throbber
Endocrine—Re/ated Cancer (1 999) 6 75-92
`
`Use of aromatase inhibitors in
`
`breast carcinoma
`
`R J Santen and H A Harvey’
`Department of Medicine, University of Virginia Health Sciences Center. Charlottesville, Virginia 22908, USA
`‘Department of Medicine, Penn State College of Medicine, Hershey, Pennsylvania 17033, USA
`(Requests for offprints should be addressed to R J Santen)
`
`Abstract
`Aromatase, a cytochrome P—45D enzyme that catalyzes the conversion of androgens to estrogens, is
`the major mechanism of estrogen synthesis in the post-menopausal woman. We review some of the
`recent scientific advances which shed light on the biologic significance, physiology, expression and
`regulation of aromatase in breast tissue.
`Inhibition of aromatase, the terminal step in estrogen
`biosynthesis, provides a way of treating hormone—dependent breast cancer in older patients.
`Aminoglutethimide was the first widely used aromatase inhibitor but had several clinical drawbacks.
`Newer agents are considerably more selective, more potent, less toxic and easier to use in the clinical
`setting. This article reviews the clinical data supporting the use of the potent, oral competitive
`aromatase inhibitors anastrozole, letrozole and vorozole and the irreversible inhibitors 4—OH andro-
`stenedione and exemestane. The more potent compounds inhibit both peripheral and intra-tumoral
`aromatase. We discuss the evidence supporting the notion that aromatase inhibitors lack cross-
`resistance with antiestrogens and suggest that the newer, more potent compounds may have a
`particular application in breast cancertreatment in a setting of adaptive hypersensitivity to estrogens.
`Currently available aromatase inhibitors are safe and effective in the management of hormone-
`dependent breast cancer in post—menopausal women failing antiestrogen therapy and should now be
`used before progestational agents. There is abundant evidence to support testing these compounds
`asfirst—line hormonal therapy for metastatic breast cancer as well as part of adjuvant regimens in older
`patients and quite possibly in chemoprevention trials of breast cancer.
`Endocrine-Related Cancer (1999) 6 75-92
`
`Introduction
`
`Epithelial cells of the normal breast undergo dramatic
`changes during various events in a woman’s life such as
`puberty, the follicular and luteal phases of the menstrual
`cycle, pregnancy and menopause. The co—ordinated
`interaction of growth factors and steroid hormones
`regulate the proliferation and differentiated function of
`epithelial and stromal cells in the normal mammary gland.
`The key growth factors are insulin-like growth factor-I,
`prolactin, insulin, the fibroblast growth factor family of
`growth factors and growth hormone, and major steroid
`hormones are estradiol, progesterone and testosterone
`(Frantz & Wilson 1998).
`For the process of inducing breast cancer, estrogens
`appear to play a predominant role. These sex steroids are
`believed to initiate and promote the process of breast
`carcinogenesis by enhancing the rate of cell division and
`reducing time available for DNA repair. An emerging new
`
`can be metabolized to
`that estrogens
`is
`concept
`catecholestrogens and then to quinones which directly
`damage DNA. These two processes —
`the estrogen
`rcecptopmcdiatcd, genomic effects on proliferation and
`the receptor-independent, genotoxic effects of estrogen
`metabolites - can act either in an additive or synergistic
`fashion to cause breast cancer (Santen er a1. 1999).
`Breast cancers which arise in patients can be divided
`into two subtypes:
`those which are dependent upon
`hormones for growth and those which grow independently
`of hormonal stimulation (Santen er‘ a1. 1990). In the
`honnone—dependent subtype,
`the role of estrogens as
`modulators ofmitogenesis overrides the influence of other
`factors. These sex steroids stimulate cell proliferation
`directly by increasing the rate of transcription of early
`response genes such as c—1nyc and indirectly through
`stimulation of growth factors which are produced largely
`in response to estrogenic regulation (Dickson & Lippman
`1995).
