biological rationale and clinical experience with hyperthermia

27
ELSEVIER Biological Rationale and Clinical Experience with Hyperthermia Kayihan Engin, MD Department of Radiation Oncology, Thomas Jefferson University, Philadelphia, Pennsylvania ABSTRACT: Hyperthermia (HT) as an adjunct to radiation therapy (RT) has been a focus of interest in cancer management. In recent years there have been numerous randomized and nonran- domized studies conducted to assess the efficacy of HT combined with either RT or chemother- apy especially in the treatment of superficially seated malignant tumors. The major impact of HT is currently on locoregional control of tumor. Heat may be directly cytotoxic to tumor cells or inhibit repair of both sublethal and potentially lethal damage after radiation. These effects are augmented by the physiological conditions in tumor that lead to states of acidosis and hypoxia. Blood flow is often impaired in tumor relative to normal tissues, and HT may lead to a further decrease in blood flow and augment heat sensitivity. Three major areas of clinical investigation have borne the greatest fruit for HT as adjunctive therapy to RT. These include recurrent and primary breast lesions, melanoma, and head and neck neoplasms. Thermal enhancement ratio was increased in all cases and is approximately 1.4 for neck nodes, 1.5 for breast, and 2 for malignant melanoma. In general, the most important prognostic factors for complete response (CR) are RT dose, tumor size and minimal thermal parameters [minimal thermal dose (fl&, mean minimal temperature (T,J or T,, i.e., temperature exceeded by 90% of thermal sensors]. The number of HT fractions administered per week appears to have no bearing on the overall response, which may be indicative of the effects of thermotolerance. The total number of HT fractions delivered also appears irrelevant provided adequate HT is delivered in one or two sessions. The major prognostic factors for the duration of local control were tumor histology, concurrent RT dose, tumor depth and Tmi,. Although numerous single institution studies showed increased CR rates and improved local control, the efficacy of HT as an adjunct to RT should be assessed with well-designed multi-institutional randomized clinical trials. Such clinical trials are underway. Controlled Clin Trials 1996:17:3X-342 KEY WORDS: Thermoradiotherapy, hyperthermia, radiation, biology, oncology, cancer treatment INTRODUCTION In the last two decades there has been a great interest in the use of HT combined with RT (thermoradiotherapy; TRT) in cancer treatment. This is associated with a This article was originally submitted and accepted for publication in Clinical Trials and Me&analysis formerly published by Elsevier Science B.V., Amsterdam. In 1995. Clinical Trials and Meta-analysis was incorporated into Controlled Clinical Trials. Address reprint requests to: Kayihan Engin, MD, Department of Radiation Oncology, Thomas jefferson University, Bodine Center for Cancer Treatment,111 South 11 th Street, Philadelphia, PA 19107-5097. Received September I, 1993. Controlled Clinical Trials 17316-342 (1996) 0 K. Engin 1996 655 Avenue of the Americas, New York, NY 10010 0197-2456/96 SSDI 0197-2456(95)00078-U

Upload: kayihan-engin

Post on 25-Aug-2016

218 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Biological rationale and clinical experience with hyperthermia

ELSEVIER

Biological Rationale and Clinical Experience with Hyperthermia

Kayihan Engin, MD Department of Radiation Oncology, Thomas Jefferson University, Philadelphia, Pennsylvania

ABSTRACT: Hyperthermia (HT) as an adjunct to radiation therapy (RT) has been a focus of interest in cancer management. In recent years there have been numerous randomized and nonran- domized studies conducted to assess the efficacy of HT combined with either RT or chemother- apy especially in the treatment of superficially seated malignant tumors. The major impact of HT is currently on locoregional control of tumor. Heat may be directly cytotoxic to tumor cells or inhibit repair of both sublethal and potentially lethal damage after radiation. These effects are augmented by the physiological conditions in tumor that lead to states of acidosis and hypoxia. Blood flow is often impaired in tumor relative to normal tissues, and HT may lead to a further decrease in blood flow and augment heat sensitivity. Three major areas of clinical investigation have borne the greatest fruit for HT as adjunctive therapy to RT. These include recurrent and primary breast lesions, melanoma, and head and neck neoplasms. Thermal enhancement ratio was increased in all cases and is approximately 1.4 for neck nodes, 1.5 for breast, and 2 for malignant melanoma. In general, the most important prognostic factors for complete response (CR) are RT dose, tumor size and minimal thermal parameters [minimal thermal dose (fl&, mean minimal temperature (T,J or T,, i.e., temperature exceeded by 90% of thermal sensors]. The number of HT fractions administered per week appears to have no bearing on the overall response, which may be indicative of the effects of thermotolerance. The total number of HT fractions delivered also appears irrelevant provided adequate HT is delivered in one or two sessions. The major prognostic factors for the duration of local control were tumor histology, concurrent RT dose, tumor depth and Tmi,. Although numerous single institution studies showed increased CR rates and improved local control, the efficacy of HT as an adjunct to RT should be assessed with well-designed multi-institutional randomized clinical trials. Such clinical trials are underway. Controlled Clin Trials 1996:17:3X-342

KEY WORDS: Thermoradiotherapy, hyperthermia, radiation, biology, oncology, cancer treatment

INTRODUCTION

In the last two decades there has been a great interest in the use of HT combined with RT (thermoradiotherapy; TRT) in cancer treatment. This is associated with a

This article was originally submitted and accepted for publication in Clinical Trials and Me&analysis formerly published by Elsevier Science B.V., Amsterdam. In 1995. Clinical Trials and Meta-analysis was incorporated into Controlled Clinical Trials.

Address reprint requests to: Kayihan Engin, MD, Department of Radiation Oncology, Thomas jefferson University, Bodine Center for Cancer Treatment, 111 South 11 th Street, Philadelphia, PA 19107-5097.

Received September I, 1993.

Controlled Clinical Trials 17316-342 (1996) 0 K. Engin 1996 655 Avenue of the Americas, New York, NY 10010

0197-2456/96 SSDI 0197-2456(95)00078-U

Page 2: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 317

greater understanding of several biological parameters such as radiosensitization, chemosensitization, direct cytotoxicity, thermotolerance, and stepdown heating, as well as of complex changes in the micromilieu involving especially the microvascu- lature and cytoskeleton. Although HT as a new and effective modality shows definite promise in the treatment of cancer, the major challenge is the translation of these benefits to clinical practice. This requires cooperative effort and relies on close interaction of biologists, physicists and oncologists specifically to determine the most important biological parameters affecting HT, to manufacture devices that can heat deeply and homogeneously, to determine temperature reliably, consis- tently and ideally noninvasively. It is obvious that the education of fellow clinicians of the utility of this modality is also very important. This paper presents an overview of the clinical applicationof local HT for superficial lesions. The practical application of regional and whole body HT remains confined to investigational protocols.

HISTORY

The application of HT in the treatment of disease has a long tradition dating to antiquity, but took its modern form only in the last 20 years. Hippocrates is credited with the statement that “un illness not cured by heat is incurable.” The modern era of HT is thus relatively recent and occurred with the advent of new technology allowing the use of radiofrequency, microwave and ultrasound.

In 1866 Busch first observed the effect of systemic HT on possible cancer eradica- tion in human subjects [l]. He reported the disappearance of a sarcoma of the face after a bout of erysipelas. In 1893 Coley reported on 10 patients with advanced malignancies who developed prolonged febrile reactions to erysipelas [2]. Using filtered toxin or mixtures of bacterial toxins, he induced pyrogen therapy by in- jecting the toxins into the tumor itself. In 1909 Schmidt suggested that HT could be added to RT as a radiosensitizer [3].

In the last two decades, there has been a great interest in this modality. The radiobiological principles are firmly established and many clinical studies have shown that the modality is effective [4]. Problems of quality assurance are, however, still unresolved and a Radiation Therapy Oncology Group (RTOG 81-04) study highlighted the problems in the application of HT in a multi-institutional setting 151. These include (1) the difficulties associated with adequate thermometry, (2) the selection of patients with heatable lesions and (3) quality assurance affecting the administration of HT. The challenge nowadays is to prove HT as an effective modality through carefully designed and performed clinical trials.

BIOLOGICAL RATIONALE

Hyperthermia Cytotoxicity The nature of HT-induced cell lethality is quite different from that of RT-induced

killing. The G,-phase of the cell cycle is the most resistant to HT. Cells heated in G, die from membrane-related damage before entering mitosis. S-phase cells are quite sensitive to HT and die of chromosome-related damage after passing through mitosis 16-81. In other words, the cell cycle response to heat complements that of low energy transfer /LET) radiation. Exhibition of cell death in asynchronous populations is much sooner than after RT. HT also sensitizes cells and tissues to

Page 3: Biological rationale and clinical experience with hyperthermia

318

RT. A time-temperature combination can be found that HT alone is without effect on cytotoxicity but will sensitize cells or tissues to RT and the maximum reaction occurs at the time characteristic of RT alone.

Cellular RT-induced killing is mediated through damage to DNA and manifest by chromosome aberrations and loss of reproductive integrity. HT effects are brought about by alteration of proteins and/or lipids [9-101. Incorporating 5-bromodeoxyuridine (BrdUrd) into DNA does not sensitize to HT damage as it does to RT damage [7]. However, HT inhibits DNA synthesis by inhibition of replicon initiation and chain elongation, thus permitting erroneous rejoining of newly synthesized regions of DNA with resultant chromosome aberrations at mito- sis [7,11,12].

Cells are sensitized to HT damage by membrane active agents such as local anesthetics (e.g., lidocaine, procaine, or ethanol) and by low pH and nutritional deprivation [8,10,13-161, whereas the response to HT is not affected by acute hyp- oxia 1171. The cell is protected against HT damage by the presence of deuterium oxide and by glycerol [18,19]. The agents that modify the HT response are those that affect the stability of proteins [9]. HT does not cause cell transformation to malignancy and is only weakly mutagenic [20]. Heated cells experience progression delay much longer than that after ionizing RT for equal levels of cell lethality 16,211. Cells accumulate and repair sublethal damage and potentially lethal damage [22-241. Heat inhibits metabolism and macromolecular synthesis, and alters cy- toskeletal relationships 125-271.

Heat-induced cytotoxicity can be quantitated by plotting cell survival as a func- tion of time at a specific temperature much as for RT dose, and the parameters of a survival curve can be calculated such as Do, n, and d,, or a and B. In-vitro HT survival curves were first demonstrated by Westra and Dewey [6]. For temperatures of <42.5”C, there is a bend in the survival curves at 2-4 hours of exposure due to the development of thermotolerance. When the slope of the HT survival curve (l/ Do) is plotted as a function of the reciprocal of the absolute temperature, the result is called an “Arrhenius plot.” The “break’ in the Arrhenius plot at 42543°C is thought to be due to the development of thermotolerance during exposure to temperature <43”C, and the inhibition of thermotolerance development by expo- sure to temperatures >43”C [21,28]. When cells and tissues return to 37°C after exposure to temperatures >43”C, thermotolerance development occurs over a pe- riod of 8-10 hours and decays over the next 60-100 hours. A typical Arrhenius plot and the effects of modifiers are shown in Figure 1.

