Next Article in Journal
Exploring Hydrogen Sources in Catalytic Transfer Hydrogenation: A Review of Unsaturated Compound Reduction
Next Article in Special Issue
Metabolic Transformation of Gentiopicrin, a Liver Protective Active Ingredient, Based on Intestinal Bacteria
Previous Article in Journal
Subcritical Water Extraction to Valorize Grape Biomass—A Step Closer to Circular Economy
Previous Article in Special Issue
Study on Purification, Identification and Antioxidant of Flavonoids Extracted from Perilla leaves
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phenylboronic Acid and Amino Bifunctional Modified Adsorbent for Quickly Separating Phenolic Acids from Crude Extract of Clerodendranthus spicatus and Evaluation of Their Antioxidant and Hypoglycemic Activities

Tianjin Key Laboratory for Modern Drug Delivery and High-Efficiency, Collaborative Innovation Center of Chemical Science and Engineering, School of Pharmaceutical Science and Technology, Tianjin University, Tianjin 300072, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(22), 7539; https://doi.org/10.3390/molecules28227539
Submission received: 10 October 2023 / Revised: 7 November 2023 / Accepted: 9 November 2023 / Published: 11 November 2023

Abstract

:
A novel phenylboronic acid and amino bifunctional modified silica gel (SiO2-NH2-FPBA) was prepared, which was 30–80 μm, had a pore size of 8.69 nm, a specific surface area of 206.89 m2/g, was stable at low temperature, and contained 0.4793 mmol/g of the phenylboronic acid group and 1.6377 mmol/g of the amino group. It was used to develop a rapid separation method for phenolic acids. The results showed that it could adsorb 93.64 mg/g caffeic acid, 89.35 mg/g protocatechuic acid and 79.66 mg/g gallic acid. The adsorption process was consistent with the pseudo-second-order model (R2 > 0.99), and fitted the Langmuir isotherm model well (R2 > 0.99). CH3COOH could effectively desorb phenolic acids (>90%) and did not destroy their structures. When SiO2-NH2-FPBA was added to crude extract of Clerodendranthus spicatus, 93.24% of the phenolic acids could be captured, and twenty-two kinds of phenolic acids were identified by Q Exactive HF LC-MS. Furthermore, the isolated phenolic acids from Clerodendranthus spicatus possessed great DPPH, ABTS, and hydroxyl radicals scavenging activities and ferric reducing power. They also demonstrated effective inhibition of α-amylase and α-glucosidase activities (IC50 = 110.63 ± 3.67 μg/mL and 64.76 ± 0.30 μg/mL, respectively). The findings indicate that SiO2-NH2-FPBA has significant potential in practical applications of separating active constituents from natural resources.

1. Introduction

Phenolic acids are widely distributed in medicinal plants and are secondary metabolites produced by the plants themselves, accounting for one-third of the polyphenolic compounds [1]. Phenolic acid compounds are mainly composed of two carbon skeletons, respectively—C6-C1 benzoic acid type (e.g., gallic acid) and C6-C3 cinnamic acid type (e.g., caffeic acid). The different positions and number of hydroxyl groups on the aromatic ring and the various substituents on the carbon positions of the aromatic ring allow phenolic acid to form different kinds and structures of compounds. Phenolic acid substances have strong biological activities, including antitumor, antioxidant, anti-inflammatory, antibacterial and antiviral activity [2,3,4], and have been extensively studied in various fields, such as agriculture, food science, medicine, and environmental science. Taking into account further research into the pharmacological activities, sources (other traditional Chinese medicines) and applications of phenolic acids, developing an effective method for the fast separation of phenolic acids is becoming an important issue.
In the past few decades, many methods have been developed for the separation of phenolic acids, such as precipitation [5], membrane separation [6], extraction [7], chromatography [8,9], solid phase adsorption [10] and resin [11]. The most commonly used extraction method is solvent extraction to obtain the crude extracts, followed by separation of the extract using chromatographic techniques through a resin adsorption column [12]. For example, Sun et al. [13] chose Sephadex LH-20 column chromatography and reverse phase C18 gel column chromatography and Zhou et al. [14] used HW-40C and LH-20 column chromatography for the separation and collection of phenolic acids. The resin adsorption method has the advantages of being a simple process, less energy consumption and green efficiency. However, this technique has presented a number of issues such as excessive loss of solvents, inadequate separation efficiency, residual solvents, and operational hazards. Furthermore, it was limited in small-scale sample preparation and was unsuitable for large-scale industrial applications [12].
Remarkably, the adsorption method has gained a lot of attention due to its high efficiency, low cost, ease of operation, and reproducibility [15]. Various types of adsorbents have been used for phenolic acids’ adsorption. For example, Cagnon [16] investigated the adsorption of gallic acid on an activated carbon oxidized by ozone. Li [17] and Fan [18] utilized molecularly imprinted materials for the isolation and purification of caffeic acid from mushrooms and fruits. Song [19] and Chai [20] prepared magnetic chitosan-modified diatomite to adsorb gallic acid and caffeic acid from a sugar solution. Nie [21] and Du [12] synthesized ionic liquid silica gel for the adsorption and separation of water-soluble phenolic acids. Moritz et al. [22] synthesized mesoporous SBA-15 and MCF silica functionalized with aminopropyltriethoxysilane (APTES) and 3-[2-(aminoethylamino) propyl] trimethoxy silane (AEAPTMS) and explored their adsorption effect on caffeic acid. Simanaviciute et al. [10] prepared cationic cross-linked starch for adsorbing caffeic, chlorogenic and rosmarinic acids. However, most of these adsorption methods rely on electrostatic attraction to exert the adsorption effect. Therefore, considering the chemical properties of phenolic acids, which have a specific structure of cis-diols that can react with boronic acids in a specific way, boronic-acid-modified adsorbents are considered ideal for preparation. For instance, Qian [23] used 4-(acryloyloxy)phenylboronic acid as a monomer to prepare a surface molecularly imprinted magnetic nanoparticle for recovering salvianolic acid. Su [24] prepared polyethyleneimine intercalated montmorillonite with phenylboronic acid groups for capturing salvianolic acid A, salvianolic acid B and rosmarinic acid. It is worth noting that the formation of borate is reversible and is affected by the pH of the solution based on the reaction mechanism of boric acid and cis-diol [23]. It is stable under weakly basic conditions (pH ≥ 8.5). Nevertheless, phenolic acids can make the solution acidic after ionization, which hinders the combination of boric acid and phenolic acids. Therefore, the addition of basic groups is particularly important, which can neutralize the acidity of phenolic acids and promote the formation of borate.
Clerodendranthus spicatus (C. spicatus), commonly known as “Shen Cha” in Chinese, is a perennial herb that grows widely in tropical and subtropical regions [14,25,26]. It has been used to treat diabetes and kidney diseases, such as urinary lithiasis, chronic nephritis, urinary tract infection, nephrotic syndrome, and chronic renal failure with a long history. More than 150 kinds of compounds have been isolated from C. spicatus, and the main chemical substances are phenolic acids, diterpenoids, triterpenoids, and their derivatives [27]. Over 60 kinds of phenolic acids and their derivatives have been separated by traditional methods, including protocatechuic acid [28], caffeic acid [28], and their derivatives, among others [13]. However, there is no information about the rapid separation of phenolic acids from crude extract of C. spicatus by special adsorbents.
In this study, we synthesized a novel phenylboronic acid and amino bifunctional modified silica gel (SiO2-NH2-FPBA) using surface modification methods. It was systematically characterized by fourier transform infrared spectroscopy (FT-IR), scanning electron microscopy (SEM), energy dispersive spectrometry (EDS), transmission electron microscopy (TEM), thermogravimetric analysis (TGA), the Brunauer–Emmett–Teller (BET) method, and inductively coupled plasma-optical emission spectrometry (ICP-OES). Three common phenolic acids—gallic acid (GA), caffeic acid (CA) and protocatechuic acid (PCA)—were selected as the model target, and the effects of different adsorption conditions were investigated and optimized. On the basis of these, the method was applied for the first time for one-step and rapid separation of phenolic acids from C. spicatus crude extract. Then, the antioxidant activity and hypoglycemic activity of the isolated phenolic acids from C. spicatus were determined by 1,1-diphenyl-2-picrylhydrazyl (DPPH), 2,2′-azinobis-(3-ethylbenthiazoline-6-sulfonate) (ABTS), hydroxyl radicals scavenging assays, ferric-reducing antioxidant activity, and α-amylase and α-glycosidase inhibition effects in vitro. This study provided a new and rapid method for the separation of phenolic acids from traditional Chinese medicine.

