Next Article in Journal
Acute Respiratory Distress Syndrome: Focus on Viral Origin and Role of Pulmonary Lymphatics
Next Article in Special Issue
tRNA Derivatives in Multiple Myeloma: Investigation of the Potential Value of a tRNA-Derived Molecular Signature
Previous Article in Journal
Association of Mutations Identified in Xanthinuria with the Function and Inhibition Mechanism of Xanthine Oxidoreductase
Previous Article in Special Issue
MicroRNAs as Critical Biomarkers of Major Depressive Disorder: A Comprehensive Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MiRNAs as New Tools in Lesion Vitality Evaluation: A Systematic Review and Their Forensic Applications

1
Department of Surgical, Medical and Molecular Pathology and Critical Care Medicine, Section of Legal Medicine, University of Pisa, 56126 Pisa, Italy
2
Department of Anatomical, Histological, Forensic and Orthopedic Sciences, Sapienza University of Rome, 00186 Rome, Italy
*
Author to whom correspondence should be addressed.
Biomedicines 2021, 9(11), 1731; https://doi.org/10.3390/biomedicines9111731
Submission received: 26 October 2021 / Revised: 12 November 2021 / Accepted: 18 November 2021 / Published: 20 November 2021

Abstract

:
Wound vitality demonstration is one of the most challenging fields in forensic pathology. In recent years, researchers focused on the application of histological and immunohistochemical staining in this sphere of study. It is based on the detection of inflammation, red cell infiltration, and tissue alterations at the histological examination, all of which are supposedly present in antemortem rather than post-mortem wounds. Nevertheless, some doubts about the reliability of those markers have arisen. Furthermore, the lack of a standardized protocol and the operator dependency of this approach make the proper interpretation of its results difficult. Moreover, a differential miRNAs expression has been demonstrated in antemortem and post-mortem wounds. Herein, a systematic review concerning the current knowledge about the use of miRNAs in lesion vitality evaluation is carried out, to encourage researchers to deepen this peculiar study area. A compendium about the potential miRNAs that may be further investigated as vitality markers is also provided. The aim is to collect all available data about this topic to direct further studies on this field and highlight the future applications of miRNAs in forensic pathology. We found 20 articles and a total of 51 miRNAs that are involved in inflammation and wound healing. Further studies are certainly needed to deepen the role of miRNAs in inflammatory processes in lesioned skin and to evaluate their reliability in distinguishing between antemortem and post-mortem lesions.

1. Introduction

Wound vitality demonstration is one of the most challenging fields in forensic pathology. It has been classically based on the presence of inflammation, red cell infiltration, and tissue alterations at the histological examination [1,2,3,4]. However, the differentiation between vital reactions and post-mortem changes is not always clear [5,6,7]. In recent years, researchers focused on the detection of inflammatory cells, cytokines, apoptosis mediators, and other markers that are supposedly present in antemortem rather than post-mortem wounds, through immunohistochemical stains [6,7,8,9,10,11]. Nevertheless, some doubts about the reliability of these methods have arisen, and the lack of a standardized protocol and the operator dependency of this approach makes it difficult to correctly interpret the results [12].
MicroRNAs (miRNAs) are short single-strand non-coding ribonucleic acids that have a fundamental role in the post-transcriptional regulation of gene expression, mainly inhibiting the mRNA translation [13,14,15]. Due to their ubiquity, they can be used as a diagnostic tool in different clinical areas, as well as in forensic pathology [16]. A differential miRNA expression in antemortem and post-mortem wounds has been observed [17,18,19]. This evidence opens new and promising possibilities to use them as lesion vitality markers. It is unlikely that a method alone could provide the trustworthiness needed in judicial cases, but the combination of histological, immunohistochemical, and genetic analysis may become a new tool in vitality diagnosis.

2. Materials and Methods

The present systematic review was carried out according to the Preferred Reporting Items for Systematic Review (PRISMA) standards [20]. A systematic literature search and a critical review of the collected studies were conducted. An electronic search of PubMed, Science Direct Scopus, Google Scholar, and Excerpta Medica Database (EMBASE) from database inception to November 2020 was performed. The search terms were “miRNAs”, “wound vitality”, and “lesions vitality” in the title, abstract, and keywords. The bibliographies of all located papers were examined and cross-referenced to further identify relevant literature. A methodological appraisal of each study was conducted according to the PRISMA standards, including an evaluation of bias. The data collection process included study selection and data extraction. Two researchers (A.C.M., E.M.) independently examined the papers with titles or abstracts that appeared to be relevant and selected those that analyzed miRNAs involved in wound vitality demonstration and miRNAs involved in wound healing that may be further investigated as vitality markers in post-mortem samples. Researchers resolved their disagreement concerning works eligibility by consensus. Only papers in English were included in the research. Two investigators performed data extraction (A.B., E.M.), and two other investigators verified them (A.B., E.M.), which were again verified by two other investigators (A.M., E.T.). This study was exempt from institutional review board approval, as it did not involve human subjects.

