Skip to main content

PERSPECTIVE article

Front. Med., 31 March 2022
Sec. Dermatology
Volume 9 - 2022 | https://doi.org/10.3389/fmed.2022.790207

Expert Recommendations on the Evaluation of Sunscreen Efficacy and the Beneficial Role of Non-filtering Ingredients

Salvador González1* José Aguilera2 Brian Berman3 Piergiacomo Calzavara-Pinton4 Yolanda Gilaberte5 Chee-Leok Goh6 Henry W. Lim7 Sergio Schalka8 Fernando Stengel9 Peter Wolf10 Flora Xiang11
  • 1Medicine and Medical Specialties Department, University of Alcalá de Henares, Madrid, Spain
  • 2Dermatological Photobiology Laboratory, Medical Research Center, School of Medicine, University of Málaga, Málaga, Spain
  • 3Department of Dermatology and Cutaneous Surgery, University of Miami-Florida, Miami, FL, United States
  • 4Department of Dermatology, University of Brescia, Brescia, Italy
  • 5Department of Dermatology, Hospital Universitario Miguel Servet, IIS Aragón, Zaragoza, Spain
  • 6National Skin Centre, Singapore, Singapore
  • 7Department of Dermatology, Henry Ford Health System, Detroit, MI, United States
  • 8Photoprotection Laboratory, Medicine Skin Research Center, São Paulo, Brazil
  • 9Buenos Aires Skin, Buenos Aires, Argentina
  • 10Department of Dermatology, Medical University of Graz, Graz, Austria
  • 11Department of Dermatology, Shanghai Medical College, Huashan Hospital, Fudan University, Shanghai, China

A variety of non-filtering agents have been introduced to enhance sunscreen photoprotection. Most of those agents have only weak erythema protective properties but may be valuable and beneficial in supporting protection against other effects of UV radiation, such as photoimmunosuppression, skin aging, and carcinogenesis, as well as photodermatoses. The question arises how to measure and evaluate this efficacy since standard SPF testing is not appropriate. In this perspective, we aim to provide a position statement regarding the actual value of SPF and UVA-PF to measure photoprotection. We argue whether new or additional parameters and scales can be used to better indicate the protection conferred by these products against the detrimental effects of natural/artificial, UV/visible light beyond sunburn, including DNA damage, photoimmunosuppression and pigmentation, and the potential benefits of the addition of other ingredients beyond traditional inorganic and organic filters to existing sunscreens. Also, we debate the overall usefulness of adding novel parameters that measure photoprotection to reach two tiers of users, that is, the general public and the medical community; and how this can be communicated to convey the presence of additional beneficial effects deriving from non-filtering agents, e.g., biological extracts. Finally, we provide a perspective on new challenges stemming from environmental factors, focusing on the role of the skin microbiome and the role of air pollutants and resulting needs for photoprotection.

Sun Protection Factor and UVA Protection Factor: Are They Still the Benchmark for Photoprotection?

Population growth, increasing awareness of the consequences of sun exposure and increased lifespan, together with the increased incidence of sun-related skin tumors have resulted in photoprotection becoming indispensable (1). The choice of sunscreens and their proper use have become extremely important, forming the basis of a corner of the pharmaceutical market with enormous economic impact. However, this investment and expense does not correlate with a decrease in the incidence of skin cancer. This means that there are fundamental gaps in the manner photoprotection is measured, communicated and applied by scientists and companies to guide sunscreen usage by the final users, i.e., the general public.

The current standards are the SPF and UVA-PF parameters. Other normatives cover related issues, e.g., sunscreen water resistance (ISO 16217:2020: Cosmetics—Sun protection test methods—Water immersion procedure for determining water resistance).

The FDA states: “SPF is a measure of how much solar energy (UV radiation) is required to produce sunburn on protected skin (i.e., in the presence of sunscreen) relative to the amount of solar energy required to produce sunburn on unprotected skin. As the SPF value increases, sunburn protection increases.” The International Organization for Standardization (ISO) dictated a regulatory norm (ISO24444:2019), “Cosmetics—Sun protection test methods—in vivo determination of the sun protection factor (SPF)” that covers the testing methods, reference values and every parameter related to the ability of a given substance to act as a photoprotector. In practice, the effectiveness of photoprotection achieved by topical formulations is much more difficult to evaluate than expected. Several parameters need to be taken into account, including seasonal, meteorological and geographical considerations; skin phototype; amount of sunscreen applied; and frequency of re-application. None of these parameters is absolutely precise in every condition, and for every individual. As a result of this variability, many end users are much less protected than they believe.

A key value for in vivo determinations is the Minimal Erythema Dose (MED), which is defined as the threshold dose of solar radiation that produces sunburn. In fact, the SPF (Sun Protection Factor) that appears as a number (usually between 3 and 50+, or in some markets even more) on every sunscreen container is a ratio between the MED of skin treated with sunscreen divided by the MED of untreated skin. For example, if a sunscreen has a SPF value of 10, it means that it takes 10 times longer to induce erythema in treated skin compared to unprotected skin. It is thus obvious that the dose is a crucial factor to define the given MED of a photoprotector. The required dose of sunscreen applied for testing procedures is 2 mg/cm2.

Another important parameter is the UVA protection Factor (UVA-PF), which is obtained from in vitro measurements. Based on the determination of UVA (320–400 nm) transmittance in methacrylate or PMMA plates coated with 1.3 mg/cm2 sunscreen (ISO24443:2012, updated recently to ISO24443:2020), UVA-PF reflects in vivo protection conferred by the sunscreen against UVA-induced persistent pigment darkening (at 2–4 h after exposure). A sunscreen with an Ultraviolet A (UVA)-PF of 10 indicates that, e.g., in a subject with skin phototype IV it takes 10 times longer to develop PPD in sunscreen-protected skin compared to unprotected skin.

However, the manner in which some ISO-standardized measurements are made allows predicting the SPF and UVA-PF. The procedure consists of applying the sunscreen on artificial skin models and measuring the spectral transmittance at different wavelengths of the Ultraviolet (UV) spectrum. Although these systems are not included in ISO24444:2019, they provide measurements sufficiently comparable to those obtained from human volunteers, with some caveats described below. There is normalized precedent for this, as ISO 24443:2012 describes the in vitro determination of photoprotection against UVA using a hybrid method that relies on in vitro measurements but requires the SPF in vivo data for its determination. Although the FDA accepts the critical wavelength method, also in vitro and with an execution method similar to ISO 24443, the method has not yet been shown to be reproducible and interchangeable with SPF in vivo, especially in high SPF formulations.

Other alternative examples include the COLIPA method (Guidance drafted by The European Cosmetics Association). This method assesses UV transmittance of a thin layer of sunscreen on a PMMA roughened substrate after exposure to a controlled dose of UV radiation from a defined UV source (2). However, it has been shown that SPF fluctuates in a roughness-dependent manner (3). A similar method developed by the National Institute of Public Health (NIPH) measures attenuation of UVB intensity on a defined layer of a sunscreen product irradiated with an UVA/UVB source, a sheet of Mikelanta covering paper with 2 mg/cm2 of product and assessed by a radiometer (4). The VUOS method employs surgical tape affixed to a quartz layer with 1.2 mg/cm2 of product applied, and SPF calculation from transmittance measurements (5). Another method uses diffusing plates made of quartz fixed with surgical adhesive tape on human skin biopsies (6). This enables obtaining very precise spectra that define the photoprotective efficiency of a sunscreen at different UV wavelengths. The FDA (United States) uses the critical wavelength as a means to assess broad spectrum protection. Critical wavelength is the wavelength below which 90% of the area under the absorption curve resides. For products to be eligible for “broad spectrum” label, the critical wavelength must be ≥370 nm. Finally, other methods could be incorporated into future ISO covering the determination of photoprotective properties, for example, Hybrid Diffuse Reflectance Spectroscopy (HDRS) is a promising in vivo alternative due to the fact that it does not require erythema induction, measuring instead skin reflectance. HDRS displays excellent correlation with ISO24444:2019 (7).

