Next Article in Journal
Synthesis, Physical Properties and Application of a Series of New Polyoxometalate-Based Ionic Liquids
Next Article in Special Issue
The Potential of Selected Agri-Food Loss and Waste to Contribute to a Circular Economy: Applications in the Food, Cosmetic and Pharmaceutical Industries
Previous Article in Journal
Flavonoid Glycosides in Brassica Species Respond to UV-B Depending on Exposure Time and Adaptation Time
Previous Article in Special Issue
Prebiotic Impacts of Soybean Residue (Okara) on Eubiosis/Dysbiosis Condition of the Gut and the Possible Effects on Liver and Kidney Functions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Secondary Metabolites: An Opportunity for Circular Economy

Department of Pharmacy and Biotechnology, Alma Mater Studiorum—University of Bologna, Via Irnerio 42, 40126 Bologna, Italy
*
Author to whom correspondence should be addressed.
Submission received: 29 December 2020 / Revised: 12 January 2021 / Accepted: 15 January 2021 / Published: 18 January 2021

Abstract

:
Moving toward a more sustainable development, a pivotal role is played by circular economy and a smarter waste management. Industrial wastes from plants offer a wide spectrum of possibilities for their valorization, still being enriched in high added-value molecules, such as secondary metabolites (SMs). The current review provides an overview of the most common SM classes (chemical structures, classification, biological activities) present in different plant waste/by-products and their potential use in various fields. A bibliographic survey was carried out, taking into account 99 research articles (from 2006 to 2020), summarizing all the information about waste type, its plant source, industrial sector of provenience, contained SMs, reported bioactivities, and proposals for its valorization. This survey highlighted that a great deal of the current publications are focused on the exploitation of plant wastes in human healthcare and food (including cosmetic, pharmaceutical, nutraceutical and food additives). However, as summarized in this review, plant SMs also possess an enormous potential for further uses. Accordingly, an increasing number of investigations on neglected plant matrices and their use in areas such as veterinary science or agriculture are expected, considering also the need to implement “greener” practices in the latter sector.

Graphical Abstract

1. Introduction

According to the United Nations [1], the global population is expected to increase from 7.7 billion (2019) to 9.7 billion in 2050. This prospect raises several concerns about the global consumption of biomass, fossil fuels, metals, and minerals, which should double [2], and annual waste production, following the current trend, will increase by 70% in the next 40 years [3]. Undoubtedly, these premises challenge the move toward a more sustainable development.
In this scenario, wastes and by-products from the food and agricultural industries are gaining international attention, not only for the issues associated with pollution, but also to overcome the paradox of 820 million people suffering from hunger and malnutrition while others are dealing with food over-consumption and related diseases, together with increasing food waste production [4,5]. Moreover, food waste and overproduction imply unnecessary exploitation of the environment and natural resources such as carbon, water, and land. The land footprint estimated by the FAO (Food and Agriculture Organization of the United Nations) in 2013 [6] revealed that almost 30% of the world’s agricultural lands are used to produce food that is ultimately lost or wasted, determining additional pollution and greenhouse gases emissions to no actual purpose. Furthermore, food overproduction and bad waste management also have a negative impact on the economy, with money loss at different levels of the supply chain. Conversely, the reduction of food loss and waste generation has the potential to generate considerable economic value.
ReFED (Rethink Food waste through Economics and Data), a multi-stakeholder nonprofit committed to reducing food waste, identified 27 solutions, which were grouped into three categories: prevention, recovery, and recycling [7]. Although the associated incomes cannot be generalized for all countries, following these solutions, USD 100 billion is expected over 10 years in U.S. According to this program, the solutions focused on preventing waste production account for over 75% of the total, while 23% relate to waste recovery, and only 2% to recycling [5]. This proposal is consistent with the EPA (Environmental Protection Agency) “waste hierarchy,” which grades waste prevention as the most preferable option (Figure 1) [8].
In this scenario, the idea of circular economy took shape, promoting the shift from a linear economic scheme of “take-make-use-dispose” to a circular model, employing reuse, sharing, repair, refurbishment, remanufacturing, and recycling to create a closed-loop system, attempting to minimize the use of resource inputs and the generation of waste, pollution, and carbon emissions [9,10]. In Europe, the most relevant strategy launched in order to attain a sustainable development is the Circular Economy Action Plan, according to which waste materials and energy become input for other industrial processes or regenerative resources for nature (i.e., compost) [11]. In order to facilitate and stimulate the implementation of these guidelines, the scientific community is paying increasing attention to the valorization of industrial and agricultural wastes and by-products, in particular, those derived from plants. These neglected matrices are often manufactured for biofuel production, such as methane or ethanol [12]. However, according to the proposed “waste hierarchy” (Figure 1), energy recovery is a less preferable strategy of waste management, compared to others. Moreover, although biofuel is an alternative to petroleum-derived fuels, its sustainability is quite controversial [13,14].
Actually, the high added-value molecules still contained in plant wastes and by-products offer a wide spectrum of possibilities for their valorization and reuse, as foreseen by circular economy [15,16]. For example, from agri-food wastes it is still possible to extract macromolecules such as nucleic acids [17], pectins [18], cellulose material [19], and enzymes such as bromelain, which is derived from pineapple residues and extensively used as a pharmaceutical and meat tenderizer [20]. Primary metabolites (i.e., organic acids, amino acids, carbohydrates) can also be obtained from plant waste material and used for different purposes [21]. Plants in particular produce secondary metabolites (SMs), which are not directly involved in the basic functions of growth, development, and reproduction of the organism, but are essential for long-term survival and play multiple roles, including defense against predators or attraction of pollinators [22]. SMs are endowed with numerous biological activities, making them also extremely important for human health and well-being. Moreover, due to their chemical and biological properties, SMs have also found application in many other fields, serving as pigments, cosmetics, antifeedants and so on [23].
SMs are usually classified according to their biosynthetic pathways in three principal groups: phenolics, terpenes, and alkaloids [24]. These phytochemicals are characterized by enormous chemical and biological diversity and, in addition to being species-specific and organ-specific, their production depends on many biotic and abiotic factors.
One of the most common examples of SM recovery from plant by-products is from fruit peel generated from industrial processes [25], as in the case of citrus peel, from which essential oils [26] as well as phenolic compounds [27] are extracted. Phenolic acids, flavonols, and catecholamines are obtained also from banana peel [28], while carotenoids such as lycopene are usually obtained from tomato peel and other industrial tomato by-products [29].
Interestingly, SMs are potentially present in all plant organs, offering several possibilities for the valorization of wastes from plant cultivation. In fact, in the agricultural sector, only a few plant organs are harvested and fully consumed, generating numerous wastes at different levels of the supply chain. In this context, scientific investigations aimed at identifying the bioactive compounds contained in these neglected matrices play a pivotal role in laying the basis for their valorization.
Aimed at facilitating and encouraging research projects focused on plant wastes/by-products and circular economy implementation, the current review provides an overview of the most common SMs present in plant matrices, their classification and bioactivities, and the consequent potential application in different fields of the waste material containing these compounds. Consequently, in this work 99 publications (from 2006 to 2020) focused on plant waste/by-product valorization were reviewed and tabulated in order to schematize the state of the art on this topic and offer the opportunity to easily extrapolate information for the design of new studies on neglected plant material and its reuse in a circular economy perspective.

2. Polyphenols

2.1. Chemical Structure and Classification

Polyphenols are one of the largest and most complex classes of SMs produced by plants, derived from shikimate biosynthesis pathway, which provides precursors for aromatic molecules. Based on the biosynthetic pathway, the number of aromatic rings, carbon atoms, and hydroxyl groups, they are divided into different sub-classes such as: simple phenols, phenolic acids, flavonoids and tannins (Figure 2).
Simple phenols are constituted by a single benzenic ring (C6) linked to a hydroxyl group such as resorcinol, orcinol, catechol, guaiacol, hydroquinone, and phloroglucinol.
Phenolic acids present a carboxylic group among the substituents on the benzene ring, and they are generally divided into benzoic acid derivates (C6-C1) (i.e., gallic acid, vanillic acid, syringic acid) and hydroxycinnamic derivates (C6-C3) (i.e., caffeic acid, ferulic acid and coumaric acid).
Flavonoids (C6-C3-C6) are generally constituted by two benzenic rings (A ring and B ring) linked by a chain made of three carbons, often condensed into a pyranic ring (C ring). Given their complexity, flavonoids are divided into other sub-classes such as: chalcones, dihydrochalcones, aurones, flavones, flavonols, dihydroflavonol, flavanones, flavanol, flavandiol or leucoanthocyanidin, anthocyanidin, isoflavonoids. Moreover, flavonoids exist in the aglycone form or as glycoside derivatives [30].
Tannins are high molecular weight polyphenols, usually distinct into condensed tannins (proanthocyanidins), which are polymers constituted by flavonoids units, and hydrolysable tannins characterized by a monosaccharide, normally D-glucose, esterified with one or more molecules of gallic acid (gallotannins), or ellagic acid (ellagitannins). Hydrolysable tannins are more labile in acid, alkali, or hot water than condensed tannins [30].

2.2. Distribution in Plants and Biological Activities

Polyphenols are widely distributed in all plant organs [31]. In general, phenolic acids are found in seeds, leaves, roots, and stems [32], while flavonoids are prominently found in aerial parts, and tannins in roots, bark, and seeds [33].
This heterogeneous group of SMs plays different roles in plants. For instance, flavonoids confer color to the flowers, attracting insects and promoting pollination [34]; some polyphenols are deterrents for herbivores [35], others are very important for UV-protections and for their antioxidant properties [36,37].
Polyphenols, especially flavonoids, are considered important molecules for the human diet and, consequently, are often proposed as ingredients in food supplements and nutraceuticals. In fact, in addition to being antioxidant agents, they are endowed with several other biological activities [38]. For instance, flavonoids were proven active as anti-inflammatory [39], anticancer [40,41], antihypertensive [42], microcirculation improving [43], and hypolipidemic agents [44]. They proved also interesting as active ingredients in the cosmetic field [45,46,47] and as natural dyes [48].
Phenolic acids are naturally found in fruits and vegetables, and are endowed with a wide spectrum of bioactivities such as: antidepressant [49], antihypertensive [50], anti-inflammatory [51], neuroprotective [51], antihyperglycemic [52], anticancer and anti-diarrheal [53].
Regarding tannins, they are used in the veterinary field as anthelmintic and antimicrobial agents [54,55], as well as in the leather industry for their tanning properties [56]. However, tannins should be used carefully, since in addition to their health-promoting properties, some toxic effects have also been reported [57,58]. Moreover, although tannin-rich ingredients are often added to ruminants’ feed, there is still a lack of information about the interaction between hydrolysable tannins and ruminants’ gastrointestinal microbiota. The fate of hydrolysable tannin metabolites derived from gastrointestinal microbial activity in the animal is still underexplored. It is known that some metabolites derived from hydrolysable tannins (i.e., pyrogallol) have adverse effects on gastrointestinal microbiota and the host animal [59]. These data point out the need of deeper investigation on tannin uses and effects, taking into account that risks and benefits depend on the specific situation and concentration used.

2.3. Polyphenols from Agro-Industrial Wastes and By-Products

Polyphenols being present in a wide spectrum of plant organs, they are easily found in numerous agro-industrial wastes/by-products, offering several possibilities for their valorization. First of all, an excellent polyphenol source is represented by the main by-products of wine production, namely: pomace, skins, and seeds. In particular, molasses seeds have a high flavonoids content, and molasses pomace is rich in tannins [60]. Grape pomace contains gallic acid, syringic acid, vanillic acid, catechin, isoquercitrin, and epicatechin [61]. The presence of these antioxidant metabolites [60,61] makes the wine by-products useful additives for ruminant feed [62]. In addition to the antioxidant potential, pomace also shows anti-cholesterol activity [63], while tannins from pomace are also used as wood adhesive [64].
Another polyphenol source is olive pomace [65], a by-product of the olive oil supply chain. This matrix is rich in tyrosol and its derivatives, and it also contains flavonoids such as rutin, apigenin, luteolin, taxifolin, diosmetin, and quercetin, and phenolic acids such as cinnamic, p-coumaric, caffeic, vanillic, and ferulic acid [66,67]. Due to its polyphenol content, olive pomace is also added to ruminant feed [68].
Moreover, olive wastewater contains polyphenols such as hydroxytyrosol [69,70], whose antioxidant and antibacterial properties make it a useful ingredient for cosmetic formulations [71].
Juice industry by-products such as pomace, skins, and seeds are another potential source of polyphenols. Among them, apple pomace is an example [72], since it contains catechin, epicatechin, chlorogenic acid, procyanidin B2, phlorizin, and gallic acid [73].
Among by-products derived from juice production, noteworthy are also strawberries, blueberries, carrots, and pears. In particular, black currants and chokeberries are the richest in anthocyanins, which are suitable ingredients for animal feed [74,75] and textile dyes [76].
Polyphenol-rich wastes derive also from agricultural remains after harvesting [75], canning, and liquor industries [77].
Regarding the aromatic herb industry, basil, sage, and rosemary generate several kind of wastes resulting from pruning, packaging, or distillation processes. It was proved that wastewater from the distillation of these herbs contains important metabolites such as glycosylated flavonoids and caffeic acid derivatives, above all, rosmarinic acid [78]. Wastewater generated from aescin (a saponin from horse chestnuts) production contains kaempferol and quercetin [79].
By-products containing polyphenols are also generated from the production of soluble coffee. For example, coffee grounds, which is the principal by-product, contain many polyphenols such as condensed tannins [80], chlorogenic acid, p-coumaric acid, ferulic acid, rutin, naringin, resveratrol [81]. By virtue of these SMs, coffee grounds proved active to inhibit seed germination [82]. Another by-product from this supply chain is the silverskin, a thin tegument covering coffee beans that is removed in the roasting process. Silverskin is rich in chlorogenic acid derivates such as dicaffeoylquinic acids and feruloylquinic acids [83,84,85].
Regarding the agricultural sector, chestnuts wastes from Castanea sativa Miller, such as spiny burs, are noteworthy for their polyphenol content [86,87,88,89], including gallic acid and ellagic acid derivatives, together with glycosylated flavonoid [90,91]. By-products derived from chestnut flour production such as chestnut peels are sources of polyphenols, in particular, tannins [86,88,92]. For this reason, chestnut spiny burs and peels are valuable as natural antioxidants and often used in animal feed [86]. Also bud pomace of C. sativa contains cinnamic acids, benzoic acids, flavonols, and catechin [93]. Regarding C. sativa cultivation for wood production, the principal waste is the bark, which is also an important source of tannins [94], similar to other tree barks [95].
Several other polyphenol-rich by-products are known, such as Cocoa shell, the waste of chocolate production, which contains catechin, epicatechin, and gallic acid [96]; the waste of black tea, which shows antioxidant and antimicrobial activity [97]; melon peels, which contain a great amount of polyphenols [98]; larch bark, a source of proanthocyanidin B7, a well-known antioxidant [99]; and maize bran containing ferulic acid [100].

3. Terpenes

3.1. Chemical Structure and Classification

Terpenes, also termed terpenoids or isoprenoids, constitute a large family of natural products extremely diversified in their structure, functions, and properties. Terpenes derive from mevalonic acid biosynthetic pathway. However, since their decomposition generates isoprene units (C5), this compound has been defined as terpene’s basic constituent. For this reason, these SMs are classified based on the number of isoprene units present in their structure, and generally condensed head-to-tail. Following this rule they are divided into hemiterpenes (C5H8), monoterpenes (C5H8)2, sesquiterpenes (C5H8)3, diterpenes (C5H8)4, sesterterpenes (C5H8)5, triterpenes (C5H8)6, etc. [101] (Figure 3).
The broader term of terpenoids is applied to terpene-related molecules that, in addiction to isoprene units, contain other substituents, for instance, oxygenated functional groups [102]. Many other natural substances, such as alkaloids, phenolics, and vitamins, despite deriving from biosynthetic pathways other than mevalonate (i.e., acetate or shikimate), are sometimes classified as meroterpenoids since they present isoprenic moieties in their structure.
Hemiterpenes are quite rare; one example is isoprene itself, which is a volatile compound present in several plants, especially trees. Monoterpenes are generally responsible for essential oils fragrance, some examples are: limonene, borneol, camphor, pinene, cineole, menthol, thymol, carvacrol. Diterpenes have as a precursor geranylgeranyl diphosphate; examples are phytol, which forms the lateral chains of chlorophylls, and taxol isolated from Taxus brevifolia Nutt, which is an important natural antitumor agent. Other renowned diterpenes are labdanes, which are the major components of the resin produced by plants of the Cistaceae family [103], and the gibberellins, important phytohormones regulating plant growth and development [104]. Squalene is triterpenes’ precursor, deriving from numerous compounds, including tetracyclic triterpenes like dammarenes, and pentacyclic triterpenes like lupanes and oleanes. The latter are often found in saponin skeletons, where they are glycosylated with one or more sugar moieties. Saponins are phytogenic biosurfactants, which induce foam formation in aqueous solution, reducing the viscosity of heavy crude oil-in-water emulsions [105]. Triterpenic saponins are found in numerous plants, the most renowned among them are Saponaria officinalis L. (soapwort), Quillaja saponaria Molina, and Gypsophila arrostii Guss. [106], which were, in fact, traditionally used for their soap properties. Saponins are classified on the basis of the differences occurring in their aglycone structure or sugar moiety. For example, on the basis of the sugar moiety they are classified into mono, bi, and tridesmosidic. The type of aglycone and the number of linked sugar residues determine the foam properties and the entity of adsorption at the interface [107].
Steroids can be considered modified triterpenes containing the lanosterol tetracyclic system without the two methyl substituents in C-4 and C-14 positions. Cholesterol is the basic structure of this class. They can be found also as saponins (steroidal saponins), such as cardioactive glycosides.
Tetraterpenes include carotenoids, while high molecular weight isoprenoid polymers (higher than C40) are found in natural rubbers produced, for example, by trees belonging to the Euphorbiaceae and Sapotaceae families [108].