3. Results

The search performed as described above identified 197 articles, which were screened to exclude duplicates. The resulting 163 reference lists were then screened based on their title and abstract, which left 87 articles for further consideration. Non-English papers were excluded. The following inclusion criteria were used: (1) original research articles, (2) reviews and mini-reviews, and (3) case reports/series. These publications were carefully evaluated, considering the main aims of the review. This evaluation left 20 scientific papers comprising original research articles, case reports, and case series.
Figure 1 illustrates our search strategy.
Table 1 shows the studies included in this review with a brief description.
As shown in Table 2, which summarizes the main characteristics of the articles included in this review, most of the founded studies are about miRNAs involved in wound healing and therefore they are based on in vitro and/or in vivo experiments [17,18,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37].
Three studies used human skin lesion samples collected post-mortem (2) and antemortem (1) [19,29,37]. Animal studies were based on antemortem samples, mainly collected within a precise interval after wounding (days), while antemortem human skin samples were hypertrophic skin scars specimens collected during surgical procedures. In vitro and antemortem studies investigated the role of miRNAs in wound healing and, as previously mentioned, they were included in this review because they could suggest new miRNAs to study for wound vitality demonstration in post-mortem samples. Moreover, antemortem studies provide new knowledge about the expression of miRNAs in wound healing in different timeframes, and consequently, they may be used in further studies to determine the age of a lesion. Concerning the type of skin lesion, the articles investigated miRNAs expression in incisional [18] and excisional wound [21,22,23,26,27,30,31,33,34,35], ligature mark [19], burned skin [37], chronic ulcer [32], and hypertrophic scar [29]. There is heterogeneity among researchers concerning the type of adopted investigation. Real-time polymerase chain reaction (RT-PCR) was performed in 17 studies out of 20. In 11 studies, Western blot was used for protein analysis [21,22,24,25,26,27,28,30,32,33,36]. Histological examination of the wounded skin was performed in 13 out of 20 studies, but only in 8 studies, this evaluation was supplemented by immunohistochemical analysis and in 5 by hybridization in situ [17,18,19,21,22,23,24,25,27,28,29,32,33,34,35,36]. Immunofluorescence analysis was performed in two studies, while a morphometric analysis appeared in the work of Long and coll [31]. Flow cytometry analysis was used by Lin et al. to isolate fat cells [28]. Phase-contrast morphometry was performed in one study [36]. In all the studies included in this review, a statistical analysis on obtained results was performed, except for Viticchiè et al., Yang et al., and Yu et al. [33,35,36].
As a result of our research, we found that 51 miRNAs are implied in wound healing and wound vitality demonstration. To understand the molecular pathways in which these miRNAs are involved, we collected their target genes and/or proteins, when specified by the included papers, as shown in Table 3.
Some miRNAs have a differential expression over time after wounding. This “day-by-day” variation may help determine the age of a lesion. According to the studies performed by Lin et al. and Yu et al., miR-205 shows an increase during the first days in epithelial cells, while on the contrary, Etich et al. found that it is downregulated from day 1 to day 10 after wounding [23,28,36]. MiR-483-3p, studied by Bertero et al., starts to increase 3 days after the production of the wound and leads to its peak at 6–7 days [21]. MiR-21 seems to rise, and Yang et al. demonstrated its increase over three days [35]. In the presence of H2O2, both miR-19b and miR-124 levels decline [22,24]. Etich et al. proved the downregulation of miR-204 and miR-205 (respectively, from day 5 to 10, and from day 1 to day 7), and the contemporaneous overexpression of miR-31 [23]. Moreover, Wang et al. attested mir-31 increase, along with miR-21, and miR-203; miR-249 is the only one that decreases 7 days after injuring [34]. Another investigation about miR-16, miR-20a, miR-106a, miR-130a, and miR-203 in chronic ulcers revealed their growth in the epidermis [32]. In the same investigation, miR-21 levels were found overexpressed in all skin layers. Similarly, miR-152, miR-365, miR-125a-5p, miR-181d, miR-99, miR-100, miR-30c, and miR-125b-5p were downexpressed from day 1, returning to normal levels after five days [26]. A similar trend has been described for miR-149, miR-203a, miR-222, and miR-122 [29]. Lyu et al. proved the overexpression of 19 different miRNAs and the simultaneous decrease in five miRNAs in burned skin [17]. In another experiment conducted on burned skin, the authors established the upregulation of miR-711 and miR-183-3p 48 h after excision in human skin [37]. The same experiment, performed on mice burned skin by the same Authors, revealed that their rising occurred up to 120 h after wounding. MiR-203 has been investigated in five studies, which showed it is increased after wounding except in one case, where a decrease in the epidermis, dermis, and subcutaneous fat over four days was demonstrated [29,30,32,33,34]. Neri et al. studied the expression of five different miRNAs in ligature marks, showing the elevation of miR125a-5, miR125b-5, miR130a-3, miR214-3p, and miR92a-3p [19].
In summary, during the first day after wounding, it is possible to find the overexpression of at least 10 different miRNAs (miR-205, miR-152, miR-365, miR-125a-5p, miR-181d, MiR-99, miR-100, miR-30c, miR-125b-5p, and miR203) and the contemporaneous decreasing of miR-21. On day 2, miR-205, miR-21, miR-711, and miR-183-3p levels increase. The expression of miR-483-3p, miR-203, and miR-21 rises on day 3. From the fifth day, miR-31 and miR-203 levels rise, while miR-204 expression decreases. Eventually, from day 7, miR-249 starts to decrease.

