Allowed Publications
Slot System
Featured Buckets
Featured Buckets Admin

Racial Limitations of Fitzpatrick Skin Type

Article Type
Changed
Thu, 10/29/2020 - 14:46
Display Headline
Racial Limitations of Fitzpatrick Skin Type
In Collaboration With the Skin of Color Society

Fitzpatrick skin type (FST) is the most commonly used classification system in dermatologic practice. It was developed by Thomas B. Fitzpatrick, MD, PhD, in 1975 to assess the propensity of the skin to burn during phototherapy.1 Fitzpatrick skin type also can be used to assess the clinical benefits and efficacy of cosmetic procedures, including laser hair removal, chemical peel and dermabrasion, tattoo removal, spray tanning, and laser resurfacing for acne scarring.2 The original FST classifications included skin types I through IV; skin types V and VI were later added to include individuals of Asian, Indian, and African origin.1 As a result, FST often is used by providers as a means of describing constitutive skin color and ethnicity.3

How did FST transition from describing the propensity of the skin to burn from UV light exposure to categorizing skin color, thereby becoming a proxy for race? It most likely occurred because there has not been another widely adopted classification system for describing skin color that can be applied to all skin types. Even when the FST classification scale is used as intended, there are inconsistencies with its accuracy; for example, self-reported FSTs have correlated poorly with sunburn risk as well as physician-reported FSTs.4,5 Although physician-reported FSTs have been demonstrated to correlate with race, race does not consistently correlate with objective measures of pigmentation or self-reported FSTs.5 For example, Japanese women often self-identify as FST type II, but Asian skin generally is considered to be nonwhite.1 Fitzpatrick himself acknowledged that race and ethnicity are cultural and political terms with no scientific basis.6 Fitzpatrick skin type also has been demonstrated to correlate poorly with constitutive skin color and minimal erythema dose values.7

We conducted an anonymous survey of dermatologists and dermatology trainees to evaluate how providers use FST in their clinical practice as well as how it is used to describe race and ethnicity.

Methods

The survey was distributed electronically to dermatologists and dermatology trainees from March 13 to March 28, 2019, using the Association of Professors of Dermatology listserv, as well as in person at the annual Skin of Color Society meeting in Washington, DC, on February 28, 2019. The 8-item survey included questions about physician demographics (ie, primary practice setting, board certification, and geographic location); whether the respondent identified as an individual with skin of color; and how the respondent utilized FST in clinical notes (ie, describing race/ethnicity, skin cancer risk, and constitutive [baseline] skin color; determining initial phototherapy dosage and suitability for laser treatments, and likelihood of skin burning). A t test was used to determine whether dermatologists who identified as having skin of color utilized FST differently.

Results

A total of 141 surveys were returned, and 140 respondents were included in the final analysis. Given the methods used to distribute the survey, a response rate could not be calculated. The respondents included more board-certified dermatologists (70%) than dermatology trainees (30%). Ninety-three percent of respondents indicated an academic institution as their primary practice location. Notably, 26% of respondents self-identified as having skin of color.

Forty-one percent of all respondents agreed that FST should be included in their clinical documentation. In response to the question “In what scenarios would you refer to FST in a clinical note?” 31% said they used FST to describe patients’ race or ethnicity, 47% used it to describe patients’ constitutive skin color, and 22% utilized it in both scenarios. Respondents who did not identify as having skin of color were more likely to use FST to describe constitutive skin color, though this finding was not statistically significant (P=.063). Anecdotally, providers also included FST in clinical notes on postinflammatory hyperpigmentation, melasma, and treatment with cryotherapy.

 

 

Comment

The US Census Bureau has estimated that half of the US population will be of non-European descent by 2050.8 As racial and ethnic distinctions continue to be blurred, attempts to include all nonwhite skin types under the umbrella term skin of color becomes increasingly problematic. The true number of skin colors is unknown but likely is infinite, as Brazilian artist Angélica Dass has demonstrated with her photographic project “Humanae” (Figure). Given this shift in demographics and the limitations of the FST, alternative methods of describing skin color must be developed.

Artist Angélica Dass rethinks the concept of race by showing the diversity of human skin colors in her global photographic mosaic.
© Angélica Dass | Humanae Work in Progress (Courtesy of the artist).

The results of our survey suggest that approximately one-third to half of academic dermatologists/dermatology trainees use FST to describe race/ethnicity and/or constitutive skin color. This misuse of FST may occur more frequently among physicians who do not identify as having skin of color. Additionally, misuse of FST in academic settings may be problematic and confusing for medical students who may learn to use this common dermatologic tool outside of its original intent.



We acknowledge that the conundrum of how to classify individuals with nonwhite skin or skin of color is not simply answered. Several alternative skin classification models have been proposed to improve the sensitivity and specificity of identifying patients with skin of color (Table). Refining FST classification is one approach. Employing terms such as skin irritation, tenderness, itching, or skin becoming darker from sun exposure rather than painful burn or tanning may result in better identification.1,4 A study conducted in India modified the FST questionnaire to acknowledge cultural behaviors.15 Because lighter skin is culturally valued in this population, patient experience with purposeful sun exposure was limited; thus, the questionnaire was modified to remove questions on the use of tanning booths and/or creams as well as sun exposure and instead included more objective questions regarding dark brown eye color, black and dark brown hair color, and dark brown skin color.15 Other studies have suggested that patient-reported photosensitivity assessed via a questionnaire is a valid measure for assessing FST but is associated with an overestimation of skin color, known as “the dark shift.”20



Sharma et al15 utilized reflectance spectrophotometry as an objective measure of melanin and skin erythema. The melanin index consistently showed a positive correlation with FSTs as opposed to the erythema index, which correlated poorly.15 Although reflectance spectrometry accurately identifies skin color in patients with nonwhite skin,21,22 it is an impractical and cost-prohibitive tool for daily practice. A more practical tool for the clinical setting would be a visual color scale with skin hues spanning FST types I to VI, including bands of increasingly darker gradations that would be particularly useful in assessing skin of color. Once such tool is the Taylor Hyperpigmentation Scale.17 Although currently not widely available, this tool could be further refined with additional skin hues.

Conclusion

Other investigators have criticized the various limitations of FST, including physician vs patient assessment, interview vs questionnaire, and phrasing of questions on skin type.23 Our findings suggest that medical providers should be cognizant of conflating race and ethnicity with FST. Two authors of this report (O.R.W. and J.E.D.) are medical students with skin of color and frequently have observed the addition of FST to the medical records of patients who were not receiving phototherapy as a proxy for race. We believe that more culturally appropriate and clinically relevant methods for describing skin of color need to be developed and, in the interim, the original intent of FST should be emphasized and incorporated in medical school and resident education.

Acknowledgment
The authors thank Adewole Adamson, MD (Austin, Texas), for discussion and feedback.

References
  1. Goldsmith LA, Katz SI, Gilchrest BA, et al, eds. Fitzpatrick’s Dermatology in General Medicine. 8th ed. New York, NY: The McGraw-Hill Companies; 2012.
  2. Sachdeva S. Fitzpatrick skin typing: applications in dermatology. Indian J Dermatol Venereol Leprol. 2009;75:93-96.
  3. Everett JS, Budescu M, Sommers MS. Making sense of skin color in clinical care. Clin Nurs Res. 2012;21:495-516.
  4. Eilers S, Bach DQ, Gaber R, et al. Accuracy of self-report in assessingFitzpatrick skin phototypes I through VI. JAMA Dermatol. 2013;149:1289-1294.
  5. He SY, McCulloch CE, Boscardin WJ, et al. Self-reported pigmentary phenotypes and race are significant but incomplete predictors of Fitzpatrick skin phototype in an ethnically diverse population. J Am Acad Dermatol. 2014;71:731-737.
  6. Fitzpatrick TB. The validity and practicality of sun-reactive skin types I through VI. Arch Dermatol. 1988;124:869-871.
  7. Leenutaphong V. Relationship between skin color and cutaneous response to ultraviolet radiation in Thai. Photodermatol Photoimmunol Photomed. 1996;11:198-203.
  8. Colby SL, Ortman JM. Projections of the Size and Composition of the US Population: 2014 to 2060. Washington, DC: US Census Bureau; 2015.
  9. Baumann L. Understanding and treating various skin types: the Baumann Skin Type Indicator. Dermatol Clin. 2008;26:359-373.
  10. Fanous N. A new patient classification for laser resurfacing and peels: predicting responses, risks, and results. Aesthetic Plast Surg. 2002;26:99-104.
  11. Glogau RG. Chemical peeling and aging skin. J Geriatric Dermatol. 1994;2:30-35.
  12. Goldman M. Universal classification of skin type. In: Shiffman M, Mirrafati S, Lam S, et al, eds. Simplified Facial Rejuvenation. Berlin, Heidelberg, Germany: Springer; 2008:47-50.
  13. Kawada A. UVB-induced erythema, delayed tanning, and UVA-induced immediate tanning in Japanese skin. Photodermatol. 1986;3:327-333.
  14. Lancer HA. Lancer Ethnicity Scale (LES). Lasers Surg Med. 1998;22:9.
  15. Sharma VK, Gupta V, Jangid BL, et al. Modification of the Fitzpatrick system of skin phototype classification for the Indian population, and its correlation with narrowband diffuse reflectance spectrophotometry. Clin Exp Dermatol. 2018;43:274-280.
  16. Roberts WE. The Roberts Skin Type Classification System. J Drugs Dermatol. 2008;7:452-456.
  17. Taylor SC, Arsonnaud S, Czernielewski J. The Taylor hyperpigmentation scale: a new visual assessment tool for the evaluation of skin color and pigmentation. Cutis. 2005;76:270-274.
  18. Treesirichod A, Chansakulporn S, Wattanapan P. Correlation between skin color evaluation by skin color scale chart and narrowband reflectance spectrophotometer. Indian J Dermatol. 2014;59:339-342.
  19. Willis I, Earles RM. A new classification system relevant to people of African descent. J Cosmet Dermatol. 2005;18:209-216.
  20. Reeder AI, Hammond VA, Gray AR. Questionnaire items to assess skin color and erythemal sensitivity: reliability, validity, and “the dark shift.” Cancer Epidemiol Biomarkers Prev. 2010;19:1167-1173.
  21. Dwyer T, Muller HK, Blizzard L, et al. The use of spectrophotometry to estimate melanin density in Caucasians. Cancer Epidemiol Biomarkers Prev. 1998;7:203-206.
  22. Pershing LK, Tirumala VP, Nelson JL, et al. Reflectance spectrophotometer: the dermatologists’ sphygmomanometer for skin phototyping? J Invest Dermatol. 2008;128:1633-1640. 
  23. Trakatelli M, Bylaite-Bucinskiene M, Correia O, et al. Clinical assessment of skin phototypes: watch your words! Eur J Dermatol. 2017;27:615-619.
Article PDF
Author and Disclosure Information

Ms. Ware is from the Howard University College of Medicine, Washington, DC. Ms. Dawson is from the University of Washington School of Medicine, Seattle. Dr. Shinohara is from the Division of Dermatology, Department of Medicine, and the Division of Dermatopathology, Department of Pathology, University of Washington. Dr. Taylor is from the Department of Dermatology, Perelman School of Medicine, University of Pennsylvania, Philadelphia.

The authors report no conflict of interest.

This article was funded by the American Academy of Dermatology Diversity Mentorship Program.

Correspondence: Olivia R. Ware, BA, Howard University College of Medicine, 520 W St NW, Washington, DC 20059 (olivia.ware@bison.howard.edu).

Issue
Cutis - 105(2)
Publications
Topics
Page Number
77-80
Sections
Author and Disclosure Information

Ms. Ware is from the Howard University College of Medicine, Washington, DC. Ms. Dawson is from the University of Washington School of Medicine, Seattle. Dr. Shinohara is from the Division of Dermatology, Department of Medicine, and the Division of Dermatopathology, Department of Pathology, University of Washington. Dr. Taylor is from the Department of Dermatology, Perelman School of Medicine, University of Pennsylvania, Philadelphia.

The authors report no conflict of interest.

This article was funded by the American Academy of Dermatology Diversity Mentorship Program.

Correspondence: Olivia R. Ware, BA, Howard University College of Medicine, 520 W St NW, Washington, DC 20059 (olivia.ware@bison.howard.edu).

Author and Disclosure Information

Ms. Ware is from the Howard University College of Medicine, Washington, DC. Ms. Dawson is from the University of Washington School of Medicine, Seattle. Dr. Shinohara is from the Division of Dermatology, Department of Medicine, and the Division of Dermatopathology, Department of Pathology, University of Washington. Dr. Taylor is from the Department of Dermatology, Perelman School of Medicine, University of Pennsylvania, Philadelphia.

The authors report no conflict of interest.

This article was funded by the American Academy of Dermatology Diversity Mentorship Program.

Correspondence: Olivia R. Ware, BA, Howard University College of Medicine, 520 W St NW, Washington, DC 20059 (olivia.ware@bison.howard.edu).

Article PDF
Article PDF
In Collaboration With the Skin of Color Society
In Collaboration With the Skin of Color Society

Fitzpatrick skin type (FST) is the most commonly used classification system in dermatologic practice. It was developed by Thomas B. Fitzpatrick, MD, PhD, in 1975 to assess the propensity of the skin to burn during phototherapy.1 Fitzpatrick skin type also can be used to assess the clinical benefits and efficacy of cosmetic procedures, including laser hair removal, chemical peel and dermabrasion, tattoo removal, spray tanning, and laser resurfacing for acne scarring.2 The original FST classifications included skin types I through IV; skin types V and VI were later added to include individuals of Asian, Indian, and African origin.1 As a result, FST often is used by providers as a means of describing constitutive skin color and ethnicity.3

How did FST transition from describing the propensity of the skin to burn from UV light exposure to categorizing skin color, thereby becoming a proxy for race? It most likely occurred because there has not been another widely adopted classification system for describing skin color that can be applied to all skin types. Even when the FST classification scale is used as intended, there are inconsistencies with its accuracy; for example, self-reported FSTs have correlated poorly with sunburn risk as well as physician-reported FSTs.4,5 Although physician-reported FSTs have been demonstrated to correlate with race, race does not consistently correlate with objective measures of pigmentation or self-reported FSTs.5 For example, Japanese women often self-identify as FST type II, but Asian skin generally is considered to be nonwhite.1 Fitzpatrick himself acknowledged that race and ethnicity are cultural and political terms with no scientific basis.6 Fitzpatrick skin type also has been demonstrated to correlate poorly with constitutive skin color and minimal erythema dose values.7

We conducted an anonymous survey of dermatologists and dermatology trainees to evaluate how providers use FST in their clinical practice as well as how it is used to describe race and ethnicity.

Methods

The survey was distributed electronically to dermatologists and dermatology trainees from March 13 to March 28, 2019, using the Association of Professors of Dermatology listserv, as well as in person at the annual Skin of Color Society meeting in Washington, DC, on February 28, 2019. The 8-item survey included questions about physician demographics (ie, primary practice setting, board certification, and geographic location); whether the respondent identified as an individual with skin of color; and how the respondent utilized FST in clinical notes (ie, describing race/ethnicity, skin cancer risk, and constitutive [baseline] skin color; determining initial phototherapy dosage and suitability for laser treatments, and likelihood of skin burning). A t test was used to determine whether dermatologists who identified as having skin of color utilized FST differently.

Results

A total of 141 surveys were returned, and 140 respondents were included in the final analysis. Given the methods used to distribute the survey, a response rate could not be calculated. The respondents included more board-certified dermatologists (70%) than dermatology trainees (30%). Ninety-three percent of respondents indicated an academic institution as their primary practice location. Notably, 26% of respondents self-identified as having skin of color.

Forty-one percent of all respondents agreed that FST should be included in their clinical documentation. In response to the question “In what scenarios would you refer to FST in a clinical note?” 31% said they used FST to describe patients’ race or ethnicity, 47% used it to describe patients’ constitutive skin color, and 22% utilized it in both scenarios. Respondents who did not identify as having skin of color were more likely to use FST to describe constitutive skin color, though this finding was not statistically significant (P=.063). Anecdotally, providers also included FST in clinical notes on postinflammatory hyperpigmentation, melasma, and treatment with cryotherapy.

 

 

Comment

The US Census Bureau has estimated that half of the US population will be of non-European descent by 2050.8 As racial and ethnic distinctions continue to be blurred, attempts to include all nonwhite skin types under the umbrella term skin of color becomes increasingly problematic. The true number of skin colors is unknown but likely is infinite, as Brazilian artist Angélica Dass has demonstrated with her photographic project “Humanae” (Figure). Given this shift in demographics and the limitations of the FST, alternative methods of describing skin color must be developed.

Artist Angélica Dass rethinks the concept of race by showing the diversity of human skin colors in her global photographic mosaic.
© Angélica Dass | Humanae Work in Progress (Courtesy of the artist).

The results of our survey suggest that approximately one-third to half of academic dermatologists/dermatology trainees use FST to describe race/ethnicity and/or constitutive skin color. This misuse of FST may occur more frequently among physicians who do not identify as having skin of color. Additionally, misuse of FST in academic settings may be problematic and confusing for medical students who may learn to use this common dermatologic tool outside of its original intent.



We acknowledge that the conundrum of how to classify individuals with nonwhite skin or skin of color is not simply answered. Several alternative skin classification models have been proposed to improve the sensitivity and specificity of identifying patients with skin of color (Table). Refining FST classification is one approach. Employing terms such as skin irritation, tenderness, itching, or skin becoming darker from sun exposure rather than painful burn or tanning may result in better identification.1,4 A study conducted in India modified the FST questionnaire to acknowledge cultural behaviors.15 Because lighter skin is culturally valued in this population, patient experience with purposeful sun exposure was limited; thus, the questionnaire was modified to remove questions on the use of tanning booths and/or creams as well as sun exposure and instead included more objective questions regarding dark brown eye color, black and dark brown hair color, and dark brown skin color.15 Other studies have suggested that patient-reported photosensitivity assessed via a questionnaire is a valid measure for assessing FST but is associated with an overestimation of skin color, known as “the dark shift.”20



Sharma et al15 utilized reflectance spectrophotometry as an objective measure of melanin and skin erythema. The melanin index consistently showed a positive correlation with FSTs as opposed to the erythema index, which correlated poorly.15 Although reflectance spectrometry accurately identifies skin color in patients with nonwhite skin,21,22 it is an impractical and cost-prohibitive tool for daily practice. A more practical tool for the clinical setting would be a visual color scale with skin hues spanning FST types I to VI, including bands of increasingly darker gradations that would be particularly useful in assessing skin of color. Once such tool is the Taylor Hyperpigmentation Scale.17 Although currently not widely available, this tool could be further refined with additional skin hues.

Conclusion

Other investigators have criticized the various limitations of FST, including physician vs patient assessment, interview vs questionnaire, and phrasing of questions on skin type.23 Our findings suggest that medical providers should be cognizant of conflating race and ethnicity with FST. Two authors of this report (O.R.W. and J.E.D.) are medical students with skin of color and frequently have observed the addition of FST to the medical records of patients who were not receiving phototherapy as a proxy for race. We believe that more culturally appropriate and clinically relevant methods for describing skin of color need to be developed and, in the interim, the original intent of FST should be emphasized and incorporated in medical school and resident education.

Acknowledgment
The authors thank Adewole Adamson, MD (Austin, Texas), for discussion and feedback.

Fitzpatrick skin type (FST) is the most commonly used classification system in dermatologic practice. It was developed by Thomas B. Fitzpatrick, MD, PhD, in 1975 to assess the propensity of the skin to burn during phototherapy.1 Fitzpatrick skin type also can be used to assess the clinical benefits and efficacy of cosmetic procedures, including laser hair removal, chemical peel and dermabrasion, tattoo removal, spray tanning, and laser resurfacing for acne scarring.2 The original FST classifications included skin types I through IV; skin types V and VI were later added to include individuals of Asian, Indian, and African origin.1 As a result, FST often is used by providers as a means of describing constitutive skin color and ethnicity.3

How did FST transition from describing the propensity of the skin to burn from UV light exposure to categorizing skin color, thereby becoming a proxy for race? It most likely occurred because there has not been another widely adopted classification system for describing skin color that can be applied to all skin types. Even when the FST classification scale is used as intended, there are inconsistencies with its accuracy; for example, self-reported FSTs have correlated poorly with sunburn risk as well as physician-reported FSTs.4,5 Although physician-reported FSTs have been demonstrated to correlate with race, race does not consistently correlate with objective measures of pigmentation or self-reported FSTs.5 For example, Japanese women often self-identify as FST type II, but Asian skin generally is considered to be nonwhite.1 Fitzpatrick himself acknowledged that race and ethnicity are cultural and political terms with no scientific basis.6 Fitzpatrick skin type also has been demonstrated to correlate poorly with constitutive skin color and minimal erythema dose values.7

We conducted an anonymous survey of dermatologists and dermatology trainees to evaluate how providers use FST in their clinical practice as well as how it is used to describe race and ethnicity.

Methods

The survey was distributed electronically to dermatologists and dermatology trainees from March 13 to March 28, 2019, using the Association of Professors of Dermatology listserv, as well as in person at the annual Skin of Color Society meeting in Washington, DC, on February 28, 2019. The 8-item survey included questions about physician demographics (ie, primary practice setting, board certification, and geographic location); whether the respondent identified as an individual with skin of color; and how the respondent utilized FST in clinical notes (ie, describing race/ethnicity, skin cancer risk, and constitutive [baseline] skin color; determining initial phototherapy dosage and suitability for laser treatments, and likelihood of skin burning). A t test was used to determine whether dermatologists who identified as having skin of color utilized FST differently.

Results

A total of 141 surveys were returned, and 140 respondents were included in the final analysis. Given the methods used to distribute the survey, a response rate could not be calculated. The respondents included more board-certified dermatologists (70%) than dermatology trainees (30%). Ninety-three percent of respondents indicated an academic institution as their primary practice location. Notably, 26% of respondents self-identified as having skin of color.

Forty-one percent of all respondents agreed that FST should be included in their clinical documentation. In response to the question “In what scenarios would you refer to FST in a clinical note?” 31% said they used FST to describe patients’ race or ethnicity, 47% used it to describe patients’ constitutive skin color, and 22% utilized it in both scenarios. Respondents who did not identify as having skin of color were more likely to use FST to describe constitutive skin color, though this finding was not statistically significant (P=.063). Anecdotally, providers also included FST in clinical notes on postinflammatory hyperpigmentation, melasma, and treatment with cryotherapy.

 

 

Comment

The US Census Bureau has estimated that half of the US population will be of non-European descent by 2050.8 As racial and ethnic distinctions continue to be blurred, attempts to include all nonwhite skin types under the umbrella term skin of color becomes increasingly problematic. The true number of skin colors is unknown but likely is infinite, as Brazilian artist Angélica Dass has demonstrated with her photographic project “Humanae” (Figure). Given this shift in demographics and the limitations of the FST, alternative methods of describing skin color must be developed.

Artist Angélica Dass rethinks the concept of race by showing the diversity of human skin colors in her global photographic mosaic.
© Angélica Dass | Humanae Work in Progress (Courtesy of the artist).

The results of our survey suggest that approximately one-third to half of academic dermatologists/dermatology trainees use FST to describe race/ethnicity and/or constitutive skin color. This misuse of FST may occur more frequently among physicians who do not identify as having skin of color. Additionally, misuse of FST in academic settings may be problematic and confusing for medical students who may learn to use this common dermatologic tool outside of its original intent.



We acknowledge that the conundrum of how to classify individuals with nonwhite skin or skin of color is not simply answered. Several alternative skin classification models have been proposed to improve the sensitivity and specificity of identifying patients with skin of color (Table). Refining FST classification is one approach. Employing terms such as skin irritation, tenderness, itching, or skin becoming darker from sun exposure rather than painful burn or tanning may result in better identification.1,4 A study conducted in India modified the FST questionnaire to acknowledge cultural behaviors.15 Because lighter skin is culturally valued in this population, patient experience with purposeful sun exposure was limited; thus, the questionnaire was modified to remove questions on the use of tanning booths and/or creams as well as sun exposure and instead included more objective questions regarding dark brown eye color, black and dark brown hair color, and dark brown skin color.15 Other studies have suggested that patient-reported photosensitivity assessed via a questionnaire is a valid measure for assessing FST but is associated with an overestimation of skin color, known as “the dark shift.”20



Sharma et al15 utilized reflectance spectrophotometry as an objective measure of melanin and skin erythema. The melanin index consistently showed a positive correlation with FSTs as opposed to the erythema index, which correlated poorly.15 Although reflectance spectrometry accurately identifies skin color in patients with nonwhite skin,21,22 it is an impractical and cost-prohibitive tool for daily practice. A more practical tool for the clinical setting would be a visual color scale with skin hues spanning FST types I to VI, including bands of increasingly darker gradations that would be particularly useful in assessing skin of color. Once such tool is the Taylor Hyperpigmentation Scale.17 Although currently not widely available, this tool could be further refined with additional skin hues.

Conclusion

Other investigators have criticized the various limitations of FST, including physician vs patient assessment, interview vs questionnaire, and phrasing of questions on skin type.23 Our findings suggest that medical providers should be cognizant of conflating race and ethnicity with FST. Two authors of this report (O.R.W. and J.E.D.) are medical students with skin of color and frequently have observed the addition of FST to the medical records of patients who were not receiving phototherapy as a proxy for race. We believe that more culturally appropriate and clinically relevant methods for describing skin of color need to be developed and, in the interim, the original intent of FST should be emphasized and incorporated in medical school and resident education.

Acknowledgment
The authors thank Adewole Adamson, MD (Austin, Texas), for discussion and feedback.

References
  1. Goldsmith LA, Katz SI, Gilchrest BA, et al, eds. Fitzpatrick’s Dermatology in General Medicine. 8th ed. New York, NY: The McGraw-Hill Companies; 2012.
  2. Sachdeva S. Fitzpatrick skin typing: applications in dermatology. Indian J Dermatol Venereol Leprol. 2009;75:93-96.
  3. Everett JS, Budescu M, Sommers MS. Making sense of skin color in clinical care. Clin Nurs Res. 2012;21:495-516.
  4. Eilers S, Bach DQ, Gaber R, et al. Accuracy of self-report in assessingFitzpatrick skin phototypes I through VI. JAMA Dermatol. 2013;149:1289-1294.
  5. He SY, McCulloch CE, Boscardin WJ, et al. Self-reported pigmentary phenotypes and race are significant but incomplete predictors of Fitzpatrick skin phototype in an ethnically diverse population. J Am Acad Dermatol. 2014;71:731-737.
  6. Fitzpatrick TB. The validity and practicality of sun-reactive skin types I through VI. Arch Dermatol. 1988;124:869-871.
  7. Leenutaphong V. Relationship between skin color and cutaneous response to ultraviolet radiation in Thai. Photodermatol Photoimmunol Photomed. 1996;11:198-203.
  8. Colby SL, Ortman JM. Projections of the Size and Composition of the US Population: 2014 to 2060. Washington, DC: US Census Bureau; 2015.
  9. Baumann L. Understanding and treating various skin types: the Baumann Skin Type Indicator. Dermatol Clin. 2008;26:359-373.
  10. Fanous N. A new patient classification for laser resurfacing and peels: predicting responses, risks, and results. Aesthetic Plast Surg. 2002;26:99-104.
  11. Glogau RG. Chemical peeling and aging skin. J Geriatric Dermatol. 1994;2:30-35.
  12. Goldman M. Universal classification of skin type. In: Shiffman M, Mirrafati S, Lam S, et al, eds. Simplified Facial Rejuvenation. Berlin, Heidelberg, Germany: Springer; 2008:47-50.
  13. Kawada A. UVB-induced erythema, delayed tanning, and UVA-induced immediate tanning in Japanese skin. Photodermatol. 1986;3:327-333.
  14. Lancer HA. Lancer Ethnicity Scale (LES). Lasers Surg Med. 1998;22:9.
  15. Sharma VK, Gupta V, Jangid BL, et al. Modification of the Fitzpatrick system of skin phototype classification for the Indian population, and its correlation with narrowband diffuse reflectance spectrophotometry. Clin Exp Dermatol. 2018;43:274-280.
  16. Roberts WE. The Roberts Skin Type Classification System. J Drugs Dermatol. 2008;7:452-456.
  17. Taylor SC, Arsonnaud S, Czernielewski J. The Taylor hyperpigmentation scale: a new visual assessment tool for the evaluation of skin color and pigmentation. Cutis. 2005;76:270-274.
  18. Treesirichod A, Chansakulporn S, Wattanapan P. Correlation between skin color evaluation by skin color scale chart and narrowband reflectance spectrophotometer. Indian J Dermatol. 2014;59:339-342.
  19. Willis I, Earles RM. A new classification system relevant to people of African descent. J Cosmet Dermatol. 2005;18:209-216.
  20. Reeder AI, Hammond VA, Gray AR. Questionnaire items to assess skin color and erythemal sensitivity: reliability, validity, and “the dark shift.” Cancer Epidemiol Biomarkers Prev. 2010;19:1167-1173.
  21. Dwyer T, Muller HK, Blizzard L, et al. The use of spectrophotometry to estimate melanin density in Caucasians. Cancer Epidemiol Biomarkers Prev. 1998;7:203-206.
  22. Pershing LK, Tirumala VP, Nelson JL, et al. Reflectance spectrophotometer: the dermatologists’ sphygmomanometer for skin phototyping? J Invest Dermatol. 2008;128:1633-1640. 
  23. Trakatelli M, Bylaite-Bucinskiene M, Correia O, et al. Clinical assessment of skin phototypes: watch your words! Eur J Dermatol. 2017;27:615-619.
References
  1. Goldsmith LA, Katz SI, Gilchrest BA, et al, eds. Fitzpatrick’s Dermatology in General Medicine. 8th ed. New York, NY: The McGraw-Hill Companies; 2012.
  2. Sachdeva S. Fitzpatrick skin typing: applications in dermatology. Indian J Dermatol Venereol Leprol. 2009;75:93-96.
  3. Everett JS, Budescu M, Sommers MS. Making sense of skin color in clinical care. Clin Nurs Res. 2012;21:495-516.
  4. Eilers S, Bach DQ, Gaber R, et al. Accuracy of self-report in assessingFitzpatrick skin phototypes I through VI. JAMA Dermatol. 2013;149:1289-1294.
  5. He SY, McCulloch CE, Boscardin WJ, et al. Self-reported pigmentary phenotypes and race are significant but incomplete predictors of Fitzpatrick skin phototype in an ethnically diverse population. J Am Acad Dermatol. 2014;71:731-737.
  6. Fitzpatrick TB. The validity and practicality of sun-reactive skin types I through VI. Arch Dermatol. 1988;124:869-871.
  7. Leenutaphong V. Relationship between skin color and cutaneous response to ultraviolet radiation in Thai. Photodermatol Photoimmunol Photomed. 1996;11:198-203.
  8. Colby SL, Ortman JM. Projections of the Size and Composition of the US Population: 2014 to 2060. Washington, DC: US Census Bureau; 2015.
  9. Baumann L. Understanding and treating various skin types: the Baumann Skin Type Indicator. Dermatol Clin. 2008;26:359-373.
  10. Fanous N. A new patient classification for laser resurfacing and peels: predicting responses, risks, and results. Aesthetic Plast Surg. 2002;26:99-104.
  11. Glogau RG. Chemical peeling and aging skin. J Geriatric Dermatol. 1994;2:30-35.
  12. Goldman M. Universal classification of skin type. In: Shiffman M, Mirrafati S, Lam S, et al, eds. Simplified Facial Rejuvenation. Berlin, Heidelberg, Germany: Springer; 2008:47-50.
  13. Kawada A. UVB-induced erythema, delayed tanning, and UVA-induced immediate tanning in Japanese skin. Photodermatol. 1986;3:327-333.
  14. Lancer HA. Lancer Ethnicity Scale (LES). Lasers Surg Med. 1998;22:9.
  15. Sharma VK, Gupta V, Jangid BL, et al. Modification of the Fitzpatrick system of skin phototype classification for the Indian population, and its correlation with narrowband diffuse reflectance spectrophotometry. Clin Exp Dermatol. 2018;43:274-280.
  16. Roberts WE. The Roberts Skin Type Classification System. J Drugs Dermatol. 2008;7:452-456.
  17. Taylor SC, Arsonnaud S, Czernielewski J. The Taylor hyperpigmentation scale: a new visual assessment tool for the evaluation of skin color and pigmentation. Cutis. 2005;76:270-274.
  18. Treesirichod A, Chansakulporn S, Wattanapan P. Correlation between skin color evaluation by skin color scale chart and narrowband reflectance spectrophotometer. Indian J Dermatol. 2014;59:339-342.
  19. Willis I, Earles RM. A new classification system relevant to people of African descent. J Cosmet Dermatol. 2005;18:209-216.
  20. Reeder AI, Hammond VA, Gray AR. Questionnaire items to assess skin color and erythemal sensitivity: reliability, validity, and “the dark shift.” Cancer Epidemiol Biomarkers Prev. 2010;19:1167-1173.
  21. Dwyer T, Muller HK, Blizzard L, et al. The use of spectrophotometry to estimate melanin density in Caucasians. Cancer Epidemiol Biomarkers Prev. 1998;7:203-206.
  22. Pershing LK, Tirumala VP, Nelson JL, et al. Reflectance spectrophotometer: the dermatologists’ sphygmomanometer for skin phototyping? J Invest Dermatol. 2008;128:1633-1640. 
  23. Trakatelli M, Bylaite-Bucinskiene M, Correia O, et al. Clinical assessment of skin phototypes: watch your words! Eur J Dermatol. 2017;27:615-619.
Issue
Cutis - 105(2)
Issue
Cutis - 105(2)
Page Number
77-80
Page Number
77-80
Publications
Publications
Topics
Article Type
Display Headline
Racial Limitations of Fitzpatrick Skin Type
Display Headline
Racial Limitations of Fitzpatrick Skin Type
Sections
Inside the Article

Practice Points

  • Medical providers should be cognizant of conflating race and ethnicity with Fitzpatrick skin type (FST).
  • Misuse of FST may occur more frequently among physicians who do not identify as having skin of color.
  • Although alternative skin type classification systems have been proposed, more clinically relevant methods for describing skin of color need to be developed.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Atopic Dermatitis in Adolescents With Skin of Color

Article Type
Changed
Thu, 10/29/2020 - 14:48
Display Headline
Atopic Dermatitis in Adolescents With Skin of Color
In Collaboration With the Skin of Color Society

Data are limited on the management of atopic dermatitis (AD) in adolescents, particularly in patients with skin of color, making it important to identify factors that may improve AD management in this population. Comorbid conditions (eg, acne, postinflammatory hyperpigmentation [PIH]), extracurricular activities (eg, athletics), and experimentation with cosmetics in adolescents, all of which can undermine treatment efficacy and medication adherence, make it particularly challenging to devise a therapeutic regimen in this patient population. We review the management of AD in black adolescents, with special consideration of concomitant treatment of acne vulgaris (AV) as well as lifestyle and social choices (Table).

Prevalence and Epidemiology

Atopic dermatitis affects 13% to 25% of children and 2% to 10% of adults.1,2 Population‐based studies in the United States show a higher prevalence of AD in black children (19.3%) compared to European American (EA) children (16.1%).3,4

AD in Black Adolescents

Atopic dermatitis is a common skin condition that is defined as a chronic, pruritic, inflammatory dermatosis with recurrent scaling, papules, and plaques (Figure) that usually develop during infancy and early childhood.3 Although AD severity improves for some patients in adolescence, it can be a lifelong issue affecting performance in academic and occupational settings.5 One US study of 8015 children found that there are racial and ethnic disparities in school absences among children (age range, 2–17 years) with AD, with children with skin of color being absent more often than white children.6 The same study noted that black children had a 1.5-fold higher chance of being absent 6 days over a 6-month school period compared to white children. It is postulated that AD has a greater impact on quality of life (QOL) in children with skin of color, resulting in the increased number of school absences in this population.6

Atopic dermatitis on the neck with lichenification and excoriations.

The origin of AD currently is thought to be complex and can involve skin barrier dysfunction, environmental factors, microbiome effects, genetic predisposition, and immune dysregulation.1,4 Atopic dermatitis is a heterogeneous disease with variations in the prevalence, genetic background, and immune activation patterns across racial groups.4 It is now understood to be an immune-mediated disease with multiple inflammatory pathways, with type 2–associated inflammation being a primary pathway. Patients with AD have strong helper T cell (TH2) activation, and black patients with AD have higher IgE serum levels as well as absent TH17/TH1 activation.4



Atopic dermatitis currently is seen as a defect of the epidermal barrier, with variable clinical manifestations and expressivity.7 Filaggrin is an epidermal barrier protein, encoded by the FLG gene, and plays a major role in barrier function by regulating pH and promoting hydration of the skin.4 Loss of function of the FLG gene is the most well-studied genetic risk factor for developing AD, and this mutation is seen in patients with more severe and persistent AD in addition to patients with more skin infections and allergic sensitizations.3,4 However, in the skin of color population, FLG mutations are 6 times less common than in the EA population, despite the fact that AD is more prevalent in patients of African descent.4 Therefore, the role of the FLG loss-of-function mutation and AD is not as well defined in black patients, and some researchers have found no association.3 The FLG loss-of-function mutation seems to play a smaller role in black patients than in EA patients, and other genes may be involved in skin barrier dysfunction.3,4 In a small study of patients with mild AD compared to nonaffected patients, those with AD had lower total ceramide levels in the stratum corneum of affected sites than normal skin sites in healthy individuals.8

Particular disturbances in the gut microbiome have the possibility of impacting the development of AD.9 Additionally, the development of AD may be influenced by the skin microbiome, which can change depending on body site, with fungal organisms thought to make up a large proportion of the microbiome of patients with AD. In patients with AD, there is a lack of microbial diversity and an overgrowth of Staphylococcus aureus.9

 

 

Diagnosis

Clinicians diagnose AD based on clinical characteristics, and the lack of objective criteria can hinder diagnosis.1 Thus, diagnosing AD in children with dark skin can pose a particular challenge given the varied clinical presentation of AD across skin types. Severe cases of AD may not be diagnosed or treated adequately in deeply pigmented children because erythema, a defining characteristic of AD, may be hard to identify in darker skin types.10 Furthermore, clinical erythema scores among black children may be “strongly” underestimated using scoring systems such as Eczema Area and Severity Index and SCORing Atopic Dermatitis.4 It is estimated that the risk for severe AD may be 6 times higher in black children compared to white children.10 Additionally, patients with skin of color can present with more treatment-resistant AD.4

Treatment of AD

Current treatment is focused on restoring epidermal barrier function, often with topical agents, such as moisturizers containing different amounts of emollients, occlusives, and humectants; corticosteroids; calcineurin inhibitors; and antimicrobials. Emollients such as glycol stearate, glyceryl stearate, and soy sterols function as lubricants, softening the skin. Occlusive agents include petrolatum, dimethicone, and mineral oil; they act by forming a layer to slow evaporation of water. Humectants including glycerol, lactic acid, and urea function by promoting water retention.11 For acute flares, mid- to high-potency topical corticosteroids are recommended. Also, topical calcineurin inhibitors such as tacrolimus and pimecrolimus may be used alone or in combination with topical steroids. Finally, bleach baths and topical mupirocin applied to the nares also have proved helpful in moderate to severe AD with secondary bacterial infections.11 Phototherapy can be used in adult and pediatric patients with acute and chronic AD if traditional treatments have failed.2

Systemic agents are indicated and recommended for the subset of adult and pediatric patients in whom optimized topical regimens and/or phototherapy do not adequately provide disease control or when QOL is substantially impacted. The systemic agents effective in the pediatric population include cyclosporine, azathioprine, mycophenolate mofetil, and possibly methotrexate.11 Dupilumab recently was approved by the US Food and Drug Administration for patients 12 years and older with moderate to severe AD whose disease is not well controlled with topical medications.

Patients with AD are predisposed to secondary bacterial and viral infections because of their dysfunctional skin barrier; these infections most commonly are caused by S aureus and herpes simplex virus, respectively.2 Systemic antibiotics are only recommended for patients with AD when there is clinical evidence of bacterial infection. In patients with evidence of eczema herpeticum, systemic antiviral agents should be used to treat the underlying herpes simplex virus infection.2 Atopic dermatitis typically has been studied in white patients; however, patients with skin of color have higher frequencies of treatment-resistant AD. Further research on treatment efficacy for AD in this patient population is needed, as data are limited.4

Treatment of AV in Patients With AD

Two of the most prevalent skin diseases affecting the pediatric population are AD and AV, and both can remarkably impact QOL.12 Acne is one of the most common reasons for adolescent patients to seek dermatologic care, including patients with skin of color (Fitzpatrick skin types IV to VI).13 Thus, it is to be expected that many black adolescents with AD also will have AV. For mild to moderate acne in patients with skin of color, topical retinoids and benzoyl peroxide typically are first line.13 These medications can be problematic for patients with AD, as retinoids and many other acne treatments can cause dryness, which may exacerbate AD.

 

 

Moisturizers containing ceramide can be a helpful adjunctive therapy in treating acne,14 especially in patients with AD. Modifications to application of acne medications, such as using topical retinoids every other night or mixing them with moisturizers to minimize dryness, may be beneficial to these patients. Dapsone gel 7.5% used daily also may be an option for adolescents with AD and AV. A double-blind, vehicle-controlled study demonstrated that dapsone is safe and effective for patients 12 years and older with moderate acne, and patients with Fitzpatrick skin types IV to VI rated local scaling, erythema, dryness, and stinging/burning as “none” in the study.15 Another potentially helpful topical agent in patients with AD and AV is sulfacetamide, as it is not likely to cause dryness of the skin. In a small study, sodium sulfacetamide 10% and sulfur 5% in an emollient foam vehicle showed no residual film or sulfur smell and resulted in acne reduction of 50%.16



Patients with skin of color often experience PIH in AD and acne or hypopigmentation from inflammatory dermatoses including AD.17,18 In addition to the dryness from AD and topical retinoid use, patients with skin of color may develop irritant contact dermatitis, thus leading to PIH.13 Dryness and irritant contact dermatitis also can be seen with the use of benzoyl peroxide in black patients. Because of these effects, gentle moisturizers are recommended, and both benzoyl peroxide and retinoids should be initiated at lower doses in patients with skin of color.13

For patients with severe nodulocystic acne, isotretinoin is the treatment of choice in patients with skin of color,13 but there is a dearth of clinical studies addressing complications seen in black adolescents on this treatment, especially with respect to those with AD. Of note, systemic antibiotics typically are initiated before isotretinoin; however, this strategy is falling out of favor due to concern for antibiotic resistance with long-term use.19

Impact of Athletics on AD in Black Adolescents

Because of the exacerbating effects of perspiration and heat causing itch and irritation in patients with AD, it is frequently advised that pediatric patients limit their participation in athletics because of the exacerbating effects of strenuous physical exercise on their disease.12 In one study, 429 pediatric patients or their parents/guardians completed QOL questionnaires; 89% of patients 15 years and younger with severe AD reported that their disease was impacted by athletics and outdoor activities, and 86% of these pediatric patients with severe AD responded that their social lives and leisure activities were impacted.20 Because adolescents often are involved in athletics or have mandatory physical education classes, AD may be isolating and may have a severe impact on self-esteem.

Aggressive treatment of AD with topical and systemic medications may be helpful in adolescents who may be reluctant to participate in sports because of teasing, bullying, or worsening of symptoms with heat or sweating.21 Now that dupilumab is available for adolescents, there is a chance that patients with severe and/or recalcitrant disease managed on this medication can achieve better control of their symptoms without the laboratory requirement of methotrexate and the difficulties of topical medication application, allowing them to engage in mandatory athletic classes as well as desired organized sports.

 

 

Use of Cosmetics for AD

Many adolescents experiment with cosmetics, and those with AD may use cosmetic products to cover hyperpigmented or hypopigmented lesions.18 In patients with active AD or increased sensitivity to allergens in cosmetic products, use of makeup can be a contributing factor for AD flares. Acne associated with cosmetics is especially important to consider in darker-skinned patients who may use makeup that is opaque and contains oil to conceal acne or PIH.

Allergens can be present in both cosmetics and pharmaceutical topical agents, and a Brazilian study found that approximately 89% of 813 prescription and nonprescription products (eg, topical drugs, sunscreens, moisturizers, soaps, cleansing lotions, shampoos, cosmeceuticals) contained allergens.22 Patients with AD have a higher prevalence of contact sensitization to fragrances, including balsam of Peru.23 Some AD treatments that contain fragrances have caused further skin issues in a few patients. In one case series, 3 pediatric patients developed allergic contact dermatitis to Myroxylon pereirae (balsam of Peru) when using topical treatments for their AD, and their symptoms of scalp inflammation and alopecia resolved with discontinuation.23

In a Dutch study, sensitization to Fragrance Mix I and M pereirae as well as other ingredients (eg, lanolin alcohol, Amerchol™ L 101 [a lanolin product]) was notably more common in pediatric patients with AD than in patients without AD; however, no data on patients with skin of color were included in this study.24



Because of the increased risk of sensitization to fragrances and other ingredients in patients with AD as well as the high percentage of allergens in prescription and nonprescription products, it is important to discuss all personal care products that patients may be using, not just their cosmetic products. Also, patch testing may be helpful in determining true allergens in some patients. Patch testing is recommended for patients with treatment-resistant AD, and a recent study suggested it should be done prior to long-term use of immunosuppressive agents.25 Increased steroid phobia and a push toward alternative medicines are leading both patients with AD and guardians of children with AD to look for other forms of moisturization, such as olive oil, coconut oil, sunflower seed oil, and shea butter, to decrease transepidermal water loss.26,27 An important factor in AD treatment efficacy is patient acceptability in using what is recommended.27 One study showed there was no difference in efficacy or acceptability in using a cream containing shea butter extract vs the ceramide-precursor product.27 Current data show olive oil may exacerbate dry skin and AD,26 and recommendation of any over-the-counter oils and butters in patients with AD should be made with great caution, as many of these products contain fragrances and other potential allergens.

Alternative Therapies for AD

Patients with AD often seek alternative or integrative treatment options, including dietary modifications and holistic remedies. Studies investigating the role of vitamins and supplements in treating AD are limited by sample size.28 However, there is some evidence that may support supplementation with vitamins D and E in addressing AD symptoms. The use of probiotics in treating AD is controversial, but there are studies suggesting that the use of probiotics may prove beneficial in preventing infantile AD.28 Additionally, findings from an ex vivo and in vitro study show that some conditions, including AD and acne, may benefit from the same probiotics, despite the differences in these two diseases. Both AD and acne have inflammatory and skin dysbiosis characteristics, which may be the common thread leading to both conditions potentially responding to treatment with probiotics.29

 

 

Preliminary evidence indicates that supplements containing fatty acids such as docosahexaenoic acid, sea buckthorn oil, and hemp seed oil may decrease the severity of AD.28 In a 20-week, randomized, single-blind, crossover study published in 2005, dietary hemp seed oil showed an improvement of clinical symptoms, including dry skin and itchiness, in patients with AD.30



In light of recent legalization in several states, patients may turn to use of cannabinoid products to manage AD. In a systematic review, cannabinoid use was reportedly a therapeutic option in the treatment of AD and AV; however, the data are based on preclinical work, and there are no randomized, placebo-controlled studies to support the use of cannabinoids.31 Furthermore, there is great concern that use of these products in adolescents is an even larger unknown.

Final Thoughts

Eighty percent of children diagnosed with AD experience symptom improvement before their early teens32; for those with AD during their preteen and teenage years, there can be psychological ramifications, as teenagers with AD report having fewer friends, are less socially involved, participate in fewer sports, and are absent from classes more often than their peers.5 In black patients with AD, school absences are even more common.6 Given the social and emotional impact of AD on patients with skin of color, it is imperative to treat the condition appropriately.33 There are areas of opportunity for further research on alternate dosing of existing treatments for AV in patients with AD, further recommendations for adolescent athletes with AD, and which cosmetic and alternative medicine products may be beneficial for this population to improve their QOL.

Providers should discuss medical management in a broader context considering patients’ extracurricular activities, treatment vehicle preferences, expectations, and personal care habits. It also is important to address the many possible factors that may influence treatment adherence early on, particularly in adolescents, as these could be barriers to treatment. This article highlights considerations for treating AD and comorbid conditions that may further complicate treatment in adolescent patients with skin of color. The information provided should serve as a guide in initial counseling and management of AD in adolescents with skin of color.

References
  1. Feldman SR, Cox LS, Strowd LC, et al. The challenge of managing atopic dermatitis in the United States. Am Health Drug Benefits. 2019;12:83-93.
  2. Sidbury R, Davis DM, Cohen DE, et al. Guidelines of care for the management of atopic dermatitis: section 3. management and treatment with phototherapy and systemic agents. J Am Acad Dermatol. 2014;71:327-349.
  3. Kaufman BP, Guttman-Yassky E, Alexis AF. Atopic dermatitis in diverse racial and ethnic groups—variations in epidemiology, genetics, clinical presentation and treatment. Exp Dermatol. 2018;27:340-357.
  4. Brunner PM, Guttman-Yassky E. Racial differences in atopic dermatitis. Ann Allergy Asthma Immunol. 2019;122:449-455.
  5. Vivar KL, Kruse L. The impact of pediatric skin disease on self-esteem. Int J Womens Dermatol. 2018;4:27-31.
  6. Wan J, Margolis DJ, Mitra N, et al. Racial and ethnic differences in atopic dermatitis–related school absences among US children [published online May 22, 2019]. JAMA Dermatol. doi:10.1001/jamadermatol.2019.0597.
  7. Weidinger S, Novak N. Atopic dermatitis. Lancet. 2016;387:1109-1122.
  8. Ishikawa J, Narita H, Kondo N, et al. Changes in the ceramide profile of atopic dermatitis patients. J Invest Dermatol. 2010;130:2511-2514.
  9. Chernikova D, Yuan I, Shaker M. Prevention of allergy with diverse and healthy microbiota: an update. Curr Opin Pediatr. 2019;31:418-425.
  10. Ben-Gashir MA, Hay RJ. Reliance on erythema scores may mask severe atopic dermatitis in black children compared with their white counterparts. Br J Dermatol. 2002;147:920-925.
  11. Eichenfield LF, Tom WL, Berger TG, et al. Guidelines of care for the management of atopic dermatitis: section 2. management and treatment of atopic dermatitis with topical therapies. J Am Acad Dermatol. 2014;71:116-132.
  12. Nguyen CM, Koo J, Cordoro KM. Psychodermatologic effects of atopic dermatitis and acne: a review on self-esteem and identity. Pediatr Dermatol. 2016;33:129-135.
  13. Davis EC, Callender VD. A review of acne in ethnic skin: pathogenesis, clinical manifestations, and management strategies. J Clin Aesthet Dermatol. 2010;3:24-38.
  14. Lynde CW, Andriessen A, Barankin B, et al. Moisturizers and ceramide-containing moisturizers may offer concomitant therapy with benefits. J Clin Aesthet Dermatol. 2014;7:18-26.
  15. Taylor SC, Cook-Bolden FE, McMichael A, et al. Efficacy, safety, and tolerability of topical dapsone gel, 7.5% for treatment of acne vulgaris by Fitzpatrick skin phototype. J Drugs Dermatol. 2018;17:160-167.
  16. Draelos ZD. The multifunctionality of 10% sodium sulfacetamide, 5% sulfur emollient foam in the treatment of inflammatory facial dermatoses. J Drugs Dermatol. 2010;9:234-236.
  17. Vachiramon V, Tey HL, Thompson AE, et al. Atopic dermatitis in African American children: addressing unmet needs of a common disease. Pediatr Dermatol. 2012;29:395-402.
  18. Heath CR. Managing postinflammatory hyperpigmentation in pediatric patients with skin of color. Cutis. 2018;102:71-73.
  19. Nagler AR, Milam EC, Orlow SJ. The use of oral antibiotics before isotretinoin therapy in patients with acne. J Am Acad Dermatol. 2016;74:273-279.
  20. Paller AS, McAlister RO, Doyle JJ, et al. Perceptions of physicians and pediatric patients about atopic dermatitis, its impact, and its treatment. Clin Pediatr. 2002;41:323-332.
  21. Sibbald C, Drucker AM. Patient burden of atopic dermatitis. Dermatol Clin. 2017;35:303-316.
  22. Rocha VB, Machado CJ, Bittencourt FV. Presence of allergens in the vehicles of Brazilian dermatological products. Contact Dermatitis. 2017;76:126-128.
  23. Admani S, Goldenberg A, Jacob SE. Contact alopecia: improvement of alopecia with discontinuation of fluocinolone oil in individuals allergic to balsam fragrance. Pediatr Dermatol. 2017;34:e57-e60.
  24. Uter W, Werfel T, White IR, et al. Contact allergy: a review of current problems from a clinical perspective. Int J Environ Res Public Health. 2018;15:E1108.
  25. López-Jiménez EC, Marrero-Alemán G, Borrego L. One-third of patients with therapy-resistant atopic dermatitis may benefit after patch testing [published online May 13, 2019]. J Eur Acad Dermatol Venereol. doi:10.1111/jdv.15672.
  26. Karagounis TK, Gittler JK, Rotemberg V, et al. Use of “natural” oils for moisturization: review of olive, coconut, and sunflower seed oil. Pediatr Dermatol. 2019;36:9-15.
  27. Hon KL, Tsang YC, Pong NH, et al. Patient acceptability, efficacy, and skin biophysiology of a cream and cleanser containing lipid complex with shea butter extract versus a ceramide product for eczema. Hong Kong Med J. 2015;21:417-425.
  28. Reynolds KA, Juhasz MLW, Mesinkovska NA. The role of oral vitamins and supplements in the management of atopic dermatitis: a systematic review [published online March 20, 2019]. Int J Dermatol. doi:10.1111/ijd.14404.
  29. Mottin VHM, Suyenaga ES. An approach on the potential use of probiotics in the treatment of skin conditions: acne and atopic dermatitis. Int J Dermatol. 2018;57:1425-1432.
  30. Callaway J, Schwab U, Harvima I, et al. Efficacy of dietary hempseed oil in patients with atopic dermatitis. J Dermatol Treat. 2005;16:87-94.
  31. Eagleston LRM, Kalani NK, Patel RR, et al. Cannabinoids in dermatology: a scoping review [published June 15, 2018]. Dermatol Online J. 2018;24.
  32. Kim JP, Chao LX, Simpson EL, et al. Persistence of atopic dermatitis (AD): a systematic review and meta-analysis. J Am Acad Dermatol. 2016;75:681-687.e611.
  33. de María Díaz Granados L, Quijano MA, Ramírez PA, et al. Quality assessment of atopic dermatitis clinical practice guidelines in ≤ 18 years. Arch Dermatol Res. 2018;310:29-37.
Article PDF
Author and Disclosure Information

From the Department of Dermatology, Wake Forest Baptist Medical Center, Winston-Salem, North Carolina.

The authors report no conflict of interest.

Correspondence: Amy J. McMichael, MD, Department of Dermatology, Wake Forest Baptist Medical Center, Medical Center Blvd, Winston-Salem, NC 27104 (amcmicha@wakehealth.edu).

Issue
Cutis - 104(3)
Publications
Topics
Page Number
164-168
Sections
Author and Disclosure Information

From the Department of Dermatology, Wake Forest Baptist Medical Center, Winston-Salem, North Carolina.

The authors report no conflict of interest.

Correspondence: Amy J. McMichael, MD, Department of Dermatology, Wake Forest Baptist Medical Center, Medical Center Blvd, Winston-Salem, NC 27104 (amcmicha@wakehealth.edu).

Author and Disclosure Information

From the Department of Dermatology, Wake Forest Baptist Medical Center, Winston-Salem, North Carolina.

The authors report no conflict of interest.

Correspondence: Amy J. McMichael, MD, Department of Dermatology, Wake Forest Baptist Medical Center, Medical Center Blvd, Winston-Salem, NC 27104 (amcmicha@wakehealth.edu).

Article PDF
Article PDF
In Collaboration With the Skin of Color Society
In Collaboration With the Skin of Color Society

Data are limited on the management of atopic dermatitis (AD) in adolescents, particularly in patients with skin of color, making it important to identify factors that may improve AD management in this population. Comorbid conditions (eg, acne, postinflammatory hyperpigmentation [PIH]), extracurricular activities (eg, athletics), and experimentation with cosmetics in adolescents, all of which can undermine treatment efficacy and medication adherence, make it particularly challenging to devise a therapeutic regimen in this patient population. We review the management of AD in black adolescents, with special consideration of concomitant treatment of acne vulgaris (AV) as well as lifestyle and social choices (Table).

Prevalence and Epidemiology

Atopic dermatitis affects 13% to 25% of children and 2% to 10% of adults.1,2 Population‐based studies in the United States show a higher prevalence of AD in black children (19.3%) compared to European American (EA) children (16.1%).3,4

AD in Black Adolescents

Atopic dermatitis is a common skin condition that is defined as a chronic, pruritic, inflammatory dermatosis with recurrent scaling, papules, and plaques (Figure) that usually develop during infancy and early childhood.3 Although AD severity improves for some patients in adolescence, it can be a lifelong issue affecting performance in academic and occupational settings.5 One US study of 8015 children found that there are racial and ethnic disparities in school absences among children (age range, 2–17 years) with AD, with children with skin of color being absent more often than white children.6 The same study noted that black children had a 1.5-fold higher chance of being absent 6 days over a 6-month school period compared to white children. It is postulated that AD has a greater impact on quality of life (QOL) in children with skin of color, resulting in the increased number of school absences in this population.6

Atopic dermatitis on the neck with lichenification and excoriations.

The origin of AD currently is thought to be complex and can involve skin barrier dysfunction, environmental factors, microbiome effects, genetic predisposition, and immune dysregulation.1,4 Atopic dermatitis is a heterogeneous disease with variations in the prevalence, genetic background, and immune activation patterns across racial groups.4 It is now understood to be an immune-mediated disease with multiple inflammatory pathways, with type 2–associated inflammation being a primary pathway. Patients with AD have strong helper T cell (TH2) activation, and black patients with AD have higher IgE serum levels as well as absent TH17/TH1 activation.4



Atopic dermatitis currently is seen as a defect of the epidermal barrier, with variable clinical manifestations and expressivity.7 Filaggrin is an epidermal barrier protein, encoded by the FLG gene, and plays a major role in barrier function by regulating pH and promoting hydration of the skin.4 Loss of function of the FLG gene is the most well-studied genetic risk factor for developing AD, and this mutation is seen in patients with more severe and persistent AD in addition to patients with more skin infections and allergic sensitizations.3,4 However, in the skin of color population, FLG mutations are 6 times less common than in the EA population, despite the fact that AD is more prevalent in patients of African descent.4 Therefore, the role of the FLG loss-of-function mutation and AD is not as well defined in black patients, and some researchers have found no association.3 The FLG loss-of-function mutation seems to play a smaller role in black patients than in EA patients, and other genes may be involved in skin barrier dysfunction.3,4 In a small study of patients with mild AD compared to nonaffected patients, those with AD had lower total ceramide levels in the stratum corneum of affected sites than normal skin sites in healthy individuals.8

Particular disturbances in the gut microbiome have the possibility of impacting the development of AD.9 Additionally, the development of AD may be influenced by the skin microbiome, which can change depending on body site, with fungal organisms thought to make up a large proportion of the microbiome of patients with AD. In patients with AD, there is a lack of microbial diversity and an overgrowth of Staphylococcus aureus.9

 

 

Diagnosis

Clinicians diagnose AD based on clinical characteristics, and the lack of objective criteria can hinder diagnosis.1 Thus, diagnosing AD in children with dark skin can pose a particular challenge given the varied clinical presentation of AD across skin types. Severe cases of AD may not be diagnosed or treated adequately in deeply pigmented children because erythema, a defining characteristic of AD, may be hard to identify in darker skin types.10 Furthermore, clinical erythema scores among black children may be “strongly” underestimated using scoring systems such as Eczema Area and Severity Index and SCORing Atopic Dermatitis.4 It is estimated that the risk for severe AD may be 6 times higher in black children compared to white children.10 Additionally, patients with skin of color can present with more treatment-resistant AD.4

Treatment of AD

Current treatment is focused on restoring epidermal barrier function, often with topical agents, such as moisturizers containing different amounts of emollients, occlusives, and humectants; corticosteroids; calcineurin inhibitors; and antimicrobials. Emollients such as glycol stearate, glyceryl stearate, and soy sterols function as lubricants, softening the skin. Occlusive agents include petrolatum, dimethicone, and mineral oil; they act by forming a layer to slow evaporation of water. Humectants including glycerol, lactic acid, and urea function by promoting water retention.11 For acute flares, mid- to high-potency topical corticosteroids are recommended. Also, topical calcineurin inhibitors such as tacrolimus and pimecrolimus may be used alone or in combination with topical steroids. Finally, bleach baths and topical mupirocin applied to the nares also have proved helpful in moderate to severe AD with secondary bacterial infections.11 Phototherapy can be used in adult and pediatric patients with acute and chronic AD if traditional treatments have failed.2

Systemic agents are indicated and recommended for the subset of adult and pediatric patients in whom optimized topical regimens and/or phototherapy do not adequately provide disease control or when QOL is substantially impacted. The systemic agents effective in the pediatric population include cyclosporine, azathioprine, mycophenolate mofetil, and possibly methotrexate.11 Dupilumab recently was approved by the US Food and Drug Administration for patients 12 years and older with moderate to severe AD whose disease is not well controlled with topical medications.

Patients with AD are predisposed to secondary bacterial and viral infections because of their dysfunctional skin barrier; these infections most commonly are caused by S aureus and herpes simplex virus, respectively.2 Systemic antibiotics are only recommended for patients with AD when there is clinical evidence of bacterial infection. In patients with evidence of eczema herpeticum, systemic antiviral agents should be used to treat the underlying herpes simplex virus infection.2 Atopic dermatitis typically has been studied in white patients; however, patients with skin of color have higher frequencies of treatment-resistant AD. Further research on treatment efficacy for AD in this patient population is needed, as data are limited.4

Treatment of AV in Patients With AD

Two of the most prevalent skin diseases affecting the pediatric population are AD and AV, and both can remarkably impact QOL.12 Acne is one of the most common reasons for adolescent patients to seek dermatologic care, including patients with skin of color (Fitzpatrick skin types IV to VI).13 Thus, it is to be expected that many black adolescents with AD also will have AV. For mild to moderate acne in patients with skin of color, topical retinoids and benzoyl peroxide typically are first line.13 These medications can be problematic for patients with AD, as retinoids and many other acne treatments can cause dryness, which may exacerbate AD.

 

 

Moisturizers containing ceramide can be a helpful adjunctive therapy in treating acne,14 especially in patients with AD. Modifications to application of acne medications, such as using topical retinoids every other night or mixing them with moisturizers to minimize dryness, may be beneficial to these patients. Dapsone gel 7.5% used daily also may be an option for adolescents with AD and AV. A double-blind, vehicle-controlled study demonstrated that dapsone is safe and effective for patients 12 years and older with moderate acne, and patients with Fitzpatrick skin types IV to VI rated local scaling, erythema, dryness, and stinging/burning as “none” in the study.15 Another potentially helpful topical agent in patients with AD and AV is sulfacetamide, as it is not likely to cause dryness of the skin. In a small study, sodium sulfacetamide 10% and sulfur 5% in an emollient foam vehicle showed no residual film or sulfur smell and resulted in acne reduction of 50%.16



Patients with skin of color often experience PIH in AD and acne or hypopigmentation from inflammatory dermatoses including AD.17,18 In addition to the dryness from AD and topical retinoid use, patients with skin of color may develop irritant contact dermatitis, thus leading to PIH.13 Dryness and irritant contact dermatitis also can be seen with the use of benzoyl peroxide in black patients. Because of these effects, gentle moisturizers are recommended, and both benzoyl peroxide and retinoids should be initiated at lower doses in patients with skin of color.13

For patients with severe nodulocystic acne, isotretinoin is the treatment of choice in patients with skin of color,13 but there is a dearth of clinical studies addressing complications seen in black adolescents on this treatment, especially with respect to those with AD. Of note, systemic antibiotics typically are initiated before isotretinoin; however, this strategy is falling out of favor due to concern for antibiotic resistance with long-term use.19

Impact of Athletics on AD in Black Adolescents

Because of the exacerbating effects of perspiration and heat causing itch and irritation in patients with AD, it is frequently advised that pediatric patients limit their participation in athletics because of the exacerbating effects of strenuous physical exercise on their disease.12 In one study, 429 pediatric patients or their parents/guardians completed QOL questionnaires; 89% of patients 15 years and younger with severe AD reported that their disease was impacted by athletics and outdoor activities, and 86% of these pediatric patients with severe AD responded that their social lives and leisure activities were impacted.20 Because adolescents often are involved in athletics or have mandatory physical education classes, AD may be isolating and may have a severe impact on self-esteem.

Aggressive treatment of AD with topical and systemic medications may be helpful in adolescents who may be reluctant to participate in sports because of teasing, bullying, or worsening of symptoms with heat or sweating.21 Now that dupilumab is available for adolescents, there is a chance that patients with severe and/or recalcitrant disease managed on this medication can achieve better control of their symptoms without the laboratory requirement of methotrexate and the difficulties of topical medication application, allowing them to engage in mandatory athletic classes as well as desired organized sports.

 

 

Use of Cosmetics for AD

Many adolescents experiment with cosmetics, and those with AD may use cosmetic products to cover hyperpigmented or hypopigmented lesions.18 In patients with active AD or increased sensitivity to allergens in cosmetic products, use of makeup can be a contributing factor for AD flares. Acne associated with cosmetics is especially important to consider in darker-skinned patients who may use makeup that is opaque and contains oil to conceal acne or PIH.

Allergens can be present in both cosmetics and pharmaceutical topical agents, and a Brazilian study found that approximately 89% of 813 prescription and nonprescription products (eg, topical drugs, sunscreens, moisturizers, soaps, cleansing lotions, shampoos, cosmeceuticals) contained allergens.22 Patients with AD have a higher prevalence of contact sensitization to fragrances, including balsam of Peru.23 Some AD treatments that contain fragrances have caused further skin issues in a few patients. In one case series, 3 pediatric patients developed allergic contact dermatitis to Myroxylon pereirae (balsam of Peru) when using topical treatments for their AD, and their symptoms of scalp inflammation and alopecia resolved with discontinuation.23

In a Dutch study, sensitization to Fragrance Mix I and M pereirae as well as other ingredients (eg, lanolin alcohol, Amerchol™ L 101 [a lanolin product]) was notably more common in pediatric patients with AD than in patients without AD; however, no data on patients with skin of color were included in this study.24



Because of the increased risk of sensitization to fragrances and other ingredients in patients with AD as well as the high percentage of allergens in prescription and nonprescription products, it is important to discuss all personal care products that patients may be using, not just their cosmetic products. Also, patch testing may be helpful in determining true allergens in some patients. Patch testing is recommended for patients with treatment-resistant AD, and a recent study suggested it should be done prior to long-term use of immunosuppressive agents.25 Increased steroid phobia and a push toward alternative medicines are leading both patients with AD and guardians of children with AD to look for other forms of moisturization, such as olive oil, coconut oil, sunflower seed oil, and shea butter, to decrease transepidermal water loss.26,27 An important factor in AD treatment efficacy is patient acceptability in using what is recommended.27 One study showed there was no difference in efficacy or acceptability in using a cream containing shea butter extract vs the ceramide-precursor product.27 Current data show olive oil may exacerbate dry skin and AD,26 and recommendation of any over-the-counter oils and butters in patients with AD should be made with great caution, as many of these products contain fragrances and other potential allergens.