The FDA recently proposed that UV filters approved in the US to be categorized into three categories, based on determination of GRASE (Generally Recognized as Safe and Effective). The list of FDA-approved GRASE products is very short, including only TiO2 and ZnO (8). Two filters are categorized as not GRASE, while additional safety data have been requested for the remaining filters (9) in order to be considered GRASE. Conversely, the EU considers sunscreens as cosmetics, thus additional substances are allowed.

However, at the consumer level these considerations are much less relevant. The most recognized parameter by the general public is the SPF. However, higher SPF leads to a false sense of protection because quantities applied are usually much lower than needed and therefore the “true” SPF is much lower. Considering that a normal-sized adult skin measures 2 m2, an adult whose height is 1.73 m would require 35 g of sunscreen per application (10), which a very small percentage of the population actually employs. Furthermore, the actual amount of effective sunscreen on the skin is reduced due to the routine habits that accompany the usual need for photoprotection, e.g., friction with clothing, sweat, water immersion, etc., which decreases the de facto amount of sunscreen that remains on the skin post-application. Therefore, the consumer acts according to this number, generally thinking that if the number is high enough, they can disregard additional safety issues, including avoiding midday exposure, seeking shade when outdoors, limiting exposure time, wearing additional protective measures (clothes, hats, etc.), and the need for frequent reapplication of the product.

In addition to this crucial issue, SPF does not account for additional damage caused to the skin beyond MED-related measurements. The most crucial issues are DNA damage, photoaging and immunosuppression. There is a large (and growing) body of evidence indicating that sub-MED doses of UV radiation, particularly by longer exposure to low energy (UVA) photons, accelerate photoaging and mediate immunosuppression and DNA damage. Some of these effects are related to DNA-induced damage, and some are due to deletion of specific cell subpopulations. Although parameters to measure these effects do exist (see below), they are not incorporated into the ISO standard SPF or UVA-PF parameters. Finally, neither the ISO rules nor the FDA monograph on the topic take into consideration the effects of visible (VL) and infra-red (IR) light, and a growing body of evidence indicates that these wavelengths also produce biological effects on different cellular skin populations (1113). In this article, we posit that there is an emergent need for complementary methods that address photoprotective needs and measures beyond erythema, which is the focus of SPF measurements, particularly given the steady increase in diagnosed skin cancer cases over the past 10 years, which are not necessarily related to erythematous reactions. In addition, future determinations of photoprotective ability would ideally be performed in vitro, which would reduce variability and prevent human subjects from receiving high doses of light in the UV and other ranges. Other wavelengths are important given the increasing cases of photodermatosis and photosensibilization, which are important at visible and IR ranges. The major challenge will be to integrate the information conveyed by SPF measurements with these new methods. Ideally, the research community should strive to provide a comprehensive final index that is easily understandable by end-users. Such index would include ISO-compliant SPF measurements as well as additional information regarding the positive biological effects of other components of the sunscreen formulation.

Sunscreens may also contribute to the process of photo-adaptation, which consists to decreased erythema and inflammation in response to acclimation during repeated exposure (14). This is a common phenomenon that may reduce the danger of developing cancer, but may also affect specific responses to therapy (15, 16).

UV-Induced DNA Damage

UV radiation causes DNA damage (17). The best characterized effects of UV photons on DNA include the formation of thymine and pyrimidine-pyrimidone dimers (18, 19) and 8-hydroxy-2′-deoxyguanosine (8-oxo-dG), which is an oxidized derivative of deoxyguanosine also involved in carcinogenesis (20). DNA damage is often measured as the appearance of the H2A histone family member X (H2AX) (21). Normally, the onset of these types of DNA damage would cause apoptosis, but the targeting of specific tumor suppressor such as p53 enables the survival of these cells (22, 23), which become tumor seeds. UV radiation also induces telomere shortening and degradation (24). These effects may cause transforming DNA mutations (25), leading to the emergence of photo-induced carcinogenesis.

Although high energy UV photons are the most important triggers of these alterations (26), lower energy UVA photons, or even visible light photons, may also induce these effects (27, 28). This is due, at least in part, to the oxidative stress (generation of ROS) induced by UV and visible photons (27, 29). In addition to damaging DNA, ROS also trigger diverse signaling pathways involved in cell proliferation, e.g., the MAPK pathway, JNK/p38, expression of AP1 and COX2 and activation of the NF-kB pathway (3032). In addition, UV radiation affects the function of Nrf2, which is a controlling hub for antioxidant response by determining the expression of natural antioxidant enzymes, including Glu-6-phosphate dehydrogenase, thioredoxin reductase, glutathione S-transferase, and peroxidase (33). Finally, mitochondria play active and passive roles in ROS-mediated damage. ROS decrease mitochondrial functionality, decreasing O2 usage and ATP generation; and mitochondrial DNA acts as an “in vivo dosimeter,” measuring the exposure of a given cell to oxidative damage. Furthermore, mitochondria may also actively produce ROS (34).

Due to the central role of oxidation in the processes mentioned above, anti-oxidant and other absorption mechanisms protect against photoaging and photocarcinogenesis. One such absorptive mechanism is the isomerization of trans-urocanic acid into the cis- form, which has immunosuppressive properties (35). As stated above, different enzymatic systems (SOD, catalases, peroxidases, GSH and GST) also quench free radicals in different states, reducing UV-mediated oxidation and decreasing their impact on the cell’s DNA (34, 36, 37). The current measurements regulated by the ISO normative to determine the SPF of a given sunscreen do not contemplate the accumulative effect of oxidative damage, particularly in light of the fact that SPF mainly considers the effect of UVB photons, whereas oxidative damage from UVB is much lower than that from UVA radiation.

UV-Induced Skin Photoaging

The relationship between UV-induced skin erythema and photoaging is well established, and previous reviews in the field cover this aspect in minute detail (38, 39). Repeated erythema causes a wound healing-like behavior, including the onset of scarring-like events that promote ECM remodeling. Some of these events are MMP (matrix metalloproteases) secretion, collagen cross-linking and elastin degradation (17, 40, 41). All these events promote massive tissue remodeling, the onset of wrinkling and the induction of “visible aging.” Photo aging scoring is a key clinical aspect of this process that has been reviewed elsewhere (4244). The most frequent and best characterized mutation in mitochondrial DNA (mtDNA) and marker of UVA light induced photoaging is a deletion of 4,977 base pairs, called the “common deletion.” UVA radiation generates the common deletion in human fibroblasts through an oxidative mechanism, which depends on the generation of singlet oxygen, other ROS and RNS.