2. Results and Discussion

2.1. Characterization

FT-IR was used to characterize the presence of the corresponding functional groups on the SiO2-NH2-FPBA surface (Figure 1a). The peaks at 3450 cm−1, 1090 cm−1 and 800 cm−1 were the characteristic absorptions formed by the stretching vibration of -OH, asymmetric stretching vibration of Si-O-Si and bending bands of Si-OH, respectively [29]. The peaks at 2945 cm−1 and 2879 cm−1 were attributed to the asymmetric and symmetric stretching vibration of methylene [30,31], indicating that the amino groups were successfully attached and providing the amino sites for the next reaction [32]. The absorption peak at 1425 cm−1 indicated the skeleton vibration of the benzene ring and 1357 cm−1 was formed by B-O stretching vibration, which indicated that the phenylboronic acid had been successfully attached.
The surface was elementally analyzed by EDS and is shown in Figure 1b, containing 36.91% Si element, 39.09% O element, 3.94% N element and 20.05% C element. The response of the B element was low using this method and could not be clearly shown. The morphology and structure of the SiO2-NH2-FPBA were observed by SEM and TEM (Figure 1c,d). It could be seen that the material showed an irregular shape and the particle size was 30–80 μm. When enlarged, it could be found that the surface was rough and uneven with pores and concavities, which effectively provided a large number of adsorption sites and facilitated the adsorption. The light and dark areas in Figure 1e indicate the existence of mesopores in the SiO2-NH2-FPBA.
The adsorption and desorption isotherms of N2 and the pore size distribution curve of SiO2-NH2-FPBA are shown in Figure 1f. The N2 adsorption of SiO2-NH2-FPBA was a typical IV isotherm [33,34]. Its hysteresis loop appeared between 0.40 and 0.90, which belonged to the H1 type with disordered pores [33]. The pore size of 8.69 nm indicated SiO2-NH2-FPBA with mesopores [31]. The BET specific surface area of SiO2-NH2-FPBA calculated from the nitrogen adsorption analysis was 206.89 m2/g.
The TGA curve of SiO2-NH2-FPBA is shown in Figure 1g. A small weight loss of 0.94% for the material was found due to the evaporation of water [35]. At slightly greater than the boiling point of phenylboronic acid (265.9 °C), a weight loss of 3.16% at 286.51 °C was produced, presumably due to phenylboronic acid degradation. A weight loss of 5.54% in the range of 286.51–443.02 °C and 7.68% loss when temperature was up to 600 °C were observed, which may be attributed to the decomposition of the amino-propyl group on the surface and the dehydration condensation of the -OH on the silica gel surface [36]. The SiO2-NH2-FPBA was thermostable at a low temperature.
ICP-OES data showed that the amount of B in SiO2-NH2-FPBA was 0.4793 mmol/g, which indicated that the B element was attached to the material surface. The content of the residual amino group in SiO2-NH2-FPBA reduced from 2.1170 mmol/g in SiO2-NH2 to 1.6377 mmol/g. The ratio of phenylboronic acid group to amino group was about 3.4:1.

2.2. Evaluation of the Adsorption Ability of SiO2-NH2-FPBA to GA, CA and PCA

2.2.1. Effect of pH on Absorption

To explore the effect of pH, the GA, CA and PCA solutions were adjusted to 3, 4, 5, 6, 7, and 8 using HCl or NaOH. SiO2-NH2-FPBA (20 mg) was added into 8 mL of 0.5 mg/mL solution of GA, CA and PCA at different pH values. Batch adsorption experiments were conducted in a temperature-controlled water bath oscillator at 200 rpm at 20 °C for 4 h. At the end of the adsorption, the solid was separated from the liquid by centrifugation at 4100 rpm for 3 min. The concentration of each solution in the supernatant was measured by UV, and the results are shown in Figure 2a. It was found that the adsorption effect was best at pH 3, with 93.64 mg/g for CA, 89.35 mg/g for PCA and 79.66 mg/g for GA. When the pH was increased to 4, the adsorption amount changed slightly. The adsorption produced a significant decreasing trend when the pH was increased from 4 to 8. As is known, the pKa values of GA, CA and PCA were 4.41, 4.58 and 4.48, respectively, which existed in ionic form and formed sodium phenol at high pH. The instability of the three phenolic acids at higher pH seemed to be the reason for their significantly lower adsorption [19,24]. The pH values of the aqueous solutions of the three phenolic acids at 0.5 mg/mL were determined to be 3.39, 3.43 and 3.44, respectively, which were between the optimum pH of 3 and 4. Therefore, aqueous solutions of phenolic acids were used for adsorption in the next experiments.

2.2.2. Effect of Temperature on Absorption

Adsorption temperature may affect the adsorption performance of phenolic acid substances. To explore the effect of temperature, 20 mg of SiO2-NH2-FPBA was added into 8 mL of 0.5 mg/mL of aqueous solution of GA, CA, or PCA. The centrifuge tubes were placed in a water bath with different temperatures (20, 30, 40, 50, and 60 °C). From Figure 2b, it can be seen that with the increase in adsorption temperature from 20 °C to 60 °C, their adsorption amount decreased from 99.75 mg/g to 67.78 mg/g, 85.48 mg/g to 62.19 mg/g and 79.66 to 61.36 mg/g, respectively, indicating that the increase in temperature was unfavorable for the adsorption process. Thus, all the following experiments were conducted at 20 °C.

2.2.3. Effect of Time on Absorption

Adsorption kinetics can provide valuable information on solute uptake rate and additional details on the overall adsorption process [37]. To explore the adsorption kinetics, 20 mg of SiO2-NH2-FPBA was added into 20 mL of 0.5 mg/mL aqueous solution of phenolic acids at 20 °C. The supernatant was sampled at 5, 15, 30, 60, 120, 180, 240, 300, 360, and 420 min and the adsorption amount was measured. The results shown in Figure 3a indicate that the adsorption started at a fast rate, then the adsorption rate gradually slowed down. With the extension of adsorption time, the adsorption capacity kept increasing. From 0 to 120 min, the adsorption amount gradually increased over time due to the presence of a large number of adsorption sites for adsorption in the initial stage. At the 120–420 min stage, the adsorption gradually tended toward equilibrium and reached saturation with the adsorption amounts of 92.58 mg/g for CA, 85.54 mg/g for PCA and 76.25 mg/g for GA.
To further analyze the adsorption of SiO2-NH2-FPBA to phenolic acids, the pseudo-first-order and pseudo-second-order kinetic models were exploited to fit the experimental data. The equations of the two models were as follows:
The pseudo-first-order equation was
l n Q e Q t = l n Q e K 1 t
The pseudo-second-order equation was
t Q t = t Q e + 1 K 2 × Q e 2
where K1 (min−1) and K2 (g(mg/min)) are the pseudo-first-order rate constant and pseudo-second-order rate. Qt (mg/g) and Qe (mg/g) represent the adsorption amount for each phenolic acid at t time and equilibrium, respectively.
The fitting curves are shown in Figure 3b,c and the constants are summarized in Table 1. It was found that the pseudo-second-order model had better agreement; all four R2 values were above 0.999. The adsorption processes of the three phenolic acids could be well represented by pseudo-second-order kinetics, which demonstrated that there was a rate-limiting step in the adsorption process. This was a chemical adsorption process.

