Open access peer-reviewed chapter

Trans-Resveratrol: From Phytonutrient Supplement, to Novel Nanotherapeutic Agent

Written By

Tracey Lynn Harney

Submitted: 08 August 2022 Reviewed: 06 October 2022 Published: 07 November 2022

DOI: 10.5772/intechopen.108496

From the Edited Volume

Periodontology - New Insights

Edited by Gokul Sridharan

Chapter metrics overview

129 Chapter Downloads

View Full Metrics

Abstract

Trans-resveratrol (3,5,4′-trihydroxy-trans-stilbene) (RES) is a plant polyphenol that has been well documented for its anti-oxidant, antimicrobial, anti-inflammatory, and anti-aging properties. Moreover, compelling evidence presented in the abundance of pre-clinical studies using ligature-induced periodontitis models has positioned RES as a theoretically viable candidate for the reduction of the chronic inflammation, oxidative stress, and tissue destruction seen in periodontitis (PD). However, the instability of RES under physiological conditions, as well as its rapid hepatic clearance, has presented as a challenge to its ubiquitous application as an oral therapeutic in clinical practice. Fortunately, with the application of nanotechnology, the pharmacological profile of RES repositions the phytochemical from an herb-based supplement, useful as an adjunct therapy, to a stable and potent nanomedicine, demonstrating efficacy for the prevention and treatment of PD and its associated systemic diseases. This chapter explores the details of the potential for nano-RES as a viable therapeutic for PD.

Keywords

  • periodontitis
  • polyphenols
  • trans-resveratrol (RES)
  • pharmacognosy
  • nanotechnology

1. Introduction

Periodontal Disease (PD) is a complex and multifactorial chronic systemic inflammatory condition that manifests via the interplay between dysbiosis of the normal microbiota of the oral cavity, and the dysregulation of the host immune response [1] The shift in the microbiota toward pathogenic organisms, activates host inflammation which persists if left untreated, ultimately progressing to continuous tissue destruction of the supporting tissues around the tooth. The clinical manifestations that result are bleeding gums, gradual tooth detachment and bone loss, which can be painful, and potentially result in edentulism, if left unchecked [1].

Since the global prevalence of PD of some form, is positioned at approximately 50% of the adult population, the growing concern around this approximation is particularly necessary, when one reflects upon the extensive list of immune-related systemic diseases with which PD is associated [2, 3, 4]. That is, PD has been associated with cardiovascular [5, 6, 7, 8, 9, 10, 11], gastrointestinal, respiratory, renal, hepatic, neurodegenerative, and autoimmune disease; obesity, diabetes, viral infections, adverse pregnancy outcomes and some cancers have also been reported to be associated with PD. Moreover, due to the systemic nature of immune dysregulation and pathogenic dysbiosis, the relationship between PD and many of these comorbidities is bidirectional [12].

One cannot emphasize more, the multifaceted pathogenesis of PD, which involves the interchange between the host tissues and the accompanying microbial communities of the oral cavity, and the microenvironment therein. For this reason, clarification of the details of the pathogenic mechanisms continues to unfold. Fortunately, knowledge gained from studies on the oral microbiome, has recently emerged [13, 14, 15].

In this way, understanding further details of the disease process of PD from the perspective of the oral microbiome can assist in the creation of novel preventative and therapeutic applications.

The oral microbiome (OM), which has been estimated to consist of greater than 108 microbes per milliliter of saliva, is considered by many, as one of the most clinically important collections of microbial organisms in the human body [16]. Both prokaryotic and eukaryotic microorganisms make up the estimated 1000 microbial species of the OM. However, more than 700 microbes are prokaryotic, giving investigators a rationale for focusing on the bacterial taxa of the oral cavity when seeking to answer questions about the nature of oral health and disease [17, 18, 19].

Practicably, by gaining a deeper understanding of the mechanisms employed by “keystone” bacterial essential players, which are Gram negative anaerobes (e.g., red-complex pathogens, Porphyromonas gingivalis (Pg), Treponema denticola and Tannerella forsythia) in the inflammation process, one may be in a better position to produce an effective and up-to-date arsenal of remedies for the prevention, management, and resolution of PD [20].

It is worth noting that part of the survival mechanism for Gram negative bacteria is the dissemination of outer membrane vesicles containing various virulence factors which result in a distorted immunological dysregulation and tissue destruction [21, 22]. This strategy is a likely explanation for the discovery of pathogenic bacterial DNA at sites distal from the oral cavity. Further to this, Gram negative bacteria contain the virulence factor lipopolysaccharide (LPS) in their outer membrane, which acts as a constant inflammatory trigger via its pathogen-associated molecular patterning (PAMP). Interestingly, antibiotic therapy is not a straight forward solution, since dead Gram negatives also release LPS [23].

The treatment principle of PD is founded in the elimination of the pathogenic dysbiotic biofilm, whilst addressing the dysregulated immune response of the host. These treatment goals are first addressed via the mechanical removal of subgingival pathogenic dysbiotic plaque and tartar buildup, which is accomplished by the gold standard procedure, scaling and root planning (SRP). As a general rule, SRP is followed by lifelong comprehensive care, which may include subgingival irrigation with antibiotic solutions (www.NHS.uk; www.ADA.org). However, instances involving extensive tissue destruction require surgical intervention, which is more invasive and can be costly. Additionally, as a supplementary treatment, host modulation therapy (HMT) may be employed to address the dysregulation of the immune system that was triggered by dysbiosis [24].

Some phytochemicals have been identified as potential modulators of dysbiotic oral biofilms as well as mitigators of host inflammatory responses (e.g., Curcumin, Hesperidin, Silymarin, Resveratrol) [25, 26, 27]. Pharmacognosy, therefore, has been explored as a contributing modality for the mitigation of inflammatory disease.

Indeed, there is a body of evidence in the scientific literature outlining the potential of various forms of phyto-therapeutic intervention for the prevention and treatment of PD. Among these phytochemicals is the trans-isomer of the polyphenol, resveratrol (RES), which has been reported to alleviate many inflammatory conditions, including those associated with PD. RES has also demonstrated favorable synergistic effects on inflammation when combined with other bioactive plant-based compounds such as curcumin [28].

Advertisement

2. Attenuation of the Dysbiotic biofilm by RES

Since PD is initiated by a dysbiotic biofilm, modulation of the oral microbial ecosystem is a viable upstream therapeutic approach. Also, the reduction of pathogenic microbial load is a logical and necessary start to treatment, which is the primary intent of SRP and CC (www.NHS.uk; www.ADA.org).

Studies assessing the potential for addressing the pathogenic oral biofilm, reported that RES demonstrated antimicrobial action against PD pathogens. For example, O’Connor and colleagues conducted an in vitro study on different prokaryotic and eukaryotic microbes using the gold standard antiseptic rinse used in dentistry, chlorohexidine (CHX), as a positive control. They tested the inhibitory effect of RES on 15 different microbial species and only A. actinomycetemcomitans (Aa) and P. gingivalis (Pg) (ATCC2533277) showed sensitivity to RES. The authors also pointed out that the other 13 microbes were aerobic, suggesting that RES is specifically antimicrobial to anaerobes. In fact, Pg showed a much higher sensitivity to RES than Aa did in this study, suggesting that Pg may be especially vulnerable to RES, for reasons other than its anaerobic nature. However, since this was a study based on spectrophotometric optical density readings of broth cultures, the bacteria tested were planktonic and therefore not comparable to those participating in the biofilm seen in PD. Hence, although this study was a helpful preliminary survey, the definitive demonstration of the action of RES against Pg in vivo, was not solidified [29].

