Next Article in Journal
Induction of Monoterpenoid Oxindole Alkaloids Production and Related Biosynthetic Gene Expression in Response to Signaling Molecules in Hamelia patens Plant Cultures
Next Article in Special Issue
Oxidative Stress Response Mechanisms Sustain the Antibacterial and Antioxidant Activity of Quercus ilex
Previous Article in Journal
Phytochemical Profiles of Plant Materials: From Extracts to Added-Value Ingredients
Previous Article in Special Issue
The Different Composition of Coumarins and Antibacterial Activity of Phlojodicarpus sibiricus and Phlojodicarpus villosus Root Extracts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Separation Methods of Phenolic Compounds from Plant Extract as Antioxidant Agents Candidate

by
Ike Susanti
1,
Rimadani Pratiwi
1,
Yudi Rosandi
2 and
Aliya Nur Hasanah
1,3,*
1
Pharmaceutical Analysis and Medicinal Chemistry Department, Faculty of Pharmacy, Universitas Padjadjaran, Jl Raya Bandung Sumedang KM 21 r, Sumedang 45363, Indonesia
2
Faculty of Mathematics and Natural Sciences, Universitas Padjadjaran, Jl. Raya Bandung Sumedang KM 21, Sumedang 45363, Indonesia
3
Drug Development Study Center, Faculty of Pharmacy, Universitas Padjadjaran, Jl. Raya Bandung Sumedang KM 21, Sumedang 45363, Indonesia
*
Author to whom correspondence should be addressed.
Plants 2024, 13(7), 965; https://doi.org/10.3390/plants13070965
Submission received: 27 February 2024 / Revised: 18 March 2024 / Accepted: 22 March 2024 / Published: 27 March 2024

Abstract

:
In recent years, discovering new drug candidates has become a top priority in research. Natural products have proven to be a promising source for such discoveries as many researchers have successfully isolated bioactive compounds with various activities that show potential as drug candidates. Among these compounds, phenolic compounds have been frequently isolated due to their many biological activities, including their role as antioxidants, making them candidates for treating diseases related to oxidative stress. The isolation method is essential, and researchers have sought to find effective procedures that maximize the purity and yield of bioactive compounds. This review aims to provide information on the isolation or separation methods for phenolic compounds with antioxidant activities using column chromatography, medium-pressure liquid chromatography, high-performance liquid chromatography, counter-current chromatography, hydrophilic interaction chromatography, supercritical fluid chromatography, molecularly imprinted technologies, and high-performance thin layer chromatography. For isolation or purification, the molecularly imprinted technologies represent a more accessible and more efficient procedure because they can be applied directly to the extract to reduce the complicated isolation process. However, it still requires further development and refinement.

Graphical Abstract

1. Introduction

Herbal medicine, also known as a phytopharmaceutical preparation is made exclusively from a whole plant or parts of plants. It can be manufactured in a crude form or as a purified pharmaceutical formulation [1]. Herbal products are readily available in the market. In Saudi Arabia, herbal medicine use was reported to range from 10.3% to 75.0% in 2019 [2]. In Indonesia, the most commonly used traditional medicine is for cancer or malignant tumors, with a prevalence of 14.4%. Joints/rheumatism and high cholesterol have the same prevalence of 11.3%. This is followed by stroke (10.2%), diabetes (9.9%), kidney disease (9.7%), and liver (8.0%). In China, pregnant individuals have been using Chinese herbal medicines for a long time. The usage rate is relatively high, with 65.7% of the population using them. Of this, 6.1% use them during pregnancy, while 55.6% use them after delivery [3].
In herbal medicine, secondary metabolites play a crucial role as they are responsible for the clinical effects [4]. Due to their diverse and specific biological activities, secondary metabolites are considered a valuable source of lead molecules for developing new drugs. Therefore, they are continuously being studied and explored for their potential in drug development. Some of these compounds even have the ability to act in additive or synergistic ways [5]. Phenolic compounds are secondary metabolites that have potent biological activity and are commonly found in various types of plants [6]. Phenolic compounds have many activities, such as antimicrobial [7,8,9], and anti-inflammatory effects and can aid in treating diseases like obesity [10], cancer [11], and diabetes [12], and are antioxidant [13,14,15,16,17].
Antioxidants play a crucial role in preventing the process of oxidation. Oxidation is a chemical reaction that can produce free radicals and cause chain reactions, which can lead to significant damage to cells in organisms [18,19]. As antioxidants, phenolic compounds can act as radical scavengers. The hydroxyl group on the phenolic ring can transfer its hydrogen atom to a free radical, forming a delocalized and stabilized unpaired electron, phenoxy radical, across the phenolic ring [20]. The stabilization by the resonance effect of the aromatic nucleus prevents the continuation of the free radical chain reaction [21] (Figure 1).
Free radicals, Reactive oxygen species (ROS) and nitrogen species (RNS), such as superoxide, hydroxyl, and nitric oxide radicals can cause DNA damage and oxidize lipids and proteins in cells in the biological system [22,23,24]. Excessive levels of ROS can lead to oxidative stress, which can cause changes in various organ systems. Oxidative stress occurs when there is an imbalance between ROS or highly reactive compounds and antioxidants, leading to disrupted redox processes, and molecular damage because of insufficient antioxidant function [25]. Excessive levels of ROS can cause neurotoxicity [26], myocardial hypertrophy and fibrosis [27], hepatocyte dysfunction and apoptosis [28], and insulin resistance [29]. An abnormal level of ROS was identified as a predisposing factor for cell transformation, triggering pro-oncogenic signaling pathways, altering gene expression, and causing genomic instability and DNA damage [25]. The DNA damage or mutation can lead to cancer [30]. The ROS/RNS induce inflammatory cells to damaged tissue sites, which can contribute to the progression of cancer (Supporting Information Figure S1). ROS/RNS cause oxidative stress and nitrosative stress, which are also activated by chronic inflammation that induces inflammatory cells and activates gene expression through various pathways [31]. Inflammatory cells discharge cytokines that initiate the oxidation and nitration of lipids, proteins, and carbohydrates. Furthermore, the transcription factors and cytokines elicit apoptosis, causing an imbalance between pro-apoptotic and apoptotic gene surroundings [31]. This imbalance results in regeneration and cell death, eventually leading to cancer through gene modifications, mutations, proliferation, angiogenesis, and other associated mechanisms [31].
Natural antioxidants, such as polyphenols or phenolic compounds, are widely distributed in food and medicinal plants and are known for their anti-inflammatory, anti-aging, anti-atherosclerosis, anticancer and antioxidant properties [32]. Hibiscetin-3- glucoside is a flavonoid compound isolated from the petals of Hibiscus rosa sinensis. Hibiscetin-3- glucoside has excellent great antioxidant activities, as compared to the standard ascorbic acid. The Hibiscetin-3- glucoside could be utilized for scavenging free radicals, preventing the formation of toxic products, and maintaining the shelf life of food and pharmaceuticals [33]. The ethyl acetate fraction of Anacardium occidentale L. (Anacardiaceae) leaf contained antioxidant compounds, such as agathisflavone, and a mixture of quercetin 3-O-rutinoside and quercetin 3-O-rhamnoside. The mixture of quercetin 3-O-rutinoside and quercetin 3-O-rhamnoside (2:1) was the most effective in scavenging free radicals in the DPPH assay, with an IC50 value of 0.96 ± 0.01 µg/mL. That mixture also exhibited the highest activities in the total antioxidant capacity (TAC) and ferric-reducing antioxidant power (FRAP) assay [34].
The process of developing bioactive compounds from natural products into drugs has remained challenging, starting from the screening of natural products, the isolation of bioactive compounds, the characterization and optimization of the bioactive compound, the determination of the mechanism of action, and pharmaceutical development [35,36]. The isolation of secondary metabolites is a critical step before biological characterization. It is also an essential method for obtaining compounds that are hard to synthesize or for which there are no commercial standards. Many isolation methods are used to isolate phenolic compounds that have antioxidant activity, including column chromatography [37], high-performance liquid chromatography (HPLC) [38], medium-pressure liquid chromatography (MPLC) [39], centrifugal partition chromatography (CPC) [40], high-speed counter-current chromatography (HSCCC) [41], and high-performance counter-current chromatography (HPCCC) [42]. Molecularly imprinted polymer techniques (MITs) have also been developed to selectively isolate phenolic compounds from extracts producing synthetic polymeric materials with homologous adsorption sites to a template molecule [43].
Phenolic compounds are known for their antioxidant properties and are widely distributed in natural products. Many reviews from 2021 to 2023 have highlighted the extraction of phenolic compounds in natural products [44,45] or phenolic compounds with antioxidant activity [46]. However, there are still only a few reviews that focus on the isolation methods of antioxidant phenolic compounds.
One such review was conducted by Shi et al. in 2022, in which they discussed the extraction, separation, and characterization methods, as well as the determination of antioxidant activity of phenolic compounds. The separation and characterization methods are membrane filtration, solid-phase extraction (SPE)-GC/LC (gas chromatography/liquid chromatography), liquid chromatography–mass spectrometry (LC-MS/MS), HPLC, capillary electrophoresis (CE), CCC, and CPC [21]. However, based on articles published in 2017–2023, several methods have not been discussed, including hydrophilic interaction chromatography (HILIC), supercritical fluid chromatography (SFC), MPLC, high-performance thin layer chromatography (HPTLC), and MITs methods. Therefore, this review aims to focus on the isolation methods for phenolic antioxidant compounds, including those not discussed in previous reviews.
This review will highlight the improvement in isolation methods for phenolic compounds with antioxidant activities using column chromatography, MPLC, HPLC, HILIC, CCC, SFC, MITs, and HPTLC as well as the combination of the methods to achieve more selective and efficient isolation methods. During the isolation process, a large amount of organic solvents like methanol, n-hexane, acetone, chloroform, benzene, and petroleum ether are required. However, the use of these solvents has various disadvantages, such as being flammable, explosive, poorly biodegradable, and toxic for the final product [47]. Thus, it is necessary to find an environmentally friendly and safe solvent to be used for the isolation process. This review also discusses using environmentally friendly green solvents for isolating, such as deep eutectic solvents (DES).

2. Phenolic Compound

Phenolic compounds are natural metabolites that arise from the shikimate/phenylpropanoid pathway. This pathway directly provides phenylpropanoids, which are characterized by an aromatic ring with one or more hydroxyl substituents [48]. Figure 2 illustrates the various classes of phenolic compounds.
Simple phenolic compounds are substituted phenolic compounds with a C6 skeleton (Figure 2). The group, denoted by “R”, can be an organic group such as alkyl, alkenyl, aryl, hydroxy, alkoxy, amino, etc. It can be present in ortho (o), meta (m), or para (p) positions of the aromatic ring. There are three groups of simple phenolic compounds: simple phenolic, phenolic acids (hydroxybenzoic acids and hydroxycinnamic acids), and coumarins [49]. Simple phenolic compounds can be hydroxy phenols, dihydroxy benzenes, or trihydroxy benzenes. The compounds that belong to this classification are resorcinol, catechol, and pyrogallol [49]. Pyrogallol is a trihydroxy phenol isolated from the stem bark of Barringtonia asiatica. The pure compound exhibited significant biological activity, cytotoxicity, and antioxidant potential [50].
Phenolic acids are phenolic compounds that contain a carboxylic acid group [49]. Phenolic acids can be divided into two subgroups:
  • Hydroxybenzoic acids (Figure 2), the carboxylic acid group is directly attached to the phenol ring, the resulting phenolic compound (C6-C1). Examples: salicylic acid, protocatechuic acid, gallic acid.
  • Hydroxycinnamic acids (Figure 2), the carboxylic acid group and the phenol ring are separated by two doubly bonded carbon atoms (C6-C3). Examples: sinapic acid, ferulic acid, and caffeic acid [49,51].
Scopoletin or 7-hydroxy-6-methoxy coumarin is an example of a phenolic coumarin that contains two aromatic rings with a hydroxyl and methoxy group, as well as an oxo group [52]. Scopoletin can be isolated from many plants such as Eupatorium laevigatum [53], the root of Hypochaeris radicata [54], and Lasianthus lucidus Blume [54].
Flavonoids are a group of compounds that have (C6–C3–C6) as their basic skeleton, consisting of two aromatic rings connected to each other through a central three-carbon bridge [51]. Flavonoids are a diverse group of natural substances with phenolic structures that can be found in various sources including fruits, vegetables, grains, bark, roots, stems, flowers, tea, and wine [55]. The subclasses of flavonoids found in nature include anthocyanins, flavones, flavonols, flavanones, isoflavones, and flavanonols (Figure 2) [51]. For example, quercetin and kaempferol are the most common compounds in natural products. They are compounds that belong to the flavonols subclass. Quercetin and kaempferol are commonly found in polyphenols in fruits and vegetables. They are typically conjugated with sugar molecules in plants [56].
Tannins are present in various species across the plant kingdom, where they serve the purpose of safeguarding the plant against predators and potentially aiding in the regulation of plant growth. These tannins can be divided into two primary groups: hydrolyzable tannins and condensed tannins [57]. Gallotannins are the simplest hydrolyzable tannins. They contain gallic acid substituents esterified with a polyol residue, primarily D-glucose [58]. Condensed tannins are the most abundant polyphenols derived from plants. Condensed tannins are polymeric phenolic compounds made of catechin units and yield anthocyanidin when depolymerized, and are commonly known as proanthocyanidins [49]. Tannins exhibit various pharmacological effects, such as antioxidant and free radical scavenging activities, antimicrobial, anti-cancer, anti-nutritional, and cardio-protective properties [58].
Stilbenes are polyphenolic compounds with a C6-C2-C6 structure that are derived from the secondary metabolism of plants [59]. The most well-known stilbenoid is resveratrol, but there are other interesting compounds like astringin and isorhapontin that are derived from forest biomass and could potentially serve as starting materials for new products [60]. Resveratrol has antioxidant activity that affects the cardiovascular system [61]. Stilbenes have been successfully isolated from the bark of Norway Spruce roots [60] and the grape cane of Vitis vinifera L. [62].
Lignans are dimers of phenylpropanoid units linked by the central carbons of their side chains [63]. Lignans are widely distributed throughout the plant kingdom and can be found in various plant parts, including flowers, fruits, roots, rhizomes, leaves, seeds, stems, xylem, and resins [64]. Lignans have various pharmacological activities such as antioxidant [65], and anti-inflammation [66], and can treat breast cancer [67].

3. Isolation and Purification Method

The extraction of phenolic compounds from natural products has been conducted with several methods, such as maceration [68], percolation [69], Soxhlet extraction [68], reflux extraction [69], decoction [69], ultrasound-assisted extraction (UAE) [70], pulsed electric filed extraction (PEF) [71], enzyme-assisted extraction (EAE) [72], microwave-assisted extraction (MAE) [73], homogenate-assisted extraction (HAE) [74], and subcritical water extraction (SWE) [75]. An extract from plant material is a complex mixture containing different types of natural compounds with different polarities. Therefore, further separation and purification are required to obtain pure bioactive compounds [36]. Several methods have been used to isolate or purify the phenolic compounds that have antioxidant activity. HPLC, MPLC, CCC, HILIC, SFC, column chromatography, and MITs were the methods used for separating active compounds from natural products, with new improvements from 2017 to 2023. In addition, molecularly imprinted polymers also have been developed to isolate compounds with antioxidant activity.

3.1. Medium Pressure Liquid Chromatography

MPLC is one of a wide variety of preparative column chromatography techniques. Separation under pressure allows the use of smaller particle-size supports, increasing the variety of stationary phases that can be used [76]. The instrumentation of MPLC is present in Supporting Information Figure S2, which consists of a pump for the mobile phase, a sample injection system, and a self-packed stationary phase (column). Compound separation can be followed automatically by detectors and recorders connected to the outlet of the column or monitored manually by thin-layer chromatography, then collected via a fraction collector [76].
Although MPLC can be used to isolate a compound [77], it is generally used to enrich bioactive compounds (secondary metabolites) from natural products before further purification because of its low cost, high sample load, and high throughput [78]. Therefore, MPLC can be used to remove non-target compounds. The operation pressure used in MPLC is 75–300 psi [79].
The MPLC system can be applied to adsorption, partition, affinity, or ion-exchange chromatography [80]. The materials used in the stationary phase include silica gel, MCI GEL® CHP20P, and Sephadex LH-20 [81,82,83]. The choice of column type is critical in MPLC because it can affect the purity and yield. One of the MPLC developments is the use of polyamide columns and MCI GEL® CHP20P. Polyamide column chromatography is a widely used method for separating polyphenols. Polyamide has the ability to adsorb anions through electrostatic interactions [84]. MCI GEL® CHP20P is an adsorption resin that consists of a reversed-phase resin bound to a polystyrene matrix, offering excellent hydrophobicity and efficient separation for polar compounds [39]. The application of polyamide coupled with MCI GEL® CHP20P in MPLC was carried out by Dang et al. [39] to purify bergenin from Saxifraga atrata extract. Bergenin is a compound that is derived from trihydroxybenzoic acid (phenolic acid) and has a glycoside attached to it [85]. In the first step, the bergenin fraction was concentrated on a polyamide column (15 × 460 mm) and eluted using a water/acetonitrile (ACN) gradient system. Then, the desired fractions were purified in the second step using MCI GEL® CHP20P (15 × 460 mm). The elution used water/ACN in isocratic mode (5% ACN, 60 min). The authors obtained 1.2 g of bergenin from 180 g of S. atrata dry plant material. After purification using MCI GEL® CHP20P, they obtained 714.2 mg of bergenin with >99% purity [39]. Other methods for isolating the bergenin are listed in Table 1.
Based on Table 1, when viewed based on the percentage yield and purity, the polyamide column coupled with MCI GEL® CHP20P is quite promising, but this method requires instruments during the isolation process. The vacuum liquid chromatography column method could also be chosen: it has a good yield, but the purity is unknown because the authors did not mention it in the article.
Sometimes MPLC is combined with HPLC or preparative HPLC. In this case, MPLC was used for pre-treatment to enrich bioactive compounds in the extract, while HPLC was used to screen and purify the compound. Dawa et al. [89] were the first to combine MPLC and HPLC coupled with online HPLC-1,1-diphenyl-2-picrylhydrazyl (DPPH) detection. They isolated DPPH inhibitors from S. atrata. They pre-treated the extract with MPLC using MCI GEL® CHP20P (49 × 460 mm) as a stationary phase and eluted with a mixture of methanol and water, yielding 1.4 g of the target DPPH inhibitors (11.9% recovery). Then, the authors purified the compounds by using HPLC with RP-C18 followed by HILIC column separation. They obtained four phenolic compounds, ethyl gallate, 11-O-galloylbergenin, rutin, and isoquercitrin with >95% purity [89].
Four DPPH inhibitors have been also isolated from the methanol extract of Saxifraga sinomontana, including 3-methoxy-4-hydroxyphenol-(60-O-galloyl)-1-O-β-D-glucopyranoside (1), 3,4,5-trimethoxyphenyl-(60-O-galloyl)-1-O-β-D-glucopyranoside (2), Saximonsin A (3), and Saximonsin B (4). Compounds 1 and 2 were phenylpropanoid glycoside, while compounds 3 and 4 were phenolic acid. They exhibited potent antioxidant activity with IC50 values of 39.6 mM, 46.9 mM, 11.4 mM, and 20.6 mM, respectively [90]. The studies that have applied MPLC to isolate the phenolic antioxidant compounds are described in Table 2.

