Next Article in Journal
Application of Digital Image Correlation for Strain Mapping of Structural Elements and Materials
Previous Article in Journal
Optimizing the Utilization of Steel Slag in Cement-Stabilized Base Layers: Insights from Freeze–Thaw and Fatigue Testing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparison of Physicochemical, Antioxidant, and Cytotoxic Properties of Caffeic Acid Conjugates

by
Grzegorz Świderski
*,
Ewelina Gołębiewska
,
Monika Kalinowska
,
Renata Świsłocka
,
Natalia Kowalczyk
,
Agata Jabłońska-Trypuć
and
Włodzimierz Lewandowski
Department of Chemistry, Biology and Biotechnology, Institute of Civil Engineering and Energetics, Faculty of Civil Engineering and Environmental Science, Bialystok University of Technology, Wiejska 45E Street, 15-351 Bialystok, Poland
*
Author to whom correspondence should be addressed.
Materials 2024, 17(11), 2575; https://doi.org/10.3390/ma17112575
Submission received: 15 April 2024 / Revised: 18 May 2024 / Accepted: 22 May 2024 / Published: 27 May 2024

Abstract

:
Spectroscopic studies (FT-IR, Raman, 1H, and 13C NMR, UV-VIS) of caffeic acid (CFA) and its conjugates, i.e., caftaric acid (CTA), cichoric acid (CA), and cynarin (CY), were carried out. The antioxidant activity of these compounds was determined by a superoxide dismutase (SOD) activity assay and the hydroxyl radical (HO) inhibition assay. The cytotoxicity of these compounds was performed on DLD-1 cell lines. The molecules were theoretically modeled using the B3LYP-6-311++G(d,p) method. Aromaticity indexes (HOMA, I6, BAC, Aj), HOMO and LUMO orbital energies and reactivity descriptors, NBO electron charge distribution, EPS electrostatic potential maps, and theoretical IR and NMR spectra were calculated for the optimized model systems. The structural features of these compounds were discussed in terms of their biological activities.

1. Introduction

Caffeic acid (CFA) is a natural phenolic compound (a secondary metabolite of plants), belonging to the family of hydroxycinnamic acids (HCAs). CFA is biosynthesized in plant tissues via the endogenous shikimate pathway, which is known to be responsible for the production of aromatic amino acids from glucose [1]. Phenylalanine is a precursor for the synthesis of CFA [2]. CFA can be found in many products consumed daily, such as coffee beans, green tea, tomatoes [3], potatoes [4], artichokes, carrots, lettuces, dark plums, cherries, gooseberries, blackcurrants, grapes [5], and herbs (basil, rosemary, oregano) (Figure 1) [6,7]. CFA and other phenolic compounds are involved in plants’ defense mechanism against insects, pathogens, animals (biotic stresses), and environmental conditions, such as excess water, drought, low and high temperatures, salinity, heavy metals, and ultraviolet radiation (abiotic stresses) [8,9]. Numerous in vitro and in vivo studies have shown that CFA has many biological properties, including anti-inflammatory [10,11,12], anticancer [1,13,14,15], antibacterial [16,17,18], antiviral [19,20], antidiabetic [21,22], hepatoprotective [23,24,25], and cardioprotective activity [26,27]. The presence of a catechol group with a chain of α,β-unsaturated carboxylic acids in the chemical structure of CFA affects its antioxidant properties. This antioxidant mechanism of action is based on the generation of an o-chinone group after the electron donation. Conjugation of the catechol group with the double side binding of o-quinone causes electron delocalization, increasing the stability of the o-quinone radical and the antiradical activity of CFA [28]. CFA can also form complexes with metals (e.g., with iron or copper), inhibiting the decomposition of peroxides, which limits the formation of free radicals and their negative impact on the organism. The excess of free radicals in the organism contributes to unfavorable changes/damage in the structure of proteins, lipids, carbohydrates, and DNA and triggers a number of diseases (e.g., atherosclerosis, asthma, cancer, diabetes, Alzheimer’s, and Parkinson’s diseases) [1,29,30].
Compounds such as 2-caffeoyl-L-tartaric acid (caftaric acid, CTA), 2,3-dicaffeoyl-L-tartaric acid (cichoric acid, CA), and 1,5-dicaffeoylquinic acid (cynarin, CY) are natural conjugates of caffeic acid that have gained popularity in recent years because of their promising antioxidant properties (Figure 2). Structurally, these compounds are conjugates of tartaric acid with tartaric acid or quinic acid. Caftaric acid (CTA) is a major HCA, which can be found in all types of grape seeds and grape juice [31]. CTA is characterized by high bioavailability which has been confirmed in numerous in vitro studies [32]. In rats, CTA is quickly absorbed in the stomach, and can be detected in blood, kidneys, and in the brain, but not in the liver. CTA can also be found in urine as a conversion product—trans-fertaric acid [33,34]. Koriem and Soliman reported that CTA can alleviate methamphetamine (METH)-induced oxidative stress in male albino rats by preventing the accumulation of lipid peroxidation and by restoring liver superoxide dismutase (SOD), and glutathione peroxidase (GPx) activities [35]. In the study by Koriem et al. [33], CTA was found to exhibit an antioxidant effect by inhibiting LA (lead acetate)-induced oxidative damage in rat kidneys. In addition, CTA restored the LA-induced changes in p53 (tumor suppression gene) and bcl-2 (apoptosis inhibitory factor) gene expression to approximately normal levels [33]. CTA is one of the four major ingredients in sweet basil (Ocimum basilicum L.) extract (3.8 mg/g dry extract). In the work of Harnafi et al. [36], this extract exhibited a significant hypolipidemic effect by reducing plasma total cholesterol, triglycerides, and LDL (low-density lipoprotein) cholesterol, by 42%, 42%, and 86%, respectively. Moreover, the extract reduced the atherogenic index and LDL/HDL (high-density lipoprotein) cholesterol ratio by 88% and 94%, respectively [36].
Cichoric acid (CA) has several pro-health activities, including anti-diabetic [37], antiviral, anti-inflammatory [38], and anticancer activity. The study conducted by Tsai et al. [39] showed that CA may be a potential chemotherapeutic agent; the obtained results proved that CA significantly inhibited the activity of telomerase and induced apoptosis in human colon cancer cells (HCT-116) [39]. In the study by Xiao et al. [40], CA showed significant anti-obesity activity in vivo by lowering the serum lipid parameters and reducing the body weight of the tested mice [40]. Zhang et al. [41] reported the anti-hepatotoxic activity of CA isolated from the leaves of Cichorium intybus. A dose of 10–100 µg/mL CA reduced significantly the hepatitis B virus (HBV) surface and envelope antigen levels in HepG2.2.15 human hepatoblastoma cells, and produced the maximum inhibition rates of 79.94% and 76.41%, respectively. At a higher tested dose (50–100 µg/mL), CA significantly inhibited HBV DNA replication [41]. Moreover, the literature data indicate that CA is a potent inhibitor of the HIV-1 IN virus [42,43].
Cynarin (CY) is a major derivative of caffeoylquinic acids found in artichokes leaves and heads [44]. CY has been known to possess biological properties, including antioxidant [45], anti-diabetic [46], anti-atherosclerotic [47,48], hepatoprotective [49,50], anti-tumor [51,52], anti-HIV [53], choleretic [54], and immuno-suppressive activity [55]. In the study by Xia et al. [47], the treatment of HCASMC (human coronary artery smooth muscle cells) with CY led to a downregulation of iNOS (inducible nitric oxide synthase) mRNA and protein expression [47].
Caffeic acid is characterized by high biological activity. Due to the fact that in the plant world, it usually occurs in the form of combinations with other compounds, we wanted to answer the following questions:
(1)
How does the electronic charge distribution within the caffeic acid moiety change after it creates the selected conjugates?
(2)
How do these structural changes affect the activity of caffeic acid and its derivatives?
In this study, the physicochemical and biological properties of caffeic acid and its conjugates (caftaric acid, cichoric acid, and cynarin) were investigated. The molecular structures of compounds were studied using spectroscopic methods (FT-IR, Raman, UV-VIS, 1H, and 13C NMR), and quantum chemical calculations using the Gaussian 09W program. Antioxidant activity was evaluated by SOD-mimic activity and HO radical scavenging activity assays. The cytotoxicity of cichoric acid, caftaric acid, caffeic acid, and cynarin was tested on DLD-1 cell lines.

2. Materials and Methods

2.1. Materials

Cichoric acid, caftaric acid, caffeic acid, cynarin, KBr, XTT sodium salt (2,3-bis(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5-carboxanilide inner salt), KO2, FeSO4∙7H2O, H2O2, and salicylic acid were purchased from Sigma-Aldrich Co. (St. Louis, MO, USA). Dimethyl sulfoxide (DMSO), hydrochloric acid (35%), methanol, and ethanol (analytical grade) were purchased from Chempur (Piekary Śląskie, Poland). All chemicals had an analytical purity and were used without further purification.
Dulbecco’s modified Eagle’s medium (DMEM) with 4.5 mg/mL (25 mM) of glucose with Glutamax, penicillin, streptomycin, trypsin–EDTA, FBS (fetal bovine serum) Gold, and PBS (phosphate-buffered saline) (without Ca and Mg) were provided by Gibco (San Diego, CA, USA). CellTiter-Glo™ 2.0 Assay was purchased from Promega (Madison, WI, USA).