4. Discussion

MiRNAs have attracted the attention of researchers in various fields [38,39,40,41]. They influence many signaling pathways involved in inflammation [42,43,44,45]. Due to this property, miRNAs could be used in vitality wound demonstration. In this review, we present many miRNAs that have a role in wound healing and, therefore, may be investigated as post-mortem vitality markers.
Regarding the molecular pathways involved in the inflammation and healing processes highlighted in this work, miR-19-b regulates the transforming growth factor-β (TGF-β) signaling pathway by targeting C-C motif chemokine ligand 1 (CCL1) [46]. The TGF-β pathway, determinant in wound contraction, can also be modified by miR-21 and miR-149 [27,34]. Under inflammatory conditions, miR-149 also inhibits the interleukin (IL) 1α, IL-1β, and IL-6 genes, as well as the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) pathway [47]. NFκB pathway is activated by the tumor necrosis factor (TNFα) phosphorylating the inhibitor of nuclear factor kappa B (IκB) [48]. In keratinocytes, MiR-205 and miR-184 interact in regulating the SH2-containing phosphatidylinositol 3,4,5-trisphosphate 5-phosphatase (SHIP2) levels and phosphorous protein kinase B (AKT) signaling [49]. Specifically, miR-184 maintains SHIP2 levels through miR-205 expression [50]. This action assesses its effects on cytoskeletal remodeling, cell adhesion, and spreading, owing to the downregulation of miR-205. MiR-205 raises filamentous actin and p-cofilin via Rho activation; therefore, its downregulation results in central control of actin dynamics and cell motility [51]. It is also involved in keratinocytes’ survival function through two signaling pathways acting on p-Akt: inactivation of SHIP2 and activation of the phosphor-ezrin/radixin/moesin (p-ERM) [36,52]. MiR-122 is another RNA involved in cells migration owing to its influence on both the mitogen-activated protein kinase (MAPK) pathway and focal adhesion (CAM, cells adhesion molecules) [29,53]. Numerous 99- family miRNAs influence mTOR (mechanistic target of rapamycin kinase) by PI3K/AKT regulation, the upstream pathway [54]. PI3K/AKT pathway is also induced by mi-RNA26a [55]. Jiang et al. reported that miR-26a induces the expression of ITGA5 by upregulating the PI3K/AKT pathway [25]. Principally, through the integrin subunit alpha 5 (ITGA5) downregulation, miR-26a inhibits keratinocytes migration’s decreasing and therefore wound healing. This system permits the triggering of TGF-β1, thus cellular migration. Remarkably, the integrin subunit alpha 5 (ITGA5) overexpression could oppose the inhibitory effect of miR-26a [56]. miR-99 family members regulate cell proliferation, apoptosis, and migration [57]. Li et al. found that miR-222 influences the Wingless-related integration site (Wnt) signal, by targeting multiple genes, for example, Dickkopf-related protein 2 (DDK2), axin-like protein 2 (AXIN2), and FRAT regulator of WNT signaling pathway 2 (FRAT2) [29]. The MAPK pathway is induced by both miR-222 and miR-149 [58]. MiR-21 has been correlated with repression of growth, increased apoptosis, and cell cycle by its downregulation [59]. Moreover, miR-21 is involved with miR-203 in leptin receptor (LepR) regulation and, together with MiR-130a and miR-203, regulates early growth response 3 (EGR3) expression [32]. Other authors highlighted the action of miR-203 on inhibiting the proliferation process of epidermal stem cells in chronic ulcers [29,30,32,33,34,60]. Probably when miR-203 is expressed, tumor protein p63 (P63) and protein K15 (K15) and integrin B levels decreased remarkably [13]. Furthermore, overexpression of this miRNA decreases the expression of Rho-associated coiled-coil-containing protein kinase 2 (ROCK2), MAPK8, MAPK9, and protein kinase C alpha (PRKCA) by altering the Wnt/B-catenin signaling pathway [61]. Yang et al. demonstrate that miR-21 promotes keratinocyte migration and re-epithelialization of wounding inhibiting, respectively, the expression of metalloproteinase-3 (TIMP3) and T-cell lymphoma invasion and metastasis 1 (TIAM1) [35]. TGF-β-induced miR-21 expression leads to a significant tissue contraction, and in addition, it accelerates re-epithelialization [14]. Moreover, miR-21 has been linked to the activation of keloid fibroblasts and fibrosis [62]. MiR-203 has an important role in early inflammation and angiogenesis by controlling pro-migratory and pro-proliferative factors such as p63, LASP1 (LIM and SH3 protein 1), ras-related nuclear protein (RAN), RAPH1 (Ras association (RalGDS/AF-6) and pleckstrin homology domains 1) [33,63]. Jin et al. proved a correlation between PI3K/AKT, mTOR, insulin-like growth factor 1 receptor (IGF1R), and multiple miRNAs: miR-152, miR-365, miR-125a, and b, miR-181d, miR-100, and miR-30c [26]. These miRNAs are the fundamental regulator of the mTOR and MAPK signaling pathway in the skin [64].
In conclusion, by comparing the difference in the expression of miRNAs in the wound healing process, these biomarkers could become useful for the chronological diagnosis of the lesions [65], as shown in Figure 2.
Nevertheless, the reader should bear in mind that our work has some limitations. The pool of papers is heterogeneous, and it includes different kinds of studies, counting in vitro, in vivo, and human samples. Moreover, the small sample size (only 20 papers included) could interfere with the reliability of our findings. Certainly, more studies are needed in this field. Once the role of miRNAs in the wound healing process has been completely clarified, we hope to implement this knowledge in forensic applications soon thereafter. For future applications, it is conceivable to combine miRNAs analysis and histological and immunohistochemical tissue investigation, which have already been widely investigated in the forensic literature [1,2,3,4,6,7,8,9,10,11]. As in many other fields, miRNAs investigation carries great potential in forensic science [66,67,68]. A possible project could be an interlaboratory study to investigate the miRNAs differential expression in the same lesion and with the same analytical procedure, in addition to histological and immunohistochemical evaluation. The aim is to create a standardized protocol and start implementing these analyses since the reliability and reproducibility of analytical methodologies are the foundation for acceptance in legal proceedings.

5. Conclusions

Even if in the last years, some studies have provided new evidence about the forensic applications of miRNAs, much evidence still needs to be disclosed. MiRNAs are involved in many molecular pathways, and, owing to their ubiquity, they have been considered as markers in several fields. Herein, a compendium of the potential miRNAs that should be further investigated as wound vitality and age markers was provided. We found 20 articles on miRNAs involvement in skin wound and healing, but only 2 of them specifically referred to vitality demonstration in the forensic field. Further studies are certainly needed to deepen the role of miRNAs in inflammatory processes in wounded skin and to evaluate their reliability in distinguishing between antemortem and post-mortem lesions.