`
`Endocrine-Related Cancer (1999) 6 75-92
`1351-0088/99/006-075 © 1999 Society for Endocrinology Printed in Great Britain
`
`Online version via http://vwvw.endocrino|ogy.org
`
`Astrazeneca Ex. 2047 p. 1
`Mylan Pharrns. Inc. v. Astrazeneca AB IPR20l6-01324
`
`

`
`Santen and Harvey.‘ Use of aromatase inhibitors in breast carcinoma
`
`Based upon the concept that estrogen is the proximate
`regulator of cell proliferation, two general strategies were
`developed for treatment of hormone-dependent breast
`cancer: blockade of estrogen receptor action and inhibition
`of estradiol biosynthesis. Antiestrogens such as tamoxifen
`bind to the estrogen receptor and interfere with trans-
`cription of estrogen—induced genes involved in regulating
`cell proliferation Clinical trials showed tamoxifen to be
`effective in inducing objective tumor regressions and to be
`associated with minimal side-effects and toxicity. The
`second strategy, blockade of estradiol biosynthesis, was
`demonstrated to be feasible using the steroidogenesis
`inhibitor, aminoglutethimide, which produced tumor
`regressions equivalent to those observed with tamoxifen
`(Santen ef al. 1990). However, side-effects from amino-
`glutethimide were considerable and its effects on several
`steroidogenie enzymes required concomitant use of a
`glucocorticoid (Santen er
`a1.
`1982). Consequently,
`tamoxifen became the preferred, first—line endocrine agent
`with which to treat advanced breast cancer. However, the
`clinical efficacy of arninoglutethiiude focused attention
`upon the need to develop more potent, better tolerated, and
`more specific inhibitors of estrogen biosynthesis.
`
`Inhibition of estradiol biosynthesis
`
`Multiple strategies could be used to inhibit estradiol
`biosynthesis as a treatment for estrogen-dependent breast
`cancer. Inhibition of several enzymes in the steroidogenie
`pathway,
`including cholesterol side-chain cleavage, 3
`beta—ol—dehydrogenase—delta 4-5 isomerase,
`17—alpha
`hydroxylase,
`l7-beta hydroxysteroid dehydrogenase,
`estrone sulfatase, and aromatase, could be used to reduce
`the biosynthesis of estradiol
`and potentially cause
`horn1one—dependent breast tumor regression. An addition—
`al strategy is the use of exogenous glucocorticoid to inhibit
`release of adrenocoiticotropin (ACTH) and suppress
`estrogen production. Finally, synthetic progestins such as
`megestrol acetate and medroxy—progesterone acetate exert
`glucocorticoid effects and inhibit estradiol synthesis by
`suppressing ACTH.
`
`The ideal strategy would be to block the synthesis of
`estrogen without inhibiting production of other important
`steroids or giving pharmacological amounts of progestins
`or glucocorticoids. For this reason, blockade of the
`tenninal step in estradiol biosynthesis catalyzed by the
`enzyme aromatase is considered a more specific and
`therefore preferable strategy. Several pharmaceutical
`companies sought to develop potent aromatase inhibitors
`designed to specifically block estrogen biosynthesis with-
`out altering glucocorticoid and mineralocorticoid syn-
`thesis, and without requiring addition of large amounts of
`progestins or exogenous glucoeorticoid.
`
`76
`
`Physiology and regulation of aromatase
`Aromatase is a cytochrome l-’-450 enzyme which catalyzes
`the rate—limiting step in estrogen biosynthe sis,
`the
`conversion of androgens to estrogens (Simpson 61‘ al.
`1997, Sasano & Harada 1998). Two major androgens,
`androstenedionc and testosterone, serve as substrates for
`aromatase. The aromatase enzyme consists of a complex
`containing a cytochrome P—450 protein as well as the
`flavoprotein NADPH cytochrome P-450
`reductase
`(Simpson at
`a].
`1997). The gene
`coding for
`the
`cytochrome P-450 protein (P-450 AROM) exceeds 70 kb
`and is
`the largest of the cytochrome P-450 family
`(Simpson er a1. 1993). The cDNA of the aromatase gene
`contains 3.4 kb and encodes a polypeptide of 503 amino
`acids with a molecular weight of 55 kDa. Approximately
`30% homology exists with other cytochrome P-450
`proteins. Because its overall homology to other members
`of the F-450 superfamily is low, aromatase belongs to a
`separate gene family designated CYPI 9.
`Recent studies indicate that the transcription of the
`aromatase gene is highly regulated (Simpson er a]. 1989,
`1993, 1997). The first exon of the aromatase gene is
`transcribed into aromatase message but not translated into
`protein. There exist nine alternative first exons which can
`initiate the transcription of aromatase. Each of these
`alternate exons contains up stream DNA sequences which
`can either enhance or silence the transcription of arom-
`atase. Different tissues utilize specific alternate exons to
`initiate transcription. For example, the placenta utilizes
`alternate exon 1.1, the testis alternate exon H, adipose
`tissue 1.3 and l.4 and brain lf. Enhancers which react with
`upstream elements of these alternate exons markedly
`stimulate the rate of transcription of the aromatase gene.
