164
Views
11
CrossRef citations to date
0
Altmetric
Review

Review of photodynamic therapy in actinic keratosis and basal cell carcinoma

, &
Pages 1-9 | Published online: 28 Dec 2022

Abstract

The number of non-melanoma skin cancers is increasing worldwide, and so also the demand for effective treatment modalities. Topical photodynamic therapy (PDT) using aminolaevulinic acid or its methyl ester has recently become good treatment options for actinic keratosis and basal cell carcinoma; especielly when treating large areas and areas with field cancerization. The cure rates are usually good, and the cosmetic outcomes excellent. The only major side effect reported is the pain experienced by the patients during treatment. This review covers the fundamental aspects of topical PDT and its application for treatment of actinic keratosis and basal cell carcinoma. Both potentials and limitations will be reviewed, as well as some recent development within the field.

Introduction

The number of skin cancers has increased annually for many years (CitationMarks 1995). One of the most important etiologic factors is considered to be sun exposure. Staying out in the sun during a prolonged time and repeated sun burns are clear risk factors for the development of different types of skin cancer. This is especially applicable on persons having lighter complexion (skin type I-II).

There are mainly three forms of malignant tumors of the skin. These are malignant melanoma, basal cell carcinoma (BCC) and squamous cell carcinoma (SCC). BCC and SCC belong to the group of non-melanoma skin cancer (NMSC), which is the most common skin malignancy worldwide (CitationMiller 1991; CitationGreen 1992). Actinic keratosis (AK) and squamous cell carcinoma in situ (SCCIS), also called Morbus Bowen, are precursors of SCC and therefore also normally are ascribed the NMSCs. The incidence of NMSC is increasing steadily (CitationGreen 1992), and the association between skin type, solar habits and NMSC is well known. Furthermore, once a patient has had a NMSC, additional lesions are common (CitationFrankel et al 1992). Also, organ transplant recipients (OTR) run an extremely high risk of contracting NMSC (CitationAdami et al 2003; CitationBouwes Bavinck et al 2007). Consequently, an increasing number of patients seek care with sun damaged skin and skin tumors and the patients must be taken care of in an effective manner.

During recent years, several therapeutic modalities have been available for superficial skin cancer. One of those is photodynamic therapy (PDT), which involves the activation of a photosensitizer using visible light (CitationHenderson and Dougherty 1992). This results in the formation of reactive and cytotoxic singlet oxygen. PDT is a relatively new therapeutic method and it has become a good complement in this respect to already established treatments of skin cancer of non-melanoma type, particularly because of the good cosmetic outcome which is increasingly important, as recently reviewed by other authors (CitationMorton 2004; CitationMarmur et al 2004; CitationKormeili et al 2004; CitationLehmann 2007; CitationBraathen et al 2007). The current review will address PDT of BCC and AK, which are the most common applications of topical-PDT within dermatology today. Also the fundamental aspects and recent developments of the treatment will be covered.

Actinic keratosis (AK)

SCC and AK represent the same disease process at different stages of evolution (CitationCockerell 2000), ie, AKs are proliferations of transformed, neoplastic keratinocytes confined to the epidermis, whereas SCC extend more deeply including dermis. Thus AK lesions are considered as pre-cancerous or pre-malignant, and a few of the lesions progress to SCCIS or SCC. AKs are extremely prevalent and strongly related to sun exposure. The lesions occur on the dorsum of the hands, the face, scalp and other sun exposed sites. AKs are often multiple and recently the concept of ‘field cancerization’ has been discussed, meaning that the clinically normal appearing skin around AKs provides the basis for clonal expansion of genetically altered neoplastic cells (CitationBraakhuis et al 2003). This means that an entire area has undergone solar damage with precancerous and cancerous lesions. Field cancerization can, for example, occur in the face. It is clinically difficult to distinguish between a proliferative AK and an early invasive SCC. As field cancerization often occurs, it is necessary with a treatment that not only involves overt AK, but also treats subclinical lesions nearby. For this reason, topical PDT which is a non-invasive method, has become an important treatment modality.

Basal cell carcinoma

Of the NMSCs, BCCs are the most common (CitationMiller 1991). Of all BCCs, 85% appear in the head and neck region. BCCs almost never metastasize but can destroy the tissue locally and thus have to be treated. Particularly this has to be kept in mind when located in the H-zone. The BCCs can be divided into three variants according to histopathological pattern and degree of aggressiveness (CitationMiller 1991; CitationChampion et al 1998). The most common is nodular BCC, which grows in large, normally well delineated, rounded, nests pushing into dermis. Nodular BCC accounts for approximately 60% of the BCCs, and is most frequently observed in the face. Superficial BCC has a growth pattern where the tumor nests are restricted to epidermis or superficial parts of hair follicles. The most common location of the superficial BCCs is on the trunk. The third and most aggressive type is morpheic BCC, which has an infiltrative growth pattern. The tumor nests are diffuse and irregularly spread, and are often found in subcutis. This type of BCC is often located on the central face, ie, on the nose, lips, ears and around the eyes, with a higher risk of recurrence after treatment compared to other locations. Transitions between the different types of BCC may occur. Regarding therapy, the H-zone location is very important to keep in mind concerning the high recurrence risk, while, usually, superficial BCC outside this area run a low risk. Topical PDT is today an accepted treatment for superficial BCC.

Non-melanoma skin cancer in organ transplant recipients(OTRs)

OTRs have been reported to run a more than fifty times increased risk of contracting SCC because of their immunosuppressive therapy (CitationAdami et al 2003). Many of these patients obtain widespread AK. Heart transplant patients are particularly susceptible because of their older mean age and the requirement for more aggressive immunosuppressive therapy. Therefore, efficient treatment modalities are required to be able to treat these patients, and topical-PDT has recently become an interesting treatment option.