Interestingly, in 2019, Kugaji and colleagues assessed the effect of RES on an experimental biofilm both directly and indirectly. Firstly, direct assays were conducted via the determination of the minimal inhibitory and minimum bactericidal RES concentrations (MIC and MBC, respectively) applied to cultures of commercially available (e.g., ATCC 33277) and clinical strains (CS02) of Pg. Then, the calculated indices for MIC (156.25 μl/ml) and MBC (312.5 μl/ml) of RES, were applied to commercial strains of Pg to determine kill-time, adhesion to substrate, and morphological changes (determined by SEM). Next, RES was evaluated indirectly through the examination of the genetic expression of gingipains (i.e., Kgp and rgpA) and fimbriae (i.e., fimA), virulence factors from Pg that have been found to be instrumental in promoting the formation of the pathogenic biofilm in PD. The findings led the authors to conclude that RES displays antimicrobial action against keystone periodontal pathogen, Pg, as well as the ability to reduce the pathogenic biofilm indirectly, by decreasing the expression of its proteolytic virulence factors. It is also worth noting that in this study, the clinical Pg strain was more sensitive to RES than the commercially available strain. Regarding adhesion assays, the response to RES was also strain-dependent, which adds another layer of complexity regarding the in vivo translation of these findings [30].

Similarly, both antibacterial and anti-adherent action of RES on Pg (i.e., ATCC 33277) were reported by Lagha and colleagues in 2019. They determined an MIC and MBC of 250 μL/mL and 500 μL/mL, respectively, and applied a luminescence assay as an indicator of ATP production, to show the bioactivity of the bacteria in the experimental biofilm. The calculated MIC of RES in this study reduced biofilm viability by over 50% [31].

Conversely, Millhouse and colleagues reported that RES did not demonstrate any antimicrobial action, after they developed an experimental model of a periodontal biofilm, which consisted of a co-culture of epithelial and microbes. This intricate co-culture included simulated saliva, and was designed to mimic the host-biofilm interface to assess the antimicrobial and anti-inflammatory properties of Chlorohexidine (CHX), compared with those of RES [32].

Since it has been reported that Pg demonstrates significant differences in genetic expression when planktonic compared to being engaged in a multispecies biofilm [33], evolving such a model is an important step toward in vivo translation [32].

However, the results from Milhouse et al. cannot be attributed to the co-cultured biofilm model alone because the RES was suspended in water, while the other three studies, which reported antimicrobial action from RES, used ethanol or dimethyl sulfoxide (DMSO). Hence, the comparability between the studies is diminished by the fact that RES has virtually no solubility in water, while DMSO and ethanol are more successful as solvents [34], but often have unpredictable biological effects of their own, which must be appropriately controlled for.

Further to this, an in vivo study conducted by Cirano and colleagues, employed a modified ligature-induced rat model to evaluate the action of RES against key pathogens, Pg, T. forsythia (Tf) and Aa. This study applied ligatures pre-treated with the individual pathogens and evaluated the antimicrobial action of RES using RT PCR of all three microbial species, accompanied by a single daily gavage of a 10 mg/Kg dose of RES over a 30-day period. In this case, no statistically significant antimicrobial activity was demonstrated by RES. However, although the RES stock was prepared using ethanol, further dilutions were performed in water, which decreases the solubility of RES. The oral delivery route (gavage) in this study suggests that low oral bioavailability [35] and instability [34] of RES may have affected the results. Consequently, the authors recognize that for future studies, applying RES via topical/transmucosal administration in the in vivo models may overcome this limitation [36].

Although these studies cannot definitively conclude that RES demonstrates antimicrobial activity in vivo, further studies may bring to light the antimicrobial status of RES, as more information emerges about the oral microbiome and the mechanisms of the intercellular interactions that govern dysbiotic biofilm formation. What is clear from this analysis is that more consideration needs to be given to the solubility, formulation, and route of delivery of RES. Indeed, in vivo experimental designs may benefit from routes of administration of RES that bypass its low bioavailability and stability, as new models are evolved that provide more consistent and translatable data.

Advertisement

3. Modulation of host immune response by RES

Investigations using ligature-induced PD animal models, have consistently reported the favorable modification of many pro-inflammatory markers as well as the reduction and/or arrest of tissue destruction. Table 1 outlines the results of a collection of animal studies and the molecules affected by RES [28, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46].

Inflammatory Mediator and RES Action on MediatorAction of MoleculeFormulation and Administration of RESType of Assay Used to Detect the Inflammatory Mediator/MarkerTissue Sampled
IL-1β DecreasedPro-inflammatoryGavage 10 mg/Kg in 2% EtOH [37]
PO 10 mg/Kg Melinjo seed extract [38]
Gavage 20 mg/Kg [39]
RT-PCR [37]
RT-PCR [38]
RT-PCR [39]
Gingival tissue [37]
Gingival tissue [38]
Gingival tissue [39]
IL-4 IncreasedAnti-inflammatoryGavage 10 mg/Kg in Tween-80 /ddH2O [28]
Gavage 10 mg/Kg in Tween-80/ddH2O [40]
Gavage 10 mg/Kg in Tween 80/ddH2O [41]
ELISA [28]
ELISA [40]
ELISA [41]
Gingival tissue [28]
Gingival tissue /Blood [40]
Gingival tissue [41]
IL-6 DecreasedPro-inflammatoryInto socket 50 μM in 0.1% DMSO [42]
Gavage 10 mg/kg in 2% EtOH [37]
PO 10 mg/Kg Melinjo seed extract [38]
Gavage 20 mg/Kg [39]
Immunohistochemistry [42]
RT-PCR [37]
RT-PCR [38]
RT-PCR [39]
Periodontium [42]
Gingival tissue [37]
Gingival tissue [38]
Gingival tissue [39]
IL-8 DecreasedPro-inflammatoryGavage 20 mg/Kg [39] diabetic miceRT-PCR [39]Gingival tissue [39]
IL-17 DecreasedPro-inflammatoryGavage 10 mg/Kg in EtOH/ddH2O [43]ELISA [43]Gingival tissue [43]
TNF-α DecreasedPro-inflammatoryInto socket 50 μM in 0.1% DMSO [42]
Gavage 10 mg/kg in 2% EtOH [37]
PO 10 mg/Kg Melinjo seed extract [38]
Gavage 20 mg/Kg [39]
Immunohistochemistry [42]
RT-PCR [37]
RT-PCR [38]
Serological [39]
Gingival tissue [37, 42]
Gingival tissue [38]
Blood [39]
IFN-ɣ DecreasedPro-inflammatoryGavage 10 mg/Kg in Tween-80 /ddH2O [28]ELISA [28]Gingival tissue [28]
RF DecreasedPro-inflammatoryGavage 10 mg/Kg in Tween-80/ddH2O [40]ELISA [40]Blood [40]
SOD IncreasedAntioxidantGingival injection/5 mg/kg in DMSO [44]
Gavage 10 mg/kg in Tween −80 [45]
Serological [44]
ELISA [45]
Blood [44]
Gingival tissue [45]
HO-1 IncreasedROS/RNS stress defenseGingival injection/5 mg/kg in DMSO [44]Western Blot [44]
Immunohistochemistry [44]
Periodontium [44]
COX-2 DecreasedPro-inflammatoryGingival injection/5 mg/kg in DMSO [44]Immunohistochemistry [44]Periodontium [44]
MMP-2 DecreasedDestructive proteaseGingival injection/5 mg/kg in DMSO [44]Immunohistochemistry [44]Periodontium [44]
MMP-9 DecreasedDestructive proteaseGingival injection/5 mg/kg in DMSO [44]Immunohistochemistry [44]Periodontium [44]
Nrf2 IncreasedAntioxidantGingival injection/5 mg/kg in DMSO [44]
PO 10 mg/Kg Melinjo seed extract [38]
Western Blot [44]
Immunohistochemistry [44]
RT-PCR [38]
Periodontium [44]
Gingival tissue [38]
SIRT-1 IncreasedAntioxidantGavage 10 mg/Kg in Tween-80/ddH2O [45]
PO 10 mg/Kg Melinjo seed extract in [38]
RT-PCR [45]
RT-PCR and Western blot [38]
Gingival tissue [45]
Gingival tissue [38]
ACCPA DecreasedDestructive protease
Antibody
Gavage 10 mg/Kg in Tween-80/ddH2O [40]ELISA [40]Blood and Gingival tissue [40]
NADPH-OX DecreasedSource of ROS damageGavage 10 mg/Kg in Tween-80/ddH2O [45]
Gavage 10 mg/Kg in Tween-80/ddH2O [46]
ELISA [45]
ELISA [46]
Gingival tissue [45]
Gingival tissue [46]
NF-κB DecreasedPro-inflammatory Transcription FactorPO 10 mg/Kg Melinjo seed extract [38]
Gavage 20 mg/Kg [39]
RT-PCR [38]
Western blot [39]
Gingival tissue [38, 39]
NQO-1 IncreasedAnti-oxidant genePO 10 mg/Kg Melinjo seed extract [38]RT-PCR [38]Gingival tissue [38]
AMPK Induced PhosphorylationAnti-inflammatoryPO 10 mg/Kg Melinjo seed extract [38]Western blot [38]Gingival tissue [38]
P38/MAPK Decreased PhosphorylationPro-inflammatoryPO 10 mg/Kg Melinjo seed extract [38]
Gavage 20 mg/Kg [39]
Western blot [38]
Western blot [39]
Gingival tissue [38]
Gingival tissue [39]
STAT3 Decreased PhosphorylationPossible Inducer of Diabetic PDGavage 20 mg/Kg [39]Western blot [39]Gingival tissue [39]
p65/NF-κB Decreased PhosphorylationPro-inflammatory Transcription FactorGavage 20 mg/Kg [39]Western blot [39]Gingival tissue [39]
Inos DecreasedPro-inflammatoryPO 10 mg/Kg Melinjo seed extract [38]RT-PCR [38]Gingival tissue [38]
8-OHdG DecreasedOxidative stress markerPO 10 mg/Kg Melinjo seed extract [38]ELISA [38]Urine [38]
Dityrosine DecreasedOxidative stress markerPO 10 mg/Kg Melinjo seed extract in [38]ELISA [38]Urine [38]
RANKL DecreasedIncreases bone resorptionGavage 10 mg/kg in 2% EtOH [37]
Gavage 10 mg/Kg in Tween-80/ddH2O [41]
RT-PCR [37]
RT-PCR [41]
Gingival tissue [37]
Gingival tissue [41]
NOx DecreasedNitrosative stress markerPO 10 mg/Kg Melinjo seed extract [38]Serological [38]Serum [38]
Nitrotyrosine DecreasedNitrosative stress markerPO 10 mg/Kg Melinjo seed extract in [38]ELISA [38]Serum [38]
RUNX-2 DecreasedOsteoblastic markerInto socket 50 μM in 0.1% DMSO [42]Immuno-histochemistry [42]Periodontium [42]
OCN IncreasedOsteoblastic hormoneInto socket 50 μM in 0.1% DMSO [42]Immuno-histochemistry [42]Periodontium [42]
Ki67 DecreasedProliferationInto socket 50 μM in 0.1% DMSO [42]Immuno-histochemistry {37}Periodontium [42]