2.2. Theoretical Studies

The optimal geometrical structures of CFA, CA, CTA, and CY and their frequencies of infrared vibrations were calculated using the B3LYP/6-311+G(d,p) method. The theoretical values of chemical shifts were calculated by the GIAO method in B3LYP/6-311+G(d,p) using DMSO as a solvent. The electronic charge distribution of the studied molecules was calculated using the NBO [56] and CHelpG [57] methods. The electrostatic potential (ESP) distribution maps were calculated using the CHelpG method [57]. The energy of HOMO and LUMO orbitals was calculated using the B3LYP/6-311+G(d,p) method. Based on the obtained HOMO and LUMO orbitals’ energy values, other reactivity descriptors, such as energy gap, ionization potential, electron affinity, electronegativity, chemical potential, hardness and softness, and electrophilicity index were calculated. The aromaticity indices were calculated from the length of the bonds in the aromatic ring of the optimized structures. All calculations were performed using the Gaussian 09W software package [58].

2.3. Spectroscopic Studies

The IR spectra of CFA, CA, CTA, and CY were performed by pressing the samples within KBr, and the ATR multi-reflection technique was also used. The spectra were recorded in the range of 4000–400 cm−1 by the use of an Alfa spectrometer (Bruker, Billerica, MA, USA). The Raman spectra were recorded using a Multi-Raman spectrophotometer (Bruker, Bremen, Germany) in the range of 4000–400 cm−1, with a laser power of 250 mW. The 1H and 13CNMR spectra of DMSO solution of the compounds were recorded with a Bruker Avance II 400 MHz unit at room temperature. Tetramethylsilane (TMS) was used as an internal reference. The UV-VIS spectra were recorded for the aqueous solution of the compounds at a concentration of 5.10−5 M in the range of 190–400 nm, using the UV/VIS/NIR Agilent Carry 5000 spectrophotometer (Santa Clara, CA, USA).

2.4. Antioxidant Assays

2.4.1. SOD Activity

The use of the SOD-mimic in vitro test is relevant because of the importance of the SOD (superoxide dismutase) enzyme in the antiradical defense mechanism. The used method was based on the competitive reaction of the compounds, XTT dye, and KO2. The formation of orange XTT-formazane was the result of the interaction of XTT dye formed during the reaction of the superoxide anion radical. The SOD-mimic activity assay was performed based on the method described in [59]. Tested substances were dissolved in DMSO. The reaction mixture consisted of the following: 100 µL of tested substance in a concentration range of 0.05–0.4 mM, 2 mL of phosphate buffer (pH = 7.4; 0.01 M), 50 µL of XTT dye DMSO solution, and 100 µL of saturated KO2 in DMSO. Then, the samples were incubated for 30 min and the absorbance was measured at λ = 480 nm. Control samples without tested compounds were prepared in parallel. The inhibition level (I%) was calculated according to Formula (1):
I% = ((Ac − At)/Ac)) · 100%
where Ac is the absorbance of the control sample and At is the absorbance of the tested sample.

2.4.2. HO Radical Inhibition Activity

Hydroxyl radical inhibition assay was performed according to [60]. A total volume of 0.3 mL of FeSO4 (8 mM), 1 mL of salicylic acid ethanol solution (3 mM), and 0.25 mL of H2O2 (20 mM) were added to 1 mL of the tested compound in the concentration range of 0.1 mM–1 µM. Control sample consisted the same amounts of compounds but H2O was used instead of H2O2. Blank sample consisted of DMSO instead of tested compounds. All samples were incubated at 37 °C for 30 min, then 0.5 mL of H2O was added, and the absorbance was measured at λ = 510 nm. The inhibition level (I%) was calculated according to the following formula:
I% = (1 − ((Ac − At)/Ab)) · 100%
where Ac is the absorbance of the control sample, At is the absorbance of the tested sample, and Ab is the absorbance of the blank sample.
The concentration of the tested compounds was plotted against the %I, and the IC50 values (antioxidant concentration that inhibited 50% of radicals) were calculated from the obtained scavenging curves.

2.5. Cytotoxic Study

The influence of CFA, CA, CTA, and CY was studied in relation to a colorectal adenocarcinoma DLD-1 cell line, which was obtained from the American Type Culture Collection (ATCC). DLD-1 cells were cultured in DMEM (Gibco) supplemented with 10% FBS (Gibco), penicillin (100 U/mL), and streptomycin (100 μg/mL) at 37 °C in a humified atmosphere of 5% CO2 in the air. The cells viability in tested cell lines was examined at the concentrations of 0.5 µM, 1 µM, 5 µM, 10 µM, 20 µM, 50 µM, 100 µM, 200 µM, 300 µM, and 500 µM for every studied compound.

2.5.1. Chemical Treatment of Cells

CFA, CA, CTA, and CY were stored in a refrigerator at a temperature of 4 °C, and the stock solution was prepared by dissolving it in TrisHCl buffer. Compounds were added to the cultured cells for a final concentration in the range of 0.5 µM to 500 µM. The control cells were incubated without test compounds.

2.5.2. Cichoric Acid, Caftaric Acid, Caffeic Acid, and Cynarin Cytotoxicity

CFA, CA, CTA, and CY cytotoxicity was measured with the use of CellTiter-Glo™ 2.0 Assay (Promega) according to manufacturer’s protocol. DLD-1 cells were seeded on a 96-well white plate at a density of 1 × 104 cells/well, and after 24 h, the cells intended to be attached to the plate surface cells were treated with CFA, CA, CTA, and CY in a concentration range from 0.5 µM to 500 µM. After 24 h and 48 h, the cells were subjected to CellTiter-Glo™ 2.0 Assay (Promega). Luminescence was measured with a plate reader GloMax®-Multi Microplate Multimode Reader. The study was performed in triplicate to ensure consistent results were obtained.

2.5.3. Statistical Analysis

All data are given as mean values ±SD (standard deviation). Differences between treatments and untreated control human cells were analyzed by one-way ANOVA, followed by Dunnett’s procedure for multiple comparisons. Significant effects are represented by p ≤ 0.05 (*), p ≤ 0.01 (**), p ≤ 0.001 (***).

3. Results

3.1. Theoretical Calculations

The optimized structures of the analyzed compounds calculated using the B3LYP/6-311++G(d,p) method are shown in Figure 3. Table 1 presents the geometric indices of the studied compounds (energy, dipole moment, energy of HOMO and LUMO orbitals, values of aromaticity indices). The calculated values of the chemical shifts of protons and carbons from the 1H and 13C NMR spectra were analyzed and presented in the chapter “NMR study”, while the vibrational frequencies were theoretically analyzed and presented in the chapter “IR and Raman study”. The aromaticity indices were calculated on the basis of the bond lengths of the structures optimized by the DFT method. The calculated values of the π-electron systems make it possible to assess the aromaticity of the tested compounds, which is related to the stabilization of the aromatic ring and its reactivity. The aromatic ring of CFA exhibits a similar level or aromaticity in each of the presented structures. CY shows a slightly lower value of the HOMA index, the idea of which is based on the alternation of bonds in the aromatic ring, in relation to the other examined structures. The other values of the calculated indices for CTA, CA, and CY are at a similar level and slightly higher than the values for CFA. It can be concluded that in each of the tested systems, the ability to substitute in the aromatic ring will be similar and higher than in the CFA molecule. These compounds are characterized by high aromaticity; therefore, they will not be susceptible to substitution in the benzene ring. CFA is characterized by a slightly higher aromaticity index than its derivatives. The reactivity of the tested compounds is largely related to the ability of the hydroxyl groups to react with other chemicals.
Figure 4 shows the shapes of the orbitals for CFA, CA, CTA, and CY. The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) play an important role in predicting the charge transfer in a molecule, chemical reactivity/ bioactivity, and compound stability [61]. The chemical potential  μ = ( I P + E A ) 2  expresses the ability to detach and escape electrons from a stable system. Chemical hardness  η = ( I P E A ) 2  determines the resistance to the deformation of the electron cloud of a molecule under the influence of disturbances occurring in chemical reactions. Molecules with higher hardness are less susceptible to changes in the electronic charge distribution caused by the attachment of substituents, e.g., to the aromatic ring. The hardness of a molecule determines its resistance to changes in the distribution of electronic charge due to the disturbance of this charge. The HOMO-LUMO energy difference (energy GAP) determines the reactivity of the molecule. The larger the energy gap (HOMO-LUMO), the less reactive a given molecule is (the molecule is hard in local terms). The inverse of the hardness of a molecule is the softness described by the equation S = 1/2η.
The energy gap values (∆E) of the studied compounds decrease in the following series: CY > CA > CFA > CTA. This indicates a decrease in kinetic stability and an increase in their reactivity in the following order. CFA shows a lower reactivity than CY and CA. On the other hand, caftaric acid shows the lowest reactivity according to the calculated energy values of the HOMO and LUMO orbitals. The electrophilicity index (ω) provides information not only about the reactivity but also about the toxicity of the molecule. This parameter is related to energy stabilization when the system receives an additional electrostatic charge from the environment and quantifies the global electrophilic force of the molecule [62]. The electrophilicity index varies in the following series: CY > CFA > CA > CTA. This indicates that cynarin has the greatest electrophilic power.