Basic aspects of PDT

The fundamental approach of PDT is initial photosensitization of the treatment site, followed by irradiation with visible light, which initiates a tissue-toxic photochemical reaction (CitationHenderson and Dougherty 1992). The first attempts at PDT were reported by Tappeiner and Jesionek as early as in 1903 (CitationTappeiner and Jesionek 1903). But it was not until in 1972, Diamond and co-workers found that tumor cells were destroyed by visible light after sensitization using hematoporphyrin derivative (CitationDiamond et al 1972). Thereafter CitationDougherty et al (1978) initiated clinical studies of PDT of various malignant tumors with promising results. Today, most commonly, the photosensitization for PDT of superficial skin lesions is obtained by topical application of δ-5-aminolaevulinic acid (ALA) or its methyl ester (MAL) (CitationPeng et al 1997; CitationSalva 2002; CitationBraathen et al 2007). The main advantages of topical PDT are that the method is non-invasive and effective, and generally gives a good cosmetic outcome. The application of ALA or MAL enhances the formation of protoporphyrin IX (PpIX) in the skin. The tumor is irradiated with light matching the absorption of PpIX, which initiates the photochemical reaction, in which reactive singlet oxygen is formed. The singlet oxygen is generally believed to be one of the key factors for the desired therapeutic effect of PDT (CitationWeishaupt et al 1976; CitationMoan and Sommer 1985). Consequently, the presence of appropriate concentration of sensitizer, light matching the absorption of the sensitizer and molecular oxygen in the tissue, is crucial for the efficiency of PDT.

Photosensitizing drugs for topical PDT

The early photosensitisers for PDT were based on hematoporphyrin derivatives, which were directly injected into the tumors (CitationPeng et al 1997). The main drawback however was prolonged and systemic photosensitivity, persisting around 4–6 weeks. Kennedy et al investigated the possibility to use endogenously formed PpIX obtained after application of ALA, a precursor in the haem synthesis (CitationKennedy et al 1990; CitationKennedy and Pottier 1992). The results were promising, and the drawbacks of systemic and prolonged photosensitization were eliminated. Instead, a local photosensitization of the application site was found up to 48 hours. Since then, topical ALA-PDT has been widely investigated for treatment of superficial skin lesions, eg, BCCs and AK. In addition to ALA, a number of ALA esters have been investigated (CitationLopez et al 2004). The methylester, ie, 16% MAL, has gained drug approval and is the sole commercially available drug for topical PDT in Europe with the trade name Metvix® (Galderma & Photocure ASA). While in the USA, 20% ALA in ethanol solution is the approved drug, ie, Levulan® (DUSA Pharmaceuticals, Inc.).

It has been known for a long time that certain abnormalities in the haem synthesis, ie, porphyria, lead to extensive photosensitization because of the accumulation of porphyrins particularly PpIX (CitationBottomley and Muller-Eberhard 1988). But the group of CitationPottier et al (1986) was the first to demonstrate the accumulation of PpIX after injection of exogenous ALA in mice. The formation of haem takes place in the mitochondrial and the cytosolic compartments of the cells. The primary substrates in the production of haem are glycine and succinyl CoA, from which ALA is formed. This is the rate-limiting step in the pathway, due to negative feedback control by haem. By applying exogenous ALA this step is bypassed leading to an accumulation of PpIX in the tissue. Two enzymes have been reported to be of particular importance for the accumulation of PpIX in tissue after topical application of ALA. These enzymes are porphobilinogen deaminase (PBGD) and ferrochelatase (FC). Increased activity of PBGD, in connection with decreased FC activity in neoplastic tissue, have been reported to be possible factors behind the selective accumulation of PpIX (CitationDailey and Smith 1984; CitationLeibovici et al 1988; CitationKondo et al 1993; CitationHinnen et al 1998; CitationGibson et al 1998). The metabolic pathway of MAL is generally believed to be the same as ALA, but the transportmechanism into the cells have been shown to differ (CitationRud et al 2000; CitationGederaas et al 2001; CitationRodriguez et al 2006). In addition, the the methyl-ester must be dehydrolyzed at some stage, but it is not clear in which step the ester hydrolysis is undertaken.

Selective accumulation between tumor and normal tissue has been observed using ALA-induced PpIX sensitization (CitationEl-Far et al 1990; CitationKennedy and Pottier 1994; CitationAbels et al 1994; CitationAndersson-Engels et al 1995). When ALA is applied systemically, the selectivity is most likely explained by altered enzymatic activity in the tumor cells, as described above. Although, the stratum corneum, ie, the outer skin barrier, seems to be the factor of major importance for the selectivity, since it has a large impact on the penetration of ALA through the skin (CitationMoan et al 2001). The abnormal keratin layer that is produced by BCCs or SCCs is rapidly penetrated by ALA, while the adjacent normal skin with intact stratum corneum is less permeable. This has been verified in vivo, where the ALA penetration was found to be higher in BCC compared to the surrounding normal skin by microdialysis studies (CitationWennberg et al 2000). Consequently, higher concentrations of ALA will result in higher amounts of PpIX in the tumor. The fluorescence contrast after application of ALA has been found to be time dependent with an optimum around 3 hours in BCCs (CitationEricson et al 2003). There are indications that MAL is more selective towards neoplastic tissue compared to ALA (CitationAngell-Petersen et al 2006), although there is a lack of comparing studies verifying this relationship.

When ALA or MAL is applied topically, sufficient amounts of drug have been found to be present in both epidermis and dermis, although the dermal cells do not develop significant PpIX levels to become photosensitized (CitationDivaris et al 1990). Hence, only epidermis becomes sensitized. This makes it possible to treat epidermal cancers without damaging dermis, which might be the reason why scarring is uncommon in topical ALA-PDT. It has been shown that the cellular localization of PpIX after ALA application is restricted to the mitochondria (CitationPeng et al 1992; CitationIinuma et al 1994; CitationMalik et al 1996). Hence, the initial photodynamic damage will be localized to this organelle. The subsequent apoptosis of the cells has been reported to occur within 10 hours (CitationWebber et al 1996).