Table 1.

Ligature-induced PD studies with the RES formulation, type of assay and tissue sampled for the inflammation-mediating molecules examined.

Moreover, several reports from in vitro studies have also reliably concluded that RES mitigates the dysregulated immune response seen in PD [47, 48]. For example, a study by Qu et al. used Porphyromonas endodontalis (Pe)LPS-challenged osteoblast-like mouse cells (MC3T3-E1) to identify the role of silent mating type information regulation 2 homolog 1 (SIRT-1) in the resolution of inflammation and mitigation of bone loss. Their model employed the transfection of the cells with SIRT-1 siRNA, as well as the use of SIRT-1 inhibitor, EX-527, to silence SIRT-1 activity, whilst RES was used as a SIRT-1 activator [49].

Notably, elevated levels of matrix metaloprteinase-13 (MMP-13) expression, as determined by RT-PCR, ELISA and Western blot, were evidenced upon induction by LPS. Furthermore, this induction was increased by knockdown of SIRT-1 activity, either by use of siRNA SIRT-1, or by the addition of the inhibitor, EX-527. However, pre-treatment with RES (50 μM, no solvent was specified) for 1 hour, significantly (p < 0.05) suppressed the mRNA expression (as determined by RT-PCR) and protein production (assayed by Western blotting and ELISA) of MMP-13 in the LPS-challenged cells. A chromatin immunoprecipitation assay (ChIP) then determined that SIRT-1 prevents the activation-stimulating binding of NFkB-p65 to the MMP-13 promoter [49], which aligns with the findings reporting RES as a SIRT-1 activator [50, 51, 52].

An early human in vitro study reported the use of RT-PCR and ELISA, to determine that RES decreased the mRNA levels of a range of cytokines regarded to be proinflammatory, including IL-1β, IL-6, IL-8, IL-12 and TNF-α, in Pg lipopolysaccharide-(LPS)-challenged human periodontal ligamental cells. The effect was reported to be both dose-(25 μM, 50 μM and 100 μM) and time-(0–72 hrs) dependent [53].

Similarly, more recent in vitro studies employing ELISA, found that RES significantly decreased IL-6 and IL-8, in human gingival fibroblasts (HGFs) that were stimulated by LPS [26, 54]. RES was also combined with the triterpene, Celastrol (CEL), and loaded into a collagen film to enhance the effectiveness of coated dental implants by Wang et al. [55]. Using SEM and histochemistry, it was determined that the RES-loaded collagen films stimulated the most proliferation of the human periodontal ligamental fibroblasts (HPLFs), whilst the CEL-loaded films, demonstrated the lowest bone marrow macrophage-mediated osteogenesis. Hence, it was suggested that using both RES and CEL in the collagen film was a promising approach to the development of an efficacious dental regenerative agent [55].

In 2020, Ashour et al., used Hesperetin (HESP) as a glyoxalase 1(Glo1) inducer to assess its modulation of the damage caused to HPLFs from high glucose. This experimental in vitro model showed that HPLFs overloaded with glucose, increased the cellular concentration of methylglyoxal (MG) and its modified proteins, which resulted in HPLF dysfunction and inability to bind to collagen-I. By adding RES (10 μM) to the HESP, the dysfunction was corrected through the attenuation of deregulated glucose metabolism [56].

An additional study compared the anti-inflammatory activity of Tetracycline (TC), Minocycline (MC), Quercetin (QU), and RES, in LPS-stimulated macrophage-like mouse cells, (RAW264.7). Upon the application of differential scanning calorimetry, and RT-PCR, the anti-inflammatory activity was determined as an index of COX-2 inhibition and it was concluded that the order from highest anti-inflammatory action to lowest was as follows: QU > RES > MC > TC [57].