3.2. High Performance Liquid Chromatography

HPLC is a separation and purification method used to obtain compounds with a high purity. One application is the isolation of phenolic compounds. The recycle-HPLC has advantages in separating and isolating compounds, such as increased throughput, shorter separation times, reduced solvent consumption, and full compound recovery [91]. The applications of HPLC in separating or purifying phenolic compounds are listed in Table 3.
Reverse-phase HPLC (RP-HPLC) is the most commonly used HPLC mode; the stationary phase is less polar than the eluting solvent. A common RP-HPLC stationary phase is surface-modified silica. Usually, modifications were made using RMe2SiCl, where R is a linear alkyl group [92]. The eluent used in RP-HPLC is usually a mixture of water and a miscible organic solvent (ACN, methanol, or tetrahydrofuran [THF]) [93]. Semi-preparative RP-HPLC has been used to improve the purification of the isolated compound. Rutin is widely recognized as a powerful natural antioxidant that effectively reduces oxidative stress [94]. Rutin is a glycoside of quercetin, consisting of glucose and rhamnose sugars attached to position C-3 hydroxy group [95]. Yingyue et al. [96] used Sephadex column chromatography and semi-preparative RP-HPLC to purify rutin after extraction from banana leaves. After the liquid–liquid extraction process, the authors purified the fraction by using a Sephadex column. As a result, the purity of rutin was 74–84%. Then, they further purified rutin by using semi-preparative RP-HPLC. The final rutin was 98.4% pure. The scheme of this method can be seen in Supporting Information Figure S3 (isolation process 1). Based on these results, semi-preparative RP-HPLC can enhance the purity of isolated compounds [96]. Rutin has also been isolated from S. atrata by using a combination of MPLC and HPLC coupled with online HPLC-DPPH detection [89]. The scheme is shown as isolation process 2 in Supporting Information Figure S3. Both methods 1 and 2 produce products with high purity (>95%); however, method 1 is more effective and easier to operate because it only requires HPLC instruments. Method 2 can be helpful when the compound isolated is only a DPPH inhibitor (such as rutin), so the isolation focuses on compounds with DPPH inhibitory activity.
Recycle HPLC has been developed to increase the phenolic separation efficiency of the process while keeping the peak dispersion as low as possible. This goal is achieved by installing a recycle valve on the HPLC system (at the preparative or semi-preparative scale) to return the unresolved peaks to the column (Supporting Information Figure S4). No new solvent is required during the recycling period, which is another advantage of this method [97]. Recycle HPLC is an effective method for separating compounds that elute close together, and the system’s ability to remove contaminants between chromatography cycles ensures the high purity of the isolated compounds [98]. Molo et al. [38] used C18, GS-320 columns to purify phenolic compounds from a subfraction of Chaerophyllum bulbosum extract. For the C18 column, they purified luteolin-7-O-β-D-glucopyranoside, a flavonoid compound, with methanol and water (50:50, v:v) as a mobile phase. GS-320 columns were used to purify quercetin-3-O-β- D-glucopyranoside. Luteolin-7-O-β-D-glucopyranoside and quercetin-3-O-β- D-glucopyranoside exhibited higher DPPH and ABTS•+ scavenging activities compared to BHA and α-tocopherol standards [38].
Preparative high-performance/high-pressure liquid chromatography (prep-HPLC) usually implies the use of large columns, large sample loading volumes, and high flow rates on HPLC systems to purify or separate compounds in large volumes [99]. Prep-HPLC was developed to isolate quercetin-3-O-α-L-rhamnopyranosyl-(1 → 6)-β-D-glucopyranoside a quercetin glycosides using a gradient system (15–25% acetonitrile) for 45 min. This compound showed the highest antioxidant activity in DPPH, OH radicals scavenging, and CUPRAC assay with IC50 28.8, 145.8, and 13.9 μM, respectively [100]. The DPPH activity of this compound was higher than quercetin (1.6 times) and ascorbic acid (2.0 times) [100].
Table 3. Application HPLC to isolate or purify phenolic compounds from natural products.
Table 3. Application HPLC to isolate or purify phenolic compounds from natural products.
SampleCompoundSystemStationary PhaseMobile PhaseYield (%) *Ref.
Theobroma cacao(+)-Catechin, d (−)-epicatechin, B-type dimer of flavan-3-ols, epicatechin, trimer flavan-3-olsSemi-
preparative HPLC
NMNMNM[101]
Chaerophyllum bulbosumQuercetin-3-O-β- D-glucopyranosideRecycle HPLC GS-320 column100% Methanol0.011[38]
Luteolin-7-O β-D-glucopyranosideC1850% Methanol/50% water0.007
Banana leaves (Musa balbisiana)Rutin Semi-
preparative RP-HPLC for purification
C1850% Methanol/50% water3.24[96]
Hippocrepis emerus
flowers
quercetin-3-O-α-L-rhamnopyranosyl-(1 → 6)-β-D-glucopyranosidePrep-HPLCNMgradient system (15–25% acetonitrile) for 45 min 0.517[100]
quercetin-7-O-α-L-rhamnopyranosideSemipreparative-HPLCNMgradient system (30–35% acetonitrile) fo 20 min)0.138
Smilax glabra Roxb(-)-EpicatechinPreparative HPLCWaters Sunfire Prep C18 OBDTM 250 × 19 mm, 5 μmAcetonitrile (A) and water with 0.3% formic acid (B) with gradient elution1.77[102]
Neoastilbin 11.04
Astilbin 18.10
Neoisoastilbin 4.09
Isoastilbin 5.03
Pleioblastus amarus shoots3-O-feruloylquinic acidsemi-preparative HPLCC-18Methanol-0.1% acetic
acid (60:40, v/v)
0.03[103]
Frankenia pulverulentaGallic acidPrep-HPLCC-18, 5 µmGradient system using solvent A (water:
0.1% TFA) and solvent B (ACN/0.1% TFA)
NM[104]
Catechin
Quercetin
Magnolia officinalisSyringinSemi-preparative HPLCC-18, 5 µmGradient system using solvent A (water containing 0.2% acetic acid, v/v) and B (methanol)NM[105]
Magnoloside B
Magnoloside A
Magnoloside F
Magnololisocratic system using solvent A (water containing 0.2% acetic acid, v/v) and B (methanol), with 80% B
Obvatol
Honokiol
Origanum
minutiflorum
EriodictyolSemi-preparative HPLCNMNMNM[106]
Luteolin
Rosmarinic acid
Teucrium hyrcanicum L.ActeosideSemi preparative HPLCNMGradient system using water:acetic acid (99:1) (solvent A) and acetonitrile (solvent B)NM[107]
Moringa oleifera leavesIsoquercitrinSemipreparative HPLCC-18Gradient system using
water (solvent A) and an acetonitrile and water mixture (40:60, v/v)
(solvent B)
0.02[13]
Astragalin0.002
3-O-caffeoylquinic acid0.003
Schinopsis brasiliensisgallic acid 4-O-b-D-(60-Ogalloyl)-glucopyranosideSemi-preparative HPLC NM Methanol:water (3:7)0.18[108]
2-Hydroxy-4-methoxyphenol 1-O-b-D-(60-O-galloyl)
glucopyranoside
0.16
4,9-Dihydroxypropiophenone-9-O-(60-O-galloyl)-b-Dglucopyranoside0.17
3,4-di-O-galloyl-quinic acid0.31
4-hydroxy-3-methoxyphenol-1-O-(60-O-galloyl)-
b-D-glucopyranoside
Methanol:water (1:3)0.17
4-hydroxy-2-
methoxyphenol-1-O-b-D-(60-O-galloyl) glucopyranoside
0.19
Artocarpus
elasticus
Artonin WPrep-HPLCC1830% Methanol in water0.092[109]
Artorigidinone B0.088
Cycloartobiloxanthone0.078
Berberis baluchistanicaPakistaninePreparative recycling HPLCC18Acetonitrile:water (60:40)0.34[110]
Malus prunifolia (Willd.) Borkh.SachalisideSemi-preparative HPLCXbridge® (250 mm × 4.6 mm, 10 μm,)Methanol–0.1% formic acid-water (15:85)NM [111]
Chlorogenic acidACN-0.1% formic acid-water (15:85)
EpicatechinACN-0.1% formic acid-water (8:92)
Procyanidin B2ACN-0.1% formic acid-water (14:86)
Nitraria tangutorumTyrosol 8-O-β-d’glucopyranosidePrep-HPLCXCharge 185%−55% of acetonitrile with 0.2% formic acidNM[112]
Querceitn 3-O-(2G-rhamnosylrutinoside)
Vanillic acid5%−35% of acetonitrile with 0.2% formic acid
Tithonia diversifolia (Hemsl.) A. Gray(E)-3-(((3-(3,4-dihydroxyphenyl)acryloyl)oxy)methyl)-2-methyloxyrane-2-carboxylic acid Prep-HPLC for purificationC18Phosphorous acid 0.05%:isopropyl alcohol (93:7)0.003[113]
NM, not mentioned in the article; RP-HPLC, reverse-phase high-performance liquid chromatography. * Yield (%): ratio of isolate mass with crude extract mass.

3.3. Counter Current Chromatography

CCC is a form of liquid–liquid partition chromatography that uses two immiscible liquids. One phase is maintained as the stationary phase in the absence of an adsorption matrix. The second phase passes through the stationary phase and is efficiently equilibrated by utilizing a hydrodynamic or turbulent mixture [114]. The difference compared with other chromatographic systems is that this method does not require solid support: the stationary phase is held in the column by gravity or centrifugal force [115].
CCC has also been used to isolate bioactive compounds. Several CCC methods have been developed for isolation, one of which is the use of the pH zone refinement CPC. In addition, a new approach using a DES with HSCCC has been developed in an effort to use a more environmentally friendly solvent for isolation. Here, we will discuss the CCC method with its performance in compound isolation and compare the results of different CCC systems to evaluate the best one.
CCC is divided into hydrostatic and hydrodynamic equilibrium systems (Figure 3) [116]. A hydrostatic system uses a stable force field to hold the stationary phase on the column while the mobile phase flows through the column. Helical CCC (toroidal coil CCC), droplet CCC, and CPC are hydrostatic systems. The hydrodynamic systems use the Archimedes screw effect, which promotes the constant mixing of the two phases while maintaining one of the phases as the stationary phase [116]. HSCCC and HPCCC are hydrodynamic systems.
CPC is widely used to purify natural products by partitioning a sample between two immiscible phases. Several compounds have been isolated using this method, one of which is the flavonoid in Bryophyllum pinnatum (Lam.) Oken (Crassulaceae) [40], and phenolic compounds from Anogeissus leiocarpus Guill. and Perr. (Combretaceae) [117].
As mentioned above, CPC is a form of hydrostatic CCC [118] in which a stationary liquid phase is supplied to the rotor while it rotates at a moderate rotational speed and is maintained in the rotor by the resulting centrifugal force. Then, the mobile phase, which contains the solute to be extracted is delivered under pressure to the rotor and pumped through the stationary phase [119]. The partition coefficient (Kd) indicates the separation of constituents, which is determined by the concentration of the target compound in the stationary phase divided by the concentration in the mobile phase. Therefore, components with a high affinity for the mobile phase elute early, and components with a high affinity for the stationary phase elute later [120,121,122,123]. The advantages of CPC over the droplet CCC (DCCC) were a faster movement of the mobile phase past the stationary phase than DCCC because CPC generates a centrifugal force of the rotor, a greater flow rate, and a more efficient method [124]. CPC has been applied to isolate three flavonols from the bark of Weinmannia trichosperma Cav., such as isoastilbin, neoisoastilbin, and neoastilbin using the HEMWAT system (hexane-ethyl acetate-methanol-water; ratio 1:9:1:9) [125]. These compounds exhibited potent antioxidant activity not only in DPPH and the 3-ethylbenzothiazoline-6-sulfonic acid (ABTS) radical scavenging but also in the ferric-reducing ability of the plasma (FRAP) system [102].
HSCCC uses hydrodynamic equilibrium. The instrument consists of a spindle, a planetary axis, and a couple of rotating axes that use a hydrodynamic device where separation occurs in a multilayer coil. It consists of a long piece of endless tubing wrapped in multiple layers around a holder. Several of the resulting coils can be connected in series to increase the total volume of the instrument. The coil is subjected to a centrifugal force field and rotates around its axis while at the same time rotating around the central axis of the system. This motion results in the retention of the stationary phase and partition of the analyte between the two liquid phases [61,68].
As with forms of CCC, the choice of the solvent system is critical in this method. The solvent system often consists of organic solvents, including n-hexane, n-butanol, dichloromethane, methanol, and ethyl acetate. However, these solvents are dangerous for researchers and the environment [126,127,128]. A DES for HSCCC has been developed to overcome this drawback. The combination of DES and HSCCC is recent in the development of methods to isolate bioactive compounds from natural products. DES, composed of hydrogen bond donors and acceptors, are a new generation of room-temperature liquid salts that have been widely used to extract bioactive compounds [129,130,131]. Cai et al. [41] developed an HSCCC system using DES to extract and separate flavonoids from Malus hupehensis. Based on their optimization experiment, they chose choline chloride/glucose, water, and ethyl acetate (1:1:2, v/v) for HSCCC separation. They successfully isolated three phenolic compounds with this method, namely avicularin, phloridzin, and sieboldin, each with >92% purity. Overall, the authors demonstrated that two HSCCC separations with DES are a valuable and environmentally friendly way to separate pure compounds from extracts [41]. In another study, avicularin, phloridzin, and sieboldin were identified as antioxidant compounds [132,133]. The activity of avicularin was determined using DPPH and hydroxyl (OH) scavenging assays. At a concentration of 100 mg/L, avicularin exhibited effective antioxidant activity, with OH radical scavenging rates reaching 87.54% [132]. The antioxidant capacity of sieboldin was demonstrated by its ability to prevent vasoconstriction and inhibit advanced glycation end-products (AGEs) formation [133].
HSCCC was also successful in separating and purifying the antioxidant phenolic glycoside in an extract of Castanopsis chinensis Hance. In this study, preliminary separation by multistep column chromatography was applied to obtain the phenolic glycoside fraction that contains the mixture of compounds. The HSCCC successfully separated two antioxidant phenolic compounds with higher purity, chinensin D (93.0%) and chinensin E (95.7%) [134].
HPCCC works much the same way as HSCCC. Separation using HSCCC is characterized by long separation times, typically 3–8 h. So, HPCCC instruments have been developed to enable high-resolution separations with 20–60 min of elution. It is achieved by providing columns that maintain >75% steady-state retention at semi-preparative mobile phase flow rates of ≥6 mL/min. Flow rates of 20–100 mL/min are used for preparative-scale separations, maintaining >75% stationary phase retention [135]. The studies for separating phenolic compounds with antioxidant activity using HSCCC and HPCCC are presented in Table 4.
In general, the CCC systems have advantages compared with HPLC and MPLC: they do not require a solid column, there is low solvent consumption, there is no irreversible loss of sample because chemosorption can be avoided, there is higher sample recovery, and a high load capacity [147,148,149]. However, CCC systems require time, especially to optimize the solvent system, and instability of the solvent system may occur.

3.4. Hydrophilic Interaction Liquid Chromatography

Hydrophilic Interaction Liquid Chromatography or HILIC is a type of partition chromatography that uses a polar stationary phase. The partition occurs between a non-polar or organic region in the mobile phase and a polar water-enriched layer at the surface of the polar stationary phase [150]. Several stationary phases, including silanol-derivatized phases such as amino-, amide-, cyanopropyl-, carbamate-, diol-, and polyol-, have been developed for HILIC [151]. HILIC has successfully applied for the enrichment of the compound in the extract in isolation. In isolation, HILIC is usually combined with RPLC, often called two-dimensional liquid chromatography (2D-LC). 2D-LC is widely used for separating complex samples due to its high resolution and large peak capacity [152].
Dang et al. (2018) have isolated the antioxidant phenolic compound from Dracocephalum heterophyllum using an offline two-dimensional reversed-phase/hydrophilic interaction liquid chromatography (2D RP/HILIC) technique guided by on-line HPLC-DPPH [153]. During the isolation process, the C18 preparative column was used for first-dimensional (1D) separation which resulted in six antioxidative fractions with a recovery rate of 61.4% out of the ethyl acetate fraction. For the second-dimensional (2D) separation, a HILIC XAmide preparative column was used. A total of eight antioxidants (caffeoyl-β-D glucopyranoside, ferruginoside B, verbascoside, 2′-O-acetylplantamajoside, sibiricin A, luteolin, rosmarinic acid, and methyl rosmarinate) were isolated from D. heterophyllum, with a purity of over 95% [153]. The application of separation using HILIC in the isolation process can be seen in Table 5.

3.5. Column Chromatography

Column chromatography is used to separate impurities and purify biological mixtures. It is also used to isolate active molecules and extract metabolites from various samples [157]. The solid and liquid samples can be separated and purified by this method. The stationary phase of column chromatography is placed inside a narrow tube (column), and the stationary adsorbed and separated passing compounds with the help of a liquid mobile phase. Due to their chemical nature, compounds are adsorbed, and elution is based on the differential adsorption of substances by the stationary phase [158]. The studies that used column chromatography for phenolic compound isolation can be seen in Table 6.

3.5.1. Silica Gel Chromatography

Silica gel has a silanol group, and it is a polar absorbent. Molecules are held in silica gel by hydrogen bonding and dipole-dipole interactions. For example, polar natural substances are retained longer on silica gel columns than non-polar ones [159]. Pyrogallol, rutin, and morin are phenolic compounds isolated from the ethyl acetate fraction of Bergenia ciliata by silica gel column chromatography. Solvent systems, including ethyl acetate and n-hexane were employed, with polarities ranging from 1% to 50%. Pyrogallol, rutin, and morin have been demonstrated to be effective against free radicals ABTS and DPPH. Notably, pyrogallol has exhibited the highest efficacy among them [160].
Isolation using silica gel column chromatography is still used because it boasts several advantages, including its ease of use; the stationary phase is stable and does not readily decompose [30]. However, the drawback of this method is that it is time-consuming and requires a large amount of solvent.