In order to improve topical drug delivery of ALA or MAL, some attempts have been made. CitationSoler et al (2000) used a formulation including DMSO for delivery of 20% ALA. Other recent publications report on the use of cubic-lipid-systems (CitationBender et al 2005), and a bioadhesive patch (CitationMcCarron et al 2006). However, so far there is a lack of studies verifying the clinical response using these methods.

Light sources and dosimetry

Whatever the choice of light source for PDT, two criteria have to be fulfilled. Firstly, the wavelength must match the absorption of the sensitizer in order to induce the desired photochemical reaction. The second criterion is that the light must be able to penetrate the tissue, so that the depth of the tumor can be treated. Due to the presence of melanin, hemoglobin, oxyhemoglobin and water, the absorption of visible light is high in biological tissue (CitationTuchin 2000). But the absorption has a minimum in the wavelength region 600–1000 nm, ie, the “optical window” of tissue (CitationRichards-Kortum and Sevick-Muraca 1996). In this region, the light propagation is dominated by scattering, and the penetration depth is around 8–10 mm. Therefore, normally red light centered around 635 nm is used to ensure therapeutic fluence rates for topical PDT (CitationPeng et al 1997; CitationSalva 2002; CitationBraathen et al 2007), even though the maximum absorption of PpIX is around 400 nm. But successful treatment of AK with ALA-PDT and blue light has been reported (CitationJeffes et al 2001; CitationZelickson et al 2005).

Both coherent and incoherent light sources have been applied for PDT of various diseases. Lasers offer significant advantages whenever fibre optics are needed. In addition lasers have the advantage of producing monochromatic light, exactly matching the absorption band of the sensitizer. In this way excessive heating is avoided. Various lasers have been applied in PDT, eg, CitationSvanberg et al (1994), but the major drawbacks are their bulkiness and the fact that they are expensive. Moreover, the area of irradiation is limited, so the beam has to be scanned in order to treat larger areas.

Filtered broadband lamps have proven to be very useful for the treatment of superficial lesions (CitationWennberg et al 1996; CitationVarma et al 2001). The main advantages are that they are relatively inexpensive, but the wavelength bands are usually quite broad, 50–130 nm, and they can be quite bulky. The wavelengths outside the absorption spectrum of the photosensitiser were earlier thought to have no therapeutic effect, other than hyperthermia. On the other hand, it has been discussed that the excitation of photoproducts absorbing around 670 nm might actually contribute to the photodynamic reaction (CitationPeng et al 1997); but the results are contradictory (CitationSoler et al 2000; CitationClark et al 2003). Recently, so-called intense pulsed light (IPL) has been evaluated for PDT (CitationKim et al 2005; CitationGold et al 2006), although there is a report on dramatically decreased PDT reaction using IPL compared to broad-band light source (CitationStrasswimmer and Grande 2006).

Light-emitting diodes (LED) can also be used as light sources for PDT (CitationThompson et al 2001; CitationYang et al 2003). The wavelength band is narrower compared with filtered lamps. The earlier problem with LEDs has been their relatively low intensity, although the recent generation of LEDs seems to provide sufficient intensities for PDT of superficial skin lesions. These light sources are very compact and cheap, which is of great preference for clinical practice.

Light dosimetry and oxygen depletion

Despite the fact that PDT is a rather well accepted treatment modality for treatment of NMSC, there exists no common light dosimetry guide. Because of the wide variety of light sources, the reported light doses and fluence rates used for PDT have a broad span, ranging from 50 to 500 J/cm2 and 50 to 200 mW/cm2 respectively (CitationDougherty et al 1998; CitationRadakovic-Fijan et al 2005). Fluence rates below 150 mW/cm2 should be used to avoid hyperthermia (CitationPeng et al 1997). In addition, oxygen depletion is of concern at fluence rates above 50 mW/cm2 for PDT of AK (CitationEricson et al 2004). This is verified by experimental studies in rodents (CitationSitnik et al 1998; CitationRobinson et al 1999; CitationIinuma et al 1999; CitationBissonnette et al 2004), cell spheroids (CitationFoster et al 1993), and mathematical modeling (CitationFoster et al 1991), showing significantly better results when low fluence rates are applied. These results are most likely related to oxygen depletion at high fluence rates (CitationTromberg et al 1990).

Since the presence of singlet oxygen is crucial for the photodynamic reaction, it would be desirable to be able to monitor its formation in vivo. Unfortunately, there exists no simple method of direct detection of singlet oxygen in biological tissue (CitationGorman and Rodgers 1992), but there have been attempts with some success in keratinocytes (CitationBilski et al 1998). Different oxygen sensors have been investigated in pre-clinical studies for monitoring the oxygen tension during PDT (CitationCurnow et al 2000), although this has so far not been investigated in patients. Another possibility is in vivo monitoring of the photobleaching of the sensitizer. In the clinical study presented by CitationEricson et al (2004) it was reported that the treatment outcome indeed was found to be fluence-rate-dependent and correlated to the rate of photobleaching, where low fluence rates were found preferable.

Light fractionation

It has been suggested that light fractionation should be performed to minimize the effect of hypoxia during PDT. Splitting up the light dose in minor fractions would allow oxygen to diffuse back into the cells, increasing the efficiency of the treatment. Theoretical calculations have predicted that effective reoxygenation takes place already after 45 s (CitationFoster et al 1991). In vivo studies in rodents state that dark interval of 150 s is required (CitationCurnow et al 2000). Other reports with 2 hours’ dark period show similar results (CitationRobinson et al 2000; CitationRobinson et al 2003). For this reason, various illumination schemes have been suggested. For example, Citationde Haas et al (2006) suggest that 20 + 80 J/cm2 given at a fluence rate of 50 mW/cm2 is more effective than a single treatment of 75 J/cm2 when treating superficial BCCs. Earlier reports based on results from an animal model recommend 5 + 95 J/cm2 (CitationRobinson et al 2000; CitationRobinson et al 2003). Other investigators have failed to show an effect when using fractionated PDT (CitationBabilas et al 2003), which implies that the choice of illumination scheme is important in order to obtain an effect of the fractionation minimizing hypoxia during PDT.

Clinical studies with PDT of AK

AK typically appears as field cancerization of large areas in the head and face region. Thus PDT has proven to be an excellent treatment modality for this subgroup of patients, because of the ability to treat large areas and preferable cosmetic outcome. Both ALA and MAL have been applied as photosensitizisers for PDT of AK. No comparable study currently exists, although the cure rates seem to be similar by reviewing the literature.

The long-term cure rate for PDT using 20% ALA solution (14–18 hours application), blue light (10 mW/cm2, 10 J/ cm2), and two treatments was reported to be 78% after 12 months follow up (CitationTschen et al 2006). Another study reports on 89% cure rate at 3 month follow up using a similar treatment protocol (CitationPiacquadio et al 2004). In a light-dose ranging study complete response using 20% ALA cream (3 hours application), red light (30 mW/cm2, 100 J/cm2), and single treatment was found to be 89% at two months follow up in the low fluence rate group (CitationEricson et al 2004). Similar treatment protocol, but two treatments and red light (70 mW/cm2, 70 J/cm2), resulted in 85% complete response (CitationSandberg et al 2006). A study has been performed comparing 5-fluorouracil (5-FU) to a single treatment of ALA-PDT, with comparable results (CitationKurwa et al 1999).

Cure rates for PDT with 16% MAL (3 hours’ application), red light (75 J/cm2), and two treatments were reported to be 89%–91% at 3-month follow-up (CitationPariser et al 2003; CitationFreeman et al 2003). When treating thin actinic keratosis, it seems that one treatment is sufficient using MAL-PDT, although two treatments are recommended for more hyperkeratotic lesions (CitationTarstedt et al 2005). Studies comparing the outcome by two consecutive topical MAL-PDTs versus cryotherapy have been performed for AKs (CitationSzeimies et al 2002; CitationMorton et al 2006). The cure rates were found similar although the cosmetic results were superior using PDT. In addition, the patients were found to prefer PDT over cryotherapy.

Clinical studies with PDT of BCC

Superficial BCC

Several studies have been performed investigating PDT for superficial BCCs. For example, CitationWennberg et al (1996) report a 91% clearance of superficial BCCs treated with single sessions of 20% ALA-gel (4 hours’ application) and red light (125–166 mW/cm2, 75–100 J/cm2). CitationSoler et al (2000) found no differences in cure rates, ie, 82%–86%, comparing laser (630 nm, 120–150 mW/cm2, 100–150 J/cm2) and a broad band light source (100–180 mW/cm2, 150–200 J/cm2) when treating superficial BCCs with ALA-PDT. CitationHorn et al (2003) demonstrated clearance rates at 92% when using two consecutive treatments of MAL-PDT (16% MAL, red light). The cosmetic results are usually excellent and relatively large areas can be treated.

Nodular basal carcinoma

There is some evidence for using topical PDT in nBCCs (CitationHorn et al 2003; CitationVinciullo et al 2005). However, a careful curettage was carried out before application of the MAL-cream. CitationRhodes et al (2004) demonstrated potential problems with long-term recurrence rates when treating nodular BCC with MAL-PDT. In order to improve cure rates for PDT of nodular BCCs, intralesional treatment protocol has been suggested (CitationCappugi et al 2004). Again, the cosmetic outcomes with PDT are generally impressive. Thus, topical-PDT can be a preferable treatment choice for nodular BCCs in ‘difficult-to-treat’ areas, although special pre-operative handling is necessary.

Morpheic basal cell carcinomas

Morpheic BCCs shall not be treated with PDT. Instead, a Mohs’ micrographic surgery is the golden standard (CitationWennberg et al 1999).

PDT in organ transplanted recipients

Recently, topical PDT has become an interesting treatment modality for dealing with NMSC among OTRs, who exhibit an increased risk of contracting NMSC due to their immune suppression. Both ALA-PDT and MAL-PDT seem to be able to deal with AK and other epidermal dysplasia among OTRs (CitationDragieva et al 2004; CitationWulf et al 2006; CitationWennberg et al 2005). It appears that the therapeutic results from PDT are superior to that from 5-FU (CitationPerrett et al 2007). Topical PDT in OTRs has been suggested to have a prevention potential prohibiting the development of new lesions (CitationWulf et al 2006), although contradicting results exist (Citationde Graaf et al 2006).

Side effects of PDT

Acute

PDT is generally well tolerated. Immediately after treatment an erythema and a slight oedema is often seen. Crusts and superficial erosions are typically seen after a few days. Severe ulceration is rare. The most bothersome acute side effect is pain (CitationKennedy and Pottier 1992; CitationFijan et al 1995). Most patients experience a burning sensation, also described as ‘stinging’ or ‘prickling’, in the ALA-treated area during light exposure. The mechanism for this reaction is not clear. A possible explanation is hyperthermia of the tissue; however, CitationOrenstein et al (1995) obtained performed IR imaging and could not relate the pain sensation with a temperature increase. This implies that the pain sensation is a consequence of the photochemical reaction in the tissue, and related to the presence of reactive singlet oxygen. This is consistent with the clinical experience that the peak pain is obtained after a few minutes of irradiation when the photodynamic activity is high, and then gradually decreased towards the background level (CitationKennedy and Pottier 1992; CitationEricson et al 2004).

It has been reported that the site of the lesion is of importance for the pain sensation (CitationGrapengiesser et al 2002; CitationKapur et al 2003). Treatment of facial lesions, and lesions on the scalp result in more pain. Also, the amount of pain seems to be related to the size of the treated area and type of lesion, since patients with AKs have been reported to experience more pain than patients with BCCs (CitationGrapengiesser et al 2002). There are reports that lower pain scores are obtained with MAL-PDT in comparison with ALA-PDT performed on tape-stripped normal skin (CitationWiegell et al 2003) and AKs (CitationKasche et al 2006). But even still, 4 out of the 28 AK patients included in the study by CitationKasche et al (2006) were not able to complete the treatment due to unbearable pain. Pain management most commonly involves local anesthetics, premedication, fan-cooling or spraying the treated area with water. However, the pain-relieving effects are diverging. Cold air analgesia has shown some effect (CitationPagliaro et al 2004), while tetracaine gel (CitationHolmes et al 2004), capsaicin cream (CitationSandberg et al 2006), and morphine-gel (CitationSkiveren et al 2006) seem to have no effect. Currently, no standard protocol for pain relief during PDT exists and further studies on the matter are desired.

Chronic side effects

Chronic side effects of topical PDT are rare. Only three case reports on contact allergy exist. One case report concerns contact allergy to ALA and the other two on MAL (CitationWulf and Philipsen 2004; CitationHarries et al 2007). Thus, topical PDT can be considered to be a safe treatment.

Conclusion

Topical PDT, both with ALA and MAL, seems to offer a good therapeutic alternative to standard therapies in treating superficial NMSC, especially if widespread areas or field cancerization are involved. Treatment results are generally very good and the cosmetic results are excellent. Large areas of AK can easily be treated by topical-PDT, especially in the head and neck area, such as the scalp of old men. However, pain in these locations can sometimes be cumbersome to deal with and new pain-relieving strategies are required.

For BCCs, topical-PDT has proven especially suited for the superficial form, particularly for treatment of thin and multiple superficial BCCs. Also in this case, both clinical and cosmetic outcomes are excellent. Nodular BCCs are normally excised if possible. Surgery can easily be performed on the cheek, the forehead and the lips, but for lesions located on the nose, eyelids and external ear, simple excision is complicated. For these locations topical-PDT has shown some potiential, although thorough pre-treatment before application of ALA or MAL is necessary, and there is a risk for recurrences. In these instances, also cryotherapy may be good alternative to surgery (CitationLindgren and Larko 1997).

The OTRs constitute a patient group which suffer from widespread NMSC, and field cancerization constitute a major problem. For these patients, topical-PDT has shown several advantages. In addition, compliance is less of a problem using topical-PDT as the physician has total control over the treatment as opposed to topical treatments, eg, 5-FU, which should be used for a prolonged period (several weeks) and often lead to compliance problems.