UV-Induced Immunosuppression

UV-induced immunosuppression is a multi-pronged mechanism that affects different cell types, especially myeloid subtypes, e.g., Langerhans cells (45, 46) and langerin + dendritic cells (9). It also induces the production of immunosuppressive cytokines by keratinocytes (47) and skin macrophages (48). Irradiated myeloid cells undergo abnormal maturation and cannot migrate normally to the lymph nodes, decreasing the skin’s defenses against pathogens (49) which may also lead to imbalanced homeostasis of the skin with resident commensal bacteria (50). Despite the immediate anti-bacterial and anti-viral effect of UV, long-term exposure causes photoimmunosuppression, decreasing the immune system’s ability to promote pathogen clearance, including fungal pathogens, virus and bacteria. There is evidence that UV-mediated LC depletion promotes the recruitment of monocytes and immature dendritic cells that try to compensate the function of LC in the skin (51).

Immunosuppression is a crucial hallmark of cancer (52). Hence, it is possible that a sunscreen with high SPF may not completely prevent photocarcinogenesis. This would be due to the combination of skin damage occurring below MED and immunosuppression, particularly in individuals with genetic susceptibility. It has been well-demonstrated that the ability of sunscreens to prevent immunosuppression is not related to the MED, which is more influenced by UVB than by UVA, whereas UVA is at least as potent as an immunosuppressive agent as UVB [see below and Ref. (53)]; this indicates that MED measurements, which are the basis of SPF determination as per ISO24444:2019, do not correlate with the ability of a sunscreen to prevent immunosuppression.

It is worth noting that approaches that combine light measurements and immunosuppression have been carried out. For example, contact hypersensitivity assays using 2-chloro-l,3,5-trinitrobenzene (TNCB) or 1-fluoro-2,4-dinitrobenzene (DNFB) as irritants were carried out in skin of mice pre-irradiated with different UV wavelengths (54, 55). The efficacy of sunscreens containing UV filters to protect against suppression of CHS induction has been investigated in several studies and revealed that photo-immunoprotection correlated with UVA protection (56, 57). Similar measurements have been performed in humans using nickel as contact allergen on the efferent immune response, which constitute the basis of the Human Immunoprotection Factor, or HIF (58). In humans, two bands that caused immunosuppression were identified, one between 290 and 310 nm (UVB) and one that peaked around 370 nm (UVA). The 370 nm peak, which some authors extend to 380 nm (59), is particularly interesting when discussing oxidative damage.

Hyperpigmentation and Other Skin Alterations Induced by Visible Light

Pigmentation is induced by UV radiation, but also by visible light (60). In higher skin phototypes, visible light induces more durable photo-pigmentation than UVA irradiation (12). At a mechanistic level, blue light induces melanin production by activating the photoreceptor opsin-3, which acts on the transcription factor Mitf (61), controlling tyrosinase expression and thus melanin production (62). In addition, blue light also generates ROS, which causes photo-oxidative damage to DNA and cellular structures, for example, inducing MMP secretion (27). Photoprotection against visible light cannot be conjoined with UV protection since canonical, FDA-approved UV blockers, e.g., TiO2 or ZnO become whitish upon irradiation with visible photons (60), thus becoming unappealing from a cosmetic standpoint. This means that different filters are needed, or the cosmetic formulations of TiO2/ZnO need additional inactive ingredients that disguise the whitening of these oxides. Different approaches could be used, from iron oxide (FeO) to natural extracts (see below). Furthermore, sunscreens containing higher concentrations of UV blockers increase the whitish appearance of the skin under visible light, which has unpleasant effects on the customer and a tendency to “sacrifice” better protection for aesthetic reasons.

Beyond Sun Protection Factor: Biological Actuators vs. Photon Blockers

Accepting the limitations of SPF is a risky proposition due to the exclusion of DNA damage, photoaging and immunosuppressive effects of UV light, as detailed above. However, the current ISO has no wiggle room to incorporate additional parameters. Current organic and inorganic filters do provide considerable protection against such damage, but recent studies have shown that significant additional protection can be achieved by adding other ingredients to sunscreen formulations. These biological filters and/or non-filtering modulating biological molecules can induce additional biological effects, e.g., protection against DNA damage, photoaging or immunosuppression. The major limitation of this approach is that the FDA describes sunscreens as “drugs”,1 hence the threshold for the demonstration of safety and efficacy is very high (see discussion of GRASE list in section “Sun Protection Factor and UVA Protection Factor: Are They Still the Benchmark for Photoprotection”). The FDA even acknowledges the limitation of SPF by forcing producing companies to include the following: “Skin Cancer/Skin Aging Alert: Spending time in the sun increases your risk of skin cancer and early skin aging. This product has been shown only to help prevent sunburn, not skin cancer or early skin aging.” In agreement with the previous statement, United States labeling distinguishes clearly between active ingredients (normally ZnO and/or TiO2) and inactive ingredients, which usually include oxide solvents, moisturizers, aromatic substances and vitamins.

Conversely, non-US regulations are more flexible because sunscreens are considered cosmetics.2 This results in a somewhat wider list of accepted ingredients. However, claims of efficacy need to be backed up by scientific evidence, particularly when it comes to the prevention of immunosuppression.

The figure of “biological filter” emerges from these regulatory differences. A biological filter could be defined as a biological molecule or mixture endowed with or without direct sunscreen ability and able to provide additional beneficial effects. Classical examples include botanical extracts containing anti-oxidant moieties. However, the efficacy of these types of ingredients in preventing photoaging and photoimmunosuppression is usually poorly documented. Topical DNA repair enzymes such as photolyase, endonuclease and glycosylase have also been demonstrated to provide some additional protection, particularly with respect to photoimmunosuppression (63, 64), carcinogenesis (65) as well as polymorphic light eruption (66), though those enzymes provide almost no protection against sunburn.

Over the past 25 years, relevant scientific evidence has grown regarding the efficacy of diverse families of natural compounds, e.g., botanical and non-botanical extracts (Table 1). Non-botanical extracts include fatty acid preparations and probiotics (67). Botanical extracts, e.g., red fruit juices, green tea, coffee, and cocoa preparations, fern leaves extracts, etc., contain vitamin derivatives and large amounts of antioxidant moieties, which reduce the impact of oxidation and inflammation, reducing the onset of photoaging and cancer. Potential active principles include carotenes and lycopenes, xantophylls, vitamins (C, D, and E) and various types of polyphenols. Green tea polyphenols –GTPP– include several species, mainly epigallocatechin –EGC–, epigallocatechin-3-gallate –EGCG–, epicatechin –EC–, and epicatechin-3-gallate –ECG–. They all scavenge reactive oxygen species and boost immunity. ROS scavenging and immune enhancement are related, since oxidation suppresses T cell proliferation and DC function in several contexts (49, 68, 69); hence a decrease in the oxidative threshold enhances immune responses, thereby increasing immune surveillance and reinforcing local anti-tumor responses. In addition, GTPP also decrease oxidation-enhanced protein expression, e.g., metalloproteases, which contribute to skin aging and damage (70). In general, most botanical components endowed with antioxidant capabilities promote skin health (Table 1). Very-well characterized examples include diverse hydrophilic extract of the Mesoamerican fern Polypodium leucotomos (PL). PL extracts are effective and safe both topically and orally (71). Assays in mice and humans have demonstrated that one of such extracts (Fernblock®) increases MED in spite of having a modest filtering ability (72). The extract also counters the biological effects underlying UV-induced photoaging when administered orally. In mice, it decreases erythema and prevents inflammation, increasing the levels of systemic antioxidant systems, e.g., GSH and GSSG (73). Furthermore, it also decreases inflammation in harsher experimental conditions, e.g., human patients irradiated with UVB light (74). In this regard, oral treatment with PL extracts decreased the deleterious effects of UVB irradiation in human volunteers, including appearance of sunburn cells, DNA damage and inflammatory markers (75). Such treatment also decreased inflammation in human patients undergoing psoralens-UVA therapy (69), a form of treatment for psoriasis, dermatitis, vitiligo, polymorphic light eruption and cutaneous T-cell lymphoma (76, 77). In this context, oral treatment with PL extracts also decreased Langerhans cell depletion (74). Treatment also inhibited the appearance of “common deletion” (CD), which is a mitochondrial DNA deletion that denotes UVA-mediated damage (78). More recent evidence has demonstrated that this type of treatment also protects against the effects of visible (blue) light, notably against pigmentation and skin darkening (79). Treatment also regulates opsin-3 and prevents melanin-dependent photo-oxidation induced by digital screen-generated blue light (80) and prevents visible and infrared-induced skin damage (81).