3.2. Distribution in Plants and Biological Activities

Terpenes may serve a wide spectrum of functions in plants, including attracting pollinators or protecting injured tissues from herbivores, insects, and parasites attack. According to this biological role various monoterpenes are toxic to insects [109], fungi [110], and bacteria [111]. In addition, steroidal saponins such as cardenolides are toxic to many animals through their inhibition of Na+/K+-ATPases. However, the same property makes them useful as therapeutic agents, in carefully regulated doses, to slow down and strengthen the heartbeat.
Due to their numerous bioactivities, terpenes find various applications in several industrial sectors, such as pharmaceutical, food, cosmetic, perfumery, and agricultural, and are used as drugs, food supplements, flavors, fragrances, biopesticides, etc.
First of all, the peculiar fragrance of many monoterpenes, present principally in aromatic plant essential oils, makes this class of compounds extremely important for food, aromatherapy, and perfumes. In addition to the fragrance, monoterpenes from essential oils, such as thymol, thymine, and carvacrol (found principally in plants belonging to the Lamiaceae family) proved interesting also for their numerous biological activities, for instance, their potential in the treatment of disorders affecting respiratory, nervous, and cardiovascular systems, and as antimicrobial and antioxidant agents [112].
Also labdane-type diterpenes are useful for the perfume industry, finding application as fixatives in high-end perfumes. Specifically, a fixative is a material with low volatility that provides long-term scent, aids in mixing with the other materials, and extends the shelf life of the perfume. The resin obtained from some plants of the Cistaceae family is one of the most common sources of labdane diterpenes used as perfume fixatives [113].
Saponins are extremely important for the food industry. In fact, many of the processed foods, including baked goods, ice creams, sauces, desserts, and drinks, contain dispersions such as emulsions and foams used to stabilize, determine, and control texture and rheological properties of these products. Saponins, due to their amphiphilic properties, proved to stabilize food emulsions with less sensitivity to pH, ionic strength, and high temperatures (up to 90 °C) than currently used emulsifiers [114]. Moreover, the increasing consumer demand for plant-based and sustainable emulsifiers and foaming agents makes saponins much requested for the food industry nowadays.
Despite saponins being toxic at high dosages, small quantities have been approved as food additives. The two main commercial sources are Q. saponaria from Chile, whose saponins have a prominently triterpenoid structure, and Yucca schidigera Roezl ex Ortgies from Mexico, which contains saponins with a steroid structure [106].
Moreover, saponins at low dosage also showed numerous biological activities important for human healthcare, which are summarized in several review papers [115,116,117]; most relevant among them are the anticancer, cholesterol-lowering, and antiviral properties [118,119,120,121]. Some examples of the therapeutic potential of triterpenic saponins are provided by boswellic and betulinic acids. In fact, the extracts of the resin obtained from incense trees (Boswellia serrata Roxb.), containing the pentacyclic triterpenoid boswellic acid, have been employed as an anti-inflammatory drug [122], and the clinical trials on gum-resin from B. serrata have shown an improvement in the symptoms in patients with osteoarthritis and rheumatoid arthritis [123,124]. Betulinic acid, a naturally occurring pentacyclic triterpene, exhibited a high variety of biological activities [125], including potent antiviral effects [126].
Moreover, saponins are antimicrobial agents active also against bacteria and fungi invading plants [127,128]. The mechanism of these activities is likely based on saponins’ ability to form complexes with sterols present in the membrane of microorganisms and to cause, consequently, membrane perturbation [129,130]. Saponins also exert insecticidal [131] and molluscicidal [132] activities, as well as allelopathic activity toward different plant species [133]. These properties, together with their biological role in plant defense, confer to saponins an enormous potential as natural biopesticides useful for “green” agriculture practices. For example, Trdá et al. [134] found that saponin aescin, in addition to its antifungal effect against crop pathogens, is also able to activate plant immunity (in two different plant species) and to provide salicylic acid-dependent resistance against both fungal and bacterial pathogens.
In the pharmaceutical industry, terpenes are used as excipients to enhance skin penetration of active principles [135] and as therapeutic agents endowed with numerous bioactivities including, as mentioned above, chemo-preventive, antimicrobial, antifungal, antiviral, antihyperglycemic, analgesic, anti-inflammatory, and antiparasitic activities [110,111,118,122]. Among the pharmaceuticals, the anticancer paclitaxel and antimalarial artemisinin are two of the most renowned terpene-based drugs.

3.3. Terpenes and Terpenoids from Agro-Industrial Wastes and By-Products

Due to their numerous bioactivities, terpenes are particularly interesting in the context of waste requalification. Monoterpenes such as thymol and carvacrol are still present in discrete amounts in several by-products derived both from essential oil distillation and from the harvesting of some aromatic plants. For instance, the solid waste residues left after the distillation of leaves and stems of Mexican oregano (Poliomintha longiflora A. Gray) contain thymol and carvacrol, and are, as a result, endowed with antimicrobial activity [136]. Similarly, the monoterpenes limonene and nerol were found in fennel (Foeniculum vulgare Mill.) horticultural wastes [137].
The inedible part (stones, husks, kernels, seeds) from the fruit processing supply chain constitutes a huge portion of the consequent solid waste, which remains underexploited. For example, about one-third of citrus fruit production is industrially processed, with more than 80% used for orange juice production, which generates a huge amount of peel waste [138]. Orange essential oil mostly contains the monoterpene d-limonene (3.8% of orange peel dry weight) [139,140]. This molecule has been used as an ingredient in bio-based functional food, as preservatives for food [141], as well as in cosmetics and aromatherapy massage [142,143].
Moreover, the presence of d-limonene, an anti-fungal and antibacterial agent, makes orange oil a useful ingredient also for bio-pesticide formulations [144]. Finally, it is interesting to notice that waste orange peel, in addition to d-limonene, contains also other bioactive terpenes like linalool, and myrcene [145]. Besides orange peel, these terpenes can be found also in the peel of other citrus fruits such as lemon and several lime species [146], providing a good basis for the exploitation of this kind of wastes.
Regarding diterpenes, more than one million tons a year [147] of residue is produced after steam distillation of pine resin to recover the volatile fraction called turpentine. The consequent by-product is the gum rosin, which is a mixture of resin acids (90–95%) and other neutral compounds. The resin acids, most of which are isomers of each other, can be classified into two main categories: abietic-type (including abietic, neoabietic, palustric, and levopimaric), and pimaric-type (including pimaric, isopimaric, and sandaracopimaric) [148]. Gum rosin is a high-value-added residue, in fact, it is a natural alternative to fossil-based polymers obtained from the heating and evaporation of pine resin [149], as well as a producer of organocatalysts to promote complicated asymmetric industrial synthesis [150].
Several waste matrices contain saponins, which offer an enormous potential for their valorization. An example is provided by sisal (Agave sisalana Perrine), which is the main hard fiber produced worldwide. From its leaves, only the hard fibers (3–5% of total weight) are removed. The remaining 95–97% of the biomass is considered sisal waste, although it contains steroidal saponins, potentially useful for foods, cosmetics, and pharmaceuticals formulations, as well as for soil bioremediation [151].
Saponins from onion skin were found useful as a new natural emulsifier to formulate oil-in-water nanoemulsions by a high-pressure homogenizer [152].
Triterpenic acids such as oleanolic, betulinic, and ursolic acids provide another example of high-value terpenes that can be extracted from agricultural wastes prior to burning for energy production.
For example, oleanolic acid is found in agroforestry waste streams, such as in olive trees (Olea europaea L.), from which tons of wastes and by-products are generated on annual basis, including olive wood and leaves, cake, pomace, kernel, paste. During the production process, assuming a maximum content up to 3.1% of oleanolic acid in the leaves of O. europaea, a large amount of this high-value compound can potentially be extracted, contributing to the integrated valorization of the olive oil production chain [153]. In this context, the terpenes contained in the ethanol extract of olive milled residue showed anti-allergic activity on the cell line of rat basophilic leukemia, supporting the potential valorization of this other olive by-product [154].
Humulus lupulus L. (hops) flowers are used to preserve and give flavor to beer, while hops leaves are usually discarded as a waste. However, hops leaves contain β-caryophyllene, phytol, fatty acids, terpenes, and C5-ring bitter compounds, and the oil obtained from the leaves contains bitter acids, in particular cohumulinic, dehydrocohumulinic, and humulinic acid, possessing antibacterial activity [155].
Melon (Cucumis melo L.) is one of the most popular fruit cultivated in tropical countries and is industrially processed to obtain a wide spectrum of products such as juices, jams, dehydrated pulp, and salads or snacks, with consequent generation of a large number of by-products. However, these matrices (prominently pulp, seed, and peel) are still a good source of carotenoids (C40 tetraterpenoid pigments) like β-carotene, lutein, β-cryptoxanthin, phytoene, violaxanthin, neoxanthin, and zeaxanthin [156]. These substances are used to develop health-promoting functional food since they play an important role in eye photoprotection (provitamin A), improving immune functions, and preventing chronic diseases [157].

4. Alkaloids

4.1. Chemical Structure and Classification

A large number (more than 10,000 molecules) of plant SMs are classified as alkaloids [158]. The presence of nitrogen in the structure is the peculiar chemical feature of alkaloids. However, due to the huge structural diversity, alkaloid classification is extremely challenging. More recent classifications are based on carbon skeletons and/or biochemical precursor. However, this requires compromises in borderline cases, for example, the alkaloid nicotine (from Nicotiana spp.), which contains a pyridine fragment from nicotinamide and a pyrrolidine part from ornithine, could be correctly assigned to two different classes [108].
Historically, alkaloids have been defined as metabolites containing one or more nitrogen atom(s) within heterocyclic ring(s) [159]. However, N-containing compounds where the N atom is not heterocyclic, such as hordenine, ephedrine, colchicine and capsaicin, were further included into this SM group and classified as proto-alkaloids or amino-alkaloids. For this reason heterocyclic N-containing compounds are often regarded as “true alkaloids.” Since “true alkaloids” biosynthetically derive from amino acids, they are classified on the basis of the biogenetic origin (Figure 4).
In particular, the majority of them derive from ornithine, leucine, lysine, tyrosine, tryptophan, histidine, and phenylalanine. More specifically, pyrrole alkaloids derive from leucine; pyrrolidine, tropane, and pyrrolizidine alkaloids from ornithine; quinolizidine, and indolizidine alkaloids derive from lysine; catecholamines, isoquinoline, tetrahydroisoquinoline, and benzyltetrahydroisoquinoline alkaloids originate from tyrosine; indolamines, indole, carboline, quinoline, pyrrolindole and ergot alkaloids come from tryptophan; and imidazole alkaloids from histidine.
As the “true alkaloids,” proto-alkaloids derive from amino acids, and on this basis they are subsequently divided into phenylethylamino alkaloids, pyrrolizidine alkaloids, terpenoid indole alkaloids.
Following this classification criterion, another class of alkaloid, namely “pseudo-alkaloids,” was constituted, including compounds that do not originate from amino acids, while having a nitrogen atom inserted into the molecule by transamination or amination reactions (Figure 4). This letter class includes aromatic alkaloids, ephedra alkaloids, purine alkaloids, sesquiterpene alkaloids such as isoprenoid alkaloids including mono- (from geraniol), di- (from geranylgeranyl-PP), and triterpene (from cholesterol) derivatives, these latter called steroidal alkaloids [158].
Regarding steroidal “pseudo-alkaloids,” they are often glycosylated (glycoalkaloids). These peculiar compounds are produced in more than 350 plant species, mainly from Solanaceae and Liliaceae families [160]. They consist of a C27 cholestane skeleton (aglycone), where the –OH in position -3 is glycosylated by one to five monosaccharides, such as D-glucose, D-galactose, D-xylose and L-rhamnose.
Other peculiar alkaloids are the polyamine alkaloids (derivatives of putrescine, spermidine, and spermine), peptide and cyclopeptide alkaloids [161,162].

4.2. Distribution in Plants and Biological Activities

Alkaloids are an enormous group of phytochemicals of ecological importance, and possess a number of toxicological, pharmacological, nutritional, and cosmetic activities. Alkaloids are extremely abundant in flowering plants (Angiospermae), with a wide distribution in all organs such as leaves, flowers, seeds, roots, stems, fruits, bark, and bulbs. However, the presence and the distribution of these metabolites depend on the phase of plant life cycle, and strongly vary according to plant species, which produce different types of alkaloids, accumulated in various organs [163,164].
Alkaloids play numerous roles in plants, due to their involvement in defense [165,166,167], allelopathy [168], seed dispersal, and pollinator attraction [169,170]. Consistent with their defensive role, the highest alkaloid content is often found in plant reproductive organs [171]. In fact, many alkaloids are toxic to different organisms, protecting plants from pathogens and preventing non-specialist herbivore grazing [165,167]. In contrast, other alkaloids are essential for plant–pollinator interactions, increasing the number of pollinator visits, thus favoring plant reproduction [169,170].
Alkaloids have been historically used as drugs, and they remain very important in this context [172]; an example is provided by morphine from poppy straw, which is one of the most used analgesics today.
The biological activities of the principal sub-classes of alkaloids have been well summarized by Debnath et al. [159]. To cite some examples, quinine and quinidine, two quinolone alkaloids obtained from the bark of Cinchona officinalis L. (Rubiaceae), are very important historically used antimalarial drugs [173]; ephedrine, an adrenergic amine from the plants of genus Ephedra (Ephedraceae family), is used in many pharmaceutical preparations such as bronchodilators for asthmatic and allergic conditions, and to prevent low blood pressure during spinal anesthesia [174]; vinblastine and vincristine, two indole alkaloids extracted from Catharanthus roseus (L.) G. Don (Apocynaceae) [175], are renowned antitumor drugs.
Many other alkaloids have been studied for their promising bioactivities, for example, catuabine, a tropane alkaloid obtained from the bark of Trichilia catigua A. Juss. (Meliaceae) is endowed with antidepressant-like effects on the forced swim model of depression in mice and rats [176]; berberine, occurring in roots and stem-bark of different species of Berberis (Berberidaceae), showed anti-diabetic effect in rodent models of insulin resistance [177], and anti-hypertensive, anti-inflammatory, antioxidant, antidepressant, hepatoprotective activity, and anti-cancer activity [178,179].
Moreover, since some alkaloids possess psychotropic properties, they have found a role in social and ceremonial activities, as well as being important for popular spices and drinks, like the alkaloid caffeine, which is present in coffee [180]. Specifically, caffeine is a methyl-xanthine alkaloid, and its most important biological sources are Coffea arabica L. and Camelia sinensis (L.) Kuntze (leaves) (Theaceae). Caffeine is the most widely consumed stimulant drug in the world. It is also used in cold medications, analgesics, slimming agents, and cosmetics.
Moreover, consistent with their defense role, many alkaloids exhibit insecticidal [181,182], and fungicidal activity. For example, a piperidine alkaloid, pipernonaline, isolated from the hexane fraction of Piper longum L., showed potent fungicidal activity against the phytopathogen Puccinia recondita [183]; Coptis japonica (Thunb.) Makino extracts and the contained alkaloids (isoquinoline alkaloids, berberine chloride, palmatine iodide, and coptisine chloride) expressed fungicidal activities against several phytophatogens, namely: Botrytis cineria, Erysiphe graminis, Phytophthora infestans, Puccinia recondita, Pyricularia grisea, and Rhizoctonia solani in in vivo plant models [184].
Relevant in this context are also the alkaloids produced by the Solanaceae family, which have an enormous potential to deliver new chemicals for crop protection. In fact, more and more of these compounds, or mixtures of them, are being identified as pest control agents, especially against insects, fungi, and mites [185]. The Solanaceae family belongs to the most important plant taxa, particularly in terms of food production (i.e., tomatoes and potatoes). Tomato and potato are the best known and most widely used plants of this group, and they constitutively synthesize low levels of many different glycoalkaloids. These natural toxicants (stress metabolites) have insecticidal and fungicidal properties and, since naturally occurring pesticides are often biosynthesized when plants are under stress, injuries on plant tissues promote the synthesis of higher concentrations of these compounds.

4.3. Alkaloids from Agro-Industrial Wastes and By-Products

The neglected matrices containing alkaloids have a high requalification potential by virtue of the numerous bioactivities possessed by these metabolites.
For instance, caffeine, a methyl-xanthine alkaloid, besides being an important ingredient for energy drink industry, it is also relevant for cosmetics. Caffeine is, in fact, used for cellulitis reduction [186,187], and to prevent skin aging through both antioxidant activity and inhibition of skin remodeling enzymes [188].
Among the industrial by-products, a source of caffeine is represented by spent coffee grounds (from coffee bars), which still have an amount of caffeine in the range of 5.99–11.50 mg/g of dry matter [189].
Another source of methyl-xanthines (including caffeine and theobromine) are cocoa shells [96,190]. This is of particular interest, considering that the high amount of cocoa bean shell produced per year is generally disposed as waste and underutilized as fuel for boilers, animal feed, or fertilizer [191].
Another class of alkaloids, extremely relevant in view of waste valorization, is represented by glycoalkaloids, mainly produced by plants of the Solanaceae family. The major components of the glycoalkaloid family are α-solanine and α-chaconine found in potato plants (Solanum tuberosum L.), and solasonine and solamargine found in eggplants (Solanum melongena L.), whereas α-tomatine and dehydrotomatine are spirosolane-type glycoalkaloids found in tomato plants (Lycopersicon esculentum Mill.) [185].
These compounds are agrochemically important, in fact, their defensive role in plants makes many of them (i.e., α-tomatine, α-chaconine, α-solanine, and various Solanum spp. extracts) endowed with insecticidal activity against various insect species. In particular, both α-chaconine and α-solanine decrease insect feeding, delay their development, affect reproduction, and alter insect enzyme activity [185].
Although the majority of the toxicity studies have been focused on tomatoes and potatoes due to the economic importance and availability of these species, acute toxicity to insects has also been reported in plant extracts belonging to other genera, such as Piper, Datura, and Withania [185]. Notably, from Datura stramonium L. it is possible to extract a great variety of bioactive alkaloids, saponins, sterols, and polyphenols. This plant, as well as several other plants containing alkaloids, is often considered an agricultural waste [155], since it is invasive and its presence in cultivated fields is undesired.
Potato peels, a by-product of the industrial production of potato fries, chips, and flour, are a significant part of the annual worldwide production of about 1.3 billion tons of food waste [192]. In this context, alkaloids extracted from potato peels proved to be antioxidant [193] and antiprotozoal against pathogenic Trichomonad strains that infect humans, farm animals, and felines [192].
Alkaloids from plant wastes have been also tested for biological activities eventually useful for human healthcare, for instance, tomato (L. esculentum, Solanaceae) leaves alkaloids proved promising for Alzheimer disease treatment [194].