Author Contributions

Conceptualization, A.B.; methodology, A.M. and P.F.; validation, A.M. and E.T.; data curation, E.M.; writing—original draft preparation, A.C.M. and E.M.; writing—review and editing, A.M. and A.B.; supervision, V.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saukko, P.; Knight, B. Knight’s Forensic Pathology, 4th ed.; CRC Press: Boca Raton, FL, USA, 2016. [Google Scholar]
  2. Grellner, W.; Georg, T.; Wilske, J. Quantitative analysis of proinflammatory cytokines (IL-1beta, IL-6, TNF-alpha) in human skin wounds. Forensic Sci. Int. 2000, 113, 251–264. [Google Scholar] [CrossRef]
  3. Bonelli, A.; Bacci, S.; Norelli, G.A. Affinity cytochemistry analysis of mast cells in skin lesions: A possible tool to assess the timing of lesions after death. Int. J. Legal. Med. 2003, 117, 331–334. [Google Scholar] [CrossRef] [PubMed]
  4. Casse, J.M.; Martrille, L.; Vignaud, J.M.; Gauchotte, G. Skin wounds vitality markers in forensic pathology: An updated review. Med. Sci. Law. 2016, 56, 128–137. [Google Scholar] [CrossRef] [PubMed]
  5. Oehmichen, M. Vitality and time course of wounds. Forensic Sci. Int. 2004, 144, 221–231. [Google Scholar] [CrossRef]
  6. Langlois, N.E.; Gresham, G.A. The ageing of bruises: A review and study of the colour changes with time. Forensic Sci. Int. 1991, 50, 227–238. [Google Scholar] [CrossRef]
  7. Balandiz, H.; Pehlivan, S.; Çiçek, A.F.; Tuğcu, H. Evaluation of Vitality in the Experimental Hanging Model of Rats by Using Immunohistochemical IL-1β Antibody Staining. Am. J. Forensic Med. Pathol. 2015, 36, 317–322. [Google Scholar] [CrossRef]
  8. Focardi, M.; Puliti, E.; Grifoni, R.; Palandrini, M.; Bugelli, V.; Pinchi, V.; Norelli, G.A.; Bacci, S. Immunohistochemical localization of Langerhans cells as a tool for vitality in -hanging mark wounds: A pilot study. Aust. J. Forensic Sci. 2019, 52, 393–405. [Google Scholar] [CrossRef]
  9. Ishida, Y.; Kuninaka, Y.; Nosaka, M.; Shimada, E.; Hata, S.; Yamamoto, H.; Hashizume, Y.; Kimura, A.; Furukawa, F.; Kondo, T. Forensic application of epidermal AQP3 expression to determination of wound vitality in human compressed neck skin. Int. J. Legal. Med. 2018, 132, 1375–1380. [Google Scholar] [CrossRef]
  10. Turillazzi, E.; Vacchiano, G.; Luna-Maldonado, A.; Neri, M.; Pomara, C.; Rabozzi, R.; Riezzo, I.; Fineschi, V. Tryptase, CD15 and IL-15 as reliable markers for the determination of soft and hard ligature marks vitality. Histol. Histopathol. 2010, 25, 1539–1546. [Google Scholar] [PubMed]
  11. De Matteis, A.; dell’Aquila, M.; Maiese, A.; Frati, P.; La Russa, R.; Bolino, G.; Fineschi, V. The Troponin-I fast skeletal muscle is reliable marker for the determination of vitality in the suicide hanging. Forensic Sci. Int. 2019, 301, 284–288. [Google Scholar] [CrossRef]
  12. Pérez, I.L.; Falcón, M.; Gimenez, M.; Diaz, F.M.; Pérez-Cárceles, M.D.; Osuna, E.; Nuno-Vieira, D.; Luna, A. Diagnosis of Vitality in Skin Wounds in the Ligature Marks Resulting From Suicide Hanging. Am. J. Forensic Med. Pathol. 2017, 38, 211–218. [Google Scholar] [CrossRef] [PubMed]
  13. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, J.; Chen, J.; Sen, S. MicroRNA as Biomarkers and Diagnostics. J. Cell. Physiol. 2016, 231, 25–30. [Google Scholar] [CrossRef]
  15. Maiese, A.; Scatena, A.; Costantino, A.; Di Paolo, M.; La Russa, R.; Turillazzi, E.; Frati, P.; Fineschi, V. MicroRNAs as Useful Tools to Estimate Time Since Death. A Systematic Review of Current Literature. Diagnostics 2021, 11, 64. [Google Scholar] [CrossRef] [PubMed]
  16. Rocchi, A.; Chiti, E.; Maiese, A.; Turillazzi, E.; Spinetti, I. MicroRNAs: An Update of Applications in Forensic Science. Diagnostics 2020, 11, 32. [Google Scholar] [CrossRef]
  17. Lyu, H.P.; Cheng, M.; Liu, J.C.; Ye, M.Y.; Xu, D.; He, J.T.; Xie, X.L.; Wang, Q. Differentially expressed microRNAs as potential markers for vital reaction of burned skin. Forensic Sci. Med. 2018, 4, 135–141. [Google Scholar]
  18. Ibrahim, S.F.; Ali, M.M.; Basyouni, H.; Laila, A.; Rashed, E.; Amer, A.E.; Abd El-Kareem, D. Histological and miRNAs postmortem changes in incisional wound. Egypt J. Forensic Sci. 2019, 9, 37. [Google Scholar] [CrossRef] [Green Version]
  19. Neri, M.; Fabbri, M.; D’Errico, S.; Di Paolo, M.; Frati, P.; Gaudio, R.M.; La Russa, R.; Maiese, A.; Marti, M.; Pinchi, E.; et al. Regulation of miRNAs as new tool for cutaneous vitality lesions demonstration in ligature marks in deaths by hanging. Sci. Rep. 2019, 9, 20011. [Google Scholar] [CrossRef]
  20. Liberati, A.; Altman, D.G.; Tetzlaff, J.; Mulrow, C.; Gøtzsche, P.C.; Ioannidis, J.P.; Clarke, M.; Devereaux, P.J.; Kleijnen, J.; Moher, D. The PRISMA statement for reporting systematic reviews and meta-analyses of studies that evaluate healthcare interventions: Explanation and elaboration. BMJ. 2009, 339, b2700. [Google Scholar] [CrossRef] [Green Version]
  21. Bertero, T.; Gastaldi, C.; Bourget-Ponzio, I.; Imbert, V.; Loubat, A.; Selva, E.; Busca, R.; Mari, B.; Hofman, P.; Barbry, P.; et al. miR-483-3p controls proliferation in wounded epithelial cells. FASEB J. 2011, 25, 3092–3105. [Google Scholar] [CrossRef]
  22. Cao, G.; Chen, B.; Zhang, X.; Chen, H. Human Adipose-Derived Mesenchymal Stem Cells-Derived Exosomal microRNA-19b Promotes the Healing of Skin Wounds Through Modulation of the CCL1/TGF-β Signaling Axis. Clin. Cosmet. Investig. Dermatol. 2020, 13, 957–971. [Google Scholar] [CrossRef]
  23. Etich, J.; Bergmeier, V.; Pitzler, L.; Brachvogel, B. Identification of a reference gene for the quantification of mRNA and miRNA expression during skin wound healing. Connect. Tissue. Res. 2017, 58, 196–207. [Google Scholar] [CrossRef]
  24. He, L.; Zhu, C.; Jia, J.; Hao, X.Y.; Yu, X.Y.; Liu, X.Y.; Shu, M.G. ADSC-Exos containing MALAT1 promotes wound healing by targeting miR-124 through activating Wnt/β-catenin pathway. Biosci. Rep. 2020, 40, BSR20192549. [Google Scholar] [CrossRef]
  25. Jiang, Z.; Wei, J.; Yang, W.; Li, W.; Liu, F.; Yan, X.; Yan, X.; Hu, N.; Li, J. MicroRNA-26a inhibits wound healing through decreased keratinocytes migration by regulating ITGA5 through PI3K/AKT signaling pathway. Biosci. Rep. 2020, 40, BSR20201361. [Google Scholar] [CrossRef]
  26. Jin, Y.; Tymen, S.D.; Chen, D.; Fang, Z.J.; Zhao, Y.; Dragas, D.; Dai, Y.; Marucha, P.T.; Zhou, X. MicroRNA-99 family targets AKT/mTOR signaling pathway in dermal wound healing. PLoS ONE 2013, 8, e64434. [Google Scholar] [CrossRef]