TABLE 1
www.frontiersin.org

Table 1. Photoprotective effects of various natural extracts.

Evidence of the efficacy of fern extracts has been obtained at a molecular level, as their antioxidant moiety also inhibited trans-urocanic isomerization (82). At a cellular level, treatment with PL extracts blocked the activation of oxidative pathways, blocking DNA damage and promoting its repair (83), and impaired MMPs secretion and matrix remodeling (84). Finally, it efficiently prevented Langerhans cells depletion in humans and rodents (69, 73, 85), leading to decreased inflammation when administered orally (reviewed in Ref. 86).

Recent evidence indicates that the addition of PL extracts to topical formulations with traditional inorganic/organic filters can increase the formulations’ MED, DNA protection, immunoprotection (LC and HIF), and ability to reduce photoaging and pigmentation response (72, 87).

It is important to note that most of these compounds have limited effect on the SPF per se; hence they would be considered “inactive ingredients” from a FDA perspective. However, the increasing amount of scientific evidence supporting the efficacy of this type of approach suggests a potential benefit from their inclusion in sunscreens. The peer-reviewed studies regarding the efficacy of natural compounds as photoprotectors have reached a scientific standard that would support the planning and execution of another meeting, in which a large community of experts in the field, including molecular biologists, immunologists, and clinical practitioners (mainly dermatologists and oncologists) review the possible advantages of including natural extracts in topical sunscreen formulations. A new consensus parameter for evaluating photoprotection efficacy could emerge from such a meeting to complement SPF, providing additional information in an easy to understand manner that would allow consumers to make more informed decisions regarding photoprotection. This is particularly important if we consider that photoprotective measures have two tiers of potential users: the mainly healthy individual who purchases sunscreen for outdoor activities; and the patients under specialized care due to pre-existing skin conditions, e.g., vitiligo, hyperpigmentation, photodermatoses, psoriasis, actinic keratoses and skin cancer, rosacea, atopic dermatitis, advanced skin aging, etc. Sunscreens are necessary for every individual, but formulations with effective activity against a wide range of damage are particularly important for patients suffering from chronic skin diseases and sun-related conditions, which comprise a significant percentage of society. Some examples of prevalence are 0.1–2% for vitiligo (88); 0.2–5% for psoriasis (89); and 1–2% for non-melanoma skin cancer.3 Most of these patients are under dermatologist care. Thus, the decision to recommend one sunscreen or another can be guided by the specialist, who has a better understanding of the different factors and/or beneficial effects of inactive ingredients. In order to achieve this, it is essential that care providers have specific training (“photo-education”) that enables them to interpret the pertinent information regarding the additional beneficial effects of the composition of sunscreens in order to make informed recommendations. In this context, providing additional information regarding the biological effects of inactive ingredients can ultimately contribute to the wellbeing of patients.

Future Challenges

In addition to the evident need to generate a photoprotection index that is more inclusive than SPF, the field faces additional challenges. One is the growing evidence that UV- and visible light-mediated skin oxidation alters the balance of skin homeostasis with symbiotic microbiome. UV and ROS generation induce alterations in the composition of the skin microbiome, destabilizing skin immune homeostasis and increasing the prevalence of other skin infections, e.g., herpesvirus (90). Such an imbalance may also accelerate photocarcinogenesis (91). For example, Cutibacterium acnes produce less porphyrins when irradiated with UVB photons (92), which may lead to decreased inflammatory responses (93). Other symbiotic yeast and bacteria are also affected by UV radiation, unbalancing the skin microbiome and favoring the onset of skin cancer, particularly those related to virus, e.g., HPV and Merkel cell polyomavirus (94). Intriguingly, recent evidence indicates that a commensal strain of Staphylococcus epidermidis protects against skin carcinogenesis in a mouse model (95) and a normal skin microbiota reduces photoimmunosuppression (50).

Another important issue is that demographic studies indicate that more and more population is concentrating in large cities, which increases skin exposure contaminants, e.g., combustion derivatives. The WHO has estimated that over 4 million people die yearly as a result of air contamination featuring diverse types of pollutants, e.g., particulate matter, nitrogen dioxide, sulfur dioxide, black carbon, carbon monoxide, and ground-level ozone (96). These pollutants act at three levels: they generate free radicals and oxidative damage; they cause, or amplify, inflammation; and they compromise the integrity of the skin barrier function. Other epithelia are also compromised, e.g., the lungs. Particulate matter is particularly harmful in this respect, including ultrafine particles smaller than 100 nm. These can penetrate skin barriers and have systemic effects, including oxidation and inflammation. Likewise, polycyclic aromatic hydrocarbons, dioxins and ozone can produce oxidative damage and inflammation, amplifying the effects of skin unprotected against to UV radiation. Exposure to these pollutants can have additive or exponential effects that would increase UV-induced photo-damage, either during exposure (UV may ionize some of these compounds, increasing their toxicity) or post-exposure (pollutants on UV-induced damaged skin may contribute to infection, slow healing or local immunosuppression). Again, SPF does not account for these amplifying/accumulative effects, which nonetheless will inevitably affect sunscreen end users. These novel effects will also need to be taken into account when developing the next generation of photoprotective sunscreens and the measurements that will define their activity.

Additional challenges remain, including patient education and the incorporation of nanotechnologies to skin care, which constitute exciting new frontiers in the field.

Conclusion

Despite increased awareness and use of photoprotection sunscreens, cases of skin cancer continue to grow worldwide. Photoprotection efficacy of sunscreens is measured and communicated primarily by SPF, and to a lesser extent UVA-PF and critical wavelength. These measurements do not, however, take into consideration damage caused to the skin beyond MED-related measurements such as DNA damage, photoaging and immunosuppression. In addition, there is a large and growing body of evidence indicating that sub-MED doses of UV radiation can indeed accelerate photoaging and mediate immunosuppression and DNA damage. Current SPF and UVA-PF ratings do not contemplate damage from other wavelengths such as visible and IR ranges.

Given the limitations of current measurements, there is an emergent need for complementary methods to assist specialists in determining the most suitable sunscreen for their patients, especially those suffering from chronic skin diseases and sun related conditions. In addition a new comprehensive final index including ISO-compliant SPF measurements but also determination of other levels of protection against solar radiation could facilitate better overall understanding of sunscreen efficacy amongst the general public.