Alternative Therapies for AD

Patients with AD often seek alternative or integrative treatment options, including dietary modifications and holistic remedies. Studies investigating the role of vitamins and supplements in treating AD are limited by sample size.28 However, there is some evidence that may support supplementation with vitamins D and E in addressing AD symptoms. The use of probiotics in treating AD is controversial, but there are studies suggesting that the use of probiotics may prove beneficial in preventing infantile AD.28 Additionally, findings from an ex vivo and in vitro study show that some conditions, including AD and acne, may benefit from the same probiotics, despite the differences in these two diseases. Both AD and acne have inflammatory and skin dysbiosis characteristics, which may be the common thread leading to both conditions potentially responding to treatment with probiotics.29

 

 

Preliminary evidence indicates that supplements containing fatty acids such as docosahexaenoic acid, sea buckthorn oil, and hemp seed oil may decrease the severity of AD.28 In a 20-week, randomized, single-blind, crossover study published in 2005, dietary hemp seed oil showed an improvement of clinical symptoms, including dry skin and itchiness, in patients with AD.30



In light of recent legalization in several states, patients may turn to use of cannabinoid products to manage AD. In a systematic review, cannabinoid use was reportedly a therapeutic option in the treatment of AD and AV; however, the data are based on preclinical work, and there are no randomized, placebo-controlled studies to support the use of cannabinoids.31 Furthermore, there is great concern that use of these products in adolescents is an even larger unknown.

Final Thoughts

Eighty percent of children diagnosed with AD experience symptom improvement before their early teens32; for those with AD during their preteen and teenage years, there can be psychological ramifications, as teenagers with AD report having fewer friends, are less socially involved, participate in fewer sports, and are absent from classes more often than their peers.5 In black patients with AD, school absences are even more common.6 Given the social and emotional impact of AD on patients with skin of color, it is imperative to treat the condition appropriately.33 There are areas of opportunity for further research on alternate dosing of existing treatments for AV in patients with AD, further recommendations for adolescent athletes with AD, and which cosmetic and alternative medicine products may be beneficial for this population to improve their QOL.

Providers should discuss medical management in a broader context considering patients’ extracurricular activities, treatment vehicle preferences, expectations, and personal care habits. It also is important to address the many possible factors that may influence treatment adherence early on, particularly in adolescents, as these could be barriers to treatment. This article highlights considerations for treating AD and comorbid conditions that may further complicate treatment in adolescent patients with skin of color. The information provided should serve as a guide in initial counseling and management of AD in adolescents with skin of color.

Data are limited on the management of atopic dermatitis (AD) in adolescents, particularly in patients with skin of color, making it important to identify factors that may improve AD management in this population. Comorbid conditions (eg, acne, postinflammatory hyperpigmentation [PIH]), extracurricular activities (eg, athletics), and experimentation with cosmetics in adolescents, all of which can undermine treatment efficacy and medication adherence, make it particularly challenging to devise a therapeutic regimen in this patient population. We review the management of AD in black adolescents, with special consideration of concomitant treatment of acne vulgaris (AV) as well as lifestyle and social choices (Table).

Prevalence and Epidemiology

Atopic dermatitis affects 13% to 25% of children and 2% to 10% of adults.1,2 Population‐based studies in the United States show a higher prevalence of AD in black children (19.3%) compared to European American (EA) children (16.1%).3,4

AD in Black Adolescents

Atopic dermatitis is a common skin condition that is defined as a chronic, pruritic, inflammatory dermatosis with recurrent scaling, papules, and plaques (Figure) that usually develop during infancy and early childhood.3 Although AD severity improves for some patients in adolescence, it can be a lifelong issue affecting performance in academic and occupational settings.5 One US study of 8015 children found that there are racial and ethnic disparities in school absences among children (age range, 2–17 years) with AD, with children with skin of color being absent more often than white children.6 The same study noted that black children had a 1.5-fold higher chance of being absent 6 days over a 6-month school period compared to white children. It is postulated that AD has a greater impact on quality of life (QOL) in children with skin of color, resulting in the increased number of school absences in this population.6

Atopic dermatitis on the neck with lichenification and excoriations.

The origin of AD currently is thought to be complex and can involve skin barrier dysfunction, environmental factors, microbiome effects, genetic predisposition, and immune dysregulation.1,4 Atopic dermatitis is a heterogeneous disease with variations in the prevalence, genetic background, and immune activation patterns across racial groups.4 It is now understood to be an immune-mediated disease with multiple inflammatory pathways, with type 2–associated inflammation being a primary pathway. Patients with AD have strong helper T cell (TH2) activation, and black patients with AD have higher IgE serum levels as well as absent TH17/TH1 activation.4



Atopic dermatitis currently is seen as a defect of the epidermal barrier, with variable clinical manifestations and expressivity.7 Filaggrin is an epidermal barrier protein, encoded by the FLG gene, and plays a major role in barrier function by regulating pH and promoting hydration of the skin.4 Loss of function of the FLG gene is the most well-studied genetic risk factor for developing AD, and this mutation is seen in patients with more severe and persistent AD in addition to patients with more skin infections and allergic sensitizations.3,4 However, in the skin of color population, FLG mutations are 6 times less common than in the EA population, despite the fact that AD is more prevalent in patients of African descent.4 Therefore, the role of the FLG loss-of-function mutation and AD is not as well defined in black patients, and some researchers have found no association.3 The FLG loss-of-function mutation seems to play a smaller role in black patients than in EA patients, and other genes may be involved in skin barrier dysfunction.3,4 In a small study of patients with mild AD compared to nonaffected patients, those with AD had lower total ceramide levels in the stratum corneum of affected sites than normal skin sites in healthy individuals.8

Particular disturbances in the gut microbiome have the possibility of impacting the development of AD.9 Additionally, the development of AD may be influenced by the skin microbiome, which can change depending on body site, with fungal organisms thought to make up a large proportion of the microbiome of patients with AD. In patients with AD, there is a lack of microbial diversity and an overgrowth of Staphylococcus aureus.9

 

 

Diagnosis

Clinicians diagnose AD based on clinical characteristics, and the lack of objective criteria can hinder diagnosis.1 Thus, diagnosing AD in children with dark skin can pose a particular challenge given the varied clinical presentation of AD across skin types. Severe cases of AD may not be diagnosed or treated adequately in deeply pigmented children because erythema, a defining characteristic of AD, may be hard to identify in darker skin types.10 Furthermore, clinical erythema scores among black children may be “strongly” underestimated using scoring systems such as Eczema Area and Severity Index and SCORing Atopic Dermatitis.4 It is estimated that the risk for severe AD may be 6 times higher in black children compared to white children.10 Additionally, patients with skin of color can present with more treatment-resistant AD.4

Treatment of AD

Current treatment is focused on restoring epidermal barrier function, often with topical agents, such as moisturizers containing different amounts of emollients, occlusives, and humectants; corticosteroids; calcineurin inhibitors; and antimicrobials. Emollients such as glycol stearate, glyceryl stearate, and soy sterols function as lubricants, softening the skin. Occlusive agents include petrolatum, dimethicone, and mineral oil; they act by forming a layer to slow evaporation of water. Humectants including glycerol, lactic acid, and urea function by promoting water retention.11 For acute flares, mid- to high-potency topical corticosteroids are recommended. Also, topical calcineurin inhibitors such as tacrolimus and pimecrolimus may be used alone or in combination with topical steroids. Finally, bleach baths and topical mupirocin applied to the nares also have proved helpful in moderate to severe AD with secondary bacterial infections.11 Phototherapy can be used in adult and pediatric patients with acute and chronic AD if traditional treatments have failed.2

Systemic agents are indicated and recommended for the subset of adult and pediatric patients in whom optimized topical regimens and/or phototherapy do not adequately provide disease control or when QOL is substantially impacted. The systemic agents effective in the pediatric population include cyclosporine, azathioprine, mycophenolate mofetil, and possibly methotrexate.11 Dupilumab recently was approved by the US Food and Drug Administration for patients 12 years and older with moderate to severe AD whose disease is not well controlled with topical medications.

Patients with AD are predisposed to secondary bacterial and viral infections because of their dysfunctional skin barrier; these infections most commonly are caused by S aureus and herpes simplex virus, respectively.2 Systemic antibiotics are only recommended for patients with AD when there is clinical evidence of bacterial infection. In patients with evidence of eczema herpeticum, systemic antiviral agents should be used to treat the underlying herpes simplex virus infection.2 Atopic dermatitis typically has been studied in white patients; however, patients with skin of color have higher frequencies of treatment-resistant AD. Further research on treatment efficacy for AD in this patient population is needed, as data are limited.4

Treatment of AV in Patients With AD

Two of the most prevalent skin diseases affecting the pediatric population are AD and AV, and both can remarkably impact QOL.12 Acne is one of the most common reasons for adolescent patients to seek dermatologic care, including patients with skin of color (Fitzpatrick skin types IV to VI).13 Thus, it is to be expected that many black adolescents with AD also will have AV. For mild to moderate acne in patients with skin of color, topical retinoids and benzoyl peroxide typically are first line.13 These medications can be problematic for patients with AD, as retinoids and many other acne treatments can cause dryness, which may exacerbate AD.

 

 

Moisturizers containing ceramide can be a helpful adjunctive therapy in treating acne,14 especially in patients with AD. Modifications to application of acne medications, such as using topical retinoids every other night or mixing them with moisturizers to minimize dryness, may be beneficial to these patients. Dapsone gel 7.5% used daily also may be an option for adolescents with AD and AV. A double-blind, vehicle-controlled study demonstrated that dapsone is safe and effective for patients 12 years and older with moderate acne, and patients with Fitzpatrick skin types IV to VI rated local scaling, erythema, dryness, and stinging/burning as “none” in the study.15 Another potentially helpful topical agent in patients with AD and AV is sulfacetamide, as it is not likely to cause dryness of the skin. In a small study, sodium sulfacetamide 10% and sulfur 5% in an emollient foam vehicle showed no residual film or sulfur smell and resulted in acne reduction of 50%.16



Patients with skin of color often experience PIH in AD and acne or hypopigmentation from inflammatory dermatoses including AD.17,18 In addition to the dryness from AD and topical retinoid use, patients with skin of color may develop irritant contact dermatitis, thus leading to PIH.13 Dryness and irritant contact dermatitis also can be seen with the use of benzoyl peroxide in black patients. Because of these effects, gentle moisturizers are recommended, and both benzoyl peroxide and retinoids should be initiated at lower doses in patients with skin of color.13

For patients with severe nodulocystic acne, isotretinoin is the treatment of choice in patients with skin of color,13 but there is a dearth of clinical studies addressing complications seen in black adolescents on this treatment, especially with respect to those with AD. Of note, systemic antibiotics typically are initiated before isotretinoin; however, this strategy is falling out of favor due to concern for antibiotic resistance with long-term use.19

Impact of Athletics on AD in Black Adolescents

Because of the exacerbating effects of perspiration and heat causing itch and irritation in patients with AD, it is frequently advised that pediatric patients limit their participation in athletics because of the exacerbating effects of strenuous physical exercise on their disease.12 In one study, 429 pediatric patients or their parents/guardians completed QOL questionnaires; 89% of patients 15 years and younger with severe AD reported that their disease was impacted by athletics and outdoor activities, and 86% of these pediatric patients with severe AD responded that their social lives and leisure activities were impacted.20 Because adolescents often are involved in athletics or have mandatory physical education classes, AD may be isolating and may have a severe impact on self-esteem.

Aggressive treatment of AD with topical and systemic medications may be helpful in adolescents who may be reluctant to participate in sports because of teasing, bullying, or worsening of symptoms with heat or sweating.21 Now that dupilumab is available for adolescents, there is a chance that patients with severe and/or recalcitrant disease managed on this medication can achieve better control of their symptoms without the laboratory requirement of methotrexate and the difficulties of topical medication application, allowing them to engage in mandatory athletic classes as well as desired organized sports.

 

 

Use of Cosmetics for AD

Many adolescents experiment with cosmetics, and those with AD may use cosmetic products to cover hyperpigmented or hypopigmented lesions.18 In patients with active AD or increased sensitivity to allergens in cosmetic products, use of makeup can be a contributing factor for AD flares. Acne associated with cosmetics is especially important to consider in darker-skinned patients who may use makeup that is opaque and contains oil to conceal acne or PIH.

Allergens can be present in both cosmetics and pharmaceutical topical agents, and a Brazilian study found that approximately 89% of 813 prescription and nonprescription products (eg, topical drugs, sunscreens, moisturizers, soaps, cleansing lotions, shampoos, cosmeceuticals) contained allergens.22 Patients with AD have a higher prevalence of contact sensitization to fragrances, including balsam of Peru.23 Some AD treatments that contain fragrances have caused further skin issues in a few patients. In one case series, 3 pediatric patients developed allergic contact dermatitis to Myroxylon pereirae (balsam of Peru) when using topical treatments for their AD, and their symptoms of scalp inflammation and alopecia resolved with discontinuation.23

In a Dutch study, sensitization to Fragrance Mix I and M pereirae as well as other ingredients (eg, lanolin alcohol, Amerchol™ L 101 [a lanolin product]) was notably more common in pediatric patients with AD than in patients without AD; however, no data on patients with skin of color were included in this study.24



Because of the increased risk of sensitization to fragrances and other ingredients in patients with AD as well as the high percentage of allergens in prescription and nonprescription products, it is important to discuss all personal care products that patients may be using, not just their cosmetic products. Also, patch testing may be helpful in determining true allergens in some patients. Patch testing is recommended for patients with treatment-resistant AD, and a recent study suggested it should be done prior to long-term use of immunosuppressive agents.25 Increased steroid phobia and a push toward alternative medicines are leading both patients with AD and guardians of children with AD to look for other forms of moisturization, such as olive oil, coconut oil, sunflower seed oil, and shea butter, to decrease transepidermal water loss.26,27 An important factor in AD treatment efficacy is patient acceptability in using what is recommended.27 One study showed there was no difference in efficacy or acceptability in using a cream containing shea butter extract vs the ceramide-precursor product.27 Current data show olive oil may exacerbate dry skin and AD,26 and recommendation of any over-the-counter oils and butters in patients with AD should be made with great caution, as many of these products contain fragrances and other potential allergens.

Alternative Therapies for AD

Patients with AD often seek alternative or integrative treatment options, including dietary modifications and holistic remedies. Studies investigating the role of vitamins and supplements in treating AD are limited by sample size.28 However, there is some evidence that may support supplementation with vitamins D and E in addressing AD symptoms. The use of probiotics in treating AD is controversial, but there are studies suggesting that the use of probiotics may prove beneficial in preventing infantile AD.28 Additionally, findings from an ex vivo and in vitro study show that some conditions, including AD and acne, may benefit from the same probiotics, despite the differences in these two diseases. Both AD and acne have inflammatory and skin dysbiosis characteristics, which may be the common thread leading to both conditions potentially responding to treatment with probiotics.29

 

 

Preliminary evidence indicates that supplements containing fatty acids such as docosahexaenoic acid, sea buckthorn oil, and hemp seed oil may decrease the severity of AD.28 In a 20-week, randomized, single-blind, crossover study published in 2005, dietary hemp seed oil showed an improvement of clinical symptoms, including dry skin and itchiness, in patients with AD.30



In light of recent legalization in several states, patients may turn to use of cannabinoid products to manage AD. In a systematic review, cannabinoid use was reportedly a therapeutic option in the treatment of AD and AV; however, the data are based on preclinical work, and there are no randomized, placebo-controlled studies to support the use of cannabinoids.31 Furthermore, there is great concern that use of these products in adolescents is an even larger unknown.

Final Thoughts

Eighty percent of children diagnosed with AD experience symptom improvement before their early teens32; for those with AD during their preteen and teenage years, there can be psychological ramifications, as teenagers with AD report having fewer friends, are less socially involved, participate in fewer sports, and are absent from classes more often than their peers.5 In black patients with AD, school absences are even more common.6 Given the social and emotional impact of AD on patients with skin of color, it is imperative to treat the condition appropriately.33 There are areas of opportunity for further research on alternate dosing of existing treatments for AV in patients with AD, further recommendations for adolescent athletes with AD, and which cosmetic and alternative medicine products may be beneficial for this population to improve their QOL.

Providers should discuss medical management in a broader context considering patients’ extracurricular activities, treatment vehicle preferences, expectations, and personal care habits. It also is important to address the many possible factors that may influence treatment adherence early on, particularly in adolescents, as these could be barriers to treatment. This article highlights considerations for treating AD and comorbid conditions that may further complicate treatment in adolescent patients with skin of color. The information provided should serve as a guide in initial counseling and management of AD in adolescents with skin of color.

References
  1. Feldman SR, Cox LS, Strowd LC, et al. The challenge of managing atopic dermatitis in the United States. Am Health Drug Benefits. 2019;12:83-93.
  2. Sidbury R, Davis DM, Cohen DE, et al. Guidelines of care for the management of atopic dermatitis: section 3. management and treatment with phototherapy and systemic agents. J Am Acad Dermatol. 2014;71:327-349.
  3. Kaufman BP, Guttman-Yassky E, Alexis AF. Atopic dermatitis in diverse racial and ethnic groups—variations in epidemiology, genetics, clinical presentation and treatment. Exp Dermatol. 2018;27:340-357.
  4. Brunner PM, Guttman-Yassky E. Racial differences in atopic dermatitis. Ann Allergy Asthma Immunol. 2019;122:449-455.
  5. Vivar KL, Kruse L. The impact of pediatric skin disease on self-esteem. Int J Womens Dermatol. 2018;4:27-31.
  6. Wan J, Margolis DJ, Mitra N, et al. Racial and ethnic differences in atopic dermatitis–related school absences among US children [published online May 22, 2019]. JAMA Dermatol. doi:10.1001/jamadermatol.2019.0597.
  7. Weidinger S, Novak N. Atopic dermatitis. Lancet. 2016;387:1109-1122.
  8. Ishikawa J, Narita H, Kondo N, et al. Changes in the ceramide profile of atopic dermatitis patients. J Invest Dermatol. 2010;130:2511-2514.
  9. Chernikova D, Yuan I, Shaker M. Prevention of allergy with diverse and healthy microbiota: an update. Curr Opin Pediatr. 2019;31:418-425.
  10. Ben-Gashir MA, Hay RJ. Reliance on erythema scores may mask severe atopic dermatitis in black children compared with their white counterparts. Br J Dermatol. 2002;147:920-925.
  11. Eichenfield LF, Tom WL, Berger TG, et al. Guidelines of care for the management of atopic dermatitis: section 2. management and treatment of atopic dermatitis with topical therapies. J Am Acad Dermatol. 2014;71:116-132.
  12. Nguyen CM, Koo J, Cordoro KM. Psychodermatologic effects of atopic dermatitis and acne: a review on self-esteem and identity. Pediatr Dermatol. 2016;33:129-135.
  13. Davis EC, Callender VD. A review of acne in ethnic skin: pathogenesis, clinical manifestations, and management strategies. J Clin Aesthet Dermatol. 2010;3:24-38.
  14. Lynde CW, Andriessen A, Barankin B, et al. Moisturizers and ceramide-containing moisturizers may offer concomitant therapy with benefits. J Clin Aesthet Dermatol. 2014;7:18-26.
  15. Taylor SC, Cook-Bolden FE, McMichael A, et al. Efficacy, safety, and tolerability of topical dapsone gel, 7.5% for treatment of acne vulgaris by Fitzpatrick skin phototype. J Drugs Dermatol. 2018;17:160-167.
  16. Draelos ZD. The multifunctionality of 10% sodium sulfacetamide, 5% sulfur emollient foam in the treatment of inflammatory facial dermatoses. J Drugs Dermatol. 2010;9:234-236.
  17. Vachiramon V, Tey HL, Thompson AE, et al. Atopic dermatitis in African American children: addressing unmet needs of a common disease. Pediatr Dermatol. 2012;29:395-402.
  18. Heath CR. Managing postinflammatory hyperpigmentation in pediatric patients with skin of color. Cutis. 2018;102:71-73.
  19. Nagler AR, Milam EC, Orlow SJ. The use of oral antibiotics before isotretinoin therapy in patients with acne. J Am Acad Dermatol. 2016;74:273-279.
  20. Paller AS, McAlister RO, Doyle JJ, et al. Perceptions of physicians and pediatric patients about atopic dermatitis, its impact, and its treatment. Clin Pediatr. 2002;41:323-332.
  21. Sibbald C, Drucker AM. Patient burden of atopic dermatitis. Dermatol Clin. 2017;35:303-316.
  22. Rocha VB, Machado CJ, Bittencourt FV. Presence of allergens in the vehicles of Brazilian dermatological products. Contact Dermatitis. 2017;76:126-128.
  23. Admani S, Goldenberg A, Jacob SE. Contact alopecia: improvement of alopecia with discontinuation of fluocinolone oil in individuals allergic to balsam fragrance. Pediatr Dermatol. 2017;34:e57-e60.
  24. Uter W, Werfel T, White IR, et al. Contact allergy: a review of current problems from a clinical perspective. Int J Environ Res Public Health. 2018;15:E1108.
  25. López-Jiménez EC, Marrero-Alemán G, Borrego L. One-third of patients with therapy-resistant atopic dermatitis may benefit after patch testing [published online May 13, 2019]. J Eur Acad Dermatol Venereol. doi:10.1111/jdv.15672.
  26. Karagounis TK, Gittler JK, Rotemberg V, et al. Use of “natural” oils for moisturization: review of olive, coconut, and sunflower seed oil. Pediatr Dermatol. 2019;36:9-15.
  27. Hon KL, Tsang YC, Pong NH, et al. Patient acceptability, efficacy, and skin biophysiology of a cream and cleanser containing lipid complex with shea butter extract versus a ceramide product for eczema. Hong Kong Med J. 2015;21:417-425.
  28. Reynolds KA, Juhasz MLW, Mesinkovska NA. The role of oral vitamins and supplements in the management of atopic dermatitis: a systematic review [published online March 20, 2019]. Int J Dermatol. doi:10.1111/ijd.14404.
  29. Mottin VHM, Suyenaga ES. An approach on the potential use of probiotics in the treatment of skin conditions: acne and atopic dermatitis. Int J Dermatol. 2018;57:1425-1432.
  30. Callaway J, Schwab U, Harvima I, et al. Efficacy of dietary hempseed oil in patients with atopic dermatitis. J Dermatol Treat. 2005;16:87-94.
  31. Eagleston LRM, Kalani NK, Patel RR, et al. Cannabinoids in dermatology: a scoping review [published June 15, 2018]. Dermatol Online J. 2018;24.
  32. Kim JP, Chao LX, Simpson EL, et al. Persistence of atopic dermatitis (AD): a systematic review and meta-analysis. J Am Acad Dermatol. 2016;75:681-687.e611.
  33. de María Díaz Granados L, Quijano MA, Ramírez PA, et al. Quality assessment of atopic dermatitis clinical practice guidelines in ≤ 18 years. Arch Dermatol Res. 2018;310:29-37.
References
  1. Feldman SR, Cox LS, Strowd LC, et al. The challenge of managing atopic dermatitis in the United States. Am Health Drug Benefits. 2019;12:83-93.
  2. Sidbury R, Davis DM, Cohen DE, et al. Guidelines of care for the management of atopic dermatitis: section 3. management and treatment with phototherapy and systemic agents. J Am Acad Dermatol. 2014;71:327-349.
  3. Kaufman BP, Guttman-Yassky E, Alexis AF. Atopic dermatitis in diverse racial and ethnic groups—variations in epidemiology, genetics, clinical presentation and treatment. Exp Dermatol. 2018;27:340-357.
  4. Brunner PM, Guttman-Yassky E. Racial differences in atopic dermatitis. Ann Allergy Asthma Immunol. 2019;122:449-455.
  5. Vivar KL, Kruse L. The impact of pediatric skin disease on self-esteem. Int J Womens Dermatol. 2018;4:27-31.
  6. Wan J, Margolis DJ, Mitra N, et al. Racial and ethnic differences in atopic dermatitis–related school absences among US children [published online May 22, 2019]. JAMA Dermatol. doi:10.1001/jamadermatol.2019.0597.
  7. Weidinger S, Novak N. Atopic dermatitis. Lancet. 2016;387:1109-1122.
  8. Ishikawa J, Narita H, Kondo N, et al. Changes in the ceramide profile of atopic dermatitis patients. J Invest Dermatol. 2010;130:2511-2514.
  9. Chernikova D, Yuan I, Shaker M. Prevention of allergy with diverse and healthy microbiota: an update. Curr Opin Pediatr. 2019;31:418-425.
  10. Ben-Gashir MA, Hay RJ. Reliance on erythema scores may mask severe atopic dermatitis in black children compared with their white counterparts. Br J Dermatol. 2002;147:920-925.
  11. Eichenfield LF, Tom WL, Berger TG, et al. Guidelines of care for the management of atopic dermatitis: section 2. management and treatment of atopic dermatitis with topical therapies. J Am Acad Dermatol. 2014;71:116-132.
  12. Nguyen CM, Koo J, Cordoro KM. Psychodermatologic effects of atopic dermatitis and acne: a review on self-esteem and identity. Pediatr Dermatol. 2016;33:129-135.
  13. Davis EC, Callender VD. A review of acne in ethnic skin: pathogenesis, clinical manifestations, and management strategies. J Clin Aesthet Dermatol. 2010;3:24-38.
  14. Lynde CW, Andriessen A, Barankin B, et al. Moisturizers and ceramide-containing moisturizers may offer concomitant therapy with benefits. J Clin Aesthet Dermatol. 2014;7:18-26.
  15. Taylor SC, Cook-Bolden FE, McMichael A, et al. Efficacy, safety, and tolerability of topical dapsone gel, 7.5% for treatment of acne vulgaris by Fitzpatrick skin phototype. J Drugs Dermatol. 2018;17:160-167.
  16. Draelos ZD. The multifunctionality of 10% sodium sulfacetamide, 5% sulfur emollient foam in the treatment of inflammatory facial dermatoses. J Drugs Dermatol. 2010;9:234-236.
  17. Vachiramon V, Tey HL, Thompson AE, et al. Atopic dermatitis in African American children: addressing unmet needs of a common disease. Pediatr Dermatol. 2012;29:395-402.
  18. Heath CR. Managing postinflammatory hyperpigmentation in pediatric patients with skin of color. Cutis. 2018;102:71-73.
  19. Nagler AR, Milam EC, Orlow SJ. The use of oral antibiotics before isotretinoin therapy in patients with acne. J Am Acad Dermatol. 2016;74:273-279.
  20. Paller AS, McAlister RO, Doyle JJ, et al. Perceptions of physicians and pediatric patients about atopic dermatitis, its impact, and its treatment. Clin Pediatr. 2002;41:323-332.
  21. Sibbald C, Drucker AM. Patient burden of atopic dermatitis. Dermatol Clin. 2017;35:303-316.
  22. Rocha VB, Machado CJ, Bittencourt FV. Presence of allergens in the vehicles of Brazilian dermatological products. Contact Dermatitis. 2017;76:126-128.
  23. Admani S, Goldenberg A, Jacob SE. Contact alopecia: improvement of alopecia with discontinuation of fluocinolone oil in individuals allergic to balsam fragrance. Pediatr Dermatol. 2017;34:e57-e60.
  24. Uter W, Werfel T, White IR, et al. Contact allergy: a review of current problems from a clinical perspective. Int J Environ Res Public Health. 2018;15:E1108.
  25. López-Jiménez EC, Marrero-Alemán G, Borrego L. One-third of patients with therapy-resistant atopic dermatitis may benefit after patch testing [published online May 13, 2019]. J Eur Acad Dermatol Venereol. doi:10.1111/jdv.15672.
  26. Karagounis TK, Gittler JK, Rotemberg V, et al. Use of “natural” oils for moisturization: review of olive, coconut, and sunflower seed oil. Pediatr Dermatol. 2019;36:9-15.
  27. Hon KL, Tsang YC, Pong NH, et al. Patient acceptability, efficacy, and skin biophysiology of a cream and cleanser containing lipid complex with shea butter extract versus a ceramide product for eczema. Hong Kong Med J. 2015;21:417-425.
  28. Reynolds KA, Juhasz MLW, Mesinkovska NA. The role of oral vitamins and supplements in the management of atopic dermatitis: a systematic review [published online March 20, 2019]. Int J Dermatol. doi:10.1111/ijd.14404.
  29. Mottin VHM, Suyenaga ES. An approach on the potential use of probiotics in the treatment of skin conditions: acne and atopic dermatitis. Int J Dermatol. 2018;57:1425-1432.
  30. Callaway J, Schwab U, Harvima I, et al. Efficacy of dietary hempseed oil in patients with atopic dermatitis. J Dermatol Treat. 2005;16:87-94.
  31. Eagleston LRM, Kalani NK, Patel RR, et al. Cannabinoids in dermatology: a scoping review [published June 15, 2018]. Dermatol Online J. 2018;24.
  32. Kim JP, Chao LX, Simpson EL, et al. Persistence of atopic dermatitis (AD): a systematic review and meta-analysis. J Am Acad Dermatol. 2016;75:681-687.e611.
  33. de María Díaz Granados L, Quijano MA, Ramírez PA, et al. Quality assessment of atopic dermatitis clinical practice guidelines in ≤ 18 years. Arch Dermatol Res. 2018;310:29-37.
Issue
Cutis - 104(3)
Issue
Cutis - 104(3)
Page Number
164-168
Page Number
164-168
Publications
Publications
Topics
Article Type
Display Headline
Atopic Dermatitis in Adolescents With Skin of Color
Display Headline
Atopic Dermatitis in Adolescents With Skin of Color
Sections
Inside the Article

Practice Points

  • Atopic dermatitis (AD) can be a lifelong issue that affects academic and occupational performance, with higher rates of absenteeism seen in black patients.
  • The FLG loss-of-function mutation seems to play a smaller role in black patients, and other genes may be involved in skin barrier dysfunction, which could be why there is a higher rate of skin of color patients with treatment-resistant AD.
  • Diagnosing AD in skin of color patients can pose a particular challenge, and severe cases of AD may not be diagnosed or treated adequately in deeply pigmented children because erythema, a defining characteristic of AD, may be hard to identify in darker skin tones.
  • There are several areas of opportunity for further research to better treat AD in this patient population and improve quality of life.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Establishing the Diagnosis of Rosacea in Skin of Color Patients

Article Type
Changed
Thu, 10/29/2020 - 14:49
Display Headline
Establishing the Diagnosis of Rosacea in Skin of Color Patients
In Collaboration With the Skin of Color Society

Rosacea is a chronic inflammatory cutaneous disorder that affects the vasculature and pilosebaceous units of the face. Delayed and misdiagnosed rosacea in the SOC population has led to increased morbidity in this patient population. 1-3 It is characterized by facial flushing and warmth, erythema, telangiectasia, papules, and pustules. The 4 major subtypes include erythematotelangiectatic, papulopustular, phymatous, and ocular rosacea. 4 Granulomatous rosacea is considered to be a unique variant of rosacea. Until recently, rosacea was thought to predominately affect lighter-skinned individuals of Celtic and northern European origin. 5,6 A paucity of studies and case reports in the literature have contributed to the commonly held belief that rosacea occurs infrequently in patients with skin of color (SOC). 1 A PubMed search of articles indexed for MEDLINE revealed 32 results using the terms skin of color and rosacea vs 3786 using the term rosacea alone. It is possible that the nuance involved in appreciating erythema or other clinical manifestations of rosacea in SOC patients has led to underdiagnosis. Alternatively, these patients may be unaware that their symptoms represent a disease process and do not seek treatment. Many patients with darker skin will have endured rosacea for months or even years because the disease has been unrecognized or misdiagnosed. 6-8 Another factor possibly accounting for the perception that rosacea occurs infrequently in patients with SOC is misdiagnosis of rosacea as other diseases that are known to occur more commonly in the SOC population. Dermatologists should be aware that rosacea can affect SOC patients and that there are several rosacea mimickers to be considered and excluded when making the rosacea diagnosis in this patient population. To promote accurate and timely diagnosis of rosacea, we review several possible rosacea mimickers in SOC patients and highlight the distinguishing features.

Epidemiology

In 2018, a meta-analysis of published studies on rosacea estimated the global prevalence in all adults to be 5.46%.9 A multicenter study across 6 cities in Colombia identified 291 outpatients with rosacea; of them, 12.4% had either Fitzpatrick skin types IV or V.10 A study of 2743 Angolan adults with Fitzpatrick skin types V and VI reported that only 0.4% of patients had a diagnosis of rosacea.11 A Saudi study of 50 dark-skinned female patients with rosacea revealed 40% (20/50), 18% (9/50), and 42% (21/50) were Fitzpatrick skin types IV, V, and VI, respectively.12 The prevalence of rosacea in SOC patients in the United States is less defined. Data from the US National Ambulatory Medical Care Survey (1993-2010) of 31.5 million rosacea visits showed that 2% of rosacea patients were black, 2.3% were Asian or Pacific Islander, and 3.9% were Hispanic or Latino.8

Clinical Features

Each of the 4 major rosacea subtypes can present in the SOC population. The granulomatous variant has been predominantly reported in black patients.13 This predilection has been attributed to either an increased susceptibility in black patients to develop this variant or a delay in diagnosis of earlier phases of inflammatory rosacea.7

In a Saudi study (N=50), severe erythematotelangiectatic rosacea was diagnosed in 42% (21/50) of patients, with the majority having Fitzpatrick skin type IV. The severe papulopustular subtype was seen in 14% (7/50) of patients, with 20% (10/50) and 14% (7/50) having Fitzpatrick skin types IV and VI, respectively.12 In a Tunisian study (N=244), erythematotelangiectatic rosacea was seen in 12% of patients, papulopustular rosacea in 69%, phymatous rosacea in 4%, and ocular rosacea in 16%. Less frequently, the granulomatous variant was seen in 3% of patients, and steroid rosacea was noted in 12% patients.14

Recognizing the signs of rosacea may be a challenge, particularly erythema and telangiectasia. Tips for making an accurate diagnosis include use of adequate lighting, blanching of the skin (Figure 1), photography of the affected area against a dark blue background, and dermatoscopic examination.3 Furthermore, a thorough medical history, especially when evaluating the presence of facial erythema and identifying triggers, may help reach the correct diagnosis. Careful examination of the distribution of papules and pustules as well as the morphology and color of the papules in SOC patients also may provide diagnostic clues.

Figure 1. Diascopy revealed telangiectasia associated with rosacea on the central face of a woman with Fitzpatrick skin type IV. Photograph courtesy of Jennifer David, DO (Philadelphia, Pennsylvania).

Differential Diagnosis and Distinguishing Features

Several disorders are included in the differential diagnosis of rosacea and may confound a correct rosacea diagnosis, including systemic lupus erythematosus (SLE), seborrheic dermatitis, dermatomyositis, acne vulgaris, sarcoidosis, and steroid dermatitis. Many of these disorders also occur more commonly in patients with SOC; therefore, it is important to clearly distinguish these entities from rosacea in this population.

Systemic Lupus Erythematosus
Systemic lupus erythematosus is an autoimmune disease that commonly presents with erythema as well as erythematous inflammatory facial lesions similar to rosacea. The classic clinical appearance of SLE is the butterfly or malar rash, an erythematous macular eruption on the malar region of the face that also may involve the nose. This rash can appear similar to rosacea; however, the malar rash classically spares the nasolabial folds, while erythema of rosacea often involves this anatomic boundary. Although the facial erythema in both SLE and early stages of rosacea may be patchy and similar in presentation, the presence of papules and pustules rarely occurs in SLE and may help to differentiate SLE from certain variants of rosacea.15

 

 


Both SLE and rosacea may be exacerbated by sun exposure, and patients may report burning and stinging.16-18 Performing a complete physical examination, performing a skin biopsy with hematoxylin and eosin and direct immunofluorescence, and checking serologies including antinuclear antibody (ANA) can assist in making the diagnosis. It is important to note that elevated ANA, albeit lower than what is typically seen in SLE, has been reported in rosacea patients.19 If ANA is elevated, more specific SLE antibodies should be tested (eg, double-stranded DNA). Additionally, SLE can be differentiated on histology by a considerably lower CD4:CD8 ratio, fewer CD4+CD25+ regulatory T cells, and more CD123+ plasmacytoid dendritic cells compared to rosacea.20



Seborrheic Dermatitis
Seborrheic dermatitis is a frequent cause of facial erythema linked to the Malassezia yeast species in susceptible individuals. Seborrheic dermatitis has a notable prevalence in women of African descent and often is considered normal by these patients.21 Rosacea and seborrheic dermatitis are relatively common dermatoses and therefore can present concurrently. In both diseases, facial erythema may be difficult to discern upon cursory inspection. Seborrheic dermatitis may be distinguished from rosacea by the clinical appearance of erythematous patches and plaques involving the scalp, anterior and posterior hairlines, preauricular and postauricular areas, and medial eyebrows. Both seborrheic dermatitis and rosacea may involve the nasolabial folds, but the presence of scale in seborrheic dermatitis is a distinguishing feature. Scale may vary in appearance from thick, greasy, and yellowish to fine, thin, and whitish.22 In contrast to rosacea, the erythematous lesions of seborrheic dermatitis often are annular in configuration. Furthermore, postinflammatory hypopigmentation and, to a lesser extent, postinflammatory hyperpigmentation are key clinical components of seborrheic dermatitis in SOC patients but are not as commonly observed in rosacea.

Dermatomyositis
Dermatomyositis is a systemic autoimmune disease characterized by progressive and symmetric proximal musculoskeletal weakness and cutaneous findings. Facial erythema in the malar and nasolabial folds can be seen in patients with dermatomyositis18; however, the facial erythema seen in dermatomyositis, known as heliotrope rash, has a violaceous dusky quality and also involves the periorbital region. The violaceous hue and periorbital involvement are distinguishing features from rosacea. Okiyama et al23 described facial macular violaceous erythema with scale and edema in Japanese patients with dermatomyositis on the nasolabial folds, eyebrows, chin, cheeks, and ears; they also described mild atrophy with telangiectasia. Other clinical signs to help distinguish rosacea from dermatomyositis are the presence of edema of the face and extremities, Gottron papules, and poikiloderma. Dermatomyositis is a disease that affects all races; however, it is 4 times more common in black vs white patients,24 making it even more important to be able to distinguish between these conditions.

Acne Vulgaris
Acne vulgaris, the most commonly diagnosed dermatosis in patients with SOC, is characterized by papules, pustules, cysts, nodules, open and closed comedones, and hyperpigmented macules on the face, chest, and back.25,26 The absence of comedonal lesions and the presence of hyperpigmented macules distinguishes acne vulgaris from rosacea in this population.1 In addition, the absence of telangiectasia and flushing are important distinguishing factors when making the diagnosis of acne vulgaris.

Sarcoidosis
Sarcoidosis is a multisystem inflammatory disease characterized histologically by the presence of noncaseating granulomas in sites such as the lungs, lymph nodes, eyes, nervous system, liver, spleen, heart, and skin.27 Cutaneous sarcoidosis is known as a great mimicker of many other dermatoses, as it may present with multiple morphologic features. Cutaneous sarcoidosis most typically presents as papules, but nodules, plaques, lupus pernio, subcutaneous infiltrates, and infiltration of scars also have been identified.28 Sarcoid papules typically are 1 to 5 mm in size on the face, neck, and periorbital skin29; they are initially orange or yellow-brown in color, turn brownish red or violaceous, then involute to form faint macules.30 Papular lesions may either resolve or evolve into plaques, particularly on the extremities, face, scalp, back, and buttocks. Additionally, there are a few case reports of patients with cutaneous sarcoidosis presenting with large bulbous nasal masses initially thought to be rhinophyma.31-33 Finally, it may be difficult to distinguish sarcoidosis from granulomatous rosacea, which is characterized by firm yellow, brown, violaceous, red, or flesh-colored monomorphic papules or nodules affecting the perioral, periocular, medial, and/or lateral areas of the face (Figure 2).4,34 Patients also can have unilateral disease.35 Patients with granulomatous rosacea lack flushing and erythema as seen in more characteristic presentations of rosacea. They may report pain, pruritus, or burning, or they may be asymptomatic.36 Features that distinguish granulomatous rosacea from sarcoidosis include the absence of nodules, plaques, lupus pernio, subcutaneous infiltrates, and infiltration of scars. Clinical, histological, and radiographic evaluation are necessary to make the diagnosis of sarcoidosis over rosacea.

Figure 2. Granulomatous rosacea in a woman with Fitzpatrick skin type VI. Photograph courtesy of Susan C. Taylor, MD (Philadelphia, Pennsylvania).


Steroid Dermatitis
Steroid dermatitis involving the face may mimic rosacea. It is caused by the application of a potent corticosteroid to the facial skin for a prolonged period of time. In a report from a teaching hospital in Baghdad, the duration of application was 0.25 to 10 years on average.37 Reported characteristics of steroid dermatitis included facial erythema, telangiectasia, papules, pustules, and warmth to the touch. Distinguishing features from rosacea may be the presence of steroid dermatitis on the entire face, whereas rosacea tends to occur on the center of the face. Diagnosis of steroid dermatitis is made based on a history of chronic topical steroid use with rebound flares upon discontinuation of steroid.

Final Thoughts

Rosacea has features common to many other facial dermatoses, making the diagnosis challenging, particularly in patients with SOC. This difficulty in diagnosis may contribute to an underestimation of the prevalence of this disease in SOC patients. An understanding of rosacea, its nuances in clinical appearance, and its mimickers in SOC patients is important in making an accurate diagnosis.

References

References
  1. Alexis AF. Rosacea in patients with skin of color: uncommon but not rare. Cutis. 2010;86:60-62. 
  2. Kim NH, Yun SJ, Lee JB. Clinical features of Korean patients with rhinophyma. J Dermatol. 2017;44:710-712. 
  3. Hua TC, Chung PI, Chen YJ, et al. Cardiovascular comorbidities in patients with rosacea: a nationwide case-control study from Taiwan. J Am Acad Dermatol. 2015;73:249-254. 
  4. Wilkin J, Dahl M, Detmar M, et al. Standard classification of rosacea: report of the National Rosacea Society Expert Committee on the Classification and Staging of Rosacea. J Am Acad Dermatol. 2002;46:584-587. 
  5. Elewski BE, Draelos Z, Dreno B, et al. Global diversity and optimized outcome: proposed international consensus from the Rosacea International Expert Group. J Eur Acad Dermatol Venereol. 2011;25:188-200. 
  6. Alexis AF, Callender VD, Baldwin HE, et al. Global epidemiology and clinical spectrum of rosacea, highlighting skin of color: review and clinical practice experience [published online September 19, 2018]. J Am Acad Dermatol. 2019;80:1722-1729.e7. 
     
  7. Dlova NC, Mosam A. Rosacea in black South Africans with skin phototypes V and VI. Clin Exp Dermatol. 2017;42:670-673. 
  8. Al-Dabagh A, Davis SA, McMichael AJ, et al. Rosacea in skin of color: not a rare diagnosis [published online October 15, 2014]. Dermatol Online J. 2014;20. pii:13030/qt1mv9r0ss. 
  9. Gether L, Overgaard LK, Egeberg A, et al. Incidence and prevalence of rosacea: a systematic review and meta-analysis. Br J Dermatol. 2018;179:282-289. 
  10. Rueda LJ, Motta A, Pabon JG, et al. Epidemiology of rosacea in Colombia. Int J Dermatol. 2017;56:510-513. 
  11. De Luca DA, Maianski Z, Averbukh M. A study of skin disease spectrum occurring in Angola phototype V-VI population in Luanda. Int J Dermatol. 2018;57:849-855. 
  12. Al Balbeesi AO, Halawani MR. Unusual features of rosacea in Saudi females with dark skin. Ochsner J. 2014;14:321-327. 
  13. Rosen T, Stone MS. Acne rosacea in blacks. J Am Acad Dermatol. 1987;17:70-73. 
  14. Khaled A, Hammami H, Zeglaoui F, et al. Rosacea: 244 Tunisian cases. Tunis Med. 2010;88:597-601. 
  15. Usatine RP, Smith MA, Chumley HS, et al. The Color Atlas of Family Medicine. 2nd ed. New York, NY: The McGraw-Hill Companies; 2013.  
  16. O'Gorman SM, Murphy GM. Photoaggravated disorders. Dermatol Clin. 2014;32:385-398, ix. 
  17. Foering K, Chang AY, Piette EW, et al. Characterization of clinical photosensitivity in cutaneous lupus erythematosus. J Am Acad Dermatol. 2013;69:205-213. 
  18. Saleem MD, Wilkin JK. Evaluating and optimizing the diagnosis of erythematotelangiectatic rosacea. Dermatol Clin. 2018;36:127-134. 
  19. Black AA, McCauliffe DP, Sontheimer RD. Prevalence of acne rosacea in a rheumatic skin disease subspecialty clinic. Lupus. 1992;1:229-237. 
  20. Brown TT, Choi EY, Thomas DG, et al. Comparative analysis of rosacea and cutaneous lupus erythematosus: histopathologic features, T-cell subsets, and plasmacytoid dendritic cells. J Am Acad Dermatol. 2014;71:100-107. 
  21. Taylor SC, Barbosa V, Burgess C, et al. Hair and scalp disorders in adult and pediatric patients with skin of color. Cutis. 2017;100:31-35. 
  22. Gary G. Optimizing treatment approaches in seborrheic dermatitis. J Clin Aesthet Dermatol. 2013;6:44-49. 
  23. Okiyama N, Kohsaka H, Ueda N, et al. Seborrheic area erythema as a common skin manifestation in Japanese patients with dermatomyositis. Dermatology. 2008;217:374-377. 
  24. Taylor SC, Kyei A. Defining skin of color. In: Taylor SC, Kelly AP, Lim HW, et al, eds. Taylor and Kelly's Dermatology for Skin of Color. 2nd ed. New York, NY: McGraw-Hill; 2016:9-15. 
  25. Davis SA, Narahari S, Feldman SR, et al. Top dermatologic conditions in patients of color: an analysis of nationally representative data. J Drugs Dermatol. 2012;11:466-473. 
  26. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl understanding):S98-S106. 
  27. Wick MR. Granulomatous & histiocytic dermatitides. Semin Diagn Pathol. 2017;34:301-311. 
  28. Ball NJ, Kho GT, Martinka M. The histologic spectrum of cutaneous sarcoidosis: a study of twenty-eight cases. J Cutan Pathol. 2004;31:160-168. 
  29. Marchell RM, Judson MA. Chronic cutaneous lesions of sarcoidosis. Clin Dermatol. 2007;25:295-302. 
  30. Mahajan VK, Sharma NL, Sharma RC, et al. Cutaneous sarcoidosis: clinical profile of 23 Indian patients. Indian J Dermatol Venereol Leprol. 2007;73:16-21. 
  31. Goldenberg JD, Kotler HS, Shamsai R, et al. Sarcoidosis of the external nose mimicking rhinophyma. case report and review of the literature. Ann Otol Rhinol Laryngol. 1998;107:514-518. 
  32. Gupta-Elera G, Lam C, Chung C, et al. Violaceous plaque on the nose referred for rhinophyma surgery. Int J Dermatol. 2015;54:1011-1013. 
  33. Leonard AL. A case of sarcoidosis mimicking rhinophyma. J Drugs Dermatol. 2003;2:333-334. 
  34. Kelati A, Mernissi FZ. Granulomatous rosacea: a case report. J Med Case Rep. 2017;11:230. 
  35. Crawford GH, Pelle MT, James WD. Rosacea: I. etiology, pathogenesis, and subtype classification. J Am Acad Dermatol. 2004;51:327-341; quiz 342-324. 
  36. Reinholz M, Ruzicka T, Steinhoff M, et al. Pathogenesis and clinical presentation of rosacea as a key for a symptom-oriented therapy. J Dtsch Dermatol Ges. 2016;14(suppl 6):4-15. 
  37. Hameed AF. Steroid dermatitis resembling rosacea: a clinical evaluation of 75 patients. ISRN Dermatol. 2013;2013:491376.
Article PDF
Author and Disclosure Information

Ms. Onalaja is from the University of Rochester School of Medicine and Dentistry. Dr. Lester is from the Department of Dermatology, University of California San Francisco. Dr. Taylor is from the Department of Dermatology, Perelman School of Medicine, University of Pennsylvania, Philadelphia.

The authors report no conflict of interest.

Correspondence: Susan C. Taylor, MD, Perelman Center for Advanced Medicine, 3400 Civic Center Blvd, South Pavilion 768, Philadelphia, PA 19104 (Susan.Taylor@pennmedicine.upenn.edu).

Issue
Cutis - 104(1)
Publications
Topics
Page Number
38-41
Sections
Author and Disclosure Information

Ms. Onalaja is from the University of Rochester School of Medicine and Dentistry. Dr. Lester is from the Department of Dermatology, University of California San Francisco. Dr. Taylor is from the Department of Dermatology, Perelman School of Medicine, University of Pennsylvania, Philadelphia.

The authors report no conflict of interest.

Correspondence: Susan C. Taylor, MD, Perelman Center for Advanced Medicine, 3400 Civic Center Blvd, South Pavilion 768, Philadelphia, PA 19104 (Susan.Taylor@pennmedicine.upenn.edu).

Author and Disclosure Information

Ms. Onalaja is from the University of Rochester School of Medicine and Dentistry. Dr. Lester is from the Department of Dermatology, University of California San Francisco. Dr. Taylor is from the Department of Dermatology, Perelman School of Medicine, University of Pennsylvania, Philadelphia.

The authors report no conflict of interest.

Correspondence: Susan C. Taylor, MD, Perelman Center for Advanced Medicine, 3400 Civic Center Blvd, South Pavilion 768, Philadelphia, PA 19104 (Susan.Taylor@pennmedicine.upenn.edu).

Article PDF
Article PDF
In Collaboration With the Skin of Color Society
In Collaboration With the Skin of Color Society

Rosacea is a chronic inflammatory cutaneous disorder that affects the vasculature and pilosebaceous units of the face. Delayed and misdiagnosed rosacea in the SOC population has led to increased morbidity in this patient population. 1-3 It is characterized by facial flushing and warmth, erythema, telangiectasia, papules, and pustules. The 4 major subtypes include erythematotelangiectatic, papulopustular, phymatous, and ocular rosacea. 4 Granulomatous rosacea is considered to be a unique variant of rosacea. Until recently, rosacea was thought to predominately affect lighter-skinned individuals of Celtic and northern European origin. 5,6 A paucity of studies and case reports in the literature have contributed to the commonly held belief that rosacea occurs infrequently in patients with skin of color (SOC). 1 A PubMed search of articles indexed for MEDLINE revealed 32 results using the terms skin of color and rosacea vs 3786 using the term rosacea alone. It is possible that the nuance involved in appreciating erythema or other clinical manifestations of rosacea in SOC patients has led to underdiagnosis. Alternatively, these patients may be unaware that their symptoms represent a disease process and do not seek treatment. Many patients with darker skin will have endured rosacea for months or even years because the disease has been unrecognized or misdiagnosed. 6-8 Another factor possibly accounting for the perception that rosacea occurs infrequently in patients with SOC is misdiagnosis of rosacea as other diseases that are known to occur more commonly in the SOC population. Dermatologists should be aware that rosacea can affect SOC patients and that there are several rosacea mimickers to be considered and excluded when making the rosacea diagnosis in this patient population. To promote accurate and timely diagnosis of rosacea, we review several possible rosacea mimickers in SOC patients and highlight the distinguishing features.

Epidemiology

In 2018, a meta-analysis of published studies on rosacea estimated the global prevalence in all adults to be 5.46%.9 A multicenter study across 6 cities in Colombia identified 291 outpatients with rosacea; of them, 12.4% had either Fitzpatrick skin types IV or V.10 A study of 2743 Angolan adults with Fitzpatrick skin types V and VI reported that only 0.4% of patients had a diagnosis of rosacea.11 A Saudi study of 50 dark-skinned female patients with rosacea revealed 40% (20/50), 18% (9/50), and 42% (21/50) were Fitzpatrick skin types IV, V, and VI, respectively.12 The prevalence of rosacea in SOC patients in the United States is less defined. Data from the US National Ambulatory Medical Care Survey (1993-2010) of 31.5 million rosacea visits showed that 2% of rosacea patients were black, 2.3% were Asian or Pacific Islander, and 3.9% were Hispanic or Latino.8

Clinical Features

Each of the 4 major rosacea subtypes can present in the SOC population. The granulomatous variant has been predominantly reported in black patients.13 This predilection has been attributed to either an increased susceptibility in black patients to develop this variant or a delay in diagnosis of earlier phases of inflammatory rosacea.7

In a Saudi study (N=50), severe erythematotelangiectatic rosacea was diagnosed in 42% (21/50) of patients, with the majority having Fitzpatrick skin type IV. The severe papulopustular subtype was seen in 14% (7/50) of patients, with 20% (10/50) and 14% (7/50) having Fitzpatrick skin types IV and VI, respectively.12 In a Tunisian study (N=244), erythematotelangiectatic rosacea was seen in 12% of patients, papulopustular rosacea in 69%, phymatous rosacea in 4%, and ocular rosacea in 16%. Less frequently, the granulomatous variant was seen in 3% of patients, and steroid rosacea was noted in 12% patients.14

Recognizing the signs of rosacea may be a challenge, particularly erythema and telangiectasia. Tips for making an accurate diagnosis include use of adequate lighting, blanching of the skin (Figure 1), photography of the affected area against a dark blue background, and dermatoscopic examination.3 Furthermore, a thorough medical history, especially when evaluating the presence of facial erythema and identifying triggers, may help reach the correct diagnosis. Careful examination of the distribution of papules and pustules as well as the morphology and color of the papules in SOC patients also may provide diagnostic clues.

Figure 1. Diascopy revealed telangiectasia associated with rosacea on the central face of a woman with Fitzpatrick skin type IV. Photograph courtesy of Jennifer David, DO (Philadelphia, Pennsylvania).

Differential Diagnosis and Distinguishing Features

Several disorders are included in the differential diagnosis of rosacea and may confound a correct rosacea diagnosis, including systemic lupus erythematosus (SLE), seborrheic dermatitis, dermatomyositis, acne vulgaris, sarcoidosis, and steroid dermatitis. Many of these disorders also occur more commonly in patients with SOC; therefore, it is important to clearly distinguish these entities from rosacea in this population.

Systemic Lupus Erythematosus
Systemic lupus erythematosus is an autoimmune disease that commonly presents with erythema as well as erythematous inflammatory facial lesions similar to rosacea. The classic clinical appearance of SLE is the butterfly or malar rash, an erythematous macular eruption on the malar region of the face that also may involve the nose. This rash can appear similar to rosacea; however, the malar rash classically spares the nasolabial folds, while erythema of rosacea often involves this anatomic boundary. Although the facial erythema in both SLE and early stages of rosacea may be patchy and similar in presentation, the presence of papules and pustules rarely occurs in SLE and may help to differentiate SLE from certain variants of rosacea.15

 

 


Both SLE and rosacea may be exacerbated by sun exposure, and patients may report burning and stinging.16-18 Performing a complete physical examination, performing a skin biopsy with hematoxylin and eosin and direct immunofluorescence, and checking serologies including antinuclear antibody (ANA) can assist in making the diagnosis. It is important to note that elevated ANA, albeit lower than what is typically seen in SLE, has been reported in rosacea patients.19 If ANA is elevated, more specific SLE antibodies should be tested (eg, double-stranded DNA). Additionally, SLE can be differentiated on histology by a considerably lower CD4:CD8 ratio, fewer CD4+CD25+ regulatory T cells, and more CD123+ plasmacytoid dendritic cells compared to rosacea.20



Seborrheic Dermatitis
Seborrheic dermatitis is a frequent cause of facial erythema linked to the Malassezia yeast species in susceptible individuals. Seborrheic dermatitis has a notable prevalence in women of African descent and often is considered normal by these patients.21 Rosacea and seborrheic dermatitis are relatively common dermatoses and therefore can present concurrently. In both diseases, facial erythema may be difficult to discern upon cursory inspection. Seborrheic dermatitis may be distinguished from rosacea by the clinical appearance of erythematous patches and plaques involving the scalp, anterior and posterior hairlines, preauricular and postauricular areas, and medial eyebrows. Both seborrheic dermatitis and rosacea may involve the nasolabial folds, but the presence of scale in seborrheic dermatitis is a distinguishing feature. Scale may vary in appearance from thick, greasy, and yellowish to fine, thin, and whitish.22 In contrast to rosacea, the erythematous lesions of seborrheic dermatitis often are annular in configuration. Furthermore, postinflammatory hypopigmentation and, to a lesser extent, postinflammatory hyperpigmentation are key clinical components of seborrheic dermatitis in SOC patients but are not as commonly observed in rosacea.

Dermatomyositis
Dermatomyositis is a systemic autoimmune disease characterized by progressive and symmetric proximal musculoskeletal weakness and cutaneous findings. Facial erythema in the malar and nasolabial folds can be seen in patients with dermatomyositis18; however, the facial erythema seen in dermatomyositis, known as heliotrope rash, has a violaceous dusky quality and also involves the periorbital region. The violaceous hue and periorbital involvement are distinguishing features from rosacea. Okiyama et al23 described facial macular violaceous erythema with scale and edema in Japanese patients with dermatomyositis on the nasolabial folds, eyebrows, chin, cheeks, and ears; they also described mild atrophy with telangiectasia. Other clinical signs to help distinguish rosacea from dermatomyositis are the presence of edema of the face and extremities, Gottron papules, and poikiloderma. Dermatomyositis is a disease that affects all races; however, it is 4 times more common in black vs white patients,24 making it even more important to be able to distinguish between these conditions.

Acne Vulgaris
Acne vulgaris, the most commonly diagnosed dermatosis in patients with SOC, is characterized by papules, pustules, cysts, nodules, open and closed comedones, and hyperpigmented macules on the face, chest, and back.25,26 The absence of comedonal lesions and the presence of hyperpigmented macules distinguishes acne vulgaris from rosacea in this population.1 In addition, the absence of telangiectasia and flushing are important distinguishing factors when making the diagnosis of acne vulgaris.

Sarcoidosis
Sarcoidosis is a multisystem inflammatory disease characterized histologically by the presence of noncaseating granulomas in sites such as the lungs, lymph nodes, eyes, nervous system, liver, spleen, heart, and skin.27 Cutaneous sarcoidosis is known as a great mimicker of many other dermatoses, as it may present with multiple morphologic features. Cutaneous sarcoidosis most typically presents as papules, but nodules, plaques, lupus pernio, subcutaneous infiltrates, and infiltration of scars also have been identified.28 Sarcoid papules typically are 1 to 5 mm in size on the face, neck, and periorbital skin29; they are initially orange or yellow-brown in color, turn brownish red or violaceous, then involute to form faint macules.30 Papular lesions may either resolve or evolve into plaques, particularly on the extremities, face, scalp, back, and buttocks. Additionally, there are a few case reports of patients with cutaneous sarcoidosis presenting with large bulbous nasal masses initially thought to be rhinophyma.31-33 Finally, it may be difficult to distinguish sarcoidosis from granulomatous rosacea, which is characterized by firm yellow, brown, violaceous, red, or flesh-colored monomorphic papules or nodules affecting the perioral, periocular, medial, and/or lateral areas of the face (Figure 2).4,34 Patients also can have unilateral disease.35 Patients with granulomatous rosacea lack flushing and erythema as seen in more characteristic presentations of rosacea. They may report pain, pruritus, or burning, or they may be asymptomatic.36 Features that distinguish granulomatous rosacea from sarcoidosis include the absence of nodules, plaques, lupus pernio, subcutaneous infiltrates, and infiltration of scars. Clinical, histological, and radiographic evaluation are necessary to make the diagnosis of sarcoidosis over rosacea.

Figure 2. Granulomatous rosacea in a woman with Fitzpatrick skin type VI. Photograph courtesy of Susan C. Taylor, MD (Philadelphia, Pennsylvania).


Steroid Dermatitis
Steroid dermatitis involving the face may mimic rosacea. It is caused by the application of a potent corticosteroid to the facial skin for a prolonged period of time. In a report from a teaching hospital in Baghdad, the duration of application was 0.25 to 10 years on average.37 Reported characteristics of steroid dermatitis included facial erythema, telangiectasia, papules, pustules, and warmth to the touch. Distinguishing features from rosacea may be the presence of steroid dermatitis on the entire face, whereas rosacea tends to occur on the center of the face. Diagnosis of steroid dermatitis is made based on a history of chronic topical steroid use with rebound flares upon discontinuation of steroid.

Final Thoughts

Rosacea has features common to many other facial dermatoses, making the diagnosis challenging, particularly in patients with SOC. This difficulty in diagnosis may contribute to an underestimation of the prevalence of this disease in SOC patients. An understanding of rosacea, its nuances in clinical appearance, and its mimickers in SOC patients is important in making an accurate diagnosis.

References

Rosacea is a chronic inflammatory cutaneous disorder that affects the vasculature and pilosebaceous units of the face. Delayed and misdiagnosed rosacea in the SOC population has led to increased morbidity in this patient population. 1-3 It is characterized by facial flushing and warmth, erythema, telangiectasia, papules, and pustules. The 4 major subtypes include erythematotelangiectatic, papulopustular, phymatous, and ocular rosacea. 4 Granulomatous rosacea is considered to be a unique variant of rosacea. Until recently, rosacea was thought to predominately affect lighter-skinned individuals of Celtic and northern European origin. 5,6 A paucity of studies and case reports in the literature have contributed to the commonly held belief that rosacea occurs infrequently in patients with skin of color (SOC). 1 A PubMed search of articles indexed for MEDLINE revealed 32 results using the terms skin of color and rosacea vs 3786 using the term rosacea alone. It is possible that the nuance involved in appreciating erythema or other clinical manifestations of rosacea in SOC patients has led to underdiagnosis. Alternatively, these patients may be unaware that their symptoms represent a disease process and do not seek treatment. Many patients with darker skin will have endured rosacea for months or even years because the disease has been unrecognized or misdiagnosed. 6-8 Another factor possibly accounting for the perception that rosacea occurs infrequently in patients with SOC is misdiagnosis of rosacea as other diseases that are known to occur more commonly in the SOC population. Dermatologists should be aware that rosacea can affect SOC patients and that there are several rosacea mimickers to be considered and excluded when making the rosacea diagnosis in this patient population. To promote accurate and timely diagnosis of rosacea, we review several possible rosacea mimickers in SOC patients and highlight the distinguishing features.

Epidemiology

In 2018, a meta-analysis of published studies on rosacea estimated the global prevalence in all adults to be 5.46%.9 A multicenter study across 6 cities in Colombia identified 291 outpatients with rosacea; of them, 12.4% had either Fitzpatrick skin types IV or V.10 A study of 2743 Angolan adults with Fitzpatrick skin types V and VI reported that only 0.4% of patients had a diagnosis of rosacea.11 A Saudi study of 50 dark-skinned female patients with rosacea revealed 40% (20/50), 18% (9/50), and 42% (21/50) were Fitzpatrick skin types IV, V, and VI, respectively.12 The prevalence of rosacea in SOC patients in the United States is less defined. Data from the US National Ambulatory Medical Care Survey (1993-2010) of 31.5 million rosacea visits showed that 2% of rosacea patients were black, 2.3% were Asian or Pacific Islander, and 3.9% were Hispanic or Latino.8

Clinical Features

Each of the 4 major rosacea subtypes can present in the SOC population. The granulomatous variant has been predominantly reported in black patients.13 This predilection has been attributed to either an increased susceptibility in black patients to develop this variant or a delay in diagnosis of earlier phases of inflammatory rosacea.7

In a Saudi study (N=50), severe erythematotelangiectatic rosacea was diagnosed in 42% (21/50) of patients, with the majority having Fitzpatrick skin type IV. The severe papulopustular subtype was seen in 14% (7/50) of patients, with 20% (10/50) and 14% (7/50) having Fitzpatrick skin types IV and VI, respectively.12 In a Tunisian study (N=244), erythematotelangiectatic rosacea was seen in 12% of patients, papulopustular rosacea in 69%, phymatous rosacea in 4%, and ocular rosacea in 16%. Less frequently, the granulomatous variant was seen in 3% of patients, and steroid rosacea was noted in 12% patients.14

Recognizing the signs of rosacea may be a challenge, particularly erythema and telangiectasia. Tips for making an accurate diagnosis include use of adequate lighting, blanching of the skin (Figure 1), photography of the affected area against a dark blue background, and dermatoscopic examination.3 Furthermore, a thorough medical history, especially when evaluating the presence of facial erythema and identifying triggers, may help reach the correct diagnosis. Careful examination of the distribution of papules and pustules as well as the morphology and color of the papules in SOC patients also may provide diagnostic clues.

Figure 1. Diascopy revealed telangiectasia associated with rosacea on the central face of a woman with Fitzpatrick skin type IV. Photograph courtesy of Jennifer David, DO (Philadelphia, Pennsylvania).

Differential Diagnosis and Distinguishing Features

Several disorders are included in the differential diagnosis of rosacea and may confound a correct rosacea diagnosis, including systemic lupus erythematosus (SLE), seborrheic dermatitis, dermatomyositis, acne vulgaris, sarcoidosis, and steroid dermatitis. Many of these disorders also occur more commonly in patients with SOC; therefore, it is important to clearly distinguish these entities from rosacea in this population.

Systemic Lupus Erythematosus
Systemic lupus erythematosus is an autoimmune disease that commonly presents with erythema as well as erythematous inflammatory facial lesions similar to rosacea. The classic clinical appearance of SLE is the butterfly or malar rash, an erythematous macular eruption on the malar region of the face that also may involve the nose. This rash can appear similar to rosacea; however, the malar rash classically spares the nasolabial folds, while erythema of rosacea often involves this anatomic boundary. Although the facial erythema in both SLE and early stages of rosacea may be patchy and similar in presentation, the presence of papules and pustules rarely occurs in SLE and may help to differentiate SLE from certain variants of rosacea.15

 

 


Both SLE and rosacea may be exacerbated by sun exposure, and patients may report burning and stinging.16-18 Performing a complete physical examination, performing a skin biopsy with hematoxylin and eosin and direct immunofluorescence, and checking serologies including antinuclear antibody (ANA) can assist in making the diagnosis. It is important to note that elevated ANA, albeit lower than what is typically seen in SLE, has been reported in rosacea patients.19 If ANA is elevated, more specific SLE antibodies should be tested (eg, double-stranded DNA). Additionally, SLE can be differentiated on histology by a considerably lower CD4:CD8 ratio, fewer CD4+CD25+ regulatory T cells, and more CD123+ plasmacytoid dendritic cells compared to rosacea.20



Seborrheic Dermatitis
Seborrheic dermatitis is a frequent cause of facial erythema linked to the Malassezia yeast species in susceptible individuals. Seborrheic dermatitis has a notable prevalence in women of African descent and often is considered normal by these patients.21 Rosacea and seborrheic dermatitis are relatively common dermatoses and therefore can present concurrently. In both diseases, facial erythema may be difficult to discern upon cursory inspection. Seborrheic dermatitis may be distinguished from rosacea by the clinical appearance of erythematous patches and plaques involving the scalp, anterior and posterior hairlines, preauricular and postauricular areas, and medial eyebrows. Both seborrheic dermatitis and rosacea may involve the nasolabial folds, but the presence of scale in seborrheic dermatitis is a distinguishing feature. Scale may vary in appearance from thick, greasy, and yellowish to fine, thin, and whitish.22 In contrast to rosacea, the erythematous lesions of seborrheic dermatitis often are annular in configuration. Furthermore, postinflammatory hypopigmentation and, to a lesser extent, postinflammatory hyperpigmentation are key clinical components of seborrheic dermatitis in SOC patients but are not as commonly observed in rosacea.