2.2.4. Effect of Concentration on Absorption

The adsorption isotherm can be used to not only assess the adsorption capacity of adsorbents, but also can describe how the adsorbate interacts with the adsorbent [38]. Thus, the effect of concentration on adsorption was investigated. Each of the three solutions was prepared as 0.1, 0.25, 0.5, 1, 2.5, and 5 mg/mL. Then, 20 mg of each adsorbent was added into 8 mL of different solutions at 20 °C for 2 h. To understand the relationship between Qe and Ce under equilibrium conditions, the adsorption isotherms were analyzed by obtaining the best fits to the data using the extensively used Langmuir and Freundlich models. The Langmuir model assumes a monolayer adsorption, which means that adsorption molecules adsorbed on the surface of the adsorbent have the same adsorption activation energy. It assumes that the adsorbate and adsorbent are in an ideal state and the model is deployed for homogenous surfaces. The Freundlich model assumes a multilayer adsorption, which means there are many adsorption sites on the adsorbent, and the adsorption sites have different free energy values, which can adsorb multiple molecules. This model is applicable in the study of adsorption on rough and multisite (heterogonous) surfaces. These two models can be expressed using Equations (3) and (4):
C e Q e = 1 b Q 0 + C e Q 0
ln Q e = ln K f + 1 n ln C e
where Qe is the adsorption capacity at equilibrium (mg/g), Ce is the concentration of the solution at equilibrium (mg/mL), Q0 is the saturation adsorption capacity (mg/g), b represents the equilibrium adsorption constant and the adsorption affinity, n is the Freundlich constant, and Kf is the binding constant.
From Figure 4a, it is obvious that the adsorption amount increased with the increase in concentration in the initial stage. However, with the further increase in concentration, the adsorption amount gradually stabilized, indicating the adsorption reached the equilibrium. The fitting curves and the parameters obtained from fitting two isotherm models with the experimental isotherm data are shown in Figure 4b,c and Table 2. The results indicated that the adsorption data better fit the Langmuir isotherm (R2 = 0.99), which means that the uptake of the acids was dependent on the monolayer adsorption model.

2.2.5. Exploration of Desorption Conditions

Four pH 2 acidic solutions of hydrochloric acid, formic acid, acetic acid, and benzoic acid were selected to investigate the desorption effect and the desorption rate was chosen as the evaluation criterion. After 20 mg of adsorbent was added to the 8 mL mixture solution (0.5 mg/mL) for 2 h, the solid and liquid were separated and the solid was washed 3 times by adding 10 mL of water. Then, 8 mL each of different desorption solutions was added and shaken for 2 h. The HPLC method was used to detect the concentrations of various phenolic acids in the supernatant, and the desorption rate was calculated. As the results show in Figure 5a, the desorption rate was in the following order: CH3COOH, HCOOH, HCl, and benzoic acid. Among them, CH3COOH desorbed more than 90% of the phenolic acids and was used as the desorption solution.
The three phenolic acids were also analyzed using HPLC-MS before adsorption, after adsorption, and after desorption. The HPLC chromatograms in Figure 5b show that the peak areas of all the substances decreased after adsorption, indicating the presence of good adsorption. The total ion chromatography (TIC) of the MS in Figure 5c shows consistent peak times for the three substances before adsorption, after adsorption and desorption, indicating that all three substances retained their original structures. The extracted ion chromatograms (EIC) with m/z 125, m/z 109 and m/z 135 in Figure 5d were extracted, respectively, corresponding to the retention times on HPLC chromatograms, which indicated that all three substances were effectively desorbed. The TIC and EIC images obtained by MS analysis proved that the desorption components still kept their structures.

2.3. Comparison of SiO2-NH2-FPBA with the Reported Adsorbents

The adsorption capacities of SiO2-NH2-FPBA to phenolic acids were compared with amino functionalized silica gel (SiO2-NH2). The results are shown in Table 3. Under the same adsorption conditions, the adsorption capacities of SiO2-NH2 to CA, PCA, GA were apparently lower than those of SiO2-NH2-FPBA. This indicated that both the amino group and phenylboronic acid group contributed to the absorption of phenolic acids. They were also compared with other reported adsorbents (see Table 3). The results showed that SiO2-NH2-FPBA had faster adsorption kinetics and higher adsorption capacity than these reported adsorbents, which indicated its high potential for capturing phenolic acids.

2.4. Evaluation of the Adsorption of SiO2-NH2-FPBA to Phenolic Acids in C. spicatus Crude Extract

2.4.1. Effect of pH on Absorption

To explore the effect of pH, C. spicatus solutions were adjusted to 3, 4, 5, 6, 7, and 8 using HCl or NaOH, respectively. SiO2-NH2-FPBA absorbents of 50 mg were added into 4 mL of 2.5 mg/mL of each solution at different pH values. Batch adsorption experiments were conducted in a temperature-controlled water bath oscillator at 200 rpm at 20 °C for 4 h. At the end of the adsorption, the supernatant was filtered using a 0.45 μm filter membrane and the total phenol content was determined according to the Folin–Ciocalteu method. The adsorption effect at different pH values is shown in Figure 6a. It could be observed that the adsorption effect showed an increasing trend with increasing pH. The adsorption increased when the pH increased from 3 to 4, and reached a maximum value of 11.85 mg/g at pH 4. Then, as the pH increased from 4 to 8, the adsorption showed a decreasing trend. The pH of the aqueous solution of C. spicatus crude extract was 6.28. The adsorption amount at pH 6 was about 11.63 mg/g, which was only slightly different from the maximum value. Due to the complex phenolic acid composition in the crude extracts of C. spicatus, in order to better ensure its structure, all of the following adsorptions were performed using aqueous solutions of C. spicatus crude extracts.

2.4.2. Effect of Temperature on Absorption

To explore the effect of adsorption temperature, 50 mg of SiO2-NH2-FPBA was added into 4 mL of crude extracts of aqueous C. spicatus (2.5 mg/mL). The centrifuge tubes were transferred into a water bath at different temperatures (20, 30, 40, 50, or 60 °C). After adsorption, the supernatant samples were filtered and measured, and the results are shown in Figure 6b. The results showed a significant decrease in adsorption with increasing temperature, which was presumed to be due to the fact that high temperature decomposed some of the substances in the crude extracts solution, thus reducing the adsorption effect. Finally, 20 °C was chosen for adsorption.

2.4.3. Effect of Time on Absorption

To explore the adsorption kinetics, 50 mg of SiO2-NH2-FPBA was added into 8 mL of 2.5 mg/mL of each solution at 20 °C. The supernatant was sampled at 5, 15, 30, 60, 120, 180, 240, and 360 min. The effect of time on adsorption was investigated by detecting the concentration in the supernatant at different adsorption times, and the results are shown in Figure 6c. It could be seen that the adsorption amount tended to increase with time from 0 to 120 min, but as the time increased up to 360 min, the adsorption amount stabilized and reached 11.92 mg/g, indicating that the adsorption reached the maximum amount. Therefore, the optimal adsorption time was set to 2 h.
Pseudo-first-order and pseudo-second-order kinetic models were utilized to fit the adsorption process and fitting curves are shown in Figure S1a,b. It was found that the pseudo-second-order model had a better agreement (R2 = 0.99), which showed the process was a chemical adsorption.

2.4.4. Effect of Concentration on Absorption

To explore the adsorption equilibrium, the solution was prepared as 0.5, 1, 2.5, 5, 7.5, 10, 15, and 20 mg/mL. Adsorbents of 50 mg were added into 4 mL of different concentrations of solution at 20 °C for 2 h, and the results are shown in Figure 6d. We determined that the adsorption amount tended to increase gradually with the increase in initial concentration. However, when the initial concentration increased from 15 mg/mL to 20 mg/mL, the adsorption amount stabilized, indicating that the adsorption reached saturation at this concentration.
The Langmuir and Freundlich models were adopted to describe the adsorption of SiO2-NH2-FPBA to phenolic acid substances. The fitting curves obtained from fitting two isotherm models with the experimental isotherm data are shown in Figure S2a,b. The results indicated the adsorption data better fit the Langmuir isotherm (R2 = 0.99) and the maximum theoretical adsorption capacity was 34.97 mg/g, which meant that the uptake of the phenolic acid substances was dependent on the monolayer adsorption model.