Current organic and inorganic filters provide considerable protection against sun damage, but recent studies have shown that significant additional protection can be achieved by adding other ingredients to sunscreen formulations. These additional components do not need to be physical filters, but they must induce additional biological effects, e.g., protection against DNA damage, photoaging or immunosuppression. Additional components take the form of botanical or non-botanical extracts endowed with antioxidant, DNA protective and immunostimulating properties. A plethora of scientific evidence supports the potential benefit of inclusion of these types of extracts in commercial sunscreens and oral supplements with no additional risk to the overall health of the end users.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author.

Author Contributions

SG conceptualized and prepared the agenda that was discussed during the meeting. All rest of the authors contributed equally to this article and their names are listed in alphabetical order.

Conflict of Interest

SG is a consultant for Cantabrialabs, which produces Fernblock®. C-LG and FX are members of Asian Medical Advisory Board for Cantabrialabs. BB is a consultant for Ferndale. HL has served as an investigator for Incyte, La Roche-Posay, Pfizer, PCORI, as a consultant for Pierre Fabre, ISDIN, Ferndale, La Roche-Posay, Beiersdorf, and as a non-product specific speaker for La Roche-Posay and Cantabria Labs. FX is a consultant for Skinceuticals, La Roche-Posay, Avene, Bioderma, CeraVe, Cantabria Labs, and Winona.

The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

We thank Azahara Rodriguez-Luna (Cantabrialabs) for contributing materials used to prepare the discussion contained herein, and Miguel Vicente-Manzanares for editorial preparation of the manuscript.

Footnotes

  1. ^ https://www.fda.gov/regulatory-information/search-fda-guidance-documents/labeling-and-effectiveness-testing-sunscreen-drug-products-over-counter-human-use-small-entity#_Toc281970892
  2. ^ https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX:02009R1223-20190813
  3. ^ https://www.cancer.net/cancer-types/skin-cancer-non-melanoma/statistics

References

1. Laughter MR, Maymone MBC, Karimkhani C, Rundle C, Hu S, Wolfe S, et al. The burden of skin and subcutaneous diseases in the United States from 1990 to 2017. JAMA Dermatol. (2020) 156:874–81. doi: 10.1001/jamadermatol.2020.1573

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Matts PJ, Alard V, Brown MW, Ferrero L, Gers-Barlag H, Issachar N, et al. The COLIPA in vitro UVA method: a standard and reproducible measure of sunscreen UVA protection. Int J Cosmet Sci. (2010) 32:35–46. doi: 10.1111/j.1468-2494.2009.00542.x

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Gers-Barlag H, Klette E, Bimczok R, Springob C, Finkel P, Rudolph T, et al. In vitro testing to assess the UVA protection performance of sun care products. Int J Cosmet Sci. (2001) 23:3–14. doi: 10.1046/j.1467-2494.2001.00048.x

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Bendova H, Akrman J, Krejci A, Kubac L, Jirova D, Kejlova K, et al. In vitro approaches to evaluation of sun protection factor. Toxicol In Vitro. (2007) 21:1268–75. doi: 10.1016/j.tiv.2007.08.022

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Akrman J, Kubac L, Bendova H, Jirova D, Kejlova K. Quartz plates for determining sun protection in vitro and testing photostability of commercial sunscreens. Int J Cosmet Sci. (2009) 31:119–29. doi: 10.1111/j.1468-2494.2008.00482.x

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Bleasel MD, Aldous S. In vitro evaluation of sun protection factors of sunscreen agents using a novel UV spectrophotometric technique. Int J Cosmet Sci. (2008) 30:259–70. doi: 10.1111/j.1468-2494.2008.00453.x

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Rohr M, Ernst N, Schrader A. Hybrid diffuse reflectance spectroscopy: non-erythemal in vivo testing of sun protection factor. Skin Pharmacol Physiol. (2018) 31:220–8. doi: 10.1159/000488249

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Lyons AB, Trullas C, Kohli I, Hamzavi IH, Lim HW. Photoprotection beyond ultraviolet radiation: a review of tinted sunscreens. J Am Acad Dermatol. (2021) 84:1393–7. doi: 10.1016/j.jaad.2020.04.079

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Bursch LS, Wang L, Igyarto B, Kissenpfennig A, Malissen B, Kaplan DH, et al. Identification of a novel population of Langerin+ dendritic cells. J Exp Med. (2007) 204:3147–56. doi: 10.1084/jem.20071966

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Isedeh P, Osterwalder U, Lim HW. Teaspoon rule revisited: proper amount of sunscreen application. Photodermatol Photoimmunol Photomed. (2013) 29:55–6. doi: 10.1111/phpp.12017

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Mahmoud BH, Hexsel CL, Hamzavi IH, Lim HW. Effects of visible light on the skin. Photochem Photobiol. (2008) 84:450–62.

Google Scholar

12. Mahmoud BH, Ruvolo E, Hexsel CL, Liu Y, Owen MR, Kollias N, et al. Impact of long-wavelength UVA and visible light on melanocompetent skin. J Invest Dermatol. (2010) 130:2092–7. doi: 10.1038/jid.2010.95

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Sklar LR, Almutawa F, Lim HW, Hamzavi I. Effects of ultraviolet radiation, visible light, and infrared radiation on erythema and pigmentation: a review. Photochem Photobiol Sci. (2013) 12:54–64. doi: 10.1039/c2pp25152c

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Garmyn M, Young AR, Miller SA. Mechanisms of and variables affecting UVR photoadaptation in human skin. Photochem Photobiol Sci. (2018) 17:1932–40. doi: 10.1039/c7pp00430c

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Damiani G, Pacifico A, Chu S, Chi CC. Frequency of phototherapy for treating psoriasis: a systematic review. Ital J Dermatol Venereol. (2021) [Epub ahead of print]. doi: 10.23736/S2784-8671.21.06975-3

PubMed Abstract | CrossRef Full Text | Google Scholar

16. Pacifico A, Damiani G, Iacovelli P, Conic RRZ. Photoadaptation to ultraviolet B TL01 in psoriatic patients. J Eur Acad Dermatol Venereol. (2020) 34:1750–4. doi: 10.1111/jdv.16209

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Yeager DG, Lim HW. What’s new in photoprotection: a review of new concepts and controversies. Dermatol Clin. (2019) 37:149–57. doi: 10.1016/j.det.2018.11.003

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Mao P, Wyrick JJ, Roberts SA, Smerdon MJ. UV-Induced DNA damage and mutagenesis in chromatin. Photochem Photobiol. (2017) 93:216–28. doi: 10.1111/php.12646

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Moriwaki S, Takahashi Y. Photoaging and DNA repair. J Dermatol Sci. (2008) 50:169–76. doi: 10.1016/j.jdermsci.2007.08.011

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Floyd RA. The role of 8-hydroxyguanine in carcinogenesis. Carcinogenesis. (1990) 11:1447–50. doi: 10.1093/carcin/11.9.1447

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Bonner WM, Redon CE, Dickey JS, Nakamura AJ, Sedelnikova OA, Solier S, et al. GammaH2AX and cancer. Nat Rev Cancer. (2008) 8:957–67.