`Thus, each tissue can regulate the amount of aromatase
`transcribed in a highly specific manner (Simpson er a1.
`1993 ).
`Aroinatase expression occurs in many organs, includ-
`ing ovary, placenta, hypothalamus, liver, muscle, adipose
`tissue, and breast cancer itself. Aromatase eatalyyes three
`separate steroid hydroxylations which are involved in the
`conversion of androstenedione to estrone or testosterone
`to estradiol. The first two give rise to l9—hydroxy and 19-
`aldehyde structures and the third, although still contro-
`versial, probably also involves the C-19 methyl group with
`release of formic acid (Fishrnan & Hahn 1987). This
`enzymatic action results in the saturation of the A-ring of
`the steroid molecule to produce an aromatic structure,
`hence the tenn aromatization.
`
`the major source of
`In the premenopausal state,
`aromatase and of its substrates is the ovary. However,
`extra—glandular aromatization of adrenal substrates in
`peripheral
`sites
`such as
`fat,
`liver and muscle also
`contributes substantially to the estrogen pool in the early
`
`Astrazeneca Ex. 2047 p. 2
`
`

`
`Endocrine-Re/ated Cancer( 1999) 6 75-92
`
`AROMATASE INHIBITORS
`
`
`
`Potency
`
`
`-Fiidiloigl
`" 93%
`
`
`
`
`
`
`First
`
`Generation
`
`Second
`
`Generation
`
`Third
`
`Generation
`
`Figure 1 Diagrammatic representation ofthe potency of aromatase inhibitors as reflected by the isotopic kinetic method for
`determining degree ofaromatase inhibition The percent conversion of androstenedione to estrone is measured isotopically,
`correcting for losses ofestrone by giving 14[C] estrone tracer. Values indicated represent percent inhibition oftotal body aromatase.
`
`follicular and late luteal phases of the riieristiual cycle. 111
`the postmenopausal state, the ovary loses its complement
`of aromatase enzyme although it does continue to secrete
`androstenedione. The adrenal subsumes the primary role
`of providing substrate for aromatase by directly secreting
`testosterone and androstenedione. In addition, dehydro-
`epiaiidrosteroiie and its sulfate are secreted by the adrenal
`and convened into the aromatase substrates, andros—
`tenedione and testosterone,
`in peripheral
`tissues. The
`major source of the aromatase enzyme in postmenopausal
`women is peripheral
`tissues and particularly fat and
`muscle.
`Recent studies identified an additional, important site
`of estrogen production, breast tissue itself. Two-thirds of
`breast carcinomas contain aromatase and synthesiae
`biologically significant amounts of estrogen locally in the
`tumor (Abul-Hajj et a1. 1979, Miller’ & O’Neil 1987,
`Santen er a1. 1994). Proof of local estradiol synthesis
`includes measurement of tumor aromatase activity by
`radiometric or product
`isolation assays, by immuno-
`histochemistry, by demonstration of aromatase mRNA in
`tissue, and by arorriatase enzyme assays performed on
`cells isolated from human tumors and grown in cell
`culture. The expression of aroiriatase is highest in the
`strom al eompartm ent of breast tumors (Santen eta]. l 994)
`but is present in epithelial cells as well. In breast tissue
`
`surrounding the tumors, preadipoeyte fibroblasts contain
`aromatase activity that can be detected by biochemiea
`assay or immunohistocheiiiical staining (Miller & O’Nei
`1987, Santen er a1. 1994). Normal breast tissue also
`contains aromatase as documented by iiiimuiioliisto-
`chemistry, by demonstration of aromatase message, an:
`by enzyme assays of cultured cells (Mor et al. 1998,
`Brodie eta], 1999).
`The biologic relevance of in situ estrogen production
`by aromatase has been demonstrated by xenograf
`experiments which compare tumors containing and no
`containing aromatase (Yue et a1. 1998). Human breas
`cancer cells transfected peniianently with the aromatase
`enzyme
`are
`compared with cells
`transfccted with
`irrelevant DNA. In these experiments, tumors containing
`the transfeetcd aromatase enzyme have higher amounts o ‘
`estrogen and grow faster’ than those with transfeetion 0 ‘
`irrelevant DNA. Further,
`these experiments show that
`local production of estradiol in the tumor is a greater
`source of estrogen than uptake from plasma (Yue et al.