Dermatomyositis
Dermatomyositis is a systemic autoimmune disease characterized by progressive and symmetric proximal musculoskeletal weakness and cutaneous findings. Facial erythema in the malar and nasolabial folds can be seen in patients with dermatomyositis18; however, the facial erythema seen in dermatomyositis, known as heliotrope rash, has a violaceous dusky quality and also involves the periorbital region. The violaceous hue and periorbital involvement are distinguishing features from rosacea. Okiyama et al23 described facial macular violaceous erythema with scale and edema in Japanese patients with dermatomyositis on the nasolabial folds, eyebrows, chin, cheeks, and ears; they also described mild atrophy with telangiectasia. Other clinical signs to help distinguish rosacea from dermatomyositis are the presence of edema of the face and extremities, Gottron papules, and poikiloderma. Dermatomyositis is a disease that affects all races; however, it is 4 times more common in black vs white patients,24 making it even more important to be able to distinguish between these conditions.

Acne Vulgaris
Acne vulgaris, the most commonly diagnosed dermatosis in patients with SOC, is characterized by papules, pustules, cysts, nodules, open and closed comedones, and hyperpigmented macules on the face, chest, and back.25,26 The absence of comedonal lesions and the presence of hyperpigmented macules distinguishes acne vulgaris from rosacea in this population.1 In addition, the absence of telangiectasia and flushing are important distinguishing factors when making the diagnosis of acne vulgaris.

Sarcoidosis
Sarcoidosis is a multisystem inflammatory disease characterized histologically by the presence of noncaseating granulomas in sites such as the lungs, lymph nodes, eyes, nervous system, liver, spleen, heart, and skin.27 Cutaneous sarcoidosis is known as a great mimicker of many other dermatoses, as it may present with multiple morphologic features. Cutaneous sarcoidosis most typically presents as papules, but nodules, plaques, lupus pernio, subcutaneous infiltrates, and infiltration of scars also have been identified.28 Sarcoid papules typically are 1 to 5 mm in size on the face, neck, and periorbital skin29; they are initially orange or yellow-brown in color, turn brownish red or violaceous, then involute to form faint macules.30 Papular lesions may either resolve or evolve into plaques, particularly on the extremities, face, scalp, back, and buttocks. Additionally, there are a few case reports of patients with cutaneous sarcoidosis presenting with large bulbous nasal masses initially thought to be rhinophyma.31-33 Finally, it may be difficult to distinguish sarcoidosis from granulomatous rosacea, which is characterized by firm yellow, brown, violaceous, red, or flesh-colored monomorphic papules or nodules affecting the perioral, periocular, medial, and/or lateral areas of the face (Figure 2).4,34 Patients also can have unilateral disease.35 Patients with granulomatous rosacea lack flushing and erythema as seen in more characteristic presentations of rosacea. They may report pain, pruritus, or burning, or they may be asymptomatic.36 Features that distinguish granulomatous rosacea from sarcoidosis include the absence of nodules, plaques, lupus pernio, subcutaneous infiltrates, and infiltration of scars. Clinical, histological, and radiographic evaluation are necessary to make the diagnosis of sarcoidosis over rosacea.

Figure 2. Granulomatous rosacea in a woman with Fitzpatrick skin type VI. Photograph courtesy of Susan C. Taylor, MD (Philadelphia, Pennsylvania).


Steroid Dermatitis
Steroid dermatitis involving the face may mimic rosacea. It is caused by the application of a potent corticosteroid to the facial skin for a prolonged period of time. In a report from a teaching hospital in Baghdad, the duration of application was 0.25 to 10 years on average.37 Reported characteristics of steroid dermatitis included facial erythema, telangiectasia, papules, pustules, and warmth to the touch. Distinguishing features from rosacea may be the presence of steroid dermatitis on the entire face, whereas rosacea tends to occur on the center of the face. Diagnosis of steroid dermatitis is made based on a history of chronic topical steroid use with rebound flares upon discontinuation of steroid.

Final Thoughts

Rosacea has features common to many other facial dermatoses, making the diagnosis challenging, particularly in patients with SOC. This difficulty in diagnosis may contribute to an underestimation of the prevalence of this disease in SOC patients. An understanding of rosacea, its nuances in clinical appearance, and its mimickers in SOC patients is important in making an accurate diagnosis.

References

References
  1. Alexis AF. Rosacea in patients with skin of color: uncommon but not rare. Cutis. 2010;86:60-62. 
  2. Kim NH, Yun SJ, Lee JB. Clinical features of Korean patients with rhinophyma. J Dermatol. 2017;44:710-712. 
  3. Hua TC, Chung PI, Chen YJ, et al. Cardiovascular comorbidities in patients with rosacea: a nationwide case-control study from Taiwan. J Am Acad Dermatol. 2015;73:249-254. 
  4. Wilkin J, Dahl M, Detmar M, et al. Standard classification of rosacea: report of the National Rosacea Society Expert Committee on the Classification and Staging of Rosacea. J Am Acad Dermatol. 2002;46:584-587. 
  5. Elewski BE, Draelos Z, Dreno B, et al. Global diversity and optimized outcome: proposed international consensus from the Rosacea International Expert Group. J Eur Acad Dermatol Venereol. 2011;25:188-200. 
  6. Alexis AF, Callender VD, Baldwin HE, et al. Global epidemiology and clinical spectrum of rosacea, highlighting skin of color: review and clinical practice experience [published online September 19, 2018]. J Am Acad Dermatol. 2019;80:1722-1729.e7. 
     
  7. Dlova NC, Mosam A. Rosacea in black South Africans with skin phototypes V and VI. Clin Exp Dermatol. 2017;42:670-673. 
  8. Al-Dabagh A, Davis SA, McMichael AJ, et al. Rosacea in skin of color: not a rare diagnosis [published online October 15, 2014]. Dermatol Online J. 2014;20. pii:13030/qt1mv9r0ss. 
  9. Gether L, Overgaard LK, Egeberg A, et al. Incidence and prevalence of rosacea: a systematic review and meta-analysis. Br J Dermatol. 2018;179:282-289. 
  10. Rueda LJ, Motta A, Pabon JG, et al. Epidemiology of rosacea in Colombia. Int J Dermatol. 2017;56:510-513. 
  11. De Luca DA, Maianski Z, Averbukh M. A study of skin disease spectrum occurring in Angola phototype V-VI population in Luanda. Int J Dermatol. 2018;57:849-855. 
  12. Al Balbeesi AO, Halawani MR. Unusual features of rosacea in Saudi females with dark skin. Ochsner J. 2014;14:321-327. 
  13. Rosen T, Stone MS. Acne rosacea in blacks. J Am Acad Dermatol. 1987;17:70-73. 
  14. Khaled A, Hammami H, Zeglaoui F, et al. Rosacea: 244 Tunisian cases. Tunis Med. 2010;88:597-601. 
  15. Usatine RP, Smith MA, Chumley HS, et al. The Color Atlas of Family Medicine. 2nd ed. New York, NY: The McGraw-Hill Companies; 2013.  
  16. O'Gorman SM, Murphy GM. Photoaggravated disorders. Dermatol Clin. 2014;32:385-398, ix. 
  17. Foering K, Chang AY, Piette EW, et al. Characterization of clinical photosensitivity in cutaneous lupus erythematosus. J Am Acad Dermatol. 2013;69:205-213. 
  18. Saleem MD, Wilkin JK. Evaluating and optimizing the diagnosis of erythematotelangiectatic rosacea. Dermatol Clin. 2018;36:127-134. 
  19. Black AA, McCauliffe DP, Sontheimer RD. Prevalence of acne rosacea in a rheumatic skin disease subspecialty clinic. Lupus. 1992;1:229-237. 
  20. Brown TT, Choi EY, Thomas DG, et al. Comparative analysis of rosacea and cutaneous lupus erythematosus: histopathologic features, T-cell subsets, and plasmacytoid dendritic cells. J Am Acad Dermatol. 2014;71:100-107. 
  21. Taylor SC, Barbosa V, Burgess C, et al. Hair and scalp disorders in adult and pediatric patients with skin of color. Cutis. 2017;100:31-35. 
  22. Gary G. Optimizing treatment approaches in seborrheic dermatitis. J Clin Aesthet Dermatol. 2013;6:44-49. 
  23. Okiyama N, Kohsaka H, Ueda N, et al. Seborrheic area erythema as a common skin manifestation in Japanese patients with dermatomyositis. Dermatology. 2008;217:374-377. 
  24. Taylor SC, Kyei A. Defining skin of color. In: Taylor SC, Kelly AP, Lim HW, et al, eds. Taylor and Kelly's Dermatology for Skin of Color. 2nd ed. New York, NY: McGraw-Hill; 2016:9-15. 
  25. Davis SA, Narahari S, Feldman SR, et al. Top dermatologic conditions in patients of color: an analysis of nationally representative data. J Drugs Dermatol. 2012;11:466-473. 
  26. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl understanding):S98-S106. 
  27. Wick MR. Granulomatous & histiocytic dermatitides. Semin Diagn Pathol. 2017;34:301-311. 
  28. Ball NJ, Kho GT, Martinka M. The histologic spectrum of cutaneous sarcoidosis: a study of twenty-eight cases. J Cutan Pathol. 2004;31:160-168. 
  29. Marchell RM, Judson MA. Chronic cutaneous lesions of sarcoidosis. Clin Dermatol. 2007;25:295-302. 
  30. Mahajan VK, Sharma NL, Sharma RC, et al. Cutaneous sarcoidosis: clinical profile of 23 Indian patients. Indian J Dermatol Venereol Leprol. 2007;73:16-21. 
  31. Goldenberg JD, Kotler HS, Shamsai R, et al. Sarcoidosis of the external nose mimicking rhinophyma. case report and review of the literature. Ann Otol Rhinol Laryngol. 1998;107:514-518. 
  32. Gupta-Elera G, Lam C, Chung C, et al. Violaceous plaque on the nose referred for rhinophyma surgery. Int J Dermatol. 2015;54:1011-1013. 
  33. Leonard AL. A case of sarcoidosis mimicking rhinophyma. J Drugs Dermatol. 2003;2:333-334. 
  34. Kelati A, Mernissi FZ. Granulomatous rosacea: a case report. J Med Case Rep. 2017;11:230. 
  35. Crawford GH, Pelle MT, James WD. Rosacea: I. etiology, pathogenesis, and subtype classification. J Am Acad Dermatol. 2004;51:327-341; quiz 342-324. 
  36. Reinholz M, Ruzicka T, Steinhoff M, et al. Pathogenesis and clinical presentation of rosacea as a key for a symptom-oriented therapy. J Dtsch Dermatol Ges. 2016;14(suppl 6):4-15. 
  37. Hameed AF. Steroid dermatitis resembling rosacea: a clinical evaluation of 75 patients. ISRN Dermatol. 2013;2013:491376.
References
  1. Alexis AF. Rosacea in patients with skin of color: uncommon but not rare. Cutis. 2010;86:60-62. 
  2. Kim NH, Yun SJ, Lee JB. Clinical features of Korean patients with rhinophyma. J Dermatol. 2017;44:710-712. 
  3. Hua TC, Chung PI, Chen YJ, et al. Cardiovascular comorbidities in patients with rosacea: a nationwide case-control study from Taiwan. J Am Acad Dermatol. 2015;73:249-254. 
  4. Wilkin J, Dahl M, Detmar M, et al. Standard classification of rosacea: report of the National Rosacea Society Expert Committee on the Classification and Staging of Rosacea. J Am Acad Dermatol. 2002;46:584-587. 
  5. Elewski BE, Draelos Z, Dreno B, et al. Global diversity and optimized outcome: proposed international consensus from the Rosacea International Expert Group. J Eur Acad Dermatol Venereol. 2011;25:188-200. 
  6. Alexis AF, Callender VD, Baldwin HE, et al. Global epidemiology and clinical spectrum of rosacea, highlighting skin of color: review and clinical practice experience [published online September 19, 2018]. J Am Acad Dermatol. 2019;80:1722-1729.e7. 
     
  7. Dlova NC, Mosam A. Rosacea in black South Africans with skin phototypes V and VI. Clin Exp Dermatol. 2017;42:670-673. 
  8. Al-Dabagh A, Davis SA, McMichael AJ, et al. Rosacea in skin of color: not a rare diagnosis [published online October 15, 2014]. Dermatol Online J. 2014;20. pii:13030/qt1mv9r0ss. 
  9. Gether L, Overgaard LK, Egeberg A, et al. Incidence and prevalence of rosacea: a systematic review and meta-analysis. Br J Dermatol. 2018;179:282-289. 
  10. Rueda LJ, Motta A, Pabon JG, et al. Epidemiology of rosacea in Colombia. Int J Dermatol. 2017;56:510-513. 
  11. De Luca DA, Maianski Z, Averbukh M. A study of skin disease spectrum occurring in Angola phototype V-VI population in Luanda. Int J Dermatol. 2018;57:849-855. 
  12. Al Balbeesi AO, Halawani MR. Unusual features of rosacea in Saudi females with dark skin. Ochsner J. 2014;14:321-327. 
  13. Rosen T, Stone MS. Acne rosacea in blacks. J Am Acad Dermatol. 1987;17:70-73. 
  14. Khaled A, Hammami H, Zeglaoui F, et al. Rosacea: 244 Tunisian cases. Tunis Med. 2010;88:597-601. 
  15. Usatine RP, Smith MA, Chumley HS, et al. The Color Atlas of Family Medicine. 2nd ed. New York, NY: The McGraw-Hill Companies; 2013.  
  16. O'Gorman SM, Murphy GM. Photoaggravated disorders. Dermatol Clin. 2014;32:385-398, ix. 
  17. Foering K, Chang AY, Piette EW, et al. Characterization of clinical photosensitivity in cutaneous lupus erythematosus. J Am Acad Dermatol. 2013;69:205-213. 
  18. Saleem MD, Wilkin JK. Evaluating and optimizing the diagnosis of erythematotelangiectatic rosacea. Dermatol Clin. 2018;36:127-134. 
  19. Black AA, McCauliffe DP, Sontheimer RD. Prevalence of acne rosacea in a rheumatic skin disease subspecialty clinic. Lupus. 1992;1:229-237. 
  20. Brown TT, Choi EY, Thomas DG, et al. Comparative analysis of rosacea and cutaneous lupus erythematosus: histopathologic features, T-cell subsets, and plasmacytoid dendritic cells. J Am Acad Dermatol. 2014;71:100-107. 
  21. Taylor SC, Barbosa V, Burgess C, et al. Hair and scalp disorders in adult and pediatric patients with skin of color. Cutis. 2017;100:31-35. 
  22. Gary G. Optimizing treatment approaches in seborrheic dermatitis. J Clin Aesthet Dermatol. 2013;6:44-49. 
  23. Okiyama N, Kohsaka H, Ueda N, et al. Seborrheic area erythema as a common skin manifestation in Japanese patients with dermatomyositis. Dermatology. 2008;217:374-377. 
  24. Taylor SC, Kyei A. Defining skin of color. In: Taylor SC, Kelly AP, Lim HW, et al, eds. Taylor and Kelly's Dermatology for Skin of Color. 2nd ed. New York, NY: McGraw-Hill; 2016:9-15. 
  25. Davis SA, Narahari S, Feldman SR, et al. Top dermatologic conditions in patients of color: an analysis of nationally representative data. J Drugs Dermatol. 2012;11:466-473. 
  26. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl understanding):S98-S106. 
  27. Wick MR. Granulomatous & histiocytic dermatitides. Semin Diagn Pathol. 2017;34:301-311. 
  28. Ball NJ, Kho GT, Martinka M. The histologic spectrum of cutaneous sarcoidosis: a study of twenty-eight cases. J Cutan Pathol. 2004;31:160-168. 
  29. Marchell RM, Judson MA. Chronic cutaneous lesions of sarcoidosis. Clin Dermatol. 2007;25:295-302. 
  30. Mahajan VK, Sharma NL, Sharma RC, et al. Cutaneous sarcoidosis: clinical profile of 23 Indian patients. Indian J Dermatol Venereol Leprol. 2007;73:16-21. 
  31. Goldenberg JD, Kotler HS, Shamsai R, et al. Sarcoidosis of the external nose mimicking rhinophyma. case report and review of the literature. Ann Otol Rhinol Laryngol. 1998;107:514-518. 
  32. Gupta-Elera G, Lam C, Chung C, et al. Violaceous plaque on the nose referred for rhinophyma surgery. Int J Dermatol. 2015;54:1011-1013. 
  33. Leonard AL. A case of sarcoidosis mimicking rhinophyma. J Drugs Dermatol. 2003;2:333-334. 
  34. Kelati A, Mernissi FZ. Granulomatous rosacea: a case report. J Med Case Rep. 2017;11:230. 
  35. Crawford GH, Pelle MT, James WD. Rosacea: I. etiology, pathogenesis, and subtype classification. J Am Acad Dermatol. 2004;51:327-341; quiz 342-324. 
  36. Reinholz M, Ruzicka T, Steinhoff M, et al. Pathogenesis and clinical presentation of rosacea as a key for a symptom-oriented therapy. J Dtsch Dermatol Ges. 2016;14(suppl 6):4-15. 
  37. Hameed AF. Steroid dermatitis resembling rosacea: a clinical evaluation of 75 patients. ISRN Dermatol. 2013;2013:491376.
Issue
Cutis - 104(1)
Issue
Cutis - 104(1)
Page Number
38-41
Page Number
38-41
Publications
Publications
Topics
Article Type
Display Headline
Establishing the Diagnosis of Rosacea in Skin of Color Patients
Display Headline
Establishing the Diagnosis of Rosacea in Skin of Color Patients
Sections
Inside the Article

Practice Points

  • The clinical signs of rosacea may be similar in all skin types; however, dermatologists must have a high clinical index of suspicion for rosacea in patients with skin of color (SOC).
  • Dermatologists should consider a wide differential diagnosis when presented with an SOC patient with facial erythema and/or papules and pustules.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Dermoscopic Patterns of Acral Melanocytic Lesions in Skin of Color

Article Type
Changed
Thu, 10/29/2020 - 14:49
Display Headline
Dermoscopic Patterns of Acral Melanocytic Lesions in Skin of Color
In Collaboration With the Skin of Color Society

Acral lentiginous melanoma (ALM) is a rare subtype of melanoma that occurs on the palms, soles, and nail apparatus. Unlike more common types of melanoma, ALM occurs on sun-protected areas of the skin and has distinct clinical, histologic, and genetic features. Acral lentiginous melanoma accounts for a larger proportion of melanomas in individuals with skin of color and has a worse prognosis and recurrence rate than other forms of melanoma.

Population Trends in Skin of Color

Much of the literature on malignant melanoma historically has involved non-Hispanic white patients, but the incidence in lighter-skinned populations has been increasing steadily over the last few decades.1 Although ALM can occur in any race, it disproportionately affects skin of color populations; ALM accounts for only 0.8% to 1% of all melanomas in white populations, but it constitutes 4% to 58% of melanomas in ethnic populations and is the most common melanoma subtype among black Americans.2-5 Acral lentiginous melanoma also is associated with a worse prognosis compared to other subtypes, which may indicate a more aggressive biological nature6 but also may point toward socioeconomic and cultural barriers (eg, low income or education levels, lack of insurance, lower health literacy), leading to disparities in access to care and diagnosis at advanced stages.5

Similarly, the distribution of acral melanocytic nevi appears to demonstrate an association with ethnicity and skin pigmentation. Although skin of color patients have fewer nevi than non-Hispanic whites, the proportion of acral melanocytic nevi tends to be greater.6,7 Given its grim prognosis, accurately differentiating ALM from acral nevi is of utmost importance.

Diagnostic Challenges of Acral Lesions

Due to the unique nature of the surfaces of acral sites, melanocytic lesions on the palms, soles, and nail apparatus present many diagnostic challenges. It can be difficult to distinguish acral melanoma from benign lesions using the naked eye alone. Volar surfaces are characterized by the presence of dermatoglyphics, and pigment deposition along ridges and furrows create particular dermoscopic patterns exclusive to these sites.8 Thus, dermoscopy can be useful on acral surfaces, but the dermoscopic features are different from those on the rest of the body and must be learned separately.

In addition, nearly half of patients are unaware of their acral lesions.6 Acral surfaces may not always be examined by clinicians during total-body skin examinations, leading to further possibility of overlooking a lesion. Obtaining biopsies on glabrous skin or nails also is challenging because they can be more painful and hemostasis can be more difficult, especially in the nail. Acral melanomas also may be amelanotic, including those at subungual sites. Although the overall incidence of amelanotic ALM is low, approximately 20% to 28% of amelanotic melanomas in Asian patients are located on acral sites.9 Due to these challenges, acral lesions may be overlooked or misdiagnosed as warts,10 tinea pedis,11 or traumatic ulcers.12

 

 

Dermoscopic Patterns of Acral Volar Skin

Dermoscopy is a useful noninvasive tool for distinguishing between benign and malignant acral melanocytic lesions, and its efficacy in improving diagnostic accuracy and decreasing unnecessary biopsies is well-established in the literature.13,14 Acral dermoscopy allows for visualization of pigment along the dermatoglyphics that constitute the characteristic dermoscopic patterns.

Acral Lentiginous Melanoma
The hallmark dermoscopic pattern and most important finding of ALM is the parallel ridge pattern, characterized by parallel linear pigmentation along the ridges of dermatoglyphics. In the early phases of malignancy, the pattern appears light brown and involves most of the lesion; as the tumor develops, increasing melanin production results in focal areas of the parallel ridge pattern with darker bands.15,16 The sensitivity and specificity of a parallel ridge pattern for diagnosing early ALM has been shown to be 86% and 99%, respectively.15,16

A pattern of irregular diffuse pigmentation also can be observed in more advanced ALM. Dermoscopy may reveal a structureless pattern (ie, lack of identifiable structures or patterns) in a background of tan-black coloration due to more exuberant melanocyte proliferation along the epidermis.15 Sensitivity and specificity of this dermoscopic finding for invasive lesions is high at 94% and 97%, respectively.16,17 Interestingly, once ALM lesions have advanced even further, conventional melanoma-associated structures (ie, blue-white veil, polymorphous blood vessels, ulceration, irregular dots/globules or streaks) or atypical forms of typically benign acral dermoscopic patterns may be observed.15

Per a 3-step diagnostic algorithm created by Koga and Saida,18 a suspected acral lesion should first be evaluated for a parallel ridge pattern to determine the need for biopsy, as it is seen in approximately two-thirds of ALMs.19 If no parallel ridge pattern is observed, the lesion should then be checked for any of the typical dermoscopic patterns seen in benign acral nevi (eg, parallel furrow, latticelike, or fibrillar patterns).18 The maximum diameter should be measured only if the lesion does not exhibit any of the typical dermoscopic patterns. If the lesion’s diameter is greater than 7 mm in diameter, it should be biopsied; if the diameter is less than 7 mm, it should have regular clinical and dermoscopic follow-up.18

In 2015, Lallas et al20 developed the BRAAFF checklist, a scoring system of 6 variables: blotches, ridge pattern, asymmetry of structures, asymmetry of colors, parallel furrow pattern, and fibrillar pattern. The checklist also was shown to substantially improve diagnostic accuracy of dermoscopy for ALM, with sensitivity and specificity at 93.1% and 86.7%, respectively.20

Acquired Acral Nevi
Three classic dermoscopic patterns are associated with acquired acral nevi: parallel furrow pattern, latticelike pattern, and fibrillar pattern.15,21 Approximately three-quarters of all acquired acral nevi exhibit one of these patterns, roughly half exhibiting parallel furrow with tan-brown bandlike pigmentation along dermatoglyphic grooves.16,17

Latticelike patterns also are characterized by brown parallel lines along the sulci of dermatoglyphics but additionally have multiple intersecting lines. Thus, this pattern can be considered a variant of the parallel furrow pattern.15 The crisscross markings can be predominantly found in the plantar arch.22 This dermoscopic pattern comprises 15% to 25% of all acral nevi.21

Fibrillar pattern accounts for 10% to 20% of all acral melanocytic nevi.21 Dermoscopically, these lesions demonstrate parallel filamentous streaks that cross dermatoglyphics obliquely. The fibrillar pattern is predominantly found on weight-bearing areas of the sole,22 which likely is explained by pressure causing slanting of melanin columns in the horny layer.23 The fibrillar pattern has been shown to be the benign acral dermoscopic pattern that is most commonly misdiagnosed, with higher reported rates of biopsy.24

Acral Congenital Melanocytic Nevi
Congenital melanocytic nevi (CMN) present at birth or appear during the first few weeks of life. Congenital melanocytic nevi can vary widely in size, shape, and color, and they are occasionally biopsied in cases of larger diameter or dermoscopic atypia to differentiate from melanoma.25 Congenital melanocytic nevi also can occur on acral volar surfaces. Possible dermoscopic patterns include parallel furrow or fibrillar patterns as well as a crista dotted pattern, defined as evenly spaced dots/globules on the ridges near the openings of eccrine ducts.26 A more commonly observed dermoscopic pattern in acral CMN is a combination of the crista dotted and parallel furrow patterns, known as the peas-in-a-pod pattern. Changes in the clinical appearance and dermoscopic features of an acral CMN are possible over time; some lesions also may fade with age.26

Final Thoughts

Acral lentiginous melanoma is a rare but potentially aggressive melanoma subtype that accounts for a larger proportion of melanomas in patients with skin of color than in white patients. Dermoscopy of acral volar skin provides invaluable diagnostic information and allows for better management of acral melanocytic lesions. Dermoscopic patterns such as the parallel ridge, parallel furrow, latticelike, fibrillar, and peas-in-a-pod patterns are unique to acral sites and can be used to differentiate between ALMs, acquired nevi, or CMNs.

References
  1. Whiteman DC, Green AC, Olsen CM. The growing burden of invasive melanoma: projections of incidence rates and numbers of new cases in six susceptible populations through 2031. J Invest Dermatol. 2016;136:1161-1171.
  2. Bradford PT, Goldstein AM, McMaster ML, et al. Acral lentiginous melanoma: incidence and survival patterns in the United States, 1986-2005. Arch Dermatol. 2009;145:427-434.
  3. Nakamura Y, Fujisawa Y. Diagnosis and management of acral lentiginous melanoma. Curr Treat Options Oncol. 2018;19:42.
  4. Cormier JN, Xing Y, Ding M, et al. Ethnic differences among patients with cutaneous melanoma. Arch Intern Med. 2006;166:1907-1914.
  5. Wang Y, Zhao Y, Ma S. Racial differences in six major subtypes of melanoma: descriptive epidemiology. BMC Cancer. 2016;16:691.
  6. Madankumar R, Gumaste PV, Martires K, et al. Acral melanocytic lesions in the United States: prevalence, awareness, and dermoscopic patterns in skin-of-color and non-Hispanic white patients. J Am Acad Dermatol. 2016;74:724.e1-730.e1.
  7. Palicka GA, Rhodes AR. Acral melanocytic nevi: prevalence and distribution of gross morphologic features in white and black adults. Arch Dermatol. 2010;146:1085-1094.
  8. Thomas L, Phan A, Pralong P, et al. Special locations dermoscopy: facial, acral, and nail. Dermatol Clin. 2013;31:615-624.
  9. Gong HZ, Zheng HY, Li J. Amelanotic melanoma [published online January 21, 2019]. Melanoma Res. doi:10.1097/CMR.0000000000000571.
  10. Ise M, Yasuda F, Konohana I, et al. Acral melanoma with hyperkeratosis mimicking a pigmented wart. Dermatol Pract Concept. 2013;3:37-39.
  11. Serarslan G, Akçaly CM, Atik E. Acral lentiginous melanoma misdiagnosed as tinea pedis: a case report. Int J Dermatol. 2004;43:37-38.
  12. Gumaste P, Penn L, Cohen N, et al. Acral lentiginous melanoma of the foot misdiagnosed as a traumatic ulcer. a cautionary case. J Am Podiatr Med Assoc. 2015;105:189-194.
  13. Carli P, de Giorgi V, Chiarugi A, et al. Addition of dermoscopy to conventional naked-eye examination in melanoma screening: a randomized study. J Am Acad Dermatol. 2004;50:683-689.
  14. Carli P, de Giorgi V, Crocetti E, et al. Improvement of malignant/benign ratio in excised melanocytic lesions in the ‘dermoscopy era’: a retrospective study 1997-2001. Br J Dermatol. 2004;150:687-692.
  15. Saida T, Koga H, Uhara H. Key points in dermoscopic differentiation between early acral melanoma and acral nevus. J Dermatol. 2011;38:25-34.
  16. Ishihara Y, Saida T, Miyazaki A, et al. Early acral melanoma in situ: correlation between the parallel ridge pattern on dermoscopy and microscopic features. Am J Dermatopathol. 2006;28:21-27.
  17. Saida T, Miyazaki A, Oguchi S, et al. Significance of dermoscopic patterns in detecting malignant melanoma on acral volar skin: results of a multicenter study in Japan. Arch Dermatol. 2004;140:1233-1238.
  18. Koga H, Saida T. Revised 3-step dermoscopic algorithm for the management of acral melanocytic lesions. Arch Dermatol. 2011;147:741-743.
  19. Lallas A, Sgouros D, Zalaudek I, et al. Palmar and plantar melanomas differ for sex prevalence and tumor thickness but not for dermoscopic patterns. Melanoma Res. 2014;24:83-87.
  20. Lallas A, Kyrgidis A, Koga H, et al. The BRAAFF checklist: a new dermoscopic algorithm for diagnosing acral melanoma. Br J Dermatol. 2015;173:1041-1049.
  21. Saida T, Koga H. Dermoscopic patterns of acral melanocytic nevi: their variations, changes, and significance. Arch Dermatol. 2007;143:1423-1426.
  22. Miyazaki A, Saida T, Koga H, et al. Anatomical and histopathological correlates of the dermoscopic patterns seen in melanocytic nevi on the sole: a retrospective study. J Am Acad Dermatol. 2005;53:230-236.
  23. Watanabe S, Sawada M, Ishizaki S, et al. Comparison of dermatoscopic images of acral lentiginous melanoma and acral melanocytic nevus occurring on body weight-bearing areas. Dermatol Pract Concept. 2014;4:47-50.
  24. Costello CM, Ghanavatian S, Temkit M, et al. Educational and practice gaps in the management of volar melanocytic lesions. J Eur Acad Dermatol Venereol. 2018;32:1450-1455.
  25. Alikhan A, Ibrahimi OA, Eisen DB. Congenital melanocytic nevi: where are we now? part I. clinical presentation, epidemiology, pathogenesis, histology, malignant transformation, and neurocutaneous melanosis. J Am Acad Dermatol. 2012;67:495.e1-495.e17; quiz 512-514.
  26. Minagawa A, Koga H, Saida T. Dermoscopic characteristics of congenital melanocytic nevi affecting acral volar skin. Arch Dermatol. 2011;147:809-813.
Article PDF
Author and Disclosure Information

From The Ronald O. Perelman Department of Dermatology, New York University School of Medicine, New York.

The authors report no conflict of interest.

Correspondence: Jennifer A. Stein, MD, PhD, The Ronald O. Perelman Department of Dermatology, New York University School of Medicine, 240 E 38th St, New York, NY 10016 (Jennifer.Stein@nyulangone.org).

Issue
Cutis - 103(5)
Publications
Topics
Page Number
274-276
Sections
Author and Disclosure Information

From The Ronald O. Perelman Department of Dermatology, New York University School of Medicine, New York.

The authors report no conflict of interest.

Correspondence: Jennifer A. Stein, MD, PhD, The Ronald O. Perelman Department of Dermatology, New York University School of Medicine, 240 E 38th St, New York, NY 10016 (Jennifer.Stein@nyulangone.org).

Author and Disclosure Information

From The Ronald O. Perelman Department of Dermatology, New York University School of Medicine, New York.

The authors report no conflict of interest.

Correspondence: Jennifer A. Stein, MD, PhD, The Ronald O. Perelman Department of Dermatology, New York University School of Medicine, 240 E 38th St, New York, NY 10016 (Jennifer.Stein@nyulangone.org).

Article PDF
Article PDF
In Collaboration With the Skin of Color Society
In Collaboration With the Skin of Color Society

Acral lentiginous melanoma (ALM) is a rare subtype of melanoma that occurs on the palms, soles, and nail apparatus. Unlike more common types of melanoma, ALM occurs on sun-protected areas of the skin and has distinct clinical, histologic, and genetic features. Acral lentiginous melanoma accounts for a larger proportion of melanomas in individuals with skin of color and has a worse prognosis and recurrence rate than other forms of melanoma.

Population Trends in Skin of Color

Much of the literature on malignant melanoma historically has involved non-Hispanic white patients, but the incidence in lighter-skinned populations has been increasing steadily over the last few decades.1 Although ALM can occur in any race, it disproportionately affects skin of color populations; ALM accounts for only 0.8% to 1% of all melanomas in white populations, but it constitutes 4% to 58% of melanomas in ethnic populations and is the most common melanoma subtype among black Americans.2-5 Acral lentiginous melanoma also is associated with a worse prognosis compared to other subtypes, which may indicate a more aggressive biological nature6 but also may point toward socioeconomic and cultural barriers (eg, low income or education levels, lack of insurance, lower health literacy), leading to disparities in access to care and diagnosis at advanced stages.5

Similarly, the distribution of acral melanocytic nevi appears to demonstrate an association with ethnicity and skin pigmentation. Although skin of color patients have fewer nevi than non-Hispanic whites, the proportion of acral melanocytic nevi tends to be greater.6,7 Given its grim prognosis, accurately differentiating ALM from acral nevi is of utmost importance.

Diagnostic Challenges of Acral Lesions

Due to the unique nature of the surfaces of acral sites, melanocytic lesions on the palms, soles, and nail apparatus present many diagnostic challenges. It can be difficult to distinguish acral melanoma from benign lesions using the naked eye alone. Volar surfaces are characterized by the presence of dermatoglyphics, and pigment deposition along ridges and furrows create particular dermoscopic patterns exclusive to these sites.8 Thus, dermoscopy can be useful on acral surfaces, but the dermoscopic features are different from those on the rest of the body and must be learned separately.

In addition, nearly half of patients are unaware of their acral lesions.6 Acral surfaces may not always be examined by clinicians during total-body skin examinations, leading to further possibility of overlooking a lesion. Obtaining biopsies on glabrous skin or nails also is challenging because they can be more painful and hemostasis can be more difficult, especially in the nail. Acral melanomas also may be amelanotic, including those at subungual sites. Although the overall incidence of amelanotic ALM is low, approximately 20% to 28% of amelanotic melanomas in Asian patients are located on acral sites.9 Due to these challenges, acral lesions may be overlooked or misdiagnosed as warts,10 tinea pedis,11 or traumatic ulcers.12

 

 

Dermoscopic Patterns of Acral Volar Skin

Dermoscopy is a useful noninvasive tool for distinguishing between benign and malignant acral melanocytic lesions, and its efficacy in improving diagnostic accuracy and decreasing unnecessary biopsies is well-established in the literature.13,14 Acral dermoscopy allows for visualization of pigment along the dermatoglyphics that constitute the characteristic dermoscopic patterns.

Acral Lentiginous Melanoma
The hallmark dermoscopic pattern and most important finding of ALM is the parallel ridge pattern, characterized by parallel linear pigmentation along the ridges of dermatoglyphics. In the early phases of malignancy, the pattern appears light brown and involves most of the lesion; as the tumor develops, increasing melanin production results in focal areas of the parallel ridge pattern with darker bands.15,16 The sensitivity and specificity of a parallel ridge pattern for diagnosing early ALM has been shown to be 86% and 99%, respectively.15,16

A pattern of irregular diffuse pigmentation also can be observed in more advanced ALM. Dermoscopy may reveal a structureless pattern (ie, lack of identifiable structures or patterns) in a background of tan-black coloration due to more exuberant melanocyte proliferation along the epidermis.15 Sensitivity and specificity of this dermoscopic finding for invasive lesions is high at 94% and 97%, respectively.16,17 Interestingly, once ALM lesions have advanced even further, conventional melanoma-associated structures (ie, blue-white veil, polymorphous blood vessels, ulceration, irregular dots/globules or streaks) or atypical forms of typically benign acral dermoscopic patterns may be observed.15

Per a 3-step diagnostic algorithm created by Koga and Saida,18 a suspected acral lesion should first be evaluated for a parallel ridge pattern to determine the need for biopsy, as it is seen in approximately two-thirds of ALMs.19 If no parallel ridge pattern is observed, the lesion should then be checked for any of the typical dermoscopic patterns seen in benign acral nevi (eg, parallel furrow, latticelike, or fibrillar patterns).18 The maximum diameter should be measured only if the lesion does not exhibit any of the typical dermoscopic patterns. If the lesion’s diameter is greater than 7 mm in diameter, it should be biopsied; if the diameter is less than 7 mm, it should have regular clinical and dermoscopic follow-up.18

In 2015, Lallas et al20 developed the BRAAFF checklist, a scoring system of 6 variables: blotches, ridge pattern, asymmetry of structures, asymmetry of colors, parallel furrow pattern, and fibrillar pattern. The checklist also was shown to substantially improve diagnostic accuracy of dermoscopy for ALM, with sensitivity and specificity at 93.1% and 86.7%, respectively.20

Acquired Acral Nevi
Three classic dermoscopic patterns are associated with acquired acral nevi: parallel furrow pattern, latticelike pattern, and fibrillar pattern.15,21 Approximately three-quarters of all acquired acral nevi exhibit one of these patterns, roughly half exhibiting parallel furrow with tan-brown bandlike pigmentation along dermatoglyphic grooves.16,17

Latticelike patterns also are characterized by brown parallel lines along the sulci of dermatoglyphics but additionally have multiple intersecting lines. Thus, this pattern can be considered a variant of the parallel furrow pattern.15 The crisscross markings can be predominantly found in the plantar arch.22 This dermoscopic pattern comprises 15% to 25% of all acral nevi.21

Fibrillar pattern accounts for 10% to 20% of all acral melanocytic nevi.21 Dermoscopically, these lesions demonstrate parallel filamentous streaks that cross dermatoglyphics obliquely. The fibrillar pattern is predominantly found on weight-bearing areas of the sole,22 which likely is explained by pressure causing slanting of melanin columns in the horny layer.23 The fibrillar pattern has been shown to be the benign acral dermoscopic pattern that is most commonly misdiagnosed, with higher reported rates of biopsy.24

Acral Congenital Melanocytic Nevi
Congenital melanocytic nevi (CMN) present at birth or appear during the first few weeks of life. Congenital melanocytic nevi can vary widely in size, shape, and color, and they are occasionally biopsied in cases of larger diameter or dermoscopic atypia to differentiate from melanoma.25 Congenital melanocytic nevi also can occur on acral volar surfaces. Possible dermoscopic patterns include parallel furrow or fibrillar patterns as well as a crista dotted pattern, defined as evenly spaced dots/globules on the ridges near the openings of eccrine ducts.26 A more commonly observed dermoscopic pattern in acral CMN is a combination of the crista dotted and parallel furrow patterns, known as the peas-in-a-pod pattern. Changes in the clinical appearance and dermoscopic features of an acral CMN are possible over time; some lesions also may fade with age.26

Final Thoughts

Acral lentiginous melanoma is a rare but potentially aggressive melanoma subtype that accounts for a larger proportion of melanomas in patients with skin of color than in white patients. Dermoscopy of acral volar skin provides invaluable diagnostic information and allows for better management of acral melanocytic lesions. Dermoscopic patterns such as the parallel ridge, parallel furrow, latticelike, fibrillar, and peas-in-a-pod patterns are unique to acral sites and can be used to differentiate between ALMs, acquired nevi, or CMNs.

Acral lentiginous melanoma (ALM) is a rare subtype of melanoma that occurs on the palms, soles, and nail apparatus. Unlike more common types of melanoma, ALM occurs on sun-protected areas of the skin and has distinct clinical, histologic, and genetic features. Acral lentiginous melanoma accounts for a larger proportion of melanomas in individuals with skin of color and has a worse prognosis and recurrence rate than other forms of melanoma.

Population Trends in Skin of Color

Much of the literature on malignant melanoma historically has involved non-Hispanic white patients, but the incidence in lighter-skinned populations has been increasing steadily over the last few decades.1 Although ALM can occur in any race, it disproportionately affects skin of color populations; ALM accounts for only 0.8% to 1% of all melanomas in white populations, but it constitutes 4% to 58% of melanomas in ethnic populations and is the most common melanoma subtype among black Americans.2-5 Acral lentiginous melanoma also is associated with a worse prognosis compared to other subtypes, which may indicate a more aggressive biological nature6 but also may point toward socioeconomic and cultural barriers (eg, low income or education levels, lack of insurance, lower health literacy), leading to disparities in access to care and diagnosis at advanced stages.5

Similarly, the distribution of acral melanocytic nevi appears to demonstrate an association with ethnicity and skin pigmentation. Although skin of color patients have fewer nevi than non-Hispanic whites, the proportion of acral melanocytic nevi tends to be greater.6,7 Given its grim prognosis, accurately differentiating ALM from acral nevi is of utmost importance.

Diagnostic Challenges of Acral Lesions

Due to the unique nature of the surfaces of acral sites, melanocytic lesions on the palms, soles, and nail apparatus present many diagnostic challenges. It can be difficult to distinguish acral melanoma from benign lesions using the naked eye alone. Volar surfaces are characterized by the presence of dermatoglyphics, and pigment deposition along ridges and furrows create particular dermoscopic patterns exclusive to these sites.8 Thus, dermoscopy can be useful on acral surfaces, but the dermoscopic features are different from those on the rest of the body and must be learned separately.

In addition, nearly half of patients are unaware of their acral lesions.6 Acral surfaces may not always be examined by clinicians during total-body skin examinations, leading to further possibility of overlooking a lesion. Obtaining biopsies on glabrous skin or nails also is challenging because they can be more painful and hemostasis can be more difficult, especially in the nail. Acral melanomas also may be amelanotic, including those at subungual sites. Although the overall incidence of amelanotic ALM is low, approximately 20% to 28% of amelanotic melanomas in Asian patients are located on acral sites.9 Due to these challenges, acral lesions may be overlooked or misdiagnosed as warts,10 tinea pedis,11 or traumatic ulcers.12

 

 

Dermoscopic Patterns of Acral Volar Skin

Dermoscopy is a useful noninvasive tool for distinguishing between benign and malignant acral melanocytic lesions, and its efficacy in improving diagnostic accuracy and decreasing unnecessary biopsies is well-established in the literature.13,14 Acral dermoscopy allows for visualization of pigment along the dermatoglyphics that constitute the characteristic dermoscopic patterns.

Acral Lentiginous Melanoma
The hallmark dermoscopic pattern and most important finding of ALM is the parallel ridge pattern, characterized by parallel linear pigmentation along the ridges of dermatoglyphics. In the early phases of malignancy, the pattern appears light brown and involves most of the lesion; as the tumor develops, increasing melanin production results in focal areas of the parallel ridge pattern with darker bands.15,16 The sensitivity and specificity of a parallel ridge pattern for diagnosing early ALM has been shown to be 86% and 99%, respectively.15,16

A pattern of irregular diffuse pigmentation also can be observed in more advanced ALM. Dermoscopy may reveal a structureless pattern (ie, lack of identifiable structures or patterns) in a background of tan-black coloration due to more exuberant melanocyte proliferation along the epidermis.15 Sensitivity and specificity of this dermoscopic finding for invasive lesions is high at 94% and 97%, respectively.16,17 Interestingly, once ALM lesions have advanced even further, conventional melanoma-associated structures (ie, blue-white veil, polymorphous blood vessels, ulceration, irregular dots/globules or streaks) or atypical forms of typically benign acral dermoscopic patterns may be observed.15

Per a 3-step diagnostic algorithm created by Koga and Saida,18 a suspected acral lesion should first be evaluated for a parallel ridge pattern to determine the need for biopsy, as it is seen in approximately two-thirds of ALMs.19 If no parallel ridge pattern is observed, the lesion should then be checked for any of the typical dermoscopic patterns seen in benign acral nevi (eg, parallel furrow, latticelike, or fibrillar patterns).18 The maximum diameter should be measured only if the lesion does not exhibit any of the typical dermoscopic patterns. If the lesion’s diameter is greater than 7 mm in diameter, it should be biopsied; if the diameter is less than 7 mm, it should have regular clinical and dermoscopic follow-up.18

In 2015, Lallas et al20 developed the BRAAFF checklist, a scoring system of 6 variables: blotches, ridge pattern, asymmetry of structures, asymmetry of colors, parallel furrow pattern, and fibrillar pattern. The checklist also was shown to substantially improve diagnostic accuracy of dermoscopy for ALM, with sensitivity and specificity at 93.1% and 86.7%, respectively.20

Acquired Acral Nevi
Three classic dermoscopic patterns are associated with acquired acral nevi: parallel furrow pattern, latticelike pattern, and fibrillar pattern.15,21 Approximately three-quarters of all acquired acral nevi exhibit one of these patterns, roughly half exhibiting parallel furrow with tan-brown bandlike pigmentation along dermatoglyphic grooves.16,17

Latticelike patterns also are characterized by brown parallel lines along the sulci of dermatoglyphics but additionally have multiple intersecting lines. Thus, this pattern can be considered a variant of the parallel furrow pattern.15 The crisscross markings can be predominantly found in the plantar arch.22 This dermoscopic pattern comprises 15% to 25% of all acral nevi.21

Fibrillar pattern accounts for 10% to 20% of all acral melanocytic nevi.21 Dermoscopically, these lesions demonstrate parallel filamentous streaks that cross dermatoglyphics obliquely. The fibrillar pattern is predominantly found on weight-bearing areas of the sole,22 which likely is explained by pressure causing slanting of melanin columns in the horny layer.23 The fibrillar pattern has been shown to be the benign acral dermoscopic pattern that is most commonly misdiagnosed, with higher reported rates of biopsy.24

Acral Congenital Melanocytic Nevi
Congenital melanocytic nevi (CMN) present at birth or appear during the first few weeks of life. Congenital melanocytic nevi can vary widely in size, shape, and color, and they are occasionally biopsied in cases of larger diameter or dermoscopic atypia to differentiate from melanoma.25 Congenital melanocytic nevi also can occur on acral volar surfaces. Possible dermoscopic patterns include parallel furrow or fibrillar patterns as well as a crista dotted pattern, defined as evenly spaced dots/globules on the ridges near the openings of eccrine ducts.26 A more commonly observed dermoscopic pattern in acral CMN is a combination of the crista dotted and parallel furrow patterns, known as the peas-in-a-pod pattern. Changes in the clinical appearance and dermoscopic features of an acral CMN are possible over time; some lesions also may fade with age.26

Final Thoughts

Acral lentiginous melanoma is a rare but potentially aggressive melanoma subtype that accounts for a larger proportion of melanomas in patients with skin of color than in white patients. Dermoscopy of acral volar skin provides invaluable diagnostic information and allows for better management of acral melanocytic lesions. Dermoscopic patterns such as the parallel ridge, parallel furrow, latticelike, fibrillar, and peas-in-a-pod patterns are unique to acral sites and can be used to differentiate between ALMs, acquired nevi, or CMNs.

References
  1. Whiteman DC, Green AC, Olsen CM. The growing burden of invasive melanoma: projections of incidence rates and numbers of new cases in six susceptible populations through 2031. J Invest Dermatol. 2016;136:1161-1171.
  2. Bradford PT, Goldstein AM, McMaster ML, et al. Acral lentiginous melanoma: incidence and survival patterns in the United States, 1986-2005. Arch Dermatol. 2009;145:427-434.
  3. Nakamura Y, Fujisawa Y. Diagnosis and management of acral lentiginous melanoma. Curr Treat Options Oncol. 2018;19:42.
  4. Cormier JN, Xing Y, Ding M, et al. Ethnic differences among patients with cutaneous melanoma. Arch Intern Med. 2006;166:1907-1914.
  5. Wang Y, Zhao Y, Ma S. Racial differences in six major subtypes of melanoma: descriptive epidemiology. BMC Cancer. 2016;16:691.
  6. Madankumar R, Gumaste PV, Martires K, et al. Acral melanocytic lesions in the United States: prevalence, awareness, and dermoscopic patterns in skin-of-color and non-Hispanic white patients. J Am Acad Dermatol. 2016;74:724.e1-730.e1.
  7. Palicka GA, Rhodes AR. Acral melanocytic nevi: prevalence and distribution of gross morphologic features in white and black adults. Arch Dermatol. 2010;146:1085-1094.
  8. Thomas L, Phan A, Pralong P, et al. Special locations dermoscopy: facial, acral, and nail. Dermatol Clin. 2013;31:615-624.
  9. Gong HZ, Zheng HY, Li J. Amelanotic melanoma [published online January 21, 2019]. Melanoma Res. doi:10.1097/CMR.0000000000000571.
  10. Ise M, Yasuda F, Konohana I, et al. Acral melanoma with hyperkeratosis mimicking a pigmented wart. Dermatol Pract Concept. 2013;3:37-39.
  11. Serarslan G, Akçaly CM, Atik E. Acral lentiginous melanoma misdiagnosed as tinea pedis: a case report. Int J Dermatol. 2004;43:37-38.
  12. Gumaste P, Penn L, Cohen N, et al. Acral lentiginous melanoma of the foot misdiagnosed as a traumatic ulcer. a cautionary case. J Am Podiatr Med Assoc. 2015;105:189-194.
  13. Carli P, de Giorgi V, Chiarugi A, et al. Addition of dermoscopy to conventional naked-eye examination in melanoma screening: a randomized study. J Am Acad Dermatol. 2004;50:683-689.
  14. Carli P, de Giorgi V, Crocetti E, et al. Improvement of malignant/benign ratio in excised melanocytic lesions in the ‘dermoscopy era’: a retrospective study 1997-2001. Br J Dermatol. 2004;150:687-692.
  15. Saida T, Koga H, Uhara H. Key points in dermoscopic differentiation between early acral melanoma and acral nevus. J Dermatol. 2011;38:25-34.
  16. Ishihara Y, Saida T, Miyazaki A, et al. Early acral melanoma in situ: correlation between the parallel ridge pattern on dermoscopy and microscopic features. Am J Dermatopathol. 2006;28:21-27.
  17. Saida T, Miyazaki A, Oguchi S, et al. Significance of dermoscopic patterns in detecting malignant melanoma on acral volar skin: results of a multicenter study in Japan. Arch Dermatol. 2004;140:1233-1238.
  18. Koga H, Saida T. Revised 3-step dermoscopic algorithm for the management of acral melanocytic lesions. Arch Dermatol. 2011;147:741-743.
  19. Lallas A, Sgouros D, Zalaudek I, et al. Palmar and plantar melanomas differ for sex prevalence and tumor thickness but not for dermoscopic patterns. Melanoma Res. 2014;24:83-87.
  20. Lallas A, Kyrgidis A, Koga H, et al. The BRAAFF checklist: a new dermoscopic algorithm for diagnosing acral melanoma. Br J Dermatol. 2015;173:1041-1049.
  21. Saida T, Koga H. Dermoscopic patterns of acral melanocytic nevi: their variations, changes, and significance. Arch Dermatol. 2007;143:1423-1426.
  22. Miyazaki A, Saida T, Koga H, et al. Anatomical and histopathological correlates of the dermoscopic patterns seen in melanocytic nevi on the sole: a retrospective study. J Am Acad Dermatol. 2005;53:230-236.
  23. Watanabe S, Sawada M, Ishizaki S, et al. Comparison of dermatoscopic images of acral lentiginous melanoma and acral melanocytic nevus occurring on body weight-bearing areas. Dermatol Pract Concept. 2014;4:47-50.
  24. Costello CM, Ghanavatian S, Temkit M, et al. Educational and practice gaps in the management of volar melanocytic lesions. J Eur Acad Dermatol Venereol. 2018;32:1450-1455.
  25. Alikhan A, Ibrahimi OA, Eisen DB. Congenital melanocytic nevi: where are we now? part I. clinical presentation, epidemiology, pathogenesis, histology, malignant transformation, and neurocutaneous melanosis. J Am Acad Dermatol. 2012;67:495.e1-495.e17; quiz 512-514.
  26. Minagawa A, Koga H, Saida T. Dermoscopic characteristics of congenital melanocytic nevi affecting acral volar skin. Arch Dermatol. 2011;147:809-813.
References
  1. Whiteman DC, Green AC, Olsen CM. The growing burden of invasive melanoma: projections of incidence rates and numbers of new cases in six susceptible populations through 2031. J Invest Dermatol. 2016;136:1161-1171.
  2. Bradford PT, Goldstein AM, McMaster ML, et al. Acral lentiginous melanoma: incidence and survival patterns in the United States, 1986-2005. Arch Dermatol. 2009;145:427-434.
  3. Nakamura Y, Fujisawa Y. Diagnosis and management of acral lentiginous melanoma. Curr Treat Options Oncol. 2018;19:42.
  4. Cormier JN, Xing Y, Ding M, et al. Ethnic differences among patients with cutaneous melanoma. Arch Intern Med. 2006;166:1907-1914.
  5. Wang Y, Zhao Y, Ma S. Racial differences in six major subtypes of melanoma: descriptive epidemiology. BMC Cancer. 2016;16:691.
  6. Madankumar R, Gumaste PV, Martires K, et al. Acral melanocytic lesions in the United States: prevalence, awareness, and dermoscopic patterns in skin-of-color and non-Hispanic white patients. J Am Acad Dermatol. 2016;74:724.e1-730.e1.
  7. Palicka GA, Rhodes AR. Acral melanocytic nevi: prevalence and distribution of gross morphologic features in white and black adults. Arch Dermatol. 2010;146:1085-1094.
  8. Thomas L, Phan A, Pralong P, et al. Special locations dermoscopy: facial, acral, and nail. Dermatol Clin. 2013;31:615-624.
  9. Gong HZ, Zheng HY, Li J. Amelanotic melanoma [published online January 21, 2019]. Melanoma Res. doi:10.1097/CMR.0000000000000571.
  10. Ise M, Yasuda F, Konohana I, et al. Acral melanoma with hyperkeratosis mimicking a pigmented wart. Dermatol Pract Concept. 2013;3:37-39.
  11. Serarslan G, Akçaly CM, Atik E. Acral lentiginous melanoma misdiagnosed as tinea pedis: a case report. Int J Dermatol. 2004;43:37-38.
  12. Gumaste P, Penn L, Cohen N, et al. Acral lentiginous melanoma of the foot misdiagnosed as a traumatic ulcer. a cautionary case. J Am Podiatr Med Assoc. 2015;105:189-194.
  13. Carli P, de Giorgi V, Chiarugi A, et al. Addition of dermoscopy to conventional naked-eye examination in melanoma screening: a randomized study. J Am Acad Dermatol. 2004;50:683-689.
  14. Carli P, de Giorgi V, Crocetti E, et al. Improvement of malignant/benign ratio in excised melanocytic lesions in the ‘dermoscopy era’: a retrospective study 1997-2001. Br J Dermatol. 2004;150:687-692.
  15. Saida T, Koga H, Uhara H. Key points in dermoscopic differentiation between early acral melanoma and acral nevus. J Dermatol. 2011;38:25-34.
  16. Ishihara Y, Saida T, Miyazaki A, et al. Early acral melanoma in situ: correlation between the parallel ridge pattern on dermoscopy and microscopic features. Am J Dermatopathol. 2006;28:21-27.
  17. Saida T, Miyazaki A, Oguchi S, et al. Significance of dermoscopic patterns in detecting malignant melanoma on acral volar skin: results of a multicenter study in Japan. Arch Dermatol. 2004;140:1233-1238.
  18. Koga H, Saida T. Revised 3-step dermoscopic algorithm for the management of acral melanocytic lesions. Arch Dermatol. 2011;147:741-743.
  19. Lallas A, Sgouros D, Zalaudek I, et al. Palmar and plantar melanomas differ for sex prevalence and tumor thickness but not for dermoscopic patterns. Melanoma Res. 2014;24:83-87.
  20. Lallas A, Kyrgidis A, Koga H, et al. The BRAAFF checklist: a new dermoscopic algorithm for diagnosing acral melanoma. Br J Dermatol. 2015;173:1041-1049.
  21. Saida T, Koga H. Dermoscopic patterns of acral melanocytic nevi: their variations, changes, and significance. Arch Dermatol. 2007;143:1423-1426.
  22. Miyazaki A, Saida T, Koga H, et al. Anatomical and histopathological correlates of the dermoscopic patterns seen in melanocytic nevi on the sole: a retrospective study. J Am Acad Dermatol. 2005;53:230-236.
  23. Watanabe S, Sawada M, Ishizaki S, et al. Comparison of dermatoscopic images of acral lentiginous melanoma and acral melanocytic nevus occurring on body weight-bearing areas. Dermatol Pract Concept. 2014;4:47-50.
  24. Costello CM, Ghanavatian S, Temkit M, et al. Educational and practice gaps in the management of volar melanocytic lesions. J Eur Acad Dermatol Venereol. 2018;32:1450-1455.
  25. Alikhan A, Ibrahimi OA, Eisen DB. Congenital melanocytic nevi: where are we now? part I. clinical presentation, epidemiology, pathogenesis, histology, malignant transformation, and neurocutaneous melanosis. J Am Acad Dermatol. 2012;67:495.e1-495.e17; quiz 512-514.
  26. Minagawa A, Koga H, Saida T. Dermoscopic characteristics of congenital melanocytic nevi affecting acral volar skin. Arch Dermatol. 2011;147:809-813.
Issue
Cutis - 103(5)
Issue
Cutis - 103(5)
Page Number
274-276
Page Number
274-276
Publications
Publications
Topics
Article Type
Display Headline
Dermoscopic Patterns of Acral Melanocytic Lesions in Skin of Color
Display Headline
Dermoscopic Patterns of Acral Melanocytic Lesions in Skin of Color
Sections
Inside the Article

Practice Points

  • Dermatologists should be familiar with common dermoscopic patterns seen at acral sites in patients with skin of color as well as the most up-to-date diagnostic algorithms.
  • Acral lentiginous melanoma should be strongly suspected if dermoscopy reveals a parallel ridge pattern or if dermoscopy of volar skin reveals a lack of typical dermoscopic patterns in lesions with a diameter greater than 7 mm.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Managing Postinflammatory Hyperpigmentation in Pediatric Patients With Skin of Color

Article Type
Changed
Thu, 10/29/2020 - 14:50
Display Headline
Managing Postinflammatory Hyperpigmentation in Pediatric Patients With Skin of Color
In Collaboration With the Skin of Color Society

Postnflammatory hyperpigmentation (PIH) is an acquired hypermelanosis that can occur in children and adults following an inflammatory cutaneous disease or trauma. Postinflammatory hyperpigmentation may last for months to even years. Although PIH may occur in all skin types, it is more common and presents with greater severity and intensity in individuals with skin of color. By the year 2050, 1 in 3 US residents is projected to be Hispanic.1 It is projected that by 2044, non-Hispanic white individuals (all ages) will make up less than 50% of the US population.2 Currently, the majority of the US residents younger than 18 years are minorities. The majority minority population in the United States already exists in those younger than 18 years and is predicted to occur in the adult population by 2044.2

Effective treatment options and management strategies for PIH in adults with skin of color have been described in the literature.3 Due to a paucity of research, the approach to management of PIH in children with skin of color has been based on clinical experience and lessons learned from adult patients. This article focuses on management of PIH in pediatric patients with skin of color, which includes black/African American, African-Caribbean, Hispanic, Asian, Pacific Islander, and American Indian individuals.

Underlying Inflammatory Dermatoses Resulting in PIH

There are numerous conditions that may result in PIH, including but not limited to atopic dermatitis (AD), acne, arthropod bites, and injuries to the skin. Postinflammatory hyperpigmentation may have more of a psychological impact than the inciting disease or injury itself. The most important step in the approach to managing PIH is treating the underlying inflammatory condition that caused the pigmentation.

Parents/guardians may report a chief concern of dark spots, manchas (stains), blemishes, or stains on the skin, often with no mention of a coexisting inflammatory dermatosis. Parents/guardians of children with skin of color often have personally experienced PIH and may be determined to shield their children from similar angst associated with the condition. Although physicians may see just another pediatric patient with PIH, the child’s parents/guardians may see a condition that will be readily perceptible during major life events, such as the child’s prom or even his/her wedding day. Promptly diagnosing and instituting early treatment of inflammatory conditions associated with PIH may accelerate resolution and prevent worsening of the pigmentation.3

Select inflammatory dermatoses that are common in children with skin of color and may lead to PIH are highlighted below. Although this list is not comprehensive, the approach and management strategies should prompt creation of plans that keep PIH in mind when treating primary inflammatory skin diseases.

Atopic Dermatitis
Atopic dermatitis may induce PIH or hypopigmentation of the skin in children with skin of color. Developing a plan for AD flare prevention, as well as management of mild, moderate, and severe AD flares, is imperative in pediatric patients. Prevention plans should include gentle skin care, twice-daily application of emollients to the full body, and reduction of Staphylococcus aureus loads on the skin. The treatment action plan for mild to moderate flares may include topical corticosteroids, immunomodulators, and nonsteroidal agents. Treatment options for severe AD or patients who were unsuccessfully treated with other therapies may include phototherapy, biologics, and methotrexate, among others.4 Creating action plans for AD flares is a vital step in the prevention of PIH in patients with skin of color. Additionally, PIH should not be considered a sign of AD treatment failure.

 

 

Acne
Acne is a common skin disorder seen in patients with skin of color.5 A prospective observational study found that 34.3% of 683 children aged 9 to 14 years in a pediatric ambulatory clinic had acne.6 The number of preadolescents with acne is growing. Most cases are not associated with underlying endocrinopathy.7 With the growing population of children with skin of color in the United States along with the increasing childhood acne rate and subsequent inherent risk for hyperpigmentation, early acne interventions should be considered in pediatric acne patients with skin of color to reduce the impact of PIH in those at risk.

In a survey study of 313 adult acne patients with skin of color, 37.2% reported the presence of dark marks lasting 4 months or longer.5 Regardless of the severity of the acne, treatment should be initiated as tolerated in those with PIH. Adolescent acne patients with skin of color may develop PIH that is more severe and longer lasting than the acne itself.

The foundation for treatment of acne in adolescent skin of color patients is the same as those without skin of color, including topical retinoids, topical antibiotics, oral antibiotics, and isotretinoin when needed. Topical tretinoin, adapalene, azelaic acid, and tazarotene not only treat acne but also are a valuable part of the treatment armamentarium for PIH. Several studies in adults with skin of color have demonstrated improvement of PIH from the use of topical retinoids alone.8-10 Despite wanting to treat the acne aggressively, special guidance should be given to prevent retinoid dermatitis, which may lead to PIH.10 Demonstrating the application of the topical acne medications, discussing how to avoid potential side effects, and giving permission to skip applications, if needed, may empower families to make adjustments between visits to limit irritation that might prompt further PIH. Incorporating α-hydroxy acid–based cleansers, α-hydroxy acid–based chemical peels, or salicylic acid chemical peels may be warranted in the setting of intense PIH. Selecting treatments that not only help the inflammatory disease leading to the PIH but also can help improve the pigmentation are preferred; however, the risks and benefits have to be weighed because many treatments that work well for PIH also may cause irritation, leading to new or worsening PIH.

Arthropod Bites
Arthropod bites cause inflamed pruritic papules and nodules, and the resulting PIH in those with darker skin types may be quite dramatic. Parents/guardians should be instructed to have a low-potency topical corticosteroid on hand to use on bites for a few days when they appear, which will not only help with the inflammation associated with the bite but also will help decrease pruritus and subsequently skin injury from scratching. In homes with pets, checking animals routinely for fleas and other infestations is helpful. In the setting of repeated arthropod bites in the spring and summer, applying bug repellant with 10% to 30% DEET (N,N-diethyl-meta-toluamide) on the child’s clothing and exposed body areas before playing outside or in the morning before school or camp may prevent some bites. There are DEET alternatives, such as picaridin, that may be used. Product instructions should be followed when using insect repellants in the pediatric population.11

PIH Management Strategies

Gentle Skin Care Routine
There are misconceptions that areas of hyperpigmentation on the skin are caused by dirt and that scrubbing the skin harder may help to lighten the affected areas. Parents/guardians may report that the child’s skin looks dirty or, in the setting of acne, view dirt as the cause of the skin condition, which may prompt the patient to scrub the skin and the friction further worsens the PIH. Use of daily gentle cleansers and moisturizers is advised to keep the skin moisturized and free of further potential irritation and dryness that may prompt scratching or flares of the underlying condition.

Photoprotection
During the treatment course for PIH, using sun protection is helpful to prevent further darkening of the PIH areas. Sun protection may be in the form of broad-spectrum sunscreen, hats, or sun-protective clothing. Patients should be encouraged to apply sunscreen daily and to reapply every 2 hours and after water-based activities.12 For pediatric and adolescent populations, practicing sun-protective behaviors before school or outdoor activities also should be advised, as many families only think about sun protection in the setting of sunny vacation activities. Research has demonstrated that individuals with skin of color may not realize that they can be affected by skin cancer,13 thus they may not have any experience selecting, applying, or regularly using sunscreens. Products that do not leave a white hue on the skin are suggested for adolescents who may be sensitive about their appearance following sunscreen application.

 

 

Skin Lightening Treatments

Although the most important therapy for PIH is to treat the underlying inflammatory conditions, some parents/guardians may desire additional options due to the extent of involvement of the PIH, its psychological impact on the child, or adverse effect on the child’s quality of life.14 In adolescents, incorporating an α-hydroxy acid–based cleanser, glycolic acid chemical peels, salicylic acid chemical peels, and topical cosmeceuticals may be warranted in the setting of intense PIH and acne. However, irritation may lead to further dyspigmentation.

Topical ammonium lactate 12% is lactic acid neutralized with ammonium hydroxide that is formulated as a lotion or a cream. It is used to hydrate dry skin and may decrease corneocyte cohesion.15 Topical ammonium lactate also has been used anecdotally for PIH on the body during periods of watchful waiting.

Topical hydroquinone, the gold standard for treating hyperpigmentation,3,16 is not approved in children, but some parents/guardians elect to utilize hydroquinone off label to accelerate the clearing of distressing PIH in adolescents. Careful consideration including a discussion of potential risks and alternatives (eg, watchful waiting) should be highlighted.

In the setting of chronic inflammatory conditions that recur and remit, potentially irritating topical treatments should be used only during periods when symptoms of inflammation such as itching or erythema are absent.

Conclusion

Despite the best management efforts, PIH in some patients with skin of color may be present for months to years. In the pediatric skin of color population, treatment of the underlying inflammatory condition, gentle skin care, use of photoprotection, and time may be all that is needed for PIH resolution. With their parent/guardians’ consent, adolescents distressed by PIH may decide to pursue more aggressive, potentially irritating treatments. Above all, the most important management in the setting of PIH is to treat the underlying inflammatory condition causing the PIH and set reasonable expectations. For challenging cases, pediatric dermatologists with special expertise in treating pediatric and adolescent patients with skin of color may be consulted.