  27. Lang, H.; Zhao, F.; Zhang, T.; Liu, X.; Wang, Z.; Wang, R.; Shi, P.; Pang, X. MicroRNA-149 contributes to scarless wound healing by attenuating inflammatory response. Mol. Med. Rep. 2017, 16, 2156–2162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Lin, D.; Halilovic, A.; Yue, P.; Bellner, L.; Wang, K.; Wang, L.; Zhang, C. Inhibition of miR-205 impairs the wound-healing process in human corneal epithelial cells by targeting KIR4.1 (KCNJ10). Investig. Ophthalmol. Vis. Sci. 2013, 54, 6167–6178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Li, P.; He, Q.; Luo, C.; Qian, L. Differentially expressed miRNAs in acute wound healing of the skin: A pilot study. Medicine 2015, 94, 458. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, J.; Shu, B.; Zhou, Z.; Xu, Y.; Liu, Y.; Wang, P.; Xiong, K.; Xie, J. Involvement of miRNA203 in the proliferation of epidermal stem cells during the process of DM chronic wound healing through Wnt signal pathways. Stem. Cell Res. Ther. 2020, 11, 348. [Google Scholar] [CrossRef]
  31. Long, S.; Zhao, N.; Ge, L.; Wang, G.; Ran, X.; Wang, J.; Su, Y.; Wang, T. MiR-21 ameliorates age-associated skin wound healing defects in mice. J. Gene Med. 2018, 20, e3022. [Google Scholar] [CrossRef]
  32. Pastar, I.; Khan, A.A.; Stojadinovic, O.; Lebrun, E.A.; Medina, M.C.; Brem, H.; Kirsner, R.S.; Jimenez, J.J.; Leslie, C.; Tomic-Canic, M. Induction of specific microRNAs inhibits cutaneous wound healing. J. Biol. Chem. 2012, 287, 29324–29335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Viticchiè, G.; Lena, A.M.; Cianfarani, F.; Odorisio, T.; Annicchiarico-Petruzzelli, M.; Melino, G.; Candi, E. MicroRNA-203 contributes to skin re-epithelialization. Cell Death Dis. 2012, 3, e435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, T.; Feng, Y.; Sun, H.; Zhang, L.; Hao, L.; Shi, C.; Wang, J.; Li, R.; Ran, X.; Su, Y.; et al. miR-21 regulates skin wound healing by targeting multiple aspects of the healing process. Am. J. Pathol. 2012, 181, 1911–1920. [Google Scholar] [CrossRef] [PubMed]
  35. Yang, X.; Wang, J.; Guo, S.L.; Fan, K.J.; Li, J.; Wang, Y.L.; Teng, Y.; Yang, X. miR-21 promotes keratinocyte migration and re-epithelialization during wound healing. Int. J. Biol. Sci. 2011, 7, 685–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Yu, J.; Peng, H.; Ruan, Q.; Fatima, A.; Getsios, S.; Lavker, R.M. MicroRNA-205 promotes keratinocyte migration via the lipid phosphatase SHIP2. FASEB J. 2010, 24, 3950–3959. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, K.; Cheng, M.; Xu, J.; Chen, L.; Li, J.; Li, Q.; Xie, X.; Wang, Q. MiR-711 and miR-183-3p as potential markers for vital reaction of burned skin. Forensic Sci. Res. 2020. [Google Scholar] [CrossRef]
  38. Pinchi, E.; Frati, A.; Cantatore, S.; D’Errico, S.; La Russa, R.; Maiese, A.; Palmieri, M.; Pesce, A.; Viola, R.V.; Frati, P.; et al. Acute Spinal Cord Injury: A Systematic Review Investigating miRNA Families Involved. Int. J. Mol. Sci. 2019, 20, 1841. [Google Scholar] [CrossRef] [Green Version]
  39. Pinchi, E.; Frati, P.; Aromatario, M.; Cipolloni, L.; Fabbri, M.; La Russa, R.; Maiese, A.; Neri, M.; Santurro, A.; Scopetti, M.; et al. miR-1, miR-499 and miR-208 are sensitive markers to diagnose sudden death due to early acute myocardial infarction. J. Cell. Mol. Med. 2019, 23, 6005–6016. [Google Scholar] [CrossRef] [Green Version]
  40. Manetti, A.C.; Maiese, A.; Di Paolo, M.; De Matteis, A.; La Russa, R.; Turillazzi, E.; Frati, P.; Fineschi, V. MicroRNAs and Sepsis-Induced Cardiac Dysfunction: A Systematic Review. Int. J. Mol. Sci. 2020, 22, 321. [Google Scholar] [CrossRef]
  41. Chiti, E.; Di Paolo, M.; Turillazzi, E.; Rocchi, A. MicroRNAs in Hypertrophic, Arrhythmogenic and Dilated Cardiomyopathy. Diagnostics 2021, 11, 1720. [Google Scholar] [CrossRef]
  42. Tahamtan, A.; Teymoori-Rad, M.; Nakstad, B.; Salimi, V. Anti-Inflammatory MicroRNAs and Their Potential for Inflammatory Diseases Treatment. Front. Immunol. 2018, 9, 1377. [Google Scholar] [CrossRef] [Green Version]
  43. Marques-Rocha, J.L.; Samblas, M.; Milagro, F.I.; Bressan, J.; Martínez, J.A.; Marti, A. Noncoding RNAs, cytokines, and inflammation-related diseases. FASEB J. 2015, 29, 3595–3611. [Google Scholar] [CrossRef] [Green Version]
  44. Brudecki, L.; Ferguson, D.A.; McCall, C.E.; El Gazzar, M. MicroRNA-146a and RBM4 form a negative feed-forward loop that disrupts cytokine mRNA translation following TLR4 responses in human THP-1 monocytes. Immunol. Cell Biol. 2013, 91, 532–540. [Google Scholar] [CrossRef]
  45. Sessa, F.; Salerno, M.; Bertozzi, G.; Cipolloni, L.; Messina, G.; Aromatario, M.; Polo, L.; Turillazzi, E.; Pomara, C. miRNAs as Novel Biomarkers of Chronic Kidney Injury in Anabolic-Androgenic Steroid Users: An Experimental Study. Front. Pharmacol. 2020, 11, 563756. [Google Scholar] [CrossRef]
  46. Chang, R.; Yi, S.; Tan, X.; Huang, Y.; Wang, Q.; Su, G.; Zhou, C.; Cao, Q.; Yuan, G.; Kijlstra, A.; et al. MicroRNA-20a-5p suppresses IL-17 production by targeting OSM and CCL1 in patients with Vogt-Koyanagi-Harada disease. Br. J. Ophthalmol. 2018, 102, 282–290. [Google Scholar] [CrossRef] [Green Version]
  47. Pasparakis, M. Regulation of tissue homeostasis by NF-kappaB signalling: Implications for inflammatory diseases. Nat. Rev. Immunol. 2009, 9, 778–788. [Google Scholar] [CrossRef]
  48. Balkwill, F. TNF-alpha in promotion and progression of cancer. Cancer Metastasis Rev. 2006, 25, 409–416. [Google Scholar] [CrossRef] [PubMed]
  49. Pullen, N.; Thomas, G. The modular phosphorylation and activation of p70s6k. FEBS Lett. 1997, 410, 78–82. [Google Scholar] [CrossRef] [Green Version]
  50. Suwa, A.; Kurama, T.; Shimokawa, T. SHIP2 and its involvement in various diseases. Expert Opin. Ther. Targets 2010, 14, 727–737. [Google Scholar] [CrossRef] [PubMed]
  51. Jiang, L.; Liu, X.; Kolokythas, A.; Yu, J.; Wang, A.; Heidbreder, C.E.; Shi, F.; Zhou, X. Downregulation of the Rho GTPase signaling pathway is involved in the microRNA-138-mediated inhibition of cell migration and invasion in tongue squamous cell carcinoma. Int. J. Cancer 2010, 127, 505–512. [Google Scholar] [CrossRef] [PubMed]
  52. Su, N.; Qiu, H.; Chen, Y.; Yang, T.; Yan, Q.; Wan, X. miR-205 promotes tumor proliferation and invasion through targeting ESRRG in endometrial carcinoma. Oncol. Rep. 2013, 29, 2297–2302. [Google Scholar] [CrossRef]
  53. Wang, Z.; Wang, X. miR-122-5p promotes aggression and epithelial-mesenchymal transition in triple-negative breast cancer by suppressing charged multivesicular body protein 3 through mitogen-activated protein kinase signaling. J. Cell. Physiol. 2020, 235, 2825–2835. [Google Scholar] [CrossRef] [PubMed]
  54. Lima, M.H.; Caricilli, A.M.; de Abreu, L.L.; Araújo, E.P.; Pelegrinelli, F.F.; Thirone, A.C.; Tsukumo, D.M.; Pessoa, A.F.; dos Santos, M.F.; de Moraes, M.A.; et al. Topical insulin accelerates wound healing in diabetes by enhancing the AKT and ERK pathways: A double-blind placebo-controlled clinical trial. PLoS ONE 2012, 7, e36974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Liang, Z.; Chi, Y.J.; Lin, G.Q.; Luo, S.H.; Jiang, Q.Y.; Chen, Y.K. MiRNA-26a promotes angiogenesis in a rat model of cerebral infarction via PI3K/AKT and MAPK/ERK pathway. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 3485–3492. [Google Scholar] [PubMed]
  56. Eniafe, J.; Jiang, S. MicroRNA-99 family in cancer and immunity. Wiley Interdiscip. Rev. RNA 2021, 12, e1635. [Google Scholar] [CrossRef]