`1998). Taken together,
`these
`studies
`support
`the
`importance of in situ estrogen production by breast tumors
`and suggest that aromatase inhibitors in patients must be
`sufficiently potent to block intra-tumoral aromatase.
`Breast tumor tissue aromatase can be regulated by
`several enhancers of aromatase transcription (Simpson et
`
`77
`
`Astrazeneca Ex. 2047 p. 3
`
`

`
`Santen and Harvey.‘ Use of aromatase inhibitors in breast carcinoma
`
`AROMATASE INHIBITORS
`
`Spectrum of Action
`
`
`
`Third
`
`Generation
`
`
`
`First
`
`Second
`
`_.a!s19S*9r9.ne '
`
`Generation
`
`Generation
`
`Figure 2 Diagrammatic representation ofthe spectrum of action of first through third generation aromatase inhibitors. Vtfith
`development of newer inhibitors, the spectrum of action narrows. The third generation aromatase inhibitors act exclusively on the
`aromatase enzyme and do not appearto exert additional effects.
`
`a]. 1997). Dexaniethasone, phorbol esters, cyclic AM?,
`interleukin 6, and prostaglandins can all stimulate aroma-
`tase transcription in cultured breast cancer cells and
`specifically in the stromal components.
`Interestingly,
`products secreted by epithelial cells in the breast ttunors
`appear to stimulate aromatase in the stroma and provide a
`means
`for autoregulation of tumor growth through
`estrogen production. A rather novel means ofregulation of
`aromatase levels was also recently described —
`the
`stabilization of degradation of enzyme (Harada et a1.
`1999). Aromatase inhibitors bind to the active site of the
`enzyme
`and,
`through mechanisms not
`completely
`understood, prevent proteoly sis of the aromatase protein.
`Each of these mechanisms may enhance the amount of
`aromatase in tumor tissue and increase the need for very
`potent aromatase inhibitors.
`
`Development of aromatase inhibitors
`The first aromatase inhibitors were discovered nearly 30
`years ago and included aminoglutethimide and testolo-
`lactone (Santen er al. 1990). Testololaetone was not very
`potent as an inhibitor, and aminoglutethiniide blocked
`several P-450-mediated enzymatic reactions and was
`
`associated with troublesome side—effects. On the other
`hand, aminoglutethimide appeared to be quite effective in
`causing tumor regressions in patients with breast cancer.
`For this reason, pharmaceutical companies and individual
`investigators focused upon developing more potent and
`specific inhibitors. Second and third generation inhibitors
`were developed with 10- to l00OO-fold greater potency
`than aminoglutethimide and greater specificity (Figs l and
`2). The half—lives of the inhibitors increased with synthesis
`of more potent inhibitors. The third generation aromatase
`inhibitors are capable of decreasing the levels of circu-
`lating estrogens to a greater extent than the first and
`second generation inhibitors in postmenopausal women
`with hormone-dependent breast cancer. Hypothetically,
`these highly potent agents could also reduce levels of
`intra-tumoral aromatase activity to a greater extent than
`the earlier inhibitors but this has not yet been examined.
`
`Pharmacologic classification of
`aromatase inhibitors
`into
`inhibitors
`A convenient
`classification divides
`mechanism based or ‘suicide inhibitors’ (Type 1) and
`competitive inhibitors (Type ll) (Brodie 1993). Suicide
`
`78
`
`Astrazeneca Ex. 2047 p. 4
`
`
`
`- 4—0iii-\¥
`
`-Inhibition
`
`aromatase
`
`

`
`Endocrine-Re/ated Cancer( 1999) 6 75-92
`
`DJ0
`
`20
`
`10
`
`
`
`(pmollLestradiolequivalents)
`
`Estrogen
`
`RIA
`
`I Bioassay
`
`
`
`Baseline
`
`Week 6
`
`Week 1 2
`
`Time of Treatment
`
`Figure 3 Inhibition of plasma estrogen levels as assessed by RIA and by an ultrasensitive, recombinant DNA-based bioassay
`(Jones et al. 1992). Basal estradiol levels are approximately threefold lower when measured by the ultrasensitive assay. During
`administration of the aromatase inhibitor, levels fall to 0.05-0.07 pmolll as assessed by the ultrasensitive assay and to 2-5 pg/ml
`with the standard RIA.