References
  1. Broughton A. Minorities expected to be majority in 2050. CNN. August 13, 2008.  Accessed January 2, 2019.
  2. Colby SL, Ortman JM. Projections of the Size and Composition of the US Population: 2014 to 2060. Washington, DC: US Census Bureau; 2014. Current Population Reports, P25-1143. Published March 2015. Accessed January 23, 2019.
  3. Davis EC, Callender VD. Postinflammatory hyperpigmentation: a review of the epidemiology, clinical features, and treatment options in skin of color. J Clin Aesthet Dermatol2010;3:20-31.
  4. Eichenfield LF, Ahluwalia J, Waldman A, et al. Current guidelines for the evaluation and management of atopic dermatitis: a comparison of the Joint Task Force Practice Parameter and American Academy of Dermatology guidelines. J Allergy Clin Immunol. 2017;139(4S):S49-S57.
  5. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl):S98-S106.
  6. Napolitano M, Ruggiero G, Monfrecola G, et al. Acne prevalence in 9 to 14-year-old patients attending pediatric ambulatory clinics in Italy. Int J Dermatol. 2018;57:1320-1323.
  7. Mancini AJ, Baldwin HE, Eichenfield LF. Acne life cycle: the spectrum of pediatric disease. Semin Cutan Med Surg 2011;30:2-5.
  8. Lowe NJ, Rizk D, Grimes P, et al. Azelaic acid 20% cream in the treatment of facial hyperpigmentation in darker-skinned patients. Clin Ther. 1998;20:945-959.
  9. Grimes P, Callender V. Tazarotene cream for postinflammatory hyperpigmentation and acne vulgaris in darker skin: a double-blind, randomized, vehicle-controlled study. Cutis2006;77:45-50.
  10. Bulengo-Ransby SM, Griffiths CE, Kimbrough-Green CK, et al. Topical tretinoin (retinoid acid) therapy for hyperpigmented lesions caused by inflammation of the skin in black patients. N Engl J Med. 1993;328:1438-1443.
  11. American Academy of Pediatrics. Choosing an insect repellent for your child. Healthy Children website. Updated July 18, 2018. Accessed January 8, 2019.
  12. Agbai ON, Buster K, Sanchez M, et al. Skin cancer and photoprotection in people of color: a review and recommendations for physicians and the public. J Am Acad Dermatol. 2014;70:312-317.
  13. Buster KJ, You Z, Fouad M, et al. Skin cancer risk perceptions: a comparison across ethnicity, age, education, gender, and income. J Am Acad Dermatol. 2012;66:771-779.
  14. Downie J. Help prevent and reverse post-inflammatory hyperpigmentation. Pract Dermatol Pediatr. May/June 2011:12-14. Accessed January 18, 2019.
  15. Ammonium lactate lotion 12% [package insert]. Bronx, New York: Perrigo New York, Inc; 2006.
  16. Grimes PE. Management of hyperpigmentation in darker racial ethnic groups. Semin Cutan Med Surg. 2009;28:77-85.
Article PDF
Author and Disclosure Information

From the Department of Dermatology, Lewis Katz School of Medicine, Temple University Hospital, Philadelphia, Pennsylvania.

Dr. Heath is a consultant for Unilever, a former advisory board member and speaker for Pfizer Inc, and owner of Heath Health.

Correspondence: Candrice R. Heath, MD, 1316 W Ontario St, Jones Hall, Philadelphia, PA 19140 (Candrice.Heath@Temple.edu).

Issue
Cutis - 103(2)
Publications
Topics
Page Number
71-73
Sections
Author and Disclosure Information

From the Department of Dermatology, Lewis Katz School of Medicine, Temple University Hospital, Philadelphia, Pennsylvania.

Dr. Heath is a consultant for Unilever, a former advisory board member and speaker for Pfizer Inc, and owner of Heath Health.

Correspondence: Candrice R. Heath, MD, 1316 W Ontario St, Jones Hall, Philadelphia, PA 19140 (Candrice.Heath@Temple.edu).

Author and Disclosure Information

From the Department of Dermatology, Lewis Katz School of Medicine, Temple University Hospital, Philadelphia, Pennsylvania.

Dr. Heath is a consultant for Unilever, a former advisory board member and speaker for Pfizer Inc, and owner of Heath Health.

Correspondence: Candrice R. Heath, MD, 1316 W Ontario St, Jones Hall, Philadelphia, PA 19140 (Candrice.Heath@Temple.edu).

Article PDF
Article PDF
In Collaboration With the Skin of Color Society
In Collaboration With the Skin of Color Society

Postnflammatory hyperpigmentation (PIH) is an acquired hypermelanosis that can occur in children and adults following an inflammatory cutaneous disease or trauma. Postinflammatory hyperpigmentation may last for months to even years. Although PIH may occur in all skin types, it is more common and presents with greater severity and intensity in individuals with skin of color. By the year 2050, 1 in 3 US residents is projected to be Hispanic.1 It is projected that by 2044, non-Hispanic white individuals (all ages) will make up less than 50% of the US population.2 Currently, the majority of the US residents younger than 18 years are minorities. The majority minority population in the United States already exists in those younger than 18 years and is predicted to occur in the adult population by 2044.2

Effective treatment options and management strategies for PIH in adults with skin of color have been described in the literature.3 Due to a paucity of research, the approach to management of PIH in children with skin of color has been based on clinical experience and lessons learned from adult patients. This article focuses on management of PIH in pediatric patients with skin of color, which includes black/African American, African-Caribbean, Hispanic, Asian, Pacific Islander, and American Indian individuals.

Underlying Inflammatory Dermatoses Resulting in PIH

There are numerous conditions that may result in PIH, including but not limited to atopic dermatitis (AD), acne, arthropod bites, and injuries to the skin. Postinflammatory hyperpigmentation may have more of a psychological impact than the inciting disease or injury itself. The most important step in the approach to managing PIH is treating the underlying inflammatory condition that caused the pigmentation.

Parents/guardians may report a chief concern of dark spots, manchas (stains), blemishes, or stains on the skin, often with no mention of a coexisting inflammatory dermatosis. Parents/guardians of children with skin of color often have personally experienced PIH and may be determined to shield their children from similar angst associated with the condition. Although physicians may see just another pediatric patient with PIH, the child’s parents/guardians may see a condition that will be readily perceptible during major life events, such as the child’s prom or even his/her wedding day. Promptly diagnosing and instituting early treatment of inflammatory conditions associated with PIH may accelerate resolution and prevent worsening of the pigmentation.3

Select inflammatory dermatoses that are common in children with skin of color and may lead to PIH are highlighted below. Although this list is not comprehensive, the approach and management strategies should prompt creation of plans that keep PIH in mind when treating primary inflammatory skin diseases.

Atopic Dermatitis
Atopic dermatitis may induce PIH or hypopigmentation of the skin in children with skin of color. Developing a plan for AD flare prevention, as well as management of mild, moderate, and severe AD flares, is imperative in pediatric patients. Prevention plans should include gentle skin care, twice-daily application of emollients to the full body, and reduction of Staphylococcus aureus loads on the skin. The treatment action plan for mild to moderate flares may include topical corticosteroids, immunomodulators, and nonsteroidal agents. Treatment options for severe AD or patients who were unsuccessfully treated with other therapies may include phototherapy, biologics, and methotrexate, among others.4 Creating action plans for AD flares is a vital step in the prevention of PIH in patients with skin of color. Additionally, PIH should not be considered a sign of AD treatment failure.

 

 

Acne
Acne is a common skin disorder seen in patients with skin of color.5 A prospective observational study found that 34.3% of 683 children aged 9 to 14 years in a pediatric ambulatory clinic had acne.6 The number of preadolescents with acne is growing. Most cases are not associated with underlying endocrinopathy.7 With the growing population of children with skin of color in the United States along with the increasing childhood acne rate and subsequent inherent risk for hyperpigmentation, early acne interventions should be considered in pediatric acne patients with skin of color to reduce the impact of PIH in those at risk.

In a survey study of 313 adult acne patients with skin of color, 37.2% reported the presence of dark marks lasting 4 months or longer.5 Regardless of the severity of the acne, treatment should be initiated as tolerated in those with PIH. Adolescent acne patients with skin of color may develop PIH that is more severe and longer lasting than the acne itself.

The foundation for treatment of acne in adolescent skin of color patients is the same as those without skin of color, including topical retinoids, topical antibiotics, oral antibiotics, and isotretinoin when needed. Topical tretinoin, adapalene, azelaic acid, and tazarotene not only treat acne but also are a valuable part of the treatment armamentarium for PIH. Several studies in adults with skin of color have demonstrated improvement of PIH from the use of topical retinoids alone.8-10 Despite wanting to treat the acne aggressively, special guidance should be given to prevent retinoid dermatitis, which may lead to PIH.10 Demonstrating the application of the topical acne medications, discussing how to avoid potential side effects, and giving permission to skip applications, if needed, may empower families to make adjustments between visits to limit irritation that might prompt further PIH. Incorporating α-hydroxy acid–based cleansers, α-hydroxy acid–based chemical peels, or salicylic acid chemical peels may be warranted in the setting of intense PIH. Selecting treatments that not only help the inflammatory disease leading to the PIH but also can help improve the pigmentation are preferred; however, the risks and benefits have to be weighed because many treatments that work well for PIH also may cause irritation, leading to new or worsening PIH.

Arthropod Bites
Arthropod bites cause inflamed pruritic papules and nodules, and the resulting PIH in those with darker skin types may be quite dramatic. Parents/guardians should be instructed to have a low-potency topical corticosteroid on hand to use on bites for a few days when they appear, which will not only help with the inflammation associated with the bite but also will help decrease pruritus and subsequently skin injury from scratching. In homes with pets, checking animals routinely for fleas and other infestations is helpful. In the setting of repeated arthropod bites in the spring and summer, applying bug repellant with 10% to 30% DEET (N,N-diethyl-meta-toluamide) on the child’s clothing and exposed body areas before playing outside or in the morning before school or camp may prevent some bites. There are DEET alternatives, such as picaridin, that may be used. Product instructions should be followed when using insect repellants in the pediatric population.11

PIH Management Strategies

Gentle Skin Care Routine
There are misconceptions that areas of hyperpigmentation on the skin are caused by dirt and that scrubbing the skin harder may help to lighten the affected areas. Parents/guardians may report that the child’s skin looks dirty or, in the setting of acne, view dirt as the cause of the skin condition, which may prompt the patient to scrub the skin and the friction further worsens the PIH. Use of daily gentle cleansers and moisturizers is advised to keep the skin moisturized and free of further potential irritation and dryness that may prompt scratching or flares of the underlying condition.

Photoprotection
During the treatment course for PIH, using sun protection is helpful to prevent further darkening of the PIH areas. Sun protection may be in the form of broad-spectrum sunscreen, hats, or sun-protective clothing. Patients should be encouraged to apply sunscreen daily and to reapply every 2 hours and after water-based activities.12 For pediatric and adolescent populations, practicing sun-protective behaviors before school or outdoor activities also should be advised, as many families only think about sun protection in the setting of sunny vacation activities. Research has demonstrated that individuals with skin of color may not realize that they can be affected by skin cancer,13 thus they may not have any experience selecting, applying, or regularly using sunscreens. Products that do not leave a white hue on the skin are suggested for adolescents who may be sensitive about their appearance following sunscreen application.

 

 

Skin Lightening Treatments

Although the most important therapy for PIH is to treat the underlying inflammatory conditions, some parents/guardians may desire additional options due to the extent of involvement of the PIH, its psychological impact on the child, or adverse effect on the child’s quality of life.14 In adolescents, incorporating an α-hydroxy acid–based cleanser, glycolic acid chemical peels, salicylic acid chemical peels, and topical cosmeceuticals may be warranted in the setting of intense PIH and acne. However, irritation may lead to further dyspigmentation.

Topical ammonium lactate 12% is lactic acid neutralized with ammonium hydroxide that is formulated as a lotion or a cream. It is used to hydrate dry skin and may decrease corneocyte cohesion.15 Topical ammonium lactate also has been used anecdotally for PIH on the body during periods of watchful waiting.

Topical hydroquinone, the gold standard for treating hyperpigmentation,3,16 is not approved in children, but some parents/guardians elect to utilize hydroquinone off label to accelerate the clearing of distressing PIH in adolescents. Careful consideration including a discussion of potential risks and alternatives (eg, watchful waiting) should be highlighted.

In the setting of chronic inflammatory conditions that recur and remit, potentially irritating topical treatments should be used only during periods when symptoms of inflammation such as itching or erythema are absent.

Conclusion

Despite the best management efforts, PIH in some patients with skin of color may be present for months to years. In the pediatric skin of color population, treatment of the underlying inflammatory condition, gentle skin care, use of photoprotection, and time may be all that is needed for PIH resolution. With their parent/guardians’ consent, adolescents distressed by PIH may decide to pursue more aggressive, potentially irritating treatments. Above all, the most important management in the setting of PIH is to treat the underlying inflammatory condition causing the PIH and set reasonable expectations. For challenging cases, pediatric dermatologists with special expertise in treating pediatric and adolescent patients with skin of color may be consulted.

Postnflammatory hyperpigmentation (PIH) is an acquired hypermelanosis that can occur in children and adults following an inflammatory cutaneous disease or trauma. Postinflammatory hyperpigmentation may last for months to even years. Although PIH may occur in all skin types, it is more common and presents with greater severity and intensity in individuals with skin of color. By the year 2050, 1 in 3 US residents is projected to be Hispanic.1 It is projected that by 2044, non-Hispanic white individuals (all ages) will make up less than 50% of the US population.2 Currently, the majority of the US residents younger than 18 years are minorities. The majority minority population in the United States already exists in those younger than 18 years and is predicted to occur in the adult population by 2044.2

Effective treatment options and management strategies for PIH in adults with skin of color have been described in the literature.3 Due to a paucity of research, the approach to management of PIH in children with skin of color has been based on clinical experience and lessons learned from adult patients. This article focuses on management of PIH in pediatric patients with skin of color, which includes black/African American, African-Caribbean, Hispanic, Asian, Pacific Islander, and American Indian individuals.

Underlying Inflammatory Dermatoses Resulting in PIH

There are numerous conditions that may result in PIH, including but not limited to atopic dermatitis (AD), acne, arthropod bites, and injuries to the skin. Postinflammatory hyperpigmentation may have more of a psychological impact than the inciting disease or injury itself. The most important step in the approach to managing PIH is treating the underlying inflammatory condition that caused the pigmentation.

Parents/guardians may report a chief concern of dark spots, manchas (stains), blemishes, or stains on the skin, often with no mention of a coexisting inflammatory dermatosis. Parents/guardians of children with skin of color often have personally experienced PIH and may be determined to shield their children from similar angst associated with the condition. Although physicians may see just another pediatric patient with PIH, the child’s parents/guardians may see a condition that will be readily perceptible during major life events, such as the child’s prom or even his/her wedding day. Promptly diagnosing and instituting early treatment of inflammatory conditions associated with PIH may accelerate resolution and prevent worsening of the pigmentation.3

Select inflammatory dermatoses that are common in children with skin of color and may lead to PIH are highlighted below. Although this list is not comprehensive, the approach and management strategies should prompt creation of plans that keep PIH in mind when treating primary inflammatory skin diseases.

Atopic Dermatitis
Atopic dermatitis may induce PIH or hypopigmentation of the skin in children with skin of color. Developing a plan for AD flare prevention, as well as management of mild, moderate, and severe AD flares, is imperative in pediatric patients. Prevention plans should include gentle skin care, twice-daily application of emollients to the full body, and reduction of Staphylococcus aureus loads on the skin. The treatment action plan for mild to moderate flares may include topical corticosteroids, immunomodulators, and nonsteroidal agents. Treatment options for severe AD or patients who were unsuccessfully treated with other therapies may include phototherapy, biologics, and methotrexate, among others.4 Creating action plans for AD flares is a vital step in the prevention of PIH in patients with skin of color. Additionally, PIH should not be considered a sign of AD treatment failure.

 

 

Acne
Acne is a common skin disorder seen in patients with skin of color.5 A prospective observational study found that 34.3% of 683 children aged 9 to 14 years in a pediatric ambulatory clinic had acne.6 The number of preadolescents with acne is growing. Most cases are not associated with underlying endocrinopathy.7 With the growing population of children with skin of color in the United States along with the increasing childhood acne rate and subsequent inherent risk for hyperpigmentation, early acne interventions should be considered in pediatric acne patients with skin of color to reduce the impact of PIH in those at risk.

In a survey study of 313 adult acne patients with skin of color, 37.2% reported the presence of dark marks lasting 4 months or longer.5 Regardless of the severity of the acne, treatment should be initiated as tolerated in those with PIH. Adolescent acne patients with skin of color may develop PIH that is more severe and longer lasting than the acne itself.

The foundation for treatment of acne in adolescent skin of color patients is the same as those without skin of color, including topical retinoids, topical antibiotics, oral antibiotics, and isotretinoin when needed. Topical tretinoin, adapalene, azelaic acid, and tazarotene not only treat acne but also are a valuable part of the treatment armamentarium for PIH. Several studies in adults with skin of color have demonstrated improvement of PIH from the use of topical retinoids alone.8-10 Despite wanting to treat the acne aggressively, special guidance should be given to prevent retinoid dermatitis, which may lead to PIH.10 Demonstrating the application of the topical acne medications, discussing how to avoid potential side effects, and giving permission to skip applications, if needed, may empower families to make adjustments between visits to limit irritation that might prompt further PIH. Incorporating α-hydroxy acid–based cleansers, α-hydroxy acid–based chemical peels, or salicylic acid chemical peels may be warranted in the setting of intense PIH. Selecting treatments that not only help the inflammatory disease leading to the PIH but also can help improve the pigmentation are preferred; however, the risks and benefits have to be weighed because many treatments that work well for PIH also may cause irritation, leading to new or worsening PIH.

Arthropod Bites
Arthropod bites cause inflamed pruritic papules and nodules, and the resulting PIH in those with darker skin types may be quite dramatic. Parents/guardians should be instructed to have a low-potency topical corticosteroid on hand to use on bites for a few days when they appear, which will not only help with the inflammation associated with the bite but also will help decrease pruritus and subsequently skin injury from scratching. In homes with pets, checking animals routinely for fleas and other infestations is helpful. In the setting of repeated arthropod bites in the spring and summer, applying bug repellant with 10% to 30% DEET (N,N-diethyl-meta-toluamide) on the child’s clothing and exposed body areas before playing outside or in the morning before school or camp may prevent some bites. There are DEET alternatives, such as picaridin, that may be used. Product instructions should be followed when using insect repellants in the pediatric population.11

PIH Management Strategies

Gentle Skin Care Routine
There are misconceptions that areas of hyperpigmentation on the skin are caused by dirt and that scrubbing the skin harder may help to lighten the affected areas. Parents/guardians may report that the child’s skin looks dirty or, in the setting of acne, view dirt as the cause of the skin condition, which may prompt the patient to scrub the skin and the friction further worsens the PIH. Use of daily gentle cleansers and moisturizers is advised to keep the skin moisturized and free of further potential irritation and dryness that may prompt scratching or flares of the underlying condition.

Photoprotection
During the treatment course for PIH, using sun protection is helpful to prevent further darkening of the PIH areas. Sun protection may be in the form of broad-spectrum sunscreen, hats, or sun-protective clothing. Patients should be encouraged to apply sunscreen daily and to reapply every 2 hours and after water-based activities.12 For pediatric and adolescent populations, practicing sun-protective behaviors before school or outdoor activities also should be advised, as many families only think about sun protection in the setting of sunny vacation activities. Research has demonstrated that individuals with skin of color may not realize that they can be affected by skin cancer,13 thus they may not have any experience selecting, applying, or regularly using sunscreens. Products that do not leave a white hue on the skin are suggested for adolescents who may be sensitive about their appearance following sunscreen application.

 

 

Skin Lightening Treatments

Although the most important therapy for PIH is to treat the underlying inflammatory conditions, some parents/guardians may desire additional options due to the extent of involvement of the PIH, its psychological impact on the child, or adverse effect on the child’s quality of life.14 In adolescents, incorporating an α-hydroxy acid–based cleanser, glycolic acid chemical peels, salicylic acid chemical peels, and topical cosmeceuticals may be warranted in the setting of intense PIH and acne. However, irritation may lead to further dyspigmentation.

Topical ammonium lactate 12% is lactic acid neutralized with ammonium hydroxide that is formulated as a lotion or a cream. It is used to hydrate dry skin and may decrease corneocyte cohesion.15 Topical ammonium lactate also has been used anecdotally for PIH on the body during periods of watchful waiting.

Topical hydroquinone, the gold standard for treating hyperpigmentation,3,16 is not approved in children, but some parents/guardians elect to utilize hydroquinone off label to accelerate the clearing of distressing PIH in adolescents. Careful consideration including a discussion of potential risks and alternatives (eg, watchful waiting) should be highlighted.

In the setting of chronic inflammatory conditions that recur and remit, potentially irritating topical treatments should be used only during periods when symptoms of inflammation such as itching or erythema are absent.

Conclusion

Despite the best management efforts, PIH in some patients with skin of color may be present for months to years. In the pediatric skin of color population, treatment of the underlying inflammatory condition, gentle skin care, use of photoprotection, and time may be all that is needed for PIH resolution. With their parent/guardians’ consent, adolescents distressed by PIH may decide to pursue more aggressive, potentially irritating treatments. Above all, the most important management in the setting of PIH is to treat the underlying inflammatory condition causing the PIH and set reasonable expectations. For challenging cases, pediatric dermatologists with special expertise in treating pediatric and adolescent patients with skin of color may be consulted.

References
  1. Broughton A. Minorities expected to be majority in 2050. CNN. August 13, 2008.  Accessed January 2, 2019.
  2. Colby SL, Ortman JM. Projections of the Size and Composition of the US Population: 2014 to 2060. Washington, DC: US Census Bureau; 2014. Current Population Reports, P25-1143. Published March 2015. Accessed January 23, 2019.
  3. Davis EC, Callender VD. Postinflammatory hyperpigmentation: a review of the epidemiology, clinical features, and treatment options in skin of color. J Clin Aesthet Dermatol2010;3:20-31.
  4. Eichenfield LF, Ahluwalia J, Waldman A, et al. Current guidelines for the evaluation and management of atopic dermatitis: a comparison of the Joint Task Force Practice Parameter and American Academy of Dermatology guidelines. J Allergy Clin Immunol. 2017;139(4S):S49-S57.
  5. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl):S98-S106.
  6. Napolitano M, Ruggiero G, Monfrecola G, et al. Acne prevalence in 9 to 14-year-old patients attending pediatric ambulatory clinics in Italy. Int J Dermatol. 2018;57:1320-1323.
  7. Mancini AJ, Baldwin HE, Eichenfield LF. Acne life cycle: the spectrum of pediatric disease. Semin Cutan Med Surg 2011;30:2-5.
  8. Lowe NJ, Rizk D, Grimes P, et al. Azelaic acid 20% cream in the treatment of facial hyperpigmentation in darker-skinned patients. Clin Ther. 1998;20:945-959.
  9. Grimes P, Callender V. Tazarotene cream for postinflammatory hyperpigmentation and acne vulgaris in darker skin: a double-blind, randomized, vehicle-controlled study. Cutis2006;77:45-50.
  10. Bulengo-Ransby SM, Griffiths CE, Kimbrough-Green CK, et al. Topical tretinoin (retinoid acid) therapy for hyperpigmented lesions caused by inflammation of the skin in black patients. N Engl J Med. 1993;328:1438-1443.
  11. American Academy of Pediatrics. Choosing an insect repellent for your child. Healthy Children website. Updated July 18, 2018. Accessed January 8, 2019.
  12. Agbai ON, Buster K, Sanchez M, et al. Skin cancer and photoprotection in people of color: a review and recommendations for physicians and the public. J Am Acad Dermatol. 2014;70:312-317.
  13. Buster KJ, You Z, Fouad M, et al. Skin cancer risk perceptions: a comparison across ethnicity, age, education, gender, and income. J Am Acad Dermatol. 2012;66:771-779.
  14. Downie J. Help prevent and reverse post-inflammatory hyperpigmentation. Pract Dermatol Pediatr. May/June 2011:12-14. Accessed January 18, 2019.
  15. Ammonium lactate lotion 12% [package insert]. Bronx, New York: Perrigo New York, Inc; 2006.
  16. Grimes PE. Management of hyperpigmentation in darker racial ethnic groups. Semin Cutan Med Surg. 2009;28:77-85.
References
  1. Broughton A. Minorities expected to be majority in 2050. CNN. August 13, 2008.  Accessed January 2, 2019.
  2. Colby SL, Ortman JM. Projections of the Size and Composition of the US Population: 2014 to 2060. Washington, DC: US Census Bureau; 2014. Current Population Reports, P25-1143. Published March 2015. Accessed January 23, 2019.
  3. Davis EC, Callender VD. Postinflammatory hyperpigmentation: a review of the epidemiology, clinical features, and treatment options in skin of color. J Clin Aesthet Dermatol2010;3:20-31.
  4. Eichenfield LF, Ahluwalia J, Waldman A, et al. Current guidelines for the evaluation and management of atopic dermatitis: a comparison of the Joint Task Force Practice Parameter and American Academy of Dermatology guidelines. J Allergy Clin Immunol. 2017;139(4S):S49-S57.
  5. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl):S98-S106.
  6. Napolitano M, Ruggiero G, Monfrecola G, et al. Acne prevalence in 9 to 14-year-old patients attending pediatric ambulatory clinics in Italy. Int J Dermatol. 2018;57:1320-1323.
  7. Mancini AJ, Baldwin HE, Eichenfield LF. Acne life cycle: the spectrum of pediatric disease. Semin Cutan Med Surg 2011;30:2-5.
  8. Lowe NJ, Rizk D, Grimes P, et al. Azelaic acid 20% cream in the treatment of facial hyperpigmentation in darker-skinned patients. Clin Ther. 1998;20:945-959.
  9. Grimes P, Callender V. Tazarotene cream for postinflammatory hyperpigmentation and acne vulgaris in darker skin: a double-blind, randomized, vehicle-controlled study. Cutis2006;77:45-50.
  10. Bulengo-Ransby SM, Griffiths CE, Kimbrough-Green CK, et al. Topical tretinoin (retinoid acid) therapy for hyperpigmented lesions caused by inflammation of the skin in black patients. N Engl J Med. 1993;328:1438-1443.
  11. American Academy of Pediatrics. Choosing an insect repellent for your child. Healthy Children website. Updated July 18, 2018. Accessed January 8, 2019.
  12. Agbai ON, Buster K, Sanchez M, et al. Skin cancer and photoprotection in people of color: a review and recommendations for physicians and the public. J Am Acad Dermatol. 2014;70:312-317.
  13. Buster KJ, You Z, Fouad M, et al. Skin cancer risk perceptions: a comparison across ethnicity, age, education, gender, and income. J Am Acad Dermatol. 2012;66:771-779.
  14. Downie J. Help prevent and reverse post-inflammatory hyperpigmentation. Pract Dermatol Pediatr. May/June 2011:12-14. Accessed January 18, 2019.
  15. Ammonium lactate lotion 12% [package insert]. Bronx, New York: Perrigo New York, Inc; 2006.
  16. Grimes PE. Management of hyperpigmentation in darker racial ethnic groups. Semin Cutan Med Surg. 2009;28:77-85.
Issue
Cutis - 103(2)
Issue
Cutis - 103(2)
Page Number
71-73
Page Number
71-73
Publications
Publications
Topics
Article Type
Display Headline
Managing Postinflammatory Hyperpigmentation in Pediatric Patients With Skin of Color
Display Headline
Managing Postinflammatory Hyperpigmentation in Pediatric Patients With Skin of Color
Sections
Inside the Article

Practice Points

  • The US population of children with skin of color is growing rapidly.
  • Treating the underlying inflammatory dermatosis is the most important step in managing postinflammatory hyperpigmentation (PIH); however, many pediatric PIH patients and their parents/guardians presenting with a chief concern of pigmentary changes are unaware of the associated inflammatory condition.
  • When appropriate, choose treatments for the underlying inflammatory condition that can simultaneously improve any existing PIH. Gentle skin care, avoidance of rubbing and scrubbing the skin, and photoprotection are essential to halt worsening of PIH.
  • Patients’ parents/guardians may consent to more aggressive PIH treatment in select cases (eg, emotional distress in adolescents).
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Frontal Fibrosing Alopecia: Cutaneous Associations in Women With Skin of Color

Article Type
Changed
Thu, 10/29/2020 - 14:51
Display Headline
Frontal Fibrosing Alopecia: Cutaneous Associations in Women With Skin of Color
In Collaboration with the Skin of Color Society

Frontal fibrosing alopecia (FFA) has been reported in association with lichen planus pigmentosus (LPP) and facial papules.1-3 Lichen planus pigmentosus is a variant of lichen planus that causes hyperpigmentation of the face, neck, and/or intertriginous areas that may be useful as a clinical indicator in the development of FFA.1 Facial papules in association with FFA are secondary to fibrosed vellus hairs.2,3 Currently, reports of concomitant FFA, LPP, and facial papules in women with skin of color are limited in the literature. This case series includes 5 women of color (Hispanic and black) who presented to our clinic with FFA and various cutaneous associations. A review of the current literature on cutaneous associations of FFA also is provided.

Case Reports

Patient 1
A 50-year-old Hispanic woman who was previously presumed to have melasma by an outside physician presented with pruritus of the scalp and eyebrows of 1 month’s duration. Physical examination revealed decreased frontal scalp hair density with perifollicular erythema and scale with thinning of the lateral eyebrows. Hyperpigmented coalesced macules (Figure 1A) and erythematous perifollicular papules were noted along the temples and on the perioral skin. Depressed forehead and temporal veins also were noted (Figure 1B). A biopsy of the scalp demonstrated perifollicular and perivascular lymphocytic inflammation and fibrosed hair follicles, and a biopsy of the perioral skin demonstrated perivascular lymphocytic inflammation with melanophages in the papillary dermis. A diagnosis of FFA with LPP was established with these biopsies.

Figure1
Figure 1. Frontal fibrosing alopecia with hyperpigmented coalesced macules around the mouth (A). Perifollicular papules on the temples (black arrow), erythematous perifollicular papules at the frontal hairline (blue arrow), and depressed veins on the forehead and temples (yellow arrows) also were noted (B).

Patient 2
A 61-year-old black woman presented with asymptomatic hair loss along the frontal hairline for an unknown duration. On physical examination the frontal scalp and lateral eyebrows demonstrated decreased hair density with loss of follicular ostia. Fine, flesh-colored, monomorphic papules were scattered along the forehead and temples, and ill-defined brown pigmentation was present along the forehead, temples, and cheeks. Biopsy of the frontal scalp demonstrated patchy lichenoid inflammation with decreased number of follicles with replacement by follicular scars, confirming the diagnosis of FFA.

Patient 3
A 47-year-old Hispanic woman presented with hair loss of the frontal scalp and bilateral eyebrows with associated burning of 2 years’ duration. Physical examination demonstrated recession of the frontotemporal hairline with scattered lone hairs and thinning of the eyebrows. Innumerable flesh-colored papules were present on the forehead and temples (Figure 2A). Glabellar and eyebrow erythema was noted (Figure 2B). Biopsy of the frontal scalp demonstrated decreased terminal anagen hair follicles with perifollicular lymphoid infiltrate and fibrosis, consistent with a diagnosis of FFA. The patient was started on oral hydroxychloroquine 400 mg once daily, and 3 months later hyperpigmentation of the forehead and perioral skin was noted. The patient reported that she had facial hyperpigmentation prior to starting hydroxychloroquine and declined a biopsy.

Figure2
Figure 2. Frontal fibrosing alopecia with recession of the temporal hairline with visible lone hairs (red arrow) and scattered flesh-colored papules on the temples (black arrows)(A). Glabellar and eyebrow erythema also was noted with flesh-colored papules on the forehead (black arrow)(B). The eyebrows were notably drawn in due to decreased hair density, and the central frontal hairline was recessed.

Patient 4
A 40-year-old black woman presented with brown pruritic macles of the face, neck, arms, and forearms of 4 years’ duration. She also reported hair loss on the frontal and occipital scalp, eyebrows, and arms. On physical examination, ill-defined brown macules and patches were noted on the neck (Figure 3), face, arms, and forearms. Decreased hair density was noted on the frontal and occipital scalp with follicular dropout and perifollicular hyperpigmentation. Biopsy of the scalp demonstrated perivascular lymphocytic inflammation with sparse anagen follicles and fibrous tracts, and biopsy of the neck revealed superficial perivascular inflammation with numerous melanophages in the upper dermis; these histopathologic findings were consistent with FFA and LPP, respectively.

Figure3
Figure 3. Diffuse and coalescing brown-gray macules and patches on the neck consistent with lichen planus pigmentosus.

Patient 5
A 46-year-old black woman with history of hair loss presented with hyperpigmentation of the face and neck of 2 years’ duration. On physical examination decreased hair density of the frontal and vertex scalp and lateral eyebrows was noted. Flesh-colored papules were noted on the forehead and cheeks, and confluent dark brown patches were present on the temples and neck. Three punch biopsies were performed. Biopsy of the scalp revealed lymphocytic inflammation with surrounding fibroplasia with overlapping features of FFA and central centrifugal cicatricial alopecia (Figure 4). Biopsy of the neck revealed vacuolar interface dermatitis. Additionally, biopsy of a facial papule revealed lichenoid inflammation involving a vellus hair follicle. Clinical and histopathological correlation confirmed the diagnosis of FFA with LPP and facial papules.

Figure4
Figure 4. Representative photograph demonstrating a diminished number of hair follicles with partial loss of sebaceous glands. There was perifollicular fibroplasia and interface inflammation along the basement membrane of the follicular epithelium with exocytosis of lymphocytes. Low-grade vacuolar alteration also was seen along the dermoepidermal junction (H&E, original magnification ×100).

 

 

Comment

Current understanding of FFA as a progressive, lymphocytic, scarring alopecia has expanded in recent years. Clinical observation suggests that the incidence of FFA is increasing4; however more epidemiologic data are needed. Frontal fibrosing alopecia presents clinically with symmetrical frontotemporal hair loss with lone hairs. Trichoscopy reveals perifollicular hyperkeratosis, perifollicular erythema, and follicular plugging in 72%, 66%, and 44% of cases, respectively.5 In one study (N=242), patients were classified into 3 clinical patterns of FFA: pattern I (linear) showed bandlike loss of frontal hair with normal density directly behind the hairline; pattern II (diffuse) showed loss of density behind the frontal hairline; and pattern III (double line) showed a pseudo–“fringe sign” appearance. The majority of patients were classified as either pattern I or II, with pattern II predicting a poorer prognosis.6

rontal fibrosing alopecia is increasingly recognized in men, with prevalence as high as 5%.1 Facial hair involvement, particularly of the upper lip and sideburns, is an important consideration in men.7 Most studies suggest that 80% to 90% of affected women are postmenopausal,8 though a case series presented by Dlova1 identified 27% of affected women as postmenopausal. The coexistence of premature menopause and hysterectomy in FFA patients suggests a hormonal contribution, but this association is still poorly understood.8 Epidemiologic data on ethnicity in FFA are sparse but suggest that white individuals are more likely to be affected. Frontal fibrosing alopecia also may be misdiagnosed as traction alopecia in Hispanic and black patients.8

It is prudent for physicians to assess for and recognize clinical clues to severe forms of FFA. A 2014 multicenter review of 355 patients identified 3 clinical entities that predicted more severe forms of FFA: eyelash loss (madarosis), loss of body hair, and facial papules.8 Madarosis occurs due to perifollicular inflammation and fibrosis of eyelash hair follicles. Similarly, perifollicular inflammation of body hair was present in 24% of patients (N=86), most commonly of the axillary and pubic hair. Facial papules form due to facial vellus hair inflammation and fibrosis and were identified in 14% of patients (N=49).8 These clinical findings may allow providers to predict more extensive clinical involvement of FFA.

Frontal fibrosing alopecia and LPP occur concomitantly in up 54% of patients, more commonly in darker-skinned patients.1,9,10 Lichen planus pigmentosus frequently occurs on the face and neck, most commonly in a diffuse pattern, though reticulated and macular patterns also have been identified.11 In some patients, LPP precedes the development of FFA and may be useful as a herald sign1; therefore, it is important for dermatologists to evaluate for signs of FFA when evaluating those with LPP. Thorough evaluation in patients with skin of color also is important because FFA may be misdiagnosed as traction alopecia.

Additional cutaneous associations of FFA include eyebrow loss, glabellar red dots, and prominent frontal veins. Eyebrow loss occurs secondary to fibrosis of eyebrow hair follicles and has been found in 40% to 80% of patients with FFA; it is thought to be associated with milder forms of FFA.8 Glabellar red dots correlate with histopathologic lymphocytic inflammation of vellus hair follicles.12 Additionally, frontal vein prominence has been described in FFA and is thought to be secondary to atrophy in this scarring process, perhaps worsened by local steroid treatments.13 Mucocutaneous lichen planus, rosacea, thyroid disease, vitiligo, and other autoimmune disorders also have been reported in patients with FFA.14

Conclusion

Concomitant FFA, LPP, and facial papules have been rarely reported and exemplify the spectrum of cutaneous associations with FFA, particularly in individuals with skin of color. Clinical variants and associations of FFA are broad, including predictors of poorer prognosis such as eyelash loss and vellus hair involvement seen as facial papules. Lichen planus pigmentosus is well described in association with FFA and may serve as a herald sign that frontal hair loss should not be mistaken for traction alopecia in early stages. Eyebrow loss is thought to represent milder disease. It is important for dermatologists to be aware of these findings to understand the breadth of this disease and for appropriate evaluation and management of patients with FFA.

References
  1. Dlova NC. Frontal fibrosing alopecia and lichen planus pigmentosus: is there a link? Br J Dermatol. 2013;168:439-432.
  2. Donati A, Molina L, Doche I, et al. Facial papules in frontal fibrosing alopecia: evidence of vellus follicle involvement. Arch Dermatol. 2011;147:1424-1427.
  3. Tan KT, Messenger AG. Frontal fibrosing alopecia: clinical presentations and prognosis. Br J Dermatol. 2009;160:75-79.
  4. Rudnicka L, Rakowska A. The increasing incidence of frontal fibrosing alopecia. in search of triggering factors. J Eur Acad Dermatol Venereol. 2017;31:1579-1580.
  5. Toledo-Pastrana T, Hernández MJ, Camacho Martínez FM. Perifollicular erythema as a trichoscopy sign of progression in frontal fibrosing alopecia. Int J Trichology. 2013;5:151-153.
  6. Moreno-Arrones OM, Saceda-Corralo D, Fonda-Pascual P, et al. Frontal fibrosing alopecia: clinical and prognostic classification. J Eur Acad Dermatol Venereol. 2017;31:1739-1745.
  7. Tolkachjov SN, Chaudhry HM, Camilleri MJ, et al. Frontal fibrosing alopecia among men: a clinicopathologic study of 7 cases. J Am Acad Dermatol. 2017;77:683-690.e2.
  8. Vañó-Galván S, Molina-Ruiz AM, Serrano-Falcón C, et al. Frontal fibrosing alopecia: a multicenter review of 355 patients. J Am Acad Dermatol. 2014;70:670-678.
  9. Berliner JG, McCalmont TH, Price VH, et al. Frontal fibrosing alopecia and lichen planus pigmentosus. J Am Acad Dermatol. 2014;71:E26-E27.
  10. Rao R, Sarda A, Khanna R, et al. Coexistence of frontal fibrosing alopecia with lichen planus pigmentosus. Int J Dermatol. 2014;53:622-624.
  11. Pirmez R, Duque-Estrada B, Donati A, et al. Clinical and dermoscopic features of lichen planus pigmentosus in 37 patients with frontal fibrosing alopecia. Br J Dermatol. 2016;175:1387-1390.
  12. Pirmez R, Donati A, Valente NS, et al. Glabellar red dots in frontal fibrosing alopecia: a further clinical sign of vellus follicle involvement. Br J Dermatol. 2014;170:745-746.
  13. Vañó-Galván S, Rodrigues-Barata AR, Urech M, et al. Depression of the frontal veins: a new clinical sign of frontal fibrosing alopecia. J Am Acad Dermatol. 2015;72:1087-1088.
  14. Pindado-Ortega C, Saceda-Corralo D, Buendía-Castaño D, et al. Frontal fibrosing alopecia and cutaneous comorbidities: a potential relationship with rosacea. J Am Acad Dermatol. 2018;78:596-597.e1.
Article PDF
Author and Disclosure Information

Drs. Krueger and Brinster as well as Ms. Svigos are from New York University School of Medicine, New York. Drs. Krueger and Brinster are from the Ronald O. Perelman Department of Dermatology. Dr. Elbuluk is from the Department of Dermatology, University of Southern California, Los Angeles.

The authors report no conflict of interest.

Correspondence: Loren Krueger, MD, 240 E 38th St, 11th Floor, New York, NY 10016 (loren.krueger@nyumc.org).

Issue
Cutis - 102(5)
Publications
Topics
Page Number
335-338
Sections
Author and Disclosure Information

Drs. Krueger and Brinster as well as Ms. Svigos are from New York University School of Medicine, New York. Drs. Krueger and Brinster are from the Ronald O. Perelman Department of Dermatology. Dr. Elbuluk is from the Department of Dermatology, University of Southern California, Los Angeles.

The authors report no conflict of interest.

Correspondence: Loren Krueger, MD, 240 E 38th St, 11th Floor, New York, NY 10016 (loren.krueger@nyumc.org).

Author and Disclosure Information

Drs. Krueger and Brinster as well as Ms. Svigos are from New York University School of Medicine, New York. Drs. Krueger and Brinster are from the Ronald O. Perelman Department of Dermatology. Dr. Elbuluk is from the Department of Dermatology, University of Southern California, Los Angeles.

The authors report no conflict of interest.

Correspondence: Loren Krueger, MD, 240 E 38th St, 11th Floor, New York, NY 10016 (loren.krueger@nyumc.org).

Article PDF
Article PDF
In Collaboration with the Skin of Color Society
In Collaboration with the Skin of Color Society

Frontal fibrosing alopecia (FFA) has been reported in association with lichen planus pigmentosus (LPP) and facial papules.1-3 Lichen planus pigmentosus is a variant of lichen planus that causes hyperpigmentation of the face, neck, and/or intertriginous areas that may be useful as a clinical indicator in the development of FFA.1 Facial papules in association with FFA are secondary to fibrosed vellus hairs.2,3 Currently, reports of concomitant FFA, LPP, and facial papules in women with skin of color are limited in the literature. This case series includes 5 women of color (Hispanic and black) who presented to our clinic with FFA and various cutaneous associations. A review of the current literature on cutaneous associations of FFA also is provided.

Case Reports

Patient 1
A 50-year-old Hispanic woman who was previously presumed to have melasma by an outside physician presented with pruritus of the scalp and eyebrows of 1 month’s duration. Physical examination revealed decreased frontal scalp hair density with perifollicular erythema and scale with thinning of the lateral eyebrows. Hyperpigmented coalesced macules (Figure 1A) and erythematous perifollicular papules were noted along the temples and on the perioral skin. Depressed forehead and temporal veins also were noted (Figure 1B). A biopsy of the scalp demonstrated perifollicular and perivascular lymphocytic inflammation and fibrosed hair follicles, and a biopsy of the perioral skin demonstrated perivascular lymphocytic inflammation with melanophages in the papillary dermis. A diagnosis of FFA with LPP was established with these biopsies.

Figure1
Figure 1. Frontal fibrosing alopecia with hyperpigmented coalesced macules around the mouth (A). Perifollicular papules on the temples (black arrow), erythematous perifollicular papules at the frontal hairline (blue arrow), and depressed veins on the forehead and temples (yellow arrows) also were noted (B).

Patient 2
A 61-year-old black woman presented with asymptomatic hair loss along the frontal hairline for an unknown duration. On physical examination the frontal scalp and lateral eyebrows demonstrated decreased hair density with loss of follicular ostia. Fine, flesh-colored, monomorphic papules were scattered along the forehead and temples, and ill-defined brown pigmentation was present along the forehead, temples, and cheeks. Biopsy of the frontal scalp demonstrated patchy lichenoid inflammation with decreased number of follicles with replacement by follicular scars, confirming the diagnosis of FFA.

Patient 3
A 47-year-old Hispanic woman presented with hair loss of the frontal scalp and bilateral eyebrows with associated burning of 2 years’ duration. Physical examination demonstrated recession of the frontotemporal hairline with scattered lone hairs and thinning of the eyebrows. Innumerable flesh-colored papules were present on the forehead and temples (Figure 2A). Glabellar and eyebrow erythema was noted (Figure 2B). Biopsy of the frontal scalp demonstrated decreased terminal anagen hair follicles with perifollicular lymphoid infiltrate and fibrosis, consistent with a diagnosis of FFA. The patient was started on oral hydroxychloroquine 400 mg once daily, and 3 months later hyperpigmentation of the forehead and perioral skin was noted. The patient reported that she had facial hyperpigmentation prior to starting hydroxychloroquine and declined a biopsy.

Figure2
Figure 2. Frontal fibrosing alopecia with recession of the temporal hairline with visible lone hairs (red arrow) and scattered flesh-colored papules on the temples (black arrows)(A). Glabellar and eyebrow erythema also was noted with flesh-colored papules on the forehead (black arrow)(B). The eyebrows were notably drawn in due to decreased hair density, and the central frontal hairline was recessed.

Patient 4
A 40-year-old black woman presented with brown pruritic macles of the face, neck, arms, and forearms of 4 years’ duration. She also reported hair loss on the frontal and occipital scalp, eyebrows, and arms. On physical examination, ill-defined brown macules and patches were noted on the neck (Figure 3), face, arms, and forearms. Decreased hair density was noted on the frontal and occipital scalp with follicular dropout and perifollicular hyperpigmentation. Biopsy of the scalp demonstrated perivascular lymphocytic inflammation with sparse anagen follicles and fibrous tracts, and biopsy of the neck revealed superficial perivascular inflammation with numerous melanophages in the upper dermis; these histopathologic findings were consistent with FFA and LPP, respectively.

Figure3
Figure 3. Diffuse and coalescing brown-gray macules and patches on the neck consistent with lichen planus pigmentosus.

Patient 5
A 46-year-old black woman with history of hair loss presented with hyperpigmentation of the face and neck of 2 years’ duration. On physical examination decreased hair density of the frontal and vertex scalp and lateral eyebrows was noted. Flesh-colored papules were noted on the forehead and cheeks, and confluent dark brown patches were present on the temples and neck. Three punch biopsies were performed. Biopsy of the scalp revealed lymphocytic inflammation with surrounding fibroplasia with overlapping features of FFA and central centrifugal cicatricial alopecia (Figure 4). Biopsy of the neck revealed vacuolar interface dermatitis. Additionally, biopsy of a facial papule revealed lichenoid inflammation involving a vellus hair follicle. Clinical and histopathological correlation confirmed the diagnosis of FFA with LPP and facial papules.

Figure4
Figure 4. Representative photograph demonstrating a diminished number of hair follicles with partial loss of sebaceous glands. There was perifollicular fibroplasia and interface inflammation along the basement membrane of the follicular epithelium with exocytosis of lymphocytes. Low-grade vacuolar alteration also was seen along the dermoepidermal junction (H&E, original magnification ×100).

 

 

Comment

Current understanding of FFA as a progressive, lymphocytic, scarring alopecia has expanded in recent years. Clinical observation suggests that the incidence of FFA is increasing4; however more epidemiologic data are needed. Frontal fibrosing alopecia presents clinically with symmetrical frontotemporal hair loss with lone hairs. Trichoscopy reveals perifollicular hyperkeratosis, perifollicular erythema, and follicular plugging in 72%, 66%, and 44% of cases, respectively.5 In one study (N=242), patients were classified into 3 clinical patterns of FFA: pattern I (linear) showed bandlike loss of frontal hair with normal density directly behind the hairline; pattern II (diffuse) showed loss of density behind the frontal hairline; and pattern III (double line) showed a pseudo–“fringe sign” appearance. The majority of patients were classified as either pattern I or II, with pattern II predicting a poorer prognosis.6

rontal fibrosing alopecia is increasingly recognized in men, with prevalence as high as 5%.1 Facial hair involvement, particularly of the upper lip and sideburns, is an important consideration in men.7 Most studies suggest that 80% to 90% of affected women are postmenopausal,8 though a case series presented by Dlova1 identified 27% of affected women as postmenopausal. The coexistence of premature menopause and hysterectomy in FFA patients suggests a hormonal contribution, but this association is still poorly understood.8 Epidemiologic data on ethnicity in FFA are sparse but suggest that white individuals are more likely to be affected. Frontal fibrosing alopecia also may be misdiagnosed as traction alopecia in Hispanic and black patients.8

It is prudent for physicians to assess for and recognize clinical clues to severe forms of FFA. A 2014 multicenter review of 355 patients identified 3 clinical entities that predicted more severe forms of FFA: eyelash loss (madarosis), loss of body hair, and facial papules.8 Madarosis occurs due to perifollicular inflammation and fibrosis of eyelash hair follicles. Similarly, perifollicular inflammation of body hair was present in 24% of patients (N=86), most commonly of the axillary and pubic hair. Facial papules form due to facial vellus hair inflammation and fibrosis and were identified in 14% of patients (N=49).8 These clinical findings may allow providers to predict more extensive clinical involvement of FFA.

Frontal fibrosing alopecia and LPP occur concomitantly in up 54% of patients, more commonly in darker-skinned patients.1,9,10 Lichen planus pigmentosus frequently occurs on the face and neck, most commonly in a diffuse pattern, though reticulated and macular patterns also have been identified.11 In some patients, LPP precedes the development of FFA and may be useful as a herald sign1; therefore, it is important for dermatologists to evaluate for signs of FFA when evaluating those with LPP. Thorough evaluation in patients with skin of color also is important because FFA may be misdiagnosed as traction alopecia.

Additional cutaneous associations of FFA include eyebrow loss, glabellar red dots, and prominent frontal veins. Eyebrow loss occurs secondary to fibrosis of eyebrow hair follicles and has been found in 40% to 80% of patients with FFA; it is thought to be associated with milder forms of FFA.8 Glabellar red dots correlate with histopathologic lymphocytic inflammation of vellus hair follicles.12 Additionally, frontal vein prominence has been described in FFA and is thought to be secondary to atrophy in this scarring process, perhaps worsened by local steroid treatments.13 Mucocutaneous lichen planus, rosacea, thyroid disease, vitiligo, and other autoimmune disorders also have been reported in patients with FFA.14

Conclusion

Concomitant FFA, LPP, and facial papules have been rarely reported and exemplify the spectrum of cutaneous associations with FFA, particularly in individuals with skin of color. Clinical variants and associations of FFA are broad, including predictors of poorer prognosis such as eyelash loss and vellus hair involvement seen as facial papules. Lichen planus pigmentosus is well described in association with FFA and may serve as a herald sign that frontal hair loss should not be mistaken for traction alopecia in early stages. Eyebrow loss is thought to represent milder disease. It is important for dermatologists to be aware of these findings to understand the breadth of this disease and for appropriate evaluation and management of patients with FFA.

Frontal fibrosing alopecia (FFA) has been reported in association with lichen planus pigmentosus (LPP) and facial papules.1-3 Lichen planus pigmentosus is a variant of lichen planus that causes hyperpigmentation of the face, neck, and/or intertriginous areas that may be useful as a clinical indicator in the development of FFA.1 Facial papules in association with FFA are secondary to fibrosed vellus hairs.2,3 Currently, reports of concomitant FFA, LPP, and facial papules in women with skin of color are limited in the literature. This case series includes 5 women of color (Hispanic and black) who presented to our clinic with FFA and various cutaneous associations. A review of the current literature on cutaneous associations of FFA also is provided.

Case Reports

Patient 1
A 50-year-old Hispanic woman who was previously presumed to have melasma by an outside physician presented with pruritus of the scalp and eyebrows of 1 month’s duration. Physical examination revealed decreased frontal scalp hair density with perifollicular erythema and scale with thinning of the lateral eyebrows. Hyperpigmented coalesced macules (Figure 1A) and erythematous perifollicular papules were noted along the temples and on the perioral skin. Depressed forehead and temporal veins also were noted (Figure 1B). A biopsy of the scalp demonstrated perifollicular and perivascular lymphocytic inflammation and fibrosed hair follicles, and a biopsy of the perioral skin demonstrated perivascular lymphocytic inflammation with melanophages in the papillary dermis. A diagnosis of FFA with LPP was established with these biopsies.

Figure1
Figure 1. Frontal fibrosing alopecia with hyperpigmented coalesced macules around the mouth (A). Perifollicular papules on the temples (black arrow), erythematous perifollicular papules at the frontal hairline (blue arrow), and depressed veins on the forehead and temples (yellow arrows) also were noted (B).

Patient 2
A 61-year-old black woman presented with asymptomatic hair loss along the frontal hairline for an unknown duration. On physical examination the frontal scalp and lateral eyebrows demonstrated decreased hair density with loss of follicular ostia. Fine, flesh-colored, monomorphic papules were scattered along the forehead and temples, and ill-defined brown pigmentation was present along the forehead, temples, and cheeks. Biopsy of the frontal scalp demonstrated patchy lichenoid inflammation with decreased number of follicles with replacement by follicular scars, confirming the diagnosis of FFA.

Patient 3
A 47-year-old Hispanic woman presented with hair loss of the frontal scalp and bilateral eyebrows with associated burning of 2 years’ duration. Physical examination demonstrated recession of the frontotemporal hairline with scattered lone hairs and thinning of the eyebrows. Innumerable flesh-colored papules were present on the forehead and temples (Figure 2A). Glabellar and eyebrow erythema was noted (Figure 2B). Biopsy of the frontal scalp demonstrated decreased terminal anagen hair follicles with perifollicular lymphoid infiltrate and fibrosis, consistent with a diagnosis of FFA. The patient was started on oral hydroxychloroquine 400 mg once daily, and 3 months later hyperpigmentation of the forehead and perioral skin was noted. The patient reported that she had facial hyperpigmentation prior to starting hydroxychloroquine and declined a biopsy.

Figure2
Figure 2. Frontal fibrosing alopecia with recession of the temporal hairline with visible lone hairs (red arrow) and scattered flesh-colored papules on the temples (black arrows)(A). Glabellar and eyebrow erythema also was noted with flesh-colored papules on the forehead (black arrow)(B). The eyebrows were notably drawn in due to decreased hair density, and the central frontal hairline was recessed.

Patient 4
A 40-year-old black woman presented with brown pruritic macles of the face, neck, arms, and forearms of 4 years’ duration. She also reported hair loss on the frontal and occipital scalp, eyebrows, and arms. On physical examination, ill-defined brown macules and patches were noted on the neck (Figure 3), face, arms, and forearms. Decreased hair density was noted on the frontal and occipital scalp with follicular dropout and perifollicular hyperpigmentation. Biopsy of the scalp demonstrated perivascular lymphocytic inflammation with sparse anagen follicles and fibrous tracts, and biopsy of the neck revealed superficial perivascular inflammation with numerous melanophages in the upper dermis; these histopathologic findings were consistent with FFA and LPP, respectively.

Figure3
Figure 3. Diffuse and coalescing brown-gray macules and patches on the neck consistent with lichen planus pigmentosus.

Patient 5
A 46-year-old black woman with history of hair loss presented with hyperpigmentation of the face and neck of 2 years’ duration. On physical examination decreased hair density of the frontal and vertex scalp and lateral eyebrows was noted. Flesh-colored papules were noted on the forehead and cheeks, and confluent dark brown patches were present on the temples and neck. Three punch biopsies were performed. Biopsy of the scalp revealed lymphocytic inflammation with surrounding fibroplasia with overlapping features of FFA and central centrifugal cicatricial alopecia (Figure 4). Biopsy of the neck revealed vacuolar interface dermatitis. Additionally, biopsy of a facial papule revealed lichenoid inflammation involving a vellus hair follicle. Clinical and histopathological correlation confirmed the diagnosis of FFA with LPP and facial papules.

Figure4
Figure 4. Representative photograph demonstrating a diminished number of hair follicles with partial loss of sebaceous glands. There was perifollicular fibroplasia and interface inflammation along the basement membrane of the follicular epithelium with exocytosis of lymphocytes. Low-grade vacuolar alteration also was seen along the dermoepidermal junction (H&E, original magnification ×100).

 

 

Comment

Current understanding of FFA as a progressive, lymphocytic, scarring alopecia has expanded in recent years. Clinical observation suggests that the incidence of FFA is increasing4; however more epidemiologic data are needed. Frontal fibrosing alopecia presents clinically with symmetrical frontotemporal hair loss with lone hairs. Trichoscopy reveals perifollicular hyperkeratosis, perifollicular erythema, and follicular plugging in 72%, 66%, and 44% of cases, respectively.5 In one study (N=242), patients were classified into 3 clinical patterns of FFA: pattern I (linear) showed bandlike loss of frontal hair with normal density directly behind the hairline; pattern II (diffuse) showed loss of density behind the frontal hairline; and pattern III (double line) showed a pseudo–“fringe sign” appearance. The majority of patients were classified as either pattern I or II, with pattern II predicting a poorer prognosis.6

rontal fibrosing alopecia is increasingly recognized in men, with prevalence as high as 5%.1 Facial hair involvement, particularly of the upper lip and sideburns, is an important consideration in men.7 Most studies suggest that 80% to 90% of affected women are postmenopausal,8 though a case series presented by Dlova1 identified 27% of affected women as postmenopausal. The coexistence of premature menopause and hysterectomy in FFA patients suggests a hormonal contribution, but this association is still poorly understood.8 Epidemiologic data on ethnicity in FFA are sparse but suggest that white individuals are more likely to be affected. Frontal fibrosing alopecia also may be misdiagnosed as traction alopecia in Hispanic and black patients.8

It is prudent for physicians to assess for and recognize clinical clues to severe forms of FFA. A 2014 multicenter review of 355 patients identified 3 clinical entities that predicted more severe forms of FFA: eyelash loss (madarosis), loss of body hair, and facial papules.8 Madarosis occurs due to perifollicular inflammation and fibrosis of eyelash hair follicles. Similarly, perifollicular inflammation of body hair was present in 24% of patients (N=86), most commonly of the axillary and pubic hair. Facial papules form due to facial vellus hair inflammation and fibrosis and were identified in 14% of patients (N=49).8 These clinical findings may allow providers to predict more extensive clinical involvement of FFA.

Frontal fibrosing alopecia and LPP occur concomitantly in up 54% of patients, more commonly in darker-skinned patients.1,9,10 Lichen planus pigmentosus frequently occurs on the face and neck, most commonly in a diffuse pattern, though reticulated and macular patterns also have been identified.11 In some patients, LPP precedes the development of FFA and may be useful as a herald sign1; therefore, it is important for dermatologists to evaluate for signs of FFA when evaluating those with LPP. Thorough evaluation in patients with skin of color also is important because FFA may be misdiagnosed as traction alopecia.

Additional cutaneous associations of FFA include eyebrow loss, glabellar red dots, and prominent frontal veins. Eyebrow loss occurs secondary to fibrosis of eyebrow hair follicles and has been found in 40% to 80% of patients with FFA; it is thought to be associated with milder forms of FFA.8 Glabellar red dots correlate with histopathologic lymphocytic inflammation of vellus hair follicles.12 Additionally, frontal vein prominence has been described in FFA and is thought to be secondary to atrophy in this scarring process, perhaps worsened by local steroid treatments.13 Mucocutaneous lichen planus, rosacea, thyroid disease, vitiligo, and other autoimmune disorders also have been reported in patients with FFA.14

Conclusion

Concomitant FFA, LPP, and facial papules have been rarely reported and exemplify the spectrum of cutaneous associations with FFA, particularly in individuals with skin of color. Clinical variants and associations of FFA are broad, including predictors of poorer prognosis such as eyelash loss and vellus hair involvement seen as facial papules. Lichen planus pigmentosus is well described in association with FFA and may serve as a herald sign that frontal hair loss should not be mistaken for traction alopecia in early stages. Eyebrow loss is thought to represent milder disease. It is important for dermatologists to be aware of these findings to understand the breadth of this disease and for appropriate evaluation and management of patients with FFA.

References
  1. Dlova NC. Frontal fibrosing alopecia and lichen planus pigmentosus: is there a link? Br J Dermatol. 2013;168:439-432.
  2. Donati A, Molina L, Doche I, et al. Facial papules in frontal fibrosing alopecia: evidence of vellus follicle involvement. Arch Dermatol. 2011;147:1424-1427.
  3. Tan KT, Messenger AG. Frontal fibrosing alopecia: clinical presentations and prognosis. Br J Dermatol. 2009;160:75-79.
  4. Rudnicka L, Rakowska A. The increasing incidence of frontal fibrosing alopecia. in search of triggering factors. J Eur Acad Dermatol Venereol. 2017;31:1579-1580.
  5. Toledo-Pastrana T, Hernández MJ, Camacho Martínez FM. Perifollicular erythema as a trichoscopy sign of progression in frontal fibrosing alopecia. Int J Trichology. 2013;5:151-153.
  6. Moreno-Arrones OM, Saceda-Corralo D, Fonda-Pascual P, et al. Frontal fibrosing alopecia: clinical and prognostic classification. J Eur Acad Dermatol Venereol. 2017;31:1739-1745.
  7. Tolkachjov SN, Chaudhry HM, Camilleri MJ, et al. Frontal fibrosing alopecia among men: a clinicopathologic study of 7 cases. J Am Acad Dermatol. 2017;77:683-690.e2.
  8. Vañó-Galván S, Molina-Ruiz AM, Serrano-Falcón C, et al. Frontal fibrosing alopecia: a multicenter review of 355 patients. J Am Acad Dermatol. 2014;70:670-678.
  9. Berliner JG, McCalmont TH, Price VH, et al. Frontal fibrosing alopecia and lichen planus pigmentosus. J Am Acad Dermatol. 2014;71:E26-E27.
  10. Rao R, Sarda A, Khanna R, et al. Coexistence of frontal fibrosing alopecia with lichen planus pigmentosus. Int J Dermatol. 2014;53:622-624.
  11. Pirmez R, Duque-Estrada B, Donati A, et al. Clinical and dermoscopic features of lichen planus pigmentosus in 37 patients with frontal fibrosing alopecia. Br J Dermatol. 2016;175:1387-1390.
  12. Pirmez R, Donati A, Valente NS, et al. Glabellar red dots in frontal fibrosing alopecia: a further clinical sign of vellus follicle involvement. Br J Dermatol. 2014;170:745-746.
  13. Vañó-Galván S, Rodrigues-Barata AR, Urech M, et al. Depression of the frontal veins: a new clinical sign of frontal fibrosing alopecia. J Am Acad Dermatol. 2015;72:1087-1088.
  14. Pindado-Ortega C, Saceda-Corralo D, Buendía-Castaño D, et al. Frontal fibrosing alopecia and cutaneous comorbidities: a potential relationship with rosacea. J Am Acad Dermatol. 2018;78:596-597.e1.
References
  1. Dlova NC. Frontal fibrosing alopecia and lichen planus pigmentosus: is there a link? Br J Dermatol. 2013;168:439-432.
  2. Donati A, Molina L, Doche I, et al. Facial papules in frontal fibrosing alopecia: evidence of vellus follicle involvement. Arch Dermatol. 2011;147:1424-1427.
  3. Tan KT, Messenger AG. Frontal fibrosing alopecia: clinical presentations and prognosis. Br J Dermatol. 2009;160:75-79.
  4. Rudnicka L, Rakowska A. The increasing incidence of frontal fibrosing alopecia. in search of triggering factors. J Eur Acad Dermatol Venereol. 2017;31:1579-1580.
  5. Toledo-Pastrana T, Hernández MJ, Camacho Martínez FM. Perifollicular erythema as a trichoscopy sign of progression in frontal fibrosing alopecia. Int J Trichology. 2013;5:151-153.
  6. Moreno-Arrones OM, Saceda-Corralo D, Fonda-Pascual P, et al. Frontal fibrosing alopecia: clinical and prognostic classification. J Eur Acad Dermatol Venereol. 2017;31:1739-1745.
  7. Tolkachjov SN, Chaudhry HM, Camilleri MJ, et al. Frontal fibrosing alopecia among men: a clinicopathologic study of 7 cases. J Am Acad Dermatol. 2017;77:683-690.e2.
  8. Vañó-Galván S, Molina-Ruiz AM, Serrano-Falcón C, et al. Frontal fibrosing alopecia: a multicenter review of 355 patients. J Am Acad Dermatol. 2014;70:670-678.
  9. Berliner JG, McCalmont TH, Price VH, et al. Frontal fibrosing alopecia and lichen planus pigmentosus. J Am Acad Dermatol. 2014;71:E26-E27.
  10. Rao R, Sarda A, Khanna R, et al. Coexistence of frontal fibrosing alopecia with lichen planus pigmentosus. Int J Dermatol. 2014;53:622-624.
  11. Pirmez R, Duque-Estrada B, Donati A, et al. Clinical and dermoscopic features of lichen planus pigmentosus in 37 patients with frontal fibrosing alopecia. Br J Dermatol. 2016;175:1387-1390.
  12. Pirmez R, Donati A, Valente NS, et al. Glabellar red dots in frontal fibrosing alopecia: a further clinical sign of vellus follicle involvement. Br J Dermatol. 2014;170:745-746.
  13. Vañó-Galván S, Rodrigues-Barata AR, Urech M, et al. Depression of the frontal veins: a new clinical sign of frontal fibrosing alopecia. J Am Acad Dermatol. 2015;72:1087-1088.
  14. Pindado-Ortega C, Saceda-Corralo D, Buendía-Castaño D, et al. Frontal fibrosing alopecia and cutaneous comorbidities: a potential relationship with rosacea. J Am Acad Dermatol. 2018;78:596-597.e1.
Issue
Cutis - 102(5)
Issue
Cutis - 102(5)
Page Number
335-338
Page Number
335-338
Publications
Publications
Topics
Article Type
Display Headline
Frontal Fibrosing Alopecia: Cutaneous Associations in Women With Skin of Color
Display Headline
Frontal Fibrosing Alopecia: Cutaneous Associations in Women With Skin of Color
Sections
Inside the Article

Practice Points

  • Frontal fibrosing alopecia (FFA) is associated with lichen planus pigmentosus, especially in patients with skin of color.
  • Patients with FFA should be evaluated for additional cutaneous features including facial papules, glabellar red dots, and depressed frontal veins.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Surgical Procedures for Hidradenitis Suppurativa

Article Type
Changed
Thu, 01/10/2019 - 13:51
Display Headline
Surgical Procedures for Hidradenitis Suppurativa
In Collaboration with the Skin of Color Society

Hidradenitis suppurativa (HS) is a chronic inflammatory skin disease that has a social and psychosocial impact on patients with skin of color.1 It is characterized by recurrent abscesses, draining sinus tracts, and scarring in the intertriginous skin folds. The lesions are difficult to treat and present with considerable frustration for both patients and physicians. Although current treatment ladders can delay procedures and surgical intervention,1 some believe that surgery should be introduced earlier in HS management.2 In this article, we review current procedures for the management of HS, including cryoinsufflation, incision and drainage, deroofing, skin tissue–saving excision with electrosurgical peeling, and wide surgical excision, along with various closure techniques.

Cryoinsufflation

First described in 2014, cryoinsufflation is a novel method for treating sinus tracts.3 Lesions initially are identified on physical examination. Prior to the procedure, local anesthesia is administered to the lesion.3 A 21-gauge needle is mounted onto a cryosurgical unit and inserted into the opening of the sinus tract. Liquid nitrogen is sprayed into the tract for 5 seconds, followed by a 3-second pause; the process is repeated 3 times. Patients return for treatment sessions monthly until the tract is obliterated. This procedure was first performed on 2 patients with satisfactory results.3

Since the initial report, the investigators made 2 changes to refine the procedure.4 First, systemic antibiotics should be prescribed 2 months prior to the procedure to clear the sinus tracts of infection. Furthermore, a 21-gauge, olive-tipped cannula is recommended in lieu of a 21-gauge needle to mitigate the risk of adverse events such as air embolism.4

Incision and Drainage

Incision and drainage provides rapid pain relief for tense fluctuant abscesses, but recurrence is common and the procedure costs are high.5 For drainage, wide circumferential local anesthesia is administered followed by incision.6 Pus is eliminated using digital pressure or saline rinses.2 Following the elimination of pus, the wound may need gauze packing or placement of a wick for a few days.6 The general belief is that incision and drainage should be used, if necessary, to rapidly relieve the patient’s pain; however, other surgical options should be considered if the patient has had multiple incision and drainage procedures.7 Currently there are no randomized controlled trials (RCTs) on incision and drainage procedures in HS abscesses.