References

  • AbelsCHeilPDellianMIn vivo kinetics and spectra of 5-aminolaevulinic acid-induced fluorescence in an amelanotic melanoma of the hamsterBr J Cancer199470826337947087
  • AdamiJGabelHLindelofBCancer risk following organ transplantation: a nationwide cohort study in SwedenBr J Cancer2003891221714520450
  • Andersson-EngelsSBergRSvanbergKMulti-colour fluorescence imaging in connection with photodynamic therapy of delta-amino levulinic acid (ALA) sensitised skin malignanciesBioimaging1995313443
  • Angell-PetersenESorensenRWarloeTPorphyrin formation in actinic keratosis and basal cell carcinoma after topical application of methyl 5-aminolevulinateJ Invest Dermatol20061262657116374471
  • BabilasPSchachtVLiebschGEffects of light fractionation and different fluence rates on photodynamic therapy with 5-aminolaevulinic acid in vivoBr J Cancer2003881462912778078
  • BenderJEricsonMBMerclinNLipid cubic phases for improved topical drug delivery in photodynamic therapyJ Control Release200510635036015967535
  • BilskiPKukielczakBMChignellCFPhotoproduction and direct spectral detection of singlet molecular oxygen (O-1(2)) in keratinocytes stained with rose bengalPhotochem Photobiol19986867589825697
  • BissonnetteRSharfaeiSViauGIrradiance and light dose influence histological localization of photodamage induced by photodynamic therapy with aminolaevulinic acidBr J Dermatol2004151653515377353
  • BottomleySSMuller-EberhardUPathophysiology of the heme synthesisSemin Hematol1988252823023064310
  • Bouwes BavinckJNEuvrardSNaldiLKeratotic skin lesions and other risk factors are associated with skin cancer in organ-transplant recipients: a case-control study in The Netherlands, United Kingdom, Germany, France, and ItalyJ Invest Dermatol2007127164756 Epub 2007 Mar 2217380113
  • BraakhuisBJTaborMPKummerJAA genetic explanation of Slaughter’s concept of field cancerization: evidence and clinical implicationsCancer Res20036317273012702551
  • BraathenLRSzeimiesRMBasset-SeguinNGuidelines on the use of photodynamic therapy for nonmelanoma skin cancer: An international consensusJ Am Acad Dermatol2007561254317190630
  • CappugiPMaviliaLCampolmiPNew proposal for the treatment of nodular basal cell carcinoma with intralesional 5-aminolevulinic acidJ Chemother200416491315565918
  • ChampionRHRookAWilkinsonDSTextbook of dermatology1998OxfordBlackwell Science
  • ClarkCBrydenADaweRTopical 5-aminolaevulinic acid photodynamic therapy for cutaneous lesions: outcome and comparison of light sourcesPhotodermatol Photoimmunol Photomed2003191344112914598
  • CockerellCJHistopathology of incipient intraepidermal squamous cell carcinoma (“actinic keratosis”)J Am Acad Dermatol200042S1117
  • CurnowAHallerJCBownSGOxygen monitoring during 5-aminolaevulinic acid induced photodynamic therapy in normal rat colon – Comparison of continuous and fractionated light regimesJ Photochem Photobiol B-Biology20005814955
  • DaileyHASmithADifferential interaction of porphyrins used in photoradiation therapy with ferrochelataseBiochem J198422344156497856
  • de GraafYGLKennedyCWolterbeekRPhotodynamic therapy does not prevent cutaneous squamous-cell carcinoma in organ-transplant recipients: Results of a randomized-controlled trialJ Invest Dermatol20061265697416374480
  • de HaasERMKruijtBSterenborgHFractionated illumination significantly improves the response of superficial basal cell carcinoma to aminolevulinic acid photodynamic therapyJ Invest Dermatol200612626798616841035
  • DiamondIGranelliSGMcDonaghAFPhotodynamic therapy of malignant tumoursLancet19722117574117595
  • DivarisDXGKennedyJCPottierRHPhototoxic damage to sebaceous glands and hair-follicles of mice after systemic administration of 5-aminolevulinic acid correlates with localized protoporphyrin-Ix fluorescenceAm J Pathol19901368918972327473
  • DoughertyTJGomerCJHendersonBWPhotodynamic therapyJ Natl Cancer Inst1998908899059637138
  • DoughertyTJKaufmanJEGoldfarbAPhotoradiation therapy for the treatment of malignant tumorsCancer Res197838262835667856
  • DragievaGHafnerJDummerRTopical photodynamic therapy in the treatment of actinic keratoses and Bowen’s disease in transplant recipientsTransplantation2004771152114724445
  • El-FarMGhoneimMIbraheimEBiodistribution and selective in vivo tumor localization of endogenous porphyrins induced and stimulated by 5-aminolevulinic acid: a newly developed techniqueJ Tumor Marker Oncol199052734
  • EricsonMBSandbergCGudmundsonFFluorescence contrast and threshold limit: implications for photodynamic diagnosis of basal cell carcinomaJ Photochem Photobiol B, Biol2003691217
  • EricsonMBSandbergCStenquistBPhotodynamic therapy of actinic keratosis at varying fluence rates: assessment of photobleaching, pain and primary clinical outcomeBr J Dermatol200415112041215606516
  • FijanSHonigsmannHOrtelBPhotodynamic therapy of epithelial skin tumours using delta-aminolaevulinic acid and desferrioxamineBr J Dermatol199513328287547399
  • FosterTHHartleyDFNicholsMGFluence rate effects in photodynamic therapy of multicell tumor spheroidsCancer Res1993531249548443805
  • FosterTHMurantRSBryantRGOxygen consumption and diffusion effects in photodynamic therapyRadiat Res19911262963032034787
  • FrankelDHHanusaBHZitelliJANew primary nonmelanoma skin cancer in patients with a history of squamous cell carcinoma of the skin. Implications and recommendations for follow-upJ Am Acad Dermatol19922672061583171
  • FreemanMVinciulloCFrancisDA comparison of photodynamic therapy using topical methyl aminolevulinate (Metvix) with single cycle cryotherapy in patients with actinic keratosis: a prospective, randomized studyJ Dermatolog Treat2003149910612775317
  • GederaasOAHolroydABrownSB5-aminolaevulinic acid methyl ester transport on amino acid carriers in a human colon adenocarcinoma cell linePhotochem Photobiol200173164911272730
  • GibsonSLCupriksDJHavensJJA regulatory role for porphobilinogen deaminase (PBGD) in delta-aminolaevulinic acid (delta-ALA)-induced photosensitization?Br J Cancer199877235439460994
  • GoldMHBradshawVLBoringMMSplit-face comparison of photodynamic therapy with 5-aminolevulinic acid and intense pulsed light versus intense pulsed light alone for photodamageDermatol Surg20063279580316792644
  • GormanAARodgersMAJCurrent perspectives of singlet oxygen detection in biological environmentsJ Photochem Photobiol B Biol19921415976
  • GrapengiesserSEricsonMGudmundssonFPain caused by photodynamic therapy of skin cancerClin Exp Dermatol200227493712372093
