed a high-fat, high-fructose diet, suggesting renal and glycaemic benefits. De Godoy et al. [143] reported that fish oil supplementation in lean Beagles increased ghrelin, stabilised body weight, and reduced oxidative stress, cholesterol, and glucose levels. Similarly, Riyadi et al. [144] demonstrated that tilapia viscera hydrolysate lowered oxidative markers and improved kidney structure in hypertensive rats. In working dogs, Ravić et al. [145] found that fish-based diets decreased blood glucose, LDL cholesterol, and oxidative stress while enhancing lipid profiles, supporting cellular health under physical strain. Additionally, Theysgeur et al. [146] used a canine gastrointestinal model to show that tilapia hydrolysates promoted secretion of appetite-regulating hormones (CCK and GLP-1) and inhibited DPP-IV activity, indicating potential for appetite control and obesity prevention in pets.

3.5. Cognitive and Emotional Health in Ageing Pets

Age-related memory decline and anxiety are increasingly recognised as significant concerns in ageing pets, particularly dogs and cats. Cognitive deterioration, often manifested through disorientation and behavioural changes, is linked to oxidative stress and neuro-inflammation—mechanisms similar to those seen in humans [147]. Functional diets enriched with FPHs, which are rich in neuroprotective and antioxidant bioactive peptides, are gaining attention for their potential to mitigate oxidative damage, support brain health, and slow cognitive decline.
Zicker et al. [148] demonstrated that beagle puppies fed DHA-rich fish oil from weaning to one year showed improved memory, learning, coordination, visual function, and immune response. Similarly, Pan et al. [149] found that fish oil supplementation in senior cats with cognitive dysfunction enhanced memory and reduced signs of brain ageing. Further research by Pan [150] showed that a Brain Protection Blend (BPB)—combining fish oil and L-arginine—improved task performance in older dogs, especially in spatial challenges. In ageing mice, Chataigner et al. [151] reported that a fish hydrolysate diet enriched with omega-3s enhanced memory, reduced anxiety, balanced stress hormones, and increased neurotrophic factors like NGF and BDNF.
Anxiety, often triggered by separation or loud noises, is another prevalent issue in dogs, presenting as excessive barking, destructive behaviour, and nervousness [152]. FPHs have demonstrated anxiolytic properties by modulating the hypothalamic–pituitary–adrenal (HPA) axis and enhancing GABA activity, mirroring pharmaceutical interventions [153]. Bernet et al. [153] showed that Gabolysat® PC60 significantly reduced stress hormone release in rats, with Freret et al. [154] noting anxiolytic effects comparable to diazepam. Landsberg et al. [152] found that sustainable white fish-derived hydrolysates reduced anxiety and cortisol in dogs during simulated thunderstorms. Titeux et al. [155] observed improved stress resilience and sociability in fearful dogs following GABOLYSAT® PTP 55 supplementation.
In cats, Jeusette et al. [156] reported that a diet containing sardine peptides, lemon balm, oligo fructose, and L-tryptophan reduced urinary cortisol more effectively than L-tryptophan alone. Porcheron et al. [157] demonstrated that Zylkene Plus, a white fish hydrolysate, significantly decreased separation-related anxiety in dogs, with 49% showing behavioural improvement. Ephraim et al. [158] linked fish oil to reduced anxiety-related metabolites and increased beneficial gut bacteria in senior dogs. Dinel et al. [159] showed that Peptidyss®, a sardine-derived hydrolysate, lowered corticosterone and regulated stress-related gene expression in mice. Collectively, these studies highlight the dual role of FPHs in supporting cognitive function and emotional well-being in pets, making them a promising natural strategy for improving the quality of life in ageing companion animals. Commercialised products marketed as functional ingredients/products for pet health are shown in Table 2.

4. FPHs as Pet Palatants

In 2023, the worldwide pet food industry was valued at USD 120.87 billion. Forecasts indicate an expansion from USD 126.66 billion in 2024 to USD 193.65 billion by 2032, reflecting an annual growth rate of 5.45%. North America led the market in 2023, accounting for 42.55% of the market share (https://www.fortunebusinessinsights.com/industry-reports/pet-food-market-100554 (accessed on 23 September 2024)) [161]. Palatants also substantiate a crucial vertical of the pet food industry as this class of products play a vital role in influencing pets’ acceptance and preference for their food. These additives are incorporated into pet foods to enhance their taste and aroma, making them more appealing to animals [162]. The effectiveness of palatants is crucial in ensuring consistent consumption, especially among precise eaters or pets with specific dietary needs. In today’s competitive market, where pet owners increasingly prioritise high-quality, nutritious, and appealing food options for their pets, the inclusion of palatants can determine a product’s success [163]. The science behind palatants involves understanding animals’ sensory preferences and developing formulations that improve the overall palatability of pet food. This approach supports better nutrition and health outcomes for pets by encouraging them to consume food enriched with essential nutrients [163,164]. Overall, palatants play a pivotal role in the pet food industry, enhancing the quality and attractiveness of pet food products while promoting better dietary habits and overall well-being for pets. The market for pet food palatants was valued at $1981.2 million in 2023 and is expected to grow at a compound annual growth rate of 6.5% from 2024 to 2030 [165]. The appeal of pet food is influenced by numerous elements, such as the freshness of its components, the texture, size, and shape of the kibble, and particularly the use of palatants. Palatability is not only a critical component of pet nutrition, but also a key driver of brand loyalty. Palatants often consist of volatile and soluble compounds, including amino acids, peptides, and lipids [163]. These substances are typically obtained from animal by-products and fish through processes like acid hydrolysis, enzymatic treatment, or thermal liquefaction. Animal digests or hydrolysates that are partially hydrolysed animal parts in both dry and liquid forms are used as palatability enhancers or palatants in pet feeds. They are defined by the Association of American Feed Control Officials (AAFCO) (AAFCO) as “a material which results from chemical and/or enzymatic hydrolysis of clean and undecomposed animal tissue. The animal tissues shall be exclusive of hair, horns, teeth, hooves, and feathers, except in such trace amounts as might occur unavoidably in good factory practice and shall be suitable for animal feed. If it bears a name descriptive of its kind or flavour(s), it must correspond thereto” (Animal Proteins Prohibited in Ruminant Feed and Cattle Materials Prohibited in All Animal Feed, WSDA). Indeed, salmon protein hydrolysate is a well-known palatant product with documented evidence suggesting that it improves food consumption by dogs and cats [166].
Palatability is the physical and/or chemical attribute(s) of the diet of an animal, involving the promotion or suppression of feeding behaviour during the pre-absorptive period. It relates to pleasure perception or liking during consumption. Palatable foods/feeds are ones that acceptable to the companion animal. Palatants incorporate many different macro- and micro-molecules including proteins, amino acids, carbohydrates, fatty acids, peptides, vitamins, and minerals. The aim of these ingredients is to enhance the sensory experience of the animal, and this relates to the umami T1R1 and T1R3 taste receptors [167].
Cats have affinity for umami compounds. In the pet food industry, animal protein hydrolysates can create palatability enhancers via the Maillard reaction. Digests or palatants are proteins that are enzymatically broken down and applied to dry feeds to provide a sensory impact (usually meat flavoured). The Maillard reaction is the chemical reaction between amino acids and reducing sugars to create melanoidins (high molecular weight polymers or sugars). Melanoidins give food and feeds their distinctive flavour [168]. Additionally animal proteins, as well as specific amino acids, animal fats, and emulsified meats, are important flavours that are highly palatable to cats. Hydrolysis releases compounds directly from raw materials like proteins, which are not volatile and/or sapid. These compounds (like peptides derived from proteins) can react with sugars and other peptides if a thermal treatment is applied. The Maillard reaction requires temperatures > 80 °C. This reaction generates flavour compounds. Their nature, origin and chemical formulas determine the specific aroma of the product. Modification of the process and the raw material used to produce the palatability enhancers directly affects the generation of tastes and flavours. Palatants are delivered to the consumer (pet food manufacturers or directly to pet owners) in liquid or powder form.

4.1. Palatant Form

Palatants exist in both dry and liquid forms and their addition to kibbles following extrusion can enhance the flavour of pet feeds. They are used widely, and large markets exist in the USA, Australia, France, Japan, and Chile. Wet pet foods have higher palatability than dry foods due to their higher moisture content and processing techniques. The inclusion level of palatants in wet pet feed are generally lower than those in dry pet feed [169]. They can be poured or dusted on the surface of cat or dog kibbles [164]. The palatability of the final product is evaluated directly by the animal. Pet preferences for one or another product is determined by a trained panel of pets, i.e., cats or dogs, using different methodologies of food presentation, as could be done with humans [163,164]. Adopting palatants in emerging pet food markets are beneficial to pet feed manufacturers and their brands (as well as pets). As consumption of pre-packaged pet food grows, flavour requirements for the food become more important.

4.2. The Difference Between Cats and Dogs

An animal’s sense of taste can help them assess the nutrient content of feed and helps protect them against eating things that might harm them. Moisture or water content of a food plays a key role in palatability for both dogs and cats, both of which generally prefer moist foods (i.e., wet, fresh) rather than dry foods (kibble) as the moisture content of wet foods is more like fresh meat [170]. Cats are obligate carnivores and require protein in their diet to thrive and survive. They prefer foods with high protein, especially animal proteins, which ensures that they are meeting their dietary requirements. Cats have less preference for carbohydrates and fat and usually prefer higher protein containing foods/low carbohydrate diets when given the option. Most commercial dry pet feeds contain prominent levels of carbohydrates since they are critical for successful extrusion processing, whereas wet feeds and emerging fresh-cooked formats typically contain higher amounts of meat, making them naturally more palatable compared to kibble. Dogs are omnivores and are more opportunistic with their feed selection. Importantly, like humans, dogs require amino acids from protein to survive. Dogs (like humans) often show preference for foods with simple sugars and higher fat content for energy. Where there is a crude fibre increase in the diet there is a decrease in palatability of pet feeds for dogs [170].
Cats have a preference for umami flavours and respond to nucleotides and L-amino acids. L-histidine when metabolised leads to the production of glutamate and it is known that in combination with nucleotides acts as a palatant enhancer. Other L-amino acids are thought to decrease palatability while L-phenylalanine, L-tyrosine, L-tryptophan, L-methionine, L-arginine, L-isoleucine, L-leucine, L-serine, and any combination of these in an effective amount of between 0.001 and 0.8 wt. percent on dry matter enhance palatability of dog food for dogs.

4.3. Marine Hydrolysates as Pet Palatants

Recently, Guilherme-Fernandez and colleagues explored the use of squid meal and shrimp hydrolysate as a protein source for dogs [171]. Chemical composition, antioxidant activity, and palatability was evaluated by comparing a commercial diet with the inclusion of 150 g kg−1 of squid meal or shrimp hydrolysate in diets fed to Beagle dogs (2.2 ± 0.03 years). The Shrimp hydrolysate displayed a greater antioxidant activity than squid meal and the preference in terms of the First approach and taste were not affected by the inclusion of these hydrolysates, but dogs showed a preference for the basal diet. Kemin produce a pet palatant from salmon and the French company Copalis® produce CPSP® soluble fish protein concentrates that claim to provide optimum palatability and contribute to animal well-being. Other palatant products, some derived from fish, are listed in Table 3.