  57. Clevers, H.; Nusse, R. Wnt/β-catenin signaling and disease. Cell 2012, 149, 1192–1205. [Google Scholar] [CrossRef] [Green Version]
  58. Gorostizaga, A.; Brion, L.; Maloberti, P.; Cornejo Maciel, F.; Podestá, E.J.; Paz, C. Heat shock triggers MAPK activation and MKP-1 induction in Leydig testicular cells. Biochem. Biophys. Res. Commun. 2005, 327, 23–28. [Google Scholar] [CrossRef]
  59. Song, B.; Wang, C.; Liu, J.; Wang, X.; Lv, L.; Wei, L.; Xie, L.; Zheng, Y.; Song, X. MicroRNA-21 regulates breast cancer invasion partly by targeting tissue inhibitor of metalloproteinase 3 expression. J. Exp. Clin. Cancer Res. 2010, 29, 29. [Google Scholar] [CrossRef] [Green Version]
  60. Manetti, A.C.; Baronti, A.; Bosetti, C.; Costantino, A.; Di Paolo, M.; Turillazzi, E.; Maiese, A. Bleeding varicose veins’ ulcer as a cause of death: A case report and review of the current literature. Clin. Ter. 2021, 172, 395–406. [Google Scholar]
  61. Whyte, J.L.; Smith, A.A.; Helms, J.A. Wnt signaling and injury repair. Cold. Spring Harb. Perspect. Biol. 2012, 4, a008078. [Google Scholar] [CrossRef]
  62. Wang, T.; Zhang, L.; Shi, C.; Sun, H.; Wang, J.; Li, R.; Zou, Z.; Ran, X.; Su, Y. TGF-β-induced miR-21 negatively regulates the antiproliferative activity but has no effect on EMT of TGF-β in HaCaT cells. Int. J. Biochem. Cell Biol. 2012, 44, 366–376. [Google Scholar] [CrossRef]
  63. Vanbokhoven, H.; Melino, G.; Candi, E.; Declercq, W. p63, a story of mice and men. J. Investig. Dermatol. 2011, 131, 1196–1207. [Google Scholar] [CrossRef] [PubMed]
  64. Bonni, A.; Brunet, A.; West, A.E.; Datta, S.R.; Takasu, M.A.; Greenberg, M.E. Cell survival promoted by the Ras-MAPK signaling pathway by transcription-dependent and -independent mechanisms. Science 1999, 286, 1358–1362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Grellner, W.; Madea, B. Demands on scientific studies: Vitality of wounds and wound age estimation. Forensic Sci. Int. 2007, 165, 150–154. [Google Scholar] [CrossRef]
  66. Grellner, W.; Vieler, S.; Madea, B. Transforming growth factors (TGF-alpha and TGF-beta1) in the determination of vitality and wound age: Immunohistochemical study on human skin wounds. Forensic Sci. Int. 2005, 153, 174–180. [Google Scholar] [CrossRef] [PubMed]
  67. Madea, B.; Doberentz, E.; Jackowski, C. Vital reactions—An updated overview. Forensic Sci. Int. 2019, 305, 110029. [Google Scholar] [CrossRef] [PubMed]
  68. Courts, C.; Madea, B. Micro-RNA—A potential for forensic science? Forensic Sci. Int. 2010, 203, 106–111. [Google Scholar] [CrossRef]
Figure 1. Search strategy following PRISMA standards, including an evaluation of bias.
Figure 1. Search strategy following PRISMA standards, including an evaluation of bias.
Biomedicines 09 01731 g001
Figure 2. This figure represents the progressive expression of different miRNAs in time (days) after wounding. Only those miRNAs that are expressed according to a precise timeframe were considered. Only the first day of increase is shown. * miR-183-3p shows a different behavior between humans and rats; it was detected overexpressed in humans over 2 days and in rats over the course of 120 days. ↑ indicates up regulation, ↓ indicates down regulation.
Figure 2. This figure represents the progressive expression of different miRNAs in time (days) after wounding. Only those miRNAs that are expressed according to a precise timeframe were considered. Only the first day of increase is shown. * miR-183-3p shows a different behavior between humans and rats; it was detected overexpressed in humans over 2 days and in rats over the course of 120 days. ↑ indicates up regulation, ↓ indicates down regulation.
Biomedicines 09 01731 g002
Table 1. This table resumes all the studies implied in our revision. RT-PCR, when reported, is quantitative. Statistical analysis was performed in all studies except in Viticchiè et al., 2012, Yang et al., 2011, and Yu et al., 2010. AM, antemortem; H, histology; HaCaT, human keratinocyte cell line; HCECs, corneal epithelial cells; HADSCs, human subcutaneous adipose tissue; HEKs, human keratinocytes; HSF, human skin fibroblasts; IHC, immunohistochemistry; IHF, immunohistofluorescence; ISH, hybridization in situ; MP, morphometric analysis; NP, not performed; NS, not specified; PM, post-mortem; PMC, phase contrast morphometry; RT-PCR, reverse transcription–polymerase chain reaction; WB, Western blot. Excisional wounds refer to the removal of the skin surface, while incisional wounds represent a linear cut on a cell layer. ↑ indicates up regulation, ↓ indicates down regulation.
Table 1. This table resumes all the studies implied in our revision. RT-PCR, when reported, is quantitative. Statistical analysis was performed in all studies except in Viticchiè et al., 2012, Yang et al., 2011, and Yu et al., 2010. AM, antemortem; H, histology; HaCaT, human keratinocyte cell line; HCECs, corneal epithelial cells; HADSCs, human subcutaneous adipose tissue; HEKs, human keratinocytes; HSF, human skin fibroblasts; IHC, immunohistochemistry; IHF, immunohistofluorescence; ISH, hybridization in situ; MP, morphometric analysis; NP, not performed; NS, not specified; PM, post-mortem; PMC, phase contrast morphometry; RT-PCR, reverse transcription–polymerase chain reaction; WB, Western blot. Excisional wounds refer to the removal of the skin surface, while incisional wounds represent a linear cut on a cell layer. ↑ indicates up regulation, ↓ indicates down regulation.
ReferencesN. CasesModelCells and/or AnimalsKind of LesionsControlsmiRNAsmiRNAs
Expression
Performed Analysis RNA Isolation/cDNA Synthesis
Bertero et al., 2011 [21]6In vitro +
in vivo
NHKs, HaCat + MiceScratch wound + Excisional woundNormal skinmiR-483-3p
3d after injury infliction
Peak: 6–7 days
Normalization: at day 10
In epidermis (PM)
H + IHC + WB + RT-PCRTRIzol reagent/NP
Cao et al., 2020 [22]15In vitro +
in vivo
HaCaT, HSF + Mice H2O2-induced wound + Excisional woundNormal skinmiR-19b
In epidermis if H2O2
H + IHC + WB + RT-PCRTRIzol reagent/NP
Etich et al., 2017 [23]15In vivoMiceExcisional woundNormal skinmiR-204
from day 5 to day 10
H + RT-PCRmiRNeasy mini kit/miScript II RT kit
miR-205
from day 1 to day 7
miR-31
In epidermis from day 5 to 14
after injury infliction (PM samples)
He et al., 2020 [24]NSIn vitroHADSCs exosome, MALAT1 knockdown HaCaT and
HSF
H2O2 induced scratch woundNormal HaCaT and
HSF
miR-124
In epidermis if H2O2
WB + RT-PCRTRIzol reagent/cDNA Synthesis SuperMix
Ibrahim et al., 2019 [18]18In vivoRatsIncisional woundNormal skinmiR-205, miR-21No statistical significance
In epidermis
0, 24, and 48 h after death
H + RT-PCRmirVana PARIS kit/NP
Jiang et al., 2020 [25]3In vitroTGF-β1-treated
HaCaT
Scratch woundHaCaTmiR-26a
In keratinocytes
WB + RT-PCREZ-Magna RIP kit/miScript II RT kit
Jin et al., 2013 [26]5In vivo +
in vitro
HaCaT + MiceScratch wound + Excisional wound Normal skinmiR-152, miR-365,
miR-125a/b-5p, miR-181d,
miR-99, miR-100,
miR-30c