Thermal Dose

Since the shape of the Arrhenius plot and the inactivation energy are similar for various cell lines and tissues, it provides a means of comparing the response of cells and tissues to various time-temperature combinations 1291. A comparative unit of response to different time-temperature combinations based on isoeffect relationships could be the basis for a thermal dose. A unit of thermal dose is critical to comparing results from different clinics. Saparetto and Dewey [301 have defined equivalent minutes at 43°C (tti; min E+& by the formula: Eq43”C = tR(43 - 7’) where t = time, T = temperature, and R = 0.5 for T >43”C and R = 0.25 for T (43°C.

Page 4: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment

4

41 42 43 44 45

Temperature (0C)

Figure 1 A typical Arrhenius plot and the effects of modifiers.

Thermotolerance

319

Thermotolerance is a nonheritable resistance to HT induced by exposure to HT and other cytotoxic agents such as arsenic, adriamycin, or ethanol [31,32]. Thermotolerance develops within 2-3 hours during exposure to temperatures <43”C, but not during exposure to temperatures ~43°C. When the cell is returned to 37°C after exposure to temperatures >43”C, thermotolerance develops over the next B-10 hours. The amount of thermotolerance that develops depends on the temperature and the amount of HT damage induced. Thermotolerance decays over the next 60-100 hours depending on the cell line or tissue and the amount of thermotolerance induced [33]. Thermotolerance leads to an increase of the Do of the survival curve by a factor of 4-10 which can result in a difference of cell survival from 10e5 in normotolerant cells to 10-l in maximally thermotolerant cells. Li and Hahn 1281 proposed a model of thermotolerance incorporating an induction phase, a development phase, and a decay phase. Each of these components may have its own temperature dependence as well as dependence on other factors such as pH and nutrients. Induction and decay are permissive at all temperatures, whereas development is permissive at temperatures ~43°C and restricted at temperatures 243°C.

The inhibition of thermotolerance development at temperatures >43”C is due to HT-induced inhibition of protein synthesis [26,34]. Heat shock proteins (HSPs) are the first proteins synthesized in response to HT and the synthesis of HSPs is intimately related to the development of thermotolerance 135-371. The HSPs found in most mammalian cells have molecular weights of 22, 26, 70, 87, and 110 kD. The decay of the three heavier HSPs correlates with the decay of thermotolerance and a good correlation exists between the residual levels of HSP 70, 87, and 110 and cell survival during the decay of thermotolerance 1381. Tissue or tumor levels of HSPs may provide a prognostic indicator of HT resistance and the synthesis of HSPs could provide a monitor of tissue thermotolerance development after HT.

Page 5: Biological rationale and clinical experience with hyperthermia

320 K. Engin

Thermotolerance develops in normal tissues after single or multiple HT fractions similarly to cells in culture 139,401. Since maximal thermotolerance occurs by 24 hours, daily fractionation would completely waste any cumulative effect of HT. All experimental normal tissues studied to date develop thermotolerance and tu- mors are no exception [32]. There is no reason to suspect that human normal tissues and well-perfused tumor cells are not perfectly capable of developing thermotoler- ante just as in murine tissues. Three clinical studies indicated that heating with two HT fractions per week combined with RT was no more effective than heating with one fraction per week [41-431. It is possible that thermotolerance could be used to influence the HT response of normal vs. tumor tissues. The magnitude of thermotolerance is reduced at low pH. Thermotolerance may decay faster in cycling than in noncycling cells, and it may be possible to protect transit normal tissues by inducing thermotolerance in them by heating with poorly penetrating micro- waves 24 hours before heating the tumor [44,45].

Stepdown Heating

Sensitization of cells to exposures to temperatures below 43°C after exposure to temperatures 243°C for a brief period is called “stepdown heating” (SDH), which results from the inhibition of thermotolerance development [28,4648]. Although temperatures of 3941°C have very little effect on cell lethality, 15 min exposure of at 45°C before exposure to 3941°C makes these temperatures quite cytotoxic.

PH Poor tumor blood flow, chronic hypoxia, and high lactic acid levels are commonly

observed in tumors and lead to decreased extracellular pH(pHJ [49-521. Although there have been many reports on pH, in animal tumors, data on human tumor pH, have been relatively scarce [53-571. A wide range of measurements has been reported; 5.80-7.52 in rodents and 5.85-7.68 in human tumors [58]. Engin et al. showed that the tumor volume and the tumor histology were the most important factors determining the range of pH,. values in human tumors 1591. The mean tumor pH,. for the entire group of tumors was 7.06 -t 0.05 (range of 5.66-7.78). The mean tumor pH, was 6.93 2 0.08 in adenocarcinomas, 7.01 2 0.21 in soft-tissue sarcomas, 7.16 + 0.08 in squamous cell carcinomas, and 7.36 + 0.1 in malignant melanomas. Four histological groups were significantly different from each other. When histolo- gies were grouped together, the mean pY was 6.94 2 0.08 in adenocarcinoma and soft tissue sarcoma lesions vs. 7.20 + 0.07 in squamous cell carcinoma and malignant melanoma lesions. The difference in pw between squamous cell carcinoma plus malignant melanoma and adenocarcinoma plus soft-tissue sarcoma was statistically significant. This finding corresponds with a previous report by Engin et al. indicat- ing that the duration of local control after TRT was longer in adenocarcinomas and soft tissue sarcomas than in squamous cell cancers and malignant melanomas [60]. These observations further support the conclusion that tumor pH, is inversely correlated with CR rate after TRT [61].

Busse et al. 1621 showed that as murine transplantable tumors increased in size from 2.1 to 10.5 g, the mean pH fell from 6.7 to 6.5. When tumors were very necrotic the pH increased with increasing size rising from 7.2 to 7.4 as tumors increased from 1.7 to 25.7 g. Vaupel et al [63] also found alkaline values in very necrotic

Page 6: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 321

tumors and more acid values in large ulcerating tumors. Gullino et al [64] were not able to show any difference between small and large tumors. They also found that faster growing tumors were more acidic. Engin et al showed that there was a weak but significant correlation between tumor pH, and tumor volume [591. Tumor pH, increased as a function of logarithm of tumor volume.

Variation in measurements between tumors was greater than the variation in measurements within tumor. One of the main causes for intratumoral variation is probably the proximity of the electrode tip to blood vessels. Greater distance from the blood supply results in lower pH due to restricted oxygen availability.

Although the experimental literature suggests that an increase in the acidity of tumor tissue should enhance the therapeutic effect of HT, two clinical studies were unable to demonstrate that low pH indicated a favorable response to TRT in human tumors [57,65]. On the contrary, Engin et al. found a significant correlation of the pretreatment tumor pH, with tumor sensitivity to TRT [61].

Tumor cells at suboptimal PI-I may not be HT sensitive as suggested by Hahn and Shiu 1661. Chu et al [67] and Cook and Fox [68] have shown that cells adapt to an environment at low pH,, reestablish a near normal pH, and are no longer HT-sensitive. pH, is the important factor for HT sensitization [67,69,70]. However, when cells are adapted to low pH,., a further slight reduction in pH, by as little as 0.2 pH unit may have a similar effect as a 0.8 pH unit decrease in cells at physiologic pH [66]. Therefore, an acute reduction of tumor pH, as induced by hyperglycemia may provide a therapeutic advantage for TRT. Experimental results have shown that acute reduction of pH,. to below 7.0-7.2 sensitizes cells to HT damage [44,66,71] and inhibits the development of thermotolerance ]58,72]. Acute acidification also inhibits cell proliferation and decreases glycolysis and DNA synthesis [73]. How- ever, little evidence exists that these factors affect the human tumor response to TRT. One of the possibilities for increasing the therapeutic ratio in the clinic is the manipulation of pH, by inducing hyperglycemia [74].

Blood Flow

The neovasculature in tumors does not respond to increased temperatures as do blood vessels in normal tissues, and these differences in blood flow may lead to selective tumor heating. Song et al [75] have shown that rat muscle and skin have a greatly enhanced blood flow at temperatures above 42°C whereas transplantable tumors have a greatly reduced blood flow at these temperatures. On the other hand, Waterman et al [76] have shown that human tumors exposed to temperatures of 43.5-45”C have a slightly increased blood flow during heating for 60 min, but nonetheless, tumor blood flow does not respond to HT with the significant increase as seen in normal tissues. Interestingly, Kang et al 1771 have shown in transplantable mouse tumors that vascular volume decreases to a minimum between 4 and 6 hours postheating at 43.5”C; and as might be expected, pH and pOz also fall. Recovery of vascular volume and blood flow occurs between 24 and 48 hours postheating. If such a decrease in blood flow did occur after heating in human tumors, it would certainly dictate the timing of a subsequent HT treatment.

Radiosensitization

In addition to HT-induced cytotoxicity, HT sensitizes cells to low LET ionizing radiation. Unlike the response to ionizing RT, HT cytotoxicity is influenced by

Page 7: Biological rationale and clinical experience with hyperthermia

322 K. Engin

thermotolerance, low pH, and nutritional deprivation but is independent of acute hypoxia. Blood flow also influences the heating characteristics of a tumor relative to normal tissue. Thermal radiosensitization is also influenced by these parameters, and they may affect the response of human tumors to the combined modalities. The cell cycle-dependent sensitivities to HT and RT complement each other [6,78]. The predominant effect of temperatures a43”C on the radiation response is sensiti- zation, i.e., a decrease in D,,; whereas the predominant effect of exposure to tempera- tures <42.5”C for an hour or less is removal of the shoulder of the radiation survival curve [48]. Modest HT also inhibits the repair of sublethal and potentially lethal radiation damage and inhibits the repair of radiation-induced DNA damage [79-811. The mechanism of HT radiosensitization is protein denaturation resulting in changes in DNA-protein structural relationships and preventing accessibility of repair and synthetic enzymes [ll].

Interaction between HT and RT is dependent on sequence and the time between exposures. The greatest interaction results when irradiation occurs in the middle of the heating interval 1821. In the clinical setting RT is generally followed by heating in order to prevent any possible effect of HT inducing hypoxic cells via decreased blood flow. Theoretically, there should be less radiosensitization by HT with low-dose RT fractions or high LET radiation, since most of the damage consists of single-hit, irreparable lesions [83,841. Gerner and Leith 1851 demonstrated that 43°C for 30 min does not sensitize CHO cells to high LET radiation. In fact, HT plus low LET radiation has been called “a poor man’s high LET.” Clinically, the greatest effect of thermoradiosensitization is expected the higher the HT dose, the larger the RT fraction, and the closer in time the two are together.

Denekamp and Stewart [86] and Overgaard [87] examined the effect of sequence and interval between RT and HT on mouse tumors and skin. For the sequence of RT followed by HT, the repair of RT damage in skin was complete by 24 hours, and the thermal enhancement ratio (TER) fell to 1; whereas between 4 and 8 hours, the TER for tumor remained at 1.5. Therefore, the therapeutic gain factor, when using large RT doses and HT and comparing the response of tumor to that of skin at the same temperature, was enhanced by separating RT and HT by a 4-hour interval compared to simultaneous treatment.