2.4.5. Desorption Conditions

After adsorption of 50 mg of SiO2-NH2-FPBA to 4 mL of mixture solution (2.5 mg/mL) for 2 h, the solid and liquid were separated and the solid was washed 3 times by adding 10 mL of water. Then, 10 mL each of different desorption solutions—HCl, HCOOH, CH3COOH, and benzoic acid—was added and shaken for 2 h to calculate the desorption rate. The desorption rates were 72.60%, 56.43%, 42.21%, and 9.31% in the order of CH3COOH, HCOOH, HCl and benzoic acid, which showed the same trend as the previous desorption of phenolic acid simulants. However, the desorption rate was not as high as that of the previous. It was speculated that due to the complexity of the phenolic acid species in the crude extracts of C. spicatus, it was not guaranteed that each phenolic acid substance could be desorbed under the current conditions. A high desorption rate (>90%) needed more desorption solution (>20 mL).
A solution of 10 mg/mL of C. spicatus crude extracts was prepared. Then, 200 mg of SiO2-NH2-FPBA was added to 2.5 mL of C. spicatus solution (10 mg/mL) and the adsorption was shaken at room temperature for 2 h. The solutions before and after adsorption were determined using the Folin–Ciocalteu and HPLC methods. After adsorption, the removal rate was 93.24%, indicating that the vast majority of phenolic acids were adsorbed by SiO2-NH2-FPBA. As can be seen from Figure 7a, the color of C. spicatus before adsorption was brown and the supernatant after adsorption was almost transparent, indicating that SiO2-NH2-FPBA had a good adsorption effect on phenolic acids in the crude extracts of C. spicatus.
The solutions before and after adsorption were analyzed by using the HPLC method, and the results are shown in Figure 7b. It could be seen that several peak areas were significantly reduced, which were adsorbed by SiO2-NH2-FPBA. The peaks at 10.283 min, 15.933 min, and 18.394 min disappeared, and the peaks at 11.947 min, 13.133 min, 32.474 min, and 42.952 min were significantly reduced, indicating that the phenolic acid substances in the crude extracts of C. spicatus were successfully isolated by SiO2-NH2-FPBA.

2.5. Identification of the Separated Phenolic Acids by Q Exactive HF LC-MS

The compounds separated by SiO2-NH2-FPBA were analyzed by Q Exactive HF LC-MS. The results showed that twenty-two kinds of phenolic acids were successfully isolated (see Table 4). SiO2-NH2-FPBA could avoid most of the other hydroxy acids and could conveniently and quickly capture phenolic acids from the crude extract of C. spicatus. Thus, it is a good absorbent as the critical first step in the separation of phenolic acids.

2.6. Antioxidant and Hypoglycemic Activities of the Isolated Phenolic Acids from C. spicatus

The antioxidant and hypoglycemic activities of the isolated phenolic acids from C. spicatus were studied by in vitro models. The results of the DPPH, ABTS and hydroxyl radicals scavenging assays and ferric-reducing antioxidant power (FRAP) assay are shown in Figure 8a,b.
The half maximal inhibitory concentration (IC50) values of the phenolic acids on DPPH, ABTS, and hydroxyl radicals were 36.07 ± 1.23 μg/mL, 32.63 ± 0.41 μg/mL and 600.21 ± 10.84 μg/mL, respectively, while the IC50 values of the positive group (ascorbic acid) were 4.63 ± 0.26 μg/mL, 6.02 ± 0.04 μg/mL and 298.13 ± 7.32 μg/mL, respectively. Furthermore, the phenolic acids presented a dose-dependent FRAP as well as ascorbic acid. In addition, the FRAP of phenolic acids (1 mg/mL) was approximately identical to that of ascorbic acid (123.33 μg/mL).
The results of the α-amylase and α-glucosidase inhibitory assays are shown in Figure 8c. The results showed that phenolics acids from C. spicatus presented a great inhibitory effect against α-amylase and α-glucosidase, with IC50 values of 110.63 ± 3.67 μg/mL and 64.76 ± 0.30 μg/mL, respectively. As for the positive group (acarbose), the IC50 values were 0.342 ± 0.019 mg/mL and 8.986 ± 0.464 ng/mL, respectively.

3. Materials and Methods

3.1. Materials

Silica gel for thin-layer chromatography, used as the support material of the adsorbent, was purchased from a branch of Qingdao Haiyang Chemical Plant (Qingdao, China). 3-Aminopropyl triethoxysilane (3-APTES, 98%), 4-formylbenzoic acid (4-FPBA, 97%), caffeic acid (CA, 98%), protocatechuic acid (PCA, 97%), disodium hydrogen phosphate (Na2HPO4, 99%), and sodium dihydrogen phosphate (NaH2PO4, 97%) were obtained from Heowns Biochemical Technology (Tianjin, China). Gallic acid (GA, 99%) was from Macklin Biochemical (Shanghai, China). Sodium cyanoborohydride (NaBH3CN, 95%) was from Meryer Chemical Technology Co., Ltd. (Shanghai, China). Anhydrous ethanol and sodium carbonate (Na2CO3, 99.8%) were obtained from Jiangtian Chemical Technology (Tianjin, China). Methanol was from Aladdin Chemical Technology (Shanghai, China). Formic acid (HCOOH, 88%) was from Kermel Chemical Technology (Tianjin, China). Acetic acid (CH3COOH, 99.5%) was from Jieerzheng Chemical (Tianjin, China). Hydrochloric acid (HCl, 36–38%) was from Rionlon Pharmaceutical Chemical (Tianjin, China). Benzoic acid (C7H6O2, 99.5%) and sodium hydroxide (NaOH, 96%) were obtained from Damao Chemical Reagents (Tianjin, China). Catechol (99.5%) was from Rhawn Chemical Technology (Shanghai, China). Folin–Ciocalteu reagent (1 mol/L) was from Yuanye Biochemical (Shanghai, China). Clerodendranthus spicatus (C. spicatus) was obtained from the Xishuangbanna region (Yunnan, China.) 1,1-Diphenyl-2-picrylhydrazyl (DPPH) and 2,2′-azinobis-(3-ethylbenthiazoline-6-sulfonate) (ABTS) were from Sigma-Aldrich (St. Louis, MO, USA). α-Amylase (3700 U/g) was from Solarbio Science & Technology Co., Ltd. (Beijing, China). α-Glucosidase (50 U/mg) was from Yuanye Biochemical (Shanghai, China). Other reagents and chemicals were of analytical grade and obtained locally.

3.2. Preparation of Phenylboronic Acid and Amino Bifunctional Modified Absorbent

3.2.1. Preparation of Amino Modified Silica Gel (SiO2-NH2)

Fifty grams of silica gel was dispersed in 200 mL of HCl (6 mol/L) and stirred at 60 °C. After 6 h, it was taken out and washed with ultrapure water to neutral pH and dried in a vacuum oven overnight. The silica gel was activated in a vacuum oven at 130 °C for 3 h. Ten grams of activated silica gel was dispersed in 50 mL anhydrous toluene. After adding 12 mL of 3-APTES, the mixture was stirred for 24 h at 110 °C. The product was washed by toluene, anhydrous ethanol, and acetone, in turn, three times. After drying, SiO2-NH2 was obtained.

3.2.2. Preparation of Phenylboronic Acid and Amino Modified Adsorbent (SiO2-NH2-FPBA)

One gram of SiO2-NH2 was weighed and transferred into 50 mL methanol. After adding 0.48 g of 4-FPBA (3.18 mmol) and 0.4 g of NaCNBH3 (6.36 mmol), the mixture was stirred at 500 rpm for 6 h at 20 °C. After the reaction, the solid product was filtered and washed with ethanol and water, and dried under vacuum at 60 °C for 6 h to obtain the SiO2-NH2-FPBA. The optimization of the reaction conditions, including the amount of 4-FPBA input, the amount of reducing agent, the reaction time and temperature, is shown in Table S1. The synthesis route is shown in Figure 9.