3.5.2. Size Exclusion Chromatography

Size-exclusion chromatography is a type of partition chromatography, used to separate molecules based on their sizes. Sephadex®LH-20 is a stationary phase in size exclusion chromatography widely utilized for isolating the bioactive compounds in natural products. Sephadex®LH-20 is a size exclusion column prepared by hydroxypropylated dextran gel [161] that has both hydrophilic and lipophilic properties, and is stable in all solvents except strong acid, and contains strong oxidizing agents [162]. Naringin belongs to the flavonoid class known as flavanones that act as antioxidant and anticancer [163]. Naringin has been successfully isolated from pomelo peel extract using Sephadex®LH-20 with higher purity (95.7 ± 0.23%). Other phenolic compounds have also been successfully separated with Sephadex®LH-20, including quercetin and (2R)-eriodictyol in mulberry fruit extract [164]. The advantage of using Sephadex®LH-20 for isolating is that it allows for separating a wide range of natural products using either aqueous or non-aqueous solvents [159]. The disadvantage of using Sephadex®LH-20 is that it requires choosing the right solvent because the particle size and exclusion limit differ depending on the solvent used for swelling.
Table 6. A list of studies that used column chromatography.
Table 6. A list of studies that used column chromatography.
SampleCompoundType of SorbentMobile Phase Yield (%) *Ref.
Pistacia integerrima gallQuercetin and pyrogallolSilica gel Mixture of ethyl acetate: n-hexane with different concentrations (1–60%)NM[37]
Rhizomes of the Bergenia ciliatapyrogallol, rutin and morinSilica gel n-hexane at the first and followed by increase in polarity of n-Hexane/ethyl acetate gradients up to 50% ethyl acetate/n-hexane (1:1) gradient NM[160]
Endopleura uchiBergeninsilica gel LC60A (70–200 μm)Chloroform/ethanol (7:3) isocratic5.4% (Leave extract),
5.73% (Twigs extract), and 6.09%
(Bark extract)
[165]
Alseodaphne semecarpifolia NeesIcariin Silica gel Gradient elution: n-hexane: ethyl acetate (100:0 → 0:100), then ethyl acetate: petroleum ether (100:0 → 0:100), then petroleum ether: chloroform (100:0 → 0:100)1.34[166]
Baicalein 1.23
Litsea glaucescensEpicatechinSilica gel 60 column (100 cm × 5 cm)Gradient elution using mixture of hexane-ethyl acetate-methanol mixturesNM [167]
Quercitrin
Kaempferol
Boesenbergia rotunda2′,
4′-dihydroxy-6-methoxychalcone
Silica gel Hexane-ethyl acetate (6:4)0.125[168]
5-hydroxy-7-methoxyflavanone0.35
5, 7-dihydroxyflavanone0.2
Apocynum venetum tea(−)-epicatechinSilica gelHexane 0.003[169]
Jatropha podagricaFraxetinSilica gelHexane and ethyl acetate at 8:2, 7:3, and 6:4 ratios,0.059[170]
Euphorbia balsamiferaQuercetin-3-O-glucopyranosideSilica gel Methanol 0.003[171]
IsoorientinSilica gel chloroform/methanol (7:3)0.004
Hibiscus rosa sinensisHibiscetin-3-glucosideSilica Gel G-60NMNM[33]
Cordia sebestena flowerHesperitinSilica gel Chloroform/methanol (60:40)NM[172]
Ipomoea
pes-caprea (Convolvulaceae) leaves
(5,7-dihydroxy-4-phenyl-2H-chromen-2-
one)
Silica gel 100% hexane followed by a gradient mixture of hexane: methanol (95:5–100).1.02[173]
Afzelia africana3,3′ -di-O-methyl ellagic acid.Silica gel Petroleum ether-ethyl acetate (85:15)0.029[174]
Zygophyllum simplex L.MyricitrinSilica gel 100 C18Methanol:water (1:9)0.105[175]
Luteolin-7- O-β-D-glucosideMethanol:water (2:8)0.084
Calendula tripterocarpa RuprQuercetin Silica gel Ethyl acetate-methanol–water (90:5:4)0.32[176]
Scopoletin0.23
Ferulago cassiaPeucedanolSilica gel Hexane:ethyl acetate (76:24)0.1[177]
UmbelliferoneHexane:ethyl acetate (63:35)0.16
Perilla frutescens (L.) Britt.Ferulic acidSilica gel Mixture of chloroform and methanol0.042[178]
LuteolinSilica gel Mixture of chloroform and methanol0.033
Apigenin Silica gel Chloroform:methanol mixture (12:1 to 4:1)0.042
Caffeic acidCombination of silica gel and Sephadex LH-20 NM 0.024
Rosmarinic acidSephadex LH-2090% of methanol 0.16
Retama monosperma (L.) Boiss.QuercetinSilica gel Mixture of Hexane, diethyl ether, and ethyl acetate with gradient
elution
NM[179]
6-methoxykaempferol
Kaempferol
Origanum rotundifoliumApigeninSilica gel Solvent system with increasing polarity from hexane to ethyl acetate and ethyl acetate- methanolNM[180]
Ferulic acid
Vitexin
Rosmarinic acid
Globoidnan A
Palmyra palm (Borassus flabellifer Linn.) syrup2,3,4-trihydroxy-5 methylacetophenoneSilica gel Mixture of dichloromethane and methanol1.82[181]
Prunus mahaleb L.Gallic acid Silica gel Mixture of chloroform and methanol with different ratio0.0067[182]
Odontites serotina (Lam.) DumActeosideSilica gelNMNM [183]
Euphorbia geniculataGallic acid Sephadex LH-2020% of methanol in waterNM [184]
Ellagic acid
Rutin40% of methanol in water
Quercetin 100% of methanol
Pomelo peelsNaringinSephadex LH-2030% of ethanolNM [185]
Desmodium caudatumDescaudatine ACombination of silica gel, Sephadex LH-20 and C-18NM NM [186]
8-Dimethylallyltaxifolin
Arisaema heterophyllum tubers6,7-dihydroxy-2-(4-hydroxyphenyl)-4 H-chromen-4-oneSephadex LH-20Methanol0.047[187]
(E)-4-(3-hydroxypropyl-1-en-1-yl)phenol0.005
Mulberry fruit
(Morus alba L.)
(2R)-eriodictyolSephadex LH-20Mixture of methanol/water, 50:50 to 70:30 (v/v)0.006[164]
Quercetin0.0004
Ombrophytum subterraneum
(Aspl.) B. Hansen (Balanophoraceae)
3′,5,5′,7-tetrahydroxyflavanone 7-O-β-D-1 → 6 diglucosideSephadex LH-20Mixture of methanol/water, (8:2)17.82[188]
Mulberry leavesRutin Sephadex LH-20Methanol:water (3:7)0.041[189]
IsoquercetinMethanol:water (1:9)0.039
Manilkara hexandra fruitGallic acidSephadex LH-20n-Butanol–Isopropyl alcohol–Water0.133[190]
Taxifolin0.066
Myricetin10% MeOH and n-Butanol–Isopropyl alcohol–Water0.133
Quercetin0.133
NM, not mentioned in the article. * Yield (%): ratio of isolate mass and crude extract mass.

3.6. Supercritical Fluid Chromatography (SFC)

Supercritical fluid chromatography (SFC) is a separation method that uses compressed gas above the critical temperature/pressure instead of organic solvent [191]. SFC is a green separation technique that uses mostly supercritical carbon dioxide as the mobile phase, with the eluting power controlled by the addition of organic solvent as an organic modifier. This method is known for its low operational cost and is considered an environmentally friendly alternative [192]. SFC is the preferred method for separating natural products on an analytical or preparative scale (Table 7). This method has been applied to separate curcumin, demethoxycurcumin, and bisdemethoxycurcumin directly from turmeric on a preparative scale. After a single step of supercritical fluid chromatography separation, 20.8 mg of curcumin (97.9% purity), 7.0 mg of demethoxycurcumin (91.1%), and 4.6 mg of bisdemethoxycurcumin (94.8%) were obtained with a mean recovery of 76.6% [193].
A two-dimensional offline SFC/RPLC system was developed to separate and prepare lignans from Fructus Arctii. SFC was utilized in the first dimension to prepare the lignin fractions, while RPLC was employed in the second dimension to produce high-purity lignin compounds. This method has advantages such as high loading ability, and short time analysis [194].
Table 7. A list of studies that used the SFC.
Table 7. A list of studies that used the SFC.
Sample Compound Mode Separation Purity Yield (%) *Ref.
Fructus ArctiiMatairesinol2D-SFC/RPLC>90%NM [194]
Arctigenin
Lappaol C
Fructus CnidiiOsthole and Semi-preparative SFC98.9%19.6[195]
Imperatorin98.2%24.4
Alpinia officinarumPinocembrinSFC/preparative SFC 99.9%NM [196]
Galangin99.5%
Kaempferide98.5%
* Yield (%): ratio of isolate mass and crude extract mass.

3.7. Molecularly Imprinted Technology

MITs provide a versatile, tailor-made technique to separate and purify specific target molecules. MIPs are synthetic polymers with excellent properties due to their low cost, ease of fabrication, high selectivity, and good reusability [197]. MIPs have three-dimensional (3D) structures. They are synthesized by copolymerizing functional monomers and crosslinkers in the presence of template molecules [198,199]. The template molecules are then removed to obtain a polyporous polymer with complementary cavities to the shape, size, and functional groups of the template molecule [200,201]. Despite many methods existing for MIP synthesis, only several methods have been developed to synthesize MIPs for separating or isolating phenolic compounds in extracts, including in situ polymerization [202,203], Pickering emulsion polymerization [204], bulk polymerization [205,206], precipitation polymerization [207,208], and the surface molecular imprinting technique (SMIT) [209,210].
Bulk polymerization is conventionally used for MIP synthesis. The template molecules, functional monomer, crosslinker, and initiator are mixed in a non-polar solvent in specific ratios. Polymerization is initiated by light or thermal irradiation. The solid polymers must be ground and sieved, and then template molecules are removed from the polymer to obtain specific cavities [211]. Precipitation polymerization is a simple and popular method to produce the MIP. This method can make high-quality monodisperse MIP microspheres without stabilizers or emulsifiers. Unlike bulk polymerization, grinding is not carried out because it could damage a specific surface area or imprinting site, thus reducing the adsorption capacity [212]. MIPs produced using this method have uniform shapes and sizes [213]. The surface molecularly imprinted technique (SMIT) is a polymerization method that occurs on the surface of solid matrixes. The binding sites are distributed on the outer layer of the surface of the solid matrixes. One of the solid matrixes used in this method is magnetic Fe3O4 [214]; it is called magnetic molecularly imprinted polymer (MMIP). In contrast to the previously mentioned techniques, SMIT has advantages such as faster binding kinetic, higher separation efficiency, and minimizing the embedded phenomenon [215].
MIPs have mostly been used in sample pre-treatment as a sorbent for solid-phase extraction (MISPE) and dispersive solid-phase extraction (MIDSPE). In MISPE, the MIP is packed in an SPE cartridge. The process includes conditioning MISPE, loading the sample, washing to clear interference, and eluting the analyte [216]. In MIDSPE, the MIP is directly applied to the liquid volume of the sample solution. The entire procedure involves shaking and centrifugation for separation. There are two types of sorbents used in MIDSPE, namely non-magnetic MIP and MMIP [217]. An external magnet can be used in the separation process when MMIP is applied as the MIDSPE sorbent [218]. Figure 4 shows the schemes for MISPE and MIDSPE. In addition, MIPs can be used for chromatographic separation (monolithic column). Monolithic column MIP can be prepared directly by in situ free-radical polymerization within the chromatographic column. A grinding, sieving, and packing column are not needed in this stage [219]. The application of MIPs for extraction, separation, or purification are listed in Table 8.
Hosny et al. [205] synthesized MIPs using bulk polymerization to isolate a phenolic compound which is sinapic acid from broccoli. This phenolic compound has tremendous antioxidant potential [220] and could be used to treat several pathologies, including diabetes [221], infection [222], inflammation [223], and cancer [224]. The authors determined the optimum MIP for isolating sinapic acid, synthesized using 1:4:20 as the molar ratio of template to the functional monomer (4-vinyl pyridine) to the crosslinker (ethylene glycol dimethacrylate [EGDMA]). 4-Vinyl pyridine acts as a hydrogen bond acceptor to form hydrogen bonding with sinapic acid and its role by π–π stacking [205]. The authors used dimethyl sulfoxide (DMSO) as a porogen because of its low polarity, which avoids hydrogen bonding with template molecules [225]. They determined the binding isotherm by incubating 15 mg of MIP with 2 mL of different concentrations of sinapic acid prepared in pure water (1 × 10−2 to 2.5 × 10−4 M) for 2 h. They calculated the amount of template molecule bound and constructed a Scatchard plot, which showed two linear regions based on the ratio of bound to free template (Q/F) against the amount of bound template (Q). This is because the MIP has heterogenous binding sites [205]. The maximum number of binding sites available for binding MIP (Qmax) was 117.51 µmol/g for the regions of high-affinity areas and 572.098 µmol/g for the low-affinity areas. They evaluated the MIP with the total extract and the ethyl acetate fraction of Botrytis italica L. (broccoli) by using HPLC. The total extract contained sinapic acid and caffeic acid, while the ethyl acetate fraction contained sinapic acid, ferulic acid, and caffeic acid (the latter two are analogues of sinapic acid). The authors also evaluated the selectivity of their synthesized MIP. After binding between the total extract with MIP, the peak area of sinapic acid decreased by 90% while the peak area of caffeic acid decreased by 65%. Furthermore, after binding between the ethyl acetate fraction and the MIP, the chromatogram showed no peak for sinapic acid, the peak area of ferulic acid had decreased by 60%, and the peak area of caffeic acid had decreased by 11%. The authors concluded that the MIP has an excellent binding affinity and good selectivity toward sinapic acid [205].
Table 8. The application of molecularly imprinted polymers (MIPs) for extraction, separation, or purification of bioactive compounds from natural products.
Table 8. The application of molecularly imprinted polymers (MIPs) for extraction, separation, or purification of bioactive compounds from natural products.
Synthesis MethodSampleCompoundSample Pre-Treatment MethodAdsorption Capacity Yield (%)Ref.
Bulk polymerization Rosmarinus officinalis LRosmarinic acidMISPE15.49 mg/g49.11 ± 4.58 mg/g[43]
Rhodiola crenulataSalidrosideMISPE28.13 mg/gNM [226]
green coffee bean extractCaffeic acid MISPE1.03 mg/g42%[227]
Chlorogenic acidNM49%
Pickering emulsion polymerizationSpina GleditsiaeQuercetin NM 0.521 mg/gNM [204]
Precipitation polymerizationCarthamus tinctorius L. and Abelmoschus manihot (Linn.)MyricetinMISPE11.80 mg/g79.82–83.91% and 81.50–84.32%,[228]
Salvia officinalis leavesRosmarinic acidUA-DSPENM77.80%[229]
Surface molecular imprinting Citrus reticulata BlancoHesperetinMIDSPE with MMIP7.316 mg/gNM [209]
Apple sample KaempferolMIDSPE with MMIP3.84 mg/g NM [230]
Spiked sample in Larix griffithianaDihydroquercetinMIDSPE with MMIP77.72 ± 3.56 mg/gNM [231]
Polygonum cuspidatum.ResveratrolMIDSPE11.56 mg/g23.74%[232]
Note: MMIP, magnetic molecularly imprinted polymer; MIDSPE, molecularly imprinted dispersive solid-phase extraction; MISPE, molecularly imprinted solid-phase extraction; UA-DSPE, ultrasonic-assisted dispersive solid phase extraction; NM, not mentioned in the article.
Wang et al. [209] synthesized MMIP for the selective separation of hesperetin from the dried pericarp of Citrus reticulata Blanco. Hesperetin is a flavanone derivate that has myriad pharmacological activities such as anti-cancer [233], anti-inflammation [234], and anti-hyperglycemia [235]. They added 40 mg of MMIP into 2.5 mL of extract solution and then shook the mixture at 30 °C for 2 h to separate hesperetin. They separated the MMIP by using an external magnetic field. They used a mixture of methanol and acetic acid (9:1) to wash the MMIP for 2 h. The eluent contained hesperetin was evaporated and dissolved. The solution was then analyzed by HPLC. They found that the MMIP effectively and selectively separates hesperetin from the dried pericarp extract of C. reticulata and highlighted that it could be used for the rapid enrichment and isolation of hesperetin from other natural plants [209].
Several special strategies have been developed to improve the analytical performance of MIPs and to increase the efficiency of the separation process. A dummy template is used to avoid template leakage when using a target analyte as a template molecule. Dummy templates have an analogous structure to the target analyte molecules, with similar shape, size, and functional group, but they do not interfere with analytical determinations [236,237]. Eidi et al. [86] developed a dummy template molecularly imprinted polymer (DMIP) for the selective isolation of sesquiterpene coumarins from asafoetida. They used 7-hydroxycomarin (7-HC), the parent compound of sesquiterpene coumarins, as a dummy template. Bulk polymerization was used to synthesize the DMIP. The optimized polymer was synthesized using methacrylic acid (MMA) as a functional monomer with a 1:6 molar ratio of dummy template to functional monomer. The DMIP was used as a sorbent for SPE (DMISPE). After extraction using DMISPE, the peak area in the chromatogram of asafoetida extract was on average three times more than that without DMISPE. These results show that DMISPE was efficient for selective extraction and clean-up of the sesquiterpene coumarin from the asafoetida extract [206].
Overall, MITs have advantages for separating bioactive compounds: a rapid and easy procedure, selective isolation, fewer impurities, and recoveries of isolates. In contrast, this method’s drawbacks are that the mass of the isolate depends on the binding capacity of the MIP sorbent and requires the additional isolation method when using MIP multi-templates.

3.8. High-Performance Thin Layer Chromatography

HPTLC is an enhanced version of TLC (Thin Layer Chromatography). HPTLC employs various techniques to achieve better separation and analysis of compounds. For instance, HPTLC uses TLC plates with finer particle sizes in the stationary phase, resulting in better resolution [238]. The HPTLC cannot only be applied for separating or isolating the active compounds [238,239], but can also applied for the estimation of the isolated compound [240]. Jug et al. (2021) used the offline multidimensional HPTLC for fractionation and isolation of flavan-3-ols, proanthocyanidins, and anthraquinones derivates from Japanese knotweed rhizome bark extract. A combination of stationary plates was used in this study. Preparative TLC silica gel was used in first-dimension fractionation, HPTLC cellulose plate and HPTLC silica gel were used in second-dimension fractionation, and the HPTLC silica gel was used in third-dimension fractionation. In the isolation process, post-chromatographic derivatization was used to identify the analyte target using 4-dimethylaminocinnamaldehyde (DMACA) to distinguish the flavan-3-ols and proanthocyanidins derivates from other compounds. HPTLC-MS was also used to characterize the isolated compounds. In this study, HPTLC methodology was successfully isolated (+)-catechin, (-)-epicatechin, (-)-epicatechin- gallate, procyanidin B1, procyanidin B2, procyanidin B3, proanthocyanidin B dimer gallate, emodin, emodin-8-O-glucoside, and emodin-8-O-malonyl-glucoside [239]. Fractionation or isolation using HPTLC has advantages such as a short separation time, pre- and post-chromatographic in-situ analyte derivatization, and crude extract can be applied without sample preparation [238,239]. The disadvantages of HPTLC are that it is relatively more expensive than TLC, and separation occurs only up to a certain length due to the limited plate length.
To summarize the discussion on various separation methods used in the isolation of active compounds in natural products, Figure 5 provides the advantages and disadvantages of each method.

4. Conclusions

Developing isolation methods for phenolic compounds with antioxidant activity is still exciting. Several methods that have been developed, such as MPLC, HPLC, HILIC, 2D-LC, CCC, column chromatography, SFC, MITs and HPTLC, have shown promising results with high compound purity >90%. Based on the advantages and disadvantages of different methods, it has been found that HPTLC, CCC, and MITs use less solvent than other methods. In terms of time, HPTLC and MITs are faster than other methods. Despite this, chromatography columns are still commonly used for the separation of phenolic compounds as they are cost-effective, easy to use, and do not require additional instruments. HPLC and MPLC can be combined with online DPPH detection to isolate phenolic compounds that have antioxidant activity. This combination helps to simplify the identification and isolation process of target compounds. Moreover, the CCC instrument can be guided by online DPPH HPLC to help isolate the antioxidant compound. Separation and purification using MITs may be further developed because this method is straightforward and selective. MIPs can be used directly on sample extracts; therefore, the fractionation process can be eliminated.
Further investigation still needs to be conducted for a more efficient method with high purity and yield. There are several areas of potential research foci: compare several isolation methods for bioactive compounds in the same sample to determine the best and most effective method and studies related to the cost of isolation from each method to decide which method has the lowest cost with the highest % yield and %purity. In addition, it is necessary to develop a one-step procedure to streamline the isolation processing time. Finally, to promote the sustainable growth of green chemistry, it is required to modify the extraction and isolation solvents using less toxic, less hazardous, and environmentally friendly materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants13070965/s1, Figure S1: The mechanism linking inflammation to cancer; Figure S2: The Schematic of MPLC; Figure S3: Two schemes of isolation process of rutin. DPPH, 1,1-diphenyl-2-picrylhydrazyl; MPLC, medium-pressure liquid chromatography; RP-HPLC, reverse-phase high-performance liquid chromatography; Figure S4: Instrumentation of closed loop recycle-HPLC.