Deroofing

In 1959, Mullins et al8 first described the deroofing procedure, which was refined to preserve the floor of the sinus tract in the 1980s.9,10 Culp10 and Brown et al9 theorized that preservation of the exposed floor of the sinus tract allowed for the epithelial cells from sweat glands and hair follicle remnants to rapidly reepithelialize the wound. In 2010, van der Zee et al11 performed a prospective study of 88 deroofed lesions in which the investigators removed keratinous debris and epithelial remnants of the floor due to concern for recurrence in this area if the tissues remained. Only 17% (15/88) of the lesions recurred at a median follow-up of 34 months.11

In Hurley stage I or II HS, deroofing remains the primary procedure for persistent nodules and sinus tracts.2 The lesion is identified on physical examination and local anesthesia is administered, first to the area surrounding the lesion, then to the lesion itself.11 A blunt probe is used to identify openings and search for connecting fistulas. After defining the sinus tract, the roof and wings created by the incision are removed.11,12 The material on the floor of the tract is scraped away, and the wound is left to heal by secondary intention.11 In general, deroofed lesions heal with cosmetically acceptable scars. We have used this procedure in skin of color patients with good results and no difficulties with healing. Controlled trials with long-term follow-up are lacking in this population.

Skin Tissue–Saving Excision With Electrosurgical Peeling

Skin tissue–saving excision with electrosurgical peeling was first introduced in 2015.13 Blok et al14 described the procedure as a promising alternative to wide surgical excision for Hurley stage II or III HS. The procedure saves healthy tissue while completely removing lesional tissue, leading to rapid wound healing, excellent cosmesis, and a low risk of contractures2,14; however, recurrence rates are higher than those seen in wide surgical excision.15 There are no known RCTs with long-term follow-up for HS patients treated with skin tissue–saving excision with electrosurgical peeling.

The procedure typically is performed under general anesthesia.14 First, the sinus tract is palpated on physical examination and probed to delineate the extent of the tract. Next, the roof of the tract is incised electrosurgically with a wire loop tip coupled to an electrosurgical generator.14 Consecutive tangential excisions are made until the floor of the sinus tract is reached. The process of incising sinus tracts followed by tangential peeling off of tissue continues until the entire area is clear of lesional and fibrotic tissue. The wound margins are probed for the presence and subsequent removal of residual sinus tracts. Lastly, the electrosurgical generator is used to achieve hemostasis, steroids are injected to prevent the formation of hypergranulation tissue, and the wound is left to heal by secondary intention.14 Following intervention, recurrence rates appear to be similar to wide surgical excision.13,14

 

 

Wide Surgical Excision

Wide excision is a widely established technique consisting of surgical excision of a lesion plus an area of surrounding disease-free tissue such as subcutaneous fat or a lateral margin of intertriginous skin.15 Similar to other surgical techniques, wide excision is considered in cases of severe disease when pharmacologic management cannot remedy extensive fibrosis or architectural loss. It typically is performed in Hurley stage II and III HS, with pathology extending to involve deeper structures inaccessible to more superficial surgical methods.2 Prominent areas of use include gluteal, axillary, perineal, and perianal HS lesions on which conservative treatments have little effect and depend on wide excision to provide successful postoperative results.16 Although retrospective and prospective studies exist on wide excision in HS, there continues to be a dearth of RCTs. Based on the available literature, the primary motive for wide excision is lower recurrence rates (13% overall compared to 22% and 27% for local excision and deroofing, respectively) and longer asymptomatic periods compared to more local techniques.7,17 Wide excision combined with continued aggressive medical management and dietary modifications currently is an efficacious treatment in providing functional long-term results.6 These benefits, however, are not without their drawbacks, as the more extensive nature of wide surgical excision predisposes patients to larger wounds, surgery-induced infection, and prolonged recovery periods.6,15 If preoperative measurements are not wisely assessed, the excision also can extend to involve neurovascular bundles and other vital structures, contributing to greater postoperative morbidity.15 Ultrasonography provides useful anatomic information in HS, such as location and extent of fistulous tracts and fluid collections; these findings can assist in guiding the width and depth of the excision itself to ensure the entire area of HS involvement is removed.18 Published data revealed that 204 of 255 (80%) patients were markedly satisfied with postoperative outcomes of wide excision,19 which gives credence to the idea that although the complications of wide excision may not be as favorable, the long-lasting improvements in quality of life make wide surgical excision a suggested first-line treatment in all stages of HS.16,20

Closure Techniques

The best skin closure method following surgical excision is controversial and not well established in literature. Options include healing by secondary intention, primary (suture-based) closure, skin grafts, and skin flaps. Each of these methods has had moderate success in multiple observational studies, and the choice should be made based on individualized assessment of the patient’s HS lesion characteristics, ability to adhere to recovery protocols, and relevant demographics. A systematic review by Mehdizadeh et al17 provided the following recurrence rates for techniques utilized after wide excision: primary closure, 15%; flaps, 8%; and grafting, 6%. Despite conflicting evidence, allowing wounds to heal by secondary intention is best, based on the author’s experience (I.H.H.).

Secondary Intention
Healing by secondary intention refers to a wound that is intentionally left open to be filled in with granulation tissue and eventual epithelization over time rather than being approximated and closed via sutures or staples as in primary intention. It is a well-established option in wound management and results in a longer but more comfortable period of convalescence in postsurgical HS management.20 Patients can add regular moist wound dressings (eg, silastic foam dressing) to manage the wound at home and continue normal activities for most of the healing period; however, the recovery period can become excessively long and painful, and there is a high risk of formation of retractile scar bands at and around the healing site.12 Strict adherence to wound-healing protocols is paramount to minimizing unwanted complications.21 Secondary intention often is used after wide local excision and has been demonstrated to yield positive functional and aesthetic results in multiple studies, especially in the more severe Hurley stage II or III cases.21,22 It can be successfully employed after laser treatment and in surgical defects of all sizes with little to no contractures or reduced range of motion.6 Ultimately, the choice to heal via secondary intention should be made after thorough assessment of patient needs and with ample education to ensure compliance.

Primary Closure
Primary closure is the suture-mediated closing technique that is most often used in wound closure for lower-grade HS cases, especially smaller excisions. However, it is associated with potential complications. If HS lesions are not effectively excised, disease can then recur at the periphery of the excision and wound dehiscence can manifest more readily, especially as wound size increases.23 Consequently, primary closure is associated with the highest recurrence rates among closure techniques.17 Avoiding primary closure in active disease also is recommended due to the potential of burying residual foci of inflammation.6 Finally, primary closures lack skin coverage and thus often are not viable options in most perianal and genital lesions that require more extensive reconstruction. Retrospective case series and case reports exist on primary excision, but further study is needed.

Skin Grafts
Skin grafting is a technique of surgically transplanting a piece of healthy skin from one body site to another. Skin grafts typically are used when primary closure or skin flaps are not feasible (eg, in large wounds) and also when shorter time to wound closure is a greater concern in patient recovery.2,24 Additionally, skin grafts can be employed on large flat surfaces of the body, such as the buttocks or thighs, for timely wound closure when wound contraction is less effective or wound healing is slow via epithelization. Types of skin graft techniques include split-thickness skin graft (STSG), full-thickness skin graft, and recycled skin graft. All 3 types have demonstrated acceptable functional and aesthetic results in observational studies and case reports, and thus deciding which technique to use should include individualized assessment.2,25 The STSG has several advantages over the full-thickness skin graft, including hairlessness (ie, without hair follicles), ease of harvest, and a less complicated transfer to contaminated lesional areas such as those in HS.26 Additionally, STSGs allow for closure of even the largest wounds with minimal risk of serious infection. Split-thickness skin grafts are considered one of the most efficacious tools for axilla reconstruction; however, they require prolonged immobilization of the arm, result in sequelae in donor sites, and do not always prevent retractile scars.26 The recycled skin graft technique can be used to treat chronic gluteal HS, but reliability and outcomes have not been reported. Skin grafting after excision is associated with increased pain, immobilization, prolonged hospitalization, and longer healing times compared to skin flaps.19 In a systematic review of wound healing techniques following wide excision, grafting was shown to have the lowest recurrence rate (6%) compared to skin flaps (8%) and primary closure (15%).17 The absence of hair follicles and sweat glands in STSGs may be advantageous in HS because both hair follicles and sweat glands are thought to play more roles in the pathogenesis of HS.18,24 Most studies on skin grafts are limited to case reports.

Skin Flaps
Skin flaps are similar to skin grafts in that healthy skin is transplanted from one site to another; the difference is that flaps maintain an intact blood supply, whereas skin grafts depend on growth of new blood vessels.12,13 The primary advantage of skin flaps is that they provide the best quality of skin due to the thick tissue coverage, which is an important concern, especially in aesthetic scenarios. Additionally, they have been shown to provide shorter healing times than grafts, primary closure, and secondary healing, which can be especially important when functional disability is a concern in the postoperative period.26 However, their use should be limited due to several complications owing to their blood supply, as there is a high risk of ischemia to distant portions of flaps, which often can progress to necrosis and hemorrhage during the harvesting process.2 Thus, skin flaps are incredibly difficult to use in larger wounds and often require debulking due to their thickness. Additionally, skin flaps are definitive by nature, which can pose an issue if HS recurs locally. Skin flaps are recommended only when their use is mandatory, such as in the coverage of important anatomic structures (eg, exposed neurovascular bundles and large vessels).2 Advances have been made in flap construction, and now several types of flaps are employed in several body areas with differing indications and recommendations.2,21 As with skin grafts, most studies in the literature are case reports; therefore, further investigation is needed.

Combination Reconstructions
Combination reconstructions refer to the simultaneous use of multiple closure or healing techniques. By combining 2 or more methods, surgeons can utilize the advantages of each technique to provide an individualized approach that can substantially diminish wound surface area and accelerate wound healing.2 For example, with the starlike technique, 5 equilateral triangles bordering a foci of axillary disease are excised in addition to the central foci, and the edges of each triangle are then sutured together to create a final scar of considerably smaller size. The starlike technique allows the wound to be partially sutured while leaving the remaining area to heal by secondary intention.2 There are a small number of case series and prospective studies on combined reconstructions in HS but no RCTs.

Conclusion

Many procedures exist as options for treatment of patients with HS. Deroofing and cryoinsufflation are options for localized Hurley stage I or II disease. For more severe Hurley stage II or III disease, skin tissue–saving excision with electrosurgical peeling or wide surgical excisions are preferred. Following excision, there are many options for wound closure, but our preference is to allow the wound to heal by secondary intention. It is imperative that dermatologists are informed on the different techniques for treating this disease to determine the best route of care for their patients.

References
  1. Smith MK, Nicholson CL, Parks-Miller A, et al. Hidradenitis suppurativa: an update on connecting the tracts. F1000Res. 2017;6:1272.
  2. Janse I, Bieniek A, Horvath B, et al. Surgical procedures in hidradenitis suppurativa. Dermatol Clin. 2016;34:97-109.
  3. Pagliarello C, Fabrizi G, Feliciani C, et al. Cryoinsufflation for Hurley stage II hidradenitis suppurativa: a useful treatment option when systemic therapies should be avoided. JAMA Dermatol. 2014;150:765-766.
  4. Pagliarello C, Fabrizi G, di Nuzzo S. Cryoinsufflation for hidradenitis suppurativa: technical refinement to prevent complications. Dermatol Surg. 2016;42:130-132.
  5. Ritz JP, Runkel N, Haier J, et al. Extent of surgery and recurrence rate of hidradenitis suppurativa. Int J Colorectal Dis. 1998;13:164-168.
  6. Danby FW, Hazen PG, Boer J. New and traditional surgical approaches to hidradenitis suppurativa. J Am Acad Dermatol. 2015;73(5, suppl 1):S62-S65.
  7. Ellis LZ. Hidradenitis suppurativa: surgical and other management techniques. Dermatol Surg. 2012;38:517-536.
  8. Mullins JF, McCash WB, Boudreau RF. Treatment of chronic hidradenitis suppurativa: surgical modification. Postgrad Med. 1959;26:805-808.
  9. Brown SC, Kazzazi N, Lord PH. Surgical treatment of perineal hidradenitis suppurativa with special reference to recognition of the perianal form. Br J Surg. 1986;73:978-980.
  10. Culp CE. Chronic hidradenitis suppurativa of the anal canal. a surgical skin disease. Dis Colon Rectum. 1983;26:669-676.
  11. van der Zee HH, Prens EP, Boer J. Deroofing: a tissue-saving surgical technique for the treatment of mild to moderate hidradenitis suppurativa lesions. J Am Acad Dermatol. 2010;63:475-480.
  12. Lin CH, Chang KP, Huang SH. Deroofing: an effective method for treating chronic diffuse hidradenitis suppurativa. Dermatol Surg. 2016;42:273-275.
  13. Blok JL, Boersma M, Terra JB, et al. Surgery under general anaes-thesia in severe hidradenitis suppurativa: a study of 363 primary operations in 113 patients. J Eur Acad Dermatol Venereol. 2015;29:1590-1597.
  14. Blok JL, Spoo JR, Leeman FW, et al. Skin-Tissue-sparing Excision with Electrosurgical Peeling (STEEP): a surgical treatment option for severe hidradenitis suppurativa Hurley stage II/III. J Eur Acad Dermatol Venereol. 2015;29:379-382.
  15. Saunte DML, Jemec GBE. Hidradenitis suppurativa: advances in diagnosis and treatment. JAMA. 2017;318:2019-2032.
  16. Maghsoudi H, Almasi H, Miri Bonjar MR. Men, main victims of hidradenitis suppurativa (a prospective cohort study). Int J Surg. 2018;50:6-10.
  17. Mehdizadeh A, Hazen PG, Bechara FG, et al. Recurrence of hidradenitis suppurativa after surgical management: a systematic review and meta-analysis. J Am Acad Dermatol. 2015;73(5, suppl 1):S70-S77.
  18. Wortsman X, Moreno C, Soto R, et al. Ultrasound in-depth characterization and staging of hidradenitis suppurativa. Dermatol Surg. 2013;39:1835-1842.
  19. Kofler L, Schweinzer K, Heister M, et al. Surgical treatment of hidradenitis suppurativa: an analysis of postoperative outcome, cosmetic results and quality of life in 255 patients [published online February 17, 2018]. J Eur Acad Dermatol Venereol. doi:10.1111/jdv.14892.
  20. Dini V, Oranges T, Rotella L, et al. Hidradenitis suppurativa and wound management. Int J Low Extrem Wounds. 2015;14:236-244.
  21. Humphries LS, Kueberuwa E, Beederman M, et al. Wide excision and healing by secondary intent for the surgical treatment of hidradenitis suppurativa: a single-center experience. J Plast Reconstr Aesthet Surg. 2016;69:554-566.
  22. Wollina U, Langner D, Heinig B, et al. Comorbidities, treatment, and outcome in severe anogenital inverse acne (hidradenitis suppurativa): a 15-year single center report. Int J Dermatol. 2017;56:109-115.
  23. Watson JD. Hidradenitis suppurativa—a clinical review. Br J Plast Surg. 1985;38:567-569.
  24. Sugio Y, Tomita K, Hosokawa K. Reconstruction after excision of hidradenitis suppurativa: are skin grafts better than flaps? Plast Reconstr Surg Glob Open. 2016;4:E1128.
  25. Burney RE. 35-year experience with surgical treatment of hidradenitis suppurativa. World J Surg. 2017;41:2723-2730.
  26. Nail-Barthelemy R, Stroumza N, Qassemyar Q, et al. Evaluation of the mobility of the shoulder and quality of life after perforator flaps for recalcitrant axillary hidradenitis [published online February 13, 2018]. Ann Chir Plast Esthet. pii:S0294-1260(18)30005-0. doi:10.1016/j.anplas.2018.01.003.
Article PDF
Author and Disclosure Information

 

Mr. Vellaichamy is from the Department of Dermatology, Wayne State University, Detroit, Michigan. Drs. Braunberger, Nahhas, and Hamzavi are from the Department of Dermatology, Henry Ford Hospital, Detroit.

Mr. Vellaichamy and Drs. Braunberger and Nahhas report no conflict of interest. Dr. Hamzavi received honoraria as an advisory board member for Aclaris Therapeutics, Inc. He also is a consultant for Pfizer Inc and an investigator for Allergan; The Estée Lauder Companies; Ferndale Pharma Group, Inc; and Johnson & Johnson Consumer Inc.

Correspondence: Iltefat H. Hamzavi, MD, Department of Dermatology, Henry Ford Hospital, 3031 W Grand Blvd, Detroit, MI 48202 (Ihamzav1@hfhs.org).

Issue
Cutis - 102(1)
Publications
Topics
Page Number
13-16
Sections
Author and Disclosure Information

 

Mr. Vellaichamy is from the Department of Dermatology, Wayne State University, Detroit, Michigan. Drs. Braunberger, Nahhas, and Hamzavi are from the Department of Dermatology, Henry Ford Hospital, Detroit.

Mr. Vellaichamy and Drs. Braunberger and Nahhas report no conflict of interest. Dr. Hamzavi received honoraria as an advisory board member for Aclaris Therapeutics, Inc. He also is a consultant for Pfizer Inc and an investigator for Allergan; The Estée Lauder Companies; Ferndale Pharma Group, Inc; and Johnson & Johnson Consumer Inc.

Correspondence: Iltefat H. Hamzavi, MD, Department of Dermatology, Henry Ford Hospital, 3031 W Grand Blvd, Detroit, MI 48202 (Ihamzav1@hfhs.org).

Author and Disclosure Information

 

Mr. Vellaichamy is from the Department of Dermatology, Wayne State University, Detroit, Michigan. Drs. Braunberger, Nahhas, and Hamzavi are from the Department of Dermatology, Henry Ford Hospital, Detroit.

Mr. Vellaichamy and Drs. Braunberger and Nahhas report no conflict of interest. Dr. Hamzavi received honoraria as an advisory board member for Aclaris Therapeutics, Inc. He also is a consultant for Pfizer Inc and an investigator for Allergan; The Estée Lauder Companies; Ferndale Pharma Group, Inc; and Johnson & Johnson Consumer Inc.

Correspondence: Iltefat H. Hamzavi, MD, Department of Dermatology, Henry Ford Hospital, 3031 W Grand Blvd, Detroit, MI 48202 (Ihamzav1@hfhs.org).

Article PDF
Article PDF
In Collaboration with the Skin of Color Society
In Collaboration with the Skin of Color Society

Hidradenitis suppurativa (HS) is a chronic inflammatory skin disease that has a social and psychosocial impact on patients with skin of color.1 It is characterized by recurrent abscesses, draining sinus tracts, and scarring in the intertriginous skin folds. The lesions are difficult to treat and present with considerable frustration for both patients and physicians. Although current treatment ladders can delay procedures and surgical intervention,1 some believe that surgery should be introduced earlier in HS management.2 In this article, we review current procedures for the management of HS, including cryoinsufflation, incision and drainage, deroofing, skin tissue–saving excision with electrosurgical peeling, and wide surgical excision, along with various closure techniques.

Cryoinsufflation

First described in 2014, cryoinsufflation is a novel method for treating sinus tracts.3 Lesions initially are identified on physical examination. Prior to the procedure, local anesthesia is administered to the lesion.3 A 21-gauge needle is mounted onto a cryosurgical unit and inserted into the opening of the sinus tract. Liquid nitrogen is sprayed into the tract for 5 seconds, followed by a 3-second pause; the process is repeated 3 times. Patients return for treatment sessions monthly until the tract is obliterated. This procedure was first performed on 2 patients with satisfactory results.3

Since the initial report, the investigators made 2 changes to refine the procedure.4 First, systemic antibiotics should be prescribed 2 months prior to the procedure to clear the sinus tracts of infection. Furthermore, a 21-gauge, olive-tipped cannula is recommended in lieu of a 21-gauge needle to mitigate the risk of adverse events such as air embolism.4

Incision and Drainage

Incision and drainage provides rapid pain relief for tense fluctuant abscesses, but recurrence is common and the procedure costs are high.5 For drainage, wide circumferential local anesthesia is administered followed by incision.6 Pus is eliminated using digital pressure or saline rinses.2 Following the elimination of pus, the wound may need gauze packing or placement of a wick for a few days.6 The general belief is that incision and drainage should be used, if necessary, to rapidly relieve the patient’s pain; however, other surgical options should be considered if the patient has had multiple incision and drainage procedures.7 Currently there are no randomized controlled trials (RCTs) on incision and drainage procedures in HS abscesses.

Deroofing

In 1959, Mullins et al8 first described the deroofing procedure, which was refined to preserve the floor of the sinus tract in the 1980s.9,10 Culp10 and Brown et al9 theorized that preservation of the exposed floor of the sinus tract allowed for the epithelial cells from sweat glands and hair follicle remnants to rapidly reepithelialize the wound. In 2010, van der Zee et al11 performed a prospective study of 88 deroofed lesions in which the investigators removed keratinous debris and epithelial remnants of the floor due to concern for recurrence in this area if the tissues remained. Only 17% (15/88) of the lesions recurred at a median follow-up of 34 months.11

In Hurley stage I or II HS, deroofing remains the primary procedure for persistent nodules and sinus tracts.2 The lesion is identified on physical examination and local anesthesia is administered, first to the area surrounding the lesion, then to the lesion itself.11 A blunt probe is used to identify openings and search for connecting fistulas. After defining the sinus tract, the roof and wings created by the incision are removed.11,12 The material on the floor of the tract is scraped away, and the wound is left to heal by secondary intention.11 In general, deroofed lesions heal with cosmetically acceptable scars. We have used this procedure in skin of color patients with good results and no difficulties with healing. Controlled trials with long-term follow-up are lacking in this population.

Skin Tissue–Saving Excision With Electrosurgical Peeling

Skin tissue–saving excision with electrosurgical peeling was first introduced in 2015.13 Blok et al14 described the procedure as a promising alternative to wide surgical excision for Hurley stage II or III HS. The procedure saves healthy tissue while completely removing lesional tissue, leading to rapid wound healing, excellent cosmesis, and a low risk of contractures2,14; however, recurrence rates are higher than those seen in wide surgical excision.15 There are no known RCTs with long-term follow-up for HS patients treated with skin tissue–saving excision with electrosurgical peeling.

The procedure typically is performed under general anesthesia.14 First, the sinus tract is palpated on physical examination and probed to delineate the extent of the tract. Next, the roof of the tract is incised electrosurgically with a wire loop tip coupled to an electrosurgical generator.14 Consecutive tangential excisions are made until the floor of the sinus tract is reached. The process of incising sinus tracts followed by tangential peeling off of tissue continues until the entire area is clear of lesional and fibrotic tissue. The wound margins are probed for the presence and subsequent removal of residual sinus tracts. Lastly, the electrosurgical generator is used to achieve hemostasis, steroids are injected to prevent the formation of hypergranulation tissue, and the wound is left to heal by secondary intention.14 Following intervention, recurrence rates appear to be similar to wide surgical excision.13,14

 

 

Wide Surgical Excision

Wide excision is a widely established technique consisting of surgical excision of a lesion plus an area of surrounding disease-free tissue such as subcutaneous fat or a lateral margin of intertriginous skin.15 Similar to other surgical techniques, wide excision is considered in cases of severe disease when pharmacologic management cannot remedy extensive fibrosis or architectural loss. It typically is performed in Hurley stage II and III HS, with pathology extending to involve deeper structures inaccessible to more superficial surgical methods.2 Prominent areas of use include gluteal, axillary, perineal, and perianal HS lesions on which conservative treatments have little effect and depend on wide excision to provide successful postoperative results.16 Although retrospective and prospective studies exist on wide excision in HS, there continues to be a dearth of RCTs. Based on the available literature, the primary motive for wide excision is lower recurrence rates (13% overall compared to 22% and 27% for local excision and deroofing, respectively) and longer asymptomatic periods compared to more local techniques.7,17 Wide excision combined with continued aggressive medical management and dietary modifications currently is an efficacious treatment in providing functional long-term results.6 These benefits, however, are not without their drawbacks, as the more extensive nature of wide surgical excision predisposes patients to larger wounds, surgery-induced infection, and prolonged recovery periods.6,15 If preoperative measurements are not wisely assessed, the excision also can extend to involve neurovascular bundles and other vital structures, contributing to greater postoperative morbidity.15 Ultrasonography provides useful anatomic information in HS, such as location and extent of fistulous tracts and fluid collections; these findings can assist in guiding the width and depth of the excision itself to ensure the entire area of HS involvement is removed.18 Published data revealed that 204 of 255 (80%) patients were markedly satisfied with postoperative outcomes of wide excision,19 which gives credence to the idea that although the complications of wide excision may not be as favorable, the long-lasting improvements in quality of life make wide surgical excision a suggested first-line treatment in all stages of HS.16,20

Closure Techniques

The best skin closure method following surgical excision is controversial and not well established in literature. Options include healing by secondary intention, primary (suture-based) closure, skin grafts, and skin flaps. Each of these methods has had moderate success in multiple observational studies, and the choice should be made based on individualized assessment of the patient’s HS lesion characteristics, ability to adhere to recovery protocols, and relevant demographics. A systematic review by Mehdizadeh et al17 provided the following recurrence rates for techniques utilized after wide excision: primary closure, 15%; flaps, 8%; and grafting, 6%. Despite conflicting evidence, allowing wounds to heal by secondary intention is best, based on the author’s experience (I.H.H.).

Secondary Intention
Healing by secondary intention refers to a wound that is intentionally left open to be filled in with granulation tissue and eventual epithelization over time rather than being approximated and closed via sutures or staples as in primary intention. It is a well-established option in wound management and results in a longer but more comfortable period of convalescence in postsurgical HS management.20 Patients can add regular moist wound dressings (eg, silastic foam dressing) to manage the wound at home and continue normal activities for most of the healing period; however, the recovery period can become excessively long and painful, and there is a high risk of formation of retractile scar bands at and around the healing site.12 Strict adherence to wound-healing protocols is paramount to minimizing unwanted complications.21 Secondary intention often is used after wide local excision and has been demonstrated to yield positive functional and aesthetic results in multiple studies, especially in the more severe Hurley stage II or III cases.21,22 It can be successfully employed after laser treatment and in surgical defects of all sizes with little to no contractures or reduced range of motion.6 Ultimately, the choice to heal via secondary intention should be made after thorough assessment of patient needs and with ample education to ensure compliance.

Primary Closure
Primary closure is the suture-mediated closing technique that is most often used in wound closure for lower-grade HS cases, especially smaller excisions. However, it is associated with potential complications. If HS lesions are not effectively excised, disease can then recur at the periphery of the excision and wound dehiscence can manifest more readily, especially as wound size increases.23 Consequently, primary closure is associated with the highest recurrence rates among closure techniques.17 Avoiding primary closure in active disease also is recommended due to the potential of burying residual foci of inflammation.6 Finally, primary closures lack skin coverage and thus often are not viable options in most perianal and genital lesions that require more extensive reconstruction. Retrospective case series and case reports exist on primary excision, but further study is needed.

Skin Grafts
Skin grafting is a technique of surgically transplanting a piece of healthy skin from one body site to another. Skin grafts typically are used when primary closure or skin flaps are not feasible (eg, in large wounds) and also when shorter time to wound closure is a greater concern in patient recovery.2,24 Additionally, skin grafts can be employed on large flat surfaces of the body, such as the buttocks or thighs, for timely wound closure when wound contraction is less effective or wound healing is slow via epithelization. Types of skin graft techniques include split-thickness skin graft (STSG), full-thickness skin graft, and recycled skin graft. All 3 types have demonstrated acceptable functional and aesthetic results in observational studies and case reports, and thus deciding which technique to use should include individualized assessment.2,25 The STSG has several advantages over the full-thickness skin graft, including hairlessness (ie, without hair follicles), ease of harvest, and a less complicated transfer to contaminated lesional areas such as those in HS.26 Additionally, STSGs allow for closure of even the largest wounds with minimal risk of serious infection. Split-thickness skin grafts are considered one of the most efficacious tools for axilla reconstruction; however, they require prolonged immobilization of the arm, result in sequelae in donor sites, and do not always prevent retractile scars.26 The recycled skin graft technique can be used to treat chronic gluteal HS, but reliability and outcomes have not been reported. Skin grafting after excision is associated with increased pain, immobilization, prolonged hospitalization, and longer healing times compared to skin flaps.19 In a systematic review of wound healing techniques following wide excision, grafting was shown to have the lowest recurrence rate (6%) compared to skin flaps (8%) and primary closure (15%).17 The absence of hair follicles and sweat glands in STSGs may be advantageous in HS because both hair follicles and sweat glands are thought to play more roles in the pathogenesis of HS.18,24 Most studies on skin grafts are limited to case reports.

Skin Flaps
Skin flaps are similar to skin grafts in that healthy skin is transplanted from one site to another; the difference is that flaps maintain an intact blood supply, whereas skin grafts depend on growth of new blood vessels.12,13 The primary advantage of skin flaps is that they provide the best quality of skin due to the thick tissue coverage, which is an important concern, especially in aesthetic scenarios. Additionally, they have been shown to provide shorter healing times than grafts, primary closure, and secondary healing, which can be especially important when functional disability is a concern in the postoperative period.26 However, their use should be limited due to several complications owing to their blood supply, as there is a high risk of ischemia to distant portions of flaps, which often can progress to necrosis and hemorrhage during the harvesting process.2 Thus, skin flaps are incredibly difficult to use in larger wounds and often require debulking due to their thickness. Additionally, skin flaps are definitive by nature, which can pose an issue if HS recurs locally. Skin flaps are recommended only when their use is mandatory, such as in the coverage of important anatomic structures (eg, exposed neurovascular bundles and large vessels).2 Advances have been made in flap construction, and now several types of flaps are employed in several body areas with differing indications and recommendations.2,21 As with skin grafts, most studies in the literature are case reports; therefore, further investigation is needed.

Combination Reconstructions
Combination reconstructions refer to the simultaneous use of multiple closure or healing techniques. By combining 2 or more methods, surgeons can utilize the advantages of each technique to provide an individualized approach that can substantially diminish wound surface area and accelerate wound healing.2 For example, with the starlike technique, 5 equilateral triangles bordering a foci of axillary disease are excised in addition to the central foci, and the edges of each triangle are then sutured together to create a final scar of considerably smaller size. The starlike technique allows the wound to be partially sutured while leaving the remaining area to heal by secondary intention.2 There are a small number of case series and prospective studies on combined reconstructions in HS but no RCTs.

Conclusion

Many procedures exist as options for treatment of patients with HS. Deroofing and cryoinsufflation are options for localized Hurley stage I or II disease. For more severe Hurley stage II or III disease, skin tissue–saving excision with electrosurgical peeling or wide surgical excisions are preferred. Following excision, there are many options for wound closure, but our preference is to allow the wound to heal by secondary intention. It is imperative that dermatologists are informed on the different techniques for treating this disease to determine the best route of care for their patients.

Hidradenitis suppurativa (HS) is a chronic inflammatory skin disease that has a social and psychosocial impact on patients with skin of color.1 It is characterized by recurrent abscesses, draining sinus tracts, and scarring in the intertriginous skin folds. The lesions are difficult to treat and present with considerable frustration for both patients and physicians. Although current treatment ladders can delay procedures and surgical intervention,1 some believe that surgery should be introduced earlier in HS management.2 In this article, we review current procedures for the management of HS, including cryoinsufflation, incision and drainage, deroofing, skin tissue–saving excision with electrosurgical peeling, and wide surgical excision, along with various closure techniques.

Cryoinsufflation

First described in 2014, cryoinsufflation is a novel method for treating sinus tracts.3 Lesions initially are identified on physical examination. Prior to the procedure, local anesthesia is administered to the lesion.3 A 21-gauge needle is mounted onto a cryosurgical unit and inserted into the opening of the sinus tract. Liquid nitrogen is sprayed into the tract for 5 seconds, followed by a 3-second pause; the process is repeated 3 times. Patients return for treatment sessions monthly until the tract is obliterated. This procedure was first performed on 2 patients with satisfactory results.3

Since the initial report, the investigators made 2 changes to refine the procedure.4 First, systemic antibiotics should be prescribed 2 months prior to the procedure to clear the sinus tracts of infection. Furthermore, a 21-gauge, olive-tipped cannula is recommended in lieu of a 21-gauge needle to mitigate the risk of adverse events such as air embolism.4

Incision and Drainage

Incision and drainage provides rapid pain relief for tense fluctuant abscesses, but recurrence is common and the procedure costs are high.5 For drainage, wide circumferential local anesthesia is administered followed by incision.6 Pus is eliminated using digital pressure or saline rinses.2 Following the elimination of pus, the wound may need gauze packing or placement of a wick for a few days.6 The general belief is that incision and drainage should be used, if necessary, to rapidly relieve the patient’s pain; however, other surgical options should be considered if the patient has had multiple incision and drainage procedures.7 Currently there are no randomized controlled trials (RCTs) on incision and drainage procedures in HS abscesses.

Deroofing

In 1959, Mullins et al8 first described the deroofing procedure, which was refined to preserve the floor of the sinus tract in the 1980s.9,10 Culp10 and Brown et al9 theorized that preservation of the exposed floor of the sinus tract allowed for the epithelial cells from sweat glands and hair follicle remnants to rapidly reepithelialize the wound. In 2010, van der Zee et al11 performed a prospective study of 88 deroofed lesions in which the investigators removed keratinous debris and epithelial remnants of the floor due to concern for recurrence in this area if the tissues remained. Only 17% (15/88) of the lesions recurred at a median follow-up of 34 months.11

In Hurley stage I or II HS, deroofing remains the primary procedure for persistent nodules and sinus tracts.2 The lesion is identified on physical examination and local anesthesia is administered, first to the area surrounding the lesion, then to the lesion itself.11 A blunt probe is used to identify openings and search for connecting fistulas. After defining the sinus tract, the roof and wings created by the incision are removed.11,12 The material on the floor of the tract is scraped away, and the wound is left to heal by secondary intention.11 In general, deroofed lesions heal with cosmetically acceptable scars. We have used this procedure in skin of color patients with good results and no difficulties with healing. Controlled trials with long-term follow-up are lacking in this population.

Skin Tissue–Saving Excision With Electrosurgical Peeling

Skin tissue–saving excision with electrosurgical peeling was first introduced in 2015.13 Blok et al14 described the procedure as a promising alternative to wide surgical excision for Hurley stage II or III HS. The procedure saves healthy tissue while completely removing lesional tissue, leading to rapid wound healing, excellent cosmesis, and a low risk of contractures2,14; however, recurrence rates are higher than those seen in wide surgical excision.15 There are no known RCTs with long-term follow-up for HS patients treated with skin tissue–saving excision with electrosurgical peeling.

The procedure typically is performed under general anesthesia.14 First, the sinus tract is palpated on physical examination and probed to delineate the extent of the tract. Next, the roof of the tract is incised electrosurgically with a wire loop tip coupled to an electrosurgical generator.14 Consecutive tangential excisions are made until the floor of the sinus tract is reached. The process of incising sinus tracts followed by tangential peeling off of tissue continues until the entire area is clear of lesional and fibrotic tissue. The wound margins are probed for the presence and subsequent removal of residual sinus tracts. Lastly, the electrosurgical generator is used to achieve hemostasis, steroids are injected to prevent the formation of hypergranulation tissue, and the wound is left to heal by secondary intention.14 Following intervention, recurrence rates appear to be similar to wide surgical excision.13,14

 

 

Wide Surgical Excision

Wide excision is a widely established technique consisting of surgical excision of a lesion plus an area of surrounding disease-free tissue such as subcutaneous fat or a lateral margin of intertriginous skin.15 Similar to other surgical techniques, wide excision is considered in cases of severe disease when pharmacologic management cannot remedy extensive fibrosis or architectural loss. It typically is performed in Hurley stage II and III HS, with pathology extending to involve deeper structures inaccessible to more superficial surgical methods.2 Prominent areas of use include gluteal, axillary, perineal, and perianal HS lesions on which conservative treatments have little effect and depend on wide excision to provide successful postoperative results.16 Although retrospective and prospective studies exist on wide excision in HS, there continues to be a dearth of RCTs. Based on the available literature, the primary motive for wide excision is lower recurrence rates (13% overall compared to 22% and 27% for local excision and deroofing, respectively) and longer asymptomatic periods compared to more local techniques.7,17 Wide excision combined with continued aggressive medical management and dietary modifications currently is an efficacious treatment in providing functional long-term results.6 These benefits, however, are not without their drawbacks, as the more extensive nature of wide surgical excision predisposes patients to larger wounds, surgery-induced infection, and prolonged recovery periods.6,15 If preoperative measurements are not wisely assessed, the excision also can extend to involve neurovascular bundles and other vital structures, contributing to greater postoperative morbidity.15 Ultrasonography provides useful anatomic information in HS, such as location and extent of fistulous tracts and fluid collections; these findings can assist in guiding the width and depth of the excision itself to ensure the entire area of HS involvement is removed.18 Published data revealed that 204 of 255 (80%) patients were markedly satisfied with postoperative outcomes of wide excision,19 which gives credence to the idea that although the complications of wide excision may not be as favorable, the long-lasting improvements in quality of life make wide surgical excision a suggested first-line treatment in all stages of HS.16,20

Closure Techniques

The best skin closure method following surgical excision is controversial and not well established in literature. Options include healing by secondary intention, primary (suture-based) closure, skin grafts, and skin flaps. Each of these methods has had moderate success in multiple observational studies, and the choice should be made based on individualized assessment of the patient’s HS lesion characteristics, ability to adhere to recovery protocols, and relevant demographics. A systematic review by Mehdizadeh et al17 provided the following recurrence rates for techniques utilized after wide excision: primary closure, 15%; flaps, 8%; and grafting, 6%. Despite conflicting evidence, allowing wounds to heal by secondary intention is best, based on the author’s experience (I.H.H.).

Secondary Intention
Healing by secondary intention refers to a wound that is intentionally left open to be filled in with granulation tissue and eventual epithelization over time rather than being approximated and closed via sutures or staples as in primary intention. It is a well-established option in wound management and results in a longer but more comfortable period of convalescence in postsurgical HS management.20 Patients can add regular moist wound dressings (eg, silastic foam dressing) to manage the wound at home and continue normal activities for most of the healing period; however, the recovery period can become excessively long and painful, and there is a high risk of formation of retractile scar bands at and around the healing site.12 Strict adherence to wound-healing protocols is paramount to minimizing unwanted complications.21 Secondary intention often is used after wide local excision and has been demonstrated to yield positive functional and aesthetic results in multiple studies, especially in the more severe Hurley stage II or III cases.21,22 It can be successfully employed after laser treatment and in surgical defects of all sizes with little to no contractures or reduced range of motion.6 Ultimately, the choice to heal via secondary intention should be made after thorough assessment of patient needs and with ample education to ensure compliance.

Primary Closure
Primary closure is the suture-mediated closing technique that is most often used in wound closure for lower-grade HS cases, especially smaller excisions. However, it is associated with potential complications. If HS lesions are not effectively excised, disease can then recur at the periphery of the excision and wound dehiscence can manifest more readily, especially as wound size increases.23 Consequently, primary closure is associated with the highest recurrence rates among closure techniques.17 Avoiding primary closure in active disease also is recommended due to the potential of burying residual foci of inflammation.6 Finally, primary closures lack skin coverage and thus often are not viable options in most perianal and genital lesions that require more extensive reconstruction. Retrospective case series and case reports exist on primary excision, but further study is needed.

Skin Grafts
Skin grafting is a technique of surgically transplanting a piece of healthy skin from one body site to another. Skin grafts typically are used when primary closure or skin flaps are not feasible (eg, in large wounds) and also when shorter time to wound closure is a greater concern in patient recovery.2,24 Additionally, skin grafts can be employed on large flat surfaces of the body, such as the buttocks or thighs, for timely wound closure when wound contraction is less effective or wound healing is slow via epithelization. Types of skin graft techniques include split-thickness skin graft (STSG), full-thickness skin graft, and recycled skin graft. All 3 types have demonstrated acceptable functional and aesthetic results in observational studies and case reports, and thus deciding which technique to use should include individualized assessment.2,25 The STSG has several advantages over the full-thickness skin graft, including hairlessness (ie, without hair follicles), ease of harvest, and a less complicated transfer to contaminated lesional areas such as those in HS.26 Additionally, STSGs allow for closure of even the largest wounds with minimal risk of serious infection. Split-thickness skin grafts are considered one of the most efficacious tools for axilla reconstruction; however, they require prolonged immobilization of the arm, result in sequelae in donor sites, and do not always prevent retractile scars.26 The recycled skin graft technique can be used to treat chronic gluteal HS, but reliability and outcomes have not been reported. Skin grafting after excision is associated with increased pain, immobilization, prolonged hospitalization, and longer healing times compared to skin flaps.19 In a systematic review of wound healing techniques following wide excision, grafting was shown to have the lowest recurrence rate (6%) compared to skin flaps (8%) and primary closure (15%).17 The absence of hair follicles and sweat glands in STSGs may be advantageous in HS because both hair follicles and sweat glands are thought to play more roles in the pathogenesis of HS.18,24 Most studies on skin grafts are limited to case reports.

Skin Flaps
Skin flaps are similar to skin grafts in that healthy skin is transplanted from one site to another; the difference is that flaps maintain an intact blood supply, whereas skin grafts depend on growth of new blood vessels.12,13 The primary advantage of skin flaps is that they provide the best quality of skin due to the thick tissue coverage, which is an important concern, especially in aesthetic scenarios. Additionally, they have been shown to provide shorter healing times than grafts, primary closure, and secondary healing, which can be especially important when functional disability is a concern in the postoperative period.26 However, their use should be limited due to several complications owing to their blood supply, as there is a high risk of ischemia to distant portions of flaps, which often can progress to necrosis and hemorrhage during the harvesting process.2 Thus, skin flaps are incredibly difficult to use in larger wounds and often require debulking due to their thickness. Additionally, skin flaps are definitive by nature, which can pose an issue if HS recurs locally. Skin flaps are recommended only when their use is mandatory, such as in the coverage of important anatomic structures (eg, exposed neurovascular bundles and large vessels).2 Advances have been made in flap construction, and now several types of flaps are employed in several body areas with differing indications and recommendations.2,21 As with skin grafts, most studies in the literature are case reports; therefore, further investigation is needed.

Combination Reconstructions
Combination reconstructions refer to the simultaneous use of multiple closure or healing techniques. By combining 2 or more methods, surgeons can utilize the advantages of each technique to provide an individualized approach that can substantially diminish wound surface area and accelerate wound healing.2 For example, with the starlike technique, 5 equilateral triangles bordering a foci of axillary disease are excised in addition to the central foci, and the edges of each triangle are then sutured together to create a final scar of considerably smaller size. The starlike technique allows the wound to be partially sutured while leaving the remaining area to heal by secondary intention.2 There are a small number of case series and prospective studies on combined reconstructions in HS but no RCTs.

Conclusion

Many procedures exist as options for treatment of patients with HS. Deroofing and cryoinsufflation are options for localized Hurley stage I or II disease. For more severe Hurley stage II or III disease, skin tissue–saving excision with electrosurgical peeling or wide surgical excisions are preferred. Following excision, there are many options for wound closure, but our preference is to allow the wound to heal by secondary intention. It is imperative that dermatologists are informed on the different techniques for treating this disease to determine the best route of care for their patients.

References
  1. Smith MK, Nicholson CL, Parks-Miller A, et al. Hidradenitis suppurativa: an update on connecting the tracts. F1000Res. 2017;6:1272.
  2. Janse I, Bieniek A, Horvath B, et al. Surgical procedures in hidradenitis suppurativa. Dermatol Clin. 2016;34:97-109.
  3. Pagliarello C, Fabrizi G, Feliciani C, et al. Cryoinsufflation for Hurley stage II hidradenitis suppurativa: a useful treatment option when systemic therapies should be avoided. JAMA Dermatol. 2014;150:765-766.
  4. Pagliarello C, Fabrizi G, di Nuzzo S. Cryoinsufflation for hidradenitis suppurativa: technical refinement to prevent complications. Dermatol Surg. 2016;42:130-132.
  5. Ritz JP, Runkel N, Haier J, et al. Extent of surgery and recurrence rate of hidradenitis suppurativa. Int J Colorectal Dis. 1998;13:164-168.
  6. Danby FW, Hazen PG, Boer J. New and traditional surgical approaches to hidradenitis suppurativa. J Am Acad Dermatol. 2015;73(5, suppl 1):S62-S65.
  7. Ellis LZ. Hidradenitis suppurativa: surgical and other management techniques. Dermatol Surg. 2012;38:517-536.
  8. Mullins JF, McCash WB, Boudreau RF. Treatment of chronic hidradenitis suppurativa: surgical modification. Postgrad Med. 1959;26:805-808.
  9. Brown SC, Kazzazi N, Lord PH. Surgical treatment of perineal hidradenitis suppurativa with special reference to recognition of the perianal form. Br J Surg. 1986;73:978-980.
  10. Culp CE. Chronic hidradenitis suppurativa of the anal canal. a surgical skin disease. Dis Colon Rectum. 1983;26:669-676.
  11. van der Zee HH, Prens EP, Boer J. Deroofing: a tissue-saving surgical technique for the treatment of mild to moderate hidradenitis suppurativa lesions. J Am Acad Dermatol. 2010;63:475-480.
  12. Lin CH, Chang KP, Huang SH. Deroofing: an effective method for treating chronic diffuse hidradenitis suppurativa. Dermatol Surg. 2016;42:273-275.
  13. Blok JL, Boersma M, Terra JB, et al. Surgery under general anaes-thesia in severe hidradenitis suppurativa: a study of 363 primary operations in 113 patients. J Eur Acad Dermatol Venereol. 2015;29:1590-1597.
  14. Blok JL, Spoo JR, Leeman FW, et al. Skin-Tissue-sparing Excision with Electrosurgical Peeling (STEEP): a surgical treatment option for severe hidradenitis suppurativa Hurley stage II/III. J Eur Acad Dermatol Venereol. 2015;29:379-382.
  15. Saunte DML, Jemec GBE. Hidradenitis suppurativa: advances in diagnosis and treatment. JAMA. 2017;318:2019-2032.
  16. Maghsoudi H, Almasi H, Miri Bonjar MR. Men, main victims of hidradenitis suppurativa (a prospective cohort study). Int J Surg. 2018;50:6-10.
  17. Mehdizadeh A, Hazen PG, Bechara FG, et al. Recurrence of hidradenitis suppurativa after surgical management: a systematic review and meta-analysis. J Am Acad Dermatol. 2015;73(5, suppl 1):S70-S77.
  18. Wortsman X, Moreno C, Soto R, et al. Ultrasound in-depth characterization and staging of hidradenitis suppurativa. Dermatol Surg. 2013;39:1835-1842.
  19. Kofler L, Schweinzer K, Heister M, et al. Surgical treatment of hidradenitis suppurativa: an analysis of postoperative outcome, cosmetic results and quality of life in 255 patients [published online February 17, 2018]. J Eur Acad Dermatol Venereol. doi:10.1111/jdv.14892.
  20. Dini V, Oranges T, Rotella L, et al. Hidradenitis suppurativa and wound management. Int J Low Extrem Wounds. 2015;14:236-244.
  21. Humphries LS, Kueberuwa E, Beederman M, et al. Wide excision and healing by secondary intent for the surgical treatment of hidradenitis suppurativa: a single-center experience. J Plast Reconstr Aesthet Surg. 2016;69:554-566.
  22. Wollina U, Langner D, Heinig B, et al. Comorbidities, treatment, and outcome in severe anogenital inverse acne (hidradenitis suppurativa): a 15-year single center report. Int J Dermatol. 2017;56:109-115.
  23. Watson JD. Hidradenitis suppurativa—a clinical review. Br J Plast Surg. 1985;38:567-569.
  24. Sugio Y, Tomita K, Hosokawa K. Reconstruction after excision of hidradenitis suppurativa: are skin grafts better than flaps? Plast Reconstr Surg Glob Open. 2016;4:E1128.
  25. Burney RE. 35-year experience with surgical treatment of hidradenitis suppurativa. World J Surg. 2017;41:2723-2730.
  26. Nail-Barthelemy R, Stroumza N, Qassemyar Q, et al. Evaluation of the mobility of the shoulder and quality of life after perforator flaps for recalcitrant axillary hidradenitis [published online February 13, 2018]. Ann Chir Plast Esthet. pii:S0294-1260(18)30005-0. doi:10.1016/j.anplas.2018.01.003.
References
  1. Smith MK, Nicholson CL, Parks-Miller A, et al. Hidradenitis suppurativa: an update on connecting the tracts. F1000Res. 2017;6:1272.
  2. Janse I, Bieniek A, Horvath B, et al. Surgical procedures in hidradenitis suppurativa. Dermatol Clin. 2016;34:97-109.
  3. Pagliarello C, Fabrizi G, Feliciani C, et al. Cryoinsufflation for Hurley stage II hidradenitis suppurativa: a useful treatment option when systemic therapies should be avoided. JAMA Dermatol. 2014;150:765-766.
  4. Pagliarello C, Fabrizi G, di Nuzzo S. Cryoinsufflation for hidradenitis suppurativa: technical refinement to prevent complications. Dermatol Surg. 2016;42:130-132.
  5. Ritz JP, Runkel N, Haier J, et al. Extent of surgery and recurrence rate of hidradenitis suppurativa. Int J Colorectal Dis. 1998;13:164-168.
  6. Danby FW, Hazen PG, Boer J. New and traditional surgical approaches to hidradenitis suppurativa. J Am Acad Dermatol. 2015;73(5, suppl 1):S62-S65.
  7. Ellis LZ. Hidradenitis suppurativa: surgical and other management techniques. Dermatol Surg. 2012;38:517-536.
  8. Mullins JF, McCash WB, Boudreau RF. Treatment of chronic hidradenitis suppurativa: surgical modification. Postgrad Med. 1959;26:805-808.
  9. Brown SC, Kazzazi N, Lord PH. Surgical treatment of perineal hidradenitis suppurativa with special reference to recognition of the perianal form. Br J Surg. 1986;73:978-980.
  10. Culp CE. Chronic hidradenitis suppurativa of the anal canal. a surgical skin disease. Dis Colon Rectum. 1983;26:669-676.
  11. van der Zee HH, Prens EP, Boer J. Deroofing: a tissue-saving surgical technique for the treatment of mild to moderate hidradenitis suppurativa lesions. J Am Acad Dermatol. 2010;63:475-480.
  12. Lin CH, Chang KP, Huang SH. Deroofing: an effective method for treating chronic diffuse hidradenitis suppurativa. Dermatol Surg. 2016;42:273-275.
  13. Blok JL, Boersma M, Terra JB, et al. Surgery under general anaes-thesia in severe hidradenitis suppurativa: a study of 363 primary operations in 113 patients. J Eur Acad Dermatol Venereol. 2015;29:1590-1597.
  14. Blok JL, Spoo JR, Leeman FW, et al. Skin-Tissue-sparing Excision with Electrosurgical Peeling (STEEP): a surgical treatment option for severe hidradenitis suppurativa Hurley stage II/III. J Eur Acad Dermatol Venereol. 2015;29:379-382.
  15. Saunte DML, Jemec GBE. Hidradenitis suppurativa: advances in diagnosis and treatment. JAMA. 2017;318:2019-2032.
  16. Maghsoudi H, Almasi H, Miri Bonjar MR. Men, main victims of hidradenitis suppurativa (a prospective cohort study). Int J Surg. 2018;50:6-10.
  17. Mehdizadeh A, Hazen PG, Bechara FG, et al. Recurrence of hidradenitis suppurativa after surgical management: a systematic review and meta-analysis. J Am Acad Dermatol. 2015;73(5, suppl 1):S70-S77.
  18. Wortsman X, Moreno C, Soto R, et al. Ultrasound in-depth characterization and staging of hidradenitis suppurativa. Dermatol Surg. 2013;39:1835-1842.
  19. Kofler L, Schweinzer K, Heister M, et al. Surgical treatment of hidradenitis suppurativa: an analysis of postoperative outcome, cosmetic results and quality of life in 255 patients [published online February 17, 2018]. J Eur Acad Dermatol Venereol. doi:10.1111/jdv.14892.
  20. Dini V, Oranges T, Rotella L, et al. Hidradenitis suppurativa and wound management. Int J Low Extrem Wounds. 2015;14:236-244.
  21. Humphries LS, Kueberuwa E, Beederman M, et al. Wide excision and healing by secondary intent for the surgical treatment of hidradenitis suppurativa: a single-center experience. J Plast Reconstr Aesthet Surg. 2016;69:554-566.
  22. Wollina U, Langner D, Heinig B, et al. Comorbidities, treatment, and outcome in severe anogenital inverse acne (hidradenitis suppurativa): a 15-year single center report. Int J Dermatol. 2017;56:109-115.
  23. Watson JD. Hidradenitis suppurativa—a clinical review. Br J Plast Surg. 1985;38:567-569.
  24. Sugio Y, Tomita K, Hosokawa K. Reconstruction after excision of hidradenitis suppurativa: are skin grafts better than flaps? Plast Reconstr Surg Glob Open. 2016;4:E1128.
  25. Burney RE. 35-year experience with surgical treatment of hidradenitis suppurativa. World J Surg. 2017;41:2723-2730.
  26. Nail-Barthelemy R, Stroumza N, Qassemyar Q, et al. Evaluation of the mobility of the shoulder and quality of life after perforator flaps for recalcitrant axillary hidradenitis [published online February 13, 2018]. Ann Chir Plast Esthet. pii:S0294-1260(18)30005-0. doi:10.1016/j.anplas.2018.01.003.
Issue
Cutis - 102(1)
Issue
Cutis - 102(1)
Page Number
13-16
Page Number
13-16
Publications
Publications
Topics
Article Type
Display Headline
Surgical Procedures for Hidradenitis Suppurativa
Display Headline
Surgical Procedures for Hidradenitis Suppurativa
Sections
Inside the Article

Practice Points

  • Surgical intervention currently is the only definitive treatment for hidradenitis suppurativa (HS).
  • There is no consensus on the best surgical intervention for long-term outcomes in HS; rather, approach is based on clinical judgment dependent upon the location and severity of lesions.
  • After wide excision, allow wounds to heal by secondary intention.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Article PDF Media

Melanoma in US Hispanics: Recommended Strategies to Reduce Disparities in Outcomes

Article Type
Changed
Thu, 10/29/2020 - 14:52
Display Headline
Melanoma in US Hispanics: Recommended Strategies to Reduce Disparities in Outcomes
In Collaboration with the Skin of Color Society

Cutaneous melanoma is a considerable public health concern. In the United States, an estimated 87,110 cases were diagnosed in 2017, and more than 9000 deaths are expected as result of this disease in 2018.1 Early diagnosis of melanoma is associated with favorable survival rates (5-year overall survival rates for melanoma in situ and stage IA melanoma, 99% and 97%, respectively).2 In contrast, the prognosis for advanced-stage melanoma is poor, with a 5-year survival rate of 16% for patients with stage IV disease. Therefore, early detection is critical to reducing mortality in melanoma patients.3

The term Hispanic refers to a panethnic category primarily encompassing Mexican-Americans, Cubans, and Puerto Ricans, as well as individuals from the Caribbean and Central and South America. These populations are diverse in birth origin, primary language, acculturation, distinct ethnic traditions, education level, and occupation. Hispanics in the United States are heterogeneous in many dimensions related to health risks, health care use, and health outcomes.4 Genetic predisposition, lifestyle risks, and access to and use of health care services can shape melanoma diagnosis, treatment, and progression across Hispanic populations differently than in other populations.

In this review, the epidemiology and clinical presentation of melanoma in US Hispanics is summarized, and recommendations for a research agenda to advance understanding of this disease in the most rapidly growing segment of the US population is provided.

Melanoma Incidence, Presentation, and Outcomes in US Hispanics

In the period from 2008 to 2012, the age-adjusted incidence of melanoma in US Hispanics (4.6 per 100,00 men and 4.2 per 100,00 women) was lower than in NHWs.5 Garnett et al5 reported a decline in melanoma incidence in US Hispanics between 2003 and 2012—an observation that stands in contrast to state-level studies in California and Florida, in which small but substantial increases in melanoma incidence among Hispanics were reported.6,7 The rising incidence of melanomas thicker than 1.5 mm at presentation among Hispanic men living in California is particularly worrisome.6 Discrepancies in incidence trends might reflect changes in incidence over time or differences in state-level registry reporting of melanoma.5

Despite a lower overall incidence of melanoma in US Hispanics, those who do develop the disease are 2.4 times more likely (age-adjusted odds ratio) to present with stage III disease (confidence interval, 1.89-3.05)8 and are 3.64 times more likely to develop distant metastases (confidence interval, 2.65-5.0) than NHWs.3,7,9-13 Disparities also exist in the diagnosis of childhood melanoma: Hispanic children and adolescents who have a diagnosis of melanoma are 3 times more likely to present with advanced disease than NHW counterparts.14 Survival analyses by age and stage show considerably lower survival among Hispanic patients compared to NHW patients with stage I and II disease. In part, worse survival outcomes among Hispanics are the result of the pattern of more advanced disease at presentation.8,14,15

Late presentation for evaluation of melanomas in Hispanics has been attributed to a number of variables, including a lack of skin cancer awareness and knowledge,9,16 a lower rate of self- and physician-performed skin examinations,10 differences in tumor biology,9 and socioeconomic forces.7,17

In a previous study investigating the relationship between neighborhood characteristics and tumor stage at melanoma diagnosis in Hispanic men in California, Texas, and Florida, several key findings emerged.17 First, residency in a census tract with a high density of immigrants (California, Texas) and a high composition of Hispanics (California, Florida) was an important predictor of a late-stage melanoma diagnosis in fully adjusted models. Additionally, the strength of association between measures of socioeconomic status (ie, poverty and education) and tumor stage at melanoma diagnosis was attenuated in multivariate models when enclaves and availability of primary care resources were taken into account. Hispanic melanoma cases in areas with a low density of primary care physicians had an increased likelihood of late-stage diagnosis in California and Texas. The probability of late-stage diagnosis was concentrated in specific regions along the United States–Mexico border, in south central California, and along the southeastern coast of Florida. Lastly, in Texas, Hispanic men aged 18 to 34 years and 35 to 49 years were at an increased risk of late-stage melanoma diagnosis compared to men 65 years and older.17

 

 

Demographic and Clinical Characteristics of Melanoma in Hispanic Patients

Among Hispanics, white Hispanics comprise the majority of melanoma cases.5 Median age at diagnosis is younger in Hispanics compared to whites.5,6 Hispanic men typically are older (median age, 61 years) than Hispanic women (median age, 52 years) at diagnosis.5 Similar to what is seen in NHWs, young Hispanic women experience a higher melanoma incidence than young Hispanic men.5 Among older Hispanics, melanoma is more common in men.5,8

Melanomas located on the lower extremities and hips are more prevalent in Hispanics than in NHWs.5,8,18 Among Hispanics, there are age- and sex-based variations in the anatomic location of primary tumors: in Hispanic men, truncal tumors predominate, and in Hispanic women, tumors of the lower extremities are most common across all age groups.5 The incidence of melanomas located in the head and neck region increases with age for both Hispanic men and women.

For melanomas in which the histologic type is known, superficial spreading melanoma is the most common subtype among Hispanics.5,17,19 Acral lentiginous melanomas and nodular melanomas are more common among Hispanics than among NHWs.5,17,19

The observation that Hispanics with melanoma are more prone to lower-extremity tumors and nodular and acral lentiginous melanoma subtypes than NHWs suggests that UV exposure history may be of less importance in this population. Although numerous studies have explored melanoma risk factors in NHWs, there is a striking paucity of such studies in Hispanics. For example, there are conflicting data regarding the role of UV exposure in melanoma risk among Hispanics. Hu et al20 found that UV index and latitude correlated with melanoma risk in this population, whereas Eide et al21 found no association between UV exposure and melanoma incidence in Hispanics. A prospective study involving a multiethnic cohort (of whom 40 of the 107 participants were Hispanic) found no clear association between a history of sunburn and melanoma risk in Hispanics.18

Strategies for Reducing Disparities in Outcomes

Our knowledge of melanoma epidemiology in Hispanics derives mainly from secondary analyses of state-level and national cancer registry data sets.5-8,13-15,17,19,20 These administrative data sources often are limited by missing data (eg, tumor thickness, histologic subtype) or lack important patient-level information (eg, self-identified race and ethnicity, health insurance status). Additionally, the manner in which data are collected and integrated into research varies; for example, socioeconomic measures often are reported as either area-based or composite measures. Thus, there is a need to improve the consistency of reporting on demographic and socioeconomic measures across studies. Polite et al22 recommended standardization of reporting criteria and that a standard set of demographic and socioeconomic status measures be included in clinical registries and research protocols.22 Researchers should strive to collect self-reported information on race and ethnicity, as well as the most granular level of detail on health insurance status, ancestry, and immigration status.

The host phenotypic characteristics of melanoma in NHWs are well understood, but the biological and environmental determinants of melanoma risk in Hispanics and other minorities are unknown. For example, fair complexion, red hair, blue eyes, increased freckling density, and the presence of numerous dysplastic and common melanocytic nevi indicate a propensity toward cutaneous melanoma.23,24 However, the relevance of such risk factors in Hispanics is unknown and has not been widely investigated in this patient population. Park et al18 found that a person’s sunburn susceptibility phenotype (defined as hair and eye color, ability to tan, and skin reaction to sunlight) was associated with an increased risk of melanoma among nonwhite, multiracial individuals. However, this study was limited by a small number of minority cases, which included only 40 Hispanic participants with melanoma.18 There is a need for rigorous observational studies to clearly define the phenotypic characteristics, sun-exposure behavior patterns, and genetic contributors to melanoma genesis in Hispanics.

The biologic determinants of postdiagnosis survival in Hispanics with melanoma are not well understood. It is unknown if genetic predisposition modifies melanoma risk in Hispanics. For example, the frequency of BRAF gene mutation or other driver mutations in US Hispanics has been understudied. It is important to know if mutation frequency patterns differ in Hispanics patients compared to NHWs because this knowledge could have considerable implications for treatment. Several recommendations should be considered to address these knowledge gaps. First, there is a need for development or enhancement of melanoma biorepositories, which should include tumor and nontumor specimens from a diverse sample of melanoma patients. Additionally, multi-institutional and multidisciplinary consortiums need to be created in order to amass a number of Hispanic melanoma patients to identify genetic, biologic, and behavioral risk factors specific to this subgroup of patients. The AMBER Consortium, which focuses on breast cancer epidemiology and risk in black women, is a model for the type of consortium needed for the study of melanoma in Hispanics.25 Lastly, community engagement will be central to developing sustainable recruitment and data-collection efforts.26 Involvement of key stakeholders will provide an in-depth assessment of community needs as well as real-time feedback on the process and practicality of research questions. Buy-in from affected communities also may facilitate dissemination of research findings to affected communities.

Conclusion

Hispanics are more likely to present with an advanced stages of disease and have higher melanoma-specific mortality rates than NHWs. Regrettably, a huge knowledge gap exists regarding contributors and solutions to melanoma disparities among this fast-growing, understudied segment of the US population. Accordingly, critical research is needed to address the most pressing questions regarding melanoma risk and poor outcomes among Hispanics to foster implementation of interventional efforts in prevention, early detection, and treatment. A multi-institutional and multidisciplinary approach across multiple levels is needed to eliminate disparate outcomes. Although melanoma is relatively uncommon among Hispanics, studies of melanoma in Hispanics (given their diverse genetic ancestry and migration) provide a unique backdrop against which researchers can explicate melanoma etiology—thus benefiting Hispanics and non-Hispanics alike.

References
  1. American Cancer Society. Key statistics for melanoma skin cancer. www.cancer.org/cancer/melanoma-skin-cancer/about/key-statistics.html. Accessed January 13, 2018.
  2. Balch CM, Gershenwald JE, Soong S, et al. Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol. 2009;27:6199-6206.
  3. Katalinic A, Waldmann A, Weinstock MA, et al. Does skin cancer screening save lives? Cancer. 2012;118:5395-5402.
  4. Bergad LW, Klein HS. Hispanics in the United States: A Demographic, Social, and Economic History, 1980-2005. New York, NY: Cambridge University Press; 2010.
  5. Garnett E, Townsend J, Steele B, et al. Characteristics, rates, and trends of melanoma incidence among Hispanics in the USA. Cancer Causes Control. 2016;27:647-659.
  6. Pollitt RA, Clarke CA, Swetter SM, et al. The expanding melanoma burden in California Hispanics: importance of socioeconomic distribution, histologic subtype, and anatomic location. Cancer. 2011;117:152-161.
  7. Hu S, Parmet, Y, Allen G, et al. Disparity in melanoma: a trend analysis of melanoma incidence and stage at diagnosis among whites,Hispanics, and blacks in Florida. JAMA Dermatology. 2010;145:1369-1374.
  8. Cormier JN, Xing Y, Ding M, et al. Ethnic differences among patients with cutaneous melanoma. Arch Intern Med. 2006;166:1907-1914.
  9. Pollitt RA, Swetter SM, Johnson TM, et al. Examining the pathways linking lower socioeconomic status and advanced melanoma. Cancer. 2012;118:4004-4013.
  10. Ortiz CA, Goodwin JS, Freeman JL. The effect of socioeconomic factors on incidence, stage at diagnosis and survival of cutaneous melanoma. Med Sci Monit. 2005;11:RA163-RA172.
  11. Singh SD, Ajani UA, Johnson CJ, et al. Association of cutaneous melanoma incidence with area-based socioeconomic indicators-United States, 2004-2006. J Am Acad Dermatol. 2011;65(5 suppl 1):S58-S68.
  12. Pollitt RA, Clarke CA, Shema SJ, et al. California Medicaid enrollment and melanoma stage at diagnosis: a population-based study. Am J Prev Med. 2008;35:7-13.
  13. Clairwood M, Ricketts J, Grant-Kels J, et al. Melanoma in skin of color in Connecticut: an analysis of melanoma incidence and stage at diagnosis in non-Hispanic blacks, non-Hispanic whites, and Hispanics. Int J Dermatol. 2014;53:425-433.
  14. Hamilton EC, Nguyen HT, Chang YC, et al. Health disparities influence childhood melanoma stage at diagnosis and outcome. J Pediatr. 2016;175:182-187.
  15. Dawes SM, Tsai S, Gittleman H, et al. Racial disparities in melanoma survival. J Am Acad Dermatol. 2016;75:983-991.
  16. Imahiyerobo-Ip J, Ip I, Jamal S, et al. Skin cancer awareness in communities of color. J Am Acad Dermatol. 2011;64:198-200.
  17. Harvey VM, Enos CW, Chen JT, et al. The role of neighborhood characteristics in late stage melanoma diagnosis among Hispanic men in California, Texas, and Florida, 1996-2012 [published online June 18, 2017]. J Cancer Epidemiol. 2017;2017:8418904.
  18. Park SL, Le Marchand L, Wilkens LR, et al. Risk factors for malignant melanoma in white and non-white/non-African American populations: the multiethnic cohort. Cancer Prev Res. 2012;5:423-434.
  19. Wu XC, Eide MJ, King J, et al. Racial and ethnic variations in incidence and survival of cutaneous melanoma in the United States, 1999-2006. J Am Acad Dermatol. 2011;65(5 suppl 1):S26-S37.
  20. Hu S, Ma F, Collado-Mesa F, et al. UV radiation, latitude, and melanoma in US Hispanics and blacks. Arch Dermatol. 2004;140:819-824.
  21. Eide MJ, Weinstock MA. Association of UV index, latitude, and melanoma incidence in nonwhite populations—US Surveillance, Epidemiology, and End Results (SEER) program, 1992 to 2001. Arch Dermatol. 2005;141:477-481.
  22. Polite BN, Adams-Campbell LL, Brawley OW, et al. Charting the future of cancer health disparities research: a position statement from the American Association for Cancer Research, the American Cancer Society, the American Society of Clinical Oncology, and the National Cancer Institute. Cancer Res. 2017;77:4548-4555.
  23. Gandini S, Sera F, Cattaruzza MS, et al. Meta-analysis of risk factors for cutaneous melanoma: III. family history, actinic damage and phenotypic factors. Eur J Cancer. 2005;41:2040-2059.
  24. Chang YM, Newton-Bishop JA, Bishop DT, et al. A pooled analysis of melanocytic nevus phenotype and the risk of cutaneous melanoma at different latitudes. Int J Cancer. 2009;124:420-428.
  25. Palmer JR, Ambrosone CB, Olshan AF. A collaborative study of the etiology of breast cancer subtypes in African American women: the AMBER consortium. Cancer Causes Control. 2014;25:309-319.
  26. Rapkin BD, Weiss E, Lounsbury D, et al. Reducing disparities in cancer screening and prevention through community-based participatory research partnerships with local libraries: a comprehensive dynamic trial. Am J Community Psychol. 2017;60:145-159.
Article PDF
Author and Disclosure Information

From the Hampton University Skin of Color Research Institute, Virginia, and the Hampton Roads Center for Dermatology, Newport News, Virginia.