5. FPHs—Based Aquaculture Feeds

Aquaculture feeds are specifically designed to address the nutritional needs of aquatic species, combining various ingredients such as marine by-products, fish oils, plant-based components, live feeds, and vegetable proteins [172]. These feeds offer a balanced diet, supplying essential nutrients like proteins, vitamins, minerals, carbohydrates, and lipids that are vital for the growth, development, and reproductive health of fish and other aquatic organisms, ensuring their overall well-being [173]. The global aquaculture feed market has experienced remarkable growth, driven by the rising demand for seafood and the need for sustainable, high-quality feed solutions across a broad spectrum of species, from finfish to crustaceans [174]. In 2023, the global aquaculture feed market reached 49.70 million tonnes and is expected to rise to 78.93 million tonnes by 2032, with a steady CAGR of 5.3% from 2024 to 2032 [175]. The expanding aquaculture industry, heightened health awareness, increased seafood consumption, and a growing focus on sustainable fish production drive this market growth. As people recognise the health benefits of fish, including its rich protein content, omega-3 fatty acids, vitamins, and minerals, the demand for aquaculture feed has risen sharply [175,176]. Additionally, fish’s reputation as a healthier option compared to red meat due to its lower saturated fat levels has further fuelled the market’s expansion [177].
Manufacturers are increasingly turning to FPHs as a key ingredient due to their exceptional palatability, digestibility, and bioavailability, and ability to enhance growth, immune function, and nutrient absorption in aquatic species especially during early developmental stages or under stress conditions [178]. The primary mechanisms underlying FPH efficacy relate to their rapid absorption and multifaceted biological activity. Due to their short peptide length, FPHs bypass the need for extensive enzymatic digestion, allowing for faster uptake through peptide transporters (e.g., PepT1) in the intestinal epithelium [179,180]. This rapid nutrient assimilation contributes to improved feed efficiency, weight gain, and muscle accretion. Moreover, bioactive peptides in FPHs exert antimicrobial, antioxidant, immunomodulatory, and anti-inflammatory effects [180].
In terms of product development, the quality and functionality of FPHs depend on the choice of raw materials (e.g., fish offal, heads, viscera), hydrolysis conditions (enzyme type, temperature, pH, duration), and downstream processing techniques such as ultrafiltration and spray drying [181]. Commercial products such as Symrise Aqua Feed’s hydrolysates (derived from tuna, krill, and shrimp) and CPSP (Copalis Sea Solutions®) have shown consistent efficacy in improving gut histomorphology, nutrient digestibility, and pathogen resistance [182,183,184,185]. Furthermore, FPHs have enabled partial replacement of traditional protein sources like fish meal or soy protein concentrate by up to 50–60% without compromising performance, contributing to circular bio economy models in the aquafeed industry [186]. As a result, FPHs are emerging as a preferred choice for both economic and environmental reasons in the aquaculture industry.
Early evidence by Refstie et al. [187] showed that replacing part of the fish meal with FPHs made from fresh Pollock body parts (Denofa Fish Peptides, Denofa, Fredrikstad, Norway) in Atlantic salmon diets improved feed intake, growth, nutrient retention, and digestibility in a dose-dependent manner—likely due to increased palatability and peptide-driven nutrient assimilation. This pattern was echoed in subsequent studies, where the molecular weight and peptide composition of hydrolysates played a critical role in physiological outcomes. For instance, Kotzamanis et al. [185] found that a commercial hydrolysate (CPSP, Copalis Sea Solutions®) with peptides ranging from 500 to 2500 Da significantly enhanced growth and intestinal development in sea bass larvae, likely through stimulation of digestive enzyme activity and improved gut maturation.
Tang et al. [188] and Zheng et al. [189] further linked FPH inclusion to enhanced growth and immunity in large yellow croaker and Japanese flounder, respectively. Improved weight gain and immune parameters (e.g., lysozyme, immunoglobulin M) were associated with increased bioavailability of peptides and their immunomodulatory effects. Notably, ultra-filtered fractions rich in low molecular weight peptides were most effective, suggesting that smaller peptides more readily activate endocrine and metabolic pathways such as IGF-I, crucial for growth regulation.
This metabolic efficiency is evident in species like barramundi. Chaklader et al. [180] tested FPHs derived from yellowtail kingfish, carp, and tuna (Specialty Feeds, Glen Forrest, Australia), while Siddik et al. [190] evaluated tuna-based FPHs (SPF Diana Co. Ltd.), both reporting significant improvements in growth, fillet quality, and immune function. These benefits were linked to elevated polyunsaturated fatty acid levels, enhanced gut barrier integrity, and robust systemic immune responses, including greater resistance to Vibrio harveyi infection. Additionally, reduced hepatic and renal stress markers highlighted the hepatoprotective and anti-inflammatory potential of FPH supplementation. Similar benefits were recorded in shrimp, where Gunathilaka et al. [182] demonstrated that FPHs (Symrise Aqua Feed, Specialities Pet Food, France) improved feed efficiency, digestibility, carcass lipid content, and antioxidant status. The bioactive peptides likely improved intestinal absorptive capacity and cellular oxidative defence, supporting tissue repair and energy metabolism.
In Channa striata, Siddaiah et al. [191] and Suratip et al. [192] reported enhanced weight gain, blood profiles, and immune activity, including increased lysozyme and myeloperoxidase levels. These improvements were linked to elevated haemoglobin and serum proteins, reflecting systemic immune stimulation and better nutrient transport—key for robust disease resistance and survival. Tuna hydrolysate supplementation in Asian sea bass diets showed improved growth, feed efficiency, and protein utilisation (Tola et al. [193]). The observed enhancement in diet palatability and nutrient digestibility enabled a substantial replacement (up to 55%) of fish meal with plant protein sources without compromising performance. This underscores FPH’s role as a functional dietary enhancer that mitigates the antinutritional effects of plant ingredients.
Suma et al. [194] demonstrated similar outcomes in Ompok pabda, where a 2% inclusion of FPH yielded optimal growth, survival, liver function, and immune responses. Improved hepatosomatic index and condition factor were attributed to enhanced feed intake and efficient metabolic conversion. In salmonids, Sandbakken et al. [195] confirmed that enzymatically produced salmon hydrolysates improved protein and amino acid digestibility, mineral absorption, and overall growth. These outcomes reflect the hydrolysate’s ability to bypass typical digestive constraints and supply essential nutrients in highly available forms.
Recent studies in 2024 reinforced these trends. Sezu et al. [183] reported that a 7% inclusion of FPH (Symrise Aqua-Feed, Specialities Pet Food, France) in Ompok pabda broodstock diets not only improved growth and feed conversion but also significantly enhanced reproductive performance—including fecundity, spawning response, and egg fertilisation—likely due to improved energy partitioning and endocrine support. Kabir et al. [184] showed that Nile tilapia fed hydrolysate-enriched (Symrise Aqua Feed, Specialities Pet Food, France) diets displayed superior liver and gut morphology, stronger immune cell populations, lower excretory waste, and enhanced resistance to Streptococcus iniae, translating into improved farm economics.
Despite promising outcomes, several research gaps remain. The specific peptide sequences responsible for the observed biological effects are often poorly characterised, limiting the precision of functional feed design. Furthermore, the interaction between FPHs and host microbiota is not fully understood—though evidence suggests FPHs can shape microbial communities in ways that enhance disease resistance and nutrient metabolism [180].
Another area requiring attention is species-specific response variability; for instance, FPHs that work well in carnivorous marine fish may not yield similar benefits in omnivorous freshwater species. Additionally, while short-term trials demonstrate clear advantages, long-term effects on fish health, reproductive performance, and fillet quality need further elucidation. There is also a need to develop standardised testing protocols and molecular biomarkers to assess the functional efficacy of FPHs in vivo. Finally, cost-effectiveness and scalability remain critical concerns for commercial uptake [178]. Advanced bioprocessing technologies, such as membrane separation and fermentation-aided hydrolysis, may help address these challenges by increasing peptide yield and functional consistency [179]. Overall, FPH-based feeds hold immense potential, but fully leveraging their benefits requires deeper mechanistic understanding, product optimisation, and integrative research approaches. Table 4 details fish hydrolysates used in aquaculture commercially, their claimed bioactivities and manufacturers.

6. FPHs as Biostimulants

Several groups throughout history including the Incas and Norwegians have used fish by-products as fertilisers. When fishery by-products are not suitable for use as fishmeal, they can be converted to biostimulants by applying enzymes, chemicals, or endogenous enzymes using hydrolysis methods/autolysis, as reviewed recently by [196]. Nowadays, the use of FPHs as biostimulants in Europe must comply with European Regulation 2021/1165 that prohibits the application of fish viscera, for example, on the edible part of plants produced using organic farm practices. Reports state that there are approximately 154 commercial fish fertilisers and biostimulants that exist as pellets, liquids, emulsions or powder products [197].
Biostimulants play a crucial role in modern day agricultural practices and can be defined as “any substance or microorganism, when applied to plants, that can improve nutrition efficiency, abiotic stress tolerance, and crop quality traits, regardless of its nutrient content” [198]. Protein hydrolysates are a primary category of biostimulants and can be made from plant or animal materials. The biostimulants effect of FPHs are due to the amino acids and peptides that can impart “hormone-like” activities to the plant. They can also enhance tolerance to abiotic stress and can stimulate antioxidant and osmoregulation in plants but a lot concerning the mode of action of FPH as biostimulants is still unknown. Figure 1 details the different modes of action of biostimulants and FPHs. Madende and colleagues also recently reviewed this topic [199].

6.1. Mechanisms of Action of FPH Biostimulants

The composition of FPHs impact the biostimulant impact on plants as well as where the FPH is applied (i.e., either soil or to leaves or roots). Results that follow application of biostimulants to plants are species and product specific [200]. Leaf permeability and penetration of the biostimulant are important factors that determine the efficacy of biostimulants. It is known, for example, that foliar application of FPHs increases amino acid and peptide availability by the plant as this form of application reduces competition with soil microbes for the amino acids and peptides present in biostimulants products.
Amino-acid-based biostimulants from FPHs have shown promising benefits and this market is likely to grow in the coming years. The synthesis of amino acids by plants requires a lot of energy consumption and application of amino-acids in FPHs via foliar application allows the plant to save energy by regulation of nitrogen in roots that increases the speed of production during transplantation or climate stress conditions like drought. Amino acids also bind to or chelate metal ions and use of micronutrients with amino acids accelerates absorption of micronutrients and their transport within the plant which results in increased growth. Nitrogen-containing compounds have a positive effect on photosynthetic activity in plants and proline and glutamic acid are known to play a positive role in this activity which can result in greater yields of fruits from plants. Glutamic acid is abundant in FPHs.

6.2. Modes of Action of FPH as Biostimulants

FPH contain a range of amino acids and of these, glutamic acid is a primary component. A myriad of studies have reported that a supply of isolated or combined amino acids is beneficial in the vegetative phase of different plants like beet, lettuce, and tomatoes [196]. Histidine and ornithine are normally absent from FPHs but they can also contain phytohormones, carbohydrates, and minerals. FPHs usually do not contain phenolics like seaweed or plant-based biostimulants. Peptides may have hormonal activities and can help to modulate biochemical processes in plants via biotic and abiotic stimulus [201]. Examples of peptides used as biostimulants include polypeptide families like the systemins and the peptide NOD40, which can promote root nodulation in plants including rice and soybean) [202]. The mechanism of action of FPHs is due to amino acids and peptides mimicking the action of natural peptides with hormone activities in the plants. Fat free amino acids like aspartic acid, hydroxyproline, threonine, serine, glutamic acid, proline, glycine, alanine, methionine, isoleucine, leucine, tyrosine, melatonin, organic matter, short-chain peptides, and proteins are considered the active, biostimulants agents in FPHs. These compounds can improve the utilisation of nutrients by the plants and induce morphological changes in root architecture [203]. Other bioactivities include anti-drought effects and stimulation of beneficial microbes as well as increased antioxidant benefits. These benefits improve the biology of the plant and improve root growth and development, flowering and improved fruit setting, and reduced fruit drop [203]. The biostimulant market in the EU is valued at USD 2.91 billion. Table 5 details commercial examples of biostimulants available currently, including those derived from animal protein sources.
The composition of FPHs impact the biostimulant impact on plants as well as where the FPH is applied (i.e., either soil or to leaves or roots). Results that follow application of biostimulants to plants are species and product specific [204]. Leaf permeability and penetration of the biostimulant are important factors that determine the efficacy of biostimulants. It is known, for example, that foliar application of FPHs increases amino acid and peptide availability by the plant as this form of application reduces competition with soil microbes for the amino acids and peptides present in biostimulants products.
Amino-acid-based biostimulants from FPHs have shown promising benefits and this market is likely to grow in the coming years. The synthesis of amino acids by plants requires a lot of energy consumption and application of amino-acids in FPHs via foliar application allows the plant to save energy by regulation of nitrogen in roots that increases the speed of production during transplantation or climate stress conditions like drought. Amino acids also bind to or chelate metal ions and use of micronutrients with amino acids accelerates absorption of micronutrients and their transport within the plant, which results in increased growth. Nitrogen containing compounds have a positive effect on photosynthetic activity in plants and proline and glutamic acid are known to play a positive role in this activity which can result in greater yields of fruits from plants. Glutamic acid is abundant in FPHs [205].

6.3. Product Development

Recently, Venslauskas and colleagues looked at the economic feasibility of using fish viscera as biostimulants [160]. This review suggests that producing biostimulants using a biorefinery approach where fish oil can also be recovered can make the process economically feasible. Key inhibiting factors in terms of costs include enzyme costs and energy consumption costs. Liquid production can reduce costs significantly compared to a dry product, but the shelf life of dry products is greater. Conversion of fish-products to hydrolysates for biostimulant use is also feasible using ensiling or autolysis with endogenous enzymes; however, reproducibility of the end product can vary significantly if the hydrolysis process is not carefully controlled with enzymes or acid. The production strategies of FPHs were recently described by [206].

6.4. Current Research Gaps

FPHs are used today as supplementary products to improve plant growth and survival and increase crop yields. However, there are still several research gaps that exist concerning their mechanism of action on plants and effectiveness. It is necessary to define the underlying genetic, molecular, and physiological mechanisms that contribute to observed field trial results. This would likely increase use of FPH biostimulant products in agriculture. Further, according to Ruzzi et al. figuring out the best combination of “biostimulant × plant species × environment interaction × agricultural practices” is essential to select optimal combinations for improved yields, quantity, and quality of products and optimising the application process in terms of time and rate are necessary to improve the impacts of FPHs as biostimulants [207].

7. Conclusions

The utilisation of fish protein hydrolysates (FPHs) derived from underutilised species and fish processing by-products represents a strategic intersection of sustainability, innovation, and market expansion. As highlighted throughout this review, FPHs are proving to be highly versatile bioactive ingredients with broad applications across human nutrition, companion animal wellness, aquaculture, and agriculture. Their functional value lies primarily in the presence of bioactive peptides and amino acids, which offer targeted health benefits, performance enhancement, and disease resistance across species and systems. In agriculture, their role as biostimulants aligns with the global movement towards reduced reliance on chemical fertilisers and pesticides, while in food and feed industries, they offer both nutritional and sensory improvements.
Advancements in hydrolysate production technologies, especially through enzymatic methods and data-driven approaches like in silico modelling, are enhancing both the efficiency and specificity of FPHs. These tools, once considered auxiliary, are now central to optimising yield, reducing costs, and tailoring bioactivity to meet specific market demands. Moreover, integrated analytical techniques such as plant phenomics and metabolomics are beginning to clarify the mechanisms through which FPHs exert their beneficial effects, especially in agricultural applications.
The growing recognition of FPHs as preventive and functional dietary components across human and animal health sectors points to their expanding commercial and therapeutic significance. As such, continued investment in research and refinement of production methodologies will be key to unlocking their full potential. However, to fully realise these benefits, further work is needed to standardise production, enhance bioactivity targeting, and navigate regulatory and consumer acceptance hurdles—particularly in regions with stringent guidelines such as the European Union.
Overall, the findings of this review underscore the potential of FPHs for sustainable food production, both crop and animal based, as well as promotion of animal and human health, and collates the potential market value that can be garnered from the described scientific studies related to these benefits.

Author Contributions

Conceptualization, M.H. and D.B.; methodology, M.H. and D.B.; software, M.H. and D.B.; validation, M.H. and D.B.; formal analysis, M.H.; investigation, D.B.; resources, M.H.; data curation, M.H.; writing—original draft preparation, M.H. and D.B.; writing—review and editing, M.H. and D.B.; visualisation, M.H.; supervision, M.H.; project administration, M.H.; funding acquisition, M.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by The European Maritime Fisheries Fund (EMFF) and Bord Iascaigh Mhara (BIM), under Grant number reference 2023-018, 2023.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets generated during this study are available from the corresponding author upon reasonable request.