3.3. Adsorption Experiments on GA, CA and PCA

GA, CA and PCA were selected as the target compounds in the adsorption experiments (structures are shown in Figure 9). In order to optimize the phenolic acids’ adsorption conditions, the effects of initial pH, temperature, time and concentration were investigated. The GA, CA, and PCA solution was, respectively, mixed with 20 mg of adsorbents at different pH values (3–8) with different concentrations (0.1–5 mg/mL) and temperature (20–60 °C) for different durations (5–420 min). At the end of the adsorption, the solid was separated from the liquid by centrifugation at 4100 rpm for 3 min. The concentration of the GA, CA, and PCA solution in the supernatant was measured by UV–vis spectrometry at 264 nm, 290 nm and 320 nm, respectively.
The amount of each phenolic acid adsorbed (mg/g) was calculated by the following Equation (5):
Q = C 0 C e × V m
where C0 and Ce are the initial and equilibrium phenolic acid solutions (mg/mL), respectively, V is the volume (mL) of initial solution and m is the mass (g) of adsorbent.
To simulate the environment of the phenolic acids mixture, the solution was prepared by mixing GA, CA, and PCA. The mixed solution contained 0.5 mg/mL of GA, 0.5 mg/mL of CA and 0.5 mg/mL of PCA, respectively. The concentration of each substance in the mixture before adsorption, after adsorption and the desorption solution was determined by HPLC, which was performed on a Chuangxintongheng LC3000 system (Beijing, China); HPLC-MS was performed on a 1260 Infinity UHPLC-6420 system (Agilent, Santa Clara, CA, USA). HPLC conditions were generally as follows: The column was a Baulo C18 (250 × 4.6 mm, 5 μm) from Tianjin Aumi Science & Technology Co., Ltd. (Tianjin, China) The column temperature was room temperature. The mobile phase was methanol and 0.1% formic acid (v/v, 30:70). The flow rate was 1 mL/min. The sample injection volume was 5 μL. The detection wavelength was at 280 nm. The HPLC conditions in HPLC-MS were the same as above. Mass spectrum conditions were as follows: Effluents were determined by ESI combined with multiple reaction monitoring (MRM) in negative ion mode. The fragmentation voltage was 135 V, the nebulizer gas was nitrogen, and the gas flow rate was 7 mL/min. The quantitative ion pairs of the GA, CA, and PCA to be measured were m/z 169 → 125, m/z 179 → 135, and m/z 153 → 109. After detecting the concentrations of various phenolic acids in the supernatant, the desorption rate was calculated according to the following Equation (6):
D e s o r p t i o n   r a t e = C d × V d V × C 0 C e × 100 %
where Cd and Vd are the concentration (mg/mL) and the desorption volume of each phenolic acid (mL), respectively. V, C0 and Ce are the same as Equation (5).

3.4. Adsorption of SiO2-NH2-FPBA to Phenolic Acids from C. spicatus Extract Solutions

The dried C. spicatus was ground with a grinding machine and screened through a 30-mesh sifter. The powder was mixed with 50% (v/v) ethanol at the solid–liquid ratio of 1:25. Ultrasonic extraction was carried out for 45 min twice. After filtration, the supernatant was concentrated with a rotary evaporator to obtain the crude extracts of C. spicatus. The effects of initial pH, temperature, time and concentration were investigated to optimize the adsorption of SiO2-NH2-FPBA to phenolic acids from crude extract solutions of C. spicatus. Then, 50 mg of SiO2-NH2-FPBA was added into C. spicatus crude extract solutions at different pH values (3–8) with different concentrations (0.5–20 mg/mL) and temperatures (20–60 °C) for different durations (5–360 min). The adsorption amount was calculated according to Equation (5), where C was the concentration of the total phenolic acids content. The Folin–Ciocalteu method was used to determine the total phenolic acids content.
The Folin–Ciocalteu reagent can oxidize the phenolic acids to show a blue color, and the shade of blue is proportional to the phenolic acids content. GA was selected as a standard substance. GA solutions of 1 mg/mL were diluted to obtain 0.01, 0.015, 0.02, 0.025, and 0.03 mg/mL of GA solutions. Then, 1 mL of the above solutions were added to 0.5 mL of Folin–Ciocalteu reagent, 1 mL of 12% Na2CO3 solution and 2.5 mL of water. After mixing well and being left to stand for 45 min in a 45 °C water bath, the absorbance of the solution at 765 nm was measured to establish the standard curve. The C. spicatus crude extract solution to be tested after adsorption was processed by the same method and measured at 765 nm.
Solutions of crude extract of C. spicatus before and after adsorption were analyzed by HPLC. The column was a Baulo C18 (250 × 4.6 mm, 5 μm) and the column temperature was at room temperature. The mobile phase was methanol and 0.1% formic acid solution (v/v, 15:85). The flow rate of the mobile phase was 1 mL/min and the sample injection volume was 5 μL. The column effluent was monitored at a wavelength of 280 nm.

3.5. Identification of the Separated Phenolic Acids by Q Exactive HF LC-MS

The isolated substances were determined by Q Exactive HF LC-MS (Thermo Fisher Scientific, Waltham, MA, USA). The conditions were as follows: The column was an Agilent C18 (4.6 ID × 250 mm). The column temperature was at room temperature. The mobile phase was methanol and 0.1% formic acid. The gradient procedure was 40–95% methanol at 0–60 min and 95% methanol at 60–80 min at a flow rate of 0.4 mL/min. The injection volume was 20 μL. The detection wavelength was 280 nm. ESI combined with the anion mode was used for the determination of MS.

3.6. The Antioxidant and Hypoglycemic Activities of the Isolated Phenolic Acids In Vitro

The antioxidant capacity of the isolated phenolic acids from C. spicatus was determined by DPPH, ABTS and hydroxyl radicals scavenging assays and FRAP.
DPPH, ABTS and hydroxyl radicals scavenging activities were determined according to our previous methods with ascorbic acid as the positive control [42]. The results were assessed as follows:
Inhibition (%) = [(Ablank − Asample)/Ablank] × 100%
where Asample and Ablank are the UV absorbance of the sample and blank, respectively.
FRAP was evaluated according to our previous study [43]. The hypoglycemic capacities of the phenolic acids from C. spicatus were measured by α-amylase and α-glucosidase inhibitory assays with acarbose as the positive control [44]. The calculations of inhibition on α-amylase and α-glucosidase were the same as Equation (7).

3.7. Statistical Analysis

Statistical analysis was used to analyze the results. Data were expressed as mean ± standard deviation (SD) and IC50 calculations were performed using SPSS version 22.0. A one-way ANOVA test (Tukey’s test) was used to determine significant statistical differences. The means were regarded as significantly different at p < 0.05.

4. Conclusions

A phenylboronic acid and amino bifunctional modified silica gel SiO2-NH2-FPBA was synthesized and applied to the separation of GA, CA, PCA and phenolic acids in C. spicatus crude extracts. FT-IR, BET, SEM, EDS, TEM, TGA and ICP-OES characterizations were conducted to verify its structure. The adsorption experiments showed that the adsorption amounts of 5 mg/mL of gallic acid, caffeic acid and protocatechuic acid aqueous solution at 20 °C were 103.22 mg/g, 115.49 mg/g, and 109.97 mg/g. The three substances all followed the pseudo-second-order kinetic model, and reached adsorption equilibrium after 120 min. The adsorption isotherm was fit the Langmuir model, which indicated a monolayer adsorption. CH3COOH at pH 2 could desorb 90% of the phenolic acids. HPLC-MS results showed that the eluted substances still maintained their original structures. The adsorption amount of SiO2-NH2-FPBA to phenolic acids in crude extracts of C. spicatus could reach 31.20 mg/g for 2 h at 20 °C. The adsorption kinetics followed the pseudo-second-order model and the adsorption isotherm followed the Langmuir model. The HPLC analysis of the crude extract solutions of C. spicatus before and after the adsorption showed that the peak areas at different retention times were observably reduced and the adsorption rate was 93.24%. Twenty-two kinds of phenolic acids were identified by Q Exactive HF LC-MS, which indicated that SiO2-NH2-FPBA had a significant effect on the one-step isolation of phenolic acid substances from C. spicatus. Furthermore, the isolated phenolic acids from C. spicatus possessed great antioxidant activity with DPPH, ABTS, and hydroxyl radicals scavenging activity (IC50 = 36.07 ± 1.23 μg/mL, 32.63 ± 0.41 μg/mL and 600.21 ± 10.84 μg/mL, respectively) and ferric-reducing power. Furthermore, the phenolic acids showed effective inhibition on α-amylase and α-glucosidase (IC50 = 110.63 ± 3.67 μg/mL and 64.76 ± 0.30 μg/mL, respectively). The findings demonstrate that SiO2-NH2-FPBA has significant potential in the practical application of separating active ingredients’ constituents from natural resources for traditional Chinese medicine.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28227539/s1, Figure S1: (a) Fitted curves of pseudo-first-order model; (b) pseudo-second-order model of the adsorption of C. spicatus; Figure S2: (a) Fitted curves of Langmuir model; (b) Freundlich model of the adsorption of C. spicatus. Table S1: Preparation and conditions optimization of SiO2-NH2-FPBA;

Author Contributions

Conceptualization, Q.S., Y.L. and H.C.; methodology, M.C., Q.S. and X.Z.; validation, M.C. and P.Z.; formal analysis, Q.S., X.Z. and R.Z.; investigation, M.C., P.Z. and R.Z.; resources, Y.L. and H.C.; data curation, M.C., Q.S. and X.Z.; writing—original draft preparation, M.C., Q.S. and X.Z.; writing—review and editing, M.C., X.Z., Y.L. and H.C.; visualization, P.Z. and R.Z.; supervision, Y.L. and H.C.; project administration, Y.L. and H.C.; funding acquisition, Y.L. and H.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2021YFE0110000, and the Tianjin Municipal Science and Technology Foundation, grant number 22JCYBJC00160.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are available from the authors.