In keratinocytes
Peak: day 1
Normalization: day 5
WB + RT-PCRTRIzol reagent/CyQUANT assay
Lang et al., 2017 [27]9In vitro +
In vivo
HaCaT, HSF + RatsScratch wound + Excisional woundNo- TNFα-exposed cells miR-149
In epidermis
H + IHC + WBmirVana miRNA isolation kit/NP
Lin et al., 2013 [28]NSIn vitroHCECs and HEKsScratch woundNormal miR-16 levelsmiR-205
In corneal epithelial cells 24h after incision
WB + RT-PCRTRIzol reagent/cDNA Synthesis SuperMix
Li et al., 2015 [29]9In vivoHuman (surgical samples)Hypertrophic scarNormal skinmiR-149, miR-203a, miR-222, miR-122
In epidermis
ISHmiRcute RNA Isolation kit/NP
Liu et al., 2020 [30]24In vivoRats and DM ratsExcisional woundNo-diabetic ratsmir-203
Over the course of the first 4 days in non-diabetic rats, over the course of the first 6 days in DM rats, in epidermis, dermis, and subcutaneous fat
H + IHC+ IHF + WB + RT-PCRTRIzol rea-gent/RevertAid
H minus first-strand cDNA synthesis kit
Lyu et al., 2018 [17]22In vivoMiceBurned skinNormal skinmiR-135a-1-3p miR-183-3p miR-188-5p miR-3081-5p
miR-5103
miR-6378
miR-6385
miR-6391
miR-6769b-5p
miR-6969-5p
miR-7005-5p
miR-7036a-5p
miR-7044-5p
miR-710
miR-711
miR-7118-5p
miR-7668-3p
miR-8090
mmu-miR-874-3p
H + RT-PCRNS/PrimeScript RT reagent
miR-155-5p
miR-28a-3p
miR-467b-3p
miR-5132-5p
miR-6924-3p
Long et al., 2018 [31]NSIn vivomiR-21 knock-in miceExcisional woundNormal skinmiR-21H + MP + ISH + RT-PCRTaqMan
microR-NA assay kit/NP
Neri et al., 2019 [19]64Autopsy casuistryHumanLigature marksNormal skinmiR125a/b-5p miR130a-3p miR214-3p miR92a-3pH + IHC + RT-PCRmiScript miRNA PCR Array- Human Cell Differentiation/miScript II RT Kit
Pastar et al., 2012 [32]21In vivo +
in vitro
HEK, HSF + Rats (6) + Humans (15)Chronic ulcers + Excisional wound Normal skinmiR-16, miR-20a,
miR-21,
miR-106a, miR-130a, miR-203