Google Scholar

22. Borras C, Gomez-Cabrera MC, Vina J. The dual role of p53: DNA protection and antioxidant. Free Radic Res. (2011) 45:643–52. doi: 10.3109/10715762.2011.571685

PubMed Abstract | CrossRef Full Text | Google Scholar

23. You YH, Szabo PE, Pfeifer GP. Cyclobutane pyrimidine dimers form preferentially at the major p53 mutational hotspot in UVB-induced mouse skin tumors. Carcinogenesis. (2000) 21:2113–7. doi: 10.1093/carcin/21.11.2113

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Rochette PJ, Brash DE. Human telomeres are hypersensitive to UV-induced DNA damage and refractory to repair. PLoS Genet. (2010) 6:e1000926. doi: 10.1371/journal.pgen.1000926

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Basu AK. DNA damage, mutagenesis and cancer. Int J Mol Sci. (2018) 19:970. doi: 10.3390/ijms19040970

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Pfeifer GP. Mechanisms of UV-induced mutations and skin cancer. Genome Instab Dis. (2020) 1:99–113. doi: 10.1007/s42764-020-00009-8

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Liebel F, Kaur S, Ruvolo E, Kollias N, Southall MD. Irradiation of skin with visible light induces reactive oxygen species and matrix-degrading enzymes. J Invest Dermatol. (2012) 132:1901–7. doi: 10.1038/jid.2011.476

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Strickland PT. Photocarcinogenesis by near-ultraviolet (UVA) radiation in Sencar mice. J Invest Dermatol. (1986) 87:272–5. doi: 10.1111/1523-1747.ep12696669

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Fang X, Ide N, Higashi S, Kamei Y, Toyooka T, Ibuki Y, et al. Somatic cell mutations caused by 365 nm LED-UVA due to DNA double-strand breaks through oxidative damage. Photochem Photobiol Sci. (2014) 13:1338–46. doi: 10.1039/c4pp00148f

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Ding Y, Jiratchayamaethasakul C, Lee SH. Protocatechuic aldehyde attenuates UVA-induced photoaging in human dermal fibroblast cells by suppressing MAPKs/AP-1 and NF-κB signaling pathways. Int J Mol Sci. (2020) 21:4619. doi: 10.3390/ijms21134619

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Mahns A, Wolber R, Stab F, Klotz LO, Sies H. Contribution of UVB and UVA to UV-dependent stimulation of cyclooxygenase-2 expression in artificial epidermis. Photochem Photobiol Sci. (2004) 3:257–62. doi: 10.1039/b309067a

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Silvers AL, Bachelor MA, Bowden TG. The role of JNK and p38 MAPK activities in UVA-induced signaling pathways leading to AP-1 activation and c-Fos expression. Neoplasia. (2003) 5:319–29. doi: 10.1016/S1476-5586(03)80025-8

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Schafer M, Dutsch S, auf dem Keller U, Werner S. Nrf2: a central regulator of UV protection in the epidermis. Cell Cycle. (2010) 9:2917–8. doi: 10.4161/cc.9.15.12701

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Brand RM, Wipf P, Durham A, Epperly MW, Greenberger JS, Falo LD Jr. Targeting mitochondrial oxidative stress to mitigate UV-Induced skin damage. Front Pharmacol. (2018) 9:920. doi: 10.3389/fphar.2018.00920

PubMed Abstract | CrossRef Full Text | Google Scholar

35. de fine Olivarius F, Wulf HC, Crosby J, Norval M. The sunscreening effect of urocanic acid. Photodermatol Photoimmunol Photomed. (1996) 12:95–9. doi: 10.1111/j.1600-0781.1996.tb00183.x

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Meloni M, Nicolay JF. Dynamic monitoring of glutathione redox status in UV-B irradiated reconstituted epidermis: effect of antioxidant activity on skin homeostasis. Toxicol In Vitro. (2003) 17:609–13. doi: 10.1016/s0887-2333(03)00114-0

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Rinnerthaler M, Bischof J, Streubel MK, Trost A, Richter K. Oxidative stress in aging human skin. Biomolecules. (2015) 5:545–89. doi: 10.3390/biom5020545

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Pedic L, Pondeljak N, Situm M. Recent information on photoaging mechanisms and the preventive role of topical sunscreen products. Acta Dermatovenerol Alpina Pannonica Adriat. (2020) 29:201–7.

PubMed Abstract | Google Scholar

39. Rabe JH, Mamelak AJ, McElgunn PJ, Morison WL, Sauder DN. Photoaging: mechanisms and repair. J Am Acad Dermatol. (2006) 55:1–19. doi: 10.1016/j.jaad.2005.05.010

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Fisher GJ, Wang ZQ, Datta SC, Varani J, Kang S, Voorhees JJ. Pathophysiology of premature skin aging induced by ultraviolet light. N Engl J Med. (1997) 337:1419–28. doi: 10.1056/NEJM199711133372003

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Pittayapruek P, Meephansan J, Prapapan O, Komine M, Ohtsuki M. Role of matrix metalloproteinases in photoaging and photocarcinogenesis. Int J Mol Sci. (2016) 17:868. doi: 10.3390/ijms17060868

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Buranasirin P, Pongpirul K. Development of a global subjective skin aging assessment score from the perspective of dermatologists. BMC Res Notes. (2019) 12:364. doi: 10.1186/s13104-019-4404-z

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Kappes UP, Elsner P. Clinical and photographic scoring of skin aging. Skin Pharmacol Appl Skin Physiol. (2003) 16:100–7. doi: 10.1159/000069024

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Vierkötter A, Ranft U, Krämer U, Sugiri D, Reimann V, Krutmann J. The SCINEXA: a novel, validated score to simultaneously assess and differentiate between intrinsic and extrinsic skin ageing. J Dermatol Sci. (2009) 53:207–11. doi: 10.1016/j.jdermsci.2008.10.001

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Toews GB, Bergstresser PR, Streilein JW. Epidermal Langerhans cell density determines whether contact hypersensitivity or unresponsiveness follows skin painting with DNFB. J Immunol. (1980) 124:445–53.

PubMed Abstract | Google Scholar

46. Yan B, Liu N, Li J, Li J, Zhu W, Kuang Y, et al. The role of Langerhans cells in epidermal homeostasis and pathogenesis of psoriasis. J Cell Mol Med. (2020) 24:11646–55. doi: 10.1111/jcmm.15834

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Grewe M, Gyufko K, Krutmann J. Interleukin-10 production by cultured human keratinocytes: regulation by ultraviolet B and ultraviolet A1 radiation. J Invest Dermatol. (1995) 104:3–6. doi: 10.1111/1523-1747.ep12613446

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Kang K, Gilliam AC, Chen G, Tootell E, Cooper KD. In human skin, UVB initiates early induction of IL-10 over IL-12 preferentially in the expanding dermal monocytic/macrophagic population. J Invest Dermatol. (1998) 111:31–8. doi: 10.1046/j.1523-1747.1998.00121.x

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Mittelbrunn M, Tejedor R, de la Fuente H, Garcia-Lopez MA, Ursa A, Penas PF, et al. Solar-simulated ultraviolet radiation induces abnormal maturation and defective chemotaxis of dendritic cells. J Invest Dermatol. (2005) 125:334–42. doi: 10.1111/j.0022-202X.2005.23824.x

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Patra V, Wagner K, Arulampalam V, Wolf P. Skin microbiome modulates the effect of ultraviolet radiation on cellular response and immune function. iScience. (2019) 15:211–22. doi: 10.1016/j.isci.2019.04.026

PubMed Abstract | CrossRef Full Text | Google Scholar

51. Achachi A, Vocanson M, Bastien P, Peguet-Navarro J, Grande S, Goujon C, et al. UV radiation induces the epidermal recruitment of dendritic cells that compensate for the depletion of Langerhans cells in human skin. J Invest Dermatol. (2015) 135:2058–67. doi: 10.1038/jid.2015.118

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. (2011) 144:646–74. doi: 10.1016/j.cell.2011.02.013