`‘P< 0.01 vs baseline.
`
`inhibitors initially compete with natural substrates (i.e.
`androstenedione and testosterone) for binding to the active
`site of the enzyme. The enzyme, then, specifically acts
`upon the inhibitor to yield reactive alkylating species
`which form covalent bonds at or near the active site of the
`enzyme. Through this mechanism, the enzyme is irrever-
`sibly inactivated. Competitive inhibitors, on the other
`hand, bind reversibly to the active site of the enzyme and
`prevent product formation only as long as the inhibitor
`occupies the catalytic site. Whereas mechanism—based
`inhibitors are exclusively steroidal in type, competitive
`inhibitors consist both of steroidal and non-steroidal
`compounds (Brodie 1993).
`
`Methods used to demonstrate
`aromatase inhibition
`
`The standard method to study aromatase inhibitors in
`patien s is to measure either plasma or urinary estrogen by
`RIA. Early studies demonstrated 50-80% inhibition of
`plasma or urinary estrone or estradiol (Santen e1 21]. 1978,
`
`1981, 1982). Another method involved measurement of
`each estrogen metabolite in urine with calculation of total
`aromatized product. This technique provided results
`similar to those from measurements of urinary estrone or
`estradiol (Lipton 61 a1. 1995). Using these plasma or
`urinary methods, each agent appeared to suppress estrogen
`levels to concentrations approaching the sensitivity of the
`RlAs used. To gain greater specificity and sensitivity,
`investigators utilized the isotopic kinetic technique of
`Siiteri el al. to measure total body aromatase (Grodin ei a1.
`1973, Santcn er a1. 1978, Jones et a1. 1992, Dowsctt er a1.
`1995). This required adn1inist1’ation of tritiated andros-
`tenedione and 14[C]-estrone to patients under steady-state
`conditions and measurement of radiochemically pure
`tritiated estrone and estradiol (Santen er a]. 1978). The
`14[C]-estrone
`allowed correction for
`losses during
`multiple purification steps. Using this technique,
`the
`degree of inhibition with various inhibitors ranged from
`90 to 99%.
`From these observations, it was recognized that more
`sensitive plasma assays of estradiol were needed. One
`
`79
`
`Astrazeneca Ex. 2047 p. 5
`
`

`
`Santen and Harvey.‘ Use of aromatase inhibitors in breast carcinoma
`
`approach was tl1e use of the plasma estrone sulfate assay
`since basal levels of this conjugate in postmenopausal
`women are tenfold higher than the levels of uneonjugated
`estrone and estradiol (Sarnojlik eta]. 1982, Lonning et a1.
`1997). With this measurement, suppression to 85% of
`basal values was observed with most inhibitors. Finally, an
`ultrasensitive bioassay of plasma estradiol which was 50-
`to 100-fold more sensitive than RIA was developed
`(Oerter—Kleir1 el al. 1995). Surprisingly, with this assay,
`one could demonstrate suppression to levels of estradiol of
`0.05-0.07 pg/ml, concentrations substantially lower than
`the 2-5 pg/ml suppressed levels detected by RIA (Fig. 3).
`As observed with use of other highly sensitive plasma
`hormone assays, for example for luteinizing hormone
`(TH), follicle-stimulating hormone (FSH),
`thyrotropin
`(TSH), and growth hormone, the levels measured under
`basal conditions and during suppression with these assays
`reveals much lower values than with insensitive RIAS.
`This probably reflects the fact that insensitive assays are
`measuring a substantial fraction of ‘blank’ or non-specific
`assay artifact. With the use ofhighly sensitive assays, this
`artifactual measurement
`is eliminated and the actual
`values measured are much lower. Thus with the ultra-
`sensitive estradiol bioassay,
`the basal
`levels in post-
`menopausal women average 1-3 pg/1111 (vs 5-20 pg/ml
`with RIA) (Oerter-Klein et a1. 1995). During development
`of the second and third generation aromatase inhibitors,
`each of these methods has been used to demonstrate the
`magnitude of suppression of enzymatic activity. For these
`measurements, the isotopic kinetic technique is consid-
`ered the “gold standard’ since it is highly sensitive and
`allows comparison among various inhibitors (Fig. 1).