Hypoxic Cell Sensitizers

HT dramatically enhances the cytotoxicity of the electron affinic radiosensitizers in hypoxic cells 1881. Particularly noticeable is the enhancement by temperatures that by themselves are nontoxic. The combination of misonidazole and HT in vivo shows only a small beneficial effect. However, when HT and misonidazole are combined with RT, there is a marked therapeutic gain. Falk et al. have used metroni- dazole in combination with HT and chemotherapy 1891. Patients treated without metronidazole showed little or no tumor response, whereas the inclusion of an oral dose of 1.5 g metronidazole just prior to heating caused regression of some tumors. Metronidazole apparently caused up to a 2.5”C rise in the temperature of the heated tumors. The reason is not clear, but could be related to the known ability of nitro compounds to cause a breakdown of temperature regulation in vivo [90]. If these results are confirmed, they could have important implications for the future use of nitro compounds with HT [91].

Page 8: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 323

Chemosensitization

HT sensitizes the cell to many cytotoxic agents and even converts some drugs that are innocuous to highly toxic [92-941. HT chemosensitization may occur by an increased reaction rate as for the alkylating agents, increased permeability as for the intercalating antibiotics such as adriamycin, or decreased repair as for bleomycin. Some compounds that are relatively nontoxic at 37°C have greatly enhanced cytotoxicity at 43°C. Two examples are the polyene antibiotic amphoteri- tin B and the sulfhydral radioprotectors such as mercaptoethylamine (MEA) or AET [S-(2-aminoethyl) isothiouronium bromide hydrochloride]. The metabolic inhibitors such as 5-fluorouracil (5-FU) seem not to be enhanced by HT. The most promising chemosensitization by HT would seem to be with alkylating agents and cis-platinum since they are enhanced at all elevated temperatures. Bleomycin, on the other hand, seems to have an enhancement threshold at or about 43°C. The intercalating antibiotics such as adriamycin and actinomycin-D do not appear to be good candi- dates for thermochemotherapy because of HT-induced drug resistance that may be related to thermotolerance. Simultaneous treatment by heating and by drugs was found to be much more effective in inducing tumor growth delay than when HT and chemotherapy were separated by 24 hr. Although there is precious little clinical experience with HT and chemotherapy, the in-vitro and in-vivo data suggest that thermochemotherapy could be a very useful treatment modality.

CLINICAL RATIONALE Locoregional failure remains a significant cause of cancer death as well as of

considerable patient morbidity even when the the problem of disseminated metasta- ses is more life threatening [95]. The major impact of HT is currently on locoregional control of tumor. Heat may be directly cytotoxic to tumor cells or inhibit repair of both sublethal and potentially lethal damage after RT 179,961. These effects are augmented by the physiological conditions in tumor that lead to states of acidosis and hypoxia. Blood flow is often impaired in tumor relative to normal tissue, and HT may lead to a further decrease in blood flow and augment HT sensitivity [97,98]. At lower, more realistic temperatures (~43”C), blood flow is not affected in human tumors by HT as it is in rapidly growing experimental tumors 1991.

HT-induced radiosensitization is a general phenomenon but, in particular, cells in the radioresistant S-phase are sensitized to the greatest extent to RT by HT [6,100,101]. More recently, the interaction of HT and chemotherapy has gained increasing interest 11021. HT may increase the cytotoxic effects of certain chemother- apeutic agents, e.g., the alkylating agents, and help to overcome acquired drug resistance 11031. Furthermore, it can improve drug delivery to the tumor and possibly modify drug pharmacokinetics [104].

CLINICAL EXPERIENCE

Randomized Studies

Despite the fact that more than 11,000 patients had already been treated with HT by 1984 141, there have been few prospective randomized multi-institutional studies in HT, and to date a definitive study is still lacking. The RTOG reported on a randomized study [5] of superficial recurrent tumors that were treated with

Page 9: Biological rationale and clinical experience with hyperthermia

K. Engin

Table 1 Randomized Trials of Hyperthermia Adjuvant to the Radiation Therapy Number of

Reference Type of Tumors Patients Results Perez et al 151 Misc. 300 RT alone; CR = 28%

TRT; CR = 32% Arcangeli et al [106] Misc. 123 RT alone; CR = 42%

TRT; CR = 73% Valdagni et al [108] Head and neck 44 RT alone; CR = 41%

TRT; CR = 83% RT alone; 5-year nodal control = 24% TRT; 5-year nodal control = 69%

Overgaard et al [109] Melanoma 134 RT alone; 2-year local control = 28% TRT; 2-year local control = 46%

Vernon et al [110] Breast N300 RT alone; CR = 40% TRT; CR = 60%

either RT alone or TRT (see Table 1). The positive feature of this study was that lesions that were ~3 cm in diameter and had received 22 sessions of 42.5”C HT showed an improvement in CR in the category of breast and extremity lesions. There was also a greater probability of maintaining a response over 12 months (82% vs. 12%) in those receiving HT, indicating a more durable CR. A prospective randomized study was initiated by Valdagni et al. who treated N3 neck disease arising from TI.z primary lesions or unknown primary sites [105]. Although the number of patients accrued is small, the trial was closed prematurely for ethical reasons. A CR was noted at 3 months in 82% of patients receiving TRT compared to 37% who were treated with RT alone. However, the difference in total response (complete and partial response) is not statistically significant (88% vs. 79%). This indicates that many of the patients achieving a partial response with RT were converted to complete responders by the HT. A clear distinction between clinical tumor regression and overall tumor control must also be made. HT appears to augment tumor control even though initial tumor response may not be significantly improved, which suggests that the clonogenicity of residual viable cells is affected by HT despite the absence of outright cell death. Engin et al. [42] showed that RT does and tumor volume, but not tumor temperature, were effective in achieving a CR (i.e., first one to two decades of cell kill), whereas the duration of local control (total cell kill) was more influenced by tumor temperature and tumor volume than by RT dose. Earlier work by Arcangeli et al. [106,107] involved a series of 77 patients with two or more superficial lesions, treated with various protocols in separate randomized studies. The best therapeutic effect was obtained by restricting the heat to the tumor site and using large fraction sizes and immediate heat, a schedule designed to produce maximal cytotoxicity. They also observed that tumor control was improved by the use of a larger number of HT fractions and increased thermal dose (tu, minE+&.

Recently, Valdagni and Amichetti [lo81 reported on inoperable metastatic lymph nodes in patients with head and neck cancer, which were randomized to receive conventionally fractionated radical radiation therapy alone or combined with local hyperthermia. The addition of hyperthermia significantly improved early response rate and 5-year actuarial nodal control in the absence of severe toxicity.

Page 10: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 325

Table 2 Effect of Hyperthermia Adjuvant to the Radiation Therapy on Clinical Tumor Response

Reference RT Alone” TRT* TER’

Kim et al [112] 39 72 1.8 Overgaard et al [113] 39 62 1.6 Arcangeli et al [107] 38 74 2.0 Scott et al [114] 39 87 2.2 Li et al [115] 29 54 1.9 Hiraoka et al [116] 25 71 2.8 Valdagni et al 11171 36 73 2.0 Perez et al [118] 41 69 1.7 Valdagni et al [105] 36 73 2.0 van der Zee et al [119] 5 27 5.4 Steeves et al [120] 31 65 2.1 Dunlop et al [121] 50 60 1.2 Uozumi et al [122] 63 88 1.3 Hornback et al 11231 35 72 2.1 Fuwa et al [124] 63 83 1.3 Gonzalez et al [125] 33 50 1.5 Lindholm et al [126] 25 46 1.8 Goldobenko et al [127] 86 100 1.2 Muratkhodzhaev et al [128] 25 63 2.5 Datta et al [129] 46 66 1.4

* RT alone: radiation therapy alone (Lib). b TRT: thermoradiotherapy (%). ’ TER: thermal enhancement ratio.

Preliminary results of two major European multi-institutional studies appear to be very promising [109,110]. In ESHO protocol 3-85 [109], 134 metastatic or recurrent malignant melanoma lesions were randomized to receive radiotherapy alone (3 fractions in 8 days; 3 X 8 Gy or 3 X 9 Gy) or each fraction followed by hyperthermia. Two-year actuarial local tumor control was 28% in radiation alone group and 46% in thermoradiotherapy group (p = 0.008). Cox multivariate regression analysis showed that the most important prognostic parameters were hyperthermia (p = 0.02), tumor size (p = 0.05) and radiation dose (p = 0.05). Addition of hyperthermia did not significantly increase the acute or late radiation reactions. In a collaborative phase III superficial hyperthemia trial in primary and recurrent breast cancer lesions [llO], conducted by several institutions (MRC, ESHO, PMH), it was indicated that complete response rate could be increased from 40% (radiation alone) to 60% with the addition of hyperthermia. The difference was statistically significant.

Nonrandomized Studies

In multiple nonrandomized single-institution studies, the benefit of HT has been clearly demonstrated and is remarkably consistent in showing an advantage in favor of HT for all histologies in all sites tested [ill]. A therapeutic gain is suggested in spite of the wide diversity of techniques used for heating, thermometry, and criteria for assessment of response. The thermal enhancement ratio ranged between 1.2 and 5.4 (see Table 2). Several studies of matched and paired lesion analysis in the same patient similarly confirmed the therapeutic benefit of HT 1120--1261. The

Page 11: Biological rationale and clinical experience with hyperthermia

326

Table 3 Complete Response Rate as a Function of Tumor Histology (%) No. of

Author Lesions SCC” MMb AC‘

Bicher et al [130] 82 37 47 73 Corry et al [131] 21 100 64 57 Kim et al [112] 86 82 90 Hiraoka et al [116] 33 50 67 43 Hofman et al [132] 61 83 67 30-61* Luk et al [133] 63 61 57 40 Matsuda et al [134] 20 83 64 Bicher et al 11351 125 49 33 64 Gonzales et al 11251 88 50 50-lo@ Arcangeli et al [136] 59 79 76 Howard et al [102] 13 100 80 100 Perez et al [137] 199 58 79 46 Valdagni et al [105] 101 50-78’ 75 58 Engin et al [6O]~y <3 cm depth 76 81 50 71

>3 cm depth 50 17 9 38

’ SCC: squamous cell carcinoma.

‘MM: malignant melanoma.

’ AC: adenocarcinoma.

’ Inoperable breast cancer: CR = 30%; chest wall recurrence: CR = 61%.

’ Inoperable breast cancer: CR = 100%; chest wall recurrence: CR = 50%.