Acknowledgments

We acknowledge financial support from Michael Gallagher, Bord Iascaigh Mhara (BIM) and the European Maritime Fisheries Fund (EMFF), 2023–2025.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mora, L.; Toldrá, F. Advanced enzymatic hydrolysis of food proteins for the production of bioactive peptides. Curr. Opin. Food Sci. 2023, 49, 100973. [Google Scholar] [CrossRef]
  2. de Carvalho Oliveira, L.; Martinez-Villaluenga, C.; Frias, J.; Cartea, M.E.; Francisco, M.; Cristianini, M.; Peñas, E. High pressure-assisted enzymatic hydrolysis potentiates the production of quinoa protein hydrolysates with antioxidant and ACE-inhibitory activities. Food Chem. 2024, 447, 138887. [Google Scholar] [CrossRef] [PubMed]
  3. Fadimu, G.J.; Le, T.T.; Gill, H.; Farahnaky, A.; Olatunde, O.O.; Truong, T. Enhancing the Biological Activities of Food Protein-Derived Peptides Using Non-Thermal Technologies: A Review. Foods 2022, 11, 1823. [Google Scholar] [CrossRef] [PubMed]
  4. Suryaningtyas, I.T.; Je, J.Y. Bioactive peptides from food proteins as potential anti-obesity agents: Mechanisms of action and future perspectives. Trends Food Sci. Technol. 2023, 138, 141–152. [Google Scholar] [CrossRef]
  5. Zaky, A.A.; Simal-Gandara, J.; Eun, J.B.; Shim, J.H.; Abd El-Aty, A.M. Bioactivities, Applications, Safety, and Health Benefits of Bioactive Peptides from Food and By-Products: A Review. Front. Nutr. 2022, 8, 815640. [Google Scholar] [CrossRef]
  6. Du, Z.; Li, Y. Review and perspective on bioactive peptides: A roadmap for research, development, and future opportunities. J. Agric. Food Res. 2022, 9, 100353. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Zhang, H.; Ghosh, D.; Williams, R.O., 3rd. Just how prevalent are peptide therapeutic products? A critical review. Int. J. Pharm. 2020, 587, 119491. [Google Scholar] [CrossRef]
  8. Ribeiro, T.B.; Maia, M.R.G.; Fonseca, A.J.M.; Marques, B.; Caleja, C.; Rosa, A.; Martins, R.; Almeida, A.; Mota, M.J.; Aires, T.; et al. A Comprehensive Review of Fish Protein Hydrolysates Targeting Pet Food Formulations. Food Rev. Int. 2024, 1–39. [Google Scholar] [CrossRef]
  9. Hayes, M. Maximizing Use of Pelagic Capture Fisheries for Global Protein Supply: Potential and Caveats Associated with Fish and Co-Product Conversion into Value-Add Ingredients. Glob. Chall. 2023, 7, 2200098. [Google Scholar] [CrossRef]
  10. Kurnianto, M.A.; Defri, I.; Syahbanu, F.; Aulia, S.S. Fish-derived bioactive peptide: Bioactivity potency, structural characteristics, and conventional and bioinformatics approaches for identification. Future Foods 2024, 9, 100386. [Google Scholar] [CrossRef]
  11. Hasler, C.M. Functional foods: Benefits, concerns and challenges—A position paper from the American Council on Science and Health. J. Nutr. 2002, 132, 3772–3781. [Google Scholar] [CrossRef] [PubMed]
  12. Venkatakrishnan, K.; Chiu, H.F.; Wang, C.K. Impact of functional foods and nutraceuticals on high blood pressure with a special focus on meta-analysis: Review from a public health perspective. Food Funct. 2020, 11, 2792–2804. [Google Scholar] [CrossRef] [PubMed]
  13. Lu, C.C.; Yen, G.C. Antioxidative and anti-inflammatory activity of functional foods. Curr. Opin. Food Sci. 2015, 2, 1–8. [Google Scholar] [CrossRef]
  14. Faienza, M.F.; Giardinelli, S.; Annicchiarico, A.; Chiarito, M.; Barile, B.; Corbo, F.; Brunetti, G. Nutraceuticals and Functional Foods: A Comprehensive Review of Their Role in Bone Health. Int. J. Mol. Sci. 2024, 25, 5873. [Google Scholar] [CrossRef]
  15. Mohamed, S. Functional foods against metabolic syndrome (obesity, diabetes, hypertension and dyslipidemia) and cardiovascular disease. Trends Food Sci. Technol. 2014, 35, 114–128. [Google Scholar] [CrossRef]
  16. Vignesh, A.; Amal, T.C.; Sarvalingam, A.; Vasanth, K. A review on the influence of nutraceuticals and functional foods on health. Food Chem. Adv. 2024, 5, 100749. [Google Scholar] [CrossRef]
  17. Henriques, A.; Vázquez, J.A.; Valcarcel, J.; Mendes, R.; Bandarra, N.M.; Pires, C. Characterization of Protein Hydrolysates from Fish Discards and By-Products from the North-West Spain Fishing Fleet as Potential Sources of Bioactive Peptides. Mar. Drugs 2021, 19, 338. [Google Scholar] [CrossRef]
  18. Aziz, T.; Hussain, N.; Hameed, Z.; Lin, L. Elucidating the role of diet in maintaining gut health to reduce the risk of obesity, cardiovascular and other age-related inflammatory diseases: Recent challenges and future recommendations. Gut Microbes 2024, 16, 2297864. [Google Scholar] [CrossRef]
  19. Whitaker, R.D.; Altintzoglou, T.; Lian, K.; Fernandez, E.N. Marine bioactive peptides in supplements and functional foods—A commercial perspective. Curr. Pharm. Des. 2021, 27, 1353–1364. [Google Scholar] [CrossRef]
  20. Gevaert, B.; Veryser, L.; Verbeke, F.; Wynendaele, E.; De Spiegeleer, B. Fish hydrolysates: A regulatory perspective of bioactive peptides. Protein Pept. Lett. 2016, 23, 1052–1060. [Google Scholar] [CrossRef]
  21. Rivero-Pino, F.; Villanueva, Á.; Montserrat-de-la-Paz, S.; Sanchez-Fidalgo, S.; Millán-Linares, M.C. Evidence of Immunomodulatory Food-Protein Derived Peptides in Human Nutritional Interventions: Review on the Outcomes and Potential Limitations. Nutrients 2023, 15, 2681. [Google Scholar] [CrossRef] [PubMed]
  22. Arias, L.; Sánchez-Henao, C.P.; Zapata, J.E. Optimization of the separation conditions of antioxidant peptides from red tilapia (Oreochromis spp.) viscera on cross-flow filtration ceramic membranes. S. Afr. J. Chem. Eng. 2023, 45, 100–110. [Google Scholar] [CrossRef]
  23. La Manna, S.; Di Natale, C.; Florio, D.; Marasco, D. Peptides as therapeutic agents for inflammatory-related diseases. Int. J. Mol. Sci. 2018, 19, 2714. [Google Scholar] [CrossRef] [PubMed]
  24. Kemp, D.C.; Kwon, J.Y. Fish and Shellfish-Derived Anti-Inflammatory Protein Products: Properties and Mechanisms. Molecules 2021, 26, 3225. [Google Scholar] [CrossRef] [PubMed]
  25. Lessa, R.C.; Ebrahimi, B.; Li, H.; Guan, X.; Li, Y.; Lu, J. Investigation of the In Vitro Immunomodulatory Effects of Extracts from Green-Lipped Mussels (Perna canaliculus). Nutraceuticals 2024, 4, 127–146. [Google Scholar] [CrossRef]
  26. Phadke, G.G.; Rathod, N.B.; Ozogul, F.; Elavarasan, K.; Karthikeyan, M.; Shin, K.H.; Kim, S.K. Exploiting of Secondary Raw Materials from Fish. Processing Industry as a Source of Bioactive Peptide-Rich Protein Hydrolysates. Mar. Drugs 2021, 19, 480. [Google Scholar] [CrossRef]
  27. Hayes, M.; Gargan, C.; Milovanovic, I. Antihypertensive Food Ingredients for Companion Animal Applications. US Patent Application 18/250/822, 5 May 2022. [Google Scholar]
  28. Kawasaki, T.; Seki, E.; Osajima, K.; Yoshida, M.; Asada, K.; Matsui, T.; Osajima, Y. Antihypertensive effect of valyl-tyrosine, a short chain peptide derived from sardine muscle hydrolyzate, on mild hypertensive subjects. J. Hum. Hypertens. 2000, 14, 519–523. [Google Scholar] [CrossRef]
  29. Zhang, L.; Zhao, G.X.; Zhao, Y.Q.; Qiu, Y.T.; Chi, C.F.; Wang, B. Identification and Active Evaluation of Antioxidant Peptides from Protein Hydrolysates of Skipjack Tuna (Katsuwonus pelamis) Head. Antioxidants 2019, 8, 318. [Google Scholar] [CrossRef]
  30. Abdelhedi, O.; Nasri, M. Basic and recent advances in marine antihypertensive peptides: Production, structure-activity relationship and bioavailability. Trends Food Sci. Technol. 2019, 88, 543–557. [Google Scholar] [CrossRef]
  31. Wang, X.; Wu, S.; Xu, D.; Xie, D.; Guo, H. Inhibitor and substrate binding by angiotensin-converting enzyme: Quantum mechanical/molecular mechanical molecular dynamics studies. J. Chem. Inf. Model. 2011, 51, 1074–1082. [Google Scholar] [CrossRef]
  32. Naik, A.S.; Mora, L.; Hayes, M. Characterisation of Seasonal Mytilus edulis By-Products and Generation of Bioactive Hydrolysates. Appl. Sci. 2020, 10, 6892. [Google Scholar] [CrossRef]
  33. Girgih, A.T.; Nwachukwu, I.D.; Hasan, F.; Fagbemi, T.N.; Gill, T.; Aluko, R.E. Kinetics of the inhibition of renin and angiotensin I-converting enzyme by cod (Gadus morhua) protein hydrolysates and their antihypertensive effects in spontaneously hypertensive rats. Food Nutr. Res. 2015, 59, 29788. [Google Scholar] [CrossRef] [PubMed]
  34. Power, O.; Nongonierma, A.B.; Jakeman, P.; FitzGerald, R.J. Food protein hydrolysates as a source of dipeptidyl peptidase IV inhibitory peptides for the management of type 2 diabetes. Proc. Nutr. Soc. 2014, 73, 34–46. [Google Scholar] [CrossRef] [PubMed]
  35. Zhou, X.; Chai, L.; Wu, Q.; Wang, Y.; Li, S.; Chen, J. Anti-diabetic properties of bioactive components from fish and milk. J. Funct. Foods 2021, 85, 104669. [Google Scholar] [CrossRef]
  36. Cudennec, B.; Fouchereau-Peron, M.; Ferry, F.; Duclos, E.; Ravallec, R. In vitro and in vivo evidence for a satiating effect of fish protein hydrolysate obtained from blue whiting (Micromesistius poutassou) muscle. J. Funct. Foods 2012, 4, 271–277. [Google Scholar] [CrossRef]
  37. Lees, M.J.; Carson, B.P. The potential role of fish-derived protein hydrolysates on metabolic health, skeletal muscle mass and function in ageing. Nutrients 2020, 12, 2434. [Google Scholar] [CrossRef]
  38. Pilon, G.; Ruzzin, J.; Rioux, L.E.; Lavigne, C.; White, P.J.; Frøyland, L.; Jacques, H.; Bryl, P.; Beaulieu, L.; Marette, A. Differential effects of various fish proteins in altering body weight, adiposity, inflammatory status, and insulin sensitivity in high-fat–fed rats. Metabolism 2011, 60, 1122–1130. [Google Scholar] [CrossRef]
  39. Li-Chan, E.C.; Hunag, S.L.; Jao, C.L.; Ho, K.P.; Hsu, K.C. Peptides derived from Atlantic salmon skin gelatin as dipeptidyl-peptidase IV inhibitors. J. Agric. Food Chem. 2012, 60, 973–978. [Google Scholar] [CrossRef]
  40. Rivero-Pino, F.; Espejo-Carpio, F.J.; Guadix, E.M. Production and identification of dipeptidyl peptidase IV (DPP-IV) inhibitory peptides from discarded Sardine pilchardus protein. Food Chem. 2020, 328, 127096. [Google Scholar] [CrossRef]
  41. Liu, D.; Ren, Y.; Zhong, S.; Xu, B. New Insight into Utilization of Fish By-Product Proteins and Their Skin Health Promoting Effects. Mar. Drugs 2024, 22, 215. [Google Scholar] [CrossRef]
  42. Zhang, X.; Zhuang, H.; Wu, S.; Mao, C.; Dai, Y.; Yan, H. Marine Bioactive Peptides: Anti-Photoaging Mechanisms and Potential Skin Protective Effects. Curr. Issues Mol. Biol. 2024, 46, 990–1009. [Google Scholar] [CrossRef] [PubMed]
  43. He, X.; Wan, F.; Su, W.; Xie, W. Research Progress on Skin Aging and Active Ingredients. Molecules 2023, 28, 5556. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, T.; Hou, H.; Fan, Y.; Wang, S.; Chen, Q.; Si, L.; Li, B. Protective effect of gelatin peptides from pacific cod skin against photoaging by inhibiting the expression of MMPs via MAPK signaling pathway. J. Photochem. Photobiol. 2016, 165, 34–41. [Google Scholar] [CrossRef]
  45. Li, H.-L.; Li, M.-J.; Xiong, G.-Q.; Cai, J.; Liao, T.; Zu, X.-Y. Silver Carp (Hypophthalmichthys molitrix) Scale Collagen Peptides-1 (SCPs1) Inhibit Melanogenesis through Downregulation of the cAMP-CREB Signaling Pathway. Nutrients 2023, 15, 2449. [Google Scholar] [CrossRef]
  46. Iosageanu, A.; Ilie, D.; Craciunescu, O.; Seciu-Grama, A.-M.; Oancea, A.; Zarnescu, O.; Moraru, I.; Oancea, F. Effect of Fish Bone Bioactive Peptides on Oxidative, Inflammatory and Pigmentation Processes Triggered by UVB Irradiation in Skin Cells. Molecules 2021, 26, 2691. [Google Scholar] [CrossRef]
  47. Kong, J.; Hu, X.-M.; Cai, W.-W.; Wang, Y.-M.; Chi, C.-F.; Wang, B. Bioactive Peptides from Skipjack Tuna Cardiac Arterial Bulbs (II): Protective Function on UVB-Irradiated HaCaT Cells through Antioxidant and Anti-Apoptotic Mechanisms. Mar. Drugs 2023, 21, 105. [Google Scholar] [CrossRef]
  48. Azizi, M.; Sharifan, A.; Hoseini, E.; Ghavami, A. Enzymatic Hydrolysis of Whitefish Viscera of Caspian Sea (Caspian kutum) and Evaluation of Antioxidant Properties of hydrolyzed protein. Iran. J. Biosyst. Eng. 2021, 52, 379–390. [Google Scholar] [CrossRef]
  49. Shahidi, F.; Saeid, A. Bioactivity of Marine-Derived Peptides and Proteins: A Review. Mar. Drugs 2025, 23, 157. [Google Scholar] [CrossRef]
  50. Halim, N.R.; Yusof, H.M.; Sarbon, N.M. Functional and bioactive properties of fish protein hydolysates and peptides: A comprehensive review. Trends Food Sci. Technol. 2016, 51, 24–33. [Google Scholar] [CrossRef]