3.2. NBO and ESP

The reactivity of chemical compounds was also assessed on the basis of the calculated values of electron charges using the NBO method, and maps of electrostatic potential distribution. Table 2 shows the calculated values of the electron charges using the NBO and CHelp methods. Figure 5 shows the maps of the molecular electrostatic potential distribution (EPS). The electrostatic potential map shows the areas of a molecule related to its electrophilic (red) and nucleophilic (blue) reactivity (Figure 5). In CFA, it was observed that the oxygen atom of the carbonyl group was susceptible to electrophilic attack, while the areas of the molecule around the hydrogen atoms in the hydroxyl groups are characterized by increased susceptibility to nucleophilic attack. In the case of CTA and CA, additional centers susceptible to nucleophilic attack appear. These are the hydroxyl groups of the aromatic ring and the hydroxyl groups of a part of the molecule found in tartaric acid. The hydroxyl groups of CTA and CA are more susceptible to nucleophilic attack than in CFA. Fragments of caffeic acid included in the CY molecules show a similar distribution of electrophilic and nucleophilic susceptibility as in other molecules of the tested compounds. The hydroxyl groups of quinic acid present in the CY molecule show low nucleophilic activity.
Calculations of the electron charge distribution using the NBO method showed that the electron density around the atoms of the aromatic ring carbons labeled as C1, C2, and C3 increases in the CFA conjugates compared to the pure acid. In the case of the C4 carbon bonded to the hydroxyl group, in the series of compounds studied, the electron density around this atom is at a similar level. The electron density around the C5 atom is the highest in caffeic acid, while it takes on lower values in its conjugates. The electron densities around the aliphatic atoms C7 and C8 increase slightly in the conjugates with respect to CFA. In contrast, the electron density on the carboxyl group carbon C9 is highest in CFA compared to the conjugates of this acid. The changes in electron charge calculated by the ChelpG method do not coincide in every case with the results obtained by the NBO method.
The electron densities around the protons of the aromatic system H1, H2, and H3 calculated by the NBO method do not change significantly in the series of CFA derivatives. Larger changes were observed for the protons of the hydroxyl groups H4 and H5 (a significant decrease in the electron density at the conjugates with respect to CFA was noted). In the case of aliphatic protons H6 and H7, the observed changes in density around these atoms were also negligible.

3.3. IR and Raman Spectra

The wavenumbers, intensities, and assignments of the selected bands occurring in the FT-IR (recorded in a KBr pellet, and using the ATR technique), the Raman spectra of the tested compounds, and the theoretical infrared vibrational frequencies are presented in Table 3. The FT-IR spectra of CFA, CA, CTA, and CY are presented in Figure 6. In the spectra of the studied compounds, there are characteristic bands derived from vibration bands of the caffeic, tartaric, and quinic acid carboxyl groups. The CFA spectra show characteristic bands assigned to the stretching vibrations of the carbonyl group ν(C=O) located at 1644 cm−1 (IR), and 1640 cm−1 (Raman). In the spectra of conjugates of CFA with tartaric acid, i.e., in CTA and CA, these bands are shifted toward higher values, i.e., up to 1647 cm−1 and 1682 cm−1 in the IR spectra, and 1648 cm−1, 1681 cm−1 in the Raman spectra. In the CY (caffeic acid and quinic acid conjugate) spectra, the wavenumbers of stretching ν(C=O) occurs at 1637 cm−1 (IR) and at 1638 cm−1 (Raman). In the CTA spectra, the bands assigned to the stretching vibrations of the tartaric acid carbonyl group are located at 1758 cm−1, 1707 cm−1 (IR), and at 1757 cm−1, 1706 cm−1 (Raman). These bands are slightly shifted to the values of 1748 cm−1, 1718 cm−1 in the IR spectrum of CA. CY is a conjugate of tartaric acid and quinic acid. Characteristic bands of stretching vibrations of the carbonyl group of quinic acid are observed in the CY spectra. They are located at 1716 cm−1, 1692 cm−1 in the IR spectra, and at 1715 cm−1, 1698 cm−1 in the Raman spectra. The bands resulting from stretching vibrations of C-OH bonds of the carboxyl group of CFA are observed in the spectra of all tested compounds.
A number of characteristic bands related to the vibrations of the aromatic system appear on the spectrum of CFA. The aromatic bands have been assigned according to Versanyi [63]. The formation of conjugates with tartaric or quinic acid causes changes in the spectrum of CFA. By observing the number of bands, position or intensity, it is possible to determine how the formation of conjugates affects the electron charge distribution in the ligand molecule. If the intensity of the bands decreases, the number of bands decreases or the wavenumber values of the aromatic system bands decrease, we are dealing with a disruption of the electron charge distribution in the molecule. Changes in the distribution of electronic charge are associated with a decrease in bond strength constants, which affects the position and intensity of bands on the IR spectra.
In the spectra of the studied conjugates, an increase in the intensity of some bands related to the vibrations of the aromatic system was observed compared to the spectrum of CFA. The wavenumbers of many bands shift toward higher values. These include the stretching bands ν(CH)ar labeled 20a, the stretching bands ν(CC)ar labeled 14, the out-of-plane bending bands γ(CH)ar labeled 17a and 17b, and the deformation bands of the aromatic ring γ(CC)ar, defring ou labeled 16b. Bands that were not present in the spectra of CFA conjugates appear on the spectra of CFA. These are the bands labeled 9a and 4. There was an increase in the wavenumbers of the bands labeled 8a and 5 on the spectra of CTA and CA relative to those bands observed on the spectrum of CFA.
A decrease in the values of the wavenumbers of some bands on the spectra of CFA conjugates compared to the spectrum of CFA was observed. These are bands 19b, 18a, and 6a and 6b (Table 3) On the basis of the analysis of the IR and Raman spectra, it can be concluded that CTA, CA, and CY have higher aromaticity than CFA. However, it should be noted that the stabilization of the electron charge distribution of the aromatic ring of CY is lower in CTA and CA.

3.4. NMR Spectra

As shown in Table 4 and Figure 7, the chemical shifts of the aromatic protons in the 1H NMR spectra of caffeic acid are 6.96, 6.75, and 7.02 ppm. The values of the signals of these protons in the spectra of CTA and CA shift toward higher values, indicating a decrease in the electron density around the nuclei of H1, H2, and H3 and an increase in the aromaticity of the CFA ring in its derivatives. In the case of CY, in which the CFA is linked to the quinic acid molecule, the opposite trend is observed. The values of the aromatic proton signals are lower as compared to CFA. The aromatic rings of CFA in CY show lower aromaticity than in pure CFA. Tartaric acid forming conjugates with CFA in derivatives (CTA, CA) has a stabilizing effect on the electron system of the aromatic ring of caffeic acid, whereas quinic acid (in CY) destabilizes the electron system. Also, calculations of the HOMA aromaticity index for the theoretically modeled structures showed that CY had a lower aromaticity than CFA, while in CTA and CA, the values of this index were higher. Around the protons of the hydroxyl groups H4 and H5, a decrease in electron density (an increase in signal shifts on the spectra) is observed in the conjugates of CFA compared to pure acid. This affects the reactivity of CFA derivatives in proton transfer-based free radical reactions. CFA conjugates are better free radical scavengers than pure CFA. The chemical shifts of the aliphatic protons H6 and H7 are lower in the conjugates compared to CFA. There is an increase in the electron density around the nuclei of these atoms when tartaric or quinic acid is attached to the CFA molecule.
The changes in the values of the chemical shifts of the carbons in the 13C NMR spectra of the studied compounds are related to the change in the electron density around the nuclei of the carbons. The values of the signals of the C1 and C2 aromatic ring carbons are lower in the CFA conjugates compared to the pure acid. This indicates an increase in electronic density around these atoms. In the case of the signals derived from the C4 and C5 carbons (attached to hydroxyl groups), there is an increase in the chemical shift values of these nuclei in the spectra of the CFA conjugates relative to the pure acid. In the case of the C6-labeled carbon, the values of the chemical shifts in all the structures studied are at similar levels. The electronic density around the aliphatic carbon labeled C8 in CFA conjugates is higher than in CFA as evidenced by the lower ppm value of the signal in the 13C NMR spectra of CTA, CA, and CY compared to CFA. In the case of the aliphatic carbon C7, the electronic density around this atomic nucleus is higher in CY than in CFA, while it has a lower value in CTA and CA. The carbon signals of the C9 carboxyl group take on lower values in the spectra of CY, CTA, and CA than in CFA, indicating an increase in the electronic density around this carbon after the formation of the CFA conjugates. The changes in electronic density around carbon atoms in the CFA conjugates relative to the pure acid, observed as changes in signal values on the 13C NMR spectra, follow a similar pattern to the data obtained by theoretical calculations carried out using the NBO method.

3.5. UV-VIS Spectra

In the UV-VIS spectrum recorded for the aqueous solution of CFA, three bands associated with π→π* electron transitions are observed at the wavelengths λ1 = 214.5 nm, λ2 = 289.0 nm, and λ3 = 313.5 nm (Figure 8, Table 5). A bathochromic shift of the λ2 and λ3 bands is observed in the spectra of CFA conjugates. The bathochromic shift of the aromatic system bands in the spectrum indicates an increase in the aromaticity of the molecule. The λ2 band also undergoes a bathochromic shift in the UV-VIS spectrum of CY compared to that of CFA. This band flattens out on the spectral range of CA and CTA and the position of its maximum cannot be determined.