Dr. Harvey serves on the speakers’ bureau for Aclaris Therapeutics, Inc.

Correspondence: Valerie M. Harvey, MD, MPH, Hampton University Skin of Color Research Institute, P.O. Box 6035, Hampton University, Hampton VA 23668 (Valerieharvey10@gmail.com).

Issue
Cutis - 101(4)
Publications
Topics
Page Number
243-246
Sections
Author and Disclosure Information

From the Hampton University Skin of Color Research Institute, Virginia, and the Hampton Roads Center for Dermatology, Newport News, Virginia.

Dr. Harvey serves on the speakers’ bureau for Aclaris Therapeutics, Inc.

Correspondence: Valerie M. Harvey, MD, MPH, Hampton University Skin of Color Research Institute, P.O. Box 6035, Hampton University, Hampton VA 23668 (Valerieharvey10@gmail.com).

Author and Disclosure Information

From the Hampton University Skin of Color Research Institute, Virginia, and the Hampton Roads Center for Dermatology, Newport News, Virginia.

Dr. Harvey serves on the speakers’ bureau for Aclaris Therapeutics, Inc.

Correspondence: Valerie M. Harvey, MD, MPH, Hampton University Skin of Color Research Institute, P.O. Box 6035, Hampton University, Hampton VA 23668 (Valerieharvey10@gmail.com).

Article PDF
Article PDF
In Collaboration with the Skin of Color Society
In Collaboration with the Skin of Color Society

Cutaneous melanoma is a considerable public health concern. In the United States, an estimated 87,110 cases were diagnosed in 2017, and more than 9000 deaths are expected as result of this disease in 2018.1 Early diagnosis of melanoma is associated with favorable survival rates (5-year overall survival rates for melanoma in situ and stage IA melanoma, 99% and 97%, respectively).2 In contrast, the prognosis for advanced-stage melanoma is poor, with a 5-year survival rate of 16% for patients with stage IV disease. Therefore, early detection is critical to reducing mortality in melanoma patients.3

The term Hispanic refers to a panethnic category primarily encompassing Mexican-Americans, Cubans, and Puerto Ricans, as well as individuals from the Caribbean and Central and South America. These populations are diverse in birth origin, primary language, acculturation, distinct ethnic traditions, education level, and occupation. Hispanics in the United States are heterogeneous in many dimensions related to health risks, health care use, and health outcomes.4 Genetic predisposition, lifestyle risks, and access to and use of health care services can shape melanoma diagnosis, treatment, and progression across Hispanic populations differently than in other populations.

In this review, the epidemiology and clinical presentation of melanoma in US Hispanics is summarized, and recommendations for a research agenda to advance understanding of this disease in the most rapidly growing segment of the US population is provided.

Melanoma Incidence, Presentation, and Outcomes in US Hispanics

In the period from 2008 to 2012, the age-adjusted incidence of melanoma in US Hispanics (4.6 per 100,00 men and 4.2 per 100,00 women) was lower than in NHWs.5 Garnett et al5 reported a decline in melanoma incidence in US Hispanics between 2003 and 2012—an observation that stands in contrast to state-level studies in California and Florida, in which small but substantial increases in melanoma incidence among Hispanics were reported.6,7 The rising incidence of melanomas thicker than 1.5 mm at presentation among Hispanic men living in California is particularly worrisome.6 Discrepancies in incidence trends might reflect changes in incidence over time or differences in state-level registry reporting of melanoma.5

Despite a lower overall incidence of melanoma in US Hispanics, those who do develop the disease are 2.4 times more likely (age-adjusted odds ratio) to present with stage III disease (confidence interval, 1.89-3.05)8 and are 3.64 times more likely to develop distant metastases (confidence interval, 2.65-5.0) than NHWs.3,7,9-13 Disparities also exist in the diagnosis of childhood melanoma: Hispanic children and adolescents who have a diagnosis of melanoma are 3 times more likely to present with advanced disease than NHW counterparts.14 Survival analyses by age and stage show considerably lower survival among Hispanic patients compared to NHW patients with stage I and II disease. In part, worse survival outcomes among Hispanics are the result of the pattern of more advanced disease at presentation.8,14,15

Late presentation for evaluation of melanomas in Hispanics has been attributed to a number of variables, including a lack of skin cancer awareness and knowledge,9,16 a lower rate of self- and physician-performed skin examinations,10 differences in tumor biology,9 and socioeconomic forces.7,17

In a previous study investigating the relationship between neighborhood characteristics and tumor stage at melanoma diagnosis in Hispanic men in California, Texas, and Florida, several key findings emerged.17 First, residency in a census tract with a high density of immigrants (California, Texas) and a high composition of Hispanics (California, Florida) was an important predictor of a late-stage melanoma diagnosis in fully adjusted models. Additionally, the strength of association between measures of socioeconomic status (ie, poverty and education) and tumor stage at melanoma diagnosis was attenuated in multivariate models when enclaves and availability of primary care resources were taken into account. Hispanic melanoma cases in areas with a low density of primary care physicians had an increased likelihood of late-stage diagnosis in California and Texas. The probability of late-stage diagnosis was concentrated in specific regions along the United States–Mexico border, in south central California, and along the southeastern coast of Florida. Lastly, in Texas, Hispanic men aged 18 to 34 years and 35 to 49 years were at an increased risk of late-stage melanoma diagnosis compared to men 65 years and older.17

 

 

Demographic and Clinical Characteristics of Melanoma in Hispanic Patients

Among Hispanics, white Hispanics comprise the majority of melanoma cases.5 Median age at diagnosis is younger in Hispanics compared to whites.5,6 Hispanic men typically are older (median age, 61 years) than Hispanic women (median age, 52 years) at diagnosis.5 Similar to what is seen in NHWs, young Hispanic women experience a higher melanoma incidence than young Hispanic men.5 Among older Hispanics, melanoma is more common in men.5,8

Melanomas located on the lower extremities and hips are more prevalent in Hispanics than in NHWs.5,8,18 Among Hispanics, there are age- and sex-based variations in the anatomic location of primary tumors: in Hispanic men, truncal tumors predominate, and in Hispanic women, tumors of the lower extremities are most common across all age groups.5 The incidence of melanomas located in the head and neck region increases with age for both Hispanic men and women.

For melanomas in which the histologic type is known, superficial spreading melanoma is the most common subtype among Hispanics.5,17,19 Acral lentiginous melanomas and nodular melanomas are more common among Hispanics than among NHWs.5,17,19

The observation that Hispanics with melanoma are more prone to lower-extremity tumors and nodular and acral lentiginous melanoma subtypes than NHWs suggests that UV exposure history may be of less importance in this population. Although numerous studies have explored melanoma risk factors in NHWs, there is a striking paucity of such studies in Hispanics. For example, there are conflicting data regarding the role of UV exposure in melanoma risk among Hispanics. Hu et al20 found that UV index and latitude correlated with melanoma risk in this population, whereas Eide et al21 found no association between UV exposure and melanoma incidence in Hispanics. A prospective study involving a multiethnic cohort (of whom 40 of the 107 participants were Hispanic) found no clear association between a history of sunburn and melanoma risk in Hispanics.18

Strategies for Reducing Disparities in Outcomes

Our knowledge of melanoma epidemiology in Hispanics derives mainly from secondary analyses of state-level and national cancer registry data sets.5-8,13-15,17,19,20 These administrative data sources often are limited by missing data (eg, tumor thickness, histologic subtype) or lack important patient-level information (eg, self-identified race and ethnicity, health insurance status). Additionally, the manner in which data are collected and integrated into research varies; for example, socioeconomic measures often are reported as either area-based or composite measures. Thus, there is a need to improve the consistency of reporting on demographic and socioeconomic measures across studies. Polite et al22 recommended standardization of reporting criteria and that a standard set of demographic and socioeconomic status measures be included in clinical registries and research protocols.22 Researchers should strive to collect self-reported information on race and ethnicity, as well as the most granular level of detail on health insurance status, ancestry, and immigration status.

The host phenotypic characteristics of melanoma in NHWs are well understood, but the biological and environmental determinants of melanoma risk in Hispanics and other minorities are unknown. For example, fair complexion, red hair, blue eyes, increased freckling density, and the presence of numerous dysplastic and common melanocytic nevi indicate a propensity toward cutaneous melanoma.23,24 However, the relevance of such risk factors in Hispanics is unknown and has not been widely investigated in this patient population. Park et al18 found that a person’s sunburn susceptibility phenotype (defined as hair and eye color, ability to tan, and skin reaction to sunlight) was associated with an increased risk of melanoma among nonwhite, multiracial individuals. However, this study was limited by a small number of minority cases, which included only 40 Hispanic participants with melanoma.18 There is a need for rigorous observational studies to clearly define the phenotypic characteristics, sun-exposure behavior patterns, and genetic contributors to melanoma genesis in Hispanics.

The biologic determinants of postdiagnosis survival in Hispanics with melanoma are not well understood. It is unknown if genetic predisposition modifies melanoma risk in Hispanics. For example, the frequency of BRAF gene mutation or other driver mutations in US Hispanics has been understudied. It is important to know if mutation frequency patterns differ in Hispanics patients compared to NHWs because this knowledge could have considerable implications for treatment. Several recommendations should be considered to address these knowledge gaps. First, there is a need for development or enhancement of melanoma biorepositories, which should include tumor and nontumor specimens from a diverse sample of melanoma patients. Additionally, multi-institutional and multidisciplinary consortiums need to be created in order to amass a number of Hispanic melanoma patients to identify genetic, biologic, and behavioral risk factors specific to this subgroup of patients. The AMBER Consortium, which focuses on breast cancer epidemiology and risk in black women, is a model for the type of consortium needed for the study of melanoma in Hispanics.25 Lastly, community engagement will be central to developing sustainable recruitment and data-collection efforts.26 Involvement of key stakeholders will provide an in-depth assessment of community needs as well as real-time feedback on the process and practicality of research questions. Buy-in from affected communities also may facilitate dissemination of research findings to affected communities.

Conclusion

Hispanics are more likely to present with an advanced stages of disease and have higher melanoma-specific mortality rates than NHWs. Regrettably, a huge knowledge gap exists regarding contributors and solutions to melanoma disparities among this fast-growing, understudied segment of the US population. Accordingly, critical research is needed to address the most pressing questions regarding melanoma risk and poor outcomes among Hispanics to foster implementation of interventional efforts in prevention, early detection, and treatment. A multi-institutional and multidisciplinary approach across multiple levels is needed to eliminate disparate outcomes. Although melanoma is relatively uncommon among Hispanics, studies of melanoma in Hispanics (given their diverse genetic ancestry and migration) provide a unique backdrop against which researchers can explicate melanoma etiology—thus benefiting Hispanics and non-Hispanics alike.

Cutaneous melanoma is a considerable public health concern. In the United States, an estimated 87,110 cases were diagnosed in 2017, and more than 9000 deaths are expected as result of this disease in 2018.1 Early diagnosis of melanoma is associated with favorable survival rates (5-year overall survival rates for melanoma in situ and stage IA melanoma, 99% and 97%, respectively).2 In contrast, the prognosis for advanced-stage melanoma is poor, with a 5-year survival rate of 16% for patients with stage IV disease. Therefore, early detection is critical to reducing mortality in melanoma patients.3

The term Hispanic refers to a panethnic category primarily encompassing Mexican-Americans, Cubans, and Puerto Ricans, as well as individuals from the Caribbean and Central and South America. These populations are diverse in birth origin, primary language, acculturation, distinct ethnic traditions, education level, and occupation. Hispanics in the United States are heterogeneous in many dimensions related to health risks, health care use, and health outcomes.4 Genetic predisposition, lifestyle risks, and access to and use of health care services can shape melanoma diagnosis, treatment, and progression across Hispanic populations differently than in other populations.

In this review, the epidemiology and clinical presentation of melanoma in US Hispanics is summarized, and recommendations for a research agenda to advance understanding of this disease in the most rapidly growing segment of the US population is provided.

Melanoma Incidence, Presentation, and Outcomes in US Hispanics

In the period from 2008 to 2012, the age-adjusted incidence of melanoma in US Hispanics (4.6 per 100,00 men and 4.2 per 100,00 women) was lower than in NHWs.5 Garnett et al5 reported a decline in melanoma incidence in US Hispanics between 2003 and 2012—an observation that stands in contrast to state-level studies in California and Florida, in which small but substantial increases in melanoma incidence among Hispanics were reported.6,7 The rising incidence of melanomas thicker than 1.5 mm at presentation among Hispanic men living in California is particularly worrisome.6 Discrepancies in incidence trends might reflect changes in incidence over time or differences in state-level registry reporting of melanoma.5

Despite a lower overall incidence of melanoma in US Hispanics, those who do develop the disease are 2.4 times more likely (age-adjusted odds ratio) to present with stage III disease (confidence interval, 1.89-3.05)8 and are 3.64 times more likely to develop distant metastases (confidence interval, 2.65-5.0) than NHWs.3,7,9-13 Disparities also exist in the diagnosis of childhood melanoma: Hispanic children and adolescents who have a diagnosis of melanoma are 3 times more likely to present with advanced disease than NHW counterparts.14 Survival analyses by age and stage show considerably lower survival among Hispanic patients compared to NHW patients with stage I and II disease. In part, worse survival outcomes among Hispanics are the result of the pattern of more advanced disease at presentation.8,14,15

Late presentation for evaluation of melanomas in Hispanics has been attributed to a number of variables, including a lack of skin cancer awareness and knowledge,9,16 a lower rate of self- and physician-performed skin examinations,10 differences in tumor biology,9 and socioeconomic forces.7,17

In a previous study investigating the relationship between neighborhood characteristics and tumor stage at melanoma diagnosis in Hispanic men in California, Texas, and Florida, several key findings emerged.17 First, residency in a census tract with a high density of immigrants (California, Texas) and a high composition of Hispanics (California, Florida) was an important predictor of a late-stage melanoma diagnosis in fully adjusted models. Additionally, the strength of association between measures of socioeconomic status (ie, poverty and education) and tumor stage at melanoma diagnosis was attenuated in multivariate models when enclaves and availability of primary care resources were taken into account. Hispanic melanoma cases in areas with a low density of primary care physicians had an increased likelihood of late-stage diagnosis in California and Texas. The probability of late-stage diagnosis was concentrated in specific regions along the United States–Mexico border, in south central California, and along the southeastern coast of Florida. Lastly, in Texas, Hispanic men aged 18 to 34 years and 35 to 49 years were at an increased risk of late-stage melanoma diagnosis compared to men 65 years and older.17

 

 

Demographic and Clinical Characteristics of Melanoma in Hispanic Patients

Among Hispanics, white Hispanics comprise the majority of melanoma cases.5 Median age at diagnosis is younger in Hispanics compared to whites.5,6 Hispanic men typically are older (median age, 61 years) than Hispanic women (median age, 52 years) at diagnosis.5 Similar to what is seen in NHWs, young Hispanic women experience a higher melanoma incidence than young Hispanic men.5 Among older Hispanics, melanoma is more common in men.5,8

Melanomas located on the lower extremities and hips are more prevalent in Hispanics than in NHWs.5,8,18 Among Hispanics, there are age- and sex-based variations in the anatomic location of primary tumors: in Hispanic men, truncal tumors predominate, and in Hispanic women, tumors of the lower extremities are most common across all age groups.5 The incidence of melanomas located in the head and neck region increases with age for both Hispanic men and women.

For melanomas in which the histologic type is known, superficial spreading melanoma is the most common subtype among Hispanics.5,17,19 Acral lentiginous melanomas and nodular melanomas are more common among Hispanics than among NHWs.5,17,19

The observation that Hispanics with melanoma are more prone to lower-extremity tumors and nodular and acral lentiginous melanoma subtypes than NHWs suggests that UV exposure history may be of less importance in this population. Although numerous studies have explored melanoma risk factors in NHWs, there is a striking paucity of such studies in Hispanics. For example, there are conflicting data regarding the role of UV exposure in melanoma risk among Hispanics. Hu et al20 found that UV index and latitude correlated with melanoma risk in this population, whereas Eide et al21 found no association between UV exposure and melanoma incidence in Hispanics. A prospective study involving a multiethnic cohort (of whom 40 of the 107 participants were Hispanic) found no clear association between a history of sunburn and melanoma risk in Hispanics.18

Strategies for Reducing Disparities in Outcomes

Our knowledge of melanoma epidemiology in Hispanics derives mainly from secondary analyses of state-level and national cancer registry data sets.5-8,13-15,17,19,20 These administrative data sources often are limited by missing data (eg, tumor thickness, histologic subtype) or lack important patient-level information (eg, self-identified race and ethnicity, health insurance status). Additionally, the manner in which data are collected and integrated into research varies; for example, socioeconomic measures often are reported as either area-based or composite measures. Thus, there is a need to improve the consistency of reporting on demographic and socioeconomic measures across studies. Polite et al22 recommended standardization of reporting criteria and that a standard set of demographic and socioeconomic status measures be included in clinical registries and research protocols.22 Researchers should strive to collect self-reported information on race and ethnicity, as well as the most granular level of detail on health insurance status, ancestry, and immigration status.

The host phenotypic characteristics of melanoma in NHWs are well understood, but the biological and environmental determinants of melanoma risk in Hispanics and other minorities are unknown. For example, fair complexion, red hair, blue eyes, increased freckling density, and the presence of numerous dysplastic and common melanocytic nevi indicate a propensity toward cutaneous melanoma.23,24 However, the relevance of such risk factors in Hispanics is unknown and has not been widely investigated in this patient population. Park et al18 found that a person’s sunburn susceptibility phenotype (defined as hair and eye color, ability to tan, and skin reaction to sunlight) was associated with an increased risk of melanoma among nonwhite, multiracial individuals. However, this study was limited by a small number of minority cases, which included only 40 Hispanic participants with melanoma.18 There is a need for rigorous observational studies to clearly define the phenotypic characteristics, sun-exposure behavior patterns, and genetic contributors to melanoma genesis in Hispanics.

The biologic determinants of postdiagnosis survival in Hispanics with melanoma are not well understood. It is unknown if genetic predisposition modifies melanoma risk in Hispanics. For example, the frequency of BRAF gene mutation or other driver mutations in US Hispanics has been understudied. It is important to know if mutation frequency patterns differ in Hispanics patients compared to NHWs because this knowledge could have considerable implications for treatment. Several recommendations should be considered to address these knowledge gaps. First, there is a need for development or enhancement of melanoma biorepositories, which should include tumor and nontumor specimens from a diverse sample of melanoma patients. Additionally, multi-institutional and multidisciplinary consortiums need to be created in order to amass a number of Hispanic melanoma patients to identify genetic, biologic, and behavioral risk factors specific to this subgroup of patients. The AMBER Consortium, which focuses on breast cancer epidemiology and risk in black women, is a model for the type of consortium needed for the study of melanoma in Hispanics.25 Lastly, community engagement will be central to developing sustainable recruitment and data-collection efforts.26 Involvement of key stakeholders will provide an in-depth assessment of community needs as well as real-time feedback on the process and practicality of research questions. Buy-in from affected communities also may facilitate dissemination of research findings to affected communities.

Conclusion

Hispanics are more likely to present with an advanced stages of disease and have higher melanoma-specific mortality rates than NHWs. Regrettably, a huge knowledge gap exists regarding contributors and solutions to melanoma disparities among this fast-growing, understudied segment of the US population. Accordingly, critical research is needed to address the most pressing questions regarding melanoma risk and poor outcomes among Hispanics to foster implementation of interventional efforts in prevention, early detection, and treatment. A multi-institutional and multidisciplinary approach across multiple levels is needed to eliminate disparate outcomes. Although melanoma is relatively uncommon among Hispanics, studies of melanoma in Hispanics (given their diverse genetic ancestry and migration) provide a unique backdrop against which researchers can explicate melanoma etiology—thus benefiting Hispanics and non-Hispanics alike.

References
  1. American Cancer Society. Key statistics for melanoma skin cancer. www.cancer.org/cancer/melanoma-skin-cancer/about/key-statistics.html. Accessed January 13, 2018.
  2. Balch CM, Gershenwald JE, Soong S, et al. Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol. 2009;27:6199-6206.
  3. Katalinic A, Waldmann A, Weinstock MA, et al. Does skin cancer screening save lives? Cancer. 2012;118:5395-5402.
  4. Bergad LW, Klein HS. Hispanics in the United States: A Demographic, Social, and Economic History, 1980-2005. New York, NY: Cambridge University Press; 2010.
  5. Garnett E, Townsend J, Steele B, et al. Characteristics, rates, and trends of melanoma incidence among Hispanics in the USA. Cancer Causes Control. 2016;27:647-659.
  6. Pollitt RA, Clarke CA, Swetter SM, et al. The expanding melanoma burden in California Hispanics: importance of socioeconomic distribution, histologic subtype, and anatomic location. Cancer. 2011;117:152-161.
  7. Hu S, Parmet, Y, Allen G, et al. Disparity in melanoma: a trend analysis of melanoma incidence and stage at diagnosis among whites,Hispanics, and blacks in Florida. JAMA Dermatology. 2010;145:1369-1374.
  8. Cormier JN, Xing Y, Ding M, et al. Ethnic differences among patients with cutaneous melanoma. Arch Intern Med. 2006;166:1907-1914.
  9. Pollitt RA, Swetter SM, Johnson TM, et al. Examining the pathways linking lower socioeconomic status and advanced melanoma. Cancer. 2012;118:4004-4013.
  10. Ortiz CA, Goodwin JS, Freeman JL. The effect of socioeconomic factors on incidence, stage at diagnosis and survival of cutaneous melanoma. Med Sci Monit. 2005;11:RA163-RA172.
  11. Singh SD, Ajani UA, Johnson CJ, et al. Association of cutaneous melanoma incidence with area-based socioeconomic indicators-United States, 2004-2006. J Am Acad Dermatol. 2011;65(5 suppl 1):S58-S68.
  12. Pollitt RA, Clarke CA, Shema SJ, et al. California Medicaid enrollment and melanoma stage at diagnosis: a population-based study. Am J Prev Med. 2008;35:7-13.
  13. Clairwood M, Ricketts J, Grant-Kels J, et al. Melanoma in skin of color in Connecticut: an analysis of melanoma incidence and stage at diagnosis in non-Hispanic blacks, non-Hispanic whites, and Hispanics. Int J Dermatol. 2014;53:425-433.
  14. Hamilton EC, Nguyen HT, Chang YC, et al. Health disparities influence childhood melanoma stage at diagnosis and outcome. J Pediatr. 2016;175:182-187.
  15. Dawes SM, Tsai S, Gittleman H, et al. Racial disparities in melanoma survival. J Am Acad Dermatol. 2016;75:983-991.
  16. Imahiyerobo-Ip J, Ip I, Jamal S, et al. Skin cancer awareness in communities of color. J Am Acad Dermatol. 2011;64:198-200.
  17. Harvey VM, Enos CW, Chen JT, et al. The role of neighborhood characteristics in late stage melanoma diagnosis among Hispanic men in California, Texas, and Florida, 1996-2012 [published online June 18, 2017]. J Cancer Epidemiol. 2017;2017:8418904.
  18. Park SL, Le Marchand L, Wilkens LR, et al. Risk factors for malignant melanoma in white and non-white/non-African American populations: the multiethnic cohort. Cancer Prev Res. 2012;5:423-434.
  19. Wu XC, Eide MJ, King J, et al. Racial and ethnic variations in incidence and survival of cutaneous melanoma in the United States, 1999-2006. J Am Acad Dermatol. 2011;65(5 suppl 1):S26-S37.
  20. Hu S, Ma F, Collado-Mesa F, et al. UV radiation, latitude, and melanoma in US Hispanics and blacks. Arch Dermatol. 2004;140:819-824.
  21. Eide MJ, Weinstock MA. Association of UV index, latitude, and melanoma incidence in nonwhite populations—US Surveillance, Epidemiology, and End Results (SEER) program, 1992 to 2001. Arch Dermatol. 2005;141:477-481.
  22. Polite BN, Adams-Campbell LL, Brawley OW, et al. Charting the future of cancer health disparities research: a position statement from the American Association for Cancer Research, the American Cancer Society, the American Society of Clinical Oncology, and the National Cancer Institute. Cancer Res. 2017;77:4548-4555.
  23. Gandini S, Sera F, Cattaruzza MS, et al. Meta-analysis of risk factors for cutaneous melanoma: III. family history, actinic damage and phenotypic factors. Eur J Cancer. 2005;41:2040-2059.
  24. Chang YM, Newton-Bishop JA, Bishop DT, et al. A pooled analysis of melanocytic nevus phenotype and the risk of cutaneous melanoma at different latitudes. Int J Cancer. 2009;124:420-428.
  25. Palmer JR, Ambrosone CB, Olshan AF. A collaborative study of the etiology of breast cancer subtypes in African American women: the AMBER consortium. Cancer Causes Control. 2014;25:309-319.
  26. Rapkin BD, Weiss E, Lounsbury D, et al. Reducing disparities in cancer screening and prevention through community-based participatory research partnerships with local libraries: a comprehensive dynamic trial. Am J Community Psychol. 2017;60:145-159.
References
  1. American Cancer Society. Key statistics for melanoma skin cancer. www.cancer.org/cancer/melanoma-skin-cancer/about/key-statistics.html. Accessed January 13, 2018.
  2. Balch CM, Gershenwald JE, Soong S, et al. Final version of 2009 AJCC melanoma staging and classification. J Clin Oncol. 2009;27:6199-6206.
  3. Katalinic A, Waldmann A, Weinstock MA, et al. Does skin cancer screening save lives? Cancer. 2012;118:5395-5402.
  4. Bergad LW, Klein HS. Hispanics in the United States: A Demographic, Social, and Economic History, 1980-2005. New York, NY: Cambridge University Press; 2010.
  5. Garnett E, Townsend J, Steele B, et al. Characteristics, rates, and trends of melanoma incidence among Hispanics in the USA. Cancer Causes Control. 2016;27:647-659.
  6. Pollitt RA, Clarke CA, Swetter SM, et al. The expanding melanoma burden in California Hispanics: importance of socioeconomic distribution, histologic subtype, and anatomic location. Cancer. 2011;117:152-161.
  7. Hu S, Parmet, Y, Allen G, et al. Disparity in melanoma: a trend analysis of melanoma incidence and stage at diagnosis among whites,Hispanics, and blacks in Florida. JAMA Dermatology. 2010;145:1369-1374.
  8. Cormier JN, Xing Y, Ding M, et al. Ethnic differences among patients with cutaneous melanoma. Arch Intern Med. 2006;166:1907-1914.
  9. Pollitt RA, Swetter SM, Johnson TM, et al. Examining the pathways linking lower socioeconomic status and advanced melanoma. Cancer. 2012;118:4004-4013.
  10. Ortiz CA, Goodwin JS, Freeman JL. The effect of socioeconomic factors on incidence, stage at diagnosis and survival of cutaneous melanoma. Med Sci Monit. 2005;11:RA163-RA172.
  11. Singh SD, Ajani UA, Johnson CJ, et al. Association of cutaneous melanoma incidence with area-based socioeconomic indicators-United States, 2004-2006. J Am Acad Dermatol. 2011;65(5 suppl 1):S58-S68.
  12. Pollitt RA, Clarke CA, Shema SJ, et al. California Medicaid enrollment and melanoma stage at diagnosis: a population-based study. Am J Prev Med. 2008;35:7-13.
  13. Clairwood M, Ricketts J, Grant-Kels J, et al. Melanoma in skin of color in Connecticut: an analysis of melanoma incidence and stage at diagnosis in non-Hispanic blacks, non-Hispanic whites, and Hispanics. Int J Dermatol. 2014;53:425-433.
  14. Hamilton EC, Nguyen HT, Chang YC, et al. Health disparities influence childhood melanoma stage at diagnosis and outcome. J Pediatr. 2016;175:182-187.
  15. Dawes SM, Tsai S, Gittleman H, et al. Racial disparities in melanoma survival. J Am Acad Dermatol. 2016;75:983-991.
  16. Imahiyerobo-Ip J, Ip I, Jamal S, et al. Skin cancer awareness in communities of color. J Am Acad Dermatol. 2011;64:198-200.
  17. Harvey VM, Enos CW, Chen JT, et al. The role of neighborhood characteristics in late stage melanoma diagnosis among Hispanic men in California, Texas, and Florida, 1996-2012 [published online June 18, 2017]. J Cancer Epidemiol. 2017;2017:8418904.
  18. Park SL, Le Marchand L, Wilkens LR, et al. Risk factors for malignant melanoma in white and non-white/non-African American populations: the multiethnic cohort. Cancer Prev Res. 2012;5:423-434.
  19. Wu XC, Eide MJ, King J, et al. Racial and ethnic variations in incidence and survival of cutaneous melanoma in the United States, 1999-2006. J Am Acad Dermatol. 2011;65(5 suppl 1):S26-S37.
  20. Hu S, Ma F, Collado-Mesa F, et al. UV radiation, latitude, and melanoma in US Hispanics and blacks. Arch Dermatol. 2004;140:819-824.
  21. Eide MJ, Weinstock MA. Association of UV index, latitude, and melanoma incidence in nonwhite populations—US Surveillance, Epidemiology, and End Results (SEER) program, 1992 to 2001. Arch Dermatol. 2005;141:477-481.
  22. Polite BN, Adams-Campbell LL, Brawley OW, et al. Charting the future of cancer health disparities research: a position statement from the American Association for Cancer Research, the American Cancer Society, the American Society of Clinical Oncology, and the National Cancer Institute. Cancer Res. 2017;77:4548-4555.
  23. Gandini S, Sera F, Cattaruzza MS, et al. Meta-analysis of risk factors for cutaneous melanoma: III. family history, actinic damage and phenotypic factors. Eur J Cancer. 2005;41:2040-2059.
  24. Chang YM, Newton-Bishop JA, Bishop DT, et al. A pooled analysis of melanocytic nevus phenotype and the risk of cutaneous melanoma at different latitudes. Int J Cancer. 2009;124:420-428.
  25. Palmer JR, Ambrosone CB, Olshan AF. A collaborative study of the etiology of breast cancer subtypes in African American women: the AMBER consortium. Cancer Causes Control. 2014;25:309-319.
  26. Rapkin BD, Weiss E, Lounsbury D, et al. Reducing disparities in cancer screening and prevention through community-based participatory research partnerships with local libraries: a comprehensive dynamic trial. Am J Community Psychol. 2017;60:145-159.
Issue
Cutis - 101(4)
Issue
Cutis - 101(4)
Page Number
243-246
Page Number
243-246
Publications
Publications
Topics
Article Type
Display Headline
Melanoma in US Hispanics: Recommended Strategies to Reduce Disparities in Outcomes
Display Headline
Melanoma in US Hispanics: Recommended Strategies to Reduce Disparities in Outcomes
Sections
Inside the Article

Practice Points

  • Although the age-adjusted incidence of melanoma among US Hispanics is lower than among non-Hispanic whites, Hispanics with melanoma are more likely to present with stage III disease and have distant metastases.
  • Late presentation of melanoma in Hispanics is not completely understood but may be attributed to socioeconomic factors, lack of skin cancer awareness and knowledge, lower rate of self- and physician-performed skin examinations, and differences in tumor biology, among other variables.
  • Research is needed to address gaps in knowledge about the risk of melanoma and comparatively poor outcomes among Hispanics so interventional efforts for prevention, early detection, and treatment can be implemented.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

No Sulfates, No Parabens, and the “No-Poo” Method: A New Patient Perspective on Common Shampoo Ingredients

Article Type
Changed
Thu, 10/29/2020 - 14:53
Display Headline
No Sulfates, No Parabens, and the “No-Poo” Method: A New Patient Perspective on Common Shampoo Ingredients
In Collaboration with the Skin of Color Society

Shampoo is a staple in hair grooming that is ever-evolving along with cultural trends. The global shampoo market is expected to reach an estimated value of $25.73 billion by 2019. A major driver of this upward trend in market growth is the increasing demand for natural and organic hair shampoos.1 Society today has a growing fixation on healthy living practices, and as of late, the ingredients in shampoos and other cosmetic products have become one of the latest targets in the health-consciousness craze. In the age of the Internet where information—and misinformation—is widely accessible and dispersed, the general public often strives to self-educate on specialized matters that are out of their expertise. As a result, individuals have developed an aversion to using certain shampoos out of fear that the ingredients, often referred to as “chemicals” by patients due to their complex names, are unnatural and therefore unhealthy.1,2 Product developers are working to meet the demand by reformulating shampoos with labels that indicate sulfate free or paraben free, despite the lack of proof that these formulations are an improvement over traditional approaches to hair health. Additionally, alternative methods of cleansing the hair and scalp, also known as the no-shampoo or “no-poo” method, have begun to gain popularity.2,3


It is essential that dermatologists acknowledge the concerns that their patients have about common shampoo ingredients to dispel the myths that may misinform patient decision-making. This article reviews the controversy surrounding the use of sulfates and parabens in shampoos as well as commonly used shampoo alternatives. Due to the increased prevalence of dry hair shafts in the skin of color population, especially black women, this group is particularly interested in products that will minimize breakage and dryness of the hair. To that end, this population has great interest in the removal of chemical ingredients that may cause damage to the hair shafts, despite the lack of data to support sulfates and paraben damage to hair shafts or scalp skin. Blogs and uninformed hairstylists may propagate these beliefs in a group of consumers who are desperate for new approaches to hair fragility and breakage.

Surfactants and Sulfates

The cleansing ability of a shampoo depends on the surface activity of its detergents. Surface-active ingredients, or surfactants, reduce the surface tension between water and dirt, thus facilitating the removal of environmental dirt from the hair and scalp,4 which is achieved by a molecular structure containing both a hydrophilic and a lipophilic group. Sebum and dirt are bound by the lipophilic ends of the surfactant, becoming the center of a micelle structure with the hydrophilic molecule ends pointing outward. Dirt particles become water soluble and are removed from the scalp and hair shaft upon rinsing with water.4

Surfactants are classified according to the electric charge of the hydrophilic polar group as either anionic, cationic, amphoteric (zwitterionic), or nonionic.5 Each possesses different hair conditioning and cleansing qualities, and multiple surfactants are used in shampoos in differing ratios to accommodate different hair types. In most shampoos, the base consists of anionic and amphoteric surfactants. Depending on individual product requirements, nonionic and cationic surfactants are used to either modify the effects of the surfactants or as conditioning agents.4,5

One subcategory of surfactants that receives much attention is the group of anionic surfactants known as sulfates. Sulfates, particularly sodium lauryl sulfate (SLS), recently have developed a negative reputation as cosmetic ingredients, as reports from various unscientific sources have labeled them as hazardous to one’s health; SLS has been described as a skin and scalp irritant, has been linked to cataract formation, and has even been wrongly labeled as carcinogenic.6 The origins of some of these claims are not clear, though they likely arose from the misinterpretation of complex scientific studies that are easily accessible to laypeople. The link between SLS and ocular irritation or cataract formation is a good illustration of this unsubstantiated fear. A study by Green et al7 showed that corneal exposure to extremely high concentrations of SLS following physical or chemical damage to the eye can result in a slowed healing process. The results of this study have since been wrongly quoted to state that SLS-containing products lead to blindness or severe corneal damage.8 A different study tested for possible ocular irritation in vivo by submerging the lens of an eye into a 20% SLS solution, which accurately approximates the concentration of SLS in rinse-off consumer products.9 However, to achieve ocular irritation, the eyes of laboratory animals were exposed to SLS constantly for 14 days, which would not occur in practical use.9 Similarly, a third study achieved cataract formation in a laboratory only by immersing the lens of an eye into a highly concentrated solution of SLS.10 Such studies are not appropriate representations of how SLS-containing products are used by consumers and have unfortunately been vulnerable to misinterpretation by the general public.

There is no known study that has shown SLS to be carcinogenic. One possible origin of this idea may be from the wrongful interpretation of studies that used SLS as a vehicle substance to test agents that were deemed to be carcinogenic.11 Another possible source of the idea that SLS is carcinogenic comes from its association with 1,4-dioxane, a by-product of the synthesis of certain sulfates such as sodium laureth sulfate due to a process known as ethoxylation.6,12 Although SLS does not undergo this process in its formation and is not linked to 1,4-dioxane, there is potential for cross-contamination of SLS with 1,4-dioxane, which cannot be overlooked. 1,4-Dioxane is classified as “possibly carcinogenic to humans (Group 2B)” by the International Agency for Research on Cancer,13 but screening of SLS for this substance prior to its use in commercial products is standard.

Sulfates are inexpensive detergents that are responsible for lather formation in shampoos as well as in many household cleaning agents.5 Sulfates, similar to all anionic surfactants, are characterized by a negatively charged hydrophilic polar group. The best-known and most commonly used anionic surfactants are sulfated fatty alcohols, alkyl sulfates, and their polyethoxylated analogues alkyl ether sulfates.5,6 Sodium lauryl sulfate (also known as sodium laurilsulfate or sodium dodecyl sulfate) is the most common of them all, found in shampoo and conditioner formulations. Ammonium lauryl sulfate and sodium laureth sulfate are other sulfates commonly used in shampoos and household cleansing products. Sodium lauryl sulfate is a nonvolatile, water-soluble compound. Its partition coefficient (P0), a measure of a substance’s hydrophilic or lipophilic nature, is low at 1.6, making it a rather hydrophilic substance.6 Hydrophilic substances tend to have low bioaccumulation profiles in the body. Additionally, SLS is readily biodegradable. It can be derived from both synthetic and naturally occurring sources; for example, palm kernel oil, petrolatum, and coconut oil are all sources of lauric acid, the starting ingredient used to synthesize SLS. Sodium lauryl sulfate is created by reacting lauryl alcohol with sulfur trioxide gas, followed by neutralization with sodium carbonate (also a naturally occurring compound).6 Sodium lauryl sulfate and other sulfate-containing shampoos widely replaced the usage of traditional soaps formulated from animal or vegetable fats, as these latter formations created a film of insoluble calcium salts on the hair strands upon contact with water, resulting in tangled, dull-appearing hair.5 Additionally, sulfates were preferred to the alkaline pH of traditional soap, which can be harsh on hair strands and cause irritation of the skin and mucous membranes.14 Because they are highly water soluble, sulfates enable the formulation of clear shampoos. They exhibit remarkable cleaning properties and lather formation.5,14

Because sulfates are potent surfactants, they can remove dirt and debris as well as naturally produced healthy oils from the hair and scalp. As a result, sulfates can leave the hair feeling dry and stripped of moisture.4,5 Sulfates are used as the primary detergents in the formulation of deep-cleaning shampoos, which are designed for people who accumulate a heavy buildup of dirt, sebum, and debris from frequent use of styling products. Due to their potent detergency, these shampoos typically are not used on a daily basis but rather at longer intervals.15 A downside to sulfates is that they can have cosmetically unpleasant properties, which can be compensated for by including appropriate softening additives in shampoo formulations.4 A number of anionic surfactants such as olefin sulfonate, alkyl sulfosuccinate, acyl peptides, and alkyl ether carboxylates are well tolerated by the skin and are used together with other anionic and amphoteric surfactants to optimize shampoo properties. Alternatively, sulfate-free shampoos are cleansers compounded by the removal of the anionic group and switched for surfactants with less detergency.4,5

 

 

Preservatives and Parabens

Parabens refer to a group of esters of 4-hydroxybenzoic acid commonly used as preservatives in foods, pharmaceuticals, and cosmetics whose widespread use dates back to 1923.16 Concerns over the presence of parabens in shampoos and other cosmetics have been raised by patients for their reputed estrogenic and antiandrogenic effects and suspected involvement in carcinogenesis via endocrine modulation.16,17 In in vitro studies done on yeast assays, parabens have shown weak estrogenic activity that increases in proportion to both the length and increased branching of the alkyl side chains in the paraben’s molecular structure.18 They are 10,000-fold less potent than 17β-estradiol. In in vivo animal studies, parabens show weak estrogenic activity and are 100,000-fold less potent than 17β-estradiol.18 4-Hydroxybenzoic acid, a common metabolite, showed no estrogenic activity when tested both in vitro and in vivo.19 Some concerning research has implicated a link between parabens used in underarm cosmetics, such as deodorants and antiperspirants, and breast cancer16; however, the studies have been conflicting, and there is simply not enough data to assert that parabens cause breast cancer.

The Cosmetic Ingredient Review expert panel first reviewed parabens in 1984 and concluded that “methylparaben, ethylparaben, propylparaben, and butylparaben are safe as cosmetic ingredients in the present practices of use.”20 They extended this statement to include isopropylparaben and isobutylparaben in a later review.21 In 2005, the Scientific Committee on Consumer Products (now known as the Scientific Committee for Consumer Safety) in Europe stated that methylparaben and ethylparaben can be used at levels up to 0.4% in products.22 This decision was reached due to reports of decreased sperm counts and testosterone levels in male juvenile rats exposed to these parabens; however, these reults were not successfully replicated in larger studies.16,22 In 2010, the Scientific Committee for Consumer Safety revisited its stance on parabens, and they then revised their recommendations to say that concentrations of propylparaben and butylparaben should not exceed concentrations of 0.19%, based on “the conservative choice for the calculation of the [Margin-of-Safety] of butyl- and propylparaben.”23 However, in 2011 the use of propylparaben and butylparaben was banned in Denmark for cosmetic products used in children 3 years or younger,16 and the European Commission subsequently amended their directive in 2014, banning isopropylparaben, isobutylparaben, phenylparaben, benzylparaben, and pentylparaben due to lack of data available to evaluate the human risk of these products.24

Contrary to the trends in Europe, there currently are no regulations against the use of parabens in shampoos or other cosmetics in the United States. The American Cancer Society found that there is no evidence to suggest that the current levels of parabens in cosmetic products (eg, antiperspirants) increase one’s risk of breast cancer.25 Parabens are readily absorbed into the body both transdermally and through ingestion but also are believed to be rapidly transformed into harmless and nonspecific metabolites; they are readily metabolized by the liver and excreted in urine, and there is no measured accumulation in tissues.17

Parabens continue to be the most widely used preservatives in personal care products, usually in conjunction with other preservatives. Parabens are good biocides; short-chain esters (eg, methylparabens, ethylparabens) are effective against gram-positive bacteria and are weakly effective against gram-negative bacteria. Long-chain paraben esters (eg, propylparabens, butylparabens) are effective against mold and yeast. The addition of other preservatives creates a broad spectrum of antimicrobial defense in consumer products. Other preservatives include formaldehyde releasers or phenoxyethanol, as well as chelating agents such as EDTA, which improve the stability of these cosmetic products when exposed to air.16 Parabens are naturally occurring substances found in foods such as blueberries, barley, strawberries, yeast, olives, and grapes. As a colorless, odorless, and inexpensive substance, their use has been heavily favored in cosmetic and food products.16

 

 

Shampoo Alternatives and the No-Poo Method

Although research has not demonstrated any long-term danger to using shampoo, certain chemicals found in shampoos have the potential to irritate the scalp. Commonly cited allergens in shampoos include cocamidopropyl betaine, propylene glycol, vitamin E (tocopherol), parabens, and benzophenones.5 Additionally, the rising use of formaldehyde-releasing preservatives and isothiazolinones due to mounting pressures to move away from parabens has led to an increase in cases of allergic contact dermatitis (ACD).16 However, the irritability (rather than allergenicity) of these substances often is established during patch testing, a method of detecting delayed-type allergic reactions, which is important to note because patch testing requires a substance to be exposed to the skin for 24 to 48 hours, whereas exposure to shampoo ingredients may last a matter of minutes at most and occur in lesser concentrations because the ingredients are diluted by water in the rinsing process. Given these differences, it is unlikely that a patient would develop a true allergic response from regular shampoo use. Nevertheless, in patients who are already sensitized, exposure could conceivably trigger ACD, and patients must be cognizant of the composition of their shampoos.16

The no-poo method refers to the avoidance of commercial shampoo products when cleansing the hair and scalp and encompasses different methods of cleansing the hair, such as the use of household items (eg, baking soda, apple cider vinegar [ACV]), the use of conditioners to wash the hair (also known as conditioner-only washing or co-washing), treating the scalp with tea tree oil, or simply rinsing the hair with water. Proponents of the no-poo method believe that abstaining from shampoo use leads to healthier hair, retained natural oils, and less exposure to supposedly dangerous chemicals such as parabens or sulfates.2,3,26-28 However, there are no known studies in the literature that assess or support the hypotheses of the no-poo method.

Baking Soda and ACV
Baking soda (sodium bicarbonate) is a substance commonly found in the average household. It has been used in toothpaste formulas and cosmetic products and is known for its acid-neutralizing properties. Baking soda has been shown to have some antifungal and viricidal properties through an unknown mechanism of action.28 It has gained popularity for its use as a means of reducing the appearance of excessive greasiness of the hair shafts. Users also have reported that when washing their hair with baking soda, they are able to achieve a clean scalp and hair that feels soft to the touch.2,3,26,27,29 Despite these reports, users must beware of using baking soda without adequately diluting it with water. Baking soda is a known alkaline irritant.26,30 With a pH of 9, baking soda causes the cuticle layer of the hair fiber to open, increasing the capacity for water absorption. Water penetrates the scales that open, breaking the hydrogen bonds of the keratin molecule.31 Keratin is a spiral helical molecule that keeps its shape due to hydrogen, disulfide, and ionic bonds, as well as Van der Waals force.30 Hydrolysis of these bonds due to exposure to baking soda lowers the elasticity of the hair and increases the negative electrical net charge of the hair fiber surface, which leads to increased friction between fibers, cuticle damage, hair fragility, and fiber breakage.32,33

Apple cider vinegar is an apple-derived acetic acid solution with a pH ranging from 3.1 to 5.28 The pH range of ACV is considered to be ideal for hair by no-poo proponents, as it is similar to the natural pH of the scalp. Its acidic properties are responsible for its antimicrobial abilities, particularly its effectiveness against gram-negative bacteria.30 The acetic acid of ACV can partially interrupt oil interfaces, which contributes to its mild ability to remove product residue and scalp buildup from the hair shaft; the acetic acid also tightens the cuticles on hair fibers.33 Apple cider vinegar is used as a means of cleansing the hair and scalp by no-poo proponents2,3,26; other uses for ACV include using it as a rinse following washing and/or conditioning of the hair or as a means of preserving color in color-treated hair. There also is evidence that ACV may have antifungal properties.28 However, consumers must be aware that if it is not diluted in water, ACV may be too caustic for direct application to the hair and may lead to damage; it can be irritating to eyes, mucus membranes, and acutely inflamed skin. Also, vinegar rinses used on processed or chemically damaged hair may lead to increased hair fragility.2,3

Hair fibers have a pH of 3.67, while the scalp has a pH between 4.5 and 6.2. This slightly acidic film acts as a barrier to viruses, bacteria, and other potential contaminants.33 Studies have shown that the pH of skin increases in proportion to the pH of the cleanser used.34 Therefore, due to the naturally acidic pH of the scalp, acid-balanced shampoos generally are recommended. Shampoos should not have a pH higher than 5.5, as hair shafts can swell due to alkalinization, which can be prevented by pH balancing the shampoo through the addition of an acidic substance (eg, glycolic acid, citric acid) to lower the pH down to approximately 5.5. Apple cider vinegar often is used for this purpose. However, one study revealed that 82% of shampoos already have an acidic pH.34

Conditioner-Only Washing (Co-washing)
Conditioner-only washing, or co-washing, is a widely practiced method of hair grooming. It is popular among individuals who find that commercial shampoos strip too much of the natural hair oils away, leaving the hair rough or unmanageable. Co-washing is not harmful to the hair; however, the molecular structure and function of a conditioner and that of a shampoo are very different.5,35,36 Conditioners are not formulated to remove dirt and buildup in the hair but rather to add substances to the hair, and thus cannot provide extensive cleansing of the hair and scalp; therefore, it is inappropriate to use co-washing as a replacement for shampooing. Quaternary conditioning agents are an exception because they contain amphoteric detergents comprised of both anionic and cationic groups, which allow them both the ability to remove dirt and sebum with its anionic group, typically found in shampoos, as well as the ability to coat and condition the hair due to the high affinity of the cationic group for the negatively charged hair fibers.36,37 Amphoteric detergents are commonly found in 2-in-1 conditioning cleansers, among other ingredients, such as hydrolyzed animal proteins that temporarily plug surface defects on the hair fiber, and dimethicone, a synthetic oil that creates a thin film over the hair shaft, increasing shine and manageability. Of note, these conditioning shampoos are ideal for individuals with minimal product buildup on the hair and scalp and are not adequate scalp cleansers for individuals who either wash their hair infrequently or who regularly use hairstyling products.36,37

Tea Tree Oil
Tea tree oil is an essential oil extracted from the Melaleuca alternifolia plant of the Myrtaceae family. It is native to the coast of northeastern Australia. A holy grail of natural cosmetics, tea tree oil is widely known for its antiviral, antifungal, and antiseptic properties.38 Although not used as a stand-alone cleanser, it is often added to a number of cosmetic products, including shampoos and co-washes. Although deemed safe for topical use, it has been shown to be quite toxic when ingested. Symptoms of ingestion include nausea, vomiting, hallucinations, and coma. The common concern with tea tree oil is its ability to cause ACD. In particular, it is believed that the oxidation products of tea tree oil are allergenic rather than the tea tree oil itself. The evaluation of tea tree oil as a potential contact allergen has been quite difficult; it consists of more than 100 distinct compounds and is often mislabeled, or does not meet the guidelines of the International Organization for Standardization. Nonetheless, the prevalence of ACD due to tea tree oil is low (approximately 1.4%). Despite its low prevalence, tea tree oil should remain in the differential as an ACD-inducing agent. Patch testing with the patient’s supply of tea tree oil is advised when possible.38

Conclusion

It is customary that the ingredients used in shampoos undergo periodic testing and monitoring to assure the safety of their use. Although it is encouraging that patients are proactive in their efforts to stay abreast of the literature, it is still important that cosmetic scientists, dermatologists, and other experts remain at the forefront of educating the public about these substances. Not doing so can result in the propagation of misinformation and unnecessary fears, which can lead to the adaptation of unhygienic or even unsafe hair care practices. As dermatologists, we must ensure that patients are educated about the benefits and hazards of off-label use of household ingredients to the extent that evidence-based medicine permits. Patients must be informed that not all synthetic substances are harmful, and likewise not all naturally occurring substances are safe.

References
  1. The global shampoo market 2014-2019 trends, forecast, and opportunity analysis [press release]. New York, NY: Reportlinker; May 21, 2015.
  2. Is the ‘no shampoo’ trend healthy or harmful? Mercola website. Published January 16, 2016. Accessed December 8, 2017.
  3. Feltman R. The science (or lack thereof) behind the ‘no-poo’ hair trend. Washington Post. March 10, 2016. https://www.washingtonpost.com/news/speaking-of-science/wp/2016/03/10/the-science-or-lack-thereof-behind-the-no-poo-hair-trend/?utm_term=.9a61edf3fd5a. Accessed December 11, 2017.
  4. Bouillon C. Shampoos. Clin Dermatol. 1996;14:113-121.
  5. Trueb RM. Shampoos: ingredients, efficacy, and adverse effects. J Dtsch Dermatol Ges. 2007;5:356-365.
  6. Bondi CA, Marks JL, Wroblewski LB, et al. Human and environmental toxicity of sodium lauryl sulfate (SLS): evidence for safe use in household cleaning products. Environ Health Insights. 2015;9:27-32.
  7. Green K, Johnson RE, Chapman JM, et al. Preservative effects on the healing rate of rabbit corneal epithelium. Lens Eye Toxic Res. 1989;6:37-41.
  8. Sodium lauryl sulphate. Healthy Choices website. http://www.healthychoices.co.uk/sls.html. Accessed December 8, 2017.
  9. Tekbas¸ ÖF, Uysal Y, Og˘ur R, et al. Non-irritant baby shampoos may cause cataract development. TSK Koruyucu Hekimlik Bülteni. 2008;1:1-6.
  10. Cater KC, Harbell JW. Prediction of eye irritation potential of surfactant-based rinse-off personal care formulations by the bovine corneal opacity and permeability (BCOP) assay. Cutan Ocul Toxicol. 2006;25:217-233.
  11. Birt DF, Lawson TA, Julius AD, et al. Inhibition by dietary selenium of colon cancer induced in the rat by bis(2-oxopropyl) nitrosamine. Cancer Res. 1982;42:4455-4459.
  12. Rastogi SC. Headspace analysis of 1,4-dioxane in products containing polyethoxylated surfactants by GC-MS. Chromatographia. 1990;29:441-445.
  13. 1,4-Dioxane. IARC Monogr Eval Carcinog Risks Hum. 1999;71, pt 2:589-602.
  14. Trueb RM. Dermocosmetic aspects of hair and scalp. J Investig Dermatol Symp Proc. 2005;10:289-292.
  15. D’Souza P, Rathi SK. Shampoo and conditioners: what a dermatologist should know? Indian J Dermatol. 2015;60:248-254.
  16. Sasseville D, Alfalah M, Lacroix JP. “Parabenoia” debunked, or “who’s afraid of parabens?” Dermatitis. 2015;26:254-259.
  17. Krowka JF, Loretz L, Geis PA, et al. Preserving the facts on parabens: an overview of these important tools of the trade. Cosmetics & Toiletries. http://www.cosmeticsandtoiletries.com/research/chemistry/Preserving-the-Facts-on-Parabens-An-Overview-of-These-Important-Tools-of-the Trade-425784294.html. Published June 1, 2017. Accessed December 20, 2017.
  18. Routledge EJ, Parker J, Odum J, et al. Some alkyl hydroxy benzoate preservatives (parabens) are estrogenic. Toxicol Appl Pharmacol. 1998;153:12Y19.
  19. Hossaini A, Larsen JJ, Larsen JC. Lack of oestrogenic effects of food preservatives (parabens) in uterotrophic assays. Food Chem Toxicol. 2000;38:319-323.
  20. Cosmetic Ingredient Review. Final report on the safety assessment of methylparaben, ethylparaben, propylparaben and butylparaben. J Am Coll Toxicol. 1984;3:147-209.
  21. Cosmetic Ingredient Review. Final report on the safety assessment of isobutylparaben and isopropylparaben. J Am Coll Toxicol. 1995;14:364-372.
  22. Scientific Committee on Consumer Products. Extended Opinion on the Safety Evaluation of Parabens. European Commission website. https://ec.europa.eu/health/ph_risk/committees/04_sccp/docs/sccp_o_019.pdf. Published January 28, 2005. Accessed December 20, 2017.
  23. Scientific Committee on Consumer Products. Opinion on Parabens. European Commission website. http://ec.europa.eu/health/scientific_committees/consumer_safety/docs/sccs_o_041.pdf. Revised March 22, 2011. Accessed December 20, 2017.
  24. European Commission. Commission Regulation (EU) No 258/2014 of 9 April 2014 amending Annexes II and V to Regulation (EC) No 1223/2009 of the European Parliament and of the Council on cosmetic products. EUR-Lex website. http://eur-lex.europa.eu/legal-content/EN/TXT/?uri=uriserv:OJ.L_.2014.107.01.0005.01.ENG. Accessed December 20, 2017.
  25. American Cancer Society. Antiperspirants and breast cancer risk. https://www.cancer.org/cancer/cancer-causes/antiperspirants-and-breast-cancer-risk.html#references. Revised October 14, 2014. Accessed January 2, 2018.
  26. MacMillan A. Cutting back on shampoo? 15 things you should know. Health. February 25, 2014. http://www.health.com/health/gallery/0,,20788089,00.html#should-you-go-no-poo--1. Accessed December 10, 2017.
  27. The ‘no poo’ method. https://www.nopoomethod.com/. Accessed December 10, 2017.
  28. Fong, D, Gaulin C, Le M, et al. Effectiveness of alternative antimicrobial agents for disinfection of hard surfaces. National Collaborating Centre for Environmental Health website. http://www.ncceh.ca/sites/default/files/Alternative_Antimicrobial_Agents_Aug_2014.pdf. Published August 2014. Accessed December 10, 2017.
  29. Is baking soda too harsh for natural hair? Black Girl With Long Hair website. http://blackgirllonghair.com/2012/02/is-baking-soda-too-harsh-for-hair/2/. Published February 5, 2012. Accessed December 12, 2017.
  30. O’Lenick T. Anionic/cationic complexes in hair care. J Cosmet Sci. 2011;62:209-228.
  31. Gavazzoni Dias MF, de Almeida AM, Cecato PM, et al. The shampoo pH can affect the hair: myth or reality? Int J Trichology. 2014;6:95-99.
  32. Goodman H. The acid mantle of the skin surface. Ind Med Surg. 1958;27:105-108.
  33. Korting HC, Kober M, Mueller M, et al. Influence of repeated washings with soap and synthetic detergents on pH and resident flora of the skin of forehead and forearm. results of a cross-over trial in health probationers. Acta Derm Venereol. 1987;67:41-47.
  34. Tarun J, Susan J, Suria J, et al. Evaluation of pH of bathing soaps and shampoos for skin and hair care. Indian J Dermatol. 2014;59:442-444.
  35. Corbett JF. The chemistry of hair-care products. J Soc Dyers Colour. 1976;92:285-303.
  36. McMichael AJ, Hordinsky M. Hair Diseases: Medical, Surgical, and Cosmetic Treatments. New York, NY: Taylor & Francis; 2008:59-72.
  37. Allardice A, Gummo G. Hair conditioning: quaternary ammonium compounds on various hair types. Cosmet Toiletries. 1993;108:107-109.
  38. Larson D, Jacob SE. Tea tree oil. Dermatitis. 2012;23:48-49.
Article PDF
Author and Disclosure Information

 

Dr. Cline is from Augusta University Medical Center, Georgia. Drs. Uwakwe and McMichael are from the Department of Dermatology, Wake Forest School of Medicine, Winston-Salem, North Carolina.

Drs. Cline and Uwakwe report no conflict of interest. Dr. McMichael is a consultant for Johnson & Johnson and Procter & Gamble. She also is an investigator for and has received research grants from Procter & Gamble.

Correspondence: Amy J. McMichael, MD, Department of Dermatology, Wake Forest Baptist Medical Center, 4618 Country Club Rd, Winston-Salem, NC 27104 (amcmicha@wakehealth.edu).

Issue
Cutis - 101(1)
Publications
Topics
Page Number
22-26
Sections
Author and Disclosure Information

 

Dr. Cline is from Augusta University Medical Center, Georgia. Drs. Uwakwe and McMichael are from the Department of Dermatology, Wake Forest School of Medicine, Winston-Salem, North Carolina.

Drs. Cline and Uwakwe report no conflict of interest. Dr. McMichael is a consultant for Johnson & Johnson and Procter & Gamble. She also is an investigator for and has received research grants from Procter & Gamble.

Correspondence: Amy J. McMichael, MD, Department of Dermatology, Wake Forest Baptist Medical Center, 4618 Country Club Rd, Winston-Salem, NC 27104 (amcmicha@wakehealth.edu).

Author and Disclosure Information

 

Dr. Cline is from Augusta University Medical Center, Georgia. Drs. Uwakwe and McMichael are from the Department of Dermatology, Wake Forest School of Medicine, Winston-Salem, North Carolina.

Drs. Cline and Uwakwe report no conflict of interest. Dr. McMichael is a consultant for Johnson & Johnson and Procter & Gamble. She also is an investigator for and has received research grants from Procter & Gamble.

Correspondence: Amy J. McMichael, MD, Department of Dermatology, Wake Forest Baptist Medical Center, 4618 Country Club Rd, Winston-Salem, NC 27104 (amcmicha@wakehealth.edu).

Article PDF
Article PDF
Related Articles
In Collaboration with the Skin of Color Society
In Collaboration with the Skin of Color Society

Shampoo is a staple in hair grooming that is ever-evolving along with cultural trends. The global shampoo market is expected to reach an estimated value of $25.73 billion by 2019. A major driver of this upward trend in market growth is the increasing demand for natural and organic hair shampoos.1 Society today has a growing fixation on healthy living practices, and as of late, the ingredients in shampoos and other cosmetic products have become one of the latest targets in the health-consciousness craze. In the age of the Internet where information—and misinformation—is widely accessible and dispersed, the general public often strives to self-educate on specialized matters that are out of their expertise. As a result, individuals have developed an aversion to using certain shampoos out of fear that the ingredients, often referred to as “chemicals” by patients due to their complex names, are unnatural and therefore unhealthy.1,2 Product developers are working to meet the demand by reformulating shampoos with labels that indicate sulfate free or paraben free, despite the lack of proof that these formulations are an improvement over traditional approaches to hair health. Additionally, alternative methods of cleansing the hair and scalp, also known as the no-shampoo or “no-poo” method, have begun to gain popularity.2,3


It is essential that dermatologists acknowledge the concerns that their patients have about common shampoo ingredients to dispel the myths that may misinform patient decision-making. This article reviews the controversy surrounding the use of sulfates and parabens in shampoos as well as commonly used shampoo alternatives. Due to the increased prevalence of dry hair shafts in the skin of color population, especially black women, this group is particularly interested in products that will minimize breakage and dryness of the hair. To that end, this population has great interest in the removal of chemical ingredients that may cause damage to the hair shafts, despite the lack of data to support sulfates and paraben damage to hair shafts or scalp skin. Blogs and uninformed hairstylists may propagate these beliefs in a group of consumers who are desperate for new approaches to hair fragility and breakage.

Surfactants and Sulfates

The cleansing ability of a shampoo depends on the surface activity of its detergents. Surface-active ingredients, or surfactants, reduce the surface tension between water and dirt, thus facilitating the removal of environmental dirt from the hair and scalp,4 which is achieved by a molecular structure containing both a hydrophilic and a lipophilic group. Sebum and dirt are bound by the lipophilic ends of the surfactant, becoming the center of a micelle structure with the hydrophilic molecule ends pointing outward. Dirt particles become water soluble and are removed from the scalp and hair shaft upon rinsing with water.4

Surfactants are classified according to the electric charge of the hydrophilic polar group as either anionic, cationic, amphoteric (zwitterionic), or nonionic.5 Each possesses different hair conditioning and cleansing qualities, and multiple surfactants are used in shampoos in differing ratios to accommodate different hair types. In most shampoos, the base consists of anionic and amphoteric surfactants. Depending on individual product requirements, nonionic and cationic surfactants are used to either modify the effects of the surfactants or as conditioning agents.4,5

One subcategory of surfactants that receives much attention is the group of anionic surfactants known as sulfates. Sulfates, particularly sodium lauryl sulfate (SLS), recently have developed a negative reputation as cosmetic ingredients, as reports from various unscientific sources have labeled them as hazardous to one’s health; SLS has been described as a skin and scalp irritant, has been linked to cataract formation, and has even been wrongly labeled as carcinogenic.6 The origins of some of these claims are not clear, though they likely arose from the misinterpretation of complex scientific studies that are easily accessible to laypeople. The link between SLS and ocular irritation or cataract formation is a good illustration of this unsubstantiated fear. A study by Green et al7 showed that corneal exposure to extremely high concentrations of SLS following physical or chemical damage to the eye can result in a slowed healing process. The results of this study have since been wrongly quoted to state that SLS-containing products lead to blindness or severe corneal damage.8 A different study tested for possible ocular irritation in vivo by submerging the lens of an eye into a 20% SLS solution, which accurately approximates the concentration of SLS in rinse-off consumer products.9 However, to achieve ocular irritation, the eyes of laboratory animals were exposed to SLS constantly for 14 days, which would not occur in practical use.9 Similarly, a third study achieved cataract formation in a laboratory only by immersing the lens of an eye into a highly concentrated solution of SLS.10 Such studies are not appropriate representations of how SLS-containing products are used by consumers and have unfortunately been vulnerable to misinterpretation by the general public.

There is no known study that has shown SLS to be carcinogenic. One possible origin of this idea may be from the wrongful interpretation of studies that used SLS as a vehicle substance to test agents that were deemed to be carcinogenic.11 Another possible source of the idea that SLS is carcinogenic comes from its association with 1,4-dioxane, a by-product of the synthesis of certain sulfates such as sodium laureth sulfate due to a process known as ethoxylation.6,12 Although SLS does not undergo this process in its formation and is not linked to 1,4-dioxane, there is potential for cross-contamination of SLS with 1,4-dioxane, which cannot be overlooked. 1,4-Dioxane is classified as “possibly carcinogenic to humans (Group 2B)” by the International Agency for Research on Cancer,13 but screening of SLS for this substance prior to its use in commercial products is standard.

Sulfates are inexpensive detergents that are responsible for lather formation in shampoos as well as in many household cleaning agents.5 Sulfates, similar to all anionic surfactants, are characterized by a negatively charged hydrophilic polar group. The best-known and most commonly used anionic surfactants are sulfated fatty alcohols, alkyl sulfates, and their polyethoxylated analogues alkyl ether sulfates.5,6 Sodium lauryl sulfate (also known as sodium laurilsulfate or sodium dodecyl sulfate) is the most common of them all, found in shampoo and conditioner formulations. Ammonium lauryl sulfate and sodium laureth sulfate are other sulfates commonly used in shampoos and household cleansing products. Sodium lauryl sulfate is a nonvolatile, water-soluble compound. Its partition coefficient (P0), a measure of a substance’s hydrophilic or lipophilic nature, is low at 1.6, making it a rather hydrophilic substance.6 Hydrophilic substances tend to have low bioaccumulation profiles in the body. Additionally, SLS is readily biodegradable. It can be derived from both synthetic and naturally occurring sources; for example, palm kernel oil, petrolatum, and coconut oil are all sources of lauric acid, the starting ingredient used to synthesize SLS. Sodium lauryl sulfate is created by reacting lauryl alcohol with sulfur trioxide gas, followed by neutralization with sodium carbonate (also a naturally occurring compound).6 Sodium lauryl sulfate and other sulfate-containing shampoos widely replaced the usage of traditional soaps formulated from animal or vegetable fats, as these latter formations created a film of insoluble calcium salts on the hair strands upon contact with water, resulting in tangled, dull-appearing hair.5 Additionally, sulfates were preferred to the alkaline pH of traditional soap, which can be harsh on hair strands and cause irritation of the skin and mucous membranes.14 Because they are highly water soluble, sulfates enable the formulation of clear shampoos. They exhibit remarkable cleaning properties and lather formation.5,14

Because sulfates are potent surfactants, they can remove dirt and debris as well as naturally produced healthy oils from the hair and scalp. As a result, sulfates can leave the hair feeling dry and stripped of moisture.4,5 Sulfates are used as the primary detergents in the formulation of deep-cleaning shampoos, which are designed for people who accumulate a heavy buildup of dirt, sebum, and debris from frequent use of styling products. Due to their potent detergency, these shampoos typically are not used on a daily basis but rather at longer intervals.15 A downside to sulfates is that they can have cosmetically unpleasant properties, which can be compensated for by including appropriate softening additives in shampoo formulations.4 A number of anionic surfactants such as olefin sulfonate, alkyl sulfosuccinate, acyl peptides, and alkyl ether carboxylates are well tolerated by the skin and are used together with other anionic and amphoteric surfactants to optimize shampoo properties. Alternatively, sulfate-free shampoos are cleansers compounded by the removal of the anionic group and switched for surfactants with less detergency.4,5

 

 

Preservatives and Parabens

Parabens refer to a group of esters of 4-hydroxybenzoic acid commonly used as preservatives in foods, pharmaceuticals, and cosmetics whose widespread use dates back to 1923.16 Concerns over the presence of parabens in shampoos and other cosmetics have been raised by patients for their reputed estrogenic and antiandrogenic effects and suspected involvement in carcinogenesis via endocrine modulation.16,17 In in vitro studies done on yeast assays, parabens have shown weak estrogenic activity that increases in proportion to both the length and increased branching of the alkyl side chains in the paraben’s molecular structure.18 They are 10,000-fold less potent than 17β-estradiol. In in vivo animal studies, parabens show weak estrogenic activity and are 100,000-fold less potent than 17β-estradiol.18 4-Hydroxybenzoic acid, a common metabolite, showed no estrogenic activity when tested both in vitro and in vivo.19 Some concerning research has implicated a link between parabens used in underarm cosmetics, such as deodorants and antiperspirants, and breast cancer16; however, the studies have been conflicting, and there is simply not enough data to assert that parabens cause breast cancer.