‘Palliative RT: CR = 50%; radical RT: CR = 78%.

r Soft tissue sarcoma; <3 cm depth; CR = SO%, >3 cm depth; CR = 0%.

difficulty of uniformly heating apparently favorable, superficial sites in areas where problems of contour and shape are minimal has been well described. Dunlop et al [121] reported that 58% of the patients treated with TRT reached a minimum dose of 20 minutes at 43”C, and 24% reached 60 rninEqUoc and could achieve 20 minegUc at each occasion. However, 23% of tumors could not be heated effectively at any time. The tumors that were effectively heated had a CR rate of 86% vs. 35% for those inadequately heated. The CR rate of lesions receiving RT alone was also 35%. This study emphasizes the need to heat to minimum threshold temperatures and to improving intratumoral temperature distribution. Additional difficulties in heating capability arise from the site being treated. The head-and-neck region is more difficult to heat because of several factors, including changing contour and poor applicator apposition, the extent and depth of the tumor and the tolerance of the patient.

PROGNOSTIC FACTORS Three major areas of clinical investigation have borne the greatest fruit for HT

as adjunctive therapy to RT. These include recurrent and primary breast lesions, melanoma, and head-and-neck neoplasms. The general experience of HT with regard to various prognostic factors is summarized in Tables 3-5. The TER was increased in all cases and is approximately 1.4 for neck nodes, 1.5 for breast, and 2 for malignant melanoma. Because of the apparent lack of histological specificity to thermosensitivity, sarcomas and other traditionally radioresistant tumors are

Page 12: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 327

Table 4 Clinical Tumor Response as a Function of the Number of Total and Weekly Heat Treatments

No. of Total No. of No. of Weekly CR Author Randomized Lesions HT Sessions HT Sessions (%o)

Arcangeli et al [107] No

Hiraoka et al 11161 NO

Kim et al [112] No

Valdagni et al [105] No

Engin et al [42] Yes

Arcangeli et al [136] No

Kapp et al [43] Yes

Emami et al [41] Yes

23 5 1 64 10 2 78

40 2-7 2 50 8-12 2 53

50 6 1 74 10 2 59

<6 2 or 3 83 6 2 or 3 50

>6 2 or 3 71 44 4 1 59

8 2 55 21 5 2 75

8 2 77 179 2 1 68

6 2 63 240 4 1 55

8 2 58

also eminently suitable for this modality, provided they are accessible and of a size that is heatable. 11501 In general, the most important prognostic factors for CR are RT dose, tumor size and minimal thermal parameters [minimal thermal dose (t8min), mean minimal temperature (T,,,), or T,, i.e., temperature exceeded by 90% of thermal sensors], Table 4). The number of HT fractions administered per week appears to have no bearing on the overall response, which may be indicative of the effects of thermotolerance (Table 3). The total number of HT fractions delivered also appears irrelevant provided adequate heat is delivered in one or two sessions. The administration of multiple fractions of HT is advantageous only in leading to a greater probability of achieving at least one good HT session and of allowing full coverage of the tumor if a “patchwork’ approach is adopted [151]. The major prognostic factors for the duration of local control were tumor histology, concurrent TR dose, tumor depth, and T,, [42,43,60,143,149,152].

CRITERIA FOR SUITABILITY FOR SUPERFICIAL LOCALIZED HEATING

The selection of the sites and types of lesions that are appropriate for this approach have been described by Kapp 11531. The most important indication for HT is the determination that a poor response is likely if RT alone was to be applied. This would apply to radioresistant histologies, recurrent disease, or excessive tumor bulk. The combination of HT and RT is an attempt to improve the therapeutic ratio and is akin to surgery and RT in being a form of localized combined modality treatment. The most important criterion is accessibility both in size and depth of lesion. Flat surfaces such as the chest wall and the limbs are ideally suited for external applicator HT. The depth of the lesion is critical and for practical purposes is limited to a maximum of 3-4 cm with 915 MHz microwave and 7-8 cm if 1 MHz ultrasound is used. The presence of bone and air-tissue interfaces in the treatment volume precludes the use of ultrasound. Cystic lesions may be heated to a greater

Page 13: Biological rationale and clinical experience with hyperthermia

K. Engin

Table 5 Thermal Parameters that Correlate with Complete Response Rate Author Thermal Parameter

Dewhirst et al 11381 Dewhirst and Sim 11391 Hiraoka et al [116] Luk et al [133] Oleson et al 11401 Sim et al 11411 Van der Zee et al 11191

Arcangeli et al [142] Kapp et al [43,143] Dewhirst et al [138]

Sapozink et al [144] Van der Zee et al [145]

Dunlop et al 11211

Arcangeli et al [136] Storm et al [97] Seegenschmiedt et al 11461

Emami et al 11471 Scott et al 11141 Oleson et al 11481 Engin et al [42,149]

Minimum Eq43”C at first treatment Average minimum Eq43”C Average tumor center temperature Lowest daily average temperature Minimum tumor temperature Estimated treated volume (ETV) obtaining >42.5”C Mean temperature Mean of maximum temperature Average overall Eq42.5”C

% temperature <4O”C % temperature <42.5”C (first treatment) Minimum temperature Average temperature % temperature <4O”C % temperature >42.5”C Average all treatments minimum Eq43”C

No. of HT with any tumor temperature >42”C Mean minimum T mean Mean minimum T max

Total muumum tusc No. of HT of Tmin > 20 minEqooc

Eq42.5”C Lowest tumor temperature/first treatment Eq43”C/first treatment Cumulative Eq43”C for all treatments Total/mean minimum Eq43”C

>400 min Eq43”C Tave > 43°C Tedge plus slope (dT/dr) outer shell Mean minimum temperature Thermal dose/first session

depth because of increased penetrability of energy and transmission of heat. In general, patients should have a reasonable performance rating, sufficient to with- stand a 60-minute session of heat therapy, and should have a life expectancy of 22 months for the response to be monitored.

CARCINOMA OF BREAST

Locoregional failure in the breast still constitutes an important cause of morbidity and suffering, even when overshadowed by the problem of distant metastases. Even after radical or modified mastectomy the local recurrence rate is 1040% depending on size, tumor characteristics of the primary, underlying fixation, the presence of grave signs, and the number of regional lymph nodes involved [154].

There are few randomized studies of HT in chest wall disease. At least 40% of the 236 patients evaluated in the RTOG study (U-04) had carcinoma of the breast [5]. Although the overall result for all patients showed no significant difference between the two arms (30% and 32%, respectively), 62% of patients with lesions ~3 cm had a CR vs. 40% with RT alone. Besides size, RT dose was also critical.

Page 14: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 329

Van der Zee et al [155] treated 97 patients with recurrent carcinoma of the breast in combination with HT. Based on this experience, the authors suggested that a total dose of 32 Gy in 8 fractions administered with twice-weekly fractionation and accompanied by HT is safe and efficacious. Engin et al [151] reported excellent results by using patchwork HT fields in extensive chest wall recurrences due to advanced carcinoma of breast.

Although most studies have concentrated on locoregionally recurrent lesions, advanced primary breast carcinoma is also an appropriate situation for evaluating the role of HT, especially when the presence of grave adverse prognostic features makes local control distinctly remote. Results with RT alone are poor, with 4565% of tumors recurring locally [156]. Vora et al [157] treated 11 patients with inoperable and inflammatory carcinoma of the breast with a combination of external and interstitial RT and interstitial HT. With a follow-up of 11-36 months, local control was obtained in 80% of patients. Similarly, Puthawala et al [158] reported 88%, local control with RT using combined external and interstitial RT to total doses of 80-90 Gy and accompanied by interstitial HT.

The role of HT in an adjuvant setting where there is no gross evidence of residual disease is undefined. Because of the presence of large amounts of normal tissue and minimal tumoral elements, there would seem to be little indication for HT. However, this must be viewed in context, and factors of patients’ age and general condition must be considered as well as the risk of recurrence with conventional RT. Kapp et al [159] recently reported their experience with TRT for residual microscopic cancer.

Various attempts to address the problem of tumor thickness and extent on heatability have included the use of multiple applicators, overlapping patchwork HT [151], combined interstitial and external applicator techniques to heat more deeply and uniformly 11601, and the use of multiple separately powered heating elements, which can be tailored to encompass wide surface areas [161].

MALIGNANT MELANOMA

Local control is an important determinant of length of survival in this disease even though distant metastases is the major cause of death. Overgaard showed that local control was improved by locoregional HT [162]. He also found that the RT fraction size was an important prognostic factor. Emami et al [147] reported on 49 superficial recurrent, primary and metastatic malignant melanoma lesions [147]. They found that a thermal dose of 3500 minhganc was associated with a 100% CR rate. Engin et al [149] observed a similar correlation between CR and the minimal thermal dose of first HT session. In a large study, Overgaard and Overgaard [163] reported an improvement in locoregional control in patients receiving HT. The EDSo (RT dose required for 50% control) was reduced by HT with a TER of 1.43 for simultaneous treatment and 1.24 for sequential treatment, i.e., RT followed 4 hours later by HT. Tumor control was also more durable with HT. The rate of persistent local control at 18 months was increased by the HT (86% vs. 56%). Although a high TER was obtained with simultaneous treatment, normal skin tissue reaction was also increased, thus yielding no improvement in therapeutic ratio. With sequential treatment, i.e., HT given at least 4 hours after RT, normal tissue sensitization did not occur, resulting in a therapeutic gain of 1.22. The favorable impact on larger tumors was reported earlier by Kim et al [164] which

Page 15: Biological rationale and clinical experience with hyperthermia

330 K. Engin

was attributed to greater heatability of tumors to minimum tumor temperatures of 42°C. RT dose per fraction was demonstrated to be critically important in tumor control. Smaller tumors treated with high doses per fraction showed no beneficial effect of HT. With larger tumors, HT increased the rate of response and accelerated tumor regression with both large and small RT doses per fraction. HT appeared to aid in overcoming the adverse effects of large volume disease an the use of low RT fraction sizes. It was associated with a TER of 1.5.

HEAD AND NECK TUMORS

Despite the aggressive use of combined modality therapy in the head-and-neck region, locoregional disease is the most important mode of failure. HT has been applied to the head-and-neck region mainly for metastatic deposits. In most cases, because of depth of tumor, the primary site cannot be adequately heated with currently available external equipment, and an interstitial implant is required for HT to be effective. This region presents considerable technical challenge because of rapidly changing contours and difficulties with coupling the external applicator to the tumor. Lesions are frequently extremely large and irregular in shape and cannot be well encompassed by available external applicators. Often the tumor is endophytic, extending deeply to the parapharyngeal space and base of the skull region, or situated deeply beneath muscle and other normal tissue.

Primary Treatment

Arcangeli et al, [106,142] have reported on the use of multiple daily RT fractions and HT on N2-N3 multiple cervical lymph nodes. Matched lymph nodes were used in the same patient. A higher CR rate was obtained for the combined treatment compared to hyperfractionation alone or to a historical series of conventionally fractionated RT. A TER of 1.88 was obtained. When thermal dose was above 200 mmEqQ.PC, all responses were complete irrespective of volume. The combined use of hyperfractionation and HT in primary disease of the head and neck in a nonran- domized trial setting also makes this experience distinct. In a prospective random- ized study with N3 disease arising from T,-TZ lesions or unknown primaries, Val- dagni et al [105] reported that the results were highly significant in favor of TRT.