5. Extraction Techniques

5.1. Conventional Extraction Procedures and New Prospective for Solvents

Natural products are characterized by great diversity, which implies the necessity to develop specific extraction methods according to the starting raw material and selected metabolite(s). Generally, the raw matrix is subjected to an air drying process, dried in an oven or freeze-dried, and subsequently ground to create a homogeneous sample before the extraction [195,196,197]. Sometimes, before polyphenol extraction, the matrix is defatted with a non-polar solvent like n-hexane [70].
One of the most common extraction methods is solid–liquid extraction using water or organic solvents [65,84]. In some cases the extraction is performed through a Soxhlet extractor, which allows an efficient recycle of the solvent, which can be used in small amounts to extract a significant quantity of plant material [195,198].
If the starting material is a liquid, as for wastewaters, a liquid–liquid extraction, for instance, using ethyl acetate, is preferable [70,71]. Liquid/liquid partition has often been employed also to extract alkaloids and terpenes [199].
Sometimes the extraction can be facilitated by acidic or basic conditions under heating, as in the case of tannins, which can be extracted by maceration under reflux, using aqueous solvents, slightly alkaline at 85 °C. In particular, hydrolysable tannins are well extracted using a blend of 1% NaOH for 240 min, while condensed tannins are extracted with a blend of 1% Na2SO3 for 960 min [92].
Essential oils are obtained by stem distillation, which extracts preferentially volatile monoterpenes, while resins, which are generally more enriched in diterpenes, are extracted using organic solvents.
However, toxic solvents, like n-hexane or methanol, are unsuitable to extract products conceived for food or health uses. An option is to replace methanol with ethanol, which makes safer and “greener” the process. For example, an excellent polyphenol yield is obtained thought maceration in an hydro-alcoholic blend of 50 or 60% ethanol for 30 min at a temperature between 60 °C and 80 °C, under reflux or simply under continuous stirring at room temperature [61,189].
However, in search of more efficient and environmental friendly alternatives to extract high-value molecules, ionic liquids (ILs) and natural deep eutectic solvents (NADES) have been proposed [200].
ILs are ionic species (organic salts), fluids or solids at room temperature, consisting of an organic cation (i.e., ammonium, imidazolium, pyridinium, phosphonium) and an anion (i.e., bromide, chloride, tetrafluoroborate, hexafluorophosphate). Due to their ionic nature, ILs possess negligible vapor pressure and high solvation ability, and they offer a wide spectrum of extraction abilities and selectivity [201]. Compounds such as flavonoids, alkaloids, phenolics, terpenoids, phenylpropanoids, and polysaccharides have been successfully extracted by ILs [202]. Moreover, recently, the possibility of using aqueous solutions of ILs instead of their pure forms led to a substantial improvement in their extraction efficiency and cost reduction. These solvents have already been used for the extraction of waste from natural products. For instance, solutions of surface-active ILs in water were used to efficiently extract triterpenic acids from apple peels [203], oleanolic acid from O. europaea [138], anthocyanins from grape pomace and peel of eggplant [204,205].
Natural deep eutectic solvents (NADES) are considered as a specific class of liquids present in living cells, where they play an important role in biosynthesis, transport, and storage of compounds with intermediate polarity [206]. NADES are composed of hydrogen bond donors (HBDs) and hydrogen bond acceptors (HBAs) mixed together. The usual HBAs are nontoxic quaternary ammonium salts or amino acids, while HBDs are organic acids or carbohydrates. Alcohol, amine, aldehyde, ketone, and carboxylic groups can be used as both HBAs and HBDs. There are a huge number of natural metabolites, which can be combined to prepare NADES, making the latter a high versatility tailor-made class of solvents. NADES have numerous favorable properties, they are liquid state within a wide temperature range, manifest chemical and thermal stability, are non-flammable and non-volatile, nontoxic as well as having sustainable “green” properties. On this basis, NADES have been used to extract a wide range of natural compounds, including phenolics, alkaloids, saponins, anthraquinones, terpenoids, polyunsaturated fatty acids, and photosynthetic pigments [207,208]. NADES and their perspectives in the agri-food sector were extensively reviewed by Mišan et al. [206]. Regarding the employment of NADES in the extraction of plant-driven industrial wastes, they have been used to efficiently obtain phenols from by-products of the olive oil industry, and the onion, tomato, potato, orange, and pear canning industries [209,210].

5.2. Non-Conventional Extraction Methods

Different extraction methods have been proposed in order to shorten, improve, and obtain greener procedures for natural metabolite extraction. Several works have been dedicated to summarize these techniques and their use to extract bioactive metabolites from natural wastes [211,212]. Ultrasonic-assisted extraction (UAE) is an effective extraction technique for a wide range of compounds from different types of matrices. Due to the cavitational effects leading to cell wall disruption the release of the target compound(s) from the biomass is favored [213]. In addition, ultrasounds also use the oxidative energy of radicals created during sonolysis to make more efficient the extraction process [214].
This results in a shortening of the extraction process with low consumption of solvents and high product yields. This method has been used to efficiently extract different classes of metabolites from several plants [213].
Microwave-assisted extraction (MAE) reduces extraction time with a minor consumption of solvents and minimum degradation of target compounds. It is based on microwave energy, which heats the solvents and increases the internal pressure inside the cell, helping the disruption of the cellular wall and the release of active compounds to the solvent [215]. Combined with inert atmosphere to avoid polyphenol degradation, this is a sustainable technique [216] that allows to work at high temperature (150 °C), obtaining very good yields of extracts endowed with a high antioxidant power [217]. This method can be optimized using a solution at pH 12, which facilitates the extraction of polyphenols [195].
Supercritical fluid technology (SFT) is based on supercritical fluids, among them the most used is supercritical carbon dioxide (CO2), generated by increasing the pressure and the temperature of the liquid/gas above the critical point. These fluids have liquid-like solvent power and gas-like diffusivity, resulting optimal to extract compounds from plant matrices, and yield solvent-free extracts by the reduction of CO2 pressure, which allows to easily remove it. This technique has been extensively applied to extract plant metabolites from waste. For instance, STF was used to efficiently extract tocopherols and carotenoids rich oil guava seeds, which are endowed with good antioxidant activity [218].
Supercritical fluid extraction (SFE) was proven not suitable for polyphenol extraction [198], however it could be useful in a first phase where fatty acids are removed from the matrix, and polyphenols might be extracted subsequently with the help of a co-solvent [197].
Among the green technologies, pressurized liquid extraction (PLE), also called accelerated solvent extraction (ASE), is a fully automated technique that combines high temperature and pressure with the use of liquid solvents. This technique proved efficient to extract different matrices such as food [219] medicinal plants [220], and environmental samples [221]. PLE improves extraction yield, significantly reducing time and solvent consumption, [222] and it can be used to extract molecules of different polarity, such as phenolic compounds, carotenoids, and essential oils [222]. An optimized PLE protocol for the extraction of water-soluble molecules is represented by hydrothermal extraction, which uses water at high temperature and high pressure and has been employed to extract spent coffee grounds [223].

6. Trends in Publications Focused on Plant Wastes Valorization

The number of scientific publications dealing with the valorization of waste derived from plants have notably increased in the last three years (Figure 5) [224]. This trend reflects the rising global interest in circular economy and requalification of neglected plant matrices.
This review carried out a bibliographic survey, taking into account 99 scientific articles on this topic published from 2006 to 2020. Table 1 summarizes the information about waste type, its plant source, industrial sector of provenience, contained secondary metabolites, reported bioactivity, and potential use for its valorization. The survey included 64 plants generating wastes, mainly derived from the food and beverage industries, followed by the herbal, agriculture, and forestry industries, and only one report was dedicated to a plant waste from the perfume industry. Moreover, according to this survey, wastes from 16 plant species originated from more than one industrial sector. This is easily understandable taking into account that the majority of the industrial products undergo different steps in the production chain. For example, different kinds of wastes are generated from Castanea sativa cultivated for nuts production, and these wastes originate at different level of the production chain. In particular, spiny burs are the main agricultural residue from tree cultivation, while husks are discarded by the food industry when producing chestnut flour. The beverage industry generates also a considerable amount of wastes/by-products from plant origin, most prominent among them fruit pomace and or/peel. Nevertheless, a conspicuous number of wastes are also produced during the previous steps of the supply chain, namely, the fruit production itself leads to several agricultural residues. An example is provided by the wine production chain, where grape pomace is the main final waste, even though the cultivation of Vitis vinifera L. itself generates residuals like green prunings. In this case, as highlighted also by our survey, due to their content in polyphenols, flavonoids, and tannins, grape pomace and skin are strongly valorized by-products (12 publications were focused on the requalification of these wastes). However, as pointed out by Acquadro et al. [225], green pruning might be exploited for its phenolic content.
Focusing on the specific classes of SMs considered in the examined papers, polyphenols emerged as the most investigated one. Out of 64 plants mentioned in the reviewed articles, the generic presence of polyphenols was reported in 31 of them. Moreover, several investigations were focused also on specific polyphenol sub-classes, such as flavonoids (found in 21 plants) and phenolic acids (found in 18 plants).
Conversely, terpenes and alkaloids resulted as less investigated; in fact, they were mentioned as important compounds in only 12 and 7 plant species, respectively. In the context of waste valorization, the main bioactivity related to polyphenols was the in vitro antioxidant activity, which is common also to terpenes [29] and alkaloids [189]; this encourages further works dedicated to these latter classes of plant metabolites, often contained in wastes/by-products.
According to this bibliographic survey, terpene saponins resulted as the less explored metabolite from plant waste matrices. Their presence was, in fact, reported only in waste constituted by onion skin and A. sisalana leaves. As mentioned above, saponins have numerous applications and bioactivities, among them their possible use in soil bioremediation. Moreover, this survey pointed out that only few waste matrices were studied for their potential use in the agriculture sector (Figure 6).
In particular, Anacardium occidentale L. nut shells and Castanea sativa burs proved good sources of molecules endowed with pesticide and anti-fungal properties, respectively. On this basis they could find application as crop protection agents. Spent coffee ground showed allelopathic activity, which is useful for herbicide products.
As it is evident in Figure 6, the majority of plant wastes were valorized proposing potential exploitation in the area here called “human healthcare and food” (HHF) (including cosmetic, pharmaceutical, nutraceutical, and food additive use). However, since plant SMs are endowed with a wide range of chemical and biological properties (as summarized in this review), it is expected that, following the actual trend, wastes and by-products of plant origin will be increasingly investigated also for their potential use in other areas of application, especially in agriculture, considering the need to implement “greener” practices in this sector.