  51. Hasani, K.; Ariaii, P.; Ahmadi, M. Antimicrobial, antioxidant and anti-cancer properties of protein hydrolysates from indian mackerel (Rastrelliger kanagurta) waste prepared using commercial enzyme. Int. J. Pept. Res. Ther. 2022, 28, 86. [Google Scholar] [CrossRef]
  52. Pezeshk, S.; Ojagh, S.M.; Rezaei, M.; Shabanpour, B. Fractionation of Protein Hydrolysates of Fish Waste Using Membrane Ultrafiltration: Investigation of Antibacterial and Antioxidant Activities. Probiotics Antimicrob. Proteins 2019, 11, 1015–1022. [Google Scholar] [CrossRef] [PubMed]
  53. Tejpal, C.S.; Vijayagopal, P.; Elavarasan, K.; Linga Prabu, D.; Lekshmi, R.G.K.; Asha, K.K.; Anandan, R.; Chatterjee, S.; Mathew, S. Antioxidant, functional properties and amino acid composition of pepsin-derived protein hydrolysates from whole tilapia waste as influenced by pre-processing ice storage. J. Food Sci. Technol. 2017, 54, 4257–4267. [Google Scholar] [CrossRef] [PubMed]
  54. Hamzeh, A.; Rezaei, M.; Khodabandeh, S.; Motamedzadegan, A.; Noruzinia, M.; Regenstein, J.M. Optimization of Antioxidant Peptides Production from the Mantle of Cuttlefish (Sepia pharaonis) Using RSM and Fractionation. J. Aquat. Food Product. Technol. 2019, 28, 392–401. [Google Scholar] [CrossRef]
  55. Tadesse, S.A.; Emire, S.A. Production and processing of antioxidant bioactive peptides: A driving force for the functional food market. Heliyon 2020, 6, e04765. [Google Scholar] [CrossRef] [PubMed]
  56. Daniele, F.; Marasco, D.; La Manna, S. Approaches for developing peptide-and metal complexes-or chelators-based leads for anti-amyloid drugs. Inorganica Chim. Acta 2024, 577, 122474. [Google Scholar] [CrossRef]
  57. Liao, X.; Zhu, Z.; Wu, S.; Chen, M.; Huang, R.; Wang, J.; Wu, Q.; Ding, Y. Preparation of Antioxidant Protein Hydrolysates from Pleurotus geesteranus and Their Protective Effects on H2O2 Oxidative Damaged PC12 Cells. Molecules 2020, 25, 5408. [Google Scholar] [CrossRef]
  58. Sohaib, M.; Anjum, F.M.; Sahar, A.; Arshad, M.S.; Rahman, U.U.; Imran, A.; Hussain, S. Antioxidant proteins and peptides to enhance the oxidative stability of meat and meat products: A comprehensive review. Int. J. Food Prop. 2017, 20, 2581–2593. [Google Scholar] [CrossRef]
  59. Brogden, N.K.; Mehalick, L.; Fischer, C.L.; Wertz, P.W.; Brogden, K.A. The emerging role of peptides and lipids as antimicrobial epidermal barriers and modulators of local inflammation. Skin Pharmacol. Physiol. 2012, 25, 167–181. [Google Scholar] [CrossRef]
  60. Bi, J.; Tian, C.; Jiang, J.; Zhang, G.L.; Hao, H.; Hou, H.M. Antibacterial Activity and Potential Application in Food Packaging of Peptides Derived from Turbot Viscera Hydrolysate. J. Agric. Food Chem. 2020, 68, 9968–9977. [Google Scholar] [CrossRef]
  61. Perez Espitia, P.J.; de Fátima Ferreira Soares, N.; Dos Reis Coimbra, J.S.; de Andrade, N.J.; Souza Cruz, R.; Alves Medeiros, E.A. Bioactive Peptides: Synthesis, Properties, and Applications in the Packaging and Preservation of Food. Compr. Rev. Food Sci. Food Saf. 2012, 11, 187–204. [Google Scholar] [CrossRef]
  62. Da Rocha, M.; Alemán, A.; Romani, V.P.; López-Caballero, M.; Gómez-Guillén, M.; Montero, P.; Prentice, C. Effects of agar films incorporated with fish protein hydrolysate or clove essential oil on flounder (Paralichthys orbignyanus) fillets shelf-life. Food Hydrocoll. 2018, 81, 351–363. [Google Scholar] [CrossRef]
  63. Chen, M.F.; Gong, F.; Zhang, Y.Y.; Li, C.; Zhou, C.; Hong, P.; Sun, S.; Qian, Z.J. Preventive Effect of YGDEY from Tilapia Fish Skin Gelatin Hydrolysates against Alcohol-Induced Damage in HepG2 Cells through ROS-Mediated Signaling Pathways. Nutrients 2019, 11, 392. [Google Scholar] [CrossRef] [PubMed]
  64. Choi, F.D.; Sung, C.T.; Juhasz, M.L.; Mesinkovsk, N.A. Oral Collagen Supplementation: A Systematic Review of Dermatological Applications. J. Drugs Dermatol. 2019, 18, 9–16. [Google Scholar]
  65. Jahandideh, F.; Bourque, S.L.; Wu, J. A comprehensive review on the glucoregulatory properties of food-derived bioactive peptides. Food Chem. 2022, 13, 100222. [Google Scholar] [CrossRef]
  66. Borges, S.; Odila, J.; Voss, G.; Martins, R.; Rosa, A.; Couto, J.A.; Almeida, A.; Pintado, M. Fish By-Products: A Source of Enzymes to Generate Circular Bioactive Hydrolysates. Molecules 2023, 28, 1155. [Google Scholar] [CrossRef]
  67. Jensen, C.; Fjeldheim Dale, H.; Hausken, T.; Hatlebakk, J.G.; Brønstad, I.; Lied, G.A.; Hoff, D.A.L. Supplementation with Low Doses of a Cod Protein Hydrolysate on Glucose Regulation and Lipid Metabolism in Adults with Metabolic Syndrome: A Randomized, Double-Blind Study. Nutrients 2020, 12, 1991. [Google Scholar] [CrossRef] [PubMed]
  68. Guénard, F.; Jacques, H.; Gagnon, C.; Marette, A.; Vohl, M.C. Acute Effects of Single Doses of Bonito Fish Peptides and Vitamin D on Whole Blood Gene Expression Levels: A Randomized Controlled Trial. Int. J. Mol. Sci. 2019, 20, 1944. [Google Scholar] [CrossRef]
  69. Le Poncin, M.; Lamproglou, I. Experimental study: Stress and memory. Eur. Neuropsychopharmacol. 1996, 6, 110. [Google Scholar] [CrossRef]
  70. Nesse, K.O.; Nagalakshmi, A.P.; Marimuthu, P.; Singh, M. Efficacy of a fish protein hydrolysate in malnourished children. Indian J. Clin. Biochem. 2011, 26, 360–365. [Google Scholar] [CrossRef]
  71. Heffernan, S.; Nunn, L.; Harnedy-Rothwell, P.A.; Gite, S.; Whooley, J.; Giblin, L.; FitzGerald, R.J.; O’Brien, N.M. Blue Whiting (Micromesistius poutassou) Protein Hydrolysates Increase GLP-1 Secretion and Proglucagon Production in STC-1 Cells Whilst Maintaining Caco-2/HT29-MTX Co-Culture Integrity. Mar. Drugs 2022, 20, 112. [Google Scholar] [CrossRef]
  72. Harnedy, P.A.; Parthsarathy, V.; McLaughlin, C.M.; O’Keeffe, M.B.; Allsopp, P.J.; McSorley, E.M.; O’Harte, F.P.M.; FitzGerald, R.J. Blue whiting (Micromesistius poutassou) muscle protein hydrolysate with in vitro and in vivo antidiabetic properties. J. Funct. Foods 2018, 40, 137–145. [Google Scholar] [CrossRef]
  73. Available online: https://www.futuremarketinsights.com/reports/fish-based-pet-food-market (accessed on 3 September 2024).
  74. Zhang, W.; Cao, H.; Lin, L. Analysis of the future development trend of the pet industry. In Proceedings of the 2022 7th International Conference on Financial Innovation and Economic Development (ICFIED 2022), Harbin, China, 21–23 January 2022; Atlantis Press: Zhengzhou, China, 2022; pp. 1682–1689. [Google Scholar] [CrossRef]
  75. Di Cerbo, A.; Morales-Medina, J.C.; Palmieri, B.; Pezzuto, F.; Cocco, R.; Flores, G.; Iannitti, T. Functional foods in pet nutrition: Focus on dogs and cats. Res. Vet. Sci. 2017, 112, 161–166. [Google Scholar] [CrossRef]
  76. Swanson, K.S.; Carter, R.A.; Yount, T.P.; Aretz, J.; Buff, P.R. Nutritional sustainability of pet foods. Adv. Nutr. 2013, 4, 141–150. [Google Scholar] [CrossRef]
  77. Gasco, L.; Acuti, G.; Bani, P.; Dalle Zotte, A.; Danieli, P.P.; De Angelis, A.; Fortina, R.; Marino, R.; Parisi, G.; Piccolo, G.; et al. Insect and fish by-products as sustainable alternatives to conventional animal proteins in animal nutrition. Ital. J. Anim. Sci. 2020, 19, 360–372. [Google Scholar] [CrossRef]
  78. Meeker, D.L.; Meisinger, J.L. Companion Animals Symposium: Rendered ingredients significantly influence sustainability, quality, and safety of pet food. J. Anim. Sci. 2015, 93, 835–847. [Google Scholar] [CrossRef]
  79. Ashraf, S.A.; Adnan, M.; Patel, M.; Siddiqui, A.J.; Sachidanandan, M.; Snoussi, M.; Hadi, S. Fish-Based Bioactives as Potent Nutraceuticals: Exploring the Therapeutic Perspective of Sustainable Food from the Sea. Mar. Drugs 2020, 18, 265. [Google Scholar] [CrossRef] [PubMed]
  80. Naik, A.S.; Whitaker, R.D.; Albrektsen, S.; Solstad, R.G.; Thoresen, L.; Hayes, M. Mesopelagic fish protein hydrolysates and extracts: A source of novel anti-hypertensive and anti-diabetic peptides. Front. Mar. Sci. 2021, 8, 719608. [Google Scholar] [CrossRef]
  81. Geirsdottir, M.; Sigurgisladottir, S.; Hamaguchi, P.Y.; Thorkelsson, G.; Johannsson, R.; Kristinsson, H.G.; Kristjansson, M.M. Enzymatic hydrolysis of blue whiting (Micromesistius poutassou); functional and bioactive properties. J. Food Sci. 2011, 76, C14–C20. [Google Scholar] [CrossRef]
  82. de Oliveira Coelho, L.; Pinto, R. Amino Acids as Key Mediators of Immune Status and Nutritional Condition in Fish; University of Porto: Porto, Portugal, 2020. [Google Scholar]
  83. Cave, N.J. Hydrolyzed protein diets for dogs and cats. Vet. Clin. Small Anim. Pract. 2006, 36, 1251–1268. [Google Scholar] [CrossRef]
  84. Noli, C.; Varina, A.; Barbieri, C.; Pirola, A.; Olivero, D. Analysis of Intestinal Microbiota and Metabolic Pathways before and after a 2-Month-Long Hydrolyzed Fish and Rice Starch Hypoallergenic Diet Trial in Pruritic Dogs. Vet. Sci. 2023, 10, 478. [Google Scholar] [CrossRef]
  85. Martínez-Alvarez, O.; Chamorro, S.; Brenes, A. Protein hydrolysates from animal processing by-products as a source of bioactive molecules with interest in animal feeding: A review. Food Res. Int. 2015, 73, 204–212. [Google Scholar] [CrossRef]
  86. Baco, N.; Oslan, S.N.H.; Shapawi, R.; Mohhtar, R.A.M.; Noordin, W.N.M.; Huda, N. Antibacterial Activity of Functional Bioactive Peptides Derived from Fish Protein Hydrolysate; IOP Conference Series: Earth and Environmental Science; IOP Publishing: Bristol, UK, 2022; Volume 967, p. 012019. [Google Scholar] [CrossRef]
  87. Da Rocha, M.; Alemán, A.; Baccan, G.C.; López-Caballero, M.E.; Gómez-Guillén, C.; Montero, P.; Prentice, C. Anti-inflammatory, antioxidant, and antimicrobial effects of underutilized fish protein hydrolysate. J. Aquat. Food Prod. Technol. 2018, 27, 592–608. [Google Scholar] [CrossRef]
  88. Hatab, S.; Chen, M.-L.; Miao, W.; Lin, J.; Wu, D.; Wang, C.; Yuan, P.; Deng, S. Protease hydrolysates of filefish (Thamnaconus modestus) byproducts effectively inhibit foodborne pathogens. Foodborne Pathog. Dis. 2017, 14, 656–666. [Google Scholar] [CrossRef]
  89. Trang, H.T.H.; Pasuwan, P. Screening antimicrobial activity against pathogens from protein hydrolysate of rice bran and Nile Tilapia by-products. Int. Food Res. J. 2018, 25, 2157–2163. [Google Scholar]
  90. Lima, K.O.; da Costa de Quadros, C.; Rocha, M.D.; Jocelino Gomes de Lacerda, J.T.; Juliano, M.A.; Dias, M.; Mendes, M.A.; Prentice, C. Bioactivity and bioaccessibility of protein hydrolyzates from industrial byproducts of Stripped weakfish (Cynoscion guatucupa). LWT-Food Sci. Technol. 2019, 111, 408–413. [Google Scholar] [CrossRef]
  91. Hou, Y.; Wu, Z.; Dai, Z.; Wang, G.; Wu, G. Protein hydrolysates in animal nutrition: Industrial production, bioactive peptides, and functional significance. J. Anim. Sci. Biotechnol. 2022, 8, 209–232. [Google Scholar] [CrossRef] [PubMed]
  92. Chen, Y.P.; Liang, C.H.; Wu, H.T.; Pang, H.Y.; Chen, C.; Wang, G.H.; Chan, L.P. Antioxidant and anti-inflammatory capacities of collagen peptides from milkfish (Chanos chanos) scales. J. Food Sci. Technol. 2018, 55, 2310–2317. [Google Scholar] [CrossRef]
  93. Giannetto, A.; Esposito, E.; Lanza, M.; Oliva, S.; Riolo, K.; Di Pietro, S.; Abbate, J.M.; Briguglio, G.; Cassata, G.; Cicero, L.; et al. Protein Hydrolysates from Anchovy (Engraulis encrasicolus) Waste: In Vitro and In Vivo Biological Activities. Mar. Drugs 2020, 18, 86. [Google Scholar] [CrossRef]
  94. Narayanasamy, A.; Balde, A.; Raghavender, P.; Shashanth, D.; Abraham, J.; Joshi, I.; Nazeer, R.A. Isolation of marine crab (Charybdis natator) leg muscle peptide and its anti-inflammatory effects on macrophage cells. Biocatal. Agric. Biotechnol. 2020, 25, 101577. [Google Scholar] [CrossRef]
  95. Zhu, W.; Ren, L.; Zhang, L.; Qiao, Q.; Farooq, M.Z.; Xu, Q. The Potential of Food Protein-Derived Bioactive Peptides against Chronic Intestinal Inflammation. Mediat. Inflamm. 2020, 2020, 6817156. [Google Scholar] [CrossRef]
  96. Zamora-Sillero, J.; Gharsallaoui, A.; Prentice, C. Peptides from Fish By-product Protein Hydrolysates and Its Functional Properties: An Overview. Mar. Biotechnol. 2018, 20, 118–130. [Google Scholar] [CrossRef] [PubMed]
  97. Larder, C.E.; Iskandar, M.M.; Kubow, S. Collagen Hydrolysates: A Source of Bioactive Peptides Derived from Food Sources for the Treatment of Osteoarthritis. Medicines 2023, 10, 50. [Google Scholar] [CrossRef] [PubMed]