`
`First generation aromatase inhibitors
`The first aromatase inhibitor to be widely used in the
`treatment of metastatic breast cancer i11 postmenopausal
`women was the drug aminoglutethimide (Santen ct ai.
`1978,
`1981,
`1982,
`1990).
`lsotopic kinetic
`studies
`demonstrated a 90-95% inhibition of aromatase activity
`(Santen er a1. 1978). Plasma estrone and estradiol levels
`and urinary estrogens fell by 50-80% in response to this
`aromatase inhibitor. An additional effect, described by
`Lom1ing
`and colleagues, was
`the
`acceleration of
`metabolism of estrogen sulfate (Geisler et a1. 1997). This
`effect resulted in further lowering of free estrogen levels
`in plasma and in urine. With further study of a1nino—
`glutethimide, multiple metabolic effects were demon-
`strated,
`including inhibition of 11-beta hydroxylase,
`aldosterone synthase, and thyrcxine synthesis as well as
`induction of enzymes metabolizing synthetic g1ucocorti-
`coids and aminoglutethimide itself (Santen er a1. 1990).
`When aminoglutethimide was combined with a
`corticosteroid such as hydrocortisone,
`the regimen
`
`80
`
`produced durable clinical responses in 30-50% of patients
`(Santen er a1. 1990). This approach, however, had several
`important drawbacks. First, aminoglutethimide was asso-
`ciated with troublesome side-effects, including drowsi-
`ness, skin rash, and ataxia. Secondly, standard doses of
`1000 mg aminoglutethimide daily could also inhibit other
`cytochrome P-450-mediated steroid hydroxylations, par-
`tieularly those involving the cholesterol
`side-chain
`cleavage enzymes (Santen et a1. 1990, Cocconi 1994).
`This non-selectivity for aromatase led to inhibition of the
`biosynthesis of cortisol, aldosterone and also of thyroid
`hormone. This necessitated co-administration of the
`glueocorticoid, hydroeortisone, and in about 5% of
`patients, thyroxine.
`Four randomized. controlled clinical trials compared
`aminoglutethimide in combination with hydrocortisone
`with tamoxifen in advanced breast cancer. (Smith eta].
`1981, Lipton et a1. 1982, Alonso—Mun0z et a]. 1988, Gale
`et a1. 1994). The antiestrogen tamoxifen and the inhibitor
`of estrogen biosynthesis, aminoglutethimide/hydroeorti-
`sone produced similar rates of objective disease regression
`and duration of response (Santen et a1. 1990, Gale et a1.
`1994). Tamoxifen produced many fewer side-effects than
`did aminoglutethimide/hydrocortisone. Cross-over
`re-
`sponses to aminoglutethimide/hydrocortisone in patients
`relapsing on tamoxifen were substantial, ranging from 25
`to 50% and 36% in the largest randomized study (Gale at
`al. 1994). In marked contrast, patients initially treated with
`aminoglutethimide/hydrocortisone responded less
`fre-
`quently when crossed over to tamoxifen (19%) (Gale eta].
`1994). This observation reinforced the concept that the
`antiestrogens be used as
`first-line
`agents and the
`aromatase inhibitors as second- or third-line therapies.
`With the development of better aromatase inhibitors,
`aminoglutethimide is now of historical interest only.
`
`Second generation aromatase inhibitors
`
`Fadrozole
`
`4—(5,6,7,8—tetrahy dro-
`l6949A',
`(CGS
`Fadrozole
`imidazo[1,5a]-pyridin-5yl)
`bcnzonitrile monohydro-
`chloride) is a fairly potent inhibitor of aromatase witl1 an
`inhibitory constant
`(IQ) of 0.19 nM (vs 600 nM for
`aminoglutethimide) (Harvey et a1. 1994, Harvey 1996).
`Cholesterol side-chain cleavage activity is minimal but C-
`11 hydroxylase inhibitory effects are observed in I/Itno at
`high drug concentrations.