  • GreenAChanging patterns in incidence of non-melanoma skin cancerEpithelial Cell Biol1992147511307938
  • HarriesMJStreetGGilmourEAllergic contact dermatitis to methyl aminolevulinate (Metvix(R)) cream used in photodynamic therapyPhotodermatol Photoimmunol Photomed20072335617254035
  • HendersonBWDoughertyTJHow does photodynamic therapy work?Photochem Photobiol199255145571603846
  • HinnenPde RooijFWMvan VelthuysenMLFBiochemical basis of 5-aminolaevulinic acid induced protoporphyrin IX accumulation: a study in patients with (pre)malignant lesions of the oesophagusBr J Cancer199878679829744510
  • HolmesMVDaweRSFergusonJA randomized, double-blind, placebo-controlled study of the efficacy of tetracaine gel (Ametop(R)) for pain relief during topical photodynamic therapyBr J Dermatol20041503374014996106
  • HornMWolfPWulfHCTopical methyl aminolaevulinate photodynamic therapy in patients with basal cell carcinoma prone to complications and poor cosmetic outcome with conventional treatmentBr J Dermatol20031491242914674903
  • IinumaSFarshiSSOrtelBA mechanistic study of cellular photodestruction with 5-aminolevulinic acid-induced porphyrinBr J Cancer1994702188018536
  • IinumaSSchomackerKTWagnieresGIn vivo fluence rate and fractionation effects on tumor response and photobleaching: photodynamic therapy with two photosensitizers in an orthotopic rat tumor modelCancer Res19995961647010626808
  • JeffesEWMcCulloughJLWeinsteinGDPhotodynamic therapy of actinic keratoses with topical aminolevulinic acid hydrochloride and fluorescent blue lightJ Am Acad Dermatol2001459610411423841
  • KapurNKernlandKBraathenLRPhotodynamic therapy-induced pain: a patient-centred surveyBr J Dermatol2003149478
  • KascheALuderschmidtSRingJPhotodynamic therapy induces less pain in patients treated with methyl aminolevulinate compared to aminolevulinic acidJ Drugs Dermatol20065353616673803
  • KennedyJCPottierRHEndogenous protoporphyrin IX, a clinically useful photosensitizer for photodynamic therapyJ Photochem Photobiol B Biol19921427592
  • KennedyJCPottierRHDukeSORebeizCAACS symposium series: Porphyric pesticides chemistry, toxicology, and pharmaceutical applications1994559Washington D.CAmerican Chemical Society291302
  • KennedyJCPottierRHProssDCPhotodynamic therapy with endogenous protoporphyrin ix. basic principles and present clinical experienceJ Photochem Photobiol B Biol199061438
  • KimHSYooJYChoKHTopical photodynamic therapy using intense pulsed light for treatment of actinic keratosis: Clinical and histopathologic evaluationDermatol Surg20053133715720093
  • KondoMHirotaNTakaokaTHeme-biosynthetic enzyme activities and porphyrin accumulation in normal liver and hepatoma cell lines of ratCell Biol Toxicol19939951058390914
  • KormeiliTYamauchiPSLoweNJTopical photodynamic therapy in clinical dermatologyBr J Dermatol20041501061915214890
  • KurwaHAYong-GeeSASeedPTA randomized paired comparison of photodynamic therapy and. topical 5-fluorouracil in the treatment of actinic keratosesJ Am Acad Dermatol1999414141810459115
  • LehmannPMethyl aminolaevulinate-photodynamic therapy: a review of clinical trials in the treatment of actinic keratoses and nonmelanoma skin cancerBr J Dermatol200715679380117419691
  • LeiboviciLSchoenfeldNYehoshuaHAActivity of porphobilinogen deaminase in peripheral-blood mononuclear-cells of patients with metastatic cancerCancer19886222973003179945
  • LindgrenGLarkoOLong-term follow-up of cryosurgery of basal cell carcinoma of the eyelidJ Am Acad Dermatol19973674269146537
  • LopezRFVLangeNGuyRPhotodynamic therapy of skin cancer: controlled drug delivery of 5-ALA and its estersAdv Drug Deliv Rev200456779414706446
  • MalikZDishiMGariniYFourier transform multipixel spectroscopy and spectral imaging of protoporphyrin in single melanoma cellsPhotochem Photobiol199663608148628752
  • MarksRAn overview of skin cancers - incidence and causationCancer199575607127804986
  • MarmurESSchmultsCDGoldbergDJA review of laser and photodynamic therapy for the treatment of nonmelanoma skin cancerDermatol Surg2004302647114871220
  • McCarronPADonnellyRRZawislakADesign and evaluation of a water-soluble bioadhesive patch formulation for cutaneous delivery of 5-aminolevulinic acid to superficial neoplastic lesionsEur J Pharm Sci20062726827916330192
  • MillerSJBiology of basal cell carcinoma (Part I)J Am Acad Dermatol1991241131999506
  • MoanJSommerSOxygen dependence of the photosensitizing effect of hematoporphyrin derivative in NHIK 3025 cellsCancer Res1985451608103978628
  • MoanJVan den AkkerJTHMJuzenasPOn the basis for tumor selectivity in the 5-aminolevulinic acid-induced synthesis of protoporphyrin IXJournal of Porphyrins and Phthalocyanines20015170176
  • MortonCCampbellSGuptaGIntraindividual, right-left comparison of topical methyl aminolaevulinate-photodynamic therapy and cryotherapy in subjects with actinic keratoses: a multicentre, randomized controlled studyBr J Dermatol200615510293617034536
  • MortonCAPhotodynamic therapy for nonmelanoma skin cancer – and more?Arch Dermatol20041401162014732670
  • OrensteinAKostenichGTsurHTemperature monitoring during photodynamic therapy of skin tumors with topical 5-aminolevulinic acid applicationCancer Lett199593227327621433
  • PagliaroJElliottTBulsaraMCold air analgesia in photodynamic therapy of basal cell carcinomas and Bowen’s disease: An effective addition to treatment: A pilot studyDermatol Surg20043063614692930
  • PariserDMLoweNJStewartDMPhotodynamic therapy with topical methyl aminolevulinate for actinic keratosis: Results of a prospective randomized multicenter trialJ Am Acad Dermatol2003482273212582393
  • PengQBergKMoanJ5-aminolevulinic acid-based photodynamic therapy: Principles and experimental researchPhotochem Photobiol199765235519066303
  • PengQMoanJWarloeTDistribution and photosensitizing efficiency of porphyrins induced by application of exogenous 5-aminolevulinic acid in mice bearing mammary carcinomaInt J Cancer199252433431399120
  • PerrettCMMcGregorJMWarwickJTreatment of post-transplant premalignant skin disease: a randomized intrapatient comparative study of 5-fluorouracil cream and topical photodynamic therapyBr J Dermatol2007156320817223873
  • PiacquadioDJChenDMFarberHFPhotodynamic therapy with aminolevulinic acid topical solution and visible blue, light in the treatment of multiple actinic keratoses of the face and scalp – investigator-blinded, phase 3, multicenter trialsArch Dermatol14041614732659
  • PottierRHChowYFALaplanteJPNoninvasive technique for obtaining fluorescence excitation and emission-spectra in vivoPhotochem Photobiol198644679873809258