Author Contributions

Conceptualization, I.C. and M.M.; writing—original draft preparation, I.C., M.M., P.T., L.M., F.P.; writing—review and editing, I.C., M.M., F.P.; supervision, M.M and F.P.; funding acquisition, F.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are openly available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. United Nations. World Population Prospects 2019; United Nations: New York, NY, USA, 2019. [Google Scholar]
  2. OECD. Global Material Resources Outlook to 2060: Economic Drivers and Environmental Consequences; OECD Publishing: Paris, France, 2019. [Google Scholar]
  3. Kaza, S.; Yao, L.; Bhada-Tata, P.; Van Woerden, F. What a Waste 2.0: A Global Snapshot of Solid Waste Management to 2050; Urban Development; World Bank: Washington, DC, USA, 2018; p. 168. [Google Scholar]
  4. Mirabella, N.; Castellani, V.; Sala, S. Current options for the valorization of food manufacturing waste: A review. J. Clean. Prod. 2014, 65, 28–41. [Google Scholar] [CrossRef] [Green Version]
  5. FAO (Food and Agriculture Organization of the United Nations). The State of Food and Agriculture 2019; FAO: Rome, Italy, 2019. [Google Scholar]
  6. FAO (Food and Agriculture Organization of the United Nations). Food Wastage Footprint: Impacts on Natural Resources–Summary Report; FAO: Rome, Italy, 2013. [Google Scholar]
  7. ReFED (Rethink Food Waste through Economics and Data). A Roadmap to Reduce US Food Waste by 20 Percent; ReFED: Berkeley, CA, USA, 2016. [Google Scholar]
  8. The NSW Environment Protection Authority Website. Available online: https://www.epa.nsw.gov.au/your-environment/recycling-and-reuse/warr-strategy/the-waste-hierarchy (accessed on 1 October 2020).
  9. Barreiro-Gen, M.; Lozano, R. How circular is the circular economy? Analysing the implementation of circular economy in organisations. Bus. Strateg. Environ. 2020, 29, 3484–3494. [Google Scholar] [CrossRef]
  10. Morseletto, P. Targets for a circular economy. Resour. Conserv. Recy. 2020, 153, 104553. [Google Scholar] [CrossRef]
  11. European Commission Website. Available online: https://ec.europa.eu/environment/circular-economy/ (accessed on 9 January 2021).
  12. Momayez, F.; Karimi, K.; Taherzadeh, M.J. Energy recovery from industrial crop wastes by dry anaerobic digestion: A review. Ind. Crops Prod. 2019, 129, 673–687. [Google Scholar] [CrossRef]
  13. Solomon, B.D. Biofuels and sustainability. Ann. N. Y. Acad. Sci. 2010, 1185, 119–134. [Google Scholar] [CrossRef]
  14. Baudry, G.; Delrue, F.; Legrand, J.; Pruvost, J.; Vallée, T. The challenge of measuring biofuel sustainability: A stakeholder-driven approach applied to the Fr.0ench case. Renew Sust. Energ. Rev. 2017, 69, 933–947. [Google Scholar] [CrossRef]
  15. Ben-Othman, S.; Jõudu, I.; Bhat, R. Bioactives from Agri-Food Wastes: Present Insights and Future Challenges. Molecules 2020, 25, 510. [Google Scholar] [CrossRef] [Green Version]
  16. Sagar, N.A.; Pareek, S.; Sharma, S.; Yahia, E.M.; Lobo, M.G. Fruit and vegetable waste: Bioactive compounds, their extraction, and possible utilization. Compr. Rev. Food Sci. Food Saf. 2018, 17, 512–531. [Google Scholar] [CrossRef] [Green Version]
  17. Bosco, F.; Casale, A.; Gribaudo, G.; Mollea, C.; Malucelli, G. Nucleic acids from agro-industrial wastes: A green recovery method for fire retardant applications. Ind. Crops Prod. 2017, 108, 208–218. [Google Scholar] [CrossRef]
  18. Nagel, A.; Winkler, C.; Carle, R.; Endress, H.U.; Rentschler, C.; Neidhart, S. Processes involving selective precipitation for the recovery of purified pectins from mango peel. Carbohydr. Polym. 2017, 174, 1144–1155. [Google Scholar] [CrossRef]
  19. Ibrahim, M.M.; El-Zawawy, W.K.; Jüttke, Y.; Koschella, A.; Heinze, T. Cellulose and microcrystalline cellulose from rice straw and banana plant waste: Preparation and characterization. Cellulose 2013, 20, 2403–2416. [Google Scholar] [CrossRef]
  20. Banerjee, S.; Ranganathan, V.; Patti, A.; Arora, A. Valorisation of pineapple wastes for food and therapeutic applications. Trends Food Sci. Tech. 2018, 82, 60–70. [Google Scholar] [CrossRef]
  21. Patsalou, M.; Menikea, K.K.; Makri, E.; Vasquez, M.I.; Drouza, C.; Koutinas, M. Development of a citrus peel-based biorefinery strategy for the production of succinic acid. J. Clean. Prod. 2017, 166, 706–716. [Google Scholar] [CrossRef]
  22. Demain, A.L.; Fang, A. The natural functions of secondary metabolites. In History of Modern Biotechnology I, Advances in Biochemical Engineering/Biotechnology; Fiechter, A., Ed.; Springer: Berlin/Heidelberg, Germany, 2000; Volume 69, pp. 1–39. [Google Scholar]
  23. Kabera, J.N.; Semana, E.; Mussa, A.R.; He, X. Plant secondary metabolites: Biosynthesis, classification, function and pharmacological properties. J. Pharm. Pharmacol. 2014, 2, 377–392. [Google Scholar]
  24. Bourgaud, F.; Gravot, A.; Milesi, S.; Gontier, E. Production of plant secondary metabolites: A historical perspective. Plant Sci. 2001, 161, 839–851. [Google Scholar] [CrossRef]
  25. Pathak, P.D.; Mandavgane, S.A.; Kulkarni, B.D. Fruit peel waste: Characterization and its potential uses. Curr. Sci. 2017, 113, 444–454. [Google Scholar] [CrossRef]
  26. Mahato, N.; Sharma, K.; Sinha, M.; Cho, M.H. Citrus waste derived nutra-/pharmaceuticals for health benefits: Current trends and future perspectives. J. Funct. Foods 2018, 40, 307–316. [Google Scholar] [CrossRef]
  27. Singh, B.; Singh, J.P.; Kaur, A.; Singh, N. Phenolic composition, antioxidant potential and health benefits of citrus peel. Food Res. Int. 2020, 109–114. [Google Scholar] [CrossRef]
  28. Vu, H.T.; Scarlett, C.J.; Vuong, Q.V. Phenolic compounds within banana peel and their potential uses: A review. J. Funct. Foods 2018, 40, 238–248. [Google Scholar] [CrossRef]
  29. Szabo, K.; Cătoi, A.F.; Vodnar, D.C. Bioactive compounds extracted from tomato processing by-products as a source of valuable nutrients. Plant Foods Hum. Nutr. 2018, 73, 268–277. [Google Scholar] [CrossRef]
  30. Bravo, L. Polyphenols: Chemistry, Dietary Sources, Metabolism, and Nutritional Significance. Nutr. Rev. 1998, 56, 317–333. [Google Scholar] [CrossRef] [PubMed]
  31. Harborne, J.B. General procedures and measurement of total phenolics. In Methods in Plant Biochemistry; Academic Press Limited: Cambridge, MA, USA, 1989; Volume 1, pp. 1–28. [Google Scholar]
  32. Robbins, R.J. Phenolic Acids in Foods: An Overview of Analytical Methodology. J. Agric. Food Chem. 2003, 51, 2866–2887. [Google Scholar] [CrossRef] [PubMed]
  33. Tuominen, A.; Toivonen, E.; Mutikainen, P.; Salminen, J.P. Defensive strategies in Geranium sylvaticum. Part 1: Organ-specific distribution of water-soluble tannins, flavonoids and phenolic acids. Phytochemistry 2013, 95, 394–407. [Google Scholar] [CrossRef] [PubMed]
  34. He, Q.; Shen, Y.; Wang, M.; Huang, M.; Yang, R.; Zhu, S.; Wang, L.; Xu, Y.; Wu, R. Natural variation in petal color in Lycoris longituba revealed by anthocyanin components. PLoS ONE 2011, 6, e22098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Boeckler, G.A.; Gershenzon, J.; Unsicker, S.B. Phenolic glycosides of the Salicaceae and their role as anti-herbivore defenses. Phytochemistry 2011, 72, 1497–1509. [Google Scholar] [CrossRef] [PubMed]
  36. Agati, G.; Azzarello, E.; Pollastri, S.; Tattini, M. Flavonoids as antioxidants in plants: Location and functional significance. Plant Sci. 2012, 196, 67–76. [Google Scholar] [CrossRef]
  37. Landry, L.G.; Chapple, C.C.; Last, R.L. Arabidopsis mutants lacking phenolic sunscreens exhibit enhanced ultraviolet-B injury and oxidative damage. Plant Physiol. 1995, 109, 1159–1166. [Google Scholar] [CrossRef] [Green Version]
  38. Pengfei, L.; Tiansheng, D.; Xianglin, H.; Jianguo, W. Antioxidant properties of isolated isorhamnetin from the sea buckthorn marc. Plant Foods Hum. Nutr. 2009, 64, 141–145. [Google Scholar] [CrossRef]
  39. Jang, J.H.; Lee, S.H.; Jung, K.; Yoo, H.; Park, G. Inhibitory effects of myricetin on lipopolysaccharide-induced neuroinflammation. Brain Sci. 2020, 10, 32. [Google Scholar] [CrossRef] [Green Version]
  40. Tavsan, Z.; Kayali, H.A. Flavonoids showed anticancer effects on the ovarian cancer cells: Involvement of reactive oxygen species, apoptosis, cell cycle and invasion. Biomed. Pharmacother. 2019, 116, 109004. [Google Scholar] [CrossRef]
  41. Rosa, L.D.S.; Jordão, N.A.; Da Costa Pereira Soares, N.; DeMesquita, J.F.; Monteiro, M.; Teodoro, A.J. Pharmacokinetic, antiproliferative and apoptotic effects of phenolic acids in human colon adenocarcinoma cells using in vitro and in silico approaches. Molecules 2018, 23, 2569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Hou, S.Z.; Xu, S.J.; Jiang, D.X.; Chen, S.X.; Wang, L.L.; Huang, S.; Lai, X.P. Effect of the flavonoid fraction of Lithocarpus polystachyus Rehd. on spontaneously hypertensive and normotensive rats. J. Ethnopharmacol. 2012, 143, 441–447. [Google Scholar] [CrossRef] [PubMed]
  43. Mastantuono, T.; Di Maro, M.; Chiurazzi, M.; Battiloro, L.; Muscariello, E.; Nasti, G.; Starita, N.; Colantuoni, A.; Lapi, D. Rat pial microvascular changes during cerebral blood flow decrease and recovery: Effects of cyanidin administration. Front. Physiol. 2018, 9, 540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Bae, C.R.; Park, Y.K.; Cha, Y.S. Quercetin-rich onion peel extract suppresses adipogenesis by down-regulating adipogenic transcription factors and gene expression in 3T3-L1 adipocytes. J. Sci. Food Agric. 2014, 94, 2655–2660. [Google Scholar] [CrossRef] [PubMed]
  45. Shin, S.W.; Jung, E.; Kim, S.; Kim, J.H.; Kim, E.G.; Lee, J.; Park, D. Antagonizing effects and mechanisms of afzelin against UVB-induced cell damage. PLoS ONE 2013, 8, e61971. [Google Scholar] [CrossRef] [PubMed]
  46. Lianza, M.; Mandrone, M.; Chiocchio, I.; Tomasi, P.; Marincich, L.; Poli, F. Screening of ninety herbal products of commercial interest as potential ingredients for phytocosmetics. J. Enzyme Inhib. Med. Chem. 2020, 35, 1287–1291. [Google Scholar] [CrossRef]
  47. Chiocchio, I.; Mandrone, M.; Sanna, C.; Maxia, A.; Tacchini, M.; Poli, F. Screening of a hundred plant extracts as tyrosinase and elastase inhibitors, two enzymatic targets of cosmetic interest. Ind. Crops Prod. 2018, 122, 498–505. [Google Scholar] [CrossRef]
  48. Rose, P.M.; Cantrill, V.; Benohoud, M.; Tidder, A.; Rayner, C.M.; Blackburn, R.S. Application of anthocyanins from blackcurrant (Ribes nigrum L.) fruit waste as renewable hair dyes. J. Agric. Food Chem. 2018, 66, 6790–6798. [Google Scholar] [CrossRef]
  49. Barauna, S.C.; Delwing-Dal Magro, D.; Brueckheimer, M.B.; Maia, T.P.; Sala, G.A.B.N.; Döhler, A.W.; Campestrini Harger, M.; Machado de Melo, D.F.; De Gasper, A.L.; Debiasi Alberton, M.; et al. Antioxidant and antidepressant-like effects of Eugenia catharinensis D. Legrand in an animal model of depression induced by corticosterone. Metab. Brain Dis. 2018, 33, 1985–1994. [Google Scholar] [CrossRef]
  50. Agunloye, O.M.; Oboh, G.; Ademiluyi, A.O.; Ademosun, A.O.; Akindahunsi, A.A.; Oyagbemi, A.A.; Omobowale, T.O.; Ajibade, T.O.; Adedapo, A.A. Cardio-protective and antioxidant properties of caffeic acid and chlorogenic acid: Mechanistic role of angiotensin converting enzyme, cholinesterase and arginase activities in cyclosporine induced hypertensive rats. Biomed. Pharmacother. 2019, 109, 450–458. [Google Scholar] [CrossRef]
  51. Zaitone, S.A.; Ahmed, E.; Elsherbiny, N.M.; Mehanna, E.T.; El-Kherbetawy, M.K.; ElSayed, M.H.; Alshareef, D.; Moustafa, Y.M. Caffeic acid improves locomotor activity and lessens inflammatory burden in a mouse model of rotenone-induced nigral neurodegeneration: Relevance to Parkinson’s disease therapy. Pharmacol. Rep. 2019, 71, 32–41. [Google Scholar] [CrossRef] [PubMed]
  52. Sanchez, M.B.; Miranda-Perez, E.; Verjan, J.C.G.; Barrera, M.D.L.A.F.; Perez-Ramos, J.; Alarcon-Aguilar, F.J. Potential of the chlorogenic acid as multitarget agent: Insulin-secretagogue and PPAR α/γ dual agonist. Biomed. Pharmacother. 2017, 94, 169–175. [Google Scholar] [CrossRef] [PubMed]
  53. Frauches, N.S.; do Amaral, T.O.; Largueza, C.B.D.; Teodoro, A.J. Brazilian myrtaceae fruits: A review of anticancer proprieties. J. Pharm. Res. Int. 2016, 12, 1–15. [Google Scholar] [CrossRef]
  54. Del Carmen Acevedo-Ramírez, P.M.; Hallal-Calleros, C.; Flores-Pérez, I.; Alba-Hurtado, F.; Mendoza-Garfías, M.B.; del Campo, N.C.; Barajas, R. Anthelmintic effect and tissue alterations induced in vitro by hydrolysable tannins on the adult stage of the gastrointestinal nematode Haemonchus contortus. Vet. Parasitol. 2019, 266, 1–6. [Google Scholar] [CrossRef]
  55. Redondo, L.M.; Chacana, P.A.; Dominguez, J.E.; Fernandez Miyakawa, M.E.D. Perspectives in the use of tannins as alternative to antimicrobial growth promoter factors in poultry. Front. Microbiol. 2014, 5, 118. [Google Scholar] [CrossRef] [Green Version]
  56. Mutiar, S.; Asben, A. Quality of Leather Using Vegetable Tannins Extract of Acacia Mangium Bark from Waste of Industrial Plantation. In IOP Conference Series: Earth and Environmental Science; IOP Publishing Ltd.: West Sumatera, Indonesia, 2018; Volume 327, pp. 13–14. [Google Scholar]
  57. Labieniec, M.; Gabryelak, T. Effects of tannins on Chinese hamster cell line B14. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2003, 539, 127–135. [Google Scholar] [CrossRef]
  58. Labieniec, M.; Gabryelak, T.; Falcioni, G. Antioxidant and pro-oxidant effects of tannins in digestive cells of the freshwater mussel Unio tumidus. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2003, 539, 19–28. [Google Scholar] [CrossRef]
  59. Lotfi, R. A commentary on methodological aspects of hydrolysable tannins metabolism in ruminant: A perspective view. Lett. Appl. Microbiol. 2020, 71, 466–478. [Google Scholar] [CrossRef]
  60. Gülcü, M.; Uslu, N.; Özcan, M.M.; Gökmen, F.; Özcan, M.M.; Banjanin, T.; Gezgin, S.; Dursun, N.; Geçgel, Ü.; Ceylan, D.A.; et al. The investigation of bioactive compounds of wine, grape juice and boiled grape juice wastes. J. Food Process. Preserv. 2019, 43, e13850. [Google Scholar] [CrossRef] [Green Version]
  61. Moldovan, M.L.; Iurian, S.; Puscas, C.; Silaghi-Dumitrescu, R.; Hanganu, D.; Bogdan, C.; Vlase, L.; Oniga, I.; Benedec, D. A Design of Experiments strategy to enhance the recovery of polyphenolic compounds from Vitis vinifera by-products through heat reflux extraction. Biomolecules 2019, 9, 529. [Google Scholar] [CrossRef] [Green Version]
  62. Ishida, K.; Kishi, Y.; Oishi, K.; Hirooka, H.; Kumagai, H. Effects of feeding polyphenol-rich winery wastes on digestibility, nitrogen utilization, ruminal fermentation, antioxidant status and oxidative stress in wethers. Anim. Sci. J. 2015, 86, 260–269. [Google Scholar] [CrossRef] [PubMed]
  63. Ferri, M.; Bin, S.; Vallini, V.; Fava, F.; Michelini, E.; Roda, A.; Minnucci, G.; Bucchi, G.; Tassoni, A. Recovery of polyphenols from red grape pomace and assessment of their antioxidant and anti-cholesterol activities. New Biotechnol. 2016, 33, 338–344. [Google Scholar] [CrossRef] [PubMed]
  64. Ping, L.; Pizzi, A.; Guo, Z.D.; Brosse, N. Condensed tannins extraction from grape pomace: Characterization and utilization as wood adhesives for wood particleboard. Ind. Crops Prod. 2011, 34, 907–914. [Google Scholar] [CrossRef] [Green Version]
  65. Nunes, M.A.; Páscoa, R.N.; Alves, R.C.; Costa, A.S.; Bessada, S.; Oliveira, M.B.P. Fourier transform near infrared spectroscopy as a tool to discriminate olive wastes: The case of monocultivar pomaces. Waste Manag. 2020, 103, 378–387. [Google Scholar] [CrossRef]
  66. Peralbo-Molina, A.; Priego-Capote, F.; Luque de Castro, M.D. Tentative identification of phenolic compounds in olive pomace extracts using liquid chromatography–tandem mass spectrometry with a quadrupole–quadrupole-time-of-flight mass detector. J. Agric. Food Chem. 2012, 60, 11542–11550. [Google Scholar] [CrossRef]