`lnitial dose-seeking studies conducted in patients
`demonstrated effective aromatase inhibition at doses of
`1.8-4.0 mg daily (Harvey et a1. 1994). A phase 11 study
`then compared doses of 0 6 mg three times daily,
`1 mg
`twice daily, and 2 mg twice daily. Maximal suppression of
`plasma and urinary estrogens occurred at a dose of 1.0 mg
`
`Astrazeneca Ex. 2047 p. 6
`
`

`
`Table 1 Comparison ofthird generation aromatase inhibitors with progestin therapy
`
`Endocrine-Re/ated Cancer( 1999) 6 75-92
`
`Meg ace vs vo rozo|e*
`
`Megace vs anastrozole (1 mg)
`
`Megace vs letrozole (2.5 mg)
`
`Response
`parameters Megace
`Overall
`28.7
`survival
`months
`
`Vorozole
`26
`months
`
`P
`NS
`
`Megace
`22.5
`months
`
`Anastrozole
`26.7
`months
`
`P
`0.02
`
`Megace
`215
`months
`
`Letrozole
`25.3
`months
`
`P
`0.15
`
`Objective
`response
`rates
`(CR+PR)
`
`Clinical
`benefit
`(CR+PR+
`stab|e>6
`months)
`
`Timeto
`progression
`
`Number in
`study
`
`7.6%
`
`10.5%
`
`NS
`
`7.9%
`
`10.3%
`
`NS
`
`16.4%
`
`23.6%
`
`0.04
`
`Not
`reported
`
`Not
`reported
`
`26.1%
`
`35%
`
`NS
`
`32%
`
`35%
`
`NS
`
`3.6
`months
`
`2.7
`months
`
`452
`
`NS
`
`5
`months
`
`5
`months
`
`764
`
`NS
`
`5.5
`months
`
`0.07
`
`5.6
`months
`
`551
`
`NS, not significant.
`"Goss (1998).
`Megace, megestrol acetate
`
`twice daily and minimal effects on cortisol secretion were
`observed. Basal cortisol and AC TH levels were unaffected
`and cortisol
`levels increased appropriately after exog-
`enous syntlietic ACTH (cortrosyn) administration in all
`patients. Basal levels of aldosterone also remained stable
`following administration of all three drug doses. There
`were no changes
`in urinary or plasma sodium or
`potassium, nor in standing blood pressure to suggest a
`clinical
`state of aldosterone
`deficiency. However,
`cortrosyn- stimulated aldo sterone levels were significantly
`blunted at all three doses. (Santen 61 a1. 1991). Based on
`several phase 11 trials, toxicity attributed to this agent was
`mild and consisted mainly of nausea, anorexia, fatigue,
`and hot
`flashes. The potency of the compound,
`its
`relatively specific effects on aromatase and its lack of
`toxicity
`suggested that
`it might provide
`a major
`improvement over aniinoglutcthimidc for treatment of
`patients with breast cancer.
`Two large multicenter phase 111 trials in the USA
`comparing fadrozole hydro chloride to niegestrol acetate in
`patients who had received only tamoxifen as prior
`hormonal therapy have now been completed (Buzdar er a].
`19961), Trunet er a1. 1997). These two studies accrued a
`total of 672 patients. Final clinical results showed that
`there were no significant differences between the two
`treatment arms of the trials with respect
`to time to
`progression, objective response rates, response duration or
`
`overall survival. In these two trials, responses to mcgestrol
`acetate were somewhat lower than expected from previous
`studies with objective response rates of 11 and 13%
`respectively. Randomized patients receiving fadrozole
`experienced objective responses of 11 and 16% which did
`not differ significantly from those with megestrol. Stable
`disease for more than 6 months occurred in 25% of
`patients receiving fadrozole and 20% taking inegestrol
`acetate. Nausea was more frequent for fadrozole than
`megestrol acetate in both trials (22 vs 13% and 36% vs
`11% respectively). lncontrast, edema was commoner with
`megestrol acetate (21 vs 12% and 19 vs 12%) as was
`weight gain.
`compared fadrozole with tamoxifen
`Two trials
`(Falkson & Falkson 1996, Thurlimann et al. 1996). In the
`first, 1 mg fadrozole twice daily was compared with 20 mg
`tamoxifen daily in 212 postmenopausal patients with
`riietastatic breast cancer. Response rates to tamoxifen
`(27%) and to fadrozole (20%) did not differ significantly
`nor did response durations (20 months vs 15 months).
`However, tamoxifen achieved a significantly longer time
`to treatment failure (8.5 months vs 6 months, P<0.05). In
`the second study. fadrozole was compared with tamoxifen
`as first—line therapy in a randomized, controlled trial
`conducted in South Africa. Response rates to tamoxifen
`were 48% vs 43% with fadrozole (P=riot significant).