Advertisement

4. Restoration of damaged periodontal tissues by RES

An extensive in vitro and in situ investigation by Wang et al., using human periodontal ligamental stem cells (HPLSCs) treated with TNF-α, found that RES conserved their osteo-differentiation. Here, histological analysis showed that RES preserved the formation of cell aggregates and made the cells more resilient to TNF-α—induced inhibition of alkaline phosphatase (PhoA) whilst promoting mineralization. Further to this, RT-PCR was used to show the partial restoration of mRNA expression of the osteo-differentiation drivers, OCN, RUNX2, PhoA and Collagen-1. Moreover, Western blotting showed restored values for OCN, Collagen-1, and RUNX2, an osteoblastic modulator, confirming the mRNA findings at the level of protein. The in vitro action of RES on the ontogenetic capabilities of HPLSCs originating from PD patients was also determined to be beneficial through histological analysis, RT-PCR and Western blot. Additionally, an ectopic regeneration experiment was conducted, involving transplantation of normal, PD alone, or RES-treated PD-induced HPLSC aggregates into nude mice, which allowed the researchers to assess the regenerative action of RES in situ [58].

An additional in vitro study on HPLSCs, employing histochemical analysis, RT-PCR, Western blot, as well as ELISA, reported that the osteogenic suppression induced by treatment with TNF-α was mitigated by RES. Also, the results of ERK1/2 pathway inhibition and activation assays showed that the anti-inflammatory action of RES was diminished by ERK1/2 pathway inhibition. This suggests that the mechanism employed by RES, involves activation of the ERK1/2 pathway in TNF-α-challenged HPLSCs [59].

Interestingly, Kudo et al., examined the role of SIRT1 in angiogenesis in PD by treating human umbilical vein endothelial cells (HUVECs) with E. coli LPS plus RES (i.e., a SIRT1 activator) or sirtinol (i.e., a SIRT1 inhibitor). Also, in this study, periapical granulomas were obtained from PD patients, whereas healthy individuals provided normal periodontal tissues, which were also assayed. Immunofluorescent imaging using antibodies against SIRT1 and Ki67, showed that the periapical granulomas had higher expression of both SIRT1 and Ki67, compared to healthy gingival tissue. Additionally, the quantification of the mRNA of SIRT1, VEGF (which stimulates endothelial growth and differentiation into vessel morphology) and VE-cadherin (an adhesion protein for endothelia) in the LPS-induced HUVECS indicated that the mRNA expression statistically increased for all three proteins (P = 0.0019 (SIRT-1), 0.00005(VEGF) and 0.0045 (VE-cadherin)) in the RES-treated groups compared to the group treated with sirtinol [60].

Advertisement

5. Mitigation of systemic conditions associated with PD by RES

Studies have been conducted integrating RA, DM, cigarette smoking or osteoporosis (OP) into their induced-PD models. For example, two studies employed a ligature-induced PD rat model which incorporated three, eight-minute exposures per day, to the equivalent of 10 cigarettes, for 37 days, and examined the effects of RES on bone loss and oxidative stress, respectively [40, 41]. In this study, they found that the gingival tissue of the RES-treated group displayed higher expression of anti-inflammatory cytokine IL-4, and the antioxidant SIRT-1 [40], suggesting that RES plays a beneficial role when the added risk factor of smoking is involved. Further to this, RT-PCR determined that the mRNA expression of the osteoclastic inducer, RANKL was diminished by RES [41] and ELISA showed that the ROS inducer, NADPH oxidase, was also significantly diminished in the RES-treated group (i.e., RES + Cigarette Smoke) whilst SOD was significantly higher (p < 0.05) (Table 1) [37].

Moreover, the administration of RES was found to decrease the TH17/TH2 cell ratio in diseased gingival tissue, suggesting that it may modulate the overstimulation of TH17 (i.e., TH lymphocytes purported be stimulated by the outer membrane protein of Pg) production, which contributes to tissue destruction [38, 61]. These results also indicate that RES may assist in the mitigation of the periodontal damage contributed to, by the modifiable risk factor, smoking [38]. Additionally, a study which assessed the effect of RES on experimental PD in diabetic mice, found that RES decreased the mRNA expression of LPS—induced inflammatory cytokines IL-1β, IL-6, IL-8 and TNF-α. In addition, the LPS-induced phosphorylation, and therefore activation, of transcription factors downstream to TLR4, p65 NF-κB, p38 MAPK, and STAT3, was also suppressed by RES, suggesting that RES works through the TLR4 signaling pathway [39]. Moreover, RES was also found to reduce alveolar bone loss and attenuate hyperglycemia [39, 56], demonstrating its potential as an attenuator of diabetic PD.

An additional study conducted by Correa and colleagues, examined the effect of RES compared to Ibuprofen on experimental PD and RA combined, using the ligature-induced PD rat model. Here, they determined by way of ELISA, that the administration RES or Ibuprofen resulted in the reduction of anti-cyclic citrullinated peptide antibody (ACCPA) in the tissues by 72% and 99%, respectively (p < 0.05), and, RES alone was reported to reduce serum rheumatoid factor (RF)(p < 0.05). In this study, RES also demonstrated sustained reduction in paw edema compared to Ibuprofen. Moreover, the RES treated group demonstrated reduced articular damage in the histological analysis, indicating its ability to modulate RA-induced damage in the context of experimental PD in rats [40].

Similarly, in an animal study, using ovariectomized rats and PD-induction via ligature installment, RES was compared to Zoledronate, a drug used to treat osteoporosis (OP). Here, ELISA showed that RES downregulated NADPH oxidase and reduced alveolar bone loss, but not as drastically as in the Zoledronate-treated group. Nonetheless, the results suggest that RES may attenuate alveolar bone loss in estrogen-deficient rats via the attenuation of NADPH oxidase, making NADPH oxidase a potential drug target for RES. Using a human in vitro neuroinflammation model, it was determined that several pathways that promote oxidative stress (via the decrease in AKT1, FOS, IKBKB, IRF1, JAK2, NFKB1PIK3RI, RELA, STAT1, TNF-α and TNFRSF1 mRNA) as well as iNOS (via the decrease in FOS, IKBKB, IRAK1, IRAK2, IRF1, JAK2, NFKB1, RELA and STAT1 mRNA) was attenuated by RES. The researchers in this study used qPCR and biochemical pathway analysis to examine 96 genes, present in human neuroblastoma cells, that were induced into inflammation via the application of LPS originating from Pg. It was concluded that RES decreased NF-κB-mediated acute inflammation pathways in human neuroblastoma cell culture. Further to this, RES was also reported to inhibit IGF-1 and Insulin receptor while activating the PTEN, PPARa/RXRa, PPAR metabolic pathways. This extensive in vitro study showed the potential for RES to address Pg-related disease, with a particular focus on the prevention of AD [62].

Advertisement

6. Clinical translation of RES

Two randomized clinical trials were conducted by Javid et al. (IRCT ID: IRCT2015012420765N1). One study evaluated the effect of RES on blood glucose, Insulin, Insulin resistance and periodontal markers in 43 type two DM patients aged 30–60 years old, with chronic moderate PD. This double-blind and placebo-controlled study, evaluated the impact of RES by supplementing the experimental group with 480 mg capsules containing RES, once a day, for 28 days. The subjects in the placebo group were given identical capsules containing 480 mg of starch. Although no significant differences in fasting blood glucose or TGs between the groups were evident, the RES-supplemented group showed significantly lower (p < 0.05) periodontal pocket depth, serum fasting glucose and insulin resistance [63].

The second study, also a randomized clinical trial, assessed the serum levels of inflammatory cytokines IL-6 and TNF-α, as well as total antioxidant capacity (TAC) and clinical attachment loss (CAL) in 43 patients with type two DM and PD. After 28 days of supplementation with 480 mg/d of RES, serum levels of IL-6 were significantly lower(p = 0.039), but the other parameters measured, showed no significant change compared to control [64]. The results of both studies do not fully align with those of the animal studies, which may be partly due to the formulation and posology used. That is, Javid et al., applied a 480 mg capsule of Polygonum cuspidatum reported to contain 240 mg of RES. In general, the animal studies used 10–20 mg/kg of pure RES, delivered via gavage, which would calibrate to approximately 730–1460 mg of RES in this study, considering the average mass of the human subjects in the RES-treated group (mean mass of subjects =73.8 ± 10.2 Kg) [63, 64].