  67. Cea Pavez, I.; Lozano-Sánchez, J.; Borrás-Linares, I.; Nuñez, H.; Robert, P.; Segura-Carretero, A. Obtaining an Extract Rich in Phenolic Compounds from Olive Pomace by Pressurized Liquid Extraction. Molecules 2019, 24, 3108. [Google Scholar] [CrossRef] [Green Version]
  68. Mannelli, F.; Cappucci, A.; Pini, F.; Pastorelli, R.; Decorosi, F.; Giovannetti, L.; Mele, M.; Minieri, S.; Conte, G.; Pauselli, M.; et al. Effect of different types of olive oil pomace dietary supplementation on the rumen microbial community profile in Comisana ewes. Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef]
  69. De Toffoli, A.; Monteleone, E.; Bucalossi, G.; Veneziani, G.; Fia, G.; Servili, M.; Zanoni, B.; Pagliarini, E.; Toschi, G.; Dinnella, C. Sensory and chemical profile of a phenolic extract from olive mill waste waters in plant-base food with varied macro-composition. Food Res. Int. 2019, 119, 236–243. [Google Scholar] [CrossRef]
  70. El-Abbassi, A.; Kiai, H.; Hafidi, A. Phenolic profile and antioxidant activities of olive mill wastewater. Food Chem. 2012, 132, 406–412. [Google Scholar] [CrossRef]
  71. Aissa, I.; Kharrat, N.; Aloui, F.; Sellami, M.; Bouaziz, M.; Gargouri, Y. Valorization of antioxidants extracted from olive mill wastewater. Biotechnol. Appl. Biochem. 2017, 64, 579–589. [Google Scholar] [CrossRef]
  72. Sudha, M.L.; Baskaran, V.; Leelavathi, K. Apple pomace as a source of dietary fiber and polyphenols and its effect on the rheological characteristics and cake making. Food Chem. 2007, 104, 686–692. [Google Scholar] [CrossRef]
  73. Krawitzky, M.; Arias, E.; Peiro, J.M.; Negueruela, A.I.; Val, J.; Oria, R. Determination of color, antioxidant activity, and phenolic profile of different fruit tissue of Spanish ‘Verde Doncella’apple cultivar. Int. J. Food Prop. 2014, 17, 2298–2311. [Google Scholar] [CrossRef] [Green Version]
  74. Pieszka, M.; Gogol, P.; Pietras, M.; Pieszka, M. Valuable components of dried pomaces of chokeberry, black currant, strawberry, apple and carrot as a source of natural antioxidants and nutraceuticals in the animal diet. Ann. Anim. Sci. 2015, 15, 475–491. [Google Scholar] [CrossRef] [Green Version]
  75. Peschel, W.; Sánchez-Rabaneda, F.; Diekmann, W.; Plescher, A.; Gartzía, I.; Jiménez, D.; Lamuela-Raventós, R.; Buxaderas, S.; Codina, C. An industrial approach in the search of natural antioxidants from vegetable and fruit wastes. Food Chem. 2006, 97, 137–150. [Google Scholar] [CrossRef]
  76. Phan, K.; Van Den Broeck, E.; Van Speybroeck, V.; De Clerck, K.; Raes, K.; De Meester, S. The potential of anthocyanins from blueberries as a natural dye for cotton: A combined experimental and theoretical study. Dyes Pigment 2020, 176, 108180. [Google Scholar] [CrossRef]
  77. Maurício, E.M.; Rosado, C.; Duarte, M.P.; Fernando, A.L.; Díaz-Lanza, A.M. Evaluation of industrial sour cherry liquor wastes as an ecofriendly source of added value chemical compounds and energy. Waste Biomass Valorization 2020, 11, 201–210. [Google Scholar] [CrossRef]
  78. Celano, R.; Piccinelli, A.L.; Pagano, I.; Roscigno, G.; Campone, L.; De Falco, E.; Russo, M.; Rastrelli, L. Oil distillation wastewaters from aromatic herbs as new natural source of antioxidant compounds. Food Res. Int. 2017, 99, 298–307. [Google Scholar] [CrossRef]
  79. Kapusta, I.; Janda, B.; Szajwaj, B.; Stochmal, A.; Piacente, S.; Pizza, C.; Franceschi, F.; Franz, C.; Oleszek, W. Flavonoids in horse chestnut (Aesculus hippocastanum) seeds and powdered waste water byproducts. J. Agric. Food Chem. 2007, 55, 8485–8490. [Google Scholar] [CrossRef]
  80. Mussatto, S.I.; Ballesteros, L.F.; Martins, S.; Teixeira, J.A. Extraction of antioxidant phenolic compounds from spent coffee grounds. Sep. Purif. Technol. 2011, 83, 173–179. [Google Scholar] [CrossRef] [Green Version]
  81. Ramón-Gonçalves, M.; Gómez-Mejía, E.; Rosales-Conrado, N.; León-González, M.E.; Madrid, Y. Extraction, identification and quantification of polyphenols from spent coffee grounds by chromatographic methods and chemometric analyses. Waste Manag. 2019, 96, 15–24. [Google Scholar] [CrossRef]
  82. Sant’Anna, V.; Biondo, E.; Kolchinski, E.M.; da Silva, L.F.S.; Corrêa, A.P.F.; Bach, E.; Brandelli, A. Total polyphenols, antioxidant, antimicrobial and allelopathic activities of spend coffee ground aqueous extract. Waste Biomass Valorization 2017, 8, 439–442. [Google Scholar] [CrossRef]
  83. Bessada, S.M.; Alves, R.C.; Costa, A.S.; Nunes, M.A.; Oliveira, M.B.P. Coffea canephora silverskin from different geographical origins: A comparative study. Sci. Total Environ. 2018, 645, 1021–1028. [Google Scholar] [CrossRef] [PubMed]
  84. Regazzoni, L.; Saligari, F.; Marinello, C.; Rossoni, G.; Aldini, G.; Carini, M.; Orioli, M. Coffee silver skin as a source of polyphenols: High resolution mass spectrometric profiling of components and antioxidant activity. J. Funct. Foods 2016, 20, 472–485. [Google Scholar] [CrossRef]
  85. Bresciani, L.; Calani, L.; Bruni, R.; Brighenti, F.; Del Rio, D. Phenolic composition, caffeine content and antioxidant capacity of coffee silverskin. Food Res. Int. 2014, 61, 196–201. [Google Scholar] [CrossRef]
  86. Zamuz, S.; López-Pedrouso, M.; Barba, F.J.; Lorenzo, J.M.; Domínguez, H.; Franco, D. Application of hull, bur and leaf chestnut extracts on the shelf-life of beef patties stored under MAP: Evaluation of their impact on physicochemical properties, lipid oxidation, antioxidant, and antimicrobial potential. Food Res. Int. 2018, 112, 263–273. [Google Scholar] [CrossRef] [PubMed]
  87. Vella, F.M.; De Masi, L.; Calandrelli, R.; Morana, A.; Laratta, B. Valorization of the agro-forestry wastes from Italian chestnut cultivars for the recovery of bioactive compounds. Eur. Food Res. Technol. 2019, 245, 2679–2686. [Google Scholar] [CrossRef]
  88. Flórez-Fernández, N.; Torres, M.D.; Gómez, S.; Couso, S.; Domínguez, H. Potential of Chestnut Wastes for Cosmetics and Pharmaceutical Applications. Waste Biomass Valorization 2019, 11, 1–10. [Google Scholar] [CrossRef]
  89. Pinto, D.; Braga, N.; Rodrigues, F.; Oliveira, M.B.P.P. Castanea sativa bur: An undervalued by-product but a promising cosmetic ingredient. Cosmetics 2017, 4, 50. [Google Scholar] [CrossRef] [Green Version]
  90. Esposito, T.; Celano, R.; Pane, C.; Piccinelli, A.L.; Sansone, F.; Picerno, P.; Zaccardelli, M.; Aquino, R.P.; Mencherini, T. Chestnut (Castanea sativa Miller.) burs extracts and functional compounds: Uhplc-uv-hrms profiling, antioxidant activity, and inhibitory effects on phytopathogenic fungi. Molecules 2019, 24, 302. [Google Scholar] [CrossRef] [Green Version]
  91. Chiocchio, I.; Prata, C.; Mandrone, M.; Ricciardiello, F.; Marrazzo, P.; Tomasi, P.; Angeloni, C.; Fiorentini, D.; Malaguti, M.; Poli, F.; et al. Leaves and Spiny Burs of Castanea Sativa from an Experimental Chestnut Grove: Metabolomic Analysis and Anti-Neuroinflammatory Activity. Metabolites 2020, 10, 408. [Google Scholar] [CrossRef]
  92. Aires, A.; Carvalho, R.; Saavedra, M.J. Valorization of solid wastes from chestnut industry processing: Extraction and optimization of polyphenols, tannins and ellagitannins and its potential for adhesives, cosmetic and pharmaceutical industry. Waste Manag. 2016, 48, 457–464. [Google Scholar] [CrossRef] [PubMed]
  93. Turrini, F.; Donno, D.; Boggia, R.; Beccaro, G.L.; Zunin, P.; Leardi, R.; Pittaluga, A.M. An innovative green extraction and re-use strategy to valorize food supplement by-products: Castanea sativa bud preparations as case study. Food Res. Int. 2019, 115, 276–282. [Google Scholar] [CrossRef] [PubMed]
  94. Comandini, P.; Lerma-García, M.J.; Simó-Alfonso, E.F.; Toschi, T.G. Tannin analysis of chestnut bark samples (Castanea sativa Mill.) by HPLC-DAD–MS. Food Chem. 2014, 157, 290–295. [Google Scholar] [CrossRef] [PubMed]
  95. Seabra, I.J.; Chim, R.B.; Salgueiro, P.; Braga, M.E.; de Sousa, H.C. Influence of solvent additives on the aqueous extraction of tannins from pine bark: Potential extracts for leather tanning. J. Chem. Technol. Biotechnol. 2018, 93, 1169–1182. [Google Scholar] [CrossRef]
  96. Grillo, G.; Boffa, L.; Binello, A.; Mantegna, S.; Cravotto, G.; Chemat, F.; Dizhbite, T.; Lauberte, L.; Telysheva, G. Cocoa bean shell waste valorisation; extraction from lab to pilot-scale cavitational reactors. Food Res. Int. 2019, 115, 200–208. [Google Scholar] [CrossRef] [PubMed]
  97. Üstündağ, Ö.G.; Erşan, S.; Özcan, E.; Özan, G.; Kayra, N.; Ekinci, F.Y. Black tea processing waste as a source of antioxidant and antimicrobial phenolic compounds. Eur. Food Res. Technol. 2016, 242, 1523–1532. [Google Scholar] [CrossRef]
  98. Vella, F.M.; Cautela, D.; Laratta, B. Characterization of polyphenolic compounds in cantaloupe melon by-products. Foods 2019, 8, 196. [Google Scholar] [CrossRef] [Green Version]
  99. Kakumu, Y.; Yamauchi, K.; Mitsunaga, T. Identification of chemical constituents from the bark of Larix kaempferi and their tyrosinase inhibitory effect. Holzforschung 2019, 73, 637–643. [Google Scholar] [CrossRef]
  100. Tilay, A.; Bule, M.; Kishenkumar, J.; Annapure, U. Preparation of ferulic acid from agricultural wastes: Its improved extraction and purification. J. Agric. Food Chem. 2008, 56, 7644–7648. [Google Scholar] [CrossRef]
  101. Guimarães, A.G.; Serafini, M.R.; Quintans-Júnior, L.J. Terpenes and derivatives as a new perspective for pain treatment: A patent review. Expert Opin. Ther. Pat. 2014, 24, 243–265. [Google Scholar] [CrossRef]
  102. Schwab, W.; Fuchs, C.; Huang, F.C. Transformation of terpenes into fine chemicals. Eur. J. Lipid Sci. Technol. 2013, 115, 3–8. [Google Scholar] [CrossRef]
  103. Salomé-Abarca, L.F.; Mandrone, M.; Sanna, C.; Poli, F.; van der Hondel, C.A.; Klinkhamer, P.G.; Choi, Y.H. Metabolic variation in Cistus monspeliensis L. ecotypes correlated to their plant-fungal interactions. Phytochemistry 2020, 176, 112402. [Google Scholar] [CrossRef] [PubMed]
  104. Fleet, C.M.; Sun, T.P. A DELLAcate balance: The role of gibberellin in plant morphogenesis. Curr. Plant Biol. 2005, 8, 77–85. [Google Scholar] [CrossRef] [PubMed]
  105. Hajimohammadi, R.; Hosseini, M.; Amani, H.; Najafpour, G.D. Production of saponin biosurfactant from Glycyrrhiza glabra as an agent for upgrading heavy crude oil. J. Surfactants Deterg. 2016, 19, 1251–1261. [Google Scholar] [CrossRef]
  106. Gonzalez, P.J.; Sörensen, P.M. Characterization of saponin foam from Saponaria officinalis for food applications. Food Hydrocoll. 2020, 101, 105541. [Google Scholar] [CrossRef]
  107. Tippel, J.; Gies, K.; Harbaum-Piayda, B.; Steffen-Heins, A.; Drusch, S. Composition of Quillaja saponin extract affects lipid oxidation in oil-in-water emulsions. Food Chem. 2017, 221, 386–394. [Google Scholar] [CrossRef]
  108. Dewick, P.M. Medicinal Natural Products: A Biosynthetic Approach; John Wiley & Sons: New York, NY, USA, 2002. [Google Scholar]
  109. Lee, S.; Peterson, C.J.; Coats, J.R. Fumigation toxicity of monoterpenoids to several stored product insects. J. Stored Prod. Res. 2003, 39, 77–85. [Google Scholar] [CrossRef]
  110. Hammer, K.A.; Carson, C.F.; Riley, T.V. Antifungal activity of the components of Melaleuca alternifolia (tea tree) oil. J. Appl. Microbiol. 2003, 95, 853–860. [Google Scholar] [CrossRef] [Green Version]
  111. Friedman, M.; Henika, P.R.; Mandrell, R.E. Bactericidal activities of plant essential oils and some of their isolated constituents against Campylobacter jejuni, Escherichia coli, Listeria monocytogenes, and Salmonella enterica. J. Food Prot. 2002, 65, 1545–1560. [Google Scholar] [CrossRef]
  112. Salehi, B.; Mishra, A.P.; Shukla, I.; Sharifi-Rad, M.; Contreras, M.D.M.; Segura-Carretero, A.; Sharifi-Rad, J. Thymol, thyme, and other plant sources: Health and potential uses. Phytother. Res. 2018, 32, 1688–1706. [Google Scholar] [CrossRef]
  113. Raimundo, J.R.; Frazão, D.F.; Domingues, J.L.; Quintela-Sabarís, C.; Dentinho, T.P.; Anjos, O.; Delgado, F. Neglected Mediterranean plant species are valuable resources: The example of Cistus ladanifer. Planta 2018, 248, 1351–1364. [Google Scholar] [CrossRef] [PubMed]
  114. McClements, D.J.; Gumus, C.E. Natural emulsifiers—Biosurfactants, phospholipids, biopolymers, and colloidal particles: Molecular and physicochemical basis of functional performance. Adv. Colloid Interface Sci. 2016, 234, 3–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. del Hierro, J.N.; Herrera, T.; Fornari, T.; Reglero, G.; Martin, D. The gastrointestinal behavior of saponins and its significance for their bioavailability and bioactivities. J. Funct. Food 2018, 40, 484–497. [Google Scholar] [CrossRef]
  116. Rehan, M.; Mir, S.A. Structural diversity, natural sources, and pharmacological potential of plant-based saponins with special focus on anticancer activity: A review. Med. Chem. Res. 2020, 29, 1707–1722. [Google Scholar] [CrossRef]
  117. Singh, B.; Singh, J.P.; Singh, N.; Kaur, A. Saponins in pulses and their health promoting activities: A review. Food Chem. 2017, 233, 540–549. [Google Scholar] [CrossRef]
  118. Jesch, E.D.; Carr, T.P. Food ingredients that inhibit cholesterol absorption. Prev. Nutr. Food Sci. 2017, 22, 67–80. [Google Scholar]
  119. Marrelli, M.; Conforti, F.; Araniti, F.; Statti, G.A. Effects of saponins on lipid metabolism: A review of potential health benefits in the treatment of obesity. Molecules 2016, 21, 1404. [Google Scholar] [CrossRef] [Green Version]
  120. Vinarova, L.; Vinarov, Z.; Atanasov, V.; Pantcheva, I.; Tcholakova, S.; Denkov, N.; Stoyanov, S. Lowering of cholesterol bioaccessibility and serum concentrations by saponins: In vitro and in vivo studies. Food Funct. 2015, 6, 501–512. [Google Scholar] [CrossRef]
  121. Zhao, Y.L.; Cai, G.M.; Hong, X.; Shan, L.M.; Xiao, X.H. Anti-hepatitis B virus activities of triterpenoid saponin compound from potentilla anserine L. Phytomedicine 2008, 15, 253–258. [Google Scholar] [CrossRef]
  122. Anthoni, C.; Laukoetter, M.G.; Rizken, E. Mechanism underlying the anti-inflammation actions of boswellic acid derivatives in experimental colitis. Am. J Physiol. Gastrointest. Liver Physiol. 2006, 290, G1131–G1137. [Google Scholar] [CrossRef] [Green Version]
  123. Poeckel, D.; Tausch, L.; Altmann, A.; Feißt, C.; Klinkhardt, U.; Graff, J.; Werz, O. Induction of central signalling pathways and select functional effects in human platelets by β-boswellic acid. Br. J. Pharmacol. 2005, 146, 514–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Poeckel, D.; Tausch, L.; George, S.; Jauch, J.; Werz, O. 3-O-acetyl -11-keto-boswellic acid decreases basal intracellular Ca2+ levels and inhibits agonist-induced Ca2+ mobilization and mitogen-activated protein kinase activation in human monocyte cells. J. Pharmacol. Exp. Ther. 2006, 316, 224–232. [Google Scholar] [CrossRef] [PubMed]
  125. Singh, B.; Sharma, R.A. Plant terpenes: Defense responses, phylogenetic analysis, regulation and clinical applications. 3 Biotech 2015, 5, 129–151. [Google Scholar] [CrossRef] [Green Version]
  126. Alakurtti, S.; Mäkelä, T.; Koskimies, S.; Yli-Kauhaluoma, J. Pharmacological properties of the ubiquitous natural product betulinic acid. Eur. J. Pharm. Sci. 2006, 29, 1–13. [Google Scholar] [CrossRef] [PubMed]
  127. Hoagland, R.E.; Zablotowicz, R.M.; Reddy, K.N. Studies of the phytotoxicity of saponins on weed and crop plants. In Saponins Used in Food and Agriculture; Waller, G.R., Yamasaki, K., Eds.; Springer: Boston, MA, USA, 1996; Volume 405, pp. 57–73. [Google Scholar]
  128. Moses, T.; Papadopoulou, K.K.; Osbourn, A. Metabolic and functional diversity of saponins, biosynthetic intermediates and semi-synthetic derivatives. Crit. Rev. Biochem. Mol. 2014, 49, 439–462. [Google Scholar] [CrossRef] [Green Version]
  129. Augustin, J.M.; Kuzina, V.; Andersen, S.B.; Bak, S. Molecular activities, biosynthesis and evolution of triterpenoid saponins. Phytochemistry 2011, 72, 435–457. [Google Scholar] [CrossRef]