The Cosmetic Ingredient Review expert panel first reviewed parabens in 1984 and concluded that “methylparaben, ethylparaben, propylparaben, and butylparaben are safe as cosmetic ingredients in the present practices of use.”20 They extended this statement to include isopropylparaben and isobutylparaben in a later review.21 In 2005, the Scientific Committee on Consumer Products (now known as the Scientific Committee for Consumer Safety) in Europe stated that methylparaben and ethylparaben can be used at levels up to 0.4% in products.22 This decision was reached due to reports of decreased sperm counts and testosterone levels in male juvenile rats exposed to these parabens; however, these reults were not successfully replicated in larger studies.16,22 In 2010, the Scientific Committee for Consumer Safety revisited its stance on parabens, and they then revised their recommendations to say that concentrations of propylparaben and butylparaben should not exceed concentrations of 0.19%, based on “the conservative choice for the calculation of the [Margin-of-Safety] of butyl- and propylparaben.”23 However, in 2011 the use of propylparaben and butylparaben was banned in Denmark for cosmetic products used in children 3 years or younger,16 and the European Commission subsequently amended their directive in 2014, banning isopropylparaben, isobutylparaben, phenylparaben, benzylparaben, and pentylparaben due to lack of data available to evaluate the human risk of these products.24

Contrary to the trends in Europe, there currently are no regulations against the use of parabens in shampoos or other cosmetics in the United States. The American Cancer Society found that there is no evidence to suggest that the current levels of parabens in cosmetic products (eg, antiperspirants) increase one’s risk of breast cancer.25 Parabens are readily absorbed into the body both transdermally and through ingestion but also are believed to be rapidly transformed into harmless and nonspecific metabolites; they are readily metabolized by the liver and excreted in urine, and there is no measured accumulation in tissues.17

Parabens continue to be the most widely used preservatives in personal care products, usually in conjunction with other preservatives. Parabens are good biocides; short-chain esters (eg, methylparabens, ethylparabens) are effective against gram-positive bacteria and are weakly effective against gram-negative bacteria. Long-chain paraben esters (eg, propylparabens, butylparabens) are effective against mold and yeast. The addition of other preservatives creates a broad spectrum of antimicrobial defense in consumer products. Other preservatives include formaldehyde releasers or phenoxyethanol, as well as chelating agents such as EDTA, which improve the stability of these cosmetic products when exposed to air.16 Parabens are naturally occurring substances found in foods such as blueberries, barley, strawberries, yeast, olives, and grapes. As a colorless, odorless, and inexpensive substance, their use has been heavily favored in cosmetic and food products.16

 

 

Shampoo Alternatives and the No-Poo Method

Although research has not demonstrated any long-term danger to using shampoo, certain chemicals found in shampoos have the potential to irritate the scalp. Commonly cited allergens in shampoos include cocamidopropyl betaine, propylene glycol, vitamin E (tocopherol), parabens, and benzophenones.5 Additionally, the rising use of formaldehyde-releasing preservatives and isothiazolinones due to mounting pressures to move away from parabens has led to an increase in cases of allergic contact dermatitis (ACD).16 However, the irritability (rather than allergenicity) of these substances often is established during patch testing, a method of detecting delayed-type allergic reactions, which is important to note because patch testing requires a substance to be exposed to the skin for 24 to 48 hours, whereas exposure to shampoo ingredients may last a matter of minutes at most and occur in lesser concentrations because the ingredients are diluted by water in the rinsing process. Given these differences, it is unlikely that a patient would develop a true allergic response from regular shampoo use. Nevertheless, in patients who are already sensitized, exposure could conceivably trigger ACD, and patients must be cognizant of the composition of their shampoos.16

The no-poo method refers to the avoidance of commercial shampoo products when cleansing the hair and scalp and encompasses different methods of cleansing the hair, such as the use of household items (eg, baking soda, apple cider vinegar [ACV]), the use of conditioners to wash the hair (also known as conditioner-only washing or co-washing), treating the scalp with tea tree oil, or simply rinsing the hair with water. Proponents of the no-poo method believe that abstaining from shampoo use leads to healthier hair, retained natural oils, and less exposure to supposedly dangerous chemicals such as parabens or sulfates.2,3,26-28 However, there are no known studies in the literature that assess or support the hypotheses of the no-poo method.

Baking Soda and ACV
Baking soda (sodium bicarbonate) is a substance commonly found in the average household. It has been used in toothpaste formulas and cosmetic products and is known for its acid-neutralizing properties. Baking soda has been shown to have some antifungal and viricidal properties through an unknown mechanism of action.28 It has gained popularity for its use as a means of reducing the appearance of excessive greasiness of the hair shafts. Users also have reported that when washing their hair with baking soda, they are able to achieve a clean scalp and hair that feels soft to the touch.2,3,26,27,29 Despite these reports, users must beware of using baking soda without adequately diluting it with water. Baking soda is a known alkaline irritant.26,30 With a pH of 9, baking soda causes the cuticle layer of the hair fiber to open, increasing the capacity for water absorption. Water penetrates the scales that open, breaking the hydrogen bonds of the keratin molecule.31 Keratin is a spiral helical molecule that keeps its shape due to hydrogen, disulfide, and ionic bonds, as well as Van der Waals force.30 Hydrolysis of these bonds due to exposure to baking soda lowers the elasticity of the hair and increases the negative electrical net charge of the hair fiber surface, which leads to increased friction between fibers, cuticle damage, hair fragility, and fiber breakage.32,33

Apple cider vinegar is an apple-derived acetic acid solution with a pH ranging from 3.1 to 5.28 The pH range of ACV is considered to be ideal for hair by no-poo proponents, as it is similar to the natural pH of the scalp. Its acidic properties are responsible for its antimicrobial abilities, particularly its effectiveness against gram-negative bacteria.30 The acetic acid of ACV can partially interrupt oil interfaces, which contributes to its mild ability to remove product residue and scalp buildup from the hair shaft; the acetic acid also tightens the cuticles on hair fibers.33 Apple cider vinegar is used as a means of cleansing the hair and scalp by no-poo proponents2,3,26; other uses for ACV include using it as a rinse following washing and/or conditioning of the hair or as a means of preserving color in color-treated hair. There also is evidence that ACV may have antifungal properties.28 However, consumers must be aware that if it is not diluted in water, ACV may be too caustic for direct application to the hair and may lead to damage; it can be irritating to eyes, mucus membranes, and acutely inflamed skin. Also, vinegar rinses used on processed or chemically damaged hair may lead to increased hair fragility.2,3

Hair fibers have a pH of 3.67, while the scalp has a pH between 4.5 and 6.2. This slightly acidic film acts as a barrier to viruses, bacteria, and other potential contaminants.33 Studies have shown that the pH of skin increases in proportion to the pH of the cleanser used.34 Therefore, due to the naturally acidic pH of the scalp, acid-balanced shampoos generally are recommended. Shampoos should not have a pH higher than 5.5, as hair shafts can swell due to alkalinization, which can be prevented by pH balancing the shampoo through the addition of an acidic substance (eg, glycolic acid, citric acid) to lower the pH down to approximately 5.5. Apple cider vinegar often is used for this purpose. However, one study revealed that 82% of shampoos already have an acidic pH.34

Conditioner-Only Washing (Co-washing)
Conditioner-only washing, or co-washing, is a widely practiced method of hair grooming. It is popular among individuals who find that commercial shampoos strip too much of the natural hair oils away, leaving the hair rough or unmanageable. Co-washing is not harmful to the hair; however, the molecular structure and function of a conditioner and that of a shampoo are very different.5,35,36 Conditioners are not formulated to remove dirt and buildup in the hair but rather to add substances to the hair, and thus cannot provide extensive cleansing of the hair and scalp; therefore, it is inappropriate to use co-washing as a replacement for shampooing. Quaternary conditioning agents are an exception because they contain amphoteric detergents comprised of both anionic and cationic groups, which allow them both the ability to remove dirt and sebum with its anionic group, typically found in shampoos, as well as the ability to coat and condition the hair due to the high affinity of the cationic group for the negatively charged hair fibers.36,37 Amphoteric detergents are commonly found in 2-in-1 conditioning cleansers, among other ingredients, such as hydrolyzed animal proteins that temporarily plug surface defects on the hair fiber, and dimethicone, a synthetic oil that creates a thin film over the hair shaft, increasing shine and manageability. Of note, these conditioning shampoos are ideal for individuals with minimal product buildup on the hair and scalp and are not adequate scalp cleansers for individuals who either wash their hair infrequently or who regularly use hairstyling products.36,37

Tea Tree Oil
Tea tree oil is an essential oil extracted from the Melaleuca alternifolia plant of the Myrtaceae family. It is native to the coast of northeastern Australia. A holy grail of natural cosmetics, tea tree oil is widely known for its antiviral, antifungal, and antiseptic properties.38 Although not used as a stand-alone cleanser, it is often added to a number of cosmetic products, including shampoos and co-washes. Although deemed safe for topical use, it has been shown to be quite toxic when ingested. Symptoms of ingestion include nausea, vomiting, hallucinations, and coma. The common concern with tea tree oil is its ability to cause ACD. In particular, it is believed that the oxidation products of tea tree oil are allergenic rather than the tea tree oil itself. The evaluation of tea tree oil as a potential contact allergen has been quite difficult; it consists of more than 100 distinct compounds and is often mislabeled, or does not meet the guidelines of the International Organization for Standardization. Nonetheless, the prevalence of ACD due to tea tree oil is low (approximately 1.4%). Despite its low prevalence, tea tree oil should remain in the differential as an ACD-inducing agent. Patch testing with the patient’s supply of tea tree oil is advised when possible.38

Conclusion

It is customary that the ingredients used in shampoos undergo periodic testing and monitoring to assure the safety of their use. Although it is encouraging that patients are proactive in their efforts to stay abreast of the literature, it is still important that cosmetic scientists, dermatologists, and other experts remain at the forefront of educating the public about these substances. Not doing so can result in the propagation of misinformation and unnecessary fears, which can lead to the adaptation of unhygienic or even unsafe hair care practices. As dermatologists, we must ensure that patients are educated about the benefits and hazards of off-label use of household ingredients to the extent that evidence-based medicine permits. Patients must be informed that not all synthetic substances are harmful, and likewise not all naturally occurring substances are safe.

Shampoo is a staple in hair grooming that is ever-evolving along with cultural trends. The global shampoo market is expected to reach an estimated value of $25.73 billion by 2019. A major driver of this upward trend in market growth is the increasing demand for natural and organic hair shampoos.1 Society today has a growing fixation on healthy living practices, and as of late, the ingredients in shampoos and other cosmetic products have become one of the latest targets in the health-consciousness craze. In the age of the Internet where information—and misinformation—is widely accessible and dispersed, the general public often strives to self-educate on specialized matters that are out of their expertise. As a result, individuals have developed an aversion to using certain shampoos out of fear that the ingredients, often referred to as “chemicals” by patients due to their complex names, are unnatural and therefore unhealthy.1,2 Product developers are working to meet the demand by reformulating shampoos with labels that indicate sulfate free or paraben free, despite the lack of proof that these formulations are an improvement over traditional approaches to hair health. Additionally, alternative methods of cleansing the hair and scalp, also known as the no-shampoo or “no-poo” method, have begun to gain popularity.2,3


It is essential that dermatologists acknowledge the concerns that their patients have about common shampoo ingredients to dispel the myths that may misinform patient decision-making. This article reviews the controversy surrounding the use of sulfates and parabens in shampoos as well as commonly used shampoo alternatives. Due to the increased prevalence of dry hair shafts in the skin of color population, especially black women, this group is particularly interested in products that will minimize breakage and dryness of the hair. To that end, this population has great interest in the removal of chemical ingredients that may cause damage to the hair shafts, despite the lack of data to support sulfates and paraben damage to hair shafts or scalp skin. Blogs and uninformed hairstylists may propagate these beliefs in a group of consumers who are desperate for new approaches to hair fragility and breakage.

Surfactants and Sulfates

The cleansing ability of a shampoo depends on the surface activity of its detergents. Surface-active ingredients, or surfactants, reduce the surface tension between water and dirt, thus facilitating the removal of environmental dirt from the hair and scalp,4 which is achieved by a molecular structure containing both a hydrophilic and a lipophilic group. Sebum and dirt are bound by the lipophilic ends of the surfactant, becoming the center of a micelle structure with the hydrophilic molecule ends pointing outward. Dirt particles become water soluble and are removed from the scalp and hair shaft upon rinsing with water.4

Surfactants are classified according to the electric charge of the hydrophilic polar group as either anionic, cationic, amphoteric (zwitterionic), or nonionic.5 Each possesses different hair conditioning and cleansing qualities, and multiple surfactants are used in shampoos in differing ratios to accommodate different hair types. In most shampoos, the base consists of anionic and amphoteric surfactants. Depending on individual product requirements, nonionic and cationic surfactants are used to either modify the effects of the surfactants or as conditioning agents.4,5

One subcategory of surfactants that receives much attention is the group of anionic surfactants known as sulfates. Sulfates, particularly sodium lauryl sulfate (SLS), recently have developed a negative reputation as cosmetic ingredients, as reports from various unscientific sources have labeled them as hazardous to one’s health; SLS has been described as a skin and scalp irritant, has been linked to cataract formation, and has even been wrongly labeled as carcinogenic.6 The origins of some of these claims are not clear, though they likely arose from the misinterpretation of complex scientific studies that are easily accessible to laypeople. The link between SLS and ocular irritation or cataract formation is a good illustration of this unsubstantiated fear. A study by Green et al7 showed that corneal exposure to extremely high concentrations of SLS following physical or chemical damage to the eye can result in a slowed healing process. The results of this study have since been wrongly quoted to state that SLS-containing products lead to blindness or severe corneal damage.8 A different study tested for possible ocular irritation in vivo by submerging the lens of an eye into a 20% SLS solution, which accurately approximates the concentration of SLS in rinse-off consumer products.9 However, to achieve ocular irritation, the eyes of laboratory animals were exposed to SLS constantly for 14 days, which would not occur in practical use.9 Similarly, a third study achieved cataract formation in a laboratory only by immersing the lens of an eye into a highly concentrated solution of SLS.10 Such studies are not appropriate representations of how SLS-containing products are used by consumers and have unfortunately been vulnerable to misinterpretation by the general public.

There is no known study that has shown SLS to be carcinogenic. One possible origin of this idea may be from the wrongful interpretation of studies that used SLS as a vehicle substance to test agents that were deemed to be carcinogenic.11 Another possible source of the idea that SLS is carcinogenic comes from its association with 1,4-dioxane, a by-product of the synthesis of certain sulfates such as sodium laureth sulfate due to a process known as ethoxylation.6,12 Although SLS does not undergo this process in its formation and is not linked to 1,4-dioxane, there is potential for cross-contamination of SLS with 1,4-dioxane, which cannot be overlooked. 1,4-Dioxane is classified as “possibly carcinogenic to humans (Group 2B)” by the International Agency for Research on Cancer,13 but screening of SLS for this substance prior to its use in commercial products is standard.

Sulfates are inexpensive detergents that are responsible for lather formation in shampoos as well as in many household cleaning agents.5 Sulfates, similar to all anionic surfactants, are characterized by a negatively charged hydrophilic polar group. The best-known and most commonly used anionic surfactants are sulfated fatty alcohols, alkyl sulfates, and their polyethoxylated analogues alkyl ether sulfates.5,6 Sodium lauryl sulfate (also known as sodium laurilsulfate or sodium dodecyl sulfate) is the most common of them all, found in shampoo and conditioner formulations. Ammonium lauryl sulfate and sodium laureth sulfate are other sulfates commonly used in shampoos and household cleansing products. Sodium lauryl sulfate is a nonvolatile, water-soluble compound. Its partition coefficient (P0), a measure of a substance’s hydrophilic or lipophilic nature, is low at 1.6, making it a rather hydrophilic substance.6 Hydrophilic substances tend to have low bioaccumulation profiles in the body. Additionally, SLS is readily biodegradable. It can be derived from both synthetic and naturally occurring sources; for example, palm kernel oil, petrolatum, and coconut oil are all sources of lauric acid, the starting ingredient used to synthesize SLS. Sodium lauryl sulfate is created by reacting lauryl alcohol with sulfur trioxide gas, followed by neutralization with sodium carbonate (also a naturally occurring compound).6 Sodium lauryl sulfate and other sulfate-containing shampoos widely replaced the usage of traditional soaps formulated from animal or vegetable fats, as these latter formations created a film of insoluble calcium salts on the hair strands upon contact with water, resulting in tangled, dull-appearing hair.5 Additionally, sulfates were preferred to the alkaline pH of traditional soap, which can be harsh on hair strands and cause irritation of the skin and mucous membranes.14 Because they are highly water soluble, sulfates enable the formulation of clear shampoos. They exhibit remarkable cleaning properties and lather formation.5,14

Because sulfates are potent surfactants, they can remove dirt and debris as well as naturally produced healthy oils from the hair and scalp. As a result, sulfates can leave the hair feeling dry and stripped of moisture.4,5 Sulfates are used as the primary detergents in the formulation of deep-cleaning shampoos, which are designed for people who accumulate a heavy buildup of dirt, sebum, and debris from frequent use of styling products. Due to their potent detergency, these shampoos typically are not used on a daily basis but rather at longer intervals.15 A downside to sulfates is that they can have cosmetically unpleasant properties, which can be compensated for by including appropriate softening additives in shampoo formulations.4 A number of anionic surfactants such as olefin sulfonate, alkyl sulfosuccinate, acyl peptides, and alkyl ether carboxylates are well tolerated by the skin and are used together with other anionic and amphoteric surfactants to optimize shampoo properties. Alternatively, sulfate-free shampoos are cleansers compounded by the removal of the anionic group and switched for surfactants with less detergency.4,5

 

 

Preservatives and Parabens

Parabens refer to a group of esters of 4-hydroxybenzoic acid commonly used as preservatives in foods, pharmaceuticals, and cosmetics whose widespread use dates back to 1923.16 Concerns over the presence of parabens in shampoos and other cosmetics have been raised by patients for their reputed estrogenic and antiandrogenic effects and suspected involvement in carcinogenesis via endocrine modulation.16,17 In in vitro studies done on yeast assays, parabens have shown weak estrogenic activity that increases in proportion to both the length and increased branching of the alkyl side chains in the paraben’s molecular structure.18 They are 10,000-fold less potent than 17β-estradiol. In in vivo animal studies, parabens show weak estrogenic activity and are 100,000-fold less potent than 17β-estradiol.18 4-Hydroxybenzoic acid, a common metabolite, showed no estrogenic activity when tested both in vitro and in vivo.19 Some concerning research has implicated a link between parabens used in underarm cosmetics, such as deodorants and antiperspirants, and breast cancer16; however, the studies have been conflicting, and there is simply not enough data to assert that parabens cause breast cancer.

The Cosmetic Ingredient Review expert panel first reviewed parabens in 1984 and concluded that “methylparaben, ethylparaben, propylparaben, and butylparaben are safe as cosmetic ingredients in the present practices of use.”20 They extended this statement to include isopropylparaben and isobutylparaben in a later review.21 In 2005, the Scientific Committee on Consumer Products (now known as the Scientific Committee for Consumer Safety) in Europe stated that methylparaben and ethylparaben can be used at levels up to 0.4% in products.22 This decision was reached due to reports of decreased sperm counts and testosterone levels in male juvenile rats exposed to these parabens; however, these reults were not successfully replicated in larger studies.16,22 In 2010, the Scientific Committee for Consumer Safety revisited its stance on parabens, and they then revised their recommendations to say that concentrations of propylparaben and butylparaben should not exceed concentrations of 0.19%, based on “the conservative choice for the calculation of the [Margin-of-Safety] of butyl- and propylparaben.”23 However, in 2011 the use of propylparaben and butylparaben was banned in Denmark for cosmetic products used in children 3 years or younger,16 and the European Commission subsequently amended their directive in 2014, banning isopropylparaben, isobutylparaben, phenylparaben, benzylparaben, and pentylparaben due to lack of data available to evaluate the human risk of these products.24

Contrary to the trends in Europe, there currently are no regulations against the use of parabens in shampoos or other cosmetics in the United States. The American Cancer Society found that there is no evidence to suggest that the current levels of parabens in cosmetic products (eg, antiperspirants) increase one’s risk of breast cancer.25 Parabens are readily absorbed into the body both transdermally and through ingestion but also are believed to be rapidly transformed into harmless and nonspecific metabolites; they are readily metabolized by the liver and excreted in urine, and there is no measured accumulation in tissues.17

Parabens continue to be the most widely used preservatives in personal care products, usually in conjunction with other preservatives. Parabens are good biocides; short-chain esters (eg, methylparabens, ethylparabens) are effective against gram-positive bacteria and are weakly effective against gram-negative bacteria. Long-chain paraben esters (eg, propylparabens, butylparabens) are effective against mold and yeast. The addition of other preservatives creates a broad spectrum of antimicrobial defense in consumer products. Other preservatives include formaldehyde releasers or phenoxyethanol, as well as chelating agents such as EDTA, which improve the stability of these cosmetic products when exposed to air.16 Parabens are naturally occurring substances found in foods such as blueberries, barley, strawberries, yeast, olives, and grapes. As a colorless, odorless, and inexpensive substance, their use has been heavily favored in cosmetic and food products.16

 

 

Shampoo Alternatives and the No-Poo Method

Although research has not demonstrated any long-term danger to using shampoo, certain chemicals found in shampoos have the potential to irritate the scalp. Commonly cited allergens in shampoos include cocamidopropyl betaine, propylene glycol, vitamin E (tocopherol), parabens, and benzophenones.5 Additionally, the rising use of formaldehyde-releasing preservatives and isothiazolinones due to mounting pressures to move away from parabens has led to an increase in cases of allergic contact dermatitis (ACD).16 However, the irritability (rather than allergenicity) of these substances often is established during patch testing, a method of detecting delayed-type allergic reactions, which is important to note because patch testing requires a substance to be exposed to the skin for 24 to 48 hours, whereas exposure to shampoo ingredients may last a matter of minutes at most and occur in lesser concentrations because the ingredients are diluted by water in the rinsing process. Given these differences, it is unlikely that a patient would develop a true allergic response from regular shampoo use. Nevertheless, in patients who are already sensitized, exposure could conceivably trigger ACD, and patients must be cognizant of the composition of their shampoos.16

The no-poo method refers to the avoidance of commercial shampoo products when cleansing the hair and scalp and encompasses different methods of cleansing the hair, such as the use of household items (eg, baking soda, apple cider vinegar [ACV]), the use of conditioners to wash the hair (also known as conditioner-only washing or co-washing), treating the scalp with tea tree oil, or simply rinsing the hair with water. Proponents of the no-poo method believe that abstaining from shampoo use leads to healthier hair, retained natural oils, and less exposure to supposedly dangerous chemicals such as parabens or sulfates.2,3,26-28 However, there are no known studies in the literature that assess or support the hypotheses of the no-poo method.

Baking Soda and ACV
Baking soda (sodium bicarbonate) is a substance commonly found in the average household. It has been used in toothpaste formulas and cosmetic products and is known for its acid-neutralizing properties. Baking soda has been shown to have some antifungal and viricidal properties through an unknown mechanism of action.28 It has gained popularity for its use as a means of reducing the appearance of excessive greasiness of the hair shafts. Users also have reported that when washing their hair with baking soda, they are able to achieve a clean scalp and hair that feels soft to the touch.2,3,26,27,29 Despite these reports, users must beware of using baking soda without adequately diluting it with water. Baking soda is a known alkaline irritant.26,30 With a pH of 9, baking soda causes the cuticle layer of the hair fiber to open, increasing the capacity for water absorption. Water penetrates the scales that open, breaking the hydrogen bonds of the keratin molecule.31 Keratin is a spiral helical molecule that keeps its shape due to hydrogen, disulfide, and ionic bonds, as well as Van der Waals force.30 Hydrolysis of these bonds due to exposure to baking soda lowers the elasticity of the hair and increases the negative electrical net charge of the hair fiber surface, which leads to increased friction between fibers, cuticle damage, hair fragility, and fiber breakage.32,33

Apple cider vinegar is an apple-derived acetic acid solution with a pH ranging from 3.1 to 5.28 The pH range of ACV is considered to be ideal for hair by no-poo proponents, as it is similar to the natural pH of the scalp. Its acidic properties are responsible for its antimicrobial abilities, particularly its effectiveness against gram-negative bacteria.30 The acetic acid of ACV can partially interrupt oil interfaces, which contributes to its mild ability to remove product residue and scalp buildup from the hair shaft; the acetic acid also tightens the cuticles on hair fibers.33 Apple cider vinegar is used as a means of cleansing the hair and scalp by no-poo proponents2,3,26; other uses for ACV include using it as a rinse following washing and/or conditioning of the hair or as a means of preserving color in color-treated hair. There also is evidence that ACV may have antifungal properties.28 However, consumers must be aware that if it is not diluted in water, ACV may be too caustic for direct application to the hair and may lead to damage; it can be irritating to eyes, mucus membranes, and acutely inflamed skin. Also, vinegar rinses used on processed or chemically damaged hair may lead to increased hair fragility.2,3

Hair fibers have a pH of 3.67, while the scalp has a pH between 4.5 and 6.2. This slightly acidic film acts as a barrier to viruses, bacteria, and other potential contaminants.33 Studies have shown that the pH of skin increases in proportion to the pH of the cleanser used.34 Therefore, due to the naturally acidic pH of the scalp, acid-balanced shampoos generally are recommended. Shampoos should not have a pH higher than 5.5, as hair shafts can swell due to alkalinization, which can be prevented by pH balancing the shampoo through the addition of an acidic substance (eg, glycolic acid, citric acid) to lower the pH down to approximately 5.5. Apple cider vinegar often is used for this purpose. However, one study revealed that 82% of shampoos already have an acidic pH.34

Conditioner-Only Washing (Co-washing)
Conditioner-only washing, or co-washing, is a widely practiced method of hair grooming. It is popular among individuals who find that commercial shampoos strip too much of the natural hair oils away, leaving the hair rough or unmanageable. Co-washing is not harmful to the hair; however, the molecular structure and function of a conditioner and that of a shampoo are very different.5,35,36 Conditioners are not formulated to remove dirt and buildup in the hair but rather to add substances to the hair, and thus cannot provide extensive cleansing of the hair and scalp; therefore, it is inappropriate to use co-washing as a replacement for shampooing. Quaternary conditioning agents are an exception because they contain amphoteric detergents comprised of both anionic and cationic groups, which allow them both the ability to remove dirt and sebum with its anionic group, typically found in shampoos, as well as the ability to coat and condition the hair due to the high affinity of the cationic group for the negatively charged hair fibers.36,37 Amphoteric detergents are commonly found in 2-in-1 conditioning cleansers, among other ingredients, such as hydrolyzed animal proteins that temporarily plug surface defects on the hair fiber, and dimethicone, a synthetic oil that creates a thin film over the hair shaft, increasing shine and manageability. Of note, these conditioning shampoos are ideal for individuals with minimal product buildup on the hair and scalp and are not adequate scalp cleansers for individuals who either wash their hair infrequently or who regularly use hairstyling products.36,37

Tea Tree Oil
Tea tree oil is an essential oil extracted from the Melaleuca alternifolia plant of the Myrtaceae family. It is native to the coast of northeastern Australia. A holy grail of natural cosmetics, tea tree oil is widely known for its antiviral, antifungal, and antiseptic properties.38 Although not used as a stand-alone cleanser, it is often added to a number of cosmetic products, including shampoos and co-washes. Although deemed safe for topical use, it has been shown to be quite toxic when ingested. Symptoms of ingestion include nausea, vomiting, hallucinations, and coma. The common concern with tea tree oil is its ability to cause ACD. In particular, it is believed that the oxidation products of tea tree oil are allergenic rather than the tea tree oil itself. The evaluation of tea tree oil as a potential contact allergen has been quite difficult; it consists of more than 100 distinct compounds and is often mislabeled, or does not meet the guidelines of the International Organization for Standardization. Nonetheless, the prevalence of ACD due to tea tree oil is low (approximately 1.4%). Despite its low prevalence, tea tree oil should remain in the differential as an ACD-inducing agent. Patch testing with the patient’s supply of tea tree oil is advised when possible.38

Conclusion

It is customary that the ingredients used in shampoos undergo periodic testing and monitoring to assure the safety of their use. Although it is encouraging that patients are proactive in their efforts to stay abreast of the literature, it is still important that cosmetic scientists, dermatologists, and other experts remain at the forefront of educating the public about these substances. Not doing so can result in the propagation of misinformation and unnecessary fears, which can lead to the adaptation of unhygienic or even unsafe hair care practices. As dermatologists, we must ensure that patients are educated about the benefits and hazards of off-label use of household ingredients to the extent that evidence-based medicine permits. Patients must be informed that not all synthetic substances are harmful, and likewise not all naturally occurring substances are safe.

References
  1. The global shampoo market 2014-2019 trends, forecast, and opportunity analysis [press release]. New York, NY: Reportlinker; May 21, 2015.
  2. Is the ‘no shampoo’ trend healthy or harmful? Mercola website. Published January 16, 2016. Accessed December 8, 2017.
  3. Feltman R. The science (or lack thereof) behind the ‘no-poo’ hair trend. Washington Post. March 10, 2016. https://www.washingtonpost.com/news/speaking-of-science/wp/2016/03/10/the-science-or-lack-thereof-behind-the-no-poo-hair-trend/?utm_term=.9a61edf3fd5a. Accessed December 11, 2017.
  4. Bouillon C. Shampoos. Clin Dermatol. 1996;14:113-121.
  5. Trueb RM. Shampoos: ingredients, efficacy, and adverse effects. J Dtsch Dermatol Ges. 2007;5:356-365.
  6. Bondi CA, Marks JL, Wroblewski LB, et al. Human and environmental toxicity of sodium lauryl sulfate (SLS): evidence for safe use in household cleaning products. Environ Health Insights. 2015;9:27-32.
  7. Green K, Johnson RE, Chapman JM, et al. Preservative effects on the healing rate of rabbit corneal epithelium. Lens Eye Toxic Res. 1989;6:37-41.
  8. Sodium lauryl sulphate. Healthy Choices website. http://www.healthychoices.co.uk/sls.html. Accessed December 8, 2017.
  9. Tekbas¸ ÖF, Uysal Y, Og˘ur R, et al. Non-irritant baby shampoos may cause cataract development. TSK Koruyucu Hekimlik Bülteni. 2008;1:1-6.
  10. Cater KC, Harbell JW. Prediction of eye irritation potential of surfactant-based rinse-off personal care formulations by the bovine corneal opacity and permeability (BCOP) assay. Cutan Ocul Toxicol. 2006;25:217-233.
  11. Birt DF, Lawson TA, Julius AD, et al. Inhibition by dietary selenium of colon cancer induced in the rat by bis(2-oxopropyl) nitrosamine. Cancer Res. 1982;42:4455-4459.
  12. Rastogi SC. Headspace analysis of 1,4-dioxane in products containing polyethoxylated surfactants by GC-MS. Chromatographia. 1990;29:441-445.
  13. 1,4-Dioxane. IARC Monogr Eval Carcinog Risks Hum. 1999;71, pt 2:589-602.
  14. Trueb RM. Dermocosmetic aspects of hair and scalp. J Investig Dermatol Symp Proc. 2005;10:289-292.
  15. D’Souza P, Rathi SK. Shampoo and conditioners: what a dermatologist should know? Indian J Dermatol. 2015;60:248-254.
  16. Sasseville D, Alfalah M, Lacroix JP. “Parabenoia” debunked, or “who’s afraid of parabens?” Dermatitis. 2015;26:254-259.
  17. Krowka JF, Loretz L, Geis PA, et al. Preserving the facts on parabens: an overview of these important tools of the trade. Cosmetics & Toiletries. http://www.cosmeticsandtoiletries.com/research/chemistry/Preserving-the-Facts-on-Parabens-An-Overview-of-These-Important-Tools-of-the Trade-425784294.html. Published June 1, 2017. Accessed December 20, 2017.
  18. Routledge EJ, Parker J, Odum J, et al. Some alkyl hydroxy benzoate preservatives (parabens) are estrogenic. Toxicol Appl Pharmacol. 1998;153:12Y19.
  19. Hossaini A, Larsen JJ, Larsen JC. Lack of oestrogenic effects of food preservatives (parabens) in uterotrophic assays. Food Chem Toxicol. 2000;38:319-323.
  20. Cosmetic Ingredient Review. Final report on the safety assessment of methylparaben, ethylparaben, propylparaben and butylparaben. J Am Coll Toxicol. 1984;3:147-209.
  21. Cosmetic Ingredient Review. Final report on the safety assessment of isobutylparaben and isopropylparaben. J Am Coll Toxicol. 1995;14:364-372.
  22. Scientific Committee on Consumer Products. Extended Opinion on the Safety Evaluation of Parabens. European Commission website. https://ec.europa.eu/health/ph_risk/committees/04_sccp/docs/sccp_o_019.pdf. Published January 28, 2005. Accessed December 20, 2017.
  23. Scientific Committee on Consumer Products. Opinion on Parabens. European Commission website. http://ec.europa.eu/health/scientific_committees/consumer_safety/docs/sccs_o_041.pdf. Revised March 22, 2011. Accessed December 20, 2017.
  24. European Commission. Commission Regulation (EU) No 258/2014 of 9 April 2014 amending Annexes II and V to Regulation (EC) No 1223/2009 of the European Parliament and of the Council on cosmetic products. EUR-Lex website. http://eur-lex.europa.eu/legal-content/EN/TXT/?uri=uriserv:OJ.L_.2014.107.01.0005.01.ENG. Accessed December 20, 2017.
  25. American Cancer Society. Antiperspirants and breast cancer risk. https://www.cancer.org/cancer/cancer-causes/antiperspirants-and-breast-cancer-risk.html#references. Revised October 14, 2014. Accessed January 2, 2018.
  26. MacMillan A. Cutting back on shampoo? 15 things you should know. Health. February 25, 2014. http://www.health.com/health/gallery/0,,20788089,00.html#should-you-go-no-poo--1. Accessed December 10, 2017.
  27. The ‘no poo’ method. https://www.nopoomethod.com/. Accessed December 10, 2017.
  28. Fong, D, Gaulin C, Le M, et al. Effectiveness of alternative antimicrobial agents for disinfection of hard surfaces. National Collaborating Centre for Environmental Health website. http://www.ncceh.ca/sites/default/files/Alternative_Antimicrobial_Agents_Aug_2014.pdf. Published August 2014. Accessed December 10, 2017.
  29. Is baking soda too harsh for natural hair? Black Girl With Long Hair website. http://blackgirllonghair.com/2012/02/is-baking-soda-too-harsh-for-hair/2/. Published February 5, 2012. Accessed December 12, 2017.
  30. O’Lenick T. Anionic/cationic complexes in hair care. J Cosmet Sci. 2011;62:209-228.
  31. Gavazzoni Dias MF, de Almeida AM, Cecato PM, et al. The shampoo pH can affect the hair: myth or reality? Int J Trichology. 2014;6:95-99.
  32. Goodman H. The acid mantle of the skin surface. Ind Med Surg. 1958;27:105-108.
  33. Korting HC, Kober M, Mueller M, et al. Influence of repeated washings with soap and synthetic detergents on pH and resident flora of the skin of forehead and forearm. results of a cross-over trial in health probationers. Acta Derm Venereol. 1987;67:41-47.
  34. Tarun J, Susan J, Suria J, et al. Evaluation of pH of bathing soaps and shampoos for skin and hair care. Indian J Dermatol. 2014;59:442-444.
  35. Corbett JF. The chemistry of hair-care products. J Soc Dyers Colour. 1976;92:285-303.
  36. McMichael AJ, Hordinsky M. Hair Diseases: Medical, Surgical, and Cosmetic Treatments. New York, NY: Taylor & Francis; 2008:59-72.
  37. Allardice A, Gummo G. Hair conditioning: quaternary ammonium compounds on various hair types. Cosmet Toiletries. 1993;108:107-109.
  38. Larson D, Jacob SE. Tea tree oil. Dermatitis. 2012;23:48-49.
References
  1. The global shampoo market 2014-2019 trends, forecast, and opportunity analysis [press release]. New York, NY: Reportlinker; May 21, 2015.
  2. Is the ‘no shampoo’ trend healthy or harmful? Mercola website. Published January 16, 2016. Accessed December 8, 2017.
  3. Feltman R. The science (or lack thereof) behind the ‘no-poo’ hair trend. Washington Post. March 10, 2016. https://www.washingtonpost.com/news/speaking-of-science/wp/2016/03/10/the-science-or-lack-thereof-behind-the-no-poo-hair-trend/?utm_term=.9a61edf3fd5a. Accessed December 11, 2017.
  4. Bouillon C. Shampoos. Clin Dermatol. 1996;14:113-121.
  5. Trueb RM. Shampoos: ingredients, efficacy, and adverse effects. J Dtsch Dermatol Ges. 2007;5:356-365.
  6. Bondi CA, Marks JL, Wroblewski LB, et al. Human and environmental toxicity of sodium lauryl sulfate (SLS): evidence for safe use in household cleaning products. Environ Health Insights. 2015;9:27-32.
  7. Green K, Johnson RE, Chapman JM, et al. Preservative effects on the healing rate of rabbit corneal epithelium. Lens Eye Toxic Res. 1989;6:37-41.
  8. Sodium lauryl sulphate. Healthy Choices website. http://www.healthychoices.co.uk/sls.html. Accessed December 8, 2017.
  9. Tekbas¸ ÖF, Uysal Y, Og˘ur R, et al. Non-irritant baby shampoos may cause cataract development. TSK Koruyucu Hekimlik Bülteni. 2008;1:1-6.
  10. Cater KC, Harbell JW. Prediction of eye irritation potential of surfactant-based rinse-off personal care formulations by the bovine corneal opacity and permeability (BCOP) assay. Cutan Ocul Toxicol. 2006;25:217-233.
  11. Birt DF, Lawson TA, Julius AD, et al. Inhibition by dietary selenium of colon cancer induced in the rat by bis(2-oxopropyl) nitrosamine. Cancer Res. 1982;42:4455-4459.
  12. Rastogi SC. Headspace analysis of 1,4-dioxane in products containing polyethoxylated surfactants by GC-MS. Chromatographia. 1990;29:441-445.
  13. 1,4-Dioxane. IARC Monogr Eval Carcinog Risks Hum. 1999;71, pt 2:589-602.
  14. Trueb RM. Dermocosmetic aspects of hair and scalp. J Investig Dermatol Symp Proc. 2005;10:289-292.
  15. D’Souza P, Rathi SK. Shampoo and conditioners: what a dermatologist should know? Indian J Dermatol. 2015;60:248-254.
  16. Sasseville D, Alfalah M, Lacroix JP. “Parabenoia” debunked, or “who’s afraid of parabens?” Dermatitis. 2015;26:254-259.
  17. Krowka JF, Loretz L, Geis PA, et al. Preserving the facts on parabens: an overview of these important tools of the trade. Cosmetics & Toiletries. http://www.cosmeticsandtoiletries.com/research/chemistry/Preserving-the-Facts-on-Parabens-An-Overview-of-These-Important-Tools-of-the Trade-425784294.html. Published June 1, 2017. Accessed December 20, 2017.
  18. Routledge EJ, Parker J, Odum J, et al. Some alkyl hydroxy benzoate preservatives (parabens) are estrogenic. Toxicol Appl Pharmacol. 1998;153:12Y19.
  19. Hossaini A, Larsen JJ, Larsen JC. Lack of oestrogenic effects of food preservatives (parabens) in uterotrophic assays. Food Chem Toxicol. 2000;38:319-323.
  20. Cosmetic Ingredient Review. Final report on the safety assessment of methylparaben, ethylparaben, propylparaben and butylparaben. J Am Coll Toxicol. 1984;3:147-209.
  21. Cosmetic Ingredient Review. Final report on the safety assessment of isobutylparaben and isopropylparaben. J Am Coll Toxicol. 1995;14:364-372.
  22. Scientific Committee on Consumer Products. Extended Opinion on the Safety Evaluation of Parabens. European Commission website. https://ec.europa.eu/health/ph_risk/committees/04_sccp/docs/sccp_o_019.pdf. Published January 28, 2005. Accessed December 20, 2017.
  23. Scientific Committee on Consumer Products. Opinion on Parabens. European Commission website. http://ec.europa.eu/health/scientific_committees/consumer_safety/docs/sccs_o_041.pdf. Revised March 22, 2011. Accessed December 20, 2017.
  24. European Commission. Commission Regulation (EU) No 258/2014 of 9 April 2014 amending Annexes II and V to Regulation (EC) No 1223/2009 of the European Parliament and of the Council on cosmetic products. EUR-Lex website. http://eur-lex.europa.eu/legal-content/EN/TXT/?uri=uriserv:OJ.L_.2014.107.01.0005.01.ENG. Accessed December 20, 2017.
  25. American Cancer Society. Antiperspirants and breast cancer risk. https://www.cancer.org/cancer/cancer-causes/antiperspirants-and-breast-cancer-risk.html#references. Revised October 14, 2014. Accessed January 2, 2018.
  26. MacMillan A. Cutting back on shampoo? 15 things you should know. Health. February 25, 2014. http://www.health.com/health/gallery/0,,20788089,00.html#should-you-go-no-poo--1. Accessed December 10, 2017.
  27. The ‘no poo’ method. https://www.nopoomethod.com/. Accessed December 10, 2017.
  28. Fong, D, Gaulin C, Le M, et al. Effectiveness of alternative antimicrobial agents for disinfection of hard surfaces. National Collaborating Centre for Environmental Health website. http://www.ncceh.ca/sites/default/files/Alternative_Antimicrobial_Agents_Aug_2014.pdf. Published August 2014. Accessed December 10, 2017.
  29. Is baking soda too harsh for natural hair? Black Girl With Long Hair website. http://blackgirllonghair.com/2012/02/is-baking-soda-too-harsh-for-hair/2/. Published February 5, 2012. Accessed December 12, 2017.
  30. O’Lenick T. Anionic/cationic complexes in hair care. J Cosmet Sci. 2011;62:209-228.
  31. Gavazzoni Dias MF, de Almeida AM, Cecato PM, et al. The shampoo pH can affect the hair: myth or reality? Int J Trichology. 2014;6:95-99.
  32. Goodman H. The acid mantle of the skin surface. Ind Med Surg. 1958;27:105-108.
  33. Korting HC, Kober M, Mueller M, et al. Influence of repeated washings with soap and synthetic detergents on pH and resident flora of the skin of forehead and forearm. results of a cross-over trial in health probationers. Acta Derm Venereol. 1987;67:41-47.
  34. Tarun J, Susan J, Suria J, et al. Evaluation of pH of bathing soaps and shampoos for skin and hair care. Indian J Dermatol. 2014;59:442-444.
  35. Corbett JF. The chemistry of hair-care products. J Soc Dyers Colour. 1976;92:285-303.
  36. McMichael AJ, Hordinsky M. Hair Diseases: Medical, Surgical, and Cosmetic Treatments. New York, NY: Taylor & Francis; 2008:59-72.
  37. Allardice A, Gummo G. Hair conditioning: quaternary ammonium compounds on various hair types. Cosmet Toiletries. 1993;108:107-109.
  38. Larson D, Jacob SE. Tea tree oil. Dermatitis. 2012;23:48-49.
Issue
Cutis - 101(1)
Issue
Cutis - 101(1)
Page Number
22-26
Page Number
22-26
Publications
Publications
Topics
Article Type
Display Headline
No Sulfates, No Parabens, and the “No-Poo” Method: A New Patient Perspective on Common Shampoo Ingredients
Display Headline
No Sulfates, No Parabens, and the “No-Poo” Method: A New Patient Perspective on Common Shampoo Ingredients
Sections
Inside the Article

Practice Points

  • The ingredients in shampoos and other cosmetic products have become scrutinized by the general public and the Internet has contributed to misinformation about certain shampoos.
  • Dermatologists must be prepared to acknowledge the concerns that their patients have about common shampoo ingredients to dispel the myths that may misinform patient decision-making.
  • This article reviews the controversy surrounding the use of sulfates and parabens in shampoos, as well as commonly used shampoo alternatives, often called the “no-poo” method.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media

Approach to Treatment of Medical and Cosmetic Facial Concerns in Skin of Color Patients

Article Type
Changed
Thu, 10/29/2020 - 14:55
Display Headline
Approach to Treatment of Medical and Cosmetic Facial Concerns in Skin of Color Patients
In Collaboration With the Skin of Color Society

The approach to the treatment of common skin disorders and cosmetic concerns in patients with skin of color (SOC) requires the clinician to understand the biological differences, nuances, and special considerations that are unique to patients with darker skin types.1-3 This article addresses 4 common facial concerns in SOC patients—acne, rosacea, facial hyperpigmentation, and cosmetic enhancement—and provides treatment recommendations and management pearls to assist the clinician with optimal outcomes for SOC patients.

Acne in SOC Patients

Acne vulgaris is one of the most common conditions that dermatologists treat and is estimated to affect 40 to 50 million individuals in the United States.1 Many of these acne patients are individuals with SOC.2-4 A study of 2835 females (aged 10–70 years) conducted in 4 different cities—Los Angeles, California; London, United Kingdom; Akita, Japan; and Rome, Italy—demonstrated acne prevalence of 37% in blacks, 32% in Hispanics, 30% in Asians, 24% in whites, and 23% in Continental Indians.5 Blacks, Hispanics, and Continental Indians demonstrated equal prevalence with comedonal and inflammatory acne. Asians displayed more inflammatory acne lesions than comedones. In contrast, whites demonstrated more comedones than inflammatory acne. Dyspigmentation, postinflammatory hyperpigmentation (PIH), and atrophic scars were more common in black and Hispanic females than other ethnicities.5 This study illustrated that acne-induced PIH is a common sequela in SOC patients and is the main reason they seek treatment.6,7

The pathogenesis of acne is the same in all racial and ethnic groups: (1) follicular hyperkeratinization and the formation of a microcomedone caused by abnormal desquamation of the keratinocytes within the sebaceous follicle, (2) production of sebum by circulating androgens, (3) proliferation of Propionibacterium acnes, and (4) inflammation. Subclinical inflammation is present throughout all stages of acne, including normal-appearing skin, inflammatory lesions, comedones, and scarring, and may contribute to PIH in acne patients with SOC (Figure 1).8 A thorough history should be obtained from acne patients, including answers to the following questions7:

  • What skin and hair care products do you use?
  • Do you use sunscreen daily?
  • What cosmetic products or makeup do you use?
  • Do you use any ethnic skin care products, including skin lightening creams?
  • Do you have a history of keloids?

Figure 1. Acne and postinflammatory hyperpigmentation in a patient with skin of color (Fitzpatrick skin type V).

It is important to ask these questions to assess if the SOC patient has developed pomade acne,9 acne cosmetica,10 or a potential risk of skin irritation from the use of skin care practices. It is best to take total control of the patient’s skin care regimen and discontinue use of toners, astringents, witch hazel, exfoliants, and rubbing alcohol, which may lead to skin dryness and irritation, particularly when combined with topical acne medications.

Treatment
Treatment of acne in SOC patients is similar to generally recommended treatments, with special considerations. Consider the following key points when treating acne in SOC patients:

  • Treat acne early and aggressively to prevent or minimize subsequent PIH and acne scarring.
  • Balance aggressive treatment with nonirritating topical skin care.
  • Most importantly, target PIH in addition to acne and choose a regimen that limits skin irritation that might exacerbate existing PIH.7

Develop a maintenance program to control future breakouts. Topical agents can be used as monotherapy or in fixed combinations and may include benzoyl peroxide, antibiotics, dapsone, azelaic acid (AZA), and retinoids. Similar to white patients, topical retinoids remain a first-line treatment for acne in patients with SOC.11,12

Tolerability must be managed in SOC acne patients. Therapeutic maneuvers that can be instituted should include a discussion on using gentle skin care, initiating therapy with a retinoid applied every other night starting with a low concentration and gradually titrating up, and applying a moisturizer before or after applying acne medication. Oral therapies consist of antibiotics (doxycycline, minocycline), retinoids (isotretinoin), and hormonal modulators (oral contraceptives, spironolactone). Isotretinoin, recommended for patients with nodulocystic acne, may play a possible role in treating acne-induced PIH.13

Two common procedural therapies for acne include comedone extraction and intralesional corticosteroid injection. A 6- to 8-week course of a topical retinoid prior to comedonal extraction may facilitate the procedure and is recommended in SOC patients to help reduce cutaneous trauma and PIH.11 Inflammatory acne lesions can be treated with intralesional injection of triamcinolone acetonide 2.5 or 5.0 mg/mL, which usually reduces inflammation within 2 to 5 days.11

Treatment of acne-induced PIH includes sun protection, topical and oral medications, chemical peels, lasers, and energy devices. Treatment of hypertrophic scarring and keloids involves intralesional injection of triamcinolone acetonide 20, 30, or 40 mg/mL every 4 weeks until the lesion is flat.11

Superficial chemical peels can be used to treat acne and PIH in SOC patients,14 such as salicylic acid (20%–30%), glycolic acid (20%–70%), trichloroacetic acid (15%–30%), and Jessner peels.

Acne Scarring
Surgical approaches to acne scarring in patients with SOC include elliptical excision, punch excision, punch elevation, punch autografting, dermal grafting, dermal planning, subcutaneous incision (subcision), dermabrasion, microneedling, fillers, and laser skin resurfacing. The treatment of choice depends on the size, type, and depth of the scar and the clinician’s preference.

Lasers
Fractional photothermolysis has emerged as a treatment option for acne scars in SOC patients. This procedure produces microscopic columns of thermal injury in the epidermis and dermis, sparing the surrounding tissue and minimizing downtime and adverse events. Because fractional photothermolysis does not target melanin and produces limited epidermal injury, darker Fitzpatrick skin types (IV–VI) can be safely and effectively treated with this procedure.15

 

 

Rosacea in SOC Patients

Rosacea is a chronic inflammatory disorder that affects the vasculature and pilosebaceous units of the face. It commonly is seen in Fitzpatrick skin types I and II; however, rosacea can occur in all skin types (Figure 2). Triggers include emotional stress, extreme environmental temperatures, hot and spicy foods, red wine or alcohol, and topical irritants or allergens found in common cosmetic products.16

Figure 2. Rosacea in a patient with skin of color (Fitzpatrick skin type IV).

Data suggest that 4% of rosacea patients in the United States are of African, Latino, or Asian descent.11 National Ambulatory Medical Care Survey data revealed that of 31.5 million rosacea visits, 2% of patients were black, 2.3% were Asian or Pacific Islander, and 3.9% were Hispanic or Latino. In a 5-year longitudinal study of 2587 rosacea patients enrolled in Medicaid in North Carolina who were prescribed at least 1 topical treatment for rosacea, 16.27% were black and 10% were of a race other than white.17

Although the pathogenesis of rosacea is unclear, hypotheses include immune system abnormalities, neurogenic dysregulation, presence of microorganisms (eg, Demodex folliculorum), UV damage, and skin barrier dysfunction.18

The 4 major subtypes of rosacea are erythematotelangiectatic, papulopustular, phymatous, and ocular rosacea.16 Interestingly, rosacea in SOC patients may present with hypopigmentation surrounding the borders of the facial erythema. For phymatous rosacea, isotretinoin may reduce incipient rhinophyma but must be carefully monitored and pregnancy must be excluded. Surgical or laser therapy may be indicated to recontour the nose if severe.

There are several skin conditions that can present with facial erythema in patients with SOC, including seborrheic dermatitis, systemic lupus erythematosus, and contact dermatitis. It is important to note that the detection of facial erythema in darker skin types may be difficult; therefore, laboratory evaluation (antinuclear antibodies), patch testing, and skin biopsy should be considered if the clinical diagnosis is unclear.

Treatment
Treatment of rosacea in SOC patients does not differ from other racial groups. Common strategies include gentle skin care, sun protection (sun protection factor 30+), and barrier repair creams. Topical agents include metronidazole, AZA, sodium sulfacetamide/sulfur, ivermectin, and retinoids.16 Oral treatments include antibiotics in the tetracycline family (eg, subantimicrobial dose doxycycline) and isotretinoin.16 Persistent erythema associated with rosacea can be treated with brimonidine19 and oxymetazoline.20 Vascular lasers and intense pulsed light may be used to address the vascular components of rosacea21; however, the latter is not recommended in Fitzpatrick skin types IV through VI.

Facial Hyperpigmentation in SOC Patients

Hyperpigmentation disorders can be divided into conditions that affect Fitzpatrick skin types I through III and IV though VI. Mottled hyperpigmentation (photodamage) and solar lentigines occur in patients with lighter skin types as compared to melasma, PIH, and age-related (UV-induced) hyperpigmentation, which occur more commonly in patients with darker skin types. Facial hyperpigmentation is a common concern in SOC patients. In a survey of cosmetic concerns of 100 women with SOC, hyperpigmentation or dark spots (86%) and blotchy uneven skin (80%) were the top concerns.22 In addition, facial hyperpigmentation has been shown to negatively impact quality of life.23

Postinflammatory hyperpigmentation occurs from a pathophysiological response to inflammation, cutaneous irritation or injury, and subsequent melanocyte lability. Postinflammatory hyperpigmentation is a common presenting concern in patients with SOC and is seen as a result of many inflammatory skin disorders (eg, acne, eczema) and dermatologic procedures (eg, adverse reaction to electrodesiccation, microdermabrasion, chemical peels, laser surgery).24

Melasma is an acquired idiopathic disorder of hyperpigmentation and often referred to as the mask of pregnancy (Figure 3). It occurs on sun-exposed areas of skin, mainly in women with Fitzpatrick skin types III through V. Associated factors or triggers include pregnancy, hormonal treatments, exposure to UV radiation, and medications.25 Hereditary factors play a role in more than 40% of cases.26

Figure 3. Facial hyperpigmentation consistent with melasma in a patient with skin of color (Fitzpatrick skin type IV).

Other not-so-common facial dyschromias include contact dermatitis, acanthosis nigricans, exogenous ochronosis, lichen planus pigmentosus (associated with frontal fibrosing alopecia),27 drug-induced hyperpigmentation (associated with minocycline or diltiazem),28,29 and UV-induced (age-related) hyperpigmentation.

Treatment
The treatment of hyperpigmentation should provide the following: (1) protection from sun exposure; (2) inhibition of tyrosinase, the enzyme responsible for the conversion of tyrosine to melanin; (3) inhibition of melanosome transfer from the melanocyte to the keratinocyte; (4) removal of melanin from the epidermis through exfoliation; and (5) destruction or disruption of melanin in the dermis.30 Therapies for facial hyperpigmentation are listed in Table 1.

Topical therapies include prescription medications and nonprescription cosmeceuticals. Prescription medications include hydroquinone (HQ), topical retinoids, and AZA. Hydroquinone, a tyrosinase inhibitor, is the gold standard for skin lightening and often is used as a first-line therapy. It is used as a monotherapy (HQ 4%) or as a fixed combination with tretinoin 0.05% and fluocinolone 0.01%.31 Use caution with HQ in high concentrations (6% and higher) and low concentrations (2% [over-the-counter strength]) used long-term due to the potential risk of exogenous ochronosis.

Topical retinoids have been shown to be effective therapeutic agents for melasma and PIH. Tretinoin,32 tazarotene,33 and adapalene34 all have demonstrated efficacy for acne and acne-induced PIH in SOC patients. Patients must be monitored for the development of retinoid dermatitis and worsening of hyperpigmentation.

Azelaic acid is a naturally occurring dicarboxylic acid obtained from cultures of Malassezia furfur. Azelaic acid inhibits tyrosinase activity, DNA synthesis, and mitochondrial enzymes, thus blocking direct cytotoxic effects toward melanocytes. Azelaic acid is approved by the US Food and Drug Administration for acne in a 20% cream formulation and rosacea in 15% gel and foam formulations, and it is used off label for melasma and PIH.35

Oral tranexamic acid is currently used as a hemostatic agent due to its ability to inhibit the plasminogen-plasmin pathway. In melasma, it blocks the interaction between melanocytes and keratinocytes in the epidermis and modulates the vascular component of melasma in the dermis. In an open-label study, 561 Asian melasma patients were treated with oral tranexamic acid 250 mg twice daily for 4 months. Results demonstrated improvement in 90% of patients, and 7.1% reported adverse effects (eg, abdominal bloating and pain, nausea, vomiting, headache, tinnitus, numbness, menstrual irregularities).36 Coagulation screening should be monitored monthly, and any patient with a history of clotting abnormalities should be excluded from off-label treatment with oral tranexamic acid.

Nonprescription cosmeceuticals are available over-the-counter or are office dispensed.37 For optimal results, cosmeceutical agents for skin lightening are used in combination. Most of these combinations are HQ free and have additive benefits such as a multimodal skin lightening agent containing key ingredients that correct and prevent skin pigmentation via several pathways affecting melanogenesis.38 It is an excellent alternative to HQ for mottled and diffuse UV-induced hyperpigmentation and can be used for maintenance therapy in patients with melasma.

Photoprotection is an essential component of therapy for melasma and PIH, but there is a paucity of data on the benefits for SOC patients. Halder et al39 performed a randomized prospective study of 89 black and Hispanic patients who applied sunscreen with a sun protection factor of 30 or 60 daily for 8 weeks. Clinical grading, triplicate L*A*B chromameter, and clinical photography were taken at baseline and weeks 4 and 8. The results demonstrated skin lightening in both black and Hispanic patients and support the use of sunscreen in the prevention and management of dyschromia in SOC patients.39 Visible light also may play a role in melasma development, and thus use of sunscreens or makeup containing iron oxides are recommended.40

Procedural treatments for facial hyperpigmentation include microdermabrasion, chemical peels, lasers, energy-based devices, and microneedling. There are many types and formulations of chemical peeling agents available; however, superficial and medium-depth chemical peels are recommended for SOC patients (Table 2). Deep chemical peels are not recommended for SOC patients due to the potential increased risk for PIH and scarring.

 

 

Cosmetic Enhancement in SOC Patients

Cosmetic procedures are gaining popularity in the SOC population and account for more than 20% of cosmetic procedures in the United States.41 Facial cosmetic concerns in SOC include dyschromia, benign growths (dermatosis papulosa nigra), hyperkinetic facial lines, volume loss, and skin laxity.42 Key principles to consider when treating SOC patients are the impact of ethnicity on aging and facial structure, the patient’s desired cosmetic outcome, tissue reaction to anticipated treatments, and the patient’s expectations for recommended therapies.

Aging in SOC Patients
Skin aging can be classified as intrinsic aging or extrinsic aging. Intrinsic aging is genetic and involves subsurface changes such as volume loss, muscle atrophy, and resorption of bony structure. Extrinsic aging (or photoaging) involves surface changes of the epidermis/dermis and manifests as mottled pigmentation, textural changes, and fine wrinkling. Due to the photoprotection of melanin (black skin=SPF 13.4), skin aging in SOC patients is delayed by 10 to 20 years.43 In addition, SOC patients have more reactive collagen and can benefit from noninvasive cosmetic procedures such as fillers and skin-tightening procedures.42

Cosmetic Treatments and Procedures
Dermatosis papulosa nigra (benign growths of skin that have a genetic predisposition)44 occur mainly on the face but can involve the entire body. Treatment modalities include electrodesiccation, cryotherapy, scissor excision, and laser surgery.45

Treatment of hyperkinetic facial lines with botulinum toxin type A is a safe and effective procedure in patients with SOC. Grimes and Shabazz46 performed a 4-month, randomized, double-blind study that evaluated the treatment of glabellar lines in women with Fitzpatrick skin types V and VI. The results demonstrated that the duration of effects was the same in the patients who received either 20 or 30 U of botulinum toxin type A.46 Dynamic rhytides (furrows and frown/scowl lines arising from laughing, frowning, or smiling) can be treated safely in patients with SOC using botulinum toxin type A off label for relaxation of the upper and lower hyperkinetic muscles that result in these unwanted signs of aging. Botulinum toxin type A often is used for etched-in crow’s-feet, which rarely are evident in SOC patients.47 Facial shaping also can be accomplished by injecting botulinum toxin type A in combination with soft-tissue dermal fillers.47

Although black individuals do not experience perioral rhytides at the frequency of white individuals, they experience a variety of other cosmetic issues related to skin sagging and sinking. Currently available hyaluronic acid (HA) fillers have been shown to be safe in patients with Fitzpatrick skin types IV through VI.48 Two studies evaluated fillers in patients with SOC, specifically HA49 and calcium hydroxylapatite,50 focused on treatment of the nasolabial folds and the potential risk for dyspigmentation and keloidal scarring. Taylor et al49 noted that the risk of hyperpigmentation was 6% to 9% for large- and small-particle HA, respectively, and was associated with the serial or multiple puncture injection technique. No hypertrophic or keloidal scarring occurred in both studies.49,50

Facial contouring applications with fillers include glabellar lines, temples, nasal bridge, tear troughs, malar and submalar areas, nasolabial folds, radial lines, lips, marionette lines, mental crease, and chin. Hyaluronic acid fillers also can be used for lip enhancement.47 Although white women are looking to increase the size of their lips, black women are seeking augmentation to restore their lip size to that of their youth. Black individuals do not experience the same frequency of perioral rhytides as white patients, but they experience a variety of other issues related to skin sagging and sinking. Unlike white women, enhancement of the vermilion border rarely is performed in black women due to development of rhytides, predominantly in the body of the lip below the vermilion border in response to volume loss in the upper lip while the lower lip usually maintains its same appearance.47

Facial enhancement utilizing poly-L-lactic acid can be used safely in SOC patients.51 Poly-L-lactic acid microparticles induce collagen formation, leading to dermal thickening over 3 to 6 months; however, multiple sessions are required to achieve optimal aesthetic results.

Patients with more reactive collagen can benefit from noninvasive cosmetic procedures such as skin-tightening procedures.52 Radiofrequency and microfocused ultrasound are cosmetic procedures used to provide skin tightening and facial lifting. They are safe and effective treatments for patients with Fitzpatrick skin types IV to VI.53 Histologically, there is less thinning of collagen bundles and elastic tissue in ethnic skin. Due to stimulation of collagen by these procedures, most SOC patients will experience a more enhanced response, requiring fewer treatment sessions than white individuals.

Conclusion

Medical and aesthetic facial concerns in SOC patients vary and can be a source of emotional and psychological distress that can negatively impact quality of life. The approach to the treatment of SOC patients should be a balance between tolerability and efficacy, considering the potential risk for PIH.