Recurrent Tumor

It is clear that CR rates in the order of 50-85% may be obtained in advanced nodal disease in the head and neck with the combination of high-dose RT and HT. The effectiveness of standard RT may thus be improved without increased serious morbidity. This approach may help certain patients with advanced tumor to over- come the prognostic implications of inoperability by producing sufficient tumor shrinkage to allow resection to be performed. The major constraint is the ability to heat tumor adequately. The major limitation at this time is largely technical and related to the difficulty in delivering heat sufficiently deeply and to the whole tumor. In attempts to treat large nodal masses that are recurrent after full conven- tional therapy, a combination of interstitial and external applicator HT has been applied [160]. This approach allows catheters that are inserted into the base of the

Page 16: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 331

lesion to be used for interstitial heating and interstitial RT as well as for thermome- try. Treatment of such masses presents considerable technical difficulties due to the greater medial extension than is appreciated clinically. For this reason, CT scanning is crucial in planning the placement of the temperature catheters [165]. Interstitial HT in combination with external beam and interstitial RT has also been applied to advanced and locally recurrent primary disease [166].

SKIN REACTIONS Several randomized and nonrandomized studies comparing RT alone with the

combination of RT and HT have shown that skin reactions were not increased significantly by the addition of HT. Gonzalez Gonzalez et al 11251 observed enhance- ment of acute skin reactions although the enhancement did not constitute a clinical problem. Lindholm et al 11261 reported that local pain and normal tissue reactions were. a problem with 2450 MHz microwaves but were reduced when 915 MHz microwaves were used. Gabriele et al [167] treated 57 patients with HT alone. They observed skin burns in two cases (3%). Other side effects were blisters with moist desquamation (10%) and local infection after invasive procedures of thermometry (13%). In their randomized study of the effects of TRT on spontaneous tumors in pets, Dewhirst and Sim [139] found that the incidence of direct thermal injury to skin was positively correlated with maximum intratumor equivalent minutes at 43°C. They also concluded that little or no enhancement of early or late radiation effects occurred with the addition of HT. In studies on human tumors, Luk et al [168] showed a positive correlation between the incidence of burns and blisters and the intratumoral temperature maxima, Howard et al [169] showed that the average maximum skin HT dose per treatment correlated with the incidence of severe skin reactions, and Seegenschmiedt et al [146] showed a correlation between total mean and maximum tumor Eq43 minutes and acute complications. Kapp et al [170] showed that the average of the maximum tumor temperature and the number of HT treatments per field were the most important predictive factors for complications in TRT. Engin et al [171] showed that the presence or absence of previous RT, previous RT dose, concurrent RT factors, i.e., concurrent RT dose, number of elapsed days, the minimal tumor temperature factors of T,,,, and f,lm,n,, and the number of HT treatments all correlated with occurrence of skin reactions.

HYPERTHERMIA TRIALS Despite the strong scientific and rational basis for HT, its true efficacy when

given as an adjuvant to primary RT still needs to be fully evaluated in objective randomized multi-institutional studies. In this same context, the early and late tolerance of normal tissues, also requires assessment. The identification of the technical, biological and clinical variables affecting tumor response and normal tissue tolerance remains a critical issue that will eventually lead to standardization of techniques including thermometry, method of heating, fractionation schedules, and quality assurance. Various international societies have been actively involved in conducting clinical trials to achieve these objectives. In investigating the role of HT in primary treatment, a strong priority is to minimize the risk of severe late effects. With this in mind, the general principle is to recommend conventional fractionation and to supplement this with HT either once or twice weekly. External

Page 17: Biological rationale and clinical experience with hyperthermia

K. Engin

HT combined with RT can be applied to superficial lesions of head and neck, chest wall recurrences, and melanoma lesions with depth ~3 cm. For deeper lesions (i.e., 3-6 cm), external and intersititial techniques can be combined.

Interest has also developed in the use of endocavitary and whole-body HT [172]. Whole-body HT has most often been combined with chemotherapy 11731, frequently in the treatment of advanced small-cell carcinoma of the lung, but is associated with considerable morbidity and is being investigated in few institutions. More widespread is the use of hyperthermic perfusion of the lower limbs for the treatment of malignant melanoma.

The combination of external applicator HT and chemotherapy, especially cisplati- num for the treatment of refractory superficial malignancies is also under investiga- tion [174]. This is designed for patients who have already received full-dosage RT often in multiple courses of treatment. The use of all three modalities delivered concurrently (trimodality therapy) is also being investigated in single-institution studies [103]. This form of therapy holds considerable promise because of the synergistic interaction of any two of the three modalities and has been found to be efficacious in vitro [175]. Cisplatinum in combination with RT is being closely investigated in several RT studies and seems to produce less sensitization of normal tissues compared to other agents. Maximum cytotoxicity occurs when these modal- ities are administered simultaneously, but optimal dose scheduling and sequencing of therapy needs to be investigated in clinical studies.

QUALITY ASSURANCE

The need for quality assurance in HT is particularly great at this time because of the lack of standardization of equipment and techniques used for heating and monitoring temperatures. There are few objective criteria with which to judge the quality of HT treatment and quantitative assessment of patients derived from different institutions is fouled by variability in patient treatment and temperature data [176-1781. Of particular importance is ensuring the adequacy of heating pat- terns for a given tumor volume while limiting normal tissue heating. Because there is no direct relationship between the sulfarsphenamine (SAR) pattern derived from phantom studies and temperatures achieved in perfused tissue, considerable experi- ence is required to choose the appropriate applicator for a given tumor size and site. In general, the tumor volume and margin should be encompassed by the 50% iso-SAR and placement arranged to maximize power coupling and minimize reflected and stray electromagnetic leakage. Unfortunately, this often requires the use of an applicator roughly twice the surface area of the lesion, which may not always be feasible depending on the location of the tumor. Moreover, the SAR pattern rapidly constricts with depth and a lesion of appreciable thickness may not be adequately heated even with a generous applicator [179-1811.

An area of particular concern is the standardization of thermometer placement within the tumor. The need for ample multipoint thermometry or thermal mapping cannot be overemphasized. The RTOG Thermometry Task Force has formulated recommendations and guidelines for the insertion of catheters for monitoring tumor temperature for both microwave [180] and untrasound [182] modalities. Verification of thermometry catheter placement with computed tomography scans is necessary in tumors of >1.5 cm depth, both to confirm that the catheters in fact traverse tumor and to determine the depth of placement [165]. Documentation of catheter

Page 18: Biological rationale and clinical experience with hyperthermia

Thcrmoradiotherapy in Cancer Treatment 333

locations relative to tumor margins is mandatory. Thermal mapping should be employed by using multisensor thermal probes, preferably performed mechanically under computer guidance.

AREAS OF FUTURE DEVELOPMENT

The major limitation at this time is the inadequacy of current equipment to deliver power homogeneously to tissue and allow inclusion of tumor volumes of greater extent and depth. No less important is the availability of temperature monitoring systems that will accurately reflect the tumor temperature with good resolution and without significant artefact. A severe drawback at this time is the necessity to resort to invasive thermometry and the need to insert multiple multiarray temperature sensors in order to assess tumor temperature adequately. Improvement in protocol design and analysis, taking into account all important prognostic factors and stratifying them also needs to be achieved. The combination of whole-body HT in addition to locoregional HT is being investigated both as a means of producing greater homogeneity in the heating pattern and raising the minimum tumor temperature to therapeutic levels 11831. The interaction of chemo- therapy and HT is also being explored in early phase l/II studies separately and in combination with RT and has enormous potential for clinical use.

REFERENCES 1. Busch W. Uber den einfluf3, welchen heftigere erysipele zuweilen auf organisierte

neubildungen ausuben. Verhandl. des naturhistorischen vereines der preupishchen rheinlande und westphalens. 23:28-30.

2. Coley WB. The treatment of malignant tumors by repeated inoculations of crysipe- las: with a report of ten original cases. Am J Med Sci 1893;105:487-511.

3. Schmidt HE. Zur rontgenbehandlung tiefliegender tumoren. Fortschr Roent,pstr 1909;14:134-136.

4. Overgaard J. Rationale and problems in the design of clinical studies. In Hyperther- mic Oncology, Vol. I, Summary Papers, Overgaard J, ed. London, Taylor & Fran- cis, 1984;325-338.

5. Perez CA, Gillespie B, Pajak T, et al. Quality assurance problems in clinical hyper- thermia and their impact on therapeutic outcome: A report by the RTOG. lnt J Rad Orzcol Biol Phys 1989;16:551-558.

6. Westra A, Dewey WC. Heat shock during the cell cycle of Chinese hamster cells in vitro. Int ] Radiut Biol 1971;19:467-477.

7. Dewey WC, Westra A, Miller HH. Heat-induced lethality and chromosomal dam- age in synchronized Chinese hamster cells treated with 5-bromodeoxyuridine. Int J Radint Biol 1971;20:50%520.

8. Coss RA, Dewey WC. Mechanism of heat sensitization of Gl and S-phase cells by procaine-HCl. In: Broerse JJ, et al., eds. Proc VII Int Cong Radiat Res Martinus Amsterdam, Nijhoff; 1983;D6-05-06.

9. Lepock JR, Cheng KH, Al-Qysi H, Kruuv J. Thermotropic lipid and protein transi- tions in Chinese hamster lung cell membranes: Relationship to hyperthermic cell killing. Can J Biochern Cell Biol 1983;61:421-427.

10. Yatvin MB. The influence of membrane lipid composition and procaine on hyper- thermia death of cells. Inf I Radiat Biol 1977;32:513-521.

Page 19: Biological rationale and clinical experience with hyperthermia

334 K. Engin

11. Wong RSL, Dewey WC: Effect of hyperthermia on DNA synthesis. In: Anghileri LJ, Robert J, eds. Hyperthermia in cuncer treatment. vol. I, Boca Raton, FL: CRC Press; 1986;79-92.

12. Warters RL, Stone OL. The effects of hyperthemia on DNA replication in HeLa cells. Radint Res 1983;93:71-77.

13. Coss RA, Dewey WC. Heat sensitization of G,- and S-phase cells by procaine hydrochloride. Radiat Res 1982;92:615-617.

14. Li GC, Shiu EC, Hahn GM. Similarities in cellular inactivation by hyperthermia or by ethanol. Radiat Res 1980;82:257-268.

15. Gerweck LE. Modification of cell lethality at elevated temperatures: the pH effect. Radiat Res 1977;70:224-235.

16. Hahn GM. Metabolic aspects of the role of hyperthermia in mammalian cell inacti- vation and their possible relevance to cancer treatment. Cancer Res 1974;34:3117-3123.

17. Gerweck LE, Nygaard TG, Burlett M. Response of cells to hyperthermia under acute and chronic hypoxic conditions. Cancer Res 1979;39:966-972.

18. Fisher GA, Li GC, Hahn GM. Modification of the thermal response by D20: Cell survival and the temperature shift. Radiat Res 1982;92:530-540.

19. Henle KJ, Peck JW, Higashikubo R. Protection against heat-induced cell killing by polyols in vitro. Cancer Res 1983;43:1624-1629.