  • Radakovic-FijanSBlecha-ThalhammerUKittlerHEfficacy of 3 different light doses join the treatment of actinic keratosios with 5-aminolevulinic acid photodynamic therapy: A randomized, observer-blinded, intrapatient, comparison studyJ Am Acad Dermatol200553823716243131
  • RhodesLEde RieMEnstromYPhotodynamic therapy using topical methyl aminolevulinate vs surgery for nodular basal cell carcinoma – Results of a multicenter randomized prospective trialArch Dermatol2004140172314732655
  • Richards-KortumRSevick-MuracaEQuantitative optical spectroscopy for tissue diagnosisAnnu Rev Phys Chem1996475556068930102
  • RobinsonDJde BruijnHSde WolfWJTopical 5-aminolevulinic acid-photodynamic therapy of hairless mouse skin using two-fold illumination schemes: PpIX fluorescence kinetics, photobleaching and biological effectPhotochem Photobiol20007279480211140268
  • RobinsonDJde BruijnHSStarWMDose and timing of the first light fraction in two-fold illumination schemes for topical ALA-mediated photodynamic therapy of hairless mouse skinPhotochem Photobiol2003773192312685661
  • RobinsonDJde BruijnHSvan der VeenNProtoporphyrin IX fluorescence photobleaching during ALA-mediated photodynamic therapy of UVB-induced tumors in hairless mouse skinPhotochem Photobiol199969617010063801
  • RodriguezLBatlleADi VenosaGMechanisms of 5-aminolevulic acid ester uptake in mammalian cellsBr J Pharmacol20061478253316432502
  • RudEGederaasOHogsetA5-aminolevulinic acid, but not 5-aminolevulinic acid esters, is transported into adenocarcinoma cells by system BETA transportersPhotochem Photobiol200071640710818796
  • SalvaKAPhotodynamic therapy: Unapproved uses, dosages, or indicationsClin Dermatol2002205718112435528
  • SandbergCStenquistBRosdahlIImportant factors for pain during photodynamic therapy for actinic keratosisActa Dermato-Venereologica200686404816955183
  • SitnikTMHamptonJAHendersonBWReduction of tumour oxygenation during and after photodynamic therapy in vivo: effects of fluence rateBr J Cancer1998771386949652753
  • SkiverenJHaedersdalMPhilipsenPAMorphine gel 0.3% does not relieve pain during topical photodynamic therapy: A randomized, double-blind, placebo controlled studyActa Dermato-Venereologica2006864091116955184
  • SolerAMAngell-PetersenEWarloeTPhotodynamic therapy of superficial basal cell carcinoma with 5-aminolevulinic acid with dimethylsulfoxide and ethylendiaminetetraacetic acid: a comparison of two light sourcesPhotochem Photobiol200071724910857368
  • StrasswimmerJGrandeDJDo pulsed lasers produce an effective photodynamic therapy response?Lasers Surg Med20063822516392149
  • SvanbergKAnderssonTKillanderDPhotodynamic therapy of nonmelanoma malignant-tumors of the skin using topical delta-amino levulinic acid sensitization and laser irradiationBr J Dermatol1994130743518011500
  • SzeimiesRMKarrerSRadakovic-FijanSPhotodynamic therapy using topical methyl 5-aminolevulinate compared with cryotherapy for actinic keratosis: A prospective, randomized studyJ Am Acad Dermatol2002472586212140473
  • TappeinerHJesionekATherapeutische versuche mit fluoreszierenden stoffenMünch Med Wochenschr19034720424
  • TarstedtMRosdahlIBerneBA randomized multicenter study to compare two treatment regimens of topical methyl aminolevulinate (Metvix (R))-PDT in actinic keratosis of the face and scalpActa Dermato-Venereologica200585424816159735
  • ThompsonMSGustafssonLPalssonSPhotodynamic therapy and diagnostic measurements of basal cell carcinomas using esterified and non-esterified delta-aminolevulinic acidJournal of Porphyrins and Phthalocyanines2001514753
  • TrombergBJOrensteinAKimelSIn vivo tumor oxygen tension measurements for the evaluation of the efficiency of photodynamic therapyPhotochem Photobiol199052375852145595
  • TschenEHWongDSPariserDMPhotodynamic therapy using aminotaevulinic acid for patients with nonhyperkeratotic actinic keratoses of the face and scalp: phase IV multicentre clinical trial with 12-month follow upBr J Dermatol20061551262917107399
  • TuchinVTissue optics: light scattering methods and instruments for medical diagnosis2000Bellingham, WashingtonSPIE – The International Society for optical Engineering
  • VarmaSWilsonHKurwaHABowen’s disease, solar keratoses and superficial basal cell carcinomas treated by photodynamic therapy using a large-field incoherent light sourceBr J Dermatol20011445677411260016
  • WebberJLuoYCrillyRAn apoptotic response to photodynamic therapy with endogenous protoporphyrin in vivoJ Photochem Photobiol B, Biol19963520911
  • WeishauptKRGomerCJDoughertyTJIdentification of singlet oxygen as the cytotoxic agent in photoinactivation of a murine tumorCancer Res197636232691277137
  • WennbergA-MKeohaneSLearJA multicenter study of photodynamic therapy with methyl aminolevulinate (MAL-PDT) cream in immuno-compromised organ transplant recipients with non-melanoma skin cancerPresented at the 10th World Congress on Cancers of the SkinVienna, Austria
  • WennbergAMLarköOLönnrothPDelta-aminolevulinic acid in superficial basal cell carcinomas and normal skin – a microdialysis and perfusion studyClin Exp Dermatol2000253172210971495
  • WennbergAMLarköOStenquistBFive-year results of Mohs‘ micrographic surgery for aggressive facial basal cell carcinoma in SwedenActa Derm Venereol199979370210494714
  • WennbergAMLindholmLEAlpstenMTreatment of superficial basal cell carcinomas using topically applied delta-aminolaevulinic acid and a filtered xenon lampArch Dermatol Res199628856148919036
  • WiegellSRStenderIMNaRHPain associated with photodynamic therapy using 5-aminolevulinic acid or 5-aminolevulinic acid methylester on tape-stripped normal skinArch Dermatol20031391173712975159
  • VinciulloCElliottTFrancisDPhotodynamic therapy with topical methyl aminolaevulinate for ‘difficult-to-treat’ basal cell carcinomaBr J Dermatol20051527657215840111
  • WulfHCPavelSStenderITopical photodynamic therapy for prevention of new skin lesions in renal transplant recipientsActa Dermato-Venereologica20068625816585985
  • WulfHCPhilipsenPAllergic contact dermatitis to 5-aminolaevulinic acid methylester but not to 5-aminolaevulinic acid after photodynamic therapyBr J Dermatol2004150143514746630
  • YangCHLeeJCChenCHPhotodynamic therapy for bowenoid papulosis using a novel incoherent light-emitting diode deviceBr J Dermatol20031491297914674917
  • ZelicksonBCountersJColesCLight patch: preliminary report of a novel form of blue light delivery for the treatment of actinic keratosisDermatol Surg200531375815841647