Acknowledgments

Thanks to the staff of the Tianjin Key Laboratory of Modern Drug Delivery and Efficiency and Instrumental Analysis Center at School of Pharmaceutical Science and Technology for their help in this research.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Robbins, R.J. Phenolic acids in foods: An overview of analytical methodology. J. Agric. Food Chem. 2003, 51, 2866–2887. [Google Scholar] [CrossRef]
  2. Brautigan, D.L.; Gielata, M.; Heo, J.; Kubicka, E.; Wilkins, L.R. Selective toxicity of caffeic acid in hepatocellular carcinoma cells. Biochem. Biophys. Res. Commun. 2018, 505, 612–617. [Google Scholar] [CrossRef]
  3. Kim, K.W.; Lee, H.J.; Hwang, J.S. Anti-inflammatory effects of chlorogenic acid in lipopolysaccharide-stimulated RAW 264.7 cells. Inflamm. Res. 2014, 63, 81–90. [Google Scholar]
  4. Zhao, H.; Jia, J.; Bai, M.; Miao, M. Updated pharmacological effects of total phenolic acid on the meningeal microcirculation in mice. J. Healthc. Eng. 2022, 2022, 8441050–8441055. [Google Scholar] [CrossRef]
  5. Villanueva-Bermejo, D.; Zahran, F.; Troconis, D.; Villalva, M.; Reglero, G.; Fornari, T. Selective precipitation of phenolic compounds from Achillea millefolium L. extracts by supercritical anti-solvent technique. J. Supercrit. Fluids 2017, 120, 52–58. [Google Scholar] [CrossRef]
  6. Luo, J.; Zeuner, B.; Morthensen, S.T.; Meyer, A.S.; Pinelo, M. Separation of phenolic acids from monosaccharides by low-pressure nanofiltration integrated with laccase pre-treatments. J. Membr. Sci. 2015, 482, 83–91. [Google Scholar] [CrossRef]
  7. Lehmann, M.L.; Counce, R.M.; Counce, R.W.; Watson, J.S.; Labbé, N.; Tao, J. Recovery of Phenolic Compounds from Switchgrass Extract. ACS Sustain. Chem. Eng. 2018, 6, 374–379. [Google Scholar] [CrossRef]
  8. Sun, Y.; Zhu, H.; Wang, J.; Liu, Z.; Bi, J. Isolation and purification of salvianolic acid A and salvianolic acid B from Salvia miltiorrhiza by high-speed counter-current chromatography and comparison of their antioxidant activity. J. Chromatogr. B 2009, 877, 733–737. [Google Scholar] [CrossRef]
  9. Byun, H.-S.; Chun, D. Adsorption and separation properties of gallic acid imprinted polymers prepared using supercritical fluid technology. J. Supercrit. Fluids 2017, 120, 249–257. [Google Scholar] [CrossRef]
  10. Simanaviciute, D.; Klimaviciute, R.; Rutkaite, R. Equilibrium adsorption of caffeic, chlorogenic and rosmarinic acids on cationic cross-linked starch with quaternary ammonium groups. Int. J. Biol. Macromol. 2017, 95, 788–795. [Google Scholar] [CrossRef]
  11. Víctor-Ortega, M.D.; Airado-Rodríguez, D. Revalorization of agro-industrial effluents based on gallic acid recovery through a novel anionic resin. Process Saf. Environ. Prot. 2018, 115, 17–26. [Google Scholar] [CrossRef]
  12. Du, N.; Cao, S.; Yu, Y. Research on the adsorption property of supported ionic liquids for ferulic acid, caffeic acid and salicylic acid. J. Chromatogr. B 2011, 879, 1697–1703. [Google Scholar] [CrossRef]
  13. Sun, Z.; Zheng, Q.; Ma, G.; Zhang, X.; Yuan, J.; Wu, H.; Liu, H.; Yang, J.; Xu, X. Four new phenolic acids from Clerodendranthus spicatus. Phytochem. Lett. 2014, 8, 16–21. [Google Scholar] [CrossRef]
  14. Zhou, H.; Yang, L.; Guo, R.; Li, J. Phenolic acid derivatives with neuroprotective effect from the aqueous extract of Clerodendranthus spicatus. J. Asian Nat. Prod. Res. 2017, 19, 974–980. [Google Scholar] [CrossRef]
  15. Chai, W.S.; Cheun, J.Y.; Kumar, P.S.; Mubashir, M.; Majeed, Z.; Banat, F.; Ho, S.-H.; Show, P.L. A review on conventional and novel materials towards heavy metal adsorption in wastewater treatment application. J. Clean. Prod. 2021, 296, 126589–126604. [Google Scholar] [CrossRef]
  16. Cagnon, B.; Chedeville, O.; Cherrier, J.F.; Caqueret, V.; Porte, C. Evolution of adsorption kinetics and isotherms of gallic acid on an activated carbon oxidized by ozone: Comparison to the raw material. J. Taiwan Inst. Chem. Eng. 2011, 42, 996–1003. [Google Scholar] [CrossRef]
  17. Li, N.; Ng, T.B.; Wong, J.H.; Qiao, J.X.; Zhang, Y.N.; Zhou, R.; Chen, R.R.; Liu, F. Separation and purification of the antioxidant compounds, caffeic acid phenethyl ester and caffeic acid from mushrooms by molecularly imprinted polymer. Food Chem. 2013, 139, 1161–1167. [Google Scholar] [CrossRef]
  18. Fan, D.; Li, H.; Shi, S.; Chen, X. Hollow molecular imprinted polymers towards rapid, effective and selective extraction of caffeic acid from fruits. J. Chromatogr. A 2016, 1470, 27–32. [Google Scholar] [CrossRef] [PubMed]
  19. Song, X.; Chai, Z.; Zhu, Y.; Li, C.; Liang, X. Preparation and characterization of magnetic chitosan-modified diatomite for the removal of gallic acid and caffeic acid from sugar solution. Carbohydr. Polym. 2019, 219, 316–327. [Google Scholar] [CrossRef]
  20. Chai, Z.; Li, C.; Zhu, Y.; Song, X.; Chen, M.; Yang, Y.; Chen, D.; Liang, X.; Wu, J. Arginine-modified magnetic chitosan: Preparation, characterization and adsorption of gallic acid in sugar solution. Int. J. Biol. Macromol. 2020, 165, 506–516. [Google Scholar] [CrossRef]
  21. Nie, L.; Lu, J.; Zhang, W.; He, A.; Yao, S. Ionic liquid-modified silica gel as adsorbents for adsorption and separation of water-soluble phenolic acids from Salvia militiorrhiza Bunge. Sep. Purif. Technol. 2015, 155, 2–12. [Google Scholar] [CrossRef]
  22. Moritz, M.; Geszke-Moritz, M. Amine-modified SBA-15 and MCF mesoporous molecular sieves as promising sorbents for natural antioxidant. Modeling of caffeic acid adsorption. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 61, 411–421. [Google Scholar] [CrossRef]
  23. Qian, J.; Xu, X.; Su, J.; Zeng, W.; Han, B.; Hao, X.; Kai, G. A strategy for effective recovery of salvianolic acid a from Salvia miltiorrhiza (Danshen) through multiple interactions. Compos. B Eng. 2022, 231, 109563–109574. [Google Scholar] [CrossRef]
  24. Su, J.J.; Qian, J.; Zeng, W.H.; Wang, Y.; Kai, G.Y. Effective adsorption of salvianolic acids with phenylboronic acid functionalized polyethyleneimine-intercalated montmorillonite. Sep. Purif. Technol. 2023, 311, 123304–123315. [Google Scholar] [CrossRef]
  25. Guan, S.C.; Fan, G.Y. Diterpenoids from aerial parts of Clerodendranthus spicatus and their cytotoxic activity. Helvetica. Chimica. Acta. 2015, 97, 1708–1713. [Google Scholar] [CrossRef]
  26. Shafaei, A.; Halim, N.; Zakaria, N. Analysis of free amino acids in different extracts of orthosiphon stamineus leaves by high-performance liquid chromatography combined with solid-phase extraction. Pharmacogn. Mag. 2017, 13, 385–391. [Google Scholar]
  27. Luo, Y.; Cheng, L.Z.; Luo, Q.; Yan, Y.M.; Wang, S.M.; Sun, Q.; Cheng, Y.X. New ursane-type triterpenoids from Clerodendranthus spicatus. Fitoterapia 2017, 119, 69–74. [Google Scholar] [CrossRef]
  28. Chen, X.; Ma, G.; Huang, Z.; Tongyu, W.; Xu, X.; Zhong, X. Identification of water-soluble phenolic acids from Clerodendranthus spicatus. Chin. Trad. Herb. Drugs 2017, 48, 2614–2618. [Google Scholar]
  29. David, A.E.; Wang, N.S.; Yang, V.C.; Yang, A.J. Chemically surface modified gel (CSMG): An excellent enzyme-immobilization matrix for industrial processes. J. Biotechnol. 2006, 125, 395–407. [Google Scholar] [CrossRef]
  30. Esmi, F.; Nematian, T.; Salehi, Z.; Khodadadi, A.A.; Dalai, A.K. Amine and aldehyde functionalized mesoporous silica on magnetic nanoparticles for enhanced lipase immobilization, biodiesel production, and facile separation. Fuel 2021, 291, 120126–120133. [Google Scholar] [CrossRef]
  31. Jang, E.H.; Pack, S.P.; Kim, I.; Chung, S. A systematic study of hexavalent chromium adsorption and removal from aqueous environments using chemically functionalized amorphous and mesoporous silica nanoparticles. Sci. Rep. 2020, 10, 5558–5579. [Google Scholar] [CrossRef] [PubMed]
  32. Li, Y.; He, J.; Zhang, K.; Liu, T.; Hu, Y.; Chen, X.; Wang, C.; Huang, X.; Kong, L.; Liu, J. Super rapid removal of copper, cadmium and lead ions from water by NTA-silica gel. RSC Adv. 2018, 9, 397–407. [Google Scholar] [CrossRef] [PubMed]
  33. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  34. Balbuena, P.B.; Gubbins, K.G. Theoretical interpretation of adsorption behavior of simple fluids in slit pores. Langmuir 1993, 10, 1801–1814. [Google Scholar]
  35. Zheng, X.; Qin, Y.; Meng, X.; Jin, Z.; Fan, L.; Wang, J. Synthesis of polyethylene glycol functional bonded silica gel for selective recognition and separation of alpha-cyclodextrin. J. Chromatogr. A 2021, 1639, 461917–461924. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, Z.; Liu, M.C.; Chang, Z.Y.; Li, H.B. Study on the graft modification mechanism of macroporous silica gel surface based on silane coupling agent vinyl triethoxysilane. RSC Adv. 2021, 11, 25158–25169. [Google Scholar] [CrossRef] [PubMed]
  37. Jafri, N.A.A.; Rahman, R.A.; Yusof, A.H.M.; Sulaiman, N.J.; Sukmawati, D.; Syukri, M.S.M. Adsorption kinetics of immobilized laccase on magnetically-separable hierarchically-ordered mesocellular mesoporous silica. Bioresour. Technol. Rep. 2023, 22, 101445–101449. [Google Scholar] [CrossRef]
  38. Mishra, P.; Singh, K.; Dixit, U. Adsorption, kinetics and thermodynamics of phenol removal by ultrasound-assisted sulfuric acid-treated pea (Pisum sativum) shells. Sustain. Chem. Pharm. 2021, 22, 100491–100499. [Google Scholar] [CrossRef]