  130. Sreij, R.; Dargel, C.; Schweins, R.; Prévost, S.; Dattani, R.; Hellweg, T. Aescin-cholesterol complexes in DMPC model membranes: A DSC and temperature-dependent scattering study. Sci. Rep. 2019, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  131. Nielsen, J.K.; Nagao, T.; Okabe, H.; Shinoda, T. Resistance in the plant, Barbarea vulgaris, and counter-adaptations in flea beetles mediated by saponins. J. Chem. Ecol. 2010, 36, 277–285. [Google Scholar] [CrossRef]
  132. Huang, H.C.; Liao, S.C.; Chang, F.R.; Kuo, Y.H.; Wu, Y.C. Molluscicidal saponins from Sapindus mukorossi, inhibitory agents of golden apple snails, Pomacea canaliculata. J. Agric. Food Chem. 2003, 51, 4916–4919. [Google Scholar] [CrossRef]
  133. Waller, G.R.; Jurzysta, M.; Thorne, R.L.Z. Allelopathic activity of root saponins from alfalfa (Medicago sativa L.) on weeds and wheat. Bot. Bull. Acad. Sinica 1993, 34, 1–11. [Google Scholar]
  134. Trdá, L.; Janda, M.; Macková, D.; Pospíchalová, R.; Dobrev, P.; Burketová, L.; Matušinsky, P. Dual Mode of the Saponin Aescin in Plant Protection: Antifungal Agent and Plant Defense Elicitor. Front. Plant Sci. 2019, 10, 1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Fox, L.T.; Gerber, M.; Plessis, J.D.; Hamman, J.H. Transdermal drug delivery enhancement by compounds of natural origin. Molecules 2011, 16, 10507–10540. [Google Scholar] [CrossRef]
  136. Cid-Pérez, T.S.; Ávila-Sosa, R.; Ochoa-Velasco, C.E.; Rivera-Chavira, B.E.; Nevárez-Moorillón, G.V. Antioxidant and antimicrobial activity of Mexican Oregano (Poliomintha longiflora) essential oil, hydrosol and extracts from waste solid residues. Plants 2019, 8, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Cautela, D.; Vella, F.M.; Castaldo, D.; Laratta, B. Characterization of essential oil recovered from fennel horticultural wastes. Nat. Prod. Res. 2019, 33, 1964–1968. [Google Scholar] [CrossRef] [PubMed]
  138. Negro, V.; Mancini, G.; Ruggeri, B.; Fino, D. Citrus waste as feedstock for bio-based products recovery: Review on limonene case study and energy valorization. Bioresour. Technol. 2016, 214, 806–815. [Google Scholar] [CrossRef]
  139. Pourbafrani, M.; Forgács, G.; Horváth, I.S.; Niklasson, C.; Taherzadeh, M.J. Production of biofuels, limonene and pectin from citrus wastes. Bioresour. Technol. 2010, 101, 4246–4250. [Google Scholar] [CrossRef]
  140. Ciriminna, R.; Lomeli-Rodriguez, M.; Cara, P.D.; Lopez-Sanchez, J.A.; Pagliaro, M. Limonene: A versatile chemical of the bioeconomy. Chem. Commun. 2014, 50, 15288–15296. [Google Scholar] [CrossRef]
  141. Fitzgerald, C.; Hossain, M.; Rai, D.K. Waste/By-Product Utilisations. In Innovative Technologies in Beverage Processing; John Wiley & Sons; Wiley-Blackwell: Hoboken, NJ, USA, 2017; Volume 7, pp. 297–309. [Google Scholar]
  142. Edris, A.E. Pharmaceutical and therapeutic potentials of essential oils and their individual volatile constituents: A review. Phytother. Res. 2007, 21, 308–323. [Google Scholar] [CrossRef]
  143. González-Molina, E.; Domínguez-Perles, R.; Moreno, D.A.; García-Viguera, C. Natural bioactive compounds of Citrus limon for food and health. J. Pharm. Biomed. Anal. 2010, 51, 327–345. [Google Scholar] [CrossRef]
  144. Hollingsworth, R.G. Limonene, a citrus extract, for control of mealybugs and scale insects. J. Econ. Entomol. 2005, 98, 772–779. [Google Scholar] [CrossRef]
  145. Meneguzzo, F.; Brunetti, C.; Fidalgo, A.; Ciriminna, R.; Delisi, R.; Albanese, L.; Ferrini, F. Real-Scale Integral Valorization of Waste Orange Peel via Hydrodynamic Cavitation. Processes 2019, 7, 581. [Google Scholar] [CrossRef] [Green Version]
  146. Lota, M.L.; de Rocca Serra, D.; Tomi, F.; Jacquemond, C.; Casanova, J. Volatile components of peel and leaf oils of lemon and lime species. J. Agric. Food Chem. 2002, 50, 796–805. [Google Scholar] [CrossRef] [PubMed]
  147. Celli, A.; Colonna, M.; Gandini, A.; Gioia, C.; Lacerda, T.M.; Vannini, M. Polymers from monomers derived from biomass. In Chemicals and Fuels from Bio-Based Building Blocks; Cavani, F., Albonetti, S., Basile, F., Gandini, A., Eds.; Wiley-VCH: Weinheim, Germany, 2016; pp. 297–308. [Google Scholar]
  148. Llevot, A.; Grau, E.; Carlotti, S.; Grelier, S.; Cramail, H. Dimerization of abietic acid for the design of renewable polymers by ADMET. Eur. Polym. J. 2015, 67, 409–417. [Google Scholar] [CrossRef] [Green Version]
  149. Aldas, M.; Pavon, C.; López-Martínez, J.; Arrieta, M.P. Pine Resin Derivatives as Sustainable Additives to Improve the Mechanical and Thermal Properties of Injected Moulded Thermoplastic Starch. Appl. Sci. 2020, 10, 2561. [Google Scholar] [CrossRef] [Green Version]
  150. Meninno, S. Valorization of Waste: Sustainable Organocatalysts from Renewable Resources. ChemSusChem 2020, 13, 439–468. [Google Scholar] [CrossRef]
  151. Ribeiro, B.D.; Barreto, D.W.; Coelho, M.A.Z. Use of micellar extraction and cloud point preconcentration for valorization of saponins from sisal (Agave sisalana) waste. Food Bioprod. Process. 2015, 94, 601–609. [Google Scholar] [CrossRef]
  152. Dahlawi, S.M.; Nazir, W.; Iqbal, R.; Asghar, W.; Khalid, N. Formulation and characterization of oil-in-water nanoemulsions stabilized by crude saponins isolated from onion skin waste. RSC Adv. 2020, 10, 39700–39707. [Google Scholar] [CrossRef]
  153. Cláudio, A.F.M.; Cognigni, A.; de Faria, E.L.; Silvestre, A.J.; Zirbs, R.; Freire, M.G.; Bica, K. Valorization of olive tree leaves: Extraction of oleanolic acid using aqueous solutions of surface-active ionic liquids. Sep. Purif. Technol. 2018, 204, 30–37. [Google Scholar] [CrossRef]
  154. Kishikawa, A.; Amen, Y.; Shimizu, K. Anti-allergic triterpenes isolated from olive milled waste. Cytotechnology 2017, 69, 307–315. [Google Scholar] [CrossRef] [Green Version]
  155. Campalani, C.; Chioggia, F.; Amadio, E.; Gallo, M.; Rizzolio, F.; Selva, M.; Perosa, A. Supercritical CO2 extraction of natural antibacterials from low value weeds and agro-waste. J. CO2 Util. 2020, 40, 101–198. [Google Scholar] [CrossRef]
  156. Gómez-García, R.; Campos, D.A.; Aguilar, C.N.; Madureira, A.R.; Pintado, M. Valorization of melon fruit (Cucumis melo L.) by-products: Phytochemical and Biofunctional properties with Emphasis on Recent Trends and Advances. Trends Food Sci. Technol. 2020, 25, 1–22. [Google Scholar] [CrossRef]
  157. Arscott, S.A. Food sources of carotenoids. In Carotenoids and Human Health; Humana Press: Totowa, NJ, USA, 2013; pp. 3–19. [Google Scholar]
  158. Aniszewski, T. Alkaloids: Chemistry, Biology, Ecology, and Applications, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2015. [Google Scholar]
  159. Debnath, B.; Singh, W.S.; Das, M.; Goswami, S.; Singh, M.K.; Maiti, D.; Manna, K. Role of plant alkaloids on human health: A review of biological activities. Mater. Today Chem. 2018, 9, 56–72. [Google Scholar] [CrossRef]
  160. Roddick, J.G. Steroidal glycoalkaloids: Nature and consequences of bioactivity. Adv. Exp. Med. Biol. 1996, 404, 277–295. [Google Scholar] [PubMed]
  161. Guggisberg, A.; Hesse, M. Putrescine, spermidine, spermine, and related polyamine alkaloids. In The Alkaloids: Chemistry and Pharmacology; Academic Press: New York, NY, USA, 1984; Volume 22, pp. 85–188. [Google Scholar]
  162. Tschesche, R.; Kaussmann, E.U. Alkaloids: Chemistry and Physiology, 1st ed.; Academic Press: Cambridge, MA, USA, 1975; pp. 165–205. [Google Scholar]
  163. Bednarz, H.; Roloff, N.; Niehaus, K. Mass spectrometry imaging of the spatial and temporal localization of alkaloids in Nightshades. J. Agric. Food Chem. 2019, 67, 13470–13477. [Google Scholar] [CrossRef] [PubMed]
  164. van Dam, N.M.; Vuister, L.W.; Bergshoeff, C.; de Vos, H.; Van der Meijden, E.D. The “Raison D’être” of pyrrolizidine alkaloids in Cynoglossum officinale: Deterrent effects against generalist herbivores. J. Chem. Ecol. 1995, 21, 507–523. [Google Scholar] [CrossRef]
  165. Mattocks, A.R. Toxicity of pyrrolizidine alkaloids. Nature. 1968, 217, 723–728. [Google Scholar] [CrossRef]
  166. Steppuhn, A.; Gase, K.; Krock, B.; Halitschke, R.; Baldwin, I.T. Nicotine’s defensive function in nature. PLoS Biol. 2004, 2, e217. [Google Scholar] [CrossRef]
  167. Trigo, J.R. Effects of pyrrolizidine alkaloids through different trophic levels. Phytochem. Rev. 2011, 10, 83–98. [Google Scholar] [CrossRef]
  168. Liu, D.L.; Lovett, J.V. Biologically active secondary metabolites of barley. II. Phytotoxicity of barley allelochemicals. J. Chem. Ecol. 1993, 19, 2231–2244. [Google Scholar] [CrossRef]
  169. Kessler, D.; Bhattacharya, S.; Diezel, C.; Rothe, E.; Gase, K.; Schöttner, M.; Baldwin, I.T. Unpredictability of nectar nicotine promotes outcrossing by hummingbirds in Nicotiana attenuata. Plant J. 2012, 71, 529–538. [Google Scholar] [CrossRef]
  170. Mustard, J.A. Neuroactive nectar: Compounds in nectar that interact with neurons. Arthropod Plant Interact. 2020, 14, 151–159. [Google Scholar] [CrossRef]
  171. Prada, F.; Stashenko, E.E.; Martínez, J.R. LC/MS study of the diversity and distribution of pyrrolizidine alkaloids in Crotalaria species growing in Colombia. J. Sep. Sci. 2020, 43, 4322–4337. [Google Scholar] [CrossRef] [PubMed]
  172. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef] [PubMed]
  173. D’Alessandro, S.; Scaccabarozzi, D.; Signorini, L.; Perego, F.; Ilboudo, D.P.; Ferrante, P.; Delbue, S. The use of antimalarial drugs against viral infection. Microorganisms 2020, 8, 85. [Google Scholar] [CrossRef] [Green Version]
  174. Ma, G.; Bavadekar, S.A.; Davis, Y.M.; Lalchandani, S.G.; Nagmani, R.; Schaneberg, B.T.; Feller, D.R. Pharmacological effects of ephedrine alkaloids on human α1-and α2-adrenergic receptor subtypes. J. Pharmacol. Exp. Ther. 2007, 322, 214–221. [Google Scholar] [CrossRef]
  175. Ronghe, M.; Burke, G.A.A.; Lowis, S.P.; Estlin, E.J. Remission induction therapy for childhood acute lymphoblastic leukaemia: Clinical and cellular pharmacology of vincristine, corticosteroids, L-asparaginase and anthracyclines. Cancer Treat. Rev. 2001, 27, 327–337. [Google Scholar] [CrossRef]
  176. Campos, M.M.; Fernandes, E.S.; Ferreira, J.; Santos, A.R.; Calixto, J.B. Antidepressant-like effects of Trichilia catigua (Catuaba) extract: Evidence for dopaminergic-mediated mechanisms. Psychopharmacology 2005, 182, 45–53. [Google Scholar] [CrossRef]
  177. Turner, N.; Li, J.Y.; Gosby, A.; To, S.W.C.; Cheng, Z.; Miyoshi, H.; Taketo, M.M.; Cooneyl, G.J.; Kraegen, E.W.; James, D.E.; et al. Berberine and its more biologically available derivative, dihydroberberine, inhibit mitochondrial respiratory complex I: A mechanism for the action of berberine to activate AMP-activated protein kinase and improve insulin action. Diabetes 2008, 57, 1414–1418. [Google Scholar] [CrossRef] [Green Version]
  178. Amritpal, S.; Sanjiv, D.; Navpreet, K.; Jaswinder, S. Berberine: Alkaloid with wide spectrum of pharmacological activities. J. Nat. Prod. 2010, 3, 64–75. [Google Scholar]
  179. Palmieri, A.; Scapoli, L.; Iapichino, A.; Mercolini, L.; Mandrone, M.; Poli, F.; Giannì, A.B.; Baserga, C.; Martinelli, M. Berberine and Tinospora cordifolia exert a potential anticancer effect on colon cancer cells by acting on specific pathways. Int. J. Immunopathol. Pharmacol. 2019, 33, 2058738419855567. [Google Scholar] [CrossRef] [Green Version]
  180. Crozier, A.; Clifford, M.; Ashihara, H. Plant Secondary Metabolites: Occurrence, Structure and Role in the Human Diet; Blackwell: Oxford, UK, 2006. [Google Scholar]
  181. Liu, Z.L.; Liu, Q.Z.; Du, S.S.; Deng, Z.W. Mosquito larvicidal activity of alkaloids and limonoids derived from Evodia rutaecarpa unripe fruits against Aedes albopictus (Diptera: Culicidae). Parasitol. Res. 2012, 111, 991–996. [Google Scholar] [CrossRef] [PubMed]
  182. Yang, Y.C.; Lee, S.G.; Lee, H.K.; Kim, M.K.; Lee, S.H.; Lee, H.S. A piperidine amide extracted from Piper longum L. fruit shows activity against Aedes aegypti mosquito larvae. J. Agric. Food Chem. 2002, 50, 3765–3767. [Google Scholar] [PubMed]
  183. Yogendra, K.N.; Dhokane, D.; Kushalappa, A.C.; Sarmiento, F.; Rodriguez, E.; Mosquera, T. StWRKY8 transcription factor regulates benzylisoquinoline alkaloid pathway in potato conferring resistance to late blight. Plant Sci. 2017, 256, 208–216. [Google Scholar] [CrossRef] [PubMed]
  184. Lee, C.H.; Lee, H.J.; Jeon, J.H.; Lee, H.S. In vivo antifungal effects of Coptis japonica root-derived isoquinoline alkaloids against phytopathogenic fungi. J. Microbiol. Biotechnol. 2005, 15, 1402–1407. [Google Scholar]
  185. Chowański, S.; Adamski, Z.; Marciniak, P.; Rosiński, G.; Büyükgüzel, E.; Büyükgüzel, K.; Falabella, P.; Scrano, L.; Ventrella, E.; Lelario, F.; et al. A review of bioinsecticidal activity of Solanaceae alkaloids. Toxins 2016, 8, 60. [Google Scholar]
  186. Bertin, C.; Zunino, H.; Pittet, J.C.; Beau, P.; Pineau, P.; Massonneau, M.; Robert, C.; Hopkins, J. A double-blind evaluation of the activity of an anti-cellulite product containing retinol, caffeine, and ruscogenine by a combination of several non-invasive methods. J. Cosmet. Sci. 2001, 52, 199–210. [Google Scholar]
  187. Velasco, M.V.R.; Tano, C.T.N.; Machado-Santelli, G.M.; Consiglieri, V.O.; Kaneko, T.M.; Baby, A.R. Effects of caffeine and siloxanetriol alginate caffeine, as anticellulite agents, on fatty tissue: Histological evaluation. J. Cosmet. Dermatol. 2008, 7, 23–29. [Google Scholar] [CrossRef]
  188. Eun Lee, K.; Bharadwaj, S.; Yadava, U.; Gu Kang, S. Evaluation of caffeine as inhibitor against collagenase, elastase and tyrosinase using in silico and in vitro approach. J. Enzyme Inhib. Med. Chem. 2019, 34, 927–936. [Google Scholar] [CrossRef] [Green Version]
  189. Panusa, A.; Zuorro, A.; Lavecchia, R.; Marrosu, G.; Petrucci, R. Recovery of natural antioxidants from spent coffee grounds. J. Agric. Food Chem. 2013, 61, 4162–4168. [Google Scholar] [CrossRef]
  190. Okiyama, D.C.; Soares, I.D.; Cuevas, M.S.; Crevelin, E.J.; Moraes, L.A.; Melo, M.P.; Oliveira, A.L.; Rodrigues, C.E. Pressurized liquid extraction of flavanols and alkaloids from cocoa bean shell using ethanol as solvent. Food Res. Int. 2018, 114, 20–29. [Google Scholar] [CrossRef]
  191. Arlorio, M.; Coïsson, J.D.; Travaglia, F.; Varsaldi, F.; Miglio, G.; Lombardi, G.; Martelli, A. Antioxidant and biological activity of phenolic pigments from Theobroma cacao hulls extracted with supercritical CO2. Food Res. Int. 2005, 38, 1009–1014. [Google Scholar] [CrossRef]
  192. Friedman, M.; Huang, V.; Quiambao, Q.; Noritake, S.; Liu, J.; Kwon, O.; Chintalapati, S.; Young, J.; Levin, C.E.; Tam, C.; et al. Potato peels and their bioactive glycoalkaloids and phenolic compounds inhibit the growth of pathogenic trichomonads. J. Agric. Food Chem. 2018, 66, 7942–7947. [Google Scholar] [CrossRef] [PubMed]
  193. Wu, Z.G.; Xu, H.Y.; Ma, Q.; Cao, Y.; Ma, J.N.; Ma, C.M. Isolation, identification and quantification of unsaturated fatty acids, amides, phenolic compounds and glycoalkaloids from potato peel. Food Chem. 2012, 135, 2425–2429. [Google Scholar] [CrossRef] [PubMed]
  194. Figueiredo-Gonzalez, M.; Valentao, P.; Andrade, P.B. Tomato plant leaves: From by-products to the management of enzymes in chronic diseases. Ind. Crops Prod. 2016, 94, 621–629. [Google Scholar] [CrossRef]
  195. Mellinas, C.; Ramos, M.; Jiménez, A.; Garrigós, M.C. Recent Trends in the Use of Pectin from Agro-Waste Residues as a Natural-Based Biopolymer for Food Packaging Applications. Materials 2020, 13, 673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Xu, M.; Ran, L.; Chen, N.; Fan, X.; Ren, D.; Yi, L. Polarity-dependent extraction of flavonoids from citrus peel waste using a tailor-made deep eutectic solvent. Food Chem. 2019, 297, 124970. [Google Scholar] [CrossRef]
  197. Aizpurua-Olaizola, O.; Ormazabal, M.; Vallejo, A.; Olivares, M.; Navarro, P.; Etxebarria, N.; Usobiaga, A. Optimization of supercritical fluid consecutive extractions of fatty acids and polyphenols from Vitis vinifera grape wastes. J. Food Sci. 2015, 80, E101–E107. [Google Scholar] [CrossRef]