3.6. Antioxidant Activity

Theoretical studies (HOMO and LUMO energy calculations) show that CTA is a better electron acceptor molecule (better antioxidant) than CY. The lower the value of ∆E, the easier the electron enters the excited state in the molecule, and the better its antioxidant activity. Compounds with the lowest EA value have the highest electron transfer capacity, thus giving the highest SOD activity. According to the calculated theoretical parameters, CTA—with the highest energy of HOMO orbital (−8.52886 eV), and with the lowest values of ionization potential (8.52886 eV) and electronegativity (7.322444 eV)—should demonstrate the best antioxidant properties from the tested compounds. The antioxidant activity of CFA, CA, CTA, and CY has been investigated by hydroxyl radical (HO) and superoxide radical (O2) scavenging potentials (Table 6).
As shown in Figure 9, out of four tested compounds, CY was the most active in the inhibition of superoxide anion formation, with an IC50 value equal to 6.880 ± 0.31 µM. CA and CTA were less effective, with IC50 8.062 ± 0.59 and 9.510 ± 0.96 µM, respectively. CFA, with the least complex structure, had an IC50 equal to 10.491 ± 1.20 µM.
The antioxidant activity against HO radical of four tested compounds is shown in Figure 10. IC50 value increased in the following order: CA (20.77 ± 0.86 µM) < CY (31.74 ± 5.51 µM) < CTA (39.18 ± 3.12 µM) < CFA (45.714 ± 3.15 µM). The obtained results prove the fact that CFA is the weakest antioxidant of the series. In all of the performed assays, CFA with the least complex structure was the weakest antioxidant. The presence of the two CFA moieties in the CA and CY molecules determines their high antioxidant properties. Substitution of the aromatic ring in ortho- or para-position can enhance the antioxidant activity because of the possible resonance structures leading to increased stability of the antioxidant radical formed upon the scavenging of other radicals. CFA has one ortho-dihydroxy phenyl group only, while CA and CTA are composed of two molecules of CFA. Moreover, among all the tested compounds, CY and CA have the highest number of hydroxyl groups in their structures, which may also affect their biological activity. The results of a previous study conducted by another research group are consistent with the results of the present work and showed that the antioxidant activity of CY and CA was superior to that of other tested compounds. Liu and coworkers [64], evaluated the HO free radical scavenging ability of CFA, CA, and CTA. They found that at compound concentrations of 500 µM, the degree of radical scavenging activity by CA was about 15.7% and 20.5% higher than that by CTA and CFA, respectively [64]. Many studies have shown that CA and CY have antibacterial, anti-inflammatory, and anti-HIV properties, which could be linked to their antioxidant activity [65,66].

3.7. Cytotoxicity

We evaluated the cytotoxicity of CFA, CA, CTA, and CY (0.5–500 µM) after 24 and 48 h of incubation on DLD-1 cancer cell lines using CellTiter-Glo™ 2.0 assay. As illustrated in Figure 11, all tested polyphenolic compounds exert a cytotoxic effect on an analyzed cell line. DLD-1 colorectal adenocarcinoma cells treated with CA exhibited a statistically significant decrease in relative cell viability even in low concentrations such as 10 µM and 20 µM after 24 h of incubation, but an IC50 value was obtained for the highest analyzed concentration—500 µM. In the case of CY activity in DLD-1 cells, the most inhibitory effect on cell viability was noticed for 500 µM of CY, causing a decrease higher than 70%. An inhibition of DLD-1 cell viability by about 50% was observed in a 300 µM CY 48-hour treatment. In the case of CFA influence on DLD-1 cells, a tendency in cell viability decreasing simultaneously with an increase in the studied compound concentration was noticed. Similar results were obtained for CTA.

4. Conclusions

The formation of the conjugates of caffeic acid (CFA) causes changes in the electronic charge distribution within the CFA molecule, which in turn affects the biological activity of molecules. Spectroscopic studies (FT-IR, Raman, UV-VIS, 1H NMR, 13C NMR) showed higher aromaticity of conjugates of CFA (i.e., CFA, CA, and CY) compared to the unconjugated molecule of CFA. CFA, CA, and CY are characterized by an increased stability of the electronic arrangement in the aromatic ring compared to CFA. Theoretical calculations (NBO, electrostatic potential map) and experimental studies (1H NMR, 13C NMR) showed that the electronic density around the protons of the hydroxyl groups in the conjugate molecules (CTA, CA, and CY) is higher than in free CFA, which causes the increase in the antioxidant capacity of these molecules. The antioxidant assays showed that CFA conjugates possessed higher antiradical activity toward superoxide radical O2 and the hydroxyl radical HO than CFA. Theoretical calculations (including the calculation of the HOMO and LUMO orbitals’ energy confirm these findings. CFA and its conjugates showed cytotoxic properties against DLD-1 cell lines. CA showed the best cytotoxic effect against DLD-1 cells and reduced cell viability at the lowest concentrations used in the study. Slightly weaker cytotoxic potential was exhibited by CY, CTA, and CFA.

Author Contributions

Conceptualization, G.Ś.; methodology, G.Ś.; software, G.Ś.; validation, G.Ś. and M.K.; formal analysis, G.Ś., A.J.-T., M.K., E.G., N.K. and R.Ś.; investigation, G.Ś., A.J.-T., M.K., E.G., N.K. and R.Ś.; resources, G.Ś., A.J.-T., M.K, E.G. and R.Ś.; data curation, G.Ś., A.J.-T., M.K, E.G., N.K. and R.Ś.; writing—original draft preparation, G.Ś., E.G., N.K. and A.J.-T.; writing—review and editing, G.Ś., R.Ś. and N.K.; visualization, G.Ś., E.G. and A.J.-T.; supervision, G.Ś.; project administration, G.Ś. and W.L.; funding acquisition, W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by National Science Centre, Poland, under the research project number 2018/31/B/NZ7/03083.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Monteiro Espíndola, K.M.; Ferreira, R.G.; Mosquera Narvaez, L.E.; Rocha Silva Rosario, A.C.; Machado Da Silva, A.H.; Bispo Silva, A.G.; Oliveira Vieira, A.P.; Chagas Monteiro, M. Chemical and pharmacological aspects of caffeic acid and its activity in hepatocarcinoma. Front. Oncol. 2019, 9, 541. [Google Scholar] [CrossRef] [PubMed]
  2. Hsieh, C.Y.; Huang, Y.H.; Yeh, H.H.; Hong, P.Y.; Hsiao, C.J.; Hsieh, L.S. Phenylalanine, Tyrosine, and DOPA Are bona fide Substrates for Bambusa oldhamii BoPAL4. Catalysts 2021, 11, 1263. [Google Scholar] [CrossRef]
  3. Nia, N.N.; Hadjmohammadi, M.R. Amino acids- based hydrophobic natural deep eutectic solvents as a green acceptor phase in two-phase hollow fiber-liquid microextraction for the determination of caffeic acid in coffee, green tea, and tomato samples. Microchem. J. 2021, 164, 106021. [Google Scholar] [CrossRef]
  4. Singh, B.; Singh, J.; Singh, J.P.; Kaur, A.; Singh, N. Phenolic compounds in potato (Solanum tuberosum L.) peel and their health-promoting activities. Int. J. Food Sci. Technol. 2020, 55, 2273–2281. [Google Scholar] [CrossRef]
  5. El-Seedi, H.R.; El-Said, A.M.A.; Khalifa, S.A.M.; Göransson, U.; Bohlin, L.; Borg-Karlson, A.K.; Verpoorte, R. Biosynthesis, Natural Sources, Dietary Intake, Pharmacokinetic Properties, and Biological Activities of Hydroxycinnamic Acids. J. Agric. Food Chem. 2012, 60, 10877–10895. [Google Scholar] [CrossRef]
  6. Meinhart, A.D.; Damin, F.M.; Caldeirão, L.; de Jesus Filho, M.; da Silva, L.C.; da Silva Constant, L.; Teixeira Filho, J.; Wagner, R.; Teixeira Godoy, H. Study of new sources of six chlorogenic acids and caffeic acid. J. Food Compos. Anal. 2019, 82, 103244. [Google Scholar] [CrossRef]
  7. Phenol-Explorer, Database of Polyphenol Content in Foods, Version 3.6. Available online: http://phenol-explorer.eu/contents/polyphenol/457 (accessed on 4 February 2022).
  8. Riaz, U.; Kharal, M.A.; Murtaza, G.; Zaman, Q. Prospective Roles and Mechanisms of Caffeic Acid in Counter Plant Stress: A Mini Review. Artic. Pak. J. Agric. Res. 2018, 32, 8–19. [Google Scholar] [CrossRef]
  9. Kalinowska, M.; Gołębiewska, E.; Świderski, G.; Męczyńska-Wielgosz, S.; Lewandowska, H.; Pietryczuk, A.; Cudowski, A.; Astel, A.; Świsłocka, R.; Samsonowicz, M.; et al. Plant-Derived and Dietary Hydroxybenzoic Acids—A Comprehensive Study of Structural, Anti-/Pro-Oxidant, Lipophilic, Antimicrobial, and Cytotoxic Activity in MDA-MB-231 and MCF-7 Cell Lines. Nutrients 2021, 13, 3107. [Google Scholar] [CrossRef]
  10. Choudhary, S.; Mourya, A.; Ahuja, S.; Sah, S.P.; Kumar, A. Plausible anti-inflammatory mechanism of resveratrol and caffeic acid against chronic stress-induced insulin resistance in mice. Inflammopharmacology 2016, 24, 347–361. [Google Scholar] [CrossRef]
  11. Paciello, F.; Di Pino, A.; Rolesi, R.; Troiani, D.; Paludetti, G.; Grassi, C.; Fetoni, A.R. Anti-oxidant and anti-inflammatory effects of caffeic acid: In vivo evidences in a model of noise-induced hearing loss. Food Chem. Toxicol. 2020, 143, 111555. [Google Scholar] [CrossRef]
  12. Zielińska, D.; Zieliński, H.; Laparra-Llopis, J.M.; Szawara-Nowak, D.; Honke, J.; Giménez-Bastida, J.A. Caffeic Acid Modulates Processes Associated with Intestinal Inflammation. Nutrients 2021, 13, 554. [Google Scholar] [CrossRef] [PubMed]
  13. Kanimozhi, G.; Prasad, N.R. Anticancer Effect of Caffeic Acid on Human Cervical Cancer Cells. In Coffee in Health and Disease Prevention; Elsevier: Amsterdam, The Netherlands, 2015; pp. 655–661. [Google Scholar]
  14. Kabala-Dzik, A.; Rzepecka-Stojko, A.; Kubina, R.; Jastrzȩbska-Stojko, Ż.; Stojko, R.; Wojtyczka, R.D.; Stojko, J. Migration Rate Inhibition of Breast Cancer Cells Treated by Caffeic Acid and Caffeic Acid Phenethyl Ester: An In Vitro Comparison Study. Nutrients 2017, 9, 1144. [Google Scholar] [CrossRef] [PubMed]
  15. Rezaei-Seresht, H.; Cheshomi, H.; Falanji, F.; Movahedi-Motlagh, F.; Hashemian, M.; Mireskandari, E. Cytotoxic activity of caffeic acid and gallic acid against MCF-7 human breast cancer cells: An in silico and in vitro study. Avicenna J. Phytomed. 2019, 9, 574–586. [Google Scholar] [PubMed]
  16. Lima, V.N.; Oliveira-Tintino, C.D.M.; Santos, E.S.; Morais, L.P.; Tintino, S.R.; Freitas, T.S.; Geraldo, Y.S.; Pereira, R.L.S.; Cruz, R.P.; Menezes, I.R.A.; et al. Antimicrobial and enhancement of the antibiotic activity by phenolic compounds: Gallic acid, caffeic acid and pyrogallol. Microb. Pathog. 2016, 99, 56–61. [Google Scholar] [CrossRef] [PubMed]
  17. Kȩpa, M.; Miklasińska-Majdanik, M.; Wojtyczka, R.D.; Idzik, D.; Korzeniowski, K.; Smoleń-Dzirba, J.; Wasik, T.J. Antimicrobial potential of caffeic acid against staphylococcus aureus clinical strains. BioMed Res. Int. 2018, 2018, 7413504. [Google Scholar] [CrossRef] [PubMed]
  18. Khan, F.; Bamunuarachchi, N.I.; Tabassum, N.; Kim, Y.M. Caffeic Acid and Its Derivatives: Antimicrobial Drugs toward Microbial Pathogens. J. Agric. Food Chem. 2021, 69, 2979–3004. [Google Scholar] [CrossRef] [PubMed]
  19. Wu, Z.M.; Yu, Z.J.; Cui, Z.Q.; Peng, L.Y.; Li, H.R.; Zhang, C.L.; Shen, H.Q.; Yi, P.F.; Fu, B.D. In vitro antiviral efficacy of caffeic acid against canine distemper virus. Microb. Pathog. 2017, 110, 240–244. [Google Scholar] [CrossRef] [PubMed]
  20. Shen, J.; Wang, G.; Zuo, J. Caffeic acid inhibits HCV replication via induction of IFNα antiviral response through p62-mediated Keap1/Nrf2 signaling pathway. Antivir. Res. 2018, 154, 166–173. [Google Scholar] [CrossRef] [PubMed]
  21. Matboli, M.; Eissa, S.; Ibrahim, D.; Hegazy, M.G.A.; Imam, S.S.; Habib, E.K. Caffeic Acid Attenuates Diabetic Kidney Disease via Modulation of Autophagy in a High-Fat Diet/Streptozotocin- Induced Diabetic Rat. Sci. Rep. 2017, 7, 2263. [Google Scholar] [CrossRef] [PubMed]
  22. Xu, W.; Luo, Q.; Wen, X.; Xiao, M.; Mei, Q. Antioxidant and anti-diabetic effects of caffeic acid in a rat model of diabetes. Trop. J. Pharm. Res. 2020, 19, 1227–1232. [Google Scholar] [CrossRef]
  23. Yang, S.Y.; Hong, C.O.; Lee, G.P.; Kim, C.T.; Lee, K.W. The hepatoprotection of caffeic acid and rosmarinic acid, major compounds of Perilla frutescens, against t-BHP-induced oxidative liver damage. Food Chem. Toxicol. 2013, 55, 92–99. [Google Scholar] [CrossRef] [PubMed]
  24. Bispo, V.S.; Dantas, L.S.; Chaves Filho, A.B.; Pinto, I.F.D.; da Silva, R.P.; Otsuka, F.A.M.; Santos, R.B.; Santos, A.C.; Trindade, D.J.; Matos, H.R. Reduction of the DNA damages, Hepatoprotective Effect and Antioxidant Potential of the Coconut Water, ascorbic and Caffeic Acids in Oxidative Stress Mediated by Ethanol. An. Acad. Bras. Cienc. 2017, 89, 1095–1109. [Google Scholar] [CrossRef] [PubMed]
  25. Shokouh, P.; Jeppesen, P.B.; Hermansen, K.; Nørskov, N.P.; Laustsen, C.; Hamilton-Dutoit, S.J.; Qi, H.; Stødkilde-Jørgensen, H.; Gregersen, S. A Combination of Coffee Compounds Shows Insulin-Sensitizing and Hepatoprotective Effects in a Rat Model of Diet-Induced Metabolic Syndrome. Nutrients 2018, 10, 6. [Google Scholar] [CrossRef] [PubMed]
  26. Agunloye, O.M.; Oboh, G.; Ademiluyi, A.O.; Ademosun, A.O.; Akindahunsi, A.A.; Oyagbemi, A.A.; Omobowale, T.O.; Ajibade, T.O.; Adedapo, A.A. Cardio-protective and antioxidant properties of caffeic acid and chlorogenic acid: Mechanistic role of angiotensin converting enzyme, cholinesterase and arginase activities in cyclosporine induced hypertensive rats. Biomed. Pharmacother. 2019, 109, 450–458. [Google Scholar] [CrossRef] [PubMed]
  27. Silva, H.; Lopes, N.M.F. Cardiovascular Effects of Caffeic Acid and Its Derivatives: A Comprehensive Review. Front. Physiol. 2020, 11, 595516. [Google Scholar] [CrossRef] [PubMed]
  28. Medina, I.; Undeland, I.; Larsson, K.; Storrø, I.; Rustad, T.; Jacobsen, C.; Kristinová, V.; Gallardo, J.M. Activity of caffeic acid in different fish lipid matrices: A review. Food Chem. 2012, 131, 730–740. [Google Scholar] [CrossRef]
  29. Lobo, V.; Patil, A.; Phatak, A.; Chandra, N. Free radicals, antioxidants and functional foods: Impact on human health. Pharmacogn. Rev. 2010, 4, 118–126. [Google Scholar] [CrossRef] [PubMed]
  30. Damasceno, S.S.; Dantas, B.B.; Ribeiro-Filho, J.; Antônio, M.; Araújo, D.; da Costa, G.M. Chemical Properties of Caffeic and Ferulic Acids in Biological System: Implications in Cancer Therapy. A Review. Curr. Pharm. Des. 2017, 9, 3015–3023. [Google Scholar] [CrossRef] [PubMed]
  31. Waterhouse, A.L.; Sacks, G.L.; Jeffery, D.W. Non-flavonoid Phenolics. In Understanding Wine Chemistry; Wiley: Hoboken, NJ, USA, 2016; pp. 112–116. [Google Scholar]
  32. Mohamed, K.; Koriem, M. Caftaric acid: An overview on its structure, daily consumption, bioavailability and pharmacological effects. Biointerface Res. Appl. Chem. 2020, 10, 5616–5623. [Google Scholar]
  33. Koriem, K.M.M.; Arbid, M.S. Role of caftaric acid in lead-associated nephrotoxicity in rats via antidiuretic, antioxidant and anti-apoptotic activities. J. Complement. Integr. Med. 2017, 15, 20170024. [Google Scholar] [CrossRef]
  34. Vanzo, A.; Cecotti, R.; Vrhovsek, U.; Torres, A.M.; Mattivi, F.; Passamonti, S. The fate of trans-caftaric acid administered into the rat stomach. J. Agric. Food Chem. 2007, 55, 1604–1611. [Google Scholar] [CrossRef] [PubMed]
  35. Koriem, K.M.M.; Soliman, R.E. Chlorogenic and caftaric acids in liver toxicity and oxidative stress induced by methamphetamine. J. Toxicol. 2014, 2014, 583494. [Google Scholar] [CrossRef] [PubMed]
  36. Harnafi, H.; Ramchoun, M.; Tits, M.; Wauters, J.N.; Frederich, M.; Angenot, L.; Aziz, M.; Alem, C.; Amrani, S. Phenolic acid-rich extract of sweet basil restores cholesterol and triglycerides metabolism in high fat diet-fed mice: A comparison with fenofibrate. Biomed. Prev. Nutr. 2013, 3, 393–397. [Google Scholar] [CrossRef]
  37. Azay-Milhau, J.; Ferrare, K.; Leroy, J.; Aubaterre, J.; Tournier, M.; Lajoix, A.D.; Tousch, D. Antihyperglycemic effect of a natural chicoric acid extract of chicory (Cichorium intybus L.): A comparative in vitro study with the effects of caffeic and ferulic acids. J. Ethnopharmacol. 2013, 150, 755–760. [Google Scholar]
  38. Abd El-Twab, S.M.; Hussein, O.E.; Hozayen, W.G.; Bin-Jumah, M.; Mahmoud, A.M. Chicoric acid prevents methotrexate-induced kidney injury by suppressing NF-κB/NLRP3 inflammasome activation and up-regulating Nrf2/ARE/HO-1 signaling. Inflamm. Res. 2019, 68, 511–523. [Google Scholar] [CrossRef]
  39. Tsai, Y.L.; Chiu, C.C.; Yi-Fu Chen, J.; Chan, K.C.; Lin, S.D. Cytotoxic effects of Echinacea purpurea flower extracts and cichoric acid on human colon cancer cells through induction of apoptosis. J. Ethnopharmacol. 2012, 143, 914–919. [Google Scholar] [CrossRef] [PubMed]
  40. Xiao, H.; Xie, G.; Wang, J.; Hou, X.; Wang, X.; Wu, W.; Liu, X. Chicoric acid prevents obesity by attenuating hepatic steatosis, inflammation and oxidative stress in high-fat diet-fed mice. Food Res. Int. 2013, 54, 345–353. [Google Scholar] [CrossRef]
  41. Zhang, H.L.; Dai, L.H.; Wu, Y.H.; Yu, X.P.; Zhang, Y.Y.; Guan, R.F.; Liu, T.; Zhao, J. Evaluation of hepatocyteprotective and anti-hepatitis B virus properties of Cichoric acid from Cichorium intybus leaves in cell culture. Biol. Pharm. Bull. 2014, 37, 1214–1220. [Google Scholar] [CrossRef] [PubMed]
  42. Lin, Z.; Neamati, N.; Zhao, H.; Kiryu, Y.; Turpin, J.A.; Aberham, C.; Strebel, K.; Kohn, K.; Witvrouw, M.; Pannecouque, C.; et al. Chicoric acid analogues as HIV-1 integrase inhibitors. J. Med. Chem. 1999, 42, 1401–1414. [Google Scholar] [CrossRef] [PubMed]
  43. Siwe-Noundou, X.; Musyoka, T.M.; Moses, V.; Ndinteh, D.T.; Mnkandhla, D.; Hoppe, H.; Tastan Bishop, Ö.; Krause, R.W.M. Anti-HIV-1 integrase potency of methylgallate from Alchornea cordifolia using in vitro and in silico approaches. Sci. Rep. 2019, 9, 4718. [Google Scholar] [CrossRef]
  44. Gezer, C. Potential health effects of the popular compound of artichoke: Cynarin (CYN). Prog. Nutr. 2017, 19, 5–9. [Google Scholar]
  45. Jun, N.-J.; Jang, K.-C.; Kim, S.-C.; Moon, D.-Y.; Seong, K.-C.; Kang, K.-H.; Tandang, L.; Kim, P.-H.; Cho, S.-M.K.; Park, K.-H. Radical Scavenging Activity and Content of Cynarin (1,3-dicaffeoylquinic acid) in Artichoke (Cynara scolymus L.). J. Appl. Biol. Chem. 2007, 50, 244–248. [Google Scholar]
  46. Sun, Z.; Chen, J.; Ma, J.; Jiang, Y.; Wang, M.; Ren, G.; Chen, F. Cynarin-rich sunflower (Helianthus annuus) sprouts possess both antiglycative and antioxidant activities. J. Agric. Food Chem. 2012, 60, 3260–3265. [Google Scholar] [CrossRef] [PubMed]
  47. Xia, N.; Pautz, A.; Wollscheid, U.; Reifenberg, G.; Förstermann, U.; Li, H. Artichoke, Cynarin and Cyanidin Downregulate the Expression of Inducible Nitric Oxide Synthase in Human Coronary Smooth Muscle Cells. Molecules 2014, 19, 3668. [Google Scholar] [CrossRef] [PubMed]
  48. Li, H.; Xia, N.; Brausch, I.; Yao, Y.; Förstermann, U. Flavonoids from artichoke (Cynara scolymus L.) up-regulate endothelial-type nitric-oxide synthase gene expression in human endothelial cells. J. Pharmacol. Exp. Ther. 2004, 310, 926–932. [Google Scholar] [CrossRef] [PubMed]
  49. Adzet, T.; Camarasa, J.; Laguna, J.C. Hepatoprotective activity of polyphenolic compounds from Cynara scolymus against CCl4 toxicity in isolated rat hepatocytes. J. Nat. Prod. 1987, 50, 612–617. [Google Scholar] [CrossRef] [PubMed]
  50. Gebhardt, R. Antioxidative and protective properties of extracts from leaves of the artichoke (Cynara scolymus L.) against hydroperoxide-induced oxidative stress in cultured rat hepatocytes. Toxicol. Appl. Pharmacol. 1997, 144, 279–286. [Google Scholar] [CrossRef] [PubMed]
  51. Atasever, B.; Akgün Dar, K.; Kuruca, S.; Turan, N.; Seyhanli, V.; Meriçli, A. Effects of flavonoids obtained from cynara syriaca on leukemic cells. J. Fac. Pharm. Ankara 2003, 32, 143–150. [Google Scholar]
  52. Gezer, C.; Yücecan, S.; Rattan, S.I.S. Artichoke compound cynarin differentially affects the survival, growth, and stress response of normal, immortalized, and cancerous human cells. Turk. J. Biol. 2015, 39, 299–305. [Google Scholar] [CrossRef]
  53. Slanina, J.; Táborská, E.; Bochořáková, H.; Slaninová, I.; Humpa, O.; Robinson, W.E.; Schram, K.H. New and facile method of preparation of the anti-HIV-1 agent, 1,3-dicaffeoylquinic acid. Tetrahedron Lett. 2001, 42, 3383–3385. [Google Scholar] [CrossRef]
  54. Bradley, P. British Herbal Compendium; British Herbal Medicine Association: Exeter, UK, 2006; Volume 2. [Google Scholar]
  55. Dong, G.C.; Chuang, P.H.; Chang, K.C.; Jan, P.S.; Hwang, P.I.; Wu, H.B.; Yi, M.; Zhou, H.X.; Chen, H.M. Blocking effect of an immuno-suppressive agent, cynarin, on CD28 of T-cell receptor. Pharm. Res. 2009, 26, 375–381. [Google Scholar] [CrossRef]
  56. Weinhold, F.; Landis, C.R. Natural bond orbitals and extensions of localized bonding concepts. Chem. Educ. Res. Pract. 2001, 2, 91–104. [Google Scholar] [CrossRef]
  57. Breneman, C.M.; Wiberg, K.B. Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis. J. Comput. Chem. 1990, 11, 361–373. [Google Scholar] [CrossRef]
  58. Frish, M.J.; Trucks, G.W.; Schlegel, H.; Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; et al. Gaussian 09, Revision A.02; Gaussian: Wallingford, CT, USA, 2009. [Google Scholar]
  59. Štarha, P.; Trávníček, Z.; Herchel, R.; Popa, I.; Suchý, P.; Vančo, J. Dinuclear copper(II) complexes containing 6-(benzylamino)purines as bridging ligands: Synthesis, characterization, and in vitro and in vivo antioxidant activities. J. Inorg. Biochem. 2009, 103, 432–440. [Google Scholar] [CrossRef] [PubMed]
  60. Xiong, S.L.; Li, A.; Huang, N.; Lu, F.; Hou, D. Antioxidant and immunoregulatory activity of different polysaccharide fractions from tuber of Ophiopogon japonicus. Carbohydr. Polym. 2011, 86, 1273–1280. [Google Scholar] [CrossRef]
  61. Freeman, H.C.; Hutchinson, N.D. The crystal structure of the anti-tumor agent 5-(3,3-dimethyl-1-triazenyl)imidazole-4-carboxamide (NSC-45388). Acta Crystall. Sec. B 1979, 35, 2051–2054. [Google Scholar] [CrossRef]
  62. Glossman-Mitnik, D. Computational Study of the Chemical Reactivity Properties of the Rhodamine B Molecule. Procedia Comput. Sci. 2013, 18, 816–825. [Google Scholar] [CrossRef]
  63. Varsanyi, G.; Kovner, M.A.; Lang, L. Assignments for Vibrational Spectra of 700 Benzene Derivatives (Buch, 1973) [WorldCat.org]; Akademiai Kiado: Budapest, Hungary, 1973. [Google Scholar]
  64. Liu, Q.; Wang, Y.; Xiao, C.; Wu, W.; Liu, X. Metabolism of chicoric acid by rat liver microsomes and bioactivity comparisons of chicoric acid and its metabolites. Food Funct. 2015, 6, 1928–1935. [Google Scholar] [CrossRef] [PubMed]
  65. McGrowder, D.A.; Miller, F.G.; Nwokocha, C.R.; Anderson, M.S.; Wilson-Clarke, C.; Vaz, K.; Lennox, A.J.; Brown, J. Medicinal herbs used in traditional management of breast cancer: Mechanisms of action. Medicines 2020, 7, 47. [Google Scholar] [CrossRef]
  66. Janda, K.; Gutowska, I.; Geszke-Moritz, M.; Jakubczyk, K. The common cichory (Cichorium intybus L.) as a source of extracts with health-promoting properties—A review. Molecules 2021, 26, 1814. [Google Scholar] [CrossRef]
Figure 1. Selected plant products containing caffeic acid (on the basis of data from [5]).
Figure 1. Selected plant products containing caffeic acid (on the basis of data from [5]).
Materials 17 02575 g001
Figure 2. Structural formulas of caftaric acid, cichoric acid, and cynarin.
Figure 2. Structural formulas of caftaric acid, cichoric acid, and cynarin.
Materials 17 02575 g002
Figure 3. Structures of CFA, CA, CTA, and CY calculated in B3LYP/6-311++G(d,p).
Figure 3. Structures of CFA, CA, CTA, and CY calculated in B3LYP/6-311++G(d,p).
Materials 17 02575 g003
Figure 4. HOMO and LUMO molecular orbitals of CFA, CA, CTA, and CY.
Figure 4. HOMO and LUMO molecular orbitals of CFA, CA, CTA, and CY.
Materials 17 02575 g004
Figure 5. Electrostatic potential maps (calculated via the SCF in B3LYP/6-311++G(d,p)) for CFA, CA, CTA, and CY.
Figure 5. Electrostatic potential maps (calculated via the SCF in B3LYP/6-311++G(d,p)) for CFA, CA, CTA, and CY.
Materials 17 02575 g005
Figure 6. Experimental infrared (IR) spectra of CFA, CA, CTA, and CY.
Figure 6. Experimental infrared (IR) spectra of CFA, CA, CTA, and CY.
Materials 17 02575 g006
Figure 7. NMR spectra of (A) caffeic acid, (B) caftaric acid, (C) cichoric acid, and (D) cynarin.
Figure 7. NMR spectra of (A) caffeic acid, (B) caftaric acid, (C) cichoric acid, and (D) cynarin.
Materials 17 02575 g007
Figure 8. UV-VIS spectra of aqueous solutions of CFA, CA, CTA, and CY.
Figure 8. UV-VIS spectra of aqueous solutions of CFA, CA, CTA, and CY.
Materials 17 02575 g008
Figure 9. Antioxidant activity (measured using O2 radical assay) of tested compounds.
Figure 9. Antioxidant activity (measured using O2 radical assay) of tested compounds.
Materials 17 02575 g009
Figure 10. Antioxidant activity (measured using HO radical assay) of tested compounds.
Figure 10. Antioxidant activity (measured using HO radical assay) of tested compounds.
Materials 17 02575 g010
Figure 11. Cell viability results for DLD-1 cell lines exposed to different concentrations of (A) CFA, (B) CTA, (C) CA, and (D) CY for 24 h and 48 h, calculated as a percentage of control untreated cells. Each value on the graph is the mean of three independent experiments and error bars show the standard deviation (SD). * p < 0.05, ** p < 0.01, and *** p < 0.001 represent significant effects between treatments and control followed by a Dunnett’s test.
Figure 11. Cell viability results for DLD-1 cell lines exposed to different concentrations of (A) CFA, (B) CTA, (C) CA, and (D) CY for 24 h and 48 h, calculated as a percentage of control untreated cells. Each value on the graph is the mean of three independent experiments and error bars show the standard deviation (SD). * p < 0.05, ** p < 0.01, and *** p < 0.001 represent significant effects between treatments and control followed by a Dunnett’s test.
Materials 17 02575 g011
Table 1. Energy parameters and aromaticity for CFA, CA, CTA, and CY calculated in B3LYP/6-311++G(d,p).
Table 1. Energy parameters and aromaticity for CFA, CA, CTA, and CY calculated in B3LYP/6-311++G(d,p).
CFACTACACY
Energy [hartree]−648.8686−1179.9939−1752.3902−1869.9097
Energy [eV]−17,656.6036−32,109.2506−47,684.9381−50,882.8000
Dipole moment [D]2.12853.09491.63353.7528
HOMO [eV]−8.8187−8.5289−8.7993−8.8959
LUMO [eV]−6.2477−6.1160−6.1269−6.0513
Energy gap [eV]2.57092.41282.67242.8446
Ionization potential, I = −EHOMO8.81878.52898.79938.8959
Electron affinity, A = −ELUMO6.24776.11606.12696.0513
Electronegativity, χ 7.53327.32247.46317.4736
Chemical potential, μ−6.2477−6.1160−6.1269−6.0513
Chemical hardness, η 1.28551.20641.33621.4223
Chemical softness, S 0.38900.41450.37420.3515
Electrophilicity index, ω100.354090.2539100.3204104.1770
Aromaticity indices
Aj0.9910.9930.9930.993
BAC0.8750.8850.8840.886
HOMA0.9560.9580.9580.939
EN0.0200.0250.0250.045
GEO0.0200.0160.0160.016
I692.5193.3093.2993.39
Table 2. Electron density calculated using NBO and ChelpG methods (B3LYP/6-311++G(d,p) for CFA and their conjugates.
Table 2. Electron density calculated using NBO and ChelpG methods (B3LYP/6-311++G(d,p) for CFA and their conjugates.
CFACTACACY
NBOCHelpGNBOCHelpGNBOCHelpGNBOCHelpG
H10.2070.0900.2070.1030.2080.0870.2070.127
H20.2220.1550.2040.2170.2040.2370.2050.150
H30.2050.1820.2030.2620.2020.2240.2030.199
H40.4880.3620.4690.2980.4690.3450.4690.375
H50.4820.3780.4690.3020.4680.3150.4690.369
H60.2150.0980.2140.0350.2140.1140.2160.138
H70.2110.2070.2130.1830.2140.0930.2080.092
H80.4810.359------
O1−0.662−0.513−0.649−0.452−0.649−0.554−0.650−0.529
O2−0.705−0.515−0.657−0.483−0.657−0.447−0.658−0.610
O3−0.618−0.599−0.603−0.582−0.605−0.571−0.611−0.539
O4−0.689−0.589−0.564−0.602−0.555−0.534−0.585−0.478
C1−0.1070.226−0.1120.216−0.1180.328−0.1200.259
C2−0.168−0.195−0.171−0.163−0.171−0.070−0.173−0.096
C3−0.231−0.195−0.259−0.382−0.259−0.556−0.260−0.330
C40.2950.3240.2940.3250.2940.4930.2940.251
C50.2490.2310.2750.4420.2710.3540.2720.608
C6−0.225−0.408−0.225−0.604−0.214−0.525−0.215−0.582
C7−0.095−0.042−0.0730.098−0.075−0.230−0.083−0.178
C8−0.316−0.419−0.315−0.419−0.314−0.145−0.305−0.264
C90.7600.8540.7780.9040.7770.7840.7710.743
Table 3. The wavenumbers [cm−1], intensities, and assignments of bands observed in the experimental FT-IR (KBr and ATR) and FT-Raman spectra and theoretical FT-Raman spectra (calculated by DFT-B3LYP/6-311++G(d,p) method) of CFA, CA, CTA, and CY.