  98. Gao, R.; Yu, Q.; Shen, Y.; Chu, Q.; Chen, G.; Fen, S.; Yang, M.; Yuan, L.; Mcclements, D.J.; Sun, Q. Production, bioactive properties, and potential applications of fish protein hydrolysates: Developments and challenges. Trends Food Sci. Technol. 2021, 110, 687–699. [Google Scholar] [CrossRef]
  99. Cabrita, A.R.J.; Maia, M.R.G.; Alves, A.P.; Aires, T.; Rosa, A.; Almeida, A.; Martins, R.; Fonseca, A.J.M. Protein hydrolysate and oil from fish waste reveal potential as dog food ingredients. Front. Vet. Sci. 2024, 11, 1372023. [Google Scholar] [CrossRef]
  100. Ortizo, R.G.G.; Sharma, V.; Tsai, M.L.; Wang, J.X.; Sun, P.P.; Nargotra, P.; Kuo, C.H.; Chen, C.W.; Dong, C.D. Extraction of Novel Bioactive Peptides from Fish Protein Hydrolysates by Enzymatic Reactions. Appl. Sci. 2023, 13, 5768. [Google Scholar] [CrossRef]
  101. Hayes, M.; Naik, A.; Mora, L.; Iñarra, B.; Ibarruri, J.; Bald, C.; Cariou, T.; Reid, D.; Gallagher, M.; Dragøy, R.; et al. Generation, Characterisation and Identification of Bioactive Peptides from Mesopelagic Fish Protein Hydrolysates Using in Silico and in Vitro Approaches. Mar. Drugs 2024, 22, 297. [Google Scholar] [CrossRef] [PubMed]
  102. Yathisha, U.G.; Bhat, I.; Karunasagar, I.; Mamatha, B.S. Antihypertensive activity of fish protein hydrolysates and its peptides. Crit. Rev. Food Sci. Nutr. 2019, 59, 2363–2374. [Google Scholar] [CrossRef]
  103. Kasran, M.; Mohamed, N.; Ishak, Z. Extraction of protein and antioxidant and antihypertensive properties of protein hydrolysates derived from black tilapia (Oreochromis placidus). J. Trop. Agric. Food Sci. 2023, 51, 51–60. [Google Scholar]
  104. Kubinyi, E.; Bel Rhali, S.; Sándor, S.; Szabó, A.; Felföldi, T. Gut Microbiome Composition is Associated with Age and Memory Performance in Pet Dogs. Animals 2020, 10, 1488. [Google Scholar] [CrossRef]
  105. Chalamaiah, M.; Hemalatha, R.; Jyothirmayi, T. Fish protein hydrolysates: Proximate composition, amino acid composition, antioxidant activities and applications: A review. Food Chem. 2012, 135, 3020–3038. [Google Scholar] [CrossRef]
  106. Nirmal, N.P.; Santivarangkna, C.; Benjakul, S.; Maqsood, S. Fish protein hydrolysates as a health-promoting ingredient-recent update. Nutr. Rev. 2022, 80, 1013–1026. [Google Scholar] [CrossRef] [PubMed]
  107. Inoue, M.; Hasegawa, A.; Hosoi, Y.; Sugiura, K. Association between breed, gender and age in relation to cardiovascular disorders in insured dogs in Japan. J. Vet. Med. Sci. 2016, 78, 347–350. [Google Scholar] [CrossRef] [PubMed]
  108. Svensson, M.; Selling, J.; Dirven, M. Myxomatous Mitral Valve Disease in Large Breed Dogs: Survival Characteristics and Prognostic Variables. Vet. Sci. 2024, 11, 136. [Google Scholar] [CrossRef]
  109. Druzhaeva, N.; Nemec Svete, A.; Tavčar-Kalcher, G.; Babič, J.; Ihan, A.; Pohar, K.; Krapež, U.; Domanjko Petrič, A. Effects of Coenzyme Q10 Supplementation on Oxidative Stress Markers, Inflammatory Markers, Lymphocyte Subpopulations, and Clinical Status in Dogs with Myxomatous Mitral Valve Disease. Antioxidants 2022, 11, 1427. [Google Scholar] [CrossRef]
  110. Larsen, J.A.; Fascetti, A.J. The role of taurine in cardiac health in dogs and cats. Adv. Small Anim. Care 2020, 1, 227–238. [Google Scholar] [CrossRef]
  111. Lordan, S.; Ross, R.P.; Stanton, C. Marine bioactives as functional food ingredients: Potential to reduce the incidence of chronic diseases. Mar. Drugs 2011, 9, 1056–1100. [Google Scholar] [CrossRef] [PubMed]
  112. Pasławski, R.; Kurosad, A.; Ząbek, A.; Pasławska, U.; Noszczyk-Nowak, A.; Michałek, M.; Młynarz, P. Effect of 6-Month Feeding with a Diet Enriched in EPA + DHA from Fish Meat on the Blood Metabolomic Profile of Dogs with Myxomatous Mitral Valve Disease. Animals 2021, 11, 3360. [Google Scholar] [CrossRef]
  113. Liu, H.; Yang, Y.; Liu, Y.; Cui, L.; Fu, L.; Li, B. Various bioactive peptides in collagen hydrolysate from salmo salar skin and the combined inhibitory effects on atherosclerosis in vitro and in vivo. Food Res. Int. 2022, 157, 111281. [Google Scholar] [CrossRef]
  114. Maneesai, P.; Wattanathorn, J.; Potue, P.; Khamseekaew, J.; Rattanakanokchai, S.; Thukham-Mee, W.; Muchimapura, S.; Pakdeechote, P. Cardiovascular complications are resolved by tuna protein hydrolysate supplementation in rats fed with a high-fat diet. Sci. Rep. 2023, 13, 12280. [Google Scholar] [CrossRef]
  115. Brown, S.; Atkins, C.; Bagley, R.; Carr, A.; Cowgill, L.; Davidson, M.; Egner, B.; Elliott, J.; Henik, R.; Labato, M.; et al. Guidelines for the identification, evaluation, and management of systemic hypertension in dogs and cats. J. Vet. Intern. Med. 2007, 21, 542–558. [Google Scholar] [CrossRef]
  116. Available online: https://todaysveterinarypractice.com/cardiology/systemic-hypertension-in-dogs-cats/ (accessed on 10 August 2024).
  117. Aspevik, T.; Oterhals, Å.; Rønning, S.B.; Altintzoglou, T.; Wubshet, S.G.; Gildberg, A.; Afseth, N.K.; Whitaker, R.D.; Lindberg, D. Valorization of Proteins from Co- and By-Products from the Fish and Meat Industry. Top. Curr. Chem. 2017, 375, 53. [Google Scholar] [CrossRef]
  118. Atanassova, M.R.; Stangeland, J.K.; Lausen, S.E.; Dahl, T.H.; Barnung, T.; Larssen, W.E. Antioxidants, ACE I Inhibitory Peptides, and Physicochemical Composition, with a Special Focus on Trace Elements and Pollutants, of SPRING Spawning Atlantic Herring (Clupea harengus) Milt and Hydrolysates for Functional Food Applications. Fishes 2024, 9, 456. [Google Scholar] [CrossRef]
  119. Aissaoui, N.; Abidi, F.; Marzouki, M.N. ACE inhibitory and antioxidant activities of red scorpionfish (Scorpaena notata) protein hydrolysates. J. Food Sci. Technol. 2015, 52, 7092–7102. [Google Scholar] [CrossRef]
  120. Taheri, A.; Bakhshizadeh, G.A. Antioxidant and ACE Inhibitory Activities of Kawakawa (Euthynnus affinis) Protein Hydrolysate Produced by Skipjack Tuna Pepsin. J. Aquat. Food Prod. Technol. 2019, 29, 148–166. [Google Scholar] [CrossRef]
  121. Johnson, K.A.; Lee, A.H.; Swanson, K.S. Nutrition and nutraceuticals in the changing management of osteoarthritis for dogs and cats. J. Am. Vet. Med. Assoc. 2020, 256, 1335–1341. [Google Scholar] [CrossRef]
  122. Barrouin-Melo, S.M.; Anturaniemi, J.; Sankari, S.; Griinari, M.; Atroshi, F.; Ounjaijean, S.; Hielm-Björkman, A.K. Evaluating oxidative stress, serological- and haematological status of dogs suffering from osteoarthritis, after supplementing their diet with fish or corn oil. Lipids Health Dis. 2016, 15, 139. [Google Scholar] [CrossRef]
  123. Manfredi, S.; Di Ianni, F.; Di Girolamo, N.; Canello, S.; Gnudi, G.; Guidetti, G.; Miduri, F.; Fabbi, M.; Daga, E.; Parmigiani, E.; et al. Effect of a commercially available fish-based dog food enriched with nutraceuticals on hip and elbow dysplasia in growing Labrador retrievers. Can. J. Vet. Res. 2018, 82, 154–158. [Google Scholar]
  124. Eckert, T.; Jährling-Butkus, M.; Louton, H.; Burg-Roderfeld, M.; Zhang, N.; Zhang, R.; Hesse, K.; Petridis, A.K.; Kožár, T.; Steinmeyer, J.; et al. Biophysical, Biochemical and Biomedical Evaluation of Collagen Hydrolysate in Comparison to Sulfated Glucosamine from Marine Organisms and Lipids of Fish Oil Origin Used as Chondroprotective Food Supplements. Preprints 2021. [Google Scholar]
  125. Heinrich, N.A.; Eisenschenk, M.; Harvey, R.G.; Nuttall, T. Skin Diseases of the Dog and Cat, 3rd. ed.; CRC Press: London, UK, 2018. [Google Scholar] [CrossRef]
  126. Hsu, C.; Marx, F.; Guldenpfennig, R.; Valizadegan, N.; de Godoy, M.R.C. The effects of hydrolyzed protein on macronutrient digestibility, fecal metabolites and microbiota, oxidative stress and inflammatory biomarkers, and skin and coat quality in adult dogs. J. Anim. Sci. 2024, 102, skae057. [Google Scholar] [CrossRef]
  127. Tresamol, P.V.; Nissar, B.J. Nutritional approaches for management of dermatological disorders in canine. Intas Polivet 2020, 21, 474–477. [Google Scholar]
  128. Fritsch, D.A.; Roudebush, P.; Allen, T.A.; Leventhal, P.S.; Brejda, J.; Hahn, K.A. Effect of Two Therapeutic Foods in Dogs with Chronic Nonseasonal Pruritic Dermatitis. Int. J. Appl. Res. Vet. Med. 2010, 8, 146–154. [Google Scholar]
  129. Szczepanik, M.P.; Gołyński, M.; Wilkołek, P.; Kalisz, G. Evaluation of a hydrolysed salmon and pea hypoallergenic diet application in dogs and cats with cutaneous adverse food reaction. Pol. J. Vet. Sci. 2022, 25, 67–73. [Google Scholar] [CrossRef] [PubMed]
  130. Meineri, G.; Martello, E.; Radice, E.; Bruni, N.; Saettone, V.; Atuahene, D.; Armandi, A.; Testa, G.; Ribaldone, D.G. Chronic Intestinal Disorders in Humans and Pets: Current Management and the Potential of Nutraceutical Antioxidants as Alternatives. Animals 2022, 12, 812. [Google Scholar] [CrossRef] [PubMed]
  131. Dandrieux, J.R. Inflammatory bowel disease versus chronic enteropathy in dogs: Are they one and the same? J. Small Anim. Pract. 2016, 57, 589–599. [Google Scholar] [CrossRef] [PubMed]
  132. Allenspach, K. Clinical immunology and immunopathology of the canine and feline intestine. Vet. Clin. N. Am. Small Anim. Pract. 2011, 41, 345–360. [Google Scholar] [CrossRef] [PubMed]
  133. Awuchi, C.G.; Chukwu, C.N.; Iyiola, A.O.; Noreen, S.; Morya, S.; Adeleye, A.O.; Twinomuhwezi, H.; Leicht, K.; Mitaki, N.B.; Okpala, C.O.R. Bioactive Compounds and Therapeutics from Fish: Revisiting Their Suitability in Functional Foods to Enhance Human Wellbeing. BioMed Res. Int. 2022, 2022, 3661866. [Google Scholar] [CrossRef]
  134. Jergens, A.E.; Heilmann, R.M. Canine chronic enteropathy-Current state-of-the-art and emerging concepts. Front. Vet. Sci. 2022, 9, 923013. [Google Scholar] [CrossRef]
  135. Marks, S.L.; Laflamme, D.P.; McAloose, D. Dietary trial using a commercial hypoallergenic diet containing hydrolyzed protein for dogs with inflammatory bowel disease. Vet. Ther. Res. Appl. Vet. Med. 2002, 3, 109–118. [Google Scholar]
  136. Marchesi, M.C.; Timpano, C.C.; Busechian, S.; Pieramati, C.; Rueca, F. The role of diet in managing inflamatory bowel disease affected dogs: A retrospective cohort study on 76 cases. Vet. Ital. 2017, 53, 297–302. [Google Scholar] [CrossRef]
  137. Ontsouka, E.C.; Burgener, I.A.; Luckschander-Zeller, N.; Blum, J.W.; Albrecht, C. Fish-meal diet enriched with omega-3 PUFA and treatment of canine chronic enteropathies. Eur. J. Lipid Sci. Technol. 2012, 114, 412–422. [Google Scholar] [CrossRef]
  138. Simpson, K.W.; Miller, M.L.; Loftus, J.P.; Rishniw, M.; Frederick, C.E.; Wakshlag, J.J. Randomized controlled trial of hydrolyzed fish diets in dogs with chronic enteropathy. J. Vet. Intern. Med. 2023, 37, 2334–2343. [Google Scholar] [CrossRef]
  139. Zinn, K.E.; Hernot, D.C.; Fastinger, N.D.; Karr-Lilienthal, L.K.; Bechtel, P.J.; Swanson, K.S.; Fahey, G.C., Jr. Fish protein substrates can substitute effectively for poultry by-product meal when incorporated in high-quality senior dog diets. J. Anim. Physiol. Anim. Nutr. 2009, 93, 447–455. [Google Scholar] [CrossRef]
  140. Sechi, S.; Fiore, F.; Chiavolelli, F.; Dimauro, C.; Nudda, A.; Cocco, R. Oxidative stress and food supplementation with antioxidants in therapy dogs. Can. J. Vet. Res. 2017, 81, 206–216. [Google Scholar]
  141. Brown, S.A. Oxidative stress and chronic kidney disease. Vet. Clin. N. Am. Small Anim. Pract. 2008, 38, 157–166. [Google Scholar] [CrossRef]
  142. Nasri, R.; Abdelhedi, O.; Jemil, I.; Daoued, I.; Hamden, K.; Kallel, C.; Elfeki, A.; Lamri-Senhadji, M.; Boualga, A.; Nasri, M.; et al. Ameliorating effects of goby fish protein hydrolysates on high-fat-high-fructose diet-induced hyperglycemia, oxidative stress and deterioration of kidney function in rats. Chem.-Biol. Interact. 2015, 242, 71–80. [Google Scholar] [CrossRef]
  143. de Godoy, M.R.C.; Conway, C.E.; Mcleod, K.R.; Harmon, D.L. Influence of feeding a fish oil-containing diet to young, lean, adult dogs: Effects on lipid metabolites, postprandial glycaemia and body weight. Arch. Anim. Nutr. 2015, 69, 499–514. [Google Scholar] [CrossRef]
  144. Riyadi, P.H.; Atho’illah, M.F.; Tanod, W.A.; Rahmawati, I.S. Tilapia viscera hydrolysate extract alleviates oxidative stress and renal damage in deoxycorticosterone acetate-salt induced hypertension rats. Vet. World 2020, 13, 2477–2483. [Google Scholar] [CrossRef]
  145. Ravić, B.; Debeljak-Martacić, J.; Pokimica, B.; Vidović, N.; Ranković, S.; Glibetić, M.; Stepanović, P.; Popović, T. The Effect of Fish Oil-Based Foods on Lipid and Oxidative Status Parameters in Police Dogs. Biomolecules 2022, 12, 1092. [Google Scholar] [CrossRef]