`However, response duration was significantly longer with
`
`81
`
`Astrazeneca Ex. 2047 p. 7
`
`

`
`Santen and Harvey.‘ Use of aromatase inhibitors in breast carcinoma
`
`tamoxifen (median duration not reached vs 343 days,
`P<0.009) as was overall
`survival
`(34 months
`for
`tamoxifen vs 26 months for fadrozole, I’<0.0-16).
`Taken together,
`these studies demonstrated that
`fadrozole may be inferior to tamoxifen in efficacy and no
`better tolerated than rnegestrol acetate. Based upon these
`findings,
`the second generation aromatase inhibitor,
`fadrozole, would likely find its place as third-line therapy.
`Fadrozole has been approved for
`the treatment of
`advanced breast cancer in postmenopausal women in
`Japan. This agent is not likely to be further developed in
`the United States since both anastrozole and letrozole
`appear to be more potent and more selective aroinatase
`inhibitors.
`
`Careful analysis of the fadrozole/inegestrol acetate
`trials raises the concern that responses to endocrine
`therapies appeared to be less frequent than observed in
`prior studies. For example, the randomised comparison of
`the first generation aromatase inhibitor, aminoglute—
`thimide, with surgical
`adrenalectomy demonstrated
`responses of 40-50% in patents previously treated with
`tamoxifen (Santen et ai. 1981). Other
`studies with
`megestrol acetate as second—line therapy demonstrated
`responses ranging from 30 to 50%. Several possibilities
`could explain the low response rates. In recent studies,
`more stringent criteria have been used than in previous
`trials. For example, reealcification of mixed lytic/blastic
`metastases were previously considered objective evidence
`of partial responses. Such lesions are now considered non-
`assessable, non-measurable disease. External review of
`cases probably also increases the stringency of assess-
`ment.
`lt should be noted that
`in a previous study
`comparing tamoxifen alone vs tamoxifen and fluoxy-
`mestrone, the objective response rate for tam oxifcn alone
`was only 10% (Swain et a1. 1988). These considerations
`lead to the conclusion that one can only compare new
`agents with established ones such as tamoxifen and assess
`the relative differences between them. It is inappropriate
`to compare the percent of objective responses to those
`observed in historical controls.
`
`4-Hydroxyandrostenedione (4-OHA)
`Formestane (Lentaron, 4-OHA‘, 4-l1ydroxyandrost-4-ene-
`3,17—dione) is a structural analog of androstenedione and
`is thus a highly specific aromatase inhibitor (Lonning
`1998).
`It was the first steroidal suicide—type (Type 1)
`aromatase inhibitor to enter clinical
`trials and is now
`commercially available in Europe. Using the in Vjtm
`placental aromatase assay system, 4-OHA was shown to
`be 60-fold more potent than aininoglutethimide (Ki=-1.1
`uM). Extensive studies revealed no estrogenic, anti-
`estrogenic, or antiandrogenic properties (Brodie & V1/ing
`1987). However, transformation to 4—hydroxytestosterone
`
`82
`
`occurs and androgenic effects can be demonstrated under
`certain circumstances (Brodie er a1. 1981).
`4-OHA (Lentaron) has been studied extensively in
`Europe in postmenopausal women with breast cancer.
`Data from four phase ll clinical
`trials of 4-OHA
`demonstrated a 33% objective regression rate of breast
`cancer in postmenopausal patients previously treated with
`multiple endocrine therapies. Toxicity included six
`patients with sterile abscesses due to intramuscular
`injections, two of sufficient severity to warrant discon-
`tinuation of therapy. No androgenic effects were observed
`(Goss eta]. 1986).
`Hoffken eta]. ( 1990) conducted a large trial of 4-OHA
`in postmenopausal women. Patients initially received 500
`mg intramuscularly every two weeks for 6 weeks and then
`250 rrrg every 2 weeks thereafter. Plasrrra estradiol levels
`fell from baseline values of 10-11 pg/ml to levels of
`approximately 4 pg/ml for up to 7 months of therapy. The
`drug appeared specific since no reduction of cortisol or
`symptoms of cortisol deficiency were observed. Of 86
`evaluable patients, there were 2 complete and 19 partial
`remissions (24%) and 26 with disease stabilization (30%).
`Side-effects included minor systemic symptoms in 11%
`(hot flashes, constipation, alopecia, and pruritns) and local
`symptoms in 8% (pruritus,
`local pain, and erytherna).
`These side-effects resulted in discontinuation oftherapy in
`only 2% of patients.

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket