Next Article in Journal
Recent Advances on Stimuli-Responsive Hydrogels Based on Tissue-Derived ECMs and Their Components: Towards Improving Functionality for Tissue Engineering and Controlled Drug Delivery
Next Article in Special Issue
Grape Canes (Vitis vinifera L.) Applications on Packaging and Particleboard Industry: New Bioadhesive Based on Grape Extracts and Citric Acid
Previous Article in Journal
Structural Design Calculation of Basalt Fiber Polymer-Modified RPC Beams Subjected to Four-Point Bending
Previous Article in Special Issue
Synthesis of New Chitosan from an Endemic Chilean Crayfish Exoskeleton (Parastacus Pugnax): Physicochemical and Biological Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Systematic Study of the Antioxidant Capacity of Humic Substances against Peroxyl Radicals: Relation to Structure

by
Olga I. Klein
1,
Natalia A. Kulikova
1,2,*,
Andrey I. Konstantinov
3,
Maria V. Zykova
4 and
Irina V. Perminova
3
1
Bach Institute of Biochemistry, Fundamentals of Biotechnology Federal Research Center, Russian Academy of Sciences, pr. Leninskiy 33, 119071 Moscow, Russia
2
Department of Soil Science, Lomonosov Moscow State University, Leninskiye Gory 1-12, 119991 Moscow, Russia
3
Department of Chemistry, Lomonosov Moscow State University, Leninskiye Gory 1-3, 119991 Moscow, Russia
4
Department of Chemistry, Siberian State Medical University, 634050 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(19), 3262; https://doi.org/10.3390/polym13193262
Submission received: 31 August 2021 / Revised: 21 September 2021 / Accepted: 22 September 2021 / Published: 25 September 2021
(This article belongs to the Special Issue Biobased and Natural Polymers)

Abstract

:
Humic substances (HS) are natural supramolecular systems of high- and low-molecular-weight compounds with distinct immunomodulatory and protective properties. The key beneficial biological activity of HS is their antioxidant activity. However, systematic studies of the antioxidant activity of HS against biologically relevant peroxyl radicals are still scarce. The main objective of this work was to estimate the antioxidant capacity (AOC) of a broad set of HS widely differing in structure using an oxygen radical absorption capacity (ORAC) assay. For this purpose, 25 samples of soil, peat, coal, and aquatic HS and humic-like substances were characterized using elemental analysis and quantitative 13C solution-state NMR. The Folin–Ciocalteu method was used to quantify total phenol (TP) content in HS. The determined AOC values varied in the range of 0.31–2.56 μmol Trolox eqv. mg−1, which is close to the values for ascorbic acid and vitamin E. Forward stepwise regression was used to reveal the four main factors contributing to the AOC value of HS: atomic C/N ratio, content of O-substituted methine and methoxyl groups, and TP. The results obtained clearly demonstrate the dependence of the AOC of HS on both phenolic and non-phenolic moieties in their structure, including carbohydrate fragments.

Graphical Abstract

1. Introduction

Humic substances (HS) are ubiquitous in natural and human-made environments, such as organic rocks, soil, compost, and natural water [1,2]. Consolidated resources of humic materials are deposited mostly in peat, coal, composts, and sapropel [3]. HS are produced in situ due to chemical, physical, and microbial degradation, as well as (re)polymerization of phenolic and aromatic components such as lignin, tannins, polysaccharides, lipids, and proteins [4,5,6]. They have a non-stoichiometric elemental composition, irregular structure, and heterogeneous molecular composition [7]. A single structural formula cannot be assigned to any sample of HS. As a result, HS are still operationally classified according to their solubility in acidic and alkaline solutions into two major classes: humic acids (HA), which are insoluble at pH < 2, and fulvic acids (FA), which are soluble in the whole pH range [8].
Recently, HS have attracted increasing attention from the point of view of their use in medicine [9] due to their therapeutic potential against chronic inflammatory diseases [10,11,12], antiviral activity [13,14,15], beneficial effect in accelerating cutaneous wound healing [16,17], and chelating activity toward toxic metals [18,19].
Antioxidant properties of HS against reactive oxygen species (ROS) and free radicals are generally believed to largely determine their significant potential for use in medicinal, pharmaceutical, and cosmetic application and food industries [9,19,20,21,22]. Piotrowska and co-workers reported suppression of lipid peroxidation in human placenta mitochondria in the presence of peat HS, confirmed by a decreased malondialdehyde level [23]. The reported hepatoprotective [24,25,26], neuroprotective [20], renoprotective [27], and cardioprotective [28,29] effects of HS were explained by their antioxidant activity.
ROS or high levels of free radicals cause oxidative stress, leading to degradation of DNA, cell membranes, proteins, and other cellular constituents [30,31,32]. This may induce a number of human diseases and conditions, such as atherosclerosis, rheumatoid arthritis, muscular dystrophy, cataracts, neurological disorders, some types of cancers, and aging [33]. Oxidative stress can be ameliorated by supplying effective antioxidants [34,35]. Epidemiological studies of dietary antioxidant consumption and health outcomes have demonstrated the protective effects of various antioxidant-rich foods against chronic diseases (cardiovascular disease, neurodegeneration, and cancer), myocardial infarction, and stroke. Regular dietary antioxidant intake results in increased total antioxidant capacity in the serum and decreased plasma levels of inflammatory cytokines (IL-6 and TNF-α), the oxidative stress biomarker F2-isoprostane, and a nonspecific marker of inflammation in serum, C-reactive protein (CRP) [36]. The antioxidant activity of HS has been attributed mainly to the presence of phenolic and quinoid moieties. Similar to phenols, HS can behave as electron donors or acceptors depending on the redox state of the system [21,37]. This paradigm is in good agreement with the observation that phenol moieties are structural fragments responsible for the antioxidant activity of many other naturally occurring bioactive substances [9,38,39]. Still, very little is known about the non-phenolic moieties of HS, which could also contribute to their antioxidant properties. Non-phenolic antioxidants represent an important and abundant class of natural radical scavengers [34]. Over the past decade, a variety of natural polysaccharides have attracted great interest due to their antioxidant functions, responsible for wide-ranging beneficial therapeutic effects and health-promoting properties [40,41]. The content of carbohydrates in HS can reach 10% [42], which might represent their contribution to the antioxidant activity of humic materials.
More than 40 laboratory protocols for testing antioxidant capacity and their modifications are reported in the literature [43,44,45]. Among them, the oxygen radical absorption capacity (ORAC) method is relevant to in vivo conditions because it uses a biologically relevant free radical source, peroxyl radical ROO, which is the most prevalent free radical in human biology [36,46]. Peroxyl radicals are characterized as free radicals that predominate in lipid oxidation in biological systems and in foodstuffs under physiological conditions [46]. The ORAC method was initially based on the ability of antioxidants to prevent the consumption of β-phycoerythrin mediated by peroxyl radicals generated during the aerobic thermal decomposition of 2,2′-azo-bis(2-amidinopropane)dihydrochloride (AAPH). However, the currently employed assay, based on work published by Ou and co-workers [47], in which fluorescein is proposed as the target molecule, is believed to determine the capacity of antioxidants to trap AAPH-derived peroxyl and/or AAPH-derived alkoxyl radicals [48]. Along with peroxyl radicals, alkoxyl radicals are also abundant, damaging free radicals in the human body [49].
The ORAC test is usually classified as a method that involves hydrogen atom transfer (HAT) reactions [36,46,48,50]. For non-phenolic antioxidant reactions, donation from other than H atoms can also occur with peroxyl radicals [34]. The presence of alkoxyl radicals complicates the situation, and they can be scavenged by reactions involving proton-coupled electron transfer (PCET), electron transfer–proton transfer (ET-PT), and second sequential proton loss electron transfer (SPLET) mechanisms [49,51]. Despite some uncertainty in our understanding of the mechanisms underlying the ORAC assay [48], this approach provides a direct measure of the hydrophilic and lipophilic chain-breaking antioxidant capacity [36] of both phenolic [50] and non-phenolic antioxidants, such as carbohydrates [52]. However, studies devoted to assessing the antioxidant activity of HS mainly use assays with the 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonic acid (ABTS) cation radical [37,53], superoxide radical O2 [54], reduction of the complex of ferric ions (Fe3+) [55], or Cu (II) reduction to Cu (I) [56], or they initiate oxidation of 1,4-dioxane as a model reaction [57]. The only study of the antioxidant activity of HS against peroxyl radical induced by α,α-azobisizobutyronytril (AIBN) was performed for a single sample of coal-derived HA [58].
The main objective of this work was to characterize HS antioxidant capacity (AOC) using the ORAC assay as a function of HS structure, which can be responsible for the type of activity of humic materials. To this end, we quantified total phenol content using the Folin–Ciocalteu method and carbon distribution among the structural fragments using quantitative 13C solution-state NMR of 25 samples of soil, peat, coal, and aquatic HS and humic-like substances (HLS) produced by basidiomycete fungi. To obtain insight into the antioxidant activity of HS, we developed a forward stepwise multiple linear regression. To visualize HS clustering based on physical–chemical properties responsible for antioxidant activity, we applied principal component analysis (PCA).

2. Materials and Methods

Ultrapure deionized water (Milli-Q®, 18 MΩ cm, Millipore, Burlington, MA, USA) was used to prepare all solutions. All chemicals used were of at least analytical grade. Inorganic compounds (NaOH, HCl, KCl, KH2PO4, K2HPO4) were obtained from Chimmed (Russia). Folin and Ciocalteu’s phenol reagent, 2,2′-azobis (2-amidino-propane) dihydrochloride (AAPH), fluorescein, 6-hydroxy-2,5,7,8-tetra-methylchroman-2-carboxylic acid (Trolox), hydroxybenzoic acid (HBA), vanillic acid, syringic acid, gallic acid, fumaric acid, ferulic acid, sinapic acid, ascorbic acid, and vitamin E were procured from Sigma-Aldrich (Burlington, MA, USA). Standard samples of HS, including Suwannee River humic acid (SRHA) and fulvic acid (SRFA) and dissolved organic matter (SRDOM), were kindly provided by I.V. Perminova, coordinator of the Russian chapter of the International Humic Substances Society (IHSS, St. Paul, MN, USA). Commercially available coal humic materials CHA-AGK and CHA-ALD were purchased from Biotechnology Ltd. (Moscow, Russia) and Sigma-Aldrich (Burlington, MA, USA), respectively.

2.1. Isolation of Humic Substances

Humic materials were isolated from natural sources including soil (9 samples), peat (9 samples), coal (2 samples), and natural water (3 samples). They were either fractionated into humic acid (HA) and fulvic acid (FA) or used as a non-fractionated mixture of HA and FA (HF). Additionally, humic-like substances (HLS) were used, synthesized during solid-phase cultivation of the basidiomycete Trametes maxima 0275 (2 samples). Overall, 25 samples of humic materials were used in the study. A list of the HS and their sources is presented in Table 1.
Soil HA and FA (SHA and SFA) were isolated from Albic Retisol (Moscow region, Russia) and Gleyic and Endocalcic Chernozems (Voronezh region, Russia), as described in [59]. In brief, the SHA and SFA from Albic Retisol were obtained by alkali extraction with 0.1 M NaOH. For Chernozems, before alkali extraction the sample was pretreated with 10% HCl. The alkali extract was treated with 0.3 M KCl and centrifuged to remove organomineral colloidal particles. The SHA and SFA were obtained by acidification of the supernatant to pH 1–2. The precipitated HA were desalted by dialysis. To isolate FA, the acidic supernatant was passed through XAD-2 resin. The sorbed fraction of SFA was recovered by back elution with 0.1 M NaOH, desalted on cation-exchange resin, and freeze-dried.
Peat HF (PHF) was isolated from samples of highland and lowland peat located in the Tver region (Russia). The highland peat type was sedge (T6), the lowland peat types were sedge (T3) and woody (T7). The isolation procedure was described elsewhere [60] and included preliminary treatment with an ethanol–benzene mixture (1:1) followed by alkaline extraction (0.1 M NaOH). The SHA and SFA were obtained by acidification as described for soil HA and FA.
Coal HA (CHA) was obtained from two commercial preparations, ALD (Aldrich HA) and AGK (Biotechnology Ltd., Moscow, Russia). They were desalted on a cation-exchange resin and freeze-dried.
Aquatic HS consisted of standard samples of HS produced by the International Humic Substances Society (IHSS, St. Paul, MN, USA), including Suwannee River HA (SRHA), FA (SRFA), and dissolved organic matter (SRDOM).
Two samples of HLS, HLS-45, and HLS-70 were produced in our laboratory by the basidiomycete Trametes maxima 0275 cultivated on oat straw. HLS-45 and HLS-70 were isolated after 45 and 70 days of solid-phase cultivation, as described previously [61]. In brief, after 45 and 70 days of cultivation, the flasks containing fungal cultures were supplemented with warm distilled water (60 °C) and left for 6 h under constant stirring. Then the liquid was filtered through a paper filter to separate the straw with mycelium. The filtrate was supplemented with concentrated HCl drop-by-drop to adjust the pH to 2.0. The solution was left for 24 h; as a result, a grayish-brown flocculent precipitate was formed. The precipitate was collected by centrifugation, washed with distilled water, centrifuged again, and dialyzed against distilled water.

2.2. Characterization of Humic Substances

The CHN content was determined using a model 1106 elemental analyzer (Carlo Erba Strumentazione, Milan, Italy). The oxygen content was calculated as the difference between the dry ash-free weight of the sample and the total CHN content.
Quantitative 13C solution-state NMR spectra were recorded on a Bruker Advance 400 spectrometer (Bruker BioSpin, Germany) operating at 100 MHz according to Hertkorn and co-authors [62]. To quantify the spectra obtained, the following assignments were made (in ppm): 220–189, C atoms of quinone and ketone groups (CC=O); 189–168, C atoms of carboxylic and esteric groups (CCOO); 168–145, aromatic O-substituted C atoms (CAr-O); 145–108, aromatic H- and C-substituted atoms (CAr); 108–91, anomeric double substituted aliphatic C atoms (COCO); 91–66, О-substituted methine groups (CCHO), 66–59, O-substituted methylene groups (CCH2O); 59–48, methoxyl C atoms (CCH3O), 48–0, and aliphatic H- and C substituted C atoms (CCHn). In addition to the integrals of the given ranges, the sum of O-substituted aliphatic C ΣCAlk-O and carbohydrate carbon ΣCCarb and the ΣCAr/ΣCAlk ratio were calculated. The value of ΣCAlk-O was the sum of COCHO + CCHO + CCH2O + CCH3O, ΣCCarb was the sum of COCO, CCHO, and CCH2O, and ΣCAr and ΣCAlk were the sum of CAr + CAr–O and ΣCAlk-O + CCHn, respectively.
The total phenolic (TP) content of the HS and HLS samples was determined according to the Folin–Ciocalteu method [63].

2.3. Estimation of Oxygen Radical Absorbance Capacity (ORAC) of Humic Substances

ORAC of HS was determined as described by Ou and co-authors [47]. For this purpose, 2,2′-azobis (2-amidino-propane) dihydrochloride (AAPH), fluorescein, and 6-hydroxy-2,5,7,8-tetra-methylchroman-2-carboxylic acid (Trolox) were prepared in 75 mM phosphate buffer with a pH of 7.4. For measurements, AAPH was prepared at a concentration of 600 mM and made fresh daily. A sodium fluorescein stock solution (4 μM) was stored wrapped in foil at 5 °C. Immediately prior to use, the stock solution was diluted with phosphate buffer to a final concentration 0.08 μM. Trolox concentration was 25 μM.
To perform the measurements, a Synergy HT Multi-Detection Microplate Reader (BioTek Instruments Inc., Winooski, VT, USA) was used, with 115 µL of working sodium fluorescein solution added to all experimental wells. Control wells received 15 µL of phosphate buffer, while standards received 15 µL of Trolox dilution and samples received 15 µL of HS solution at a concentration 10 mg L−1. The plate was then allowed to equilibrate by incubating for 30 min at 37 °C. Reactions were initiated by the addition of 15 µL of AAPH solution [64]. Fluorescence was then monitored kinetically with data taken every minute for 1 h. An excitation wavelength of 485 nm and an emission wavelength of 528 nm were used in the assay. The AOC value referred to the net protective area under the quenching curve of fluorescein in the presence of an antioxidant. The net area under the curve (AUC) of the standard (Trolox) and HS was calculated. The AOC related to Trolox was calculated as follows:
AOC = AUC HS AUC control AUC Trolox AUC control × [ Trolox ] ,
where [Trolox] is the concentration of Trolox in micromoles.
The results were evaluated using Origin 8.0 software (OriginLab Corp., Northampton, MA, USA) and expressed as μmol of Trolox equivalent (TE) per milligram of sample. The measurements were performed in fourfold repetition. To compare antioxidant properties of HS with other antioxidants, the AOC of several individual compounds was estimated using the ORAC approach including hydroxybenzoic acid (HBA), vanillic acid, syringic acid, gallic acid, fumaric acid, ferulic acid, sinapic acid, ascorbic acid, and vitamin E.

2.4. Statistical Data Treatment

Data on the AOC of HS are presented as the mean ± standard deviation (SD). To analyze differences among means, analysis of variance (ANOVA) was applied, followed by Tukey’s HSD test at p < 0.05. To reveal the relationship between the structural characteristics and antioxidant properties of HS, a correlation analysis, principal component analysis (PCA), was applied and a forward stepwise regression model was developed. All statistical data treatments were carried out using the Statistica 8.0 software package (StatSoft, Dell Inc., Round Rock, TX, USA).

3. Results

3.1. Structural Characteristics of HS

A set of the analyzed humic materials included HS samples isolated from different natural environments and HLS produced by the basidiomycete fungus. In addition, the HS samples were of different fractional composition, including HA, FA, and non-fractionated HA and FA. This enabled substantial variation of structural characteristics within the test set (Table 2).
The data on elemental composition show that the highest amount of nitrogen was characteristic of soil humic acids and fulvic acids: the C/N ratio values varied between 11 and 20, whereas for peat HA and HF they varied from 26 to 50, reaching absolute maximum for peat FA above 100. This is indicative of the minimal content of nitrogen in peat FA. These humic materials were characterized with maximum values of the O/C ratio and the content of aliphatic units, for both carbohydrates and non-oxidized alkyl chains. They had minimum ΣCAr/ΣCAlk values (around 0.5), which is indicative of the highly aliphatic nature of these humic materials. The most aromatic humic materials with the lowest H/C ratio (0.6–0.8) and the highest CAr values were humic acids from Mollisol and coal.
The magnitude of the H/C ratio, which indicates the degree of aromaticity or unsaturation [65], varied from 0.62 (prevalence of aromatic fragments) to 1.22 (prevalence of aliphatic fragments). On average, aromaticity increased in the following order: HLS < freshwater HS < peat HS < soil HS < coal HS. The same trend could be observed when using the ΣCAr/ΣCAlk ratio as an indicator of aromaticity (the higher this ratio, the higher the aromaticity). A statistically significant correlation (r = −0.64) between H/C and ΣCAr/ΣCAlk was calculated (Table S1).
The relative content of oxygen-bearing moieties, which can be indicated by the O/C ratio, or oxidation degree, was in the range of 0.31–0.66. The oxidation degree varied in the following descending order: freshwater HS > peat HS > soil HS > HLS > coal HS. Recently, polysaccharides as a source of novel potential antioxidants were reported [31,66,67,68,69,70]. Although carbohydrates do not possess strong antioxidant activity due to low numbers of aldehyde and ketone groups in their structure, a detailed study of the antioxidant activity of monosaccharides, oligosaccharides, and complex carbohydrates yielded non-zero AOC values for sugars from 0.001 to 0.022 μmol TE mg−1 [52]. That is why it was of interest to compare the content of oxygen-bearing fragments in HS from various sources (Figure 1).
Aquatic HS demonstrated the highest content of carboxylic groups, and coal HS had the lowest. Coal HS was characterized by its minimal content of carbohydrate fragments. In contrast, peat HS and HLS had the highest content of ΣCCarb. HLS showed the maximum content of methoxyl carbon, showing a high contribution of lignin fragments in the structure of these humic materials.
The obtained data on the elemental composition and carbon content of the structural fragments are in good agreement with those reported in the literature [65,71]. For instance, in a statistical evaluation of the elemental composition of HS by Rice and MacCarthy [65], the reported values for O/C and H/C of HA were 0.08–1.20 and 0.08–1.85, respectively, which coincide with the values of 0.31–0.60 and 0.62–1.10 found in our study (Table 2). For FA, O/C and H/C varied from 0.17 to 1.19 and from 0.77 to 2.13, while in our work the analogous values were 0.51–0.66 and 0.76–1.10. The distribution of carbon among major structural fragments of HS as measured by 13C NMR spectroscopy corroborated well with the reported data for soil HS [59,72], peat HS [73], freshwater HS [74], and coal humic materials [75].
In general, the obtained data demonstrate trends that are generally reported for HS derived from different natural resources. Coal-derived HA had the minimum O/C ratio; they were enriched with aromatic fragments ΣCAr [65,76]. HS extracted from freshwater were characterized by their maximal content of CCOO carboxylic groups due of high microbiological activity, resulting in oxidation of humic materials [74]. HS from peat had a high content of ΣCCarb carbohydrates [73] due to low intense biological turnover in peat, resulting in incomplete microbiological decomposition of hydrocarbon residues. So, it can be concluded that the HS samples used in this study were typical representatives of humic materials isolated from the corresponding environments.
Phenolic moieties are supposed to be the most important scavengers of peroxyl radicals in different antioxidant compounds, including HS [37,38,77,78]. They convert peroxyl radicals into hydroperoxides and are themselves converted into phenoxyl radicals. A direct estimation of TP in HS demonstrated a phenol content in the range of 0.51–3.53 μmol TE mg−1 (Table 2). The maximum was observed for peat HA (PHA-T698) and the minimum for soil HA (SHA-CTV94). In general, peat HS were characterized by higher phenol content than other preparations. For peat HS, the TP values were in the range of 1.239–3.353 μmol TE mg−1, while for soil, coal, and freshwater HS, they did not exceed 1.760, 2.348, and 2.504 μmol TE mg−1, respectively. For the HLS samples, the TP values were 1.381–1.481 μmol TE mg−1, close to those observed for soil HS (0.510–1.760 μmol TE mg−1). As a rule, the humic acid fractions were characterized by a higher content of phenols compared to the corresponding fulvic acids. In general, the values of this parameter for HS from different sources strongly overlap.
Thus, the analysis of structural characteristics showed that the studied HS had a wide variation in parameters that might determine their antioxidant activity.

3.2. Antioxidant Capacity of HS

The AOC values of the HS samples isolated from different environments used in this study are shown in Table 3.
The obtained data indicate that for terrestrial sources (peat, soil), fulvic acids were characterized by much higher AOC values compared to humic acid isolated from the same source. For example, for soil humic materials, the AOC value (μmol TE mg−1) was 0.75 for SHA-PW96 and 1.2 for SFA-PW96; for peat HS, the corresponding values were 1.63 for PHA-T798 and 2.25 for PFA-T798. The AOC value of the non-fractionated sample of PHF-T798 was 1.56 μmol TE mg−1, which is very close to that of PHA-T798. This might be connected to the prevalence of humic acid fractions in the non-fractionated samples of peat HS.
Since phenolic moieties are generally believed to be responsible for the antioxidant activity of HS [39,79,80], one would expect that among the studied samples, the maximum antioxidant capacity would be observed for peat HS. Indeed, the data in Table 3 show that the highest AOC values were measured for peat HS. However, ANOVA followed by Tukey’s HSD test demonstrated that some samples of soil FA, coal HA, and freshwater HA also fell within the homogeneous groups of peat HS formed according to AOC (Table S2), indicating the proximity of their values.
AOC values of HS (0.31–2.56 μmol TE mg−1) were close to those of ascorbic acid (2.32 μmol TE mg−1) and vitamin E (2.95 μmol TE mg−1), but much lower than those of the other antioxidants used in the study (Table 4).
The obtained ORAC values for HS were slightly lower compared to those determined with the Trolox equivalent antioxidant capacity (TEAC) approach [53]. For HLS-45 and HLS-70, the TEAC AOC values were 3.3 and 2.9 μmol mg−1, respectively, which are 3–5 times larger than the corresponding ORAC AOC values (Table 3). For water HS samples, IHSS standards SRHA, SFRA, and SRDOM, the TEAC AOC values were 3.0, 2.4, and 2.5 μmol mg−1, respectively, which is 1.5–2.0 times larger than the ORAC values (Table 3). Of importance is that the described inconsistency between the TEAC and ORAC values for the analyzed humic materials, with the former substantially exceeding the latter, was not observed with the individual antioxidants [52,81]. Hu and co-authors reported that monosaccharides, oligosaccharides, and complex carbohydrates showed ORAC values, but not TEAC values [52]. A similar trend of lower TEAC values compared to ORAC values was reported for the phenolic antioxidants oxyresveratrol, resveratrol, pinosylvin, and pterostilbene [81]. This inconsistency might be related to the more complex antioxidant behavior of HS than individual antioxidant compounds.
Several significant correlations were found between the AOC values and structural characteristics of HS (Table S1), including TP (r = 0.64), C/N (r = 0.65), CCH3O (r = −0.56), ΣCCarb (r = 0.46), CCHO (r = 0.46), and COCO (r = 0.39). All significant correlations between the AOC values and structural parameters of HS are shown in Figure 2.
As can be seen from the correlation dependencies presented in Figure 2, the antioxidant activity of humic materials against peroxyl radicals depended on the content of both phenolic and non-phenolic fragments in the structure of HS.

4. Discussion

The AOC values of HS against peroxyl radicals found in this study varied from 0.31 to 2.56 μmol TE mg−1. The high-end values are close to those of ascorbic acid (2.32 μmol TE mg−1) and vitamin E (2.95 μmol TE mg−1). This is in line with other studies on the antioxidant activity of HS and their comparison with other antioxidants [82,83,84]. Similar effects of ascorbic acid and HA were reported for wheat growth and attributed to their antioxidant activity [83]. Sakr and co-authors demonstrated a similar capability of ascorbic acid and HA to mitigate biotic stress in wheat plants induced by a pathogenic fungus, Fusarium graminearum [85]. Peat HS were shown to possess antioxidant activity close to that of vitamin E [23,84,86].
Significant correlations were found between the AOC values and structural characteristics of HS such as TP, C/N, CCH3O, ΣCCarb, CCHO, and COCO (Figure 2). The relationship between AOC and TP can be explained by the fact that a formal H atom donation from weak X–H bonds (where X = O, N, or S) is the major mechanism of peroxyl radical quenching. This mechanism relies on the ability of certain antioxidants (chain-breaking antioxidants) to donate H atoms to ROO [34]. This mechanism of action occurs for phenols and antioxidant compounds structurally related to phenols, such as aromatic amines, phenothiazines, pyridinols, pyrimidinols, and thiols [34].
The C/N ratio of humic materials is supposed to reflect a depletion of amino acid content along with a dominant accumulation of aromatic compounds with increasing HS age [2]. As a result, more transformed HS possess an aromatic backbone enriched with phenolic and quinonid units [87]. The calculated negative significant correlation r = −0.40 between C/N and the metoxyl carbon CCH3O content (Table S1), related to unaltered lignin fragments in the HS structure, is in line with this idea. Thus, a positive correlation between the AOC of HS and the C/N ratio most likely indicates a possible increase in the AOC of humic materials in the process of aging. The content of methoxyl fragments, in their turn, also negatively correlated (r = −0.56) with the AOC of HS (Table S1). Of particular interest is a direct correlation between AOC values and the content of anomeric aliphatic C atoms COCO and О-substituted methine groups CCHO, and the total content of carbohydrates ΣCCarb. Over the past decade, a variety of natural polysaccharides from functional and medicinal foods have attracted great interest due to their antioxidant functions, such as free radical scavenging [41], including peroxyl radicals [39,88]. One proposed model for the effects of free radical scavenging of carbohydrates is to subtract their anomeric hydrogen by free radicals and combine it to form a neutral molecule; then the generated alkoxyl radicals promote the intramolecular hydrogen abstraction reaction, which triggers the spirocyclization reaction to terminate the reactions of radical chains [40].
Some HS samples did not obey the general trends shown in Figure 2. The most dramatic examples can be seen for the AOC–TP and AOC–COCO pair content in soil HS, where similar TP or COCO values were related to very different AOC values. To define the important properties of HS that contribute the most to their antioxidant activity, a forward stepwise multiple linear regression model was developed. The AOC value (Table 3) was the dependent variable, while the structural parameters (Table 2) were independent variables. Among the variables studied, only four were responsible for the antioxidant capability of HS (Table 5).
A high coefficient of multiple determination for the developed model is illustrated in Figure 3.
The four structural parameters contributing the most to the AOC values of HS were C/N, CCHO, CCH3O, and TP. An increase in C/N, CCHO, and TP led to an increase in AOC, whereas an increase in the CCH3O content led to a decrease in AOC.
Free radical-scavenging antioxidants react with ROO peroxyl radicals by one of the following three reaction mechanisms [89]: hydrogen abstraction (hydrogen atom transfer (HAT)), radical adduct formation (RAF), or single electron transfer (SET). The relative importance of these reactions depends on the type of radicals, antioxidants, and microenvironment [35].
The ORAC method mostly involves the HAT reaction, where an antioxidant and a substrate (probe) compete for free radicals [81]. Thus, the observed direct correlation between AOC value and phenol content in HS aligns with the expected trends. Hydrogen atom transfer from phenol ArOH to peroxyl radical ROO is described as (2):
ArOH + ROO → ArO + ROOH
A similar direct relationship between AOC value and phenol content in HS was reported by numerous studies [37,53,56,90,91]. The decisive role of phenols in quenching peroxyl radicals was demonstrated for other natural antioxidants as well [39,92,93].
Though phenols are known to play a key role as chain-breaking antioxidants, mainly via the HAT mechanism [34,35], quenching of the radical by the SET mechanism can also be expected. The predominant quenching of the radical by HAT or SET depends on phenol dissociation. For neutral forms of phenols (e.g., piceatannol, resveratrol), SET is much slower than HAT, which is not the case for phenolate ions, where the SET process becomes preferable [37,89]:
ArO + ROO → ArO + ROO
A similar relationship is observed for several radicals where phenolate has much higher reactivity than their protonated forms [94]. Deprotonation of phenols is pronounced at pH > 7, at which point deprotonation of hydroxyl moieties generates semiquinone and quinone, facilitating electron transfer and boosting antioxidant activity [37,95]. Phenolate is 1000 or 1 million times more active than phenol because the H atom is heavier than the electron, and more difficult to transfer [96]. As a result, for anionic forms of phenols, the peroxyl radical scavenging activity takes place almost exclusively via the SET mechanism [89]. This phenomenon explains the observed negative value of the correlation coefficient between AOC and the carbon content of methoxyl group CCH3O (Table S1). The methoxyl group increases the pKa of the phenol group to higher values, making the phenol group less acidic and thus decreasing the concentration of phenols in phenolate form at a certain pH value [97]. As the methoxyl group content in HS decreases along with an increase in C/N ratio due to microbial transformation [2,98], the opposite relationships of AOC versus CCH3O and AOC versus C/N look rather logical.
The inverse relationship between AOC and the methoxyl group content at pH 7.4 obtained in this study is important, as methoxyl groups are usually reported to reduce the bond dissociation enthalpy (BDE) of the phenolic hydroxyl group and enhance the electron-donating ability of phenolic acids by reducing proton affinity and electron transfer enthalpy values. Altogether, these processes result in an increase in the antioxidant activity of methoxyl-substituted phenols, and an increase in methoxyl content brings about higher antioxidant activity of phenolic acids. At the same time, Chen and co-workers reported an increase in the antioxidant activity of phenolic acids along with an increase in methoxyl group content using a ferric ion reducing antioxidant power (FRAP) assay [99]. We believe that the observed inconsistency might result from a relatively high pH value used in the ORAC assay (pH 7.4) compared to other methods (pH 3.6 for FRAP assay).
Another important finding of this study is the observed significant contribution of carbohydrate moieties to the AOC value of HS against peroxyl radicals. Nominally, statistically significant direct correlations were obtained for AOC and the content of anomeric aliphatic C atoms (COCO) and О-substituted methine groups (CCHO; Table S1). Moreover, CCHO was selected as a predictive variable for AOC by forward stepwise regression (Table 5).
Although carbohydrates and polysaccharides represent an important and abundant class of free radical scavengers [6,31,34,66,67,68,69,70,100], these non-phenolic antioxidants are not considered to be strong antioxidants due to their low content of electron-donating aldehyde or ketone groups in the structure [52]. Besides, the observed antioxidant activity in this case is often related to impurities such as phenolic and/or protein components rather than to the carbohydrate moieties themselves [39,40,52].
However, Hernandez-Marin and Martínez, based on a theoretical study, concluded that direct scavenging of hydroxyl radicals OH was by carbohydrates Cm(H2O)n via the HAT mechanism [100]:
Cm(H2O)n + OH → CmHO(H2O)n−1 + H2O
There was no pronounced regioselectivity for the abstraction of the hydrogen atom. Therefore, HAT from both any C–H or O–H bond to the radical can be hypothesized. Direct scavenging of hydroperoxyl radical OOH by carbohydrates was considered less likely, since according to the modelling results, hydrogen atom abstraction was possible only from one position of a single studied carbohydrate (sucrose) [100]. Another proposed model is based on the observed increased antioxidant ability of carbohydrate polymers vs. monosaccharides. The authors speculated that the weak free radical scavenging activity of monosaccharides is due to the abstraction of anomeric hydrogen. The enhanced antioxidant activity of the polymers over the monomeric form may be due to the greater ease of abstraction of anomeric hydrogen from an internal glucose unit rather than from the reducing end [101]. However, further verification experiments to confirm the model have not been conducted [40]. Additionally, it is believed that the antioxidant activity of phenolic acids can be increased by modifying them with sugars. For instance, esterification of ferulic acid with arabinoxylan resulted in feruloyl arabinoxylo-oligosaccharide with much stronger antioxidant activity compared to free ferulic acid. The activity increased along with an increased number of sugar moieties; the glycosyl group by itself showed no activity [102].
Regardless of the mechanism determining the antioxidant activity of sugars, the results obtained in this study clearly demonstrate the dependence of the AOC of HS on the content of carbohydrate fragments in their structure.
The PCA performed for the selected structural parameters of HS to determine their antioxidant capability against peroxyl radicals resulted in a separation of all samples by the first two factors (Figure 4).
Figure 4 shows the PCA results, with 64.44% of the total variance explained by the properties determining the AOC of HS against hydroxyl radicals. The samples from the different natural sources built distinct clusters, except for aquatic HS. These HS were isolated from HLS, soil HS, and coal HS, but formed a group with peat HS. The latter most likely resulted from the fact that the Suwannee River, which was a source for the aquatic samples in this work, is a blackwater river rising in the Okefenokee Swamp [103]. The HLS were clustered and different from coal HS and aquatic HS in PC1 (41.72%) because of the high content of methoxyl carbon CCH3O derived from lignin. Aquatic and coal HS were clustered with negative values because of the high C/N values. Soil HS were clustered from peat HS in PC2 (22.72%) due to the lower content of О-substituted methine groups CCHO.
So, based on a statistical analysis of the obtained data, we believe that, apart from phenol fragments, the presence of non-phenolic moieties such as carbohydrates impart antioxidant activity to HS against peroxyl radicals. To verify this hypothesis, more experiments are required to assess the antioxidant activity of HS using biological models.

5. Conclusions

This is the first systematic study conducted on a broad set of humic materials of different origins and fractional compositions with respect to the quenching capability of HS against biologically relevant peroxyl radicals with the ORAC method. The findings demonstrate that humic materials derived from different environments possess distinct antioxidant activity against peroxyl radicals. The high measured AOC values reached those for ascorbic acid and vitamin E. They were observed for the most active representatives: peat fulvic acids enriched with both phenolic and sugar moieties. Four structural parameters of HS were identified as major contributors to antioxidant capacity: total phenol content, atomic C/N ratio, and content of O-substituted methine and methoxyl groups. The results obtained clearly demonstrate that the antioxidant activity of HS against peroxyl radicals depends on both phenolic and non-phenolic moieties in their structure. For the first time, the substantial contribution of sugar moieties to the AOC of humic materials is revealed. The possible mechanisms of these phenomena are discussed. A conclusion is made regarding the complex character of AOC dependence on HS structure due to the very complex molecular organization of these natural systems, consisting of both high- and low-molecular-weight compounds. Fine fractionation of HS beyond traditional separation into humic and fulvic acids is highly recommended in order to determine more consistent relationships between structure and antioxidant properties. Deeper insight into the molecular composition of HS using high-resolution Fourier-transform ion cyclotron mass spectrometry might also advance our understanding of the hidden relationships between structure and properties for such chemically complex molecular systems.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/polym13193262/s1: Table S1: Pearson correlation coefficients between physical–chemical HS properties and their AOC, Table S2: Homogeneous groups of HS used in the study according Tukey’s HSD test; variable AOC, μmol TE mg−1; p < 0.05.

Author Contributions

Conceptualization, N.A.K., M.V.Z. and I.V.P.; methodology, N.A.K. and O.I.K.; investigation, O.I.K. and A.I.K.; writing—original draft preparation, O.I.K. and N.A.K.; writing—review and editing, M.V.Z. and I.V.P.; visualization, N.A.K., O.I.K. and A.I.K.; funding acquisition, I.V.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 20-63-47070. and the federal budget of the Russian Federation (CIT&S no. 121041300098-7).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors express their deep gratitude to deceased Ilya V. Nikolaev for his help in conducting experiments to assess the antioxidant activity of humic substances. This research was performed according to the development program of the Interdisciplinary Scientific and Educational School of M.V. Lomonosov Moscow State University, “Future Planet and Global Environmental Change”.

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. Gaffney, J.S.; Marley, N.A.; Clarks, S.B. Humic and fulvic acids and organic colloidal materials in the environment. In Humic and Fulvic Acids: Isolation, Structure, and Environmental Role; Gaffney, J.S., Marley, N.A., Clarks, S.B., Eds.; American Chemical Society: Washington, DC, USA, 1996; pp. 2–16. [Google Scholar]
  2. Olk, D.C.; Bloom, P.R.; Perdue, E.M.; McKnight, D.M.; Chen, Y.; Farenhorst, A.; Senesi, N.; Chin, Y.-P.; Schmitt-Kopplin, P.; Hertkorn, N.; et al. Environmental and agricultural relevance of humic fractions extracted by alkali from soils and natural waters. J. Environ. Qual. 2019, 48, 217–232. [Google Scholar] [CrossRef]
  3. Quilty, J.R.; Cattle, S.R. Use and understanding of organic amendments in Australian agriculture: A review. Soil Res. 2011, 49, 1–26. [Google Scholar] [CrossRef]
  4. Field, J.A.; Cervantes, F.J.; van der Zee, F.P.; Lettinga, G. Role of quinones in the biodegradation of priority pollutants: A review. Water Sci. Technol. 2000, 42, 215–222. [Google Scholar] [CrossRef]
  5. Lehtonen, K.; Hanninen, K.; Ketola, M. Structurally bound lipids in peat humic acids. Org. Geochem. 2001, 32, 33–43. [Google Scholar] [CrossRef]
  6. van Trump, J.I.; Vega Fransheska, J.R.; Coates, J.D. Natural organic matter as global antennae for primary production. Astrobiology 2013, 3, 476–482. [Google Scholar] [CrossRef] [PubMed]
  7. Orlov, A.A.; Zherebker, A.; Eletskaya, A.A.; Chernikov, V.S.; Kozlovskaya, L.I.; Zhernov, Y.V.; Kostyukevich, Y.; Palyulin, V.A.; Nikolaev, E.N.; Osolodkin, D.I.; et al. Examination of molecular space and feasible structures of bioactive components of humic substances by FTICR MS data mining in ChEMBL database. Sci. Rep. 2019, 9, 12066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Stevenson, F.J. Humus Chemistry: Genesis, Composition, Reactions; John Wiley and Sons, Inc.: New York, NY, USA, 1994; 521p. [Google Scholar]
  9. De Melo, B.A.G.; Motta, F.L.; Santana, M.H.A. Humic acids: Structural properties and multiple functionalities for novel technological developments. Mat. Sci. Engineer. C 2016, 62, 967–974. [Google Scholar] [CrossRef] [PubMed]
  10. Jooné, G.K.; van Rensburg, C.E. An in vitro investigation of the anti-inflammatory properties of potassium humate. Inflammation 2004, 28, 169–174. [Google Scholar] [CrossRef]
  11. van Rensburg, C.E. The anti-inflammatory properties of humic substances: A mini review. Phytother. Res. 2015, 29, 791–795. [Google Scholar] [CrossRef] [Green Version]
  12. Winkler, J.; Ghosh, S. Therapeutic potential of fulvic acid in chronic inflammatory diseases and diabetes. J. Diabetes Res. 2018, 2018, 5391014. [Google Scholar] [CrossRef] [Green Version]
  13. Schols, D.; Wutzler, P.; Klöcking, R.; Helbig, B.; De Clercq, E. Selective inhibitory action of polyhydroxycarboxylates derived from phenolic compounds against human immunodeficiency virus replication. J. Acquir. Immune Defic. Syndr. 1991, 7, 677–685. [Google Scholar]
  14. Lu, F.J.; Tseng, S.N.; Li, M.L.; Shih, S.R. In vitro anti-influenza virus activity of synthetic humate analogues derived from protocatechuic acid. Arch. Virol. 2002, 147, 273–284. [Google Scholar] [CrossRef] [PubMed]
  15. Zhernov, Y.V.; Konstantinov, A.I.; Zherebker, A.; Nikolaev, E.; Orlov, A.; Savinykh, M.I.; Kornilaeva, G.V.; Karamov, E.V.; Perminova, I.V. Antiviral activity of natural humic substances and shilajit materials against HIV-1: Relation to structure. Environ. Res. 2021, 193, 110312. [Google Scholar] [CrossRef]
  16. Jurcsik, I. Possibilities of applying humic acids in medicine (wound healing and cancer therapy). In Humic Substances in the Global Environment; Senesi, N., Milano, T.M., Eds.; Elsevier: Amsterdam, The Netherlands; London, UK; New York, NY, USA; Tokyo, Japan, 1994; pp. 1331–1336. [Google Scholar]
  17. Ji, Y.; Zhang, A.; Chen, X.; Che, X.; Zhou, K.; Wang, Z. Sodium humate accelerates cutaneous wound healing by activating TGF-β/Smads signaling pathway in rats. Acta Pharm. Sin. B 2016, 6, 132–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Litvin, V.A.; Minaev, B.F. Spectroscopy study of silver nanoparticles fabrication using synthetic humic substances and their antimicrobial activity. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 108, 115–122. [Google Scholar] [CrossRef] [PubMed]
  19. Krempaská, K.; Vaško, L.; Vašková, J. Humic acids as therapeutic compounds in lead intoxication. Curr. Clin. Pharmacol. 2016, 11, 159–167. [Google Scholar] [CrossRef]
  20. Ozkan, A.; Sen, H.M.; Sehitoglu, I.; Alacam, H.; Guven, M.; Aras, A.B.; Akman, T.; Silan, C.; Cosar, M.; Karaman, H.I.O. Neuroprotective effect of humic acid on focal cerebral ischemia injury: An experimental study in rats. Inflammation 2015, 38, 32–39. [Google Scholar] [CrossRef]
  21. Zykova, M.V.; Schepetkin, I.A.; Belousov, M.V.; Krivoshchekov, S.V.; Logvinova, L.A.; Bratishko, K.A.; Yusubov, M.S.; Romanenko, S.V.; Quinn, M.T. Physicochemical characterization and antioxidant activity of humic acids isolated from peat of various origins. Molecules 2018, 23, 753. [Google Scholar] [CrossRef] [Green Version]
  22. Dell’Anno, M.; Hejna, M.; Sotira, S.; Caprarulo, V.; Reggi, S.; Pilu, R.; Miragoli, F.; Callegari, M.L.; Panseri, S.; Rossi, L. Evaluation of leonardite as a feed additive on lipid metabolism and growth of weaned piglets. Anim. Feed Sci. Technol. 2020, 266, 114519. [Google Scholar] [CrossRef]
  23. Piotrowska, D.; Dlugosz, A.; Witkiewicz, K.; Pak, J. The research on the antioxidant properties of Tolpa peat preparation and its fractions. Acta Pol. Pharm. 2000, 57, 127–129. [Google Scholar]
  24. Vetvicka, V.; Vashishta, A.; Fuentes, M.; Baigorri, R.; Jose, M.G.-M.; Yvin, J.-C. The relative abundance of oxygen alkyl-related groups in aliphatic domains is involved in the main pharmacological-pleiotropic effects of humic acids. J. Med. Food 2013, 16, 625–632. [Google Scholar] [CrossRef] [Green Version]
  25. Maslinski, C.; Fogel, W.A.; Andjewski, W. The influence of Tolpa peat preparation on rat liver regeneration. Acta Pol. Pharm. 1995, 50, 413–416. [Google Scholar]
  26. Belousov, M.V.; Akhmedzhanov, R.R.; Zykova, M.V.; Gur’ev, A.M.; Yusubov, M.S. Hepatoprotective properties of native humic acids isolated from lowland peat of Tomsk. Pharm. Chem. J. 2014, 48, 249–252. [Google Scholar] [CrossRef]
  27. Akbas, A.; Silan, C.; Gulpinar, M.T.; Sancak, E.B.; Ozkanli, S.S.; Cakir, D.U. Renoprotective effect of humic acid on renal ischemia-reperfusion injury: An experimental study in rats. Inflammation 2015, 38, 2042–2048. [Google Scholar] [CrossRef] [PubMed]
  28. Wollina, U. The response of Erythematous rosacea to ondasentron. Br. J. Dermatol. 1999, 140, 561–562. [Google Scholar] [CrossRef] [PubMed]
  29. Lasukova, T.V.; Zykova, M.V.; Belousov, M.V.; Logvinova, L.A.; Dygai, A.M. The role of NO synthase in the cardioprotective effect of substances of humic origin on the model of ischemia and reperfusion of isolated rat heart. Bull. Exp. Biol. Med. 2019, 166, 598–601. [Google Scholar] [CrossRef]
  30. Fang, Y.Z.; Yang, S.; Wu, G. Free radicals, antioxidants, and nutrition. Nutrition 2002, 18, 872–879. [Google Scholar] [CrossRef]
  31. Kim, E.-A.; Lee, S.-H.; Ko, C.; Cha, S.-H.; Kang, M.-C.; Kang, S.-M.; Ko, S.-C.; Lee, W.-W.; Ko, J.-Y.; Lee, J.-H.; et al. Protective effect of fucoidan against AAPH-induced oxidative stress in zebrafish model. Carbohydr. Polym. 2014, 102, 185–191. [Google Scholar] [CrossRef]
  32. Whayne, F.T.; Saha, P.S.; Mukherjee, D. Antioxidants in the practice of medicine; what should the clinician know? Cardiovasc. Hematol. Disord. Drug Targets 2016, 16, 13–20. [Google Scholar] [CrossRef]
  33. Ruberto, G.; Baratta, M.T.; Biondi, D.M.; Amico, V. Antioxidant activity of extracts of the marine algal genus Cystoseira in a micellar model system. J. Appl. Phycol. 2001, 13, 403–407. [Google Scholar] [CrossRef]
  34. Foti, M.C.; Amorati, R. Non-phenolic radical-trapping antioxidants. J. Pharm. Pharmacol. 2009, 61, 1435–1448. [Google Scholar] [CrossRef]
  35. Niki, E. Role of vitamin E as a lipid-soluble peroxyl radical scavenger: In vitro and in vivo evidence. Free Radic. Biol. Med. 2014, 66, 3–12. [Google Scholar] [CrossRef]
  36. Prior, R.L. Oxygen radical absorbance capacity (ORAC): New horizons in relating dietary antioxidants/bioactives and health benefits. J. Funct. Foods 2015, 18, 797–810. [Google Scholar] [CrossRef]
  37. Aeschbacher, M.; Graf, C.; Schwarzenbach, R.P.; Sander, M. Antioxidant properties of humic substances. Environ. Sci. Technol. 2012, 46, 4916–4925. [Google Scholar] [CrossRef] [PubMed]
  38. Masuda, T.; Yamada, K.; Maekawa, T.; Takeda, Y.; Yamaguchi, H. Antioxidant mechanism studies on ferulic acid: Identification of oxidative coupling products from methyl ferulate and linoleate. J. Agric. Food Chem. 2006, 54, 6069–6074. [Google Scholar] [CrossRef] [PubMed]
  39. Asikin, Y.; Takahashi, M.; Mishima, T.; Mizu, M.; Takara, K.; Wada, K. Antioxidant activity of sugarcane molasses against 2,2’-azobis(2-amidinopropane) dihydrochloride-induced peroxyl radicals. Food Chem. 2013, 141, 466–472. [Google Scholar] [CrossRef]
  40. Wang, J.; Hu, S.; Nie, S.; Yu, Q.; Xie, M. Reviews on mechanisms of in vitro antioxidant activity of polysaccharides. Oxid. Med. Cell. Longev. 2016, 2016, 5692852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Li, H.; Ding, F.; Xiao, L.; Shi, R.; Wang, H.; Han, W.; Huang, Z. Food-derived antioxidant polysaccharides and their pharmacological potential in neurodegenerative diseases. Nutrients 2017, 9, 778. [Google Scholar] [CrossRef] [Green Version]
  42. Kulikova, N.A.; Perminova, I.V. Interactions between humic substances and microorganisms and their implications for nature-like bioremediation technologies. Molecules 2021, 26, 2706. [Google Scholar] [CrossRef] [PubMed]
  43. Prior, R.L.; Wu, X.; Schaich, K. Standardized methods for the determination of antioxidant capacity and phenolics in foods and dietary supplements. J. Agric. Food Chem. 2005, 53, 4290–4302. [Google Scholar] [CrossRef]
  44. Somogyi, A.; Rosta, K.; Pusztai, P.; Tulassay, Z.; Nagy, G. Antioxidant measurements. Physiol. Meas. 2007, 28, R41. [Google Scholar] [CrossRef]
  45. Frankel, E.N.; Finley, J.W. How to standardize the multiplicity of methods to evaluate natural antioxidants. J. Agric. Food Chem. 2008, 56, 4901–4908. [Google Scholar] [CrossRef]
  46. Munteanu, I.G.; Apetrei, C. Analytical methods used in determining antioxidant activity: A review. Int. J. Mol. Sci. 2021, 22, 3380. [Google Scholar] [CrossRef]
  47. Ou, B.; Hampsch-Woodill, M.; Prior, R.L. Development and validation of an improved oxygen radical absorbance capacity assay using fluorescein as the fluorescent probe. J. Agric. Food Chem. 2001, 49, 4619–4620. [Google Scholar] [CrossRef] [PubMed]
  48. Dorta, E.; Fuentes-Lemus, E.; Aspée, A.; Atala, E.; Speisky, H.; Bridi, R.; Lissi, E.; López-Alarcón, C. The ORAC (oxygen radical absorbance capacity) index does not reflect the capacity of antioxidants to trap peroxyl radicals. RSC Adv. 2015, 5, 39899–39902. [Google Scholar] [CrossRef]
  49. Amić, A.; Lučić, B.; Stepanić, V.; Marković, Z.; Marković, S.; Dimitrić Marković, J.M.; Amić, D. Free radical scavenging potency of quercetin catecholic colonic metabolites: Thermodynamics of 2H+/2e processes. Food Chem. 2017, 218, 144–151. [Google Scholar] [CrossRef] [PubMed]
  50. Huang, D.; Ou, B.; Prior, R.L. The chemistry behind antioxidant capacity assays. J. Agric. Food Chem. 2005, 53, 1841–1856. [Google Scholar] [CrossRef]
  51. Di Meo, F.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Duroux, J.-L.; Olivier, Y.; Trouillas, P. Free radical scavenging by natural polyphenols: Atom versus electron transfer. J. Phys. Chem. A 2013, 117, 2082–2092. [Google Scholar] [CrossRef]
  52. Hu, S.; Yin, J.; Nie, S.; Wang, J.; Phillips, G.O.; Xie, M.; Cui, S.W. In vitro evaluation of the antioxidant activities of carbohydrates. Bioact. Carbohydr. Diet. Fibre 2016, 7, 19–27. [Google Scholar] [CrossRef]
  53. Klein, O.I.; Kulikova, N.A.; Filimonov, I.S.; Koroleva, O.V.; Konstantinov, A.I. Long-term kinetics study and quantitative characterization of the antioxidant capacities of humic and humic-like substances. J. Soils Sediments 2016, 18, 1355–1364. [Google Scholar] [CrossRef]
  54. Vašková, J.; Veliká, B.; Pilátová, M.; Kron, I.; Vaško, L. Effects of humic acids in vitro. In Vitro Cell. Dev. Biol. Anim. 2011, 47, 376–382. [Google Scholar] [CrossRef]
  55. Tarasova, A.S.; Stom, D.I.; Kudryasheva, N.S. Antioxidant activity of humic substances via bioluminescent monitoring in vitro. Environ. Monit. Assess. 2015, 187, 89. [Google Scholar] [CrossRef]
  56. Karadirek, Ş.; Kanmaz, N.; Balta, Z.; Demirçivi, P.; Üzer, A.; Hızal, J.; Apak, R. Determination of total antioxidant capacity of humic acids using CUPRAC, Folin–Ciocalteu, noble metal nanoparticle- and solid–liquid extraction-based methods. Talanta 2016, 153, 120–129. [Google Scholar] [CrossRef]
  57. Avvakumova, N.P.; Gerchikov, A.Y.; Khairullina, V.R.; Zhdanova, A.V. Antioxidant properties of humic substances isolated from peloids. Pharm. Chem. J. 2011, 45, 192. [Google Scholar] [CrossRef]
  58. Khil’ko, S.L.; Efimova, I.V.; Smirnova, O.V. Antioxidant properties of humic acids from brown coal. Solid Fuel Chem. 2011, 45, 367–371. [Google Scholar] [CrossRef]
  59. Kholodov, V.A.; Konstantinov, A.I.; Kudryavtsev, A.V.; Perminova, I.V. Structure of humic acids in zonal soils from 13C NMR data. Euras. Soil Sci. 2011, 44, 976–983. [Google Scholar] [CrossRef]
  60. Lowe, L.E. Studies on the nature of sulfur in peat humic acids from Frazer River delta, British Columbia. Sci. Total Environ. 1992, 113, 133–145. [Google Scholar] [CrossRef]
  61. Koroleva, O.V.; Kulikova, N.A.; Alekseeva, T.N.; Stepanova, E.V.; Davidchik, V.N.; Belyaeva, E.Y.; Tsvetkova, E.A. A comparative characterization of fungal melanin and the humin-like substances synthesized by Cerrena maxima 0275. Appl. Biochem. Microbiol. 2007, 43, 61–67. [Google Scholar] [CrossRef]
  62. Hertkorn, N.; Claus, H.; Schmitt-Kopplin, P.; Perdue, E.M.; Filip, Z. Utilization and transformation of aquatic humic substances by autochthonous microorganisms. Environ. Sci. Technol. 2002, 36, 4334–4345. [Google Scholar] [CrossRef] [PubMed]
  63. Singleton, V.L.; Rossi, J. Colorimetry of total phenolics with phospohmolibdic-phosphotungstic acid reagents. Am. J. Enol. Viticult. 1965, 16, 144–158. [Google Scholar]
  64. Prior, R.L.; Hoang, H.; Gu, L.; Wu, X.; Bacchiocca, M.; Howard, L.; Hampsch-Woodill, M.; Huang, D.; Ou, B.; Jacob, R. Assays for hydrophilic and lipophilic antioxidant capacity (oxygen radical absorbance (ORACFL)) of plasma and other biological and food samples. J. Agric. Food Chem. 2003, 51, 3273–3279. [Google Scholar] [CrossRef]
  65. Rice, J.A.; MacCarthy, P. Statistical evaluation of the elemental composition of humic substances. Org. Geochem. 1991, 17, 635–648. [Google Scholar] [CrossRef]
  66. Barahona, T.; Chandía, N.P.; Encinas, M.V.; Matsuhiro, B.; Zúñiga, E.A. Antioxidant capacity of sulfated polysaccharides from seaweeds. A kinetic approach. Food Hydrocoll. 2011, 25, 529–535. [Google Scholar] [CrossRef]
  67. Petera, B.; Delattre, C.; Pierre, G.; Wadouachi, A.; Elboutachfaiti, R.; Engel, E.; Poughon, L.; Michaud, P.; Fenoradosoa, T.A. Characterization of arabinogalactan-rich mucilage from Cereus triangularis cladodes. Carbohydr. Polym. 2015, 127, 372–380. [Google Scholar] [CrossRef] [PubMed]
  68. Ferreira, I.C.F.R.; Heleno, S.A.; Reis, F.S.; Stojkovic, D.; Queiroz, M.J.R.P.; Vasconcelos, M.H.; Sokovic, M. Chemical features of Ganoderma polysaccharides with antioxidant, antitumor and antimicrobial activities. Phytochemistry 2015, 114, 38–55. [Google Scholar] [CrossRef] [Green Version]
  69. Zhu, W.; Xue, X.; Zhang, Z. Structural, physicochemical, antioxidant and antitumor property of an acidic polysaccharide from Polygonum multiflorum. Int. J. Biol. Macromol. 2017, 96, 494–500. [Google Scholar] [CrossRef]
  70. Mooradian, A.D.; Haas, M.J.; Onstead-Haas, L.; Tani, Y.; Iida, T.; Tokuda, M. Naturally occurring rare sugars are free radical scavengers and can ameliorate endoplasmic reticulum stress. Int. J. Vitam. Nutr. Res. 2020, 90, 210–220. [Google Scholar] [CrossRef] [PubMed]
  71. García, A.C.; de Souza, L.G.A.; Pereira, M.G.; Castro, R.N.; García-Mina, J.M.; Zonta, E.; Lisboa, F.G.J.; Berbara, R.L.L. Structure-property-function relationship in humic substances to explain the biological activity in plants. Sci. Rep. 2016, 6, 20798. [Google Scholar] [CrossRef] [Green Version]
  72. Ussiri, D.A.; Johnson, C.E. Characterization of organic matter in a northern hardwood forest soil by 13C NMR spectroscopy and chemical methods. Geoderma 2003, 111, 123–149. [Google Scholar] [CrossRef]
  73. Fong, S.S.; Mohamed, M. Chemical characterization of humic substances occurring in the peats of Sarawak, Malaysia. Org. Geochem. 2007, 38, 967–976. [Google Scholar] [CrossRef]
  74. Wilson, M.A.; Barron, P.F.; Gillam, A.H. The structure of freshwater humic substances as revealed by 13C-NMR spectroscopy. Geochim. Cosmochim. Acta 1981, 45, 1743–1750. [Google Scholar] [CrossRef]
  75. Yan, S.; Zhang, N.; Li, J.; Wang, Y.; Liu, Y.; Cao, M.; Yan, Q. Characterization of humic acids from original coal and its oxidization production. Sci. Rep. 2021, 11, 15381. [Google Scholar] [CrossRef]
  76. Kurková, M.; Klika, Z.; Kliková, C.; Havel, J. Humic acids from oxidized coals. Chemosphere 2004, 54, 1237–1245. [Google Scholar] [CrossRef]
  77. Rice-Evans, C.A.; Miller, N.J.; Paganga, G. Structure antioxidant activity relationships of flavonoids and phenolic acids. Free Radic. Biol. Med. 1996, 20, 933–956. [Google Scholar] [CrossRef]
  78. Gijsman, P. Polymer Stabilization. In Handbook of Environmental Degradation of Materials, 2nd ed.; Elsevier Inc.: Amsterdam, The Netherland, 2012; pp. 673–714. [Google Scholar] [CrossRef]
  79. Rimmer, D.L.; Smith, A.M. Antioxidants in soil organic matter and in associated plant materials. Eur. J. Soil Sci. 2009, 60, 170–175. [Google Scholar] [CrossRef]
  80. Rimmer, D.L.; Abbott, G.D. Phenolic compounds in NaOH extracts of UK soils and their contribution to antioxidant capacity. Eur. J. Soil Sci. 2011, 62, 285–294. [Google Scholar] [CrossRef]
  81. Rodríguez-Bonilla, P.; Gandía-Herrero, F.; Matencio, A.; García-Carmona, F.; López-Nicolás, J.M. Comparative study of the antioxidant capacity of four stilbenes using ORAC, ABTS+, and FRAP techniques. Food Anal. Methods 2017, 10, 2994–3000. [Google Scholar] [CrossRef]
  82. Korotkova, E.I.; Karbainov, Y.A.; Avramchik, O.A. Investigation of antioxidant and catalytic properties of some biologically active substances by voltammetry. Anal. Bioanal. Chem. 2003, 375, 465–468. [Google Scholar] [CrossRef] [PubMed]
  83. Hegab, R.H.; Fawy, H.A.; Habib, A.A. Evaluates effect of amino acids, humic acid and antioxidants as foliar application on the biochemical content and productivity of wheat under North Sinai soils conditions. Am. J. Agric. For. 2020, 8, 167–174. [Google Scholar] [CrossRef]
  84. Bratishko, K.A.; Zykova, M.V.; Ivanov, V.V.; Perminova, I.V.; Belousov, M.V. Peat humic acids-prospective biologically active substances with antioxidant activity for the development of protective agents. Khimiya Rastitel’nogo Syr’ya 2021, 1, 287–298. [Google Scholar] [CrossRef]
  85. Sakr, M.T.; Sarkassy, N.M.; Fuller, M.P. Exogenously applied antioxidants and biostimulants counteract the adverse effect of biotic stress in wheat plant. Agric. Res. Technol. Open Access J. 2017, 12, 555853. [Google Scholar] [CrossRef]
  86. Drobnik, J.; Stebel, A. Central European ethnomedical and officinal uses of peat, with special emphasis on the Tołpa peat preparation (TPP): An historical review. J. Ethnopharmac. 2019, 112248. [Google Scholar] [CrossRef] [PubMed]
  87. Jouraiphy, A.; Amir, S.; Winterton, P.; El Gharous, M.; Revel, J.-C.; Hafidi, M. Structural study of the fulvic fraction during composting of activated sludge-plant matter: Elemental analysis, FTIR and 13C NMR. Bioresour. Technol. 2008, 99, 1066–1072. [Google Scholar] [CrossRef]
  88. Wang, J.; Sun, B.; Cao, Y.; Tian, Y. Protection of wheat bran feruloyl oligosaccharides against free radical-induced oxidative damage in normal human erythrocytes. Food Chem. Toxicol. 2009, 47, 1591–1599. [Google Scholar] [CrossRef] [PubMed]
  89. Cordova-Gomez, M.; Galano, A.; Alvarez-Idaboy, J.R. Piceatannol, a better peroxyl radical scavenger than resveratrol. RSC Adv. 2013, 3, 20209. [Google Scholar] [CrossRef]
  90. Efimova, I.V.; Khil’ko, S.L.; Smirnova, O.V. Antioxidant activity of humic acids in radical-chain oxidation processes. Russ. J. Appl. Chem. 2020, 85, 1351–1354. [Google Scholar] [CrossRef]
  91. Catalá, T.S.; Rossel, P.E.; Álvarez-Gómez, F.; Tebben, J.; Figueroa, F.L.; Dittmar, T. Antioxidant activity and phenolic content of marine dissolved organic matter and their relation to molecular composition. Front. Mar. Sci. 2020, 7, 603447. [Google Scholar] [CrossRef]
  92. Chandrasekara, A.; Shahidi, F. Content of insoluble bound phenolics in millets and their contribution to antioxidant capacity. J. Agric. Food Chem. 2010, 58, 6706–6714. [Google Scholar] [CrossRef] [PubMed]
  93. Fu, L.; Xu, B.-T.; Xu, X.-R.; Qin, X.-S.; Gan, R.-Y.; Li, H.-B. Antioxidant capacities and total phenolic contents of 56 wild fruits from South China. Molecules 2010, 15, 8602–8617. [Google Scholar] [CrossRef] [Green Version]
  94. Aspée, A.; Aliaga, C.; Maretti, L.; Zúñiga-Núñez, D.; Godoy, J.; Pino, E.; Cárdenas-Jirón, G.; Lopez-Alarcon, C.; Scaiano, J.C.; Alarcon, E.I. Reaction kinetics of phenolic antioxidants toward photoinduced pyranine free radicals in biological models. J. Phys. Chem. B 2017, 121, 6331–6340. [Google Scholar] [CrossRef]
  95. Jha, N.S.; Mishra, S.; Jha, S.K.; Surolia, A. Antioxidant activity and electrochemical elucidation of the enigmatic redox behavior of curcumin and its structurally modified analogues. Electrochim. Acta 2015, 151, 574–583. [Google Scholar] [CrossRef]
  96. Laguerre, M.; Lecomte, J.; Villeneuvene, P. The physico-chemical basis of phenolic antioxidant activity. Lipid Technol. 2014, 26, 59–62. [Google Scholar] [CrossRef]
  97. Acid Strength and pKa. Available online: https://chem.libretexts.org/@go/page/225794 (accessed on 19 August 2021).
  98. Mayans, B.; Pérez-Esteban, J.; Escolástico, C.; Eymar, E.; Masaguer, A. Evaluation of commercial humic substances and other organic amendments for the immobilization of copper through 13C CPMAS NMR, FT-IR, and DSC analyses. Agronomy 2019, 9, 762. [Google Scholar] [CrossRef] [Green Version]
  99. Chen, J.; Yang, J.; Ma, L.; Li, J.; Shahzad, N.; Kim, C.K. Structure-antioxidant activity relationship of methoxy, phenolic hydroxyl, and carboxylic acid groups of phenolic acids. Sci. Rep. 2020, 10, 5666. [Google Scholar] [CrossRef] [Green Version]
  100. Hernandez-Marin, E.; Martínez, A. Carbohydrates and their free radical scavenging capability: A theoretical study. J. Phys. Chem. B 2012, 116, 9668–9675. [Google Scholar] [CrossRef] [PubMed]
  101. Tsiapali, E.; Whaley, S.; Kalbfleisch, J.; Ensley, H.E.; Browder, I.W.; Williams, D.L. Glucans exhibit weak antioxidant activity, but stimulate macrophage free radical activity. Free Radic. Biol. Med. 2001, 30, 393–402. [Google Scholar] [CrossRef]
  102. Shyama Prasad Rao, R.; Muralikrishna, G. Water soluble feruloyl arabinoxylans from rice and ragi: Changes upon malting and their consequence on antioxidant activity. Phytochemistry 2006, 67, 91–99. [Google Scholar] [CrossRef]
  103. Source Materials for IHSS Samples. Available online: https://humic-substances.org/source-materials-for-ihss-samples/ (accessed on 19 August 2021).
Figure 1. Percentage of O-containing structural fragments in humic materials used in this study as measured by 13C NMR spectroscopy.