Table 3. The wavenumbers [cm−1], intensities, and assignments of bands observed in the experimental FT-IR (KBr and ATR) and FT-Raman spectra and theoretical FT-Raman spectra (calculated by DFT-B3LYP/6-311++G(d,p) method) of CFA, CA, CTA, and CY.
CFACTACACYAssignments[63]
IRKBrIRATRRamanDFTInt.IRKBrIRATRRamanDFTInt.IRKBrIRATRRamanDFTInt.IRKBrIRATRRamanDFTInt.
3433 s3434 m 385089.93487 s3485 m 383564.43412 s3408 w 3835109.23398 vs3399 m 382961.7νOHar
3236 s3228 w 3783160.63239 m3229 w 3832122.23359 s3352 w3352 w3831240.43234 w 3825119.7νOHar
3770119.9 374271.3 375592.2 381394.7νOHCOOH/νOHchin
3060 m 32036.43063 w3062 vw3069 w32025.43061 w 32035.93041 m 3043 w32062.6ν(CH)ar + ν(CH)C=C2
3025 m 31864.9 31941.7 31954.03022 w 3024 w317216.3ν(CH)ar + ν(CH)C=C20b
2924 m 31539.52953 w 2961 w315515.42954 w 2963 w315424.52925 m 2931 w316912.4ν(CH)ar20a
1758 s1758 m1757 vw1842252.11748 m1746 w 1835236.01716 vs1716 s1715 w νC=Otart /νC=Ochin
1707 vs1707 s1706 w1812332.81718 s1716 m 1824352.51692 vs1692 m1698 w1822254.4νC=Otart /νC=Ochin
1644 vs1643 s1640 m1775351.91647 vs1646 s1648 m1771276.61682 vs1679 vs1681 s1765495.51637 s1639 m1638 m1769292.2νC=Ocaff, ν(C=C)C=C
1618 vs1619 s1612 vs1679193.4 1678228.21624 m 1627 s1678474.81609 s1609 s1609 s1685318.2νC=CC=C
1600 s1600 s1594 m1643163.81616 vs1616 s1617 vs1639438.21606 s1604 m1609 vs1638901.81598 vs1598 s1599 vs1646621.1ν(CC)ar, ν(C=C)C=C8a
1530 m1531 w1531 w1631310.41518 s1518 s1518 vw163357.31515 s1515 m1515 w1632106.31532 m1534 m 164173.2νCCar, νC=CC=C, βCHar, 8b
1557187.91473 w1474 w1476 w1560130.31484 w1489 w1482 w1559266.5 1508 w 1566206.5β(CC)ar, β(CH)ar, ν(CC)ar,19a
1450 vs1449 vs1450 vw147213.11418 m1418 m 146646.91447 w 146694.71445 m1446 m 147154.5ν(CC), β(CH)tart19b
1384 w1375 w 1384 w 141536.51384 m1385 w β(CH)ar, β(CH)tart
1353 m1353 w1352 w140416.61353 s1352 s1351 w1382115.41362 s1362 m1363 w1381340.71367 s1364 s1365 w1393181.4ν(CC)ar, βOHar, β(CH)tart14
1323 m1324 w1325 vw137883.41337 w1336 w1342 vw142142.8 13789.1β(CH)tart
1296 s1296 s1298 vs138193.3 1381340.71307 s1306 m1308 w136422.8βOHar, νCCar, β(CH)C=C
135425.11293 s1294 s1286 m134710.21302 s1300 m1303 m13629.9 1337147.8defring, νCCar, βCHar, βOHar
1280 vs1277 vs1285 m1312260.51262 s1258 s1265 w1306217.51280 m1281 w1271 w1333216.71279 vs1277 vs1279 m1313397.3ν(C−OH)caff, βCHar, βCHtart
1238 s1238 vs1234 w130543.01246 s1246 s1248 m132151.21241 s1239 s 129086.6βCHtart
129394.0 1285124.1 1305440.2 126837.6βCHar, βCHtart, ν(C−O)
1217 s1217 s 1215127.9 120954.1 1209106.5 121471.5βOH, βCHar
119133.71222 s1224 m1215 vw11933.51215 s1203 s1210 w119335.11205 s1202 vs1193 m11933.6defring, βCHar, β(CH)C=C9a
1195 s1195 s1195 w1172150.6 defring, βOHtart
1174 m1174 m1186 m1171126.11163 m1164 m1164 w1151261.31166 m1168 m1169 w1190473.41171 s1170 s1168 w1189214.4βCHar, β(CH)C=C18a
11401351.9 1157 s1157 s 1132871.5βCOH, βCHtart, βOHtart
1120 m1120 m1107 w112155.31120 s1122 s1118 w1117174.91120 s1121 m1125 w1116334.2 1111 w1117192.9βCHar, βOHar 18b
102031.01069 s1069 s1071 vw101528.41077 m1076 m 101551.01082 s βOHcaff, γ(CH)C=C
96816.0989 m989 m989 vw98320.2988 m989 w981 vw98141.7 defring, β(CH)C=C, νCCtart
974 m974 m975 w9666.9966 m964 w 9756.2975 m972 m974 vw96929.6979 m979 m978 w9741.9γ(CH)C=C, γ(CH)ar17b
936 w936 w954 vw9501.7943 w943 m 9251.1926 w 9251.9940 w 962 w9281.3γ(CH)ar17a
900 m898 m 909 w909 w910 w 890 w896 w897 vw 9199.7ν(CCO)caff, defring /defchin
872 vw 888 w888 w 876 w876 w878 w 8874.3β(CH)C=C, defring /defchin
849 m850 m852 vw8907.2858 m858 m858 w8886.5866 w866 w862 w88812.3848 m848 m 8843.2γCHar, γ(CH)C=C5
817 m816 m 85251.9826 m826 s824 vw85137.2 8527.2814 m824 m 85439.3γCHar, γ(CH)C=C
801 w803 m802 w82815.2811 m813 m813 w80626.5803 m803 m808 w80747.5 814 m 80830.2γCHar, 10a
780 w780 m779 vw80818.4768 w762 m 80527.2764 w766 w765 vw78849.6787 w766 w771 w80520.0defring, βOH, β(CH)C=C12
736 w 78218.4749 w749 w751 vw78621.1736 w736 w730 w7465.9740 w 79710.7defring, ν(CC)ar
718 w718 m721 vw7439.4715 vw 7385.2 76952.8γ(C=O)tart, γ(C=O)chin
699 w700 w 75110.5 697 w 73713.9698 vw698 vw699 w7362.2707 w717 w729 w7411.4γ(C=O)caff
686 vw7104.0674 m676 m 6990.3679 w 7060.3668 w 6820.1defring, 4
648 w649 m 65721.7 642 w 6588.8658 w661 w663 vw67320.6647 w649 m 65518.3β(C=O)
616 w 61067.3617 m616 m596 vw59822.4596 w 597 vw59830.4 61382.6defringou, γOH16a
603 w603 m601 w59918.2595 w 5962.7583 m 585 w5831.8600 m 59023.9defring 6a
576 m 57138.0565 w 567 vw57841.9 58179.9568 m 57826.3defring6b
56445.1518 w 54938.8502 w 5530.6531 m 5026.9γOHcaff, γOHtart
458 vw 460 vw4588.9469 w 4524.3450 w 484 vw4547.3460 w 458 w4554.5γ(CC)ar, defringou16b
446 w43467.7435 w 38991.7431 w 436 vw4327.8421 w 429 w γOHar, γOHtart
* The intensities (Int.) of the bands were assigned as follows: wv—very weak, w—weak, m—medium, s—strong, vs—very strong and ** The types of vibrations were assigned as follows: ν—stretching vibrations, β—in-plane bending modes, γ—out-of-plane bending modes, def.—deforming vibrations, defring—deforming vibrations of the aromatic ring, ou—out-of-plane bending vibrations, caff—vibrations of the atoms of caffeic acic, tart—vibrations of the atoms of tartaric acid, al—aliphatic atoms, ar—aromatic system atoms.
Table 4. Chemical shifts observed in NMR spectra of CFA, CA, CTA, and CY.
Table 4. Chemical shifts observed in NMR spectra of CFA, CA, CTA, and CY.
CACTACACY
Exp.Calc.Exp.Calc.Exp.Calc.Exp.Calc.
H16.966.677.587.727.547.636.637.87, 8.00
H26.756.476.787.037.106.906.517.26
H37.027.537.057.057.086.907.007.21
H49.513.909.654.939.704.669.565.28
H59.124.979.214.619.184.319.425.06
H67.417.717.028.036.787.846.208.23
H76.176.176.246.706.386.576.066.86, 6.72
H81.105.25-----
C1125.71132.75125.37133.41125.23131.83125.25114.97
C2121.16126.72121.54125.35115.81123.80120.42106.43
C3115.13121.63114.93122.38115.26120.58115.85107.39
C4145.57155.83146.48155.49147.04153.81148.27134.73
C5144.59149.80145.65150.42145.62149.56145.57130.55
C6115.76124.91115.87126.68121.74125.48115.53108.94
C7148.14157.11148.72156.63148.87155.96145.33135.54
C8114.65116.09112.92116.92112.37114.05114.1999.64
C9167.91175.12165.67174.00165.53172.88165.64155.24
Table 5. The wavelengths of maximum absorbance from the UV-VIS spectra of CFA, CA, CTA, and CY.
Table 5. The wavelengths of maximum absorbance from the UV-VIS spectra of CFA, CA, CTA, and CY.
CFACTACACY
λmax1 [nm]214.5216.5217.0215.0
λmax2 [nm]289.0--299.5
λmax3 [nm]313.5324.0325.0319.0
Table 6. Results of the various antioxidant activity tests of tested compounds.
Table 6. Results of the various antioxidant activity tests of tested compounds.
IC50: HO [µM]O2 [µM]
Caffeic acid45.714 ± 3.1510.491 ± 1.20
Cichoric acid20.768 ± 0.868.062 ± 0.59
Caftaric acid39.180 ± 3.129.510 ± 0.96
Cynarin31.741 ± 5.516.880 ± 0.31
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

Świderski, G.; Gołębiewska, E.; Kalinowska, M.; Świsłocka, R.; Kowalczyk, N.; Jabłońska-Trypuć, A.; Lewandowski, W. Comparison of Physicochemical, Antioxidant, and Cytotoxic Properties of Caffeic Acid Conjugates. Materials 2024, 17, 2575. https://doi.org/10.3390/ma17112575

AMA Style

Świderski G, Gołębiewska E, Kalinowska M, Świsłocka R, Kowalczyk N, Jabłońska-Trypuć A, Lewandowski W. Comparison of Physicochemical, Antioxidant, and Cytotoxic Properties of Caffeic Acid Conjugates. Materials. 2024; 17(11):2575. https://doi.org/10.3390/ma17112575

Chicago/Turabian Style

Świderski, Grzegorz, Ewelina Gołębiewska, Monika Kalinowska, Renata Świsłocka, Natalia Kowalczyk, Agata Jabłońska-Trypuć, and Włodzimierz Lewandowski. 2024. "Comparison of Physicochemical, Antioxidant, and Cytotoxic Properties of Caffeic Acid Conjugates" Materials 17, no. 11: 2575. https://doi.org/10.3390/ma17112575

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