  39. Zhang, Q.; Zhou, D.D.; Zhang, J.W.; Gao, D.; Yang, F.Q.; Chen, H.; Xia, Z.N. Amino-terminated supramolecular cucurbit [6] uril pseudorotaxane complexes immobilized on magnetite@silica nanoparticles: A highly efficient sorbent for salvianolic acids. Talanta 2019, 195, 354–365. [Google Scholar] [CrossRef]
  40. Zhang, Y.Z.; Qin, B.; Zhang, B.; Ma, J.G.; Hu, Y.Q.; Han, L.; He, M.F.; Liu, C.Y. Specific enrichment of caffeic acid from Taraxacum mon-golicum Hand.-Mazz. by pH and magnetic dual-responsive molecularly imprinted polymers. Anal. Chim. Acta 2020, 1096, 193–202. [Google Scholar] [CrossRef]
  41. Liu, B.; Dong, B.; Yuan, X.; Kuang, Q.; Zhao, Q.; Yang, M.; Liu, J.; Zhao, B. Enrichment and separation of chlorogenic acid from the extract of Eupatorium adenophorum Spreng by macroporous resin. J. Chromatogr. B 2016, 1008, 58–64. [Google Scholar] [CrossRef] [PubMed]
  42. Jia, Y.N.; Gao, X.D.; Xue, Z.H.; Wang, Y.J.; Lu, Y.P.; Zhang, M.; Panichayupakaranant, P.; Chen, H.X. Characterization, antioxidant activities, and inhibition on α-glucosidase activity of corn silk polysaccharides obtained by different extraction methods. Int. J. Biol. Macromol. 2020, 163, 1640–1648. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, J.; Zhang, X.Y.; Liu, J.Y.; Li, R.L.; Zhou, J.N.; Li, M.Y.; Lu, J.Y.; Zhao, G.Y.; Li, X.; Sui, W.J.; et al. Steam explosion improves extractability, antioxidant activity and α-glucosidase inhibitory activity of the constituents of Java tea (Clerodendranthus spicatus). Innov. Food Sci. Emerg. 2023, 86, 103350. [Google Scholar] [CrossRef]
  44. Wang, Y.J.; Chen, Y.; Jia, Y.N.; Xue, Z.H.; Chen, Z.Q.; Zhang, M.; Panichayupakaranant, P.; Yang, S.Y.; Chen, H.X. Chrysophyllum cainito. L alleviates diabetic and complications by playing antioxidant, antiglycation, hypoglycemic roles and the chemical profile analysis. J. Ethnopharmacol. 2021, 281, 114569. [Google Scholar] [CrossRef]
Figure 1. Characterization of phenylboronic acid and amino bifunctional modified silica gel (SiO2-NH2-FPBA): (a) The FT-IR spectra of SiO2-NH2-FPBA and synthetic intermedia; (b) The EDS image; (ce) The SEM and TEM images; (f) Nitrogen adsorption–desorption isotherm; (g) Thermogravimetric curve.
Figure 1. Characterization of phenylboronic acid and amino bifunctional modified silica gel (SiO2-NH2-FPBA): (a) The FT-IR spectra of SiO2-NH2-FPBA and synthetic intermedia; (b) The EDS image; (ce) The SEM and TEM images; (f) Nitrogen adsorption–desorption isotherm; (g) Thermogravimetric curve.
Molecules 28 07539 g001
Figure 2. Effects of different factors on the adsorption capacity of three phenolic acids: (a) pH; (b) temperature.
Figure 2. Effects of different factors on the adsorption capacity of three phenolic acids: (a) pH; (b) temperature.
Molecules 28 07539 g002
Figure 3. (a) Adsorption kinetics of SiO2-NH2-FPBA for three phenolic acids and the model fitting of the pseudo-first-order and pseudo-second-order equations. Fitted curves of (b) pseudo-first-order models and (c) pseudo-second-order models of the adsorption.
Figure 3. (a) Adsorption kinetics of SiO2-NH2-FPBA for three phenolic acids and the model fitting of the pseudo-first-order and pseudo-second-order equations. Fitted curves of (b) pseudo-first-order models and (c) pseudo-second-order models of the adsorption.
Molecules 28 07539 g003
Figure 4. (a) Adsorption isotherm of SiO2-NH2-FPBA for three phenolic acids and the model fitting of the Langmuir and Freundlich models. Fitted curves of (b) Langmuir model and (c) Freundlich model of the adsorption.
Figure 4. (a) Adsorption isotherm of SiO2-NH2-FPBA for three phenolic acids and the model fitting of the Langmuir and Freundlich models. Fitted curves of (b) Langmuir model and (c) Freundlich model of the adsorption.
Molecules 28 07539 g004
Figure 5. (a) Desorption rates of different desorption solutions; (b) HPLC chromatograms; (c) TIC images of before adsorption, after adsorption and desorption solution; (d) EIC images of desorption solution (m/z = 125, 109, 135).
Figure 5. (a) Desorption rates of different desorption solutions; (b) HPLC chromatograms; (c) TIC images of before adsorption, after adsorption and desorption solution; (d) EIC images of desorption solution (m/z = 125, 109, 135).
Molecules 28 07539 g005
Figure 6. Effect of different factors on the adsorption capacity of phenolic acids from C. spicatus crude extract. (a) pH; (b) Temperature; (c) Time; (d) Concentration.
Figure 6. Effect of different factors on the adsorption capacity of phenolic acids from C. spicatus crude extract. (a) pH; (b) Temperature; (c) Time; (d) Concentration.
Molecules 28 07539 g006
Figure 7. (a) Image of C. spicatus solution before and after adsorption; (b) HPLC of C. spicatus solution before and after adsorption.
Figure 7. (a) Image of C. spicatus solution before and after adsorption; (b) HPLC of C. spicatus solution before and after adsorption.
Molecules 28 07539 g007
Figure 8. (a) The scavenging ability of the isolated phenolic acids from C. spicatus on DPPH, ABTS and hydroxyl radicals; (b) FRAP of the isolated phenolic acids from C. spicatus; (c) The α-amylase and α-glucosidase inhibition activity of the isolated phenolic acids from C. spicatus. Note: Different lowercase letters indicate significant differences (p < 0.05) among the columns with same color.
Figure 8. (a) The scavenging ability of the isolated phenolic acids from C. spicatus on DPPH, ABTS and hydroxyl radicals; (b) FRAP of the isolated phenolic acids from C. spicatus; (c) The α-amylase and α-glucosidase inhibition activity of the isolated phenolic acids from C. spicatus. Note: Different lowercase letters indicate significant differences (p < 0.05) among the columns with same color.
Molecules 28 07539 g008
Figure 9. Synthesis route of SiO2-NH2-FPBA and structures of GA, PCA and CA.
Figure 9. Synthesis route of SiO2-NH2-FPBA and structures of GA, PCA and CA.
Molecules 28 07539 g009
Table 1. Pseudo-first-order and pseudo-second-order adsorption kinetics model parameters.
Table 1. Pseudo-first-order and pseudo-second-order adsorption kinetics model parameters.
Phenolic Acid TypePseudo-First-OrderPseudo-Second-Order
k1
(min−1)
Qm
(mg/g)
R2k1
(g·mg−1min−1)
Qm
(mg/g)
R2
Gallic acid0.000724637.130.53300.00686376.450.9999
Caffeic acid0.00228027.810.60380.00273893.280.9999
Protocatechuic acid0.00168035.550.58320.00246185.940.9992
Table 2. Langmuir and Freundlich isotherm models and parameters.
Table 2. Langmuir and Freundlich isotherm models and parameters.
Phenolic Acid TypeLangmuir ParametersFreundlich Parameters
Q0KLRL2KFnRF2
Gallic acid105.78.2980.998983.194.7360.9573
Caffeic acid116.213.480.9989100.29.1090.9889
Protocatechuic acid111.68.9600.998790.256.5170.9943
Table 3. Comparison of phenolic acid adsorption properties with other reported adsorbents.
Table 3. Comparison of phenolic acid adsorption properties with other reported adsorbents.
AdsorbentCompounds *Adsorption Capacity (mg/g)Time (min)Reference
SILsFA, CA, SA53.2–72.2-[12]
Fe3O4@SiO2CA21.1035[18]
MCMDGA, CA31.95, 27.64720[19]
AMCSGA48.38720[20]
Fe3O4@NH2 (CB [6]) NH2 MNPsSAB, FA, RA, SAA, LA, CA21.96–36.3625[39]
pH-MMIPsCA11.50-[40]
NKA-II resinCGA66.86300[41]
SiO2-NH2CA, PCA, GA81.19, 38.15, 58.45120This work
SiO2-NH2-FPBACA, PCA, GA93.64, 89.35, 79.66120This work
* FA: ferulic acid; CA: caffeic acid; SA: salicylic acid; GA: gallic acid; SAB: salvianolic acid B; RA: rosmarinic acid; SAA: salvianolic acid A; LA: lithospermic acid; CGA: chlorogenic acid; PCA: protocatechuic acid.
Table 4. Identification of phenolic acid compounds adsorbed by SiO2-NH2-FPBA.
Table 4. Identification of phenolic acid compounds adsorbed by SiO2-NH2-FPBA.
No.Rt (min)Compound IdentifiedMolecular FormulaMolecular Weight[M-H] m/zErr [ppm]MS/MS
16.73Dihydroferulic acidC10H12O4196.0736195.06583.007195.05064, 59.01374
26.85Chlorogenic acidC16H18O9354.0945353.1023−0.095191.03650
36.93Salicylic acidC7H6O3138.0313137.02411.119119.02329, 93.03446
47.992,5-DihydroxybenzaldehydeC7H6O3138.0317137.02414.017137.02419, 109.02918
58.10Gallic acidC7H6O5170.0212169.12301.325125.02416
68.14Rosmarinic acidC18H16O8360.0845359.20441.475197.04518, 179.03462, 161.02415, 135.04495, 123.04496, 72.99299
78.17Succinic acidC4H6O4118.0266117.39944.574117.01916, 114.40305, 99.00867, 73.02940
88.244-HydroxybenzaldehydeC7H6O2122.0368121.02924.663121.02927, 108.02168, 93.03434
98.883,5-O-dimethyl gallic acidC9H10O5198.0528197.80752.651197.80759
1010.42Caffeic acid ethyl esterC11H12O4208.0736207.13852.834207.02722, 179.03477, 135.04494
1110.62Protocatechuic acidC7H6O4154.0267153.01904.154153.01906, 109.02934
1210.64Dihydrocaffeic acidC9H10O4182.0579181.05022.964181.05038, 137.06052, 109.02934, 59.01374
1311.26Caffeic acidC9H8O4180.0423179.19283.276135.04494
1411.29Vanillic acidC8H8O4168.0423167.04593.510167.033474, 152.01125, 138.92929, 108.02152
1512.07Protocatechuic aldehydeC7H6O3138.0317137.50614.017137.02415
1618.76Yunnaneic acid DC27H24O12540.1267539.13850.875491.10147, 297.07635, 179.03484, 161.02408, 135.04500
1720.31Rosmarinic acid-GlcC24H26O13522.1373521.12900.973359.09784, 341.08820, 323.07614, 197.04514, 179.03461
1821.38Salvianolic aid AC26H22O10494.1140493.20720.763295.06058, 267.06546, 203.03455, 185.02414, 135.04495, 109.02928
1923.22Dicaffeoyltartaric acidC22H18O12474.0798473.06891.102179.03491, 161.02411, 135.04504
2023.99Lithospermic acidC27H22O12538.1111537.10290.971493.11276, 161.02406, 135.04492
2124.35Clerodendranoic acidC29H26O12566.1424565.15530.923357.09732
2264.45Asiatic acidC30H48O5488.3493487.3419−0.668487.34143
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