Author Contributions

Conceptualisation, A.N.H. and Y.R.; methodology, A.N.H., R.P. and I.S.; writing—original draft preparation, I.S.; writing—review and editing, R.P., A.N.H. and Y.R.; visualisation, I.S.; funding acquisition, A.N.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Research and Innovation Agency (BRIN) and Educational Fund Management Institution (LPDP) trough Riset Indonesia Maju (RIM) grants 2023 number 5994/UN6.3.1/PT.00/2023.

Data Availability Statement

Data sharing is not applicable, no new data were created in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

2D-LC2-dimensional liquid chromatography
2D RP/HILIC2-dimensional reversed-phase/hydrophilic interaction liquid chromatography
ABTS3-ethylbenzothiazoline-6-sulfonic acid
ACNAcetonitrile
DESDeep eutectic solvent
DPPH1,1-diphenyl-2-picrylhydrazyl
EGDMAEthylene glycol dimethacrylate
FRAPFerric reducing antioxidant power
TACTotal antioxidant capacity
HPLChigh-performance liquid chromatography
MPLCMedium-pressure liquid chromatography
CCCCounter-current chromatography
CPCCentrifugal partition chromatography
CUPRACCupric reducing antioxidant capacity
HILICHydrophilic interaction chromatography
HSCCCHigh-speed counter-current chromatography
HPCCCHigh-performance counter-current chromatography
HPTLCHigh-performance thin layer chromatography
IC50Half-maximal inhibitory concentration
MITsMolecularly imprinted polymer techniques
MIPMolecularly imprinted polymer
MIDSPEMolecularly imprinted-dispersive solid-phase extraction
MISPEMolecularly imprinted-solid phase extraction
MMIPMagnetic Molecularly imprinted polymer
MMAMethacrylic acid
Prep-HPLCPreparative high-performance/high-pressure liquid chromatography
RNSReactive nitrogen species
ROSReactive oxygen species
RP-HPLCReverse-phase high-performance liquid chromatography
SFCSupercritical fluid chromatography
SMITSurface molecularly imprinted technique