20. Harisiadis L, Miller RC, Harisiadis S, Hall EJ. Oncogenic transformation and hyper- thermia. Brit J Radio1 1980;53:479-482.

21. Sapareto SA, Hopwood L, Dewey WC, et al. Effects of hyperthennia on survival and progression of Chinese hamster ovary cells. Cancer Res 1978;38:39%400.

22. Henle KJ, Leeper DB. Interaction of hyperthermia and radiation in CHO cells: Recovery kinetics. Radiat Res 1976;66:505518.

23. Henle KJ, Leeper DB. Interaction of sublethal and potentially lethal 45”-hyperther- mia and radiation damage at 0, 20, 37, or 40°C. Eur J Cancer 1978;15:1387-1392.

24. Li GC, Shiu EC, Hahn GM. Recovery of cells from heat-induced potentially lethal damage: Effects of pH and nutrient environment. Int 1 Radiut Oncol Biol Phys 1980;6:577-582.

25. Streffer C. Metabolic changes during and after hyperthermia. Int 1. Hyperther 1985;4:305-319.

26. Henle KJ, Leeper DB. Effects of hyperthermia (45°C) on macromolecular synthesis in Chinese hamster ovary cells. Cancer Res 1979;39:2665-2674.

27. Lin PS, Turi A, Kwock L, Lu RC. Hyperthermic effect on microtubule organization. Nat’1 Cancer lnst Monogr 1982;61:57-60.

28. Li GC, Hahn GM. A proposed model of thermotolerance based on effects of nutri- ents and the initial treatment temperature. Cancer Res 1980;40:4501-4508.

29. Henle KJ, Dethlefsen LA. Time-temperature relationships for heat-induced killing of mammalian cells. In: Jain RK, Gulliano PM, eds. Thermal characteristics of tumors: applications and detection in treatment, New York, NY Acad Sci 1980;335:234-252.

30. Sapareto SA, Dewey WC. Thermal dose determination in cancer therapy. Znt I Radiat Oncol Biol Phys 1984;10:787~800.

31. Henle KJ, Dethlefsen LA. Heat fractionation and thermotolerance: A review. Cancer Res 1978;38:1843-1851.

32. Li GC, Mivechi NF: Thermotolerance in mammalian systems: A review. In: Anghil- eri LJ, Robert J, eds. Hyperthermia in Cancer Treatment. Vol I, Boca Raton, FL, CRC Press; 1986:59-77.

33. Henle KJ, Bitner AF, Dethlefsen IA. Induction of thermotolerance by multiple fractions in Chinese hamster ovary cells. Cancer Res 1979;39:2486-2491.

Page 20: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 335

34. Henle KJ, Leeper DB. Modification of the heat response and thermotolerance by cycloheximide, hydroxyurea and lucanthone in CHO cells. Radiaf Res 1982;90:339.

35. Li GC, Laszlo A. Thermotolerance in mammalian cells: a possible role for heat shock proteins. In: Atkinson BG, Walden DB, eds. Changes in Gene Expression in Response to Environmental Stress, New York, Academic Press; 1985:227-254.

36. Subjeck JR, Shyy TT. Stress protein systems of mammalian cells. Am ] Physiol 1986;25O:Cl-C17.

37. Landry J. Heat shock proteins and thermotolerance. In: Anghileri LJ, Robert J, eds. Hyyerfhermia in Cancer Treatment. Vol. I, Boca Raton, FL, CRC Press; 1986:37-58.

38. Li GC. Elevated levels of 70,000 dalton heat shock protein in transiently thermotoler- ant Chinese hamster fibroblasts and in their stable heat-resistant variants. ht 1 Rudiuf Oncol Biol Phys 1985;11:165-177.

39. Milligan AJ, Metz MS, Leeper DB. Effect of intestinal hyperthermia in the Chinese hamster. Int 1 Rudiut Oncol Biol Phys 1984;10:259-263.

40. Law MI’, Ahier RG, Somaia S, Field SB. The induction of thermotolerance in the ear of the mouse by fractional hyperthermia. Int JRudiuf Oncol BiolPhys 1984;10:865873.

41. Emami B, Myerson RJ, Cardenes H, et al. Combined hypertherrnia and irradiation in the treatment of superficial tumors: Results of a prospective randomized trial of hyperthermia fractionation (l/wk vs 2/wk). Int J Rudiuf Oncol Biol Phys 1992;24:145152.

42. Engin K, Tupchong L. Moylan DJ, et al. Randomized trial of once weekly vs twice weekly hyperthermia treatments in thermoradiotherapy of superficial tumors. Int J Hyperfher 1993;9:327-340.

43. Kapp DS, Petersen IA, Cox RS, et al. Two or six hyperthermia treatments as an adjunct to radiation therapy yield similar tumor responses: results of a randomized trial. Int ] Radial Oncol Biol Phys 1990;19:1481-1495.

44. Goldin EM, Leeper DB. The effect of low pH on thermotolerance induction using fractionated 45°C hyperthermia. Rudiuf Res 1981;85:472-479.

45. Gerweck LE, DeLaney TF. Persistence of thermal tolerance in slowly proliferating plateau phase cells. Rudiuf Res 1984;97:365-372.

46. Henle KJ, Leeper DB. Combinations of hyperthermia (40”,45”C) with radiation. Radiology 1976;121:451454.

47. Henle KJ. Sensitization to hyperthermia below 43°C induced in Chinese hamster ovary cells by step-down heating. j Nat1 Cancer Inst 1980;64:1479-1483.

48. Henle KJ, Keeper DB. The modification of radiation damage in CHO cells by hyperthermia at 40” and 45°C. Rudiut Res 1977;70:415424.

49. Song CW. Effects of local hyperthermia on blood flow and microenvironment: a review. Cancer Res 1984;44(supp):4721s473Os.

50. Reinhold HS, Wike-Hooley JL, van den Berg AI’. Environmental factors, blood flow and microcirculation. In: Overgaard J, ed. Hyperthermiu Oncology, Vol. 11 London, Taylor & Francis; 1985:41-52.

51. Rofstad EK, Howell RL, DeMuth I’, et al. 31P NMR spectroscopy in vivo of two murine tumor lines with widely different fractions of radiobiologically hypoxic cells. lnt J Rudiut Biol 1988;54:635-649.

52. Vaupel P, Kallinowski F, Okunieff P. Blood flow, oxygen and nutrient supply, and metabolic microenvironment of human tumors: a review. Cancer Res 1989; 49:6449-6465.

53. Thistlethwaite AJ, Leeper DB, Moylan DJ, Nerlinger RE. pH distribution in human tumors. Inf J Rudiat Oncol Biol Phys 1985;11:1647-1652.

54. Ashby BS. pH studies in human malignant tumors. Luncet 1966;389:312-313. 55. Wike-Hooley JL, Haveman J, Reinhold HS. The relevance of tumour pH to the

treatment of malignant disease. Rudiother Oncol 1984;2:34%366.

Page 21: Biological rationale and clinical experience with hyperthermia

336 K. Engin

56. Kallinowski F, Vaupel P. pH distributions in spontaneous and isotransplanted rat tumors. British j Cancer 1988;58:314-321.

57. van den Berg AI’, Wike-Hooley JL, Broekmeyer Reurink MI’, et al. The relationship between the unmodified initial tissue pH of human tumors and the response to combined radiotherapy and local hyperthermia treatment. Eur J Cancer Clin Oncol 1989;25:73-78.

58. Griffiths JR. Are cancer cells acidic? British I Cancer 1991;64:425-427. 59. Engin K, Leeper DB, Thistlethwaite AJ, et al. Extracellular pH distribution in human

tumors, Int J Hyperther in press, 1994 60. Engin K, Leeper DB, Tupchong L, et al. Thermoradiotherapy for superficial malig-

nant tumors. Cancer 1993;72:287-296. 61. Engin K, Leeper DB, Thistlethwaite AJ, et al. Tumor extracellular pH as a prognostic

factor in thermoradiotherapy. Int j Radiat Oncol Biol Phys 1994;29:125-132. 62. Busse J, Muller-Klieser W, Vaupel P. Intratumor pH distributionA function of

tumor growth stage? Pflugers Arch 1981;2:55. 63. Vaupel P, Frinak S, Bicher HI. Heterogeneous oxygen partial pressure and pH

distribution in CH3 mouse mammary adenocarcinoma. Cancer Res 1981;41: 20082013.

64. Gullino PM, Grantham FH, Smith SH, Hagerty AC. Modifications of acid-base status of the internal milieu of tumors. J Nat1 Cancer Inst 1965;34:857-869.

65. Sostman HD, Prescott DH, Dewhirst MW, et al. MRI/MRS for prognostic evaluation in soft tissue sarcomas. Radiology 1994;190:269-275.

66. Hahn GM, Shiu EC. Adaptation to low pH modifies thermal and thermochemical responses to mammalian cells. Int j Hyperther 1986;2:379-387.

67. Chu CL, Wang Z, Hyun WC, et al. The role of intracellular pH and its variance in low pH sensitization of killing by hyperthermia. Radiat Res 1990;122:288-293.

68. Cook JA, Fox MH. Effects of chronic pH 6.6 on growth, intracellular pH, and response to 42°C hyperthermia of Chinese hamster ovary cells. Cancer Res 1988;48:2417-2420.

69. Fellenz MI’, Gerweck LE. Influence of extracellular pH on intracellular pH and cell energy status: relationship to hyperthermic sensitivity. Radiat Res 1988;116:305-312.

70. Lyons JC, Kim GE, Song CW. Modification of intracellular pH and thermosensiti- vity. Radial Res 1992;129:79-87.

71. Holahan EV, Highfield DP, Holahan PK, Dewey WC. Hyperthermic killing and hyperthermic radiosensitization in Chinese hamster ovary cells: effects of pH and thermotolerance. Radiat Res 1984;97:10%131.

72. Prescott DM, Charles HC, Sostman HD, et al. Manipulation of intra- and extracellu- lar pH in spontaneous canine tumors by use of hyperglycemia. Int 1 Hyperther 1993;9:745754.

73. Dickson JA, Calderwood SK. Effects of hyperglycemia and hyperthermia on the pH, glycolysis and respiration of the Yoshida sarcoma in vivo. J Nat1 Cancer Znst 1979;63:1371-1381.

74. Leeper DB, Engin K, Thistlethwaite AJ, et al. Human tumor extracellular pH as a function of blood glucose concentration. Int J Radiat Oncol BioZ Phys 1994;28:935-943.

75. Song CW, Lokshina A, Rhee JG, et al. Implication of blood flow in hyperthermia treatment of tumors. IEEE Trans BME 1984;31:9-16.

76. Waterman FM, Nerlinger RE, Moylan III DJ, Leeper DB. Response of human tumor blood flow to local hyperthermia. Int 1 Radiat Oncol Biol Phys 1987;13:7%82.

77. Kang MS, Song CW, Levit SH. Role of vascular function in response of tumors in vivo to hyperthermia. Cancer Res 1980;40:1130-1135.