Cheng, M.; Song, Q.; Zhang, X.; Zheng, P.; Zhao, R.; Li, Y.; Chen, H. Phenylboronic Acid and Amino Bifunctional Modified Adsorbent for Quickly Separating Phenolic Acids from Crude Extract of Clerodendranthus spicatus and Evaluation of Their Antioxidant and Hypoglycemic Activities. Molecules 2023, 28, 7539. https://doi.org/10.3390/molecules28227539

AMA Style

Cheng M, Song Q, Zhang X, Zheng P, Zhao R, Li Y, Chen H. Phenylboronic Acid and Amino Bifunctional Modified Adsorbent for Quickly Separating Phenolic Acids from Crude Extract of Clerodendranthus spicatus and Evaluation of Their Antioxidant and Hypoglycemic Activities. Molecules. 2023; 28(22):7539. https://doi.org/10.3390/molecules28227539

Chicago/Turabian Style

Cheng, Mengqi, Qianyi Song, Xiaoyu Zhang, Pingyi Zheng, Ran Zhao, Youxin Li, and Haixia Chen. 2023. "Phenylboronic Acid and Amino Bifunctional Modified Adsorbent for Quickly Separating Phenolic Acids from Crude Extract of Clerodendranthus spicatus and Evaluation of Their Antioxidant and Hypoglycemic Activities" Molecules 28, no. 22: 7539. https://doi.org/10.3390/molecules28227539

Article Metrics

Back to TopTop