Lastly, Shoukheba, conducted a six-month human study in 2020 consisting of 15 male smokers with moderate to severe (i.e., CAL >5 mm) PD, following SRP. Interestingly, RES gel (0.001% w/v final concentration) was directly applied to the testing sites (15 healthy, 15 with PD), applying the split mouth method at days 7,14, and 21. The clinical parameters, plaque index (PI), probing pocket depth (PPD), bleeding index (BI)and CAL were assessed. In addition, the gingival crevicular fluid was collected during the treatment visits to assess SOD levels. The SOD levels were higher in the RES group compared to control and although both groups showed a significant decrease (p < 0.05) in the clinical parameters, PPD and CAL, at the three-month mark, only the RES-treated group demonstrated significant improvement from baseline at 6 months (p < 0.05) [65].

Again, further studies using RES in the absence of SRP, need to be conducted to fully elucidate the role of RES in the attenuation of PD. In addition to this, since cigarette smokers have been reported to have higher susceptibility to an oral Pg infection (Zeller et al., 2014), perhaps more information may be gained by comparing the Pg load between groups [65].

RES has a long-term body of evidence reporting its low bioavailability, which is mainly due to its physicochemical properties. That is, RES is pH-, thermo-, and photo-sensitive, [66]. More importantly, RES is unstable under physiological pH (7.4) and temperature (37°C), thus diminishing the in vivo translatability [67, 68, 69].

There are a myriad of studies demonstrating the effective mitigation of the low bioavailability, stability and pharmacokinetic profile of RES via nano-formulation. In this way, RES may be potentially upgraded from a supplement to a nanomedicine.

Advertisement

7. Improvement of RES as a therapeutic via nanotechnology

Nanotechnology, which employs materials that measure at <100 nm in at least one dimension, has been found to improve the stability, bioavailability and activity of RES. Overall, there exist several studies exploring the action of various formulations of Nano-RES which reported improvements in stability, bioavailability and pharmacokinetics, compared to RES in bulk form [70, 71, 72, 73, 74, 75]. For example, nano-formulations employing solid lipid nanoparticles (SLNs) and nanostructured lipid carriers (NLCs), are purported to have higher stability, and a pharmacokinetic profile indicative of sustained release. Interestingly, SLNs and NLCs have more economical scalability compared to liposomal systems [70], making them attractive options as RES carriers.

Successful production of the SLNs and NLCs may be accomplished using waxes, fats, oils, or combinations thereof, combined with surfactant, and the application of high-pressure shear homogenization [71], sonication [72] and [73], or both [74, 75]. Studies involving RES-loaded SLNs and NLCs, demonstrated higher stability compared to bulk RES, as indicated by the high zeta potentials reported (i.e., ζ > 20 mV, [74] and ζ < -20 mV, [71, 72, 73, 75]).

Furthermore, Singh and Pai reported a self-nano-emulsifying drug delivery system (SNEDDS) that enhanced the pharmacokinetic profile of RES in vivo. Here, the RES-loaded SNEDDS was administered orally to rats (at 20 mg/Kg body weight dose) before withdrawing serial aliquots of blood. The SNEDDS was designed to automatically form a nano-emulsion once in contact with the fluids of the gastrointestinal tract and spontaneous emulsification into stable (ζ = −29.76 mV), 57 nm nano-globules took place in 43 seconds after being placed in 0.1 M HCl. The authors noted that the enhanced oral availability of RES-loaded SNEDDS compared to crude RES was likely due to an increase in absorbability as well as diminished first pass clearing by the liver enzyme CYP3A4 [76].

Moreover, Al-Bishri and colleagues compared the anti-diabetic activity of a commercial nano-emulsion of RES (Life Enhancement, Petaluma, California, USA) with that of chromium picolinate in streptozocin-induced diabetic in rats. Two weeks after the induction of DM, rats in the experimental groups were orally administered either (80 μg/Kg body weight) or Nano-RES emulsion (20 μg/Kg body weight) every day for 30 days when serum levels for glucose insulin, as well as biomarkers: NO, SOD, CAT, GPX, GST and GSH were determined. Both nano-RES emulsion and Chromium picolinate demonstrated inhibition of the oxidative stress induced from hyperglycemia [77].

Further to this, explorations emphasizing the fortification process involved in functional food production, demonstrate the potential of functional foods to be administered as an effective prevention, management, and treatment of PD. For example, Ahmad and Gani, assessed the biological action of RES-fortified snacks and reported that starch nano-encapsulation improved the thermostability of RES whilst enhancing anti-diabetic and anti-obesity effects compared to bulk RES, which was determined by the percent inhibition of the enzymes: α-glucosidase and, pancreatic lipase, and cholesterol esterase, respectively [78]. Inhibition of lipid peroxidase, which was also reported, demonstrated antioxidant activity of the nano-formulated food complex. Moreover, since the nano-embedded RES-fortified designer food-snacks exhibited enhanced desired activity at physiological pH (7.4) and temperature (37°C) the formulation in this study shows promise as a functional food [78].

Similarly, Jayan et al. reported the sustained release of RES from ZEIN-encapsulated nanoparticles (NPs) under physiological conditions (i.e., pH 7.4, 37°C) [79] and casein-encapsulated RES NPs, designed by Penlava et al., were found to be stable through a continuous pH range mimicking those of the gastrointestinal compartments (i.e., pH 1.2 for 2 hours and pH 6.8 for 2–24 hours). Interestingly, the latter study also demonstrated in vivo (using rats), a ten-fold increase in oral availability of casein-nano-encapsulated RES compared to the bulk form as determined by blood plasma assays over a 24-hour period following a single oral dose of 15 mg/Kg of RES (in ddH2O and PEG) or Casein-encapsulated RES NPs [80].

Additionally, Rabbani and colleagues set out to explore the fortification of foods and reported that when enhanced with a nano-encapsulated formulation of RES, mayonnaise demonstrated extended shelf-life, as suggested by the reduction in peroxide value over a six-month period. The ability of the nano-RES food complex to neutralize DPPH free radicals was investigated and the initially high number of free radicals present in the mayonnaise nano-RES complex were attributed to the successful encapsulation (and therefore protection) of RES in the nano formulation. Additionally, characterization of the mayonnaise nano-RES complex was conducted using XRD and Fourier transform infrared (FTIR) spectroscopy and RES was found to be amorphous (via XRD), demonstrating its incorporation into the complex, and it was confirmed that the molecular structure of the drug was retained while interacting with the complex via hydrogen bonds (via FTIR spectroscopy) [81].

Additionally, a promising 2021 investigation by Berta et al., involved the design of an oral Nano-RES spray via encapsulation of crude RES in 2-hydroxypropyl- β-cyclodextrins (i.e., RES-HPβCD) and tested its action on plaque formation in children. This study demonstrated that after using the spray once per day, the RES-HPβCD plaque was significantly reduced; the spray was also found to be doubled in efficacy when compared to tooth brushing alone [82].

Advertisement

8. From trash to treasure: Sources of trans-RES

RES is widely known in the Western hemisphere as a product from the skin of red grapes but there are other viable sources. In fact, the most abundant source of RES is found in roots of the plant, P. cuspidatum (i.e., Japanese knotweed), which has been used for a while in the East as a Traditional Chinese Medicine (i.e., ko-jo-kon) for the treatment of ailments such as cardiovascular disease. Unfortunately, this valuable resource has been wasted in the West, where it is perceived as an invasive species to be burned and buried. Since Japanese knotweed is edible and consists of more than 90% RES, adjustments to the current perception of the plant, could benefit many, as awareness of its medicinal (and potentially economic) value has been surfacing in the western scientific literature over recent decades [83].