References

  1. Ahmad Khan, M.S.; Chattopadhyay, D.; Ahmad, I. New Look to Phytomedicine: Advancements in Herbal Products as Novel Drug Leads; Elsevier B.V.: Amsterdam, The Netherlands, 2019. [Google Scholar] [CrossRef]
  2. Aldossary, K.M. Use of Herbal Medicines in Saudi Arabia; A Systematic Review. Int. Res. J. Pharm. 2019, 10, 9–14. [Google Scholar] [CrossRef]
  3. Xiong, Y.; Liu, C.; Li, M.; Qin, X.; Guo, J.; Wei, W.; Yao, G.; Qian, Y.; Ye, L.; Liu, H.; et al. The Use of Chinese Herbal Medicines throughout the Pregnancy Life Course and Their Safety Profiles: A Population-Based Cohort Study. Am. J. Obstet. Gynecol. MFM 2023, 5, 100907. [Google Scholar] [CrossRef] [PubMed]
  4. Li, Y.; Kong, D.; Fu, Y.; Sussman, M.R.; Wu, H. The Effect of Developmental and Environmental Factors on Secondary Metabolites in Medicinal Plants. Plant Physiol. Biochem. 2020, 148, 80–89. [Google Scholar] [CrossRef] [PubMed]
  5. Leicach, S.R.; Chludil, H.D. Plant Secondary Metabolites: Structure-Activity Relationships in Human Health Prevention and Treatment of Common Diseases, 1st ed.; Elsevier B.V.: Amsterdam, The Netherlands, 2014; Volume 42. [Google Scholar]
  6. Nurzyńska-Wierdak, R. Phenolic Compounds from New Natural Sources—Plant Genotype and Ontogenetic Variation. Molecules 2023, 28, 1731. [Google Scholar] [CrossRef] [PubMed]
  7. Schmidt, T.J.; Khalid, S.A.; Romanha, A.J.; Alves, T.M.A.; Biavatti, M.W.; Brun, R.; Da Costa, F.B.; de Castro, S.L.; Ferreira, V.F.; de Lacerda, M.V.G.; et al. The Potential of Secondary Metabolites from Plants as Drugs or Leads Against Protozoan Neglected Diseases—Part II. Curr. Med. Chem. 2012, 19, 2176–2228. [Google Scholar] [CrossRef] [PubMed]
  8. Kabir, F.; Sultana, S.; Kurnianta, H. Antimicrobial Activities of Grape (Vitis vinifera L.) Pomace Polyphenols as a Source of Naturally Occurring Bioactive Components. African J. Biotechnol. 2015, 14, 2157–2161. [Google Scholar] [CrossRef]
  9. Nazeam, J.A.; AL-Shareef, W.A.; Helmy, M.W.; El-Haddad, A.E. Bioassay-Guided Isolation of Potential Bioactive Constituents from Pomegranate Agrifood by-Product. Food Chem. 2020, 326, 126993. [Google Scholar] [CrossRef] [PubMed]
  10. Jiao, X.; Wang, Y.; Lin, Y.; Lang, Y.; Li, E.; Zhang, X.; Zhang, Q.; Feng, Y.; Meng, X.; Li, B. Blueberry Polyphenols Extract as a Potential Prebiotic with Anti-Obesity Effects on C57BL/6 J Mice by Modulating the Gut Microbiota. J. Nutr. Biochem. 2019, 64, 88–100. [Google Scholar] [CrossRef] [PubMed]
  11. Matulja, D.; Vranješević, F.; Markovic, M.K.; Pavelić, S.K.; Marković, D. Anticancer Activities of Marine-Derived Phenolic Compounds and Their Derivatives. Molecules 2022, 27, 1449. [Google Scholar] [CrossRef]
  12. Praparatana, R.; Maliyam, P.; Barrows, L.R.; Puttarak, P. Flavonoids and Phenols, the Potential Anti-Diabetic Compounds from Bauhinia strychnifolia Craib. Stem. Molecules 2022, 27, 2393. [Google Scholar] [CrossRef]
  13. Oldoni, T.L.C.; Merlin, N.; Bicas, T.C.; Prasniewski, A.; Carpes, S.T.; Ascari, J.; de Alencar, S.M.; Massarioli, A.P.; Bagatini, M.D.; Morales, R.; et al. Antihyperglycemic Activity of Crude Extract and Isolation of Phenolic Compounds with Antioxidant Activity from Moringa oleifera Lam. Leaves Grown in Southern Brazil. Food Res. Int. 2020, 141, 110082. [Google Scholar] [CrossRef] [PubMed]
  14. Kasmi, S.; Hamdi, A.; Atmani-Kilani, D.; Debbache-Benaida, N.; Jaramillo-Carmona, S.; Rodríguez-Arcos, R.; Jiménez-Araujo, A.; Ayouni, K.; Atmani, D.; Guillén-Bejarano, R. Characterization of Phenolic Compounds Isolated from the Fraxinus angustifolia Plant and Several Associated Bioactivities. J. Herb. Med. 2021, 29, 100485. [Google Scholar] [CrossRef]
  15. Hwang, I.W.; Chung, S.K. Isolation and Identification of Myricitrin, an Antioxidant Flavonoid, from Daebong Persimmon Peel. Prev. Nutr. Food Sci. 2018, 23, 341–346. [Google Scholar] [CrossRef] [PubMed]
  16. Ruiz-Ruiz, J.C.; Matus-Basto, A.J.; Acereto-Escoffié, P.; Segura-Campos, M.R. Antioxidant and Anti-Inflammatory Activities of Phenolic Compounds Isolated from Melipona beecheii Honey. Food Agric. Immunol. 2017, 28, 1424–1437. [Google Scholar] [CrossRef]
  17. Tessema, F.B.; Gonfa, Y.H.; Asfaw, T.B.; Tadesse, M.G.; Bachheti, R.K. Antioxidant Activity of Flavonoids and Phenolic Acids from Dodonaea angustifolia Flower: HPLC Profile and PASS Prediction. J. Chem. 2023, 2023, 1–11. [Google Scholar] [CrossRef]
  18. Mititelu, R.R.; Pădureanu, R.; Băcănoiu, M.; Pădureanu, V.; Docea, A.O.; Calina, D.; Barbulescu, A.L.; Buga, A.M. Inflammatory and Oxidative Stress Markers-Mirror Tools in Rheumatoid Arthritis. Biomedicines 2020, 8, 125. [Google Scholar] [CrossRef] [PubMed]
  19. Rahaman, M.M.; Hossain, R.; Herrera-Bravo, J.; Islam, M.T.; Atolani, O.; Adeyemi, O.S.; Owolodun, O.A.; Kambizi, L.; Daştan, S.D.; Calina, D.; et al. Natural Antioxidants from Some Fruits, Seeds, Foods, Natural Products, and Associated Health Benefits: An Update. Food Sci. Nutr. 2023, 11, 1657–1670. [Google Scholar] [CrossRef] [PubMed]
  20. Rungratanawanich, W.; Memo, M.; Uberti, D. Redox Homeostasis and Natural Dietary Compounds: Focusing on Antioxidants of Rice (Oryza sativa L.). Nutrients 2018, 10, 1605. [Google Scholar] [CrossRef] [PubMed]
  21. Shi, L.; Zhao, W.; Yang, Z.; Subbiah, V.; Suleria, H.A.R. Extraction and Characterization of Phenolic Compounds and Their Potential Antioxidant Activities. Environ. Sci. Pollut. Res. 2022, 29, 81112–81129. [Google Scholar] [CrossRef]
  22. Fang, Y.Z.; Yang, S.; Wu, G. Free Radicals, Antioxidants, and Nutrition. Nutrition 2002, 18, 872–879. [Google Scholar] [CrossRef] [PubMed]
  23. Li, S.; Tan, H.Y.; Wang, N.; Zhang, Z.J.; Lao, L.; Wong, C.W.; Feng, Y. The Role of Oxidative Stress and Antioxidants in Liver Diseases. Int. J. Mol. Sci. 2015, 16, 26087–26124. [Google Scholar] [CrossRef] [PubMed]
  24. Peng, C.; Wang, X.; Chen, J.; Jiao, R.; Wang, L.; Li, Y.M.; Zuo, Y.; Liu, Y.; Lei, L.; Ma, K.Y.; et al. Biology of Ageing and Role of Dietary Antioxidants. BioMed Res. Int. 2014, 2014, 831841. [Google Scholar] [CrossRef] [PubMed]
  25. Neganova, M.; Liu, J.; Aleksandrova, Y.; Klochkov, S.; Fan, R. Therapeutic Influence on Important Targets Associated with Chronic Inflammation and Oxidative Stress in Cancer Treatment. Cancers 2021, 13, 6062. [Google Scholar] [CrossRef] [PubMed]
  26. Kölker, S.; Ahlemeyer, B.; Krieglstein, J.; Hoffmann, G.F. Contribution of Reactive Oxygen Species to 3-Hydroxyglutarate Neurotoxicity in Primary Neuronal Cultures from Chick Embryo Telencephalons. Pediatr. Res. 2001, 50, 76–82. [Google Scholar] [CrossRef] [PubMed]
  27. Ryden, L.; Grant, P.J.; Anker, S.D.; Berne, C.; Cosentino, F.; Danchin, N.; Deaton, C.; Escaned, J.; Hammes, H.P.; Huikuri, H.; et al. ESC Guidelines on Diabetes, Pre-Diabetes, and Cardiovascular Diseases Developed in Collaboration with the EASD—Summary the Task Force on Diabetes, Pre-Diabetes, and Cardiovascular Diseases of the European Society of Cardiology (ESC) and Developed in Collaboration with the European Association for the Study of Diabetes (EASD). Diabetes Vasc. Dis. Res. 2014, 11, 133–173. [Google Scholar] [CrossRef]
  28. Poelstra, K.; Schuppan, D. Targeted Therapy of Liver Fibrosis/Cirrhosis and Its Complications. J. Hepatol. 2011, 55, 726–728. [Google Scholar] [CrossRef] [PubMed]
  29. Spoto, B.; Pisano, A.; Zoccali, C. Insulin Resistance in Chronic Kidney Disease: A Systematic Review. Am. J. Physiol. 2016, 311, F1087–F1108. [Google Scholar] [CrossRef] [PubMed]
  30. Sochorova, L.; Prusova, B.; Cebova, M.; Jurikova, T.; Mlcek, J.; Adamkova, A.; Nedomova, S.; Baron, M.; Sochor, J. Health Effects of Grape Seed and Skin Extracts and Their Influence on Biochemical Markers. Molecules 2020, 25, 5311. [Google Scholar] [CrossRef] [PubMed]
  31. Dharshini, L.C.P.; Rasmi, R.R.; Kathirvelan, C.; Kumar, K.M.; Saradhadevi, K.M.; Sakthivel, K.M. Regulatory Components of Oxidative Stress and Inflammation and Their Complex Interplay in Carcinogenesis. Appl. Biochem. Biotechnol. 2023, 195, 2893–2916. [Google Scholar] [CrossRef]
  32. Xu, D.P.; Li, Y.; Meng, X.; Zhou, T.; Zhou, Y.; Zheng, J.; Zhang, J.J.; Li, H. Bin. Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources. Int. J. Mol. Sci. 2017, 18, 96. [Google Scholar] [CrossRef]
  33. Rengarajan, S.; Melanathuru, V.; Govindasamy, C.; Chinnadurai, V.; Elsadek, M.F. Antioxidant Activity of Flavonoid Compounds Isolated from the Petals of Hibiscus rosa Sinensis. J. King Saud Univ.-Sci. 2020, 32, 2236–2242. [Google Scholar] [CrossRef]
  34. Ajileye, O.O.; Obuotor, E.M.; Akinkunmi, E.O.; Aderogba, M.A. Isolation and Characterization of Antioxidant and Antimicrobial Compounds from Anacardium occidentale L. (Anacardiaceae) Leaf Extract. J. King Saud Univ.-Sci. 2015, 27, 244–252. [Google Scholar] [CrossRef]
  35. Atanasov, A.G.; Zotchev, S.B.; Dirsch, V.M.; Orhan, I.E.; Banach, M.; Rollinger, J.M.; Barreca, D.; Weckwerth, W.; Bauer, R.; Bayer, E.A.; et al. Natural Products in Drug Discovery: Advances and Opportunities. Nat. Rev. Drug Discov. 2021, 20, 200–216. [Google Scholar] [CrossRef] [PubMed]
  36. Huang, M.; Lu, J.J.; Ding, J. Natural Products in Cancer Therapy: Past, Present and Future. Nat. Prod. Bioprospecting 2021, 11, 5–13. [Google Scholar] [CrossRef] [PubMed]
  37. Zahoor, M.; Zafar, R.; Rahman, N.U. Isolation and Identification of Phenolic Antioxidants from Pistacia Integerrima Gall and Their Anticholine Esterase Activities. Heliyon 2018, 4, e01007. [Google Scholar] [CrossRef] [PubMed]
  38. Molo, Z.; Tel-Çayan, G.; Deveci, E.; Öztürk, M.; Duru, M.E. Insight into Isolation and Characterization of Compounds of Chaerophyllum bulbosum Aerial Part with Antioxidant, Anticholinesterase, Anti-Urease, Anti-Tyrosinase, and Anti-Diabetic Activities. Food Biosci. 2021, 42, 101201. [Google Scholar] [CrossRef]
  39. Dang, J.; Ma, J.; Du, Y.; Dawa, Y.Z.; Wang, Q.; Chen, C.; Wang, Q.; Tao, Y.; Ji, T. Large-Scale Preparative Isolation of Bergenin Standard Substance from Saxifraga Atrata Using Polyamide Coupled with MCI GEL® CHP20P as Stationary Phases in Medium Pressure Chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2021, 1170, 122617. [Google Scholar] [CrossRef] [PubMed]
  40. Morais Fernandes, J.; Ortiz, S.; Padilha, M.; Tavares, R.; Mandova, T.; Rodrigues, D.; Araújo, E.; Anderson, A.W.; Michel, S.; Grougnet, R.; et al. Bryophyllum pinnatum Markers: CPC Isolation, Simultaneous Quantification by a Validated UPLC-DAD Method and Biological Evaluations. J. Pharm. Biomed. Anal. 2021, 193, 113682. [Google Scholar] [CrossRef] [PubMed]
  41. Cai, X.; Xiao, M.; Zou, X.; Tang, J.; Huang, B.; Xue, H. Extraction and Separation of Flavonoids from Malus Hupehensis Using High-Speed Countercurrent Chromatography Based on Deep Eutectic Solvent. J. Chromatogr. A 2021, 1641, 461998. [Google Scholar] [CrossRef]
  42. Xu, Y.; Lv, X.; Yang, G.; Zhan, J.; Li, M.; Long, T.; Ho, C.T.; Li, S. Simultaneous Separation of Six Pure Polymethoxyflavones from Sweet Orange Peel Extract by High Performance Counter Current Chromatography. Food Chem. 2019, 292, 160–165. [Google Scholar] [CrossRef]
  43. Saad, E.M.; El Gohary, N.A.; Abdel-Halim, M.; Handoussa, H.; Mohamed El Nashar, R.; Mizaikoff, B. Molecularly Imprinted Polymers for Selective Extraction of Rosmarinic Acid from Rosmarinus officinalis L. Food Chem. 2021, 335, 127644. [Google Scholar] [CrossRef] [PubMed]
  44. Alara, O.R.; Abdurahman, N.H.; Ukaegbu, C.I. Extraction of Phenolic Compounds: A Review. Curr. Res. Food Sci. 2021, 4, 200–214. [Google Scholar] [CrossRef]
  45. Pizani, R.S.; Viganó, J.; de Souza Mesquita, L.M.; Contieri, L.S.; Sanches, V.L.; Chaves, J.O.; Souza, M.C.; da Silva, L.C.; Rostagno, M.A. Beyond Aroma: A Review on Advanced Extraction Processes from Rosemary (Rosmarinus officinalis) and Sage (Salvia officinalis) to Produce Phenolic Acids and Diterpenes. Trends Food Sci. Technol. 2022, 127, 245–262. [Google Scholar] [CrossRef]
  46. Lu, X.; Gu, X.; Shi, Y. A Review on Lignin Antioxidants: Their Sources, Isolations, Antioxidant Activities and Various Applications. Int. J. Biol. Macromol. 2022, 210, 716–741. [Google Scholar] [CrossRef] [PubMed]
  47. Socas-Rodríguez, B.; Torres-Cornejo, M.V.; Álvarez-Rivera, G.; Mendiola, J.A. Deep Eutectic Solvents for the Extraction of Bioactive Compounds from Natural Sources and Agricultural By-Products. Appl. Sci. 2021, 11, 4897. [Google Scholar] [CrossRef]
  48. Tsao, R. Chemistry and Biochemistry of Dietary Polyphenols. Nutrients 2010, 2, 1231–1246. [Google Scholar] [CrossRef] [PubMed]
  49. Al Mamari, H.H. Phenolic Compounds: Classification, Chemistry, and Updated Techniques of Analysis and Synthesis; IntechOpen Limited: London, UK, 2022. [Google Scholar] [CrossRef]
  50. Umaru, I.J. Pyrogallol Isolation, Characterization, Cytotoxicity, Antioxidant and Bioactive Potentials on Selected Bacterial and Fungi. Int. J. Pharm. Biomed. Res. 2020, 7, 1–11. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Cai, P.; Cheng, G.; Zhang, Y. A Brief Review of Phenolic Compounds Identified from Plants: Their Extraction, Analysis, and Biological Activity. Nat. Prod. Commun. 2022, 17, 1934578X211069721. [Google Scholar] [CrossRef]
  52. Antika, L.D.; Tasfiyati, A.N.; Hikmat, H.; Septama, A.W. Scopoletin: A Review of Its Source, Biosynthesis, Methods of Extraction, and Pharmacological Activities. Zeitschrift fur Naturforschung-Sect. C J. Biosci. 2022, 77, 303–316. [Google Scholar] [CrossRef]
  53. Grivicich, I.; Mello, R.; Ledel, A.; Silva, T.; Dieter, D.; Conter, F.; Grossi, M.; Ferraz, A. In Vitro Cytotoxicity of Scopoletin Derived from Eupatorium laevigatum Lam. Br. J. Pharm. Res. 2016, 13, 1–7. [Google Scholar] [CrossRef]
  54. Jamuna, S.; Karthika, K.; Paulsamy, S.; Thenmozhi, K.; Kathiravan, S.; Venkatesh, R. Confertin and Scopoletin from Leaf and Root Extracts of Hypochaeris Radicata Have Anti-Inflammatory and Antioxidant Activities. Ind. Crops Prod. 2015, 70, 221–230. [Google Scholar] [CrossRef]
  55. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An Overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [PubMed]
  56. Dabeek, W.M.; Marra, M.V. Dietary Quercetin and Kaempferol: Bioavailability and Potential Cardiovascular-Related Bioactivity in Humans. Nutrients 2019, 11, 2288. [Google Scholar] [CrossRef] [PubMed]
  57. Das, A.K.; Islam, M.N.; Faruk, M.O.; Ashaduzzaman, M.; Dungani, R. Review on Tannins: Extraction Processes, Applications and Possibilities. S. Afr. J. Bot. 2020, 135, 58–70. [Google Scholar] [CrossRef]
  58. Smeriglio, A.; Barreca, D.; Bellocco, E.; Trombetta, D. Proanthocyanidins and Hydrolysable Tannins: Occurrence, Dietary Intake and Pharmacological Effects. Br. J. Pharmacol. 2017, 174, 1244–1262. [Google Scholar] [CrossRef] [PubMed]
  59. Gabaston, J.; Leborgne, C.; Valls, J.; Renouf, E.; Richard, T.; Waffo-Teguo, P.; Mérillon, J.M. Subcritical Water Extraction of Stilbenes from Grapevine By-Products: A New Green Chemistry Approach. Ind. Crops Prod. 2018, 126, 272–279. [Google Scholar] [CrossRef]
  60. Latva-Mäenpää, H.; Wufu, R.; Mulat, D.; Sarjala, T.; Saranpää, P.; Wähälä, K. Stability and Photoisomerization of Stilbenes Isolated from the Bark of Norway Spruce Roots. Molecules 2021, 26, 1036. [Google Scholar] [CrossRef] [PubMed]
  61. Xia, N.; Daiber, A.; Förstermann, U.; Li, H. Antioxidant Effects of Resveratrol in the Cardiovascular System. Br. J. Pharmacol. 2017, 174, 1633–1646. [Google Scholar] [CrossRef]
  62. Soural, I.; Vrchotová, N.; Tříska, J.; Balík, J.; Horník, Š.; Cuřínová, P.; Sýkora, J. Various Extraction Methods for Obtaining Stilbenes from Grape Cane of Vitis vinifera L. Molecules 2015, 20, 6093–6112. [Google Scholar] [CrossRef]
  63. Shahidi, F.; Ho, C.-T. Phenolics in Food and Natural Health Products: An Overview; ACS Publications: Washington, DC, USA, 2005. [Google Scholar] [CrossRef]
  64. Cui, Q.; Du, R.; Liu, M.; Rong, L. Lignans and Their Derivatives from Plants as Antivirals. Molecules 2020, 25, 183. [Google Scholar] [CrossRef]
  65. Gülçin, I.; Elias, R.; Gepdiremen, A.; Boyer, L. Antioxidant Activity of Lignans from Fringe Tree (Chionanthus virginicus L.). Eur. Food Res. Technol. 2006, 223, 759–767. [Google Scholar] [CrossRef]
  66. Liu, Y.; Zhang, Y.; Muema, F.W.; Kimutai, F.; Chen, G.; Guo, M. Phenolic Compounds from Carissa spinarum Are Characterized by Their Antioxidant, Anti-Inflammatory and Hepatoprotective Activities. Antioxidants 2021, 10, 652. [Google Scholar] [CrossRef] [PubMed]
  67. Sriwiriyajan, S.; Sukpondma, Y.; Srisawat, T.; Madla, S.; Graidist, P. (−)-Kusunokinin and Piperloguminine from Piper nigrum: An Alternative Option to Treat Breast Cancer. Biomed. Pharmacother. 2017, 92, 732–743. [Google Scholar] [CrossRef] [PubMed]
  68. Arora, S.; Itankar, P. Extraction, Isolation and Identification of Flavonoid from Chenopodium album Aerial Parts. J. Tradit. Complement. Med. 2018, 8, 476–482. [Google Scholar] [CrossRef] [PubMed]
  69. El Houda Lezoul, N.; Belkadi, M.; Habibi, F.; Guillén, F. Extraction Processes with Several Solvents on Total Bioactive Compounds in Different Organs of Three Medicinal Plants. Molecules 2020, 25, 4672. [Google Scholar] [CrossRef] [PubMed]
  70. Hao, J.; Wang, Z.; Jia, Y.; Sun, L.; Fu, Z.; Zhao, M.; Li, Y.; Yuan, N.; Cong, B.; Zhao, L.; et al. Optimization of Ultrasonic-Assisted Extraction of Flavonoids from Lactuca indica L. Cv. Mengzao and Their Antioxidant Properties. Front. Nutr. 2023, 10, 1065662. [Google Scholar] [CrossRef] [PubMed]
  71. Kim, H.S.; Ko, M.J.; Park, C.H.; Chung, M.S. Application of Pulsed Electric Field as a Pre-Treatment for Subcritical Water Extraction of Quercetin from Onion Skin. Foods 2022, 11, 1069. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, R.; Chu, X.; Su, J.; Fu, X.; Kan, Q.; Wang, X.; Zhang, X. Enzyme-Assisted Ultrasonic Extraction of Total Flavonoids from Acanthopanax senticosus and Their Enrichment and Antioxidant Properties. Processes 2021, 9, 1708. [Google Scholar] [CrossRef]