References
  1. White GM. Recent findings in the epidemiologic evidence, classification, and subtypes of acne vulgaris. J Am Acad Dermatol. 1998;39(2 pt 3):S34-S37.
  2. Halder RM, Grimes PE, McLaurin CL, et al. Incidence of common dermatoses in a predominantly black dermatologic practice. Cutis. 1983;32:388, 390.
  3. Alexis AF, Sergay AB, Taylor SC. Common dermatologic disorders in skin of color: a comparative practice survey. Cutis. 2007;80:387-394.
  4. Davis SA, Narahari S, Feldman SR, et al. Top dermatologic conditions in patients of color: an analysis of nationally representative data. J Drugs Dermatol. 2012;11:466-473.
  5. Perkins AC, Cheng CE, Hillebrand GG, et al. Comparison of the epidemiology of acne vulgaris among Caucasians, Asian, Continental Indian and African American women. J Eur Acad Dermatol Venereol. 2011;25:1054-1060.
  6. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl):S98-S106.
  7. Davis EC, Callender VD. A review of acne in ethnic skin: pathogenesis, clinical manifestations, and management strategies. J Clin Aesthet Dermatol. 2010;3:24-38.
  8. Halder RM, Holmes YC, Bridgeman-Shah S, et al. A clinicohistologic study of acne vulgaris in black females (abstract). J Invest Dermatol. 1996;106:888.
  9. Plewig G, Fulton JE, Kligman AM. Pomade acne. Arch Dermatol. 1970;101:580-584.
  10. Kligman AM, Mills OH. Acne cosmetica. Arch Dermatol. 1972;106:893-897.
  11. Halder RM, Brooks HL, Callender VD. Acne in ethnic skin. Dermatol Clin. 2003;21:609-615.
  12. Callender VD. Acne in ethnic skin: special considerations for therapy. Dermatol Ther. 2004;17:184-195.
  13. Winhoven SM. Postinflammatory hyperpigmentation in an Asian patient. a dramatic response to oral isotretinoin (13-cis-retinoic acid). Br J Med. 2005;152:368-403.
  14. Sarkar R, Bansal S, Garg VK. Chemical peels for melasma in dark-skinned patients. J Cutan Aesthet Surg. 2012;5:247-253.
  15. Alexis AF, Coley MK, Nijhawan RI, et al. Nonablative fractional laser resurfacing for acne scarring in patients with Fitzpatrick skin phototypes IV-VI. Dermatol Surg. 2016;42:392-402.
  16. Culp B, Scheinfeld N. Rosacea: a review. P T. 2009;34:38-45.
  17. Al-Dabagh A, Davis SA, McMichael AJ, et al. Rosacea in skin of color: not a rare diagnosis. Dermatol Online J. 2014:20. pii:13030/qt1mv9r0ss.
  18. Del Rosso JQ. Advances in understanding and managing rosacea: part 1: connecting the dots between pathophysiological mechanisms and common clinical features of rosacea with emphasis on vascular changes and facial erythema. J Clin Aesthet Dermatol. 2012;5:16-25.
  19. Jackson JM, Knuckles M, Minni JP, et al. The role of brimonidine tartrate gel in the treatment of rosacea. Clin Cosmet Investig Dermatol. 2015;23:529-538.
  20. Patel NU, Shukla S, Zaki J, et al. Oxymetazoline hydrochloride cream for facial erythema associated with rosacea. Expert Rev Clin Pharmacol. 2017;10:104954.
  21. Weinkle AP, Doktor V, Emer J. Update on the management of rosacea. Clin Cosmet Investig Dermatol. 2015;8:159-177.
  22. Grimes PE. Skin and hair cosmetic issues in women of color. Dermatol Clin. 2000;19:659-665.
  23. Taylor A, Pawaskar M, Taylor SL, et al. Prevalence of pigmentary disorders and their impact on quality of life: a prospective cohort study. J Cosmet Dermatol. 2008;7:164-168.
  24. Davis EC, Callender VD. Postinflammatory hyperpigmentation: a review of the epidemiology, clinical features, and treatment options in skin of color. J Clin Aesthet Dermatol. 2010;3:20-31.
  25. Grimes PE. Melasma: etiologic and therapeutic considerations. Arch Dermatol. 1995;131:1453-1457.
  26. Handel AC, Miot LD, Miot HA. Melasma: a clinical and epidemiological review. An Bras Dermatol. 2014;89:771-782.
  27. Callender VD, Reid SD, Obayan O, et al. Diagnostic clues to frontal fibrosing alopecia in patients of African descent. J Clin Aesthet Dermatol. 2016;9:45-51.
  28. Narang T, Sawatkar GU, Kumaran MS, et al. Minocycline for recurrent and/or chronic erythema nodosum leprosum. JAMA Dermatol. 2015;151:1026-1028.
  29. Boyer M, Katta R, Markus R. Diltiazem-induced photodistributed hyperpigmentation. Dermatol Online J. 2003;9:10.
  30. Pandya AG, Guevara IL. Disorders of hyperpigmentation. Dermatol Clin. 2000;18:91-98.
  31. Taylor SC, Torok H, Jones T, et al. Efficacy and safety of a new triple-combination agent for the treatment of facial melasma. Cutis. 2003;72:67-72.
  32. Bulengo-Ransby SM. Topical tretinoin (retinoic acid) therapy for hyperpigmented lesions caused by inflammation of the skin in black patients. N Engl J Med. 1993;328:1438-1443.
  33. Grimes P, Callender V. Tazarotene cream for postinflammatory hyperpigmentation and acne vulgaris in darker skin: a double-blind, randomized, vehicle-controlled study. Cutis. 2006;77:45-50.
  34. Jacyk WK. Adapalene in the treatment of African patients. J Eur Acad Dermatol Venereol. 2001;15(suppl 3):37-42.
  35. Kircik LH. Efficacy and safety of azelaic acid (AzA) gel 15% in the treatment of postinflammatory hyperpigmentation and acne: a 16-week, baseline-controlled study. J Drugs Dermatol. 2011;10:586-590.
  36. Lee HC, Thng TG, Goh CL. Oral tranexamic acid (TA) in the treatment of melasma. J Am Acad Dermatol. 2016;75:385-392.
  37. Kindred C, Okereke U, Callender VD. Skin-lightening agents: an overview of prescription, office-dispensed, and over-the-counter products. Cosmet Dermatol. 2013;26:18-26.
  38. Makino ET, Kadoya K, Sigler ML, et al. Development and clinical assessment of a comprehensive product for pigmentation control in multiple ethnic populations. J Drugs Dermatol. 2016;15:1562-1570.
  39. Halder R, Rodney I, Munhutu M, et al. Evaluation and effectiveness of a photoprotection composition (sunscreen) on subjects of skin of color. J Am Acad Dermatol. 2015;72(suppl 1):AB215.
  40. Castanedo-Cazares JP, Hernandez-Blanco D, Carlos-Ortega B, et al. Near-visible light and UV photoprotection in the treatment of melasma: a double-blind randomized trial. Photodermatol Photoimmunol Photomed. 2014;30:35-42.
  41. American Society for Aesthetic Plastic Surgery. 2016 Cosmetic Surgery National Data Bank Statistics. https://www.surgery.org/sites/default/files/ASAPS-Stats2016.pdf. Accessed November 15, 2017.
  42. Burgess CM. Soft tissue augmentation in skin of color: market growth, available fillers and successful techniques. J Drugs Dermatol. 2007;6:51-55.
  43. Davis EC, Callender VD. Aesthetic dermatology for aging ethnic skin. Dermatol Surg. 2011;37:901-917.
  44. Grimes PE, Arora S, Minus HR, et al. Dermatosis papulosa nigra. Cutis. 1983;32:385-386.
  45. Lupo M. Dermatosis papulosa nigra: treatment options. J Drugs Dermatol. 2007;6:29-30.
  46. Grimes PE, Shabazz D. A four-month randomized, double-blind evaluation of the efficacy of botulinum toxin type A for the treatment of glabellar lines in women with skin types V and VI. Dermatol Surg. 2009;35:429-435.
  47. Burgess CM, Awosika O. Ethnic and gender considerations in the use of facial injectables: African-American patients. Plast Reconstr Surg. 2015;136(5 suppl):28S-31S.
  48. Taylor SC, Kelly AP, Lim HW, et al, eds. Taylor and Kelly’s Dermatology for Skin of Color. 2nd ed. New York, NY: McGraw-Hill Education; 2016.
  49.  Taylor SC, Burgess CM, Callender VD. Safety of nonanimal stabilized hyaluronic acid dermal fillers in patients with skin of color: a randomized, evaluator-blinded comparative trial. Dermatol Surg. 2009;35(suppl 2):1653-1660.
  50. Marmur ES, Taylor SC, Grimes PE, et al. Six-month safety results of calcium hydroxylapatite for treatment of nasolabial folds in Fitzpatrick skin types IV to VI. Dermatol Surg. 2009;35(suppl 2):1641-1645.
  51. Hamilton TK, Burgess CM. Consideration for the use of injectable poly-L-lactic acid in people of color. J Drugs Dermatol. 2010;9:451-456.
  52. Fabi SG, Goldman MP. Retrospective evaluation of micro-focused ultrasound for lifting and tightening of the face and neck. Dermatol Surg. 2014;40:569-575.
  53. Harris MO, Sundaram HA. Safety of microfocused ultrasound with visualization in patients with Fitzpatrick skin phototypes III to VI. JAMA Facial Plast Surg. 2015;17:355-357.
Article PDF
Author and Disclosure Information

Dr. Callender is from Callender Dermatology and Cosmetic Center, Glenn Dale, Maryland. Dr. Barbosa is from Millennium Park Dermatology, Chicago, Illinois. Dr. Burgess is from the Center for Dermatology and Dermatologic Surgery, Washington, DC. Dr. Heath is from Premier Dermatology and Cosmetic Surgery, Newark, Delaware. Dr. McMichael is from the Department of Dermatology, Wake Forest School of Medicine, Winston-Salem, North Carolina. Drs. Ogunleye and Taylor are from the Department of Dermatology, University of Pennsylvania, Philadelphia.

Dr. Callender is a consultant for Allergan; Galderma Laboratories, LP; and Unilever. She also is a researcher for Aclaris Therapeutics, Inc; Allergan; and Revance Therapeutics Inc. Drs. Barbosa, Heath, and Ogunleye report no conflict of interest. Dr. Burgess is a clinical research investigator and stockholder and has received honorarium from Allergan. She also is a clinical research investigator for Aclaris Therapeutics, Inc; Cutanea Life Sciences; Foamix Pharmaceuticals; and Revance Therapeutics Inc, and is a clinical research investigator and speaker and has received honoraria from Merz Pharma. Dr. McMichael is a consultant for Allergan; Galderma Laboratories, LP; Johnson & Johnson; and Procter & Gamble. She also has received research grants from Allergan and Procter & Gamble. Dr. Taylor is an advisory board member for Allergan; Aclaris Therapeutics Inc; Beiersdorf Inc; Galderma Laboratories, LP; and NeoStrata Company, Inc. She also is an investigator for Aclaris Therapeutics Inc and Croma-Pharma.

Correspondence: Valerie D. Callender, MD, 12200 Annapolis Rd, Ste 315, Glenn Dale, MD 20769 (drcallender@callenderskin.com).

Issue
Cutis - 100(6)
Publications
Topics
Page Number
375-380
Sections
Author and Disclosure Information

Dr. Callender is from Callender Dermatology and Cosmetic Center, Glenn Dale, Maryland. Dr. Barbosa is from Millennium Park Dermatology, Chicago, Illinois. Dr. Burgess is from the Center for Dermatology and Dermatologic Surgery, Washington, DC. Dr. Heath is from Premier Dermatology and Cosmetic Surgery, Newark, Delaware. Dr. McMichael is from the Department of Dermatology, Wake Forest School of Medicine, Winston-Salem, North Carolina. Drs. Ogunleye and Taylor are from the Department of Dermatology, University of Pennsylvania, Philadelphia.

Dr. Callender is a consultant for Allergan; Galderma Laboratories, LP; and Unilever. She also is a researcher for Aclaris Therapeutics, Inc; Allergan; and Revance Therapeutics Inc. Drs. Barbosa, Heath, and Ogunleye report no conflict of interest. Dr. Burgess is a clinical research investigator and stockholder and has received honorarium from Allergan. She also is a clinical research investigator for Aclaris Therapeutics, Inc; Cutanea Life Sciences; Foamix Pharmaceuticals; and Revance Therapeutics Inc, and is a clinical research investigator and speaker and has received honoraria from Merz Pharma. Dr. McMichael is a consultant for Allergan; Galderma Laboratories, LP; Johnson & Johnson; and Procter & Gamble. She also has received research grants from Allergan and Procter & Gamble. Dr. Taylor is an advisory board member for Allergan; Aclaris Therapeutics Inc; Beiersdorf Inc; Galderma Laboratories, LP; and NeoStrata Company, Inc. She also is an investigator for Aclaris Therapeutics Inc and Croma-Pharma.

Correspondence: Valerie D. Callender, MD, 12200 Annapolis Rd, Ste 315, Glenn Dale, MD 20769 (drcallender@callenderskin.com).

Author and Disclosure Information

Dr. Callender is from Callender Dermatology and Cosmetic Center, Glenn Dale, Maryland. Dr. Barbosa is from Millennium Park Dermatology, Chicago, Illinois. Dr. Burgess is from the Center for Dermatology and Dermatologic Surgery, Washington, DC. Dr. Heath is from Premier Dermatology and Cosmetic Surgery, Newark, Delaware. Dr. McMichael is from the Department of Dermatology, Wake Forest School of Medicine, Winston-Salem, North Carolina. Drs. Ogunleye and Taylor are from the Department of Dermatology, University of Pennsylvania, Philadelphia.

Dr. Callender is a consultant for Allergan; Galderma Laboratories, LP; and Unilever. She also is a researcher for Aclaris Therapeutics, Inc; Allergan; and Revance Therapeutics Inc. Drs. Barbosa, Heath, and Ogunleye report no conflict of interest. Dr. Burgess is a clinical research investigator and stockholder and has received honorarium from Allergan. She also is a clinical research investigator for Aclaris Therapeutics, Inc; Cutanea Life Sciences; Foamix Pharmaceuticals; and Revance Therapeutics Inc, and is a clinical research investigator and speaker and has received honoraria from Merz Pharma. Dr. McMichael is a consultant for Allergan; Galderma Laboratories, LP; Johnson & Johnson; and Procter & Gamble. She also has received research grants from Allergan and Procter & Gamble. Dr. Taylor is an advisory board member for Allergan; Aclaris Therapeutics Inc; Beiersdorf Inc; Galderma Laboratories, LP; and NeoStrata Company, Inc. She also is an investigator for Aclaris Therapeutics Inc and Croma-Pharma.

Correspondence: Valerie D. Callender, MD, 12200 Annapolis Rd, Ste 315, Glenn Dale, MD 20769 (drcallender@callenderskin.com).

Article PDF
Article PDF
Related Articles
In Collaboration With the Skin of Color Society
In Collaboration With the Skin of Color Society

The approach to the treatment of common skin disorders and cosmetic concerns in patients with skin of color (SOC) requires the clinician to understand the biological differences, nuances, and special considerations that are unique to patients with darker skin types.1-3 This article addresses 4 common facial concerns in SOC patients—acne, rosacea, facial hyperpigmentation, and cosmetic enhancement—and provides treatment recommendations and management pearls to assist the clinician with optimal outcomes for SOC patients.

Acne in SOC Patients

Acne vulgaris is one of the most common conditions that dermatologists treat and is estimated to affect 40 to 50 million individuals in the United States.1 Many of these acne patients are individuals with SOC.2-4 A study of 2835 females (aged 10–70 years) conducted in 4 different cities—Los Angeles, California; London, United Kingdom; Akita, Japan; and Rome, Italy—demonstrated acne prevalence of 37% in blacks, 32% in Hispanics, 30% in Asians, 24% in whites, and 23% in Continental Indians.5 Blacks, Hispanics, and Continental Indians demonstrated equal prevalence with comedonal and inflammatory acne. Asians displayed more inflammatory acne lesions than comedones. In contrast, whites demonstrated more comedones than inflammatory acne. Dyspigmentation, postinflammatory hyperpigmentation (PIH), and atrophic scars were more common in black and Hispanic females than other ethnicities.5 This study illustrated that acne-induced PIH is a common sequela in SOC patients and is the main reason they seek treatment.6,7

The pathogenesis of acne is the same in all racial and ethnic groups: (1) follicular hyperkeratinization and the formation of a microcomedone caused by abnormal desquamation of the keratinocytes within the sebaceous follicle, (2) production of sebum by circulating androgens, (3) proliferation of Propionibacterium acnes, and (4) inflammation. Subclinical inflammation is present throughout all stages of acne, including normal-appearing skin, inflammatory lesions, comedones, and scarring, and may contribute to PIH in acne patients with SOC (Figure 1).8 A thorough history should be obtained from acne patients, including answers to the following questions7:

  • What skin and hair care products do you use?
  • Do you use sunscreen daily?
  • What cosmetic products or makeup do you use?
  • Do you use any ethnic skin care products, including skin lightening creams?
  • Do you have a history of keloids?

Figure 1. Acne and postinflammatory hyperpigmentation in a patient with skin of color (Fitzpatrick skin type V).

It is important to ask these questions to assess if the SOC patient has developed pomade acne,9 acne cosmetica,10 or a potential risk of skin irritation from the use of skin care practices. It is best to take total control of the patient’s skin care regimen and discontinue use of toners, astringents, witch hazel, exfoliants, and rubbing alcohol, which may lead to skin dryness and irritation, particularly when combined with topical acne medications.

Treatment
Treatment of acne in SOC patients is similar to generally recommended treatments, with special considerations. Consider the following key points when treating acne in SOC patients:

  • Treat acne early and aggressively to prevent or minimize subsequent PIH and acne scarring.
  • Balance aggressive treatment with nonirritating topical skin care.
  • Most importantly, target PIH in addition to acne and choose a regimen that limits skin irritation that might exacerbate existing PIH.7

Develop a maintenance program to control future breakouts. Topical agents can be used as monotherapy or in fixed combinations and may include benzoyl peroxide, antibiotics, dapsone, azelaic acid (AZA), and retinoids. Similar to white patients, topical retinoids remain a first-line treatment for acne in patients with SOC.11,12

Tolerability must be managed in SOC acne patients. Therapeutic maneuvers that can be instituted should include a discussion on using gentle skin care, initiating therapy with a retinoid applied every other night starting with a low concentration and gradually titrating up, and applying a moisturizer before or after applying acne medication. Oral therapies consist of antibiotics (doxycycline, minocycline), retinoids (isotretinoin), and hormonal modulators (oral contraceptives, spironolactone). Isotretinoin, recommended for patients with nodulocystic acne, may play a possible role in treating acne-induced PIH.13

Two common procedural therapies for acne include comedone extraction and intralesional corticosteroid injection. A 6- to 8-week course of a topical retinoid prior to comedonal extraction may facilitate the procedure and is recommended in SOC patients to help reduce cutaneous trauma and PIH.11 Inflammatory acne lesions can be treated with intralesional injection of triamcinolone acetonide 2.5 or 5.0 mg/mL, which usually reduces inflammation within 2 to 5 days.11

Treatment of acne-induced PIH includes sun protection, topical and oral medications, chemical peels, lasers, and energy devices. Treatment of hypertrophic scarring and keloids involves intralesional injection of triamcinolone acetonide 20, 30, or 40 mg/mL every 4 weeks until the lesion is flat.11

Superficial chemical peels can be used to treat acne and PIH in SOC patients,14 such as salicylic acid (20%–30%), glycolic acid (20%–70%), trichloroacetic acid (15%–30%), and Jessner peels.

Acne Scarring
Surgical approaches to acne scarring in patients with SOC include elliptical excision, punch excision, punch elevation, punch autografting, dermal grafting, dermal planning, subcutaneous incision (subcision), dermabrasion, microneedling, fillers, and laser skin resurfacing. The treatment of choice depends on the size, type, and depth of the scar and the clinician’s preference.

Lasers
Fractional photothermolysis has emerged as a treatment option for acne scars in SOC patients. This procedure produces microscopic columns of thermal injury in the epidermis and dermis, sparing the surrounding tissue and minimizing downtime and adverse events. Because fractional photothermolysis does not target melanin and produces limited epidermal injury, darker Fitzpatrick skin types (IV–VI) can be safely and effectively treated with this procedure.15

 

 

Rosacea in SOC Patients

Rosacea is a chronic inflammatory disorder that affects the vasculature and pilosebaceous units of the face. It commonly is seen in Fitzpatrick skin types I and II; however, rosacea can occur in all skin types (Figure 2). Triggers include emotional stress, extreme environmental temperatures, hot and spicy foods, red wine or alcohol, and topical irritants or allergens found in common cosmetic products.16

Figure 2. Rosacea in a patient with skin of color (Fitzpatrick skin type IV).

Data suggest that 4% of rosacea patients in the United States are of African, Latino, or Asian descent.11 National Ambulatory Medical Care Survey data revealed that of 31.5 million rosacea visits, 2% of patients were black, 2.3% were Asian or Pacific Islander, and 3.9% were Hispanic or Latino. In a 5-year longitudinal study of 2587 rosacea patients enrolled in Medicaid in North Carolina who were prescribed at least 1 topical treatment for rosacea, 16.27% were black and 10% were of a race other than white.17

Although the pathogenesis of rosacea is unclear, hypotheses include immune system abnormalities, neurogenic dysregulation, presence of microorganisms (eg, Demodex folliculorum), UV damage, and skin barrier dysfunction.18

The 4 major subtypes of rosacea are erythematotelangiectatic, papulopustular, phymatous, and ocular rosacea.16 Interestingly, rosacea in SOC patients may present with hypopigmentation surrounding the borders of the facial erythema. For phymatous rosacea, isotretinoin may reduce incipient rhinophyma but must be carefully monitored and pregnancy must be excluded. Surgical or laser therapy may be indicated to recontour the nose if severe.

There are several skin conditions that can present with facial erythema in patients with SOC, including seborrheic dermatitis, systemic lupus erythematosus, and contact dermatitis. It is important to note that the detection of facial erythema in darker skin types may be difficult; therefore, laboratory evaluation (antinuclear antibodies), patch testing, and skin biopsy should be considered if the clinical diagnosis is unclear.

Treatment
Treatment of rosacea in SOC patients does not differ from other racial groups. Common strategies include gentle skin care, sun protection (sun protection factor 30+), and barrier repair creams. Topical agents include metronidazole, AZA, sodium sulfacetamide/sulfur, ivermectin, and retinoids.16 Oral treatments include antibiotics in the tetracycline family (eg, subantimicrobial dose doxycycline) and isotretinoin.16 Persistent erythema associated with rosacea can be treated with brimonidine19 and oxymetazoline.20 Vascular lasers and intense pulsed light may be used to address the vascular components of rosacea21; however, the latter is not recommended in Fitzpatrick skin types IV through VI.

Facial Hyperpigmentation in SOC Patients

Hyperpigmentation disorders can be divided into conditions that affect Fitzpatrick skin types I through III and IV though VI. Mottled hyperpigmentation (photodamage) and solar lentigines occur in patients with lighter skin types as compared to melasma, PIH, and age-related (UV-induced) hyperpigmentation, which occur more commonly in patients with darker skin types. Facial hyperpigmentation is a common concern in SOC patients. In a survey of cosmetic concerns of 100 women with SOC, hyperpigmentation or dark spots (86%) and blotchy uneven skin (80%) were the top concerns.22 In addition, facial hyperpigmentation has been shown to negatively impact quality of life.23

Postinflammatory hyperpigmentation occurs from a pathophysiological response to inflammation, cutaneous irritation or injury, and subsequent melanocyte lability. Postinflammatory hyperpigmentation is a common presenting concern in patients with SOC and is seen as a result of many inflammatory skin disorders (eg, acne, eczema) and dermatologic procedures (eg, adverse reaction to electrodesiccation, microdermabrasion, chemical peels, laser surgery).24

Melasma is an acquired idiopathic disorder of hyperpigmentation and often referred to as the mask of pregnancy (Figure 3). It occurs on sun-exposed areas of skin, mainly in women with Fitzpatrick skin types III through V. Associated factors or triggers include pregnancy, hormonal treatments, exposure to UV radiation, and medications.25 Hereditary factors play a role in more than 40% of cases.26

Figure 3. Facial hyperpigmentation consistent with melasma in a patient with skin of color (Fitzpatrick skin type IV).

Other not-so-common facial dyschromias include contact dermatitis, acanthosis nigricans, exogenous ochronosis, lichen planus pigmentosus (associated with frontal fibrosing alopecia),27 drug-induced hyperpigmentation (associated with minocycline or diltiazem),28,29 and UV-induced (age-related) hyperpigmentation.

Treatment
The treatment of hyperpigmentation should provide the following: (1) protection from sun exposure; (2) inhibition of tyrosinase, the enzyme responsible for the conversion of tyrosine to melanin; (3) inhibition of melanosome transfer from the melanocyte to the keratinocyte; (4) removal of melanin from the epidermis through exfoliation; and (5) destruction or disruption of melanin in the dermis.30 Therapies for facial hyperpigmentation are listed in Table 1.

Topical therapies include prescription medications and nonprescription cosmeceuticals. Prescription medications include hydroquinone (HQ), topical retinoids, and AZA. Hydroquinone, a tyrosinase inhibitor, is the gold standard for skin lightening and often is used as a first-line therapy. It is used as a monotherapy (HQ 4%) or as a fixed combination with tretinoin 0.05% and fluocinolone 0.01%.31 Use caution with HQ in high concentrations (6% and higher) and low concentrations (2% [over-the-counter strength]) used long-term due to the potential risk of exogenous ochronosis.

Topical retinoids have been shown to be effective therapeutic agents for melasma and PIH. Tretinoin,32 tazarotene,33 and adapalene34 all have demonstrated efficacy for acne and acne-induced PIH in SOC patients. Patients must be monitored for the development of retinoid dermatitis and worsening of hyperpigmentation.

Azelaic acid is a naturally occurring dicarboxylic acid obtained from cultures of Malassezia furfur. Azelaic acid inhibits tyrosinase activity, DNA synthesis, and mitochondrial enzymes, thus blocking direct cytotoxic effects toward melanocytes. Azelaic acid is approved by the US Food and Drug Administration for acne in a 20% cream formulation and rosacea in 15% gel and foam formulations, and it is used off label for melasma and PIH.35

Oral tranexamic acid is currently used as a hemostatic agent due to its ability to inhibit the plasminogen-plasmin pathway. In melasma, it blocks the interaction between melanocytes and keratinocytes in the epidermis and modulates the vascular component of melasma in the dermis. In an open-label study, 561 Asian melasma patients were treated with oral tranexamic acid 250 mg twice daily for 4 months. Results demonstrated improvement in 90% of patients, and 7.1% reported adverse effects (eg, abdominal bloating and pain, nausea, vomiting, headache, tinnitus, numbness, menstrual irregularities).36 Coagulation screening should be monitored monthly, and any patient with a history of clotting abnormalities should be excluded from off-label treatment with oral tranexamic acid.

Nonprescription cosmeceuticals are available over-the-counter or are office dispensed.37 For optimal results, cosmeceutical agents for skin lightening are used in combination. Most of these combinations are HQ free and have additive benefits such as a multimodal skin lightening agent containing key ingredients that correct and prevent skin pigmentation via several pathways affecting melanogenesis.38 It is an excellent alternative to HQ for mottled and diffuse UV-induced hyperpigmentation and can be used for maintenance therapy in patients with melasma.

Photoprotection is an essential component of therapy for melasma and PIH, but there is a paucity of data on the benefits for SOC patients. Halder et al39 performed a randomized prospective study of 89 black and Hispanic patients who applied sunscreen with a sun protection factor of 30 or 60 daily for 8 weeks. Clinical grading, triplicate L*A*B chromameter, and clinical photography were taken at baseline and weeks 4 and 8. The results demonstrated skin lightening in both black and Hispanic patients and support the use of sunscreen in the prevention and management of dyschromia in SOC patients.39 Visible light also may play a role in melasma development, and thus use of sunscreens or makeup containing iron oxides are recommended.40

Procedural treatments for facial hyperpigmentation include microdermabrasion, chemical peels, lasers, energy-based devices, and microneedling. There are many types and formulations of chemical peeling agents available; however, superficial and medium-depth chemical peels are recommended for SOC patients (Table 2). Deep chemical peels are not recommended for SOC patients due to the potential increased risk for PIH and scarring.

 

 

Cosmetic Enhancement in SOC Patients

Cosmetic procedures are gaining popularity in the SOC population and account for more than 20% of cosmetic procedures in the United States.41 Facial cosmetic concerns in SOC include dyschromia, benign growths (dermatosis papulosa nigra), hyperkinetic facial lines, volume loss, and skin laxity.42 Key principles to consider when treating SOC patients are the impact of ethnicity on aging and facial structure, the patient’s desired cosmetic outcome, tissue reaction to anticipated treatments, and the patient’s expectations for recommended therapies.

Aging in SOC Patients
Skin aging can be classified as intrinsic aging or extrinsic aging. Intrinsic aging is genetic and involves subsurface changes such as volume loss, muscle atrophy, and resorption of bony structure. Extrinsic aging (or photoaging) involves surface changes of the epidermis/dermis and manifests as mottled pigmentation, textural changes, and fine wrinkling. Due to the photoprotection of melanin (black skin=SPF 13.4), skin aging in SOC patients is delayed by 10 to 20 years.43 In addition, SOC patients have more reactive collagen and can benefit from noninvasive cosmetic procedures such as fillers and skin-tightening procedures.42

Cosmetic Treatments and Procedures
Dermatosis papulosa nigra (benign growths of skin that have a genetic predisposition)44 occur mainly on the face but can involve the entire body. Treatment modalities include electrodesiccation, cryotherapy, scissor excision, and laser surgery.45

Treatment of hyperkinetic facial lines with botulinum toxin type A is a safe and effective procedure in patients with SOC. Grimes and Shabazz46 performed a 4-month, randomized, double-blind study that evaluated the treatment of glabellar lines in women with Fitzpatrick skin types V and VI. The results demonstrated that the duration of effects was the same in the patients who received either 20 or 30 U of botulinum toxin type A.46 Dynamic rhytides (furrows and frown/scowl lines arising from laughing, frowning, or smiling) can be treated safely in patients with SOC using botulinum toxin type A off label for relaxation of the upper and lower hyperkinetic muscles that result in these unwanted signs of aging. Botulinum toxin type A often is used for etched-in crow’s-feet, which rarely are evident in SOC patients.47 Facial shaping also can be accomplished by injecting botulinum toxin type A in combination with soft-tissue dermal fillers.47

Although black individuals do not experience perioral rhytides at the frequency of white individuals, they experience a variety of other cosmetic issues related to skin sagging and sinking. Currently available hyaluronic acid (HA) fillers have been shown to be safe in patients with Fitzpatrick skin types IV through VI.48 Two studies evaluated fillers in patients with SOC, specifically HA49 and calcium hydroxylapatite,50 focused on treatment of the nasolabial folds and the potential risk for dyspigmentation and keloidal scarring. Taylor et al49 noted that the risk of hyperpigmentation was 6% to 9% for large- and small-particle HA, respectively, and was associated with the serial or multiple puncture injection technique. No hypertrophic or keloidal scarring occurred in both studies.49,50

Facial contouring applications with fillers include glabellar lines, temples, nasal bridge, tear troughs, malar and submalar areas, nasolabial folds, radial lines, lips, marionette lines, mental crease, and chin. Hyaluronic acid fillers also can be used for lip enhancement.47 Although white women are looking to increase the size of their lips, black women are seeking augmentation to restore their lip size to that of their youth. Black individuals do not experience the same frequency of perioral rhytides as white patients, but they experience a variety of other issues related to skin sagging and sinking. Unlike white women, enhancement of the vermilion border rarely is performed in black women due to development of rhytides, predominantly in the body of the lip below the vermilion border in response to volume loss in the upper lip while the lower lip usually maintains its same appearance.47

Facial enhancement utilizing poly-L-lactic acid can be used safely in SOC patients.51 Poly-L-lactic acid microparticles induce collagen formation, leading to dermal thickening over 3 to 6 months; however, multiple sessions are required to achieve optimal aesthetic results.

Patients with more reactive collagen can benefit from noninvasive cosmetic procedures such as skin-tightening procedures.52 Radiofrequency and microfocused ultrasound are cosmetic procedures used to provide skin tightening and facial lifting. They are safe and effective treatments for patients with Fitzpatrick skin types IV to VI.53 Histologically, there is less thinning of collagen bundles and elastic tissue in ethnic skin. Due to stimulation of collagen by these procedures, most SOC patients will experience a more enhanced response, requiring fewer treatment sessions than white individuals.

Conclusion

Medical and aesthetic facial concerns in SOC patients vary and can be a source of emotional and psychological distress that can negatively impact quality of life. The approach to the treatment of SOC patients should be a balance between tolerability and efficacy, considering the potential risk for PIH.

The approach to the treatment of common skin disorders and cosmetic concerns in patients with skin of color (SOC) requires the clinician to understand the biological differences, nuances, and special considerations that are unique to patients with darker skin types.1-3 This article addresses 4 common facial concerns in SOC patients—acne, rosacea, facial hyperpigmentation, and cosmetic enhancement—and provides treatment recommendations and management pearls to assist the clinician with optimal outcomes for SOC patients.

Acne in SOC Patients

Acne vulgaris is one of the most common conditions that dermatologists treat and is estimated to affect 40 to 50 million individuals in the United States.1 Many of these acne patients are individuals with SOC.2-4 A study of 2835 females (aged 10–70 years) conducted in 4 different cities—Los Angeles, California; London, United Kingdom; Akita, Japan; and Rome, Italy—demonstrated acne prevalence of 37% in blacks, 32% in Hispanics, 30% in Asians, 24% in whites, and 23% in Continental Indians.5 Blacks, Hispanics, and Continental Indians demonstrated equal prevalence with comedonal and inflammatory acne. Asians displayed more inflammatory acne lesions than comedones. In contrast, whites demonstrated more comedones than inflammatory acne. Dyspigmentation, postinflammatory hyperpigmentation (PIH), and atrophic scars were more common in black and Hispanic females than other ethnicities.5 This study illustrated that acne-induced PIH is a common sequela in SOC patients and is the main reason they seek treatment.6,7

The pathogenesis of acne is the same in all racial and ethnic groups: (1) follicular hyperkeratinization and the formation of a microcomedone caused by abnormal desquamation of the keratinocytes within the sebaceous follicle, (2) production of sebum by circulating androgens, (3) proliferation of Propionibacterium acnes, and (4) inflammation. Subclinical inflammation is present throughout all stages of acne, including normal-appearing skin, inflammatory lesions, comedones, and scarring, and may contribute to PIH in acne patients with SOC (Figure 1).8 A thorough history should be obtained from acne patients, including answers to the following questions7:

  • What skin and hair care products do you use?
  • Do you use sunscreen daily?
  • What cosmetic products or makeup do you use?
  • Do you use any ethnic skin care products, including skin lightening creams?
  • Do you have a history of keloids?

Figure 1. Acne and postinflammatory hyperpigmentation in a patient with skin of color (Fitzpatrick skin type V).

It is important to ask these questions to assess if the SOC patient has developed pomade acne,9 acne cosmetica,10 or a potential risk of skin irritation from the use of skin care practices. It is best to take total control of the patient’s skin care regimen and discontinue use of toners, astringents, witch hazel, exfoliants, and rubbing alcohol, which may lead to skin dryness and irritation, particularly when combined with topical acne medications.

Treatment
Treatment of acne in SOC patients is similar to generally recommended treatments, with special considerations. Consider the following key points when treating acne in SOC patients:

  • Treat acne early and aggressively to prevent or minimize subsequent PIH and acne scarring.
  • Balance aggressive treatment with nonirritating topical skin care.
  • Most importantly, target PIH in addition to acne and choose a regimen that limits skin irritation that might exacerbate existing PIH.7

Develop a maintenance program to control future breakouts. Topical agents can be used as monotherapy or in fixed combinations and may include benzoyl peroxide, antibiotics, dapsone, azelaic acid (AZA), and retinoids. Similar to white patients, topical retinoids remain a first-line treatment for acne in patients with SOC.11,12

Tolerability must be managed in SOC acne patients. Therapeutic maneuvers that can be instituted should include a discussion on using gentle skin care, initiating therapy with a retinoid applied every other night starting with a low concentration and gradually titrating up, and applying a moisturizer before or after applying acne medication. Oral therapies consist of antibiotics (doxycycline, minocycline), retinoids (isotretinoin), and hormonal modulators (oral contraceptives, spironolactone). Isotretinoin, recommended for patients with nodulocystic acne, may play a possible role in treating acne-induced PIH.13

Two common procedural therapies for acne include comedone extraction and intralesional corticosteroid injection. A 6- to 8-week course of a topical retinoid prior to comedonal extraction may facilitate the procedure and is recommended in SOC patients to help reduce cutaneous trauma and PIH.11 Inflammatory acne lesions can be treated with intralesional injection of triamcinolone acetonide 2.5 or 5.0 mg/mL, which usually reduces inflammation within 2 to 5 days.11

Treatment of acne-induced PIH includes sun protection, topical and oral medications, chemical peels, lasers, and energy devices. Treatment of hypertrophic scarring and keloids involves intralesional injection of triamcinolone acetonide 20, 30, or 40 mg/mL every 4 weeks until the lesion is flat.11

Superficial chemical peels can be used to treat acne and PIH in SOC patients,14 such as salicylic acid (20%–30%), glycolic acid (20%–70%), trichloroacetic acid (15%–30%), and Jessner peels.

Acne Scarring
Surgical approaches to acne scarring in patients with SOC include elliptical excision, punch excision, punch elevation, punch autografting, dermal grafting, dermal planning, subcutaneous incision (subcision), dermabrasion, microneedling, fillers, and laser skin resurfacing. The treatment of choice depends on the size, type, and depth of the scar and the clinician’s preference.

Lasers
Fractional photothermolysis has emerged as a treatment option for acne scars in SOC patients. This procedure produces microscopic columns of thermal injury in the epidermis and dermis, sparing the surrounding tissue and minimizing downtime and adverse events. Because fractional photothermolysis does not target melanin and produces limited epidermal injury, darker Fitzpatrick skin types (IV–VI) can be safely and effectively treated with this procedure.15

 

 

Rosacea in SOC Patients

Rosacea is a chronic inflammatory disorder that affects the vasculature and pilosebaceous units of the face. It commonly is seen in Fitzpatrick skin types I and II; however, rosacea can occur in all skin types (Figure 2). Triggers include emotional stress, extreme environmental temperatures, hot and spicy foods, red wine or alcohol, and topical irritants or allergens found in common cosmetic products.16

Figure 2. Rosacea in a patient with skin of color (Fitzpatrick skin type IV).

Data suggest that 4% of rosacea patients in the United States are of African, Latino, or Asian descent.11 National Ambulatory Medical Care Survey data revealed that of 31.5 million rosacea visits, 2% of patients were black, 2.3% were Asian or Pacific Islander, and 3.9% were Hispanic or Latino. In a 5-year longitudinal study of 2587 rosacea patients enrolled in Medicaid in North Carolina who were prescribed at least 1 topical treatment for rosacea, 16.27% were black and 10% were of a race other than white.17

Although the pathogenesis of rosacea is unclear, hypotheses include immune system abnormalities, neurogenic dysregulation, presence of microorganisms (eg, Demodex folliculorum), UV damage, and skin barrier dysfunction.18

The 4 major subtypes of rosacea are erythematotelangiectatic, papulopustular, phymatous, and ocular rosacea.16 Interestingly, rosacea in SOC patients may present with hypopigmentation surrounding the borders of the facial erythema. For phymatous rosacea, isotretinoin may reduce incipient rhinophyma but must be carefully monitored and pregnancy must be excluded. Surgical or laser therapy may be indicated to recontour the nose if severe.

There are several skin conditions that can present with facial erythema in patients with SOC, including seborrheic dermatitis, systemic lupus erythematosus, and contact dermatitis. It is important to note that the detection of facial erythema in darker skin types may be difficult; therefore, laboratory evaluation (antinuclear antibodies), patch testing, and skin biopsy should be considered if the clinical diagnosis is unclear.

Treatment
Treatment of rosacea in SOC patients does not differ from other racial groups. Common strategies include gentle skin care, sun protection (sun protection factor 30+), and barrier repair creams. Topical agents include metronidazole, AZA, sodium sulfacetamide/sulfur, ivermectin, and retinoids.16 Oral treatments include antibiotics in the tetracycline family (eg, subantimicrobial dose doxycycline) and isotretinoin.16 Persistent erythema associated with rosacea can be treated with brimonidine19 and oxymetazoline.20 Vascular lasers and intense pulsed light may be used to address the vascular components of rosacea21; however, the latter is not recommended in Fitzpatrick skin types IV through VI.

Facial Hyperpigmentation in SOC Patients

Hyperpigmentation disorders can be divided into conditions that affect Fitzpatrick skin types I through III and IV though VI. Mottled hyperpigmentation (photodamage) and solar lentigines occur in patients with lighter skin types as compared to melasma, PIH, and age-related (UV-induced) hyperpigmentation, which occur more commonly in patients with darker skin types. Facial hyperpigmentation is a common concern in SOC patients. In a survey of cosmetic concerns of 100 women with SOC, hyperpigmentation or dark spots (86%) and blotchy uneven skin (80%) were the top concerns.22 In addition, facial hyperpigmentation has been shown to negatively impact quality of life.23

Postinflammatory hyperpigmentation occurs from a pathophysiological response to inflammation, cutaneous irritation or injury, and subsequent melanocyte lability. Postinflammatory hyperpigmentation is a common presenting concern in patients with SOC and is seen as a result of many inflammatory skin disorders (eg, acne, eczema) and dermatologic procedures (eg, adverse reaction to electrodesiccation, microdermabrasion, chemical peels, laser surgery).24

Melasma is an acquired idiopathic disorder of hyperpigmentation and often referred to as the mask of pregnancy (Figure 3). It occurs on sun-exposed areas of skin, mainly in women with Fitzpatrick skin types III through V. Associated factors or triggers include pregnancy, hormonal treatments, exposure to UV radiation, and medications.25 Hereditary factors play a role in more than 40% of cases.26

Figure 3. Facial hyperpigmentation consistent with melasma in a patient with skin of color (Fitzpatrick skin type IV).

Other not-so-common facial dyschromias include contact dermatitis, acanthosis nigricans, exogenous ochronosis, lichen planus pigmentosus (associated with frontal fibrosing alopecia),27 drug-induced hyperpigmentation (associated with minocycline or diltiazem),28,29 and UV-induced (age-related) hyperpigmentation.

Treatment
The treatment of hyperpigmentation should provide the following: (1) protection from sun exposure; (2) inhibition of tyrosinase, the enzyme responsible for the conversion of tyrosine to melanin; (3) inhibition of melanosome transfer from the melanocyte to the keratinocyte; (4) removal of melanin from the epidermis through exfoliation; and (5) destruction or disruption of melanin in the dermis.30 Therapies for facial hyperpigmentation are listed in Table 1.

Topical therapies include prescription medications and nonprescription cosmeceuticals. Prescription medications include hydroquinone (HQ), topical retinoids, and AZA. Hydroquinone, a tyrosinase inhibitor, is the gold standard for skin lightening and often is used as a first-line therapy. It is used as a monotherapy (HQ 4%) or as a fixed combination with tretinoin 0.05% and fluocinolone 0.01%.31 Use caution with HQ in high concentrations (6% and higher) and low concentrations (2% [over-the-counter strength]) used long-term due to the potential risk of exogenous ochronosis.

Topical retinoids have been shown to be effective therapeutic agents for melasma and PIH. Tretinoin,32 tazarotene,33 and adapalene34 all have demonstrated efficacy for acne and acne-induced PIH in SOC patients. Patients must be monitored for the development of retinoid dermatitis and worsening of hyperpigmentation.

Azelaic acid is a naturally occurring dicarboxylic acid obtained from cultures of Malassezia furfur. Azelaic acid inhibits tyrosinase activity, DNA synthesis, and mitochondrial enzymes, thus blocking direct cytotoxic effects toward melanocytes. Azelaic acid is approved by the US Food and Drug Administration for acne in a 20% cream formulation and rosacea in 15% gel and foam formulations, and it is used off label for melasma and PIH.35

Oral tranexamic acid is currently used as a hemostatic agent due to its ability to inhibit the plasminogen-plasmin pathway. In melasma, it blocks the interaction between melanocytes and keratinocytes in the epidermis and modulates the vascular component of melasma in the dermis. In an open-label study, 561 Asian melasma patients were treated with oral tranexamic acid 250 mg twice daily for 4 months. Results demonstrated improvement in 90% of patients, and 7.1% reported adverse effects (eg, abdominal bloating and pain, nausea, vomiting, headache, tinnitus, numbness, menstrual irregularities).36 Coagulation screening should be monitored monthly, and any patient with a history of clotting abnormalities should be excluded from off-label treatment with oral tranexamic acid.

Nonprescription cosmeceuticals are available over-the-counter or are office dispensed.37 For optimal results, cosmeceutical agents for skin lightening are used in combination. Most of these combinations are HQ free and have additive benefits such as a multimodal skin lightening agent containing key ingredients that correct and prevent skin pigmentation via several pathways affecting melanogenesis.38 It is an excellent alternative to HQ for mottled and diffuse UV-induced hyperpigmentation and can be used for maintenance therapy in patients with melasma.

Photoprotection is an essential component of therapy for melasma and PIH, but there is a paucity of data on the benefits for SOC patients. Halder et al39 performed a randomized prospective study of 89 black and Hispanic patients who applied sunscreen with a sun protection factor of 30 or 60 daily for 8 weeks. Clinical grading, triplicate L*A*B chromameter, and clinical photography were taken at baseline and weeks 4 and 8. The results demonstrated skin lightening in both black and Hispanic patients and support the use of sunscreen in the prevention and management of dyschromia in SOC patients.39 Visible light also may play a role in melasma development, and thus use of sunscreens or makeup containing iron oxides are recommended.40

Procedural treatments for facial hyperpigmentation include microdermabrasion, chemical peels, lasers, energy-based devices, and microneedling. There are many types and formulations of chemical peeling agents available; however, superficial and medium-depth chemical peels are recommended for SOC patients (Table 2). Deep chemical peels are not recommended for SOC patients due to the potential increased risk for PIH and scarring.

 

 

Cosmetic Enhancement in SOC Patients

Cosmetic procedures are gaining popularity in the SOC population and account for more than 20% of cosmetic procedures in the United States.41 Facial cosmetic concerns in SOC include dyschromia, benign growths (dermatosis papulosa nigra), hyperkinetic facial lines, volume loss, and skin laxity.42 Key principles to consider when treating SOC patients are the impact of ethnicity on aging and facial structure, the patient’s desired cosmetic outcome, tissue reaction to anticipated treatments, and the patient’s expectations for recommended therapies.

Aging in SOC Patients
Skin aging can be classified as intrinsic aging or extrinsic aging. Intrinsic aging is genetic and involves subsurface changes such as volume loss, muscle atrophy, and resorption of bony structure. Extrinsic aging (or photoaging) involves surface changes of the epidermis/dermis and manifests as mottled pigmentation, textural changes, and fine wrinkling. Due to the photoprotection of melanin (black skin=SPF 13.4), skin aging in SOC patients is delayed by 10 to 20 years.43 In addition, SOC patients have more reactive collagen and can benefit from noninvasive cosmetic procedures such as fillers and skin-tightening procedures.42

Cosmetic Treatments and Procedures
Dermatosis papulosa nigra (benign growths of skin that have a genetic predisposition)44 occur mainly on the face but can involve the entire body. Treatment modalities include electrodesiccation, cryotherapy, scissor excision, and laser surgery.45

Treatment of hyperkinetic facial lines with botulinum toxin type A is a safe and effective procedure in patients with SOC. Grimes and Shabazz46 performed a 4-month, randomized, double-blind study that evaluated the treatment of glabellar lines in women with Fitzpatrick skin types V and VI. The results demonstrated that the duration of effects was the same in the patients who received either 20 or 30 U of botulinum toxin type A.46 Dynamic rhytides (furrows and frown/scowl lines arising from laughing, frowning, or smiling) can be treated safely in patients with SOC using botulinum toxin type A off label for relaxation of the upper and lower hyperkinetic muscles that result in these unwanted signs of aging. Botulinum toxin type A often is used for etched-in crow’s-feet, which rarely are evident in SOC patients.47 Facial shaping also can be accomplished by injecting botulinum toxin type A in combination with soft-tissue dermal fillers.47

Although black individuals do not experience perioral rhytides at the frequency of white individuals, they experience a variety of other cosmetic issues related to skin sagging and sinking. Currently available hyaluronic acid (HA) fillers have been shown to be safe in patients with Fitzpatrick skin types IV through VI.48 Two studies evaluated fillers in patients with SOC, specifically HA49 and calcium hydroxylapatite,50 focused on treatment of the nasolabial folds and the potential risk for dyspigmentation and keloidal scarring. Taylor et al49 noted that the risk of hyperpigmentation was 6% to 9% for large- and small-particle HA, respectively, and was associated with the serial or multiple puncture injection technique. No hypertrophic or keloidal scarring occurred in both studies.49,50

Facial contouring applications with fillers include glabellar lines, temples, nasal bridge, tear troughs, malar and submalar areas, nasolabial folds, radial lines, lips, marionette lines, mental crease, and chin. Hyaluronic acid fillers also can be used for lip enhancement.47 Although white women are looking to increase the size of their lips, black women are seeking augmentation to restore their lip size to that of their youth. Black individuals do not experience the same frequency of perioral rhytides as white patients, but they experience a variety of other issues related to skin sagging and sinking. Unlike white women, enhancement of the vermilion border rarely is performed in black women due to development of rhytides, predominantly in the body of the lip below the vermilion border in response to volume loss in the upper lip while the lower lip usually maintains its same appearance.47

Facial enhancement utilizing poly-L-lactic acid can be used safely in SOC patients.51 Poly-L-lactic acid microparticles induce collagen formation, leading to dermal thickening over 3 to 6 months; however, multiple sessions are required to achieve optimal aesthetic results.

Patients with more reactive collagen can benefit from noninvasive cosmetic procedures such as skin-tightening procedures.52 Radiofrequency and microfocused ultrasound are cosmetic procedures used to provide skin tightening and facial lifting. They are safe and effective treatments for patients with Fitzpatrick skin types IV to VI.53 Histologically, there is less thinning of collagen bundles and elastic tissue in ethnic skin. Due to stimulation of collagen by these procedures, most SOC patients will experience a more enhanced response, requiring fewer treatment sessions than white individuals.

Conclusion

Medical and aesthetic facial concerns in SOC patients vary and can be a source of emotional and psychological distress that can negatively impact quality of life. The approach to the treatment of SOC patients should be a balance between tolerability and efficacy, considering the potential risk for PIH.

References
  1. White GM. Recent findings in the epidemiologic evidence, classification, and subtypes of acne vulgaris. J Am Acad Dermatol. 1998;39(2 pt 3):S34-S37.
  2. Halder RM, Grimes PE, McLaurin CL, et al. Incidence of common dermatoses in a predominantly black dermatologic practice. Cutis. 1983;32:388, 390.
  3. Alexis AF, Sergay AB, Taylor SC. Common dermatologic disorders in skin of color: a comparative practice survey. Cutis. 2007;80:387-394.
  4. Davis SA, Narahari S, Feldman SR, et al. Top dermatologic conditions in patients of color: an analysis of nationally representative data. J Drugs Dermatol. 2012;11:466-473.
  5. Perkins AC, Cheng CE, Hillebrand GG, et al. Comparison of the epidemiology of acne vulgaris among Caucasians, Asian, Continental Indian and African American women. J Eur Acad Dermatol Venereol. 2011;25:1054-1060.
  6. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl):S98-S106.
  7. Davis EC, Callender VD. A review of acne in ethnic skin: pathogenesis, clinical manifestations, and management strategies. J Clin Aesthet Dermatol. 2010;3:24-38.
  8. Halder RM, Holmes YC, Bridgeman-Shah S, et al. A clinicohistologic study of acne vulgaris in black females (abstract). J Invest Dermatol. 1996;106:888.
  9. Plewig G, Fulton JE, Kligman AM. Pomade acne. Arch Dermatol. 1970;101:580-584.
  10. Kligman AM, Mills OH. Acne cosmetica. Arch Dermatol. 1972;106:893-897.
  11. Halder RM, Brooks HL, Callender VD. Acne in ethnic skin. Dermatol Clin. 2003;21:609-615.
  12. Callender VD. Acne in ethnic skin: special considerations for therapy. Dermatol Ther. 2004;17:184-195.
  13. Winhoven SM. Postinflammatory hyperpigmentation in an Asian patient. a dramatic response to oral isotretinoin (13-cis-retinoic acid). Br J Med. 2005;152:368-403.
  14. Sarkar R, Bansal S, Garg VK. Chemical peels for melasma in dark-skinned patients. J Cutan Aesthet Surg. 2012;5:247-253.
  15. Alexis AF, Coley MK, Nijhawan RI, et al. Nonablative fractional laser resurfacing for acne scarring in patients with Fitzpatrick skin phototypes IV-VI. Dermatol Surg. 2016;42:392-402.
  16. Culp B, Scheinfeld N. Rosacea: a review. P T. 2009;34:38-45.
  17. Al-Dabagh A, Davis SA, McMichael AJ, et al. Rosacea in skin of color: not a rare diagnosis. Dermatol Online J. 2014:20. pii:13030/qt1mv9r0ss.
  18. Del Rosso JQ. Advances in understanding and managing rosacea: part 1: connecting the dots between pathophysiological mechanisms and common clinical features of rosacea with emphasis on vascular changes and facial erythema. J Clin Aesthet Dermatol. 2012;5:16-25.
  19. Jackson JM, Knuckles M, Minni JP, et al. The role of brimonidine tartrate gel in the treatment of rosacea. Clin Cosmet Investig Dermatol. 2015;23:529-538.
  20. Patel NU, Shukla S, Zaki J, et al. Oxymetazoline hydrochloride cream for facial erythema associated with rosacea. Expert Rev Clin Pharmacol. 2017;10:104954.
  21. Weinkle AP, Doktor V, Emer J. Update on the management of rosacea. Clin Cosmet Investig Dermatol. 2015;8:159-177.
  22. Grimes PE. Skin and hair cosmetic issues in women of color. Dermatol Clin. 2000;19:659-665.
  23. Taylor A, Pawaskar M, Taylor SL, et al. Prevalence of pigmentary disorders and their impact on quality of life: a prospective cohort study. J Cosmet Dermatol. 2008;7:164-168.
  24. Davis EC, Callender VD. Postinflammatory hyperpigmentation: a review of the epidemiology, clinical features, and treatment options in skin of color. J Clin Aesthet Dermatol. 2010;3:20-31.
  25. Grimes PE. Melasma: etiologic and therapeutic considerations. Arch Dermatol. 1995;131:1453-1457.
  26. Handel AC, Miot LD, Miot HA. Melasma: a clinical and epidemiological review. An Bras Dermatol. 2014;89:771-782.
  27. Callender VD, Reid SD, Obayan O, et al. Diagnostic clues to frontal fibrosing alopecia in patients of African descent. J Clin Aesthet Dermatol. 2016;9:45-51.
  28. Narang T, Sawatkar GU, Kumaran MS, et al. Minocycline for recurrent and/or chronic erythema nodosum leprosum. JAMA Dermatol. 2015;151:1026-1028.
  29. Boyer M, Katta R, Markus R. Diltiazem-induced photodistributed hyperpigmentation. Dermatol Online J. 2003;9:10.
  30. Pandya AG, Guevara IL. Disorders of hyperpigmentation. Dermatol Clin. 2000;18:91-98.
  31. Taylor SC, Torok H, Jones T, et al. Efficacy and safety of a new triple-combination agent for the treatment of facial melasma. Cutis. 2003;72:67-72.
  32. Bulengo-Ransby SM. Topical tretinoin (retinoic acid) therapy for hyperpigmented lesions caused by inflammation of the skin in black patients. N Engl J Med. 1993;328:1438-1443.
  33. Grimes P, Callender V. Tazarotene cream for postinflammatory hyperpigmentation and acne vulgaris in darker skin: a double-blind, randomized, vehicle-controlled study. Cutis. 2006;77:45-50.
  34. Jacyk WK. Adapalene in the treatment of African patients. J Eur Acad Dermatol Venereol. 2001;15(suppl 3):37-42.
  35. Kircik LH. Efficacy and safety of azelaic acid (AzA) gel 15% in the treatment of postinflammatory hyperpigmentation and acne: a 16-week, baseline-controlled study. J Drugs Dermatol. 2011;10:586-590.
  36. Lee HC, Thng TG, Goh CL. Oral tranexamic acid (TA) in the treatment of melasma. J Am Acad Dermatol. 2016;75:385-392.
  37. Kindred C, Okereke U, Callender VD. Skin-lightening agents: an overview of prescription, office-dispensed, and over-the-counter products. Cosmet Dermatol. 2013;26:18-26.
  38. Makino ET, Kadoya K, Sigler ML, et al. Development and clinical assessment of a comprehensive product for pigmentation control in multiple ethnic populations. J Drugs Dermatol. 2016;15:1562-1570.
  39. Halder R, Rodney I, Munhutu M, et al. Evaluation and effectiveness of a photoprotection composition (sunscreen) on subjects of skin of color. J Am Acad Dermatol. 2015;72(suppl 1):AB215.
  40. Castanedo-Cazares JP, Hernandez-Blanco D, Carlos-Ortega B, et al. Near-visible light and UV photoprotection in the treatment of melasma: a double-blind randomized trial. Photodermatol Photoimmunol Photomed. 2014;30:35-42.
  41. American Society for Aesthetic Plastic Surgery. 2016 Cosmetic Surgery National Data Bank Statistics. https://www.surgery.org/sites/default/files/ASAPS-Stats2016.pdf. Accessed November 15, 2017.
  42. Burgess CM. Soft tissue augmentation in skin of color: market growth, available fillers and successful techniques. J Drugs Dermatol. 2007;6:51-55.
  43. Davis EC, Callender VD. Aesthetic dermatology for aging ethnic skin. Dermatol Surg. 2011;37:901-917.
  44. Grimes PE, Arora S, Minus HR, et al. Dermatosis papulosa nigra. Cutis. 1983;32:385-386.
  45. Lupo M. Dermatosis papulosa nigra: treatment options. J Drugs Dermatol. 2007;6:29-30.
  46. Grimes PE, Shabazz D. A four-month randomized, double-blind evaluation of the efficacy of botulinum toxin type A for the treatment of glabellar lines in women with skin types V and VI. Dermatol Surg. 2009;35:429-435.
  47. Burgess CM, Awosika O. Ethnic and gender considerations in the use of facial injectables: African-American patients. Plast Reconstr Surg. 2015;136(5 suppl):28S-31S.
  48. Taylor SC, Kelly AP, Lim HW, et al, eds. Taylor and Kelly’s Dermatology for Skin of Color. 2nd ed. New York, NY: McGraw-Hill Education; 2016.
  49.  Taylor SC, Burgess CM, Callender VD. Safety of nonanimal stabilized hyaluronic acid dermal fillers in patients with skin of color: a randomized, evaluator-blinded comparative trial. Dermatol Surg. 2009;35(suppl 2):1653-1660.
  50. Marmur ES, Taylor SC, Grimes PE, et al. Six-month safety results of calcium hydroxylapatite for treatment of nasolabial folds in Fitzpatrick skin types IV to VI. Dermatol Surg. 2009;35(suppl 2):1641-1645.
  51. Hamilton TK, Burgess CM. Consideration for the use of injectable poly-L-lactic acid in people of color. J Drugs Dermatol. 2010;9:451-456.
  52. Fabi SG, Goldman MP. Retrospective evaluation of micro-focused ultrasound for lifting and tightening of the face and neck. Dermatol Surg. 2014;40:569-575.
  53. Harris MO, Sundaram HA. Safety of microfocused ultrasound with visualization in patients with Fitzpatrick skin phototypes III to VI. JAMA Facial Plast Surg. 2015;17:355-357.
References
  1. White GM. Recent findings in the epidemiologic evidence, classification, and subtypes of acne vulgaris. J Am Acad Dermatol. 1998;39(2 pt 3):S34-S37.
  2. Halder RM, Grimes PE, McLaurin CL, et al. Incidence of common dermatoses in a predominantly black dermatologic practice. Cutis. 1983;32:388, 390.
  3. Alexis AF, Sergay AB, Taylor SC. Common dermatologic disorders in skin of color: a comparative practice survey. Cutis. 2007;80:387-394.
  4. Davis SA, Narahari S, Feldman SR, et al. Top dermatologic conditions in patients of color: an analysis of nationally representative data. J Drugs Dermatol. 2012;11:466-473.
  5. Perkins AC, Cheng CE, Hillebrand GG, et al. Comparison of the epidemiology of acne vulgaris among Caucasians, Asian, Continental Indian and African American women. J Eur Acad Dermatol Venereol. 2011;25:1054-1060.
  6. Taylor SC, Cook-Bolden F, Rahman Z, et al. Acne vulgaris in skin of color. J Am Acad Dermatol. 2002;46(2 suppl):S98-S106.
  7. Davis EC, Callender VD. A review of acne in ethnic skin: pathogenesis, clinical manifestations, and management strategies. J Clin Aesthet Dermatol. 2010;3:24-38.
  8. Halder RM, Holmes YC, Bridgeman-Shah S, et al. A clinicohistologic study of acne vulgaris in black females (abstract). J Invest Dermatol. 1996;106:888.
  9. Plewig G, Fulton JE, Kligman AM. Pomade acne. Arch Dermatol. 1970;101:580-584.
  10. Kligman AM, Mills OH. Acne cosmetica. Arch Dermatol. 1972;106:893-897.
  11. Halder RM, Brooks HL, Callender VD. Acne in ethnic skin. Dermatol Clin. 2003;21:609-615.
  12. Callender VD. Acne in ethnic skin: special considerations for therapy. Dermatol Ther. 2004;17:184-195.
  13. Winhoven SM. Postinflammatory hyperpigmentation in an Asian patient. a dramatic response to oral isotretinoin (13-cis-retinoic acid). Br J Med. 2005;152:368-403.
  14. Sarkar R, Bansal S, Garg VK. Chemical peels for melasma in dark-skinned patients. J Cutan Aesthet Surg. 2012;5:247-253.
  15. Alexis AF, Coley MK, Nijhawan RI, et al. Nonablative fractional laser resurfacing for acne scarring in patients with Fitzpatrick skin phototypes IV-VI. Dermatol Surg. 2016;42:392-402.
  16. Culp B, Scheinfeld N. Rosacea: a review. P T. 2009;34:38-45.
  17. Al-Dabagh A, Davis SA, McMichael AJ, et al. Rosacea in skin of color: not a rare diagnosis. Dermatol Online J. 2014:20. pii:13030/qt1mv9r0ss.
  18. Del Rosso JQ. Advances in understanding and managing rosacea: part 1: connecting the dots between pathophysiological mechanisms and common clinical features of rosacea with emphasis on vascular changes and facial erythema. J Clin Aesthet Dermatol. 2012;5:16-25.
  19. Jackson JM, Knuckles M, Minni JP, et al. The role of brimonidine tartrate gel in the treatment of rosacea. Clin Cosmet Investig Dermatol. 2015;23:529-538.
  20. Patel NU, Shukla S, Zaki J, et al. Oxymetazoline hydrochloride cream for facial erythema associated with rosacea. Expert Rev Clin Pharmacol. 2017;10:104954.
  21. Weinkle AP, Doktor V, Emer J. Update on the management of rosacea. Clin Cosmet Investig Dermatol. 2015;8:159-177.
  22. Grimes PE. Skin and hair cosmetic issues in women of color. Dermatol Clin. 2000;19:659-665.
  23. Taylor A, Pawaskar M, Taylor SL, et al. Prevalence of pigmentary disorders and their impact on quality of life: a prospective cohort study. J Cosmet Dermatol. 2008;7:164-168.
  24. Davis EC, Callender VD. Postinflammatory hyperpigmentation: a review of the epidemiology, clinical features, and treatment options in skin of color. J Clin Aesthet Dermatol. 2010;3:20-31.
  25. Grimes PE. Melasma: etiologic and therapeutic considerations. Arch Dermatol. 1995;131:1453-1457.
  26. Handel AC, Miot LD, Miot HA. Melasma: a clinical and epidemiological review. An Bras Dermatol. 2014;89:771-782.
  27. Callender VD, Reid SD, Obayan O, et al. Diagnostic clues to frontal fibrosing alopecia in patients of African descent. J Clin Aesthet Dermatol. 2016;9:45-51.
  28. Narang T, Sawatkar GU, Kumaran MS, et al. Minocycline for recurrent and/or chronic erythema nodosum leprosum. JAMA Dermatol. 2015;151:1026-1028.
  29. Boyer M, Katta R, Markus R. Diltiazem-induced photodistributed hyperpigmentation. Dermatol Online J. 2003;9:10.
  30. Pandya AG, Guevara IL. Disorders of hyperpigmentation. Dermatol Clin. 2000;18:91-98.
  31. Taylor SC, Torok H, Jones T, et al. Efficacy and safety of a new triple-combination agent for the treatment of facial melasma. Cutis. 2003;72:67-72.
  32. Bulengo-Ransby SM. Topical tretinoin (retinoic acid) therapy for hyperpigmented lesions caused by inflammation of the skin in black patients. N Engl J Med. 1993;328:1438-1443.
  33. Grimes P, Callender V. Tazarotene cream for postinflammatory hyperpigmentation and acne vulgaris in darker skin: a double-blind, randomized, vehicle-controlled study. Cutis. 2006;77:45-50.
  34. Jacyk WK. Adapalene in the treatment of African patients. J Eur Acad Dermatol Venereol. 2001;15(suppl 3):37-42.
  35. Kircik LH. Efficacy and safety of azelaic acid (AzA) gel 15% in the treatment of postinflammatory hyperpigmentation and acne: a 16-week, baseline-controlled study. J Drugs Dermatol. 2011;10:586-590.
  36. Lee HC, Thng TG, Goh CL. Oral tranexamic acid (TA) in the treatment of melasma. J Am Acad Dermatol. 2016;75:385-392.
  37. Kindred C, Okereke U, Callender VD. Skin-lightening agents: an overview of prescription, office-dispensed, and over-the-counter products. Cosmet Dermatol. 2013;26:18-26.
  38. Makino ET, Kadoya K, Sigler ML, et al. Development and clinical assessment of a comprehensive product for pigmentation control in multiple ethnic populations. J Drugs Dermatol. 2016;15:1562-1570.
  39. Halder R, Rodney I, Munhutu M, et al. Evaluation and effectiveness of a photoprotection composition (sunscreen) on subjects of skin of color. J Am Acad Dermatol. 2015;72(suppl 1):AB215.
  40. Castanedo-Cazares JP, Hernandez-Blanco D, Carlos-Ortega B, et al. Near-visible light and UV photoprotection in the treatment of melasma: a double-blind randomized trial. Photodermatol Photoimmunol Photomed. 2014;30:35-42.
  41. American Society for Aesthetic Plastic Surgery. 2016 Cosmetic Surgery National Data Bank Statistics. https://www.surgery.org/sites/default/files/ASAPS-Stats2016.pdf. Accessed November 15, 2017.
  42. Burgess CM. Soft tissue augmentation in skin of color: market growth, available fillers and successful techniques. J Drugs Dermatol. 2007;6:51-55.
  43. Davis EC, Callender VD. Aesthetic dermatology for aging ethnic skin. Dermatol Surg. 2011;37:901-917.
  44. Grimes PE, Arora S, Minus HR, et al. Dermatosis papulosa nigra. Cutis. 1983;32:385-386.
  45. Lupo M. Dermatosis papulosa nigra: treatment options. J Drugs Dermatol. 2007;6:29-30.
  46. Grimes PE, Shabazz D. A four-month randomized, double-blind evaluation of the efficacy of botulinum toxin type A for the treatment of glabellar lines in women with skin types V and VI. Dermatol Surg. 2009;35:429-435.
  47. Burgess CM, Awosika O. Ethnic and gender considerations in the use of facial injectables: African-American patients. Plast Reconstr Surg. 2015;136(5 suppl):28S-31S.
  48. Taylor SC, Kelly AP, Lim HW, et al, eds. Taylor and Kelly’s Dermatology for Skin of Color. 2nd ed. New York, NY: McGraw-Hill Education; 2016.
  49.  Taylor SC, Burgess CM, Callender VD. Safety of nonanimal stabilized hyaluronic acid dermal fillers in patients with skin of color: a randomized, evaluator-blinded comparative trial. Dermatol Surg. 2009;35(suppl 2):1653-1660.
  50. Marmur ES, Taylor SC, Grimes PE, et al. Six-month safety results of calcium hydroxylapatite for treatment of nasolabial folds in Fitzpatrick skin types IV to VI. Dermatol Surg. 2009;35(suppl 2):1641-1645.
  51. Hamilton TK, Burgess CM. Consideration for the use of injectable poly-L-lactic acid in people of color. J Drugs Dermatol. 2010;9:451-456.
  52. Fabi SG, Goldman MP. Retrospective evaluation of micro-focused ultrasound for lifting and tightening of the face and neck. Dermatol Surg. 2014;40:569-575.
  53. Harris MO, Sundaram HA. Safety of microfocused ultrasound with visualization in patients with Fitzpatrick skin phototypes III to VI. JAMA Facial Plast Surg. 2015;17:355-357.
Issue
Cutis - 100(6)
Issue
Cutis - 100(6)
Page Number
375-380
Page Number
375-380
Publications
Publications
Topics
Article Type
Display Headline
Approach to Treatment of Medical and Cosmetic Facial Concerns in Skin of Color Patients
Display Headline
Approach to Treatment of Medical and Cosmetic Facial Concerns in Skin of Color Patients
Sections
Inside the Article

Practice Points

  • Treat acne in skin of color (SOC) patients early and aggressively to prevent or minimize subsequent postinflammatory hyperpigmentation (PIH) and acne scarring.
  • Vascular lasers and intense pulsed light may be used to address the vascular components of rosacea; however, the latter is not recommended in Fitzpatrick skin types IV to VI.
  • Hydroquinone is the gold standard for skin lightening and is often used as a first-line therapy for melasma and PIH.
  • Photoprotection is an essential component of therapy for hyperpigmented skin disorders.  
  • Cosmetic procedures are gaining popularity in the SOC population. When treating SOC patients, consider the impact of ethnicity on aging and facial structure, the patient's desired cosmetic outcome, tissue reaction to anticipated treatments, and the patient's expectations for recommended therapies.
Disallow All Ads
Content Gating
No Gating (article Unlocked/Free)
Alternative CME
Disqus Comments
Default
Use ProPublica
Hide sidebar & use full width
render the right sidebar.
Conference Recap Checkbox
Not Conference Recap
Clinical Edge
Display the Slideshow in this Article
Medscape Article
Article PDF Media