  198. Manna, L.; Bugnone, C.A.; Banchero, M. Valorization of hazelnut, coffee and grape wastes through supercritical fluid extraction of triglycerides and polyphenols. J. Supercrit. Fluids 2015, 104, 204–211. [Google Scholar] [CrossRef]
  199. Takla, S.S.; Shawky, E.; Hammoda, H.M.; Darwish, F.A. Green techniques in comparison to conventional ones in the extraction of Amaryllidaceae alkaloids: Best solvents selection and parameters optimization. J. Chromatogr. A 2018, 1567, 99–110. [Google Scholar] [CrossRef]
  200. Chemat, F.; Vian, M.A. Alternative Solvents for Natural Products Extraction. In Green Chemistry and Sustainable Technology; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar]
  201. Passos, H.; Freire, M.G.; Coutinho, J.A. Ionic liquid solutions as extractive solvents for value-added compounds from biomass. Green Chem. 2014, 16, 4786–4815. [Google Scholar] [CrossRef] [Green Version]
  202. Ventura, S.P.; e Silva, F.A.; Quental, M.V.; Mondal, D.; Freire, M.G.; Coutinho, J.A. Ionic-liquid-mediated extraction and separation processes for bioactive compounds: Past, present, and future trends. Chem. Rev. 2017, 117, 6984–7052. [Google Scholar] [CrossRef] [PubMed]
  203. De Faria, E.L.P.; Shabudin, S.V.; Claúdio, A.F.M.; Válega, M.; Domingues, F.M.J.; Freire, C.S.R.; Silvestre, A.J.D.; Freire, M.G. Aqueous solutions of surface-active ionic liquids: Remarkable alternative solvents to improve the solubility of triterpenic acids and their extraction from biomass. ACS Sustain. Chem. Eng. 2017, 5, 7344–7351. [Google Scholar] [CrossRef] [PubMed]
  204. Fan, C.; Li, N.; Cao, X.; Wen, L. Ionic liquid-modified countercurrent chromatographic isolation of high-purity delphinidin-3-rutinoside from eggplant peel. J. Food Sci. 2020, 85, 1132–1139. [Google Scholar] [CrossRef] [PubMed]
  205. Lima, Á.S.; Soares, C.M.F.; Paltram, R.; Halbwirth, H.; Bica, K. Extraction and consecutive purification of anthocyanins from grape pomace using ionic liquid solutions. Fluid Ph. Equilibria 2017, 451, 68–78. [Google Scholar] [CrossRef]
  206. Mišan, A.; Nađpal, J.; Stupar, A.; Pojić, M.; Mandić, A.; Verpoorte, R.; Choi, Y.H. The perspectives of natural deep eutectic solvents in agri-food sector. Crit. Rev. Food Sci. Nutr. 2020, 60, 2564–2592. [Google Scholar] [CrossRef] [PubMed]
  207. Jeong, K.M.; Jin, Y.; Han, S.Y.; Kim, E.M.; Lee, J. One-step sample preparation for convenient examination of volatile monoterpenes and phenolic compounds in peppermint leaves using deep eutectic solvents. Food Chem. 2018, 251, 69–76. [Google Scholar] [CrossRef]
  208. Jablonský, M.; Šima, J. Phytomass Valorization by Deep Eutectic Solvents-Achievements, Perspectives, and Limitations. Crystals 2020, 10, 800. [Google Scholar] [CrossRef]
  209. De los Ángeles Fernández, M.; Espino, M.; Gomez, F.J.; Silva, M.F. Novel approaches mediated by tailor-made green solvents for the extraction of phenolic compounds from agro-food industrial by-products. Food Chem. 2018, 239, 671–678. [Google Scholar] [CrossRef]
  210. Panzella, L.; Moccia, F.; Nasti, R.; Marzorati, S.; Verotta, L.; Napolitano, A. Bioactive phenolic compounds from agri-food wastes: An update on green and sustainable extraction methodologies. Front. Nutr. 2020, 7, 60. [Google Scholar] [CrossRef]
  211. Roselló-Soto, E.; Koubaa, M.; Moubarik, A.; Lopes, R.P.; Saraiva, J.A.; Boussetta, N.; Barba, F.J. Emerging opportunities for the effective valorization of wastes and by-products generated during olive oil production process: Non-conventional methods for the recovery of high-added value compounds. Trends. Food Sci. Technol. 2015, 45, 296–310. [Google Scholar] [CrossRef]
  212. Burlini, I.; Sacchetti, G. Secondary Bioactive Metabolites from Plant-Derived Food Byproducts through Ecopharmacognostic Approaches: A Bound Phenolic Case Study. Plants 2020, 9, 1060. [Google Scholar] [CrossRef] [PubMed]
  213. Roselló-Soto, E.; Galanakis, C.M.; Brnčić, M.; Orlien, V.; Trujillo, F.J.; Mawson, R.; Barba, F.J. Clean recovery of antioxidant compounds from plant foods, by-products and algae assisted by ultrasounds processing. Modeling approaches to optimize processing conditions. Trends. Food Sci. Technol. 2015, 42, 134–149. [Google Scholar] [CrossRef]
  214. Chemat, F.; Cravotto, G. Microwave-Assisted Extraction for Bioactive Compounds; Springer: New York, NY, USA, 2013. [Google Scholar]
  215. Barba, F.J.; Puértolas, E.; Brnčić, M.; Nedelchev Panchev, I.; Dimitrov, D.A.; Athès-Dutour, V.; Moussa, M.; Souchon, I. Emerging extraction. In Food Waste Recovery: Processing Technologies and Industrial Techniques; Galanakis, C., Ed.; Academic Press: Cambridge, MA, USA, 2015; pp. 249–272. [Google Scholar]
  216. Plazzotta, S.; Ibarz, R.; Manzocco, L.; Martín-Belloso, O. Optimizing the antioxidant biocompound recovery from peach waste extraction assisted by ultrasounds or microwaves. Ultrason. Sonochem. 2020, 63, 104954. [Google Scholar] [CrossRef] [PubMed]
  217. Casazza, A.A.; Pettinato, M.; Perego, P. Polyphenols from apple skins: A study on microwave-assisted extraction optimization and exhausted solid characterization. Sep. Purif. Technol. 2020, 240, 116640. [Google Scholar] [CrossRef]
  218. Kapoor, S.; Gandhi, N.; Tyagi, S.K.; Kaur, A.; Mahajan, B.V.C. Extraction and characterization of guava seed oil: A novel industrial byproduct. Lebensm. Wiss. Technol. 2020, 132, 109882. [Google Scholar] [CrossRef]
  219. Carabias-Martínez, R.; Rodríguez-Gonzalo, E.; Revilla-Ruiz, P.; Hernández-Méndez, J. Pressurized liquid extraction in the analysis of food and biological samples. J. Chromatogr. A 2005, 1089, 1–17. [Google Scholar] [CrossRef]
  220. Benthin, B.; Danz, H.; Hamburger, M. Pressurized liquid extraction of medicinal plants. J. Chromatogr. A 1999, 837, 211–219. [Google Scholar] [CrossRef]
  221. Schantz, M.M. Pressurized liquid extraction in environmental analysis. Anal. Bioanal. Chem. 2006, 386, 1043–1047. [Google Scholar] [CrossRef]
  222. Mustafa, A.; Turner, C. Pressurized liquid extraction as a green approach in food and herbal plants extraction: A review. Anal. Chim. Acta 2011, 703, 8–18. [Google Scholar] [CrossRef]
  223. Wu, C.T.; Agrawal, D.C.; Huang, W.Y.; Hsu, H.C.; Yang, S.J.; Huang, S.L.; Lin, Y.S. Functionality analysis of spent coffee ground extracts obtained by the hydrothermal method. J. Chem. 2019. [Google Scholar] [CrossRef] [Green Version]
  224. Scopus Preview. Available online: https://0-www-scopus-com.brum.beds.ac.uk/term/analyzer.uri?sid=a812e38dc2896c675cc2befb7b3d7df8&origin=resultslist&src=s&s=TITLE-ABS-KEY+%28+plant+AND+by-products+AND+waste+AND+valorization%29&sort=plf-f&sdt=a&sot=a&sl=65&count=204&analyzeResults=Analyze+results&txGid=ad7a147decc8a12b718c1490e34a2899 (accessed on 19 November 2020).
  225. Acquadro, S.; Appleton, S.; Marengo, A.; Bicchi, C.; Sgorbini, B.; Mandrone, M.; Gai, F.; Peiretti, P.G.; Cagliero, C.; Rubiolo, P. Grapevine Green Pruning Residues as a Promising and Sustainable Source of Bioactive Phenolic Compounds. Molecules 2020, 25, 464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Simitzis, P.E. Agro-industrial by-products and their bioactive compounds—An ally against oxidative stress and skin aging. Cosmetics 2018, 5, 58. [Google Scholar] [CrossRef] [Green Version]
  227. Borges-Argaez, R.; Chan-Balan, R.; Cetina-Montejo, L.; Ayora-Talavera, G.; Sansores-Peraza, P.; Gomez-Carballo, J.; Caceres-Farfan, M. In vitro evaluation of anthraquinones from Aloe vera (Aloe barbadensis Miller) roots and several derivatives against strains of influenza virus. Ind. Crops Prod. 2019, 132, 468–475. [Google Scholar] [CrossRef] [PubMed]
  228. Saha, A.; Basak, B.B. Scope of value addition and utilization of residual biomass from medicinal and aromatic plants. Ind. Crops Prod. 2020, 145, 111979. [Google Scholar] [CrossRef]
  229. Sharma, P.; Gaur, V.K.; Sirohi, R.; Larroche, C.; Kim, S.H.; Pandey, A. Valorization of cashew nut processing residues for industrial applications. Ind. Crops Prod. 2020, 152, 112550. [Google Scholar] [CrossRef]
  230. Sepúlveda, L.; Romaní, A.; Aguilar, C.N.; Teixeira, J. Valorization of pineapple waste for the extraction of bioactive compounds and glycosides using autohydrolysis. Innov. Food Sci. Emerg. Technol. 2018, 47, 38–45. [Google Scholar] [CrossRef] [Green Version]
  231. Toomer, O.T. A comprehensive review of the value-added uses of peanut (Arachis hypogaea) skins and by-products. Crit. Rev. Food Sci. 2020, 60, 341–350. [Google Scholar] [CrossRef]
  232. Mandalari, G.; Bennett, R.N.; Bisignano, G.; Trombetta, D.; Saija, A.; Faulds, C.B.; Gasson, M.J.; Narbad, A. Antimicrobial activity of flavonoids extracted from bergamot (Citrus bergamia Risso) peel, a byproduct of the essential oil industry. J. Appl. Microbiol. 2007, 103, 2056–2064. [Google Scholar] [CrossRef]
  233. Gavahian, M.; Chu, Y.H.; Mousavi Khaneghah, A. Recent advances in orange oil extraction: An opportunity for the valorisation of orange peel waste a review. Int. J. Food Sci. Technol. 2019, 54, 925–932. [Google Scholar] [CrossRef]
  234. Pujol, D.; Liu, C.; Gominho, J.; Olivella, M.À.; Fiol, N.; Villaescusa, I.; Pereira, H. The chemical composition of exhausted coffee waste. Ind. Crops Prod. 2013, 50, 423–429. [Google Scholar] [CrossRef]
  235. Valdez-Arjona, L.P.; Ramírez-Mella, M. Pumpkin Waste as Livestock Feed: Impact on Nutrition and Animal Health and on Quality of Meat, Milk, and Egg. Animals 2019, 9, 769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Kumar, A.; Agarwal, K.; Singh, M.; Saxena, A.; Yadav, P.; Maurya, A.K.; Yadav, A.; Tandon, S.; Chanda, D.; Bawankule, D.U. Essential oil from waste leaves of Curcuma longa L. alleviates skin inflammation. Inflammopharmacology 2018, 26, 1245–1255. [Google Scholar] [CrossRef] [PubMed]
  237. Mejri, F.; Baati, T.; Martins, A.; Selmi, S.; Serralheiro, M.L.; Falé, P.L.; Rauter, A.; Casabianca, H.; Hosni, K. Phytochemical analysis and in vitro and in vivo evaluation of biological activities of artichoke (Cynara scolymus L.) floral stems: Towards the valorization of food by-products. Food Chem. 2020, 333, 127506. [Google Scholar] [CrossRef] [PubMed]
  238. Sayed-Ahmad, B.; Straumīte, E.; Šabovics, M.; Krūma, Z.; Merah, O.; Saad, Z.; Hijazi, A.; Talou, T. Effect of addition of fennel (Foeniculum vulgare L.) on the quality of protein bread. Proceedings of the Latvian Academy of Sciences, Section B. Natural, Exact, and Applied Sciences. J. Latvian Acad. Sci. 2017, 71, 509–514. [Google Scholar]
  239. Ortiz-de Elguea-Culebras, G.; Berruga, M.I.; Santana-Méridas, O.; Herraiz-Peñalver, D.; Sánchez-Vioque, R. Chemical Composition and Antioxidant Capacities of Four Mediterranean Industrial Essential Oils and Their Resultant Distilled Solid By-Products. Eur. J. Lipid Sci. Technol. 2017, 119, 1700242. [Google Scholar] [CrossRef]
  240. Gullón, B.; Eibes, G.; Moreira, M.T.; Herrera, R.; Labidi, J.; Gullón, P. Yerba mate waste: A sustainable resource of antioxidant compounds. Ind. Crops Prod. 2018, 113, 398–405. [Google Scholar] [CrossRef]
  241. Jahanban-Esfahlan, A.; Ostadrahimi, A.; Tabibiazar, M.; Amarowicz, R. A comprehensive review on the chemical constituents and functional uses of walnut (Juglans spp.) husk. Int. J. Mol. Sci. 2019, 20, 3920. [Google Scholar] [CrossRef] [Green Version]
  242. Torras-Claveria, L.; Jauregui, O.; Bastida, J.; Codina, C.; Viladomat, F. Antioxidant activity and phenolic composition of lavandin (Lavandula x intermedia Emeric ex Loiseleur) waste. J. Agric. Food Chem. 2007, 55, 8436–8443. [Google Scholar] [CrossRef]
  243. Pagano, I.; Sánchez-Camargo, A.D.P.; Mendiola, J.A.; Campone, L.; Cifuentes, A.; Rastrelli, L.; Ibañez, E. Selective extraction of high-value phenolic compounds from distillation wastewater of basil (Ocimum basilicum L.) by pressurized liquid extraction. Electrophoresis 2018, 39, 1884–1891. [Google Scholar] [CrossRef] [Green Version]
  244. King, A.J.; Griffin, J.K.; Roslan, F. In vivo and in vitro addition of dried olive extract in poultry. J. Agric. Food Chem. 2014, 62, 7915–7919. [Google Scholar] [CrossRef]
  245. Agulló-Chazarra, L.; Borrás-Linares, I.; Lozano-Sánchez, J.; Segura-Carretero, A.; Micol, V.; Herranz-López, M.; Barrajón-Catalán, E. Sweet Cherry Byproducts Processed by Green Extraction Techniques as a Source of Bioactive Compounds with Antiaging Properties. Antioxidants 2020, 9, 418. [Google Scholar] [CrossRef] [PubMed]
  246. Yılmaz, F.M.; Görgüç, A.; Karaaslan, M.; Vardin, H.; Ersus Bilek, S.; Uygun, Ö.; Bircan, C. Sour cherry by-products: Compositions, functional properties and recovery potentials—A review. Crit. Rev. Food Sci. Nutr. 2019, 59, 3549–3563. [Google Scholar] [CrossRef] [PubMed]
  247. Piccirillo, C.; Demiray, S.; Ferreira, A.S.; Pintado, M.E.; Castro, P.M. Chemical composition and antibacterial properties of stem and leaf extracts from Ginja cherry plant. Ind. Crops Prod. 2013, 43, 562–569. [Google Scholar] [CrossRef]
  248. Prgomet, I.; Gonçalves, B.; Dominguez-Perles, R.; Santos, R.; Saavedra, M.J.; Aires, A.; Pascual-Seva, N.; Barros, A. Irrigation deficit turns almond by-products into a valuable source of antimicrobial (poly) phenols. Ind. Crops Prod. 2019, 132, 186–196. [Google Scholar] [CrossRef]
  249. Kharchoufi, S.; Licciardello, F.; Siracusa, L.; Muratore, G.; Hamdi, M.; Restuccia, C. Antimicrobial and antioxidant features of ‘Gabsi’ pomegranate peel extracts. Ind. Crops Prod. 2018, 111, 345–352. [Google Scholar] [CrossRef]
  250. Andrade, M.A.; Lima, V.; Silva, A.S.; Vilarinho, F.; Castilho, M.C.; Khwaldia, K.; Ramos, F. Pomegranate and grape by-products and their active compounds: Are they a valuable source for food applications? Trends Food Sci. Technol. 2019, 86, 68–84. [Google Scholar] [CrossRef]
  251. Legua, P.; Forner-Giner, M.Á.; Nuncio-Jáuregui, N.; Hernández, F. Polyphenolic compounds, anthocyanins and antioxidant activity of nineteen pomegranate fruits: A rich source of bioactive compounds. J. Funct. Foods 2016, 23, 628–636. [Google Scholar] [CrossRef]
  252. Fischer, U.A.; Carle, R.; Kammerer, D.R. Identification and quantification of phenolic compounds from pomegranate (Punica granatum L.) peel, mesocarp, aril and differently produced juices by HPLC-DAD–ESI/MSn. Food Chem. 2011, 127, 807–821. [Google Scholar] [CrossRef]
  253. Alexandre, E.M.; Silva, S.; Santos, S.A.; Silvestre, A.J.; Duarte, M.F.; Saraiva, J.A.; Pintado, M. Antimicrobial activity of pomegranate peel extracts performed by high pressure and enzymatic assisted extraction. Food Res. Int. 2019, 115, 167–176. [Google Scholar] [CrossRef] [Green Version]
  254. Carriço, C.; Ribeiro, H.M.; Marto, J. Converting cork by-products to ecofriendly cork bioactive ingredients: Novel pharmaceutical and cosmetics applications. Ind. Crops Prod. 2018, 125, 72–84. [Google Scholar] [CrossRef]
  255. Plainfossé, H.; Trinel, M.; Verger-Dubois, G.; Azoulay, S.; Burger, P.; Fernandez, X. Valorisation of Ribes nigrum L. Pomace, an Agri-Food By-Product to Design a New Cosmetic Active. Cosmetics 2020, 7, 56. [Google Scholar] [CrossRef]
  256. Sanchez-Vioque, R.; Izquierdo-Melero, M.E.; Quilez, M.; Herraiz-Penalver, D.; Santana-Méridas, O.; Jordan, M.J. Solid Residues from the Distillation of Salvia lavandulifolia Vahl as a Natural Source of Antioxidant Compounds. J. Am. Oil Chem. Soc. 2018, 95, 1277–1284. [Google Scholar] [CrossRef]
  257. Navarrete, A.; Herrero, M.; Martín, A.; Cocero, M.J.; Ibáñez, E. Valorization of solid wastes from essential oil industry. J. Food Eng. 2011, 104, 196–201. [Google Scholar] [CrossRef]
  258. Ma, H.; Huang, Q.; Qu, W.; Li, L.; Wang, M.; Li, S.; Chu, F. In vivo and in vitro anti-inflammatory effects of Sophora flavescens residues. J. Ethnopharmacol. 2018, 224, 497–503. [Google Scholar] [CrossRef] [PubMed]