  73. Choommongkol, V.; Punturee, K.; Klumphu, P.; Rattanaburi, P.; Meepowpan, P.; Suttiarporn, P. Microwave-Assisted Extraction of Anticancer Flavonoid, 2′,4′-Dihydroxy-6′-Methoxy-3′,5′-Dimethyl Chalcone (DMC), Rich Extract from Syzygium nervosum Fruits. Molecules 2022, 27, 1397. [Google Scholar] [CrossRef]
  74. Yang, J.; Li, N.; Wang, C.; Chang, T.; Jiang, H. Ultrasound-Homogenization-Assisted Extraction of Polyphenols from Coconut Mesocarp: Optimization Study. Ultrason. Sonochem. 2021, 78, 105739. [Google Scholar] [CrossRef]
  75. Radovanović, K.; Gavarić, N.; Švarc-Gajić, J.; Brezo-Borjan, T.; Zlatković, B.; Lončar, B.; Aćimović, M. Subcritical Water Extraction as an Effective Technique for the Isolation of Phenolic Compounds of Achillea Species. Processes 2023, 11, 86. [Google Scholar] [CrossRef]
  76. Hostettmann, K. and Terreaux, C. Medium-Pressure Liquid. In Encyclopedia of Separation Science; Poole, C.F., Cooke, M., Eds.; Academic Press: Cambridge, MA, USA, 2000; pp. 3296–3303. [Google Scholar]
  77. Ma, Z.; Wei, X.; Zhou, M.; Liu, G.; Liu, F.; Liu, Z.; Yu, X.; Zong, Z. Isolation and Purification of Carbazole Contained in Anthracene Slag by Extraction Combined with Medium Pressure Liquid Chromatography. Chin. J. Chem. Eng. 2019, 27, 2925–2929. [Google Scholar] [CrossRef]
  78. Cheng, Y.; Liang, Q.; Hu, P.; Wang, Y.; Jun, F.W.; Luo, G. Combination of Normal-Phase Medium-Pressure Liquid Chromatography and High-Performance Counter-Current Chromatography for Preparation of Ginsenoside-Ro from Panax Ginseng with High Recovery and Efficiency. Sep. Purif. Technol. 2010, 73, 397–402. [Google Scholar] [CrossRef]
  79. Rahman, M. Computational Phytochemistry: Chapter 4- Application of Computational Methods in Isolation of Plant Secondary Metabolites; Elsevier B.V.: Amsterdam, The Netherlands, 2018. [Google Scholar] [CrossRef]
  80. Hwu, J.R.; Robl, J.A.; Khoudary, K.P. Instrumentation and Separation Results of Medium Pressure Liquid Chromatography. J. Chromatogr. Sci. 1987, 25, 501–505. [Google Scholar] [CrossRef] [PubMed]
  81. Dang, J.; Wang, Q.; Wang, Q.; Yuan, C.; Li, G.; Ji, T. Preparative Isolation of Antioxidative Gallic Acid Derivatives from Saxifraga tangutica Using a Class Separation Method Based on Medium-Pressure Liquid Chromatography and Reversed-Phase Liquid Chromatography. J. Sep. Sci. 2021, 44, 3734–3746. [Google Scholar] [CrossRef] [PubMed]
  82. Ebada, S.S.; Edrada, R.A.; Lin, W.; Proksch, P. Methods for Isolation, Purification and Structural Elucidation of Bioactive Secondary Metabolites from Marine Invertebrates. Nat. Protoc. 2008, 3, 1820–1831. [Google Scholar] [CrossRef] [PubMed]
  83. Mottaghipisheh, J.; Iriti, M. Sephadex® LH-20, Isolation, and Purification of Flavonoids from Plant Species: A Comprehensive Review. Molecules 2020, 25, 4146. [Google Scholar] [CrossRef] [PubMed]
  84. Shu, X.; Wang, M.; Liu, D.; Wang, D.; Lin, X.; Liu, J.; Wang, X.; Huang, L. Preparative Separation of Polyphenols from Artichoke by Polyamide Column Chromatography and High-Speed Counter-Current Chromatography. Quim. Nova 2013, 36, 836–839. [Google Scholar] [CrossRef]
  85. Salimo, Z.M.; Yakubu, M.N.; da Silva, E.L.; de Almeida, A.C.G.; Chaves, Y.O.; Costa, E.V.; da Silva, F.M.A.; Tavares, J.F.; Monteiro, W.M.; de Melo, G.C.; et al. Chemistry and Pharmacology of Bergenin or Its Derivatives: A Promising Molecule. Biomolecules 2023, 13, 403. [Google Scholar] [CrossRef]
  86. Agber, C.T.; Tor-Anyii, T.A.; Igoli, J.O.; Anyam, J.V. Isolation and Characterisation of Bergenin from Ethyl Acetate Extract of Flueggea virosa Leaves. J. Chem. Soc. Niger. 2020, 45, 1042–1047. [Google Scholar] [CrossRef]
  87. Magaji, M.G.; Musa, A.M.; Abdullahi, M.I.; Ya ’u, J.; Hussaini, I.M. Isolation of Bergenin from the Root Bark of Securinega virosa and Evaluation of Its Potential Sleep Promoting Effect. AJP 2015, 5, 587–596. [Google Scholar] [PubMed]
  88. Subramanian, R.; Subbramaniyan, P.; Raj, V. Isolation of Bergenin from Peltophorum pterocarpum Flowers and Its Bioactivity. Beni-Suef Univ. J. Basic Appl. Sci. 2015, 4, 256–261. [Google Scholar] [CrossRef]
  89. Dawa, Y.; Du, Y.; Wang, Q.; Chen, C.; Zou, D.; Qi, D.; Ma, J.; Dang, J. Targeted Isolation of 1,1-Diphenyl-2-Picrylhydrazyl Inhibitors from Saxifraga Atrata Using Medium- and High-Pressure Liquid Chromatography Combined with Online High Performance Liquid Chromatography–1,1-Diphenyl-2-Picrylhydrazyl Detection. J. Chromatogr. A 2021, 1635, 461690. [Google Scholar] [CrossRef] [PubMed]
  90. Dang, J.; Ma, J.; Dawa, Y.; Liu, C.; Ji, T.; Wang, Q. Preparative Separation of 1,1-Diphenyl-2-Picrylhydrazyl Inhibitors Originating from: Saxifraga sinomontana Employing Medium-Pressure Liquid Chromatography in Combination with Reversed-Phase Liquid Chromatography. RSC Adv. 2021, 11, 38739–38749. [Google Scholar] [CrossRef] [PubMed]
  91. Hamid, S.; Ramli, R.; Manshoor, N. Resolving Co-Eluted Oligostilbenes Using Recycling High Performance Liquid Chromatography (R-HPLC). Aust. J. Basic Appl. Sci. Aust. J. Basic Appl. Sci 2016, 10, 111–116. [Google Scholar]
  92. Kirkland, J.J. Development of Some Stationary Phases for Reversed-Phase High-Performance Liquid Chromatography. J. Chromatogr. A 2004, 1060, 9–21. [Google Scholar] [CrossRef] [PubMed]
  93. Sarker, S.D.; Nahar, L. Applications of High Performance Liquid Chromatography in the Analysis of Herbal Products. In Evidence-Based Validation of Herbal Medicine; Elsevier Inc.: Amsterdam, The Netherlands, 2015. [Google Scholar] [CrossRef]
  94. Enogieru, A.B.; Haylett, W.; Hiss, D.C.; Bardien, S.; Ekpo, O.E. Rutin as a Potent Antioxidant: Implications for Neurodegenerative Disorders. Oxidative Med. Cell. Longev. 2018, 2018, 1–17. [Google Scholar] [CrossRef] [PubMed]
  95. Balmeh, N.; Mahmoudi, S.; Mohammadi, N.; Karabedianhajiabadi, A. Predicted Therapeutic Targets for COVID-19 Disease by Inhibiting SARS-CoV-2 and Its Related Receptors. Inform. Med. Unlocked 2020, 20, 100407. [Google Scholar] [CrossRef] [PubMed]
  96. Yingyuen, P.; Sukrong, S.; Phisalaphong, M. Isolation, Separation and Purification of Rutin from Banana Leaves (Musa balbisiana). Ind. Crops Prod. 2020, 149, 112307. [Google Scholar] [CrossRef]
  97. Sidana, J.; Joshi, L.K. Recycle HPLC: A Powerful Tool for the Purification of Natural Products. Chromatogr. Res. Int. 2013, 2013, 1–7. [Google Scholar] [CrossRef]
  98. Ramli, R.; Ismail, N.H.; Manshoor, N. Recycling HPLC for the Purification of Oligostilbenes from Dipterocarpus semivestitus and Neobalanocarpus heimii (Dipterocarpaceae). J. Liq. Chromatogr. Relat. Technol. 2017, 40, 943–949. [Google Scholar] [CrossRef]
  99. Latif, Z.; Sarker, S.D. Isolation of Natural Products by Preparative High Performance Liquid Chromatography (Prep-HPLC). Methods Mol. Biol. 2012, 864, 255–274. [Google Scholar] [CrossRef] [PubMed]
  100. Schmitt, M.; Alabdul Magid, A.; Hubert, J.; Etique, N.; Duca, L.; Voutquenne-Nazabadioko, L. Bio-Guided Isolation of New Phenolic Compounds from Hippocrepis emerus Flowers and Investigation of Their Antioxidant, Tyrosinase and Elastase Inhibitory Activities. Phytochem. Lett. 2020, 35, 28–36. [Google Scholar] [CrossRef]
  101. de la Luz Cádiz-Gurrea, M.; Fernández de las Nieves, I.; Aguilera Saez, L.M.; Fernández-Arroyo, S.; Legeai-Mallet, L.; Bouaziz, M.; Segura-Carretero, A. Bioactive Compounds from Theobroma Cacao: Effect of Isolation and Safety Evaluation. Plant Foods Hum. Nutr. 2019, 74, 40–46. [Google Scholar] [CrossRef] [PubMed]
  102. Zhao, X.; Chen, R.; Shi, Y.; Zhang, X.; Tian, C.; Xia, D. Antioxidant and Anti-Inflammatory Activities of Six Flavonoids from Smilax glabra Roxb. Molecules 2020, 25, 5295. [Google Scholar] [CrossRef] [PubMed]
  103. Ao, X.; Yan, J.; Liu, S.; Chen, S.; Zou, L.; Yang, Y.; He, L.; Li, S.; Liu, A.; Zhao, K. Extraction, Isolation and Identification of Four Phenolic Compounds from Pleioblastus amarus Shoots and Their Antioxidant and Anti-Inflammatory Properties In Vitro. Food Chem. 2022, 374, 131743. [Google Scholar] [CrossRef] [PubMed]
  104. Mansour, R.B.; Wided, M.K.; Cluzet, S.; Krisa, S.; Richard, T.; Ksouri, R. LC-MS Identification and Preparative HPLC Isolation of Frankenia pulverulenta Phenolics with Antioxidant and Neuroprotective Capacities in PC12 Cell Line. Pharm. Biol. 2017, 55, 880–887. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, M.; Cheng, S.; Liang, Y.; Mu, Y.; Yan, H.; Liu, Q.; Geng, Y.; Wang, X.; Zhao, H. Rapid Purification of Antioxidants from Magnolia officinalis by Semi-Prep-HPLC with a Two-Step Separation Strategy Guided by on-Line HPLC-Radical Scavenging Detection. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1100–1101, 140–147. [Google Scholar] [CrossRef]
  106. Elmastas, M.; Celik, S.M.; Genc, N.; Aksit, H.; Erenler, R.; Gulcin, İ. Antioxidant Activity of an Anatolian Herbal Tea—Origanum minutiflorum: Isolation and Characterization of Its Secondary Metabolites. Int. J. Food Prop. 2018, 21, 374–384. [Google Scholar] [CrossRef]
  107. Ghasemi, S.; Evazalipour, M.; Peyghanbari, N.; Zamani, E.; Bellstedt, P.; Molaee, M.; Koohi, D.E.; Yousefbeyk, F. Isolation and Structure Elucidation of the Compounds from Teucrium hyrcanicum L. and the Investigation of Cytotoxicity, Antioxidant Activity, and Protective Effect on Hydrogen Peroxide-Induced Oxidative Stress. BMC Complement. Med. Ther. 2023, 23, 447. [Google Scholar] [CrossRef] [PubMed]
  108. Santos, C.C. de S.; Masullo, M.; Cerulli, A.; Mari, A.; Estevam, C.D.S.; Pizza, C.; Piacente, S. Isolation of Antioxidant Phenolics from Schinopsis brasiliensis Based on a Preliminary LC-MS Profiling. Phytochemistry 2017, 140, 45–51. [Google Scholar] [CrossRef]
  109. Baiseitova, A.; Shah, A.B.; Khan, A.M.; Idrees, M.; Kim, J.H.; Lee, Y.H.; Kong, I.K.; Park, K.H. Antioxidant Potentials of Furanodihydrobenzoxanthones from Artocarpus Elasticus and Their Protection against OxLDL Induced Injury in SH-SY5Y Cells. Biomed. Pharmacother. 2023, 165, 115278. [Google Scholar] [CrossRef] [PubMed]
  110. Muddassir, M.; Batool, A.; Alam, M.; Abbas Miana, G.; Altaf, R.; Alghamdi, S.; Almehmadi, M.; Abdulaziz, O.; Amer Alsaiari, A.; Umar Khayam Sahibzada, M.; et al. Evaluation of In Vitro, In Silico Antidiabetic and Antioxidant Potential of Bioactivity Based Isolated “Pakistanine” from Berberis baluchistanica. Arab. J. Chem. 2022, 15, 104221. [Google Scholar] [CrossRef]
  111. Abuduaini, M.; Li, J.; Ruan, J.H.; Zhao, Y.X.; Maitinuer, M.; Aisa, H.A. Bioassay-Guided Preparation of Antioxidant, Anti-Inflammatory Active Fraction from Crabapples (Malus prunifolia (Willd.) Borkh.). Food Chem. 2023, 406, 135091. [Google Scholar] [CrossRef] [PubMed]
  112. Zhao, J.Q.; Wang, Y.M.; Yang, Y.L.; Zeng, Y.; Wang, Q.L.; Shao, Y.; Mei, L.J.; Shi, Y.P.; Tao, Y.D. Isolation and Identification of Antioxidant and α-Glucosidase Inhibitory Compounds from Fruit Juice of Nitraria tangutorum. Food Chem. 2017, 227, 93–101. [Google Scholar] [CrossRef] [PubMed]
  113. Pantoja Pulido, K.D.; Colmenares Dulcey, A.J.; Isaza Martínez, J.H. New Caffeic Acid Derivative from Tithonia diversifolia (Hemsl.) A. Gray Butanolic Extract and Its Antioxidant Activity. Food Chem. Toxicol. 2017, 109, 1079–1085. [Google Scholar] [CrossRef] [PubMed]
  114. Ito, Y. Countercurrent Chromatography—Overview. In Encyclopedia of Analytical Science: Second Edition; Elsevier Ltd.: Amsterdam, The Netherlands, 2004. [Google Scholar] [CrossRef]
  115. Shibuswa, Y.; Ito, Y. PROTEINS|High-Speed Countercurrent Chromatography. In Encyclopedia of Separation Science; Academic Press: Cambridge, MA, USA, 2000. [Google Scholar] [CrossRef]
  116. Ito, Y. Countercurrent Chromatography: Overview☆. In Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier Ltd.: Amsterdam, The Netherlands, 2016. [Google Scholar] [CrossRef]
  117. Hamzaoui, M.; Renault, J.H.; Reynaud, R.; Hubert, J. Centrifugal Partition Extraction in the PH-Zone-Refining Displacement Mode: An Efficient Strategy for the Screening and Isolation of Biologically Active Phenolic Compounds. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2013, 937, 7–12. [Google Scholar] [CrossRef] [PubMed]
  118. Bojczuk, M.; Żyżelewicz, D.; Hodurek, P. Centrifugal Partition Chromatography—A Review of Recent Applications and Some Classic References. J. Sep. Sci. 2017, 40, 1597–1609. [Google Scholar] [CrossRef]
  119. Kumar, G.M.; Neelam, I.; Ajitha, A.; Rao, V.U.M. Centrifugal Partition Chromatography: An Overview. Int. J. Pharm. Res. Anal. 2014, 4, 353–360. [Google Scholar]
  120. Schwienheer, C.; Merz, J.; Schembecker, G. Investigation, Comparison and Design of Chambers Used in Centrifugal Partition Chromatography on the Basis of Flow Pattern and Separation Experiments. J. Chromatogr. A 2015, 1390, 39–49. [Google Scholar] [CrossRef] [PubMed]
  121. Bezold, F.; Goll, J.; Minceva, M. Study of the Applicability of Non-Conventional Aqueous Two-Phase Systems in Counter-Current and Centrifugal Partition Chromatography. J. Chromatogr. A 2015, 1388, 126–132. [Google Scholar] [CrossRef] [PubMed]
  122. Lee, J.H.; Ko, J.Y.; Samarakoon, K.; Oh, J.Y.; Heo, S.J.; Kim, C.Y.; Nah, J.W.; Jang, M.K.; Lee, J.S.; Jeon, Y.J. Preparative Isolation of Sargachromanol E from Sargassum siliquastrum by Centrifugal Partition Chromatography and Its Anti-Inflammatory Activity. Food Chem. Toxicol. 2013, 62, 54–60. [Google Scholar] [CrossRef] [PubMed]
  123. Huang, X.Y.; Ignatova, S.; Hewitson, P.; Di, D.L. An Overview of Recent Progress in Elution Mode of Counter Current Chromatography. TrAC Trends Anal. Chem. 2016, 77, 214–225. [Google Scholar] [CrossRef]
  124. Abdelgadir, A.A.; Boudesocque-Delaye, L.; Thery-Koné, I.; Gueiffier, A.; Ahmed, E.M.; Enguehard-Gueiffier, C. One-Step Preparative Isolation of Aristolochic Acids by Strong Ion-Exchange Centrifugal Partition Chromatography. Sep. Purif. Technol. 2015, 156, 444–449. [Google Scholar] [CrossRef]
  125. Barrientos, R.; Fernández-Galleguillos, C.; Pastene, E.; Simirgiotis, M.; Romero-Parra, J.; Ahmed, S.; Echeverría, J. Metabolomic Analysis, Fast Isolation of Phenolic Compounds, and Evaluation of Biological Activities of the Bark From Weinmannia trichosperma Cav. (Cunoniaceae). Front. Pharmacol. 2020, 11, 780. [Google Scholar] [CrossRef] [PubMed]
  126. Bezold, F.; Weinberger, M.E.; Minceva, M. Computational Solvent System Screening for the Separation of Tocopherols with Centrifugal Partition Chromatography Using Deep Eutectic Solvent-Based Biphasic Systems. J. Chromatogr. A 2017, 1491, 153–158. [Google Scholar] [CrossRef] [PubMed]
  127. Bezold, F.; Minceva, M. Liquid-Liquid Equilibria of n-Heptane, Methanol and Deep Eutectic Solvents Composed of Carboxylic Acid and Monocyclic Terpenes. Fluid Phase Equilib. 2018, 477, 98–106. [Google Scholar] [CrossRef]
  128. Bezold, F.; Minceva, M. A Water-Free Solvent System Containing an L-Menthol-Based Deep Eutectic Solvent for Centrifugal Partition Chromatography Applications. J. Chromatogr. A 2019, 1587, 166–171. [Google Scholar] [CrossRef] [PubMed]
  129. Muley, P.D.; Mobley, J.K.; Tong, X.; Novak, B.; Stevens, J.; Moldovan, D.; Shi, J.; Boldor, D. Rapid Microwave-Assisted Biomass Delignification and Lignin Depolymerization in Deep Eutectic Solvents. Energy Convers. Manag. 2019, 196, 1080–1088. [Google Scholar] [CrossRef]
  130. Wang, S.; Su, S.; Xiao, L.P.; Wang, B.; Sun, R.C.; Song, G. Catechyl Lignin Extracted from Castor Seed Coats Using Deep Eutectic Solvents: Characterization and Depolymerization. ACS Sustain. Chem. Eng. 2020, 8, 7031–7038. [Google Scholar] [CrossRef]
  131. Cunha, S.C.; Fernandes, J.O. Extraction Techniques with Deep Eutectic Solvents. TrAC Trends Anal. Chem. 2018, 105, 225–239. [Google Scholar] [CrossRef]
  132. Lee, J.S.; Lee, A.Y.; Quilantang, N.G.; Geraldino, P.J.L.; Cho, E.J.; Lee, S. Anti-Oxidant Activity of Avicularin and Isovitexin from Lespedeza Cuneata. J. Appl. Biol. Chem. 2019, 62, 143–147. [Google Scholar] [CrossRef]
  133. Dugé de Bernonville, T.; Guyot, S.; Paulin, J.P.; Gaucher, M.; Loufrani, L.; Henrion, D.; Derbré, S.; Guilet, D.; Richomme, P.; Dat, J.F.; et al. Dihydrochalcones: Implication in Resistance to Oxidative Stress and Bioactivities against Advanced Glycation End-Products and Vasoconstriction. Phytochemistry 2010, 71, 443–452. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, Y.F.; Lin, P.; Huang, Y.L.; He, R.J.; Yang, B.Y.; Liu, Z. Bin. Isolation of Two New Phenolic Glycosides from Castanopsis Chinensis Hance by Combined Multistep CC and HSCCC Separation and Evaluation of Their Antioxidant Activity. Molecules 2023, 28, 3331. [Google Scholar] [CrossRef] [PubMed]
  135. Guzlek, H.; Wood, P.L.; Janaway, L. Performance Comparison Using the GUESS Mixture to Evaluate Counter-Current Chromatography Instruments. J. Chromatogr. A 2009, 1216, 4181–4186. [Google Scholar] [CrossRef] [PubMed]
  136. Zhou, Y.; Shan, H.; Haitao, L. Optimization, Extraction, and Purification of Three Bioactive Compounds from Entada phaseoloides by High-Speed Countercurrent Chromatography. Biomed. Chromatogr. 2021, 35, e5232. [Google Scholar] [CrossRef] [PubMed]
  137. Echiburu-Chau, C.; Pastén, L.; Parra, C.; Bórquez, J.; Mocan, A.; Simirgiotis, M.J. High Resolution UHPLC-MS Characterization and Isolation of Main Compounds from the Antioxidant Medicinal Plant Parastrephia lucida (Meyen). Saudi Pharm. J. 2017, 25, 1032–1039. [Google Scholar] [CrossRef] [PubMed]
  138. Shaheen, N.; Lu, Y.; Geng, P.; Shao, Q.; Wei, Y. Isolation of Four Phenolic Compounds from Mangifera indica L. Flowers by Using Normal Phase Combined with Elution Extrusion Two-Step High Speed Countercurrent Chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2017, 1046, 211–217. [Google Scholar] [CrossRef] [PubMed]
  139. Bianchi, S.E.; Kaiser, S.; Pittol, V.; Doneda, E.; De Souza, K.C.B.; Bassani, V.L. Semi-Preparative Isolation and Purification of Phenolic Compounds from Achyrocline satureioides (Lam) D.C. by High-Performance Counter-Current Chromatography. Phytochem. Anal. 2019, 30, 182–192. [Google Scholar] [CrossRef]
  140. Kiene, M.; Blum, S.; Jerz, G.; Winterhalter, P. A Comparison between High-Performance Countercurrent Chromatography and Fast-Centrifugal Partition Chromatography for a One-Step Isolation of Flavonoids from Peanut Hulls Supported by a Conductor like Screening Model for Real Solvents. Molecules 2023, 28, 5111. [Google Scholar] [CrossRef]