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Poon TS, Barnetson RS, Halliday GM. Prevention of immunosuppression by sunscreens in humans is unrelated to protection from erythema and dependent on protection from ultraviolet a in the face of constant ultraviolet B protection. J Invest Dermatol. (2003) 121:184–90. doi: 10.1046/j.1523-1747.2003.12317.x

PubMed Abstract | CrossRef Full Text | Google Scholar

54. De Fabo EC, Noonan FP. Mechanism of immune suppression by ultraviolet irradiation in vivo. I. Evidence for the existence of a unique photoreceptor in skin and its role in photoimmunology. J Exp Med. (1983) 158:84–98. doi: 10.1084/jem.158.1.84

PubMed Abstract | CrossRef Full Text | Google Scholar

55. Noonan FP, De Fabo EC. Immunosuppression by ultraviolet B radiation: initiation by urocanic acid. Immunol Today. (1992) 13:250–4. doi: 10.1016/0167-5699(92)90005-R

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Fourtanier A, Moyal D, Maccario J, Compan D, Wolf P, Quehenberger F, et al. Measurement of sunscreen immune protection factors in humans: a consensus paper. J Invest Dermatol. (2005) 125:403–9. doi: 10.1111/j.0022-202X.2005.23857.x

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Wolf P, Hoffmann C, Quehenberger F, Grinschgl S, Kerl H. Immune protection factors of chemical sunscreens measured in the local contact hypersensitivity model in humans. J Invest Dermatol. (2003) 121:1080–7. doi: 10.1046/j.1523-1747.2003.12361.x

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Damian DL, Matthews YJ, Phan TA, Halliday GM. An action spectrum for ultraviolet radiation-induced immunosuppression in humans. Br J Dermatol. (2011) 164:657–9. doi: 10.1111/j.1365-2133.2010.10161.x

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Diffey BL, Brown MW. The ideal spectral profile of topical sunscreens. Photochem Photobiol. (2012) 88:744–7. doi: 10.1111/j.1751-1097.2012.01084.x

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Narla S, Kohli I, Hamzavi IH, Lim HW. Visible light in photodermatology. Photochem Photobiol Sci. (2020) 19:99–104. doi: 10.1039/c9pp00425d

PubMed Abstract | CrossRef Full Text | Google Scholar

61. Regazzetti C, Sormani L, Debayle D, Bernerd F, Tulic MK, De Donatis GM, et al. Melanocytes sense blue light and regulate pigmentation through opsin-3. J Invest Dermatol. (2018) 138:171–8. doi: 10.1016/j.jid.2017.07.833

PubMed Abstract | CrossRef Full Text | Google Scholar

62. Yasumoto K, Yokoyama K, Shibata K, Tomita Y, Shibahara S. Microphthalmia-associated transcription factor as a regulator for melanocyte-specific transcription of the human tyrosinase gene. Mol Cell Biol. (1994) 14:8058–70. doi: 10.1128/mcb.14.12.8058-8070.1994

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Wolf P, Maier H, Mullegger RR, Chadwick CA, Hofmann-Wellenhof R, Soyer HP, et al. Topical treatment with liposomes containing T4 endonuclease V protects human skin in vivo from ultraviolet-induced upregulation of interleukin-10 and tumor necrosis factor-alpha. J Invest Dermatol. (2000) 114:149–56. doi: 10.1046/j.1523-1747.2000.00839.x

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Wolf P, Yarosh DB, Kripke ML. Effects of sunscreens and a DNA excision repair enzyme on ultraviolet radiation-induced inflammation, immune suppression, and cyclobutane pyrimidine dimer formation in mice. J Invest Dermatol. (1993) 101:523–7. doi: 10.1111/1523-1747.ep12365902

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Yarosh D, Klein J, O’Connor A, Hawk J, Rafal E, Wolf P. Effect of topically applied T4 endonuclease V in liposomes on skin cancer in Xeroderma pigmentosum: a randomised study. Xeroderma pigmentosum study group. Lancet. (2001) 357:926–9. doi: 10.1016/s0140-6736(00)04214-8

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Hofer A, Legat FJ, Gruber-Wackernagel A, Quehenberger F, Wolf P. Topical liposomal DNA-repair enzymes in polymorphic light eruption. Photochem Photobiol Sci. (2011) 10:1118–28. doi: 10.1039/c1pp05009e

PubMed Abstract | CrossRef Full Text | Google Scholar

67. Parrado C, Philips N, Gilaberte Y, Juarranz A, González S. Oral photoprotection: effective agents and potential candidates. Front Med (Lausanne). (2018) 5:188. doi: 10.3389/fmed.2018.00188

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Lankford KV, Mosunjac M, Hillyer CD. Effects of UVB radiation on cytokine generation, cell adhesion molecules, and cell activation markers in T-lymphocytes and peripheral blood HPCs. Transfusion. (2000) 40:361–7. doi: 10.1046/j.1537-2995.2000.40030361.x

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Middelkamp-Hup MA, Pathak MA, Parrado C, Garcia-Caballero T, Rius-Diaz F, Fitzpatrick TB, et al. Orally administered Polypodium leucotomos extract decreases psoralen-UVA-induced phototoxicity, pigmentation, and damage of human skin. J Am Acad Dermatol. (2004) 50:41–9. doi: 10.1016/s0190-9622(03)02732-4

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Mantena SK, Meeran SM, Elmets CA, Katiyar SK. Orally administered green tea polyphenols prevent ultraviolet radiation-induced skin cancer in mice through activation of cytotoxic T cells and inhibition of angiogenesis in tumors. J Nutr. (2005) 135:2871–7. doi: 10.1093/jn/135.12.2871

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Nestor MS, Berman B, Swenson N. Safety and efficacy of oral Polypodium leucotomos extract in healthy adult subjects. J Clin Aesthet Dermatol. (2015) 8:19–23.

PubMed Abstract | Google Scholar

72. Aguilera J, Vicente-Manzanares M, de Gálvez MV, Herrera-Ceballos E, Rodríguez-Luna A, González S. Booster effect of a natural extract of Polypodium leucotomos (Fernblock§) that improves the UV barrier function and immune protection capability of sunscreen formulations. Front Med. (2021) 8:790. doi: 10.3389/fmed.2021.684665

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Mulero M, Rodriguez-Yanes E, Nogues MR, Giralt M, Romeu M, Gonzalez S, et al. Polypodium leucotomos extract inhibits glutathione oxidation and prevents Langerhans cell depletion induced by UVB/UVA radiation in a hairless rat model. Exp Dermatol. (2008) 17:653–8. doi: 10.1111/j.1600-0625.2007.00684.x

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Middelkamp-Hup MA, Pathak MA, Parrado C, Goukassian D, Rius-Diaz F, Mihm MC, et al. Oral Polypodium leucotomos extract decreases ultraviolet-induced damage of human skin. J Am Acad Dermatol. (2004) 51:910–8. doi: 10.1016/j.jaad.2004.06.027

PubMed Abstract | CrossRef Full Text | Google Scholar

75. Kohli I, Shafi R, Isedeh P, Griffith JL, Al-Jamal MS, Silpa-Archa N, et al. The impact of oral Polypodium leucotomos extract on ultraviolet B response: a human clinical study. J Am Acad Dermatol. (2017) 77:33–41.e1. doi: 10.1016/j.jaad.2017.01.044

PubMed Abstract | CrossRef Full Text | Google Scholar

76. Parrish JA, Fitzpatrick TB, Tanenbaum L, Pathak MA. Photochemotherapy of psoriasis with oral methoxsalen and longwave ultraviolet light. N Engl J Med. (1974) 291:1207–11. doi: 10.1056/NEJM197412052912301

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Rathod DG, Muneer H, Masood S. Phototherapy. Treasure Island, FL: StatPearls Publishing (2021).