Page 22: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 337

78. Raaphorst GP, Azzam EI: A comparative study of heat and/or radiation sensitivity of V79 cells synchronized by three different methods. In: Overgaard J, ed. Hypertker- nlic Oncology Vol. 1, London, Taylor & Francis; 1984:301-304.

79. Li CG, Evans RG, Hahn GM. Modification and inhibition of repair of potentially lethal x-ray damage by hyperthermia. Radiat Res 1976;67:491-501.

80. Mills MD, Meyn RE. Hyperthermic potentiation of unrejoined DNA strand breaks following irradiation. Radiat Res 1983;95:327-338.

81. Jorritsma JBM, Konings AWT. Inhibition of repair of radiation-induced strand breaks by hyperthermia and its relationship to cell survival after hyperthermia alone. lnt J Radiat Biol 1983;43:505-511.

82. Meyer KR, Hopwood LE, Gillette EL. The response of mouse adenocarcinoma cells to hyperthermia and irradiation. Radiat Res 1979;78:98-107.

83. Kellerer AM, Rossi HH. The theory of dual radiation action. Curr Top Radiat Res 1971;8:85-158.

84. Chadwick KH, Keenhouts HP. The Molecular Theory of Radiation Biology. New York, Springer-Verlag, 1981.

85. Gerner EW, Leith JT: Interaction of hyperthermia with radiation of different linear energy transfer. lnt 1 Radiat Biol 1977;31:283-288.1977

86. Dennekamp J, Stewart F. Evidence for repair capacity in mouse tumors relative to skin. Int J Radial Oncol Biol Pkys 1979;5:2003--2010.

87. Overgaard J. Simultaneous and sequential hyperthermia in radiation treatment of an experimental tumor and its surrounding tissue in vivo. Znt \ Radiat Oncol Biol Plrys 1980;6:1507-1517.

88. Stratford IJ. Hyperthermia and hypoxic cell radiosensitizers in combination. In: Anghilaeri LJ, Robert J, eds. Hypertkermia in Cancer Treatment Vol. I. Boca Raton, FL, CRC Press; 1986:151-167.

89. Falk RE, Moffat FL, et al. Effect of radiofrequency hyperthermia on human neo- plasms when used with adjuvant metronidazole. Surg Gynecol Obstet 1983; 157:505-511.

90. Ovcrgaard J. Effect of misonidazole and hyperthermia on the radiosensitivity of a C3H mouse mammary carcinoma and its surrounding normal tissue. Br I Can- cer 1980;41:10-17.

91. Hahn GM. Potential for therapy of drugs and hyperthermia. Cancer RES 1979; 39:2264--2268.

92. Hahn GM. Hyperthermia and Cancer. New York, Plenum Press, 1982. 93. Marmor JB. Interactions of hyperthermia and chemotherapy in animals. Cancer Res

1979;39:2269-2276. 94. Dewey WC. Interaction of heat with radiation and chemotherapy. Cnnccr Res

1984;fsuppl) 44:4714s-4720s. 95. Suit HD. Potential for improving survival rates for the cancer patient by increasing

the efficacy of treatment of the primary lesion. Cancer 1982;50:1227-1234. 96. Ben-Hur E, Elkind MM, Bronk BV. Thermally enhanced radioresponse of cultured

Chinese hamster cells: inhibition of repair of sublethal damage and enhancement of lethal damage. Radiat Res 1974;58:3851.

97. Storm FK, Harrison FH, Elliot RS, Morton DL. Hyperthermic therapy for human neoplasms: Thermal death time. Cmcer 1980;46:1849-1854.

98. Song CW, Kang MS, Rhee JG, Levitt S. Vascular damage and delayed cell death in tumours after hyperthermia. Br J Cancer 1980;41:309-312.

99. Waterman FM, Tupchong L, Matthews J, Nerlinger RE. Blood flow in human tumors. Int J Radiat Oncol Biol Pkys 1991;20:1255-1262.

100. Streffer C. Mechanisms of heat injury: In: Overgaard J, ed. Hypertkermic Oncology London, Taylor & Francis; 1985:213--222.

Page 23: Biological rationale and clinical experience with hyperthermia

338 K. Engin

101. Leeper DB. Molecular and cellular mechanisms of hyperthermia alone or combined with other modalities. In: Overgaard J, ed. Hyperthermic Oncology, Vol. II, London, Taylor & Francis; 1985940.

102. Howard GCW, Bleehen NM. Clinical experience in the combination of hyperther- mia with chemotherapy or radiotherapy. Ret Res Cancer Res 1988;107:214-221.

103. Herman TS, Teicher BA, Jochelson M, et al. Rationale for use of local hyperthermia with radiation therapy and selected anticancer drugs in locally advanced human malignancies. Int J Hyperther 1988;4:143-158.

104. Dahl 0. Interaction of hyperthermia and chemotherapy. Ret Res Cancer Res 1988;107:157-169.

105. Valdagni R, Amichetti M, Pani G. Radical radiation alone versus radical radiation plus microwave hyperthermia for N3 (TNM-UICCO) neck nodes: A prospective randomized clinical trial. Int ] Radiat Oncol Biol Phys 1988;15:13-24.

106. Arcangeli G, Cividalli A, Creton G, et al. Tumor control and therapeutic gain with different schedules of combined radiotherapy and local hyperthermia in human cancer. Int J Radiat Oncol Biol Phys 1983;9:112!%1134.

107. Arcangeli G, Nervi C, Cividalli A, Lovisolo GA. Problem of sequence and fraction- ation in the clinical application of combined heat and radiation. Cancer Res (suppl) 1984;44:48574863.

108. Valdagni R, Amichetti M. Report of long-term follow-up in a randomized trial comparing radiation therapy and radiation therapy plus hyperthermia to metastatic lymph nodes in stage IV head and neck patients. Int J Radiat Oncol Biol Phys 1994;28:163-169.

109. Overgaard J. Thermoradiotherapy of malignant melanoma: the ESHO experience. Radiother Oncol submitted, 1994.

110. Vernon CC. Collaborative phase III superficial hypertheimia trial (MRC/ESHO- PMH). Abstracts of 42nd Ann Mg of RRS and 14th Ann Mtg of NAHS, April 29-May 4, Nashville, TN, Abs. #S-11-2, 1994;87.

111. Overgaard J. The current and potential role of hyperthermia in radiotherapy. Int J Radiat Oncol Biol Phys 1989;16:535-549.

112. Kim JH, Hahn EW, Ahmed SA, Kim YS. Clinical study of the sequence of combined hyperthermia and radiation therapy of malignant melanoma. In: Overgaard J, ed. Hyperthermic Oncology, Vol. 1, London-Philadelphia, Taylor & Francis; 1984: 387-391.

113. Overgaard J. Hyperthermia as an adjuvant to radiotherapy. Review of the random- ized multicenter studies of the ESHO. Strahlenther Onkol 1987;163:453457.

114. Scott R, Gillespie B, Perez CA, et al. Hyperthermia in combination with definitive radiation therapy: results of a phase I/II RTOG study. Int J Radiat Oncol Biol Phys 1988;15:711-716.

115. Li RY, Zhang TZ, Lin SY, Wang HP. Effect of hyperthermia combined with radiation in the treatment of superficial malignant lesion in 90 patients. In: Overgaard J, ed. Hyperfhermic Oncology Vol. 1, London and Philadelphia, Taylor & Francis; 19% 395-397.

116. Hiraoka M, Jo S, Dodo Y, et al. Clinical results of radiofrequency hyperthermia combined with radiation in the treatment of radioresistant cancers. Cancer 1984;54:2898-2904.

117. Valdagni R, Gosetti G, Amichetti M, et al. Irradiation and hyperthermia in Ns metastatic neck nodes: clinical results with an analysis of treatment parameters. StrahZenther Onkol 1985;161:551-552.

118. Perez CA, Kuske RR, Emami B, Fineberg 8. Irradiation alone or combined with hyperthermia in the treatment of recurrent carcinoma of the breast in the chest wall. A nonrandomized comparison. Int J Hyperther 1986;2:179-187.

Page 24: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 339

119. van der Zee J, van Putten WLJ, van den Berg AI’, et al. Retrospective analysis of the response of tumours in patients treated with combination of radiotherapy and hyperthermia. Int J Hyperther 1986;2:337-349.

120. Steeves RA, Severson SB, Paliwal BR, et al. Matched-pair analysis of response to local hyperthermia and megavoltage electron therapy for superficial human tu- mors. Endocuriether Hyperther Oncol 1986;2:163-170.

121. Dunlop PR, Hand JW, Dickinson RJ, Field SB. An assessment of local hyperthermia in clinical practice. Int J Hyperther 1986;2:39-50.

122. Uozumi H, Baba Y, Yasunaga T, et al.: Clinical evaluation of combined hyperther- mia and radiation therapy of superficial malignant tumors. In: Onoyama Y, ed. Hyperthermic Oncology 1986 in Japan; 1986:311-312.

123. Hornback NB, Shupe RE, Shidnia H, et al. Advanced stage IIIB cancer of the cervix treatment by hyperthermia and radiation. Gynecol Oncol 1986;23:160-167.

124. Fuwa N, Mortia K, Kimura C, et al. Combined treatment of radiotherapy and local hyperthermia using 8 MHz RF-wave for advanced carcinoma of the breast. In: Onoyama, ed. Hyperthernlic Oncology 1986 in Japan; 1986:337-338.

125. Gonzalez Gonzalez D, van Dijk JDP, Blank LECM, Rumke PH. Combined treatment with radiation and hyperthermia in metastatic malignant melanoma. Radiother Oncol 1986;6:105-113.

126. Lindholm CE, Kjellen E, Nilsson I’, Hertzman S. Microwave-induced hyperthermia and radiotherapy in human superficial tumours: clinical results with comparative study of combined treatment versus radiotherapy alone. Int J Hyperther 1987; 3:39%11.

127. Goldobenko GV, Durnov LA, Knysh VI, et al. Experience in the use of thermoradio- therapy of malignant tumors. Med Radio1 1987;32:36-37.

128. Muratkhodzhaev NK, Svetitsky PV, Kochegarov AA, et al. Hyperthermia in therapy of cancer patients. Med Radio1 1987;32:30-36.

129. Datta NR, Bose AK, Kapoor HK. Thermoradiotherapy in the management of carci- noma cervix (IIIB): A controlled clinical study. Indian Med Gazette 1987;121:68-71.

130. Bide Z, Xiaoxiong W, Diven Z, Zhonghe G. Hyperthermia in combination with radiotherapy for middle and late stage bladder carcinoma. In: Overgaard J, ed. Hyperthermic Oncology, Vol. 1, London and Philadelphia, Taylor & Francis; 1984783-784.

131. Corry PM, Spanos WJ, Tilchen EJ, et al. Combined ultrasound and radiation therapy treatment of humen superficial tumors. Radiology 1982;145:165-169.

132. Hofman I’, Lagendijk JJW, Schipper J. The combination of radiotherapy with hyper- thermia in protocolized clinical studies. In: Overgaard J, ed. Hyperthermic Oncology, Vol, l., London-Philadelphia, Taylor & Francis; 1984379-382.