  141. Liu, M.; Li, X.; Liu, Q.; Xie, S.; Zhu, F.; Chen, X. Preparative Isolation and Purification of 12 Main Antioxidants from the Roots of Polygonum multiflorum Thunb. Using High-Speed Countercurrent Chromatography and Preparative HPLC Guided by 1,1′-Diphenyl-2-Picrylhydrazyl-HPLC. J. Sep. Sci. 2020, 43, 1415–1422. [Google Scholar] [CrossRef] [PubMed]
  142. Wang, D.; Du, N.; Wen, L.; Zhu, H.; Liu, F.; Wang, X.; Du, J.; Li, S. An Efficient Method for the Preparative Isolation and Purification of Flavonoid Glycosides and Caffeoylquinic Acid Derivatives from Leaves of Lonicera japonica Thunb. Using High Speed Counter-Current Chromatography (HSCCC) and Prep-HPLC Guided by DPPH-HPLC Experiments. Molecules 2017, 22, 229. [Google Scholar] [CrossRef]
  143. Peng, J.; Li, K.; Zhu, W.; Deng, X.; Li, C. Separation and Purification of Four Phenolic Compounds from Persimmon by High-Speed Counter-Current Chromatography. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1072, 78–85. [Google Scholar] [CrossRef] [PubMed]
  144. Yang, F.; Qi, Y.; Liu, W.; Li, J.; Wang, D.; Fang, L.; Zhang, Y. Separation of Five Flavonoids from Aerial Parts of Salvia miltiorrhiza Bunge Using HSCCC and Their Antioxidant Activities. Molecules 2019, 24, 3448. [Google Scholar] [CrossRef] [PubMed]
  145. Hu, W.; Zhou, J.; Shen, T.; Wang, X. Target-Guided Isolation of Three Main Antioxidants from Mahonia bealei (Fort.) Carr. Leaves Using HSCCC. Molecules 2019, 24, 1907. [Google Scholar] [CrossRef] [PubMed]
  146. Dong, X.; Huang, H.; Wang, R.; Luo, S.; Mi, Y.; Pan, Y.; Shen, W.; Cui, J.; Hu, X.; Cheng, X.; et al. High-Speed Counter-Current Chromatography Assisted Preparative Isolation of Phenolic Compounds from the Flowers of Chrysanthemum morifolium Cv. Fubaiju. J. Sep. Sci. 2023, 46, e2300172. [Google Scholar] [CrossRef] [PubMed]
  147. Correa, D.I.; Pastene-Navarrete, E.; Bustamante, L.; Baeza, M.; Alarcón-Enos, J. Isolation of Three Lycorine Type Alkaloids from Rhodolirium speciosum (Herb.) Ravenna Using Ph-Zone-Refinement Centrifugal Partition Chromatography and Their Acetylcholinesterase Inhibitory Activities. Metabolites 2020, 10, 309. [Google Scholar] [CrossRef] [PubMed]
  148. Szabó, L.U.; Schmidt, T.J. Target-Guided Isolation of O-Tigloylcyclovirobuxeine-B from Buxus sempervirens L. By Centrifugal Partition Chromatography. Molecules 2020, 25, 4804. [Google Scholar] [CrossRef] [PubMed]
  149. Bárcenas-Pérez, D.; Lukeš, M.; Hrouzek, P.; Kubáč, D.; Kopecký, J.; Kaštánek, P.; Cheel, J. A Biorefinery Approach to Obtain Docosahexaenoic Acid and Docosapentaenoic Acid N-6 from Schizochytrium Using High Performance Countercurrent Chromatography. Algal Res. 2021, 55, 102241. [Google Scholar] [CrossRef]
  150. Hage, D.S. Chromatography. In Principles and Applications of Clinical Mass Spectrometry: Small Molecules, Peptides, and Pathogens; Elsevier Ltd.: Amsterdam, The Netherlands, 2018; pp. 1–32. [Google Scholar] [CrossRef]
  151. Berlinck, R.G.S.; Crnkovic, C.M.; Gubiani, J.R.; Bernardi, D.I.; Ióca, L.P.; Quintana-Bulla, J.I. The Isolation of Water-Soluble Natural Products—Challenges, Strategies and Perspectives. Nat. Prod. Rep. 2021, 39, 596–669. [Google Scholar] [CrossRef]
  152. Wang, S.; Cao, J.; Deng, J.; Hou, X.; Hao, E.; Zhang, L.; Yu, H.; Li, P. Chemical Characterization of Flavonoids and Alkaloids in Safflower (Carthamus tinctorius L.) by Comprehensive Two-Dimensional Hydrophilic Interaction Chromatography Coupled with Hybrid Linear Ion Trap Orbitrap Mass Spectrometry. Food Chem. X 2021, 12, 100143. [Google Scholar] [CrossRef] [PubMed]
  153. Dang, J.; Zhang, L.; Shao, Y.; Mei, L.; Liu, Z.; Yue, H.; Wang, Q.; Tao, Y. Preparative Isolation of Antioxidative Compounds from Dracocephalum heterophyllum Using Off-Line Two-Dimensional Reversed-Phase Liquid Chromatography/Hydrophilic Interaction Chromatography Guided by on-Line HPLC-DPPH Assay. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1095, 267–274. [Google Scholar] [CrossRef] [PubMed]
  154. Dang, J.; Zhang, L.; Wang, Q.; Mei, L.; Yue, H.; Liu, Z.; Shao, Y.; Gao, Q.; Tao, Y. Target Separation of Flavonoids from Saxifraga tangutica Using Two-Dimensional Hydrophilic Interaction Chromatography/Reversed-Phase Liquid Chromatography. J. Sep. Sci. 2018, 41, 4419–4429. [Google Scholar] [CrossRef] [PubMed]
  155. Dang, J.; Shao, Y.; Zhao, J.; Mei, L.; Tao, Y.; Wang, Q.; Zhang, L. Two-Dimensional Hydrophilic Interaction Chromatography × Reversed-Phase Liquid Chromatography for the Preparative Isolation of Potential Anti-Hepatitis Phenylpropanoids from Salvia prattii. J. Sep. Sci. 2016, 39, 3327–3338. [Google Scholar] [CrossRef] [PubMed]
  156. Cui, Y.; Shen, N.; Yuan, X.; Dang, J.; Shao, Y.; Mei, L.; Tao, Y.; Wang, Q.; Liu, Z. Two-Dimensional Chromatography Based on on-Line HPLC-DPPH Bioactivity-Guided Assay for the Preparative Isolation of Analogue Antioxidant Compound from Arenaria kansuensis. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2017, 1046, 81–86. [Google Scholar] [CrossRef] [PubMed]
  157. Kumari, V.B.C.; Patil, S.M.; Ramu, R.; Shirahatti, P.S.; Kumar, N.; Sowmya, B.P.; Egbuna, C.; Uche, C.Z.; Patrick-Iwuanyanwu, K.C. Chromatographic Techniques: Types, Principles, and Applications. In Analytical Techniques in Biosciences; Academic Press: Cambridge, MA, USA, 2022. [Google Scholar] [CrossRef]
  158. Srivastava, N.; Singh, A.; Kumari, P.; Nishad, J.H.; Gautam, V.S.; Yadav, M.; Bharti, R.; Kumar, D.; Kharwar, R.N. Advances in Extraction Technologies: Isolation and Purification of Bioactive Compounds from Biological Materials. In Natural Bioactive Compounds; Academic Press: Cambridge, MA, USA, 2021. [Google Scholar] [CrossRef]
  159. Zhang, Q.W.; Lin, L.G.; Ye, W.C. Techniques for Extraction and Isolation of Natural Products: A Comprehensive Review. Chin. Med. 2018, 13, 1–26. [Google Scholar] [CrossRef] [PubMed]
  160. Zafar, R.; Ullah, H.; Zahoor, M.; Sadiq, A. Isolation of Bioactive Compounds from Bergenia ciliata (Haw.) Sternb Rhizome and Their Antioxidant and Anticholinesterase Activities. BMC Complement. Altern. Med. 2019, 19, 296. [Google Scholar] [CrossRef] [PubMed]
  161. Månsson, M.; Phipps, R.K.; Gram, L.; Munro, M.H.G.; Larsen, T.O.; Nielsen, K.F. Explorative Solid-Phase Extraction (E-SPE) for Accelerated Microbial Natural Product Discovery, Dereplication, and Purification. J. Nat. Prod. 2010, 73, 1126–1132. [Google Scholar] [CrossRef] [PubMed]
  162. Murphy, B.E.P.; D’Aux, R.C.D. The Use of Sephadex LH-20 Column Chromatography to Separate Unconjugated Steroids. J. Steroid Biochem. 1975, 6, 233–237. [Google Scholar] [CrossRef]
  163. Stabrauskiene, J.; Kopustinskiene, D.M.; Lazauskas, R.; Bernatoniene, J. Naringin and Naringenin: Their Mechanisms of Action and the Potential Anticancer Activities. Biomedicines 2022, 10, 1686. [Google Scholar] [CrossRef]
  164. Xu, X.; Huang, Y.; Xu, J.; He, X.; Wang, Y. Anti-Neuroinflammatory and Antioxidant Phenols from Mulberry Fruit (Morus alba L.). J. Funct. Foods 2020, 68, 103914. [Google Scholar] [CrossRef]
  165. Muniza, M.P.; Nunomura, S.M.; Lima, E.S.; Lima, A.S.; Patrícia, D.O.d.A.; Nunomuraa, R.C.S. Quantification of bergenin, antioxidant activity and nitric oxide inhibition from bark, leaf and twig of Endopleura uchi. Quim. Nova 2020, 43, 413–418. [Google Scholar] [CrossRef]
  166. Chethankumara, G.P.; Nagaraj, K.; Krishna, V.; Krishnaswamy, G. Isolation, Characterization and In Vitro Cytotoxicity Studies of Bioactive Compounds from Alseodaphne semecarpifolia Nees. Heliyon 2021, 7, e07325. [Google Scholar] [CrossRef] [PubMed]
  167. César, L.R.J.; Javier, H.; Fernando, A.Z.J.; Carlos, V.; Enrique, R.Z.R.; Efrain, A.; Evelin, M.B.; Inocencio, H.C.; Luis, O.J.; Zaira, D.; et al. Identification of the Main Phenolic Compounds Responsible for the Antioxidant Activity of Litsea glaucescens Kunth. S. Afr. J. Bot. 2022, 147, 208–214. [Google Scholar] [CrossRef]
  168. Atun, S.; Handayani, S.; Rakhmawati, A. Potential Bioactive Compounds Isolated from Boesenbergia rotunda as Antioxidant and Antimicrobial Agents. Pharmacogn. J. 2018, 10, 513–518. [Google Scholar] [CrossRef]
  169. Sun, S.; Zhao, Y.; Wang, L.; Tan, Y.; Shi, Y.; Sedjoah, R.C.A.A.; Shao, Y.; Li, L.; Wang, M.; Wan, J.; et al. Ultrasound-Assisted Extraction of Bound Phenolic Compounds from the Residue of Apocynum venetum Tea and Their Antioxidant Activities. Food Biosci. 2022, 47, 101646. [Google Scholar] [CrossRef]
  170. Minh, T.N.; Xuan, T.D.; Tran, H.D.; Van, T.M.; Andriana, Y.; Khanh, T.D.; Van Quan, N.; Ahmad, A. Isolation and Purification of Bioactive Compounds from the Stem Bark of Jatropha podagrica. Molecules 2019, 24, 889. [Google Scholar] [CrossRef]
  171. Aljubiri, S.M.; Mahgoub, S.A.; Almansour, A.I.; Shaaban, M.; Shaker, K.H. Isolation of Diverse Bioactive Compounds from Euphorbia balsamifera: Cytotoxicity and Antibacterial Activity Studies. Saudi J. Biol. Sci. 2021, 28, 417–426. [Google Scholar] [CrossRef] [PubMed]
  172. Prakash, S.; Elavarasan, N.; Subashini, K.; Kanaga, S.; Dhandapani, R.; Sivanandam, M.; Kumaradhas, P.; Thirunavukkarasu, C.; Sujatha, V. Isolation of Hesperetin—A Flavonoid from Cordia sebestena Flower Extract through Antioxidant Assay Guided Method and Its Antibacterial, Anticancer Effect on Cervical Cancer via In Vitro and In Silico Molecular Docking Studies. J. Mol. Struct. 2020, 1207, 127751. [Google Scholar] [CrossRef]
  173. Alagesan, V.; Ramalingam, S.; Kim, M.; Venugopal, S. Antioxidant Activity Guided Isolation of a Coumarin Compound from Ipomoea Pes-Caprea (Convolvulaceae) Leaves Acetone Extract and Its Biological and Molecular Docking Studies. Eur. J. Integr. Med. 2019, 32, 100984. [Google Scholar] [CrossRef]
  174. Vigbedor, B.Y.; Akoto, C.O.; Neglo, D. Isolation and Characterization of 3,3′-Di-O-Methyl Ellagic Acid from the Root Bark of Afzelia Africana and Its Antimicrobial and Antioxidant Activities. Sci. Afr. 2022, 17, e01332. [Google Scholar] [CrossRef]
  175. Abdallah, H.M.; Esmat, A. Antioxidant and Anti-Inflammatory Activities of the Major Phenolics from Zygophyllum simplex L. J. Ethnopharmacol. 2017, 205, 51–56. [Google Scholar] [CrossRef] [PubMed]
  176. Al-Rifai, A. Identification and Evaluation of In-Vitro Antioxidant Phenolic Compounds from the Calendula tripterocarpa Rupr. South African J. Bot. 2018, 116, 238–244. [Google Scholar] [CrossRef]
  177. Karakaya, S.; Koca, M.; Sytar, O.; Dursunoglu, B.; Ozbek, H.; Duman, H.; Guvenalp, Z.; Kılıc, C.S. Antioxidant and Anticholinesterase Potential of Ferulago cassia with Farther Bio-Guided Isolation of Active Coumarin Constituents. S. Afr. J. Bot. 2019, 121, 536–542. [Google Scholar] [CrossRef]
  178. Lee, J.H.; Cho, Y.S. Assessment of Phenolic Profiles from Various Organs in Different Species of Perilla Plant (Perilla frutescens (L.) Britt.) and Their Antioxidant and Enzyme Inhibitory Potential. Ind. Crops Prod. 2021, 171, 113914. [Google Scholar] [CrossRef]
  179. Zefzoufi, M.; Fdil, R.; Bouamama, H.; Gadhi, C.; Katakura, Y.; Mouzdahir, A.; Sraidi, K. Effect of Extracts and Isolated Compounds Derived from Retama monosperma (L.) Boiss. on Anti-Aging Gene Expression in Human Keratinocytes and Antioxidant Activity. J. Ethnopharmacol. 2021, 280, 114451. [Google Scholar] [CrossRef] [PubMed]
  180. Erenler, R.; Meral, B.; Sen, O.; Elmastas, M.; Aydin, A.; Eminagaoglu, O.; Topcu, G. Bioassay-Guided Isolation, Identification of Compounds from Origanum rotundifolium and Investigation of Their Antiproliferative and Antioxidant Activities. Pharm. Biol. 2017, 55, 1646–1653. [Google Scholar] [CrossRef] [PubMed]
  181. Reshma, M.V.; Jacob, J.; Syamnath, V.L.; Habeeba, V.P.; Dileep Kumar, B.S.; Lankalapalli, R.S. First Report on Isolation of 2,3,4-Trihydroxy-5-Methylacetophenone from Palmyra Palm (Borassus flabellifer Linn.) Syrup, Its Antioxidant and Antimicrobial Properties. Food Chem. 2017, 228, 491–496. [Google Scholar] [CrossRef] [PubMed]
  182. Bayrakçeken Güven, Z.; Dogan, Z.; Saracoglu, I.; Picot, L.; Nagatsu, A.; Basaran, A.A. Food Plant with Antioxidant, Tyrosinase Inhibitory and Antimelanoma Activity: Prunus mahaleb L. Food Biosci. 2022, 48, 101804. [Google Scholar] [CrossRef]
  183. Zhang, Z.; Dai, L.; Wang, H.; Chang, X.; Ren, S.; Lai, H.; Liu, L. Phytochemical Profiles and Antioxidant, Anticholinergic, and Antidiabetic Activities of Odontites serotina (Lam.) Dum. Eur. J. Integr. Med. 2021, 44, 101340. [Google Scholar] [CrossRef]
  184. Al-Yousef, H.M.; Alqahtani, A.S.; Ghani, A.E.A.; El-Toumy, S.A.; El-Dougdoug, W.I.A.; Hassan, W.H.B.; Hassan, H.M. Nephroprotective and Antioxidant Activities of Ethyl Acetate Fraction of Euphorbia geniculata Ortega Family Euphorbiaceae. Arab. J. Chem. 2020, 13, 7843–7850. [Google Scholar] [CrossRef]
  185. Shangguan, Y.; Ni, J.; Jiang, L.; Hu, Y.; He, C.; Ma, Y.; Wu, G.; Xiong, H. Response Surface Methodology-Optimized Extraction of Flavonoids from Pomelo Peels and Isolation of Naringin with Antioxidant Activities by Sephadex LH20 Gel Chromatography. Curr. Res. Food Sci. 2023, 7, 100610. [Google Scholar] [CrossRef] [PubMed]
  186. Xu, Q.N.; Zhu, D.; Wang, G.H.; Lin, T.; Sun, C.L.; Ding, R.; Tian, W.J.; Chen, H.F. Phenolic Glycosides and Flavonoids with Antioxidant and Anticancer Activities from Desmodium caudatum. Nat. Prod. Res. 2020, 35, 4534–4541. [Google Scholar] [CrossRef] [PubMed]
  187. Sun, N.; Yang, H.; Zhou, M.; Hou, Y.; Wu, X.; Li, W.; Zhang, Y.; Sun, J.; Xiong, X.; Huang, J.; et al. Isolation and Evaluation of Antioxidants from Arisaema heterophyllum Tubers. Phytochem. Lett. 2023, 53, 106–110. [Google Scholar] [CrossRef]
  188. Nina, N.; Theoduloz, C.; Giménez, A.; Schmeda-Hirschmann, G. Phenolics from the Bolivian Highlands Food Plant Ombrophytum subterraneum (Aspl.) B. Hansen (Balanophoraceae): Antioxidant and α-Glucosidase Inhibitory Activity. Food Res. Int. 2020, 137, 109382. [Google Scholar] [CrossRef] [PubMed]
  189. Ganzon, J.G.; Chen, L.G.; Wang, C.C. 4-O-Caffeoylquinic Acid as an Antioxidant Marker for Mulberry Leaves Rich in Phenolic Compounds. J. Food Drug Anal. 2017, 26, 985–993. [Google Scholar] [CrossRef] [PubMed]
  190. Baky, M.H.; Kamal, A.M.; Haggag, E.G.; Elgindi, M.R. Flavonoids from Manilkara hexandra and Antimicrobial and Antioxidant Activities. Biochem. Syst. Ecol. 2022, 100, 104375. [Google Scholar] [CrossRef]
  191. Lau, E. Handbook of Modern Pharmaceutical Analysis: Preformulation Studies; Academic Press: Cambridge, MA, USA, 2001; Volume 3. [Google Scholar]
  192. Ul’yanovskii, N.V.; Onuchina, A.A.; Ovchinnikov, D.V.; Faleva, A.V.; Gorbova, N.S.; Kosyakov, D.S. Analytical and Preparative Separation of Softwood Lignans by Supercritical Fluid Chromatography. Separations 2023, 10, 449. [Google Scholar] [CrossRef]
  193. Song, W.; Qiao, X.; Liang, W.F.; Ji, S.; Yang, L.; Wang, Y.; Xu, Y.W.; Yang, Y.; Guo, D.A.; Ye, M. Efficient Separation of Curcumin, Demethoxycurcumin, and Bisdemethoxycurcumin from Turmeric Using Supercritical Fluid Chromatography: From Analytical to Preparative Scale. J. Sep. Sci. 2015, 38, 3450–3453. [Google Scholar] [CrossRef] [PubMed]
  194. Yang, B.; Xin, H.; Wang, F.; Cai, J.; Liu, Y.; Fu, Q.; Jin, Y.; Liang, X. Purification of Lignans from Fructus arctii Using Off-Line Two-Dimensional Supercritical Fluid Chromatography/Reversed-Phase Liquid Chromatography. J. Sep. Sci. 2017, 40, 3231–3238. [Google Scholar] [CrossRef]
  195. Zhang, L.; Sun, A.; Li, A.; Kang, J.; Wang, Y.; Liu, R. Isolation and Purification of Osthole and Imperatorin from Fructus cnidii by Semi-Preparative Supercritical Fluid Chromatography. J. Liq. Chromatogr. Relat. Technol. 2017, 40, 407–414. [Google Scholar] [CrossRef]
  196. Wu, D.; Ge, D.; Dai, Y.; Chen, Y.; Fu, Q.; Jin, Y. Extraction and Isolation of Diphenylheptanes and Flavonoids from Alpinia officinarum Hance Using Supercritical Fluid Extraction Followed by Supercritical Fluid Chromatography. J. Sep. Sci. 2023, 46, e2300156. [Google Scholar] [CrossRef] [PubMed]
  197. Belbruno, J.J. Molecularly Imprinted Polymers. Chem. Rev. 2019, 119, 94–119. [Google Scholar] [CrossRef] [PubMed]
  198. Zhang, Y.; Xie, Y.; Zhang, C.; Wu, M.; Feng, S. Preparation of Porous Magnetic Molecularly Imprinted Polymers for Fast and Specifically Extracting Trace Norfloxacin Residue in Pork Liver. J. Sep. Sci. 2019, 43, 478–485. [Google Scholar] [CrossRef] [PubMed]
  199. Kong, X.; Li, F.; Li, Y.; He, X.; Chen, L.; Zhang, Y. Molecularly Imprinted Polymer Functionalized Magnetic Fe3O4 for the Highly Selective Extraction of Triclosan. J. Sep. Sci. 2019, 43, 808–817. [Google Scholar] [CrossRef] [PubMed]
  200. Abbasi Ghaeni, F.; Karimi, G.; Mohsenzadeh, M.S.; Nazarzadeh, M.; Motamedshariaty, V.S.; Mohajeri, S.A. Preparation of Dual-Template Molecularly Imprinted Nanoparticles for Organophosphate Pesticides and Their Application as Selective Sorbents for Water Treatment. Sep. Sci. Technol. 2018, 53, 2517–2526. [Google Scholar] [CrossRef]
  201. Ostovan, A.; Ghaedi, M.; Arabi, M.; Yang, Q.; Li, J.; Chen, L. Hydrophilic Multitemplate Molecularly Imprinted Biopolymers Based on a Green Synthesis Strategy for Determination of B-Family Vitamins. ACS Appl. Mater. Interfaces 2018, 10, 4140–4150. [Google Scholar] [CrossRef] [PubMed]
  202. Huang, M.; Pang, W.; Zhang, J.; Lin, S.; Hu, J. A Target Analogue Imprinted Polymer for the Recognition of Antiplatelet Active Ingredients in Radix Salviae Miltiorrhizae by LC/MS/MS. J. Pharm. Biomed. Anal. 2012, 58, 12–18. [Google Scholar] [CrossRef] [PubMed]
  203. Fu, Q.; Fang, Q.; Feng, B.; Sun, S.; Du, W.; Amut, E.; Xiao, A.; Chang, C. Matrine-Imprinted Monolithic Stationary Phase for Extraction and Purification of Matrine from Sophorae flavescentis Ait. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2011, 879, 894–900. [Google Scholar] [CrossRef]
  204. Sun, Y.; Zhang, Y.; Ju, Z.; Niu, L.; Gong, Z.; Xu, Z. Molecularly Imprinted Polymers Fabricated by Pickering Emulsion Polymerization for the Selective Adsorption and Separation of Quercetin from Spina Gleditsiae. New J. Chem. 2019, 43, 14747–14755. [Google Scholar] [CrossRef]