Commercially available micronized versions of pure RES powder have been available for purchase for some time (e.g., www.megaresveratrol.net and www.Biotivia.com). However, nano-formulations of RES for oral use (many are topical [84]) have just began to emerge within the last year (e.g., www.oic.com.vn/en/ (Vietnam) www.nanoceuticalsolutions.com (USA), and www.hiimmune.com/product/nano-resveratrol-30ml/ (UK)).

Advertisement

9. Conclusion

According to the body of evidence explored in this chapter, the future of Nano-RES as a viable medicine for the prevention, management and treatment of PD and its associated systemic diseases, is promising.

In addition to nanomedicines for treatment, the development of functional foods fortified with Nano-RES, which demonstrate the potential for the preventative measures, may be an important addition to our dietary regimen moving forward [85].

Further to this, to realize the full potential of Nano-RES, an initiative is needed, to encourage more clinical research on the efficacy of low-cost, phytonutrient-based, novel nano-formulations, specifically aimed to address PD.

This could result in the commercial availability of high quality, ubiquitously accessible, effective RES-based Nano-formulations, and ultimately, a step toward a healthier future population.

References

  1. 1. Genco R, Sanz M. Clinical and public health implications of periodontal and systemic diseases: An overview. Periodontology 2000. 2020;83(1):7-13. DOI: 10.1111/prd.12344
  2. 2. Bernabe E, Marcenes W, Hernandez C, Bailey J, et al. Global, regional, and national levels and trends in burden of oral conditions from 1990 to 2017: A systematic analysis for the global burden of disease 2017 study. Journal of Dental Research. 2020;99(4):362-373. DOI: 10.1177/0022034520908533
  3. 3. Frencken J, Sharma P, Stenhouse L, Green D, et al. Global epidemiology of dental caries and severe periodontitis—A comprehensive review. Journal of Clinical Periodontology. 2017;44:94-105. DOI: 10.3390/ma13112631
  4. 4. Tonetti M, Jepsen S, Jin L, Otomo-Corgel J. Impact of the global burden of periodontal diseases on health, nutrition and wellbeing of mankind: A call for global action. Journal of Clinical Periodontology. 2017;44(5):456-462
  5. 5. Agarwal S. Impact of periodontitis on cardiovascular diseases. European Journal of Dental and Oral Health. 2021;2(2):1-8. DOI: 10.24018/EJDENT.2021.2.2.48
  6. 6. Bengtsson V, Persson G, Berglund J, Renvert S. Periodontitis related to cardiovascular events and mortality: A long-time longitudinal study. Clinical Oral Investigations. 2021;25(6):4085-4095. DOI: 10.1007/s00784-020-03739-x
  7. 7. Bodanese L, Louzeiro G, Magnus G, Baptista Â, et al. Association between periodontitis and myocardial infarction: Systematic review and meta-analysis. International Journal of Cardiovascular Sciences. 2021;34(5Supl.1):121-127
  8. 8. Sia S, Jan M, Wang Y, Huang Y, Wei J. Periodontitis is associated with incidental valvular heart disease: A nationwide population-based cohort study. Journal of Clinical Periodontology. 2021;48(8):1085-1092
  9. 9. Surma S, Romańczyk M, Witalińska-Łabuzek J, Czerniuk M, et al. Periodontitis, blood pressure, and the risk and control of arterial hypertension: Epidemiological, clinical, and pathophysiological aspects—Review of the literature and clinical trials. Current Hypertension Reports. 2021;23(5). Available from: https://link.springer.com/article/10.1007/s11906-021-01140-x [Accessed: 28 October 2022]
  10. 10. Tiensripojamarn N, Lertpimonchai A, Tavedhikul K, Udomsak A, et al. Periodontitis is associated with cardiovascular diseases: A 13-year study. Journal of Clinical Periodontology. 2021;48(3):348-356
  11. 11. Trindade F, Perpétuo L, Ferreira R, Leite-Moreira A, et al. Automatic text-mining as an unbiased approach to uncover molecular associations between periodontitis and coronary artery disease. Biomarkers. 2021;26(5):385-394
  12. 12. Hajishengallis G. Interconnection of periodontal disease and comorbidities: Evidence, mechanisms, and implications. Periodontology 2000. 2022;89(1):9-18
  13. 13. Chen H, Jiang W. Application of high-throughput sequencing in understanding human oral microbiome related with health and disease. Frontiers in Microbiology. 2014;5:1-6
  14. 14. Kilian M, Chapple I, Hannig M, Marsh P, et al. The oral microbiome—An update for oral healthcare professionals. British Dental Journal. 2016;221(10):657-666
  15. 15. Najmanova L, Sabova L, Lenartova M, Janatova T, et al. R/G value: A numeric index of individual periodontal health and oral microbiome dynamics. Frontiers in Cellular and Infection Microbiology. 2021;11:1-13. DOI: 10.3389/fcimb.2021.602643
  16. 16. Sedghi L, DiMassa V, Harrington A, Lynch S, Kapila Y. The Oral microbiome: Role of key organisms and complex networks in oral health and disease. Periodontology 2000. 2021;87(1):107-131
  17. 17. Berger D, Rakhamimova A, Pollack A, Loewy Z, et al. Oral biofilms: Development, control, and analysis. High-Throughput. 2018;7(3):24. DOI: 10.3390/ht7030024
  18. 18. Deo PN, Deshmukh R. Oral microbiome: Unveiling the fundamentals. Journal of Oral and Maxillofacial Pathology: JOMFP. 2019;23:122-128
  19. 19. Radaic A, Kapila Y. The Oralome and its Dysbiosis: New insights into oral microbiome-host interactions. Computational and Structural Biotechnology Journal. 2021;19:1335-1360
  20. 20. Hajishengallis G, Darveau RP, Curtis MA. The keystone-pathogen hypothesis. Nature Reviews Microbiology. 2012;10(10):717-725. DOI: 10.1038/nrmicro2873 [Accessed: 28 October 2022]
  21. 21. Lunar Silva I, Cascales E. Molecular strategies underlying Porphyromonas gingivalis virulence. Journal of Molecular Biology. 2021;433(7):166836
  22. 22. De Andrade K, Almeida-da-Silva C, Coutinho-Silva R. Immunological pathways triggered by porphyromonas gingivalis and fusobacterium nucleatum: Therapeutic possibilities? Mediators of Inflammation. 2019;2019:1-20. DOI: 10.1155/2019/7241312
  23. 23. Matsuura M. Structural modifications of bacterial lipopolysaccharide that facilitate gram-negative bacteria evasion of host innate immunity. Frontiers in Immunology. 2013;4:1-9. DOI: 10.3389/fimmu.2013.00109
  24. 24. Golub L, Lee H. Periodontal therapeutics: Current host-modulation agents and future directions. Periodontology 2000. 2019;82(1):186-204. DOI: 10.1111/prd.12315
  25. 25. Gonçalves V, Musskopf M, Rivera-Concepcion A, Yu C, Wong S, Tuin S, et al. Systemic dietary hesperidin modulation of Osteoclastogenesis, bone homeostasis and periodontal disease in mice. International Journal of Molecular Sciences. 2022;23(13):7100
  26. 26. Farzanegan A, Shokuhian M, Jafari S, Shirazi F, Shahidi M. Anti-histaminic effects of resveratrol and Silymarin on human gingival fibroblasts. Inflammation. 2019;42(5):1622-1629