Figure 1. Percentage of O-containing structural fragments in humic materials used in this study as measured by 13C NMR spectroscopy.
Polymers 13 03262 g001
Figure 2. Correlation fields between AOC values of HS and their physical–chemical properties: (a) TP; (b) C/N ratio; (c) CCH3O; (d) ΣCCarb; (e) CCHO; (f) COCO.
Figure 2. Correlation fields between AOC values of HS and their physical–chemical properties: (a) TP; (b) C/N ratio; (c) CCH3O; (d) ΣCCarb; (e) CCHO; (f) COCO.
Polymers 13 03262 g002
Figure 3. Relationship between measured and calculated AOC values of studied HS using a developed forward stepwise multiple linear regression model.
Figure 3. Relationship between measured and calculated AOC values of studied HS using a developed forward stepwise multiple linear regression model.
Polymers 13 03262 g003
Figure 4. Projection of cases on a factor plane: PCA result for classification of HS used in this study based on their physical–chemical properties, selected after forward stepwise regression (C/N, content of CCHO, CCH3O, TP).
Figure 4. Projection of cases on a factor plane: PCA result for classification of HS used in this study based on their physical–chemical properties, selected after forward stepwise regression (C/N, content of CCHO, CCH3O, TP).
Polymers 13 03262 g004
Table 1. Humic substances used in the study and their sources.
Table 1. Humic substances used in the study and their sources.
HS IndexHA TypeSource
Soil
SHA-PG96HAAlbic Retisol
SHA-PW96HAAlbic Retisol
SHA-PG94HAAlbic Retisol
SHA-CM94HAGleyic Chernozem
SHA-PW94HAAlbic Retisol
SHA-CTV94HAEndocalcic Chernozem
SFA-PP96FAAlbic Retisol
SFA-PW96FAAlbic Retisol
SFA-PG96FAAlbic Retisol
Peat
PHA-T398HALowland peat, sedge type
PHA-T798HALowland peat, woody type
PHA-T698HAHighland peat, sedge type
PFA-T398FALowland peat, sedge type
PFA-T698FAHighland peat, sedge type
PFA-T798FALowland peat, woody type
PHF-T398HA+FALowland peat, sedge type
PHF-T698HA+FAHighland peat, sedge type
PHF-T798HA+FALowland peat, woody type
Natural water
SRHAHASuwannee River (IHSS standard)
SRFAFASuwannee River (IHSS standard)
SRDOMDOMSuwannee River (IHSS standard)
Coal
СHA-AGKHABiotechnology Ltd. (RF)
CHA-ALDHAAldrich (Germany)
Humic-like substances
HLS-45HAOat straw solubilized by Trametes maxima 0275
HLS-70HAOat straw solubilized by Trametes maxima 0275
Table 2. Structural characteristics of humic materials used in this study.
Table 2. Structural characteristics of humic materials used in this study.
HS IndexAtomic Ratio 1Content of Carbon in Structural Fragments Determined by 13CNMR Spectroscopy as Integral Intensity (%) 2TP 3, μmol TE mg−1
O/CH/CC/NCC=OCCOOСArOСArCOCOCCHOCCH2OCCH3OCCHn∑СAr∑CСarb∑CAlk∑CAlk-O∑CAr/∑CAlk
Soil
SHA-PG960.480.9315.92.017.011.031.02.010.02.04.022.042.014.040.018.01.051.378 ± 0.001
SHA-PW960.461.1014.04.019.09.025.02.08.02.04.026.034.012.042.016.00.811.230 ± 0.005
SHA-PG940.451.0013.61.115.212.832.83.010.02.15.817.545.615.138.420.91.191.392 ± 0.001
SHA-CM940.340.6613.81.514.38.748.12.06.00.84.614.256.88.827.613.42.061.760 ± 0.001
SHA-PW940.460.9211.51.017.013.033.02.08.02.55.517.046.012.535.018.01.311.753 ± 0.001
SHA-CTV940.420.6215.93.015.011.043.03.07.02.03.013.054.012.028.015.01.930.510 ± 0.001
SFA-PP960.540.9111.32.023.07.019.03.015.03.04.022.026.021.047.025.00.551.167 ± 0.001
SFA-PW960.580.9417.24.019.59.823.52.010.02.53.525.133.314.543.118.00.771.523 ± 0.001
SFA-PG960.610.8819.43.118.013.128.22.011.02.44.217.641.315.437.219.61.111.176 ± 0.001
Peat
PHA-T3980.440.8725.98.715.57.432.47.44.42.04.118.239.713.836.117.91.102.743 ± 0.001
PHA-T7980.490.8746.24.111.09.431.73.714.22.73.419.841.120.643.824.00.942.986 ± 0.001
PHA-T6980.550.9137.91.78.67.235.42.612.32.74.025.742.517.647.321.60.903.353 ± 0.001
PFA-T3980.660.761203.117.09.028.14.911.82.33.820.037.119.042.822.80.871.239 ± 0.001
PFA-T6980.511.031011.610.96.924.46.926.13.11.818.331.336.156.237.90.562.368 ± 0.001
PFA-T7980.601.0074.12.211.810.924.09.624.13.5211.934.937.251.139.20.681.602 ± 0.001
PHF-T3980.491.1035.53.410.69.832.96.711.03.05.616.942.720.743.226.30.992.398 ± 0.001
PHF-T6980.540.9148.53.710.98.424.47.719.84.84.216.232.832.352.736.50.622.275 ± 0.001
PHF-T7980480.8753.03.510.48.728.18.815.23.54.417.436.827.549.331.90.752.275 ± 0.001
Natural water
SRHA0.600.9752.58.518.110.726.66.38.91.93.415.637.317.136.120.51.032.504 ± 0.001
SRFA0.620.9991.17.419.610.922.34.410.01.84.019.633.216.239.820.20.831.537 ± 0.001
SRDOM0.610.9555.66.919.98.023.75.211.32.04.019.031.818.541.522.50.761.791 ± 0.001
Coal
СHA-AGK0.320.791120.516.910.047.8001.03.020.857.81.024.84.02.332.348 ± 0.001
CHA-ALD0.310.8177.01.015.013.043.0001.06.021.056.01.028.07.02.001.521 ± 0.001
Humic-like substances
HLS-450.371.2216.13.810.911.225.66.813.74.48.415.136.824.948.433.30.761.481 ± 0.003
HLS-700.551.0323.74.015.110.926.84.812.74.29.012.537.621.743.230.70.871.381 ± 0.001
1 H/C, O/C and N/C ratios were calculated on an ash- and water-free basis. 2 Content of carbon in structural fragments was determined by 13C NMR spectroscopy as integral intensity (%) of the following spectral regions (ppm): 220–189 (CC=O), 189–168 (CCOO), 168–145 (CAr-O), 145–108 (CAr), 108–91 (COCHO), 91–66 (CCHO), 66–59 (CCH2O), 59–48 (CCH3O), 48–0 (CCHn). 3 Values are means ± SD (n = 3).
Table 3. AOC values of HS samples from different environments used in this study as measured by the ORAC method.
Table 3. AOC values of HS samples from different environments used in this study as measured by the ORAC method.
HS IndexAOC, μmol TE mg−1
Soil
SHA-PG960.87 ± 0.08 bc
SHA-PW960.74 ± 0.04 bc
SHA-PG940.75 ± 0.03 bc
SHA-CM940.97 ± 0.05 d
SHA-PW941.04 ± 0.05 de
SHA-CTV940.31 ± 0.02 a
SFA-PP961.08 ± 0.06 def
SFA-PW961.20 ± 0.06 efg
SFA-PG961.39 ± 0.06 hi
Peat
PHA-T3981.55 ± 0.03 hij
PHA-T7981.63 ± 0.05 jk
PHA-T6981.75 ± 0.03 kl
PFA-T3981.56 ± 0.08 hij
PFA-T6982.56 ± 0.06 m
PFA-T7982.25 ± 0.13 m
PHF-T3981.38 ± 0.03 gh
PHF-T6981.44 ± 0.05 hi
PHF-T7981.56 ± 0.08 ij
Natural water
SRHA2.07 ± 0.09 m
SRFA1.22 ± 0.04 fg
SRDOM1.41 ± 0.05 hi
Coal
СHA-AGK1.84 ± 0.07 l
CHA-ALD1.07 ± 0.06 def
Humic-like substances
HLS-450.70 ± 0.05 b
HLS-700.91 ± 0.08 cd
Values are means ± SD (n = 4). Values denoted with different letters within a column are significantly different at p < 0.05 according to Tukey’s HSD test.
Table 4. AOC of individual antioxidants estimated by the ORAC method.
Table 4. AOC of individual antioxidants estimated by the ORAC method.
CompoundAOC, μmol TE mg−1
Ascorbic acid2.32 ± 0.21
Vitamin E2.95 ± 0.16
Gallic acid6.35 ± 0.25
Syringic acid7.67 ± 0.39
Hydroxybenzoic acid13.11 ± 0.63
Sinapic acid15.71 ± 1.01
Vanillic acid20.46 ± 1.14
Ferulic acid23.64 ± 1.18
Coumaric acid30.58 ± 1.32
Values are means ± SD (n = 4).
Table 5. Parameters of a forward stepwise multiple linear regression model.
Table 5. Parameters of a forward stepwise multiple linear regression model.
ParameterValue−95%+95%
Intercept0.432−0.0870.952
C/N0.0060.0030.009
CCHO, %0.0260.0090.042
CCH3O, %−0.077−0.143−0.010
TP, μmol mg−10.3690.2160.522
Multiple determination coefficient0.91
Adjusted multiple determination coefficient0.80
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Klein, O.I.; Kulikova, N.A.; Konstantinov, A.I.; Zykova, M.V.; Perminova, I.V. A Systematic Study of the Antioxidant Capacity of Humic Substances against Peroxyl Radicals: Relation to Structure. Polymers 2021, 13, 3262. https://doi.org/10.3390/polym13193262

AMA Style

Klein OI, Kulikova NA, Konstantinov AI, Zykova MV, Perminova IV. A Systematic Study of the Antioxidant Capacity of Humic Substances against Peroxyl Radicals: Relation to Structure. Polymers. 2021; 13(19):3262. https://doi.org/10.3390/polym13193262

Chicago/Turabian Style

Klein, Olga I., Natalia A. Kulikova, Andrey I. Konstantinov, Maria V. Zykova, and Irina V. Perminova. 2021. "A Systematic Study of the Antioxidant Capacity of Humic Substances against Peroxyl Radicals: Relation to Structure" Polymers 13, no. 19: 3262. https://doi.org/10.3390/polym13193262

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