  259. Weng, Z.; Guo, S.; Qian, D.; Zhu, Z.; Zhang, S.; Li, A.; Lei, Z.; Duan, J. Sophora flavescens Seed as a promising high potential by-product: Phytochemical characterization and bioactivity evaluation. Ind. Crops Prod. 2017, 109, 19–26. [Google Scholar] [CrossRef]
  260. Cabrera-Barjas, G.; Quezada, A.; Bernardo, Y.; Moncada, M.; Zúñiga, E.; Wilkens, M.; Giordano, A.; Nesic, A.; Delgado, N. Chemical composition and antibacterial activity of red murta (Ugni molinae Turcz.) seeds: An undervalued Chilean resource. J. Food Meas. Charact. 2020, 14, 1–12. [Google Scholar] [CrossRef]
  261. Ribeiro, L.F.; Ribani, R.H.; Francisco, T.M.G.; Soares, A.A.; Pontarolo, R.; Haminiuk, C.W.I. Profile of bioactive compounds from grape pomace (Vitis vinifera and Vitis labrusca) by spectrophotometric, chromatographic and spectral analyses. J. Chromatogr. B 2015, 1007, 72–80. [Google Scholar] [CrossRef]
  262. Poveda, J.M.; Loarce, L.; Alarcón, M.; Díaz-Maroto, M.C.; Alañón, M.E. Revalorization of winery by-products as source of natural preservatives obtained by means of green extraction techniques. Ind. Crops Prod. 2018, 112, 617–625. [Google Scholar] [CrossRef]
  263. Leal, C.; Gouvinhas, I.; Santos, R.A.; Rosa, E.; Silva, A.M.; Saavedra, M.J.; Barros, A.I. Potential application of grape (Vitis vinifera L.) stem extracts in the cosmetic and pharmaceutical industries: Valorization of a by-product. Ind. Crops Prod. 2020, 154, 112675. [Google Scholar] [CrossRef]
  264. Pantelić, M.M.; Zagorac, D.Č.D.; Davidović, S.M.; Todić, S.R.; Bešlić, Z.S.; Gašić, U.M.; Tešić, Z.L.; Natić, M.M. Identification and quantification of phenolic compounds in berry skin, pulp, and seeds in 13 grapevine varieties grown in Serbia. Food Chem. 2016, 211, 243–252. [Google Scholar] [CrossRef]
  265. Maier, T.; Schieber, A.; Kammerer, D.R.; Carle, R. Residues of grape (Vitis vinifera L.) seed oil production as a valuable source of phenolic antioxidants. Food Chem. 2009, 112, 551–559. [Google Scholar] [CrossRef]
  266. Farhadi, K.; Esmaeilzadeh, F.; Hatami, M.; Forough, M.; Molaie, R. Determination of phenolic compounds content and antioxidant activity in skin, pulp, seed, cane and leaf of five native grape cultivars in West Azerbaijan province, Iran. Food Chem. 2016, 199, 847–855. [Google Scholar] [CrossRef] [PubMed]
  267. Soares, S.C.S.; De Lima, G.C.; Laurentiz, A.C.; Féboli, A.; Dos Anjos, L.A.; de Paula Carlis, M.S.; Filardi, R.D.S.; de Laurentiz, R.D.S. In vitro anthelmintic activity of grape pomace extract against gastrointestinal nematodes of naturally infected sheep. Int. J. Vet. Sci. 2018, 6, 243–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. World Checklist of Vascular Plants. Available online: https://wcvp.science.kew.org/ (accessed on 14 December 2020).
Figure 1. Scheme of the “waste hierarchy” proposed by the EPA [8].
Figure 1. Scheme of the “waste hierarchy” proposed by the EPA [8].
Molecules 26 00495 g001
Figure 2. Examples of polyphenols, basic nucleus and classification.
Figure 2. Examples of polyphenols, basic nucleus and classification.
Molecules 26 00495 g002
Figure 3. Some examples of terpene chemical structures and classification according to the number of isoprene units contained in the structure.
Figure 3. Some examples of terpene chemical structures and classification according to the number of isoprene units contained in the structure.
Molecules 26 00495 g003
Figure 4. Examples of alkaloids (basic skeletons) and their precursors.
Figure 4. Examples of alkaloids (basic skeletons) and their precursors.
Molecules 26 00495 g004
Figure 5. Trend of scientific publications from 2001 to 2020 reported by Scopus using as a query string: TITLE-ABS-KEY (plant AND by-products AND waste AND valorization).
Figure 5. Trend of scientific publications from 2001 to 2020 reported by Scopus using as a query string: TITLE-ABS-KEY (plant AND by-products AND waste AND valorization).
Molecules 26 00495 g005
Figure 6. Number of plant matrices investigated for potential application in four different areas: agriculture, OIA (Other Industrial Applications), HHF (Human Healthcare and Food), and veterinary science. The graphic summarizes the results of our bibliographic survey (from 2006 to 2020).
Figure 6. Number of plant matrices investigated for potential application in four different areas: agriculture, OIA (Other Industrial Applications), HHF (Human Healthcare and Food), and veterinary science. The graphic summarizes the results of our bibliographic survey (from 2006 to 2020).
Molecules 26 00495 g006
Table 1. Summary of the results emerging from the survey. The table reports: plant species from which the waste is generated, the industrial sector manufacturing the plant material, the nature of the produced waste, the class of secondary metabolites found in the waste material, the bioactivity possessed by waste extract(s), the potential use for waste valorization, and the corresponding reference source of all the information reported in a line. If no biological activity was investigated, or no potential valorization in a specific field was proposed, the wording “not specified” (n. s.) was inserted in the corresponding cell.
Table 1. Summary of the results emerging from the survey. The table reports: plant species from which the waste is generated, the industrial sector manufacturing the plant material, the nature of the produced waste, the class of secondary metabolites found in the waste material, the bioactivity possessed by waste extract(s), the potential use for waste valorization, and the corresponding reference source of all the information reported in a line. If no biological activity was investigated, or no potential valorization in a specific field was proposed, the wording “not specified” (n. s.) was inserted in the corresponding cell.
Plant Species 1Waste Producer IndustryWaste TypeContained Secondary MetabolitesReported BioactivityPotential Use for Waste ValorizationReferences
Acacia mangium Willd.Forestrybarktanninsn. s.leather tanning[56]
Aesculus hippocastanum L.Herbal industrywastewatersflavonoidsn. s.n. s.[79]
Agave sisalana PerrineFiber Industryleavessaponinsn. s.cosmetic, pharmacy[151]
soil bioremediation
Allium cepa L.Foodpeelflavonoids, saponins anti-obesitypharmacy[44]
surface activityfood additive[152]
pomacephotoprotectioncosmetic[226]
Allium sativum L.Foodn. s.flavonoidsphotoprotectioncosmetic[226]
Aloe barbadensis Mill.Agricultureroots anthraquinonesantiviralpharmacy[227,228]
Anacardium occidentale L.Foodnut shell liquid (CNSL), testa, cashew apple, and cashew apple bagassephenolic acids, alkaloids, tanninspesticide, larvicide, anti-termite, dyes, anti-cancer, anti-bacterial, antioxidant, neurotransmitter, etc. colorant[229]
crop protection
pharmacy
Ananas comosus (L.) Merr.Foodcore and skinpolyphenolsantioxidant in vitro food additive[230]
Arachis hypogaea L.Foodskin and husksphenolic acids, flavonoids polyphenolsantioxidant in vitro feed additive[231]
antibacterialpharmacy
Aronia melanocarpa (Michx.) ElliotBeveragepomaceflavonoidsantioxidant in vitro feed additive[74]
phenolic acids
Camellia sinensis (L.) KuntzeFoodblack tea wastepolyphenolsantioxidant in vitro cosmetic, food additive, pharmacy[97]
antibacterial
Castanea sativa Mill.Agriculture and Foodburspolyphenolsfungicide, enzyme inhibitor and antioxidant in vitrocosmetic, crop protection, food additive, pharmacy[86,87,88,89,90]
leavesenzyme inhibitor, antitumoral and antioxidant in vitro [86,87,88]
budn. s.[93]
husksenzyme inhibitor[86,88]
antioxidant in vitro
husks and barktanninsn. s.cosmetic, pharmacy, industrial application[92,94]
Citrus spp.Food and Beveragepeel and pulpflavonoids, essential oil, phenolic acidantioxidant in vitro, anti-inflammatory in vitro, antitumoralcosmetic, food additive, pharmacy[27,226,228]
Citrus x bergamia (Risso) Risso & Poit.Herbal Industrypeelflavonoidsantibacterialpharmacy[228,232]
Citrus x sinensis (L.) OsbeckBeveragepeelessential oiln. s.n. s.[233]
Coffea spp.Beverage and Food spent coffee grounds and silverskinpolyphenols and alkaloidsantioxidant in vitro cosmetic, food additive, nutraceutical, pharmacy[80,81,84,189]
spent coffee groundsn. s.anti-pollutants[234]
spent coffee groundspolyphenolsallelopathic activityherbicide[82]
Coffea arabica L.Foodsilverskinpolyphenolsantioxidant in vitro n. s.[85]
Cucumis melo L. Food and Beveragepeelpolyphenolsantioxidant in vitro cosmetic, nutraceutical, food additive, pharmacy[98]
Cucurbita spp.Foodpeel, seeds, and fruits not suitable for human consumptionterpenes and phenolic acidsimprovement of meat, milk, or eggsfeed additive[235]
Curcuma longa L.Agricultureleavesterpenesanti-inflammatory in vivo and in vitropharmacy[236]
Cynara scolymus L.Foodfloral stemspolyphenols, flavonoids, tanninsantioxidant in vitro, antibacterial, anti-denaturating protein, antidiabetic in vivo and anti-hyperlipidemic in vivonutraceutical[237]
Agriculturebracts of the headscosmetic[75]
Datura stramonium L.Agricultureleaves and flowersterpenesantibacterialpharmacy[155]
Daucus carota L.Beveragepomaceterpenesantioxidant in vitro feed additive[74]
Foeniculum vulgare Mill.Herbal Industryseeds solid residue from distillationflavonoids and polyphenolsantibacterial and antioxidant in vitro food additive[228,238]
Agricultureleaves, inflorescence, and pseudo stemsterpenesantioxidant in vitron. s.[137]
Fragaria spp.Beveragepomaceflavonoidsantioxidant in vitro feed additive[74]
Humulus lupulus L.Beverageleaves and flowersterpenesantibacterialpharmacy[155]
Hyssopus officinalis L.Herbal Industrysolid distillation residuephenolic acidsantioxidant in vitro food additive[228,239]
Ilex paraguariensis A. St.-Hil.Beverage exhausted leavesterpenes, flavonoids, polyphenolsantioxidant in vitro n. s.[240]
Juglans spp.Agriculture and Foodhuskstannins, flavonoids, terpenes, phenolic acids, anthraquinones, naphthoquinones removal of hazardous materials, antioxidant in vitro, antibacterial, anti-platelet, cytotoxiccosmetic, pharmacy [241]
colorant
Larix kaempferi (Lamb.) CarrièreForestrybarkflavonoidsenzyme inhibitorcosmetic[99]
Lavandula x intermedia Emeric ex Loisel.Perfume Industrysolid distillation residueflavonoids, polyphenols, phenolic acids antioxidant in vitro n. s.[228,242]
Lycopersicon esculentum Mill. Agricultureleavespolyphenols and alkaloidsenzyme inhibitorpharmacy[194]
Lycopersicon lycopersicum (L.) H. Karst. Foodpeelpolyphenolsantioxidant in vitro cosmetic, food additive, pharmacy [75]
Malus spp.Beveragepomaceflavonoids and polyphenolsantioxidant in vitro, additive cosmetic, food additive, pharmacy [74,75,217]
Malus domestica (Suckow) Borkh.Food and Beveragepomaceflavonoids and phenolic acidsantioxidant in vitro, photoprotectioncosmetic, feed additive[74,226]
Mangifera indica L. Foodpomacepolyphenolsphotoprotectioncosmetic[226]
Ocimum basilicum L.Food and Herbal Industryoil distillation wastewaterspolyphenolsantioxidant in vitro cosmetic, food additive, nutraceutical, pharmacy[78,228,243]
Olea europaea L.Fooddistillation wastewater, pomace, leavespolyphenolsantioxidant in vitro, antibacterial, decrease lipolysiscosmetic, food additive, nutraceutical, pharmacy[69,70,71,153,226,244]
leaves, fruit milled wasteterpenesantiallergicnutraceutical, pharmacy[153,154]
pomacepolyphenolsantioxidant in vitro, improvement of gut microbiotafeed additive[68]
Pinus spp.Forestryresinterpenesimprove thermal stability of bio-based materialstechnological application[149]
organocatalyst[150]
Pinus pinaster AitonAgriculturebarktanninstanningleather tanning[95]
Poliomintha longiflora A. GrayHerbal Industrysolid distillation residuepolyphenols, phenolic acids, acids, terpenesantibacterial and antioxidant in vitro n. s.[136,228]
Prunus avium (L.) L.Foodpeel and stemspolyphenolsantioxidant in vitro, enzyme inhibitor and photoprotection cosmetic[245]
Prunus cerasus L.Food and Beveragecherry liquor pomace, pomace, and seedspolyphenolsantioxidant in vitrocosmetic, nutraceutical[77,246,247]
stems, leaves, pomace, and seedsterpenesantibacterial nutraceutical
Prunus dulcis (Mill.) D. A. WebbAgriculture and Foodhusks, skins, and blanching waterpolyphenols and flavonoids antibacterial, antioxidant in vitropharmacy, technological application[248]
Punica granatum L.Food and Beveragepeel, arils, mesocarp, and pulppolyphenolsantioxidant in vitro, antibacterial, photoprotection cosmetic[226,249,250,251,252,253]
Pyrus spp.Beveragepomacepolyphenolsantioxidant in vitro cosmetic[75]
Quercus suber L.Forestryby-product of bark processingphenolic acids and tanninsantioxidant in vitro cosmetic[254]
Ribes nigrum L. Beveragepomacetannins and flavonoidsdyes, antioxidant in vitro feed additive[74]
phenolic acids and flavonoidsenzyme inhibitor, antioxidant in vitro cosmetic[48,255]
Salvia officinalis L.Herbal industry and Foodoil distillation wastewaterspolyphenols antioxidant in vitro n. s.[78]
Salvia officinalis subsp. Lavandulifolia (Vahl) GamsHerbal Industrysolid distillation residuephenolic acidsantioxidant in vitro cosmetic, food additive[228,256]
Salvia rosmarinus Spenn.Herbal Industrysolid distillation residuepolyphenols and phenolic acidsantioxidant in vitro food additive[228,257]
oil distillation wastewatersterpenes and polyphenolsantioxidant in vitro n. s.[78]
Santolina chamaecyparissus L.Herbal Industrysolid distillation residuephenolic acidsantioxidant in vitro pharmacy[228,239]
Satureja montana L.Herbal Industrysolid distillation residuephenolic acidsantioxidant in vitro pharmacy[228,239]
Solanum lycopersicum L.Foodpomaceterpenesantioxidant in vitro, photoprotection, chemoprevention cosmetic, nutraceutical[29,226]
Solanum tuberosum L.Foodpeelpolyphenols and alkaloidsantioxidant in vitro n. s.[193]
peelanti-trichomonadspharmacy[192]
Solidago virgaurea L.Herbal industrydried herbpolyphenolsantioxidant in vitro cosmetic, food additive [75]
Sophora flavescens AitonHerbal Industryall plantsflavonoids and alkaloidsanti-inflammatory in vitropharmacy[258]
Agricultureseedsalkaloidsantioxidant in vitro and antibacterial n. s.[259]
Theobroma cacao L.Foodbean shellphenolic acidsantioxidant in vitro food additive[195]
polyphenols and alkaloidsn. s.[190]
flavonoids and alkaloidsn. s.[96]
Thymus mastichina (L.) L.Herbal Industrysolid distillation residuephenolic acidsantioxidant in vitro n. s.[228,256]
Ugni molinae TurczFood and Herbal Industryseedspolyphenolsantibacterialn. s.[260]
Vaccinium myrtillus L.Beveragepomaceflavonoidsdyescolorant[76]
Vitis labrusca L.Beveragepomacepolyphenolsn. s.nutraceutical[250,261]
Vitis vinifera L.Agriculturegreen pruning polyphenolsantioxidant in vitro cosmetic, nutraceutical, pharmacy[225]
Beverageseeds, pomace, and stemsflavonoids, polyphenolsantioxidant in vitro, antibacterial[225,250,262,263,264,265]
pomaceflavonoids, polyphenolsantioxidant in vitro, photoprotection, cholesterol-lowering activities [61,63,226,250,251,261,264,266]
pomacepolyphenols, tannins, flavonoids, saponinsantioxidant in vivo, anthelmintic feed additive, pharmacy[62,267]
pomacetanninsadhesive propertiesadhesive[64]
Zea mays L.Agriculturemaize bran phenolic acidsn. s.n. s.[100]
1 Plant scientific names have been updated following the World Checklist of Vascular Plants (WCVP 2020) [268].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chiocchio, I.; Mandrone, M.; Tomasi, P.; Marincich, L.; Poli, F. Plant Secondary Metabolites: An Opportunity for Circular Economy. Molecules 2021, 26, 495. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26020495

AMA Style

Chiocchio I, Mandrone M, Tomasi P, Marincich L, Poli F. Plant Secondary Metabolites: An Opportunity for Circular Economy. Molecules. 2021; 26(2):495. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26020495

Chicago/Turabian Style

Chiocchio, Ilaria, Manuela Mandrone, Paola Tomasi, Lorenzo Marincich, and Ferruccio Poli. 2021. "Plant Secondary Metabolites: An Opportunity for Circular Economy" Molecules 26, no. 2: 495. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26020495

Article Metrics

Back to TopTop