  27. 27. Asteriou E, Gkoutzourelas A, Mavropoulos A, Katsiari C, Sakkas L, Bogdanos D. Curcumin for the management of periodontitis and early ACPA-positive rheumatoid arthritis: Killing two birds with one stone. Nutrients. 2018;10(7):908
  28. 28. Corrêa M, Pires P, Ribeiro F, Pimentel S, et al. Systemic treatment with resveratrol and/or curcumin reduces the progression of experimental periodontitis in rats. Journal of Periodontal Research. 2016;52(2):201-209
  29. 29. O’Connor D, Wong R, Rabie A. Resveratrol inhibits periodontal pathogens in vitro. Phytotherapy Research. 2011;25(11):1727-1731
  30. 30. Kugaji M, Kumbar V, Peram M, Patil S, et al. Effect of resveratrol on biofilm formation and virulence factor gene expression of Porphyromonas gingivalis in periodontal disease. APMIS. 2019;127(4):187-195
  31. 31. Lagha A, Andrian E, Grenier D. Resveratrol attenuates the pathogenic and inflammatory properties of Porphyromonas gingivalis. Molecular Oral Microbiology. 2019;34(3):118-130
  32. 32. Millhouse E, Jose A, Sherry L, Lappin D, et al. Development of an in vitro periodontal biofilm model for assessing antimicrobial and host modulatory effects of bioactive molecules. BMC Oral Health. 2014;14(1):1-11
  33. 33. Sánchez M, Romero-Lastra P, Ribeiro-Vidal H, Llama-Palacios A, et al. Comparative gene expression analysis of planktonic Porphyromonas gingivalis ATCC 33277 in the presence of a growing biofilm versus planktonic cells. BMC Microbiology. 2019;19(1):1-11
  34. 34. Robinson K, Mock C, Liang D. Pre-formulation studies of resveratrol. Drug Development and Industrial Pharmacy. 2014;41(9):1464-1469
  35. 35. Planas J, Alfaras I, Colom H, Juan M. The bioavailability and distribution of trans-resveratrol are constrained by ABC transporters. Archives of Biochemistry and Biophysics. 2012;527(2):67-73
  36. 36. Cirano F, Casarin R, Ribeiro F, Casati M, et al. Effect of resveratrol on periodontal pathogens during experimental periodontitis in rats. Brazilian Oral Research. 2016;30(1). Available from: https://www.scielo.br/j/bor/a/TSWnY94YqcYT9C93MtfCbWC/?lang=en [Accessed: 28 October 2022]
  37. 37. Chen J, Wang Y, Wang M, Wang B, et al. Resveratrol inhibited inflammation and alveolar bone loss in periodontitis. International Journal of Dentistry and Oral Science. 2015:1-5. Available from: https://www.academia.edu/19794213/Resveratrol_Inhibited_Inflammation_and_Alveolar_Bone_Loss_in_Periodontitis?auto=citations&from=cover_page. [Accessed: 5 August 2022]
  38. 38. Tamaki N, Cristina Orihuela-Campos R, Inagaki Y, Fukui M, et al. Resveratrol improves oxidative stress and prevents the progression of periodontitis via the activation of the Sirt1/AMPK and the Nrf2/antioxidant defense pathways in a rat periodontitis model. Free Radical Biology and Medicine. 2014;75:22-229
  39. 39. Zhen L, Fan D, Zhang Y, Cao X, Wang L. Resveratrol ameliorates experimental periodontitis in diabetic mice through negative regulation of TLR4 signaling. Acta Pharmacologica Sinica. 2014;36(2):221-228
  40. 40. Corrêa M, Pires P, Ribeiro F, Pimentel S, et al. Systemic treatment with resveratrol reduces the progression of experimental periodontitis and arthritis in rats. PLoS One. 2018;13(10):204414. DOI: 10.1371/journal.pone.0204414
  41. 41. Ribeiro F, Pino D, Franck F, Benatti B, et al. Resveratrol inhibits periodontitisrelated bone loss in rats subjected to cigarette smoke inhalation. Journal of Periodontology. 2017;88(8):788-798
  42. 42. Adhikari N, Prasad Aryal Y, Jung J, Ha J, et al. Resveratrol enhances bone formation by modulating inflammation in the mouse periodontitis model. Journal of Periodontal Research. 2021;56(4):735-745
  43. 43. Casati M, Algayer C, Cardoso da Cruz G, Ribeiro F, et al. Resveratrol decreases periodontal breakdown and modulates local levels of cytokines during periodontitis in rats. Journal of Periodontology. 2013;84(10):58-64
  44. 44. Bhattarai G, Poudel S, Kook S, Lee J. Resveratrol prevents alveolar bone loss in an experimental rat model of periodontitis. Acta Biomaterialia. 2016;29:398-408
  45. 45. Corrêa M, Absy S, Tenenbaum H, Ribeiro F, et al. Resveratrol attenuates oxidative stress during experimental periodontitis in rats exposed to cigarette smoke inhalation. Journal of Periodontal Research. 2019;54(3):225-232
  46. 46. Molez A, Nascimento E, Haiter Neto F, Cirano F, et al. Effect of resveratrol on the progression of experimental periodontitis in an Ovariectomized rat model of osteoporosis: Morphometric, immune-enzymatic, and gene expression analysis. Journal of Periodontal Research. 2020;55(6):840-849
  47. 47. Calabrese E. Hormesis and dental apical papilla stem cells. Chemico-Biological Interactions. 2022;357:109887. DOI: 10.1016/j.cbi.2022.109887
  48. 48. Li W, Huang X, Yu W, Xu Y, Huang R, Park J, et al. Activation of functional somatic stem cells promotes endogenous tissue regeneration. Journal of Dental Research. 2022;101(7):802-811
  49. 49. Qu L, Yu Y, Qiu L, Yang D, et al. Sirtuin 1 regulates matrix metalloproteinase-13 expression induced by Porphyromonas endodontalis lipopolysaccharide via targeting nuclear factor–κB in osteoblasts. Journal of Oral Microbiology. 2017;9(1):1317578
  50. 50. Borra M, Smith B, Denu J. Mechanism of human SIRT1 activation by resveratrol. Journal of Biological Chemistry. 2005;280(17):17187-17195
  51. 51. Chao S, Chen Y, Huang K, Kuo K, et al. Induction of sirtuin-1 signaling by resveratrol induces human chondrosarcoma cell apoptosis and exhibits antitumor activity. Scientific Reports. 2017;7(1):1-11
  52. 52. Hou X, Rooklin D, Fang H, Zhang Y. Resveratrol serves as a protein-substrate interaction stabilizer in human SIRT1 activation. Scientific Reports. 2016;6(1):1-9
  53. 53. Rizzo A, Bevilacqua N, Guida L, Annunziata M, et al. Effect of resveratrol and modulation of cytokine production on human periodontal ligament cells. Cytokine. 2012;60(1):197-204. DOI: 10.1016/j.cyto.2012.06.004
  54. 54. Shahidi M, Vaziri F, Haerian A, Farzanegan A, et al. Proliferative and AntiInflammatory effects of resveratrol and Silymarin on human gingival fibroblasts: A view to the future. Journal of dentistry (Tehran, Iran). 2017;14(4):203-211
  55. 55. Wang R, Bao B, Bao C, Wang S, et al. Resveratrol and celastrol loaded collagen dental implants regulate periodontal ligament fibroblast growth and osteoclastogenesis of bone marrow macrophages. Chemistry & Biodiversity. 2020. [Preprint] DOI: 10.1002/cbdv.202000295
  56. 56. Ashour A, Xue M, Al-Motawa M, Thornalley P, Rabbani N. Glycolytic overload-driven dysfunction of periodontal ligament fibroblasts in high glucose concentration, corrected by glyoxalase 1 inducer. BMJ Open Diabetes Research & Care. 2020;8(2):72-81