Google Scholar

78. Villa A, Viera MH, Amini S, Huo R, Perez O, Ruiz P, et al. Decrease of ultraviolet A light-induced “common deletion” in healthy volunteers after oral Polypodium leucotomos extract supplement in a randomized clinical trial. J Am Acad Dermatol. (2010) 62:511–3. doi: 10.1016/j.jaad.2009.05.045

PubMed Abstract | CrossRef Full Text | Google Scholar

79. Mohammad TF, Kohli I, Nicholson CL, Treyger G, Chaowattanapanit S, Nahhas AF, et al. Oral polypodium leucotomos extract and its impact on visible light-induced pigmentation in human subjects. J Drugs Dermatol. (2019) 18:1198–203.

PubMed Abstract | Google Scholar

80. Portillo M, Mataix M, Alonso-Juarranz M, Lorrio S, Villalba M, Rodriguez-Luna A, et al. The aqueous extract of Polypodium leucotomos (Fernblock((R))) regulates opsin 3 and prevents photooxidation of melanin precursors on skin cells exposed to blue light emitted from digital devices. Antioxidants. (2021) 10:400. doi: 10.3390/antiox10030400

PubMed Abstract | CrossRef Full Text | Google Scholar

81. Zamarron A, Lorrio S, Gonzalez S, Juarranz A. Fernblock prevents dermal cell damage induced by visible and infrared a radiation. Int J Mol Sci. (2018) 19:2250. doi: 10.3390/ijms19082250

PubMed Abstract | CrossRef Full Text | Google Scholar

82. Capote R, Alonso-Lebrero JL, Garcia F, Brieva A, Pivel JP, Gonzalez S. Polypodium leucotomos extract inhibits trans-urocanic acid photoisomerization and photodecomposition. J Photochem Photobiol B Biol. (2006) 82:173–9. doi: 10.1016/j.jphotobiol.2005.11.005

PubMed Abstract | CrossRef Full Text | Google Scholar

83. Zattra E, Coleman C, Arad S, Helms E, Levine D, Bord E, et al. Oral Polypodium leucotomos decreases UV-induced Cox-2 expression, inflammation, and enhances DNA repair in Xpc +/- mice. Am J Pathol. (2009) 175:1952–61. doi: 10.2353/ajpath.2009.090351

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Alonso-Lebrero JL, Domínguez-Jiménez C, Tejedor R, Brieva A, Pivel JP. Photoprotective properties of a hydrophilic extract of the fern Polypodium leucotomos on human skin cells. J Photochem Photobiol B. (2003) 70:31–7. doi: 10.1016/s1011-1344(03)00051-4

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Gonzalez S, Pathak MA, Cuevas J, Villarrubia VG, Fitzpatrick TB. Topical or oral administration with an extract of Polypodium leucotomos prevents acute sunburn and psoralen-induced phototoxic reactions as well as depletion of Langerhans cells in human skin. Photodermatol Photoimmunol Photomed. (1997) 13:50–60. doi: 10.1111/j.1600-0781.1997.tb00108.x

PubMed Abstract | CrossRef Full Text | Google Scholar

86. Parrado C, Nicolas J, Juarranz A, Gonzalez S. The role of the aqueous extract Polypodium leucotomos in photoprotection. Photochem Photobiol Sci. (2020) 19:831–43. doi: 10.1039/d0pp00124d

CrossRef Full Text | Google Scholar

87. Schalka S, Coelho Donato L. Evaluation of effectiveness of a sunscreen containing Polypodium leucatomos extract in reducing the sun damage to the skin. Surg Cosmet Dermatol. (2019) 11:310–8.

Google Scholar

88. Zhang Y, Cai Y, Shi M, Jiang S, Cui S, Wu Y, et al. The prevalence of vitiligo: a meta-analysis. PLoS One. (2016) 11:e0163806. doi: 10.1371/journal.pone.0163806

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Parisi R, Iskandar IYK, Kontopantelis E, Augustin M, Griffiths CEM, Ashcroft DM, et al. National, regional, and worldwide epidemiology of psoriasis: systematic analysis and modelling study. BMJ. (2020) 369:m1590. doi: 10.1136/bmj.m1590

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Stoeger T, Adler H. “Novel” triggers of herpesvirus reactivation and their potential health relevance. Front Microbiol. (2018) 9:3207. doi: 10.3389/fmicb.2018.03207

PubMed Abstract | CrossRef Full Text | Google Scholar

91. Patra V, Gallais Serezal I, Wolf P. Potential of skin microbiome, pro- and/or pre-biotics to affect local cutaneous responses to UV exposure. Nutrients. (2020) 12:1795. doi: 10.3390/nu12061795

PubMed Abstract | CrossRef Full Text | Google Scholar

92. Wang Y, Zhu W, Shu M, Jiang Y, Gallo RL, Liu YT, et al. The response of human skin commensal bacteria as a reflection of UV radiation: UV-B decreases porphyrin production. PLoS One. (2012) 7:e47798. doi: 10.1371/journal.pone.0047798

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Spittaels KJ, van Uytfanghe K, Zouboulis CC, Stove C, Crabbe A, Coenye T. Porphyrins produced by acneic Cutibacterium acnes strains activate the inflammasome by inducing K(+) leakage. iScience. (2021) 24:102575. doi: 10.1016/j.isci.2021.102575

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Gracia-Cazana T, Gonzalez S, Parrado C, Juarranz A, Gilaberte Y. Influence of the exposome on skin cancer. Actas Dermosifiliogr (Engl Ed). (2020) 111:460–70. doi: 10.1016/j.ad.2020.04.008

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Nakatsuji T, Chen TH, Butcher AM, Trzoss LL, Nam SJ, Shirakawa KT, et al. A commensal strain of Staphylococcus epidermidis protects against skin neoplasia. Sci Adv. (2018) 4:eaao4502.

Google Scholar

96. Parrado C, Mercado-Saenz S, Perez-Davo A, Gilaberte Y, Gonzalez S, Juarranz A. Environmental stressors on skin aging. Mechanistic insights. Front Pharmacol. (2019) 10:759. doi: 10.3389/fphar.2019.00759

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: sunscreen, photoprotection, UV, ISO, skin immunity, polypodium leucomotos

Citation: González S, Aguilera J, Berman B, Calzavara-Pinton P, Gilaberte Y, Goh C-L, Lim HW, Schalka S, Stengel F, Wolf P and Xiang F (2022) Expert Recommendations on the Evaluation of Sunscreen Efficacy and the Beneficial Role of Non-filtering Ingredients. Front. Med. 9:790207. doi: 10.3389/fmed.2022.790207

Received: 06 October 2021; Accepted: 04 March 2022;
Published: 31 March 2022.

Edited by:

Robert Gniadecki, University of Alberta, Canada

Reviewed by:

Giovanni Damiani, University of Milan, Italy

Copyright © 2022 González, Aguilera, Berman, Calzavara-Pinton, Gilaberte, Goh, Lim, Schalka, Stengel, Wolf and Xiang. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Salvador González, salvagonrod@gmail.com

Download