  146. Theysgeur, S.; Cudennec, B.; Deracinois, B.; Perrin, C.; Guiller, I.; Lepoudère, A.; Flahaut, C.; Ravallec, R. New Bioactive Peptides Identified from a Tilapia Byproduct Hydrolysate Exerting Effects on DPP-IV Activity and Intestinal Hormones Regulation after Canine Gastrointestinal Simulated Digestion. Molecules 2020, 26, 136. [Google Scholar] [CrossRef] [PubMed]
  147. Sordo, L.; Gunn-Moore, D.A. Cognitive Dysfunction in Cats: Update on Neuropathological and Behavioural Changes Plus Clinical Management. Vet. Rec. 2021, 188, e3. [Google Scholar] [CrossRef] [PubMed]
  148. Zicker, S.C.; Jewell, D.E.; Yamka, R.M.; Milgram, N.W. Evaluation of cognitive learning, memory, psychomotor, immunologic, and retinal functions in healthy puppies fed foods fortified with docosahexaenoic acid-rich fish oil from 8 to 52 weeks of age. J. Am. Vet. Med. Assoc. 2012, 241, 583–594. [Google Scholar] [CrossRef]
  149. Pan, Y.; Araujo, J.A.; Burrows, J.; de Rivera, C.; Gore, A.; Bhatnagar, S.; Milgram, N.W. Cognitive enhancement in middle-aged and old cats with dietary supplementation with a nutrient blend containing fish oil, B vitamins, antioxidants and arginine. Br. J. Nutr. 2013, 110, 40–49. [Google Scholar] [CrossRef]
  150. Pan, Y.; Kennedy, A.D.; Jönsson, T.J.; Milgram, N.W. Cognitive enhancement in old dogs from dietary supplementation with a nutrient blend containing arginine, antioxidants, B vitamins and fish oil. Br. J. Nutr. 2018, 119, 349–358. [Google Scholar] [CrossRef]
  151. Chataigner, M.; Mortessagne, P.; Lucas, C.; Pallet, V.; Layé, S.; Mehaignerie, A.; Bouvret, E.; Dinel, A.L.; Joffre, C. Dietary fish hydrolysate supplementation containing n-3 LC-PUFAs and peptides prevents short-term memory and stress response deficits in aged mice. Brain Behav. Immun. 2021, 91, 716–730. [Google Scholar] [CrossRef]
  152. Landsberg, G.M.; Mougeot, I.; Kelly, S.; Milgram, N.W. Assessment of noise-induced fear and anxiety in dogs: Modification by a novel fish hydrolysate supplemented diet. J. Vet. Behav. 2015, 10, 391–398. [Google Scholar] [CrossRef]
  153. Bernet, F.; Montel, V.; Noël, B.; Dupouy, J.P. Diazepam-like effects of a fish protein hydrolysate (Gabolysat PC60) on stress responsiveness of the rat pituitary-adrenal system and sympathoadrenal activity. Psychopharmacology 2000, 149, 34–40. [Google Scholar] [CrossRef] [PubMed]
  154. Freret, T.; Largilliere, S.; Nee, G.; Coolzaet, M.; Corvaisier, S.; Boulouard, M. Fast Anxiolytic-Like Effect Observed in the Rat Conditioned Defensive Burying Test, after a Single Oral Dose of Natural Protein Extract Products. Nutrients 2021, 13, 2445. [Google Scholar] [CrossRef]
  155. Titeux, E.; Padilla, S.; Paragon, B.M.; Gilbert, C. Effects of a new dietary supplement on behavioural responses of dogs exposed to mild stressors. Vet. Med. Sci. 2021, 7, 1469–1482. [Google Scholar] [CrossRef]
  156. Jeusette, I.C.; Tami, G.; Fernández, A.; Torre, C.; Tvarijonaviciute, A.; Cerón, J.J.; Salas-Mani, A.; Fatjó, J. Evaluation of a new prescription diet with lemon balm, fish peptides, oligofructose and L-tryptophan to reduce urinary cortisol, used as a marker of stress, in cats. J. Vet. Behav. 2021, 42, 30–36. [Google Scholar] [CrossRef]
  157. Porcheron, G.; Bodet, M.; Poiron, K.; Brault, J.; Massal, N.; Roy, O.; Fantini, O. Supplementation Effects of an Alpha-Casozepine and White Fish Muscle Hydrolyzed Complementary Feed on Canine Separation-Related Disorders and Quality of Life of Dogs and Their Pet Caregivers. Open J. Vet. Med. 2023, 13, 68–81. [Google Scholar] [CrossRef]
  158. Ephraim, E.; Brockman, J.A.; Jewell, D.E. A Diet Supplemented with Polyphenols, Prebiotics and Omega-3 Fatty Acids Modulates the Intestinal Microbiota and Improves the Profile of Metabolites Linked with Anxiety in Dogs. Biology 2022, 11, 976. [Google Scholar] [CrossRef] [PubMed]
  159. Dinel, A.-L.; Lucas, C.; Le Faouder, J.; Bouvret, E.; Pallet, V.; Layé, S.; Joffre, C. Supplementation with low molecular weight peptides from fish protein hydrolysate reduces acute mild stress-induced corticosterone secretion and modulates stress responsive gene expression in mice. J. Funct. Foods 2021, 76, 104292. [Google Scholar] [CrossRef]
  160. Venslauskas, K.; Navickas, K.; Nappa, M.; Kangas, P.; Mozūraitytė, R.; Šližytė, R.; Župerka, V. Energetic and Economic Evaluation of Zero-Waste Fish Co-Stream Processing. Int. J. Environ. Res. Public Health 2021, 18, 2358. [Google Scholar] [CrossRef]
  161. Available online: https://www.fortunebusinessinsights.com/industry-reports/pet-food-market-100554 (accessed on 23 September 2024).
  162. Nagodawithana, T.W.; Nelles, L.; Trivedi, N.B. Protein Hydrolysates as Hypoallergenic, Flavors and Palatants for Companion Animals. In Protein Hydrolysates in Biotechnology; Pasupuleti, V., Demain, A., Eds.; Springer: Dordrecht, The Netherlands, 2008. [Google Scholar] [CrossRef]
  163. Watson, P.E.; Thomas, D.G.; Bermingham, E.N.; Schreurs, N.M.; Parker, M.E. Drivers of Palatability for Cats and Dogs—What It Means for Pet Food Development. Animals 2023, 13, 1134. [Google Scholar] [CrossRef]
  164. Samant, S.S.; Crandall, P.G.; Jarma Arroyo, S.E.; Seo, H.S. Dry Pet Food Flavor Enhancers and Their Impact on Palatability: A Review. Foods 2021, 10, 2599. [Google Scholar] [CrossRef]
  165. Available online: https://www.precisionbusinessinsights.com/market-reports/pet-food-palatants-market (accessed on 23 September 2024).
  166. Folador, J.F.; Karr-Lilienthal, L.K.; Parsons, C.M.; Bauer, L.L.; Utterback, P.L.; Schasteen, C.S.; Bechtel, P.J.; Fahey, G.C., Jr. Fish meals, fish components, and fish protein hydrolysates as potential ingredients in pet foods. J. Anim. Sci. 2006, 84, 2752–2765. [Google Scholar] [CrossRef]
  167. Aldrich, G.C.; Koppel, K. Pet Food Palatability Evaluation: A Review of Standard Assay Techniques and Interpretation of Results with a Primary Focus on Limitations. Animals 2015, 5, 43–55. [Google Scholar] [CrossRef] [PubMed]
  168. van Rooijen, C.; Bosch, G.; van der Poel, A.F.; Wierenga, P.A.; Alexander, L.; Hendriks, W.H. The Maillard reaction and pet food processing: Effects on nutritive value and pet health. Nutr. Res. Rev. 2013, 26, 130–148. [Google Scholar] [CrossRef]
  169. Pekel, A.Y.; Mülazımoğlu, S.B.; Acar, N. Taste preferences and diet palatability in cats. J. Appl. Anim. Res. 2020, 48, 281–292. [Google Scholar] [CrossRef]
  170. Salaun, F.; Blanchard, G.; Le Paih, L.; Roberti, F.; Niceron, C. Impact of macronutrient composition and palatability in wet diets on food selection in cats. J. Anim. Physiol. Anim. Nutr. 2017, 101, 320–328. [Google Scholar] [CrossRef] [PubMed]
  171. Guilherme-Fernandes, J.; Aires, T.; Fonseca, A.J.M.; Yergaliyev, T.; Camarinha-Silva, A.; Lima, S.A.C.; Maia, M.R.G.; Cabrita, A.R.J. Squid meal and shrimp hydrolysate as novel protein sources for dog food. Front. Vet. Sci. 2024, 11, 1360939. [Google Scholar] [CrossRef] [PubMed]
  172. Cottrell, R.S.; Blanchard, J.L.; Halpern, B.S.; Metian, M.; Froehlich, H.E. Global adoption of novel aquaculture feeds could substantially reduce forage fish demand by 2030. Nat. Food 2020, 1, 301–308. [Google Scholar] [CrossRef]
  173. Sandström, V.; Chrysafi, A.; Lamminen, M.; Troell, M.; Jalava, M.; Piipponen, J.; Siebert, S.; van Hal, O.; Virkki, V.; Kummu, M. Food system by-products upcycled in livestock and aquaculture feeds can increase global food supply. Nat. Food 2022, 3, 729–740. [Google Scholar] [CrossRef] [PubMed]
  174. Naylor, R.L.; Hardy, R.W.; Buschmann, A.H.; Bush, S.R.; Cao, L.; Klinger, D.H.; Little, D.C.; Lubchenco, J.; Shumway, S.E.; Troell, M. A 20-year retrospective review of global aquaculture. Nature 2021, 591, 551–563. [Google Scholar] [CrossRef]
  175. Miao, W.; Wang, W. Trends of aquaculture production and trade: Carp, tilapia, and shrimp. Asian Fish. Sci. 2020, 33 (Suppl. 1), 1–10. [Google Scholar] [CrossRef]
  176. Boyd, C.E.; D’Abramo, L.R.; Glencross, B.D.; Huyben, D.C.; Juarez, L.M.; Lockwood, G.S.; McNevin, A.A.; Tacon, A.G.J.; Teletchea, F.; Tomasso, J.R., Jr.; et al. Achieving sustainable aquaculture: Historical and current perspectives and future needs and challenges. J. World Aquac. Soc. 2020, 51, 578–633. [Google Scholar] [CrossRef]
  177. Shahzad, S.M. Fish as a Healthy Source of Human Nutrition: An Exploratory Study. J. Naut. Eye Strateg. Stud. 2024, 4, 1–22. [Google Scholar]
  178. Thazeem, B.; Umesh, M.; Sarojini, S.; Allwyn Vyas, G.; Adhithya Sankar, S.; Sapthami, K.; Suresh, S.; Merin Stanly, L. Developments in Feeds in Aquaculture Sector: Contemporary Aspects. In Aquaculture Science and Engineering; Balasubramanian, B., Liu, W.C., Sattanathan, G., Eds.; Springer: Singapore, 2022. [Google Scholar] [CrossRef]
  179. Abdullah, F.I.; Hamid, N.H.; Abd Karim, M.M.; Ismail, M.F.; Sin, N.L.W.W.; Kamaruddin, M.S. Fish protein hydrolysate for fish health. Biocatal. Agric. Biotechnol. 2024, 60, 103292. [Google Scholar] [CrossRef]
  180. Chaklader, M.R.; Fotedar, R.; Howieson, J.; Siddik, M.A.; Foysal, M.J. The ameliorative effects of various fish protein hydrolysates in poultry by-product meal based diets on muscle quality, serum biochemistry and immunity in juvenile barramundi, Lates calcarifer. Fish Shellfish. Immunol. 2020, 104, 567–578. [Google Scholar] [CrossRef]
  181. Sampathkumar, K.; Yu, H.; Loo, S.C.J. Valorisation of industrial food waste into sustainable aquaculture feeds. Future Foods 2023, 7, 100240. [Google Scholar] [CrossRef]
  182. Gunathilaka, B.E.; Kim, S.; Herault, M.; Fournier, V.; Lee, K.-J. Marine protein hydrolysates as a substitute of squid-liver powder in diets for Pacific white shrimp (Litopenaeus vannamei). Aquac. Res. 2022, 53, 4059–4068. [Google Scholar] [CrossRef]
  183. Sezu, N.H.; Nandi, S.K.; Suma, A.Y.; Kari, Z.A.; Wei, L.S.; Seguin, P.; Herault, M.; Hossain, M.S.; Irwan Khoo, M.; El-Sayed Hemdan, I.; et al. Ameliorative effects of different dietary levels of fish protein hydrolysate (FPH) on growth and reproductive performance, feed stability, tissues biochemical composition, haematobiochemical profile, liver histology, and economic analysis of Pabda (Ompok pabda) broodstock. Aquac. Res. 2024, 2024, 6044920. [Google Scholar] [CrossRef]
  184. Kabir, M.A.; Nandi, S.K.; Suma, A.Y.; Abdul Kari, Z.; Mohamad Sukri, S.A.; Wei, L.S.; Al Mamun, A.; Seguin, P.; Herault, M.; Khoo, M.I.; et al. The Potential of Fish Protein Hydrolysate Supplementation in Nile Tilapia Diets: Effects on Growth and Health Performance, Disease Resistance, and Farm Economic Analysis. Appl. Biochem. Biotechnol. 2024, 196, 7145–7167. [Google Scholar] [CrossRef]
  185. Kotzamanis, Y.P.; Gisbert, E.; Gatesoupe, F.J.; Zambonino Infante, J.; Cahu, C. Effects of different dietary levels of fish protein hydrolysates on growth, digestive enzymes, gut microbiota, and resistance to Vibrio anguillarum in European sea bass (Dicentrarchus labrax) larvae. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2007, 147, 205–214. [Google Scholar] [CrossRef]
  186. Colombo, S.M.; Roy, K.; Mraz, J.; Wan, A.H.; Davies, S.J.; Tibbetts, S.M.; Øverland, S.M.; Francis, D.S.; Rocker, M.M.; Gasco, L.; et al. Towards achieving circularity and sustainability in feeds for farmed blue foods. Rev. Aquac. 2023, 15, 1115–1141. [Google Scholar] [CrossRef]
  187. Refstie, S.; Olli, J.J.; Standal, H. Feed intake, growth, and protein utilisation by post-smolt Atlantic salmon (Salmo salar) in response to graded levels of fish protein hydrolysate in the diet. Aquaculture 2004, 239, 331–349. [Google Scholar] [CrossRef]
  188. Tang, H.G.; Wu, T.X.; Zhao, Z.Y.; Pan, X.D. Effects of fish protein hydrolysate on growth performance and humoral immune response in large yellow croaker (Pseudosciaena crocea R.). J. Zhejiang Univ. Sci. B 2008, 9, 684–690. [Google Scholar] [CrossRef]
  189. Zheng, K.; Liang, M.; Yao, H.; Wang, J.; Chang, Q. Effect of dietary fish protein hydrolysate on growth, feed utilisation and IGF-I levels of Japanese flounder (Paralichthys olivaceus). Aquac. Nutr. 2012, 18, 297–303. [Google Scholar] [CrossRef]
  190. Siddik, M.A.; Pham, H.D.; Francis, D.S.; Vo, B.V.; Shahjahan, M. Dietary supplementation of fish protein hydrolysate in high plant protein diets modulates growth, liver and kidney health, and immunity of barramundi (Lates calcarifer). Aquac. Nutr. 2021, 27, 86–98. [Google Scholar] [CrossRef]