  57. 57. Murakami Y, Kawata A, Suzuki S, Fujisawa S. Radical-scavenging and pro-anti-inflammatory activity of tetracycline and related phenolic compounds with or without visible light irradiation. In Vivo. 2019;34(1):81-94
  58. 58. Wang Y, Zhao P, Sui B, Liu N, et al. Resveratrol enhances the functionality and improves the regeneration of mesenchymal stem cell aggregates. Experimental & Molecular Medicine. 2018;50(6):1-15
  59. 59. Yuan J, Wang X, Ma D, Gao H, et al. Resveratrol rescues TNF-α-induced inhibition of osteogenesis in human periodontal ligament stem cells via the ERK1/2 pathway. Molecular Medicine Reports. 2020;21:2085-2094. DOI: 10.3892/mmr.2020.11021
  60. 60. Kudo H, Takeichi O, Hatori K, Makino K, et al. A potential role for the silent information regulator 2 homologue 1 (SIRT1) in periapical periodontitis. International Endodontic Journal. 2018;51(7):747-757
  61. 61. Tesmer L, Lundy S, Sarkar S, Fox D. Th17 cells in human disease. Immunological Reviews. 2008;223(1):87-113
  62. 62. Bahar B, Singhrao S. An evaluation of the molecular mode of action of transresveratrol in the Porphyromonas gingivalis lipopolysaccharide challenged neuronal cell model. Molecular Biology Reports. 2020;48(1):147-156
  63. 63. Javid A, Hormoznejad R, Yousefimanesh H, Zakerkish M, et al. The impact of resveratrol supplementation on blood glucose, insulin, insulin resistance, triglyceride, and periodontal markers in type 2 diabetic patients with chronic periodontitis. Phytotherapy Research. 2016;31(1):108-114
  64. 64. Javid A, Hormoznejad R, Yousefimanesh H, Haghighi-zadeh M, Zakerkish M. Impact of resveratrol supplementation on inflammatory, antioxidant, and periodontal markers in type 2 diabetic patients with chronic periodontitis. Diabetes & Metabolic Syndrome: Clinical Research & Reviews. 2019;13(4):769-2774
  65. 65. Shoukheba M. Effect of resveratrol as an antioxidant in the treatment of smoker patients with stage III periodontitis. SAR Journal of Dentistry, Oral Surgery and Medicine. 2020;1(7):1-7. Available from: http://www.sarmedjournals.com/media/articles/SARJDOSM_11_1-7.pdf. [Accessed: 3 August 2022]
  66. 66. Zupančič Š, Lavrič Z, Kristl J. Stability and solubility of trans-resveratrol are strongly influenced by pH and temperature. European Journal of Pharmaceutics and Biopharmaceutics. 2015;93:196-204
  67. 67. Singh C, Ndiaye M, Ahmad N. Resveratrol and cancer: Challenges for clinical translation. Biochimica et Biophysica Acta (BBA)—Molecular Basis of Disease. 2015;1852(6):1178-1185
  68. 68. Smoliga J, Vang O, Baur J. Challenges of translating basic research into therapeutics: Resveratrol as an example. The Journals of Gerontology Series A: Biological Sciences and Medical Sciences. 2011;67A(2):158-167
  69. 69. Subramanian L, Youssef S, Bhattacharya S, Kenealey J, et al. Resveratrol: Challenges in translation to the clinic—A critical discussion. Clinical Cancer Research. 2010;16(24):5942-5948
  70. 70. Naseri N, Valizadeh H, Zakeri-Milani P. Solid lipid nanoparticles and nanostructured lipid carriers: Structure, preparation and application. Advanced Pharmaceutical Bulletin. 2015;5(3):305-313
  71. 71. Soeratri W, Hidayah R, Rosita N. Effect of combination soy bean oil and oleic acid to characteristic, penetration, physical stability of nanostructure lipid carrier resveratrol. Folia Medica Indonesiana. 2019;55(3):213
  72. 72. Poonia N, Kaur Narang J, Lather V, Beg S, Sharma T, Singh B, et al. Resveratrol loaded functionalized nanostructured lipid carriers for breast cancer targeting: Systematic development, characterization and pharmacokinetic evaluation. Colloids and Surfaces B: Biointerfaces. 2019;181:756-766
  73. 73. Qin L, Lu T, Qin Y, He Y, Cui N, et al. In vivo effect of resveratrol-loaded solid lipid nanoparticles to relieve physical fatigue for sports nutrition supplements. Molecules. 2020;25(22):5302. DOI: 10.3390/molecules25225302
  74. 74. Midha K, Nagpal N, Polonini, et al. Nano delivery systems of resveratrol to enhance its Oral bioavailability. American Journal of Pharmacy and Health Research. 2015;3(7). Available from: https://www.researchgate.net/publication/281062873_Nano_delivery_Systems_of_Resveratrol_to_Enhance_its_Oral_Bioavailability. [Accessed: 3 August 2022]
  75. 75. Zu Y, Overby H, Ren G, Fan Z, et al. Resveratrol liposomes and lipid Nanocarriers: Comparison of characteristics and inducing browning of white adipocytes. Colloids and Surfaces B: Biointerfaces. 2018;164:414-423
  76. 76. Singh G, Pai R. Trans-resveratrol self-nano-emulsifying drug delivery system (SNEDDS) with enhanced bioavailability potential: Optimization, pharmacokinetics and in situ single pass intestinal perfusion (spip) studies. Drug Delivery. 2014;22(4):522-530
  77. 77. Al-Bishri W. Attenuating impacts of chromium and nano resveratrol against hyperglycemia induced oxidative stress in diabetic rat. International Journal of Pharmaceutical Research & Allied Sciences. 2017;6(2):61-69
  78. 78. Ahmad M, Gani A. Development of novel functional snacks containing nanoencapsulated resveratrol with anti-diabetic, anti-obesity and antioxidant properties. Food Chemistry. 2021;352:129-323
  79. 79. Jayan H, Maria Leena M, Sivakama Sundari S, et al. Improvement of bioavailability for resveratrol through encapsulation in zein using electrospraying technique. Journal of Functional Foods. 2019;57:417-424
  80. 80. Peñalva R, Morales J, González-Navarro C, Larrañeta E, et al. Increased oral bioavailability of resveratrol by its encapsulation in casein nanoparticles. International Journal of Molecular Sciences. 2018;19(9):2816
  81. 81. Rabbani M, Pezeshki A, Ahmadi R, Mohammadi M, et al. Phytosomal Nanocarriers for encapsulation and delivery of resveratrol: Preparation, characterization, and application in mayonnaise. LWT. 2021;151:112093. DOI: 10.1016/j.jphotobiol.2020.111846
  82. 82. Berta G, Romano F, Vallone R, Abbadessa G, et al. An innovative strategy for oral biofilm control in early childhood based on a resveratrol-cyclodextrin nanotechnology approach. Materials. 2021;14(14):3801. DOI: 10.3390/ma14143801
  83. 83. Chen H, Tuck T, Ji X, Zhou X, Kelly G, Cuerrier A, et al. Quality assessment of Japanese knotweed (Fallopia japonica) grown on Prince Edward Island as a source of resveratrol. Journal of Agricultural and Food Chemistry. 2013;61(26):6383-6392
  84. 84. Gugleva V, Zasheva S, Hristova M, Andonova V. Topical use of resveratrol: Technological aspects. Pharmacia. 2020;67(2):89-94. DOI: 10.3897/pharmacia.67.e48472
  85. 85. Lim Y, Preshaw P, Lin H, Tan K. Resveratrol and its analogs as functional foods in periodontal disease management. Frontiers in Dental Medicine. 2021;2:1-13. DOI: 10.3389/fdmed.2021.636423

Written By

Tracey Lynn Harney

Submitted: 08 August 2022 Reviewed: 06 October 2022 Published: 07 November 2022