In epidermis
(miR-21: in epidermis, dermis, and blood vessels)
H + ISH + WB + RT-PCRTaqMan
Mi-croRNA As-says/NP
Viticchiè et al., 2012 [33]NSIn vitro +
in vivo
HEK and MEK + MiceExcisional woundNormal skinmiR-203
In epidermis surrounding the wound and wound margin

In keratinocytes at the migratory front
(day 3 and 5 after wounding)
H + ISH + WB + RT-PCRmirVana miRNA isolation kit/NP
Wang et al., 2012 [34]NS In vivoMiceExcisional woundNormal skinmiR-31, miR-21, miR-203
In epithelial cells
H + IHC + ISH + RT-PCRTaqMan
Mi-croRNA As-says/NP
miR-249
(day 7 after wounding)
Yang et al., 2011 [35]9In vivoMiceExcisional woundmiR-21 knockdown micemiR-21
In epidermis
after 0, 1, 2 and 3 days
H + IHC + IHF + RT-PCRNS/NS
Yu et al., 2010 [36]NSIn vitroHCECs and HEKs (miR-205 and miR-184 downregulated)Scratch woundNormal cellsmiR-205, miR-184
In epithelial cells
IHC + PCM + WBQuik-Change Site-Directed Muta-genesis kit/NP
Zhang et al., 2020 [37]35In vivo + autopsy casesMice (9) + autoptic human samples (26)Excisional wound in
burned skin
Mice unburned skin + mice PM burned skinmiR-711, miR-183-3p
48 h after excision in human burned skin (PM samples),
120 h in mice burned skin (AM samples)
RT-PCRNS/PrimeScript RT reagent kit
Tot: 20 articlesTot: 255 (at least) Tot: 51 different miRNAs
Table 2. This table shows the main characteristics of the studies included in our review. Each count refers only to the use of single method. Excisional wounds include one case of excisional wound in burned skin and 1 case of comparative study between chronic ulcers and excision. * Type of lesion refers only to in vivo studies. HE, hematoxylin and eosin staining; IHC, immunohistochemistry; IHF, immunohistofluorescence; ISH, hybridization in situ; MP, morphometric analysis; PMC, phase contrast morphometry; RT-PCR, real-time polymerase chain reaction; WB, Western blot.
Table 2. This table shows the main characteristics of the studies included in our review. Each count refers only to the use of single method. Excisional wounds include one case of excisional wound in burned skin and 1 case of comparative study between chronic ulcers and excision. * Type of lesion refers only to in vivo studies. HE, hematoxylin and eosin staining; IHC, immunohistochemistry; IHF, immunohistofluorescence; ISH, hybridization in situ; MP, morphometric analysis; PMC, phase contrast morphometry; RT-PCR, real-time polymerase chain reaction; WB, Western blot.
Study’s CharacteristicsN. of Studies (Tot. 20)References
ModelIn Vivo8[17,18,23,29,30,31,34,35]
In Vitro4[24,25,28,36]
In Vivo + In Vitro6 [21,22,26,27,32,33]
Human samplesAutoptic 2[19,37]
Surgical 1[29]
Type of lesion *Excisional wound10[21,22,23,26,27,30,31,33,34,35]
Incisional wound1[18]
Burned skin1[37]
Ligature mark1[19]
Chronic ulcer1[32]
Hypertrophic scar1[29]
Performed analysisRT-PCR17[17,18,19,21,22,23,24,25,26,28,30,31,32,33,34,35,37]
Histology (HE)13[18,19,21,22,23,24,27,30,31,32,33,34,35]
WB11[21,22,24,25,26,27,28,30,32,33,36]
IHC8[19,21,22,27,30,34,35,36]
ISH5[29,31,32,33,34]
IHF2[30,35]
MP1[31]
PMC1[36]
Table 3. The list of miRNAs pinpointed by the included papers and the respective target genes, when available.
Table 3. The list of miRNAs pinpointed by the included papers and the respective target genes, when available.
miRNAsTarget Genes and/or ProteinsReferences
miR19-bCCL1, TGFβ[22]
miR-21Factor 3, vinculin, LepR, EGR3, Collagen, TGF-β, TIMP3, TIAM1, TP53[29,32,34,35]
miR-26aITGA5, PI3K/ AKT, SMAD1, GSK3β[25]
miR-30cPI3K/AKT, mTOR, IGF1R[26]
miR-31IL-1b, PTPRC/CD45, SHIP2, RNU6B, Col1a1[23]
miR-99PI3K/AKT, mTOR, IGF1R[26]
miR-100PI3K/AKT, mTOR, IGF1R[26]
miR-122MAPK, lysosome, insulin signaling pathway, focal adhesion[29]
miR-125a-5p,
miR-125b-5p
PI3K/AKT, mTOR, IGF1R[26]
miR-130aEGR3[32]
miR-149IL-1a, IL-1b, IL-6, TGF-β3, collagen III, Nf-Kb, RelB, Rel, MAPK[27,29]
miR-152PI3K/AKT, mTOR, IGF1R[26]
miR-181dPI3K/AKT, mTOR, IGF1R[26]
miR-184SHIP2, PI3K-Akt, actin filaments, p-cofilin (via Rho)[36]
miR-203MAPK, lysosome, insulin signaling pathway, focal adhesion, K15, P63, integrin-β1, TCF-4, ID-2, CD44, VEGFA, NRCAM, C-MET, Wnt, Notch, Factor 3, vinculin, LepR, EGR3, p63, LASP1, RAN, RAPH1[29,30,32,33]
miR-204IL-1b, PTPRC/CD45, SHIP2, RNU6B, Col1a1[23]
miR-205IL-1b, PTPRC/CD45, SHIP2, RNU6B, Col1a1, KCNJ10, SHIP2, PI3K-Akt, actin filaments, p-cofilin (via Rho), p-ERM[23,28,36]
miR-222DDK2, AXIN2, FRAT2, MAPK[29]
miR-365PI3K/AKT, mTOR, IGF1R[26]
miR-483-3pMK2, YAP1, ASH2, MKI67[21]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manetti, A.C.; Maiese, A.; Baronti, A.; Mezzetti, E.; Frati, P.; Fineschi, V.; Turillazzi, E. MiRNAs as New Tools in Lesion Vitality Evaluation: A Systematic Review and Their Forensic Applications. Biomedicines 2021, 9, 1731. https://doi.org/10.3390/biomedicines9111731

AMA Style

Manetti AC, Maiese A, Baronti A, Mezzetti E, Frati P, Fineschi V, Turillazzi E. MiRNAs as New Tools in Lesion Vitality Evaluation: A Systematic Review and Their Forensic Applications. Biomedicines. 2021; 9(11):1731. https://doi.org/10.3390/biomedicines9111731

Chicago/Turabian Style

Manetti, Alice Chiara, Aniello Maiese, Arianna Baronti, Eleonora Mezzetti, Paola Frati, Vittorio Fineschi, and Emanuela Turillazzi. 2021. "MiRNAs as New Tools in Lesion Vitality Evaluation: A Systematic Review and Their Forensic Applications" Biomedicines 9, no. 11: 1731. https://doi.org/10.3390/biomedicines9111731

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