133. Luk KH, Francis ME, Perez CA, Johnson RJ. Combined radiation and hyperthermia: comparison to two treatment schedules based on data from a registry established by the radiationtherapy oncology group (RTOG). Int J Radiat Oncol Biol Phys 1984;10:801-809.

134. Matsuda T, Sugiyama A, Nakata Y: Fundamental and clinical studies of radiofre- quency hyperthermia and radiation therapy. In: Overgaard J, ed. Hypertherrnic Oncology, Vol. 1, London-Philadelphia, Taylor & Francis; 1984:349-352.

135. Bicher HI, Wolfstein RS, Lewinsky BS, et al. Microwave hyperthermia as an adjunct to radiation therapy: Summary experience of 256 multifraction treatment cases. Int j Radiat Oncol Biol Phys 1986;12:1667-1671.

136. Arcangeli G, Benassi M, Cividalli A, et al. Radiotherapy and hyperthermia: Analysis of clinical results and identification of prognostic variables. Cancer 1987;60:95&956.

137. Perez CA, Nussbaum G, Emami B, vonGerichten D. Clinical results of irradiation combined with local hyperthermia. Cancer 1983;52:1597-1603.

Page 25: Biological rationale and clinical experience with hyperthermia

340 K. Engin

138. Dewhirst MW, Sim DA, Sapareto S, Connor WG. Importance of minimum tumor temperature in determining early and long-term responses of spontaneous canine and feline tumors to heat and radiation. Cancer Res 1984;44:43-50.

139. Dewhirst MW, Sim DA. The utility of thermal dose as a predictor of tumor and normal tissue responses to combined radiation and hyperthermia. Cancer Res 1984;44(Suppl):4772s-4780s.

140. Oleson JR, Sim DA, Manning MR. Analysis of prognostic variables in hyperthen-nia treatment of 161 patients. Int J Radiut Oncol Biol Phys 1984;10:2231-2239.

141. Sim DA, Dewhirst MW, Oleson JR, Grochowski KJ: Estimating the therapeutic advantage of adequate heat. In: Overgaard J, ed. Hyperthermic Oncology, Vol. I, London-Philadelphia, Taylor & Francis; 1984359-362.

142. Arcangeli G, Guerra A, Lovisolo G, et al. Tumor response to heat and radiation: Prognostic variables in the treatment of neck node metastases from head and neck cancer. Int I Hyperther 1985;3:207-217.

143. Kapp DS, Barnett TA, Cox RS, et al. Hyperthermia and radiation therapy of local- regional recurrent breast cancer: Prognostic factors for response and local control of diffuse or nodular tumors. Int J Radiut Oncol Biol Phys 1991;20:1147-1164.

144. Sapozink MD, Gibbs FA, Egger MJ, Stewart JR. Regional hyperthermia for clinically advanced deep-seated pelvic malignancy. Am J Clin Oncol 1986;9:162-169.

145. van der Zee J, van Rhoon GC, Wike-Hooley JL, et al.: Thermal enhancement of radiotherapy in breast carcinoma. In: Overgaard J, ed. Hyperthermic Oncology, Vol. I, London-Philadelphia, Francis & Taylor; 1984:345-349.

146. Seegenschmiedt MH, Karlsson UL, Sauer R, et al. Superficial chest wall recurrences of breast cancer: Prognostic treatment factors for combined radiation therapy and hyperthermia. Radiology 1989;173:551-558.

147. Emami B, Perez CA, Konefal J, et al. Thermoradiotherapy of malignant melanoma. Int ] Hyperther 1988;4:373-381.

148. Oleson JR, Dewhirst MW, Harrelson JM, et al. Tumor temperature distributions predict hyperthermia effect. Int I Radial Oncol Biol Phys 1989;16:559-570.

149. Engin K, Tupchong L, Waterman FM, et al. Hyperthermia and radiation in ad- vanced malignant melanoma. Inl J Radiat Oncol Biol Phys 1993;25:87-94.

150. Engin K, Tupchong L, Waterman FM, et al. Thermoradiotherapy for advanced soft tissue sarcoma. Endocuriether Hyperther Oncol 1992;8:19-26.

151. Engin K, Tupchong L, Waterman FM, et al. “Patchwork” fields in the hyperthermic management of patients with extensive locoregional disease due to carcinoma of breast. Breast Cancer Res Treat 1993;27:263-270.

152. Engin K, Tupchong L, Waterman FM, et al. Thermoradiotherapy for superficial tumor deposits in the head and neck. Int J Hyperther 1994;10:153-164.

153. Kapp DS. Site and disease selection for hyperthermia clinical trials. Int ] Hyper- ther 1986;2:139-156.

154. Montague ED, Gutierrez AE, Barker JL, et al. Conservation surgery and irradiation for the treatment of favorable breast cancer. Cancer 1979;43:1058-1061.

155. van der Zee J, Treurniet-Donker AD, The SK, et al. Low dose reirradiation in combination with hyperthermia: a palliative treatment for patients with breast cancer recurring in previously irradiated areas. Irzl J Radiat Oncol Biol Phys 1988;15:1407-1413.

156. Bedwinek JM, Lee J, Fineberg B, Ocwieza M. Incidence and sites of local failure following local therapy in patients with isolated local-regional recurrence of breast cancer. Int J Rndiat Oncol Biol Phys 1981;7:581-585.

157. Vora N, Shaw S, Fore11 8, et al. Primary radiation combined with hyperthermia for advanced and inflammatory carcinoma of breast. Endocuriether Hyperther On- col 1986;2:101-106.

Page 26: Biological rationale and clinical experience with hyperthermia

Thermoradiotherapy in Cancer Treatment 341

158. Puthawala AA, Syed AMN, Sheikh KMA, et al. Interstitial hyperthermia for recur- rent malignancies. Endocurietker Hypertker Oncol 1985;1:125-131.

159. Kapp DS, Cox RS, Barnett TA, Ben-Yosef R. Thermoradiotherapy for residual microscopic cancer: Elective or post-excisional hyperthermia and radiation therapy in the management of local-regional recurrent breast cancer. Int J Radiut Oncol Biol Pkys 1992;24:261-277.

160. Engin K, Tupchong L, Waterman FM, et al. Thermoradiotherapy with combined interstitial and external hyperthermia in advanced tumors in the head and neck with depth 23 cm. Int J Hypertker 1993;9:645-654.

161. Fessenden I’, Lee E, Kapp DS, et al. The microwave blanket: a large area conformable applicator for hyperthermia of superficially located tumors. (Abstr). Int 1 Radiat Oncol Biol Pkys 1989;17(suppl):214.

162. Overgaard J. The role of radiotherapy in recurrent and metastatic malignant mela- noma: a clinical radiobiological study. Int I Radiat Oncol Biol Pkys 1986;12:867-872.

163. Overgaard J, Overgaard M. Hyperthermia as an adjuvant to radiotherapy in the treatment of malignant melanoma. Int ] Hypertker 1987;3:483-501.

164. Kim JH, Hahn EW, Ahmed SA. Combination hyperthermia and radiation therapy for malignant melanoma. Cancer 1982;50:478-482.

165. Engin K, Tupchong L, Waterman FM, et al. Optimization of hyperthermia with CT Scanning. Int J Hypertker 1992;8:855-864.

166. Seegenschmiedt MH, Sauer R, Fietkau R, et al. Interstitial thermal radiation therapy: Five-year experience with head and neck tumors. Radiology 1992;184:795-804.

167. Gabriele I’, Orecchia R, Ragona R, et al. Hyperthermia alone in the treatment of recurrences of malignant tumors: experience with 60 lesions. Cancer 1990; 66:2191-2195.

168. Luk KH, Francis ME, Perez CA, Johnson RJ. Radiation therapy and hyperthermia in the treatment of superficial lesions: Preliminary analysis. Treatment efficacy, and reactions of skin, subcutaneous tissues. Radiation Therapy Oncology Group Phase I/II Protocol 78-06. Am J Clin Oncol (CCT) 1983;6:399406.

169. Howard GCW, SathiaseelanV, Freedman L, Bleehen NM. Hyperthermia and radia- tion in the treatment of superficial malignancy: an analysis of treatment parameters, response and toxicity. lnt J Hypertker 1987;3:1-8.

170. Kapp DS, Cox RS, Fessenden I’, et al. Parameters predictive for complications of treatment with combined hyperthermia and radiation therapy. Int 1 Rndiat Oncol Biol Pkys 1992;22:999-1008.

171. Engin K, Tupchong L, Waterman FM, et al. Predictive factors for skin reactions in patients treated with thermoradiotherapy. Int J Hypertker in press, 1994.

172. Englehart R. Whole body hyperthermia. Methods and clinical results. In: Overgaard J, ed. Hypertkerrnic Oncology, Vol. II, London and Philadelphia, Taylor & Francis; 1985:263-276.

173. Bull JM, Lees D, Schuette W, et al. Whole body hypertherrnia: a Phase I trial of potential adjuvant chemotherapy. Ann Int Med 1979;90:317-322.

174. Fisher GA, Hahn GM. Enhancement of cis-platinum(I1) diamminedichloride cyto- toxicity by hyperthermia. Nat Cancer lnst Mono 1982;61:255-258.

175. Urano M, Kim MS, Kahn J, et al. Effect of thermochemotherapy (combined cyclo- phosphamide and hyperthermia) given at various temperatures with or without glucose administration in a murine fibrosarcoma. Cancer Res 1985;45:41624166.

176. Myerson RJ, Perez CA, Emami 8, et al. Tumor control in long-term survivors following superficial hyperthermia. Int J Radiat Oncol Biol Pkys 1990;18:1123-1129.

177. Myerson RJ, Emami BN, Perez CA, et al. Equilibrium temperature distributions in uniform phantoms for superficial microwave applicators: Implications for tem- perature-based standards of applicator adequacy. Int J Hypertker 1992;8:11-21.

Page 27: Biological rationale and clinical experience with hyperthermia

342 K. Engin

178. Straube WL, Myerson RJ, Emami B, Leybovich LB. SAR patterns of external 915 MHz microwave applicators. Int J Hyperther 1990;6:665-670.

179. Shrivastava P, Luk K, Oleson J, et al. Hyperthermia quality assurance guidelines. Int ] Radiat Oncol Bio2 Phys 1989;16:571-587.

180. Dewhirst MW, Phillips TL, Samulski TV, et al.: RTOG quality assurance guidelines for clinical trials using hyperthermia. Int ] Radiat OncoI Biol Phys 1990;18:1249-1259.

181. Hand JW, Lagendijk JJW, Bach Anderson J, et al. Quality assurance guidelines for ESHO protocols. Int J Hyperther 1989;5:421428.

182. Waterman FM, Dewhirst MW, Fessenden I’, et al. Quality assurance guidelines for clinical trials using hyperthermia administered by ultrasound. Int J Radiaf Oncol Biol Phys 1991;20:1099-1108.

183. Thrall DE, Dewhirst MW, Page RL, et al. A comparison of temperatures in canine solid tumors during local and whole-body hyperthermia administered alone and simultaneously. Int J Hyperther 1990;6:305-317.