  191. Siddaiah, G.M.; Kumar, R.; Kumari, R.; Damle, D.K.; Rasal, K.D.; Manohar, V.; Sundaray, J.K.; Pillai, B.R. Dietary supplementation of fish protein hydrolysate improves growth, feed efficiency and immune response in freshwater carnivore fish, Channa striata fingerlings. Aquac. Res. 2022, 53, 3401–3415. [Google Scholar] [CrossRef]
  192. Suratip, N.; Charoenwattanasak, S.; Klahan, R.; Herault, M.; Yuangsoi, B. An investigation into the effects of using protein hydrolysate in low fish meal diets on growth performance, feed utilization and health status of snakehead fish (Channa striata) fingerling. Aquac. Rep. 2023, 30, 101623. [Google Scholar] [CrossRef]
  193. Tola, S.; Sommit, N.; Seel-audom, M.; Khamtavee, P.; Waiho, K.; Boonmee, T.; Yuangsoi, B.; Munpholsri, N. Effects of dietary tuna hydrolysate supplementation on feed intake, growth performance, feed utilization and health status of Asian sea bass (Lates calcarifer) fed a low fish meal soybean meal-based diet. Aquac. Res. 2022, 53, 3898–3912. [Google Scholar] [CrossRef]
  194. Suma, A.Y.; Nandi, S.K.; Abdul Kari, Z.; Goh, K.W.; Wei, L.S.; Tahiluddin, A.B.; Seguin, P.; Herault, M.; Al Mamun, A.; Téllez-Isaías, G.; et al. Beneficial Effects of Graded Levels of Fish Protein Hydrolysate (FPH) on the Growth Performance, Blood Biochemistry, Liver and Intestinal Health, Economics Efficiency, and Disease Resistance to Aeromonas hydrophila of Pabda (Ompok pabda) Fingerling. Fishes 2023, 8, 147. [Google Scholar] [CrossRef]
  195. Sandbakken, I.S.; Five, K.K.; Bardal, T.; Knapp, J.L.; Olsen, R.E. Salmon hydrolysate as a protein source for Atlantic salmon; prion content and effects on growth, digestibility and gut health. Aquaculture 2023, 576, 739863. [Google Scholar] [CrossRef]
  196. Jaies, I.; Qayoom, I.; Saba, F.; Khan, S. Fish wastes as sources of fertiliser and manure. In Fish Wastes to Valuable Products, 1st ed.; Mqsood, S., Naseer, M.N., Benjakul, S., Zaidi, A.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2024; pp. 329–338. [Google Scholar]
  197. du Jardin, P. Plant biostimulants: Definition, concept, main categories & regulation. Sci. Hortic. 2015, 196, 3–14. [Google Scholar]
  198. Lindsey, A.J.; Thoms, A.W.; Christians, N.E. Evaluation of Humic Fertilizers on Kentucky Bluegrass Subjected to Simulated Traffic. Agronomy 2021, 11, 611. [Google Scholar] [CrossRef]
  199. Madende, M.; Hayes, M. Fish By-Product Use as Biostimulants: An Overview of the Current State of the Art, Including Relevant Legislation and Regulations within the EU and USA. Molecules 2020, 25, 1122. [Google Scholar] [CrossRef]
  200. Domínguez, H.; Iñarra, B.; Labidi, J.; Bald, C. Fish Viscera Hydrolysates and Their Use as Biostimulants for Plants as an Approach towards a Circular Economy in Europe: A Review. Sustainability 2024, 16, 8779. [Google Scholar] [CrossRef]
  201. Malécange, M.; Sergheraert, R.; Teulat, B.; Mounier, E.; Lothier, J.; Sakr, S. Biostimulant Properties of Protein Hydrolysates: Recent Advances and Future Challenges. Int. J. Mol. Sci. 2023, 24, 9714. [Google Scholar] [CrossRef]
  202. Takayama, S.; Sakagami, Y. Peptide signalling in plants. Curr. Opin. Plant Biol. 2002, 5, 382–387. [Google Scholar] [CrossRef]
  203. Bulgari, R.; Cocetta, G.; Trivellini, A.; Vernieri, P.; Ferrante, A. Biostimulants and crop responses: A review. Biol. Agric. Hortic. 2014, 31, 1–17. [Google Scholar] [CrossRef]
  204. Rajabi Hamedani, S.; Rouphael, Y.; Colla, G.; Colantoni, A.; Cardarelli, M. Biostimulants as a Tool for Improving Environmental Sustainability of Greenhouse Vegetable Crops. Sustainability 2020, 12, 5101. [Google Scholar] [CrossRef]
  205. Rouphael, Y.; Franken, P.; Schneider, C.; Schwarz, D.; Giovannetti, M.; Agnolucci, M.; De Pascale, S.; Bonini, P.; Colla, G. Arbuscular mycorrhizal fungi act as biostimulants in horticultural crops. Sci. Hortic. 2015, 196, 91–108. [Google Scholar] [CrossRef]
  206. Petrova, I.; Tolstorebrov, I.; Eikevik, T.M. Production of fish protein hydrolysates step by step: Technological aspects, equipment used, major energy costs and methods of their minimizing. Int. Aquat. Res. 2018, 10, 223–241. [Google Scholar] [CrossRef]
  207. Ruzzi, M.; Cola, G.; Rouphael, Y. Biostimulants in Agriculture II: Towards a sustainable future. Crop Prod. Physiol. 2024, 15, 1427283. [Google Scholar] [CrossRef]
Figure 1. Different modes of action of biostimulants.
Figure 1. Different modes of action of biostimulants.
Applsci 15 05769 g001
Table 2. Bioactive ingredients/hydrolysates marketed for pets available commercially.
Table 2. Bioactive ingredients/hydrolysates marketed for pets available commercially.
ProductManufacturersIngredients/Raw MaterialProduct DescriptionSpecific Features
Salmigo® Protect L60Biomega
(Skogvåg, Vestland, Norway)
[160]
Fresh, high-quality Atlantic salmon leftovers
Liquid salmon peptide
Produced using patented non-GMO enzymatic process
Food-grade, preservative-free formulation
Clean-label, natural protein source
90% protein content
Rapidly absorbable bioactive peptides and amino acids
Enhances nutrient delivery to body tissues
Supports feline-specific nutritional needs
Salmigo® ActiveFresh, high-quality Atlantic salmon leftovers
Salmon peptide powder
Made via patented non-GMO enzymatic hydrolysis
Food-grade, preservative-free and naturally processed
Promotes skin, coat, joint, and overall pet well-being
Highly digestible, absorbable protein source (>70% peptides and free amino acids)
Delivers high biological value nutrition
Dog Adult Dermal Hydrolysed FishTrovet + Plus
(De Vergert 4, 6681 Le Bemmel, Netherlands)
[126]
White FPHs, rice, potato, apple, fish oil, linseed, minerals, borage oil
Low molecular weight hydrolysed protein (<10,000 Da)
Nutritional profile: protein 8%, fat 7%, fibre 0.75%, ash 2.5%
Supports skin health in pets with dermatosis and hair loss
Reduces inflammation and allergic reactions
Promotes skin barrier strength, fur vitality, and overall dermal function
Cat Intestinal Fresh Hydrolysed White FishExtruded rice, rice protein, hydrolysed white fish, fish oil, sunflower oil, hydrolysed fish, krill Antarctic, apple fibre, vegetable fibres, psyllium, hydrolysed yeast cell wall, and inulin
Single animal protein source
Nutrient profile: 32% protein, 20% fat, 2.75% fibre, 6.50% ash
Minimises ingredient and nutrient intolerances
Supports compromised digestion
Promotes skin health and reduces excessive hair loss
Enriched with prebiotics
Multi-Purpose Treat for cats (Hydrolysed Protein)Rice, salmon protein (hydrolysate), sodium chloride, sugar, collagen (hydrolysate), cellulose, poultry fat
Hypoallergenic treat with a single hydrolysed protein (fish) source
Nutrient profile: 30% protein, 8% fat, 5.5% ash, 2.5% fibre
Prevents both food allergies and intolerances that cause skin issues
Reduces risk of digestive problems like diarrhoea and vomiting
Supports prevention of chronic kidney (renal) insufficiencies
Sensitive Soft Chews SalmonOptima Nova
(ARGANDA DEL REY, Madrid, Spain)
Salmon and salmon by-products, tapioca and potato, and fats
Grain free functional snacks for dogs
Nutrient profile: protein 35%, fat 5%, fibres 2%, ash 10.5%
Easily digestible functional snacks with hypoallergenic ingredients
Ideal for dogs with sensitive skin or digestive issues
Adult Sensitive Salmon and PotatoFresh, ground, hydrolysed salmon, dehydrated potato, potato protein, oil, sugar beet meal, yeast, FOS, MOS
Balanced and complete nutrition for dogs
Hypoallergenic
Gluten free formulation
Nutrient profile: protein 27%, fats 16%, fibre 2.75%, ash 7%
Supports dogs with food intolerances and allergies, improving skin and digestion
Provides healthy skin, shiny coat, and enhanced heart, immune, and nervous system function
Potential anti-cancer benefits
Medica Hypoallergenic PLUS Salmon for dogsSelect Gold
(Germany)
Potato flakes, salmon protein, hydrolysates, hydrolysed poultry liver fish oil, potato protein, sunflower oil, poultry fat
Single, partially hydrolysed, animal protein source
Nutritional profile: protein 23.5%, fat 16.5%, ash 6.8%, fibre 2.6%
Assists dogs with food sensitivities and allergies
Alleviates initial symptoms of intolerance
Provides balanced and complete nutrition
Medica Hypoallergenic PLUS Fish and rice for catsRice, potato protein, hydrolysed fish protein, poultry fat, potato starch, dried beet pulp, beef fat, cellulose, sardine oil, hydrolysed poultry liver, salmon oil, and linseed oil
Single, partially hydrolysed, animal protein source
Nutritional profile: protein 33%, fat 18%, ash 6%, fibre 3.3%
Minimises nutrient intolerances and food allergies in cats
Supports skin function in dermatosis
Reduce inflammation in sensitive cats
Functional treats—sensitive proteinBoxby (Berltsum, Friesland, Netherlands)Hydrolysed salmon, potato flakes, rice meal, omega-3 fatty acids, propylene glycol, omega-6 fatty acids
Hydrolysed salmon proteins
Free from wheat and soya
Nutritional Profile: protein 23%, fat 5%, fibre 1%, ash 2%
Reduces unwanted dietary allergies and intolerances in dogs
Supports skin and coat health
SEACURE®—Daily protein supplement for dogsProper Nutrition (West Chester, PA, USA).Made from Pacific whiting caught in the Pacific Northwest
Enzymatically hydrolysed proteins
Rich in amino acids and peptides
Protein supplement for overall health
Nutritional Profile: protein 60%, fibre 5%, fat 2%
Enhances overall health for puppies, adults, and sick or elderly dogs
Boosts immune function
Promotes healthy skin and coat
Alleviates food intolerance symptoms
HY Food Allergen ManagementSpecific Cat FDD (Dechra, Northwich, UK)Rice, rice protein, hydrolysed salmon protein, pork fat, cellulose, protein hydrolysate, fish oil, psyllium husks, rosemary extract
Specially formulated for sensitive cats with ingredient intolerances
Nutritional profile: protein 27.5%, fat 12.5%, fibre 3.3%, ash 7.2%
Supports a healthy immune system
Helps maintain normal urinary tract function
Relieves acute intestinal malabsorption
Aids digestion and supports pancreatic function
Veterinary HPM® Hypoallergic dog foodwith hydrolysed fish proteinVirbac UK (Woolpit, Bury, Suffolk, UK)Potato starch, hydrolysed fish protein, animal fats, lignocellulose, minerals, hydrolysed pork and poultry proteins, beet pulp, FOS, mono di and triglycerides of fatty acids, Lactobacillus acidophilus
Elimination diet for adult and senior dogs
Contains 97% peptides <10 kDa (mean molecular weight: 1.82 kDa) for optimal digestibility
Nutritional profile: protein 24%, fat 18%, fibre 4.5%, ash 7%
Provides natural dermal defences
Reduces the risk of other allergies
Maintains a healthy coat
Helps with gastrointestinal functions
Vet Life Cat Ultra hypoFarmina Pet Foods (Essex, UK)
[84]
Rice starch, hydrolysed fish protein, fish oil, potassium chloride, calcium carbonate, mono-di-calcium phosphate
Developed without vegetable proteins for food allergies or intolerances
Nutritional profile: protein 18%, fat 15%, fibre 1.2%, ash 5.3%
Hypoallergenic and highly digestible formula
Supports a healthy immune system and natural skin barrier
Feline c/d stress urinary care—ocean fishHill’s® Prescription Diet®
[124,128]
Brewer’s rice, ocean FPHs, chicken- and turkey meal, maize gluten meal, animal fats, tuna fish meal, soya bean oil, minerals, fish oil, linseeds
Nutritional Profile: protein 33.3%, fat 15.2%, fibre 0.6%, ash 5.2%
Reduces risk of urinary tract issues and urinary stone formation
Contains protein hydrolysates and L-tryptophan to support emotional balance
Canine skin support potato salmon formulaSalmon proteins, potato protein, potato starch, pork fat, soybean oil, pork flavour, fish oil, fish meal.
Nutritional Profile: protein 18.4%, fat 15.5%, fibre 1.86%, total omega-3 FA 1.41%
Supports healthy skin, coat, digestion, and immune system
Natural veterinary diet hf hydrolysed intolerance, salmon, dry cat foodBlue Buffalo
(Wilton, CT, USA)
Salmon hydrolysate, peas, potatoes, pea starch, canola oil, pea protein, flaxseed, pea fibre, fish oil, pumpkin, dried kelp, oil of rosemary.
Formula for cats with sensitive stomachs
Nutritional Profile: protein >30%, fat >14%, fibre <4%
Minimises adverse food reactions in cats with food intolerances
Supports healthy skin, coat, digestion, and immune system
Pure mackerel hydrolysateiQi—Trusted pet food ingredients (MG Amersfoort, The Netherlands)