  205. Hosny, H.; El Gohary, N.; Saad, E.; Handoussa, H.; El Nashar, R.M. Isolation of Sinapic Acid from Broccoli Using Molecularly Imprinted Polymers. J. Sep. Sci. 2018, 41, 1164–1172. [Google Scholar] [CrossRef] [PubMed]
  206. Eidi, S.; Iranshahi, M.; Mohammadinejad, A.; Mohsenzadeh, M.S.; Farhadi, F.; Mohajeri, S.A. Selective Isolation of Sesquiterpene Coumarins from Asafoetida Using Dummy Molecularly Imprinted Solid Phase Extraction Method. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2020, 1138, 121943. [Google Scholar] [CrossRef] [PubMed]
  207. Ma, X.; Lin, H.; Zhang, J.; She, Y.; Zhou, X.; Li, X.; Cui, Y.; Wang, J.; Rabah, T.; Shao, Y. Extraction and Identification of Matrine-Type Alkaloids from Sophora moorcroftiana Using Double-Templated Molecularly Imprinted Polymers with HPLC–MS/MS. J. Sep. Sci. 2018, 41, 1691–1703. [Google Scholar] [CrossRef] [PubMed]
  208. Winingsih, W.; Ibrahim, S.; Damayanti, S. Purification of Andrographolide Methanolic Extract Using Molecularly Imprinted Polymer Prepared by Precipitation Polymerization. Sci. Pharm. 2022, 90, 27. [Google Scholar] [CrossRef]
  209. Wang, D.D.; Gao, D.; Xu, W.J.; Li, F.; Yin, M.N.; Fu, Q.F.; Xia, Z.N. Magnetic Molecularly Imprinted Polymer for the Selective Extraction of Hesperetin from the Dried Pericarp of Citrus Reticulata Blanco. Talanta 2018, 184, 307–315. [Google Scholar] [CrossRef] [PubMed]
  210. Kang, X.; Deng, L.; Quan, T.; Gao, M.; Zhang, K.; Xia, Z.; Gao, D. Selective Extraction of Quinolizidine Alkaloids from Sophora flavescens Aiton Root Using Tailor-Made Deep Eutectic Solvents and Magnetic Molecularly Imprinted Polymers. Sep. Purif. Technol. 2021, 261, 118282. [Google Scholar] [CrossRef]
  211. He, J.X.; Pan, H.Y.; Xu, L.; Tang, R.Y. Application of Molecularly Imprinted Polymers for the Separation and Detection of Aflatoxin. J. Chem. Res. 2020, 45, 400–410. [Google Scholar] [CrossRef]
  212. Xia, Q.; Yun, Y.; Li, Q.; Huang, Z.; Liang, Z. Preparation and Characterization of Monodisperse Molecularly Imprinted Polymer Microspheres by Precipitation Polymerization for Kaempferol. Des. Monomers Polym. 2016, 20, 201–209. [Google Scholar] [CrossRef] [PubMed]
  213. Bakhtiar, S.; Bhawani, S.A.; Shafqat, S.R. Synthesis and Characterization of Molecular Imprinting Polymer for the Removal of 2-Phenylphenol from Spiked Blood Serum and River Water. Chem. Biol. Technol. Agric. 2019, 6, 1–10. [Google Scholar] [CrossRef]
  214. Zhao, W.R.; Kang, T.F.; Lu, L.P.; Cheng, S.Y. Magnetic Surface Molecularly Imprinted Poly(3-Aminophenylboronic Acid) for Selective Capture and Determination of Diethylstilbestrol. RSC Adv. 2018, 8, 13129–13141. [Google Scholar] [CrossRef]
  215. Dong, C.; Shi, H.; Han, Y.; Yang, Y.; Wang, R.; Men, J. Molecularly Imprinted Polymers by the Surface Imprinting Technique. Eur. Polym. J. 2020, 145, 110231. [Google Scholar] [CrossRef]
  216. Maranata, G.J.; Surya, N.O.; Hasanah, A.N. Optimising Factors Affecting Solid Phase Extraction Performances of Molecular Imprinted Polymer as Recent Sample Preparation Technique. Heliyon 2021, 7, e05934. [Google Scholar] [CrossRef]
  217. Jayasinghe, G.D.T.M.; Moreda-Piñeiro, A. Molecularly Imprinted Polymers for Dispersive (Micro)Solid Phase Extraction: A Review. Separations 2021, 8, 99. [Google Scholar] [CrossRef]
  218. Susanti, I.; Holik, H.A. Review: Application of Magnetic Solid-Phase Extraction (MSPE) in Various Types of Samples. Int. J. Appl. Pharm. 2021, 13, 59–68. [Google Scholar] [CrossRef]
  219. Chen, L.; Wang, X.; Lu, W.; Wu, X.; Li, J. Molecular Imprinting: Perspectives and Applications. Chem. Soc. Rev. 2016, 45, 2137–2211. [Google Scholar] [CrossRef] [PubMed]
  220. Chen, C. Sinapic Acid and Its Derivatives as Medicine in Oxidative Stress-Induced Diseases and Aging. Oxidative Med. Cell. Longev. 2016, 2016, 3571614. [Google Scholar] [CrossRef] [PubMed]
  221. Kanchana, G.; Shyni, W.J.; Rajadurai, M.; Periasamy, R. Evaluation of Antihyperglycemic Effect of Sinapic Acid in Normal and Streptozotocin-Induced Diabetes in Albino Rats. Glob. J. Pharmacol. 2011, 5, 33–39. [Google Scholar]
  222. Maddox, C.E.; Laur, L.M.; Tian, L. Antibacterial Activity of Phenolic Compounds against the Phytopathogen Xylella fastidiosa. Curr. Microbiol. 2010, 60, 53–58. [Google Scholar] [CrossRef] [PubMed]
  223. Zou, Y.; Kim, A.R.; Kim, J.E.; Choi, J.S.; Chung, H.Y. Peroxynitrite Scavenging Activity of Sinapic Acid (3,5-Dimethoxy-4-Hydroxycinnamic Acid) Isolated from Brassica juncea. J. Agric. Food Chem. 2002, 50, 5884–5890. [Google Scholar] [CrossRef]
  224. Hudson, E.A.; Dinh, P.A.; Kokubun, T.; Simmonds, M.S.J.; Gescher, A. Characterization of Potentially Chemopreventive Phenols in Extracts of Brown Rice That Inhibit the Growth of Human Breast and Colon Cancer Cells. Cancer Epidemiol. Biomarkers Prev. 2000, 9, 1163–1170. [Google Scholar]
  225. Shoravi, S.; Olsson, G.D.; Karlsson, B.C.G.; Nicholls, I.A. On the Influence of Crosslinker on Template Complexation in Molecularly Imprinted Polymers: A Computational Study of Prepolymerization Mixture Events with Correlations to Template-Polymer Recognition Behavior and NMR Spectroscopic Studies. Int. J. Mol. Sci. 2010, 15, 10622–10634. [Google Scholar] [CrossRef] [PubMed]
  226. Yu, H.; He, Y.; She, Y.; Wang, M.; Yan, Z.; Ren, J.H.; Cao, Z.; Shao, Y.; Wang, S.; Abd El-Aty, A.M.; et al. Preparation of Molecularly Imprinted Polymers Coupled with High-Performance Liquid Chromatography for the Selective Extraction of Salidroside from Rhodiola Crenulata. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2019, 1118–1119, 180–186. [Google Scholar] [CrossRef] [PubMed]
  227. Ersoy, Ş.K.; Tütem, E.; Başkan, K.S.; Apak, R. Preparation and Application of Caffeic Acid Imprinted Polymer. Turkish J. Chem. 2023, 47, 699–714. [Google Scholar] [CrossRef] [PubMed]
  228. Wan, Y.; Wang, M.; Fu, Q.; Wang, L.; Wang, D.; Zhang, K.; Xia, Z.; Gao, D. Novel Dual Functional Monomers Based Molecularly Imprinted Polymers for Selective Extraction of Myricetin from Herbal Medicines. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1097–1098, 1–9. [Google Scholar] [CrossRef] [PubMed]
  229. Alipour, S.; Azar, P.A.; Husain, S.W.; Rajabi, H.R. Synthesis, Characterization and Application of Spherical and Uniform Molecularly Imprinted Polymeric Nanobeads as Efficient Sorbent for Selective Extraction of Rosmarinic Acid from Plant Matrix. J. Mater. Res. Technol. 2021, 12, 2298–2306. [Google Scholar] [CrossRef]
  230. Cheng, Y.; Nie, J.; Liu, H.; Kuang, L.; Xu, G. Synthesis and Characterization of Magnetic Molecularly Imprinted Polymers for Effective Extraction and Determination of Kaempferol from Apple Samples. J. Chromatogr. A 2020, 1630, 461531. [Google Scholar] [CrossRef] [PubMed]
  231. Ma, X.; Zhang, X.; Lin, H.; Abd El-Aty, A.M.; Rabah, T.; Liu, X.; Yu, Z.; Yong, Y.; Ju, X.; She, Y. Magnetic Molecularly Imprinted Specific Solid-Phase Extraction for Determination of Dihydroquercetin from Larix griffithiana Using HPLC. J. Sep. Sci. 2020, 43, 2301–2310. [Google Scholar] [CrossRef] [PubMed]
  232. Cao, J.; Shen, C.; Wang, X.; Zhu, Y.; Bao, S.; Wu, X.; Fu, Y. A Porous Cellulose-Based Molecular Imprinted Polymer for Specific Recognition and Enrichment of Resveratrol. Carbohydr. Polym. 2021, 251, 117026. [Google Scholar] [CrossRef] [PubMed]
  233. Trivedi, P.P.; Kushwaha, S.; Tripathi, D.N.; Jena, G.B. Cardioprotective Effects of Hesperetin against Doxorubicin-Induced Oxidative Stress and DNA Damage in Rat. Cardiovasc. Toxicol. 2011, 11, 215–225. [Google Scholar] [CrossRef] [PubMed]
  234. Ding, H.W.; Huang, A.L.; Zhang, Y.L.; Li, B.; Huang, C.; Ma, T.T.; Meng, X.M.; Li, J. Design, Synthesis and Biological Evaluation of Hesperetin Derivatives as Potent Anti-Inflammatory Agent. Fitoterapia 2017, 121, 212–222. [Google Scholar] [CrossRef] [PubMed]
  235. Toumi, M.L.; Merzoug, S.; Boutefnouchet, A.; Tahraoui, A.; Ouali, K.; Guellati, M.A. Hesperidin, a Natural Citrus Flavanone, Alleviates Hyperglycaemic State and Attenuates Embryopathies in Pregnant Diabetic Mice. J. Med. Plants Res. 2009, 3, 862–869. [Google Scholar]
  236. Bagheri, A.R.; Arabi, M.; Ghaedi, M.; Ostovan, A.; Wang, X.; Li, J.; Chen, L. Dummy Molecularly Imprinted Polymers Based on a Green Synthesis Strategy for Magnetic Solid-Phase Extraction of Acrylamide in Food Samples. Talanta 2019, 195, 390–400. [Google Scholar] [CrossRef] [PubMed]
  237. Guo, L.; Ma, X.; Xie, X.; Huang, R.; Zhang, M.; Li, J.; Zeng, G.; Fan, Y. Preparation of Dual-Dummy-Template Molecularly Imprinted Polymers Coated Magnetic Graphene Oxide for Separation and Enrichment of Phthalate Esters in Water. Chem. Eng. J. 2018, 361, 245–255. [Google Scholar] [CrossRef]
  238. Golfakhrabadi, F.; Khaledi, M.; Nazemi, M.; Safdarian, M. Isolation, Identification, and HPTLC Quantification of Dehydrodeoxycholic Acid from Persian Gulf Sponges. J. Pharm. Biomed. Anal. 2021, 197, 113962. [Google Scholar] [CrossRef] [PubMed]
  239. Jug, U.; Vovk, I.; Glavnik, V.; Makuc, D.; Naumoska, K. Off-Line Multidimensional High Performance Thin-Layer Chromatography for Fractionation of Japanese Knotweed Rhizome Bark Extract and Isolation of Flavan-3-Ols, Proanthocyanidins and Anthraquinones. J. Chromatogr. A 2021, 1637, 461802. [Google Scholar] [CrossRef]
  240. Pandya, D.J.; Anand, I.S. Isolation and High-Performance Thin Layer Chromatographic Estimation of Lupeol from Oxystelma esculentum. Pharm. Methods 2011, 2, 99–105. [Google Scholar] [CrossRef]
Figure 1. Mechanism action of phenolic compounds as an antioxidant agent.
Figure 1. Mechanism action of phenolic compounds as an antioxidant agent.
Plants 13 00965 g001
Figure 2. Classification of phenolic compounds: (1) simple phenolic compound, (2) flavonoid, (3) tannins, (4) stilbenes, and (5) lignans.
Figure 2. Classification of phenolic compounds: (1) simple phenolic compound, (2) flavonoid, (3) tannins, (4) stilbenes, and (5) lignans.
Plants 13 00965 g002
Figure 3. Counter-current chromatography (CCC) systems.
Figure 3. Counter-current chromatography (CCC) systems.
Plants 13 00965 g003
Figure 4. Schemes of separation using (a) solid-phase extraction (SPE) and (b) dispersive solid-phase extraction (DSPE).
Figure 4. Schemes of separation using (a) solid-phase extraction (SPE) and (b) dispersive solid-phase extraction (DSPE).
Plants 13 00965 g004
Figure 5. The advantages and disadvantages of isolation methods with new improvement from 2017 to 2023 to isolate antioxidant phenolic compounds.
Figure 5. The advantages and disadvantages of isolation methods with new improvement from 2017 to 2023 to isolate antioxidant phenolic compounds.
Plants 13 00965 g005
Table 1. The separation and purification methods used to isolate bergenin.
Table 1. The separation and purification methods used to isolate bergenin.
SampleSeparation and Purification MethodYield (%) *Purity (%)Ref.
Saxifraga atrataPolyamide column coupled with MCI GEL® CHP20P in MPLC1.9>99[39]
Flueggea virosa leavesvacuum liquid chromatography column2NM[86]
Securinega virosaSilica column chromatography and Sephadex LH gel filtration chromatography 0.043NM[87]
Peltophorum pterocarpumCrystallization 1NM[88]
NM, not mentioned in the article, * Yield (%): ratio of isolate mass and crude extract mass.
Table 2. Application of MPLC to isolate active compounds from natural products.
Table 2. Application of MPLC to isolate active compounds from natural products.
SampleCompoundSeparation Method Yield (%) * Purity (%)Ref.
Saxifraga atrataBergeninMPLC1.9>99[39]
Saxifraga atrataEthyl gallate MPLC-HPLC0.013>95[89]
11-O-Galloylbergenin0.031
Rutin1.12
Isoquercitrin0.176
Saxifraga sinomontan3-methoxy-4-hydroxyphenol-(60-O-galloyl)-1-O-β-D-glucopyranosideMPLC-RP-HPLC0.6>95[90]
3,4,5-trimethoxyphenyl-(60-O-galloyl)-1-O-β-D-glucopyranoside1.39
Saximonsin A1.40
Saximonsin B0.24
MPLC, medium-pressure liquid chromatography; RP-HPLC, reverse-phase high-performance liquid chromatography, * Yield (%): ratio of isolate mass and crude extract mass.
Table 4. Studies that have used high-speed counter-current chromatography (HSCCC) and high-performance counter-current chromatography (HPCCC) to isolate or purify phenolic compounds.
Table 4. Studies that have used high-speed counter-current chromatography (HSCCC) and high-performance counter-current chromatography (HPCCC) to isolate or purify phenolic compounds.
SampleCompoundInstrumentElution SystemYield (%) *Purity (%)Ref.
Entada phaseoloidesPhaseoloidinHSCCCn-Butanol:acetic acid:water, 4:1:5 (v/v)7.76%99.3%[136]
Entadamide A6.97%96.4%
Entadamide A-β-D-glucopyranoside6.79%97.7%
Malus hupehensisAvicularinHSCCCDES: choline chloride/glucose:water:ethyl acetate, 1:1:2 (v/v)NM93.1%[41]
Phloridzin94.5%
Sieboldin93.6%
Sweet orange peel extractSinensetinHPCCCHexanes:ethyl acetate:methanol:water, 1.4:0.6:1.4:0.6 (v/v), for the normal phase
Hexanes:ethyl acetate:methanol:water, 0.7:1.3:0.7:1.3 (v/v), for the reverse phase
1.08%100%[42]
3,5,6,7,3′,4′-hexamethoxyflavone1.17%100%
Nobiletin14.25%99.1%
5,6,7,4′-tetramethoxyflavone2.17%96.6%
3,5,6,7,8,3′,4′-heptamethoxyflavone10.08%98.4%
Tangeretin2.75%97.6%
(+)-catechin-phloroglucinol derivative; 76%
(-)-epicatechin93%
(+)-catechin77%
(-)-epicatechin-3-O-gallate85%
(-)-epicatechin-3-O-galloyl-phloroglucinol derivative95%
Parastrephia lucida
(Meyen)
11- p-coumaroyloxyltremetoneHSCCCN-hexane:
ethyl acetate:methanol:water (6:5:6:3 v/v/v)
11%NM[137]
Mango flowersGallic acidHSCCCN-hexane-ethyl acetate-methanol-water (4:6:4:6, v/v) for normal phase1.85%98.87%[138]
Ethyl gallate1.95%99.55%
Ellagic aciddichloromethane-methanol-water (4:3:2, v/v) elution-extrusion mode2.85%99.71%
Achyrocline satureioides (Lam) D.C.QuercetinHPCCC semi-preparative HPLCN-hexane:ethyl acetate:methanol:water
(0.8:1.0:0.8:1.0) and dichloromethane:methanol:water (3.5:3.5:2.5)
60%97.5%[139]
Luteolin90.2%
3-O-methylquercetin65%97.0%
Peanut HullLuteolinHPCCCN-hexane:ethyl acetate:methanol:water (1.0:1.0:1.0:1.5)1.5%96%[140]
Eriodictyol0.8%
5,7-dihydroxychromone0.3%99%
Roots of Polygonum multiflorum ThunbGallic acidHSCCC and preparative HPLCPetroleum ether:ethyl acetate:methanol:water (1:5:1:5)
Preparative HPLC: using methanol/water
NM98.28%[141]
Epicatechin96.71%
Piceatannol96.85%
Rutin97.92%
Resveratrol96.94%
Hyperoside98.52%
Roots of Polygonum multiflorum ThunbCatechinHSCCC Petroleum ether:ethyl acetate:methanol:water (1:5:1:5)NM90.69%[141]
Polydatin94.91%
2,3,5,4′
-tetrahydroxy
stilbene-2-O-β-D-glucoside
95.23%
Leaves of Lonicera japonica Thunb.RhoifolinHSCCCCMethyl tert-butyl ether:n-butanol:acetonitrile:water (0.5% acetic acid) (2:2:1:5, v/v)2.15%94.3%[142]
Luteoloside3.19%96.1%
Chlorogenic acidHSCCC and preparative HPLCHSCCC system:methyl tert-butyl ether:n-butanol:acetonitrile:water (0.5% acetic acid) (2:2:1:5, v/v)
Preparative HPLC system: C-18 (15 μm) column as stationary phase and solution of eluent A (methanol) and eluent B (0.3%, v/v, acetic acid in water) as mobile phase
1.09%99.5%
Lonicerin3.07%98.7%
Rutin1.67%99.3%
3,4-O-dicaffeoylquinic acid2.03%97.1%
Hyperoside1.82%97.4%
3,5-O-Dicaffeoylquinic acid2.47%96.9%
4,5-O-Dicaffeoylquinic acid2.61%97.8%
PersimmonGallic acidHSCCCN-hexane:ethyl acetate:water (3:17:20, v/v/v) and ethyl acetate:methanol:water (50:1:50, v/v/v)3.13%>95%[143]
Methyl gallate29.47%
Epigallocatechin-3-gallate-(4β → 8, 2β → O → 7)-epigallocatechin-3-gallate dimer3.93%
Salvia MiltiorrhizaRutinHSCCCtert-butyl methyl ether/n-butanol/acetonitrile/water (3:1:1:20, v/v)0.14%97.3%[144]
Isoquercitrin0.17%99.5%
Mahonia bealei (Fort.) Carr. LeavesChlorogenic acidHSCCCn-hexane/ethyl acetate/methanol/water (1:5:1:5, v/v/v/v)NM>92%[145]
Quercetin-3-O-β-D-glucopyranoside
Isorhamnetin-3-O-β-D-glucopyranoside
Castanopsis chinensis HanceChinensin DCombined multi step CC and HSCCCN-Hexane/Ethyl acetate/Methanol/Water (1:6:3:4, v/v/v/v)NM93%[134]
chinensin E95.7%
Chrysanthemum morifolium cv. FubaijuLuteolin-7-O-β-D-glucosideHSCCC combined with preparative HPLCEthyl acetate-n-butanol–acetonitrile–water–acetic acid (5:0.5:2.5:5:0.25,
v/v/v/v/v)
NM97.1%[146]
Luteolin-7-O-β-Dglucuronide97.8%
Apigenin-7-O-β-D-glucoside95.8%
luteolin-
7-O-β-D-rutinoside
96.7%
3,5-dicaffeoylquinic
acid
97.8%
4,5-dicaffeoylquinic acid97.5%
NM, not mentioned in the article. * Yield (%): ratio of isolate mass and crude extract mass.
Table 5. A list of studies that used the HILIC.
Table 5. A list of studies that used the HILIC.
Sample CompoundMode Separation Purity Yield(%) *Ref.
Saxifraga tanguticaHyperoside2D HILIC/RPLC>95%NM [154]
Luteoline-glucoside
Trifolin
Salvia prattiiCaffeic Acid 2D HILIC/RPLC>98% 1.39[155]
Ethyl Rosmarinate 1.49
Methyl Rosmarinate 1.09
Rosmarinic acid 8.90
Dracocephalum heterophyllumcaffeoyl-β-D glucopyranoside2D RP/HILIC technique guided by on-line HPLC-DPPH>95%0.011[153]
ferruginoside B0.014
verbascoside0.083
2′-O-acetylplantamajoside0.039
sibiricin A0.026
luteolin0.0106
rosmarinic acid0.108
methyl rosmarinate0.017
Arenaria
kansuensis
Tricin2D RP/HILIC technique guided by on-line HPLC-DPPH>98%NM[156]
Homoeriodictyol
Luteolin
Lycium ruthenicum Murr.Anthocyanin 2D RP/HILICNMNM[156]
NM, not mentioned in the article. * Yield (%): ratio of isolate mass and crude extract mass.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Susanti, I.; Pratiwi, R.; Rosandi, Y.; Hasanah, A.N. Separation Methods of Phenolic Compounds from Plant Extract as Antioxidant Agents Candidate. Plants 2024, 13, 965. https://doi.org/10.3390/plants13070965

AMA Style

Susanti I, Pratiwi R, Rosandi Y, Hasanah AN. Separation Methods of Phenolic Compounds from Plant Extract as Antioxidant Agents Candidate. Plants. 2024; 13(7):965. https://doi.org/10.3390/plants13070965

Chicago/Turabian Style

Susanti, Ike, Rimadani Pratiwi, Yudi Rosandi, and Aliya Nur Hasanah. 2024. "Separation Methods of Phenolic Compounds from Plant Extract as Antioxidant Agents Candidate" Plants 13, no. 7: 965. https://doi.org/10.3390/plants13070965

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

Article Metrics

Back to TopTop