Fiordo Austral (Calbuco,
Chile) & ORIVO (Molde, Norway)
Mackerel processing side streams
Spray-dried hydrolysate with over 71% protein and less than 10% ash
Rich in EPA and DHA for added health benefits
Highly digestible with high protein and low ash content
Supports joint health, mobility, and promotes healthy skin and coat in cats and dogs
Pure salmon hydrolysateSalmon processing side streams
Rich in protein (minimum 72%) and essential nutrients.
Can be directly applied to pet food for enhanced nutrition
99% protein digestibility and low ash content
Meets hypoallergenic food standards with broad health benefits for cats and dogs
Pure sardine hydrolysateSardine processing side streams
Spray—dried powder
Hydrolysate that contains over 71% protein
Superior digestibility
High protein and lower ash
Contributes to healthy joints and overall mobility
Pure tuna hydrolysateTuna processing side streams
Fine spray-dried tuna hydrolysate powder offers superior quality compared to rendered meal
Ideal for low-temperature processing applications
Hypoallergenic with highly digestible protein
Rich in omega fatty acids
Supports brain development, ideal for young animals
Veterinary diet—hypoallergenic[129]Dehydrated salmon, yellow peas, hydrolysed salmon protein, buckwheat, salmon oil, hydrolysed salmon gravy, minerals, dried Ascophyllum nodosum, MOS, FOS
Hypoallergenic grain-free food
Nutritional profile: protein 27%, fat 14%, fibre 2%, ash 7.2%
Ideal for dogs with chronic diarrhoea, IBD, or skin issues
Reduces allergic reactions and supports gut health
Table 3. Palatant product produced from fish and other animal and plant sources.
Table 3. Palatant product produced from fish and other animal and plant sources.
ProductManufacturersMarket
Presence
Ingredients/Raw MaterialProduct DescriptionSpecific FeaturesAverage Price Point
(per 200 kg)
PalasuranceTMKemin
(palatants produced in (De Moines, IA, USA, (Cavriago RE), Italy, and Macuco, Valinhos, Brazil)
Europe, USA, AustraliaSalmon, lamb, beefAvailable in both dry and liquid forms
Support the unique protein trend/claim in pet food
Deliver desired level of palatability on the surface of kibble
USD 6.29 (per kg)
PalivateTMEurope, USA, AustraliaDuck, chicken, salmon, soy, lambPalatant for wet pet foods meeting flavour, pH, and texture expectations
Withstand retort process for wet pet food applications
Low inclusion rate in multiple food formats (loaf/chunk/gravy)
Minimise effects to product integrity
>USD 6.29 (per kg)
PaltevaTM & Palteva PTMEurope, USA, AustraliaDuck, chicken, salmon, turkey, lamb, plant proteinsNatural flavour addition to pet diets
Naturally, sourced flavour.
Naturally stabilised and preserved
Animal or plant-based products.
Super premium palatant in canines
>USD 6.29 (per kg)
Fish cream (cat food palatant)Matchwell Nanjing Pet Products Supply Co., Ltd.
(Nanjing, Jiangsu, China)
Available globally—online networkOcean fishBrown liquid,
Crude protein (>15%), ash (<10%), moisture (<70%)
Made from natural materials
Contains compound amino acid that boost immune system
Helps increase cats’ appetite
Enhance cat food palatability
USD 2.00/barrel
Seafood liquid palatantAvailable globally—online networkSeafood
Liquid form, brown in colour
Appeal booster innovation flavouring agent
USD 38.00 per barrel
CPSP® 90 and CPSP® GCopalis Sea Solutions ® (Hauts, France, Symrise®, Teterboro, NJ, USA)Europe and AsiaAll fish—salmon and whitefishFine powder and liquid, soluble fish protein concentrates and hydrolysates, high value ingredient
Strengthen the immune system
Consistent supply of bioavailable proteins
Improves the growth and development of young animals, pets and farmed fish
Unknown
Range of palatants from fish, animal proteinsAFB International ® (St. Louis, MO, USA).Global—US based companyFish, chicken, pork, turkey, duck, lamb, and non-meat proteinLiquid and dry no-GMO, grain free, non-soy, non-wheat, gluten free and clean label palatant products
Helps customers develop new pet food products and improve existing ones
No artificial flavours, colours, or preservatives
Responsibly sourced fish based palatant products
Depends on the form of the product
ProShore—Fish protein Concentrates (FPC) and Fish Protein Isolate (FPI)BioMarine Ingredients Ireland (BII) (Ballybay, Monaghan, Ireland).Western Europe; Eastern Europe; Middle East; Asia; Australia; North America; Africa; Central/South AmericaUnknownHigh quality liquid and powdered ingredients for premium pet food formulations
High-grade FPI and FPC containing 55%–90% protein
Proteins enables repair and building of muscle
Has excellent molecular weight profiles
Unknown
ProAtlantic—FPHUnknownSpray dried, free flowing with a neutral odour, and superior solubility
Contains 95% protein and 0.5% fat
Excellent amino acid profile
Improves growth and repair of the body and maintenance of good health
Consists of low molecular weight peptides
Unknown
PetSavio™ PPP 400Essentia Protein Solutions
(Oak Tree Court Ankeny, IA 50021 USA)
Global—US based companyPlaicePaste form with protein (39%), fat (0.6%), ash (<10%)
Enhances flavour
Brings specific flavour directions
Boosts the pet food formulation’s taste
Reduced fat content
Unknown
PetSavio™ PCP 400Global—US-based companyCodPaste form with protein (40%), fat (1%), ash (9.5%)Unknown
Seasoning Creame for cats (YCG-MT-915)Jiangsu Yichong Biotechnology Company Limited
(Suqian, Jiangsu, China)
Global—China-based companyOcean fish, chicken liver, compound amino acidsCrude protein (>10%), ash (<10%)
Ocean fish significantly trigger a cat’s taste
Helps in increasing appetite
Enhances the palatability
Compound amino acids to boost immune system
Unknown
Seasoning powder for cats (YCF-MT-02)Global—China-based companyOcean fish, chicken liver, Brewer’s Yeast, acidity regulatorCrude protein (>30%), crude fat (<20%), Ash (<45%)
Ocean fish triggers cat’s taste
Cat food flavour enhancer
A tawny powder with good fluidity
Helps in increasing appetite and palatability
Unknown
Fish cream for catsGlobal—China-based companyOcean fish, compound amino acids, Brewer’s Yeast, ProteaseDeep brown colour product, crude protein (>10%), ash (<10%)
Compound amino acids are added to boost immune system
Helps increase appetite and enhance the palatability
Unknown
Table 4. Manufactures of FPHs used in the manufacturer of aquaculture feeds.
Table 4. Manufactures of FPHs used in the manufacturer of aquaculture feeds.
ProductManufacturersIngredients/Raw MaterialProduct DescriptionSpecific FeaturesReferences
Actipal™ (Fish hydrolysate for fish and shrimp)Diana Group S.r.L (Treviso, Italy)Locally sourced, 100% fish side-stream
Made in Costa Rica
Enzymatically hydrolysed
Contains free amino acids and bioactive peptides
Bioactive peptides enhance resistance to stress and pathogens
Boosts immunity and growth hormones
Supports steady feed intake during stress, diet shifts, and growth stages
Maintains optimal health and performance
Gunathilaka et al. [182]
Suma et al. [194]
Nutri™ Tuna
(Tuna hydrolysates)
Locally sourced, 100% side-stream tuna
Made in Indonesia
Contains highly soluble free amino acids and highly bioavailable short peptides
Maintains a high level of feed intake throughout the rearing cycle
Helps in stressful events or changes in feed formula
Improves feed palatability, digestibility, efficiency and growth
Tola et al., [193]
Nutri™ Tuna Soluble extract (TSE)Locally sourced, 100% side-stream tuna
Made in Thailand
Highly digestible proteins
Attractive palatable product for fish and shrimp feeds
Produced using resources from certified fisheries through a standardised process to guarantee optimal freshness and product consistency
Soluble and highly digestible protein source for fish and shrimp feed
Contains a high level of palatable amino acids that increase attractant properties
Suratip et al. [192]
Nutri™ Tuna Liver powder (TLP)Locally sourced, 100% side-stream tuna
Made in Thailand
Offers digestible proteins and high-quality fats
Provides a full range of amino acids and high-quality fats
Naturally palatable and attractive to aquatic species.
Boost palatability and performance in the early growth stages of fish and shrimp
ProShoreBio Marine Ingredients (Ballybay, Monaghan, Ireland)Premium FPHs
Optimal for hatchery and fish farming
It has impressive protein peptide and mineral profiles with superior solubility
Boosts the growth and development rate of cultivated freshwater and saltwater populations
Contains high protein content
Enables growth enhancement and immunity enrichment
Fish Protein Hydrolysate LiquidJanatha Fish Meal & Oil Products (Kota, India) Liquid FPH with ≥40% hydrolysed protein
Enhances growth, survival, and larval development
Reduces the incidence of skeletal deformities
Improves larval quality in both freshwater and marine fish species
Act as feed attractant, thus enhancing the palatability and acceptance of the feed
Siddaiah et al. [191]
Fish Protein Hydrolysate Powder Powder FPH with ≥80% hydrolysed protein
Protein fractions with weight between 1 and 10 kDa enhance feeding activity in fish larvae
Peptides with lower molecular weight (0.2 to 2.5 kDa) beneficially affect larval growth and survival
Enhances the activity of digestive enzymes in the intestine
Fish Soluble Paste (Super attractant) Liquid FPH with ≥60% hydrolysed protein
Helps in growth simulation
Hydrolysed proteins have better palatability, nutrition and bioactivity
Works as a feed attractant and organic fertiliser
ScanPro™Scanbio Marine Group AS (Brattørkaia, Trondheim, Norway)Norwegian salmon, pelagic and white fishEnzymatically hydrolysed and concentrated fish protein is an excellent economical ingredient for feeds
Highly digestible and hypoallergenic
Assimilation of peptides offers maximum nutritional value
Helps in growth simulation of fish larvae
Natural Nautic®Omega Protein Corporation (Reedville, VA, USA)MenhadenFish meal is a high-quality protein made from menhaden and is naturally stabilised with mixed tocopherols
Complete source of highly digestible, low-allergenic protein
Has a superior amino acid profile
An excellent source of omega-3 fatty acids and natural protein for fish
Special Select® MenhadenPremium fish meal is a high-quality protein made from menhaden
SeaLac® Menhaden
Hydrolysed Salmon ProteinSociedad Pesquera Landes Sa (Talcahuano, Chile)By-products of the Chilean salmon industryA perfect ingredient for fish nutrition with 71% protein, 20% lipids, and > 98% digestible
Ideal for improving palatability
Easy to digest
Excellent ingredient with low molecular weight peptides (size < 5000 Da) for fish nutrition
Table 5. Commercial examples of biostimulants including those made from animal protein.
Table 5. Commercial examples of biostimulants including those made from animal protein.
ProductsManufacturerOriginBioactiveApplication MethodPlant TrialsBioactivity
CBio
Biostimulant
C Fish (Charlie Vial)
(Dunkineely, Co. Donegal, Ireland)
White fish/mixed fish composition autolysates and
hydrolysates
Peptides, amino acidsFoliar, pre-planting, irrigationFruits and
vegetables
Enhances plant resistance to biotic (insect, disease) and abiotic (heat, drought) stresses.
RadifarmValagro
(Atessa, Chieti, Italy)
Commercial
formulation
Amino acids, peptides, saponins, betaines, polysaccharides, vitamins, microelementsIrrigation, soil drench, foliar applicationFruits and
vegetables
Stimulates extensive root architecture by accelerating lateral and adventitious root elongation.
MegafolSyngenta
Biologicals
(Basel, Switzerland)
Commercial
formulation
Amino acids, betaines, proteins, vitamins, auxin, gibberellinIrrigation, soil drench, foliar applicationFruits and
vegetables
Supports balanced vegetative growth, boosts overall productivity, and strengthens resilience to environmental stresses such as frost, root asphyxia, weeding, and hail.
BioRootDASA’s elfer®
(Menàrguens, Lleida, Spain)
Plant and mineral derived organic humates, soybean mealPlant derived protein hydrolysateIrrigation, foliar, soil drenchIrrigation, foliar, soil drench of fruits and vegetablesImproves rooting efficiency while increasing chlorophyll and protein content in plant tissues.
ErgonfillK—Adriatica
(Loreo, Italy)
Animal derived protein hydrolysateAnimal protein hydrolysates, cysteine, folic acid, keratin derivativesFoliar, pre-planting, irrigationFruits and
vegetables
Promotes synthesis of indole-3-acetic acid (IAA) and chlorophyll, enhancing nutrient translocation and chelation of macro and micronutrients.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bhati, D.; Hayes, M. From Ocean to Market: Technical Applications of Fish Protein Hydrolysates in Human Functional Food, Pet Wellness, Aquaculture and Agricultural Bio-Stimulant Product Sectors. Appl. Sci. 2025, 15, 5769. https://doi.org/10.3390/app15105769

AMA Style

Bhati D, Hayes M. From Ocean to Market: Technical Applications of Fish Protein Hydrolysates in Human Functional Food, Pet Wellness, Aquaculture and Agricultural Bio-Stimulant Product Sectors. Applied Sciences. 2025; 15(10):5769. https://doi.org/10.3390/app15105769

Chicago/Turabian Style

Bhati, Dolly, and Maria Hayes. 2025. "From Ocean to Market: Technical Applications of Fish Protein Hydrolysates in Human Functional Food, Pet Wellness, Aquaculture and Agricultural Bio-Stimulant Product Sectors" Applied Sciences 15, no. 10: 5769. https://doi.org/10.3390/app15105769

APA Style

Bhati, D., & Hayes, M. (2025). From Ocean to Market: Technical Applications of Fish Protein Hydrolysates in Human Functional Food, Pet Wellness, Aquaculture and Agricultural Bio-Stimulant Product Sectors. Applied Sciences, 15(10), 5769. https://doi.org/10.3390/app15105769

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop