14.01.2013 Views

Somatostatin Analogs for Cancer Treatment and Diagnosis

Somatostatin Analogs for Cancer Treatment and Diagnosis

Somatostatin Analogs for Cancer Treatment and Diagnosis

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Somatostatin</strong><br />

<strong>Analogs</strong> in <strong>Cancer</strong><br />

Management<br />

Editor<br />

C. Scarpignato, Parma/Nantes<br />

17 figures, 1 in color, <strong>and</strong> 16 tables, 2001<br />

ABC Basel � Freiburg � Paris � London � New York �<br />

New Delhi � Bangkok � Singapore � Tokyo � Sydney


S. Karger<br />

Medical <strong>and</strong> Scientific Publishers<br />

Basel � Freiburg � Paris � London<br />

New York � New Delhi � Bangkok<br />

Singapore � Tokyo � Sydney<br />

ABC<br />

Fax+ 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

Drug Dosage<br />

The authors <strong>and</strong> the publisher have exerted every<br />

ef<strong>for</strong>t to ensure that drug selection <strong>and</strong> dosage set<br />

<strong>for</strong>th in this text are in accord with current recommendations<br />

<strong>and</strong> practice at the time of publication.<br />

However, in view of ongoing research,<br />

changes in government regulations, <strong>and</strong> the constant<br />

flow of in<strong>for</strong>mation relating to drug therapy<br />

<strong>and</strong> drug reactions, the reader is urged to check<br />

the package insert <strong>for</strong> each drug <strong>for</strong> any change in<br />

indications <strong>and</strong> dosage <strong>and</strong> <strong>for</strong> added warnings<br />

<strong>and</strong> precautions. This is particularly important<br />

when the recommended agent is a new <strong>and</strong>/or infrequently<br />

employed drug.<br />

All rights reserved.<br />

No part of this publication may be translated into<br />

other languages, reproduced or utilized in any<br />

<strong>for</strong>m or by any means, electronic or mechanical,<br />

including photocopying, recording, microcopying,<br />

or by any in<strong>for</strong>mation storage <strong>and</strong> retrieval<br />

system, without permission in writing from the<br />

publisher or, in the case of photocopying, direct<br />

payment of a specified fee to the Copyright Clearance<br />

Center (see ‘General In<strong>for</strong>mation’).<br />

© Copyright 2001 by S. Karger AG,<br />

P.O. Box, CH–4009 Basel (Switzerl<strong>and</strong>)<br />

Printed in Switzerl<strong>and</strong> on acid-free paper by<br />

Reinhardt Druck, Basel


ABC<br />

Fax+ 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

Vol. 46, Suppl. 2, 2000<br />

Contents<br />

Access to full text <strong>and</strong> tables of contents,<br />

including tentative ones <strong>for</strong> <strong>for</strong>thcoming issues:<br />

www.karger.com/journals/che/che_bk.htm<br />

V Foreword<br />

Lamberts, S.W.J. (Rotterdam)<br />

VIII Preface<br />

Scarpignato, C. (Parma/Nantes)<br />

1 <strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong> <strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong>:<br />

An Overview<br />

Scarpignato, C.; Pelosini, I. (Parma/Nantes)<br />

30 Antiproliferative Effect of <strong>Somatostatin</strong> <strong>and</strong> <strong>Analogs</strong><br />

Bousquet, C.; Puente, E.; Buscail, L.; Vaysse, N.; Susini, C. (Toulouse)<br />

40 Established Clinical Use of Octreotide <strong>and</strong> Lanreotide in<br />

Oncology<br />

Öberg, K. (Uppsala)<br />

54 The Palliative Effects of Octreotide in <strong>Cancer</strong> Patients<br />

Dean, A. (Nedl<strong>and</strong>s)<br />

62 Management of Breast <strong>Cancer</strong>: Is There a Role <strong>for</strong><br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong>?<br />

Boccardo, F.; Amoroso, D. (Genoa)<br />

78 <strong>Somatostatin</strong>, Its Receptors <strong>and</strong> <strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

O’Byrne, K.J. (Leicester); Schally, A.V. (New Orleans, La.); Thomas, A.<br />

(Leicester); Carney, D.N. (Dublin); Steward, W.P. (Leicester)<br />

109 Octreotide in the Management of Hormone-Refractory<br />

Prostate <strong>Cancer</strong><br />

Vainas, I.G. (Thessaloniki)


127 Gastrointestinal <strong>Cancer</strong> Refractory to Chemotherapy:<br />

A Role <strong>for</strong> Octreotide?<br />

Cascinu, S.; Catalano, V.; Giordani, P.; Baldelli, A.M.; Agostinelli, R.;<br />

Catalano, G. (Pesaro)<br />

134 Pancreatic <strong>Cancer</strong>: Does Octreotide Offer Any Promise?<br />

Rosenberg, L. (Montreal)<br />

150 Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong> Biliary Tree<br />

Kouroumalis, E.A. (Heraklion)<br />

162 <strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology: A Look to the Future<br />

Jenkins, S.A. (Swansea); Kynaston, H.G. (Cardiff); Davies, N.<br />

(London); Baxter, J.N. (Swansea); Nott, D.M. (London)<br />

197 Author Index<br />

198 Subject Index<br />

IV Contents


ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Foreword<br />

<strong>Somatostatin</strong> plays an inhibitory role in<br />

the regulation of the function of a number of<br />

organs like the brain, the anterior pituitary<br />

gl<strong>and</strong>, the gastrointestinal tract, the exocrine<br />

<strong>and</strong> endocrine pancreas, as well as lymphoid<br />

cells. It can be considered as an inhibitory<br />

(growth) factor in these organs, which mainly<br />

prevents local overreaction from a multitude<br />

of stimulatory factors.<br />

In addition to the negative role in controlling<br />

the physiological regulation of these organ<br />

systems, somatostatin also exerts inhibitory<br />

effects on the proliferation of normal <strong>and</strong><br />

neoplastic cells. <strong>Somatostatin</strong> analogs inhibit<br />

tumor growth in a wide variety of experimental<br />

models in several species, like transplantable<br />

osteo- <strong>and</strong> chondrosarcomas, transplantable<br />

acinar <strong>and</strong> ductal pancreatic carcinomas,<br />

as well as different types of rat <strong>and</strong> mouse<br />

mammary <strong>and</strong> prostatic carcinomas. Also, a<br />

number of human pancreatic, colonic, gastric<br />

<strong>and</strong> small cell lung cancer lines xenografted in<br />

nude mice are inhibited in their growth during<br />

therapy with somatostatin analogs.<br />

While most of these experimental tumors<br />

<strong>and</strong> cell lines express a dense <strong>and</strong> homogeneous<br />

distribution of somatostatin receptors,<br />

some of these tumors seem to be inhibited in<br />

growth by somatostatin analog administra-<br />

tion via indirect mechanisms involving inhibitory<br />

effects on local or general growth factors<br />

(growth hormone, insulin-like growth factor<br />

1, epidermal growth factor, gastrointestinal<br />

hormones) <strong>and</strong>/or angiogenesis.<br />

One should realize, however, that in most<br />

of the experimental models the inhibitory effects<br />

of somatostatin analogs on growth are<br />

most potent early after tumor implantation as<br />

well as early after the start of drug administration,<br />

<strong>and</strong> that complete growth curves of the<br />

(transplanted) tumors with <strong>and</strong> without somatostatin<br />

analog treatment often indicate a<br />

delay in growth only, while an escape of tumor<br />

growth from the inhibitory effects of somatostatin<br />

analogs is observed eventually. This<br />

indicates a decreased sensitivity of these tumors<br />

during long-term somatostatin analog<br />

therapy, which involves somatostatin receptor<br />

downregulation <strong>and</strong>/or the selection of<br />

somatostatin receptor-negative tumor cell<br />

clones.<br />

In clinical oncology prostate, breast <strong>and</strong><br />

endometrial carcinomas are known to be<br />

‘conditional’ cancers, which grow in certain<br />

specific hormonal environmental conditions.<br />

When these conditions are altered the tumors<br />

regress, but do not die. After a certain period<br />

of a persistent nonproliferating state, part of<br />

V


the tumor cells resume active growth. This is<br />

the simple basis <strong>for</strong> the success of <strong>and</strong>rogen,<br />

estrogen <strong>and</strong> progesterone depletion in the<br />

treatment of these cancers. The symptomatic<br />

relief of patients with metastatic prostatic,<br />

breast <strong>and</strong> endometrial cancers during treatment<br />

with anti<strong>and</strong>rogens, antiestrogens <strong>and</strong><br />

progesterone receptor-blocking agents is highly<br />

appreciated <strong>and</strong> looked <strong>for</strong> by oncologists<br />

that otherwise mainly use intensive chemotherapy.<br />

The promising data demonstrating that<br />

many experimental tumor models also seem<br />

to be ‘conditionally’ dependent on somatostatin,<br />

growth hormone, insulin-like growth factor<br />

1, prolactin <strong>and</strong> other hormones <strong>and</strong><br />

growth factors has raised hopes that direct or<br />

indirect blockade of their activity <strong>and</strong>/or receptors<br />

on metastatic human cancers with<br />

well-tolerated drugs like somatostatin analogs,<br />

bromocriptine <strong>and</strong> retinoids would also<br />

induce a transient state of dormancy or even a<br />

slight temporary regression. Also, the somatostatin<br />

analog octreotide is widely <strong>and</strong> successfully<br />

used in patients with acromegaly,<br />

metastatic islet cell tumors <strong>and</strong> carcinoids. In<br />

these patients the quality of life improves <strong>and</strong><br />

there is strong evidence <strong>for</strong> control of tumor<br />

growth as well as a clear prolongation of survival.<br />

Like most hormone-secreting tumors<br />

many human adenocarcinomas originating<br />

from the breast, colon, kidney, ovary as well<br />

as meningiomas <strong>and</strong> malignant lymphomas<br />

often express somatostatin receptors. Although<br />

somatostatin receptors are in many<br />

cases not distributed homogeneously over all<br />

tumor cells, <strong>and</strong> their numbers per tumor cell<br />

in general are lower than in hormone secreting<br />

tumors, most somatostatin receptor-bearing<br />

human cancers can be visualized by somatostatin<br />

receptor imaging. Only little evidence<br />

so far indicates that long-term therapy of patients<br />

with somatostatin receptor-positive tu-<br />

VI<br />

mors with somatostatin analogs induces a decrease<br />

or even control of tumor growth. More<br />

evidence points to a transient improvement<br />

in the quality of life in some of the patients<br />

treated.<br />

Human cancers differ in many respects<br />

from the experimental tumor models that respond<br />

so well to somatostatin analog therapy.<br />

(1) Most human cancers consist of a mixture<br />

of varying amounts of stromal tissue <strong>and</strong> different<br />

clones of epithelial tumor cells that do<br />

not uni<strong>for</strong>mly express somatostatin receptors.<br />

This sharply contrasts with the monoclonal<br />

tumor models in animals, which in most instances<br />

express somatostatin receptors on all<br />

tumor cells. (2) <strong>Somatostatin</strong> receptor expression<br />

in human breast, prostate <strong>and</strong> colonic<br />

cancers often indicates loss of differentiation<br />

of the tumors, meaning that these undifferentiated<br />

tumors have a bad prognosis. (3) The<br />

nature of new clinical trials in oncology is<br />

often such that patients are mainly included<br />

‘late’ in their disease, when tumors have already<br />

progressed considerably. Also, it remains<br />

uncertain whether somatostatin receptor subtype<br />

2 (SSTR-2)-specific analogs like octreotide<br />

are the optimal compounds to be used in<br />

the treatment of human cancer. These analogs<br />

have little activity towards the SSTR-3 subtype,<br />

which is important in mediating apoptosis.<br />

In vitro studies have demonstrated that<br />

SSTR-2-expressing tumor cell lines as well<br />

as primary cultures of human tumors internalize<br />

radiolabeled somatostatin analogs like<br />

111 In-[DTPA 0 ]octreotide <strong>and</strong> [ 90 Y-DOTA 0 ,<br />

Tyr 3 ]octreotide. Preclinical studies using experimental<br />

tumor models have now demonstrated<br />

that tumor growth can be inhibited by<br />

administration of these two radiopharmaceutical<br />

compounds. Clinical trials already demonstrated<br />

promising effects using these radiopharmaceuticals<br />

on tumor size in patients<br />

with advanced somatostatin receptor-positive<br />

neuroendocrine tumors. Also, the concept of<br />

Foreword


targeted chemotherapy to deliver chemotherapeutic<br />

compounds selectively to somatostatin<br />

receptor-positive tumor cells, thereby reducing<br />

their toxicity, has now been validated<br />

using newly developed cytotoxic somatostatin<br />

analogs in experimental mouse <strong>and</strong> rat models<br />

of human pancreatic, breast <strong>and</strong> prostate<br />

cancers.<br />

In this supplement to Chemotherapy Professor<br />

Scarpignato has succeeded in bringing<br />

together the most knowledgeable scientists in<br />

the field of oncology that have extensive personal<br />

experience with the use of somatostatin<br />

analogs in the treatment of cancer patients.<br />

Foreword<br />

They review the current status of their use in<br />

an admirable manner, while the perspectives<br />

of newer <strong>for</strong>ms of therapy with the targeted<br />

somatostatin receptor-mediated approach are<br />

also discussed.<br />

I would like to congratulate the editor <strong>and</strong><br />

the contributors on the exhaustive <strong>and</strong> balanced<br />

description of the current thoughts on<br />

the use of somatostatin analogs in patients<br />

with different <strong>for</strong>ms of metastasized cancer.<br />

S.W.J. Lamberts, MD, PhD<br />

Professor of Medicine<br />

University Hospital Dijkzigt, Rotterdam<br />

VII


ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Preface<br />

Despite the enormous advances in science<br />

<strong>and</strong> medical technology in recent times, our<br />

knowledge of the pathophysiology <strong>and</strong> treatment<br />

of neoplasia is far from complete <strong>and</strong>,<br />

<strong>for</strong> the patients at least, the myths that surround<br />

cancer remain intact. <strong>Cancer</strong> remains<br />

the ‘bogy man’ of medicine as we enter the<br />

new millennium, <strong>and</strong> the very mention of the<br />

word strikes mortal fear in patients <strong>and</strong> their<br />

families in a way not generally seen with any<br />

other disease.<br />

Although chemotherapy is very effective in<br />

the management of certain neoplasms such as<br />

testicular cancer, the efficacy of this therapeutic<br />

modality in the treatment of many<br />

common malignancies such as those of the<br />

lung, breast, prostate, bowel, pancreas <strong>and</strong><br />

kidney is limited. Cure of macroscopic metastatic<br />

disease is exceedingly rare, <strong>and</strong> palliation<br />

of symptoms of metastatic neoplasms by<br />

chemotherapy can be problematic since the<br />

toxicity of the treatment often outweighs any<br />

improvement in quality of life resulting from<br />

a temporary decrease in tumor burden. This<br />

situation has not only motivated attempts to<br />

develop novel cytotoxic agents <strong>and</strong> targeted<br />

chemotherapy, but has also stimulated research<br />

on innovative noncytotoxic therapies<br />

<strong>for</strong> cancer.<br />

Amongst the various hormonal agents, currently<br />

being evaluated <strong>for</strong> the management of<br />

neoplasia, considerable attention is directed<br />

to somatostatin analogs. This is largely due to<br />

the demonstration of antineoplastic activity<br />

of these compounds in a variety of experimental<br />

models in vitro <strong>and</strong> in vivo <strong>and</strong> to the elucidation<br />

of some aspects of the molecular<br />

mechanisms whereby they exert their cytostatic<br />

<strong>and</strong> cytotoxic effects. Furthermore,<br />

clinical experience with somatostatin analogs<br />

in the treatment of conditions such as acromegaly<br />

<strong>and</strong> GEP tumors has shown that they<br />

are well tolerated compared to other antineoplastic<br />

therapies currently in use. As a consequence,<br />

there is much ongoing clinical research<br />

to determine whether or not results<br />

from experimental studies will translate into<br />

clinically useful antineoplastic activity.<br />

In the recent years there has been extensive<br />

international media coverage of the allegedly<br />

successful treatment of a number of malignant<br />

neoplasms with the Di Bella multitherapy<br />

(DBM). This is a multidrug, individually<br />

tailored medical treatment (comprising a<br />

mixture of melatonin, bromocriptine, somatostatin,<br />

a solution of retinoids, <strong>and</strong>, depending<br />

on the kind of cancer, either cyclophosphamide<br />

or hydroxyurea) developed by Luigi


Di Bella, an Italian physician. Over the past<br />

25 years Dr. Di Bella has consistently used it<br />

on an outpatient basis claiming its effectiveness<br />

in halting the progression or completely<br />

curing most cancers. Not surprisingly, in view<br />

of the public interest in cancer therapy, a<br />

number of associations have been <strong>for</strong>med to<br />

support this treatment. As a matter of fact,<br />

these associations mounted a campaign to<br />

request that DBM be included among those<br />

cancer treatments considered to be effective<br />

<strong>and</strong> that its cost be fully reimbursed by the<br />

Italian National Health Service. Despite some<br />

debate about the methodology of the different<br />

clinical trials [1–8], a phase II study, coordinated<br />

by the Italian National Institute of<br />

Health <strong>and</strong> the National <strong>Cancer</strong> Advisory<br />

Committee, did not show sufficient efficacy<br />

in patients with advanced cancer to warrant<br />

further clinical testing [9]. Furthermore, a follow-up<br />

of treated patients did not give any<br />

evidence that DBM improves the survival of<br />

cancer patients [10].<br />

Since somatostatin or octreotide represent<br />

only one component of the DBM, the ineffectiveness<br />

of this approach does not necessarily<br />

translate into a lack of efficacy of peptide analogs<br />

in the management of neoplasia. A single<br />

component of a multidrug therapy could indeed<br />

be effective on its own [11]. However, as<br />

summarized by Jenkins et al. [12] in this<br />

issue, apart from some notable exceptions,<br />

somatostatin analog therapy has proven to be<br />

disappointing in the management of advanced<br />

malignancy. Should somatostatin analogs be<br />

ab<strong>and</strong>oned? A thorough analysis of the available<br />

literature suggests that this is not the<br />

case. Indeed, besides being used in cancer<br />

treatment (alone or in combination with other<br />

cytotoxic or hormonal agents) <strong>and</strong> palliation,<br />

radiolabelled somatostatin analogs are employed<br />

<strong>for</strong> the localization of primary <strong>and</strong><br />

metastatic tumors expressing somatostatin receptors.<br />

Indeed the so-called ‘somatostatin re-<br />

Preface<br />

ceptor scintigraphy’ is the most important<br />

clinical diagnostic investigation <strong>for</strong> patients<br />

with suspected neuroendocrine tumors. Targeted<br />

radiotherapy, which is being evaluated<br />

in clinical trials, represents an obvious extension<br />

of somatostatin scintigraphy. In addition,<br />

new receptor-selective <strong>and</strong> ‘universal’<br />

analogs are being developed <strong>and</strong> new highdose<br />

regimens are being tested. Further, somatostatin<br />

receptor-targeted chemotherapy<br />

represents an appealing approach to treatment<br />

of SSTR-expressing tumors. Finally, the<br />

genetic transfer of hSSTR-2 <strong>and</strong> hSSTR-5<br />

genes together with the genes that encode<br />

their membrane proteins to those neoplasms<br />

that do not express these receptor subtypes<br />

will translate the benefits of gene therapy to<br />

somatostatin analogue treatment of cancer.<br />

The final chapter on somatostatin <strong>and</strong> cancer<br />

has, there<strong>for</strong>e, not yet been written.<br />

Taking all the above considerations into<br />

account, we felt it timely <strong>and</strong> worthwhile to<br />

attempt a critical review of the recent developments<br />

in the field. To this end, the present<br />

issue of Chemotherapy was planned with the<br />

aim of synthesizing the massive body of evidence<br />

available on somatostatin analog therapy<br />

of cancer <strong>and</strong> the scientific basis <strong>for</strong> their<br />

antineoplastic effects, to enable oncologists to<br />

rationalize the use of these compounds in<br />

their clinical practice <strong>and</strong> to stimulate research<br />

on new therapeutic approaches.<br />

Unlike some other publications in the<br />

field, this supplement is not the result of any<br />

national or international symposium. It represents<br />

the collection of 11 commissioned<br />

monographic reviews generously offered by<br />

30 international scientists, all of whom have<br />

significantly contributed to this new knowledge,<br />

in order to provide a glimpse of what<br />

may lie ahead. I am indebted to all the contributors<br />

<strong>for</strong> having accepted to share with us<br />

their experience of somatostatin analog therapy<br />

of cancer <strong>and</strong> <strong>for</strong> providing us with excel-<br />

IX


lent manuscripts despite their many daily<br />

commitments.<br />

I would like to thank Mr. Peter Roth <strong>and</strong><br />

Mrs. Andrea Brauns of S. Karger AG <strong>for</strong><br />

their excellent cooperation during the publication<br />

of this supplement. Moreover, I am<br />

grateful to Novartis Pharma AG who backed<br />

the publication costs <strong>and</strong> to Voluntary Association<br />

<strong>for</strong> <strong>Cancer</strong> Research in Parma<br />

(A.VO.PRO.RI.T.) <strong>for</strong> financial help <strong>and</strong> <strong>for</strong><br />

spreading this volume to Italian physicians.<br />

My sincere gratitude goes also to Dr. Viktor<br />

Boerlin, Mr. Gary Cheng <strong>and</strong> Dr. Susanne<br />

Schaffert at the Strategic Marketing Depart-<br />

November 2000<br />

References<br />

1 Müllner M: Di Bella’s therapy: The<br />

last word? The evidence would be<br />

stronger if the researchers had r<strong>and</strong>omised<br />

their studies. BMJ 1999;<br />

318:208–209.<br />

2 Reyes JL: Compared to what? eBMJ<br />

1999; 22 January.<br />

3 Liberati A, Magrini N, Patoia L, Pagliaro<br />

L: R<strong>and</strong>omised controlled<br />

trials may not always be absolutely<br />

needed. BMJ 1999;318:1073.<br />

4 Raschetti R, Bruzzi P: Methodological<br />

<strong>and</strong> ethical difficulties in clinical<br />

oncology trials. Di Bella Multitreatment<br />

Italian Trial Coordinating<br />

Group. Lancet 1999;353:153–154.<br />

5 Raschetti R, Greco D, Menniti-Ippolito<br />

F, Spila-Alegiani S, Traversa<br />

G, Benagiano G, Bruzzi P: The Di<br />

Bella multitherapy trial. Criticisms<br />

ignores st<strong>and</strong>ard methodology of<br />

cancer treatments. BMJ 1999;318:<br />

1074.<br />

6 Müllner M, Evans SJW: Reply. BMJ<br />

1999;318:1073.<br />

7 Tirelli U, Di Filippo F: Debate on<br />

Di Bella therapy. Lancet 1999;354:<br />

159.<br />

8 Laderoute MP: Debate on Di Bella<br />

therapy. Lancet 1999;354:159.<br />

9 Italian Study Group <strong>for</strong> the Di Bella<br />

Multitherapy Trials: Evaluation of<br />

an unconventional cancer treatment<br />

(the Di Bella multitherapy): Results<br />

of phase II trials in Italy. BMJ 1999;<br />

318:224–228.<br />

X Preface<br />

ment, Novartis, who rendered this publication<br />

possible. They have shown great interest in the<br />

project from the very beginning <strong>and</strong> made<br />

huge ef<strong>for</strong>ts to make these proceedings available<br />

to the medical community. Last but not<br />

least, I am indebted to Dr. Spencer A. Jenkins<br />

(Departments of General Surgery & Urology,<br />

University Hospital of Wales, Cardiff, UK) <strong>for</strong><br />

his invaluable help during my editorial work.<br />

He made constructive criticism <strong>and</strong> gave useful<br />

suggestions <strong>for</strong> every paper published in<br />

this supplement. We have had many discussions<br />

from which some ideas <strong>and</strong> concepts<br />

expressed in our papers emerged.<br />

Carmelo Scarpignato<br />

MD, DSc (Hons), PharmD (h.c.), FCP, FACG<br />

Professor of Pharmacology <strong>and</strong> Therapeutics<br />

Associate Professor of Gastroenterology<br />

Universities of Parma <strong>and</strong> Nantes<br />

10 Buiatti E, Arniani S, Verdecchia A,<br />

Tomatis L: Results from a historical<br />

survey of the survival of cancer patients<br />

given Di Bella multitherapy.<br />

<strong>Cancer</strong> 1999;86:2143–2149.<br />

11 Calabresi P: Medical alternatives to<br />

alternative medicine. <strong>Cancer</strong> 1999;<br />

86:1887–1889.<br />

12 Jenkins SA, Kynaston HG, Davies<br />

N, Baxter JN, Nott DM: <strong>Somatostatin</strong><br />

analogues in oncology: A look to<br />

the future. Chemotherapy 2001;47<br />

(suppl 2):162–196.


Chemotherapy 2001;47(suppl 2):1–29<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong><br />

<strong>Cancer</strong> <strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong>:<br />

An Overview<br />

Carmelo Scarpignato a, b Iva Pelosini a<br />

a Department of Internal Medicine, School of Medicine <strong>and</strong> Dentistry, University of Parma,<br />

Italy; b Department of Gastroenterology <strong>and</strong> Hepatology, Faculty of Medicine, University of<br />

Nantes, France<br />

Key Words<br />

<strong>Somatostatin</strong> W Octreotide W Lanreotide W<br />

Vapreotide W <strong>Cancer</strong> treatment W<br />

<strong>Somatostatin</strong> receptor scintigraphy W<br />

<strong>Somatostatin</strong> receptor-targeted<br />

radiotherapy<br />

Abstract<br />

Due to the limited efficacy <strong>and</strong> considerable<br />

toxicity of conventional chemotherapy, novel<br />

cytotoxic agents <strong>and</strong> innovative noncytotoxic<br />

approaches to cancer treatment are being<br />

developed. Amongst the various hormonal<br />

agents, increasing attention is being<br />

directed to somatostatin analogs. This is<br />

largely due to the demonstration of antineoplastic<br />

activity of these compounds in a variety<br />

of experimental models in vitro <strong>and</strong> in<br />

vivo <strong>and</strong> to the elucidation of some aspects<br />

of the molecular mechanisms underlying<br />

their antineoplastic activity. On the other<br />

h<strong>and</strong>, clinical experience with somatostatin<br />

analogs in the treatment of conditions like<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0001$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

acromegaly <strong>and</strong> GEP tumors has shown that<br />

they are well tolerated compared to other<br />

antineoplastic therapies currently in use. As<br />

a consequence, there is much ongoing clinical<br />

research to determine whether or not<br />

results from experimental studies will translate<br />

into clinically useful antineoplastic activity.<br />

Besides being used in cancer treatment<br />

<strong>and</strong> palliation, radiolabelled somatostatin<br />

analogs are employed <strong>for</strong> the localization of<br />

primary <strong>and</strong> metastatic tumors expressing<br />

somatostatin receptors. The so-called ‘somatostatin<br />

receptor scintigraphy’ is indeed the<br />

most important clinical diagnostic investigation<br />

<strong>for</strong> patients with suspected neuroendocrine<br />

tumors. Targeted radiotherapy, which<br />

is being evaluated in clinical trials, represents<br />

an obvious extension of somatostatin<br />

scintigraphy. Since the short half-life of native<br />

somatostatin makes continuous intravenous<br />

infusion m<strong>and</strong>atory, several long-acting<br />

analogs have been synthesized. Amongst<br />

the hundreds of peptides synthesized, octreotide<br />

(which binds mainly to SSTR-2 <strong>and</strong><br />

Carmelo Scarpignato, MD, DSc, PharmD, FCP, FACG<br />

Laboratory of Clinical Pharmacology, Department of Internal Medicine<br />

Maggiore University Hospital, I–43100 Parma (Italy)<br />

Tel. +39 0521 903863, Fax +39 0521 292499, E-Mail scarpi@tin.it


SSTR-5 receptor subtypes) has been the<br />

most extensively investigated. A thorough<br />

analysis of the pharmacological activities<br />

<strong>and</strong> therapeutic efficacy of the native somatostatin<br />

<strong>and</strong> the synthetic analogs (octreotide,<br />

lanreotide <strong>and</strong> vapreotide) reveals that<br />

the biological actions of these peptides are<br />

not always identical. These differences appear<br />

to be related to the different affinities of<br />

the natural hormone <strong>and</strong> synthetic derivatives<br />

<strong>for</strong> the different receptor subtypes. For<br />

all the three peptides long-lasting <strong>for</strong>mulations<br />

have been developed to provide patients<br />

with the convenience of once or twice a<br />

month administration <strong>and</strong> to ensure stable<br />

drug serum concentrations between injections.<br />

Radiolabelled derivatives of octreotide,<br />

lanreotide <strong>and</strong> vapreotide have been synthesized<br />

<strong>and</strong> used as radiopharmaceuticals <strong>for</strong><br />

somatostatin receptor scintigraphy <strong>and</strong> somatostatin<br />

receptor-targeted radiotherapy.<br />

The safety profile of synthetic somatostatin<br />

analogs is well established. Most adverse<br />

reactions to these peptides are merely a consequence<br />

of their pharmacological activity<br />

<strong>and</strong> consist mainly of gastrointestinal complaints,<br />

cholelithiasis <strong>and</strong> effects on glucose<br />

metabolism. They are often of little clinical<br />

relevance, thus making somatostatin analogs<br />

safe drugs <strong>for</strong> long-term use. While immediate<br />

release preparations are the drugs<br />

of choice in the short term, long-acting <strong>for</strong>mulations<br />

are better indicated, on an outpatient<br />

basis, <strong>for</strong> the long-term management of<br />

chronic conditions. New ‘receptor-selective’<br />

<strong>and</strong> ‘universal’ somatostatin analogs are being<br />

developed <strong>and</strong> combinations of currently<br />

available derivatives with other (cytotoxic<br />

<strong>and</strong>/or hormonal) agents are being explored<br />

in the search <strong>for</strong> an efficacious <strong>and</strong> well-tolerated<br />

treatment of the various malignancies.<br />

<strong>Somatostatin</strong> receptor-targeted chemotherapy<br />

(with conjugates of somatostatin peptides<br />

with cytotoxic drugs) <strong>and</strong> gene therapy (e.g.<br />

transferring the SSTR-2 gene into neoplastic<br />

cells), which have been successfully tested in<br />

experimental studies, should be applied to<br />

human beings in a not too distant future.<br />

Introduction<br />

2 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini<br />

Copyright © 2001 S. Karger AG, Basel<br />

Although cytotoxic chemotherapy is very<br />

effective in the management of certain neoplasms<br />

such as testicular cancer, the efficacy<br />

of this therapeutic modality in the treatment<br />

of many common neoplasms such as those of<br />

the lung, breast, prostate, bowel, pancreas <strong>and</strong><br />

kidney is limited. Cure of macroscopic metastatic<br />

disease is exceedingly rare, <strong>and</strong> palliation<br />

of symptoms of metastatic neoplasms by<br />

chemotherapy can be problematic since the<br />

toxicity of the treatment often outweighs any<br />

improvement in quality of life resulting from<br />

the temporary decrease in tumor burden [1–<br />

4]. Moreover, postsurgical adjuvant chemotherapy<br />

is frequently without beneficial effect<br />

(as in the case of renal cancer), or is associated<br />

with only small improvements in diseasefree<br />

survival (as in the case of colon cancer,<br />

<strong>for</strong> instance). This situation has not only<br />

motivated attempts to develop novel cytotoxic<br />

agents, but also has stimulated the research<br />

regarding innovative noncytotoxic approaches<br />

to cancer treatment [1–4].<br />

Amongst the various hormonal agents, increasing<br />

attention is being directed to somatostatin<br />

analogs [5–12]. This is largely due<br />

to the demonstration of antineoplastic activity<br />

of these compounds in a variety of experimental<br />

models in vitro <strong>and</strong> in vivo [5, 7] <strong>and</strong><br />

to the elucidation of some aspects of the molecular<br />

mechanisms underlying their antineoplastic<br />

activity [13–16]. On the other h<strong>and</strong>,<br />

clinical experience with somatostatin analogs<br />

in the treatment of conditions like acromegaly<br />

<strong>and</strong> GEP tumors has shown that they are well


tolerated compared to other antineoplastic<br />

therapies currently in use [16, 17]. As a consequence,<br />

there is much ongoing clinical research<br />

to determine whether or not results<br />

from experimental studies will translate into<br />

clinically useful antineoplastic activity.<br />

Besides being used in cancer treatment <strong>and</strong><br />

palliation, radiolabelled somatostatin analogs<br />

are employed <strong>for</strong> the localization of primary<br />

<strong>and</strong> metastatic tumors expressing somatostatin<br />

receptors [18–21]. The so-called ‘somatostatin<br />

receptor scintigraphy’ is indeed<br />

the most important clinical diagnostic investigation<br />

<strong>for</strong> patients with suspected neuroendocrine<br />

tumors [19]. Targeted radiotherapy,<br />

which is being evaluated in clinical trials [22,<br />

23], represents an obvious extension of somatostatin<br />

scintigraphy. This paper will review<br />

the chemistry <strong>and</strong> pharmacokinetics of<br />

currently used synthetic peptides <strong>and</strong> summarize<br />

the rationale <strong>for</strong> their use in cancer treatment<br />

<strong>and</strong> diagnosis.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong>:<br />

Chemistry <strong>and</strong> Pharmacokinetics<br />

Due to its central role in the regulation of<br />

growth hormone secretion, somatostatin is often<br />

referred to as somatotropin release-inhibiting<br />

factor (SRIF) or growth hormone (GH)<br />

release-inhibiting factor [24]. This peptide<br />

displays a wide range of biological actions that<br />

can make it an appropriate drug <strong>for</strong> the treatment<br />

of a variety of human diseases.<br />

Shortly after the isolation of somatostatin,<br />

protein chemists began to synthesize peptide<br />

analogs with a similar spectrum of action but<br />

with much longer biological half-life [25]. The<br />

short half-life of the native peptide [26] makes<br />

indeed continuous intravenous infusion m<strong>and</strong>atory.<br />

To design a more stable peptide derivative<br />

one needs to strengthen the metabolic resis-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

Fig. 1. Sites of enzymatic degradation of natural somatostatin.<br />

tance of the cleavage sequences of the native<br />

peptide. In the case of somatostatin, at least 5<br />

sites of enzymatic degradation are known<br />

(fig. 1) [27]. The most dangerous cleavage can<br />

occur after Trp 8 because such a rupture leads to<br />

completely inactive fragments. Aminopeptidase<br />

attack at the N-terminal is less important<br />

since sequences which are 1 or 2 amino acids<br />

shorter are as potent as the native peptide.<br />

The usual trick to prevent or slow down<br />

enzymatic degradation of a peptide is the<br />

replacement of an L-amino acid by its D-isomer.<br />

Fortunately, the systematic replacement<br />

of L- by D-amino acids demonstrated that this<br />

exchange at position 8 is not only well tolerated<br />

but also leads to an enhanced GH-inhibitory<br />

potency [28].<br />

Further important in<strong>for</strong>mation on the active<br />

site can be deduced by systematic replacement<br />

of all amino acids in SRIF by a<br />

neutral amino acid such as Ala. The total<br />

inactivity of analogs with Ala in positions 6, 7,<br />

8 or 9 indicates that these residues are essential<br />

<strong>for</strong> its biological activity. Similarly, yet<br />

more in<strong>for</strong>mation can be gained by systematically<br />

deleting single amino acids from the natural<br />

sequence. Such studies showed that the<br />

first two amino acids, Ala-Gly, are not necessary<br />

<strong>for</strong> full biological activity [28].<br />

Chemotherapy 2001;47(suppl 2):1–29 3


Fig. 2. Primary structure of octreotide <strong>and</strong> derived peptides.<br />

Table 1. Synthetic analogs of somatostatin currently available<br />

Cyclic octapeptide analogsLinear peptide analogsHexapeptide analogs<br />

SMS 201-995 (octreotide)<br />

RC-160 (vapreotide)<br />

NC-8-12<br />

NC-4-28S<br />

DC 23-60<br />

BIM 23014 (lanreotide)<br />

BIM 23023<br />

BIM 23034<br />

BIM 23059<br />

BIM 23060<br />

D-Trp 8 somatostatin-14.<br />

Lcu 8 , D-Trp 22 , Tyr 25 somatostatin-28.<br />

Against this background, hundreds of somatostatin<br />

analogs have been synthesized in<br />

many research centers all over the world.<br />

Amongst the different peptides (table 1), the<br />

octapeptide SMS 201-995, called octreotide,<br />

has been the most extensively investigated<br />

[<strong>for</strong> reviews, see 29–32]. More recently, two<br />

BIM 23049<br />

BIM 23051<br />

BIM 23052<br />

BIM 23053<br />

BIM 23055<br />

BIM 23057<br />

BIM 23065<br />

BIM 23067<br />

BIM 23068<br />

BIM 23069<br />

4 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini<br />

MK-678<br />

BIM 23050<br />

Linear octapeptide analogs<br />

EC5-21<br />

BIM 23042<br />

BIM 23056<br />

BIM 23058<br />

additional peptides [33], namely lanreotide<br />

(BIM 23014) [34] <strong>and</strong> vapreotide (RC-160)<br />

[35], have become available <strong>for</strong> clinical use. It<br />

is worthwhile to emphasize that all the three<br />

peptides share a common feature, namely<br />

the tetrapeptide X 7 -Trp 8 -Lys 9 -Y 10 (where X<br />

could be either Phe or Tyr <strong>and</strong> Y either Thr or


Fig. 3. Stability of octreotide (SMS<br />

201-995) <strong>and</strong> SIRF against degradation<br />

by rat kidney homogenates<br />

[from 25].<br />

Fig. 4. Plasma levels of unchanged<br />

peptides in the rat: intravenous<br />

application of [ 3 H-Phe 6 ]SRIF (o)<br />

<strong>and</strong> [ 4 C-D-Trp 4 ]octreotide ()).<br />

SRIF was injected at 1.6 mg/kg<br />

(12–16 mCi/mmol), octreotide at<br />

1 mg/kg (35 ÌCi/mg). Unchanged<br />

peptides were recovered after highpressure<br />

liquid chromatography<br />

purification of plasma samples.<br />

Data are taken from Peters [42] <strong>for</strong><br />

SRIF <strong>and</strong> from Lemaire et al. [38]<br />

<strong>for</strong> octreotide. Only the most relevant<br />

alpha phases of elimination<br />

are indicated.<br />

Val, fig. 2), thus suggesting that this amino<br />

acid sequence is essential <strong>for</strong> receptor binding<br />

[36]. And indeed con<strong>for</strong>mational analysis of<br />

the bioactive analogs of somatostatin that<br />

have con<strong>for</strong>mational constraints revealed that<br />

the peptide backbone is not directly involved<br />

in binding, but serves mainly as a scaffold<br />

allowing the side chains to adopt the necessary<br />

pharmacophore spatial arrangement necessary<br />

<strong>for</strong> receptor binding [37].<br />

Compared with native somatostatin, the<br />

synthetic derivatives show a remarkable stability.<br />

Indeed, introduction of a D-amino acid (D-<br />

Phe or D-ßnal) at the N-terminus protects<br />

against exopeptidases, as does the amino-alcohol<br />

Thr(ol) at the C-terminus of octreotide <strong>and</strong><br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

lanreotide. The disulfide bridge itself offers<br />

some protection, <strong>and</strong> the D-Trp protects a position<br />

which would otherwise be cleaved by a<br />

specific endopeptidase. For instance, when incubated<br />

with kidney homogenate, a system<br />

known to degrade natural peptides within a few<br />

minutes, more than 90% of the biological activity<br />

of octreotide was still present after 20 h,<br />

whereas the natural peptide was almost completely<br />

destroyed in less than 1 h (fig. 3) [25].<br />

The pharmacokinetics of octreotide was<br />

investigated in rats after administration of<br />

unlabelled <strong>and</strong> labelled ( 3 H- <strong>and</strong> 14 C-) peptides<br />

[38], <strong>and</strong> data were compared with those<br />

from a study with 3 H-labelled SRIF [39]<br />

(fig. 4). The marked stability of octreotide<br />

Chemotherapy 2001;47(suppl 2):1–29 5<br />

3<br />

4


Fig. 5. Tissue levels of octreotide<br />

after subcutaneous administration<br />

of 1 mg/kg in the rat. Tissue levels<br />

were measured by means of<br />

a specific radioimmunoassay after<br />

extraction into a mixture of<br />

methanol-trifluoroacetic acid (80–<br />

0.1%) <strong>and</strong> subsequent lyophilization<br />

<strong>and</strong> reconstitution in buffer.<br />

For each organ the columns represent<br />

(from left to right) concentrations<br />

measured 0.5, 4, 7 <strong>and</strong> 24 h<br />

after administration [from 39].<br />

against proteolytic degradation together with<br />

a reduced hepatic clearance is responsible <strong>for</strong><br />

the dramatically improved elimination halflife<br />

in rats [39, 40]. In both rats <strong>and</strong> monkeys,<br />

the peptide is well absorbed after subcutaneous<br />

administration, the elimination rate<br />

(based on plasma levels) being slower in the<br />

latter species [41]. Significant levels of octreotide<br />

can also be detected in the rat after oral<br />

administration of the peptide [41].<br />

While extensive degradation into small<br />

fragments <strong>and</strong> amino acids is evident in all<br />

tissues within the first minutes after intravenous<br />

injection of 3 H-SRIF in rats [42], octreotide<br />

proved to be quite stable in all the tissues<br />

examined [38]. Concentrations measured by<br />

quantitative whole body autoradiography,<br />

which determines total radioactivity, <strong>and</strong> by<br />

specific radioimmunoassay were quite comparable.<br />

Thirty minutes after intravenous administration<br />

the highest concentrations were<br />

detected in the kidney, skin <strong>and</strong> liver. By following<br />

the time course of organ distribution<br />

of total radioactivity <strong>and</strong> of unchanged octreotide<br />

(fig. 5) it became apparent that elimination<br />

from presumed target organs such as<br />

pituitary <strong>and</strong> pancreas was much slower when<br />

compared with nontarget tissues such as muscle,<br />

lung <strong>and</strong> heart. This indicates high-affini-<br />

ty binding to target receptors characterized by<br />

slow off-kinetics.<br />

Pharmacokinetic investigations have also<br />

been per<strong>for</strong>med in healthy subjects <strong>and</strong> patients<br />

with pituitary tumors <strong>and</strong> the review by<br />

Chanson et al. [40], to which the reader is<br />

referred, thoroughly summarizes all the available<br />

data. Studies in healthy volunteers [43]<br />

demonstrated that octreotide plasma levels<br />

are proportional to the dose administered<br />

both after intravenous <strong>and</strong> subcutaneous administration.<br />

Plasma peak concentration values,<br />

which were reached after 30 min, were<br />

approximately half those obtained after intravenous<br />

injection of the same dose. Systemic<br />

bioavailability after subcutaneous octreotide<br />

was reported to be almost complete [44].<br />

Plotting the areas under concentration-time<br />

curves (AUCs) against the dose administered,<br />

gives a linear relationship <strong>and</strong> this suggests<br />

that the pharmacokinetic of octreotide is linear<br />

– at least in the dose range studied – irrespective<br />

of the route of administration [43].<br />

The disposition half-life ranged from 80 to<br />

100 min <strong>for</strong> both routes of administration,<br />

depending on the dose, that is more than 30<br />

times the half-life of the natural peptide.<br />

In blood, octreotide is mainly distributed<br />

in the plasma, 65% of the drug being bound to<br />

6 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


lipoprotein <strong>and</strong>, to a lesser extent, albumin,<br />

while negligible amounts are taken up by red<br />

cells [38]. No conclusive data are available<br />

concerning the tissue distribution of octreotide<br />

in humans [45], although it has been<br />

shown that the drug concentrates in many tissues<br />

in the rat [38].<br />

Pharmacokinetic data on the metabolism<br />

<strong>and</strong> elimination of octreotide are similar in<br />

acromegalic <strong>and</strong> healthy individuals [46, 47].<br />

When compared with healthy volunteers, total<br />

body clearance of octreotide was reduced<br />

to 75 ml/min (4.5 liters/h) in patients with<br />

chronic renal failure [48]. Although biliary<br />

excretion <strong>and</strong> proteolysis are also important<br />

elimination pathways in the rat <strong>and</strong> the dog,<br />

they have not been thoroughly studied in humans.<br />

The pharmacokinetics of lanreotide in<br />

healthy volunteers [49, 50] has shown a pattern<br />

similar to that observed with octreotide,<br />

i.e. a Tmax of about 30 min <strong>and</strong> an elimination<br />

half-life of 90 min after single subcutaneous<br />

injection of the peptide. The pharmacokinetics<br />

of lanreotide proved to be linear either<br />

after subcutaneous or intravenous route. The<br />

significant negative correlation between plasma<br />

GH <strong>and</strong> peptide concentrations observed<br />

in the study does suggest a dose-dependent<br />

biological effect [49]. Conversely from octreotide,<br />

lanreotide in blood is mainly bound to<br />

albumin [51].<br />

Given the very short half-life of SRIF, distribution<br />

<strong>and</strong> elimination studies become almost<br />

obsolete, <strong>and</strong> application is restricted to<br />

continuous infusion to maintain therapeutically<br />

relevant plasma concentrations. With<br />

octreotide <strong>and</strong> lanreotide, however, the highly<br />

improved metabolic stability, small volume<br />

of distribution <strong>and</strong> low clearance result in a<br />

long duration of exposure; consequently, longlasting<br />

biological activity after a single subcutaneous<br />

injection of the analogs is obtained.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

Receptor Selectivity of <strong>Somatostatin</strong><br />

<strong>Analogs</strong><br />

The action of somatostatin is mediated<br />

through specific receptors that are functionally<br />

coupled to inhibition of adenylyl cyclase via<br />

pertussis toxin-sensitive GTP binding proteins<br />

[<strong>for</strong> reviews, see 15, 52–54]. Up to five<br />

cell-surface somatostatin receptors have been<br />

characterized. They have been termed SSTR-<br />

1 through SSTR-5 according to the chronology<br />

of their discovery <strong>and</strong> because they all display<br />

the structural hallmark of the seventransmembrane-domain<br />

receptor (SSTR is indeed<br />

the acronym <strong>for</strong> somatostatin seventransmembrane-domain<br />

receptor). Their tissue<br />

distribution is depicted in figure 6. Human<br />

SSTRs (hSSTRs) are encoded by a family<br />

of 5 genes which map to separate chromosomes<br />

<strong>and</strong> which, with one exception, are<br />

intronless. SSTR-2 gives rise to spliced variants,<br />

SSTR-2A <strong>and</strong> 2B. hSSTR-1 to hSSTR-4<br />

display weak selectivity <strong>for</strong> SST-14 binding<br />

whereas hSSTR-5 is SST-28-selective. Based<br />

on structural similarity <strong>and</strong> reactivity <strong>for</strong> octapeptide<br />

<strong>and</strong> hexapeptide SST analogs (table<br />

2), hSSTR-2, 3 <strong>and</strong> 5 belong to a similar<br />

SSTR subclass. hSSTR-1 <strong>and</strong> 4 react poorly<br />

with these analogs <strong>and</strong> belong to a separate<br />

subclass [55, 56].<br />

<strong>Somatostatin</strong> receptors have also been detected<br />

in a number of tumors such as pituitary<br />

adenomas, neuroendocrine <strong>and</strong> nonendocrine<br />

tumors [57–60]. Pituitary <strong>and</strong> islet tumors<br />

express several SSTR genes suggesting<br />

that multiple SSTR subtypes are coexpressed<br />

in the same cell [56]. However, there is variability<br />

in both the number <strong>and</strong> the distribution<br />

of somatostatin receptors between tumors<br />

<strong>and</strong> from site to site in a given tumor<br />

[61, 62]. Since a variable suppression of GH<br />

plasma levels following administration of somatostatin<br />

or octreotide has been demonstrated<br />

in an acromegalic patient, it has been<br />

Chemotherapy 2001;47(suppl 2):1–29 7


Fig. 6. Localization of human somatostatin receptors. GI = Gastrointestinal.<br />

Table 2. Agonist selectivity (Ki, nM) of cloned hSSTR<br />

Compound SSTR-1 SSTR-2 SSTR-3 SSTR-4 SSTR-5<br />

<strong>Somatostatin</strong>-14<br />

1.1 1.3 1.6 0.53 0.9<br />

<strong>Somatostatin</strong>-28 2.2 4.1 6.1 1.1 0.07<br />

Octreotide 11,000 2.1 4.4 11,000 5.6<br />

Lanreotide 11,000 1.8 43 66 0.62<br />

Vapreotide 11,000 5.4 31 45 0.7<br />

Seglitide 11,000 1.5 27 127 2<br />

From Patel <strong>and</strong> Srikant [55].<br />

Table 3. Receptor selectivity of some synthetic somatostatin analogs<br />

Compound SSTR-1 SSTR-2 SSTR-3 SSTR-4 SSTR-5<br />

<strong>Somatostatin</strong>-14<br />

Synthetic analogs<br />

+++ +++ +++ +++ +++<br />

Octreotide 0 ++ + 0 +<br />

Lanreotide 0 ++ 0 0 +++<br />

Vapreotide 0 + 0 0 +++<br />

Derived from data in table 2 by using a cut-off value <strong>for</strong> agonist selectivity (Ki) of 6 nM.<br />

hypothesized that the heterogeneity of both<br />

the number <strong>and</strong> distribution of somatostatin<br />

receptors might in part explain the individual<br />

variable sensitivity to treatment with somatostatin<br />

or its analogs [61, 62].<br />

Although the actions of synthetic analogs<br />

are similar to those of native somatostatin,<br />

some differences have emerged that probably<br />

relate to the different lig<strong>and</strong> affinities <strong>for</strong><br />

SSTR subtypes (table 3). This may have sev-<br />

8 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


eral clinical consequences <strong>and</strong> the spectrum<br />

of the therapeutic efficacy of octreotide <strong>and</strong><br />

its derivatives (i.e. lanreotide <strong>and</strong> vapreotide)<br />

may not be the same as somatostatin. Indeed,<br />

cells expressing SSTR-1 or SSTR-4 will respond<br />

poorly or not at all to somatostatin analogs.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong>: Mechanisms<br />

of the Antineoplastic Action<br />

Besides having an important role in the<br />

symptomatic treatment of endocrine tumors<br />

through peptide suppression, somatostatin<br />

may also exert an antiproliferative effect,<br />

which is not limited to endocrine tumors.<br />

Nonendocrine tumors may also be affected,<br />

although some somatostatin influence may in<br />

part be hormonally mediated. Some authors<br />

have actually reported – after treatment with<br />

somatostatin <strong>and</strong> synthetic analogs – tumor<br />

regression either in patients or animals with<br />

experimentally induced neoplasms.<br />

More recent research has provided in<strong>for</strong>mation<br />

regarding mechanisms underlying the<br />

antiproliferative <strong>and</strong> apoptosis-inducing actions<br />

of somatostatin analogs. These include<br />

both direct mechanisms that are sequelae of<br />

binding of somatostatin analogs to somatostatin<br />

receptors present on neoplastic cells<br />

[57–60] <strong>and</strong> indirect mechanisms related to<br />

effects of somatostatin analogs on the host<br />

[<strong>for</strong> reviews, see 7, 14, 15].<br />

The indirect mechanism would operate<br />

through a suppression of the GH release from<br />

the pituitary <strong>and</strong> the resulting inhibition of<br />

the hepatic production of insulin-like growth<br />

factor-1 (IGF-1) [14, 15]. The fall in IGF-1<br />

could inhibit the growth of various tumors<br />

since IGF-1 <strong>and</strong> IGF-2 as well as other growth<br />

factors, including EGF, appear to be involved<br />

in the proliferation of neoplastic cells. The<br />

potential importance of these mechanisms of<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

action is emphasized by the in vivo antineoplastic<br />

activity of these compounds against<br />

somatostatin receptor-negative neoplasms.<br />

Another potential mechanism through<br />

which somatostatin may exert an antitumor<br />

effect is through its inhibition of tumor angiogenesis,<br />

which is essential <strong>for</strong> implantation<br />

<strong>and</strong> growth [63–65]. Several experimental<br />

pieces of evidence suggest that SSTR-2 preferring<br />

agonists such as octreotide do inhibit<br />

angiogenesis in vitro <strong>and</strong> in vivo [11]. Since<br />

peritumoral vessels express somatostatin receptors<br />

[66] <strong>and</strong> neovascularization is enhanced<br />

by IGF-1 [67], inhibition of angiogenesis<br />

itself might involve direct <strong>and</strong>/or indirect<br />

actions of somatostatin analogs on the nontrans<strong>for</strong>med<br />

cells comprising the microvasculature<br />

of neoplastic tissue.<br />

Studies per<strong>for</strong>med over the past 5 years<br />

have demonstrated that induction of apoptosis<br />

represents one of the mechanisms by<br />

which cytotoxic drugs exert their antineoplastic<br />

action [68]. Several lines of evidence suggest<br />

that somatostatin analogs can also induce<br />

apoptosis via interaction with SSTR-3 [14]. In<br />

this connection it is worth mentioning that<br />

IGF-1 is recognized as a potent antiapoptotic<br />

factor [69]. Thus, the inhibitory effects of<br />

somatostatin analogs on IGF-1 gene expression<br />

may enhance their direct apoptosis-inducing<br />

action <strong>and</strong> contribute to the apoptotic<br />

effect of these compounds. A recent investigation<br />

[70], per<strong>for</strong>med in patients with gut neuroendocrine<br />

tumors, did show that treatment<br />

with high-dose somatostatin analogs induces<br />

apoptosis in tumor cells, which correlated<br />

with the biochemical response (i.e. decrease in<br />

tumor markers), while low-dose somatostatin<br />

analogs do not modify the apoptotic index.<br />

Finally, somatostatin analogs stimulate the<br />

activity of the reticuloendothelial <strong>and</strong> lymphopoietic<br />

systems in the rat [6]. Changes in<br />

natural killer cell activity were reported in<br />

man [6]. It is, there<strong>for</strong>e, possible that modula-<br />

Chemotherapy 2001;47(suppl 2):1–29 9


Table 4. Possible mechanisms of the antineoplastic<br />

action of somatostatin analogs<br />

1 Direct antimitotic effects via somatostatin<br />

receptors on tumor cells<br />

2 Suppression of the release of trophic hormones<br />

(e.g. GH, insulin, prolactin <strong>and</strong> gut peptides)<br />

3 Direct or indirect inhibition of growth factors<br />

(e.g. IGF-1, EGF, PDGF)<br />

4 Inhibition of angiogenesis<br />

5 Induction of apoptosis<br />

6 Modulation of the immune response<br />

tion of immune defense mechanisms might<br />

contribute to the tumor growth inhibitory effects<br />

of these compounds.<br />

In summary, somatostatin <strong>and</strong> its analogs<br />

may have tumoricidal or antiproliferative effects<br />

mediated by suppressing promotor hormones,<br />

by inhibiting mitogens (directly suppressing<br />

cell division, protein synthesis, or<br />

translation), by inhibiting angiogenesis <strong>and</strong><br />

inducing apoptosis or by stimulating the immune<br />

system (table 4). Although the clinical<br />

relevance of some experimental models is at<br />

present unknown, the prospects of such investigations<br />

are worthy of serious consideration.<br />

Octreotide <strong>and</strong> its derivatives may thus<br />

evolve towards an adjunctive, albeit limited<br />

role, in a direct chemotherapeutic management<br />

of endocrine <strong>and</strong> nonendocrine tumors.<br />

The key to this problem appears to be in the<br />

heterogeneity of somatostatin receptor subtype.<br />

There is no doubt that not all of the<br />

receptor subtypes are responsible <strong>for</strong> growth<br />

inhibition [58, 59]. Indeed, there is evidence<br />

that some may even promote cell growth. The<br />

future of somatostatin as a clinically useful<br />

anticancer drug thus lies in the characterization<br />

of the specific receptors that mediate<br />

growth inhibition <strong>and</strong> the synthesis of analogs<br />

that bind selectively to them.<br />

Long-Lasting Formulations of<br />

<strong>Somatostatin</strong> <strong>Analogs</strong><br />

Several studies [71–75] showed that S<strong>and</strong>ostatin<br />

® administered by continuous subcutaneous<br />

pump infusion produced better suppression<br />

of GH <strong>and</strong> IGF-1 serum concentrations,<br />

rapid clinical improvement, <strong>and</strong><br />

shrinkage of GH-secreting adenomas in comparison<br />

to intermittent subcutaneous injections.<br />

Data on the very good efficacy of S<strong>and</strong>ostatin<br />

administered by pump infusion stimulated<br />

research to develop a new galenical <strong>for</strong>mulation<br />

that could ensure long-lasting, sustained<br />

<strong>and</strong> consistent drug delivery. Nasal administration<br />

provides a satisfactory control of<br />

GH hypersecretion, but because of poor local<br />

tolerability, its chronic use is not feasible at<br />

present [76]. An extended-release <strong>for</strong>mulation<br />

mimicking the continuous subcutaneous infusion<br />

of octreotide to be injected monthly<br />

would be an obvious improvement in the<br />

treatment of acromegalic patients requiring<br />

long-term S<strong>and</strong>ostatin therapy by twice daily<br />

or 3 times daily dosing. S<strong>and</strong>ostatin LAR ® ,<br />

obtained by incorporating octreotide into microspheres<br />

of a biodegradable polymer, poly(DL-lactide-co-glycolide<br />

glucose), was developed<br />

to provide patients with the convenience<br />

of a once-a-month administration <strong>and</strong><br />

to ensure a stable serum octreotide concentration<br />

between injections, sustained GH <strong>and</strong><br />

IGF-1 suppression, good clinical control of<br />

symptoms <strong>and</strong> signs of acromegaly, <strong>and</strong> improved<br />

acceptability <strong>and</strong> compliance <strong>for</strong> longterm<br />

treatment with S<strong>and</strong>ostatin [<strong>for</strong> a review,<br />

see 77].<br />

The release characteristics <strong>and</strong> toxicology<br />

of S<strong>and</strong>ostatin LAR were studied in rats <strong>and</strong><br />

rabbits [32, 78]. Single intramuscular injections<br />

of S<strong>and</strong>ostatin LAR resulted in an initial<br />

peak, attributed to drug adsorbed to the surface<br />

of microspheres, followed by low concentrations<br />

over 1–2 weeks <strong>and</strong> thereafter by sus-<br />

10 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


Fig. 7. Time course of serum octreotide<br />

concentrations after administration<br />

of single doses of<br />

10 mg (P, n = 16), 20 mg (d, n =<br />

39) or 30 mg ($, n = 37) of S<strong>and</strong>ostatin<br />

LAR to acromegalic patients.<br />

Each point is the mean of 12-hour<br />

mean concentrations per patient.<br />

Vertical bars are st<strong>and</strong>ard errors<br />

[from 78].<br />

tained octreotide plasma levels over a period<br />

of 4–6 weeks. After repeated injections at 4week<br />

intervals, consistent <strong>and</strong> stable plasma<br />

concentrations of octreotide were recorded.<br />

Toxicological studies per<strong>for</strong>med in rabbits<br />

<strong>and</strong> rats revealed only a very limited, reversible<br />

granulomatous myositis at the injection<br />

site. The biodegradation of the microspheres<br />

is completed within 10–12 weeks, <strong>and</strong> toxicological<br />

studies showed that S<strong>and</strong>ostatin LAR<br />

has low toxicity <strong>and</strong> good local tolerability<br />

[32, 78].<br />

Three different preparations (i.e. 10, 20 or<br />

30 mg) are available <strong>for</strong> clinical use. Their<br />

pharmacokinetics has been studied in acromegalic<br />

patients [78]. A consistent pattern of<br />

octreotide release from the polymer matrix of<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

S<strong>and</strong>ostatin LAR was documented in all studies<br />

<strong>and</strong> <strong>for</strong> all dose levels investigated. A rapid<br />

increase in octreotide serum concentrations<br />

was noted after intramuscular injection of<br />

S<strong>and</strong>ostatin LAR, with a peak occurring within<br />

1 h after the injection followed by a progressive<br />

decrease to low octreotide levels<br />

within 12 h. On days 2 through 7, after single<br />

doses of S<strong>and</strong>ostatin LAR, octreotide serum<br />

concentrations were at lowest levels. Thereafter,<br />

an increase in serum octreotide concentrations<br />

occurred, <strong>and</strong> dose-dependent plateau<br />

concentrations were observed between<br />

days 14 <strong>and</strong> 42 followed by a progressive<br />

decrease from day 42 on (fig. 7). In the plateau<br />

phase (days 14–42), the daily average plasma<br />

concentrations remained very stable over the<br />

Chemotherapy 2001;47(suppl 2):1–29 11


Table 5. Mean pharmacokinetic parameters of octreotide assessed over a period of 60 days<br />

[from 78]<br />

Dose of S<strong>and</strong>ostatin LAR<br />

10 mg (n = 16) 20 mg (n = 39) 30 mg (n = 37)<br />

tmax, days<br />

28B10 28B11 34B17<br />

Cmax, ng/l 387B107 1,126B749 1,935B1,430<br />

Cmax/D, ng/l 39B11 56B38 66B48<br />

AUC0–60 days, ng/l 13,412B3,417 35,737B16,243 61,494B28,245<br />

AUC0–60 days/D, ng/l 1,341B342 1,787B812 2,050B942<br />

Plateau duration, days19.3B10.2 18.5B10.1 18.5B9.8<br />

Relative bioavailability1 , % 31 39 50<br />

tmax = Time to maximum concentration; Cmax = maximum concentration; Cmax/D = maximum<br />

concentration normalized on dose; AUC0–60 days = area under the curve from day 0 to<br />

day 60; AUC0–60 days/D = AUC normalized on dose. Plateau duration is the duration during<br />

which the concentrations were above 80% of Cmax.<br />

1 Relative bioavailability with respect to subcutaneous 3 times daily treatment; values are<br />

the geometric mean.<br />

12-hour observation period, similar to those<br />

seen after subcutaneous continuous infusion.<br />

The height of the octreotide peak on day 1 <strong>for</strong><br />

all doses tested was lower than the plateau<br />

concentrations, <strong>and</strong> the area under the peak<br />

on the day of injection of S<strong>and</strong>ostatin LAR<br />

was not larger than 0.5% of the total AUC<br />

(0–60 days).<br />

A dose-dependent increase of the maximum<br />

concentration <strong>and</strong> AUC of octreotide<br />

was recorded in the dose range between 10<br />

<strong>and</strong> 30 mg. The computed key pharmacokinetic<br />

parameters are summarized in table 5.<br />

In agreement with animal data, human<br />

studies also showed good systemic <strong>and</strong> local<br />

tolerability as well as a lack of dose dumping<br />

(i.e. immediate release of significant quantities<br />

of drug). Preliminary results from studies<br />

per<strong>for</strong>med in acromegalic subjects, responsive<br />

to subcutaneous octreotide, have shown<br />

that one single injection of 30 mg S<strong>and</strong>ostatin<br />

LAR is followed by 4- to 6-week GH suppression<br />

in 80% of patients [78–80]. It is likely,<br />

there<strong>for</strong>e, that this long-lasting <strong>for</strong>mulation<br />

can replace 3 times daily subcutaneous injections<br />

by an intramuscular injection at 4-week<br />

intervals to improve the acceptability of longterm<br />

therapy in acromegalics. In addition, by<br />

releasing consistent concentrations of serum<br />

octreotide <strong>and</strong> by producing a consistent suppression<br />

of GH secretion, S<strong>and</strong>ostatin LAR<br />

appears to be as effective as subcutaneous<br />

infusions of S<strong>and</strong>ostatin <strong>and</strong> more effective<br />

than intermittent subcutaneous administration.<br />

Indeed, in the patients switched from<br />

subcutaneous treatment to S<strong>and</strong>ostatin LAR,<br />

suppression of GH secretion <strong>and</strong> serum<br />

IGF-1 concentrations <strong>and</strong> the clinical improvement<br />

have been either as good as or better<br />

than with S<strong>and</strong>ostatin administered subcutaneously<br />

[78]. Indeed, a larger number of<br />

patients showed a normalization of serum<br />

IGF-1 concentrations <strong>and</strong> a clinical improvement.<br />

Beyond the improvement/disappearance<br />

of symptoms/signs of acromegaly, some<br />

patients actually become asymptomatic. The<br />

12 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


Fig. 8. Pharmacokinetics of lanreotide<br />

<strong>and</strong> GH pattern in 21 patients<br />

at the first intramuscular injection<br />

of the drug (Somatuline SR,<br />

30 mg) [from 89].<br />

usefulness of this octreotide <strong>for</strong>mulation in<br />

the management of malignant carcinoid syndrome<br />

has been recently shown [81].<br />

A slow-release <strong>for</strong>mulation (Somatuline-<br />

SR ® ) is also available <strong>for</strong> the other somatostatin<br />

analog, lanreotide. This <strong>for</strong>mulation<br />

has been studied in healthy volunteers [82]<br />

<strong>and</strong> acromegalic patients [83]. The maximum<br />

lanreotide concentration (Cmax) in plasma<br />

(38.3 B 4.1 ng/ml) was obtained 2 h following<br />

injection. The levels then progressively decreased,<br />

remaining above 1.5 ng/ml until day<br />

11 <strong>and</strong> reaching 0.92 B 0.28 ng/ml 2 weeks<br />

after injection. The apparent plasma half-life<br />

<strong>and</strong> mean residence time were 4.52 B 0.50<br />

<strong>and</strong> 5.48 B 0.51 days, respectively [81].<br />

Studies with Somatuline-SR have been<br />

carried out by several European centers [83–<br />

88] on a few groups of acromegalic patients,<br />

often selected on the basis of their previous<br />

responsiveness to octreotide therapy. An Italian<br />

multicenter study [89] evaluated the tolerability<br />

<strong>and</strong> effectiveness of this <strong>for</strong>mulation in<br />

a large number of acromegalic patients with<br />

active disease, unselected in terms of their<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

previous responsiveness to octreotide, <strong>and</strong><br />

found that 30 mg of the compound, administered<br />

every 14 days, provided an effective<br />

treatment in the majority. After drug administration,<br />

an inverse correlation was found<br />

between lanreotide <strong>and</strong> GH plasma levels<br />

(fig. 8).<br />

Both octreotide <strong>and</strong> lanreotide slow-release<br />

<strong>for</strong>mulations, administered monthly<br />

<strong>and</strong> every 10–14 days, respectively, proved to<br />

be effective in controlling symptoms associated<br />

with neuroendocrine gut tumors, providing<br />

– in addition – a substantial improvement<br />

in patient compliance [90–95].<br />

A slow-release <strong>for</strong>mulation of vapreotide<br />

was developed more than 10 years ago [96]<br />

whereas a long-term delivery system has been<br />

produced only recently [97]. This injectable,<br />

biodegradable depot <strong>for</strong>mulation ensures satisfactory<br />

peptide blood levels in rats <strong>for</strong> over<br />

250 days. No pharmacokinetic data with<br />

these <strong>for</strong>mulations have yet been published in<br />

humans.<br />

Chemotherapy 2001;47(suppl 2):1–29 13


Radiolabelled <strong>Somatostatin</strong> <strong>Analogs</strong><br />

Radiopharmaceuticals <strong>for</strong> <strong>Somatostatin</strong><br />

Receptor Scintigraphy<br />

The diagnosis <strong>and</strong> staging of neuroendocrine<br />

tumors is often difficult <strong>and</strong> time consuming.<br />

Blood levels of hormonal markers are<br />

frequently elevated <strong>and</strong> allow a presumptive<br />

diagnosis [98] but, since tumors are frequently<br />

small, st<strong>and</strong>ard imaging techniques such as<br />

ultrasonography or computed tomography<br />

cannot accurately localize the tumor [99]. Arteriography<br />

<strong>and</strong> selective venous sampling are<br />

more specific, but technically dem<strong>and</strong>ing <strong>and</strong><br />

not always accurate [100]. <strong>Somatostatin</strong> receptor-expressing<br />

tumors <strong>and</strong> their respective<br />

metastases are attractive targets <strong>for</strong> diagnostic<br />

imaging with gamma emitter-labelled synthetic<br />

analogs. Indeed, somatostatin receptor<br />

detection can be accomplished by injecting a<br />

radiolabelled peptide analog <strong>and</strong> imaging tissue<br />

uptake of the compound via scintigraphy.<br />

Selective radioactive uptake will occur in proportion<br />

to the density <strong>and</strong> affinity of the<br />

receptor population. Octreotide binds with<br />

high affinity to the SSTR-2, while this analog<br />

has a relatively low affinity <strong>for</strong> SSTR-3 <strong>and</strong><br />

SSTR-5 <strong>and</strong> shows no binding to SSTR-1 <strong>and</strong><br />

SSTR-4 (see above). Octreotide scintigraphy<br />

(OctreoScan ® ) is, there<strong>for</strong>e, based on the visualization<br />

of (an) octreotide-binding somatostatin<br />

receptor(s), most probably SSTR-<br />

2 <strong>and</strong> SSTR-5 [18–21]. Visualization of<br />

SSTR-positive tumors is widely used in tumor<br />

staging <strong>and</strong> may also predict therapeutic response<br />

to octreotide. A number of studies<br />

[101, 102] have suggested that somatostatin<br />

receptor scintigraphy can be used to select<br />

patients with malignant carcinoid tumors<br />

suitable <strong>for</strong> somatostatin analog treatment<br />

<strong>and</strong> exclude those that will not benefit from<br />

such medication since most hormone-secreting<br />

tumors react in vitro to octreotide with an<br />

inhibition of hormone release <strong>and</strong> possibly<br />

inhibition of growth. It has been shown, <strong>for</strong><br />

instance, that in patients with carcinoids,<br />

there was a complete agreement between the<br />

presence of mRNA <strong>for</strong> SSTR-2 detected by in<br />

situ hybridization <strong>and</strong> therapeutic response to<br />

octreotide [103]. In those patients with pathological<br />

tracer accumulation without expression<br />

of somatostatin SSTR-2 mRNA, other<br />

SSTRs may be present that can bind the<br />

somatostatin analog but not inhibit hormone<br />

secretion. However, octreotide scintigraphy<br />

alone may not be sufficient in determining the<br />

patients with neuroendocrine tumors who can<br />

benefit from chronic treatment with somatostatin<br />

analogs, because almost 20% of patients<br />

with pathological somatostatin scintigraphy<br />

fail to respond to such treatment <strong>and</strong> further,<br />

in rare cases, octreotide treatment results in<br />

clinical improvement in spite of octreotide<br />

scintigraphy failure to demonstrate any tumor<br />

localization [58, 59].<br />

A radioiodinated analog of somatostatin,<br />

[ 123 I-Tyr 3 ]octreotide (fig. 9), was first used to<br />

detect somatostatin receptor-positive tumors<br />

[104]. However, despite the successful visualization<br />

with this radiopharmaceutical of a variety<br />

of somatostatin receptor-positive tumors<br />

in more than 100 patients, this method of in<br />

vivo imaging had several drawbacks, amongst<br />

which are the limited availability of chemically<br />

pure 123 I <strong>and</strong> the high abdominal background<br />

of radioactivity, caused by clearance<br />

of this analog via the liver [104]. There<strong>for</strong>e, an<br />

111 In-labelled somatostatin analog was developed.<br />

[Diethylenetriamine pentaacetic acid<br />

(DTPA)-D-Phe 1 ]-octreotide was shown to<br />

bind 111 In efficiently in a single step procedure.<br />

The binding as well as the biological<br />

activity of this new labelled peptide were<br />

shown to be similar to that of octreotide, making<br />

it a good radiopharmaceutical <strong>for</strong> in vivo<br />

imaging of somatostatin receptor-positive tumors<br />

[105]. The 111 In-labelled octreotide is<br />

excreted mainly via the kidneys, 90% of the<br />

14 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


Fig. 9. Chemical structures of octreotide,<br />

[Tyr 3 ]-octreotide, DTPA<br />

<strong>and</strong> DOTA.<br />

dose being present in the urine 24 h after<br />

injection. Because of its relatively long effective<br />

half-life, [ 111 In-DTPA-D-Phe 1 ]-octreotide<br />

is a radiopharmaceutical which can be<br />

used to visualize somatostatin receptor-bearing<br />

tumors efficiently after 24 <strong>and</strong> 48 h, when<br />

interfering background radioactivity is minimized<br />

by renal clearance [104]. The synthesis<br />

<strong>and</strong> biological properties of 99m Tc-hydrazinonicotinyl-Tyr<br />

3 -octreotide (HYNIC-TOC)<br />

using different colig<strong>and</strong>s <strong>for</strong> radiolabeling was<br />

reported quite recently by Decristo<strong>for</strong>o et al.<br />

[106]. HYNIC-TOC was radiolabelled at high<br />

specific activities using tricine, ethylenediaminediacetic<br />

acid, <strong>and</strong> tricine-nicotinic acid<br />

as colig<strong>and</strong> systems. All 99m Tc-labelled<br />

HYNIC peptides showed retained somatostatin<br />

receptor binding affinities (Kd !2.65<br />

nM). Protein binding <strong>and</strong> internalization<br />

rates were dependent on the colig<strong>and</strong> used.<br />

Specific tumor uptake between 5.8 <strong>and</strong> 9.6%<br />

of the injected dose/g was found <strong>for</strong> the 99m Tclabelled<br />

peptides compared with 4.3% injected<br />

dose/g <strong>for</strong> [ 111 In-DTPA-D-Phe 1 ]-octreotide<br />

[106]. The high specific tumor uptake,<br />

rapid blood clearance, <strong>and</strong> predomi-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

nantly renal excretion make [ 99m Tc-EDDA-<br />

HYNIC-TOC] a promising c<strong>and</strong>idate as an<br />

alternative to [ 111 In-DTPA-D-Phe 1 ]-octreotide<br />

<strong>for</strong> tumor imaging.<br />

The major limitation of somatostatin receptor<br />

scintigraphy using radiolabelled lig<strong>and</strong>s<br />

of octreotide is that the technique will<br />

only allow detection of those tumors expressing<br />

hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong> possibly those<br />

neoplasms expressing hSSTR-3 in sufficient<br />

density to allow visualization. Radiolig<strong>and</strong>s<br />

of lanreotide or vapreotide might be more<br />

useful than radiolabelled lig<strong>and</strong>s of octreotide<br />

in visualizing those tumors that express<br />

hSSTR-4, but not hSSTR-2 <strong>and</strong> hSSTR-5. Visualization<br />

of tumors by lanreotide or vapreotide<br />

scintigraphy but not by octreotide scintigraphy<br />

may provide a rationale <strong>for</strong> the selection<br />

of patients that are likely to benefit from<br />

therapy with these analogs but this hypothesis<br />

requires confirmation in prospective controlled<br />

trials.<br />

Taking the above considerations into account,<br />

radiolabelled derivative of both lanreotide<br />

[ 111 In-DOTA-lanreotide] <strong>and</strong> vapreotide<br />

[ 111 In-DTPA-D-Phe 1 ]-RC-160 have been de-<br />

Chemotherapy 2001;47(suppl 2):1–29 15


veloped [107–109]. However, while labelled<br />

lanreotide showed a high tumor uptake <strong>for</strong> a<br />

variety of different human tumor types <strong>and</strong><br />

a favorable dosimetry over labelled octreotide<br />

[107], with [ 111 In-DTPA-D-Phe 1 ]-RC-<br />

160 blood radioactivity (background) was<br />

higher, resulting in a lower tumor to blood<br />

(background) ratio [108]. This radiopharmaceutical<br />

should, there<strong>for</strong>e, have no advantage<br />

over [ 111 In-DTPA-D-Phe 1 ]-octreotide <strong>for</strong> the<br />

visualization of somatostatin receptors which<br />

bind both analogs. However, recent reports<br />

suggest the existence of different somatostatin<br />

receptor subtypes on some human cancers,<br />

which differentially bind the synthetic somatostatin<br />

analogs [60]. These tumors include<br />

cancers of the breast, ovary, exocrine<br />

pancreas, prostate <strong>and</strong> colon. Radiolabelled<br />

lanreotide or vapreotide might be of interest<br />

<strong>for</strong> future use in such cancer patients as a<br />

radiopharmaceutical <strong>for</strong> imaging somatostatin<br />

receptor-positive tumors, which do not<br />

bind octreotide. Compared with the parent<br />

peptide (i.e. lanreotide), DOTA-lanreotide<br />

seems to display a distinct binding pattern,<br />

since it binds all transfected hSSTR subtypes<br />

as well as a large variety of primary human<br />

tumors [107]. As a consequence, the radiopharmaceutical<br />

is claimed to be a ‘universal’<br />

SSRT lig<strong>and</strong>. A multicenter study (called<br />

Multicentre Analysis of a Universal Receptor<br />

Imaging <strong>and</strong> <strong>Treatment</strong> Initiative: a European<br />

Study) was recently started, <strong>for</strong> which<br />

the acronym MAURITIUS has been coined.<br />

[DOTA]-lanreotide was then renamed MAU-<br />

RITIUS. In a preliminary report 111 In-MAU-<br />

RITIUS was used in a series of 25 patients<br />

with advanced malignancies refractory to<br />

conventional antineoplastic treatment <strong>and</strong> in<br />

all of them at least one tumor site could be<br />

visualized at scintigraphy [110]. Interestingly<br />

enough, some neoplasms, which were repeatedly<br />

negative by the conventional OctreoScan,<br />

could be visualized by means of this<br />

new radiopharmaceutical, thus suggesting<br />

that somatostatin receptors other than<br />

hSSTR-2 <strong>and</strong> hSSTR-5 are responsible <strong>for</strong><br />

binding.<br />

While single photon emission computed<br />

tomography seems to improve accuracy of<br />

somatostatin receptor scintigraphy [111], intraoperative<br />

gamma detection reveals abdominal<br />

endocrine turmors more efficiently<br />

than conventional OctreoScan [112–114] <strong>and</strong><br />

may allow improvement in surgical management<br />

allowing radioimmunoguided surgery<br />

[115]. In vitro <strong>and</strong> in vivo studies [116]<br />

showed that a recently developed terbium-<br />

161-labelled derivative, i.e. [ 161 Tb-DTPA-D-<br />

Phe 1 ]-octreotide, represents a promising pharmaceutical<br />

<strong>for</strong> intraoperative scanning <strong>and</strong><br />

radiotherapy.<br />

Octreotide has also been labelled with positron-emitting<br />

67 Ga [117], 64 Cu [118, 119] or<br />

18 F [120]. Some of these radiolabelled derivatives<br />

([2 - 18 F - fluoropropionyl - D - Phe 1 ]octreotide,<br />

[ 64 Cu-TETA-D-Phe 1 ]-octreotide<br />

<strong>and</strong> [ 67 Ga]-DFO-B-succinyl-D-Phe 1 ]-octreotide)<br />

have, there<strong>for</strong>e, been used <strong>for</strong> PET imaging<br />

[120]. However, the hepatobiliary excretion<br />

of these compounds complicates the interpretation<br />

of the images arising from abdominal<br />

tumors [120]. In contrast, [ 64 Cu-<br />

TETA-D-Phe 1 ]-octreotide binds to somatostatin<br />

receptor with five times the affinity of<br />

[ 111 In-DTPA-D-Phe 1 ]-octreotide, has desirable<br />

clearance properties (renal clearance with<br />

rapid excretion) <strong>and</strong> is a potential agent <strong>for</strong><br />

PET imaging of somatostatin receptors. At<br />

present, however, other labelled compounds<br />

(e.g. 11 C-5-HTP or 11 C-labelled-L-DOPA) are<br />

preferred <strong>for</strong> PET scanning of neuroendocrine<br />

tumors [121].<br />

The detection of heterogenous metastases<br />

(with regard to the expression of different<br />

peptide receptors or the accumulation of other<br />

radiolig<strong>and</strong>s) becomes possible if a combination<br />

of different radiolabelled peptides or<br />

16 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


of a radiolabelled peptide with other radiolig<strong>and</strong>s<br />

(all labelled with different radionuclides)<br />

can be used. Simultaneous use of 111 Inoctreotide<br />

<strong>and</strong> 131 I-MIBG (metaiodobenzylguanidine)<br />

scintigraphy in patients with metastasized<br />

pheochromocytoma [22] represents<br />

a successful example of such an approach.<br />

Radiopharmaceuticals <strong>for</strong> <strong>Somatostatin</strong><br />

Receptor-Targeted Radiotherapy<br />

A new <strong>and</strong> fascinating application of radiolabelled<br />

peptides is represented by their use in<br />

the so-called peptide receptor radiotherapy<br />

[122]. The success of this therapeutic strategy<br />

relies upon the concentration of the radiolig<strong>and</strong><br />

within tumor cells which will depend on<br />

the rates of internalization, degradation <strong>and</strong><br />

recycling of both lig<strong>and</strong> <strong>and</strong> receptor.<br />

Binding of several peptide hormones to<br />

specific surface receptors is generally followed<br />

by internalization of the lig<strong>and</strong>-receptor complex<br />

via invagination of the plasma membrane<br />

[123]. The resulting intracellular vesicles,<br />

termed endosomes, rapidly acidify, thus<br />

causing dissociation of the lig<strong>and</strong> from the<br />

receptor. The lig<strong>and</strong> may be delivered to lysosomes<br />

<strong>and</strong> the receptor recycles back to plasma<br />

membrane. The whole process takes approximately<br />

15 min <strong>and</strong> a single receptor can<br />

deliver numerous lig<strong>and</strong> molecules to the lysosomes<br />

[124].<br />

Receptor-mediated endocytosis of somatostatin<br />

analogs is especially important<br />

when radiotherapy of somatostatin-positive<br />

tumors using radiolabelled analogs is considered.<br />

Human neuroendocrine tumor cells internalize<br />

the radiolig<strong>and</strong> [ 111 In-DTPA-D-<br />

Phe 1 ]-octreotide. However, this radiolig<strong>and</strong><br />

may not be the most suitable compound to<br />

carry out radiotherapy because 111 In, which<br />

emits Auger (as well as conversion) electrons,<br />

is probably not the optimal radionuclide.<br />

Moreover, since a stable coupling of ·- <strong>and</strong> ßemitting<br />

isotopes to [DTPA-D-Phe 1 ]-octreo-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

tide has not been feasible, a novel compound<br />

[tetraazacyclododecane tetraacetic acid<br />

(DOTA), Tyr 3 ]-octreotide (compound coded<br />

as SDZ-SMT 487, fig. 9) in which the DTPA<br />

molecule is replaced by another chelator,<br />

DOTA, allowing a stable bind with the ßemitter<br />

yttrium-90, has been synthesized<br />

[125]. It was recently shown that iodinated<br />

[DOTA, Tyr 3 ]-octreotide is internalized in a<br />

large amount by mouse AtT20 pituitary tumor<br />

cells as well as by human insulinoma cells<br />

[126]. The high internalization rate of this<br />

lig<strong>and</strong> in vitro was also evident from the very<br />

high uptake of this radiolig<strong>and</strong> in vivo by<br />

somatostatin receptor-positive organs in rats<br />

[126]. Along with the high internalization of<br />

the iodinated molecule, de Jong et al. [127]<br />

recently showed that the amount of [ 90 Y-<br />

DOTA, Tyr 3 ]-octreotide internalized by somatostatin<br />

receptor-positive pancreatic tumor<br />

cells was higher than that of [ 111 In-<br />

DOTA, Tyr 3 ]-octreotide <strong>and</strong> of [ 111 In-DTPA-<br />

D-Phe 1 ]-octreotide (1.8- <strong>and</strong> 3.5-fold, respectively).<br />

In vitro, SMT 487 binds selectively<br />

with nanomolar affinity to the somatostatin<br />

receptor subtype 2 (IC30 = 0.39 B 0.02 nM).<br />

In vivo, [ 90 Y-DOTA, Tyr 3 ]-octreotide shows a<br />

rapid blood clearance (t½· !5 min) <strong>and</strong> high<br />

accumulation in somatostatin subtype 2 receptor-expressing<br />

tumors [128]. The in vivo<br />

administration of this radiopharmaceutical<br />

induces a rapid tumor shrinkage in three different<br />

somatostatin receptor-positive tumor<br />

models, namely CA20948 rat pancreatic tumors<br />

grown in normal rats, AR42J rat pancreatic<br />

tumors <strong>and</strong> NCI-H69 human small<br />

cell lung cancer both grown in nude mice. The<br />

radiotherapeutic efficacy of 90 Y-SMT 487 was<br />

enhanced when used in combination with<br />

st<strong>and</strong>ard anticancer drugs, such as mitomycin<br />

C, <strong>and</strong> resulted in a tumor decrease of 70% of<br />

the initial volume. In the CA20948 syngeneic<br />

rat tumor model, a single treatment with 10<br />

ÌCi/kg [ 90 Y-DOTA, Tyr 3 ]-octreotide resulted<br />

Chemotherapy 2001;47(suppl 2):1–29 17


in the disappearance of 5 out of 7 tumors.<br />

Thus the new radiotherapeutic agent showed<br />

its curative potential <strong>for</strong> the selective treatment<br />

of SRIF receptor-expression tumors<br />

[128]. According to these data, [ 90 Y-DOTA,<br />

Tyr 3 ]-octreotide would appear to be a suitable<br />

radiopharmaceutical <strong>for</strong> somatostatin receptor-targeted<br />

radiotherapy.<br />

To achieve an optimal radiotherapeutic effect,<br />

the radiopharmaceutical should also be<br />

retained within tumor cells to allow intracellular<br />

radioactivity exerting its antineoplastic<br />

activity. There<strong>for</strong>e ‘trapping’ of radiolig<strong>and</strong>s<br />

into the tumor cells may be an additional<br />

important mechanism determining the<br />

amount of uptake of the radiopharmaceutical<br />

which is used <strong>for</strong> somatostatin receptor scintigraphy<br />

<strong>and</strong>/or targeted radiotherapy. While<br />

previous investigations [124] have shown that<br />

[ 111 In-DTPA-D-Phe 1 ]-octreotide is delivered<br />

in vivo to lysosomes of pancreatic tumor cells,<br />

the intracellular fate of [ 90 Y-DOTA, Tyr 3 ]octreotide<br />

is presently unknown.<br />

Although being not the ideal radiolig<strong>and</strong>,<br />

[ 111 In-DTPA-D-Phe 1 ]-octreotide has been<br />

used <strong>for</strong> radionuclide therapy in patients with<br />

somatostatin receptor-positive tumors <strong>and</strong><br />

proved the feasibility of the approach [129].<br />

The trial did show a tendency towards better<br />

results in patients whose tumors had a higher<br />

accumulation of the radiolig<strong>and</strong>. In a recent<br />

preliminary study, Otte et al. [130] reported<br />

encouraging results after treatment with [ 90 Y-<br />

DOTA, Tyr 3 ]-octreotide (OctreoTher ® ) in 10<br />

patients with different somatostatin receptorpositive<br />

tumors. In addition, a case report<br />

[131] described a favorable response of a metastatic<br />

gastrinoma to treatment with another<br />

90 Y-labelled somatostatin analog, namely<br />

[ 90 Y-DOTA]-lanreotide [132].<br />

The concept of targeted radiotherapy of<br />

tumors using radiolig<strong>and</strong>s of somatostatin<br />

analogs remains a very attractive approach<br />

<strong>for</strong> the treatment of neoplasia. The very fact<br />

that it is over 12 years since the concept of<br />

octreotide targeted radiotherapy of neoplasia<br />

was first proposed <strong>and</strong> we are still awaiting<br />

good phase 2 clinical trials on the efficacy <strong>and</strong><br />

tolerability of this appealing treatment highlights<br />

the practical difficulties involved in developing<br />

this technique.<br />

Safety <strong>and</strong> Tolerability of<br />

<strong>Somatostatin</strong> <strong>Analogs</strong><br />

The safety profile of S<strong>and</strong>ostatin is well<br />

established [17]. Most adverse reactions to<br />

octreotide are merely a consequence of its<br />

pharmacological activity <strong>and</strong> consist mainly<br />

of gastrointestinal complaints, cholelithiasis<br />

<strong>and</strong> effects on glucose metabolism. The reported<br />

cases of toxicity unrelated to the drug’s<br />

pharmacological profile include reactions at<br />

the injection site, allergic reactions, <strong>and</strong> a few<br />

cases of reversible hepatic dysfunction. Although<br />

the kind of adverse events associated<br />

with S<strong>and</strong>ostatin is well known, their true frequency<br />

has not been accurately estimated.<br />

34.4% of patients reported one or more side<br />

effects, most of which (93.2%) were of little<br />

clinical relevance [17]. For this reason, adverse<br />

events are only seldom mentioned in<br />

published series <strong>and</strong> are rarely reported to the<br />

manufacturer. In contrast, most reports received<br />

by the Novartis Pharmacosurveillance<br />

Unit pertained to events occurring in patients<br />

with severe underlying diseases <strong>and</strong> multiple<br />

drug treatment; there<strong>for</strong>e, a cause-effect relationship<br />

with octreotide could only seldom be<br />

established.<br />

The tolerability of octreotide LAR appears<br />

to be comparable to that of the subcutaneous<br />

<strong>for</strong>mulation [77]. Here again, gastrointestinal<br />

adverse events predominate: abdominal pain,<br />

flatulence, diarrhea, constipation, steatorrhea,<br />

nausea <strong>and</strong> vomiting occurred in up to<br />

50% of patients with acromegaly who re-<br />

18 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


Fig. 10. Adverse events observed<br />

in acromegalic patients (n = 93–<br />

101) after a single intramuscular<br />

injection of S<strong>and</strong>ostatin LAR (10–<br />

30 mg) or multiple injections (30<br />

injections at 4-week intervals) of<br />

the same <strong>for</strong>mulation (20–40 mg)<br />

[from 77].<br />

ceived 1–3 intramuscular doses of octreotide<br />

LAR (10–30 mg). Representative results from<br />

the largest clinical trial are depicted in figure<br />

10. Gastrointestinal symptoms tended to<br />

be mild to moderate <strong>and</strong> often disappeared<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

within 1–4 days of the injection. Furthermore,<br />

the incidence of these events decreased<br />

with long-term (up to 7 months) treatments.<br />

In addition, there was no evidence that tolerability<br />

worsened with increasing dose.<br />

Chemotherapy 2001;47(suppl 2):1–29 19


Injection site events (pain, burning, redness<br />

<strong>and</strong> swelling at injection site) occurred in<br />

some patients receiving intramuscular octreotide<br />

LAR, but were generally mild <strong>and</strong> of<br />

short duration [77]. These phenomena are<br />

thought to be caused by the acidic vehicle of<br />

S<strong>and</strong>ostatin <strong>for</strong>mulations <strong>and</strong> can be minimized<br />

by simple precautions, i.e. to allow refrigerator-cold<br />

vials to reach room temperature<br />

be<strong>for</strong>e administration, <strong>and</strong> to rotate the<br />

sites of injections.<br />

Although the risk of cholelithiasis increases<br />

in patients receiving octreotide [17], simultaneous<br />

bile acid administration strongly reduces<br />

its incidence [77]. Although diabetes mellitus<br />

may occur as a result of reduced glucose tolerance,<br />

the net effect of drug-induced changes is<br />

usually mild <strong>and</strong> not clinically relevant [17,<br />

77]. Finally, few patients developed moderate<br />

to severe hair loss [77]. The spectrum <strong>and</strong> incidence<br />

of adverse events reported after lanreotide,<br />

either immediate <strong>and</strong> slow release <strong>for</strong>mulations,<br />

are similar to those reported after<br />

octreotide [87, 88, 93, 94], the majority of poorly<br />

tolerant patients experiencing untoward<br />

reactions to both compounds [87].<br />

Synthetic somatostatin analogs are, there<strong>for</strong>e,<br />

safe drugs <strong>for</strong> long-term use. While immediate<br />

release preparations are the drugs of<br />

choice in the short term, long-acting <strong>for</strong>mulations<br />

are better indicated, on an outpatient<br />

basis, <strong>for</strong> the long-term management of<br />

chronic conditions.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Diagnosis</strong> <strong>and</strong> <strong>Treatment</strong>: A Look<br />

into the Future<br />

New <strong>Somatostatin</strong> <strong>Analogs</strong> <strong>and</strong> Regimens<br />

The principal challenge in somatostatin research<br />

derives from the fact that the five basic<br />

somatostatin receptor subtypes have high<br />

structural similarities <strong>and</strong> different tissue dis-<br />

tributions. The strong functional similarity<br />

among the five receptor types is exhibited in<br />

their common inhibitory effect on adenylyl<br />

cyclase activity. There<strong>for</strong>e, it is underst<strong>and</strong>able<br />

that somatostatin, which binds with high<br />

affinity to each receptor type, has multiple<br />

physiological actions. To explore the specific<br />

biological function of each subtype, receptorspecific<br />

synthetic analogs, agonists (as well as<br />

antagonists) are being developed. New compounds<br />

in the early phase of development<br />

include receptor-selective <strong>and</strong> ‘universal’ analogs.<br />

The receptor-selective analogs bind to<br />

one, possibly two somatostatin receptor subtypes<br />

[133, 134] while the universal analogs<br />

bind to most or all of the five known SSTRs.<br />

The elucidation of the 3-dimensional structures<br />

of receptor subtype-selective somatostatin<br />

agonists has aided <strong>and</strong> will considerably<br />

enhance the rational design of novel analogs<br />

including non-peptide compounds [135–137].<br />

Development of potent non-peptide somatostatin<br />

analogs is important because they may<br />

display a good bioavailabilty [138] following<br />

oral administration. Continuing research on<br />

somatostatin receptors <strong>and</strong> somatostatin analogs<br />

will help to characterize better the functional<br />

somatostatin receptor models. Recent<br />

availability of the five transfected cell lines has<br />

enabled the use of more rational research<br />

methods. This will help to develop novel drug<br />

c<strong>and</strong>idates beyond the clinically used octreotide-type<br />

analogs. Newly developed analogs<br />

(BIM 23190 <strong>and</strong> BIM 23197) show higher<br />

plasma levels, greater distribution to target tissues<br />

<strong>and</strong> longer in vivo stability [139]. They<br />

may prove to be superior to the currently available<br />

compounds <strong>for</strong> the treatment of acromegaly<br />

<strong>and</strong> some types of cancer.<br />

Future investigations should also be aimed<br />

at further exploring the use of long-acting<br />

somatostatin analogs as antineoplastic agents,<br />

either alone or in combination with other<br />

drugs. In this respect, it will be important to<br />

20 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


etter define the dose-response relationship<br />

<strong>and</strong> to ascertain whether higher doses are<br />

associated with more disease stability or with<br />

greater response rate or survival [16].<br />

<strong>Somatostatin</strong> Receptor-Targeted<br />

Chemotherapy<br />

It is now well established that chemotherapeutic<br />

compounds <strong>and</strong> toxins can be covalently<br />

attached to various carriers, including<br />

hormones, <strong>for</strong> which receptors are present on<br />

cancer cells or to antibodies that preferentially<br />

recognize tumor cells [140]. Such conjugates<br />

are designed to deliver cytotoxic agents<br />

more selectively to cancer cells. Ideally, tumor<br />

cells that bind these conjugates would be<br />

killed while normal cells that do not have the<br />

receptors would be spared [141].<br />

Like targeted radiotherapy, somatostatin<br />

receptor-targeted chemotherapy represents<br />

an appealing approach to treatment of SSTR<br />

expressing tumors. By synthesizing conjugates<br />

of somatostatin analogs <strong>and</strong> cytotoxic<br />

drugs (such as methotrexate or doxirubicin)<br />

[142, 143], selective accumulation of cytotoxic<br />

radicals in somatosatin receptor-positive<br />

tumor cells would be possible. Obviously,<br />

with the currently available somatostatin analogs,<br />

targeted chemotherapy would be limited<br />

to the treatment of SSTR-2- <strong>and</strong> SSTR-5expressing<br />

tumors. Experimental studies<br />

[142, 143] have actually shown that these<br />

derivatives are less toxic <strong>and</strong> more effective<br />

than the parent cytotoxic drugs in inhibiting<br />

tumor growth in vivo. A recent study [144]<br />

demonstrated a high efficacy of SSTR-targeted<br />

chemotherapy in a model of disseminated<br />

human <strong>and</strong>rogen-independent prostatic<br />

carcinoma. The use of cytotoxic somatostatin<br />

analog AN-238 (fig. 11) could provide an<br />

effective therapy <strong>for</strong> patients with advanced<br />

hormone-refractory prostatic carcinoma.<br />

Other studies in progress show that growth of<br />

various human pancreatic, colorectal <strong>and</strong> gas-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

Fig. 11. Molecular structure of the cytotoxic somatostatin<br />

analog AN-238. The somatostatin analog RC-<br />

121 is linked through the ·-aminogroup of it D-Phe<br />

moiety <strong>and</strong> a glutaric acid spacer to the 14-OH group<br />

of 2-pyrrolinodoxorubicin [from 142].<br />

tric cancers in nude mice as well as glioblastomas<br />

<strong>and</strong> non-SCLC can be suppressed by<br />

cytotoxic somatostatin analogs [141]. Thus<br />

these somatostatin analogs might find applications<br />

<strong>for</strong> the therapy of different types of<br />

human malignancies.<br />

Gene Therapy<br />

Gene therapy is at an early phase, but<br />

represents an exciting opportunity to prolong<br />

life in some patients with advanced malignancies<br />

[145–149]. The key problems are getting<br />

the replacement gene to the appropriate cellular<br />

target <strong>and</strong> once there persuading it to<br />

make the normal gene product in sufficient<br />

quantities to correct the defect. With respect<br />

to somatostatin analog therapy there are a<br />

number of areas in which effective gene therapy<br />

may be used to potentiate the antineoplastic<br />

effects of these drugs. The most obvious<br />

application of gene therapy to somatostatin<br />

analog treatment of neoplasia is the delivery<br />

Chemotherapy 2001;47(suppl 2):1–29 21


Fig. 12. Expression of the SSTR-2<br />

in pancreatic tumor cells suppresses<br />

clonigenicity in vitro <strong>and</strong><br />

tumorigenicity in nude mice by a<br />

feedback mechanism. According to<br />

the findings in vitro <strong>and</strong> in vivo,<br />

the expression of this receptor subtype<br />

leads to an increase in somatostatin<br />

lig<strong>and</strong> production <strong>and</strong> hence<br />

a constitutive receptor activation.<br />

Experimental data also suggest that<br />

the SSTR-2 expression is associated<br />

with an increase in the endoproteolytic<br />

processing of prosomatostatin<br />

[from 7].<br />

of hSSTR-2 <strong>and</strong> hSSTR-5 genes together with<br />

the genes that encode their membrane proteins<br />

to those cancers such as pancreatic, gastric<br />

<strong>and</strong> colorectal carcinomas that do not<br />

express these receptor subtypes. The somatostatin<br />

analogs currently available <strong>for</strong> clinical<br />

use (i.e. octreotide, lanreotide <strong>and</strong> vapreotide)<br />

all exert the majority of their antineoplastic<br />

effects via hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong> it follows,<br />

there<strong>for</strong>e, that effective transfer of genes<br />

encoding <strong>for</strong> hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong> their<br />

membrane proteins to cancers which do not<br />

express these receptor subtypes may render<br />

them responsive to the direct antineoplastic<br />

effects of the current generation of somatostatin<br />

analogs.<br />

Human pancreatic adenocarcinomas lose<br />

the ability to express SSTR-2, the somatostatin<br />

receptor, which mediates the antiproliferative<br />

effect of currently available somatostatin<br />

analogs. Reintroducing SSTR-2 into human<br />

pancreatic cancer cells by stable expression<br />

evokes an autocrine negative feedback loop<br />

leading to a constitutive activation of the<br />

SSTR-2 gene <strong>and</strong> an inhibition of cell proliferation<br />

<strong>and</strong> tumorigenicity. In vivo studies<br />

[150], per<strong>for</strong>med in athymic mice, confirmed<br />

the antitumor byst<strong>and</strong>er effects resulting from<br />

the transfer of the SSTR-2 gene into human<br />

pancreatic cancer cell line BxPC-3. Mice were<br />

separately xenografted with control cells on<br />

one flank <strong>and</strong> with SSTR-2-expressing cells<br />

on the other flank. A distant antitumor effect<br />

was induced: growth of control tumors was<br />

delayed by 33 days, the Ki67 index decreased<br />

significantly, <strong>and</strong> apoptosis increased when<br />

compared with control tumors that grew<br />

alone [150]. The distant byst<strong>and</strong>er effect may<br />

be explained in part by a significant increase<br />

in serum somatostatin-like immunoreactivity<br />

levels resulting from the autocrine feedback<br />

loop produced by SSTR-2 expressing cells<br />

(fig. 12) [151] <strong>and</strong> inducing an upregulation of<br />

the type 1 somatostatin receptor, SSTR-1,<br />

which also mediates the antiproliferative effect<br />

of somatostatin [150].<br />

Limitations to peptide receptor radiotherapy<br />

are principally due to poor tumor pene-<br />

22 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


tration of the radiolig<strong>and</strong> <strong>and</strong> insufficient accumulation<br />

of radioactivity within the neoplastic<br />

cell. In addition, low or variable expression<br />

of tumor-associated receptors may<br />

lead to poor tumor localization of radiolabelled<br />

peptide agonists. An attempt to overcome<br />

these problems consists in the use of<br />

biological response modifiers to increase target<br />

receptor expression [152]. In this connection,<br />

replication-deficient adenoviral vectors<br />

were constructed encoding the cDNA <strong>for</strong> the<br />

somatostatin receptor subtype (SSTR-2). In<br />

vitro binding <strong>and</strong> in vivo tumor localization<br />

were observed with radiolabelled octreotide<br />

analogs to cells infected with adenoviral vectors<br />

encoding the corresponding gene [152].<br />

Provided it is successful in humans, this<br />

method could be useful <strong>for</strong> increasing the<br />

therapeutic efficacy of targeted radiotherapy<br />

in cancer patients.<br />

References<br />

1 Devita VT: Principles of cancer<br />

management: Chemotherapy; in<br />

Devita VT, Rosenberg SA, Hellman<br />

S (eds): <strong>Cancer</strong>. Principles <strong>and</strong> Practice<br />

of Oncology, ed 5. Philadelphia,<br />

Lippincott Williams & Wilkins,<br />

1997, pp 333–347.<br />

2 Fisher DS, Tish Knobf M, Durivage<br />

HJ, Tish Knobf M: The <strong>Cancer</strong> Chemotherapy<br />

H<strong>and</strong>book. St Louis,<br />

Mosby-Year Book, 1997, pp 1–530.<br />

3 Baquiran DC, Gallagher J: Lippincott’s<br />

<strong>Cancer</strong> Chemotherapy H<strong>and</strong>book,<br />

ed 5. Philadelphia, Lippincott<br />

Williams & Wilkins, 1998, pp 1–<br />

384.<br />

4 Skeel RT: H<strong>and</strong>book of <strong>Cancer</strong> Chemotherapy,<br />

ed 5. Philadelphia, Lippincott<br />

Williams & Wilkins, 1999,<br />

pp 1–720.<br />

5 Schally AV: Oncological applications<br />

of somatostatin analogs. <strong>Cancer</strong><br />

Res 1988;48:6877–6885.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

Conclusions<br />

6 Lamberts SWJ, Krenning EP, Reubi<br />

JC: The role of somatostatin <strong>and</strong> its<br />

analogs in the diagnosis <strong>and</strong> treatment<br />

of tumors. Endocr Rev 1991;<br />

12:450–482.<br />

7 Weckbecker G, Stolz B, Susini C,<br />

Bruns C: Antiproliferative somatostatin<br />

analogues with potential in<br />

oncology; in Lamberts SWJ (ed):<br />

Octreotide: The Next Decade. Bristol,<br />

Bioscientifica, 1999, pp 339–<br />

352.<br />

8 Höffken K: Peptides in Oncology.<br />

II. <strong>Somatostatin</strong> Analogues <strong>and</strong><br />

Bombesin Antagonists. Berlin,<br />

Springer, 1993, pp 1–136.<br />

9 Reubi JC: Octreotide <strong>and</strong> nonendocrine<br />

tumors: Basic knowledge <strong>and</strong><br />

therapeutic potential; in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine. Basel,<br />

Karger, 1996, pp 256–269.<br />

10 Robbins RJ: <strong>Somatostatin</strong> <strong>and</strong> cancer.<br />

Metab Clin Exp 1996;45(suppl):<br />

98–100.<br />

Despite the explosion of knowledge in recent<br />

years in the somatostatin field, a great<br />

deal remains to be discovered. In particular,<br />

developing the potential of somatostatin analogs<br />

<strong>for</strong> cancer treatment will require a more<br />

complete underst<strong>and</strong>ing of their intracellular<br />

actions <strong>and</strong> interactions. Moreover, in spite of<br />

the ongoing clinical application of octreotide<br />

<strong>and</strong> its analogs in cancer management [153],<br />

the molecular mechanism <strong>and</strong> the beneficial<br />

effects of these drugs need to be elucidated.<br />

Twenty-seven years after its discovery, somatostatin<br />

is still the subject of continuing<br />

investigations by researchers in both the industry<br />

<strong>and</strong> the academia. In the future, the<br />

joint ef<strong>for</strong>t of many scientists from different<br />

fields will no doubt produce more specific<br />

therapeutic agents that are likely to result in<br />

major improvements in clinical management<br />

of various malignancies.<br />

11 Woltering EA, Watson JC, Alperin-<br />

Lea RC, Sharma C, Keenan E, Kurozawa<br />

D, Barrie R: <strong>Somatostatin</strong><br />

analogs: Angiogenesis inhibitors<br />

with novel mechanisms of action.<br />

Invest New Drugs 1997;15:77–86.<br />

12 Kath R, Höffken K: The significance<br />

of somatostatin analogues in<br />

the antiproliferative treatment of<br />

carcinomas; in Höffken K (ed): Peptides<br />

in Oncology III. Berlin, Springer,<br />

2000, pp 23–43.<br />

13 Yamada T, Creutzfeldt W, Beglinger<br />

C, Chiba T: Working Team Report:<br />

The effect of somatostatin on<br />

cellular proliferation. Gastroenterol<br />

Int 1994;7:13–23.<br />

14 Pollak MN, Schally AV: Mechanisms<br />

of antineoplastic action of somatostatin<br />

analogs. Proc Soc Exp<br />

Biol Med 1998;217:143–152.<br />

Chemotherapy 2001;47(suppl 2):1–29 23


15 Bousquet C, Puente E, Buscail L,<br />

Vaysse N, Susini C: Antiproliferative<br />

effect of somatostatin <strong>and</strong> analogs.<br />

Chemotherapy 2001;47(suppl<br />

2):30–39.<br />

16 Öberg K: Established clinical use of<br />

octreotide <strong>and</strong> lanreotide in oncology.<br />

Chemotherapy 2001;47(suppl 2):<br />

40–53.<br />

17 Scarpignato C, Camboni MG: Safety<br />

profile of octreotide; in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine. Basel,<br />

Karger, 1996, pp 296–309.<br />

18 Kwekkeboom DJ, Krenning EP,<br />

Lamberts SWJ: The role of octreotide<br />

scintigraphy in clinical diagnosis<br />

<strong>and</strong> therapy; in Scarpignato C<br />

(ed): Octreotide: From Basic Science<br />

to Clinical Medicine. Basel, Karger,<br />

1996, pp 281–294.<br />

19 Krenning EP, Kwekkeboom DJ,<br />

Pauwels S, Kvols LK, Reubi J-C:<br />

<strong>Somatostatin</strong> receptor scintigraphy.<br />

Nucl Med Ann 1995;1:1–50.<br />

20 O’Byrne KJ, Carney DN: Radiolabelled<br />

somatostatin analogue scintigraphy<br />

in oncology. Anticancer<br />

Drugs 1996;7(suppl 1):33–44.<br />

21 Virgolini I: Vasointestinal peptide<br />

<strong>and</strong> somatostatin receptor scintigraphy<br />

<strong>for</strong> diagnosis <strong>and</strong> treatment of<br />

tumor patients. Eur J Clin Invest<br />

1997;27:793–800.<br />

22 Wiseman GA, Kvols LK: Therapy<br />

of neurometastatic tumors with radiolabelled<br />

MIBG <strong>and</strong> somatostatin<br />

analogues. Semin Nucl Med 1995;<br />

25:272–278.<br />

23 Krenning EP, Valkema R, Kooij<br />

PPM, Breeman WAP, Bakker WH,<br />

deHerder WW, vanEijck CHJ,<br />

Kwekkeboom DJ, deJong M, Pauwels<br />

S: Scintigraphy <strong>and</strong> radionuclide<br />

therapy with [indium-111-labelled-diethyl<br />

triamine penta-acetic<br />

acid-D-Phe 1 ]-octreotide. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl<br />

2):S219–S223.<br />

24 Guillemin R: <strong>Somatostatin</strong>: The<br />

early days. Metabolism 1993;41<br />

(suppl 2):1–4.<br />

25 Pless J, Bauer W, Briner U, Doepner<br />

W, Marbach P, Maurer R, Petcher<br />

TJ, Reubi J-C, Vonderscher J:<br />

Chemistry <strong>and</strong> pharmacology of<br />

SMS 201-995, a long-acting octapeptide<br />

analogue of somatostatin.<br />

Sc<strong>and</strong> J Gastroenterol 1986;21<br />

(suppl 119):54–64.<br />

26 Ho LT, Chen RL, Chou TY, Fong<br />

JC, Wong PS, Chou CK: Pharmacokinetics<br />

<strong>and</strong> effects of intravenous<br />

infusion of somatostatin in normal<br />

subjects – A two-compartment open<br />

model. Clin Physiol Biochem 1986;<br />

4:257–267.<br />

27 Marks N, Stern F: Inactivation of<br />

somatostatin (GH-RIH) <strong>and</strong> its analogs<br />

by crude <strong>and</strong> partially purified<br />

rat brain extracts. FEBS Lett 1975;<br />

55:220–224.<br />

28 Pless J: Chemical structure – Pharmacological<br />

profile of S<strong>and</strong>ostatin ® ;<br />

in O’Dorisio TM (ed): S<strong>and</strong>ostatin<br />

in the <strong>Treatment</strong> of GEP Endocrine<br />

Tumors. Berlin, Springer, 1989, pp<br />

3–13.<br />

29 Vale W, Rivier J, Ling N, Brown M:<br />

Biologic <strong>and</strong> immunologic activities<br />

<strong>and</strong> applications of somatostatin<br />

analogs. Metabolism 1978;27(suppl<br />

1):1391–1401.<br />

30 Battershill PE, Clissold SP: Octreotide.<br />

A review of its pharmacodynamic<br />

<strong>and</strong> pharmacokinetic properties,<br />

<strong>and</strong> therapeutic potential in<br />

conditions associated with excessive<br />

peptide secretion. Drugs 1989;38:<br />

658–702.<br />

31 Camboni MG: Octreotide; in Braga<br />

PC, Gusl<strong>and</strong>i M, Tittobello A (eds):<br />

Drugs in Gastroenterology. New<br />

York, Raven Press, 1991, pp 318–<br />

336.<br />

32 Scarpignato C: Octreotide, the synthetic<br />

long-acting somatostatin analogue:<br />

Pharmacological profile; in<br />

Scarpignato C (ed): Octreotide:<br />

From Basic Science to Clinical Medicine.<br />

Basel, Karger, 1996, pp 54–<br />

72.<br />

33 Eriksson B, Tiensuu Janson E, Bax<br />

NDS, Mignon M, Morant R, Opolon<br />

P, Rougier P, Öberg KE: The use<br />

of new somatostatin analogs, lanreotide<br />

<strong>and</strong> octastatin, in neuroendocrine<br />

gastro-intestinal tumors. Digestion<br />

1996;57(suppl 1):77–80.<br />

34 Lamrani A, Vidon N, Sogni P, Nepveux<br />

P, Catus F, Blumberg J, Chaussade<br />

S: Effects of lanreotide, a somatostatin<br />

analogue, on postpr<strong>and</strong>ial<br />

gastric functions <strong>and</strong> biliopancreatic<br />

secretions in humans. Br J<br />

Clin Pharmacol 1997;43:65–70.<br />

35 Barthomeuf C, Pourrat H, Pourrat<br />

A, Ibrahim H, Cottier PE: Stabilization<br />

of Octastatin, a somatostatin<br />

analogue: Comparative accelerated<br />

stability studies of two <strong>for</strong>mulations<br />

<strong>for</strong> freeze-dried products. Pharm<br />

Acta Helv 1996;71:161–166.<br />

36 Veber DF, Freidlinger RM, Perlow<br />

DS, Paleveda WJ, Holly FW, Strachan<br />

RG, Nutt RF, Arison BH,<br />

Homnick C, R<strong>and</strong>all WC, Glitzer<br />

MS, Saperstein R, Hirschmann R: A<br />

potent cyclic hexapeptide analogue<br />

of somatostatin. Nature 1981;292:<br />

55–58.<br />

37 Huang Z, Probstl A, Spencer JR,<br />

Yamazaki T, Goodman M: Cyclic<br />

hexapeptide analogs of somatostatin<br />

containing bridge modifications.<br />

Syntheses <strong>and</strong> con<strong>for</strong>mational analyses.<br />

Int J Pept Protein Res 1993;42:<br />

352–365.<br />

38 Lemaire M, Azria M, Dannecker R,<br />

Marbach P, Schweitzer A, Maurer<br />

G: Disposition of S<strong>and</strong>ostatin, a<br />

new synthetic somatostatin analogue<br />

in rats. Drug Metab Dispos<br />

1989;17:699–703.<br />

39 Marbach P, Briner U, Lemaire M,<br />

Schweitzer A, Terasaki T: From somatostatin<br />

to S<strong>and</strong>ostatin ® : Pharmacodynamics<br />

<strong>and</strong> pharmacokinetics.<br />

Digestion 1993;54(suppl 1):9–<br />

13.<br />

40 Chanson P, Timsit J, Harris AG:<br />

Clinical pharmacokinetics of octreotide.<br />

Therapeutic applications in patients<br />

with pituitary tumors. Clin<br />

Pharmacokinet 1993;25:375–391.<br />

41 Pless J: Chemical structure – Pharmacological<br />

profile of S<strong>and</strong>ostatin ® ;<br />

in O’Dorisio TM (ed): S<strong>and</strong>ostatin<br />

in the <strong>Treatment</strong> of GEP Endocrine<br />

Tumors. Berlin, Springer, 1989, pp<br />

3–13.<br />

42 Peters GE: Distribution <strong>and</strong> metabolism<br />

of exogenous somatostatin in<br />

rats. Regul Peptides 1982;3:361–<br />

369.<br />

43 Kutz K, Nusch E, Rosenthaler J:<br />

Pharmacokinetics of SMS 201-995<br />

in healthy subjects. Sc<strong>and</strong> J Gastroenterol<br />

1986;21(suppl 119):65–<br />

72.<br />

44 Longnecker SM: <strong>Somatostatin</strong> <strong>and</strong><br />

octreotide: Literature review <strong>and</strong><br />

description of therapeutic activity in<br />

pancreatic neoplasia. Drug Intell<br />

Clin Pharm 1988;22:99–106.<br />

24 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


45 Wynick D, Bloom SR: The use of<br />

long-acting somatostatin analog octreotide<br />

in the treatment of gut neuroendocrine<br />

tumors. J Clin Endocr<br />

Metab 1991;73:1–3.<br />

46 Nicholls J, Wynick D, Domin J,<br />

S<strong>and</strong>ler LM, Bloom SR: Pharmacokinetics<br />

of the long-acting somatostatin<br />

analogue octreotide (SMS<br />

201-995) in acromegaly. Clin Endocrinol<br />

1990;32:545–550.<br />

47 Weeke J, Christensen SE, Orskov H,<br />

Kaal A, Pedersen MM, Illum P,<br />

Harris AG: A r<strong>and</strong>omized comparison<br />

of intranasal <strong>and</strong> injectable octreotide<br />

administration in patients<br />

with acromegaly. J Clin Endocr Metab<br />

1992;75:163–169.<br />

48 Kallivretakis N, Yotis A, Del Pozo<br />

E, Marbach P, Mountokalakis T, et<br />

al: Pharmacokinetics of SMS 201-<br />

995 in normal subjects <strong>and</strong> in patients<br />

with severe renal failure. Neuroendocrinol<br />

Lett 1985;7:92.<br />

49 Kuhn JM, Basin C, Mollard M, de<br />

Rougé B, Baudoin C, Obach R, Tolis<br />

G: Pharmacokinetic study <strong>and</strong> effects<br />

on growth hormone secretion<br />

in healthy volunteers of the new somatostatin<br />

analogue BIM 23014.<br />

Eur J Clin Pharmacol 1993;45:73–<br />

77.<br />

50 Chassard D, Barbanoj M, Català M,<br />

Hawkins F, Moreiro J, et al: Pharmacokinetics<br />

of lanreotide. J Endocrinol<br />

Invest 1997;20(suppl 7):30–<br />

32.<br />

51 Robinson C, Castañer J: Lanreotide<br />

acetate. Drugs Future 1994;19:992–<br />

999.<br />

52 Lewin MJM, Le Romancer M: <strong>Somatostatin</strong><br />

receptors; in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine. Karger,<br />

Basel, 1996, pp 23–34.<br />

53 Patel YC: <strong>Somatostatin</strong> <strong>and</strong> its receptor<br />

family. Front Neuroendocrinol<br />

1999;20:157–198.<br />

54 Schonbrunn A: <strong>Somatostatin</strong> receptors:<br />

Present knowledge <strong>and</strong> future<br />

directions. Ann Oncol 1999;10<br />

(suppl 2):S17–S21.<br />

55 Patel YC, Srikant CB: Subtype selectivity<br />

of peptide analogs <strong>for</strong> all<br />

five cloned human somatostatin receptors<br />

(hSSTR 1–5). Endocrinology<br />

1994;135:2814–2817.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

56 Patel YC, Greenwood MT, Panetta<br />

R, Demchyshyn L, Niznik H, Srikant<br />

CB: The somatostatin receptor<br />

family. Life Sci 1995;57:1249–<br />

1265.<br />

57 Reubi JC: Octreotide <strong>and</strong> nonendocrine<br />

tumors: Basic knowledge <strong>and</strong><br />

therapeutic potential; in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine. Karger,<br />

Basel, 1996, pp 246–269.<br />

58 Hofl<strong>and</strong> LJ, Lamberts SWJ: <strong>Somatostatin</strong><br />

receptors <strong>and</strong> disease:<br />

Role of receptor subtypes. Baillières<br />

Clin Endocrinol Metab 1996;10:<br />

163–176.<br />

59 Lytras A, Tolis G: Clinical significance<br />

of tumor somatostatin receptor<br />

subtype expression. Rev Clin<br />

Pharmacol Pharmacokinet 1997;11:<br />

3–12.<br />

60 Virgolini I, Pangeri T, Bischof C,<br />

Smith-Jones P, Peck-Radosavljevic<br />

M: <strong>Somatostatin</strong> receptor subtype<br />

expression in human tissues: A prediction<br />

<strong>for</strong> diagnosis <strong>and</strong> treatment<br />

of cancer? Eur J Clin Invest 1997;<br />

27:645–647.<br />

61 Reubi JC, L<strong>and</strong>olt AM: High density<br />

of somatostatin receptors in pituitary<br />

tumors from acromegalic patients.<br />

J Clin Endocr Metab 1984;<br />

59:1148–1151.<br />

62 Ikuyama S, Nawata H, Kato KI,<br />

Ibayashi H, Nakagaki H: Plasma<br />

growth hormone responses to somatostatin<br />

(SRIH) <strong>and</strong> SRIH receptors<br />

in pituitary adenomas in acromegalic<br />

patients. J Clin Endocr Metab<br />

1986;62:729–733.<br />

63 Folkman J: Tumour angiogenesis:<br />

Therapeutic implications. N Engl J<br />

Med 1971;285:1182–1186.<br />

64 Folkman J: What is the evidence<br />

that tumors are angiogenesis dependent?<br />

J Natl <strong>Cancer</strong> Inst 1990;82:4–<br />

6.<br />

65 Lichtenbeld HHC, Van Dam Mieras<br />

MCE, Hillen HFP: Tumour angiogenesis:<br />

Pathophysiology <strong>and</strong><br />

clinical significance. Neth J Med<br />

1996;49:42–51.<br />

66 Denzler B, Reubi JC: Expression of<br />

somatostatin receptors in peritumoral<br />

veins of human tumors. <strong>Cancer</strong><br />

1999;85:188–198.<br />

67 Nakao-Hayashi J, Ito H, Kanayasu<br />

T, Morita I, Murota S: Stimulatory<br />

effects of insulin <strong>and</strong> insulin-like<br />

growth factor 1 on migration <strong>and</strong><br />

tube <strong>for</strong>mation by vascular endothelial<br />

cells. Atherosclerosis 1992;92:<br />

141–149.<br />

68 Kaufmann SH, Earnshaw WC: Induction<br />

of apoptosis by cancer chemotherapy.<br />

Exp Cell Res 2000;256:<br />

42–49.<br />

69 Baserga R: The insulin-like growth<br />

factor I receptor: A key to tumor<br />

growth? <strong>Cancer</strong> Res 1995;55:249–<br />

252.<br />

70 Imam H, Eriksson B, Lukinius A,<br />

Janson ET, Lindgren PG, Wil<strong>and</strong>er<br />

E, Öberg K: Induction of apoptosis<br />

in neuroendocrine tumors of the digestive<br />

system during treatment<br />

with somatostatin analogs. Acta Oncol<br />

1977;36:607–614.<br />

71 Christensen ES, Weeke J, Ørskov H,<br />

Moller N, Flyvbjerg A, Harris AG,<br />

Lund E, Jorgensen J: Continuous<br />

subcutaneous pump infusion of<br />

somatostatin analogue SMS 201-<br />

995 versus subcutaneous injection<br />

schedule in acromegalic patients.<br />

Clin Endocrinol 1987;27:297–306.<br />

72 Timsit J, Chanson PH, Larger E,<br />

Duet M, Mosse A, Guillausseau PJ,<br />

Harris AG, Moulonguet M, Warnet<br />

A, Lubetzki J: The effect of subcutaneous<br />

infusion versus subcutaneous<br />

injections of a somatostatin analogue<br />

(SMS 201-995) on the diurnal<br />

GH profile in acromegaly. Acta Endocrinol<br />

1987;116:108–112.<br />

73 Tauber P, Babin T, Tauber MT,<br />

Vigoni F, Bonafe A, Ducasse M,<br />

Harris AG, Bayard F: Long-term effects<br />

of continuous subcutaneous infusion<br />

of the somatostatin analog octreotide<br />

in the treatment of acromegaly.<br />

J Clin Endocrinol Metab<br />

1989;68:917–924.<br />

74 Roelfsema F, Frolich M, de Boer H,<br />

Harris AG: Octreotide treatment in<br />

acromegaly: A comparison between<br />

pen-treated <strong>and</strong> pump-treated patients<br />

in a cross-over study. Acta Endocrinol<br />

1991;125:43–48.<br />

75 James RA, Chatterjee S, White MC,<br />

Hall K, Moller N, Kendall Taylor P:<br />

Comparison of octreotide delivered<br />

by continuous subcutaneous infusion<br />

with intermittent injection in<br />

the treatment of acromegaly. Eur J<br />

Clin Invest 1992;22:554–561.<br />

Chemotherapy 2001;47(suppl 2):1–29 25


76 Invitti C, Fatti LM, Cavagnini F,<br />

Ørskov H, Porcu L, Camboni MG:<br />

Octreotide nasal power in acromegalic<br />

patients: A dose-range <strong>and</strong> tolerability<br />

study. J Endocrinol Invest<br />

1994;17(suppl 2):A41.<br />

77 Gillis JC, Noble S, Goa KL: Octreotide<br />

long-acting release (LAR). A review<br />

of its pharmacological properties<br />

<strong>and</strong> therapeutic use in the management<br />

of acromegaly. Drugs 1997;<br />

63:681–699.<br />

78 Lancranjan I, Bruns C, Grass P, Jaquet<br />

P, Jervell J, Kendall-Taylor P,<br />

Lamberts SWJ, Marbach P, Ørskov<br />

H, Pagani G, Sheppard M, Simionescu<br />

L: S<strong>and</strong>ostatin LAR ® : Pharmacokinetics,<br />

pharmacodynamics,<br />

efficacy, <strong>and</strong> tolerability in acromegalic<br />

patients. Metabolism 1995;44<br />

(suppl 1):18–26.<br />

79 Simionescu L, Boanta C, Dumitrache<br />

C, Mitrea M, Popa O, Bruns<br />

C, Marbach P, Lancrajan I: S<strong>and</strong>ostatin<br />

LAR: Pharmacokinetics, tolerability<br />

<strong>and</strong> efficacy in 24 acromegalic<br />

patients. J Endocrinol Invest<br />

1993;16(suppl 1):148A.<br />

80 Helse J, Kvistborg A, Lancrajan I,<br />

Bruns C, Jervell J: S<strong>and</strong>ostatin LAR<br />

in acromegalic patients: A dose<br />

range <strong>and</strong> tolerability study. J Endocrinol<br />

Invest 1993;16(suppl 1):24A.<br />

81 Rubin J, Ajani J, Schirmer W, Venook<br />

AP, Bukowski R, Pommier R,<br />

Saltz L, D<strong>and</strong>ona P, Anthony L: Octreotide<br />

acetate long-acting <strong>for</strong>mulation<br />

versus open-label subcutaneous<br />

octreotide acetate in malignant carcinoid<br />

syndrome. J Clin Oncol<br />

1999;17:600–606.<br />

82 Kuhn JM, Legr<strong>and</strong> A, Ruiz JM,<br />

Obach R, De Ronzan J, Thomas F:<br />

Pharmacokinetic <strong>and</strong> pharmacodynamic<br />

properties of a long-acting<br />

<strong>for</strong>mulation of the new somatostatin<br />

analogue, lanreotide, in normal<br />

healthy volunteers. Br J Clin Pharmacol<br />

1994;38:213–219.<br />

83 Heron I, Thomas F, Dero M, Gancel<br />

A, Ruiz JM, Schatz B, Kuhn JM:<br />

Pharmacokinetics <strong>and</strong> efficacy of a<br />

long-acting <strong>for</strong>mulation of the new<br />

somatostatin analog BIM 23014 in<br />

patients with acromegaly. J Clin Endocrinol<br />

Metab 1993;76:721–727.<br />

84 Morange I, De Boisvilliers F, Chanson<br />

P, Lucas B, DeWailly D, Catus<br />

F, Thomas F, Jaquet P: Slow-release<br />

lanreotide treatment in acromegalic<br />

patients previously normalized by<br />

octreotide. J Clin Endocrinol Metab<br />

1994;79:145–151.<br />

85 Marek J, Hana V, Krsek M, Justova<br />

V, Catus F, Thomas F: Long-term<br />

treatment of acromegaly with the<br />

slow-release somatostatin analogue<br />

lanreotide. Eur J Endocrinol 1994;<br />

131:20–26.<br />

86 Soule S, Conway G, Hatfield A, Jacobs<br />

H, Giusti M, Gussoni G, Cuttica<br />

CM, Giordano G: Effectiveness<br />

<strong>and</strong> tolerability of slow release lanreotide<br />

treatment in active acromegaly:<br />

Six-month report on an Italian<br />

multicentre study. J Clin Endocrinol<br />

Metab 1996;81:4502–4503.<br />

87 Colao A, Marzullo P, Ferone D,<br />

Marino V, Pivonello R, Di Somma<br />

C, Di Sarno A, Giaccio A, Lombardi<br />

G: Effectiveness <strong>and</strong> tolerability of<br />

slow release lanreotide treatment in<br />

active acromegaly. J Endocrinol Invest<br />

1999;22:40–47.<br />

88 Suliman M, Jenkins R, Ross R,<br />

Powell T, Battersby R, Cullen DR:<br />

Long-term treatment of acromegaly<br />

with the somatostatin analogue<br />

SR-lanreotide. J Endocrinol Invest<br />

1999;22:409–418.<br />

89 Giusti M, Gussoni G, Cuttica CM,<br />

Giordano G, Italian Multicenter<br />

Slow Release Lanreotide Study<br />

Group: Effectiveness <strong>and</strong> tolerability<br />

of slow release lanreotide treatment<br />

in active acromegaly: Sixmonth<br />

report on an Italian multicenter<br />

study. J Clin Endocrinol Metab<br />

1996;81:2089–2097.<br />

90 Scarpignato C, Modlin IM: The<br />

place of octreotide in the medical<br />

management of neuroendocrine gut<br />

tumors; in Scarpignato C (ed): Octreotide:<br />

From Basic Science to<br />

Clinical Medicine. Basel, Karger,<br />

1996, pp 214–232.<br />

91 Ruszniewski P, Ducreux M, Chayvialle<br />

JA, Blumberg J, Cloarec D,<br />

Michel H, Raymond J-M, Dupas<br />

J-L, Gouerou H, Jian R, Genestin E,<br />

Bernades P, Rougier P: <strong>Treatment</strong><br />

of the carcinoid syndrome with the<br />

long-acting somatostatin analogue<br />

lanreotide: A prospective study in<br />

39 patients. Gut 1996;39:279–283.<br />

92 Bajetta E, Carnaghi C, Ferrari L,<br />

Spagnoli I, Mazzaferro V, Buzzoni<br />

R: The role of somatostatin analogues<br />

in the treatment of gastroenteropancreatic<br />

endocrine tumors.<br />

Digestion 1996;57(suppl 1):<br />

72–76.<br />

93 Tomassetti P, Migliori M, Gullo L:<br />

Slow-release lanreotide treatment<br />

in endocrine gastrointestinal tumors.<br />

Am J Gastroenterol 1998;<br />

93:1468–1471.<br />

94 Wymenga ANM, Eriksson B, Salmela<br />

PI, Jacobsen MB, Van Cutsem<br />

EJDG, Fiasse RH, Valimaki<br />

MJ, Renstrup J, De Vries EGE,<br />

Öberg KE: Efficacy <strong>and</strong> safety of<br />

prolonged-release lanreotide in patients<br />

with gastrointestinal neuroendocrine<br />

tumors <strong>and</strong> hormonerelated<br />

symptoms. J Clin Oncol<br />

1999;17:1111–1117.<br />

95 Anthony LB: Long-acting <strong>for</strong>mulations<br />

of somatostatin analogues.<br />

Ital J Gastroenterol Hepatol 1999;<br />

31(suppl 2):S216–S218.<br />

96 Mason-Garcia M, Vaccarella M,<br />

Horvath J, Redding TW, Groot K,<br />

Orsolini P, Schally AV: Radioimmunoassay<br />

<strong>for</strong> octapeptide analogs<br />

of somatostatin: Measurement of<br />

serum levels after administration<br />

of long-acting microcapsule <strong>for</strong>mulations.<br />

Proc Natl Acad Sci<br />

USA 1988;85:5688–5692.<br />

97 Rothen-Weinhold A, Besseghir K,<br />

De Zelicourt Y, Gurny R: Development<br />

<strong>and</strong> evaluation in vivo of a<br />

long-term delivery system <strong>for</strong> vapreotide,<br />

a somatostatin analogue.<br />

J Control Release 1998;52:205–<br />

213.<br />

98 Öberg K, Tiensuu Janson E, Eriksson<br />

B: Tumour markers in neuroendocrine<br />

tumors. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl<br />

2):S160–S162.<br />

99 Vekemans M-C, Urbain J-L,<br />

Charkes D: Advances in radioimaging<br />

of neuroendocrine tumors.<br />

Curr Opin Oncol 1995;7:<br />

63–67.<br />

100 Doppman JL, Jensen RT: Localization<br />

of gastroenteropancreatic<br />

tumors by angiography. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl<br />

2):S163–S166.<br />

26 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


101 Kolby L, Wangberg B, Ahlman H,<br />

Tisell LE, Fjalling M, Forssell-<br />

Aronsson E, Nilsson O: <strong>Somatostatin</strong><br />

receptor subtypes, octreotide<br />

scintigraphy, <strong>and</strong> clinical response<br />

to octreotide treatment in<br />

patients with neuroendocrine tumors.<br />

World J Surg 1998;22:679–<br />

683.<br />

102 Nilsson O, Kolby L, Wangberg B,<br />

Wig<strong>and</strong>er A, Billig H, William-<br />

Olsson L, Fjalling M, Forssell-<br />

Aronsson E, Ahlman H: Comparative<br />

studies on the expression of<br />

somatostatin receptor subtypes,<br />

outcome of octreotide scintigraphy<br />

<strong>and</strong> response to octreotide treatment<br />

in patients with carcinoid tumours.<br />

Br J <strong>Cancer</strong> 1998;77:632–<br />

637.<br />

103 Janson ET, Gobl A, Kalkner KM,<br />

Öberg K: A comparison between<br />

the efficacy of somatostatin receptor<br />

scintigraphy <strong>and</strong> that of in situ<br />

hybridization <strong>for</strong> somatostatin receptor<br />

subtype 2 messenger RNA<br />

to predict therapeutic outcome in<br />

carcinoid patients. <strong>Cancer</strong> Res<br />

1996;56:2561–2565.<br />

104 Krenning EP, Bakker WH, Kooij<br />

PPM, Breman WAP, Oei HY, de<br />

Jong M, Reubi JC, Visser TJ,<br />

Kwekkeboom DJ, Reijs AEM, Van<br />

Hagen PM, Koper JW, Lamberts<br />

SWJ: <strong>Somatostatin</strong> receptor<br />

scintigraphy with [ 111 In-DTPA-D-<br />

Phe 1 ]-octreotide in man: Metabolism,<br />

dosimetry <strong>and</strong> comparison<br />

with [ 123 I-Tyr 3 ]-octreotide. J Nucl<br />

Med 1992;33:652–658.<br />

105 Bakker WH, Alberts R, Bruns C,<br />

Breeman WAP, Hofl<strong>and</strong> LJ, Marbach<br />

P, Pless J, Koper JW, Lamberts<br />

SWJ, Visser TJ, Krenning<br />

EP: [ 111 In-DTPA-D-Phe 1 ]-octreotide,<br />

a potential radiopharmaceutical<br />

<strong>for</strong> imaging of somatostatin receptor-positive<br />

tumors: Synthesis,<br />

radiolabeling <strong>and</strong> in vivo validation.<br />

Life Sci 1991;49:1583–1591.<br />

106 Decristo<strong>for</strong>o C, Melendez-Ala<strong>for</strong>t<br />

L, Sosabowski JK, Mather SJ:<br />

99m Tc - HYNIC - [Tyr 3 ] - octreotide<br />

<strong>for</strong> imaging somatostatin-receptorpositive<br />

tumors: Preclinical evaluation<br />

<strong>and</strong> comparison with 111 Inoctreotide.<br />

J Nucl Med 2000;41:<br />

1114–1119.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

107 Virgolini I, Szilvasi I, Kurtaran A,<br />

Angelberger P, Raderer M, Havlik<br />

E, Vorbeck F, Bischof C, Leimer<br />

M, Dorner G, Kletter K, Niederle<br />

B, Scheithauer W, Smith-Jones<br />

P: Indium-111-DOTA-lanreotide:<br />

Biodistribution, safety <strong>and</strong> radiation<br />

absorbed dose in tumor patients.<br />

J Nucl Med 1998;39:1928–<br />

1936.<br />

108 Breeman WA, Hofl<strong>and</strong> LJ, van der<br />

Pluijm M, von Koetsveld PM, de<br />

Jong M, Setyono-Itan B, Bakker<br />

WH, Kwekkeboom DJ, Visser TJ,<br />

Lamberts SW: A new radiolabelled<br />

somatostatin analogue [ 111 In-<br />

DTPA-D-Phe 1 ]RC-160: Preparation,<br />

biological activity, receptor<br />

scintigraphy in rats <strong>and</strong> comparison<br />

with [ 111 In-DTPA-D-Phe 1 ]octreotide.<br />

Eur J Nucl Med 1994;21:<br />

328–335.<br />

109 Thakur ML, John E, Li J, Reddy<br />

HR, Halmos G, Schally AV: Tc-<br />

99m-RC-160: A somatostatin analog<br />

<strong>for</strong> imaging prostate cancer –<br />

Comparison with I-125-RC-160<br />

<strong>and</strong> In-111-octreotide. J Nucl Med<br />

1995;36:92P.<br />

110 Virgolini I, Kurtaran A, Angelberger<br />

P, Raderer M, Havlik E,<br />

Smith-Jones P: ‘MAURITIUS’:<br />

Tumor dose in patients with advanced<br />

carcinoma. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl 2):<br />

S227–S230.<br />

111 Schillaci O, Corleto VD, Annibale<br />

B, Scopinaro F, Delle Fave G: Single<br />

photon emission computed tomography<br />

procedure improves accuracy<br />

of somatostatin receptor<br />

scintigrapy in gastro-entero pancreatic<br />

tumors. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl 2):S186–<br />

S189.<br />

112 Ohrvall U, Westlin JE, Nilsson S,<br />

Juhlin C, Rastad J, Lundqvist H,<br />

Akerstrom G: Intraoperative gamma<br />

detection reveals abdominal<br />

endocrine tumors more efficiently<br />

than somatostatin receptor scintigraphy.<br />

<strong>Cancer</strong> 1997;80(suppl):<br />

2490–2494.<br />

113 Adams S, Baum RP, Hertel A, Wenisch<br />

HJC, Staib-Sebler E, Herrmann<br />

G, Encke A, Hor G: Intraoperative<br />

gamma probe detection<br />

of neuroendocrine tumors. J Nucl<br />

Med 1998;39:1155–1160.<br />

114 Benevento A, Dominioni L, Carcano<br />

G, Dionigi R: Intraoperative localization<br />

of gut endocrine tumors<br />

with radiolabeled somatostatin<br />

analogs <strong>and</strong> a gamma-detecting<br />

probe. Semin Surg Oncol 1998;15:<br />

239–244.<br />

115 Schneebaum S, Even Sapir E, Cohen<br />

M, Shacham-Lehrman H, Gat<br />

A, Brazovsky E, Livshitz G, Stadler<br />

J, Skornick Y: Clinical applications<br />

of gamma-detection probes –<br />

Radioguided surgery. Eur J Nucl<br />

Med 1999;26(suppl):S26–S35.<br />

116 de Jong M, Breeman WAP, Bernard<br />

BF, Rolleman EJ, Hofl<strong>and</strong><br />

LJ, Visser TJ, Setyono-Han B,<br />

Bakker WH, Van der Puijm ME,<br />

Krenning EP: Evaluation in vitro<br />

<strong>and</strong> in rats of 161 Tb-DTPA-octreotide,<br />

a somatostatin with potential<br />

<strong>for</strong> intraoperative scanning <strong>and</strong> radiotherapy.<br />

Eur J Nucl Med 1995;<br />

22:608–616.<br />

117 Smith-Jones PM, Stolz B, Bruns C,<br />

Albert R, Reist HW, Fridrich R,<br />

Macke HR: Gallium-67/gallium-<br />

68-[DFO]-octreotide – A potential<br />

radiopharmaceutical <strong>for</strong> PET<br />

imaging of somatostatin receptorpositive<br />

tumors: Synthesis <strong>and</strong> radiolabeling<br />

in vitro <strong>and</strong> preliminary<br />

in vivo studies. J Nucl Med<br />

1994;35:317–325.<br />

118 Anderson CJ, Pajeau TS, Edwards<br />

WB, Sherman ELC, Rogers BE,<br />

Welch-MJ: In vitro <strong>and</strong> in vivo<br />

evaluation of copper-64-octreotide<br />

conjugates. J Nucl Med 1995;36:<br />

2315–2325.<br />

119 Anderson CJ, Jones LA, Bass LA,<br />

Sherman ELC, McCarthy DW,<br />

Cutler PD, Lanahan MV, Cristel<br />

ME, Lewis JS, Schwarz SW: Radiotherapy,<br />

toxicity <strong>and</strong> dosimetry<br />

of copper-64-TETA-octreotide in<br />

tumor-bearing rats. J Nucl Med<br />

1998;39:1944–1951.<br />

120 Wester HJ, Brockmann J, Rosch F,<br />

Wutz W, Herzog H, Smith-Jones<br />

P, Stolz B, Bruns C, Stocklin G:<br />

PET-pharmacokinetics of (18)Foctreotide:<br />

A comparison with<br />

(67)Ga-DFO- <strong>and</strong> (86)Y-DTPAoctreotide.<br />

Nucl Med Biol 1997;<br />

24:275–286.<br />

Chemotherapy 2001;47(suppl 2):1–29 27


121 Eriksson B, Örle<strong>for</strong>s A, Sundin A,<br />

Skogseid B, Långström B, Bregström<br />

M, Öberg K: Positron emission<br />

tomography in neuroendocrine<br />

tumors. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl 2):S167–<br />

S171.<br />

122 Thakur ML: Radiolabelled peptides:<br />

Now <strong>and</strong> the future. Nucl<br />

Med Commun 1955;16:724–732.<br />

123 Schwartz AL, Frodovich SE, Lodish<br />

HF: Kinetics of internalization<br />

<strong>and</strong> recycling of the asialoglycoprotein<br />

receptor in a hepatoma<br />

cell line. J Biol Chem 1982;257:<br />

4230–4237.<br />

124 Duncan JR, Stephenson MT,<br />

Wu HP, Anderson CJ: Indium-<br />

111-diethylene-triaminepentaacetic<br />

acid-octreotide is delivered in<br />

vivo to pancreatic, tumor cell, renal,<br />

<strong>and</strong> hepatocyte lysosomes.<br />

<strong>Cancer</strong> Res 1997;57:659–671.<br />

125 de Jong M, Bakker WH, Krenning<br />

EP, Breeman WAP, Van der<br />

Pluijm ME, Bernard BF, Visser<br />

TJ, Jermann E, Behe M, Powell P,<br />

Macke HR: Yttrium-90 <strong>and</strong> indium-111<br />

labelling, receptor binding<br />

<strong>and</strong> biodistribution of<br />

[DOTA 0 -D-Phe1,Tyr 3 ]-octreotide,<br />

a promising somatostatin analogue<br />

<strong>for</strong> radionuclide therapy. Eur J<br />

Nucl Med 1997;24:368–371.<br />

126 Hofl<strong>and</strong> LJ, Breeman WAP, Krenning<br />

EP, de Jong M, Waaijers M,<br />

van Koetsveld PM, Mäcke HR,<br />

Lamberts SWJ: Internalization<br />

of [DOTA 0 , 125 I-Tyr 3 ]-octreotide<br />

by somatostatin receptor-positive<br />

cells in vitro <strong>and</strong> in vivo: Implications<br />

<strong>for</strong> somatostatin-targeted radioguided<br />

surgery. Proc Assoc Am<br />

Physicians 1999;111:63–69.<br />

127 de Jong M, Bernard BF, de Bruin<br />

E, van Gameren A, Bakker WH,<br />

Visser TJ, Mäcke HR, Krenning<br />

EP: Internalization of [DTPA 0 ]octreotide<br />

<strong>and</strong> of [DOTA 0 ,Tyr 3 ]octreotide:<br />

Peptides <strong>for</strong> somatostatin<br />

receptor-targeted scintigraphy <strong>and</strong><br />

radionuclide therapy. Nucl Med<br />

Commun 1998;19:283–288.<br />

128 Stolz B, Smith-Jones P, Albert R,<br />

Weckbeker G, Bruns C: New somatostatin<br />

analogues <strong>for</strong> radiotherapy<br />

of somatostatin receptor<br />

expressing tumors. Ital J Gastroenterol<br />

Hepatol 1999;31(suppl 2):<br />

S224–S226.<br />

129 Krenning EP, Valkema R, Kooij<br />

PPM, Breeman WAP, Bakker<br />

WH, deHerder WW, vanEijck<br />

CHJ, Kwekkeboom DJ, deJong M,<br />

Pauwels S: Scintigraphy <strong>and</strong> radionuclide<br />

therapy with [indium-111labelled-diethyl<br />

triamine pentaacetic<br />

acid-D-Phe 1 ]-octreotide. Ital<br />

J Gastroenterol Hepatol 1999;31<br />

(suppl 2):S219–S223.<br />

130 Otte A, Müller-Br<strong>and</strong> J, Dellas S,<br />

Nitzsche EU, Hermann R, Mäcke<br />

HR: Yttrium-90-labelled somatostatin<br />

analogue <strong>for</strong> cancer treatment.<br />

Lancet 1998;351:417–418.<br />

131 Leimer M, Kurtaran A, Smith-<br />

Jones P, Raderer M, Havlik E, Angelberger<br />

P, Vorbeck F, Niederle<br />

B, Herold C, Virgolini I: Response<br />

to treatment with yttrium-90-<br />

DOTA-lanreotide of a patient with<br />

metastatic gastrinoma. J Nucl Med<br />

1998;39:2090–2094.<br />

132 Virgolini I, Szilvasi I, Kurtaran A,<br />

Angelberger P, Raderer M, Havlik<br />

E, Vorbeck F, Bischof C, Leimer<br />

M, Dorner G, Kletter K, Niederle<br />

B, Scheithauer W, Smith-Jones<br />

P: Indium-111-DOTA-lanreotide:<br />

Biodistribution, safety <strong>and</strong> radiation<br />

absorbed dose in tumor patients.<br />

J Nucl Med 1998;39:1928–<br />

1936.<br />

133 Reubi JC, Schaer JC, Waser B,<br />

Hoeger C, Rivier J: A selective<br />

analog <strong>for</strong> the somatostatin sst1receptor<br />

subtype expressed by human<br />

tumors. Eur J Pharmacol<br />

1998;345:103–110.<br />

134 Rohrer SP, Birzin ET, Mosley RT,<br />

Berk SC, Hutchins SM, Shen DM,<br />

Xiong Y, Hayes EC, Parmar RM,<br />

Foor F, Mitra SW, Degrado SJ,<br />

Shu M, Klopp JM, Cai SJ, Blake A,<br />

Chan WW, Pasternak A, Yang L,<br />

Patchett AA, Smith RG, Chapman<br />

KT, Schaeffer JM: Rapid identification<br />

of subtype-selective agonists<br />

of the somatostatin receptor<br />

through combinatorial chemistry.<br />

Science 1998;282:737–740.<br />

135 Hirschmann R, Nicolaou KC, Pietranico<br />

S, Leaby EM, Salvino J,<br />

Arison B, Clichy MA, Spoors PG,<br />

Shakespeare WC, Sprengeler PA,<br />

Hamley P, Smith AB III, Reisine<br />

T, Raynor K, Maechler L, Donaldson<br />

C, Vale W, Freidinger RM,<br />

Cascieri MR, Strader CD: Nonpeptidal<br />

peptidomimetics with a ß-<br />

D-glucose scaffolding. A partial so-<br />

matostatin agonist bearing a close<br />

structural relationship to a potent,<br />

selective substance P agonist. J Am<br />

Chem Soc 1992;114:9217–9218.<br />

136 Hirschmann R, Nicolaou K, Pietranico<br />

S, Salvino J, Leahy EM,<br />

Sprengeler PA, Furst G, Smith AB<br />

III: De novo design <strong>and</strong> synthesis<br />

of somatostatin non-peptide peptidomimetics<br />

utilizing ß-D-glucose<br />

as a novel scaffolding. J Am Chem<br />

Soc 1993;115:12550–12568.<br />

137 Damour D, Barreau M, Blanchard<br />

J, Burgevin M-C, Doble A, Herman<br />

F, Pantel G, James-Surcouf E,<br />

Vuilhorgne M, Mignani S, Poitout<br />

L, Le Merrer Y, Depezay J-C: Design,<br />

synthesis <strong>and</strong> binding affinities<br />

of novel non-peptide mimics<br />

of somatostatin/s<strong>and</strong>ostatin. Bioorg<br />

Med Chem Lett 1996;6:1667–<br />

1672.<br />

138 Pasternak A, Pan Y, Marino D,<br />

S<strong>and</strong>erson PE, Mosley R, Rohrer<br />

SP, Birzin ET, Huskey SEW, Jacks<br />

T, Schleim KD, Cheng K, Schaeffer<br />

JM, Patchett AA, Yang L: Potent,<br />

orally bioavailable somatostatin<br />

agonists: Good absorption<br />

achieved by urea backbone cyclization.<br />

Bioorg Med Chem Lett<br />

1999;9:491–496.<br />

139 Gillespie TJ, Erenberg A, Kim S,<br />

Dong J, Taylor JE, Hau V, Davis<br />

TP: Novel somatostatin analogs<br />

<strong>for</strong> the treatment of acromegaly<br />

<strong>and</strong> cancer exhibit improved in<br />

vivo stability <strong>and</strong> distribution. J<br />

Pharmacol Exp Ther 1998;285:<br />

95–104.<br />

140 Magrath T: Targeted approaches<br />

to cancer therapy. Int J <strong>Cancer</strong><br />

1994;56:163–166.<br />

141 Schally AV, Nagy A: <strong>Cancer</strong> chemotherapy<br />

based on targeting of<br />

cytotoxic peptide conjugates to<br />

their receptors on tumors. Eur J<br />

Endocrinol 1999;141:1–14.<br />

142 Nagy A, Schally AV, Halmos G,<br />

Armatis P, Cai R-Z, Csernus V,<br />

Kovács M, Koppán M, Szepesházi<br />

K, Kahán Z: Synthesis <strong>and</strong> biological<br />

evaluation of cytotoxic analogs<br />

of somatostatin containing doxorubicin<br />

or its intensely potent derivative,<br />

2-pyrrolinodoxorubicin.<br />

Proc Natl Assoc Sci USA 1998;95:<br />

1794–1799.<br />

28 Chemotherapy 2001;47(suppl 2):1–29 Scarpignato/Pelosini


143 Radulovic S, Nagy A, Szoke B,<br />

Schally AV: Cytotoxic analog of<br />

somatostatin containing methotrexate<br />

inhibits growth of MIA<br />

PaCa-2 human pancreatic cancer<br />

xenografts in nude mice. <strong>Cancer</strong><br />

Lett 1992;62:263–271.<br />

144 Plonowski A, Schally AV, Nagy A,<br />

Sun B, Szepeshazi K: Inhibition of<br />

PC-3 human <strong>and</strong>rogen-independent<br />

prostate cancer <strong>and</strong> its metastases<br />

by cytotoxic somatostatin<br />

analog AN-238. <strong>Cancer</strong> Res 1999;<br />

59:1947–1953.<br />

145 Blum HE: Molecular biology <strong>and</strong><br />

gene therapy in gastroenterology<br />

<strong>and</strong> hepatology. Eur J Gastroenterol<br />

Hepatol 1999;11:1–7.<br />

146 Roth JA, Cristiano J: Gene therapy<br />

<strong>for</strong> cancer: What have we done<br />

<strong>and</strong> where are we going? J Natl<br />

Invest 1999;89:21–36.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> <strong>for</strong> <strong>Cancer</strong><br />

<strong>Treatment</strong> <strong>and</strong> <strong>Diagnosis</strong><br />

147 Wake N, Kondoh H, Katoh H: Recent<br />

advances <strong>and</strong> perspective insights<br />

of gene therapy. Acta Obstet<br />

Gynaecol Jpn 1999;51:715–724.<br />

148 Caplen NJ: Gene therapy: Different<br />

strategies <strong>for</strong> different applications.<br />

Mol Med Today 1998;4:<br />

374–375.<br />

149 Farzaneh F, Trefzer U, Sterry W,<br />

Walden P: Gene therapy of cancer.<br />

Immunol Today 1998;19:<br />

294–296.<br />

150 Rochaix P, Delesque N, Esteve<br />

J-P, Saint Laurent N, Voigt JJ,<br />

Vaysse N, Susini C, Buscail L:<br />

Gene therapy <strong>for</strong> pancreatic carcinoma:<br />

Local <strong>and</strong> distant antitumor<br />

effects after somatostatin receptor<br />

sst2 gene transfer. Hum Gene Ther<br />

1999;10:995–1008.<br />

151 Delesque N, Buscail L, Esteve JP,<br />

Saint Laurent N, Muller C, Weckbecker<br />

G, Bruns C, Vaysse N, Susini<br />

C: sst2 somatostatin receptor<br />

expression reverses tumorigenicity<br />

of human pancreatic cancer cells.<br />

<strong>Cancer</strong> Res 1997;57:956–962.<br />

152 Rogers BE, Garver RI, Grizzle<br />

WE, Buchsbaum DJ: Genetic induction<br />

of antigens <strong>and</strong> receptors<br />

as targets <strong>for</strong> cancer radiotherapy.<br />

Tumor Targeting 1998;3:122–<br />

137.<br />

153 Scarpignato C: <strong>Somatostatin</strong> analogs<br />

in cancer management. Chemotherapy<br />

2001;47(suppl 2):1–<br />

198.<br />

Chemotherapy 2001;47(suppl 2):1–29 29


Chemotherapy 2001;47(suppl 2):30–39<br />

Antiproliferative Effect of<br />

<strong>Somatostatin</strong> <strong>and</strong> <strong>Analogs</strong><br />

Corinne Bousquet Elena Puente Louis Buscail Nicole Vaysse<br />

Christiane Susini<br />

INSERM U 151, CHU Rangueil, IFR 31, Toulouse, France<br />

Key Words<br />

<strong>Somatostatin</strong> receptor W Proliferation W<br />

Receptor signaling<br />

Abstract<br />

Over the past decade, antiproliferative effects<br />

of somatostatin <strong>and</strong> analogs have been<br />

reported in many somatostatin receptor-positive<br />

normal <strong>and</strong> tumor cell types. Regarding<br />

the molecular mechanisms involved, somatostatin<br />

or analogs mediate their action<br />

through both indirect <strong>and</strong> direct effects. <strong>Somatostatin</strong><br />

acts through five somatostatin receptors<br />

(SSTR1–5) which are variably expressed<br />

in normal <strong>and</strong> tumor cells. These<br />

receptors regulate a variety of signal transduction<br />

pathways including inhibition of adenylate<br />

cyclase, regulation of ion channels,<br />

regulation of serine/threonine <strong>and</strong> tyrosine<br />

kinases <strong>and</strong> phosphatases. This review focuses<br />

on recent advances in biological mechanisms<br />

involved in the antineoplastic activity<br />

of somatostatin <strong>and</strong> analogs.<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

Copyright © 2001 S. Karger AG, Basel<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0030$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Introduction<br />

<strong>Somatostatin</strong> (somatotroph release-inhibiting<br />

factor) was originally discovered as a<br />

hypothalamic neurohormone that inhibited<br />

growth hormone secretion. It was subsequently<br />

demonstrated that somatostatin is a widely<br />

distributed peptide both in the central <strong>and</strong><br />

peripheral nervous systems <strong>and</strong> in peripheral<br />

tissues including the endocrine pancreas, gut,<br />

adrenals, kidney <strong>and</strong> immune cells. In mammals,<br />

two <strong>for</strong>ms of bioactive peptides, somatostatin<br />

14 <strong>and</strong> a C-terminally extended <strong>for</strong>m,<br />

somatostatin 28 are produced by tissue-specific<br />

proteolytic processing of a common precursor.<br />

<strong>Somatostatin</strong> acts on various targets<br />

including the brain, pituitary, pancreas, gut,<br />

adrenals, thyroid, kidney <strong>and</strong> on the immune<br />

system to regulate a variety of physiological<br />

functions. Its actions include inhibition of endocrine<br />

<strong>and</strong> exocrine secretions, modulation<br />

of neurotransmission, motor <strong>and</strong> cognitive<br />

functions, inhibition of intestinal motility, absorption<br />

of nutrients <strong>and</strong> ions, vascular contractility<br />

<strong>and</strong> cell proliferation.<br />

Due to the short half-life of natural somatostatin<br />

peptides (F1 min), many somato-<br />

Dr. Christiane Susini<br />

INSERM U151, Institut Louis Bugnard, IFR 31<br />

CHU Rangueil, F–31403 Toulouse Cedex 4 (France)<br />

Tel. +33 5 61 32 24 07, Fax +33 5 61 32 24 03<br />

E-Mail susinich@rangueil.inserm.fr


statin peptide analogs have been synthesized.<br />

Among them, octreotide (SMS 201-995), lanreotide<br />

(BIM 23014) <strong>and</strong> vapreotide have<br />

been intensively investigated <strong>and</strong> are in clinical<br />

use <strong>for</strong> the medical treatment of acromegaly<br />

<strong>and</strong> neuroendocrine tumors. All of these<br />

cyclic octapeptides retain amino acid residues<br />

(or substitutes) within a cyclic peptide backbone<br />

involved in the biological effect of the<br />

peptide (Phe 7 or Tyr 7 , DTrp 8 , Lys 9 <strong>and</strong> Thr 10<br />

or Val 10 ) <strong>and</strong> display markedly increased stability<br />

(half-life 190 min). The biological effects<br />

of somatostatin are mediated via high<br />

affinity plasma membrane receptors which<br />

are coupled to various signal transduction<br />

pathways including inhibition of adenylate<br />

cyclase <strong>and</strong> Ca 2+ channels <strong>and</strong> stimulation of<br />

several K + channels <strong>and</strong> protein phosphatases.<br />

<strong>Somatostatin</strong> receptors are widely distributed<br />

throughout many tissues ranging<br />

from the central nervous system to the pancreas<br />

<strong>and</strong> gut, <strong>and</strong> also in pituitary, kidney,<br />

thyroid, lung <strong>and</strong> immune cells [<strong>for</strong> a review,<br />

see 1, 2].<br />

Compelling evidence has implicated somatostatin<br />

in the inhibition of the growth <strong>and</strong><br />

development of various normal <strong>and</strong> tumor<br />

cells. Thus, somatostatin analogs show antineoplastic<br />

activity in a variety of experimental<br />

models in vivo <strong>and</strong> in vitro [<strong>for</strong> a review,<br />

see 3, 4]. <strong>Somatostatin</strong> receptors are expressed<br />

in a large variety of human tumors.<br />

Octreotide treatment induces the shrinkage of<br />

the pituitary [5] <strong>and</strong> the stabilization of neuroendocrine<br />

tumor progression [6]. However,<br />

the clinical implications of the presence of<br />

somatostatin receptors in tumors <strong>for</strong> diagnostic<br />

<strong>and</strong> therapeutic purposes will need to define<br />

the physiological role of each receptor<br />

subtype expressed in the tumors with respect<br />

to its antisecretory <strong>and</strong> antiproliferative properties.<br />

<strong>Somatostatin</strong> or its analogs may influence<br />

tumor cell growth via indirect <strong>and</strong><br />

direct mechanisms <strong>and</strong> this review focuses on<br />

Antiproliferative Effect of <strong>Somatostatin</strong><br />

<strong>and</strong> <strong>Analogs</strong><br />

recent advances in the molecular mechanisms<br />

involved in the antiproliferative effect of somatostatin<br />

<strong>and</strong> analogs.<br />

<strong>Somatostatin</strong> Receptor Family<br />

To date, five receptor subtypes, SSTR-1–<br />

SSTR-5, have been cloned. Human SSTRs are<br />

encoded by 5 genes localized on separate<br />

chromosomes. Four of these genes are intronless,<br />

the exception being SSTR-2 which is<br />

alternatively spliced in rodents to generate<br />

two iso<strong>for</strong>ms named SSTR-2A <strong>and</strong> SSTR-2B<br />

which diverge in their C-terminal sequence.<br />

The SSTRs subtypes belong to the family of<br />

G-protein-coupled receptors with seven transmembrane<br />

spanning domains <strong>and</strong> present a<br />

high degree of sequence identity (39–57%).<br />

They all bind somatostatin 14 <strong>and</strong> somatostatin<br />

28 natural peptides with similar affinity<br />

(nM range) although with a slightly higher<br />

affinity <strong>for</strong> somatostatin 14. Only SSTR-5 displays<br />

a 10-fold higher affinity <strong>for</strong> somatostatin<br />

28. However, they show major differences<br />

in their affinities <strong>for</strong> analogs. <strong>Analogs</strong><br />

exhibit a low affinity <strong>for</strong> SSTR-1 <strong>and</strong> SSTR-4<br />

(61 ÌM) whereas they bind SSTR-2, SSTR-5<br />

with a high affinity, comparable to that of<br />

somatostatin 14 (nM range) <strong>and</strong> bind SSTR-3<br />

with moderate affinity (65–10 nM) [7, 8].<br />

Using recombinant SSTRs transiently or<br />

stably expressed in various eukaryote cells,<br />

the intracellular signaling pathways coupled<br />

to SSTRs have been extensively studied. Each<br />

receptor subtype is coupled to multiple intracellular<br />

transduction pathways via pertussis<br />

toxin-sensitive heterotrimeric GTP binding<br />

proteins. Inhibition of the adenylate cyclase<br />

system leading to reduction of intracellular<br />

cAMP was one of the first signaling pathways<br />

identified to be associated with the occupation<br />

of somatostatin receptors. All five SSTRs<br />

are functionnally coupled to inhibition of ade-<br />

Chemotherapy 2001;47(suppl 2):30–39 31


nylate cyclase via a pertussis toxin-sensitive<br />

protein, G·i1 or G·i2, or G·i3 being involved in<br />

mediating this coupling [7, 8].<br />

A second key plasma membrane signaling<br />

pathway involved in somatostatin action<br />

concerns K + <strong>and</strong> Ca 2+ channels. In neuronal<br />

<strong>and</strong> neuroendocrine cells, somatostatin <strong>and</strong><br />

analogs regulate several subsets of K + channels<br />

including inward rectifying (GIRK1; via<br />

G protein G·i3), transient outward, delayed<br />

rectifying, voltage-dependent M current <strong>and</strong><br />

ATP-sensitive K + channels as well as large<br />

conductance Ca 2+ -activated BK channels.<br />

Activation of K + channels by somatostatin<br />

causes hyperpolarization of the plasma<br />

membrane <strong>and</strong> leads to decreased Ca 2+ influx<br />

through voltage-gated Ca 2+ channels <strong>and</strong><br />

consequently to reduction of intracellular<br />

Ca 2+ . Expression of GIRK1 channels together<br />

with each somatostatin receptor in Xenopus<br />

oocytes demonstrate that SSTR-2,<br />

SSTR-3, SSTR-4 <strong>and</strong> SSTR-5 can couple to<br />

inward rectifying K + channels, SSTR-2 being<br />

the most efficiently coupled [9]. The SSTR<br />

subtypes coupled to other K 2+ channels remain<br />

to be elucidated. <strong>Somatostatin</strong> can also<br />

decrease Ca 2+ influx by directly inhibiting<br />

high voltage-dependent Ca 2+ channels via<br />

G0·2. <strong>Somatostatin</strong> receptors couple to Ntype<br />

<strong>and</strong> L-type voltage-dependent calcium<br />

channels in several cell types including<br />

mouse pituitary AtT-20 cells <strong>and</strong> rat insulinoma<br />

RINm5F cells [7, 8]. Using receptor<br />

subtype-specific analogs, it has been demonstrated<br />

that SSTR-2 <strong>and</strong> SSTR-5 can negatively<br />

couple to voltage-dependent calcium<br />

channel (L-type) in mouse pituitary cells<br />

[10]. SSTR-1 is also implicated in the inhibition<br />

of Ca 2+ influx since it has been shown to<br />

mediate the inhibition of voltage-dependent<br />

Ca 2+ channels in rat insulinoma 1046-38<br />

cells [11]. Both of these two signal transduction<br />

pathways, inhibition of Ca 2+ <strong>and</strong> to a<br />

lesser extent inhibition of cAMP, mediate<br />

the negative regulation of hormone <strong>and</strong> neurotransmitter<br />

secretions induced by somatostatin.<br />

A third important membrane signaling<br />

a pathway linked to somatostatin receptors<br />

involves the regulation of protein phosphatases.<br />

<strong>Somatostatin</strong> <strong>and</strong> analogs activate a<br />

number of protein phosphatases including<br />

serine/threonine phosphatases, tyrosine phosphatases<br />

<strong>and</strong> Ca 2+ -dependent phosphatase<br />

[12–15]. SSTR-1, SSTR-2, SSTR-3 <strong>and</strong><br />

SSTR-4 have been reported to stimulate tyrosine<br />

phosphatase activity when expressed in<br />

NIH 3T3 fibroblast or CHO cells [16–18].<br />

<strong>Somatostatin</strong> receptor subtypes involved in<br />

the regulation of serine/threonine phosphatase<br />

have not yet been identified.<br />

Concerning the coupling of SSTRs with<br />

phospholipase C pathway, the results are controversial.<br />

In COS-7 cells expressing each receptor<br />

subtype, all five receptors are able to<br />

stimulate phospholipase C <strong>and</strong> increase Ca 2+<br />

mobilization via a pertussis toxin-dependent<br />

G protein, albeit at agonist concentration higher<br />

than 1 nM. However, the coupling of SSTR-<br />

2, SSTR-5 <strong>and</strong> SSTR-3 to phospholipase C is<br />

more efficient. For SSTR-3, this coupling is<br />

involved in the stimulatory effect of somatostatin<br />

on guinea pig intestinal smooth muscle<br />

cell contraction <strong>and</strong> is mediated by phospholipase<br />

C-ß3 [19]. In contrast, in transfected<br />

CHO-K1 cells, SSTR-5 inhibits phospholipase<br />

C-mediated intracellular Ca 2+ mobilization<br />

whereas SSTR-4 is without effect [20].<br />

The MAP kinase pathway is also involved<br />

in signal transduction coupled to SSTR but<br />

the modulation differs according to the receptor<br />

subtype. When expressed in CHOK1 cells,<br />

SSTR-4 activates MAP kinases Erk1/2 leading<br />

to phosphorylation <strong>and</strong> activation of<br />

phospholipase A2 <strong>and</strong> release of arachidonic<br />

acid whereas SSTR-5 inhibits MAP kinases<br />

ERK1/2 by a mechanism involving a dephosphorylation<br />

cascade including inhibition of<br />

guanylate cyclase <strong>and</strong> thus decreasing cGMP<br />

32 Chemotherapy 2001;47(suppl 2):30–39 Bousquet/Puente/Buscail/Vaysse/Susini


<strong>for</strong>mation <strong>and</strong> inhibition of cGMP-dependent<br />

protein kinase G [21].<br />

Finally, another signal transduction pathway<br />

coupled to SSTR-1 has also been reported.<br />

Indeed, in transfected mouse fibroblast<br />

Ltk– cells, SSTR-1 has been reported to<br />

be negatively coupled to Na + /H + exchanger by<br />

a mechanism independent of G protein. In<br />

the same cells, transfected SSTR-2 has no<br />

effect [22].<br />

Indirect Effects of <strong>Somatostatin</strong> on<br />

Cell Growth<br />

Indirect effects of somatostatin on tumor<br />

growth may be the result of inhibition of<br />

secretion of growth-promoting hormones <strong>and</strong><br />

growth factors which stimulate the growth of<br />

various types of cancer. It is known that<br />

tumors depend on specific growth factors <strong>for</strong><br />

their growth. For example, insulin-like growth<br />

factor-1 (IGF-1) which is produced by hepatocytes<br />

through GH-dependent <strong>and</strong> GH-independent<br />

mechanisms is an important modulator<br />

of many neoplasms that express IGF-1<br />

receptors <strong>and</strong> respond to this mitogenic factor<br />

[23]. Octreotide has been demonstrated to<br />

negatively control the serum IGF-1 level as a<br />

result of an effect on GH secretion <strong>and</strong> SSTR-<br />

2 <strong>and</strong> SSTR-5 have been demonstrated to be<br />

implicated in this effect [24]. In addition, a<br />

direct effect on IGF gene expression has also<br />

been reported [25]. The suppressive action of<br />

somatostatin analogs on the GH-IGF-1 axis<br />

has been proven effective in the treatment of<br />

GH-secreting pituitary adenomas [26]. Clinical<br />

studies have demonstrated a reduction of<br />

IGF-1 gene expression <strong>and</strong> serum levels of<br />

IGF-1 after treatment of breast cancers with<br />

octreotide <strong>and</strong> these effects are increased<br />

when combined with the antiestrogen tamoxifen<br />

[27]. <strong>Somatostatin</strong> <strong>and</strong> analogs also increase<br />

the expression <strong>and</strong> the secretion of<br />

Antiproliferative Effect of <strong>Somatostatin</strong><br />

<strong>and</strong> <strong>Analogs</strong><br />

IGF-binding protein-1, which specifically<br />

binds IGF-1 <strong>and</strong> negatively regulates plasma<br />

IGF-1 [28]. <strong>Somatostatin</strong> <strong>and</strong> analogs may<br />

also inhibit the secretion of gastrointestinal<br />

<strong>and</strong> pancreatic hormones involved in the regulation<br />

of tumor growth such as cholecystokinin,<br />

gastrin, insulin <strong>and</strong> glucagon <strong>and</strong> growth<br />

factors such as the epidermal growth factortrans<strong>for</strong>ming<br />

growth factor · family. In somatostatin<br />

SSTR-2 knockout mice, the use of<br />

new selective subtype analogs has enabled<br />

investigators to demonstrate the role of<br />

SSTR-2 in inhibiting glucagon <strong>and</strong> gastrin<br />

release from mouse pancreatic · <strong>and</strong> gastric<br />

cells, respectively, <strong>and</strong> the role of SSTR-5 as a<br />

mediator of insulin secretion from mouse<br />

pancreatic ß cells has recently been reported<br />

[24, 29]. In addition, the detection of SSTR-2<br />

receptors with specific anti-SSTR-2 antibodies<br />

in human A <strong>and</strong> B pancreatic islet cells<br />

suggests that this receptor subtype is involved<br />

in the regulation of human pancreatic hormones<br />

[30]. The role of other receptor subtypes<br />

in the regulation of hormone secretion<br />

remains to be defined.<br />

<strong>Somatostatin</strong> or its analogs can also indirectly<br />

control tumor development <strong>and</strong> metastasis<br />

by inhibition of angiogenesis. <strong>Somatostatin</strong><br />

analogs inhibit angiogenesis in vitro <strong>and</strong><br />

in vivo [31]. Overexpression of peritumoral<br />

vascular somatostatin receptors with a highaffinity<br />

<strong>for</strong> somatostatin 14, somatostatin 28<br />

<strong>and</strong> octreotide (suggesting a preferential expression<br />

of SSTR-2 subtype) has been reported<br />

in human primary colorectal carcinomas,<br />

small cell lung carcinoma of the lung,<br />

breast cancer, renal cell carcinoma <strong>and</strong> malignant<br />

lymphoma. Furthermore, the expression<br />

of somatostatin receptors in tumor vessels appears<br />

to be independent of receptor expression<br />

in the tumor [32]. This suggests that somatostatin<br />

<strong>and</strong> its receptors may play a regulatory<br />

role in hemodynamic tumor-host interactions,<br />

angiogenesis <strong>and</strong> vascular drainage.<br />

Chemotherapy 2001;47(suppl 2):30–39 33


Evidence suggests that somatostatin may<br />

influence the immune system. <strong>Somatostatin</strong><br />

<strong>and</strong> receptors are expressed in human lymphoid<br />

organs <strong>and</strong> can regulate various immune<br />

functions including lymphocyte proliferation,<br />

immunoglobulin synthesis, <strong>and</strong> cytokine<br />

production [33]. It has recently been<br />

reported that SSTR-2, SSTR-3, SSTR-4 <strong>and</strong><br />

SSTR-5 are expressed in human lymphoid<br />

cell lines <strong>and</strong> expression of SSTR-2 is upregulated<br />

following lymphocyte activation [34]. It<br />

has been well demonstrated that somatostatin<br />

<strong>and</strong> octreotide inhibit the proliferation of human<br />

<strong>and</strong> rat T lymphocytes in vitro [33, 35]<br />

<strong>and</strong> that somatostatin inhibits IFN-Á release<br />

from T lymphocytes probably via the SSTR-2<br />

subtype leading to a decrease of Ig2a levels<br />

[36]. However, no in<strong>for</strong>mation is available on<br />

the effect of somatostatin in vivo on the immune<br />

system.<br />

Direct Effect of <strong>Somatostatin</strong> on<br />

Tumor Cells<br />

<strong>Somatostatin</strong> may also directly inhibit tumor<br />

cell growth by interacting with specific<br />

somatostatin receptors located on tumor cells.<br />

Using either radiolabeled somatostatin or its<br />

analogs, somatostatin receptors have been described<br />

in a large variety of human cancer<br />

cells (including pituitary adenomas, islet tumors,<br />

carcinoids, adenocarcinomas of breast,<br />

prostate, ovary, kidney <strong>and</strong> colon origin, lymphomas<br />

as well as astrocytomas <strong>and</strong> neuroblastomas<br />

<strong>and</strong> medulloblastomas). Most tumors<br />

express the SRIF1 subtype characterized<br />

by a high affinity <strong>for</strong> natural peptides <strong>and</strong><br />

hexapeptide-stable analogs but a number of<br />

tumors express the SRIF2 subtype characterized<br />

by a low affinity <strong>for</strong> analogs. More recent<br />

analyses of SSTR mRNAs demonstrate that<br />

various human tumors from neuroendocrine<br />

<strong>and</strong> gastroenteropancreatic origin, brain tu-<br />

mors (gliomas <strong>and</strong> meningiomas), prostate,<br />

lung <strong>and</strong> breast tumors express various SSTR<br />

mRNA, each tumor expressing more than one<br />

subtype, SSTR-2 being the most frequently<br />

expressed [<strong>for</strong> a review, see 37, 38]. Indeed, in<br />

carcinoid tumors, of 87 tumors examined, approximately<br />

85% are positive <strong>for</strong> SSTR-2.<br />

The majority of these tumors also express<br />

SSTR-5, with SSTR-1, SSTR-3 <strong>and</strong> SSTR-4<br />

being less abundant. The high frequency of<br />

SSTR-2 mRNA (<strong>and</strong> probably also the presence<br />

of SSTR-5 mRNA) found in neuroendocrine<br />

tumors allows the localization of various<br />

human tumors <strong>and</strong> metastases by somatostatin<br />

receptor scintigraphy following injection<br />

of indium-111-labeled octreotide. However,<br />

the detection of specific SSTR mRNA in<br />

tumors does not necessarily reflect the presence<br />

of the protein. The recent availability of<br />

polyclonal antibodies has enabled different<br />

groups to identify the SSTR proteins. Indeed,<br />

using SSTR-2 antibodies, Dutour et al. [39]<br />

report the expression of SSTR-2 in human<br />

gliomas <strong>and</strong> meningiomas with a rich expression<br />

in both human brain tumor <strong>and</strong> peritumoral<br />

tissue, <strong>and</strong> a prominent expression<br />

in blood vessels. Immunohistochemical detection<br />

of somatostatin receptors SSTR-1,<br />

SSTR-2 <strong>and</strong> SSTR-3 using specific antibodies<br />

in 33 breast tumors allows the detection of the<br />

receptors on tumor cells but the level <strong>and</strong> the<br />

pattern of the expression of SSTR vary greatly<br />

between individual carcinomas: SSTR-2 detected<br />

at a high level in 28 tumors (85%),<br />

SSTR-1 in 17 tumors (52%) <strong>and</strong> SSTR-3 in 16<br />

tumors (48%) [40]. Of 35 patients with carcinoid<br />

tumors, the presence of SSTR-2 protein<br />

has been detected in 25 patients <strong>and</strong> there was<br />

a correlation with the presence of SSTR-2<br />

mRNA, the tracer uptake using somatostatin<br />

receptor autoradiography <strong>and</strong> the therapeutic<br />

response to somatostatin analog treatment<br />

[41]. In contrast to that observed in neuroendocrine<br />

tumors, in advanced pancreatic <strong>and</strong><br />

34 Chemotherapy 2001;47(suppl 2):30–39 Bousquet/Puente/Buscail/Vaysse/Susini


prostatic as well as colorectal adenocarcinoma,<br />

SSTR-2 expression is lost. This may have<br />

a growth advantage in these tumors <strong>and</strong> provide<br />

one explanation <strong>for</strong> the lack of therapeutic<br />

effect of somatostatin analogs in such tumors<br />

[42, 43].<br />

The presence of somatostatin receptors in<br />

tumors argues in favor of a direct role <strong>for</strong><br />

somatostatin in the regulation of tumor<br />

growth. A direct inhibitory effect of somatostatin<br />

or analogs on cell growth has been demonstrated<br />

on various cancer cell lines which<br />

express endogenous somatostatin receptors<br />

(cells of mammary, pancreatic, gastric, lung,<br />

colorectal or thyroid origin) [<strong>for</strong> a review, see<br />

4]. However, the mechanisms of cell growth<br />

arrest induced by somatostatin are still poorly<br />

understood. The fact that these tumor cells<br />

express multiple somatostatin receptors raises<br />

the questions of whether different somatostatin<br />

receptor(s) may account <strong>for</strong> these effects<br />

<strong>and</strong> which mechanisms are involved.<br />

The direct inhibitory action of somatostatin<br />

on cell growth may result from the blockade<br />

of mitogenic growth factor signal. In NIH<br />

3T3 or CHO cells expressing SSTR-1 or<br />

SSTR-2, octreotide <strong>and</strong> vapreotide inhibit<br />

cell proliferation induced by serum or insulin<br />

with an affinity which correlates with the<br />

binding affinity of analogs to each receptor<br />

subtype. This effect involves the stimulation<br />

of a tyrosine phosphatase [16, 19]. The somatostatin-sensitive<br />

tyrosine phosphatase has recently<br />

been identified as SHP-1. This tyrosine<br />

phosphatase contains two SH2 domains<br />

which are involved in protein-protein interactions.<br />

SHP-1 associates in vivo with various<br />

activated tyrosylated growth factor tyrosine<br />

kinase receptors <strong>and</strong> cytokine receptors. Recent<br />

studies have suggested a role of SHP-1 in<br />

terminating growth factor <strong>and</strong> cytokine mitogenic<br />

signals by dephosphorylating critical<br />

molecules [44]. In CHO cells expressing<br />

SSTR-2, SHP-1 is weakly associated with the<br />

Antiproliferative Effect of <strong>Somatostatin</strong><br />

<strong>and</strong> <strong>Analogs</strong><br />

SSTR-2 receptor via the protein Gi·3 at the<br />

resting level <strong>and</strong> occupation of SSTR-2 by<br />

analogs transiently increases the <strong>for</strong>mation of<br />

SSTR-2-SHP-1-Gi·3 complexes leading to the<br />

activation of the enzyme. The activated enzyme<br />

rapidly dissociates from the SSTR-2<br />

subtype, associates with the activated <strong>and</strong> tyrosylated<br />

insulin receptor, dephosphorylates<br />

it <strong>and</strong> its substrates, thus leading to a negative<br />

regulation of the insulin mitogenic signaling<br />

[45] due to G1 cell cycle arrest <strong>and</strong><br />

inhibition of insulin-induced S phase entry<br />

[Pages, pers. results]. A somatostatin-induced<br />

increase of SHP-1 translocation to the membrane<br />

is also observed in MCF-7 mammary<br />

cancer cells [46]. The use of dominant negative<br />

SHP-1 reveals that SHP-1 is required<br />

<strong>for</strong> the antiproliferative signal initiated by<br />

SSTR-2 [47]. In addition, SSTR-4 could also<br />

mediate the antiproliferative effect of somatostatin<br />

via a tyrosine phosphatase since in<br />

the rat thyroid PC13 cell line which only<br />

expresses SSTR-4 mRNA, somatostatin inhibits<br />

insulin- <strong>and</strong> insulin-plus-TSH-dependent<br />

proliferation through the stimulation of<br />

a tyrosine phosphatase <strong>and</strong> induces a block<br />

in the G1/S progression in the cell cycle [48].<br />

Concerning SSTR-5, its expression in CHO-<br />

K1 cells leads to somatostatin analog-induced<br />

inhibition of cell proliferation stimulated<br />

by serum or cholecystokinin via a pertussis<br />

toxin-dependent Gi/0 protein [20]. The<br />

signaling pathway coupled to SSTR-5 <strong>and</strong><br />

leading to inhibition of proliferation is selective<br />

<strong>for</strong> SSTR-5 <strong>and</strong> has been recently<br />

identified. It involves the inhibition of guanylate<br />

cyclase leading to a decrease of cGMP<br />

<strong>for</strong>mation, inhibition of cGMP-dependent<br />

protein kinase G <strong>and</strong> inhibition of MAP kinase<br />

ERK1/2 [21].<br />

The antiproliferative effect of somatostatin<br />

can also result from apoptosis. Apoptosis<br />

has been reported to be induced by the SSTR-<br />

3 subtype via a G protein-dependent signaling<br />

Chemotherapy 2001;47(suppl 2):30–39 35


<strong>and</strong> to be associated with an intracellular<br />

acidification <strong>and</strong> activation of endonuclease<br />

<strong>and</strong> induction of p53 <strong>and</strong> Bax [49, 50]. In<br />

human pancreatic cancer cells expressing mutated<br />

p53 <strong>and</strong> devoid of endogenous SSTR-2,<br />

expression of SSTR-2 does not result in a G1<br />

cell cycle arrest but induces an increase in cell<br />

death [Rochaix, pers. results]. This indicates<br />

that apoptosis can be signaled by other SSTR<br />

than SSTR-3 <strong>and</strong> somatostatin can induce<br />

apoptosis by p53-dependent <strong>and</strong> p53-independent<br />

mechanisms.<br />

Besides the antiproliferative effect of somatostatin<br />

due to cell growth arrest <strong>and</strong>/or<br />

apoptosis, somatostatin may directly control<br />

cell growth by inhibiting the synthesis <strong>and</strong>/or<br />

the secretion of autocrine growth factors, cytokines<br />

<strong>and</strong> hormones involved in the proliferation<br />

of tumor cells. It is well known that<br />

the aberrant expression of growth factors, cytokines<br />

or hormones <strong>and</strong> their receptors represent<br />

fundamental circuits that may spur<br />

<strong>and</strong> sustain uncontrolled growth <strong>and</strong> metastatic<br />

behavior of cancer cells. For example,<br />

EGF-related growth factors such as trans<strong>for</strong>ming<br />

growth factor-· <strong>and</strong> amphiregulin<br />

<strong>and</strong>/or their specific receptor, the EGF receptor<br />

as well as IGF-1 receptor <strong>and</strong> its lig<strong>and</strong>,<br />

have been detected in several types of human<br />

cancers, including breast, lung, pancreatic<br />

<strong>and</strong> colorectal cancers [51, 52]. Gastrin <strong>and</strong><br />

CCK-B receptor iso<strong>for</strong>ms are coexpressed<br />

in gastrointestinal cancer cells [53]. These<br />

growth factors may act by paracrine <strong>and</strong> autocrine<br />

mechanisms to exert growth promoting<br />

<strong>and</strong> metabolic effects. <strong>Somatostatin</strong> may influence<br />

the synthesis <strong>and</strong>/or the secretion of<br />

these factors <strong>and</strong>/or downregulate the expression<br />

of their receptors leading to disruption of<br />

proliferative autocrine loops. At the cellular<br />

level, blockade of secretion by somatostatin is<br />

mediated through inhibition of Ca 2+ <strong>and</strong><br />

cAMP production. Additionally, somatostatin<br />

can directly interfere with the exocytotic<br />

machinery by inhibiting the protein phosphatase<br />

calcineurin [14]. The specific SSTR subtypes<br />

involved in these processes <strong>and</strong> the<br />

mechanisms involved remain to be determined.<br />

Recent results using the patch-clamp<br />

technique indicate that in human neuroendocrine<br />

gut tumor cells, somatostatin <strong>and</strong> octreotide<br />

inhibit L-type voltage-dependent calcium<br />

channels with the same amplitude suggesting<br />

that at least SSTR-2 <strong>and</strong> SSTR-5 may<br />

be involved in the inhibition of Ca 2+ influx<br />

<strong>and</strong> thereby inhibition of tumor-produced<br />

neurotransmitters <strong>and</strong> hormone [54].<br />

The antiproliferative effects of somatostatin<br />

result from its actions via the endocrine<br />

pathway but evidence exists that somatostatin<br />

can also act via an autocrine/paracrine pathway.<br />

Indeed, immunoreactive somatostatin<br />

has been found in somatostatin receptor-positive<br />

normal <strong>and</strong> tumor cell types such as<br />

endocrine <strong>and</strong> lymphoid cells, breast cancer<br />

cells, colonic tumor cells <strong>and</strong>, additionally,<br />

somatostatin mRNA is detected in a wide<br />

variety of neuroendocrine tumors known to<br />

express somatostatin receptors [55–58]. Correction<br />

of the SSTR-2 deficit in human pancreatic<br />

cancer cells by SSTR-2 expression induces<br />

a negative autocrine loop in the absence<br />

of exogenous lig<strong>and</strong>, which is due to the<br />

SSTR-2-induced expression <strong>and</strong> secretion of<br />

endogenous SSTR-2 lig<strong>and</strong> (somatostatin 14<br />

<strong>and</strong> somatostatin 28). This results in inhibition<br />

of cancer cell proliferation <strong>and</strong> reversion<br />

of cell tumorigenicity in vitro <strong>and</strong> in vivo<br />

after xenografts in nude mice [59]. SSTR-2<br />

may function as a determinant factor in the<br />

negative control of cell growth <strong>and</strong> a loss of<br />

such an autocrine loop by the loss of expression<br />

of one component such as SSTR-2 or<br />

endogenous lig<strong>and</strong> contributes to the malignancy<br />

of cancers.<br />

As observed <strong>for</strong> other receptors coupled to<br />

G protein, somatostatin receptors are sensitive<br />

to agonist-induced desensitization <strong>and</strong>/or<br />

36 Chemotherapy 2001;47(suppl 2):30–39 Bousquet/Puente/Buscail/Vaysse/Susini


internalization <strong>and</strong>/or up- or down-regulation.<br />

Indeed, agonist-dependent internalization<br />

<strong>and</strong> upregulation of somatostatin receptors<br />

have been demonstrated in various cell<br />

lines expressing multiple SSTRs or transfected<br />

cells expressing individual subtype after<br />

prolonged agonist exposure. Although the<br />

molecular mechanisms involved in these effects<br />

are poorly understood, it appears that<br />

the response of somatostatin receptors to agonist<br />

application is agonist-, receptor- <strong>and</strong> cell<br />

type-specific. Indeed, a different pattern of<br />

internalization has been reported <strong>for</strong> the five<br />

SSTRs, in CHO-K1 cells <strong>and</strong> in human embryonic<br />

kidney (HEK) cells expressing individual<br />

subtypes. In CHO-K1 cells, SSTR-2–5<br />

undergo rapid internalization following agonist<br />

activation but SSTR-1 is not internalized.<br />

In contrast, in HEK cells, SSTR-1, SSTR-2<br />

<strong>and</strong> SSTR-3 are internalized in response to<br />

somatostatin 14 or somatostatin 28 whereas<br />

SSTR-5 is only internalized in response to<br />

somatostatin 28 <strong>and</strong> SSTR-4 is not internalized<br />

[<strong>for</strong> a review, see 8, 37]. In addition,<br />

somatostatin upregulates SSTRs in a receptor<br />

subtype-specific manner. Long-term exposure<br />

of CHO-K1 cells to somatostatin upregulates<br />

SSTR-1, SSTR-2 <strong>and</strong> SSTR-4 by 110, 26 <strong>and</strong><br />

References<br />

1 Lewin MJ: The somatostatin receptor<br />

in the GI tract. Annu Rev Physiol<br />

1992;54:455–468.<br />

2 Epelbaum J, Dournaud P, Fodor M,<br />

Viollet C: The neurobiology of somatostatin.<br />

Crit Rev Neurobiol<br />

1994;8/1–2:25–44.<br />

3 Schally AV: Oncological applications<br />

of somatostatin analogues<br />

<strong>Cancer</strong> Res 1988;48:6977–6985.<br />

4 Weckbecker G, Raulf F, Stolz B,<br />

Bruns C: <strong>Somatostatin</strong> analogs <strong>for</strong><br />

diagnosis <strong>and</strong> treatment of cancer.<br />

Pharmacol Ther 1993;60/2:245–<br />

264.<br />

Antiproliferative Effect of <strong>Somatostatin</strong><br />

<strong>and</strong> <strong>Analogs</strong><br />

5 Oppizzi G, Cozzi R, Dallabonzana<br />

D, Orl<strong>and</strong>i P, Benini Z, Petroncini<br />

M, Attanasio R, Milella M, Banfi G,<br />

Possa M: Scintigraphic imaging of<br />

pituitary adenomas: An in vivo evaluation<br />

of somatostatin receptors. J<br />

Endocrinol Invest 1998;21:512–<br />

519.<br />

6 Arnold R, Trautmann ME, Creutzfeldt<br />

W, Benning R, Benning M,<br />

Neuhaus C, Jurgensen R, Stein K,<br />

Schafer H, Bruns C, Dennler HJ:<br />

<strong>Somatostatin</strong> analogue octreotide<br />

<strong>and</strong> inhibition of tumour growth in<br />

metastatic endocrine gastroenteropancreatic<br />

tumours. Gut 1996;38:<br />

430–438.<br />

22%, respectively, whereas the levels of SSTR-<br />

3 <strong>and</strong> SSTR-5 are not modified [<strong>for</strong> a review,<br />

see 37]. These results suggest that these phenomena<br />

may be important <strong>for</strong> physiological<br />

<strong>and</strong> clinical indirect <strong>and</strong>/or direct response to<br />

somatostatin peptides. At the cellular level,<br />

one can expect that upregulation of somatostatin<br />

receptors transducing the antiproliferative<br />

effect of somatostatin by endogenous or exogenous<br />

somatostatin peptides can enhance the<br />

antiproliferative effect of the peptide.<br />

In conclusion, much progress has been<br />

made towards elucidation of the characterization,<br />

the molecular signal transduction, the expression<br />

<strong>and</strong> the regulation of SSTRs. However,<br />

the biological role as well as the cellular<br />

distribution of each receptor subtype is far<br />

from being completely understood. The development<br />

of specific antibodies, agonists <strong>and</strong><br />

antagonists as well as gene knockout models<br />

will help to define the specific role of individual<br />

receptors in physiological <strong>and</strong> pathological<br />

conditions <strong>and</strong> the significance of multiple<br />

receptor subtypes in the same cell. Finally, the<br />

recent findings of antioncogenic properties of<br />

SSTR-2 <strong>for</strong> human pancreatic cancer cells suggest<br />

that SSTR-2 gene transfer may be a possible<br />

adjuvant gene therapy <strong>for</strong> pancreatic cancer.<br />

7 Reisine T, Bell GI: Molecular properties<br />

of somatostatin receptors.<br />

Neuroscience 1995;67:777–790.<br />

8 Meyerhof W: The elucidation of somatostatin<br />

receptor functions: A<br />

current view. Rev Physiol Biochem<br />

Pharmacol 1998;133:55–108.<br />

9 Kreienkamp HJ, Honck HH, Richter<br />

D: Coupling of rat somatostatin<br />

receptor subtypes to a G-protein<br />

gated inwardly rectifying potassium<br />

channel (GIRK1). FEBS Lett 1997;<br />

419/1:92–94.<br />

Chemotherapy 2001;47(suppl 2):30–39 37


10 Tallent M, Liapakis G, O’Carroll<br />

AM, Lolait SJ, Dichter M, Reisine<br />

T: <strong>Somatostatin</strong> receptor subtypes<br />

SSTR2 <strong>and</strong> SSTR5 couple negatively<br />

to an L-type Ca 2+ current in<br />

the pituitary cell line AtT-20. Neuroscience<br />

1996;71:1073–1081.<br />

11 Roosterman D, Glassmeier G, Baumeister<br />

H, Scherubl H, Meyerhof<br />

W: A somatostatin receptor 1 selective<br />

lig<strong>and</strong> inhibits Ca 2+ currents<br />

in rat insulinoma 1046-38 cells.<br />

FEBS Lett 1998;425/1:137–140.<br />

12 Le Romancer M, Reyl-Desmars F,<br />

Cherifi Y, Pigeon C, Bottari S,<br />

Meyer O, Lewin MJ: The 86-kDa<br />

subunit of autoantigen Ku is a somatostatin<br />

receptor regulating protein<br />

phosphatase-2A activity. J Biol<br />

Chem 1994;269:17464–17468.<br />

13 White RE, Schonbrunn A, Armstrong<br />

DL: <strong>Somatostatin</strong> stimulates<br />

Ca(2+)-activated K+ channels<br />

through protein dephosphorylation.<br />

Nature 1991;351:570–573.<br />

14 Renstrom E, Ding WG, Bokvist K,<br />

Rorsman P: Neurotransmitter-induced<br />

inhibition of exocytosis in insulin-secreting<br />

beta cells by activation<br />

of calcineurin. Neuron 1996;17:<br />

513–522.<br />

15 Liebow C, Reilly C, Serrano M,<br />

Schally AV: <strong>Somatostatin</strong> analogues<br />

inhibit growth of pancreatic cancer<br />

by stimulating tyrosine phosphatase.<br />

Proc Natl Acad Sci USA 1989;<br />

86:2003–2007.<br />

16 Buscail L, Delesque N, Esteve JP,<br />

Saint-Laurent N, Prats H, Clerc P,<br />

Robberecht P, Bell GI, Liebow C,<br />

Schally AV, Vaysse N, Susini C:<br />

Stimulation of tyrosine phosphatase<br />

<strong>and</strong> inhibition of cell proliferation<br />

by somatostatin analogues: Mediation<br />

by human somatostatin receptor<br />

subtypes SSTR1 <strong>and</strong> SSTR2.<br />

Proc Natl Acad Sci USA 1994;91:<br />

2315–2319.<br />

17 Reardon DB, Dent P, Wood SL,<br />

Kong T, Sturgill TW: Activation in<br />

vitro of somatostatin receptor subtypes<br />

2, 3, or 4 stimulates protein<br />

tyrosine phosphatase activity in<br />

membranes from transfected Rastrans<strong>for</strong>med<br />

NIH 3T3 cells: Coexpression<br />

with catalytically inactive<br />

SHP-2 blocks responsiveness. Mol<br />

Endocrinol 1997;11:1062–1069.<br />

18 Florio T, Rim C, Hershberger RE,<br />

Loda M, Stork PJ: The somatostatin<br />

receptor SSTR1 is coupled to phosphotyrosine<br />

phosphatase activity in<br />

CHO-K1 cells. Mol Endocrinol<br />

1994;8:1289–1297.<br />

19 Murthy KS, Coy DH, Makhlouf<br />

GM: <strong>Somatostatin</strong> receptor-mediated<br />

signaling in smooth muscle.<br />

Activation of phospholipase C-ß3 by<br />

GßÁ <strong>and</strong> inhibition of adenylyl cyclase<br />

by G·i1 <strong>and</strong> G·0. J Biol Chem<br />

1996;271:23458–23463.<br />

20 Buscail L, Esteve JP, Saint-Laurent<br />

N, Bertr<strong>and</strong> V, Reisine T, O’Carroll<br />

AM, Bell GI, Schally AV, Vaysse N,<br />

Susini C: Inhibition of cell proliferation<br />

by the somatostatin analogue<br />

RC-160 is mediated by somatostatin<br />

receptor subtypes SSTR2 <strong>and</strong><br />

SSTR5 through different mechanisms.<br />

Proc Natl Acad Sci USA<br />

1995;92:1580–1584.<br />

21 Cordelier P, Esteve JP, Bousquet C,<br />

Delesque N, O’Carroll AM, Schally<br />

AV, Vaysse N, Susini C, Buscail L:<br />

Characterization of the antiproliferative<br />

signal mediated by the somatostatin<br />

receptor subtype SSTR-5.<br />

Proc Natl Acad Sci USA 1997;94:<br />

9343–9348.<br />

22 Hou C, Gilbert RL, Barber DL: Subtype-specific<br />

signaling mechanisms<br />

of somatostatin receptors SSTR1<br />

<strong>and</strong> SSTR2. J Biol Chem 1994;269:<br />

10357–10362.<br />

23 Macaulay VM: Insulin-like growth<br />

factors <strong>and</strong> cancer. Br J <strong>Cancer</strong><br />

1992;65:311–320.<br />

24 Rohrer SP, Birzin ET, Mosley RT,<br />

Berk SC, Hutchins SM, Shen DM,<br />

Xiong Y, Hayes EC, Parmar RM,<br />

Foor F, Mitra SW, Degrado SJ, Shu<br />

M, Klopp JM, Cai SJ, Blake A, Chan<br />

WWS, Pasternak A, Yang L, Patchett<br />

AA, Smith RG, Chapman KT,<br />

Schaeffer JM: Rapid identification<br />

of subtype-selective agonists of the<br />

somatostatin receptor through combinatorial<br />

chemistry. Science 1998;<br />

282:737–740.<br />

25 Serri O, Brazeau P, Kachra Z, Posner<br />

B: Octreotide inhibits insulin-like<br />

growth factor-I hepatic gene expression<br />

in the hypophysectomized rat:<br />

Evidence <strong>for</strong> a direct <strong>and</strong> indirect<br />

mechanism of action. Endocrinology<br />

1992;130:1816–1821.<br />

26 Newman CB, Melmed S, George A,<br />

Torigian D, Duhaney M, Snyder P,<br />

Young W, Klibanski A, Molitch<br />

ME, Gagel R, Sheeler L, Cook D,<br />

Malarkey W, Jackson I, Vance ML,<br />

Barkan A, Frohman L, Kleinberg<br />

DL: Octreotide as primary therapy<br />

<strong>for</strong> acromegaly. J Clin Endocrinol<br />

Metab 1998;83:3034–3040.<br />

27 Canobbio L, Cannata D, Miglietta<br />

L, Boccardo F: Somatuline (BIM<br />

23014) <strong>and</strong> tamoxifen treatment of<br />

postmenopausal breast cancer patients:<br />

Clinical activity <strong>and</strong> effect on<br />

insulin-like growth factor-I (IGF-I)<br />

levels. Anticancer Res 1995;15/6B:<br />

2687–2690.<br />

28 Ren SG, Ezzat S, Melmed S, Braunstein<br />

GD: <strong>Somatostatin</strong> analog induces<br />

insulin-like growth factor<br />

binding protein-1 (IGFBP-1) expression<br />

in human hepatoma cells.<br />

Endocrinology 1992;131:2479–<br />

2481.<br />

29 Martinez V, Curi AP, Torkian B,<br />

Schaeffer JM, Wilkinson HA, Walsh<br />

JH, Tache Y: High basal gastric acid<br />

secretion in somatostatin receptor<br />

subtype 2 knockout mice. Gastroenterology<br />

1998;114:1125–1132.<br />

30 Reubi JC, Kappeler A, Waser B,<br />

Schonbrunn A, Laissue J: Immunohistochemical<br />

localization of somatostatin<br />

receptor SSTR-2A in human<br />

pancreatic islets. J Clin Endocrinol<br />

Metab 1998;83:3746–3749.<br />

31 Woltering EA, Watson JC, Alperin-<br />

Lea RC, Sharma C, Keenan E, Kurozawa<br />

D, Barrie R: <strong>Somatostatin</strong><br />

analogs: Angiogenesis inhibitors<br />

with novel mechanisms of action.<br />

Invest New Drugs 1997;15/1:77–<br />

86.<br />

32 Reubi JC, Horisberger U, Laissue J:<br />

High density of somatostatin receptors<br />

in veins surrounding human<br />

cancer tissue: Role in tumor-host interaction?<br />

Int J <strong>Cancer</strong> 1994;56:<br />

681–688.<br />

33 van Hagen PM, Krenning EP,<br />

Kwekkeboom DJ, Reubi JC, Anker-<br />

Lugtenburg PJ, Lowenberg B, Lamberts<br />

SW: <strong>Somatostatin</strong> <strong>and</strong> the immune<br />

<strong>and</strong> haematopoetic system; a<br />

review. Eur J Clin Invest 1994;24/2:<br />

91–99.<br />

34 Tsutsumi A, Takano H, Ichikawa K,<br />

Kobayashi S, Koike T: Expression<br />

of somatostatin receptor subtype 2<br />

mRNA in human lymphoid cells.<br />

Cell Immunol 1997;181:44–49.<br />

38 Chemotherapy 2001;47(suppl 2):30–39 Bousquet/Puente/Buscail/Vaysse/Susini


35 Casnici C, Lattuada D, Perego C,<br />

Franco P, Marelli O: Inhibitory effect<br />

of somatostatin on human T<br />

lymphocytes proliferation. Int J Immunopharmacol<br />

1997;19:721–727.<br />

36 Elliott DE, Metwali A, Blum AM,<br />

S<strong>and</strong>or M, Lynch R, Weinstock JV:<br />

T lymphocytes isolated from the hepatic<br />

granulomas of schistosome-infected<br />

mice express somatostatin receptor<br />

subtype II (SSTR2) messenger<br />

RNA. J Immunol 1994;153:<br />

1180–1186.<br />

37 Patel YC: Molecular pharmacology<br />

of somatostatin receptor subtypes. J<br />

Endocrinol Invest 1997;20:348–<br />

367.<br />

38 Reubi JC, Schaer JC, Waser B, Mengod<br />

G: Expression <strong>and</strong> localization<br />

of somatostatin receptor SSTR1,<br />

SSTR2, <strong>and</strong> SSTR3 messenger<br />

RNAs in primary human tumors using<br />

in situ hybridization. <strong>Cancer</strong><br />

Res 1994;54:3455–3459.<br />

39 Dutour A, Kumar U, Panetta R,<br />

Ouafik L, Fina F, Sasi R, Patel YC:<br />

Expression of somatostatin receptor<br />

subtypes in human brain tumors. Int<br />

J <strong>Cancer</strong> 1998;76:620–627.<br />

40 Schulz S, Schulz S, Schmitt J, Wiborny<br />

D, Schmidt H, Olbricht S,<br />

Weise W, Roessner A, Gramsch C,<br />

Hollt V: Immunocytochemical detection<br />

of somatostatin receptors<br />

SSTR-1, SSTR-2A, SSTR-2B, <strong>and</strong><br />

SSTR-3 in paraffin-embedded<br />

breast cancer tissue using subtypespecific<br />

antibodies. Clin <strong>Cancer</strong> Res<br />

1998;4:2047–2052.<br />

41 Janson ET, Stridsberg M, Gobl A,<br />

Westlin JE, Oberg K: Determination<br />

of somatostatin receptor subtype<br />

2 in carcinoid tumors by immunohistochemical<br />

investigation with<br />

somatostatin receptor subtype 2 antibodies.<br />

<strong>Cancer</strong> Res 1998;58:2375–<br />

2378.<br />

42 Buscail L, Saint-Laurent N, Chastre<br />

E, Vaillant JC, Gespach C, Capella<br />

G, Kalthoff H, Lluis F, Vaysse N,<br />

Susini C: Loss of SSTR-2 somatostatin<br />

receptor gene expression in human<br />

pancreatic <strong>and</strong> colorectal cancer.<br />

<strong>Cancer</strong> Res 1996;56:1823–<br />

1827.<br />

43 Reubi JC, Waser B, Schaer JC,<br />

Markwalder R: <strong>Somatostatin</strong> receptors<br />

in human prostate <strong>and</strong> prostate<br />

cancer. J Clin Endocrinol Metab<br />

1995;80:2806–2814.<br />

Antiproliferative Effect of <strong>Somatostatin</strong><br />

<strong>and</strong> <strong>Analogs</strong><br />

44 Neel BG, Tonks NK: Protein tyrosine<br />

phosphatases in signal transduction.<br />

Curr Opin Cell Biol 1997;<br />

9/2:193–204.<br />

45 Bousquet C, Delesque N, Lopez F,<br />

Saint-Laurent N, Esteve JP, Bedecs<br />

K, Buscail L, Vaysse N, Susini C:<br />

SSTR-2 somatostatin receptor mediates<br />

negative regulation of insulin<br />

receptor signaling through the tyrosine<br />

phosphatase Shp-1. J Biol<br />

Chem 1998;273:7099–7106.<br />

46 Srikant CB, Shen SH: Octapeptide<br />

somatostatin analog SMS 201-995<br />

induces translocation of intracellular<br />

PTP1C to membranes in MCF-7<br />

human breast adenocarcinoma cells.<br />

Endocrinology 1996;137:3461–<br />

3468.<br />

47 Lopez F, Esteve JP, Buscail L, Delesque<br />

N, Saint-Laurent N, Theveniau<br />

M, Nahmias C, Vaysse N, Susini<br />

C: The tyrosine phosphatase<br />

SHP-1 associates with the SSTR-2<br />

somatostatin receptor <strong>and</strong> is an essential<br />

component of SSTR-2-mediated<br />

inhibitory growth signaling. J<br />

Biol Chem 1997;272:24448–24454.<br />

48 Florio T, Scorizello A, Fattore M,<br />

D’Alto V, Salzano S, Rossi G, et al:<br />

<strong>Somatostatin</strong> inhibits PC Cl3 thyroid<br />

cell proliferation through the<br />

modulation of phosphotyrosine activity.<br />

Impairment of the somatostatinergic<br />

effects by stable expression<br />

of E1A viral oncogene. J Biol Chem<br />

1996;271:6129–6136.<br />

49 Sharma K, Patel YC, Srikant CB:<br />

Subtype-selective induction of wildtype<br />

p53 <strong>and</strong> apoptosis, but not cell<br />

cycle arrest, by human somatostatin<br />

receptor 3. Mol Endocrinol 1996;10:<br />

1688–1696.<br />

50 Sharma K, Srikant CB: G protein<br />

coupled receptor signaled apoptosis<br />

is associated with activation of a cation<br />

insensitive acidic endonuclease<br />

<strong>and</strong> intracellular acidification Biochem<br />

Biophys Res Commun 1998;<br />

242:134–140.<br />

51 Quinn KA, Treston AM, Unsworth<br />

EJ, Miller MJ, Vos M, Grimley C,<br />

Battey J, Mulshine JL, Cuttitta F:<br />

Insulin-like growth factor expression<br />

in human cancer cell lines. J<br />

Biol Chem 1996;27/1:11477–<br />

11483.<br />

52 Korc M: Role of growth factors in<br />

pancreatic cancer. Surg Oncol Clin<br />

N Am 1998;71:25–41.<br />

53 McWilliams DF, Watson SA, Crosbee<br />

DM, Michaeli D, Seth R: Coexpression<br />

of gastrin <strong>and</strong> gastrin receptors<br />

(CCK-B <strong>and</strong> delta CCK-B) in<br />

gastrointestinal tumour cell lines.<br />

Gut 1998;42:795–798.<br />

54 Glassmeier G, Hopfner M, Riecken<br />

EO, Mann B, Buhr H, Neuhaus P,<br />

Meyerhof W, Scherubl H: Inhibition<br />

of L-type calcium channels by octreotide<br />

in isolated human neuroendocrine<br />

tumor cells of the gut. Biochem<br />

Biophys Res Commun 1998;<br />

250:511–515.<br />

55 Levy L, Bourdais J, Mouhieddine B,<br />

Benlot C, Villares S, Cohen P, Peillon<br />

F, Joubert D: Presence <strong>and</strong> characterization<br />

of the somatostatin precursor<br />

in ormal human pituitaries<br />

<strong>and</strong> in growth hormone secreting adenomas.<br />

J Clin Endocrinol Metab<br />

1993;76/1:85–90.<br />

56 Reubi JC, Waser B, Lamberts SW,<br />

Mengod G: <strong>Somatostatin</strong> (SRIH)<br />

messenger ribonucleic acid expression<br />

in human neuroendocrine <strong>and</strong><br />

brain tumors using in situ hybridization<br />

histochemistry: Comparison<br />

with SRIH receptor content. J Clin<br />

Endocrinol Metab 1993;76:642–<br />

647.<br />

57 Nelson J, Cremin M, Murphy RF:<br />

Synthesis of somatostatin by breast<br />

cancer cells <strong>and</strong> their inhibition by<br />

exogenous somatostatin <strong>and</strong> s<strong>and</strong>ostatin.<br />

Br J <strong>Cancer</strong> 1989;59:739–<br />

742.<br />

58 Elliott DE, Blum AM, Li J, Metwali<br />

A, Weinstock JV: Preprosomatostatin<br />

messenger RNA is expressed by<br />

inflammatory cells <strong>and</strong> induced by<br />

inflammatory mediators <strong>and</strong> cytokines.<br />

J Immunol 1998;160:3997–<br />

4003.<br />

59 Delesque N, Buscail L, Esteve JP,<br />

Saint-Laurent N, Muller C, Weckbecker<br />

G, Bruns C, Vaysse N, Susini<br />

C: SSTR-2 somatostatin receptor<br />

expression reverses tumorigenicity<br />

of human pancreatic cancer cells.<br />

<strong>Cancer</strong> Res 1997;57:956–962.<br />

Chemotherapy 2001;47(suppl 2):30–39 39


Chemotherapy 2001;47(suppl 2):40–53<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

Kjell Öberg<br />

Department of Medical Sciences, Uppsala University, Uppsala, Sweden<br />

Key Words<br />

Neuroendocrine tumors W Octreoscan W<br />

Octreotide W Lanreotide W Octastatin<br />

Abstract<br />

The diagnosis <strong>and</strong> treatment of neuroendocrine<br />

tumors have been significantly improved<br />

during the last decades. Localization<br />

<strong>and</strong> staging of the disease by somatostatin<br />

receptor scintigraphy (Octreoscan) are now<br />

the ‘gold st<strong>and</strong>ard’ <strong>for</strong> the management of<br />

these tumors. <strong>Treatment</strong> with somatostatin<br />

analogs has improved quality of life <strong>and</strong><br />

possibly also survival <strong>for</strong> patients with neuroendocrine<br />

tumors. New long-acting <strong>for</strong>mulations<br />

of the somatostatin analogs are as<br />

effective as the old regular <strong>for</strong>mulations<br />

but will further improve quality of life <strong>for</strong><br />

the patients. Tumor-targeted therapy with<br />

111 In <strong>and</strong> 90 Y coupled to somatostatin analogs<br />

show promising results but await further<br />

studies.<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

Copyright © 2001 S. Karger AG, Basel<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0040$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Introduction<br />

Octreotide <strong>and</strong> lanreotide are registered in<br />

most countries <strong>for</strong> the control of symptoms<br />

associated with metastatic carcinoid <strong>and</strong> VIPsecreting<br />

tumors. The clinical efficacy of somatostatin<br />

analogs has been established <strong>for</strong><br />

many neuroendocrine tumors including acromegaly,<br />

although the latter will not be discussed<br />

in this paper.<br />

Neuroendocrine gut <strong>and</strong> pancreatic tumors<br />

constitute about 2% of all malignant<br />

gastrointestinal tumors. Carcinoids are the<br />

most common type of neuroendocrine tumors<br />

<strong>and</strong> may be classified according to the region<br />

in which they arise into <strong>for</strong>egut, midgut <strong>and</strong><br />

hindgut carcinoids [1, 2]. Foregut tumors<br />

originate in the thymus, lung <strong>and</strong> gastroduodenal<br />

mucosa <strong>and</strong> represent about 15% of all<br />

carcinoids. Midgut carcinoids with the primary<br />

located in the jejunum, ileum <strong>and</strong> the proximal<br />

colon including the appendix comprise<br />

the most common (40–50%) of all clinically<br />

relevant carcinoids. Hindgut tumors lo-<br />

Prof. Kjell Öberg<br />

Department of Medical Sciences, Uppsala University<br />

Uppsala (Sweden)<br />

Tel. +46 18 66 49 17, Fax +46 18 51 01 33<br />

E-Mail Kjell.Oberg@medicin.uu.se


cated in the rectum, sigmoid <strong>and</strong> the distal<br />

colon comprise approximately 20–25% of all<br />

carcinoids <strong>and</strong> are normally hormonally inactive<br />

(nonfunctioning). The hormones secreted<br />

by <strong>for</strong>egut carcinoids are varied giving rise to<br />

many clinical syndromes such as acromegaly,<br />

Cushing’s disease, SIADH, Zollinger-Ellison<br />

syndrome as well as the carcinoid syndrome<br />

per se.<br />

The above classification of carcinoid tumors<br />

has been questioned since it causes considerable<br />

confusion. Consequently the term<br />

carcinoid might in the future be reserved<br />

only <strong>for</strong> classical midgut carcinoids. Other<br />

neuroendocrine tumors in the gut should be<br />

assigned the term ‘neuroendocrine tumors’<br />

followed by their primary location, e.g. neuroendocrine,<br />

lung, gastric, duodenal, pancreatic,<br />

colonic <strong>and</strong> rectal tumors. The dominant<br />

hormone produced by the tumors may<br />

also be included in the classification, e.g. gastrin-producing<br />

neuroendocrine tumor. Such a<br />

classification would certainly be helpful in<br />

communicating in<strong>for</strong>mation about these tumors<br />

<strong>and</strong> also in evaluating potential therapies<br />

[1].<br />

The carcinoid syndrome is a well-defined<br />

clinical syndrome which includes flushes, diarrhea,<br />

carcinoid heart disease with right<br />

heart failure <strong>and</strong> bronchial constriction accompanied<br />

by elevated levels of plasma serotonin<br />

<strong>and</strong>/or urinary 5-HIAA. The tumors<br />

also release tachykinins, bradykinins <strong>and</strong><br />

prostagl<strong>and</strong>ins. For some patients the syndrome<br />

is severe enough to be potentially lifethreatening<br />

with extensive flushing combined<br />

with hypotension or very frequent diarrhea,<br />

the so-called ‘carcinoid crisis’ [2, 3]. Another<br />

group of neuroendocrine tumors are the endocrine<br />

pancreatic tumors which are classified<br />

as funtioning if they are associated with a clinical<br />

syndrome related to the hormone they<br />

produce or which are considered nonfunctioning<br />

if they do not present with clinical<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

symptoms of excessive hormone release. The<br />

latter category constitutes around 30% of all<br />

endocrine pancreatic tumors <strong>and</strong> includes<br />

those which secrete pancreatic polypeptide,<br />

chromogranin A, peptide YY <strong>and</strong> neurotensin<br />

[4, 5].<br />

The two most common clinical syndromes<br />

related to endocrine pancreatic tumors are the<br />

Zollinger-Ellison syndrome resulting from<br />

gastrin overproduction [6] <strong>and</strong> the hypoglycemic<br />

syndrome which is related to high insulin/proinsulin<br />

release [7]. The Zollinger-Ellison<br />

syndrome (gastrinoma) may also arise<br />

from hypersecretion of gastrin production by<br />

duodenal carcinoids (around 40%). More<br />

than 70% of the gastrin-producing tumors are<br />

malignant with early lymph node involvement.<br />

Most neuroendocrine pancreatic tumors<br />

are malignant except <strong>for</strong> insulin-producing<br />

tumors where 80% are benign solitary<br />

tumors in the pancreas [7].<br />

Other functioning endocrine pancreatic<br />

tumors are the VIP-producing syndrome or<br />

WDHA syndrome which is characterized by<br />

extensive diarrhea, hypokalemia <strong>and</strong> achlorhydria<br />

[8]. In such patients the diarrhea volume<br />

might exceed more than 10 liters per<br />

day <strong>and</strong> they often require intensive care.<br />

Such tumors are mostly confined to the pancreas<br />

but sometimes may be located to the<br />

lung or sympathetic ganglia. Another rare tumor<br />

is the glucagonoma. The glucagonoma<br />

syndrome [9] is characterized by a typical<br />

necrolytic migratory erythema, diabetic glucose<br />

tolerance, anemia, weight loss <strong>and</strong><br />

thromboembolism which may be related because<br />

of glucagon production <strong>and</strong> its endorgan<br />

effects. Both these rare tumors have<br />

been treated successfully with somatostatin<br />

analogs.<br />

Multiple endocrine neoplasia type 1<br />

(MEN-1) is a familial disorder inherited as an<br />

autosomal dominant trait with variable penetrance<br />

patterns. In MEN-1 the pituitary, para-<br />

Chemotherapy 2001;47(suppl 2):40–53 41


thyroids <strong>and</strong> endocrine pancreas are the most<br />

commonly affected organs but the adrenal<br />

cortex <strong>and</strong> the thyroid can also be involved<br />

[10]. A specific genetic deletion has been described<br />

<strong>for</strong> MEN-1, i.e. a loss of heterozygosity<br />

of the MEN-1 locus on chromosome 11q13<br />

where the MEN-1 gene is deleted or mutated.<br />

About 80% of MEN-1 patients develop endocrine<br />

pancreatic tumors. Furthermore,<br />

about 30% of all gastrin-producing neuroendocrine<br />

gastrointestinal tumors are related to<br />

the MEN-1 syndrome. Some of these patients<br />

also develop lung, duodenal or gastric carcinoids.<br />

In fact, gastrinoma in MEN-1 patients<br />

is more often located in the duodenum than<br />

the pancreas itself <strong>and</strong> sometimes you can<br />

find the gastrinomas in both locations.<br />

The diagnosis of neuroendocrine tumors is<br />

based on histopathology, the cells showing the<br />

typical features of amine precursor uptakes<br />

<strong>and</strong> decarboxylation including positive silver<br />

staining (argyrophil, argentaffin) <strong>and</strong> positive<br />

staining <strong>for</strong> antibodies to chromogranin A,<br />

synaptophysin <strong>and</strong> NSE. The more sophisticated<br />

histopathology takes the tumor biology<br />

into account such as the proliferation markers<br />

(Ki-67, PCNA), angiogenetic factors (VEGF,<br />

b-FGF), adhesion molecules (CD-44, exon<br />

V6–V9) <strong>and</strong> finally deletions <strong>and</strong> mutations<br />

of the MEN-1 gene (menin).<br />

An essential component in the diagnosis of<br />

neuroendocrine tumors is the biochemical determination<br />

of the different peptide hormones<br />

they secrete. For example, determination<br />

of chromogranin A is the most important<br />

screening marker <strong>for</strong> the diagnosis of neuroendocrine<br />

tumors. Urinary 5-HIAA is important<br />

in diagnosing the carcinoid syndrome<br />

[1]. The localization <strong>and</strong> staging of neuroendocrine<br />

tumors may be significantly improved<br />

by somatostatin receptor scintigraphy<br />

(octreoscan), which will be discussed below in<br />

relation to treatment with somatostatin analogs.<br />

<strong>Somatostatin</strong> receptor scintigraphy is<br />

42 Chemotherapy 2001;47(suppl 2):40–53 Öberg<br />

nowadays the procedure of choice <strong>for</strong> localizing<br />

neuroendocrine tumors <strong>and</strong> supported by<br />

computed tomography <strong>and</strong> ultrasonography<br />

to determine the tumor size follow-up during<br />

therapy. Endoscopic ultrasonography has recently<br />

come into clinical use <strong>and</strong> has been particularly<br />

useful <strong>for</strong> localizing endocrine pancreatic<br />

tumors located particularly in the head<br />

of the pancreas.<br />

The prognosis of neuroendocrine tumors,<br />

particularly carcinoids, has normally been assumed<br />

to be relatively good without treatment.<br />

There<strong>for</strong>e, many physicians have been<br />

reluctant to administer medical treatment in<br />

the early stages of the disease. However, a<br />

critical analysis of the 5-year survival rates of<br />

patients with neuroendocrine tumors suggests<br />

that the prognosis of such patients is not as<br />

good as many physicians believe. Thus, the 5year<br />

survival rates of patients with neuroendocrine<br />

tumors is less than 20% when liver<br />

metastases are present. Furthermore, the median<br />

survival <strong>for</strong> patients with malignant carcinoid<br />

tumors with the carcinoid syndrome is<br />

less than 2 years from the time of diagnosis<br />

[1, 2].<br />

The therapy <strong>for</strong> neuroendocrine tumors is<br />

aimed at providing symptomatic relief by<br />

suppression of hypersecreting hormones, inhibition<br />

of tumor growth <strong>and</strong> at improving<br />

<strong>and</strong> maintain a good quality of life. Surgery is<br />

the treatment of choice <strong>for</strong> malignant neuroendocrine<br />

tumors, <strong>and</strong> even if surgical resection<br />

is not possible, debulking <strong>and</strong> bypassing<br />

procedures should be considered [11]. To<br />

further reduce the tumor burden, embolization,<br />

with <strong>and</strong> without cytotoxic agents has<br />

been demonstrated to have beneficial effects<br />

[12]. External irradiation has been of limited<br />

value in most neuroendocrine tumors but<br />

may relieve pain in some patients [13]. Hitherto<br />

tumor-targeted treatment has been attempted<br />

using 131 I-MIBG [14] but the technique<br />

has now been replaced by 111 In-DTPA-


octreotide. More recently 90 Y label compounds<br />

have also been evaluated [15]. The<br />

medical treatment <strong>for</strong> neuroendocrine tumors<br />

include chemotherapy, somatostatin analogs<br />

<strong>and</strong> interferon-· [1, 16, 17]. Chemotherapy,<br />

particularly streptozotocin plus 5-fluorouracil,<br />

has demonstrated to be a valuable therapeutic<br />

regimen <strong>for</strong> endocrine pancreatic tumors<br />

with response rates of 50–70% being<br />

reported [1, 16, 17]; however, classical midgut<br />

carcinoids do not respond to chemotherapy,<br />

its response rates being less than 10% [1, 16,<br />

18]. The therapeutic idea of somatostatin analogs<br />

in the management of neuroendocrine<br />

tumors will be discussed in detail below. Interferon-·<br />

has been shown to be of beneficial<br />

value, particularly in patients with midgut<br />

carcinoids with a reported subjective improvement<br />

in about 50% of the patients, a<br />

reduction in tumor size in 14% <strong>and</strong> biochemical<br />

responses in 40–50% of the patients [1,<br />

19]. Patients who tolerate interferon-· therapy<br />

frequently exhibit significant long-term responses<br />

with a median survival from the diagnosis<br />

of the carcinoid syndrome of more than<br />

60 months [20].<br />

<strong>Somatostatin</strong> Receptors<br />

<strong>Somatostatin</strong> receptors are expressed in somatostatin<br />

target tissues such as brain, pituitary,<br />

pancreas, gastrointestinal tract, blood<br />

vessels as well as in various types of tumors.<br />

At present five somatostatin receptor subtypes<br />

(SSTR-1–5) have been cloned <strong>and</strong> pharmacologically<br />

characterized [21–23]. All the<br />

somatostatin receptors are coupled to G proteins<br />

<strong>and</strong> belong to the seven-transmembranespanning<br />

receptor family. All the somatostatin<br />

receptor subtypes bind to their native hormones<br />

(SRIF-28 <strong>and</strong> SRIF-14) with high affinity.<br />

In contrast, short synthetic somatostatin<br />

(SRIF) analogs used clinically such as<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

octreotide, lanreotide <strong>and</strong> octreostatin display<br />

high affinity binding only <strong>for</strong> SSTR-2 <strong>and</strong><br />

SSTR-5 receptors, intermediate binding affinity<br />

<strong>for</strong> SSTR-3 subtype <strong>and</strong> very low or no affinity<br />

<strong>for</strong> SSTR-1 <strong>and</strong> SSTR-4 subtypes. The five<br />

somatostatin receptor subtypes cloned to date<br />

show very poor homology (40–50%) which<br />

may account, at least in part <strong>for</strong> the different<br />

binding affinities [23]. Although a number of<br />

studies have been per<strong>for</strong>med over the past<br />

years to determine the distribution of the five<br />

somatostatin receptor subtypes, their physiological<br />

role is still poorly understood. This is<br />

partly due to the fact that there is no set of agonists<br />

<strong>and</strong> antagonists available with sufficient<br />

receptor selectivity <strong>for</strong> a single receptor subtype<br />

<strong>and</strong> because of coexpression of two or<br />

more SRIF receptor subtypes in a particular<br />

organ. All five somatostatin receptors are functionally<br />

coupled to inhibition of adenylyl cyclase.<br />

Likewise, the five subtypes induce phosphotyrosine<br />

phosphatase-1c in each case via<br />

pertussis toxin-sensitive GTP binding proteins.<br />

Some of the receptor subtypes are also<br />

coupled to K + <strong>and</strong> Ca 2+ channels, the Na 2+ /K +<br />

hydrogen exchanges <strong>and</strong> to phospholipase C<br />

<strong>and</strong> MAP kinases. Neuroendocrine tumors express<br />

a variety of somatostatin receptors but<br />

particularly receptor subtype 2 [22, 24, 25].<br />

Thus the SSTR-2 receptor subtype is expressed<br />

in 90% of carcinoid tumors <strong>and</strong> in 80% of<br />

endocrine pancreatic tumors. An exception are<br />

insulin-producing pancreatic tumors where<br />

less than 50% express the somatostatin receptor<br />

2 subtype. However, neuroendocrine tumors<br />

also express SSTR-5, SSTR-3 <strong>and</strong> SSTR-<br />

1 subtypes. Further, it is demonstrated that different<br />

somatostatin receptor subtypes are expressed<br />

in various patterns on different tumors<br />

[22]. This demonstration may explain the conflicting<br />

results of the therapeutic efficacy of<br />

reported somatostatin analogs in the management<br />

of neuroendocrine tumors. Activation of<br />

somatostatin receptor subtypes 1 <strong>and</strong> 2 results<br />

Chemotherapy 2001;47(suppl 2):40–53 43


in the activation of tyrosine phosphatase, the<br />

activity of which has been correlated with the<br />

antimitotic effect of somatostatin <strong>and</strong> its analogs<br />

in some types of cells [26]. In addition,<br />

inhibition of cell proliferation is mediated by<br />

the somatostatin receptor subtype 5, but via a<br />

different mechanism, which probably involves<br />

changes in the intracellular calcium<br />

fluxes [23, 27]. <strong>Somatostatin</strong> receptor type 3<br />

(SSTR-3) has been reported to mediate apoptosis.<br />

Thus, high dose treatment with octreotide,<br />

which only has a low binding affinity <strong>for</strong><br />

the SSTR-3 subtype induces apoptosis both in<br />

carcinoid tumors <strong>and</strong> BON-1 cells xenotransplanted<br />

to nude mice [28]. Octreotide <strong>and</strong><br />

other somatostatin analogs significantly inhibit<br />

the growth of the neuroendocrine-differentiated<br />

cell line BON-1, the pancreatic cancer<br />

cell line AR42J <strong>and</strong> the human breast cancer<br />

line MCF-7 in experimental animals [28,<br />

29]. Tumors treated with somatostatin analogs<br />

are less vascular than those in untreated<br />

mice suggesting that the peptide inhibits angiogenesis.<br />

This suggestion is supported in a<br />

chorioalantoic membrane model of the chick<br />

embryo where octreotide inhibits angiogenesis<br />

in a dose-related manner [26]. Furthermore,<br />

octreotide in combination with endothelial<br />

growth factor inhibits blood vessel<br />

growth. Reubi et al. [30] demonstrated a high<br />

concentration of somatostatin receptors in the<br />

veins draining some human tumors. However,<br />

the expression of somatostatin receptors<br />

in the veins was not different from that in the<br />

tumour per se.<br />

Weckbecker et al. [29] studied the effects<br />

of octreotide in combination with doxorubicin,<br />

mitomycin C <strong>and</strong> Taxol or 5-FU on the<br />

growth of the pancreatic cancer cell line<br />

AR42J in vitro. Octreotide potentiated the<br />

antiproliferative effect of three of four cytotoxic<br />

agents (mitomycin C, doxorubicin <strong>and</strong><br />

taxol) <strong>and</strong> was additive to 5-FU. The authors<br />

also investigated the temporal effects of a<br />

44 Chemotherapy 2001;47(suppl 2):40–53 Öberg<br />

combination of octreotide <strong>and</strong> doxorubicin.<br />

They observed that the antiproliferative effect<br />

was greater if the doxorubicin therapy was<br />

administered initially with the octreotide<br />

therapy added 24 h later. The in vitro data<br />

was further confirmed in vivo using nude<br />

mice bearing AR42J tumors.<br />

<strong>Somatostatin</strong> Receptor Scintigraphy<br />

<strong>Somatostatin</strong> receptor scintigraphy is the<br />

most important clinical diagnostic investigation<br />

<strong>for</strong> patients with suspected neuroendocrine<br />

tumors [31, 32]. Dynamic scintigraphy<br />

detected more than 90% of all patients with<br />

carcinoid tumors, particularly liver metastases,<br />

regional lymph node metastases, bone<br />

metastases <strong>and</strong> sometimes also the primary<br />

tumor. In patients with endocrine pancreatic<br />

tumors, the clinical use of scintigraphy in the<br />

detection of such tumors has not been as fruitful<br />

as in carcinoid tumors. However, in a<br />

recent study 122 patients with a diagnosis of<br />

the Zollinger-Ellison syndrome were evaluated<br />

with upper gastrointestinal endoscopy,<br />

computerized tomography (CT), magnetic<br />

resonance imaging (MRI), angiography <strong>and</strong><br />

bone scanning, to attempt to design a treatment<br />

plan [33]. Patients then underwent somatostatin<br />

receptor scintigraphy with 111 In-<br />

DTPA-D-Phe-octreotide. The somatostatin<br />

scintigraphy (62%) was superior to both CT<br />

(39%) <strong>and</strong> MRI (50%) in localizing primary<br />

tumors <strong>and</strong> was equal to all conventional<br />

imaging studies in combination. For metastatic<br />

disease scintigraphy (100%) was again<br />

superior to CT <strong>and</strong> MRI (64 <strong>and</strong> 80%, respectively)<br />

<strong>and</strong> equal to all conventional studies<br />

combined (96%). In 14% of the tumors, scintigraphy<br />

was the only technique which was<br />

positive. The scintigraphic results changed<br />

the clinical management in 47% of patients<br />

with gastrinoma. There<strong>for</strong>e in patients with


Table 1. Meta-analysis of<br />

symptomatic (subjective),<br />

biochemical <strong>and</strong> radiological<br />

responses to different treatments<br />

with somatostatin analogs in<br />

studies conducted over the last 10<br />

years in patients with metastatic<br />

neuroendocrine tumors<br />

the Zollinger-Ellison syndrome the investigators<br />

suggested that somatostatin receptor scintigraphy<br />

should be regarded as the imaging<br />

modality of choice. In a prospective study<br />

evaluating somatostatin receptor imaging in<br />

160 patients with confirmed gastroenteropancreatic<br />

tumors, somatostatin receptor scintigraphy<br />

was positive in 68% of patients. More<br />

interesting was the observation that in 46 patients,<br />

negative by conventional imaging, somatostatin<br />

receptor scintigraphy detected 47<br />

previously undiagnosed lesions in 36 patients.<br />

Furthermore, somatostatin receptor scintigraphy<br />

modified the tumor staging in 38 patients<br />

<strong>and</strong> changed the surgical strategy in 40 patients<br />

[34]. On the basis of these findings,<br />

somatostatin receptor scintigraphy should be<br />

per<strong>for</strong>med routinely in patients with neuroendocrine<br />

tumors both <strong>for</strong> the staging <strong>and</strong> <strong>for</strong><br />

therapeutic decisions. Small tumors, less than<br />

1 cm in diameter, such as in patients with<br />

MEN-1 still represent a diagnostic problem.<br />

Furthermore, in all primary tumors or metastases,<br />

even in the same patient, the ex-<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

Response St<strong>and</strong>ard doses<br />

of octreotide<br />

(100–1,500 Ìg/day)<br />

Slow release<br />

lanreotide<br />

(30 mg/14 day i.m.)<br />

High dose<br />

lanreotide<br />

(9–15 mg/day)<br />

Symptomatic<br />

Biochemical<br />

146/228 (64)34/66 (52) 11/26 (42)<br />

CR 6/54 (11)2/80 (2.5)1/33 (3)<br />

PR 116/211 (55)35/80 (44) 24/33 (72)<br />

SD NS 32/80 (40)7/33 (21)<br />

PD<br />

Radiological<br />

NS 11/80 (13.5)1/33 (3)<br />

CR – – 1/53 (2)<br />

PR 7/131 (5)2/42 (5) 6/53 (11)<br />

SD 50/131 (38)32/42 (76) 25/53 (47)<br />

PD 74/131 (56)8/42 (19) 21/53 (39)<br />

Figures represent numbers with the percentage in parentheses. CR =<br />

Complete response; PR = partial response; SD = stable disease; PD = progressive<br />

disease; NS = not indicated.<br />

pression of somatostatin receptor subtypes is<br />

not identical <strong>and</strong> they may not express the<br />

SSTR-2 or SSTR-5 subtypes. Since octreoscans<br />

are dependent as the expression of the<br />

SSTR-2 <strong>and</strong> SSTR-5 subtypes, positive somatostatin<br />

receptor scintigraphy has a predictive<br />

value <strong>for</strong> treatment with somatostatin analogs<br />

[35, 36]. Conversely, tumors not expressing<br />

the SSTR-2 <strong>and</strong> SSTR-5 subtypes will not be<br />

detectable by scintigraphy <strong>and</strong> are unlikely to<br />

respond to somatostatin analog therapy.<br />

<strong>Treatment</strong> with <strong>Somatostatin</strong><br />

<strong>Analogs</strong> (table 1)<br />

<strong>Somatostatin</strong> analogs, octreotide being the<br />

first available <strong>for</strong> clinical use [37], have become<br />

increasingly important in the management<br />

of patients with neuroendocrine tumors<br />

of the gut <strong>and</strong> pancreas. The rationale <strong>for</strong> the<br />

clinical efficacy of somatostatin analogs in<br />

the management of neuroendocrine tumors<br />

is related to the expression of somatostatin<br />

Chemotherapy 2001;47(suppl 2):40–53 45


eceptors in 80–90% of all such cases. All of<br />

the somatostatin analogs presently available<br />

<strong>for</strong> clinical use (octreotide, lanreotide, octastatin)<br />

bind with high affinity to the SSTR-2<br />

<strong>and</strong> SSTR-5 subtypes <strong>and</strong> with lower affinity<br />

to SSTR-3 subtype. Although somatostatin<br />

analogs have been available <strong>for</strong> more than 15<br />

years, we are still lacking r<strong>and</strong>omized controlled<br />

trials, probably because neuroendocrine<br />

tumors represent rare diseases. Furthermore,<br />

the inclusion of different types of neuroendocrine<br />

tumors in multicenter trials is<br />

not appropriate <strong>for</strong> studying the efficacy of<br />

somatostatin analogs on these indications<br />

since there are large differences between ‘classical’<br />

midgut carcinoids with a low proliferative<br />

capacity <strong>and</strong> several <strong>for</strong>ms of endocrine<br />

pancreatic tumors or other <strong>for</strong>egut tumors<br />

which progress rapidly <strong>and</strong> show a high proliferative<br />

capacity.<br />

The efficacy of somatostatin analog therapy<br />

could be divided into subjective, biochemical<br />

<strong>and</strong> tumor responses (tumor size reduction).<br />

A complete remission is rarely obtained<br />

but partial remission is more than 50% reduction<br />

of clinical symptoms, biochemical markers<br />

or tumor size; stable disease is less than<br />

50% reduction of these parameters but also<br />

less than 25% increase of the same parameters.<br />

Progressive disease is more than 25%<br />

increase of clinical symptoms, biochemical<br />

markers or tumor size.<br />

In the first trial reported by Kvols et al.<br />

[37], octreotide subcutaneously (150 Ìg t.i.d.)<br />

was observed to present symptomatic responses<br />

in 88% <strong>and</strong> biochemical responses in<br />

72% of patients with carcinoid tumors. The<br />

median duration of the biochemical response<br />

was 12 months. In 1989 Gorden et al. [38]<br />

per<strong>for</strong>med a meta-analysis of all reported<br />

cases of neuroendocrine tumors treated with<br />

somatostatin analogs. The meta-analysis indicated<br />

symptomatic improvement in 92%<br />

<strong>and</strong> a biochemical response in 66% of the<br />

46 Chemotherapy 2001;47(suppl 2):40–53 Öberg<br />

patients. A reduction of the tumor size was,<br />

however, only noted in 8% of octreotidetreated<br />

patients, whereas tumor size was unchanged<br />

in 85%. In that study tachyphylaxis<br />

was observed in 40% of the patients but a proportion<br />

did respond to increased doses of<br />

octreotide. The median dose of octreotide<br />

used in these studies was 300 Ìg/day s.c. Low<br />

dose octreotide therapy (50 Ìg b.i.d.) resulted<br />

in significantly lower biochemical response<br />

rates (30%) compared to the regular dose analog<br />

therapy [39].<br />

More than 50 patients with gastrinomas<br />

have been treated with doses of 100–1,500 Ìg<br />

octreotide/day, most of them in the short term<br />

[40, 41]. A clinical response defined as control<br />

of gastric hypersecretion, pain <strong>and</strong> diarrhea<br />

was observed in 90% of the patients <strong>and</strong> was<br />

accompanied by a significant reduction in<br />

serum gastrin <strong>and</strong> basal acid secretion. In a<br />

long-term study by Ruszniewski et al. [42], the<br />

maximal acid output decreased during 9–12<br />

months of octreotide treatment suggesting an<br />

antitrophic effect of the analog (i.e. reduction<br />

of parietal cell mass). Such an effect has also<br />

been reported by an Italian group who demonstrated<br />

that octreotide, 500 Ìg once a day,<br />

elicited a significant decrease in its ECL cell<br />

population, which was paralleled by progressive<br />

reduction in serum gastrin levels in patients<br />

with chronic atrophic gastritis [43].<br />

However, although somatostatin analog therapy<br />

is not the primary treatment <strong>for</strong> gastrinomas,<br />

many patients being initially treated<br />

with H2 receptor blockers or proton pump<br />

inhibitors combined with surgery, long-acting<br />

somatostatin analogs (S<strong>and</strong>ostatin-LAR ® ,<br />

Lanreotide PR ® ) may be beneficial <strong>for</strong> a subgroup<br />

of patients with malignant gastrinomas.<br />

Patients with insulin-producing tumors<br />

treated with somatostatin analogs should be<br />

very carefully monitored <strong>for</strong> escalation of<br />

their hypoglycemia. In some patients counterregulatory<br />

mechanisms, such as the growth


hormone, IGF-1 <strong>and</strong> glucagon may be more<br />

suppressed than the insulin secretion by the<br />

tumor [41, 44, 45]. About 50% of insulin-producing<br />

tumors do not express somatostatin<br />

receptor 2 <strong>and</strong> 5 subtypes. However, there is a<br />

subgroup of insulin-producing tumors that<br />

might benefit from somatostatin analog therapy<br />

<strong>and</strong> these are the predominantly malignant<br />

insulinomas which concomitantly hypersecrete<br />

gastrin <strong>and</strong>/or glucagon. In patients with<br />

the WDHA syndrome <strong>and</strong> VIP-producing tumors<br />

which are unoperable <strong>and</strong> in whom chemotherapy<br />

has only been transiently effective,<br />

octreotide has been suggested to be the treatment<br />

of choice [41, 45, 46]. Symptomatic<br />

improvement has been reported in more than<br />

80% of such patients treated with octreotide<br />

at doses of 100–400 Ìg/day. However, in<br />

some patients the beneficial effect of octreotide<br />

lasted only a few days requiring progressive<br />

increases in doses. Biochemical responses<br />

have been reported in about 80% of VIPoma<br />

patients. Symptomatic relief was not always<br />

related to a reduction of plasma concentration<br />

of VIP suggesting that octreotide may have a<br />

direct effect on the gut. In some patients<br />

octreotide has also been shown to change the<br />

molecular <strong>for</strong>ms of circulating VIP, possibly<br />

into less bioactive <strong>for</strong>ms. In approximately<br />

90% of patients with glucagonomas who<br />

present with a rash it disappears rapidly after<br />

octreotide therapy [41, 47]. Other symptoms<br />

characteristic of the syndrome such as weight<br />

loss <strong>and</strong> diarrhea may improve during octreotide<br />

therapy but the analog has a variable<br />

effect on diabetes. Plasma glucagon is reduced<br />

in approximately 60% of patients during octreotide<br />

therapy. The symptomatic response<br />

observed during octreotide therapy was independent<br />

of the plasma glucagon, the concentration<br />

suggesting a direct effect of the analog<br />

on the skin. Furthermore, in patients with glucagonomas,<br />

octreotide can change the circulating<br />

molecular <strong>for</strong>ms of glucagon, indicating<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

that the analogs inhibit the posttranslational<br />

processing of preproglucagon, thereby reducing<br />

the circulating levels of bioactive glucagon.<br />

<strong>Somatostatin</strong>-producing tumors, the socalled<br />

somatostatinomas, have been regarded<br />

as resistant to treatment with somatostatin<br />

analogs. However, a symptomatic <strong>and</strong> biochemical<br />

response has been reported in a<br />

small number of patients during octreotide<br />

therapy which correlated with the presence of<br />

somatostatin receptor subtypes in the tumors<br />

[48].<br />

The inhibition of tumor growth in patients<br />

with carcinoid <strong>and</strong> endocrine pancreatic tumors<br />

has been reported to be low in most<br />

studies (0–17%) [40, 41, 49]. In a study by<br />

Saltz et al. [50] in patients with neuroendocrine<br />

tumors treated with octreotide (150–<br />

250 Ìg t.i.d.), no regression was documented.<br />

However, octreotide stabilized the size assessed<br />

by CT in 50% of the patients <strong>for</strong> a<br />

median duration of 5 months. In a German<br />

multicenter trial, 52 patients with different<br />

<strong>for</strong>ms of neuroendocrine malignancies <strong>and</strong><br />

CT-documented tumor progression were<br />

treated with octreotide 200 Ìg t.i.d. [51]. Stabilization<br />

of tumor growth was achieved in 19<br />

of 52 patients (36%) <strong>for</strong> a median duration of<br />

18 months. Even though a reduction in tumor<br />

size is rarely seen with st<strong>and</strong>ard octreotide<br />

treatment, stabilization of further tumor<br />

growth suggests that octreotide has an antiproliferative<br />

effect.<br />

Harris <strong>and</strong> Redfern [49] per<strong>for</strong>med a<br />

meta-analysis of data compiled from 62 published<br />

studies on patients with carcinoids to<br />

examine the relationship between the dose of<br />

octreotide <strong>and</strong> the clinical efficacy evaluated<br />

by analyzing urinary 5-HIAA, flushing <strong>and</strong><br />

diarrhea in patients. Six dose ranges of octreotide<br />

were assessed <strong>and</strong> ranged from 100<br />

to 3,000 Ìg/day. The maximum clinical response<br />

to octreotide occurred in patients<br />

treated with 300–375 Ìg/day with some fur-<br />

Chemotherapy 2001;47(suppl 2):40–53 47


ther improvement in doses up to 1,000 Ìg/<br />

day. Doses of octreotide above 10,000 Ìg<br />

showed no additional clinical benefit. The authors’<br />

conclusion from this study is that there<br />

was a significant patient-to-patient variation<br />

in the sensitivity to octreotide treatment <strong>and</strong><br />

that it is important to titrate the dose of<br />

the analog in each patient until adequate<br />

symptoms <strong>and</strong>/or biochemical control are<br />

achieved.<br />

High-Dose <strong>Somatostatin</strong> Analog<br />

Therapy<br />

A few studies have addressed the potential<br />

value of high-dose somatostatin analog treatment<br />

in neuroendocrine gastrointestinal tumors<br />

[52–56]. A dose-related tumor response<br />

has been demonstrated in a variety of tumor<br />

models with increasing somatostatin analog<br />

treatment. In a trial of our group [53], 19<br />

patients with advanced neuroendocrine gastrointestinal<br />

tumors (13 carcinoids <strong>and</strong> 6<br />

EPT) were included in the study. The mean<br />

duration of disease prior to entry into the trial<br />

was 56 months <strong>and</strong> all except 4 patients had<br />

received a variety of treatments. Finally patients<br />

had failing st<strong>and</strong>ard dose octreotide<br />

therapy. <strong>Somatostatin</strong> receptor scintigraphy<br />

was positive in 17 of 18 patients be<strong>for</strong>e initiation<br />

of lanreotide administered in increasing<br />

doses up to 12 mg/day s.c., a dose which was<br />

maintained <strong>for</strong> 1 year or until tumor progression.<br />

High-dose lanreotide resulted in biochemical<br />

responses in 58% of the patients,<br />

stabilization of the disease was observed in<br />

70% <strong>and</strong> 1 patient (5%) showed a partial<br />

tumor response. Furthermore, patients with<br />

both a biochemical response <strong>and</strong> stabilization<br />

of their disease exhibited a progressive increase<br />

in the number of apoptotic cells in the<br />

tumor, maximal apoptosis occurring 12<br />

months following commencement of lanreo-<br />

48 Chemotherapy 2001;47(suppl 2):40–53 Öberg<br />

tide therapy [28, 53]. Apoptosis may have<br />

been mediated via the binding of lanreotide to<br />

somatostatin receptor type 3, as suggested by<br />

Patel <strong>and</strong> Srikant [23]. Moreover, in this trial<br />

positron emission tomography using the tracer<br />

11 C-L-DOPA lanreotide inhibited exocytosis<br />

more strongly than the inhibition of its<br />

synthesis of the peptide [57]. Anthony et al.<br />

[54] treated 13 patients with neuroendocrine<br />

tumors, refractory to st<strong>and</strong>ard doses of octreotide,<br />

with a high dose of the analog<br />

(6 mg/day) <strong>and</strong> reported a partial tumor response<br />

in 4 patients (31%) <strong>and</strong> stabilization of<br />

the disease in 2 (15%). The same group reported<br />

on high-dose lanreotide (9 mg/day) in<br />

13 patients with various neuroendocrine tumors<br />

finding a partial remission in 4 patients<br />

(31%) <strong>and</strong> stabilization of the disease in 2<br />

(15%). Faiss <strong>and</strong> Wiedenmann [55] treated 30<br />

patients with metastatic neuroendocrine gastrointestinal<br />

tumors with 15 mg/day of lanreotide<br />

<strong>for</strong> 1 year <strong>and</strong> reported one complete<br />

<strong>and</strong> one partial tumor response. These studies<br />

suggest that a high-dose therapy of octreotide/<br />

lanreotide can produce additional antiproliferative<br />

effects in patients deteriorating on<br />

regular dose analog therapy. These observations<br />

<strong>for</strong>m the basis <strong>for</strong> a currently ongoing<br />

study with ultrahigh doses (160 mg/month) of<br />

octreotide (Onco-LAR ® ).<br />

Continuous Infusion of <strong>Somatostatin</strong><br />

<strong>Analogs</strong><br />

Several studies of patients with acromegaly<br />

have indicated that a continuous infusion of<br />

octreotide has advantages over intermittent<br />

subcutaneous injections of the analog [58]. A<br />

more pronounced clinical <strong>and</strong> biochemical<br />

control can be achieved at lower intravenous<br />

doses of somatostatin analogs <strong>and</strong> the adverse<br />

effects may be less than with higher subcutaneous<br />

doses. To test this hypothesis we


per<strong>for</strong>med a European multicenter trial in 35<br />

patients with carcinoid syndrome (19 of them<br />

had failed st<strong>and</strong>ard doses of octreotide) with<br />

octastatin (RC-160) at a dose of 1.5 mg/day<br />

given as a continuous subcutaneous infusion<br />

via micropump <strong>for</strong> 3–6 months [56]. This was<br />

the first clinical trial of octastatin of particular<br />

interest since in vitro studies have suggested<br />

that octastatin has a stronger antiproliferative<br />

effect than both octreotide <strong>and</strong> lanreotide [58,<br />

59]. In this trial subjective improvement <strong>and</strong><br />

disease stabilization were observed in 60% of<br />

the patients. However, the biochemical response<br />

rate was rather low (23%) <strong>and</strong> there<br />

was no tumor response.<br />

Slow Release Formulations of<br />

<strong>Somatostatin</strong> <strong>Analogs</strong><br />

One of the most important improvements<br />

in somatostatin analog treatment is the development<br />

of a slow release <strong>for</strong>mulation. S<strong>and</strong>ostatin-LAR<br />

<strong>and</strong> Lanreotide-PR in which octreotide<br />

<strong>and</strong> lanreotide have been incorporated<br />

into microspheres of a biodegradable<br />

polymer were available <strong>for</strong> clinical use. S<strong>and</strong>ostatin-LAR<br />

can be administered once every<br />

4 weeks <strong>and</strong> Lanreotide-PR every 2 weeks<br />

[60]. Ruszniewski et al. [61] treated 39 patients<br />

with carcinoids with Lanreotide-PR (30<br />

mg i.m. every 2 weeks) <strong>and</strong> reported subjective<br />

responses in approximately 55% of patients,<br />

biochemical responses in 42%. However,<br />

no tumor response was observed after<br />

6 months of lanreotide treatment. We conducted<br />

a European multicenter trial <strong>and</strong> included<br />

55 patients (48 carcinoids <strong>and</strong> 7 EPT)<br />

with Lanreotide-PR (30 mg i.m.) every 2<br />

weeks <strong>for</strong> 6 months [62]. Symptomatic improvement<br />

was observed in 38% of carcinoids,<br />

66% of gastrinomas <strong>and</strong> one VIPoma.<br />

Biochemical responses were obtained in 47%<br />

<strong>and</strong> tumor responses in 2 patients (7%). Sta-<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

bilization of tumor growth was achieved in<br />

80% of patients. In this trial quality of life was<br />

studied using a validated instrument (QLC<br />

30) <strong>and</strong> assessment after 1 month showed a<br />

significant improvement of emotional <strong>and</strong><br />

cognitive function, overall health as well as<br />

sleeping disorders <strong>and</strong> diarrhea. This was the<br />

first trial to demonstrate that long-acting somatostatin<br />

analog therapy improves the quality<br />

of life of patients with neuroendocrine<br />

tumors.<br />

Combination Trial with <strong>Somatostatin</strong><br />

<strong>Analogs</strong><br />

The combination of somatostatin analogs<br />

with other agents is an interesting area <strong>for</strong><br />

future studies. We have previously reported<br />

that a combination of octreotide <strong>and</strong> interferon-·<br />

produced biochemical responses in 75%<br />

of patients resistant to either interferon-·<br />

alone or conventional doses of octreotide [63].<br />

In vitro <strong>and</strong> in vivo studies of BON-1 tumors<br />

indicate that a combination of these two compounds<br />

has a stronger antiproliferative effect<br />

than interferon or octreotide alone [28]. An<br />

Italian trial reported on 58 patients with metastatic<br />

neuroendocrine tumors who were initially<br />

treated with octreotide at a dose of<br />

500 Ìg t.i.d. [52]. The dose was increased to<br />

1,000 Ìg t.i.d. <strong>and</strong> this regimen was continued<br />

until disease progression. Twenty-five patients<br />

with progressive disease during octreotide<br />

treatment received concomitant chemotherapy<br />

(dacarbazine 200 mg/m 2 , 5-FU 250<br />

mg/m 2 , epidoxorubicin 25 mg/m 2 ) administered<br />

intravenously daily <strong>for</strong> 3 days every 3<br />

weeks. For the whole group of 58 patients the<br />

median duration of octreotide treatment was<br />

5 months. In 27 patients the disease stabilized<br />

<strong>for</strong> at least 6 months. A partial remission was<br />

achieved in 2 patients which lasted 10 <strong>and</strong> 14<br />

months, respectively. The median survival of<br />

Chemotherapy 2001;47(suppl 2):40–53 49


the entire group was 22 months. In this trial<br />

no additive or synergistic effect was obtained<br />

by adding chemotherapy. There<strong>for</strong>e, the interesting<br />

observation by Weckbecker et al.<br />

[29] that doxorubicin in combination with<br />

octreotide showed a significant synergistic effect<br />

in in vitro <strong>and</strong> in vivo studies remains to<br />

be proven in <strong>for</strong>thcoming r<strong>and</strong>omized clinical<br />

trials.<br />

Tumor-Targeted Radioactive<br />

<strong>Somatostatin</strong> Analog <strong>Treatment</strong><br />

Peptide receptor scintigraphy with a radioactive<br />

somatostatin analogue ( 111 In-<br />

DTPA-octreotide) is a sensitive <strong>and</strong> specific<br />

technique to identify in vivo the presence <strong>and</strong><br />

abundance of somatostatin receptors in various<br />

tumors. The method has now been accepted<br />

as an important tool <strong>for</strong> the staging<br />

<strong>and</strong> localization of neuroendocrine tumors.<br />

However, the technique is currently being<br />

evaluated as a possible means of targeting<br />

neuroendocrine tumors with ‘tumor-targeted’<br />

radiotherapy using a repeated administration<br />

of high doses of 111 In-DTPA-octreotide. 111 In<br />

emits Auger electrons having a tissue penetration<br />

of 0.02–10 Ìm. In a recent trial 30 endstage<br />

patients with progressive neuroendocrine<br />

tumors were treated with 111 In-DTPAoctreotide<br />

up to a maximal cumulative dose<br />

of 74 GBq in a phase 1 trial. No major clinical<br />

side effects were observed after up to 2 years’<br />

treatment, except that in a few patients there<br />

was a transient decline in platelets <strong>and</strong> lymphocytes.<br />

Of the 21 patients who received a<br />

cumulative dose of more than 20 GBq, 8<br />

patients showed stabilization of disease <strong>and</strong><br />

there was a reduction in tumor size in a further<br />

6. Furthermore, there was a tendency<br />

towards better results in patients whose<br />

tumor cells showed higher accumulation of<br />

the radiolig<strong>and</strong>. We treated 16 patients, also<br />

50 Chemotherapy 2001;47(suppl 2):40–53 Öberg<br />

end-stage neuroendocrine tumor patients, at<br />

doses up to 60 GBq <strong>and</strong> observed 30% response<br />

rates <strong>and</strong> also a tendency towards better<br />

results with higher tracer accumulation of<br />

the radiolig<strong>and</strong>. Theoretically depending on<br />

the homogeneity of the distribution of tumor<br />

cells expressing somatostatin receptor subtypes<br />

<strong>and</strong> the size of tumors, ß-emitting radionucleids,<br />

e.g. 90 Y labelled to DOTA octreotide,<br />

may be more effective than 111 In targeting<br />

radionuclear therapy. Such trials are now<br />

in progress. Recently a study of 10 patients<br />

treated with 90 Y DOTA TOC has been reported<br />

[15]. Two patients showed a significant<br />

antitumor response <strong>and</strong> another 2 developed<br />

stable disease. In 9 of the 10 patients<br />

treated renal <strong>and</strong> bone marrow toxicity did<br />

not exceed grade 1. One patient developed a<br />

persisting grade 2 thrombocytopenia at a total<br />

dose of 180 mCi. Tumor-targeted radioactive<br />

somatostatin analog therapy awaits further<br />

evaluation in patients with less advanced disease.<br />

Adverse Reaction to <strong>Somatostatin</strong><br />

Analog Therapy<br />

The most common side effects of somatostatin<br />

analog are generally mild <strong>and</strong> include<br />

nausea, transient abdominal cramps, flatulence,<br />

diarrhea <strong>and</strong> local reaction at the injection<br />

site [58]. Most of these minor side effects<br />

resolve with time. In 20–50% of patients gall<br />

stones are <strong>for</strong>med de novo, but these remain<br />

virtually always asymptomatic [64]. Rare,<br />

more severe adverse events of octreotide therapy<br />

include hypocalcemia, bradycardia, acute<br />

pancreatitis, hepatitis, jaundice, transitory,<br />

ischemic attacks, <strong>and</strong> a negative inotropic<br />

effect of the analogs.


Future Aspects<br />

<strong>Somatostatin</strong> analogs can relieve symptoms,<br />

reduce circulating hormone levels <strong>and</strong><br />

stabilize tumor growth in more than 50% of<br />

patients, which makes them a good therapy<br />

<strong>for</strong> patients with neuroendocrine tumors.<br />

There is, however, as yet no published study<br />

that shows that therapy with somatostatin<br />

analogs improves survival. However, the majority<br />

of centers working with patients with<br />

neuroendocrine tumors use a multimodal<br />

therapeutic approach. Thus, it is very unlikely<br />

that a patient with a neuroendocrine tumor<br />

will receive a somatostatin analog as the<br />

only treatment during the clinical course of<br />

the disease. There is no doubt that somatostatin<br />

analog therapy has significantly improved<br />

the quality of life in patients with<br />

malignant neuroendocrine tumors <strong>and</strong> is<br />

very important <strong>for</strong> palliative treatment. The<br />

most promising future areas are the clinical<br />

usefulness of slow release <strong>for</strong>mulations, possibly<br />

also high-dose, slow release <strong>for</strong>mulations<br />

(Onco-LAR) <strong>and</strong> the ‘tumor targeted’<br />

References<br />

1 Öberg K: Neuroendocrine tumors.<br />

Ann Oncol 1996;7:453–463.<br />

2 Norheim I, Öberg K, Theodorson-<br />

Norheim E, et al: Malignant carcinoid<br />

tumors. An analysis of 103 patients<br />

with regard to tumor localization,<br />

hormone production <strong>and</strong> survival.<br />

Ann Surg 1987;206:115–125.<br />

3 Feldman JM: Carcinoid tumors <strong>and</strong><br />

syndrome. Semin Oncol 1987;14:<br />

237.<br />

4 Eriksson B, Arnberg H, Lindgren<br />

PG, et al: Neuroendocrine pancreatic<br />

tumors: Clinical presentation,<br />

biochemical <strong>and</strong> histopathological<br />

findings in 84 patients. J Intern Med<br />

1990;228:103–113.<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

5 Solcia E, Capella C, Fiocca R, et al:<br />

The gastroenteropancreatic endocrine<br />

system <strong>and</strong> related tumors.<br />

Gastroenterol Clin North Am 1989;<br />

18:671–693.<br />

6 Jensen RT: Zollinger-Ellison syndrome:<br />

Current concepts <strong>and</strong> management.<br />

Ann Intern Med 1983;98:<br />

159–175.<br />

7 Fajan SS, Vinik AI: Insulin-producing<br />

islet cell tumors. Endocrinol Metab<br />

Clin North Am 1989;18:45.<br />

8 Long RG, Bryant MG, Mitchell SJ,<br />

et al: Clinicopathological study of<br />

pancreatic <strong>and</strong> ganglioneuroblastoma<br />

tumors secreting vasoactive intestinal<br />

polypeptide (VIPomas). Br<br />

Med J 1981;282:1767.<br />

radioactive octreotide therapy. The combination<br />

of somatostatin analogs with interferons<br />

<strong>and</strong> cytotoxic agents in clinical trials is also<br />

worthy of investigation. Furthermore, since<br />

somatostatin analogs can be coupled to chemotherapeutic<br />

drugs, internalization of the<br />

receptor/lig<strong>and</strong> complex will deliver the chemotherapeutic<br />

agent directly into the tumor<br />

cells. <strong>Somatostatin</strong> receptor subtype 5 is the<br />

receptor type which is more effectively internalized<br />

<strong>and</strong> receptor subtype 5-specific analogs<br />

alone or coupled with cytotoxic drugs or<br />

radioactivity might be of therapeutic advantage.<br />

This subtype of the somatostatin receptor<br />

is expressed in pituitary <strong>and</strong> neuroendocrine<br />

tumor cells. Combination treatments<br />

with different receptor subtype-specific analogs<br />

(cocktails) might be of benefit [65] when<br />

the determination of somatostatin receptor<br />

expression is routinely examined in all tumors.<br />

Using this concept it may be possible<br />

to design specific combinations of somatostatin<br />

analogs to treat a particular tumor.<br />

9 Stacpoole PW: The glucagonoma<br />

syndrome; clinical features, diagnosis<br />

<strong>and</strong> treatment. Endocr Rev 1981;<br />

2:347.<br />

10 Skogseid B, Eriksson B, Lundqvist<br />

G, et al: Multiple endocrine neoplasia<br />

type 1: A ten year prospective<br />

screening study in four kindreds. J<br />

Clin Endocrinol Metab 1991;73:<br />

281–287.<br />

11 Makridis C, Öberg K, Juhlin C, et al:<br />

Surgical treatment of midgut carcinoid<br />

tumors. World J Surg 1990;14:<br />

377.<br />

12 Carrasco CH, Chuang V, Wallace S:<br />

Apudoma metastatic to the liver:<br />

<strong>Treatment</strong> by hepatic artery embolization.<br />

Radiology 1983;149:79–83.<br />

Chemotherapy 2001;47(suppl 2):40–53 51


13 Schupak KP, Wallner KE: The role<br />

of radiation therapy in the treatment<br />

of locally unresectable or metastatic<br />

carcinoid tumors. Int J Radiat Oncol<br />

Biol Phys 1991;20:489.<br />

14 Hoefnagel CA, den Hartog Jager<br />

FC, Taal BG, et al: The role of 125 I-<br />

MIBG in the diagnosis <strong>and</strong> therapy<br />

of carcinoids. Eur J Nucl Med 1987;<br />

13:187.<br />

15 Olte A, Mueller-Br<strong>and</strong> J, Dellas S, et<br />

al: Yttrium–90-labelled somatostatin<br />

analoague <strong>for</strong> cancer treatment.<br />

Lancet 1998;351:417–418.<br />

16 Moertel CG, Johnson CM, McKusick<br />

MA, et al: The management of<br />

patients with advanced carcinoid tumours<br />

<strong>and</strong> islet cell carcinomas.<br />

Ann Intern Med 1994;120:302–<br />

309.<br />

17 Eriksson B, Öberg K: An update of<br />

the medical treatment of malignant<br />

endocrine pancreatic tumors. Acta<br />

Oncol 1993;32:203–208.<br />

18 Öberg K, Norheim I, Lundqvist G,<br />

Wide L: Cytotoxic treatment in patients<br />

with malignant carcinoid tumours;<br />

response to streptozocin –<br />

alone or in combination with 5-FU.<br />

Acta Oncol 1987;26:429–432.<br />

19 Öberg K, Norheim I, Lind E, et al:<br />

<strong>Treatment</strong> of malignant carcinoid<br />

tumors with human leukocyte interferon:<br />

Long-term results. <strong>Cancer</strong><br />

Treat Rep 1986;70:1297–1304.<br />

20 Öberg K, Eriksson B: The role of<br />

interferons in the management of<br />

carcinoid tumors. Acta Oncol 1991;<br />

30:519–522.<br />

21 Yamada Y, Reisine T, Law SF, et al:<br />

Cloning <strong>and</strong> functional characterization<br />

of a family of human <strong>and</strong><br />

mouse somatostatin receptors expresssed<br />

in brain, gastrointestinal<br />

tract <strong>and</strong> kidney. Proc Natl Acad Sci<br />

USA 1992;89:251–255.<br />

22 Schaer JC, Waser B, Mengod G,<br />

Reubi JC: <strong>Somatostatin</strong> receptor<br />

subtypes, SSTR-1, SSTR-2, SSTR-3<br />

<strong>and</strong> SSTR-5 expression in human<br />

pituitary, gastroenteropancreatic<br />

<strong>and</strong> mammary tumors: Comparison<br />

of mRNA analysis with receptor autoradiography.<br />

Int J <strong>Cancer</strong> 1997;<br />

70:530–537.<br />

23 Patel YC, Srikant CB: <strong>Somatostatin</strong><br />

receptors. Trends Endocrinol Metab<br />

1997;8:398–405.<br />

24 Janson ET, Stridsberg M, Gobl M,<br />

et al: Determination of somatostatin<br />

receptor subtype 2 in carcinoid tumours<br />

by immunohistochemical investigation<br />

with somatostatin receptor<br />

subtype 2 antibodies. <strong>Cancer</strong> Res<br />

1998;58:2375–2378.<br />

25 Reubi JC, Koppeler A, Waser B, et<br />

al: Immunohistochemical localization<br />

of somatostatin receptors,<br />

SSTR-2A in human tumors. Am J<br />

Pathol 1998;153:233–245.<br />

26 Fassler JE, Hughes JH, Catal<strong>and</strong> S,<br />

et al: <strong>Somatostatin</strong> analogue: An inhibitor<br />

of angiogenesis? Biomed Res<br />

1988(suppl):181–185.<br />

27 Buscail L, Estève J-P, Saint-Laurent<br />

N, et al: Inhibition of cell proliferation<br />

by the somatostatin analogue<br />

RC-160 is mediated by somatostatin<br />

receptor subtypes SSTR2 <strong>and</strong><br />

SSTR5 through different mechanisms.<br />

Proc Natl Acad Sci USA<br />

1995;92:1580–1584.<br />

28 Imam H, Eriksson B, Lukinius A, et<br />

al: Induction of apoptosis in neuroendocrine<br />

tumors of the digestive<br />

system during treatment with somatostatin<br />

analogs. Acta Oncol 1997;<br />

36:607–614.<br />

29 Weckbecker G, Raulf F, Tolcsvai L,<br />

Bruns C: Potentiation of the antiproliferative<br />

effects of anticancer<br />

drugs by octreotide in vitro <strong>and</strong> in<br />

vivo. Digestion 1996;57(suppl 1):<br />

22–28.<br />

30 Reubi JC, Horisberger K, Laissue J:<br />

High density of somatostatin receptors<br />

in veins surrounding human<br />

cancer tissue: Role in tumour-host<br />

interaction? Int J <strong>Cancer</strong> 1994;56:<br />

681–688.<br />

31 Krenning EP, Kwekkeboom DJ,<br />

Bakker WH, et al: <strong>Somatostatin</strong> receptor<br />

scintigraphy with ( 111 In-<br />

DTPA-D-Phe 1 ) <strong>and</strong> ( 123 Tyr 3 )-octreotide.<br />

The Rotterdam experience<br />

with more than 1,000 patients. Eur J<br />

Nucl Med 1993;20:716–731.<br />

32 Tiensuu Janson E, Westlin JE, Eriksson<br />

B, et al: ( 111 In-DTPA-D-<br />

Phe 1 )-Octreotide scintigraphy in patients<br />

with carcinoid tumors: The<br />

predictive value <strong>for</strong> somatostatin<br />

analog treatment. Eur J Endocrinol<br />

1994;131:577–581.<br />

52 Chemotherapy 2001;47(suppl 2):40–53 Öberg<br />

33 Termanini B, Gibril F, Reynolds JC,<br />

et al: Value of somatostatin receptor<br />

scintigraphy: A prospective study<br />

in gastrinoma of its effect on clinical<br />

management. Gastroenterology<br />

1997;112:335–347.<br />

34 Lebtai R, Cadiot G, Sarda L, et al:<br />

Clinical impact of somatostatin receptor<br />

scintigraphy in the management<br />

of patients with neuroendocrine<br />

gastroenteropancreatic tumours.<br />

J Nucl Med 1997;38:853–<br />

858.<br />

35 Janson ET, Westlin JE, Eriksson B,<br />

et al: ( 111 In-DTPA-D-Phe 1 )-Octreotide<br />

scintigraphy in patients with<br />

carcinoid tumors; the predictive value<br />

<strong>for</strong> somatostatin analogue treatment.<br />

Eur J Endocrinol 1994;131:<br />

577–581.<br />

36 Lamberts SW, van der Lely AJ, De<br />

Herder WW, Hofl<strong>and</strong> LJ: Octreotide.<br />

N Engl J Med 1996;334:246–<br />

254.<br />

37 Kvols LK, Moertel CG, O’Connell<br />

MJ, et al: <strong>Treatment</strong> of the malignant<br />

carcinoid syndrome. Evaluation<br />

of a long-acting somatostatin<br />

analogue. N Engl J Med 1986;315:<br />

663–666.<br />

38 Gorden P, Comi RJ, Maton PN, Go<br />

VLW: <strong>Somatostatin</strong> <strong>and</strong> somatostatin<br />

analogue (SMS 201-995) in treatment<br />

of hormone-secreting tumors<br />

of the pituitary <strong>and</strong> gastrointestinal<br />

tract <strong>and</strong> non-neoplastic diseases of<br />

the gut. Ann Intern Med 1989;110:<br />

35–50.<br />

39 Öberg K, Norheim I, Lundqvist G,<br />

Wide L: <strong>Treatment</strong> of the carcinoid<br />

syndrome with SMS 201-995, a somatostatin<br />

analogue. Sc<strong>and</strong> J Gastroenterol<br />

1986;119:191–192.<br />

40 Maton PN, Gardner JD, Jensen RT:<br />

Use of long-acting somatostatin analogue<br />

SMS 201-995 in patients with<br />

pancreatic islet cell tumors. Dig Dis<br />

Sci 1989;34:285–291.<br />

41 Scarpignato C: <strong>Somatostatin</strong> analogues<br />

in the management of endocrine<br />

tumors of the pancreas; in<br />

Mignon M, Jensen RT (eds): Endocrine<br />

tumors of the Pancreas. Basel,<br />

Karger, 1995, pp 385–414.<br />

42 Ruszniewski P, Ramdani A, Cadiot<br />

G, et al: Long-term treatment with<br />

octreotide in patients with Zollinger-Ellison<br />

syndrome. Eur J Clin Invest<br />

1993;23:296–301.


43 Ferraro G, Annibale B, Mariquani<br />

M, et al: Effectiveness of octreotide<br />

in controlling fasting hypergastrinemia<br />

<strong>and</strong> related enterochromaffin-like<br />

cell growth. J Clin Endocrinol<br />

Metab 1996;81:677–683.<br />

44 Lamberts SWJ, Krenning EP, Reubi<br />

JC: The role of somatostatin <strong>and</strong> its<br />

analogs in the diagnosis <strong>and</strong> treatment<br />

of tumors. Endocrinol Rev<br />

1991;12:450–482.<br />

45 Wood SM, Kraenzlin ME, Adrian<br />

TE, Bloom SR: <strong>Treatment</strong> of patients<br />

with pancreatic endocrine tumours<br />

using a new long-acting somatostatin<br />

analogue: Symptomatic<br />

<strong>and</strong> peptide response. Gut 1985;26:<br />

438–444.<br />

46 Kvols LK, Buck M, Moertel CG, et<br />

al: <strong>Treatment</strong> of metastatic islet cell<br />

carcinoma with a somatostatin analogue<br />

(SMS 2011-995). Ann Intern<br />

Med 1987;107:162–168.<br />

47 Jockenhovel S, Lederbogen S, Olbricht<br />

T, et al: The long-acting somatostatin<br />

analogue octreotide alleviates<br />

symptoms by reducing posttranslational<br />

conversion of preproglucagon<br />

to glucagon in a patient<br />

with malignant glucagonoma, but<br />

does not prevent tumor growth. Clin<br />

Invest 1994;72:127–133.<br />

48 Angeletti S, Corletto D, Schillaci O:<br />

Use of somatostatin analogue octreotide<br />

to localize <strong>and</strong> manage somatostatin-producing<br />

tumors. Gut<br />

1998;42:792–794.<br />

49 Harris A, Redfern JS: Octreotide<br />

treatment of carcinoid syndrome:<br />

Analysis of published dose-titration<br />

data. Aliment Pharmacol Ther<br />

1995;9:387–394.<br />

50 Saltz L, Trochanowski B, Buckley<br />

M, et al: Octreotide as an anti-neoplastic<br />

agent in the treatment of<br />

functional <strong>and</strong> nonfunctional neuroendocrine<br />

tumors. <strong>Cancer</strong> 1993;<br />

72:244–248.<br />

Established Clinical Use of Octreotide<br />

<strong>and</strong> Lanreotide in Oncology<br />

51 Arnold R, Trautmann ME, Creutzfeldt<br />

W, et al: <strong>Somatostatin</strong> analogue<br />

<strong>and</strong> inhibition of tumor<br />

growth in metastatic endocrine gastroentero-pancreatic<br />

tumors. Gut<br />

1996;38:430–438.<br />

52 Bartholomeo M, Bajetta E, Buzzoni<br />

R, et al: Clinical efficacy of octreotide<br />

in the treatment of metastatic<br />

neuroendocrine tumors. A study by<br />

the Italian Trials in Medical Oncology<br />

Group. <strong>Cancer</strong> 1996;77:402–<br />

408.<br />

53 Eriksson B, Renstrup J, Imam H,<br />

Öberg K: High-dose treatment with<br />

lanreotide of patients with advanced<br />

neuroendocrine gastrointestinal tumours;<br />

clinical <strong>and</strong> biological effects.<br />

Ann Oncol 1997;8:1–4.<br />

54 Anthony L, Johnson D, H<strong>and</strong>e K, et<br />

al: <strong>Somatostatin</strong> analogue phase I<br />

trials in neuroendocrine neoplasms.<br />

Acta Oncol 1993;32:217–223.<br />

55 Faiss S, Wiedenmann B: Dose-dependent<br />

<strong>and</strong> antiproliferative effects<br />

of somatostatin. J Endocrinol Invest<br />

1997;20:68–70.<br />

56 Eriksson B, Janson ET, Bax NDS, et<br />

al: The use of new somatostatin analogues,<br />

lanreotide <strong>and</strong> octastatin, in<br />

neuroendocrine gastro-intestinal tumours.<br />

Digestion 1996;57:77–80.<br />

57 Bergström M, Eriksson B, Öberg K,<br />

et al: Induction of apoptosis in neuroendocrine<br />

tumors of the digestive<br />

system during treatment with somatostatin<br />

analogs. Acta Oncol 1997;<br />

36:32–37.<br />

58 Harris A, Kokoris S, Ezzat S: Continuous<br />

versus intermittent subcutaneous<br />

infusion of octreotide in the<br />

treatment of acromegaly. J Clin<br />

Pharmacol 1995;35:59–71.<br />

59 Hofl<strong>and</strong> LJ, Koetsveld, Waaijers M,<br />

et al: Relative potencies of the somatostatin<br />

analogs octreotide, BIM–<br />

23014, <strong>and</strong> RC-160 on the inhibition<br />

of hormone release by cultured<br />

human endocrine tumor cells <strong>and</strong><br />

normal rat anterior pituitary cells.<br />

Endocrinology 1994;134:301–306.<br />

60 Lancranjan I, Bruns C, Grass P, et<br />

al: S<strong>and</strong>ostatin-LAR: Pharmaco-kinetics,<br />

pharmacodynamics, efficacy,<br />

<strong>and</strong> tolerability in acromegalic patients.<br />

Metabolism 1995;44:18–26.<br />

61 Ruszniewski P, Ducreux M, Chayvialle<br />

JA et al: <strong>Treatment</strong> of the carcinoid<br />

syndrome with the long-acting<br />

somatostatin analogue lanreotide:<br />

A prospective study in 39 patients.<br />

Gut 1996;39:279–283.<br />

62 Wymenga ANM, Eriksson B, Salmela<br />

PI, et al: Efficacy <strong>and</strong> safety of<br />

lanreotide prolonged release in patients<br />

with gastrointestinal neuroendocrine<br />

tumors with hormone related<br />

symptoms. J Clin Oncol, in<br />

press.<br />

63 Janson ET, Ahlström H, Andersson<br />

T, Öberg K: Octreotide <strong>and</strong> interferon<br />

alpha: A new combination <strong>for</strong> the<br />

treatment of malignant carcinoid tumours.<br />

Eur J <strong>Cancer</strong> 1992;28A:<br />

1647–1650.<br />

64 Trendle MC, Moertel CG, Kvols<br />

LK: Incidence <strong>and</strong> morbidity of<br />

cholelithiasis in patients receiving<br />

chronic octreotide <strong>for</strong> metastatic<br />

carcinoid <strong>and</strong> malignant islet cell tumours.<br />

<strong>Cancer</strong> 1997;79:830–834.<br />

65 Shimon I, Taylor J, Weiss MH, et al:<br />

<strong>Somatostatin</strong> receptor (SSTR) subtype-selective<br />

analogs differentially<br />

suppress in vitro growth hormone<br />

<strong>and</strong> prolactin in human pituitary adenomas.<br />

Novel potential therapy <strong>for</strong><br />

functional pituitary tumours. J Clin<br />

Invest 1997;100:2386–2392.<br />

Chemotherapy 2001;47(suppl 2):40–53 53


Chemotherapy 2001;47(suppl 2):54–61<br />

The Palliative Effects of Octreotide in<br />

<strong>Cancer</strong> Patients<br />

Andrew Dean<br />

Palliative Care Service, Sir Charles Gairdner Hospital, Nedl<strong>and</strong>s, Australia<br />

Key Words<br />

Palliative care W <strong>Somatostatin</strong> W Opioids<br />

Abstract<br />

Octreotide is an extremely useful compound<br />

<strong>for</strong> palliative care physicians. It appears to be<br />

active in a number of different pain states<br />

<strong>and</strong> may be given by the spinal <strong>and</strong> intraventricular<br />

route. Its actions in reducing gut motility<br />

<strong>and</strong> secretions make it a valuable adjunct<br />

in the management of inoperable bowel<br />

obstruction. The same actions make it a<br />

potent antidiarrheal agent. Octreotide will often<br />

succeed where other antidiarrheal agents<br />

fail. Its ability to reduce gut secretions has led<br />

to its use in the treatment of fistulae. It has<br />

also been proposed as a useful drug in the<br />

management of cachexia <strong>and</strong> ascites. Most<br />

of the existing evidence is based on small<br />

numbers of case reports <strong>and</strong> further larger<br />

trials are necessary.<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

Copyright © 2001 S. Karger AG, Basel<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0054$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Introduction<br />

Octreotide, a synthetic analog of somatostatin,<br />

is an interesting compound to many<br />

specialities. It has a number of properties<br />

which make it potentially useful in many different<br />

palliative care situations. This paper<br />

serves to address some of its applications.<br />

When considering the palliative uses of octreotide,<br />

the most pertinent quote to consider<br />

is ‘in most situations where the drug has been<br />

found to be useful, no controlled clinical trials<br />

have been per<strong>for</strong>med. At least in part this has<br />

been due to the rarity of the conditions<br />

treated. It does make it difficult, however, to<br />

assess just how effective octreotide is in many<br />

circumstances’ [1].<br />

Pain<br />

The past 5 years have seen tremendous<br />

improvements in the underst<strong>and</strong>ing of pain<br />

<strong>and</strong> analgesic neurophysiology. A distinction<br />

is made between opioid-sensitive pain <strong>and</strong> the<br />

Dr. Andrew Dean<br />

Palliative Care Service, Sir Charles Gairdner Hospital<br />

Nedl<strong>and</strong>s, WA 6009 (Australia)<br />

Tel. +61 8 9346 2551, Fax +61 8 9346 1848<br />

E-Mail <strong>and</strong>rew.dean@health.wa.gov.au


pain that exhibits partial or no response to<br />

narcotic analgesics. There are now a vast<br />

array of opioids to choose from including<br />

morphine, hydromorphone, methadone, oxycodone,<br />

<strong>and</strong> fentanyl, each having a place<br />

in mainstream analgesic pharmacotherapy.<br />

Opioids, however, are not perfect analgesics.<br />

Pain from nerve damage (neuropathic pain,<br />

neurogenic pain) exhibits a variable degree of<br />

opioid resistance. There has been substantial<br />

work done in identifying some of the mechanisms<br />

of opioid resistance. Drugs with actions<br />

on sodium channels, NMDA receptors, calcium<br />

channels <strong>and</strong> inflammatory mediators<br />

are now abundant. Many new <strong>and</strong> old drugs<br />

are thus used by the pain practitioner.<br />

Innovative routes of drug delivery are now<br />

also more common. The use of epidural <strong>and</strong><br />

intrathecal analgesia <strong>for</strong> acute, chronic <strong>and</strong><br />

cancer pain is commonplace. The number of<br />

patients who fail these now st<strong>and</strong>ard analgesic<br />

approaches are becoming fewer. The controversy<br />

regarding the value of octreotide in pain<br />

medicine should be viewed in this context of a<br />

drug with potential therapeutic benefit <strong>for</strong> a<br />

relatively small number of patients, i.e. those<br />

who fail the traditional approach.<br />

In underst<strong>and</strong>ing the actions of octreotide<br />

in pain, it is first necessary to note that somatostatin<br />

receptors are found in dorsal horn<br />

afferent neurones, spinal interneurones, <strong>and</strong><br />

ascending <strong>and</strong> descending pathways. Native<br />

somatostatin is found in the periaqueductal<br />

grey matter, substantia gelatinosa, the spinal<br />

cord <strong>and</strong> in descending pathways [2–7]. <strong>Somatostatin</strong><br />

is produced in the hypothalamus,<br />

cortex, cerebellum <strong>and</strong> spinal cord as well as<br />

in multiple organs of the gastrointestinal tract<br />

(stomach, small bowel, colon <strong>and</strong> pancreas)<br />

[2, 8–11].<br />

The effects of octreotide on pain can be<br />

considered to be a summation of its actions as<br />

a neurotransmitter <strong>and</strong> as an autocrine/paracrine<br />

regulator. These effects are mediated via<br />

The Palliative Effects of Octreotide in<br />

<strong>Cancer</strong> Patients<br />

a family of (as yet) 5 identified somatostatin<br />

receptors [12].<br />

Given the distribution of somatostatin receptors<br />

in those parts of the nervous system<br />

concerned with pain transmission <strong>and</strong> inhibition,<br />

somatostatin <strong>and</strong> its analogs would be<br />

expected to have an effect on pain. The cellular<br />

<strong>and</strong> molecular effects of somatostatin <strong>and</strong><br />

its analogs are also interesting. The somatostatin<br />

receptors are G-coupled protein receptors<br />

which exhibit many different actions.<br />

The various somatostatin analogs also exhibit<br />

differential receptor binding. Some receptors<br />

cause hyperpolarization of the cell<br />

membrane [13] <strong>and</strong> also block calcium channels<br />

[14]. Binding to some receptors regulates<br />

glutamate currents at the AMPA receptor<br />

[15]. Interestingly, opposing effects are seen<br />

on the glutamate current depending on which<br />

receptor subtype is activated [15].<br />

Further indirect evidence of a potential<br />

analgesic action is found in examining the<br />

effects of other drugs on somatostatin secretion.<br />

Opioids <strong>and</strong> GABA agonists inhibit somatostatin<br />

secretion, as do high dose steroids.<br />

Low dose steroids, however, encourage somatostatin<br />

release. These are essentially observational<br />

phenomena which do not in themselves<br />

prove that somatostatin or its analogs<br />

are directly analgesic but certainly provide<br />

circumstantial evidence that somatostatin<br />

may be involved in the pain process.<br />

What clinical evidence is there <strong>for</strong> analgesic<br />

activity? In 1990 <strong>and</strong> 1991 [16, 17] subcutaneous<br />

octreotide was reported to relieve<br />

headache in acromegaly which was not related<br />

to tumor size. Naloxone did not reverse<br />

analgesia, suggesting that the analgesic effect<br />

was not mediated by st<strong>and</strong>ard opioid mechanisms.<br />

The limitations of imaging techniques<br />

in assessing minute changes in intracranial<br />

tumor size should, however, be noted.<br />

No r<strong>and</strong>omized controlled trials have<br />

proven subcutaneous octreotide to have a spe-<br />

Chemotherapy 2001;47(suppl 2):54–61 55


Table 1. Analgesic requirements be<strong>for</strong>e <strong>and</strong> after epidural or intrathecal somatostatin treatment<br />

<strong>for</strong> the 8 patients in the study of Mollenholt et al. [20, p. 537]<br />

Patient<br />

number<br />

1<br />

Daily analgesic dose<br />

be<strong>for</strong>e somatostatin<br />

treatment<br />

Ketobemidone, 60 mg orally<br />

Ketobemidone, 20 mg i.m.<br />

Paracetamol, 3 g orally<br />

Daily concomitant<br />

analgesic dose during<br />

somatostatin treatment<br />

56 Chemotherapy 2001;47(suppl 2):54–61 Dean<br />

Response to<br />

treatment<br />

None Excellent<br />

2 Morphine, 500 mg i.v. Morphine, 20 mg s.c. a Good<br />

3 Ketobemidone, 80 mg i.v.<br />

Methadone, 30 mg orally<br />

None Excellent<br />

4 Morphine, 300 mg i.v. Morphine, 20 mg s.c. b Good<br />

5 Morphine, 40–60 mg s.c.<br />

Morphine, 160 mg orally<br />

Paracetamol, 6 g orally<br />

6 Morphine, 60 mg orally<br />

Paracetamol, 4 g orally<br />

7 Morphine, 10–30 mg s.c.<br />

Morphine, 600 mg orally<br />

Paracetamol, 3 g orally<br />

8 Morphine, 1,500 mg orally<br />

Ketobemidone, 100 mg rectally<br />

Unaltered Poor<br />

Morphine, 0–20 mg orally<br />

Paracetamol, 1 g rectally<br />

Morphine, 180 mg orally<br />

Paracetamol, 3 g orally<br />

Good<br />

Good<br />

Morphine, 60 mg i.v. c Fair<br />

Ketobemidone is a synthetic opioid that is equipotent with morphine.<br />

a This dose of morphine was used to treat withdrawal symptoms <strong>and</strong> not <strong>for</strong> pain relief.<br />

b Intravenous morphine, 300 mg daily after removal of dislodged intrathecal catheter.<br />

c Intravenous morphine, 200–1,200 mg daily after the unintentional removal of the epidural<br />

catheter.<br />

cific analgesic action. De Conno et al. [18]<br />

found pain relief in 1 of 10 patients treated<br />

with subcutaneous octreotide <strong>for</strong> pain from<br />

cancer. Although a chemical action was proposed<br />

by the authors, it should be noted that<br />

the only patient with analgesic effect received<br />

relief from severe postpr<strong>and</strong>ial pain complicating<br />

pancreatic carcinoma. Pancreatic carcinoma<br />

can produce mesenteric ischemia because<br />

of its proximity to the celiac axis. We<br />

have treated several patients with mesenteric<br />

ischemia complicating pancreatic cancer with<br />

subcutaneous octreotide. This appeared to be<br />

an effective analgesic in these circumstances.<br />

The most likely mechanism of action is a<br />

reduction in bowel motility <strong>and</strong> secretion,<br />

with consequent reduction in ischemic pain.<br />

Spinal octreotide <strong>and</strong> somatostatin have<br />

been used <strong>for</strong> analgesia. Epidural somatostatin<br />

was shown to be an effective analgesic<br />

in upper abdominal surgery [19]. Mollenholt<br />

et al. [20] reported in 1994 that of 8 patients<br />

with intractable cancer pain treated with spinal<br />

somatostatin, 6 patients reported excellent<br />

or good pain relief (table 1). Five of these were<br />

patients who were probably suffering from


neurogenic pain. Penn et al. [21] reported the<br />

results of a 6-patient pilot study on intrathecal<br />

octreotide infusion <strong>for</strong> cancer pain. The patients<br />

were suffering from a variety of different<br />

pain types but it would appear a significant<br />

component of them had neurogenic pain.<br />

Good analgesia was generally obtained. Interestingly,<br />

analgesia was achieved in patients<br />

who had demonstrated significant opioid resistance.<br />

There has been much controversy<br />

regarding the use of spinal somatostatin <strong>and</strong><br />

octreotide because of reports of neurotoxicity<br />

in some animal species [22–24].<br />

Intraventricular octreotide has also been<br />

used to provide effective analgesia [25]. This<br />

should potentially be useful in patients with<br />

head <strong>and</strong> neck cancer exhibiting local nonopioid-sensitive<br />

pain.<br />

Interestingly the study of Mollenholt et al.<br />

[20] in 1994 was able to per<strong>for</strong>m autopsies on<br />

5 patients who received spinal somatostatin.<br />

Three patients of the 5 studied showed some<br />

demyelination either of dorsal roots or dorsal<br />

columns. These effects could also be explained<br />

as being part of a paraneoplastic process.<br />

No patients had neurological deficits<br />

which could be correlated with this finding.<br />

The debate continues whether octreotide<br />

should be used by these routes but it is encouraging<br />

to read the report of Paice et al. [26] of<br />

octreotide being used <strong>for</strong> 5 years via the intrathecal<br />

route to successfully manage chronic<br />

pain. This occurred without neurotoxicity.<br />

Although there are no large r<strong>and</strong>omized<br />

double-blinded controlled trials of octreotide<br />

in these situations, the large numbers of case<br />

reports indicating that octreotide has potent<br />

analgesic activity (especially in non-opioidsensitive<br />

pain) is encouraging. My own practice<br />

is to use octreotide via these routes as an<br />

adjunct to other analgesics when conventional<br />

approaches fail but ultimately the decision on<br />

whether or not to use octreotide in difficult pain<br />

scenarios is down to the individual clinician.<br />

The Palliative Effects of Octreotide in<br />

<strong>Cancer</strong> Patients<br />

Bowel Obstruction<br />

Since Khoo et al. [27] reported the use of<br />

octreotide in bowel obstruction in 1992, there<br />

have been many publications about this subject.<br />

The effects of octreotide on the gut are to<br />

reduce gut motility <strong>and</strong> secretion. Native somatostatin<br />

has been shown to inhibit fluid<br />

secretion into the rat jejunum [28] <strong>and</strong> stimulate<br />

sodium <strong>and</strong> chloride absorption in the<br />

rabbit ileum [29]. Octreotide has been shown<br />

to prolong small intestinal transit time in humans<br />

[30]. As both gastric <strong>and</strong> pancreatic<br />

secretions are significantly decreased, there is<br />

less total fluid turnover in the gut.<br />

A consequence of reduction in gut distension<br />

is to delay the onset of edema <strong>and</strong> ischemia<br />

in the antimesenteric border of the intestine.<br />

The subsequent delay in necrosis <strong>and</strong><br />

per<strong>for</strong>ation in mice with proximal small bowel<br />

obstruction treated with octreotide was<br />

noted in 1992 [31]. If this effect is mirrored in<br />

humans, as might be expected, it raises interesting<br />

questions relating to potential improvement<br />

<strong>and</strong> survival of patients receiving octreotide<br />

<strong>for</strong> bowel obstruction. In one controlled<br />

study of jejunal ligation in rats, octreotide<br />

significantly reduced the diameter of obstructed<br />

bowel <strong>and</strong> reduced sodium <strong>and</strong> potassium<br />

losses. The histopathological ischemic<br />

changes were more prominent in the<br />

control group <strong>and</strong> anastomotic bursting pressures<br />

were higher in the treated group.<br />

In a palliative care perspective, the main<br />

benefit of octreotide use in this setting is a<br />

reduction in distressing symptoms. The original<br />

report of Khoo et al. [27] was of 5 patients<br />

with vomiting unresponsive to conventional<br />

antiemetics but who responded rapidly<br />

to octreotide 300 mg/24 h subcutaneously. A<br />

marked reduction in nasogastric tube aspirate<br />

was noted in 2 of these patients. In our own<br />

experience <strong>and</strong> that of others this observation<br />

Chemotherapy 2001;47(suppl 2):54–61 57


has been consistent [32–35]. One patient of<br />

ours with inoperable bowel obstruction survived<br />

<strong>for</strong> 3 months being treated at home with<br />

subcutaneous octreotide <strong>and</strong> nightly subcutaneous<br />

fluid. The use of octreotide in palliative<br />

care in the home setting has also been reported<br />

by Mercadante [36].<br />

Diarrhea<br />

The actions of octreotide in reducing intestinal<br />

motility <strong>and</strong> water secretion plus the<br />

effects on sodium <strong>and</strong> potassium absorption<br />

[1] make octreotide an important agent in the<br />

treatment of diarrhea.<br />

Profuse diarrhea is an accompaniment of<br />

many illnesses. In the cancer field it complicates<br />

hormonally active intestinal tumors <strong>and</strong><br />

is a feature of short bowel syndrome after<br />

multiple resections [37]. Diarrhea also complicates<br />

celiac plexus block, a technique used<br />

in cancer pain treatment [38]. Subcutaneous<br />

octreotide has successfully treated this symptom<br />

in these conditions where other treatments<br />

have failed. The diarrhea associated<br />

with AIDS can be particularly severe <strong>and</strong><br />

when seemingly intractable, octreotide offers<br />

an excellent therapeutic solution [39].<br />

Octreotide also has a role in supportive<br />

care of patients undergoing chemotherapy.<br />

Agents such as 5-fluorouracil can cause severe<br />

diarrhea unresponsive to conventional treatment.<br />

Octreotide may often work in this situation<br />

[40–42] but it does not seem to have a<br />

preventive action [43]. As with all diarrheas<br />

the etiology should be clearly established <strong>and</strong><br />

treated specifically as appropriate.<br />

Long-term subcutaneous administration of<br />

octreotide is not a cost-effective means of<br />

symptom management. If conventional measures<br />

such as use of opioids <strong>and</strong> drugs such as<br />

loperamide fail, then it is my view that octreotide<br />

is certainly worth a trial. Some sources<br />

58 Chemotherapy 2001;47(suppl 2):54–61 Dean<br />

suggest that loperamide is at least as effective<br />

as low dose octreotide (150 Ìg/day) [44]. In<br />

my experience doses from 300–600 Ìg daily<br />

are usually effective in controlling difficult<br />

diarrheal states. Although 600 Ìg is the usual<br />

ceiling dose we have used doses of up to<br />

1,200 Ìg daily with good effect <strong>and</strong> without<br />

adverse consequence.<br />

Fistulae<br />

Many patients develop fistulae after abdominal<br />

surgery <strong>for</strong> recurrent malignancy.<br />

The intensity of fistula output can be particularly<br />

distressing as both a constant reminder<br />

of the unremitting nature of malignant disease<br />

<strong>and</strong> from practical aspects such as requirements<br />

<strong>for</strong> multiple stoma bags. Octreotide<br />

can certainly reduce the fistula output<br />

[45] in these situations with consequent improvement<br />

in quality of life. Although r<strong>and</strong>omized<br />

control trials of octreotide on fistula<br />

closure have not demonstrated significant<br />

benefit, there has been some debate as to the<br />

reason <strong>for</strong> this [46–48].<br />

In an interesting report by Jenkins et al.<br />

[49] the authors measured pancreatic enzyme<br />

concentration in fistula secretion <strong>and</strong> found<br />

that enzyme concentration increases between<br />

intermittent subcutaneous injections of octreotide;<br />

they postulate that a low-volume<br />

high-enzyme concentration fistula output<br />

may be detrimental to fistula closure. It is not<br />

known whether subcutaneous infusion of octreotide<br />

would prevent this observation.<br />

Hern<strong>and</strong>ez-Ar<strong>and</strong>a et al. [50] showed an<br />

improvement in fistula closure rate in a<br />

mixed group of patients with enterocutaneous<br />

fistulae. Interestingly the total parenteral nutritional<br />

time required in the treated group<br />

was shorter in the octreotide group than in the<br />

placebo group.


Other Uses – Cachexia, Ascites <strong>and</strong><br />

Carcinoid Syndrome<br />

Some interesting laboratory data has appeared<br />

on the use of octreotide with insulin<br />

<strong>and</strong> with insulin plus growth hormone <strong>for</strong> the<br />

treatment of cancer cachexia. As increased<br />

glucagon levels are found in some situations<br />

of cancer cachexia, the experimental rationale<br />

was to reverse the low insulin/glucagon ratio.<br />

A study was carried out in rats <strong>and</strong> the combination<br />

of somatostatin plus insulin was shown<br />

to increase weight <strong>and</strong> muscle protein [51].<br />

Combination of octreotide, insulin <strong>and</strong><br />

growth was also investigated <strong>and</strong> found to<br />

have similar effects. Whether the growth hormone<br />

added anything to the regimen was not<br />

clear.<br />

Although seemingly physiologically effective,<br />

it remains to be seen whether this translates<br />

into a useful quality of life measure in<br />

human cancer-associated cachexia.<br />

Cairns <strong>and</strong> Malone [52] have recently published<br />

a report of 3 cases of patients with<br />

ascites who were treated with octreotide. For<br />

2 of these patients, the quantity of ascites<br />

appeared to decrease significantly <strong>and</strong> the<br />

need <strong>for</strong> paracentesis ceased. One of these<br />

patients was suffering from adenocarcinoma<br />

of the colon <strong>and</strong> the other adenocarcinoma of<br />

the breast. The exact reason <strong>for</strong> these observations<br />

is unclear but an effect on somatostatin<br />

receptors on tumors has been postulated.<br />

Octreotide is exceptionally useful in the<br />

palliation of symptoms due to the carcinoid<br />

syndrome. It can succesfully control flushing<br />

as well as diarrhea. Patients with the metastatic<br />

carcinoid syndrome who experience bone<br />

pain also reported alleviation of this symptom<br />

when treated with octreotide [53].<br />

One concern in long-term administration<br />

is dose tachyphylaxis. In a prospective study<br />

of octreotide’s effect on gastric function some<br />

of the physiological effects were noted to di-<br />

The Palliative Effects of Octreotide in<br />

<strong>Cancer</strong> Patients<br />

minish on day 6 <strong>and</strong> 7 of treatment [54]. Certainly<br />

it is not known whether this is of clinical<br />

importance in palliative care patients<br />

whose prognosis is often short. In our own<br />

experience, dose tachyphylaxis does not appear<br />

to be an important issue. As octreotide<br />

use in palliative situations becomes more<br />

commonplace, the literature may well answer<br />

some of these questions.<br />

Although the physiology <strong>and</strong> pharmacology<br />

of octreotide is well understood, there remains<br />

a shortage of r<strong>and</strong>omized controlled<br />

trials which prospectively study its use in<br />

many conditions. Case reports <strong>and</strong> anecdotal<br />

evidence abound. For the palliative care practitioner<br />

I believe octreotide is an important<br />

compound with many potential uses. The recent<br />

development of a long-acting slow-release<br />

depot preparation of octreotide <strong>and</strong> other<br />

somatostatin analogs is certainly exciting.<br />

The potential uses of this preparation would<br />

not necessitate subcutaneous infusion pumps,<br />

which is clearly a great convenience to the palliative<br />

care patient. I look <strong>for</strong>ward to more<br />

r<strong>and</strong>omized control trials in the <strong>for</strong>thcoming<br />

years.<br />

Acknowledgments<br />

I am indebted to Karen Mattioli <strong>and</strong> Joanne Blight<br />

<strong>for</strong> their assistance in the preparation of the manuscript<br />

<strong>and</strong> to Professor Carmelo Scarpignato (University<br />

of Parma, Italy) <strong>for</strong> his encouragement, constructive<br />

criticism <strong>and</strong> assistance with the bibliography.<br />

Chemotherapy 2001;47(suppl 2):54–61 59


References<br />

1 Maton PN: Exp<strong>and</strong>ing uses of octreotide.<br />

Baillières Clin Gastroenterol<br />

1994;8:321–337.<br />

2 Reichlin S: <strong>Somatostatin</strong>. N Engl J<br />

Med 1983;309:1495–1501.<br />

3 Terenius L: <strong>Somatostatin</strong> <strong>and</strong><br />

ACTH are peptides with partial antagonist-like<br />

selectivity <strong>for</strong> opiate receptors.<br />

Eur J Pharmacol 1976;38:<br />

211–213.<br />

4 Hökfelt T, Elde R, Johansson O,<br />

Luft R, Nilsson G, Arimura A: Immunohistochemical<br />

evidence <strong>for</strong><br />

separate populations of somatostatin-containing<br />

<strong>and</strong> substance Pcontaining<br />

primary afferent neurons<br />

in the rat. Neuroscience 1976;1:<br />

131–136.<br />

5 Elde R, Johansson O, Hökfelt T: Immunocytochemical<br />

studies of somatostatin<br />

neurons in brain. Adv<br />

Exp Med Biol 1985;196:167–181.<br />

6 Stine SM, Yang H-Y, Costa E: Evidence<br />

of ascending <strong>and</strong> descending<br />

intraspinal as well as primary sensory<br />

somatostatin projection in the rat<br />

spinal cord. J Neurochem 1982;38:<br />

1144–1150.<br />

7 Shimada S, Shiosaka S, Takami K,<br />

Yamano M, Tohyama M: <strong>Somatostatin</strong>ergic<br />

neurons in the insular<br />

cortex project to the spinal cord:<br />

Combined retrograde axonal transport<br />

<strong>and</strong> immunohistochemical<br />

study. Brain Res 1985;326:197–<br />

200.<br />

8 Newman JB, Lluis F, Townsend CM<br />

Jr: <strong>Somatostatin</strong>; in Thompson JC,<br />

Greeley GH Jr, Ray<strong>for</strong>d PL, Townsend<br />

CM Jr (eds): Gastrointestinal<br />

Endocrinology. New York,<br />

McGraw-Hill Book, 1987, pp 286–<br />

299.<br />

9 Lucey MR: Endogenous somatostatin<br />

<strong>and</strong> the gut. Gut 1986;27:<br />

457–467.<br />

10 Hökfelt T, Effendic S, Hellerstrom<br />

C, Johansson O, Luft R, Arimura A:<br />

Cellular localization of somatostatin<br />

in endocrine-like cells <strong>and</strong> neurons<br />

of the rat with special references to<br />

the A1 cells of the pancreatic islets<br />

<strong>and</strong> to the hypothalamus. Acta Endocrinol<br />

1975;80(suppl 200):5–41.<br />

11 Patel YC, Reichlin S: <strong>Somatostatin</strong><br />

in hypothalamus, extrahypothalamic<br />

brain <strong>and</strong> peripheral tissues of<br />

the rat. Endocrinology 1978;102:<br />

523–530.<br />

12 Patel YC: <strong>Somatostatin</strong> <strong>and</strong> its receptor<br />

family. Front Neuroendocrinol<br />

1999;20:157–198.<br />

13 Sims SM, Lussier BT, Kraicer J: <strong>Somatostatin</strong><br />

activates an inwardly<br />

rectifying K + conductance in freshly<br />

dispersed rat somatotrophs. J Physiol<br />

1991;441:615–637.<br />

14 Ikeda SR, Schofield GG: <strong>Somatostatin</strong><br />

blocks a calcium current in rat<br />

sympathetic ganglion neurons. J<br />

Physiol 1989;409:221–240.<br />

15 Lanneau C, Viollet C, Faivre-Bauman<br />

A, Loudes C, Kordon C, Epelbaum<br />

J, Gardette RS: <strong>Somatostatin</strong><br />

receptor subtypes sst1 <strong>and</strong> sst2 elicit<br />

opposite effects on the response to<br />

glutamate of mouse hypothalamic<br />

neurones: An electrophysiological<br />

<strong>and</strong> single cell RT-PCR study. Eur J<br />

Neurosci 1998;10:204–212.<br />

16 Musolino NR, Marino Junior R,<br />

Bronstein MD: Headache in acromegaly:<br />

Dramatic improvement<br />

with the somatostatin analogue SMS<br />

201-995. Clin J Pain 1990;6/3:243–<br />

245.<br />

17 Pascal J, Freijanes J, Berciano J,<br />

Pesquerac C: Analgesic effect of octreotide<br />

in headache associated with<br />

acromegaly is not mediated by<br />

opioid mechanisms. Case report.<br />

Pain 1991;47:341–344.<br />

18 De Conno F, Saita L, Ripamonti C,<br />

Ventafridda V: Subcutaneous octreotide<br />

in the treatment of pain in<br />

advanced cancer patients. J Pain<br />

Symptom Manage 1994;9:34–38.<br />

19 Taura P, Planella V, Balust J, Anglada<br />

T, Carrero E, Brugues S: Epidural<br />

somatostatin as an analgesic in<br />

upper body surgery: A double blind<br />

study. Pain 1994;59/1:135–140.<br />

20 Mollenholt P, Rawal N, Gordh T,<br />

Olsson Y: Intrathecal <strong>and</strong> epidural<br />

somatostatin <strong>for</strong> patients with cancer.<br />

Analgesic effects <strong>and</strong> postmortem<br />

neuropathologic investigations<br />

of spinal cord <strong>and</strong> nerve roots. Anesthesiology<br />

1994;81:534–542.<br />

60 Chemotherapy 2001;47(suppl 2):54–61 Dean<br />

21 Penn RD, Paice JA, Kroin JS: Octreotide:<br />

A potent new non-opiate<br />

analgesic <strong>for</strong> intrathecal infusion.<br />

Pain 1992;49:13–19.<br />

22 Gaumann DM, Yaksh TL: Intrathecal<br />

somatostatin in rats: Antinociception<br />

only in the presence of toxic<br />

effects. Anesthesiology 1988;69:<br />

733–742.<br />

23 Asai T: Use of epidural somatostatin<br />

<strong>for</strong> postoperative analgesia is not<br />

justifiable. Pain 1995;62:388.<br />

24 Yaksh TL: Spinal somatostatin <strong>for</strong><br />

patients with cancer: Risk-benefits<br />

assessment of an analgesic. Anesthesiology<br />

1994;81:531–533.<br />

25 C<strong>and</strong>rina R, Gallli G: Intraventricular<br />

octreotide <strong>for</strong> cancer pain. J Neurosurg<br />

1992;76:336–337.<br />

26 Paice JA, Penn RD, Croin JS: Intrathecal<br />

octreotide <strong>for</strong> relief of intractable<br />

nonmalignant pain: 5-year experience<br />

with two cases. Neurosurgery<br />

1996;38/1:203–207.<br />

27 Khoo D, Riley J, Waxman J: Control<br />

of emesis in bowel obstruction<br />

in terminally ill patients. Lancet<br />

1992;339:375–376.<br />

28 Dharmsathaphorn K, Sherwin RS,<br />

Dobbins JW: <strong>Somatostatin</strong> inhibits<br />

fluid secretion in the rat jejunum.<br />

Gastroenterology 1980;78:1554–<br />

1558.<br />

29 Dharmsathaphorn K, Binder HJ,<br />

Dobbins JW: <strong>Somatostatin</strong> stimulates<br />

sodium <strong>and</strong> chloride absorption<br />

in the rat ileum. Gastroenterology<br />

1980;78:1559–1565.<br />

30 Dueno MI, Bai JC, Santangelo WD,<br />

Krejs GJ: Effect of somatostatin analogue<br />

on water <strong>and</strong> electrolyte<br />

transport <strong>and</strong> transit time in human<br />

small bowel. Dig Dis Sci 1987;32:<br />

1092–1096.<br />

31 Gittes GK, Nelson MT, Debas HT,<br />

Mulvihill SJ: Improvement in survival<br />

of mice with proximal small<br />

bowel obstruction treated with octreotide.<br />

Am J Surg 1992;163:231–<br />

233.<br />

32 Mercadante S, Maddaloni S: Octreotide<br />

in the management of inoperable<br />

gastrointestinal obstruction<br />

in terminal cancer patients. J Pain<br />

Symptom Manage 1992;7:496–498.


33 Merc<strong>and</strong>ate S, Spoldi E, Caraceni A,<br />

Maddaloni S, Simonetti MT: Octreotide<br />

in relieving gastrointestinal<br />

symptoms due to bowel obstruction.<br />

Palliat Med 1993;7/4:295–299.<br />

34 Mangili G, Franchi M, Mariani A,<br />

Zanaboni F, Rabaiotti E, Rigerio L,<br />

Bolis PF, Ferrari A: Octreotide in<br />

the management of bowel obstruction<br />

in terminal ovarian cancer. Gynecol<br />

Oncol 1996;61:345–348.<br />

35 Fainsinger RL, Pisani A, Bruera E:<br />

Use of somatostatin analogues in<br />

terminal cancer patients. J Palliat<br />

Care 1993;9/1:56–57.<br />

36 Mercadante S: Bowel obstruction in<br />

home-care cancer patients: 4 years’<br />

experience. Support Care <strong>Cancer</strong><br />

1995;3/3:190–193.<br />

37 Sharkey MF, Kadden ML, Stabile<br />

BE: Severe posthemicolectomy diarrhoea:<br />

Evaluation <strong>and</strong> treatment<br />

with SMS 201-995. Gastroenterology<br />

1990;99:1144–1148.<br />

38 Dean AP, Reed WD: Diarrhoea –<br />

An unrecognised hazard of coeliac<br />

plexus block. Aust NZ J Med 1991;<br />

1:47–48.<br />

39 Cello JP, Grendell JH, Basuk P, Simon<br />

P, Weiss L, Wittner M, Rood<br />

RP, Wilcox CM, Forsmark CE,<br />

Read AE, et al: Effect of octreotide<br />

on refractory AIDS associated with<br />

diarrhoea. A prospective multicentre<br />

clinical trial. Ann Intern Med<br />

1991;155:705–710.<br />

40 Cascinu S, Fedeli A, Fedeli SL,<br />

Catalano G: Control of chemotherapy<br />

induce diarrhoea with octreotide<br />

in patients receiving 5-fluorouracil.<br />

Eur J <strong>Cancer</strong> 1992;28:482–483.<br />

41 Cascinu S, Fedeli A, Fedeli SL, Catalano<br />

G: Control of chemotherapyinduced<br />

diarrhoea with octreotide.<br />

A r<strong>and</strong>omised trial with placebo in<br />

patients receiving cisplatin. Oncology<br />

1994;51/1:70–73.<br />

The Palliative Effects of Octreotide in<br />

<strong>Cancer</strong> Patients<br />

42 Cascinu S, Fedeli A, Fedeli SL, Catalano<br />

G: Octreotide versus loperamide<br />

in the treatment of fluorouracil<br />

induced diarrhoea: A r<strong>and</strong>omised<br />

trial. J Clin Oncol 1993;11/1:<br />

148–151.<br />

43 Meropol NJ, Blumenoson LE,<br />

Creaven PJ: Octreotide does not<br />

prevent diarrhoea in patients<br />

treated with weekly 5-fluorouracil<br />

plus high dose leucovorin. Am J<br />

Clin Oncol 1998;21/2:135–138.<br />

44 Geller RB, Gilmore CE, Dix SP, Lin<br />

LS, Topping DL, Davidson TG,<br />

Holl<strong>and</strong> HK, Wingard JR: R<strong>and</strong>omized<br />

trial of loperamide versus dose<br />

escalation of octreotide acetate <strong>for</strong><br />

chemotherapy-induced diarrhoea in<br />

bone marrow transplant <strong>and</strong> leukaemia<br />

patients. Am J Hematol<br />

1995;50/3:167–172.<br />

45 Nubiola P, Badia JM, Martinex-<br />

Rodenas F, Gil MJ, Segura M, Sancho<br />

J, Sitges-Serra A: <strong>Treatment</strong> of<br />

27 postoperative enterocutaneous<br />

fistulas with long half-life somatostatin<br />

analogue SMS 201-995. Ann<br />

Surg 1989;210/1:56–58.<br />

46 Nubiola-Calonge P, Badia JM, Sancho<br />

J, Gil MJ, Segura M, Sitges-Serra<br />

A: Blind evaluation of the effect<br />

of octreotide (SMS 201-995) on<br />

small bowel fistulae output. Lancet<br />

1987;ii:672–674.<br />

47 Scott NA, Finnegan S, Irving MH:<br />

Octreotide <strong>and</strong> postoperative enterocutaneous<br />

fistulae: A controlled<br />

prospective study. Acta Gastroenterol<br />

Belg 1993;53/3–4:266–270.<br />

48 Sancho JJ, di Costanzo J, Nubiola P,<br />

Larrad A, Beguiristain A, Roqueta<br />

F, Franch G, Oliva A, Gubern JM,<br />

Sitges-Serra A: R<strong>and</strong>omized doubleblind<br />

placebo-controlled trial of early<br />

octreotide in patients with postoperative<br />

enterocutaneous fistula. Br J<br />

Surg 1995;82:638–641.<br />

49 Jenkins SA, Nott DM, Baxter JN:<br />

Fluctuations in the secretion of pancreatic<br />

enzymes between consecutive<br />

doses of octreotide: Implications<br />

<strong>for</strong> the management of fistulae.<br />

Eur J Gastroenterol Hepatol 1995;7/<br />

3:255–258.<br />

50 Hern<strong>and</strong>ez-Ar<strong>and</strong>a JC, Gallo-Chico<br />

B, Fores-Ramirez LA, Avalos-<br />

Huante R, Magos-Vazquez FJ,<br />

Ramirez-Barba EJ: <strong>Treatment</strong> of enterocutaneous<br />

fistula with or without<br />

octreotide <strong>and</strong> parenteral nutrition.<br />

Nutr Hosp 1996;11/4:226–<br />

229.<br />

51 Bartlett DL, Charl<strong>and</strong> SL, Torosian<br />

MH: Reversal of tumor-associated<br />

hyperglucagonemia as treatment <strong>for</strong><br />

cancer cachexia. Surgery 1995;118/<br />

1:87–97.<br />

52 Cairns W, Malone R: Octreotide as<br />

an agent <strong>for</strong> the relief of malignant<br />

ascites in palliative care patients.<br />

Palliat Med 1999;13:429–430.<br />

53 Smith S, Anthony L, Roberts LJ,<br />

Oates JA, Pincus T: Resolution of<br />

musculoskeletal symptoms in the<br />

carcinoid syndrome after treatment<br />

with the somatostatin analogue octreotide.<br />

Ann Intern Med 1990;112:<br />

66–68.<br />

54 Londong W, Angerer M, Kutz K,<br />

L<strong>and</strong>graf R, Londong V: Diminishing<br />

efficacy of octreotide (SMS<br />

201-995) on gastric functions of<br />

healthy subjects during one week<br />

administration. Gastroenterology<br />

1989;96:713–722.<br />

Chemotherapy 2001;47(suppl 2):54–61 61


Chemotherapy 2001;47(suppl 2):62–77<br />

Management of Breast <strong>Cancer</strong>:<br />

Is There a Role <strong>for</strong> <strong>Somatostatin</strong> <strong>and</strong><br />

Its <strong>Analogs</strong>?<br />

Francesco Boccardo Domenico Amoroso<br />

Academic Unit of Medical Oncology, National <strong>Cancer</strong> Institute, Genoa, Italy<br />

Key Words<br />

<strong>Somatostatin</strong> W <strong>Somatostatin</strong> analogs W<br />

Breast cancer<br />

Abstract<br />

<strong>Somatostatin</strong> <strong>and</strong> related peptides are a<br />

family of peptides which are ubiquitous <strong>and</strong><br />

function as endogenous growth inhibitors.<br />

<strong>Analogs</strong> have been developed through the<br />

introduction of a D-amino acid in the position<br />

8 of somatostatin moiety which is more resistant<br />

to the action of endogenous peptidases<br />

than the parental moiety. Both somatostatin<br />

<strong>and</strong> its analogs interact with specific receptors<br />

on the cell surface. The five receptor subtypes,<br />

SSTR-1 to SSTR-5, which have been<br />

characterized so far, have a different affinity<br />

<strong>for</strong> somatostatin <strong>and</strong> its analogs. This <strong>and</strong><br />

the fact that receptors are not homogeneously<br />

expressed in tissues account <strong>for</strong><br />

the different activity of these compounds, all<br />

of which have demonstrated tumoristatic<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0062$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

properties both in vitro <strong>and</strong> in vivo. The interaction<br />

of somatostatin <strong>and</strong> of somatostatin<br />

analogs with specific SSTR receptors is crucial<br />

to the antiproliferative mechanisms exerted<br />

by these compounds in vitro <strong>and</strong> in<br />

some animal models <strong>and</strong> the various pathways<br />

have been reviewed in detail. However,<br />

inhibition of angiogenesis <strong>and</strong> suppression<br />

of lactogenic hormones might represent alternative<br />

mechanisms, in particular in breast<br />

cancer. The rationale <strong>for</strong> the use of somatostatin<br />

<strong>and</strong> its analogs in breast cancer patients<br />

<strong>and</strong> to combine these peptides with<br />

antihormones, like antiestrogens or prolactin-lowering<br />

drugs, or cytotoxics has been<br />

reviewed together with the results obtained<br />

in phase II <strong>and</strong> comparative trials. The reasons<br />

<strong>for</strong> the limited efficacy shown by these<br />

compounds either when used alone or when<br />

used in combination with other drugs have<br />

also been critically reviewed in the perspective<br />

of new trials.<br />

Copyright © 2001 S. Karger AG, Basel<br />

Prof. F. Boccardo, Cattedra e UO Universitaria di Oncologia Medica<br />

Istituto Nazionale per la Ricerca sul Cancro, Largo R. Benzi, 10<br />

I–16132 Genova (Italy)<br />

Tel. +39 010 5600503, Fax +39 010 352753<br />

E-Mail boccardo@hp380.ist.unige.it


Introduction<br />

Breast cancer is still a leading health issue<br />

in women. In fact, this disease is the most<br />

common cancer affecting women in developed<br />

countries <strong>and</strong> a major cause of cancer<br />

death [1, 2]. The incidence of breast cancer<br />

has gradually increased during the past decades<br />

in parallel with an increase in the ageing<br />

of the population. It was estimated that in<br />

1997 about 180,000 new cases of breast cancer<br />

were diagnosed in the USA <strong>and</strong> that about<br />

44,000 women died of this disease [3]. However,<br />

in many countries incidence <strong>and</strong> mortality<br />

have recently levelled off or even decreased<br />

[4–7], as a consequence of both early<br />

diagnosis <strong>and</strong> the introduction of more effective<br />

treatments.<br />

Strategies to control breast cancer include:<br />

(1) primary prevention, to decrease the incidence<br />

of this disease, (2) early diagnosis, to<br />

ensure that more patients might be diagnosed<br />

with curable disease, <strong>and</strong> (3) improvement of<br />

treatment, to increase cure rates. Primary prevention<br />

is not possible at present due to the<br />

multifactorial etiopathogenesis of this disease.<br />

However, increasing knowledge of the<br />

genetic alterations which can trigger the neoplastic<br />

trans<strong>for</strong>mation of breast cells <strong>and</strong> the<br />

recognition of the molecular events which<br />

control the growth of trans<strong>for</strong>med cells have<br />

provided new targets <strong>for</strong> therapeutic <strong>and</strong> chemopreventive<br />

approaches [8–10]. Tamoxifen<br />

has recently been proven to be effective in<br />

decreasing the incidence of breast cancer in<br />

one study [11]. However, the lack of effectiveness<br />

in two other studies [12, 13] <strong>and</strong> the side<br />

effects produced by the long-term administration<br />

of this antiestrogen, including endometrial<br />

cancer <strong>and</strong> thromboembolic events, still<br />

question its role as a chemopreventive agent<br />

in healthy women. It is now recognized that<br />

mammographic screening results in a 25–<br />

30% decrease in the risk of women aged 50<br />

years or over dying of breast cancer [14].<br />

However, the role of mammographic screening<br />

in women between the ages of 40 <strong>and</strong> 50<br />

years is still unclear. <strong>Treatment</strong> of patients<br />

with primary breast cancer involves multiple<br />

treatment modalities, including surgery, radiotherapy<br />

<strong>and</strong> systemic therapy with combination<br />

chemotherapy, hormonal agents or<br />

both. Based on the assumption that primary<br />

breast cancer is a systemic disease in which<br />

subclinical metastases are already present at<br />

diagnosis in most patients [15], systemic adjuvant<br />

therapy has been extensively studied in<br />

several individual r<strong>and</strong>omized trials. To assess<br />

a reliable fallout from these trials, partially<br />

avoiding the biases intrinsic to each study,<br />

the statistical technique of metanalysis was<br />

employed to add value to observed or expected<br />

events from each trial <strong>and</strong> to provide<br />

an estimate of the overall effect of treatment<br />

[16]. Over the past decade there have been<br />

four metanalyses of data from prospective<br />

r<strong>and</strong>omized trials, involving more than<br />

75,000 women treated with different adjuvant<br />

therapies. Current knowledge comes<br />

from the results of these metanalyses <strong>and</strong>, in<br />

particular, from the most recent of them.<br />

These metanalyses showed: that combination<br />

chemotherapy is able to reduce the annual<br />

risk of death by about 20%, especially in<br />

women younger than 50 [17], that prolonged<br />

treatment with tamoxifen is able to reduce to<br />

a similar extent the annual risk of death in all<br />

age groups, especially in women with estrogen-receptor-positive<br />

tumors [18], <strong>and</strong> that<br />

ovarian ablation appears to produce results<br />

comparable to those obtained by tamoxifen or<br />

chemotherapy in younger women [19]. Because<br />

there was enough evidence to suggest<br />

that the combination of chemotherapy with<br />

tamoxifen (or ovarian ablation in premenopausal<br />

women) might be even more effective<br />

than either treatment modality alone, this<br />

treatment option was included among the rec-<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 63


ommendations <strong>for</strong> women with a higher risk<br />

of relapse generated by a panel of experts at a<br />

conference on breast cancer held in St. Gallen<br />

in 1998 [20]. In spite of more recent results of<br />

adjuvant therapy, at least 50% of women with<br />

primary breast cancer ultimately develop distant<br />

metastases <strong>and</strong> die of their disease, the<br />

risk being strictly correlated with the pathological<br />

status of axillary lymph nodes [21].<br />

Once metastatic disease becomes overt,<br />

treatment strategy should be realistically addressed<br />

to achieve the best palliative results,<br />

through a wise use of chemotherapy <strong>and</strong> endocrine<br />

therapy. Endocrine manipulations represent<br />

the oldest <strong>for</strong>m of systemic therapy<br />

<strong>for</strong> advanced breast cancer. Overall, approximately<br />

one third of unselected patients with<br />

metastatic breast cancer respond to these manipulations<br />

[22]. However, the discovery of<br />

steroid receptors contributed a lot to the selection<br />

of patients most likely to benefit from<br />

endocrine maneuvers <strong>and</strong> 50–75% objective<br />

response rates are usually reported in estrogen-<br />

or progesterone receptor-positive patients<br />

[23]. Tamoxifen has become the most<br />

widely used hormone therapy <strong>for</strong> advanced<br />

breast cancer, particularly after the menopause.<br />

However, other types of endocrine<br />

therapies, such as gonadal ablation in premenopausal<br />

women <strong>and</strong> progestins <strong>and</strong> aromatase<br />

inhibitors in postmenopausal women,<br />

may also provide considerable disease remission<br />

[22], especially in patients who progress<br />

after first-line treatment with tamoxifen [24].<br />

New hormonal compounds, including second<br />

generation steroidal <strong>and</strong> nonsteroidal aromatase<br />

inhibitors which are now the treatment of<br />

choice <strong>for</strong> patients who relapse following<br />

front-line or adjuvant treatment with antiestrogens,<br />

are now being evaluated as an alternative<br />

to tamoxifen <strong>and</strong> are c<strong>and</strong>idates <strong>for</strong><br />

becoming an alternative to this agent even in<br />

the adjuvant setting [24]. Finally, new steroidal<br />

antiestrogens are now being evaluated<br />

either in patients failing first-line tamoxifen<br />

treatment or as a first-line treatment option,<br />

in view of their non-cross-resistance with triphenylethylene<br />

derivatives <strong>and</strong> lack of the<br />

pure agonistic effects exerted by these compounds<br />

[25].<br />

In spite of the availability of so many therapeutic<br />

choices, the search <strong>for</strong> new compounds<br />

is still crucial <strong>and</strong> is encouraged by the<br />

new insight into the molecular biology of the<br />

neoplastic cell, with special regard to the different<br />

types of receptors, which are expressed<br />

on the surface of breast cancer cells <strong>and</strong> which<br />

interact with many signalling pathways. Since<br />

most breast cancers express a variety of somatostatin<br />

receptors, the therapeutic potential<br />

of somatostatin <strong>and</strong> its analogs has been<br />

investigated in breast cancer patients. Preclinical<br />

<strong>and</strong> clinical findings accumulated so far<br />

are the objective of this review.<br />

Structure <strong>and</strong> Function of<br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong><br />

<strong>Somatostatin</strong> <strong>and</strong> somatostatin-related<br />

peptides are a family of peptides that include<br />

two important products, somatostatin-14<br />

(SS14) <strong>and</strong> somatostatin-28 (SS28), a number<br />

of species-specific variants <strong>and</strong> even more numerous<br />

prehormone <strong>for</strong>ms. SS14 <strong>and</strong> SS28<br />

are mainly found in the gut <strong>and</strong> in various<br />

exocrine <strong>and</strong> endocrine gl<strong>and</strong>s throughout the<br />

body. However, they are ubiquitous <strong>and</strong> can<br />

be found also in the nervous system, specifically<br />

in the hypothalamus, limbic system,<br />

brain stem <strong>and</strong> spinal cord [26]. <strong>Somatostatin</strong><br />

has a broad spectrum of biological actions <strong>and</strong><br />

exerts suppressive effects on a large variety of<br />

cells, functioning as an endogenous growth<br />

inhibitor [26].<br />

Naturally occurring peptides have a plasma<br />

half-life of less than 3 min, since they are<br />

rapidly inactivated by endogenous peptidases<br />

64 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso


[27]. Therapeutic application of the native<br />

peptide has, there<strong>for</strong>e, been limited to those<br />

conditions where the use through intravenous<br />

continuous infusion is appropriate. There<strong>for</strong>e,<br />

many ef<strong>for</strong>ts have been made to develop<br />

more stable peptides. Incorporation of D-amino<br />

acids into the somatostatin backbone succeeded<br />

in slowing down enzymatic degradation<br />

[28]. In particular, the introduction of<br />

a D-amino acid after tryptophan in the position<br />

8 of the moiety led to the development of<br />

analogs with an enhanced GH inhibitory potency<br />

[29]. There<strong>for</strong>e, the three more extensively<br />

tested analogs, i.e. SMS 201–995 (octreotide),<br />

BIM 23014 (lanreotide) <strong>and</strong> RC-<br />

160 (vapreotide), are octapeptides [29].<br />

Octreotide is 45 times more potent in vitro<br />

than naive somatostatin in inhibiting the release<br />

of GH <strong>and</strong> 11 times more potent in<br />

inhibiting the release of glucagon. This analog<br />

is also more potent in inhibiting insulin secretion<br />

[30].<br />

Attempts to develop an analog with minimal<br />

effects on insulin <strong>and</strong> glucagon secretion,<br />

but a more profound effect on the release of<br />

GH <strong>and</strong> of GH-dependent growth factors,<br />

such as insulin-like growth factors (IGF), resulted<br />

in the development of lanreotide. Both<br />

octreotide <strong>and</strong> lanreotide are available as<br />

slow-release <strong>for</strong>mulations obtained through<br />

the microencapsulation of the active drug<br />

in a matrix of polylactide-glycolide microspheres.<br />

These polymers are completely biodegradable<br />

<strong>and</strong> are commonly administered<br />

at 2- to 4-week intervals [26].<br />

Mode of Action of <strong>Somatostatin</strong><br />

<strong>and</strong> Its <strong>Analogs</strong>: Overview of<br />

Receptor Functions<br />

Critical to somatostatin action is the presence<br />

of somatostatin receptors, which principally<br />

have two distinct functions: (1) to bind<br />

the lig<strong>and</strong> with high affinity <strong>and</strong> (2) to produce<br />

a transmembrane signal evoking a biological<br />

response. Up to five somatostatin receptor<br />

subtypes, SSTR-1 to SSTR-5, have<br />

been cloned <strong>and</strong> functionally characterized<br />

[31–35]. They all bind SS14 <strong>and</strong> SS28 with<br />

similar affinity but show major differences in<br />

the affinity <strong>for</strong> different somatostatin analogs<br />

[31, 34, 35]. Octreotide <strong>and</strong> vapreotide have a<br />

low affinity <strong>for</strong> SSTR-1, but have a high affinity<br />

<strong>for</strong> SSTR-2. Both of them can inhibit the<br />

proliferation of cells expressing SSTR-2 [32,<br />

34, 35]. Vapreotide <strong>and</strong> octreotide have also a<br />

moderate to high affinity <strong>for</strong> SSTR-3 <strong>and</strong><br />

SSTR-5 [32, 34]. In addition, vapreotide has a<br />

moderate affinity <strong>for</strong> SSTR-4 [35]. Noteworthy,<br />

the specificity of these analogs depends<br />

on their different affinity <strong>for</strong> receptor subtypes<br />

<strong>and</strong> on receptor heterogeneity on the<br />

surface of target cells. Indeed, there is a considerable<br />

heterogeneity between tumors <strong>and</strong><br />

within the same tumor with respect to the<br />

density of somatostatin binding sites, higher<br />

receptor levels having been found in more differentiated<br />

tumors [36]. This implies the possibility<br />

<strong>for</strong> somatostatin receptors to represent<br />

differentiation markers, at least in certain<br />

neoplastic lineages <strong>and</strong> <strong>for</strong> their loss to<br />

be implicated in tumor progression. After<br />

the chronic administration of somatostatin,<br />

downregulation of receptors can occur, as<br />

shown by studies per<strong>for</strong>med in pituitary cell<br />

lines [37]. However, this effect is not the rule.<br />

In vivo studies indicate that continuous treatment<br />

of acromegalic patients with octreotide<br />

does not desensitize the cellular responsiveness.<br />

Moreover, there is evidence that changes<br />

in gene expression or mRNA stability, rather<br />

than in receptor affinity, can be involved in<br />

both up- <strong>and</strong> downregulation of somatostatin<br />

receptors [39]. This mechanism is common<br />

with other peptides, i.e. LHRH analogs, <strong>and</strong><br />

results in the occupation of a high proportion<br />

of somatostatin receptors which are inter-<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 65


nalized into the cell <strong>and</strong> maintained so until<br />

treatment with somatostatin is continued.<br />

However, an immediate rebound effect has<br />

been observed on somatostatin discontinuation,<br />

which is associated with a rapid rise in<br />

target hormones [40]. Slow release somatostatin<br />

analogs can also be ineffective in maintaining<br />

the tissue levels necessary to downregulate<br />

the receptors <strong>and</strong> GH escape can occur<br />

in some patients if dosages are not appropriate<br />

[26].<br />

After binding with somatostatin, receptor<br />

activation occurs. This in turn is associated<br />

with several transmembrane signalling pathways<br />

via pertussis toxin-sensitive guanine nucleotide-binding<br />

proteins [41]. In particular,<br />

the activation of the receptors is associated<br />

with a prompt reduction in the intracellular<br />

level of mediators, like cAMP <strong>and</strong> Ca 2+ , due<br />

to the effects on membrane adenylyl cyclase<br />

<strong>and</strong> ion (K + , Ca 2+ ) channels [42, 43]. However,<br />

this proximal effect can explain only in<br />

part the inhibitory action of somatostatin on<br />

hormone secretion <strong>and</strong> more distal effects of<br />

the peptide in this respect have also been documented<br />

[44–46]. Another effector system,<br />

i.e. tyrosine phosphatase, has recently been<br />

shown to be stimulated by somatostatin receptor<br />

activation, leading to dephosphorylation<br />

<strong>and</strong> inactivation of the EGF receptor<br />

[47]. Emerging in<strong>for</strong>mation regarding signal<br />

transduction pathways related to somatostatin<br />

receptor activation is crucial to underst<strong>and</strong>ing<br />

the possible mechanisms of action of<br />

somatostatin <strong>and</strong> of its analogs as antiproliferative<br />

agents. The transfection of the five<br />

cloned human somatostatin receptors into<br />

somatostatin receptor-negative cell lines was<br />

shown to induce specific signal transduction<br />

pathways associated with individual somatostatin<br />

receptor subtypes. For instance, it has<br />

been demonstrated that SSTR-3-transfected<br />

CHO cells respond to octreotide with the<br />

upregulation of p53 <strong>and</strong> the subsequent in-<br />

duction of apoptosis [48]. However, the presence<br />

of a particular receptor subtype on the<br />

tumor cell does not guarantee that its binding<br />

with the lig<strong>and</strong> is followed by the activation of<br />

antiproliferative or apoptotic pathways. For<br />

instance, octreotide can be ineffective in inducing<br />

apoptosis through its binding with<br />

SSTR-3 in tumors with a mutation in the p53<br />

gene [48]. In other experimental systems, a<br />

signal transduction pathway associated with<br />

SSTR-2 implying the upregulation of phosphoprotein<br />

phosphatase activity has been<br />

demonstrated [34, 47, 49]. The activation of<br />

this pathway is supposed to represent one of<br />

the possible mechanisms through which somatostatin<br />

analogs exert their antineoplastic<br />

effects. In fact, the blockade of phosphoprotein<br />

phosphatase stimulates cellular proliferation<br />

[50], enhancing the consequences of the<br />

binding of growth factors, like EGF or IGF-1,<br />

to their receptors.<br />

Mechanisms of Antitumor Action in<br />

Breast <strong>Cancer</strong> of <strong>Somatostatin</strong><br />

<strong>and</strong> Its <strong>Analogs</strong><br />

While estradiol is recognized as the predominant<br />

hormone stimulating the growth of<br />

human breast cancer, there is also evidence<br />

that lactotrophic hormones (GH <strong>and</strong> prolactin)<br />

are involved in the growth of breast<br />

tumors, particularly in murine models [51].<br />

Indeed, prolactin receptors can be demonstrated<br />

in 20–50% of human breast cancer<br />

cells <strong>and</strong> at least some MCF-7 cell clones have<br />

been shown to be prolactin-dependent [52,<br />

53]. Inhibition of the release of lactotrophic<br />

hormones may, there<strong>for</strong>e, be one of the mechanisms<br />

through which somatostatin <strong>and</strong> its<br />

analogs could inhibit breast cancer growth.<br />

There is a substantial literature demonstrating<br />

considerable antineoplastic activity<br />

of somatostatin <strong>and</strong> its analogs in many in<br />

66 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso


vitro <strong>and</strong> in vivo experimental systems [54,<br />

55] <strong>and</strong> increasing in<strong>for</strong>mation concerning<br />

the antiproliferative effects of these compounds<br />

has emerged over the past decade.<br />

Several mechanisms have been proposed,<br />

<strong>and</strong> it is important to note that they are not<br />

mutually exclusive. The direct mechanisms<br />

refer to the inhibition of proliferation <strong>and</strong>/or<br />

the induction of apoptosis, as a consequence<br />

of the binding of somatostatin <strong>and</strong> of its analogs<br />

to specific receptors on the target neoplastic<br />

cell, as reported above. The presence<br />

of somatostatin receptors seems to be the<br />

major determinant of the antiproliferative<br />

activity of somatostatin analogs in animal<br />

models [34, 47–49, 56]. Thus, a strong antiproliferative<br />

effect of these analogs has been<br />

demonstrated in mice bearing both estrogendependent<br />

<strong>and</strong> estrogen-insensitive MXT<br />

transplantable tumors that express somatostatin<br />

receptors [27]. Conversely, no activity<br />

was shown on DMBA-induced mammary<br />

carcinoma of the rat that does not express<br />

somatostatin receptors [27]. It is worth noting<br />

that up to 60% of human breast cancers<br />

were found to express somatostatin receptors,<br />

although they may not always be homogeneously<br />

distributed in the whole tissue<br />

sample. The specificity of action of various<br />

analogs through their binding with specific<br />

type of receptor is a subject of ongoing investigation,<br />

although recent data suggest that<br />

the antiproliferative effects of somatostatin<br />

against breast cancer are mostly mediated by<br />

receptor subtypes 2 <strong>and</strong> 5 [58].<br />

<strong>Somatostatin</strong> receptors are frequently expressed<br />

in tumors with a high estrogen <strong>and</strong><br />

progesterone receptor content [59], while in<br />

contrast a negative relationship was found<br />

between their expression <strong>and</strong> EGF receptors<br />

[60], which are regarded as poor prognostic<br />

indicators [60].<br />

The indirect mechanisms of action of somatostatin<br />

<strong>and</strong> of its analogs are a direct<br />

consequence of the systemic effects of these<br />

compounds <strong>and</strong> even somatostatin receptornegative<br />

tumors might be inhibited through<br />

these mechanisms. The indirect mechanism<br />

that has received the greatest attention to date<br />

concerns the inhibitory effect of somatostatin<br />

analogs on the IGF family (IGF-1 <strong>and</strong><br />

IGF-2).<br />

IGF-1 <strong>and</strong> IGF-2 are peptides with structural<br />

similarities to insulin. IGF-1 is involved<br />

in normal growth <strong>and</strong> development processes,<br />

while the physiological function of IGF-2 remains<br />

unclear, although it appears to be essential<br />

<strong>for</strong> normal fetal growth. IGFs induce<br />

mitogenesis in cultured fibroblasts, chondroblasts,<br />

osteoblasts, neuroglial cells <strong>and</strong> erythroid<br />

progenitor cells. These peptides also<br />

appear to be involved in the control of cell<br />

proliferation in various malignant phenotypes<br />

[61–64], including breast cancer. The proliferative<br />

effects of IGF-1 <strong>and</strong> IGF-2 are mediated<br />

by IGF-1 receptors, which have been detected<br />

in 67–93% of human breast tumors [65–68].<br />

Modulation of IGF-1 <strong>and</strong> IGF-2 activity occurs<br />

through the interaction with circulating<br />

IGF-binding proteins. However, also the GH-<br />

IGF-1 axis appears to have an important<br />

influence on the biological behavior of many<br />

common neoplasms. The suppressive effect of<br />

somatostatin analogs on serum IGF-1 levels<br />

might be related either to the direct inhibition<br />

of IGF-1 gene expression or to the suppression<br />

of GH-dependent IGF-1 synthesis in the<br />

liver [69, 70]. The direct suppressive action<br />

on IGF-1 gene expression remains incompletely<br />

characterized <strong>and</strong> it is possible that<br />

both mechanisms (i.e. the suppression of IGF-<br />

1 gene expression <strong>and</strong> the suppression of IGF-<br />

1 synthesis in the liver) might contribute to<br />

somatostatin antiproliferative activity. Moreover,<br />

somatostatin analogs have been found<br />

to stimulate the secretion of certain IGFbinding<br />

proteins, an action which has been<br />

proposed to attenuate IGF-1 bioactivity<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 67


independently of the suppressive effect on<br />

IGF-1 levels [71, 72]. Since IGF-1 is recognized<br />

as a potent antiapoptotic factor [73, 74],<br />

the antiproliferative effects of somatostatin<br />

analogs could also be mediated by the suppressive<br />

effects on IGF-1 gene expression,<br />

which in turn could enhance the antiapoptotic<br />

effects exerted by these compounds in experimental<br />

models [75, 76].<br />

The interference with blood flow <strong>and</strong> nutritional<br />

support to the tumor could be a third<br />

mechanism through which somatostatin <strong>and</strong><br />

its analogs could exert their antitumor properties.<br />

In fact, somatostatin receptors, namely<br />

the SSTR-2 subtype, have been demonstrated<br />

in the peritumoral veins of various human<br />

cancers, including breast tumors [77] <strong>and</strong> it<br />

appears that the increased expression of somatostatin<br />

receptors in peritumoral vessels in<br />

comparison to normal, nontumoral, tissue,<br />

might be related to the neoplastic process<br />

itself [77].<br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong>:<br />

In vitro <strong>and</strong> in vivo Studies<br />

Inhibition of Tumor Growth with<br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong><br />

in Combination with Antiestrogens<br />

Based on the data mentioned above, it can<br />

be argued that somatostatin <strong>and</strong> its analogs<br />

might have some effect on the proliferation of<br />

breast cancer. However, unopposed estrogen<br />

action can be greater than the growth inhibitory<br />

effects of somatostatin <strong>and</strong> its analogs [78].<br />

The molecular basis <strong>for</strong> the attenuation of this<br />

antiproliferative effect has not been elucidated.<br />

However, consistent with this effect is<br />

the observation that antitumor activity of octreotide<br />

in the MXT breast tumor model was<br />

enhanced by the coadministration of an<br />

LHRH analog, which lowers estradiol levels<br />

[79]. Moreover, the activity of octreotide<br />

was maximized in the absence of estrogens<br />

[78]. Such considerations provided the rationale<br />

<strong>for</strong> the concomitant administration of<br />

somatostatin or of its analogs with other hormonal<br />

agents, namely antiestrogens. The capability<br />

of tamoxifen to suppress the GH-IGF<br />

axis through the suppression of IGF-1 gene<br />

expression <strong>and</strong> of serum IGF-1 levels [80–82]<br />

provided a further rationale <strong>for</strong> this combination<br />

therapy.<br />

Several preclinical <strong>and</strong> clinical observations<br />

have suggested an additive biological<br />

<strong>and</strong> antitumor effect by combining somatostatin<br />

analogs <strong>and</strong> tamoxifen. In short-term<br />

experiments in rats, the concomitant administration<br />

of octreotide with tamoxifen suppresses<br />

serum IGF-1 levels <strong>and</strong> IGF-1 gene<br />

expression more potently than either agent<br />

alone [83]. This combination has also been<br />

evaluated in the DMBA-induced rat mammary<br />

tumor model [84]. These experiments<br />

showed that the development <strong>and</strong> the volume<br />

of DMBA-induced tumors were significantly<br />

reduced in the animals treated with both<br />

agents as compared to the animals treated<br />

with either agent alone. In clinical studies, an<br />

enhanced suppression of serum IGF-1 levels<br />

in patients receiving a combination of octreotide<br />

or lanreotide with tamoxifen has also<br />

been demonstrated [85].<br />

It is noteworthy that in all experimental<br />

systems, tumor response to the combination<br />

of octreotide <strong>and</strong> tamoxifen was greater in<br />

smaller than in larger tumors, i.e. at an earlier<br />

stage of the disease, when the angiogenetic<br />

response to specific peptides released by the<br />

cancer cells is higher [86–88]. Since it has<br />

been suggested that part of the antitumor<br />

activity of somatostatin might be related to its<br />

inhibition of angiogenesis, these experiments<br />

have anticipated that somatostatin <strong>and</strong> its<br />

analogs would be more effective at an earlier<br />

stage of the disease in humans as well, <strong>for</strong><br />

instance by preventing the growth of micro-<br />

68 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso


metastases following surgery, <strong>and</strong> that the<br />

therapeutic effect might be much more limited<br />

in patients with overt metastatic disease.<br />

Inhibition of Tumor Growth with<br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong> in<br />

Combination with Chemotherapeutic<br />

Agents<br />

In order to augment the effects of chemotherapy,<br />

a combination of hormonal therapy<br />

<strong>and</strong> cytotoxics has been proposed. Such combination<br />

therapies are aimed at achieving additive<br />

or synergistic antitumor effects while<br />

reducing the incidence <strong>and</strong> the severity of side<br />

effects. The biological rationale <strong>for</strong> this <strong>for</strong>m<br />

of treatment is tumor cell heterogeneity, since<br />

breast cancer consists of various populations<br />

of cells with different sensitivities to cytotoxic<br />

<strong>and</strong> hormonal agents. The modulatory effect<br />

of somatostatin analogs in combination with<br />

different chemotherapeutic agents has been<br />

studied in different preclinical studies [89].<br />

Among the agents tested, combinations of mitomycin<br />

C, doxorubicin, Taxol <strong>and</strong> 5-fluorouracil<br />

have been investigated more extensively<br />

than other cytotoxics in a number of models.<br />

Thus, in AR42J cells, which express the<br />

somatostatin receptor subtype 2 <strong>and</strong> whose<br />

growth has been shown to be inhibited by<br />

octreotide [90], both mitomycin C <strong>and</strong> Taxol<br />

exerted antiproliferative effects that appeared<br />

to be synergistically enhanced by octreotide<br />

[89]. Both additive <strong>and</strong> synergistic interactions<br />

between octreotide <strong>and</strong> 5-fluorouracil<br />

were shown, depending on the 5-fluorouracil<br />

concentration used. This type of interaction<br />

was also observed with other combinations of<br />

octreotide <strong>and</strong> cytotoxics [91].<br />

The combination of doxorubicin <strong>and</strong> octreotide<br />

resulted in a clear dose-dependent<br />

synergistic interaction on AR42J cells [89].<br />

Moreover, it has been demonstrated in vitro<br />

that doxorubicin accumulation in MCF-7<br />

cells treated with octreotide was increased 3-<br />

to 4-fold [92]. This observation suggests that<br />

octreotide modulates the uptake of doxorubicin<br />

<strong>and</strong>/or interferes with the activity of P-170<br />

glycoprotein responsible <strong>for</strong> multidrug resistance.<br />

The combination of chemotherapy with<br />

somatostatin analogs also received attention<br />

because of the demonstrated ability of octreotide<br />

to reduce gastrointestinal toxicity associated<br />

with some cytotoxic agents [93].<br />

However, the evidence concerning the biological<br />

<strong>and</strong> clinical effects of combining somatostatin<br />

analogs <strong>and</strong> chemotherapeutic<br />

agents is still limited <strong>and</strong> does not make it<br />

possible to draw any conclusion about the<br />

potential role of such combinations in anticancer<br />

therapy.<br />

Clinical Studies<br />

The clinical experience with somatostatin<br />

or its analogs in advanced breast cancer is still<br />

limited, despite the promising findings observed<br />

in experimental studies which clearly<br />

indicate inhibition of tumor growth.<br />

The majority of clinical studies have been<br />

carried out in patients pretreated with a variety<br />

of therapies <strong>and</strong> combining somatostatin<br />

or somatostatin analogs with tamoxifen or<br />

with prolactin-lowering drugs such as bromocriptine<br />

(table 1).<br />

Vennin et al. [94] treated 16 postmenopausal<br />

patients with 200 Ìg octreotide daily<br />

<strong>for</strong> at least 30 days <strong>and</strong> observed a disease stabilization<br />

in 3, <strong>and</strong> a decrease of IGF-1 levels<br />

by 33% in 8. Manni et al. [95] treated 10 postmenopausal<br />

breast cancer patients with a<br />

combination of octreotide (200–400 Ìg daily)<br />

<strong>and</strong> bromocriptine (5 mg/day) <strong>and</strong> observed<br />

disease stabilization in 1 patient. Interestingly,<br />

the baseline levels of IGF-1 declined in 6<br />

out of 9 women. Moreover, following provocative<br />

tests, GH levels were suppressed in 7<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 69


Table 1. Phase II studies with somatostatin <strong>and</strong> its analogs alone or in combination with<br />

other hormonal agents<br />

Authors <strong>Treatment</strong> Dose/schedule Patients OR<br />

Vennin et al. [94] Octreotide 200 Ìg/day 16 3 SD<br />

Manni et al. [95] Octreotide <strong>and</strong> 200–400 Ìg/day<br />

1 0 1 SD<br />

bromocriptine 5 mg/day<br />

Stolfi et al. [96] Octreotide 750 Ìg <strong>for</strong> 10 days, 1<br />

then 500 Ìg <strong>for</strong> 5 days<br />

0 3 PR<br />

Anderson et al. [97] Octreotide <strong>and</strong> 200–400 Ìg/day<br />

6 4 SD<br />

bromocriptine 2.5–5 mg/day<br />

Canobbio et al. [98] Lanreotide <strong>and</strong> 20–30 mg/2 weeks 36 4 CR<br />

tamoxifen 30 mg/day<br />

12PR<br />

Di Leo et al. [99] Lanreotide 30 mg/2 weeks 10 None<br />

O’Byrne et al. [100] Vapreotide 3 mg/day, then<br />

1<br />

4.5 mg/day, then 6 mg/day<br />

4 None<br />

OR = Objective response; SD = stable disease; PR = partial response; CR = complete<br />

response.<br />

<strong>and</strong> prolactin levels in 8 of 9 patients in whom<br />

these assessments were possible. Three patients<br />

experienced nausea <strong>and</strong> 1 of them had<br />

to discontinue treatment. Stolfi et al. [96]<br />

treated 10 patients with octreotide given by<br />

intravenous infusion <strong>for</strong> 10 days at a dose of<br />

750 Ìg t.i.d. <strong>and</strong> then intramuscularly <strong>for</strong> a<br />

further 5 days at the dose of 500 Ìg b.i.d. They<br />

obtained a partial response in 3 patients.<br />

Moreover, a marked reduction in edema, cyanosis<br />

<strong>and</strong> bleeding from ulcerated tumor lesions<br />

was noted in most of the treated patients.<br />

Anderson et al. [97] treated 6 patients with<br />

octreotide (200–400 Ìg daily) <strong>and</strong> bromocriptine<br />

(2.5–5 mg/day), <strong>and</strong> observed a disease<br />

stabilization in 4. Suppression of both prolactin<br />

<strong>and</strong> IGF-1 levels were observed in all<br />

patients during the entire treatment period.<br />

Canobbio et al. [98] treated 36 postmenopausal<br />

patients with locally advanced or metastatic<br />

breast cancer with lanreotide (one 20to<br />

30-mg slow-release vial i.m. every 2 weeks)<br />

<strong>and</strong> tamoxifen (30 mg/day). None of them<br />

had received prior treatment with hormone or<br />

chemotherapy. In 4 patients a complete response<br />

was observed <strong>and</strong> a partial response in<br />

12, with an overall response rate of 52% (95%<br />

confidence interval: 35–69%). Toxicity was<br />

generally mild <strong>and</strong> treatment was well tolerated,<br />

mild diarrhea being the most common<br />

side effect. No patient, however, had to discontinue<br />

the treatment because of toxicity. A<br />

significant decrease in IGF-1 levels was observed<br />

at 3 <strong>and</strong> 6 months in the majority of<br />

patients.<br />

Di Leo et al. [99] evaluated the activity of<br />

lanreotide (BIM 23014) in 10 women with<br />

advanced breast cancer. The drug was administered<br />

at a dose of 30 mg i.m. <strong>for</strong>tnightly. No<br />

objective response was observed, <strong>and</strong> GH <strong>and</strong><br />

IGF-1 serum levels were not adequately suppressed<br />

over time.<br />

O’Byrne et al. [100] treated 14 women with<br />

advanced breast cancer with the somatostatin<br />

analog vapreotide (RC-160) at a dose of<br />

3 mg/day in week 1 increasing to 4.5 mg/day<br />

during weeks 2– 4 <strong>and</strong> then to 6 mg/day thereafter<br />

by continuous subcutaneous infusion.<br />

70 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso


Table 2. Phase III studies with<br />

somatostatin analogs<br />

No objective tumor response was observed.<br />

However, a significant reduction in serum<br />

IGF-1 <strong>and</strong> prolactin levels was observed.<br />

In 1998 in Italy, the Health Minister promoted<br />

the evaluation of the antitumor effects<br />

of a therapeutic regimen containing somatostatin<br />

or octreotide, melatonin, bromocriptine<br />

<strong>and</strong> a retinoid mixture in eleven phase II<br />

trials, involving 386 patients with various tumors,<br />

including breast cancer. Cyclophosphamide<br />

was added <strong>for</strong> the treatment of certain<br />

types of cancer. The conclusions were quite<br />

discouraging since no evidence of efficacy was<br />

observed in any the studies per<strong>for</strong>med [101].<br />

In particular, no substantial therapeutic activity<br />

was observed in breast cancer patients.<br />

The promising findings achieved by somatostatin<br />

analogs alone or in combination<br />

with tamoxifen or bromocriptine in experimental<br />

models have not been confirmed as<br />

yet even in controlled trials (table 2). One<br />

hundred <strong>and</strong> thirty-five postmenopausal patients<br />

with metastatic breast cancer were r<strong>and</strong>omized<br />

to tamoxifen (20 mg/day) alone or<br />

combined with octreotide (150 Ìg s.c. twice<br />

daily) [102]. About 30% of the patients enrolled<br />

in this study were pretreated with chemotherapy<br />

<strong>and</strong> 7% of them had received prior<br />

Authors <strong>Treatment</strong> Patients OR<br />

%<br />

Ingle et al. [102] Tamoxifen (20 mg/day)<br />

Tamoxifen (20 mg/day plus<br />

135 49<br />

octreotide (150 Ìg twice/day) 43<br />

Bajetta et al. [103] Tamoxifen plus placebo 103 21<br />

Tamoxifen plus octreotide 20<br />

Bontenbal et al. [104] Tamoxifen<br />

Tamoxifen (40 mg/day) plus<br />

22 36<br />

octreotide (0.6 mg/day) <strong>and</strong> CV<br />

205-502 (75 Ìg/day)<br />

55<br />

OR = Objective response.<br />

treatment with tamoxifen. Although a significantly<br />

greater decline in serum IGF-1 levels<br />

was observed in the group of patients treated<br />

with the combination therapy, no differences<br />

were observed with respect to either progression-free<br />

survival or overall survival. The objective<br />

response rate was 49% in patients<br />

treated with tamoxifen alone <strong>and</strong> 43% in<br />

patients treated with tamoxifen <strong>and</strong> octreotide.<br />

Most patients treated with octreotide<br />

<strong>and</strong> tamoxifen experienced side effects, such<br />

as nausea, diarrhea <strong>and</strong> steatorrhea.<br />

Bajetta et al. [103] carried out a r<strong>and</strong>omized<br />

double-blind phase III trial in previously<br />

untreated metastatic breast cancer patients<br />

who were r<strong>and</strong>omly allocated to tamoxifen<br />

combined either with placebo or with octreotide.<br />

Un<strong>for</strong>tunately, drugs dosages were not<br />

reported. Two hundred <strong>and</strong> three patients<br />

were included. An overall tumor response rate<br />

of 20% was observed in the octreotide <strong>and</strong><br />

tamoxifen arm compared to 21% in the placebo<br />

<strong>and</strong> tamoxifen arm. No difference was<br />

observed between groups in median time to<br />

progression, while more patients in the octreotide<br />

group experienced adverse effects<br />

such as diarrhea <strong>and</strong> abdominal pain.<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 71


Table 3. Toxicity of somatostatin <strong>and</strong> its analogs<br />

[105]<br />

Local (at the site of injection)<br />

Pain<br />

Redness<br />

Abscess<br />

Systemic<br />

Gastrointestinal<br />

Nausea<br />

Abdominal cramps<br />

Diarrhea<br />

Steatorrhea<br />

Malabsorption of fat<br />

Flatulence<br />

Cholesterol gallstones<br />

Glucose metabolism<br />

Reduced glucose tolerance<br />

Hyperglycemia<br />

Hypoglycemia<br />

The feasibility <strong>and</strong> the endocrine <strong>and</strong> antitumor<br />

effects of a combination of octreotide<br />

(3 ! 0.2 mg s.c.) with tamoxifen (40 mg/day)<br />

<strong>and</strong> an antiprolactin drug (CV 205–502:<br />

75 Ìg/day) was studied by Bontenbal et al.<br />

[104]. They r<strong>and</strong>omized 22 metastatic breast<br />

cancer patients who were given either 40 mg/<br />

day of tamoxifen or the above-mentioned<br />

combination. An objective response was observed<br />

in 36% of the patients treated with<br />

tamoxifen alone <strong>and</strong> in 55% of the patients<br />

treated with the combination therapy. A significant<br />

decrease of plasma IGF-1 levels was<br />

observed in both treatment arms. However,<br />

the authors reported a more uni<strong>for</strong>m suppression<br />

of IGF-1 in the combined treatment.<br />

Toxicity<br />

<strong>Treatment</strong> with somatostatin or with its<br />

analogs is usually well tolerated, the most<br />

commonly encountered side effects being<br />

listed in table 3 [105]. When the subcutaneous<br />

route of administration is used, slight pain<br />

<strong>and</strong> redness at the site of injection have been<br />

reported. Mild diarrhea, abdominal cramps<br />

<strong>and</strong> malabsorption of fat have also been observed<br />

in 5–10% of patients. The severity of<br />

these symptoms appears to be dose-dependent<br />

<strong>and</strong> they usually decrease spontaneously.<br />

In insulin-dependent diabetics a 10–20% reduction<br />

of the daily dose of insulin may be<br />

required. However, severe hypoglycemia is<br />

not usually observed. Cholesterol gallstone<br />

<strong>for</strong>mation has been observed in 20–30% of<br />

patients after prolonged treatment, but <strong>for</strong>tunately<br />

they remain asymptomatic in most<br />

cases.<br />

Conclusions<br />

<strong>Somatostatin</strong> analogs are a novel class of<br />

compounds that have an established role in<br />

the management of patients with neuroendocrine<br />

tumors but only a potential role in the<br />

treatment of other solid tumors, including<br />

breast cancer. In this tumor type in particular,<br />

somatostatin analogs showed a limited activity<br />

either when used alone or when given in<br />

combination with tamoxifen or bromocriptine.<br />

Moreover, none of the r<strong>and</strong>omized trials<br />

that compared the therapeutic value of the<br />

combination of octreotide <strong>and</strong> tamoxifen versus<br />

tamoxifen alone showed any advantage in<br />

favor of combined treatment. There<strong>for</strong>e, although<br />

the great majority of trials failed to<br />

show major side effects attributable to somatostatin<br />

analogs, there is no place as yet <strong>for</strong><br />

the use of these compounds outside of controlled<br />

trials. However, further studies need to<br />

be done in humans on the preclinical evidence<br />

of improved efficacy of somatostatin<br />

analogs, either when used alone or when<br />

given in combination with both tamoxifen<br />

<strong>and</strong> chemotherapeutic agents. Moreover, the<br />

fact that an adequate suppression of the GH-<br />

72 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso


IGF-1 axis <strong>and</strong> of lactogenic hormone secretion<br />

was also demonstrated in clinical studies,<br />

but was not associated with a substantial antitumor<br />

activity, suggests that a more appropriate<br />

setting, dosing <strong>and</strong> scheduling should<br />

be established <strong>for</strong> further trials.<br />

The limited tumor response observed in<br />

clinical studies may have several explanations.<br />

The somatostatin receptors, when<br />

present in advanced breast tumors, are not<br />

homogeneously distributed. There<strong>for</strong>e, it can<br />

be speculated that only receptor-positive cells<br />

could be inhibited in their growth by somatostatin<br />

analogs. The majority of patients included<br />

in clinical trials with somatostatin analogs<br />

had far advanced disease in which somatostatin<br />

receptors might be lacking or might<br />

be defective in their affinity <strong>for</strong> somatostatin<br />

analogs. In other cases it is possible that the<br />

tumors affecting the patients treated with<br />

such compounds did not express the receptor<br />

specifically capable of binding the analog<br />

used. This mechanism is suggested by the lack<br />

of a complete cross-resistance among the<br />

different analogs in neuroendocrine tumors.<br />

Another possible reason <strong>for</strong> the limited clinical<br />

efficacy observed in breast cancer patients<br />

treated with somatostatin or somatostatin<br />

analogs could be inadequate dosing or scheduling.<br />

In fact, the doses employed in animal<br />

models were usually higher than those employed<br />

in clinical trials. Again the possibility<br />

of achieving better symptom control with<br />

higher doses or more appropriate scheduling<br />

of the same somatostatin analog has been<br />

observed in patients affected by neuroendocrine<br />

tumors. While an adequate suppression<br />

of GH, prolactin <strong>and</strong> IGF-1 levels has been<br />

observed in most of the studies conducted in<br />

humans, there is no proof yet that circulating<br />

levels of lactogenic hormones <strong>and</strong> of insulinlike<br />

peptides can influence breast cancer<br />

growth. Rather the direct inhibitory effect on<br />

tumor growth mediated by the binding of<br />

somatostatin analogs to their specific receptors<br />

might be more crucial, as suggested by the<br />

experiments in vitro using human breast cancer<br />

cell lines. There is evidence that the concentrations<br />

which are needed at the target cell<br />

level can largely exceed those that can be produced<br />

by a dose which otherwise can adequately<br />

suppress lactogenic hormones <strong>and</strong><br />

IGF-1-circulating levels. There<strong>for</strong>e, although<br />

results of clinical trials have been quite discouraging<br />

so far, further testing of somatostatin<br />

analogs alone or in combination with other<br />

antiproliferative drugs, including cytotoxics,<br />

is warranted. Patients with tumors displaying<br />

neuroendocrine features or those with tumors<br />

positive on OctreoScan ® might be those best<br />

suitable <strong>for</strong> such new trials, provided that<br />

higher doses <strong>and</strong> more appropriate regimens<br />

of somatostatin or its analogs be used within<br />

the frame of well-designed <strong>and</strong> controlled<br />

trials.<br />

Acknowledgments<br />

The authors thank Mrs. A. Fossati <strong>for</strong> her skilful<br />

assistance.<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 73


References<br />

1Parkin DM, Pisani P, Ferlay J: Estimates<br />

of the worldwide incidence of<br />

eighteen major cancer in 1985. Int J<br />

<strong>Cancer</strong> 1993;54:594–606.<br />

2 Pisani P, Parkin DM, Ferlay J: Estimates<br />

of the worldwide mortality<br />

from eighteen major cancers in<br />

1985, implication <strong>for</strong> prevention<br />

<strong>and</strong> projection of the future burden.<br />

Int J <strong>Cancer</strong> 1993;55:891–903.<br />

3 Ries LAG, Miller BA, Hankey BF,<br />

et al: National <strong>Cancer</strong> Institute.<br />

SEER <strong>Cancer</strong> Statistics Review,<br />

1973–1991: Tables <strong>and</strong> Graphs. Bethesda,<br />

National <strong>Cancer</strong> Institute,<br />

1994, NIH Publication No 94–<br />

2789.<br />

4 Chu KC, Tarone RE, Kessler LG,<br />

Ries LAG, Hankey BF, Miller BA,<br />

Edwards BK: Recent trends in US<br />

breast cancer incidence, survival,<br />

<strong>and</strong> mortality rates. J Natl <strong>Cancer</strong><br />

Inst 1996;88:1571–1579.<br />

5 Garne JP, Aspegren K, Balldin G,<br />

Ranstam J: Increasing incidence of<br />

<strong>and</strong> declining mortality from breast<br />

carcinoma: Trends in Malmö, Sweden<br />

1961–1992. <strong>Cancer</strong> 1997;79:<br />

69–74.<br />

6 Beral V, Hermon C, Reeves G, Peto<br />

R: Sudden fall in breast cancer death<br />

rates in Engl<strong>and</strong> <strong>and</strong> Wales. Lancet<br />

1995;345:1642–1643.<br />

7 Forbes JF: The incidence of breast<br />

cancer: The global burden, public<br />

health consideration. Semin Oncol<br />

1997;24(suppl 1):20–35.<br />

8 Roth JA, Cristiano RJ: Gene therapy<br />

<strong>for</strong> cancer: What have we done<br />

<strong>and</strong> where are we going? J Natl <strong>Cancer</strong><br />

Inst 1997;89:21–39.<br />

9 Engels K, Fox SB, Harris AL: Angiogenesis<br />

as a biologic <strong>and</strong> prognostic<br />

indicator in human breast carcinoma.<br />

EXS 1997;79:113–156.<br />

10 Walker RA, Jones JL, Chappel S,<br />

Walsh T, Shaw JA: Molecular pathology<br />

of breast cancer <strong>and</strong> its application<br />

to clinical management.<br />

<strong>Cancer</strong> Metastasis Rev 1997;16:5–<br />

27.<br />

11 Fisher B, Constantino J, Wickerman<br />

DL, et al: Tamoxifen <strong>for</strong> prevention<br />

of breast cancer: Report of the National<br />

Surgical Adjuvant Breast <strong>and</strong><br />

Bowel Project P-1 study. J Natl <strong>Cancer</strong><br />

Inst 1998;90:1371–1388.<br />

12 Veronesi U, Maisonneuve P, Costa<br />

A, et al: Prevention of breast cancer<br />

with tamoxifen: Preliminary findings<br />

from the Italian r<strong>and</strong>omised<br />

trial among hysterectomised women.<br />

Lancet 1998;352:93–97.<br />

13 Powles T, Eeles R, Ashley S, et al:<br />

Interim analysis of the incidence<br />

of breast cancer in the Royal Marsden<br />

Hospital tamoxifen r<strong>and</strong>omised<br />

chemoprevention trial. Lancet 1998;<br />

352:98–101.<br />

14 Tabar L, Fagerberg D, Day NE, et<br />

al: Breast cancer treatment <strong>and</strong> natural<br />

history: New insights from results<br />

of screening. Lancet 1992;339:<br />

412–414.<br />

15 Fisher B: Biological <strong>and</strong> clinical consideration<br />

regarding the use of surgery<br />

<strong>and</strong> chemotherapy in the treatment<br />

of primary breast cancer. <strong>Cancer</strong><br />

1997;40(suppl 5):574–587.<br />

16 Simon R: Meta-analysis <strong>and</strong> cancer<br />

clinical trial. Princ Pract Oncol Update<br />

1991;6:1–10.<br />

17 Early Breast <strong>Cancer</strong> Trialists’ Collaborative<br />

Group: Systemic treatment<br />

of early breast cancer by hormonal,<br />

cytotoxic, or immune therapy: 133<br />

r<strong>and</strong>omised trials involving 31000<br />

recurrences <strong>and</strong> 24 000 deaths<br />

among 75 000 women. Lancet 1992;<br />

339:1–15, 71–85.<br />

18 Early Breast <strong>Cancer</strong> Trialists’ Collaborative<br />

Group: Tamoxifen <strong>for</strong> early<br />

breast cancer: An overview of the<br />

r<strong>and</strong>omised trials. Lancet 1998;351:<br />

1451–1467.<br />

19 Early Breast <strong>Cancer</strong> Trialists’ Collaborative<br />

Group: Ovarian ablation in<br />

early breast cancer: Overview of the<br />

r<strong>and</strong>omised trials. Lancet 1996;348:<br />

1189–1196.<br />

20 Goldhirsch A, Glick JH, Gelber RD,<br />

Senn HJ: Meeting highlights: International<br />

Consensus Panel on the<br />

<strong>Treatment</strong> of Primary Breast <strong>Cancer</strong>.<br />

J Natl <strong>Cancer</strong> Inst 1998;90:<br />

1601–1608.<br />

21Nemoto T, Vana J, Bedwani R:<br />

Management <strong>and</strong> survival of female<br />

breast cancer. <strong>Cancer</strong> 1980;45:<br />

2917–2924.<br />

22 Powles TJ, Smith IE, Coombes RC:<br />

Endocrine therapy; in Halman KE<br />

(ed): <strong>Cancer</strong> <strong>Treatment</strong>. London,<br />

Chapman & Hall, 1981, pp 103–<br />

118.<br />

74 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso<br />

23 Osborne CK: Tamoxifen in the<br />

treatment of breast cancer. Lancet<br />

1998;339:1609–1618.<br />

24 Howell A, Downey S, Anderson E:<br />

New endocrine therapies <strong>for</strong> breast<br />

cancer. Eur J <strong>Cancer</strong> 1996;32A:<br />

576–588.<br />

25 Gradishar WJ, Jordan VC: Clinical<br />

potential of new antiestrogen. J Clin<br />

Oncol 1997;15:840–852.<br />

26 Parmar H, Phillips RH, Lightman<br />

SL: <strong>Somatostatin</strong> analogues: Mechanisms<br />

of action. Recent Results<br />

<strong>Cancer</strong> Res 1993;129:1–24.<br />

27 Prevost G, Israel L: <strong>Somatostatin</strong><br />

<strong>and</strong> somatostatin analogues in human<br />

breast carcinoma. Recent Results<br />

<strong>Cancer</strong> Res 1993;129:63–70.<br />

28 Coy DH, Murphy WA, Raynor K,<br />

Reisine T: The new pharmacology<br />

of somatostatin <strong>and</strong> its multiple receptors.<br />

J Pediatr Endocrinol 1993;<br />

6:205–209.<br />

29 Lamberts SWJ, Krenning EP, Reubi<br />

JC: The role of somatostatin <strong>and</strong> its<br />

analogs in the diagnosis <strong>and</strong> treatment<br />

of tumors. Endocr Rev 1991;<br />

12:450–482.<br />

30 Lamberts SWJ, Ooestrom R, Neufeld<br />

M, del Pozo E: The somatostatin<br />

analog SMS 201–995 induces<br />

long-acting inhibition of growth hormone<br />

secretion without rebound hypersecretion<br />

in acromegalic patients.<br />

J Clin Endocrinol Metab<br />

1985;60:1161–1165.<br />

31Patel YC, Greenwood MT, Panetta<br />

R, Hukovic N, Grigorakis S, Robertson<br />

LA, Srikant CB: Molecular biology<br />

of somatostatin receptor subtypes.<br />

Metabolism 1996;45(suppl<br />

1):31–38.<br />

32 Patel YC, Greenwood MT, Panetta<br />

R, Demehyshyn L, Niznik H, Srikant<br />

CB: The somatostatin receptor family.<br />

Life Sci 1995;57:1249–1265.<br />

33 Reisine T, Bell GI: Molecular biology<br />

of somatostatin receptors. Endocr<br />

Rev 1995;16:427–442.<br />

34 Buscail L, Esteve JP, Saint-Laurent<br />

N, Bert<strong>and</strong> V, Reisine T, O’Carrol<br />

AM, Bell GI, Schally AV, Vaysse N,<br />

Susini C: Inhibition of cell proliferation<br />

by the somatostatin analogue<br />

RC-160 is mediated by somatostatin<br />

receptor subtypes SSTR2 <strong>and</strong><br />

SSTR5 through different mechanisms.<br />

Proc Natl Acad Sci USA<br />

1995;92:1580–1584.


35 Bell GI, Reisine T: Molecular biology<br />

of somatostatin receptors. Trends<br />

Neurosci 1993;16:34–38.<br />

36 Foekens JA, Portengen H, van Putten<br />

WL, Trapman AM, Reubi JC,<br />

Alexieva-Figusch J, Klijn JG: Prognostic<br />

value of receptors <strong>for</strong> insulinlike<br />

growth factor 1, somatostatin,<br />

<strong>and</strong> epidermal growth factor in human<br />

breast cancer. <strong>Cancer</strong> Res<br />

1989;49:7002–7009.<br />

37 Visser-Wisselaar HA, Hofl<strong>and</strong> LJ,<br />

van Uffelen CJC, van Koetsveld<br />

PM, Lamberts SWJ: <strong>Somatostatin</strong><br />

receptor manipulation. Digestion<br />

1996;57(suppl 1):7–10.<br />

38 Lamberts SWJ: The role of somatostatin<br />

in the regulation of anterior<br />

pituitary hormone secretion <strong>and</strong> the<br />

use of its analogues in the treatment<br />

of human pituitary tumors. Endocr<br />

Rev 1988;9:417–436.<br />

39 Bruno JF, Xu Y, Berelowitz M: <strong>Somatostatin</strong><br />

regulates somatostatin<br />

receptor subtype mRNA expression<br />

in GH3 cells. Biochem Biophys Res<br />

Commun 1994;202:1738–1743.<br />

40 Lamberts SWJ, Koper JW, Reubi<br />

JC: Potential role of somatostatin<br />

analogues in the treatment of cancer.<br />

Eur J Clin Invest 1987;17:281–<br />

287.<br />

41Koch BD, Blalock JB, Schonbrunn<br />

A: Characterization of the cyclic<br />

AMP-dependent action of somatostatin<br />

in GH cells. J Biol Chem<br />

1988;263:216–225.<br />

42 Ikeda SR, Schofield GG: <strong>Somatostatin</strong><br />

blocks a calcium current in rat<br />

sympathetic ganglion neurones. J<br />

Physiol 1989;409:221–240.<br />

43 Schonbrunn A: <strong>Somatostatin</strong> action<br />

in pituitary cells involves two independent<br />

transduction mechanisms.<br />

Metabolism 1990;39(suppl 1):96–<br />

100.<br />

44 Heisler S: Stimulation of adrenocorticotropin<br />

secretion from AtT-20<br />

cells by the calcium channel activator,<br />

BAY-K-8644, <strong>and</strong> its inhibition<br />

by somatostatin <strong>and</strong> carbachol. J<br />

Pharmacol Exp Ther 1985;235:741–<br />

748.<br />

45 Luini A, De Matteis MA: Evidence<br />

that receptor-linked G protein inhibits<br />

exocytosis by a post-secondmessenger<br />

mechanism in AtT-20<br />

cells. J Neurochem 1990;54:30–38.<br />

46 Wollheim CB, Winiger BP, Ullrich<br />

S, et al: <strong>Somatostatin</strong> inhibition of<br />

hormone release: Effects on cytosolic<br />

Ca++ <strong>and</strong> interference with<br />

distal secretory events. Metabolism<br />

1990;39(suppl 1):101–104.<br />

47 Liebow C, Reilly C, Serrano M, et<br />

al: <strong>Somatostatin</strong> analogues inhibit<br />

growth of pancreatic cancer by stimulating<br />

tyrosine phosphatase. Proc<br />

Natl Acad Sci USA 1989;86:2003–<br />

2007.<br />

48 Sharma K, Patel JC, Srikant C: Subtype<br />

specific induction of wild type<br />

p53 <strong>and</strong> apoptosis, but not cell cycle<br />

arrest, by human somatostatin receptor<br />

3. Mol Endocrinol 1996;10:<br />

1688–1696.<br />

49 Buscail L, Delesque N, Esteve JP,<br />

Saint-Laurent N, Prats H, Clerc P,<br />

Robberecht P, Bell GI, Liebow C,<br />

Schally AV, Vaysse N, Susini C:<br />

Stimulation of tyrosine phosphatase<br />

<strong>and</strong> inhibition of cell proliferation<br />

by somatostatin analogs: Mediation<br />

by human somatostatin receptor<br />

subtypes SSTR1 <strong>and</strong> SSTR2. Proc<br />

Natl Acad Sci USA 1994;91:2315–<br />

2319.<br />

50 Klarlund JK: Trans<strong>for</strong>mation of<br />

cells by an inhibitor of phosphatases<br />

acting on phosphotyrosine in proteins.<br />

Cell 1985;41:707–717.<br />

51McGuire WL, Dickson RB, Osborne<br />

CK, Salomon D: The role of<br />

growth factors in breast cancer. A<br />

panel discussion. Breast <strong>Cancer</strong> Res<br />

Treat 1988;12:159–166.<br />

52 Manni A, Wright C, Davis G, Glenn<br />

J, Joehl R, Feil P: Promotion by prolactin<br />

of the growth of human breast<br />

neoplasms cultured in vitro in the<br />

soft agar clonogenic assay. <strong>Cancer</strong><br />

Res 1986;46:1669–1672.<br />

53 Malarkey WB, Kennedy M, Allred<br />

LE, Milo G: Physiological concentrations<br />

of prolactin can promote<br />

the growth of human breast tumor<br />

cells in culture. J Clin Endocrinol<br />

Metab 1983;56:673–677.<br />

54 Schally AV: Oncological application<br />

of somatostatin analogues. <strong>Cancer</strong><br />

Res 1988;48:6877–6885.<br />

55 Weckbecker G, Raulf F, Stolz B,<br />

Bruns C: <strong>Somatostatin</strong> analogues <strong>for</strong><br />

diagnosis <strong>and</strong> treatment of cancer.<br />

Pharmacol Ther 1993;60:245–264.<br />

56 Reubi JC, Maurer R, von Werder K,<br />

Torhorst J, Klijn JG, Lamberts SW:<br />

<strong>Somatostatin</strong> receptors in human<br />

endocrine tumors. <strong>Cancer</strong> Res 1987;<br />

47:551–558.<br />

57 Srkalovic G, Cai RZ, Schally AV:<br />

Evaluation of receptors <strong>for</strong> somatostatin<br />

in various tumors using different<br />

analogues. J Clin Endocrinol<br />

Metab 1990;70:661–669.<br />

58 Reubi JC, Torhost J: The relationship<br />

between somatostatin, epidermal<br />

growth factor, <strong>and</strong> steroid hormone<br />

receptors in breast cancer.<br />

<strong>Cancer</strong> 1989;64:1254–1260.<br />

59 Reubi JC, Waser B, Foekens JA,<br />

Kljin JGM, Lamberts SWJ, Laissue<br />

J: <strong>Somatostatin</strong> receptor incidence<br />

<strong>and</strong> distribution in breast cancer using<br />

receptor autoradiography: relationship<br />

to EGF receptors. Int J<br />

<strong>Cancer</strong> 1990;46:416–420.<br />

60 Saisbury JR, Farndon R, Needham<br />

G, Malcom A, Harris A: Epidermalgrowth-factor<br />

receptor status as predictor<br />

of early recurrence <strong>and</strong> death<br />

from breast cancer. Lancet 1987;<br />

i:1398–1402.<br />

61Macaulay VM: Insulin-like growth<br />

factors <strong>and</strong> cancer. Br J <strong>Cancer</strong><br />

1992;65:311–320.<br />

62 Pollak M, Perdue JF, Margolese<br />

RG, Baer K, Richard M: Presence of<br />

somatomedin receptors on primary<br />

human breast <strong>and</strong> colon carcinomas.<br />

<strong>Cancer</strong> Lett 1987;38:223–230.<br />

63 Reeve JG, Brinkman A, Hughes S,<br />

Mitchell J, Schw<strong>and</strong>er J, Bleehan<br />

NM: Expression of insulin-like<br />

growth factor (IGF) <strong>and</strong> IGF-binding<br />

protein genes in human lung tumor<br />

cell lines. J Natl <strong>Cancer</strong> Inst<br />

1992;84:628–634.<br />

64 Yee D, Paik S, Lebovic GS, Marcus<br />

RR, Favoni RE, Cullen KJ, Lippman<br />

ME, Rosen N: Analysis of insulin-like<br />

growth factor I gene expression<br />

in malignancy: Evidence <strong>for</strong> a<br />

paracrine role in human breast cancer.<br />

Mol Endocrinol 1989;3:509–<br />

517.<br />

65 Osborne CK, Coronado E, Kitten<br />

LJ, Arteaga CI, Fuqua SAW, Ramasharma<br />

K, Marshall M, Li CH: Insulin-like<br />

growth factor-II (IGF-<br />

II): A potential autocrine/paracrine<br />

growth factor <strong>for</strong> human breast cancer<br />

acting via the IGF-I receptor.<br />

Mol Endocrinol 1989;3:1701–1709.<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 75


66 Cullen KJ, Yee D, Sly WS, Perdue J,<br />

Hampton B, Lippman ME, Rosen<br />

N: Insulin-like growth factor receptor<br />

expression <strong>and</strong> function in human<br />

breast cancer. <strong>Cancer</strong> Res<br />

1990;50:48–53.<br />

67 Peyrat JP, Bonneterre J, Beuscart B,<br />

Dijane J, Demaille A: Insulin-like<br />

growth factor-I receptors in human<br />

breast cancer <strong>and</strong> their relation to<br />

estradiol <strong>and</strong> progesterone receptors.<br />

<strong>Cancer</strong> Res 1988;48:6429–<br />

6433.<br />

68 Pekonen F, Partanen S, Makinen T,<br />

Rutanen EM: Receptors <strong>for</strong> epidermal<br />

growth factor <strong>and</strong> insulin-like<br />

growth factor-I <strong>and</strong> their relation to<br />

steroid receptors in human breast<br />

cancer. <strong>Cancer</strong> Res 1988;48:1343–<br />

1347.<br />

69 Serri O, Brazeau P, Kachra Z, Posner<br />

B: Octreotide inhibits insulinlike<br />

growth factor-I hepatic gene expression<br />

in the hypophysectomized<br />

rat: Evidence <strong>for</strong> a direct <strong>and</strong> indirect<br />

mechanism of action. Endocrinology<br />

1992;130:1816–1821.<br />

70 Ambler GR, Butler AA, Padmanabhan<br />

J, Breier BH, Gluckman PD:<br />

The effect of octreotide on GH receptor<br />

<strong>and</strong> IGF-I expression in the<br />

GH-deficient rat. J Endocrinol<br />

1996;149:223–231.<br />

71Ezzat S, Ren SG, Braunstein GD,<br />

Melmed S: Octreotide stinulates insulin-like<br />

growth factor binding protein-1:<br />

A potential pituitay independent<br />

mechanism <strong>for</strong> drug action. J<br />

Clin Endocrinol Metab 1992;75:<br />

1459–1463.<br />

72 De Herder WW, Uitterlinden P, van<br />

der Lely AJ, Hofl<strong>and</strong> LJ, Lamberts<br />

SW: Octreotide, but not bromocriptine,<br />

increases circulating insulinlike<br />

growth factor binding protein 1<br />

levels in acromegaly. Eur J Endocrinol<br />

1995;133:195–199.<br />

73 Resnicoff M, Abrallam D, Yutanaviboonchai<br />

W, Rotman HL, Kajstura<br />

J, Rubin R, Zoleck P, Baserga R:<br />

The insulin-like growth factor I receptor<br />

protects tumor cells from<br />

apoptosis in vivo. <strong>Cancer</strong> Res 1995;<br />

55:2463–2469.<br />

74 Sell C, Baserga R, Rubin R: Insulinlike<br />

growth factor I (IGF-I) <strong>and</strong> the<br />

IGF-I receptor prevent etoposideinduced<br />

apoptosis. <strong>Cancer</strong> Res<br />

1995;55:303–306.<br />

75 Szepeshazi K, Schally AV, Halmos<br />

G: Apoptosis in pancreatic cancer of<br />

hamster. Int J Pancreatol 1994;16:<br />

282–287.<br />

76 Szende B, Lapis K, Redding TW,<br />

Srkalovic G, Schally AV: Growth inhibition<br />

of MXT mammary carcinoma<br />

by enhancing programmed cell<br />

death (apoptosis) with anlog of LH-<br />

RH <strong>and</strong> somatostatin. Breast <strong>Cancer</strong><br />

Res Treat 1989;14:307–314.<br />

77 Reubi JC, Horisberger U, Laissue J:<br />

High density of somatostatin receptors<br />

in vein surrounding human cancer<br />

tissue. Role in the tumor-host<br />

interaction? Int J <strong>Cancer</strong> 1994;56:<br />

681–688.<br />

78 Setyono-Han B, Henkelman MS,<br />

Foekens JA, Klijn JGM: Direct inhibitory<br />

effects of somatostatin (analogues)<br />

on the growth of human<br />

breast cancer cells. <strong>Cancer</strong> Res<br />

1987;47:1566–1570.<br />

79 Szepeshazi K, Milovanovic S, Lapis<br />

K, Groot K, Schally AV: Growth<br />

inhibition of oestrogen independent<br />

MXT mouse mammary carcinomas<br />

in mice treated with an agonist or<br />

antagonist of LH-RH , an analog of<br />

somatostatin, or a combination.<br />

Breast <strong>Cancer</strong> Res Treat 1992;21:<br />

181–192.<br />

80 Friedl A, Jordan VC, Pollak M: Suppression<br />

of serum IGF-I levels in<br />

breast cancer patients during adjuvant<br />

tamoxifen therapy: Eur J <strong>Cancer</strong><br />

1993;29A:1368–1372.<br />

81Pollak M: Effects of adjuvant tamoxifen<br />

therapy on growth hormone<br />

<strong>and</strong> insulin-like growth factor<br />

I (IGF-I) physiology; in Salmon SE<br />

(ed): Adjuvant Therapy of <strong>Cancer</strong>.<br />

Philadelphia, Lippincott, 1993, vol<br />

7, pp 43–55.<br />

82 Huynh HT, Tetenes E, Wallace L,<br />

Pollak M: In vivo inhibition of insulin-like<br />

growth factor-I gene expression<br />

by tamoxifen. <strong>Cancer</strong> Res<br />

1993;53:1727–1730.<br />

83 Huynh HT, Pollak M: Enhancement<br />

of tamoxifen-induced suppression<br />

of insulin-like growth factor-I gene<br />

expression <strong>and</strong> serum level by a somatostatin<br />

analogue. Biochem Biophys<br />

Res Commun 1994;203:253–<br />

259.<br />

76 Chemotherapy 2001;47(suppl 2):62–77 Boccardo/Amoroso<br />

84 Weckbecker G, Tolcsvai L, Stolz B,<br />

Pollak M, Bruns S: <strong>Somatostatin</strong> analogue<br />

octreotide enhances the anti-neoplastic<br />

effects of tamoxifen<br />

<strong>and</strong> ovariectomy on 7,12-dimethylbenz(a)anthracene-induced<br />

rat mammary<br />

carcinomas. <strong>Cancer</strong> Res 1994;<br />

54:6334–6337.<br />

85 Pollak M: Enhancement of the antineoplastic<br />

effects of tamoxifen by<br />

somatostatin analogues. Digestion<br />

1996;57(suppl 1):29–33.<br />

86 Fidler IJ, Ellis LM: The implication<br />

of angiogenesis <strong>for</strong> the biology <strong>and</strong><br />

therapy of cancer metastasis. Cell<br />

1994;79:185–188.<br />

87 Rak JA, St Croix BD, Kerbel RS:<br />

Consequences of angiogenesis <strong>for</strong><br />

tumor progression, metastasis <strong>and</strong><br />

cancer therapy. Anticancer Drugs<br />

1995;6:3–18.<br />

88 Weinstat-Saslow D, Steeg SP: Angiogenesis<br />

<strong>and</strong> colonization in the<br />

tumor metastatic process: Basic <strong>and</strong><br />

applied advances. FASEB J 1994;8:<br />

401–407.<br />

89 Weckbecker G, Raulf F, Tolcsvai L,<br />

Bruns C: Potentiation of the antiproliferative<br />

effects of anti-cancer<br />

drugs by octreotide in vitro <strong>and</strong><br />

in vivo. Digestion 1996;57(suppl 1):<br />

22–28.<br />

90 Viguerie N, Tahiri-Jouti N, Ayral<br />

AM, Cambillau C, Scemama JL,<br />

Bastie MJ, Knuhtsen S, Esteve JP,<br />

Pradayrol L, Susini C, Vaysse N: Direct<br />

inhibitory effects of a somatostatin<br />

analog, SMS 201–995, on<br />

AR4–2J cell proliferation via pertussis<br />

toxin-sensitive guanosine triphosphate-bindingprotein-independent<br />

mechanism. Endocrinology<br />

1989;124:1017–1025.<br />

91Berembaum MC: What is synergy?<br />

Pharmacol Rev 1989;41:93–141.<br />

92 Lee JM, Erlich R, Bruckner HW,<br />

Roboz J, Beasley J, Ohnuma T: Octreotide<br />

acetate (S<strong>and</strong>ostatin, SMS)<br />

increases in vitro accumulation of<br />

doxorubicin (DOX). Proc Am Assoc<br />

<strong>Cancer</strong> Res 1993;34:285.<br />

93 Cascinu S, Fedeli A, Fedeli SL, Catalano<br />

G: Octreotide versus loperamide<br />

in the treatment of fluorouracil-induced<br />

diarrhea: A r<strong>and</strong>omized<br />

trial. J Clin Oncol 1993;11:148–<br />

151.


94 Vennin P, Peyrat JP, Bonneterre J,<br />

Louchez MM, Harris AG, Demaille<br />

A: Effect of the long-acting somatostatin<br />

analog SMS 201–995 (S<strong>and</strong>ostatin)<br />

in advanced breast cancer.<br />

Anticancer Res 1989;9:153–156.<br />

95 Manni A, Boucher AE, Demers LM,<br />

Harvey HA, Lipton A, Simmonds<br />

MA, Bartholomew M: Endocrine effects<br />

of combined somatostatin analog<br />

<strong>and</strong> bromocriptine therapy in<br />

women with advanced breast cancer.<br />

Breast <strong>Cancer</strong> Res Treat 1989;<br />

14:289–298.<br />

96 Stolfi R, Parisi AM, Natoli C, Iacobelli<br />

S: Advanced breast cancer: Response<br />

to somatostatin. Anticancer<br />

Res 1990;10:203–204.<br />

97 Anderson E, Ferguson JE, Morten<br />

H, Shalet SM, Robinson EL, Howell<br />

A: Serum immunoreactive <strong>and</strong><br />

bioactive lactogenic hormones in<br />

advanced breast cancer patients<br />

treated with bromocriptine <strong>and</strong> octreotide.<br />

Eur J <strong>Cancer</strong> 1993;<br />

29A:209–217.<br />

98 Canobbio L, Cannata D, Miglietta<br />

L, Boccardo F: Somatuline (BIM<br />

23014) <strong>and</strong> tamoxifen treatment of<br />

postmenopausal breast cancer patients:<br />

Clinical activity <strong>and</strong> effect on<br />

insulin-like growth factor-I (IGF-I)<br />

levels. Anticancer Res 1995;15:<br />

2687–2690.<br />

99 Di Leo A, Ferrari L, Bajetta E, Bartoli<br />

C, Vicario G, Moglia D, Miceli<br />

R, Callegari M, Bono A: Biological<br />

<strong>and</strong> clinical evaluation of lanreotide<br />

(BIM 23014), a somatostatin<br />

analogue, in the treatment of advanced<br />

breast cancer. A pilot study<br />

by the I.T.M.O. Group. Italian<br />

Trials in Medical Oncology. Breast<br />

<strong>Cancer</strong> Res Treat 1995;34:237–<br />

244.<br />

100 O’Byrne KJ, Dobbs N, Propper<br />

DJ, Braybrooke JP, Koukourakis<br />

MI, Mitchell K, Woodhull J, Talbot<br />

DC, Schally AV, Harris AL:<br />

Phase II study of RC-160 (vapreotide),<br />

an octapeptide analogue of<br />

somatostatin, in the treatment of<br />

metastatic breast cancer. Br J <strong>Cancer</strong><br />

1999;79:1413–1418.<br />

101 Italian Study Group <strong>for</strong> the Di Bella<br />

Multitherapy Trials: Evaluation<br />

of an unconventional cancer treatment<br />

(the Di Bella multitherapy):<br />

Results of phase II trials in Italy.<br />

Br J <strong>Cancer</strong> 1999;318:224–228.<br />

102 Ingle JN, Suman VJ, Kardinal CG,<br />

Krook JE, Mailliard JA, Veeder<br />

MH, Loprinzi CL, Dalton RJ,<br />

Hartmann LC, Conover CA, Pollak<br />

M: A r<strong>and</strong>omized trial of tamoxifen<br />

alone or combined with<br />

octreotide in the treatment of<br />

women with metastatic breast carcinoma.<br />

<strong>Cancer</strong> 1999;85:1284–<br />

1292.<br />

103 Bajetta E, Sommer H, Guastalla J,<br />

Szakoiczai I, Baltali E, Pinter T,<br />

Csepreghy M, Ottestad L, Boni C,<br />

Bryce C, Klijo J, Kiese B, Mietlowski<br />

W, Bone A, Kay A: Phase 3<br />

trial of octreotide pamoate (OP<br />

LAR) <strong>and</strong> tamoxifen versus placebo<br />

<strong>and</strong> tamoxifen in metastatic<br />

breast cancer. Proc ASCO 1999;<br />

18:110a.<br />

104 Bontenbal M, Foekens JA, Lamberts<br />

SWJ, de Jong FH, van Putten<br />

WLJ, Braun HJ, Burghouts JTM,<br />

van der Linden GHM, Klijn JGM:<br />

Feasibility, <strong>and</strong>ocrine <strong>and</strong> antitumour<br />

effects of a triple endocrine<br />

therapy with tamoxifen, a somatostatin<br />

anlogue <strong>and</strong> an prolactin<br />

lowering drug in postmenopausal<br />

metastatic breast cancer: A<br />

r<strong>and</strong>omized study with long-term<br />

follow-up. Br J <strong>Cancer</strong> 1998;77:<br />

115–122.<br />

105 Lamberts SWJ, Van der Lely AJ,<br />

De Herder WW, Hofl<strong>and</strong> LJ: Octreotide.<br />

N Engl J Med 1996;334:<br />

246–254.<br />

<strong>Somatostatin</strong> <strong>and</strong> Breast <strong>Cancer</strong> Chemotherapy 2001;47(suppl 2):62–77 77


Chemotherapy 2001;47(suppl 2):78–108<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

Kenneth J. O’Byrne a Andrew V. Schally b Ann Thomas a<br />

Desmond N. Carney c William P. Steward a<br />

a University Department of Oncology, Leicester Royal Infirmary, UK; b Endocrine, Polypeptide<br />

<strong>and</strong> <strong>Cancer</strong> Institute, Tulane University Medical School <strong>and</strong> Veterans Affairs Medical Centre,<br />

New Orleans, La., USA; c Department of Medical Oncology, Mater Misericordiae Hospital,<br />

Dublin, Irel<strong>and</strong><br />

Key Words<br />

<strong>Somatostatin</strong> W Receptor W Lung cancer W<br />

[ 111 In]pentetreotide W Chemotherapy<br />

Abstract<br />

Despite developments in diagnosis <strong>and</strong> treatment,<br />

lung cancer is the commonest cause of<br />

cancer death in Europe <strong>and</strong> North America.<br />

Due to increasing cigarette consumption, the<br />

incidence of the disease <strong>and</strong> resultant mortality<br />

is rising dramatically in women. Novel<br />

approaches to the management of lung cancer<br />

are urgently required. <strong>Somatostatin</strong> is a<br />

tetradecapeptide first identified in the pituitary<br />

<strong>and</strong> subsequently throughout the body<br />

particularly in neuroendocrine cells of the<br />

pancreas <strong>and</strong> gastrointestinal tract <strong>and</strong> the<br />

nervous system. The peptide has numerous<br />

functions including inhibition of hormone release,<br />

immunomodulation <strong>and</strong> neurotransmission<br />

<strong>and</strong> is an endogenous inhibitor of<br />

cell proliferation <strong>and</strong> angiogenesis. <strong>Somatostatin</strong><br />

<strong>and</strong> its analogs, including octreotide<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0078$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

(SMS 201–995), somatuline (BIM 23014) <strong>and</strong><br />

vapreotide (RC-160), act by binding to specific<br />

somatostatin receptors (SSTR) of which<br />

there are 5 principal subtypes, SSTR-1–5. Although<br />

elevated plasma somatostatin levels<br />

may be detected in 14–15% of patients, tumor<br />

cell expression appears rare. SSTR may<br />

be expressed by lung tumors, particularly<br />

small cell lung cancer <strong>and</strong> bronchial carcinoid<br />

disease. [ 111 In]pentetreotide scintigraphy<br />

may have a role to play in the localization<br />

<strong>and</strong> staging of lung cancers both be<strong>for</strong>e <strong>and</strong><br />

following treatment, <strong>and</strong> in detecting relapsed<br />

disease. The potential role of radiolabelled<br />

somatostatin analogs as radiotherapeutic<br />

agents in the management of lung<br />

cancer is currently being explored. <strong>Somatostatin</strong><br />

analog therapy results in significant<br />

growth inhibition of both SSTR-positive <strong>and</strong><br />

SSTR-negative lung tumors in vivo. Recent<br />

work indicates that these agents may enhance<br />

the efficacy of chemotherapeutic<br />

agents in the treatment of solid tumors including<br />

lung cancer.<br />

Copyright © 2001 S. Karger AG, Basel<br />

K.J. O’Byrne<br />

University Department of Oncology, Leicester Royal Infirmary<br />

Leicester LE1 5WW (UK)<br />

Tel. +44 116 258 7602<br />

E-Mail Kobyrne@Lri.org.uk


Introduction<br />

At the turn of the century lung cancer was<br />

an extremely rare tumor. Since then its incidence<br />

has increased to such an extent that it is<br />

now the commonest cause of cancer death in<br />

Europe <strong>and</strong> North America. This dramatic<br />

change is largely attributable to cigarette<br />

smoking which is responsible <strong>for</strong> over 80% of<br />

all cases. A reduction in cigarette smoking in<br />

developed countries is beginning to produce<br />

an age-adjusted decrease in the incidence of<br />

lung cancer in men. However, smoking rates<br />

<strong>and</strong> the incidence of lung cancer are rising rapidly<br />

in women. The disease has now surpassed<br />

breast cancer as the commonest cause of cancer<br />

death in women. Of great concern is the<br />

fact that the peak of the lung cancer epidemic<br />

in women has not yet been reached [1, 2].<br />

Despite improvements in the diagnostic,<br />

surgical, chemotherapeutic <strong>and</strong> radiotherapy<br />

management of lung cancer the overall results<br />

of treatment are poor. The 5-year survival<br />

rate <strong>for</strong> small cell lung cancer is approximately<br />

3%, <strong>and</strong> <strong>for</strong> non-small cell lung cancer 8–<br />

14%. Novel approaches to management are<br />

urgently required. The growth in our underst<strong>and</strong>ing<br />

of the molecular biology of lung cancer<br />

is leading to the development of new<br />

approaches to the treatment of this group of<br />

diseases including the use of growth factor<br />

antagonists, growth factor receptor antibodies<br />

<strong>and</strong> antiangiogenic agents [3–7].<br />

<strong>Somatostatin</strong><br />

<strong>Somatostatin</strong> is a tetradecapeptide hormone<br />

first identified in the hypothalamus as<br />

an inhibitor of growth hormone release. The<br />

peptide has subsequently been found throughout<br />

the body, particularly in the pancreas, gastrointestinal<br />

tract <strong>and</strong> nervous system. <strong>Somatostatin</strong><br />

inhibits the release of growth hor-<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

mone <strong>and</strong> thyroid-stimulating hormone from<br />

the anterior pituitary <strong>and</strong> hormone release<br />

from the pancreas <strong>and</strong> gastrointestinal tract.<br />

<strong>Somatostatin</strong> also functions as a neurotransmitter,<br />

immunomodulator <strong>and</strong> suppressor of<br />

angiogenesis <strong>and</strong> cell proliferation [8–14]. It<br />

acts by binding to specific receptors (SSTR) of<br />

which 5 principal subtypes have been identified:<br />

SSTR-1, SSTR-2, SSTR-3, SSTR-4 <strong>and</strong><br />

SSTR-5 [15].<br />

Growth Inhibitory Effects of<br />

<strong>Somatostatin</strong> <strong>and</strong> <strong>Somatostatin</strong><br />

<strong>Analogs</strong><br />

Through the activation of SSTR, somatostatin<br />

<strong>and</strong> its analogs exert a number of direct<br />

growth inhibitory effects on normal <strong>and</strong> malignant<br />

cells. All 5 SSTR are functionally coupled<br />

to adenylyl cyclase. Activation of SSTR results<br />

in inhibition of the intracellular accumulation<br />

of cyclic adenosine monophosphate (cAMP).<br />

This may result in direct inhibition of the<br />

growth of cells in which accumulation of<br />

cAMP, <strong>and</strong> subsequent activation of the protein<br />

kinase A pathway, results in proliferation<br />

[14–16]. This is supported by observations in<br />

the SSTR-2-positive human pancreatic cancer<br />

cell line CFPAC-1 where in vitro growth inhibition<br />

by the octapeptide somatostatin analog<br />

RC-160 is associated with inhibition of cAMP<br />

accumulation [17]. Acting through SSTR-5,<br />

the somatostatin analog RC-160 has been<br />

demonstrated to suppress the proliferative effects<br />

of cholecystokinin on CHO cells by inhibiting<br />

the accumulation of cyclic guanosine<br />

monophosphate (cGMP) [18].<br />

On binding to their extracellular receptor<br />

domains, many growth factors induce intracellular<br />

receptor tyrosine kinase activity.<br />

Likewise a number of oncogene products are<br />

truncated growth factor receptors, lacking an<br />

extracellular domain but having permanent<br />

Chemotherapy 2001;47(suppl 2):78–108 79


Table 1. Antiproliferative activity of somatostatin<br />

Direct effects<br />

Activation of phosphatase activity<br />

Inhibition of cAMP accumulation<br />

Inhibition of cGMP accumulation<br />

Inhibition of the mobilization of intracellular calcium<br />

through<br />

1 Inhibition of cell membrane calcium channels<br />

2 Interaction with phospholipase C <strong>and</strong> inositol/<br />

phospholipid growth pathways<br />

Induction of programmed cell death (apoptosis)<br />

Indirect effects<br />

Inhibition of the release of trophic factors from elsewhere,<br />

e.g. EGF <strong>and</strong> IGF-1<br />

Inhibition of angiogenesis<br />

Inhibition of cancer cell adhesion<br />

endogenous intracellular tyrosine kinase activity.<br />

This activity results in phosphorylation<br />

of a number of tyrosine residues on the intracellular<br />

portion of the receptor. The phosphorylated<br />

tyrosine residues interact with<br />

guanine nucleotide-binding proteins <strong>and</strong> subsequently<br />

ras, inducing a phosphorylation<br />

cascade which activates transcription factors<br />

<strong>and</strong> ultimately results in cell proliferation<br />

[19]. <strong>Somatostatin</strong> stimulates the activity of<br />

phosphatases including phosphotyrosine<br />

phosphatases which have been demonstrated<br />

to dephosphorylate the tyrosine residues of<br />

activated type 1 growth factor receptors, such<br />

as the epidermal growth factor receptor<br />

(EGFR) [10, 11, 14, 20, 21]. Further evidence<br />

<strong>for</strong> activation of intracellular phosphatases<br />

comes from studies which have demonstrated<br />

that somatostatin analogs can inhibit insulinlike<br />

growth factor-1 (IGF-1) <strong>and</strong> serum-stimulated<br />

MAP kinase activation in vitro [22].<br />

Activation of phosphatases is associated with<br />

inhibition of the proliferation of SSTR-positive<br />

normal <strong>and</strong> cancer cells in vitro [6, 10, 14,<br />

20–22]. The activation of phosphotyrosine<br />

phosphatases is mediated through SSTR-1<br />

<strong>and</strong> SSTR-2 [10]. <strong>Somatostatin</strong> <strong>and</strong> somatostatin<br />

analogs also inhibit the accumulation of<br />

intracellular calcium from both extracellular<br />

<strong>and</strong> intracellular sources [11, 23].<br />

<strong>Somatostatin</strong> analogs have been shown to<br />

inhibit the growth of SSTR-negative tumors<br />

in vivo demonstrating important indirect antiproliferative<br />

effects. Many tumors express<br />

IGF-1 receptors (R) <strong>and</strong> proliferate in response<br />

to exposure to IGF-1 [24, 25]. There<strong>for</strong>e<br />

blocking the release of growth factors<br />

from normal <strong>and</strong> malignant tissues may inhibit<br />

autocrine-, paracrine- <strong>and</strong> endocrineinduced<br />

cancer cell proliferation [14].<br />

Through the inhibition of the protein kinase<br />

A pathway, somatostatin <strong>and</strong> its analogs may<br />

not only modulate cell growth but also suppress<br />

the synthesis <strong>and</strong> release of hormones<br />

<strong>and</strong> trophic growth factors, including growth<br />

hormone <strong>and</strong> IGF-, from normal tissues [13,<br />

14]. In keeping with these findings the growth<br />

inhibitory effects of RC-160 in SSTR-negative<br />

tumors are associated with a decrease in<br />

serum growth hormone (GH) <strong>and</strong> IGF-1 levels.<br />

Furthermore, the growth inhibition is associated<br />

with suppression of tumor cell IGF-<br />

1R expression [14, 26–28].<br />

<strong>Somatostatin</strong> analogs are also potent antiangiogenic<br />

agents having direct growth inhibitory<br />

effects on proliferating endothelial<br />

cells [12]. As angiogenesis is essential <strong>for</strong> the<br />

growth of a tumor greater than 1–2 mm in<br />

diameter, interference with this process may<br />

result in the inhibition of cancer growth [29].<br />

<strong>Somatostatin</strong> also reduces the capacity of malignant<br />

cells to adhere to blood vessel walls<br />

thereby reducing the metastatic potential of<br />

malignant tumors [30].<br />

In keeping with both the direct <strong>and</strong> indirect<br />

antiproliferative growth inhibitory effects<br />

described, somatostatin <strong>and</strong> its analogs have<br />

been shown to induce tumor cell apoptosis<br />

both in vitro <strong>and</strong> in vivo (table 1) [13, 14].<br />

80 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


<strong>Somatostatin</strong> Expression in Lung<br />

<strong>Cancer</strong><br />

Small cell lung cancer (SCLC) accounts <strong>for</strong><br />

approximately 20% of all lung tumors. SCLC<br />

is a neuroendocrine tumor characterized by<br />

the expression of pan-neuroendocrine markers<br />

including neuron-specific enolase (NSE),<br />

creatine kinase BB <strong>and</strong> chromogranin, <strong>and</strong><br />

specific hormones <strong>and</strong> their receptors including<br />

bombesin (GRP) <strong>and</strong> IGF-1 [5, 13, 25,<br />

31]. Many of these substances may be elevated<br />

in the serum <strong>and</strong>/or plasma of patients<br />

with SCLC at presentation <strong>and</strong>, particularly<br />

in the case of NSE, may be of value as tumor<br />

markers, the levels falling if the disease responds<br />

to chemotherapy <strong>and</strong> rising if the disease<br />

progresses or relapses [13, 32, 33].<br />

A number of studies have suggested that<br />

SCLC may synthesize <strong>and</strong> secrete somatostatin.<br />

The detection of elevated serum<br />

somatostatin-like immunoreactivity in lung<br />

cancer patients was first reported in 1980<br />

[34]. Subsequent investigations revealed elevated<br />

somatostatin serum levels in 4 of 26<br />

(15%) [35] <strong>and</strong> 3 of 21 (14%) patients [36]<br />

with SCLC compared to controls. Immunoreactive<br />

tissue somatostatin has been detected<br />

in 5 of 9 [35] <strong>and</strong> 2 of 6 SCLC tumor sample<br />

extracts [37]. <strong>Somatostatin</strong> has also been detected<br />

in 1 of 13 culture media of SCLC cell<br />

lines [38] <strong>and</strong> extracted from 5 of 13 cell lines<br />

[39]. Finally Chretien <strong>and</strong> coworkers [40]<br />

reported immunohistochemically detectable<br />

somatostatin in 75% of cytology-positive<br />

bronchial brushing smear samples obtained<br />

from 24 SCLC patients. They concluded that<br />

the presence of somatostatin immunoreactivity,<br />

along with other features, was highly suggestive<br />

of SCLC. However, some studies either<br />

failed to detect, or have shown a very low<br />

incidence of somatostatin-like immunoreactivity<br />

in SCLC tumor samples. In 2 studies<br />

employing indirect immunoperoxidase tech-<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

niques only 1 of 94 [41] <strong>and</strong> 1 of 10 [42] paraffin-embedded<br />

SCLC samples were found to<br />

express somatostatin. Furthermore, in a recent<br />

study of cryostat sections from 4 SCLC<br />

tumors, in situ hybridization failed to detect<br />

evidence of somatostatin mRNA expression<br />

[43]. The significance of somatostatin expression<br />

in SCLC has not been addressed in these<br />

studies.<br />

We evaluated somatostatin immunoreactivity<br />

in plasma samples <strong>and</strong> tissue sections of<br />

biopsy <strong>and</strong> surgically resected paraffin-embedded<br />

tumor specimens of patients with<br />

SCLC. Elevated fasting plasma somatostatin<br />

levels were seen in 6/44 (13.6%) of the patients<br />

studied. Elevated somatostatin levels<br />

were also seen in 2 SCLC patients who had<br />

associated diabetes mellitus, 1 insulin-dependent<br />

<strong>and</strong> the other non-insulin-dependent.<br />

Diabetes mellitus is associated with elevated<br />

fasting somatostatin levels [44, 45]. The degree<br />

of elevation in the serum levels was similar<br />

in both the diabetic <strong>and</strong> nondiabetic SCLC<br />

patients. This observation raises the possibility<br />

that elevated somatostatin levels in SCLC<br />

patients may be secondary to the release of<br />

other hormones/peptides released by, or because<br />

of, the tumor rather than being ectopic<br />

in nature. This contention is supported by the<br />

fact that we were unable to detect definite<br />

somatostatin immunoreactivity in tissue sections<br />

from paraffin-embedded tumor samples<br />

from 163 patients with SCLC using streptavidin-biotin<br />

immunohistochemistry (unpubl.<br />

data).<br />

Further evidence to support the argument<br />

that the elevated somatostatin plasma levels<br />

are often not ectopic in nature comes from the<br />

finding of a close association between somatostatin<br />

<strong>and</strong> calcitonin gene-related peptide<br />

(CGRP) expression in the SCLC patients<br />

in our study (p = 0.019; unpubl. data). Recent<br />

work has shown that CGRP <strong>and</strong> somatostatin<br />

modulate chronic hypoxic pulmonary hyper-<br />

Chemotherapy 2001;47(suppl 2):78–108 81


tension. In in vivo experiments evaluating the<br />

role of CGRP in the regulation of pulmonary<br />

arterial pressure, infusion of this peptide into<br />

the pulmonary circulation of hypobaric hypoxic<br />

rats not only prevents pulmonary hypertension<br />

but also results in an increase in lung<br />

plasma somatostatin. Under these circumstances<br />

somatostatin has been shown to inhibit<br />

CGRP release [46]. Chronic hypoxic pulmonary<br />

hypertension may occur in patients<br />

with lung cancer <strong>and</strong> may explain, in part, the<br />

elevated CGRP <strong>and</strong>, as a result, somatostatin<br />

plasma levels observed in the SCLC patients<br />

[47].<br />

Recent work has shown that a proportion<br />

of non-SCLC (NSCLC) tumors may also express<br />

pan-neuroendocrine markers including<br />

NSE <strong>and</strong> chromogranin <strong>and</strong> specific peptide<br />

hormones including calcitonin, CGRP <strong>and</strong><br />

GRP [48–50]. The immunohistochemical expression<br />

of 2 or more neuroendocrine markers<br />

within a tumor <strong>and</strong> elevated serum NSE<br />

levels are observed in approximately 20% of<br />

patients with NSCLC <strong>and</strong> appear to define a<br />

subgroup of patients whose disease is more<br />

responsive to cytotoxic chemotherapy than<br />

neuroendocrine marker-negative disease [48,<br />

51, 52]. The immunohistochemical detection<br />

or extraction of somatostatin from NSCLC<br />

tumor tissue samples has been described in a<br />

number of small studies [42, 53, 54]. In their<br />

study of NSCLC tumors which had been<br />

found to contain dense-core granules (a marker<br />

of neuroendocrine differentiation) by electron<br />

microscopy, Wilson et al. [42] detected<br />

somatostatin immunoreactivity in all 7 cases.<br />

All had associated expression of NSE, human<br />

chorionic gonadotropin, serotonin <strong>and</strong> keratin.<br />

<strong>Somatostatin</strong> expression has also been<br />

described in a range of atypical pulmonary<br />

tumors including an adenocarcinoma with endometrioid<br />

features [55], an adenocarcinoma<br />

resembling fetal lung [56], a chondroma [57]<br />

<strong>and</strong> a blastoma [58]. We have evaluated a<br />

number of endocrine markers in serum <strong>and</strong><br />

plasma <strong>and</strong> paraffin-embedded tissue sections<br />

of patients with NSCLC. Although elevated<br />

serum NSE levels were seen in 7/25<br />

(28%) patients, somatostatin immunoreactivity<br />

was absent in all plasma <strong>and</strong> paraffinembedded<br />

tissue sections examined (unpubl.<br />

data).<br />

Bronchial carcinoids are likewise neuroendocrine<br />

tumors <strong>and</strong> express both pan-neuroendocrine<br />

markers <strong>and</strong> specific hormones,<br />

including GRP, gastrin, glucagon, calcitonin<br />

<strong>and</strong> adrenocorticotropin (ACTH) [59, 60]. In<br />

a recent study of 57 carcinoid tumors, somatostatin<br />

immunoreactivity was detectable<br />

in 43% of 30 typical <strong>and</strong> 22% of 27 atypical<br />

tumors [59].<br />

The balance of evidence of the a<strong>for</strong>ementioned<br />

findings suggests that somatostatin expression<br />

in SCLC <strong>and</strong> NSCLC is rare. Elevated<br />

plasma/serum levels may, in many instances,<br />

be a secondary response to the disease<br />

<strong>and</strong> other ectopic hormones produced by<br />

these tumors rather than as a result of ectopic<br />

production of the hormone from the tumor<br />

itself. This suggestion is supported by the<br />

close correlation between elevated somatostatin<br />

<strong>and</strong> CGRP levels observed in<br />

SCLC patients.<br />

<strong>Somatostatin</strong> Receptor (SSTR)<br />

Expression in Lung <strong>Cancer</strong><br />

The presence of SSTR on normal <strong>and</strong> malignant<br />

tumor cells has been the subject of<br />

considerable research since the early part of<br />

the last decade. SSTR are expressed by tumors<br />

derived from tissues known to be targets<br />

<strong>for</strong> somatostatin including somatotroph, thyrotroph<br />

<strong>and</strong> endocrine-inactive pituitary adenomas,<br />

central nervous system tumors including<br />

meningiomas, astrocytomas, oligodendrogliomas<br />

<strong>and</strong> medulloblastomas, gastrointesti-<br />

82 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


nal tract <strong>and</strong> pancreatic neuroendocrine<br />

(GEP) tumors, medullary thyroid carcinoma,<br />

<strong>and</strong> malignancies of the reticuloendothelial<br />

system including both Hodgkin’s <strong>and</strong> non-<br />

Hodgkin’s lymphomas. SSTR have also been<br />

detected on a proportion of breast, renal,<br />

prostatic, ovarian, exocrine pancreatic <strong>and</strong><br />

colorectal carcinomas <strong>and</strong> osteosarcomas [13,<br />

14, 61, 62].<br />

SCLC SSTR Expression<br />

The presence of SSTR on lung tumors was<br />

first demonstrated by Taylor et al. [63] in<br />

1988 when they characterized the in vitro<br />

binding of [ 125 I-Tyr 11 ]somatostatin-14 to<br />

membranes prepared from cultured human<br />

SCLC tumor cells derived from the classical<br />

SCLC cell line NCI-H69. Binding was monophasic<br />

<strong>and</strong> of high affinity with an equilibrium<br />

dissociation constant (Kd) = 0.59 B<br />

0.02 nM. The estimated maximum binding<br />

capacity (Bmax) was 173 B 2.4 fmol/mg protein.<br />

Specific binding sites were also detected<br />

on membrane preparations from solid NCI-<br />

H69 xenografts grown in athymic nude mice.<br />

However, while the calculated equilibrium Kd<br />

was similar, the Bmax was only about 10% of<br />

that observed in cell culture. The binding was<br />

specific in that biologically active analogs of<br />

somatostatin were potent inhibitors of [ 125 I-<br />

Tyr 11 ]somatostatin-14 binding whilst biologically<br />

inactive somatostatin analogs <strong>and</strong> unrelated<br />

compounds such as bombesin <strong>and</strong> gastrin-17<br />

were not.<br />

In subsequent experiments crude membrane<br />

preparations from xenografts of 2 other<br />

classic SCLC cell lines NCI-H345 <strong>and</strong> NCI-<br />

H209 were reported to express high-affinity<br />

SSTR. However, SSTR expression was not<br />

observed in the variant SCLC cell line NCI-<br />

H417 or the poorly differentiated solid SCLC<br />

tumor LX-1 [64, 65]. In vitro autoradiographic<br />

techniques have also been employed to study<br />

SSTR expression in SCLC tumor speci-<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

mens derived from cell lines <strong>and</strong> patients.<br />

SSTR have been demonstrated on fresh tumor<br />

specimens in 2 of 4 <strong>and</strong> 2 of 3 SCLC samples<br />

[66, 67]. The binding characteristics <strong>for</strong><br />

somatostatin were characterized in 1 of the 2<br />

SSTR-positive SCLC specimens in the study of<br />

Reubi et al. [66]. Specific high-affinity (Kd =<br />

0.53 nM), low-capacity (Bmax = 189 fmol/mg<br />

protein) binding sites were detected, results<br />

consistent with the previous studies in cell<br />

lines. Macaulay et al. [68] demonstrated specific<br />

SSTR on 3 of 4 SCLC cell line xenografts.<br />

Strong expression was detected in SCLC tumors<br />

derived from the SCLC cell line NCI-H69<br />

while weak, patchy expression was observed on<br />

tumors from the classic cell line HXI49 <strong>and</strong> the<br />

variant cell line ICR-SC17. Xenografts from<br />

the classic cell line HC12 were SSTR-negative.<br />

We evaluated the specific binding of radiolabelled<br />

RC-160 on membrane preparations<br />

from 9 SCLC cell lines. Specific binding sites<br />

<strong>for</strong> the octapeptide somatostatin analog were<br />

demonstrated on 6 of the 9 cell lines investigated<br />

(67%). Scatchard analysis of the displacement<br />

curve of the cell line NCI-H69<br />

revealed evidence of 2 specific binding sites, 1<br />

of high affinity <strong>and</strong> low capacity <strong>and</strong> the other<br />

of low affinity <strong>and</strong> high capacity [69]. The<br />

high-affinity binding site was in keeping with<br />

previous studies. The precise nature of the second<br />

binding site in our study was uncertain. As<br />

RC-160 binds with high affinity to SSTR-2<br />

<strong>and</strong> SSTR-5, moderate affinity to SSTR-3 <strong>and</strong><br />

low affinity to SSTR-1 <strong>and</strong> SSTR-4 [15], the<br />

results suggested that the high-affinity binding<br />

site represented SSTR-2 or SSTR-5 while the<br />

low-affinity site represented SSTR-1 or SSTR-<br />

4 or, indeed, nonspecific protein binding.<br />

Evaluation of SSTR subtype expression by<br />

RT-PCR has confirmed that SCLC tumors<br />

may express more than one SSTR subtype. In<br />

the first reported study, the classic SCLC cell<br />

lines NCI-H69 <strong>and</strong> NCI-H345 were found to<br />

express SSTR-1 <strong>and</strong> SSTR-2 mRNA [70].<br />

Chemotherapy 2001;47(suppl 2):78–108 83


This is entirely in keeping with the RC-160<br />

results presented here. A further study analyzed<br />

the binding activity of [ 125 I]somatuline<br />

<strong>and</strong> [ 125 I-Tyr 11 ]somatostatin-14 to crude homogenates<br />

of tumor xenografts from 3 SCLC<br />

cell lines, SCLC-6, SCLC-10 <strong>and</strong> SCLC-75.<br />

Employing cross-linking techniques 3 receptor<br />

proteins were identified. One major complex<br />

of 57 kD was bound by both radiolig<strong>and</strong>s in all<br />

3 tumors. Two other minor complexes were<br />

only identified by the natural lig<strong>and</strong> [ 125 I-<br />

Tyr 11 ]somatostatin-14, 90 kD in all 3 tumors<br />

<strong>and</strong> 70 kD in 2 tumors. Subsequent analysis<br />

revealed the presence of SSTR-1 mRNA in all 3<br />

tumors while SSTR-2 mRNA expression was<br />

observed in only 2. No SSTR-3 transcript was<br />

detected [71]. Finally RT-PCR analysis of the<br />

cell line COR-L103 revealed mRNA transcripts<br />

<strong>for</strong> SSTR-2, SSTR-3 <strong>and</strong> SSTR-4. Extra<br />

b<strong>and</strong>s were obtained by PCR-single str<strong>and</strong><br />

con<strong>for</strong>mation polymorphism analysis of the<br />

SSTR-2 gene. Sequence analysis of the SSTR-2<br />

gene demonstrated a point mutation in codon<br />

188 of TGG <strong>for</strong> tryptophan to TGA <strong>for</strong> a stop<br />

codon causing loss of 182 C-terminal amino<br />

residues of the receptor. The nucleotide sequences<br />

of SSTR-3 <strong>and</strong> SSTR-4 were normal.<br />

Using the radiolabelled somatostatin analog<br />

[ 125 I-Tyr 11 ]somatostatin-14, affinity crosslinking<br />

studies revealed a lack of expression of<br />

a 72-kD protein compared to the pituitary cell<br />

line AtT-20, indicating that SSTR-2 is not<br />

expressed on cell membranes of COR-L103<br />

cells due to this point mutation [72].<br />

Regarding SSTR-2 expression in SCLC,<br />

Taylor et al. [73] also detected an mRNA transcript<br />

in NCI-H69 corresponding to a truncated<br />

iso<strong>for</strong>m of SSTR-2. This was felt likely to represent<br />

a human homolog of the rodent receptor<br />

SSTR-2B. Immunoblotting analysis using the<br />

SSTR-2-specific antibody, 2e3, detected multiple<br />

immunoreactive protein species, including<br />

a predominant 150-kD molecule. The SSTR-2<br />

identity was confirmed by blocking detection<br />

of the predominant 150-kD b<strong>and</strong> by addition<br />

of the SSTR-2-derived 2e3 peptide. SSTR-2<br />

mRNA expression has also been identified in<br />

resected, snap-frozen, SCLC tumor specimens<br />

through in situ hybridization [74].<br />

In an RT-PCR study Fujita et al. [75] analyzed<br />

SSTR-1 <strong>and</strong> SSTR-2 expression in 9<br />

SCLC cell lines, including 5 classic <strong>and</strong> 4 morphologically<br />

variant cell lines. Both receptor<br />

subtypes were detectable in all SCLC cell<br />

lines. Collectively these results indicate that<br />

SSTR are present in between 50 <strong>and</strong> 100%,<br />

<strong>and</strong> specific high-affinity binding sites <strong>for</strong> radiolabelled<br />

somatostatin analogs in 50–75%<br />

of SCLC tumor samples studied.<br />

NSCLC SSTR Expression<br />

In contrast to SCLC, initial studies suggested<br />

that NSCLC tumors do not express SSTR.<br />

Using autoradiography, no specific binding<br />

of the radiolabelled somatostatin analogs<br />

[ 125 I-Tyr 3 ]octreotide or [ 125 I-Leu 8 , D-Trp 22 ,<br />

Tyr 125 ]somatostatin-28 was observed on cell<br />

pellets of the NSCLC tumors NCI-H23 or<br />

NCI-H226 [68], while only barely detectable<br />

levels of specific [ 125 I-Tyr 11 ]somatostatin-14<br />

binding were detected in membrane preparations<br />

of the NSCLC cell line H-165 [64]. Similarly<br />

in 2 studies in 1990, SSTR were not<br />

detected in 17 NSCLC fresh-frozen tumor<br />

samples using autoradiography [66, 67].<br />

Using the radiolabelled somatostatin analog<br />

[ 125 I-Tyr 11 ]somatostatin-14, we were unable<br />

to detect specific somatostatin binding<br />

sites on membrane preparations from xenografts<br />

of the adenocarcinoma NSCLC cell line<br />

NCI-H157 [26]. The results were in accord<br />

with the findings of previous NSCLC tumor<br />

sample studies [64–67]. However, we did detect<br />

a single class of specific binding sites <strong>for</strong><br />

[ 125 I]RC-160 in each of 3 NSCLC tumor samples<br />

[69]. All 3 were squamous cell tumors.<br />

The tissues examined were composed of<br />

a number of cell types including tumor<br />

84 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


cells, stroma, inflammatory cells <strong>and</strong> necrotic<br />

tissue. This observation raised the possibility<br />

that the NSCLC cells themselves were not<br />

expressing SSTR <strong>and</strong> that [ 125 I]RC-160 was<br />

binding specifically to inflammatory cells<br />

within the tumor which are known to express<br />

SSTR [62].<br />

Two recent studies support the suggestion<br />

that the [ 125 I]RC-160 may bind specifically to<br />

NSCLC cancer cells. Employing RT-PCR,<br />

Fujita et al. [75] demonstrated SSTR-1 <strong>and</strong><br />

SSTR-2 expression in a panel of squamous<br />

cell <strong>and</strong> adenocarcinoma NSCLC cell lines.<br />

The relative levels of SSTR-1 expression in<br />

the squamous cell tumors were similar to<br />

those seen in SCLC cell lines <strong>and</strong> were higher<br />

than those in adenocarcinomas. Indeed, in 2<br />

of the 5 adenocarcinomas examined, expression<br />

of SSTR-1 mRNA was very weak. Unlike<br />

SCLC, there was a positive correlation between<br />

SSTR-1 <strong>and</strong> SSTR-2 subtypes in the<br />

NSCLC tumor cell lines. In another study specific<br />

somatostatin binding sites <strong>for</strong> the radiolabelled<br />

hexapeptide somatostatin analog<br />

[ 125 I]MK-678 were found in 2 NSCLC cell<br />

lines, A 549 <strong>and</strong> H-165 [73]. This finding is of<br />

particular interest as previous studies with H-<br />

165 had shown this cell line to have only barely<br />

detectable levels of specific [ 125 I-Tyr 11 ]somatostatin-14<br />

binding [64]. As with the octapeptide<br />

somatostatin analog RC-160, MK-<br />

678 binds with high affinity to SSTR-2. Subsequent<br />

RT-PCR confirmed SSTR-2 mRNA<br />

expression within both NSCLC tumors [73].<br />

We have also evaluated somatostatin analog<br />

binding to bronchial carcinoid tissue [69].<br />

A single class of high-affinity binding site <strong>for</strong><br />

[ 125 1]RC-160 was detected in the membrane<br />

preparations of both tumors assessed. These<br />

findings were not unexpected given that<br />

SSTR expression is common to similar tumors<br />

in the gastrointestinal tract [61, 76].<br />

In summary, the available evidence indicates<br />

that the majority of SCLC tumors <strong>and</strong><br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

bronchial carcinoids express specific SSTR<br />

with high affinity <strong>for</strong> both hexa- <strong>and</strong> octapeptide<br />

somatostatin analogs. Although some<br />

controversy remains, recent results also indicate<br />

that a significant proportion of NSCLC<br />

tumors are SSTR-positive, SSTR-1 <strong>and</strong><br />

SSTR-2 mRNA having been detected in both<br />

squamous <strong>and</strong> adenocarcinomas.<br />

Radiolabelled <strong>Somatostatin</strong><br />

Analog Imaging in Lung <strong>Cancer</strong>:<br />

Therapeutic Implications<br />

In 1989, Krenning et al. [77] reported the<br />

use of the radiolabelled somatostatin analog<br />

[ 123 I-Tyr 3 ]octreotide <strong>for</strong> the detection <strong>and</strong><br />

localization of SSTR-positive tumors through<br />

scintigraphic imaging techniques. This agent<br />

was successfully applied in patients to visualize<br />

neuroendocrine GEP tumors, meningiomas<br />

<strong>and</strong> paragangliomas. Subsequent work<br />

led to the development of [ 111 In]pentreotide<br />

which is easier to make up, remains longer in<br />

the circulation <strong>and</strong> has a longer imaging halflife<br />

<strong>and</strong> reduced hepatobiliary metabolism<br />

than the <strong>for</strong>mer agent improving the quality<br />

of the images obtained, particularly in the liver.<br />

The effectiveness of [ 111 In]pentreotide has<br />

been established in the management of GEP<br />

tumors including carcinoid. The radiolabel<br />

often localizes primary GEP tumors that are<br />

undetectable by other methods including<br />

abdominal ultrasonography, computed tomography<br />

(CT) <strong>and</strong> magnetic resonance<br />

imaging (MRI), arteriography <strong>and</strong> venous<br />

sampling. [ 111 In]pentreotide may also detect<br />

otherwise unsuspected metastatic deposits.<br />

The localization of both the primary <strong>and</strong> metastatic<br />

deposits using a single technique contributes<br />

to the patient’s com<strong>for</strong>t <strong>and</strong> convenience<br />

<strong>and</strong> may lead to changes in patient<br />

management in a significant proportion of<br />

cases [61]. [ 111 In]pentreotide scintigraphy<br />

Chemotherapy 2001;47(suppl 2):78–108 85


is of proven value in predicting the responsiveness<br />

of hormone-associated conditions,<br />

arising as a result of GEP tumors, to octreotide<br />

therapy [78].<br />

SCLC [ 111 In]Pentetreotide Imaging<br />

SCLC remains a disease with a poor prognosis<br />

despite being sensitive to both chemotherapy<br />

<strong>and</strong> radiotherapy. SCLC patients are<br />

staged as having either limited or extensive<br />

disease. Limited disease, corresponding to<br />

stage I to IIIB disease in the TNM classification,<br />

is confined to a hemithorax including<br />

the mediastinum <strong>and</strong> the ipsilateral <strong>and</strong> contralateral<br />

hilar <strong>and</strong> scalene lymph nodes. Consolidation<br />

radiotherapy to the primary tumor<br />

site <strong>and</strong> mediastinum may improve long-term<br />

survival in patients with limited stage disease<br />

who have a good partial or complete response<br />

to chemotherapy. The median survival <strong>for</strong><br />

patients with limited disease is approximately<br />

15 months <strong>and</strong> 7–20% survive <strong>for</strong> 5 years or<br />

more. Extensive stage SCLC refers to the<br />

spread of the tumor beyond these boundaries<br />

<strong>and</strong> includes patients with IIIB disease with<br />

an associated malignant pleural effusion. In<br />

patients with extensive disease the median<br />

survival is approximately 9–12 months with<br />

few long-term survivors. There<strong>for</strong>e, stage has<br />

a significant impact on prognosis <strong>and</strong>, in<br />

some cases, on the management of patients [3,<br />

79].<br />

In keeping with the demonstration of<br />

SSTR in SCLC tumors, disease sites were<br />

visualized in 5 of 8 patients using [ 123 I-<br />

Tyr 3 ]octreotide [80]. On the basis of these<br />

observations we carried out a study to evaluate<br />

the efficacy of [ 111 In]pentetreotide scintigraphy<br />

as a staging modality in SCLC patients<br />

prior to chemotherapy <strong>and</strong> in the assessment<br />

of disease response after treatment.<br />

Thirteen patients with newly diagnosed<br />

SCLC were studied prior to receiving chemotherapy.<br />

Following st<strong>and</strong>ard staging investiga-<br />

tions, 6 patients were found to have limited<br />

disease, <strong>and</strong> 7 extensive disease. Of the 7<br />

patients with extensive disease, liver metastases<br />

were seen in 4, bony metastases in 4, a<br />

single large brain metastasis in 1 <strong>and</strong> an adrenal<br />

metastasis in 1. Scintigraphic imaging<br />

with [ 111 In]pentetreotide led to the detection<br />

of all primary sites of disease. This included a<br />

patient whose primary tumor, detected at<br />

bronchoscopy, was not visualized with chest<br />

x-ray (CXR) or CT of the thorax. In the 7<br />

patients with extensive disease metastatic disease<br />

was detected in 3 out of 4, <strong>and</strong> skeletal<br />

disease in 2 out of 4 patients. In 1 patient a<br />

previously undetected cerebellar metastasis<br />

was found, not suspected following routine<br />

staging. This was later confirmed with a CT<br />

brain scan. There<strong>for</strong>e the imaging procedure<br />

correctly staged 9 out of 13 patients (69%),<br />

detecting 5 out of 10 known metastases <strong>and</strong><br />

1 previously unknown disease site (56%). Disease<br />

was downstaged in 4 out of 7 patients<br />

with extensive disease <strong>and</strong> upstaged in 1 patient<br />

with limited disease. In summary,<br />

[ 111 In]pentetreotide detected secondaries in 4<br />

or 50% of the 8 patients found to have metastases<br />

at the completion of all investigations<br />

[81].<br />

We subsequently assessed a further 10 patients<br />

with SCLC. Eight of these were evaluated<br />

during chemotherapy. [ 111 In]pentetreotide<br />

failed to localize the primary tumor in 1<br />

patient <strong>and</strong> detected metastases in only 1 of 6<br />

patients with extensive disease. Of the remaining<br />

2 patients 1 was imaged at the time<br />

of disease relapse. The patient had extensive<br />

disease. All known tumor sites were localized.<br />

The other patient was imaged <strong>for</strong> what was<br />

thought to be NSCLC. Both st<strong>and</strong>ard <strong>and</strong><br />

[ 111 In]pentetreotide imaging indicated resectable<br />

disease. Histological evaluation of the<br />

resected specimen following surgery revealed<br />

a limited stage SCLC tumor (table 2) [82].<br />

86 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


Four patients imaged prior to treatment, 2<br />

with limited disease <strong>and</strong> 2 with extensive disease,<br />

were reevaluated with [ 111 In]pentetreotide.<br />

A further patient with extensive disease<br />

who was imaged with [ 111 In]pentetreotide after<br />

completion of his first cycle of chemotherapy<br />

was also assessed after completion of<br />

treatment. Pathological uptake of the radiolabel<br />

was detected in the region of the original<br />

disease in 2 patients thought to be in complete<br />

remission following st<strong>and</strong>ard staging procedures.<br />

In 1 of these patients a subsequent<br />

MRI study demonstrated an abnormality<br />

away from the bronchus suggestive of residual<br />

disease. The patient subsequently relapsed at<br />

this site. There<strong>for</strong>e the technique may allow a<br />

more accurate assessment of prognosis <strong>for</strong> the<br />

individual patient following completion of<br />

chemotherapy <strong>and</strong> aid subsequent management<br />

decisions [81].<br />

Single photon emission computed tomography<br />

(SPET) imaging was per<strong>for</strong>med in 9<br />

patients <strong>and</strong> improved the anatomical localisation<br />

of disease in the thorax but contributed<br />

little to the overall assessment.<br />

The failure of [ 111 In]pentreotide scintigraphy<br />

to visualize all disease sites in 9 of the 15<br />

patients (60%) with metastases outside the<br />

thorax, despite detecting the primary tumor<br />

in 14 of these cases, is unclear. Nonspecific<br />

uptake of the radiolabel seen in the spleen,<br />

kidneys <strong>and</strong> urinary tract, liver <strong>and</strong> gastrointestinal<br />

tract, pituitary <strong>and</strong> thyroid gl<strong>and</strong> may<br />

obscure visualization of metastases to these<br />

areas. However, the lack of uptake of radiolabel<br />

by bone <strong>and</strong> brain deposits cannot be<br />

explained in this way. This raises a number of<br />

possibilities. In a proportion of cases the metastatic<br />

disease may represent a dedifferentiated<br />

clone of the primary SCLC tumor not<br />

expressing SSTR. In individual cases local<br />

factors may downregulate SSTR expression.<br />

Furthermore, binding of the radiolig<strong>and</strong> to<br />

the receptor may be blocked if high local lev-<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

Table 2. Summary of results of SCLC patient imaging<br />

studies (n = 23)<br />

Sites of<br />

disease<br />

Number of disease sites detected<br />

st<strong>and</strong>ard<br />

staging<br />

Primary<br />

22 22a Liver 10 5<br />

Adrenal 1 0<br />

Bone 6 3<br />

Brain 1 1b [ 111 In]pentetreotide<br />

imaging<br />

a [ 111 In]pentetreotide imaging detected an intrathoracic<br />

primary SCLC not seen on CXR or CT scan of<br />

the thorax.<br />

b An otherwise unsuspected cerebellar metastasis.<br />

On completion of the study metastases were not<br />

visualized in 9 out of 15 patients with proven metastatic<br />

disease. [ 111 In]pentetreotide imaging failed to detect<br />

any site of disease in one patient with extensive<br />

SCLC.<br />

els of endogenous somatostatin are being produced.<br />

Furthermore, in those patients imaged<br />

while receiving chemotherapy the primary tumor<br />

was seen in 7 of 8 patients but metastases<br />

were detected in only 1 of 6 patients with metastatic<br />

disease. This observation suggests that<br />

the chemotherapy itself may affect uptake of<br />

the radiolabel. This suggestion is supported<br />

by the results of a recent study in which the<br />

uptake of [ 111 In]pentetreotide by primary<br />

SCLC tumors was significantly lower in patients<br />

imaged during chemotherapy compared<br />

to those evaluated be<strong>for</strong>e treatment (tumor<br />

to background ratios = 1.94 B 0.79 vs.<br />

2.35 B 0.9; p ! 0.005) [83].<br />

In image-negative metastatic disease the<br />

SCLC cells themselves may not be expressing<br />

SSTR-2 or SSTR-5, the subtypes known to<br />

bind octreotide with high affinity. Rather it<br />

may be that the primary tumor is being visualized<br />

due to uptake of the radiolabel by the local<br />

Chemotherapy 2001;47(suppl 2):78–108 87


inflammatory response, as activated immune<br />

cells are known to express SSTR [62] <strong>and</strong>/or<br />

by SSTR expressed in high concentrations on<br />

peritumoral blood vessels [12, 84]. If this is<br />

so, then those tumors where the metastases<br />

are seen may represent the true proportion of<br />

metastatic SCLC tumors which express highaffinity<br />

binding sites <strong>for</strong> radiolabelled somatostatin<br />

analogs – 6 of 15 (40%) patients with<br />

extensive disease evaluated in this study.<br />

This would be in keeping with the proportion<br />

of SCLC tumors known to express specific<br />

high-affinity binding sites <strong>for</strong> radiolabelled<br />

somatostatin analogs in vitro as discussed<br />

earlier.<br />

The effectiveness of [ 111 In]pentetreotide<br />

scintigraphy in the detection <strong>and</strong> localization<br />

of disease sites in SCLC patients has been<br />

evaluated <strong>and</strong> verified in a number of studies<br />

[83, 85–94]. Pretreatment of SCLC patients<br />

with cold somatostatin prior to administration<br />

of [ 111 In]pentetreotide may enhance the<br />

visualization of metastases through increased<br />

uptake of the radiolabel by the tumor [91, 95].<br />

However, the lack of visualization of all sites<br />

of disease indicates that the imaging technique<br />

has limited value as a single modality in<br />

the staging of SCLC prior to commencing chemotherapy.<br />

Nonetheless a recent cost-effectiveness<br />

analysis indicated that [ 111 In]pentetreotide<br />

scintigraphy should be per<strong>for</strong>med<br />

in patients with SCLC as it may alter<br />

the staging of the disease <strong>and</strong> localize otherwise<br />

undetected brain metastases. Under<br />

these circumstances the cost increase is outweighed<br />

by changes in patient management<br />

<strong>and</strong> the possibility of irradiating brain metastases<br />

at an early stage which may ultimately<br />

lead to improvements in symptom control<br />

[94].<br />

Novel approaches to therapy in SCLC are<br />

currently being assessed <strong>and</strong> include inhibiting<br />

the action of autocrine growth factors<br />

such as GRP with GRP antagonists <strong>and</strong> GRP<br />

receptor antibodies [5]. <strong>Somatostatin</strong> analogs<br />

are currently being evaluated as possible therapeutic<br />

agents in SCLC with encouraging results<br />

in vitro <strong>and</strong> in vivo. In acromegaly <strong>and</strong><br />

GEP tumors, the presence of SSTR has been<br />

shown to predict responsiveness to somatostatin<br />

analog therapy [78, 96, 97]. Imaging of<br />

SCLC tumors with [ 111 In]pentetreotide may<br />

have a role to play in identifying those patients<br />

most likely to respond to somatostatin<br />

analog treatment. Of great significance in this<br />

respect is the efficacy of the radiolabel in<br />

detecting residual SCLC disease following<br />

chemotherapy, <strong>and</strong> relapsing disease, observations<br />

which suggest that treated SCLC tumors<br />

continue to express SSTR. This lays the<br />

groundwork <strong>for</strong> evaluation of somatostatin<br />

analogs as therapeutic agents in the treatment<br />

of chemotherapeutically debulked or relapsing<br />

SCLC in the future.<br />

NSCLC [ 111 In]Pentetreotide Imaging<br />

In NSCLC the most important decisions to<br />

reach are whether the patient has resectable<br />

disease <strong>and</strong>, if so, is fit <strong>for</strong> surgery. If a patient<br />

is being considered <strong>for</strong> surgery extensive investigations<br />

are recommended including a<br />

full blood count, biochemistry profile, CXR<br />

<strong>and</strong> CT thorax, liver <strong>and</strong> adrenals. If mediastinal<br />

lymph nodes greater than 1.5 cm in the<br />

shortest transverse diameter are seen on the<br />

CT scan, mediastinoscopy <strong>and</strong>/or mediastinotomy<br />

should be per<strong>for</strong>med to exclude inoperable<br />

disease prior to definitive surgery. Any<br />

symptoms the patient has which may be attributable<br />

to metastatic disease should be<br />

thoroughly investigated, e.g. an isotope bone<br />

scan <strong>for</strong> bone pain <strong>and</strong> a CT brain scan to<br />

assess headaches. Recent evidence suggests<br />

that stage III disease in patients with an Eastern<br />

Cooperative Oncology Group (ECOG)<br />

per<strong>for</strong>mance status of 0, 1 <strong>and</strong> possibly 2 may<br />

benefit from chemotherapy followed by radical<br />

radiotherapy. Furthermore, while still<br />

88 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

Fig. 1. Anterior <strong>and</strong> posterior views of the thorax <strong>and</strong> upper abdomen of 2 patients with<br />

NSCLC. A A stage IIIB tumor with abnormal uptake of [ 111 In]pentetreotide throughout the<br />

right lung <strong>and</strong> mediastinum confirmed both radiologically <strong>and</strong> at surgery is clearly demonstrated.<br />

B Increased uptake in the right upper lobe consistent with a Pancoast tumor confirmed<br />

with CT imaging is shown.<br />

largely investigational, there is accumulating<br />

evidence that preoperative, neoadjuvant chemotherapy<br />

may downstage the disease in a<br />

proportion of patients with stage IIIA, N2 disease<br />

making resection possible. These patients<br />

likewise require intensive staging to exclude<br />

disseminated disease. If the disease is<br />

inoperable due to the patient’s health <strong>and</strong><br />

radical radiotherapy is not deemed appro-<br />

priate, or if stage IV disease is established,<br />

then staging investigations may be kept to a<br />

minimum [4, 98].<br />

We studied [ 111 In]pentetreotide scintigraphy<br />

in 23 patients with NSCLC prior to treatment.<br />

The primary tumor was detected in all<br />

cases (fig. 1). Furthermore, the resectibility of<br />

the disease was correctly established in 20<br />

cases. The value of the technique was best<br />

Chemotherapy 2001;47(suppl 2):78–108 89


exemplified in 2 patients. In the first disease<br />

spread was seen with [ 111 In]pentetreotide<br />

imaging which was not initially suspected. In<br />

the second patient a small primary squamous<br />

cell cancer found at bronchoscopy, but not<br />

visualized on CXR or CT thorax <strong>and</strong> upper<br />

abdomen, was localized. Of the 3 patients<br />

incorrectly staged, 2 were as the result of falsepositive<br />

uptake in benign lesions in the mediastinum,<br />

in a thyroid colloid cyst associated<br />

with intense lymphocytic infiltration in one<br />

case <strong>and</strong> in granulomatous lymph nodes in the<br />

other. These findings are readily explained by<br />

the fact that specific high-affinity SSTR may<br />

be expressed by immune cells, including those<br />

present in granulomas [61, 62]. In a third<br />

patient, mediastinal involvement <strong>and</strong> a malignant<br />

pleural effusion were missed. In a further<br />

patient the tumor, thought to be T2, N0/N1<br />

disease, turned out on CT of the thorax <strong>and</strong><br />

upper abdomen <strong>and</strong> at surgery to extend to<br />

<strong>and</strong> involve the chest wall pleura indicating T3<br />

disease. As with CT scanning, it is difficult to<br />

distinguish between tumor <strong>and</strong> postobstructive<br />

atelectasis or consolidation as inflammatory<br />

cells may express high-affinity SSTR.<br />

Nonetheless, this did not affect staging of disease<br />

in our patient series (unpubl. data).<br />

[ 111 In]pentetreotide SPET imaging was<br />

per<strong>for</strong>med in 16 of the 23 NSCLC patients.<br />

Unlike SCLC, SPET imaging significantly affected<br />

staging in a number of cases by improving<br />

anatomical localization of the disease, indicating,<br />

in particular, spread of disease to the<br />

pleura/chest wall (unpubl. data).<br />

Although these results are encouraging,<br />

they also demonstrate some of the potential<br />

limitations of the technique. While the radiolabel<br />

is specific <strong>for</strong> SSTR-expressing tissues, it<br />

is not specific <strong>for</strong> individual pathological processes<br />

as demonstrated by the detection of<br />

benign disease in the mediastinum in 2 patients<br />

in this series. Likewise it is well established<br />

that not all tumors of a particular<br />

pathological subtype necessarily express highaffinity<br />

binding sites <strong>for</strong> radiolabelled somatostatin<br />

analogs [61, 99]. Furthermore, it is<br />

likely that in some cases individual tumors<br />

may not express SSTR in sufficient density to<br />

be visualized by [ 111 In]pentetreotide scintigraphy.<br />

There<strong>for</strong>e, false-negative results may<br />

occur. While the primary tumor may be localized,<br />

metastases may not necessarily contain<br />

SSTR-expressing cells. Also, although not a<br />

problem in the NSCLC studies reported here,<br />

nonspecific uptake of the radiolabel seen in<br />

the spleen, kidneys <strong>and</strong> urinary tract, liver<br />

<strong>and</strong> gastrointestinal tract, pituitary <strong>and</strong> thyroid<br />

gl<strong>and</strong> may obscure visualization of metastases<br />

to these areas.<br />

Two further studies have confirmed that<br />

[ 111 In]pentetreotide scintigraphy is effective<br />

in the localization of NSCLC tumors, detecting<br />

the primary tumor in 40 of 40 [86] <strong>and</strong> 10<br />

of 13 patients, respectively [87]. However,<br />

metastatic sites of disease were not as frequently<br />

detected, thereby demonstrating the<br />

limitations of the technique. Of 15 patients<br />

with known metastases in the study of Kwekkeboom<br />

et al. [86], disease spread was detected<br />

in only 6, including 1 patient with a<br />

previously undetected brain metastasis – this<br />

was subsequently confirmed on CT brain<br />

scan. Spread to the mediastinum was detected<br />

in 5 of 15 cases <strong>and</strong> to the opposite lung in 1 of<br />

3 cases. Hepatic, adrenal <strong>and</strong> bone metastases<br />

were also missed. In the second study, while<br />

metastatic disease was detected in 9 of 10<br />

patients, spread to the liver was not detected<br />

in any of the 5 patients with known hepatic<br />

metastases [87].<br />

The finding of uptake in NSCLC tumors<br />

was, to some extent, unexpected as the in<br />

vitro experimental data available prior to<br />

commencing the [ 111 In]pentetreotide studies<br />

had failed to reveal specific binding sites <strong>for</strong><br />

a number of radiolabelled somatostatin<br />

analogs [64, 66, 67]. As discussed earlier we<br />

90 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


demonstrated the presence of high-affinity<br />

binding sites <strong>for</strong> the radiolabelled somatostatin<br />

analog [ 125 I]RC-160 in 3 squamous cell<br />

lung cancers. Two of these tumors had been<br />

localized with [ 111 In]pentreotide prior to surgical<br />

resection. Histological assessment of the<br />

resected specimens carried out be<strong>for</strong>e conducting<br />

the binding assays showed them to<br />

be comprised of tumor cells, necrotic tissue,<br />

connective tissue <strong>and</strong>/or inflammatory cells.<br />

There<strong>for</strong>e, the SSTR may have been expressed<br />

by activated immune cells [62] <strong>and</strong>/<br />

or the cancer cells themselves. The imaging<br />

study of Kwekkeboom et al. [86] is of interest<br />

in this regard. They argued that the detection<br />

<strong>and</strong> localization of NSCLC tumors was due to<br />

uptake of the radiolabel in tissues surrounding<br />

the tumor rather than to uptake by the<br />

tumor cells. This suggestion was based on<br />

observations that in a number of cases in<br />

their series (1) a halo effect was seen around<br />

the tumor on planar imaging, (2) SPET imaging<br />

showed increased radioactivity at the periphery<br />

of the tumor <strong>and</strong> (3) the region of<br />

accumulation of radioactivity during scintigraphy<br />

was greater than the tumor as measured<br />

on CT. Following [ 111 In]pentetreotide<br />

scintigraphy a number of these patients underwent<br />

surgery. Subsequent autoradiographic<br />

evaluation of the resected tumors using the<br />

radiolabelled octapeptide somatostatin analog<br />

[ 125 I-Tyr 3 ]octreotide failed to reveal SSTR<br />

on tumor cells.<br />

However, subsequent work demonstrated<br />

the presence of SSTR in NSCLC tumor cells<br />

[73, 75, 100]. As discussed earlier, in one of<br />

these studies moderate concentrations of specific<br />

binding sites <strong>for</strong> the radiolabelled hexapeptide<br />

somatostatin analog [ 125 I]MK-678,<br />

which binds SSTR-2 with high affinity, were<br />

found in 2 NSCLC cell lines. Subsequent RT-<br />

PCR confirmed SSTR-2 mRNA expression<br />

within both NSCLC tumors [73]. There<strong>for</strong>e, it<br />

would appear likely that the localization of<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

some NSCLC tumors is due, at least in part,<br />

to specific uptake of [ 111 In]pentetreotide by<br />

the malignant cells.<br />

The results of the present study suggest<br />

that [ 111 In]pentetreotide imaging may have a<br />

role to play as an adjunct in the diagnostic<br />

evaluation of patients with NSCLC. The<br />

imaging studies in SCLC <strong>and</strong> those reported<br />

<strong>for</strong> other tumors [61, 81, 101, 102] demonstrate<br />

that [ 111 In]pentetreotide scintigraphy<br />

may be of value in monitoring the response of<br />

SSTR-positive lung tumors to chemotherapy.<br />

As cytotoxic chemotherapy now plays an increasingly<br />

important role in the management<br />

of patients with NSCLC, [ 111 In]pentetreotide<br />

scintigraphy may prove to be an effective<br />

radiological technique <strong>for</strong> assessing the response<br />

of image-positive NSCLC tumors to<br />

such therapy. Furthermore, [ 111 In]pentetreotide<br />

scintigraphy may be of particular importance<br />

in the evaluation of the response of<br />

NSCLC tumors to novel treatments, including<br />

somatostatin analog treatment.<br />

Bronchial Carcinoid [ 111 In]Pentetreotide<br />

Imaging Studies<br />

We have reported the localization of bronchial<br />

carcinoid in 2 of the 3 patients studied<br />

with [ 111 In]pentetreotide scintigraphy [103].<br />

As in the NSCLC patients SPET not only<br />

improved anatomical localization of the<br />

disease but also was necessary to detect<br />

the tumor in one case. In the patient where<br />

the primary tumor was not localized with<br />

[ 111 In]pentetreotide scintigraphy, diffuse, low<br />

intensity uptake of the radiolig<strong>and</strong> was noted<br />

in the opposite lung, an area of known bronchiectasis.<br />

This false-positive localization is in<br />

keeping with the findings in the NSCLC patients<br />

described earlier where 2 false-positive<br />

results were recorded. Following resection of<br />

the carcinoid tumor in one of the imagepositive<br />

cases, membrane-binding assays revealed<br />

the presence of a single class of high-<br />

Chemotherapy 2001;47(suppl 2):78–108 91


affinity binding site <strong>for</strong> the radiolabelled somatostatin<br />

analog [ 125 I]RC-160 [103].<br />

Apart from presenting with respiratory<br />

symptoms, symptoms of metastatic disease,<br />

physical signs <strong>and</strong> associated abnormalities<br />

on CXR or CT scan, <strong>and</strong>, rarely, the carcinoid<br />

syndrome, patients with bronchial carcinoid<br />

tumors may present with ectopic hormone<br />

syndromes including Cushing’s syndrome<br />

<strong>and</strong> acromegaly. In many cases the tumor<br />

producing the ectopic-hormone-related syndrome<br />

cannot be localized on physical examination<br />

or with st<strong>and</strong>ard biochemical <strong>and</strong> radiological<br />

techniques. [ 111 In]pentetreotide<br />

scintigraphy has now been evaluated in a<br />

series of patients with Cushing’s syndrome<br />

secondary to ectopic ACTH or corticotropinreleasing<br />

factor. In keeping with the results<br />

reported <strong>for</strong> neuroendocrine tumors in general,<br />

the primary tumor or its metastases were<br />

localized in 8 of 10 patients. The failure of<br />

[ 111 In]pentetreotide scintigraphy to detect<br />

ACTH-secreting pituitary adenomas <strong>and</strong> an<br />

adrenal tumor in 9 patients is in accord with<br />

the ineffectiveness of somatostatin analog<br />

therapy in the management of Cushing’s disease<br />

[104].<br />

Subsequent studies have confirmed these<br />

results <strong>and</strong> indicate that SSTR-2 is expressed<br />

by bronchial carcinoid tumors [104–113]. The<br />

results suggest SSTR imaging may be an important<br />

diagnostic investigation in the workup<br />

of patients with suspected ectopic hormonesecreting<br />

tumors <strong>and</strong> their metastases.<br />

Radiotherapeutics<br />

The possibility of employing a radiolabelled<br />

chelated somatostatin analog as a radiotherapeutic<br />

agent in treating SSTR-positive<br />

tumors, including those of the lung, is an<br />

exciting prospect <strong>for</strong> the future. Since somatostatin<br />

analogs are based on sequences of<br />

the native hormone, they rarely induce immunization.<br />

The accumulation of [ 111 In]pentetreotide<br />

in gastrointestinal neuroendocrine tumors is<br />

between 0.0123 <strong>and</strong> 0.2% of the administered<br />

dose per gram of tumor tissue. The rapid<br />

clearance of the radiolabel from the blood, the<br />

relatively low accumulation in the liver (1.9<br />

<strong>and</strong> 2.2% of the administered dose at 4 <strong>and</strong><br />

24 h, respectively) with resultant relatively<br />

low excretion into the gastrointestinal tract<br />

<strong>and</strong> the predominant renal clearance are advantageous<br />

in this regard. However, the<br />

amount of renal accumulation <strong>and</strong> the relatively<br />

long renal effective half-life may limit<br />

the maximally applicable radiation dose [61,<br />

99]. Recent studies suggest that there are a<br />

number of approaches which may be useful in<br />

reducing the uptake of [ 111 In]pentetreotide by<br />

normal tissues.<br />

Pretreatment of patients with SCLC <strong>and</strong><br />

carcinoid disease with cold octreotide increases<br />

tumor uptake of [ 111 In]pentetreotide,<br />

whereas uptake into the liver, spleen <strong>and</strong> kidneys<br />

is reduced [91, 109]. This in vivo observation<br />

has been supported in vitro in mouse<br />

AtT20/Dl6V pituitary tumor cells <strong>and</strong> primary<br />

cultures of human GH-secreting pituitary tumor<br />

cells. After 4 h of incubation up to 8% of<br />

the added [ 125 I-Tyr 3 ]octreotide had accumulated<br />

in AtT20/Dl6V cells <strong>and</strong> between 0.24<br />

<strong>and</strong> 4.98% in 6 of the 7 human tumor cultures.<br />

This accumulation was specific <strong>and</strong> time-, temperature-<br />

<strong>and</strong> pertussis-toxin-sensitive G protein-dependent.<br />

Displacement of binding <strong>and</strong><br />

internalization of [ 125 I-Tyr 3 ]octreotide in<br />

AtT20/Dl6V cells by the addition of unlabelled<br />

octreotide showed a bell-shaped curve. At low<br />

concentrations (0.1–1 nM) unlabelled octreotide<br />

enhanced the binding <strong>and</strong> internalization<br />

of [ 125 I-Tyr 3 ]octreotide whereas at higher concentrations<br />

saturation occurred. Furthermore,<br />

after 4 h of incubation, 88% of the radioactivity<br />

present in the cells was still peptide-bound<br />

suggesting slow intracellular breakdown of the<br />

radiolig<strong>and</strong> [95].<br />

92 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


Infusion of amino acids, known to block<br />

renal tubular peptide reabsorption, significantly<br />

reduces renal parenchymal uptake of<br />

[ 111 In]pentetreotide 4 h after administration<br />

when compared to controls. In this there was<br />

a nonsignificant increase in urinary clearance<br />

of the isotope over the 4 h consistent with<br />

reduced re-uptake of the radiolabel, <strong>and</strong> a lack<br />

of effect of the amino acids or radiolabelled<br />

peptide on glomerular filtration rate [114].<br />

A number of new potential radiotherapeutic<br />

agents are currently under investigation.<br />

188 Re, a potential radiotherapeutic isotope,<br />

has been coupled to octreotide. [ 188 Re]octreotide<br />

has been assessed <strong>for</strong> the in vivo localization<br />

of NCI-H69 SCLC tumor xenografts in<br />

athymic nude mice <strong>and</strong> proved as efficient as<br />

[ 111 In]pentetreotide in this regard [115]. Likewise,<br />

l6l Tb has been coupled to octreotide. In<br />

studies in the rat, l6l Tb-DTPA-octreotide has<br />

similar distribution characteristics to [ 111 In]pentetreotide<br />

but with less uptake in the liver<br />

<strong>and</strong> other SSTR-expressing tissues such as the<br />

pancreas <strong>and</strong> adrenals, <strong>and</strong> negligible excretion<br />

into the bile. The uptake of l6l Tb-DTPAoctreotide<br />

by the renal tubules after glomerular<br />

filtration can be reduced by administration<br />

of lysine or sodium maleate [116]. The<br />

radiotherapeutic isotope 64 Cu, a reactor-produced<br />

radionuclide, has been coupled to octreotide.<br />

Apart from its radiotherapeutic potential,<br />

64 Cu is also a suitable radioisotope <strong>for</strong><br />

PET scanning. The resulting product, 64 Cu-<br />

TETA-octreotide, binds to SSTR with 5 times<br />

the affinity of [ 111 In]pentetreotide <strong>and</strong>, like<br />

[ 111 In]pentetreotide, is excreted principally<br />

through the renal system [117]. Finally, [ 90 Y]<br />

has been linked to DTPA-benzyl-acetamido-<br />

D-Phe 1 , Tyr 3 -octreotide <strong>and</strong> has proved effective<br />

in the treatment of SSTR-positive tumors<br />

in a nude mouse model [118]. Other somatostatin<br />

analogs are being linked to therapeutic<br />

radioisotopes with encouraging results.<br />

[ 188 Re]RC-160 has been evaluated in vivo in<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

nude mice bearing xenografts of the SSTRpositive<br />

prostate cancer cell lines DU-145 <strong>and</strong><br />

PC-3. In PC-3 xenografts, [ 188 Re]RC-160 induced<br />

a dose-dependent therapeutic response,<br />

with disease stabilization or shrinkage, even<br />

in animals with relatively large tumor masses<br />

[119].<br />

These results demonstrate that, in selected<br />

cases, radiolabelled somatostatin analogs may<br />

be employed as adjuncts to st<strong>and</strong>ard radiological<br />

techniques <strong>for</strong> the scintigraphic localization<br />

of primary lung tumors <strong>and</strong> their metastases.<br />

Radiotherapeutic somatostatin analogs<br />

may have a role to play in the treatment of<br />

lung tumors, in particular chemotherapeutically<br />

debulked SCLC, in the future.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in the<br />

<strong>Treatment</strong> of Lung <strong>Cancer</strong><br />

Given the mode of action of somatostatin<br />

<strong>and</strong> the fact that between 50 <strong>and</strong> 100% of<br />

SCLC cell lines <strong>and</strong> fresh frozen tissue specimens<br />

express SSTR, the role of somatostatin<br />

<strong>and</strong> somatostatin analogs as possible therapeutic<br />

agents in SCLC has been investigated<br />

in a number of studies. Initial experiments<br />

with somatostatin-14 were not encouraging<br />

[120]. In 1988 Kee et al. [121] demonstrated<br />

that somatostatin could inhibit the secretion<br />

of bombesin-like peptides from SCLC cell<br />

lines. Subsequently in 1988 Taylor et al. [122]<br />

evaluated the somatostatin analog somatuline<br />

in the treatment of NCI-H69 in vitro. Significant<br />

inhibition of cell proliferation was demonstrated.<br />

The average cell concentration of<br />

treated samples was 59% of that observed<br />

in tumor controls after 72 h. Further work<br />

with NCI-H345 revealed that somatuline<br />

could not only inhibit the clonal growth of<br />

SCLC cell lines but could also inhibit VIPinduced<br />

cAMP <strong>for</strong>mation within the tumor<br />

Chemotherapy 2001;47(suppl 2):78–108 93


cells [123]. Somatuline inhibits serum-induced<br />

MAP kinase activation in SCLC cells at<br />

a dose range similar to that required to inhibit<br />

cell proliferation. This observation suggests<br />

that phosphatase activation is important <strong>for</strong><br />

the antiproliferative effects of somatostatin<br />

analogs in SCLC as discussed earlier [10, 11,<br />

14, 20, 21]. These results suggest that the<br />

SSTR on SCLC tumors are functional <strong>and</strong><br />

interact with growth pathways within the<br />

cell.<br />

Following the promising results of the in<br />

vitro work, in vivo experiments in athymic<br />

nude mice were per<strong>for</strong>med. Suspensions of<br />

NCI-H69 cells, containing 5 ! l0 6 cells, were<br />

injected into the flanks of athymic nude mice.<br />

These tumors were treated with somatuline<br />

500 Ìg i.p. b.i.d. in one group <strong>and</strong> s.c. around<br />

the tumor in another. Both regimens resulted<br />

in delay of the mean tumor lag time compared<br />

to controls. Inhibition of tumor growth was<br />

dramatic with subcutaneous injection around<br />

the tumor site, with virtually no growth being<br />

seen at 48 days. Discontinuing treatment by<br />

any route resulted in an acceleration of tumor<br />

growth. The effects of somatuline treatment<br />

on NCI-H69 xenografts were also studied.<br />

The experiment was repeated as be<strong>for</strong>e but<br />

with a third group receiving subcutaneous<br />

somatostatin analog treatment at a site distant<br />

from the tumor. Again significant growth inhibition<br />

was seen in all treated groups [122].<br />

Subsequent work with somatuline in SSTRpositive<br />

cell lines has confirmed the earlier<br />

findings [64, 124]. Furthermore, significant<br />

growth inhibition of the SSTR-negative variant<br />

SCLC cell line NCI-H417 <strong>and</strong> poorly differentiated<br />

SCLC cell line LX-1 by somatostatin<br />

analogs was seen suggesting that they may<br />

inhibit SCLC tumor growth by indirect as<br />

well as direct means [64]. The hexapeptide<br />

somatostatin analog, MK-678, has also been<br />

demonstrated to affect NCI-H69 tumor<br />

growth in vivo. <strong>Treatment</strong> with 300 Ìg/kg s.c.<br />

t.i.d. <strong>for</strong> 46 days resulted in a 51% decrease in<br />

tumor area, a 64% decrease in tumor DNA<br />

content <strong>and</strong> a 55% decrease in RNA content.<br />

Tumor protein <strong>and</strong> weight were not affected<br />

[125].<br />

The effect of octreotide in the treatment of<br />

SCLC has also been assessed in vitro <strong>and</strong> in<br />

vivo. The SCLC lines NCI-H69, HX149,<br />

ICR-SC17 <strong>and</strong> HC12 were evaluated. The<br />

growth of the SCLC cell line HX149, found to<br />

have weak <strong>and</strong> patchy specific somatostatin<br />

binding sites at autoradiography, was significantly<br />

inhibited but that of the other cell lines,<br />

two of which were SSTR-positive, was unaltered<br />

[68].<br />

We evaluated the efficacy of RC-160 in the<br />

treatment of SCLC. In NCI-H69 SCLC xenografts<br />

significant inhibition of tumor growth<br />

accrued from day 7 (p ! 0.05) through to the<br />

end of the experiment (p ! 0.01). The mean<br />

tumor weight was reduced significantly (p !<br />

0.01) by RC-160 compared to the control<br />

group. Tumor volume doubling time in mice<br />

receiving RC-160 was extended from 7.5 to<br />

12.7 days. Histologically, the number of mitotic<br />

<strong>and</strong> apoptotic cells in treated tumors did<br />

not differ significantly from control. However,<br />

the ratio of apoptotic to mitotic indices<br />

was significantly higher in the group receiving<br />

RC-160 (p ! 0.05; table 3) [26].<br />

<strong>Somatostatin</strong> analogs have also been evaluated<br />

in the treatment of NSCLC. Somatuline<br />

has been demonstrated to inhibit the in vivo<br />

growth of the SSTR-poor NSCLC cell line H-<br />

165 supporting the contention that indirect<br />

growth inhibitory effects may be important in<br />

the antitumor activity of somatostatin analogs<br />

in solid tumors [64]. Regarding this study it is<br />

important to keep in mind the results of subsequent<br />

work which, using the radiolabelled<br />

somatostatin analog [ 125 I]MK-678 as radiolig<strong>and</strong><br />

<strong>and</strong> RT-PCR techniques, demonstrated<br />

that H-165 is SSTR-positive [73].<br />

94 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


Table 3. Summary of RC-160 in vivo growth inhibition study data<br />

Tumor volume, mm3 Initial<br />

Final<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

SCLC NCI-H69<br />

control RC-160<br />

10.5B1.6<br />

249.7B182.3<br />

9.8B1.8<br />

66.0B26.5*<br />

NSCLC NCI-H157<br />

control RC-160<br />

10.0B2.0<br />

1,580.3B455.7<br />

11.1B2.6<br />

291.0B207.8*<br />

Body weight, g 26.3B2.3 26.5B1.0 26.5B1.0 24.2B4.7<br />

Tumor weight, g 0.27B0.19 0.058B0.04* 1.9B0.3 0.64B0.3*<br />

Tumor doubling time, days 7.512.7 3.88 6.06<br />

Mitotic <strong>and</strong> apoptotic indices<br />

Mitotic index 37.9B2.9 23.0B4.6 17.3B4.0 12.9B1.2<br />

Apoptotic index 35.4B2.6 42.3B3.2 2.87B0.54.00B0.7<br />

Ratio of apoptotic to mitotic indices 0.96B0.1 2.27B0.5* 0.19B0.04 0.35B0.08<br />

GH assay<br />

GH, ng/ml 3.8B0.52.0B0.5* 6.8B0.8 3.5B0.3*<br />

Receptors<br />

EGF Kd, nM 1.3 1.6 0.7 0.5<br />

EGF Bmax, fmol/mg protein 278 174 249 100<br />

IGF-1 Kd, nM 1.0 0.9 0.50.7<br />

IGF-1 Bmax, fmol/mg protein 294 176 257 129<br />

Bombesin/GRP Kd, nM 1.1 1.3 ND ND<br />

Bombesin/GRP Bmax, fmol/mg protein 420 435ND ND<br />

<strong>Somatostatin</strong> Kd, nM 3.55.5 ND ND<br />

<strong>Somatostatin</strong> Bmax, fmol/mg protein 450 570 ND ND<br />

There were 10 animals in each group. GRP = Gastrin-releasing peptide; ND = not detected. Tumor <strong>and</strong> body<br />

weight values are means B st<strong>and</strong>ard deviation of the mean. Mitotic <strong>and</strong> apoptotic indices <strong>and</strong> GH values are<br />

means B st<strong>and</strong>ard error of the mean. * p ! 0.05 [26].<br />

Nonetheless, our work supports the contention<br />

that indirect antiproliferative effects<br />

are important in the growth inhibition of<br />

lung tumors by somatostatin analogs. RC-<br />

160, 100 Ìg s.c. once daily, significantly inhibited<br />

the growth of NCI-H157 tumors from<br />

day 14 of treatment (p ! 0.01). The final<br />

tumor volume <strong>and</strong> tumor weight were significantly<br />

reduced in animals receiving RC-160<br />

compared with controls (table 3). Tumor volume<br />

doubling time was prolonged from 3.88<br />

to 6.06 days. No significant difference in the<br />

extent of necrosis, mitoses or apoptosis or in<br />

the ratio of apoptotic to mitotic indices between<br />

control <strong>and</strong> the RC-160-treated group<br />

was seen. No specific SSTR-binding sites<br />

were found on membranes prepared from the<br />

NSCLC tumor xenografts [26]. Similar in<br />

vivo growth inhibition has been observed in<br />

other SSTR-negative tumors [27].<br />

In our studies GH levels in animals treated<br />

with RC-160 were significantly decreased<br />

[26], compared with controls (table 3), a finding<br />

confirmed in subsequent work [27]. GH<br />

induces the synthesis of IGF-1, an important<br />

growth factor <strong>for</strong> solid tumors including lung<br />

Chemotherapy 2001;47(suppl 2):78–108 95


cancer [25]. In breast cancer patients RC-160<br />

has been demonstrated to significantly reduce<br />

serum IGF-1 levels [28]. This may result in a<br />

reduced proliferative stimulus <strong>for</strong> tumors sensitive<br />

to the growth stimulatory effects of<br />

IGF-1 including lung cancers [5, 14, 28].<br />

RC-160 also reduces elevated prolactin<br />

levels to within the normal range in patients<br />

with breast cancer. Prolactin is an autocrine<br />

growth factor <strong>for</strong> breast cancer. <strong>Somatostatin</strong><br />

analogs have little or no effect on primary<br />

pituitary hyperprolactinemia. These findings<br />

suggest that the normalization of serum prolactin<br />

levels in breast cancer patients is due to<br />

a direct effect on hormone synthesis <strong>and</strong> release<br />

by the breast cancer cells themselves<br />

[28]. <strong>Somatostatin</strong> analogs inhibit the release<br />

of bombesin-like peptides from SCLC tumor<br />

cells [122] <strong>and</strong> gastrin from cultured human<br />

endocrine cells [126]. The results suggest that<br />

somatostatin analogs may downregulate autocrine<br />

feedback loops in SSTR-positive malignancies<br />

including lung tumors, in particular<br />

SCLC [5, 26].<br />

Impact of <strong>Somatostatin</strong> <strong>Analogs</strong> on SCLC<br />

<strong>and</strong> NSCLC SSTR <strong>and</strong> Growth Factor<br />

Receptor Expression<br />

In accord with previous studies [5, 14, 61,<br />

69, 127], <strong>and</strong> those mentioned above, we<br />

demonstrated specific binding sites <strong>for</strong> radiolabelled<br />

somatostatin, bombesin/GRP, IGF-1<br />

<strong>and</strong> epidermal growth factor (EGF) on membrane<br />

preparations from treated <strong>and</strong> control<br />

xenografts of the SCLC cell line NCI-H69<br />

[26]. Likewise, EGF [128] <strong>and</strong> IGF-1 [129]<br />

binding sites were found on the NSCLC xenografts,<br />

but no specific binding sites <strong>for</strong> radiolabelled<br />

somatostatin <strong>and</strong> bombesin/GRP receptors<br />

were detected [26].<br />

The effects of RC-160 on SSTR, EGFR,<br />

luteinizing hormone-releasing hormone-R<br />

<strong>and</strong> IGF-1R expression have been evaluated<br />

in a number of tumors. RC-160 appears to<br />

upregulate SSTR [130, 131] while downregulating<br />

EGFR [130, 132–136], LHRH-R [130,<br />

137] <strong>and</strong> IGF-1R [131, 137] expression. Although<br />

only per<strong>for</strong>med on one occasion, the<br />

results of the membrane binding assays from<br />

our study are consistent with the previous<br />

findings. The EGF <strong>and</strong> IGF-1 Bmax were<br />

both reduced in RC-160-treated SCLC <strong>and</strong><br />

NSCLC xenografts as compared to controls<br />

(table 3). EGFR levels in the RC-160-treated<br />

samples were 63 <strong>and</strong> 40%, <strong>and</strong> IGF-1R levels<br />

60 <strong>and</strong> 50% of control in SCLC <strong>and</strong><br />

NSCLC membranes, respectively. No change<br />

was observed in the Kd <strong>and</strong> Bmax values <strong>for</strong><br />

bombesin/GRP in the SCLC tumor xenografts<br />

[26].<br />

Both the Kd <strong>and</strong> Bmax <strong>for</strong> somatostatin<br />

increased with RC-160 treatment indicating<br />

reduced binding affinity but increased binding<br />

capacity. The Kd <strong>and</strong> Bmax values <strong>for</strong><br />

[ 125 I-Tyr 11 ]somatostatin-14 were higher in<br />

membranes prepared from NCI-H69 xenografts<br />

than the Kd <strong>and</strong> Bmax <strong>for</strong> [ 125 I]RC-160<br />

high-affinity binding site in membranes prepared<br />

from NCI-H69 cell pellets grown in<br />

vitro [26, 69]. No low-affinity binding sites<br />

were detected [26].<br />

Taken together these findings indicate that<br />

RC-160 therapy may affect the expression of a<br />

number of growth-stimulatory <strong>and</strong> potentially<br />

growth-inhibitory receptors as well as growth<br />

factors in lung cancer. These changes may<br />

result in a downregulation of autocrine, paracrine<br />

<strong>and</strong> endocrine proliferative stimuli in<br />

SCLC <strong>and</strong> NSCLC.<br />

Angiogenesis<br />

Angiogenesis is necessary <strong>for</strong> the growth<br />

of a primary tumor beyond 1–2 mm in diameter<br />

<strong>and</strong> plays an important role in tumor<br />

metastasis [29]. Recent work has demonstrated<br />

that in patients with resectable<br />

NSCLC, one of the most important prognostic<br />

factors is the degree of angiogenesis in the<br />

96 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


esected tumors as assessed by microvessel<br />

counts. Tumors with high microvessel<br />

counts are associated with spread of the tumor<br />

to locoregional lymph nodes <strong>and</strong> a significantly<br />

worse prognosis [138]. The potent<br />

antiangiogenic activity of somatostatin analogs<br />

[12] may be one of the principal mechanisms<br />

by which these agents inhibit lung cancer<br />

growth in vivo. This may be particularly<br />

so in tumors with negative or low levels of<br />

specific binding sites <strong>for</strong> radiolabelled somatostatin<br />

analogs as is the case <strong>for</strong> the<br />

NSCLC cell line NCI-H157.<br />

Patient Studies<br />

<strong>Treatment</strong> of SCLC<br />

A number of small studies have now been<br />

conducted in patients with SCLC. The first<br />

involved 20 patients including 6 newly diagnosed<br />

patients <strong>and</strong> 14 presenting with relapsed<br />

disease following first line treatment.<br />

Octreotide 250 Ìg s.c. t.i.d. resulted in a significant<br />

reduction of serum IGF-1 levels. The<br />

overall change in IGF-1 levels, with the lowest<br />

level while on treatment expressed as a percentage<br />

of the pretreatment baseline value,<br />

was median 53% <strong>and</strong> mean 62 B 7% (range<br />

23–150%). Despite the reduction in IGF-1<br />

levels, there was no evidence of antitumor<br />

activity as measured by tumor bulk or NSE<br />

levels. However, 2 of the 14 patients with<br />

relapsed SCLC had stabilization of their disease<br />

accompanied by improvements in respiratory<br />

symptoms <strong>for</strong> 8 <strong>and</strong> 15 weeks [68]. In a<br />

subsequent study, in which 13 patients with<br />

SCLC were treated with octreotide 200 Ìg s.c.<br />

t.i.d. <strong>for</strong> 1 week, a significant decrease in<br />

mean serum NSE levels from baseline values<br />

of 44.4 B 57.7 to 32.6 B 42 ng/ml was seen<br />

[139].<br />

In a phase I dose-escalating study of octreotide<br />

<strong>and</strong> somatuline in the management<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

of neuroendocrine tumors, 2 patients with<br />

SCLC were recruited to the somatuline arm.<br />

An objective tumor response was seen in 1 of<br />

these patients [140]. This was followed by a<br />

phase I dose escalation study in 18 patients<br />

with relapsing or resistant SCLC following<br />

treatment with st<strong>and</strong>ard cytotoxic chemotherapy.<br />

Patients received between 2 <strong>and</strong> 10.5 mg/<br />

day of somatuline by continuous subcutaneous<br />

infusion. The agent was well tolerated<br />

with grade 1 or 2 diarrhea occurring in 8<br />

patients <strong>and</strong> pain at the site of injection<br />

requiring a change in infusion site in 3 patients.<br />

A mean reduction of 17% in IGF-1 levels<br />

was seen. The relationship between the<br />

dose of somatuline administered <strong>and</strong> the reduction<br />

in IGF-1 levels was statistically significant,<br />

higher doses being more potent in this<br />

regard (p ! 0.01). By day 28 disease progression<br />

was seen in all 15 patients evaluable <strong>for</strong><br />

tumor response <strong>and</strong> there<strong>for</strong>e treatment was<br />

discontinued [141].<br />

Finally, the efficacy of octreotide has been<br />

studied in a number of SCLC patients with<br />

syndromes related to ectopic hormone secretion.<br />

Octreotide 100 Ìg s.c. was evaluated in<br />

2 patients with SCLC tumors <strong>and</strong> Cushing’s<br />

syndrome secondary to ectopic ACTH release.<br />

In both cases a paradoxical increase in<br />

plasma ACTH <strong>and</strong> cortisol levels was seen.<br />

One patient went on to have treatment with<br />

a slow release preparation of somatuline.<br />

Following intramuscular depot injection of<br />

30 mg of somatuline a rise in ACTH <strong>and</strong> cortisol<br />

levels similar to that seen with the octreotide<br />

challenge was found [142]. The<br />

cause <strong>for</strong> this remains unknown but raises<br />

the possibility that, in certain cases, somatostatin<br />

analogs may have a stimulatory rather<br />

than inhibitory effect on the tumor as suspected<br />

in one study in patients with prostate<br />

cancer [143].<br />

Finally, as is the case <strong>for</strong> patients with neuroendocrine<br />

GEP <strong>and</strong> medullary thyroid tu-<br />

Chemotherapy 2001;47(suppl 2):78–108 97


mors, somatostatin analogs may have a role to<br />

play in the management of SCLC-related paraneoplastic<br />

diarrhea [144].<br />

Management of Bronchial Carcinoid<br />

As outlined earlier, [ 111 In]pentetreotide<br />

imaging has an important role in the detection<br />

<strong>and</strong> staging of patients with bronchial carcinoid<br />

tumors including those presenting with<br />

ectopic hormone syndromes including Cushing’s<br />

syndrome <strong>and</strong> acromegaly [103–112]. In<br />

1988, 2 patients with bronchial carcinoidassociated<br />

Cushing’s syndrome were treated<br />

with octreotide to see if this could induce a<br />

reduction in circulating ACTH levels <strong>and</strong> an<br />

improvement in symptoms. Octreotide 50 Ìg<br />

s.c. stat produced a 50% reduction of ACTH<br />

in 1 patient whilst the other was maintained<br />

in a clinical <strong>and</strong> biochemical remission <strong>for</strong> 10<br />

weeks with octreotide 100 Ìg s.c. t.i.d. [145].<br />

Subsequent reports indicate that, in the majority<br />

of cases, octreotide is an effective agent<br />

in the management of bronchial carcinoidrelated<br />

ectopic hormone syndromes providing<br />

palliation not only in Cushing’s syndrome<br />

but also <strong>for</strong> ectopic GHRH-related acromegaly<br />

[111, 113, 146, 147].<br />

A recent case study reported the evaluation<br />

of octreotide in the treatment of a patient with<br />

multiple cerebral metastases from a bronchial<br />

carcinoid. Neurologically, the principal symptom<br />

was expressive dysphasia. <strong>Treatment</strong><br />

with octreotide, initially in combination with<br />

a corticosteroid, resulted in disease stabilization<br />

<strong>for</strong> 6 months. During that time the patient’s<br />

symptoms resolved [148].<br />

However, ACTH does not fall in response<br />

to octreotide in all cases [149, 150]. There<strong>for</strong>e,<br />

an octreotide challenge may have a role<br />

to play in determining whether or not somatostatin<br />

analogs would be of value in controlling<br />

ectopic hormone syndromes in lung<br />

cancer [99, 151].<br />

There are now a number of case reports<br />

where ACTH-secreting bronchial carcinoid<br />

tumors have been localized leading to subsequent<br />

curative resection of the disease [105–<br />

107]. Indeed radioguided surgery employing a<br />

peroperative probe may facilitate complete<br />

tumor excision whilst, at the same time, reducing<br />

the extent of resection by clearly separating<br />

involved from uninvolved tissues<br />

[113].<br />

<strong>Somatostatin</strong> analogs are well tolerated,<br />

transient diarrhea, steatorrhea, colicky abdominal<br />

pain <strong>and</strong> borborygmi being the most<br />

commonly observed problems. Other side effects<br />

of somatostatin analog therapy include<br />

pain at the injection site, glucose intolerance<br />

<strong>and</strong> gallstone <strong>for</strong>mation. Although very rare,<br />

severe complications may occur including allergic<br />

reactions varying from skin rashes to<br />

anaphylaxis, acute pancreatitis <strong>and</strong> hepatitis.<br />

<strong>Somatostatin</strong> may also have a negative inotropic<br />

effect on the heart [28, 152, 153].<br />

Future Directions<br />

Enhancing the Efficacy of Cytotoxic<br />

Agents<br />

Experimental evidence clearly demonstrates<br />

that somatostatin analogs may potentiate<br />

the cytotoxicity of a range of chemotherapeutic<br />

agents. Studies in vitro <strong>and</strong> in vivo in<br />

animal models <strong>and</strong> in patients have shown<br />

that somatostatin analogs may enhance the<br />

antitumor efficacy of 5-fluorouracil while reducing<br />

the incidence of known side effects<br />

such as neutropenia <strong>and</strong> diarrhea [154–159].<br />

Using 19 F nuclear magnetic resonance spectroscopy,<br />

octreotide has been shown to increase<br />

the <strong>for</strong>mation of fluorouridinephosphates<br />

from 5-fluorouracil in human colon<br />

HT-29 adenocarcinoma cells. Furthermore,<br />

while 5-fluorouracil arrests cells in<br />

the S phase, cotreatment with octreotide al-<br />

98 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


most eliminates the S phase cells <strong>and</strong> induces<br />

the appearance of DNA fragments [160].<br />

Perhaps of greater interest is the finding<br />

that the somatostatin analog octreotide has<br />

been shown to synergistically enhance the antitumor<br />

effects of mitomycin C, doxorubicin<br />

<strong>and</strong> paclitaxel in AR42J pancreatic cancer<br />

cells in vitro suggesting that somatostatin may<br />

have a role to play in overcoming multidrug<br />

resistance. Combination therapy with doxorubicin<br />

<strong>and</strong> octreotide has been studied in<br />

vivo <strong>for</strong> time dependency <strong>and</strong> efficacy. Pretreatment<br />

with octreotide <strong>for</strong> 24 h prior to<br />

addition of doxorubicin results in additive<br />

antitumor activity while pretreatment with<br />

doxorubicin is associated with clear synergy<br />

[159]. With specific reference to lung cancer,<br />

somatuline has been shown to enhance the<br />

cytotoxic activity of cyclophosphamide in<br />

SCLC both in vitro <strong>and</strong> in vivo [161]. These<br />

results raise the possibility that somatostatin<br />

analogs may enhance the antitumor activity<br />

of Adriamycin, mitomycin C, cyclophosphamide<br />

<strong>and</strong> paclitaxel-containing regimens in<br />

lung cancer [3, 4]. Evidence to support this<br />

observation comes from a phase II study in<br />

prostate cancer. A number of patients initially<br />

treated with octreotide received chemotherapy<br />

at disease progression. A high response rate<br />

above that normally expected <strong>for</strong> cytotoxic<br />

chemotherapy alone was seen [143].<br />

The synergy between cytotoxic agents <strong>and</strong><br />

somatostatin analogs in vivo may in part be<br />

due to the known antiangiogenic activity of<br />

somatostatin analogs. This suggestion is supported<br />

by a number of studies where the combination<br />

of known antiangiogenic drugs with<br />

cytotoxic chemotherapeutic agents has additive<br />

antitumor activity in solid tumors. For<br />

example, the antiangiogenic agent TNP-470,<br />

a derivative of fumagillin, <strong>and</strong> minocycline,<br />

an inhibitor of collagenase IV activity, potentiate<br />

the cytotoxic effects of cyclophosphamide<br />

in the treatment of Lewis lung carci-<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

noma <strong>and</strong> murine FSaIIC fibrosarcoma. The<br />

combination of both antiangiogenic agents<br />

with cyclophosphamide is additive [162].<br />

As discussed earlier, tumor angiogenesis is<br />

an important prognostic factor in surgically<br />

resected, stage I <strong>and</strong> II, NSCLC [138]. Platelet-derived<br />

endothelial cell growth factor<br />

overexpression (PD-ECGF) is observed in approximately<br />

20% of NSCLC tumors <strong>and</strong> correlates<br />

with increased tumor angiogenesis.<br />

PD-ECGF is thymidine phosphorylase, an<br />

important enzyme in the pathway leading to<br />

the conversion of 5-fluorouracil to its active<br />

metabolites including 5-fluoro-2)-deoxyuridine-5)-monophosphate<br />

[163]. Taken together,<br />

the growth inhibitory effects on NSCLC<br />

xenografts, the antiangiogenic activity <strong>and</strong> the<br />

modulation of 5-fluorouracil metabolism by<br />

somatostatin analogs suggest a role <strong>for</strong> the<br />

combination of 5-fluorouracil or other fluoropyrimidine<br />

analogs with RC-160 in the treatment<br />

of NSCLC.<br />

Early phase II studies suggest that the combination<br />

of octreotide with tamoxifen may be<br />

effective in the treatment of patients with<br />

pancreatic cancer [164]. Tamoxifen has been<br />

shown to enhance the antitumor activity of<br />

cisplatin [165]. Cisplatin is the principal cytotoxic<br />

agent in the management of NSCLC<br />

[166]. These data indicate that somatostatin<br />

analogs, with or without tamoxifen, should be<br />

evaluated in combination with cytotoxic regimens<br />

– platinum agents, the taxanes, ifosphamide,<br />

mitomycin C, vinca alkaloids, <strong>and</strong><br />

antimetabolites, such as the fluoropyrimidines<br />

<strong>and</strong> gemcitabine – in the treatment of<br />

NSCLC [4, 166, 167].<br />

A number of cytotoxic agents have been<br />

coupled to somatostatin analogs with encouraging<br />

results both in vitro <strong>and</strong> in vivo. Methotrexate<br />

has been linked, through the 7-carboxyl<br />

group of its glutamic acid moiety, to the free<br />

phenylalanine amino acid of the octapeptide<br />

somatostatin analog, RC-121. This analog,<br />

Chemotherapy 2001;47(suppl 2):78–108 99


termed AN-51, has antitumor activity in vivo<br />

against the SSTR-expressing human pancreatic<br />

cancer cell line Mia PaCa-2 [168]. Cytotoxic<br />

analogs of somatostatin-containing<br />

potent anthracyclines have recently been developed.<br />

The superactive doxorubicin derivative,<br />

2-pyrrolinodoxorubicin, has been linked<br />

to RC-121 <strong>and</strong> RC-160 yielding AN-238 <strong>and</strong><br />

AN-258, respectively [169, 170]. These agents<br />

have shown promising antitumor activity in<br />

vitro <strong>and</strong> in vivo in a number of SSTR-positive<br />

solid tumors including breast, prostate,<br />

pancreatic, gastric <strong>and</strong> lung cancer. The possibility<br />

of specifically targeting such agents to<br />

SSTR-rich lung tumors in patients holds<br />

promise <strong>for</strong> the future.<br />

Adjuvant Therapy<br />

A recent overview meta-analysis indicates<br />

that adjuvant chemotherapy, established in<br />

the management of resected breast <strong>and</strong><br />

Dukes’ C colorectal cancer, may have a role to<br />

play in NSCLC. Recent in vivo work indicates<br />

that wound healing following surgical<br />

wounding of normal tissues may stimulate the<br />

growth of malignant disease even if the primary<br />

has not been removed. This suggests that<br />

wound healing itself results in the induction<br />

<strong>and</strong> release of trophic factors that have systemic<br />

proliferative effects [171]. Application<br />

of a somatostatin analog to the wound within<br />

1 h of surgery significantly reduces the induction<br />

of tumor growth seen otherwise [172]. As<br />

such somatostatin analogs may have a role to<br />

play in the adjuvant treatment of surgically<br />

resected lung tumors in the immediate postoperative<br />

period.<br />

Chemoprevention<br />

As discussed earlier, IGF-1 is a potent trophic<br />

<strong>and</strong> survival factor <strong>and</strong> plays an important<br />

role in cell trans<strong>for</strong>mation <strong>for</strong> many normal<br />

epithelial cells. IGF-1 acts through the<br />

IGF-1R. The central role of IGF-1R in the<br />

trans<strong>for</strong>mation of many cell types is best illustrated<br />

by the effects of disruption of the IGF-<br />

1R signal transduction pathway. This reverses<br />

the trans<strong>for</strong>med phenotype <strong>and</strong>/or inhibits tumorigenesis<br />

<strong>and</strong>/or induces loss of metastatic<br />

potential in human lung, breast, ovarian <strong>and</strong><br />

melanoma tumor models amongst others<br />

[173]. Support <strong>for</strong> a central role <strong>for</strong> IGF-1 in<br />

the development of malignant disease comes<br />

from the observation that acromegalic patients<br />

are significantly more likely to develop<br />

malignant disease than the general population<br />

[174]. Furthermore, in two recently published<br />

prospective studies, high normal IGF-1 levels<br />

were associated with an increased relative risk<br />

<strong>for</strong> the development of prostate cancer in men<br />

(4.3) <strong>and</strong> breast cancer in premenopausal<br />

women (7.28 in premenopausal women aged<br />

!50 years, when adjusted <strong>for</strong> IGF-binding<br />

protein 3 levels) [175, 176].<br />

The effects of somatostatin analogs on<br />

carcinogenesis have been evaluated in vivo.<br />

Fisher 344 female rats were initially exposed<br />

<strong>for</strong> 4 weeks to the initiator carcinogen N-butyl-N-(4-hydroxybutyl)nitrosamine<br />

in drinking<br />

water. They were then exposed to the promotor<br />

carcinogen mitomycin C intravesically<br />

<strong>for</strong> 12 weeks with or without concomitant<br />

subcutaneous somatostatin analog therapy. In<br />

the untreated group 34 rats developed transitional<br />

cell carcinoma. However, in the 20 rats<br />

treated with the octapeptide somatostatin<br />

analog RC-160 only 1 in situ carcinoma was<br />

observed [177]. Subsequent studies demonstrated<br />

that RC-160, infused at 2 Ìg/day<br />

<strong>for</strong> 14 days, significantly inhibited the progression<br />

of 0.5% 9,10-dimethyl-1,2-benzanthracene<br />

(DMBA)-initiated premalignant<br />

lesions in the buccal pouch of Syrian golden<br />

hamsters, as measured by Photofrin-induced<br />

fluorescence using in vivo photometry <strong>and</strong><br />

histological evaluation of the lesions [178].<br />

Groups of animals also had 0.5-cm incisions<br />

made in one cheek over the carcinogen-ini-<br />

100 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


tiated area by CO2 laser, a cancer promotor.<br />

The development of squamous cell carcinoma<br />

in these animals was likewise inhibited by<br />

RC-160 therapy.<br />

Further work analyzing kinase activity in<br />

DMBA-induced premalignant <strong>and</strong> malignant<br />

lesions demonstrated that phosphorylation<br />

increases continuously in a linear fashion<br />

from the first application of DMBA. This is<br />

independent of stimulation by growth factors<br />

such as EGF. RC-160 reduces phosphorylation<br />

in vitro at weeks 6–10 after DMBA application<br />

to the premalignant lesions <strong>and</strong> week<br />

12 after DMBA application to the malignant<br />

tissues. The results suggest that one of the<br />

principal pathways of the inhibitory action of<br />

RC-160 on carcinogenesis is through activation<br />

of phosphotyrosine phosphatase activity<br />

<strong>and</strong> that this in turn is secondary to enhanced<br />

kinase activity [179].<br />

These findings suggest that somatostatin<br />

analogs through the sustained suppression of<br />

IGF-1 levels [28, 180, 181] <strong>and</strong> through the<br />

inhibition of cell signal transduction pathways<br />

[10–18, 19–23] may have an important<br />

role to play in the chemoprevention of solid<br />

tumors including lung cancer.<br />

Conclusion<br />

Lung cancer is a plague of the late 20th century<br />

<strong>and</strong> will remain one of the prinicipal<br />

health issues well into the next century even if<br />

cigarette smoking stopped tomorrow. Despite<br />

progress in the diagnosis <strong>and</strong> treatment of<br />

lung cancer, with improvements in surgical<br />

techniques <strong>and</strong> the development of effective<br />

radiotherapeutic <strong>and</strong> chemotherapeutic regimens,<br />

the overall prognosis is appalling <strong>and</strong><br />

novel approaches to treatment are urgently<br />

required if a significant impact is to be made<br />

on overall survival.<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

The precise role <strong>for</strong> somatostatin analogs<br />

in the evaluation <strong>and</strong> treatment of lung cancer<br />

remains to be defined. SSTR may be expressed<br />

by lung tumors, particularly SCLC<br />

<strong>and</strong> bronchial carcinoid disease. Scintigraphic<br />

imaging with [ 111 In]pentetreotide<br />

may play a role in the clinical evaluation of<br />

neoplastic lung disease both be<strong>for</strong>e <strong>and</strong> following<br />

treatment, <strong>and</strong> in detecting relapsed<br />

disease.<br />

The potential role of somatostatin analogs<br />

linked to radiotherapeutic isotopes in the<br />

management of SCLC is currently being explored.<br />

Cytotoxic somatostatin analogs containing<br />

2-pyrrolinodoxorubicin have shown<br />

encouraging antitumor activity in a range of<br />

solid tumors including breast <strong>and</strong> prostate<br />

cancer. These analogs, including AN-238 <strong>and</strong><br />

AN-258, hold promise as future therapeutic<br />

agents in the treatment of lung tumors.<br />

<strong>Somatostatin</strong> analog therapy inhibits the<br />

growth of SSTR-positive <strong>and</strong> SSTR-negative<br />

lung tumors in vivo. Furthermore, experimental<br />

evidence suggests that somatostatin<br />

analogs may enhance the efficacy of a range of<br />

chemotherapeutic agents in the treatment of<br />

solid tumors, including cisplatin, the anthracyclines<br />

<strong>and</strong> the taxanes. Based on the encouraging<br />

preclinical data we have set up a phase<br />

I/II study evaluating increasing doses of octreotide<br />

in combination with st<strong>and</strong>ard chemotherapy<br />

in the treatment of SCLC. Similar<br />

studies are indicated <strong>for</strong> NSCLC.<br />

Chemotherapy 2001;47(suppl 2):78–108 101


References<br />

1 Boyle P: <strong>Cancer</strong>, cigarette smoking<br />

<strong>and</strong> premature death in Europe: A<br />

review including the recommendations<br />

of European <strong>Cancer</strong> Experts<br />

Consensus Meeting, Helsinki, October<br />

1996. Lung <strong>Cancer</strong> 1997;17:1–<br />

60.<br />

2 Ries LAG, Hankey BF, Miller BA,<br />

Hartman AM, Edwards BK: <strong>Cancer</strong><br />

Statistics Review 1973–1988. Bethesda,<br />

National <strong>Cancer</strong> Institute<br />

1991, NIH publication No 91-2789.<br />

3 Ihde DC, Pass HI, Glatstein EJ:<br />

<strong>Cancer</strong> of the lung, section 3: Small<br />

cell lung cancer; in DeVita VT, Hellman<br />

S, Rosenberg SA (eds): <strong>Cancer</strong>,<br />

Principles <strong>and</strong> Practice of Oncology.<br />

Philadelphia, Lippincott, 1997, pp<br />

911–949.<br />

4 Ginsberg RJ, Vokes EE, Raben A:<br />

<strong>Cancer</strong> of the lung, section 2: Nonsmall<br />

cell lung cancer; in DeVita<br />

VT, Hellman S, Rosenberg SA (eds):<br />

<strong>Cancer</strong>; Principles <strong>and</strong> Practice of<br />

Oncology. Philadelphia, Lippincott,<br />

1997, pp 858–910.<br />

5Macaulay VM, Carney DN: Neuropeptide<br />

growth factors. <strong>Cancer</strong> Invest<br />

1991;9:659–673.<br />

6 O’Byrne KJ, Han C, Mitchell K,<br />

Lane D, Carmichael J, Harris AL,<br />

Talbot DC: Phase II study of liarozole<br />

in advanced non-small cell lung<br />

cancer. Eur J <strong>Cancer</strong> 1998;34:1463–<br />

1466.<br />

7 Salsbury AJ, Burrage K, Hellman K:<br />

Inhibition of metastatic spread by<br />

I.C.R.F. 159: Selective deletion of a<br />

malignant characteristic. Br Med J<br />

1970;4:344–346.<br />

8 Reichlin S: <strong>Somatostatin</strong>. N Engl J<br />

Med 1983;309:1495–1501.<br />

9 Reichlin S: <strong>Somatostatin</strong>. N Engl J<br />

Med 1983;309:1556–1563.<br />

10 Buscail L, Delesque N, Esteve J-P,<br />

Saint-Laurent N, Prats H, Clerc P,<br />

Robberecht P, Bell GI, Liebow C,<br />

Schally AV, Vaysse N, Susini C:<br />

Stimulation of tyrosine phosphatase<br />

<strong>and</strong> inhibition of cell proliferation<br />

by somatostatin analogues: Mediation<br />

by human somatostatin receptor<br />

subtypes sst1 <strong>and</strong> sst2. Proc Natl<br />

Acad Sci USA 1994;91:2315–2319.<br />

11 Buscail L, Esteve J-P, Saint-Laurent<br />

N, Bertr<strong>and</strong> V, Reisine T, O’Carroll<br />

A-M, Bell GI, Schally AV, Vaysse N,<br />

Susini C: Inhibition of cell proliferation<br />

by the somatostatin analogue<br />

RC-160 is mediated by somatostatin<br />

receptor subtypes sst2 <strong>and</strong> sst5<br />

through different mechanisms. Proc<br />

Natl Acad Sci USA 1995;92:1580–<br />

1584.<br />

12 Patel PC, Barrie R, Hill N, L<strong>and</strong>eck<br />

S, Kurozawa D, Woltering EA:<br />

Postreceptor signal transduction<br />

mechanisms involved in octreotide<br />

induced inhibition of angiogenesis.<br />

Surgery 1994;116:1148–1152.<br />

13 O’Byrne KJ, Carney DN: <strong>Somatostatin</strong><br />

<strong>and</strong> the lung. Lung <strong>Cancer</strong><br />

1993;10:151–172.<br />

14 Pollak M, Schally AV: Mechanisms<br />

of antineoplastic action of somatostatin<br />

analogs. Proc Soc Exp Biol<br />

Med 1998;217:143–152.<br />

15Bruns C, Weckbecker G, Raulf F,<br />

Kaupmann K, Schoesster P, Hoyer<br />

D, Lubbert H: Molecular pharmacology<br />

of somatostatin receptor subtypes.<br />

Ann NY Acad Sci 1994;733:<br />

138–146.<br />

16 Patel YC, Greenwood MT, Warszynska<br />

A, Panetta R, Srikant CB: All<br />

five cloned human somatostatin receptors<br />

(hsst1–5) are functionally<br />

coupled to adenylyl cyclase. Biochem<br />

Biophys Res Commun 1994;<br />

198:605–612.<br />

17 Qin Y, Eftl T, Groot K, Horvath J,<br />

Cai R-Z, Schally AV: <strong>Somatostatin</strong><br />

analog RC-160 inhibits growth of<br />

CFPAC-1 human pancreatic cancer<br />

cells in vitro <strong>and</strong> intracellular production<br />

of cyclic adenosine monophosphate.<br />

Int J <strong>Cancer</strong> 1995;60:<br />

694–700.<br />

18 Cordelier P, Esteve J-P, Bousquet C,<br />

Delesque N, O’Carroll A-M, Schally<br />

AV, Vaysse N, Susini C, Buscail L:<br />

Characterization of the anti-proliferative<br />

signal mediated by the somatostatin<br />

receptor subtype sst5.<br />

Proc Natl Acad Sci USA 1997;94:<br />

9343–9348.<br />

19 Egan SE, Weinberg RA: The pathway<br />

to signal achievement. Nature<br />

1993;365:781–783.<br />

20 Liebow C, Reilly C, Serrano M,<br />

Schally AV: <strong>Somatostatin</strong> analogs<br />

inhibit growth of pancreatic cancer<br />

by stimulating tyrosine phosphatase.<br />

Proc Natl Acad Sci USA 1988;<br />

85:1–22.<br />

21 Pan MG, Florio T, Stork PJS: G protein<br />

activation of a hormone-stimulated<br />

phosphatase in human tumor<br />

cells. Science 1992;256:1215–1217.<br />

22 Cattaneo MG, Amoroso D, Gussoni<br />

G, Sanguini AM, Vicentini LM: A<br />

somatostatin analogue inhibits<br />

MAP kinase activation <strong>and</strong> cell proliferation<br />

in human neuroblastoma<br />

<strong>and</strong> in human small cell lung carcinoma<br />

cell lines. FEBS Lett 1996;<br />

397:164–168.<br />

23 Ilondo MM, de Meyts P, Bouchelouche<br />

P: Human growth hormone<br />

increases cytosolic free calcium in<br />

cultured human IM-9 lymphocytes:<br />

A novel mechanism of growth hormone<br />

transmembrane signalling.<br />

Biochem Biophys Res Commun<br />

1994;202:391–397.<br />

24 McCarthy TL, Ji C, Casinghino S,<br />

Centrella M: Alternate signalling<br />

pathways selectively regulate binding<br />

of insulin-like growth factor I<br />

<strong>and</strong> II on fetal rat bone cells. J Biol<br />

Chem 1998;68:446–456.<br />

25Macaulay VM: Insulin-like growth<br />

factors <strong>and</strong> cancer. Br J <strong>Cancer</strong><br />

1992;65:311–320.<br />

26 Pinski J, Schally AV, Halmos G,<br />

Szepeshazi K, Groot K, O’Byrne KJ,<br />

Cai R-Z: Effects of somatostatin analogue<br />

RC-160 <strong>and</strong> bombesin/gastrin-releasing<br />

peptide antagonist on<br />

the growth of human small-cell <strong>and</strong><br />

non-small-cell lung carcinomas in<br />

nude mice. Br J <strong>Cancer</strong> 1994;70:<br />

886–892.<br />

27 Pinski J, Schally AV, Halmos G,<br />

Szepeshazi K, Groot K: <strong>Somatostatin</strong><br />

analog RC-160 inhibits the<br />

growth of human osteosarcomas in<br />

nude mice. Int J <strong>Cancer</strong> 1996;65:<br />

870–874.<br />

28 O’Byrne KJ, Dobbs N, Propper DJ,<br />

Braybrooke JP, Koukourakis MI,<br />

Mitchell K, Woodhull J, Talbot DC,<br />

Schally AV, Harris AL: Phase II<br />

study of RC-160 (vapreotide), an octapeptide<br />

analogue of somatostatin,<br />

in the treatment of metastatic breast<br />

cancer. Br J <strong>Cancer</strong>, in press.<br />

102 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


29 Folkman J: Clinical applications of<br />

research on angiogenesis. N Engl J<br />

Med 1995;333:1757–1763.<br />

30 Gastpar H, Zoltobrocki M, Weissgerber<br />

PW: The inhibition of cell<br />

stickiness by somatostatin. Res Exp<br />

Med (Berl) 1983;182:1–6.<br />

31 Carney DN: Biology of small-cell<br />

lung cancer. Lancet 1992;339:843–<br />

846.<br />

32 Bork E, Hansen M, Urdal P, Paus E,<br />

Holst JJ, Schifter S, Fenger M, Engbaek<br />

F: Early detection of response<br />

in small cell bronchogenic carcinoma<br />

by changes in serum concentrations<br />

of creatine kinase, neuron specific<br />

enolase, calcitonin, ACTH, serotonin<br />

<strong>and</strong> gastrin releasing peptide.<br />

Eur J <strong>Cancer</strong> Clin Oncol 1988;<br />

24:1033–1038.<br />

33 Splinter TAW, Carney DN, Teeling<br />

M, Peake MD, Kho GS, Oosterom<br />

R, Cooper EH: Neuron-specific enolase<br />

can be used as the sole guide to<br />

treat small-cell lung cancer patients<br />

in common clinical practice. J <strong>Cancer</strong><br />

Res Clin Oncol 1989;115:400–<br />

401.<br />

34 Penman E, Wass JAH, Besser GH,<br />

Rees LH: <strong>Somatostatin</strong> secretion by<br />

lung <strong>and</strong> thymic tumors. Clin Endocrinol<br />

(Oxf) 1980;13:613–620.<br />

35Roos BA, Lindall AW, Ells J, Elde<br />

R, Lambert PW, Birnbaum RS: Increased<br />

plasma <strong>and</strong> tumor somatostatin-like<br />

immunoreactivity in<br />

medullary thyroid carcinoma <strong>and</strong><br />

small cell lung cancer. J Clin Endocrinol<br />

Metab 1981;52:187–194.<br />

36 Noseda A, Peeters TL, Delhave M,<br />

Bormans V, Couvreur Y, V<strong>and</strong>ermoten<br />

G, De Francquen P, Rocmans<br />

P, Yernault JC: Increased<br />

plasma motilin concentrations in<br />

small cell carcinoma of the lung.<br />

Thorax 1987;42:784–789.<br />

37 Ghatei MA, Sheppard MN,<br />

O’Shaughnessy DJ, Adrian TE,<br />

McGregor GP, Polak JM, Bloom<br />

SR: Regulatory peptides in the<br />

mammalian respiratory tract. Endocrinology<br />

1982;111:1248–1254.<br />

38 Sorenson GD, Pettengill OS,<br />

Brinck-Johnsen T, Cate CC, Maurer<br />

LH: Hormone production by cultures<br />

of small-cell carcinoma of the<br />

lung. <strong>Cancer</strong> 1981;47:1289–1296.<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

39 Bepler G, Rotsch M, Jaques G,<br />

Haeder M, Heymanns J, Hartogh G,<br />

Kiefer P, Havemann K: Peptides<br />

<strong>and</strong> growth factors in small cell lung<br />

cancer: Production, binding sites,<br />

<strong>and</strong> growth effects. J <strong>Cancer</strong> Res<br />

Clin Oncol 1988;114:235–244.<br />

40 Chretien MF, Pouplard-Barthelaix<br />

A, Dubois MP, Simard C, Rebel A:<br />

<strong>Somatostatin</strong> <strong>and</strong> adrenocorticotropic<br />

hormone like immunoreactivity<br />

in small cell carcinoma of the<br />

lung. J Clin Pathol 1986;39:418–<br />

422.<br />

41 Kasurinen J, Syrjanen KJ: Peptide<br />

hormone immunoreactivity <strong>and</strong><br />

prognosis in small cell carcinoma of<br />

the lung. Respiration 1986;49:61–<br />

67.<br />

42 Wilson TS, McDowell EM, Marangos<br />

PJ, Trump BF: Histochemical<br />

studies of dense-core granulated tumors<br />

of the lung: Neuron-specific<br />

enolase as a marker <strong>for</strong> granulated<br />

cells. Arch Pathol Lab Med 1985;<br />

109:613–620.<br />

43 Reubi J-C, Waser B, Lamberts SWJ,<br />

Mengod G: <strong>Somatostatin</strong> (SRIH)<br />

messenger ribonucleic acid expression<br />

in human neuroendocrine <strong>and</strong><br />

brain tumors using in situ hybridization<br />

histochemistry: Comparison<br />

with SRIH receptor content. J Clin<br />

Endocrinol Metab 1993;76:642–<br />

647.<br />

44 Balabolkin MI, Abrikosova SLU:<br />

Hormonal function of isl<strong>and</strong>s of<br />

Langerhans in diabetes mellitus.<br />

Probl Endokrinol 1984;30:16–18.<br />

45Miyazaki K, Funakoshi A, Ibayashi<br />

H: Plasma somatostatin-like immunoreactivity<br />

responses to a mixed<br />

meal <strong>and</strong> the heterogeneity in<br />

healthy <strong>and</strong> non-insulin dependent<br />

(NIDDM) diabetics. Endocrinol Jpn<br />

1986;33:51–59.<br />

46 Tjen A, Looi S, Ekman R, Lippton<br />

H, Cary J, Keith I: CGRP <strong>and</strong> somatostatin<br />

modulate chronic hypoxic<br />

pulmonary hypertension. Am J<br />

Physiol 1992;263:H681–H690.<br />

47 Schifter S, Johannsen L, Bunker C,<br />

Brickell P, Bork E, Lindeberg H, Faber<br />

J: Calcitonin gene-related peptide<br />

in small cell lung carcinoma.<br />

Clin Endocrinol (Oxf) 1993;39:59–<br />

65.<br />

48 Graziano SL, Mazid R, Newman N,<br />

Tatum A, Oler A, Mortimer JA,<br />

Gullo JJ, DiFino SM, Scalzo AJ:<br />

The use of neuroendocrine immunoperoxidase<br />

markers to predict<br />

chemotherapy response in patients<br />

with non-small-cell lung cancer. J<br />

Clin Oncol 1989;7:1398–1406.<br />

49 Symes AJ, Craig RK, Brickell PM:<br />

Loss of transcriptional repression<br />

contributes to the ectopic expression<br />

of the calcitonin/alpha-CGRP gene<br />

in a human lung carcinoma cell line.<br />

FEBS Lett 1992;306:229–233.<br />

50 Kiriakogiani-Psaropoulou P, Malamou-Mitsi<br />

V, Martinipoulou U, Legaki<br />

S, Tamvakis N, Vrettou E,<br />

Fountzilas G, Skarlos D, Kosmidis<br />

P, Pavlidis N: The value of neuroendocrine<br />

markers in non-small cell<br />

lung cancer: A comparative immunohistopathologic<br />

study. Lung <strong>Cancer</strong><br />

1994;11:353–364.<br />

51 Rosell R, Carles J, Ariza A, Moreno<br />

I, Ribelles N, Solano V, Pellicer I,<br />

Barnadas A, Abad A: A phase II<br />

study of days 1 <strong>and</strong> 8 cisplatin <strong>and</strong><br />

recombinant alpha-2B interferon in<br />

advanced non-small cell lung cancer.<br />

<strong>Cancer</strong> 1991;67:2448–2453.<br />

52 Van Z<strong>and</strong>wijk N, Jassem E, Bonfrer<br />

JM, Mooi WJ, Tinteren H: Serum<br />

neuron-specific enolase <strong>and</strong> lactate<br />

dehydrogenase as predictors of response<br />

to chemotherapy <strong>and</strong> survival<br />

in non-small cell lung cancer.<br />

Semin Oncol 1992;19(suppl 2):37–<br />

43.<br />

53 Wood SM, Wood JR, Ghatei MA,<br />

Lee YC, O’Shaughnessy D, Bloom<br />

SR: Bombesin, somatostatin <strong>and</strong><br />

neurotensin-like immunoreactivity<br />

in bronchial carcinoma. J Clin Endocrinol<br />

Metab 1981;53:1310–<br />

1312.<br />

54 Kameya T, Shimosato Y, Kodama<br />

T, Tsumuraya M, Koide T, Yamaguchi<br />

K, Abe K: Peptide hormone<br />

production by adenocarcinomas of<br />

the lung; its morphologic basis <strong>and</strong><br />

histogenetic considerations. Virchows<br />

Arch 1983;400:245–257.<br />

55 Mardini G, Pai U, Chavez AM, Tomashefski<br />

JF Jr: Endobronchial adenocarcinoma<br />

with endometrioid<br />

features <strong>and</strong> prominent neuroendocrine<br />

differentiation. A variant of fetal<br />

adenocarcinoma. <strong>Cancer</strong> 1994;<br />

73:1383–1389.<br />

Chemotherapy 2001;47(suppl 2):78–108 103


56 Lee KG, Cho NH: Fine needle aspiration<br />

cytology of pulmonary adenocarcinoma<br />

of fetal type: Report of a<br />

case with immunohistochemical<br />

<strong>and</strong> ultrastructural studies. Diagn<br />

Cytopathol 1991;7:408–414.<br />

57 Keyhani-Rofagha S, O’Dorisio TM,<br />

Lucas JG, Fontana MB, Tuttle SE:<br />

Extra-adrenal paraganglioma <strong>and</strong><br />

pulmonary chondroma: A case report<br />

<strong>and</strong> review of the literature. J<br />

Surg Oncol 1987;35:89–95.<br />

58 Chejvec G, Cosnow I, Gould NS,<br />

Husain AN, Gould VE: Pulmonary<br />

blastoma with neuroendocrine differentiation<br />

in cell morules resembling<br />

neuroepithelial bodies. Histopathology<br />

1990;17:353–358.<br />

59 Al-Saffar N, White A, Moore M,<br />

Hasleton PS: Immunoreactivity of<br />

various peptides in typical <strong>and</strong> atypical<br />

bronchopulmonary carcinoid<br />

tumours. Br J <strong>Cancer</strong> 1988;58:762–<br />

766.<br />

60 Gould VE, Lee I, Warren WH: Immunohistochemical<br />

evaluation of<br />

neuroendocrine cells <strong>and</strong> neoplasms<br />

of the lung. Pathol Res Pract 1988;<br />

183:200–213.<br />

61 O’Byrne KJ, Carney DN: Radiolabelled<br />

somatostatin analogue scintigraphy<br />

in oncology. Anticancer<br />

Drugs 1996;7(suppl 1):33–44.<br />

62 Van Hagen PM, Krenning EP,<br />

Kwekkeboom DJ, Reubi J-C, Anker<br />

Lugtenburg PJ, Lowenberg B, Lamberts<br />

SWJ: <strong>Somatostatin</strong> <strong>and</strong> the immune<br />

<strong>and</strong> haemapoietic system; a<br />

review. Eur J Clin Invest 1994;24:<br />

91–99.<br />

63 Taylor JE, Coy DH, Moreau J-P:<br />

High affinity binding of [ 125 I-<br />

Tyr 11 ]somatostatin-14 to human<br />

small cell lung carcinoma (NCI-<br />

H69). Life Sci 1988;43:421–427.<br />

64 Bogden AE, Taylor JE, Moreau J-P,<br />

Coy DH, LePage DJ: Response of<br />

human lung tumour xenografts to<br />

treatment with a somatostatin analogue<br />

(somatuline). <strong>Cancer</strong> Res<br />

1990;50:4360–4365.<br />

65Taylor JE: <strong>Somatostatin</strong> analogues<br />

<strong>and</strong> small-cell lung carcinomas. Recent<br />

Results <strong>Cancer</strong> Res 1993;129:<br />

71–82.<br />

66 Reubi J-C, Waser B, Sheppard M,<br />

Macaulay V: <strong>Somatostatin</strong> receptors<br />

are present in small-cell but not in<br />

non-small-cell primary lung carcinomas:<br />

Relationship to EGF-receptors.<br />

Int J <strong>Cancer</strong> 1990;45:269–274.<br />

67 Sagman U, Mullen JB, Kovacs K,<br />

Kerbel R, Ginsberg R, Reubi J-C:<br />

Identification of somatostatin receptors<br />

in human small cell lung carcinoma.<br />

<strong>Cancer</strong> 1990;66:2129–<br />

2133.<br />

68 Macaulay VM, Smith IE, Everard<br />

MJ, Teale JD, Reubi J-C, Millar JL:<br />

Experimental <strong>and</strong> clinical studies<br />

with somatostatin analogue octreotide<br />

in small cell lung cancer. Br J<br />

<strong>Cancer</strong> 1991;64:451–456.<br />

69 O’Byrne KJ, Halmos G, Pinski J,<br />

Szepeshazi K, Groot K, Schally AV,<br />

Carney DN: <strong>Somatostatin</strong> receptor<br />

expression in lung cancer. Eur J<br />

<strong>Cancer</strong> 1994;30:1682–1687.<br />

70 Eden PA, Taylor JE: <strong>Somatostatin</strong><br />

receptor subtype gene expression in<br />

human <strong>and</strong> rodent tumors. Life Sci<br />

1993;53:85–90.<br />

71 Prevost G, Bourgeois Y, Mormont<br />

C, Lerrant Y, Veber N, Poupon MF,<br />

Thomas F: Characterization of somatostatin<br />

receptors <strong>and</strong> growth inhibition<br />

by the somatostatin analogue<br />

BIM23014 in small cell lung<br />

carcinoma xenograft: SCLC-6. Life<br />

Sci 1994;55:155–162.<br />

72 Zhang CY, Yokogoshi Y, Yoshimoto<br />

K, Fujinaka Y, Maysumoto K,<br />

Saito S: Point mutation of the somatostatin<br />

receptor 2 gene in the<br />

human small cell lung cancer cell<br />

line COR-LlO3. Biochem Biophys<br />

Res Commun 1995;210:805–815.<br />

73 Taylor JE, Theveniau MA, Bashirzadeh<br />

R, Reisine T, Eden PA: Detection<br />

of somatostatin receptor<br />

subtype 2 (sst2) in established tumors<br />

<strong>and</strong> tumor cell lines: Evidence<br />

<strong>for</strong> sst2 heterogeneity. Peptides<br />

1994;15:1229–1236.<br />

74 Reubi J-C, Schaer JC, Waser B,<br />

Mengod G: Expression <strong>and</strong> localization<br />

of somatostatin receptor sst1,<br />

sst2, <strong>and</strong> sst3 messenger RNAs in<br />

primary human tumors using in situ<br />

hybridisation. <strong>Cancer</strong> Res 1994;54:<br />

3455–3459.<br />

75Fujita T, Yamaji Y, Sato M, Murao<br />

K, Takahara J: Gene expression of<br />

somatostatin receptor subtypes, sst1<br />

<strong>and</strong> sst2, in human lung cancer cell<br />

lines. Life Sci 1994;55:1797–1806.<br />

76 Reubi J-C, Kvols L, Krenning E,<br />

Lamberts SWJ: Distribution of somatostatin<br />

receptors in normal <strong>and</strong><br />

tumor tissue. Metabolism 1990;<br />

39(suppl 2):78–81.<br />

77 Krenning EP, Bakker WH, Breeman<br />

WAP, Koper JW, Kooij PPM, Ausema<br />

L, Lameris JS, Reubi J-C, Lamberts<br />

SWJ: Localisation of endocrine-related<br />

tumours with radioiodinated<br />

analogue of somatostatin.<br />

Lancet 1989;i:242–244.<br />

78 Nilsson O, Kolby L, Wangberg B,<br />

Wig<strong>and</strong>er A, Billig H, William-Olsson<br />

L, Fjalling M, Forssell-Aronsson<br />

E, Ahlman H: Comparative studies<br />

on the expression of somatostatin<br />

receptor subtypes, outcome of octreotide<br />

scintigraphy <strong>and</strong> response<br />

to octreotide treatment in patients<br />

with carcinoid tumours. Br J <strong>Cancer</strong><br />

1998;77:632–637.<br />

79 Mountain CF: Revisions in the international<br />

system <strong>for</strong> staging lung<br />

cancer. Chest 1997;111:1710–1717.<br />

80 Kwekkeboom DJ, Krenning EP,<br />

Bakker WH, Oei HY, Splinter<br />

TAW, Siang Kho G, Lamberts SJW:<br />

Radioiodinated somatostatin analog<br />

scintigraphy in small-cell lung cancer.<br />

J Nucl Med 1991;32:1845–<br />

1848.<br />

81 O’Byrne KJ, Ennis JT, Freyne PJ,<br />

Clancy LJ, Prichard JS, Carney DN:<br />

Scintigraphic imaging of small cell<br />

lung cancer patients with 111 In-pentetreotide,<br />

a radiolabelled somatostatin<br />

analogue. Br J <strong>Cancer</strong> 1994;<br />

69:762–766.<br />

82 O’Byrne KJ, Ennis JT, Freyne PJ,<br />

Clancy LJ, Prichard JS, Carney DN:<br />

Imaging of small cell tumours<br />

(SCLC) with the radiolabelled somatostatin<br />

analogue, [ 111 In]pentetreotide<br />

Proc Am Soc Clin Oncol<br />

1994;13:364.<br />

83 Reisinger I, Bohuslavitzki KH,<br />

Brenner W, Braune S, Dittrich I,<br />

Geide A, Kettner B, Otto HJ,<br />

Schmidt S, Munz DL: <strong>Somatostatin</strong><br />

receptor scintigraphy in small-cell<br />

lung cancer: Results of a multicenter<br />

study. J Nucl Med 1998;39:224–<br />

227.<br />

84 Reubi J-C, Horisberger U, Laissue J:<br />

High density of somatostatin receptors<br />

in veins surrounding human<br />

cancer tissue: Role in tumor-host interaction?<br />

Int J <strong>Cancer</strong> 1994;56:<br />

681–688.<br />

104 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


85Maini CL, Tofani A, Venturo I, Pigorini<br />

F, Sciuto R, Semprebene A,<br />

Boni S, Giunta S, Lopez M: <strong>Somatostatin</strong><br />

receptor imaging in<br />

small cell lung cancer using 111 In-<br />

DTPA-octreotide: A preliminary<br />

study. Nucl Med Commun 1993;14:<br />

962–968.<br />

86 Kwekkeboom DJ, Kho GS, Lamberts<br />

SWJ, Reubi J-C, Laissue JA,<br />

Krenning EP: The value of octreotide<br />

scintigraphy in patients with<br />

lung cancer. Eur J Nucl Med 1994;<br />

21:1106–1113.<br />

87 Kirsch CM, Von Pawel J, Grau I,<br />

Tatsch K: Indium-111 pentetreotide<br />

in the diagnostic work-up of patients<br />

with bronchogenic carcinoma. Eur J<br />

Nucl Med 1994;21:1318–1325.<br />

88 Bombardieri E, Crippa F, Cataldo I,<br />

Chiti A, Seregni E, Soresi E, Boffi R,<br />

Invernizzi G, Buraggi GL: <strong>Somatostatin</strong><br />

receptor imaging of small cell<br />

lung cancer (SCLC) by means of<br />

111 ln-DTPA-octreotide scintigraphy.<br />

Eur J <strong>Cancer</strong> 1995;31A:184–<br />

188.<br />

89 Bohuslavizki KH, Brenner W,<br />

Gunther M, Eberhardt JU, Jahn N,<br />

Tinnemeyer S, Wolf H, Sippel C,<br />

Clausen M, Gatzemeier U, Henze E:<br />

<strong>Somatostatin</strong> receptor scintigraphy<br />

in the staging of small cell lung cancer.<br />

Nucl Med Commun 1996;17:<br />

191–196.<br />

90 Stokkel MPM, Kwa BH, Pauwels<br />

EKJ: Imaging <strong>and</strong> staging of smallcell<br />

lung cancer: Is there a future role<br />

<strong>for</strong> octreotide scintigraphy? Br J<br />

Clin Pract 1995;49:235–238.<br />

91 Soresi E, Bombardieri E, Chiti A,<br />

Boffi R, Invernizzi G, Crippa F,<br />

Maffioli L: Indium-111-DTPA-octreotide<br />

scintigraphy modulation by<br />

treatment with unlabelled somatostatin<br />

analogue in small-cell lung<br />

cancer. Tumori 1995;81:125–127.<br />

92 Berenger N, Moretti JL, Boaziz C,<br />

Vigneron N, Morere JF, Breau JL:<br />

<strong>Somatostatin</strong> receptor imaging in<br />

small cell lung cancer. Eur J <strong>Cancer</strong><br />

1996;32A:1429–1431.<br />

93 Hochstenbag MM, Heidendal GA,<br />

Wouters EF, ten Velde GP: In-111<br />

octreotide imaging in staging of<br />

small cell lung cancer. Clin Nucl<br />

Med 1997;22:811–816.<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

94 Kwekkeboom DJ, Lamberts SW,<br />

Habberna JD, Krenning EP: Costeffectiveness<br />

analysis of somatostatin<br />

receptor scintigraphy. J Nucl<br />

Med 1996;37:886–892.<br />

95Hofl<strong>and</strong> LJ, van Koetsveld PM,<br />

Waaijers M, Zuyderwijk J, Breeman<br />

WA, Lamberts SW: Internalization<br />

of the radioiodinated somatostatin<br />

analog ( 125 I-Tyr 3 )octreotide<br />

by mouse <strong>and</strong> human pituitary<br />

tumor cells: Increase by unlabeled<br />

octreotide. Endocrinology<br />

1995;136:3698–3706.<br />

96 Ur E, Mather SJ, Bomanji J, Ellison<br />

D, Britton KE, Grossman AB,<br />

Wass JAH, Besser GM: Pituitary<br />

imaging using a labelled somatostatin<br />

analogue in acromegaly.<br />

Clin Endocrinol 1992;36:147–<br />

150.<br />

97 Ur E, Bomanji J, Mather SJ, Britton<br />

KE, Wass JAH, Grossman AB,<br />

Besser GM: Localization of neuroendocrine<br />

tumours <strong>and</strong> insulinomas<br />

using radiolabelled somatostatin<br />

analogues, 123 I-Tyr 3 -octreotide<br />

<strong>and</strong> 111 In-pentatreotide. Clin<br />

Endocrinol 1993;38:501–506.<br />

98 American Joint Committee on<br />

<strong>Cancer</strong>: General In<strong>for</strong>mation on<br />

<strong>Cancer</strong> Staging <strong>and</strong> End Results<br />

Reporting, ed 3. Philadelphia, Lippincott,<br />

1988.<br />

99 Krenning EP, Kwekkeboom JD,<br />

Bakker WH, Breeman WAP,<br />

Kooij PPM, Oei Y, van Hagen M,<br />

Postema PTE, de Jong M, Reubi<br />

J-C, Visser TJ, Reijs AEM, Hofl<strong>and</strong><br />

LJ, Koper JW, Lamberts<br />

SWJ: <strong>Somatostatin</strong> receptor scintigraphy<br />

with [ 111 In-DTPAD-Phe 1 ]<strong>and</strong><br />

[ 123 I-Tyr 3 ]-octreotide: The<br />

Rotterdam experience with more<br />

than 1,000 patients. Eur J Nucl<br />

Med 1993;20:716–731.<br />

100 Thomas F, Brambrilla E, Friedmann<br />

A: Transcription of somatostatin<br />

receptor subtype 1 <strong>and</strong><br />

2 genes in lung cancer. Lung <strong>Cancer</strong><br />

1994;11:111–114.<br />

101 Bong SB, V<strong>and</strong>erlaan JG, Louwes<br />

H, Schuurman JJ: Clinical experience<br />

with somatostatin receptor<br />

imaging in lymphoma. Semin Oncol<br />

1994;21(suppl 13):46–50.<br />

102 O’Byrne KJ, Goggins MG,<br />

McDonald GS, Daly PA, Kelleher<br />

DP, Weir DG: A metastatic neuroendocrine<br />

anaplastic small cell<br />

tumour in a patient with MEN 1<br />

syndrome: Assessment of disease<br />

status <strong>and</strong> response to ACE chemotherapy<br />

through scintigraphic<br />

imaging with [ 111 In]pentetreotide.<br />

<strong>Cancer</strong> 1994;74:2374–2378.<br />

103 O’Byrne KJ, O’Hare NJ, Freyne<br />

PJ, Luke DA, Clancy LJ, Prichard<br />

JS, Carney DN: Imaging of bronchial<br />

carcinoid tumours with indium-111<br />

pentetreotide, a radiolabelled<br />

octapeptide analogue of somatostatin.<br />

Thorax 1994;49:284–<br />

286.<br />

104 de Herder WW, Krenning EP,<br />

Malchoff CD, Hofl<strong>and</strong> LJ, Reubi<br />

J-C, Kwekkeboom DJ, Oei HY,<br />

Pols HAP, Bruining HA, Nobels<br />

FRE, Lamberts SWJ: <strong>Somatostatin</strong><br />

receptor scintigraphy: Its value<br />

in tumor localisation in patients<br />

with Cushing’s syndrome caused<br />

by ectopic corticotropin or corticotropin-releasing<br />

hormone secretion.<br />

Am J Med 1994;96:305–312.<br />

105Phlipponneau M, Nocaudie M,<br />

Epelbaum J, de Keyzer Y, Lalau<br />

JD, March<strong>and</strong>ise X, Bertagna X:<br />

<strong>Somatostatin</strong> analogs <strong>for</strong> the localization<br />

<strong>and</strong> preoperative treatment<br />

of an adrenocorticotropin-secreting<br />

bronchial carcinoid tumor. J<br />

Clin Endocrinol Metab 1994;78:<br />

20–24.<br />

106 Iser G, Pfohl M, Dorr U, Weiss<br />

EM, Seif FJ: Ectopic ACTH secretion<br />

due to a bronchopulmonary<br />

carcinoid localized by somatostatin<br />

receptor scintigraphy. Clin<br />

Invest 1994;72:887–891.<br />

107 Weiss M, Yellin A, Husza’r M, Eisenstein<br />

Z, Bar-Ziv J, Kpausz Y:<br />

Localization of adrenocorticotropic<br />

hormone-secreting bronchial<br />

carcinoid tumor by somatostatin<br />

receptor scintigraphy. Ann Intern<br />

Med 1994;121:198–199.<br />

108 Lefebvre H, Jegou S, Leroux P,<br />

Dero M, Vaudry H, Kuhn JM:<br />

Characterization of the somatostatin<br />

receptor subtype in a bronchial<br />

carcinoid tumor responsible<br />

<strong>for</strong> Cushing’s syndrome. J Clin Endocrinol<br />

Metab 1995;80:1423–<br />

1428.<br />

Chemotherapy 2001;47(suppl 2):78–108 105


109 Kalkner KM, Janson ET, Nilsson<br />

S, Carlsson S, Oberg K, Westlin<br />

JE: <strong>Somatostatin</strong> receptor scintigraphy<br />

in patients with carcinoid<br />

tumors: Comparison between radiolig<strong>and</strong><br />

uptake <strong>and</strong> tumor markers.<br />

<strong>Cancer</strong> Res 1995;55(23 suppl):<br />

5801s–5804s.<br />

110 Orsolon P, Bagni B, Basadonna P,<br />

Geatti O, Talmassons G, Guerra<br />

UP: A case of bronchial carcinoid:<br />

<strong>Diagnosis</strong> <strong>and</strong> follow-up with<br />

111 In-DTPA-octreotide. Q J Nucl<br />

Med 1995;39:311–314.<br />

111 Christian-Maitre S, Chabbert-Buffet<br />

N, Mure A, Boukhris R, Bouchard<br />

P: Use of somatostatin analog<br />

<strong>for</strong> localization <strong>and</strong> treatment<br />

of ACTH secreting bronchial carcinoid<br />

tumor. Chest 1996;109:<br />

845–846.<br />

112 Matte J, Roufosse F, Rocmans P,<br />

Schoutens A, Jacobovitz D, Mockel<br />

J: Ectopic Cushing’s syndrome<br />

<strong>and</strong> pulmonary carcinoid tumour<br />

identified by [ 111 In-DTPA-D-<br />

Phe 1 ]octreotide. Postgrad Med J<br />

1998;74:108–110.<br />

113 Mansi L, Rambaldi PF, Panza N,<br />

Esposito V, Pastore V: <strong>Diagnosis</strong><br />

<strong>and</strong> radioguided surgery with<br />

111 In-pentreotide in a patient with<br />

paraneoplastic Cushing’s syndrome<br />

due to a bronchial carcinoid.<br />

Eur J Endocrinol 1997;137:<br />

688–690.<br />

114 Hammond PJ, Wade AF, Gwilliam<br />

ME, Peters AM, Myers MJ,<br />

Gilbey SG, Bloom SR, Calam J:<br />

Amino acid infusion blocks renal<br />

tubular uptake of an indium-labelled<br />

somatostatin analogue. Br J<br />

<strong>Cancer</strong> 1993;67:1437–1439.<br />

115Hosono M, Hosono MN, Haberberger<br />

T, Zamora PO, Guhlke B,<br />

Bender H, Knapp FF, Biersack HJ:<br />

Localization of small cell lung cancer<br />

xenografts with iodine-125-, indium-111-,<br />

<strong>and</strong> rhenium-188-somatostatin<br />

analogs. Jpn J <strong>Cancer</strong><br />

Res 1996;87:995–1000.<br />

116 de Jong M, Breeman WA, Bernard<br />

BF, Rolleman EJ, Hofl<strong>and</strong> LJ,<br />

Visser TJ, Setyono-Han B, Bakker<br />

WH, van der Pluijm ME, Krenning<br />

EP: Evaluation in vitro <strong>and</strong> in<br />

rats of l6l Tb-DTPA-octreotide, a<br />

somatostatin analogue with potential<br />

<strong>for</strong> intraoperative scanning<br />

<strong>and</strong> radiotherapy. Eur J Nucl Med<br />

1995;22:608–616.<br />

117 Anderson CJ, Pajeau TS, Edwards<br />

WB, Sherman EL, Rogers BE,<br />

Welch MJ: In vitro <strong>and</strong> in vivo<br />

evaluation of copper-64-octreotide<br />

conjugates. J Nucl Med 1995;36:<br />

2315–2325.<br />

118 Stolz B, Smith-Jones P, Albert R,<br />

Tolcvsai L, Briner U, Ruser G,<br />

Macke H, Weckbecker G, Bruns C:<br />

<strong>Somatostatin</strong> analogues <strong>for</strong> somatostatin<br />

receptor mediated radiotherapy<br />

of cancer. Digestion<br />

1996;57(suppl 1):17–21.<br />

119 Zamora PO, Guhlke S, Bender H,<br />

Diekmann D, Rhodes BA, Biersack<br />

HJ, Knapp FF Jr: Experimental<br />

radiotherapy of receptor-positive<br />

human prostate adenocarcinoma<br />

with 188 Re-RC-160, a directly<br />

radiolabeled somatostatin analogue.<br />

Int J <strong>Cancer</strong> 1996;65:214–<br />

220.<br />

120 Bepler G, Carney DN, Gazdar AF,<br />

Minna JD: In vitro growth inhibition<br />

of human small cell lung cancer<br />

by physalaemin. <strong>Cancer</strong> Res<br />

1987;47:2371–2375.<br />

121 Kee KA, Finan TM, Korman LY,<br />

Moody TW: <strong>Somatostatin</strong> inhibits<br />

the secretion of bombesin-like peptides<br />

from small cell lung cancer<br />

cells. Peptides 1988;9(suppl 1):<br />

257–261.<br />

122 Taylor JE, Bogden AE, Moreau J-<br />

P, Coy DH: In vitro <strong>and</strong> in vivo<br />

inhibition of human small cell lung<br />

carcinoma (NCI-H69) growth by a<br />

somatostatin analogue. Biochem<br />

Biophys Res Commun 1988;153:<br />

81–86.<br />

123 Taylor JE, Moreau JP, Baptiste L,<br />

Moody TW: Octapeptide analogues<br />

of somatostatin inhibit the<br />

clonal growth <strong>and</strong> vasoactive intestinal<br />

peptide-stimulated cyclic<br />

AMP <strong>for</strong>mation in human small<br />

cell lung cancer cells. Peptides<br />

1991;12:839–843.<br />

124 Prevost G, Bourgeois Y, Mormont<br />

C, Lerrant Y, Veber N, Poupon<br />

MF, Thomas F: Characterization<br />

of somatostatin receptors <strong>and</strong><br />

growth inhibition by the somatostatin<br />

analogue BIM-23014 in<br />

small cell lung carcinoma xenograft:<br />

SCLC-6. Life Sci 1994;55:<br />

155–162.<br />

125Evers M, Parekh D, Townsend<br />

CM, Thompson JC: <strong>Somatostatin</strong><br />

<strong>and</strong> analogues in the treatment of<br />

cancer. Ann Surg 1991;213:190–<br />

198.<br />

126 Hofl<strong>and</strong> LJ, van Koetsveld PM,<br />

Waaijers M, Zuyderwijk J, Lamberts<br />

SWJ: Relative potencies of<br />

the somatostatin analogs octreotide,<br />

BIM-23014, <strong>and</strong> RC-160 on<br />

the inhibition of hormone release<br />

by cultured human endocrine tumor<br />

cells <strong>and</strong> normal rat anterior<br />

pituitary cells. Endocrinology<br />

1994;134:301–306.<br />

127 Damstrup L, Rygaard K, Spang-<br />

Thomsen M, Poulsen HS: Expression<br />

of epidermal growth factor receptor<br />

in human small cell lung<br />

cancer. <strong>Cancer</strong> Res 1992;52:3089–<br />

3093.<br />

128 Veale D, Kerr N, Gibson GJ, Kelly<br />

PJ, Harris AL: The relationship of<br />

quantitative epidermal growth factor<br />

receptor expression in nonsmall<br />

cell lung cancer to long term<br />

survival. Br J <strong>Cancer</strong> 1993;68:<br />

162–165.<br />

129 Favoni RE, de Cupis A, Ravera F,<br />

Cantoni C, Pirani P, Aadizzoni A,<br />

Noonan D, Blassoni R: Expression<br />

<strong>and</strong> function of the insulin-like<br />

growth factor I system in human<br />

non-small cell lung cancer <strong>and</strong> normal<br />

lung cell lines. Int J <strong>Cancer</strong><br />

1994;56:858–866.<br />

130 Fekete M, Zalatnai A, Comaru-<br />

Schally AM, Schally AV: Membrane<br />

receptors <strong>for</strong> peptides in experimental<br />

<strong>and</strong> human pancreatic<br />

cancers. Pancreas 1989;4:521–<br />

528.<br />

131 Srkalovic G, Szende B, Redding<br />

TW, Groot K, Schally AV: Receptors<br />

<strong>for</strong> D-Trp 6 -luteinizing hormone-releasing<br />

hormone, somatostatin,<br />

<strong>and</strong> insulin-like growth factor<br />

I in MXT mouse mammary<br />

carcinoma (42987). Proc Soc Exp<br />

Biol Med 1989;192:209–218.<br />

132 Yano T, Pinski J, Szepeshazi K,<br />

Milovanovic SR, Groot K, Schally<br />

AV: Effect of microcapsules of luteinizing<br />

hormone-releasing hormone<br />

antagonist SB-75 <strong>and</strong> somatostatin<br />

analog RC-160 on endocrine<br />

status <strong>and</strong> tumor growth in<br />

the Dunning R-3327H rat prostate<br />

cancer model. Prostate 1992;20:<br />

297–310.<br />

106 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


133 Pinski J, Halmos G, Yano T, Szepeshazi<br />

K, Qin Y, Ertl T, Schally<br />

AV: Inhibition of growth of<br />

MKN45 human gastric-carcinoma<br />

xenografts in nude mice by treatment<br />

with bombesin/gastrin releasing<br />

peptide antagonist (RC-<br />

3095), <strong>and</strong> somatostatin analogue<br />

RC-160. Int J <strong>Cancer</strong> 1994;57:<br />

574–580.<br />

134 Pinski J, Reile H, Halmos G,<br />

Groot K, Schally AV: Inhibitory<br />

effects of somatostatin analogue<br />

RC-160 <strong>and</strong> bombesin/gastrin-releasing<br />

peptide antagonist RC-<br />

3095 on the growth of the <strong>and</strong>rogen-independent<br />

Dunning R-<br />

3327-AT-1 rat prostate cancer.<br />

<strong>Cancer</strong> Res 1994;54:169–174.<br />

135Pinski J, Schally AV, Halmos G,<br />

Szepeshazi K, Groot K: <strong>Somatostatin</strong><br />

analogues <strong>and</strong> bombesin/<br />

gastrin-releasing peptide antagonist<br />

RC-3095 inhibit the growth of<br />

human glioblastomas in vitro <strong>and</strong><br />

in vivo. <strong>Cancer</strong> Res 1994;54:<br />

5895–5901.<br />

136 Radulovic S, Schally AV, Reile H,<br />

Halmos G, Szepeshazi K, Groot K,<br />

Milovanovic S, Miller G, Yano T:<br />

Inhibitory effects of antagonists of<br />

bombesin/gastrin releasing peptide<br />

(GRP) <strong>and</strong> somatostatin analogue<br />

(RC-160) on growth of HT-29 human<br />

colon cancers in nude mice.<br />

Acta Oncol 1994:33:693–701.<br />

137 Szende B, Srkalovic G, Schally<br />

AV, Lapis K, Groot K: Inhibitory<br />

effects of analogs of luteinizing<br />

hormone-releasing hormone <strong>and</strong><br />

somatostatin on pancreatic cancers<br />

in hamsters. Events that accompany<br />

tumor regression. <strong>Cancer</strong><br />

1990;65:2279–2290.<br />

138 Giatromanolaki A, Koukourakis<br />

M, O’Byrne K, Fox S, Whitehouse<br />

R, Talbot DC, Harris AL, Gatter<br />

K: Prognostic value of angiogenesis<br />

in operable non-small cell lung<br />

cancer. J Pathol 1996;179:80–88.<br />

139 Soresi E, Invernizzi G, Boffi R,<br />

Borghini U, Schiraldi G, Mantellini<br />

PV, Gramegna G, Luizzi A:<br />

Effect of octreotide on neuroenolase<br />

levels in patients with small<br />

cell lung cancer. Tumori 1994;80:<br />

332–334.<br />

<strong>Somatostatin</strong>, Its Receptors <strong>and</strong><br />

<strong>Analogs</strong>, in Lung <strong>Cancer</strong><br />

140 Anthony L, Johnson D, H<strong>and</strong>e K,<br />

Shaff M, Winn S, Krozely M,<br />

Oates J: <strong>Somatostatin</strong> analogue<br />

phase I trials in neuroendocrine<br />

neoplasms. Acta Oncol 1993;32:<br />

217–223.<br />

141 Cotto C, Quoix E, Thomas F, Henane<br />

S, Trillet-Lenoir V: Phase I<br />

study of the somatostatin analogue<br />

somatuline in refractory small cell<br />

lung carcinoma. Ann Oncol 1994;<br />

5:290–291.<br />

142 Rieu M, Rosilio M, Richard A,<br />

Vannetzel JM, Kuhn JM: Paradoxical<br />

effect of somatostatin analogues<br />

on the ectopic secretion of<br />

corticotropin in two cases of small<br />

cell lung carcinoma. Horm Res<br />

1993;39:207–212.<br />

143 Logothetis CJ, Hossan EA, Smith<br />

TL: SMS 201-995 in the treatment<br />

of refractory prostatic carcinoma.<br />

Anticancer Res 1994;14:2731–<br />

2734.<br />

144 Keren-Rosenberg S, Raats JI, Rapoport<br />

BL, Falkson G: <strong>Somatostatin</strong><br />

in the treatment of paraneoplastic<br />

diarrhoea in patients with<br />

small cell lung cancer. Ann Oncol<br />

1992;3:409.<br />

145Hearn PR, Reynolds CL, Johansen<br />

K, Woodhouse NJY: Lung carcinoid<br />

with Cushing’s syndrome:<br />

Control of serum ACTH <strong>and</strong> cortisol<br />

levels using SMS 201-995 (S<strong>and</strong>ostatin).<br />

Clin Endocrinol 1988;<br />

28:181–185.<br />

146 de Rosa G, Testa A, Liberale I, Pirronti<br />

T, Granone P, Picciocchi A:<br />

Successful treatment of ectopic<br />

Cushing’s syndrome with the longacting<br />

somatostatin analog octreotide.<br />

Exp Clin Endocrinol 1993;<br />

101:319–325.<br />

147 Moller DE, Moses AC, Jones K,<br />

Thorner MO, Vance ML: Octreotide<br />

suppresses both growth hormone<br />

(GH) <strong>and</strong> GH-releasing hormone<br />

(GHRH) in acromegaly due<br />

to ectopic GHRH secretion. J Clin<br />

Endocrinol Metab 1989;68:499–<br />

504.<br />

148 Ohnsmann A, Sachsenheimer W:<br />

Intracerebral metastasis of a bronchial<br />

carcinoid tumor. Neurochirurgia<br />

(Stuttg) 1992;35:160–162.<br />

149 Cheung NW, Boyages SC: Failure<br />

of somatostatin analogue to control<br />

Cushing’s syndrome in two<br />

cases of ACTH-producing carcinoid<br />

tumours. Clin Endocrinol<br />

(Oxf) 1992;36:361–367.<br />

150 Oliaro A, Filosso PL, Casadio C,<br />

Ruffini E, Mazza E, Molinatti M,<br />

Cianci R, Porrello C, Rastelli M,<br />

Oliveri F, et al: Bronchial carcinoid<br />

associated with Cushing’s<br />

syndrome. J Cardiovascular Surg<br />

1995;36:511–514.<br />

151 O’Byrne KJ, O’Hare N, Sweeney<br />

E, Freyne PJ, Cullen MJ: <strong>Somatostatin</strong><br />

<strong>and</strong> somatostatin analogues<br />

in medullary thyroid carcinoma.<br />

Nucl Med Commun 1996;17:810–<br />

816.<br />

152 Scarpignato C, Camboni MG:<br />

Safety profile of octreotide; in<br />

Scarpigneto C (ed): Octreotide:<br />

From Basic Science to Clinical<br />

Medicine. Basel, Karger, 1996, pp<br />

296–309.<br />

153 Day SM, Gu J, Polak JM, Bloom<br />

SR: <strong>Somatostatin</strong> in the human<br />

heart <strong>and</strong> comparison with guinea<br />

pig <strong>and</strong> rat heart. Br Heart J 1985;<br />

53:153–157.<br />

154 Szepeshazi K, Lapis K, Schally<br />

AV: Effect of combination treatment<br />

with analogs of luteinizing<br />

hormone-releasing hormone (LH-<br />

RH) or somatostatin <strong>and</strong> 5-fluorouracil<br />

on pancreatic cancer in<br />

hamsters. Int J <strong>Cancer</strong> 1991;49:<br />

260–266.<br />

155 Reichert S, Truchetet F, Cuny JF,<br />

Gr<strong>and</strong>idier M: Tumeur carcinoïde<br />

à révélation cutanée. Ann Dermatol<br />

Vénéréol 1994;121:485–488.<br />

156 Romani R, Morris DL: SMS<br />

201.995 (s<strong>and</strong>ostatin) enhances in<br />

vitro effects of 5-fluorouracil in<br />

colorectal cancer. Eur J Surg Oncol<br />

1995;21:27–32.<br />

157 Lee JM, Erlich RB, Bruckner HW,<br />

Szrajer L, Ohnuma T: A somatostatin<br />

analogue (SMS 201-995) alters<br />

the toxicity of 5-fluorouracil in<br />

Swiss mice. Anticancer Res 1993;<br />

13:1453–1456.<br />

158 Wadler S, Haynes H, Wiernik PH:<br />

Phase I trial of the somatostatin<br />

analog octreotide acetate in the<br />

treatment of fluoropyrimidine-induced<br />

diarrhoea. J Clin Oncol<br />

1995;13:222–226.<br />

Chemotherapy 2001;47(suppl 2):78–108 107


159 Weckbecker G, Raulf F, Tolcsvai<br />

L, Bruns C: Potentiation of the<br />

anti-proliferative effects of anticancer<br />

drugs by octreotide in vitro<br />

<strong>and</strong> in vivo. Digestion 1996;<br />

57(suppl 1):22–28.<br />

160 Chen TB, Huzak M, Macura S,<br />

Vuk-Pavlovic S: <strong>Somatostatin</strong> analogue<br />

octreotide modulates metabolism<br />

<strong>and</strong> effects of 5-fluorouracil<br />

<strong>and</strong> 5-fluorouridine in human colon<br />

cancer spheroids. <strong>Cancer</strong> Lett<br />

1994;86:41–51.<br />

161 Thomas F, Bourgeois E, Poupon<br />

MF: Additive inhibitory effects of<br />

cyclophosphamide (CPM) <strong>and</strong> of a<br />

somatostatin analogue lanreotide<br />

(Lan) on the growth of a small cell<br />

lung cancer (SCLC) xenograft (abstract).<br />

Proc Am Assoc Clin Oncol<br />

1992;11:1036.<br />

162 Teicher BA, Holden SA, Ara G,<br />

Sotomayor EA, Huang ZD, Chen<br />

YN, Brem H: Potentiation of cytotoxic<br />

cancer therapies by TNP470<br />

alone <strong>and</strong> with other anti-angiogenic<br />

agents. Int J <strong>Cancer</strong> 1994;57:<br />

920–925.<br />

163 Koukourakis MI, Giatromanolaki<br />

A, O’Byrne KJ, Comley M, Whitehouse<br />

RM, Talbot DC, Gatter KC,<br />

Harris AL: Platelet-derived endothelial<br />

cell growth factor expression<br />

correlates with tumour angiogenesis<br />

<strong>and</strong> prognosis in non-small<br />

cell lung cancer. Br J <strong>Cancer</strong> 1996;<br />

75:477–481.<br />

164 Rosenberg L, Barkun AN, Denis<br />

MH, Pollak M: Low dose octreotide<br />

<strong>and</strong> tamoxifen in the treatment<br />

of adenocarcinoma of the<br />

pancreas. <strong>Cancer</strong> 1995;75:23–28.<br />

165McClay EF, Albright KD, Jones<br />

JA, Christen RD, Howell SB: Tamoxifen<br />

modulation of cisplatin<br />

sensitivity in human malignant<br />

melanoma cells. <strong>Cancer</strong> Res 1993;<br />

53:1571–1576.<br />

166 Non-Small Cell Lung <strong>Cancer</strong> Collaborative<br />

Group: Chemotherapy<br />

in non-small cell lung cancer: A<br />

meta-analysis using updated data<br />

on individual patients from 52<br />

r<strong>and</strong>omised clinical trials. Br Med<br />

J 1995;311:899–909.<br />

167 van Z<strong>and</strong>wijk N, Giaccone G:<br />

<strong>Treatment</strong> of metastatic non-small<br />

cell lung cancer. Curr Opin Oncol<br />

1996, 8:120–125.<br />

168 Radulovic S, Nagy A, Szoke B,<br />

Schally AV: Cytotoxic analog of<br />

somatostatin containing methotrexate<br />

inhibits growth of MIA<br />

PaCa-2 human pancreatic cancer<br />

xenografts in nude mice. <strong>Cancer</strong><br />

Lett 1992;62:263–271.<br />

169 Nagy A, Schally AV, Halmos G,<br />

Armatis P, Cai R-Z, Csernus V,<br />

Kovacs M, Koppan M, Szepeshazi<br />

K, Kahan Z: Synthesis <strong>and</strong> biological<br />

evaluation of cytotoxic analogs<br />

of somatostatin containing doxorubicin<br />

or its intensely potent derivative,<br />

2-pyrrolinodoxorubicin.<br />

Proc Natl Acad Sci 1998;95:1794–<br />

1799.<br />

170 Koppan M, Nagy A, Schally AV,<br />

Arencibia JM, Plonowski A, Halmos<br />

G: Targeted cytotoxic analogue<br />

of somatostatin AN-238 inhibits<br />

growth of <strong>and</strong>rogen-independent<br />

Dunning R-3327-AT-1<br />

prostate cancer in rats at nontoxic<br />

doses. <strong>Cancer</strong> Res 1998;58:4132–<br />

4137.<br />

171 Schaffer M, Barbul A: Lymphocyte<br />

function in wound healing <strong>and</strong> following<br />

surgery. Br J Surg 1998;85:<br />

444–460.<br />

172 Bogden AE, Moreau JP, Eden PA:<br />

Proliferation response of human<br />

<strong>and</strong> animal tumours to surgical<br />

wounding of normal tissues: Onset,<br />

duration <strong>and</strong> inhibition. Br J<br />

<strong>Cancer</strong> 1997;75:1021–1027.<br />

173 Baserga R, Hongo A, Rubini M,<br />

Prisco M, Valentinis B: The IGF-I<br />

receptor in cell growth, trans<strong>for</strong>mation<br />

<strong>and</strong> apoptosis. Biochim<br />

Biophys Acta 1997;1332:F105–<br />

126.<br />

174 Hennessey JV, Jackson IM: Clinical<br />

features <strong>and</strong> differential diagnosis<br />

of pituitary tumours with<br />

emphasis on acromegaly. Baillières<br />

Clin Endocrinol Metab 1995;9:<br />

271–314.<br />

175Chan JM, Stampfer MJ, Giovannucci<br />

E, Gann PH, Ma J, Wilkonson<br />

P, Hennekens CH, Pollak M:<br />

Plasma insulin-like growth factor-I<br />

<strong>and</strong> prostate cancer risk: A prospective<br />

study. Science 1998;279:<br />

563–565.<br />

176 Hankinson SE, Willett WC, Colditz<br />

GA, Hunter DJ, Michaud DS,<br />

Deroo B, Rosner B, Speizer FE,<br />

Pollak M: Circulating concentrations<br />

of insulin-like growth factor-I<br />

<strong>and</strong> risk of breast cancer. Lancet<br />

1998;351:1393–1396.<br />

177 Szende B, Juhasz E, Lapis K,<br />

Schally AV: Inhibition of two-step<br />

urinary bladder carcinogenesis by<br />

the somatostatin analogue RC-<br />

160. Urol Res 1992;20:383–386.<br />

178 Liebow C, Crean DH, Schally AV,<br />

Mang TS: Peptide analogues alter<br />

the progression of premalignant lesions,<br />

as measured by Photofrin<br />

fluorescence. Proc Natl Acad Sci<br />

USA 1993;90:1897–1901.<br />

179 Crean DH, Liebow C, Lee MT,<br />

Kamer AR, Schally AV, Mang TS:<br />

Alterations in receptor-mediated<br />

kinases <strong>and</strong> phosphatases during<br />

carcinogenesis. J <strong>Cancer</strong> Res Clin<br />

Oncol 1995;121:141–149.<br />

180 Helle SI, Geisler J, Poulsen JP,<br />

Hestdal K, Meadows K, Collins W,<br />

Tveit KM, Viste A, Holly JMP,<br />

Lonning PE: Microencapsulated<br />

octreotide palmoate in advanced<br />

gastrointestinal <strong>and</strong> pancreatic<br />

cancer: A phase I study. Br J <strong>Cancer</strong><br />

1998;78:14–20.<br />

181 Holly J: Insulin-like growth factor-<br />

I <strong>and</strong> new opportunities <strong>for</strong> cancer<br />

prevention. Lancet 1998;351:<br />

1373–1374.<br />

108 Chemotherapy 2001;47(suppl 2):78–108 O’Byrne/Schally/Thomas/Carney/<br />

Steward


Chemotherapy 2001;47(suppl 2):109–126<br />

Octreotide in the Management of<br />

Hormone-Refractory Prostate <strong>Cancer</strong><br />

Iraklis G. Vainas<br />

Department of Endocrinological Oncology, Theagenion <strong>Cancer</strong> Center, Thessaloniki, Greece<br />

Key Words<br />

Prostate cancer, hormone-refractory W<br />

Octreotide W Anti<strong>and</strong>rogen blockade,<br />

complete W Hormonal maneuvers, alternative<br />

Abstract<br />

Patients with advanced or metastatic prostate<br />

cancer (PC), a partially hormone-resistant<br />

disease, will require some <strong>for</strong>m of hormonal<br />

manipulations or some new therapeutic<br />

modalities. Octreotide, as somatostatin<br />

(SST) analogs, has been found to inhibit<br />

the growth of experimental PCs via several<br />

mechanisms, as indirect antihormonal <strong>and</strong><br />

direct antimitogenic actions, mainly due to<br />

inhibition of SST receptor subtypes (SSTR-1–<br />

5). Sporadic clinical trials with octreotide<br />

(alone or with a complete anti<strong>and</strong>rogen<br />

blockade) treatment of patients with advanced<br />

stage D2 PC demonstrated promising<br />

results. Un<strong>for</strong>tunately, at present these clinical<br />

trials have some disadvantages <strong>and</strong> leave<br />

some uncertainty with regard to the trial design,<br />

the SSTR subtype determination <strong>and</strong><br />

tumor localization with SSTR scintigraphy<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0109$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

be<strong>for</strong>e the start of a selective SST analog,<br />

<strong>and</strong> finally the r<strong>and</strong>omization in groups according<br />

to hormone resistance, dosage regimen<br />

<strong>and</strong> route of administration.<br />

Epidemiology<br />

Dr. Iraklis Vainas<br />

Theagenion <strong>Cancer</strong> Center<br />

2, Al. Simeonidis Street<br />

GR–54007 Thessaloniki (Greece)<br />

Tel. +30 31 82 92 12, Fax +30 31 84 55 14<br />

Copyright © 2001 S. Karger AG, Basel<br />

Prostate cancer (PC) is the most common<br />

cancer in males <strong>and</strong> is secondary in incidence<br />

only to breast cancer as the most common<br />

type of hormone-dependent neoplasm in the<br />

USA. In Europe it accounted <strong>for</strong> 20% of all<br />

newly diagnosed cancers in 1990 <strong>and</strong> <strong>for</strong><br />

160,000 patients only in 1993 [1]. Furthermore,<br />

PC is the second most common cause<br />

of death due to cancer in men, with 35,000<br />

deaths in 1993. Thus, it had been suggested<br />

that PC will become the leading cause of cancer<br />

death in men by the year 2000 [2]. In more<br />

than 50% of patients the disease is only diagnosed<br />

at a late stage, usually in men between<br />

the ages of 45 <strong>and</strong> 67, when it has already<br />

spread to distant sites, particularly to the<br />

bone, <strong>and</strong> at a time when most men have their


highest level of social responsibilities [3].<br />

Thus, advanced-stage PC is a significant<br />

health problem in the male population.<br />

Natural History – Clinical Course of<br />

the Disease<br />

The natural history of PC should be reflected<br />

by its stage, describing the burden <strong>and</strong><br />

the extent of the tumor at the time of diagnosis<br />

[1]. All staging maneuvers (Whitemore-<br />

Jewett or Tumor Node Metastasis system)<br />

determine if the disease is widespread or confined<br />

to the prostate, but understaging is a<br />

great problem clinically.<br />

The natural history of stage A1 PC remains<br />

unclear, because 16% of the patients will progress<br />

3.5–8 years after the initial diagnosis,<br />

<strong>and</strong> 61% actually have stage A2 or diffuse disease.<br />

Stage A2 or B1–3 PC patients treated<br />

with radical prostatectomy may have an<br />

equivalent risk <strong>for</strong> disease spread <strong>and</strong> death<br />

from PC [4]. Advanced-stage disease (C, D1,<br />

D2) will be identified in more than 40% of<br />

these cases [5]. About 60% of the patients<br />

with untreated stage C PC will exhibit evidence<br />

of disease progression at 5 years, at<br />

rates of 10–20% annually [5–7]. 85% of the<br />

patients with stage D1 PC demonstrate a disease<br />

progression within 5 years of diagnosis<br />

[5, 8], although diploid tumors have a biologically<br />

indolent course [9]. Finally, stage D2 PC<br />

patients (with distant metastases) have a median<br />

survival of about 30 months with a 20%<br />

5-year survival rate.<br />

Endocrinological Aspects<br />

The role of testicular <strong>and</strong>rogens in the evolution<br />

of PC has been recognized since 1941<br />

[10]. Since then surgical castration or treatment<br />

with estrogens has been reported to<br />

110 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

result in subjective-objective improvement in<br />

60–80% of patients, due to neutralization of<br />

<strong>and</strong>rogens, but <strong>for</strong> limited time intervals. The<br />

endocrine relationship between the testis, hypophysis<br />

<strong>and</strong> hypothalamus is shown in figure<br />

1. Ninety percent of the circulating <strong>and</strong>rogen<br />

pool in the normal male is derived from<br />

the testes <strong>and</strong> 10% from adrenal cortex<br />

sources, regulated by the hypothalamus (via<br />

CRF) <strong>and</strong> directly by the pituitary (via<br />

ACTH) [11]. The prostate gl<strong>and</strong> has <strong>and</strong>rogen<br />

receptors, so the circulating testosterone (T) is<br />

rapidly internalized into its cytosol <strong>and</strong> its<br />

receptors, <strong>and</strong> converted to dehydro-T (DHT)<br />

by the enzyme 5-reductase. DHT is 3 times<br />

more potent than T, with T being 5–10 times<br />

more potent than adrenal <strong>and</strong>rogens [12].<br />

DHT, from testicular or adrenal sources, is<br />

the primary (growth-promoting) intracellular<br />

messenger in <strong>and</strong>rogen-responsive target cells<br />

<strong>and</strong> is internalized into the nucleus, where an<br />

mRNA clone is transcribed, which ultimately<br />

effects translation of <strong>and</strong>rogen-dependent<br />

proteins, necessary <strong>for</strong> cellular existence [12,<br />

13].<br />

Current Therapeutic Approach in PC<br />

General Therapeutic Considerations<br />

As a general therapeutic strategy, patients<br />

with localized stages A <strong>and</strong> B PC will be managed<br />

by local external radiotherapy (ER) or<br />

radical prostatectomy [4]. In contrast, stage D<br />

(T3N0–3M0–1) disease is associated with systemic<br />

spread <strong>and</strong> should be treated with systemic<br />

cytostatic therapy. Node-positive PC<br />

(D1) is also associated with systemic disease,<br />

<strong>and</strong> there<strong>for</strong>e does not respond to ER or radical<br />

prostatectomy treatment, but can be controlled,<br />

at least <strong>for</strong> some time, with systemic<br />

therapy, in particularly early <strong>and</strong>rogen deprivation.<br />

The therapeutic strategy <strong>for</strong> stage C3<br />

PC (large volume local disease) is similar,


Fig. 1. Endocrine relationship between hypothalamus, pituitary, testis, adrenal <strong>and</strong> prostate<br />

gl<strong>and</strong>. Hormonal therapies in advanced PC. LHRH = Luteinizing hormone-releasing hormone;<br />

CRH = corticotropin-releasing hormone; LH, FSH = luteinizing <strong>and</strong> follicle-stimulating<br />

hormone; ¢4-dione = <strong>and</strong>rostenedione; DHEA = dehydroepi<strong>and</strong>rosterone; T = testosterone;<br />

DHT = dihydrotestosterone; AR = <strong>and</strong>rogen receptor.<br />

which cannot be controlled by local therapies.<br />

The decision <strong>for</strong> stage C1–2 (local extension)<br />

is more complex, combining local radiotherapy<br />

<strong>and</strong> <strong>and</strong>rogen deprivation [11].<br />

Specific Therapeutic Modalities <strong>for</strong><br />

Advanced PC<br />

Advanced or metastatic PC is a partially<br />

hormone-resistant disease, which needs systemic<br />

therapy from the outset. Adjuvant ER is<br />

attempted but is associated with undesirable<br />

side effects (13% bowel obstruction or severe<br />

cystitis or urethra strictures) <strong>and</strong> the longterm<br />

benefit of radiotherapy has yet to be<br />

demonstrated. Major surgical therapies (other<br />

than radical prostatectomy) such as cystoprostatectomy,<br />

total pelvic exenteration, or<br />

salvage surgery do not achieve a complete<br />

resection of the tumor either, nor do they<br />

delay the progression of the cancer or prolong<br />

survival [14]. These procedures are also risky<br />

being associated with a high incidence of postoperative<br />

incontinence. Almost all stage C<br />

<strong>and</strong> D PC patients will require some <strong>for</strong>m of<br />

hormonal manipulations, vitamin therapy,<br />

immune modulation, cytostatics or some new<br />

therapeutic modalities [11].<br />

Surgical or chemical castration or estrogens<br />

are the first <strong>and</strong> main endocrine therapies<br />

<strong>for</strong> patients with advanced PC (table 1).<br />

These treatments have similar efficacies with<br />

respect to survival, relief of symptoms <strong>and</strong> the<br />

prevention of metastatic spread [12, 15].<br />

However, <strong>and</strong>rogen deprivation may result in<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 111


Table 1. Conventional hormonal manipulations in<br />

advanced PC<br />

Castration<br />

Surgical (orchidectomy, hypophysectomy)<br />

Chemical (LHRH analogs, estrogens)<br />

Anti<strong>and</strong>rogens<br />

Gestagenic compounds (cyproterone acetate)<br />

Pure nongestagenic compounds (flutamide,<br />

nilutamide, Casodex)<br />

Progesterone compounds<br />

(medroxyprogesterone, megestrol)<br />

Adrenal corticolytic agents<br />

Aminoglutethimide<br />

Ketoconazole<br />

Ablative hormonal manipulations<br />

Hypophysectomy<br />

Adrenalectomy<br />

CAB<br />

Castration + anti<strong>and</strong>rogens<br />

Table 2. Recent hormonal therapies in advanced PC<br />

(under research)<br />

Reductase inhibitors (finasteride)<br />

LHRH antagonists<br />

Specific SST analogs with high antitumor activities<br />

LHRH analogs with cytotoxic radicals<br />

Bombesin/GRP antagonists<br />

CAB + chemotherapy<br />

CAB + suramin<br />

CAB + synthetic retinoids<br />

Chemotherapy following <strong>and</strong>rogen priming<br />

Specific radiolabeled ( 123 I, 131 I, 111 In) SST analogs<br />

(with ß-emitting radiation)<br />

unpleasant largely minor side effects (loss of<br />

libido, flushes, gynecomastia, impotence), but<br />

some are also serious (mainly due to estrogens),<br />

as the 7–11% incidence of thromboembolic<br />

<strong>and</strong> cardiovascular complications [14].<br />

112 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

Adrenal corticolytic agents (aminoglutethimide,<br />

ketoconazole), synthetic progesterone<br />

compounds (megestrol, medroxyprogesterone)<br />

<strong>and</strong> nonsteroidal pure anti<strong>and</strong>rogens<br />

(flutamide, nilutamide, Casodex) appear to be<br />

beneficial as palliative therapy in 20% of patients,<br />

but with a high percentage of side<br />

effects [12]. Complete <strong>and</strong>rogen blockade<br />

(CAB) combines surgical or chemical castration<br />

with a pure anti<strong>and</strong>rogen, <strong>and</strong> eliminates<br />

both testicular <strong>and</strong> adrenal <strong>and</strong>rogens in PC<br />

patients [11, 16]. Labrie et al. [16] found complete<br />

response rates of 29% at 22 months compared<br />

to the 5% response rate usually associated<br />

with monotherapies; similar results were<br />

reported by Navartil’s group in France with<br />

70% response rates following surgical <strong>and</strong><br />

chemical castration at 18 months versus 45%<br />

with monotherapies. In a large multicenter<br />

trial in the USA of 603 metastatic PC patients,<br />

CAB showed a longer disease-free survival<br />

than either drug alone <strong>and</strong> had a survival<br />

advantage of approximately 7 months, particularly<br />

among the 82 patients with a good<br />

per<strong>for</strong>mance status <strong>and</strong> with minimal disease<br />

[17]. Thus, the CAB has some advantage, the<br />

side effects are no more than with each drug<br />

alone, but the cost of the combined treatment<br />

is very high.<br />

Other Therapeutic Modalities<br />

PC is a biologically heterogeneous tumor,<br />

with an increasing number of hormone-resistant<br />

cell populations at late stages [18]. Thus,<br />

hormone treatment alone does not appear to<br />

be an effective long-term (or even curative)<br />

therapy. Un<strong>for</strong>tunately, the efficacy of several<br />

alternative modes of systemic drug treatment,<br />

including chemotherapy alone [19] or combined<br />

with hormone therapy or following <strong>and</strong>rogen<br />

priming [20], suramin [21], or natural<br />

synthetic retinoids [22], has yet to be determined.<br />

Alternative endocrine treatment (table<br />

2) with LHRH antagonists, bombesin an-


tagonists or LHRH analogs with cytotoxic<br />

radicals, or recently with somatostatin (SST)<br />

analogs are still at the research stage [23, 24].<br />

<strong>Somatostatin</strong> <strong>and</strong> <strong>Somatostatin</strong><br />

<strong>Analogs</strong> in the <strong>Treatment</strong> of<br />

Metastatic PC<br />

Introduction<br />

Natural or synthetic somatostatin (SST-<br />

14) has numerous exocrine/endocrine antisecretory<br />

effects on a variety of systems [25–<br />

27]. In addition SST-14 also acts as an endogenous<br />

growth inhibitor [28]. The growth<br />

inhibitory effects of the hormone have been<br />

documented in multiple experimental <strong>and</strong><br />

human benign <strong>and</strong> malignant tumors [26],<br />

such as hypophyseal [25], endocrine pancreatic<br />

[29], ectopic <strong>and</strong> various solid tumors<br />

[30]. SST-14 has, there<strong>for</strong>e, nonselective,<br />

multiple actions, <strong>and</strong> a short duration<br />

of antisecretory or antitumor effects, because<br />

of its short biological half-life (1–2 min).<br />

Consequently SST analogs have been designed<br />

<strong>and</strong> synthesized, which are more resistant<br />

to metabolic degradation <strong>and</strong> have a<br />

more prolonged duration of actions [29, 30].<br />

Due to substitutions <strong>and</strong> incorporation of Damino<br />

acids into the SST molecule, the resistance<br />

of SST analogs to digestion by tissue<br />

enzymes is increased [26, 31]. For example,<br />

9 of the 14 amino acids were replaced with a<br />

single proline residue to produce a long-acting<br />

analog of somatostatin [32, 33]. Veber et<br />

al. [33] retained the sequence 7–10 of SST-<br />

14 (Phe-Trp-Lys-Trp) as essential <strong>for</strong> the<br />

biological activity of native somatostatin,<br />

<strong>and</strong> incorporated the Trp residue into the D<br />

configuration (series of cystine-bridged SST<br />

analogs). Octreotide (SMS 201-995), an SST<br />

analog produced by S<strong>and</strong>oz, has a longer duration<br />

of action (up to 113 min), is more<br />

potent than SST-14 (40–70 times) <strong>and</strong> is<br />

more selective <strong>for</strong> hGH suppression than insulin<br />

or glucagon [33, 34].<br />

Although the greater potency of SST analogs<br />

is only due to resistance to degradation by<br />

enzymes <strong>and</strong> to their prolonged half-lives,<br />

there is some doubt about the selectivity of<br />

the analogs, unless the tissues express the<br />

receptor subtypes, to which they bind with<br />

approximately equal efficacy as the native<br />

hormone. The five cell surface SST receptor<br />

(SSTR) subtypes have been characterized<br />

within the past 5 years <strong>and</strong> have been termed<br />

SSTR-1 through SSTR-5 according to the<br />

chronology of their discovery. The tissue distribution<br />

of the SSTRmRNAs (by Northern<br />

blotting, RT-PCR <strong>and</strong> in situ hybridization)<br />

showed that SSTR-1smRNA <strong>and</strong><br />

SSTR-2smRNA occur in central as well as<br />

peripheral tissues (stomach, jejunum, colon<br />

<strong>and</strong> pancreatic islets), while SSTR-3smRNA<br />

<strong>and</strong> SSTR-4smRNA are limited to brain <strong>and</strong><br />

endocrine pancreas [35–38]. The SSTR-5smRNA<br />

is present in the pituitary <strong>and</strong> a variety<br />

of peripheral tissues including the small<br />

intestine [39]. SST-14 (as SST-28) binds to<br />

all SSTRs with affinities ranging from 0.2 to<br />

2.6 nM, but SST-28 shows a 10- to 20-fold<br />

higher affinity <strong>for</strong> SSTR-4 <strong>and</strong> SSTR-5.<br />

SMS 201-995 (S<strong>and</strong>ostatin) has 3- <strong>and</strong> 10fold<br />

higher affinities <strong>for</strong> SSTR-2 <strong>and</strong> SSTR-4,<br />

<strong>and</strong> a much lower affinity <strong>for</strong> the SSTR-1,<br />

SSTR-3 <strong>and</strong> SSTR-5 subtypes. Other SST<br />

analogs, such as MK-618, CGP-23996 <strong>and</strong><br />

BIM-23014, bind mainly to SSTR-3, while<br />

CGP-23996 also binds to SSTR-5 with high<br />

affinity.<br />

These different affinities of SST-14 <strong>and</strong><br />

SST-28 to the SSTR subtypes, which are consistent<br />

with tissue location of the SSTR subtypes<br />

<strong>and</strong> with their relative potency, suggest<br />

different functional properties <strong>for</strong> the various<br />

SSTRs [40]. However, the overlapping of the<br />

distribution of these five SSTR subtypes<br />

makes the ascription of a given physiological<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 113


Table 3. Antitumor effects of modern octapeptide<br />

analogs in various animal or human cancers<br />

SCLC cell lines<br />

Dunning PC in rats<br />

Ductal pancreatic cancers in hamsters<br />

Human exocrine pancreas cancer<br />

Human PC<br />

Human breast cancer<br />

Human lung cancer<br />

Bladder cancer<br />

Carcinoma of the colon<br />

Primary brain tumors<br />

function to a specific SSTR subtype difficult.<br />

In this respect, the availability of long-acting<br />

<strong>and</strong> relatively specific SST analogs would be<br />

of crucial interest. In particular, the SSTR-2<br />

agonist octreotide <strong>and</strong> other modern SST<br />

analogs, besides their antisecretory effects on<br />

gastric, pancreatic <strong>and</strong> intestinal secretions,<br />

were reported (1) to inhibit in vitro <strong>and</strong> in<br />

vivo cultured tumor cell growth, (2) to enhance<br />

cell apoptosis, <strong>and</strong> (3) to activate cell<br />

cycle-regulated phosphatases at the cytoplasmic<br />

<strong>and</strong> nuclear levels [40]. Subsequently,<br />

Schally et al. [30] synthesized (1) nearly 300<br />

octapeptide analogs with disulfide linkages,<br />

using also a C-terminal amide, (2) slow release<br />

preparations <strong>for</strong> intramuscular injections<br />

once monthly [27], <strong>and</strong> finally (3) new,<br />

modern SST analogs (RC-95-1, RC-121, RC-<br />

160-II), by incorporation of Tyr <strong>and</strong> Val in<br />

position 3 <strong>and</strong> 6 (corresponding to residues 7<br />

<strong>and</strong> 10), with high antitumor activities [41,<br />

42]. The specific antitumor activities of the<br />

modern octapeptide analogs have been studied<br />

in various animal <strong>and</strong> human tumors (table<br />

3). SST analogs, <strong>and</strong> especially octreotide,<br />

are well tolerated <strong>and</strong> have a favorable riskbenefit<br />

ratio, even in overdosage regimens<br />

with 3,000 Ìg/daily [43]. The observed main<br />

side effects were facial flushing, headache,<br />

114 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

nausea, vomiting, abdominal cramps, abdominal<br />

bloating <strong>and</strong>/or flatulence, steatorrhea<br />

<strong>and</strong> cholelithiasis [31, 43]. Although side effects<br />

of the octreotide treatment are usually<br />

mild, there is also the possibility of them<br />

being serious, <strong>for</strong>tunately in a minority of<br />

patients. Thus, occasionally steatorrhea may<br />

be serious enough to discontinue octreotide<br />

treatment, despite injecting octreotide between<br />

meals or giving pancreatic enzymes<br />

during the meals [24, 31, 43]. High-dose octreotide<br />

treatment may reduce gut motility<br />

<strong>and</strong> lead to more serious abdominal side effects<br />

mimicking gut obstruction <strong>and</strong> ileus<br />

[43]. Complications of the octreotide treatment,<br />

such as heavily symptomatic cholecystitis<br />

or acute pancreatitis, have been rare, but<br />

we observed them in 2 patients [24], which<br />

has also been observed by others [44–46].<br />

Actual liver cell damage, pulmonary dysfunction<br />

<strong>and</strong> serious anaphylactoid reactions are<br />

some other very rare, worrying side effects of<br />

octreotide treatment [43].<br />

Presence of SSTR in PC Cells<br />

Since 1986, low- <strong>and</strong> high-affinity SST-14<br />

receptors have been identified <strong>and</strong> measured<br />

in various experimental <strong>and</strong> human tumors,<br />

such as brain, pituitary, endocrine gastrointestinal,<br />

prostate, pancreatic, colon <strong>and</strong> ovarian<br />

tumors, as well as in SCLC cell lines [30,<br />

47–51]. The first studies have clearly shown<br />

the distribution of SSTRs in a wide variety of<br />

human tumors, but these SSTRs are localized<br />

in selected types of such tumors; <strong>for</strong> example,<br />

there is a very high incidence of high-affinity<br />

<strong>and</strong> specific SSTRs in meningiomas, in GRFproducing<br />

tumors, pituitary adenomas from<br />

acromegalics, but only in a small percentage<br />

of mammary cancers [48]. Interestingly,<br />

SSTRs are also present, often in high density,<br />

in tumors (e.g. meningiomas originating from<br />

tissue, arachnoidal cells of the human leptomeninx),<br />

without an established link to an


SSTR mechanism. The same is also true <strong>for</strong><br />

mammary tumors; there<strong>for</strong>e, only tumorously<br />

trans<strong>for</strong>med cells of breast cancer bear<br />

SSTRs, although it has never been shown that<br />

SST-14 plays a functional role in normal<br />

mammary gl<strong>and</strong>s.<br />

Specific SST-14 receptors of low affinity<br />

(1.3 Nm) <strong>and</strong> very high capacity (Bmax 543<br />

fmol/mg) were detected on Dunning R-<br />

3327H rat PC cell membranes [52]. Modern<br />

SST analog RC-160 decreased the weight <strong>and</strong><br />

volume of this prostate tumor, when given<br />

combined with the LHRH analog. This effect<br />

was due to a significant suppression of the<br />

SSTR capacity, as well as of the capacity <strong>and</strong><br />

number of PRL receptors. In contrast, Fekete<br />

et al. [53] did not find a binding capacity to<br />

SST-14 on the Dunning rat or on normal <strong>and</strong><br />

tumorous human prostate cell membranes.<br />

Reubi et al. [54] using 125 I-(Tyr3)-octreotide,<br />

with in vitro receptor autoradiography<br />

techniques, suggested the presence of various<br />

SSTR subtypes on PC cells, expressing highaffinity<br />

receptors <strong>for</strong> SST-14 <strong>and</strong> SST-28, but<br />

low affinity <strong>for</strong> octreotide. Thus, human PC<br />

may be a target <strong>for</strong> SST therapy. However,<br />

SST analogues with different selectivities,<br />

particularly binding affinities, <strong>for</strong> SSTR subtypes<br />

cloned in PC would be required to<br />

achieve a response [55–57]. The cloning of 5<br />

SSTR subtypes, using in situ hybridization, in<br />

a large variety of human tumors initiated a<br />

number of studies on SSTR subtypes in PC<br />

[55]. Reubi et al. [55] first detected, via in situ<br />

hybridization studies, that PC preferentially<br />

expressed the SSTR-1 subtype compared with<br />

the SSTR-2 or SSTR-3 subtypes. Furthermore,<br />

the presence of SSTR-2- or SSTR-<br />

3mRNAs generally correlated with the presence<br />

of octreotide binding sites. It was<br />

pointed out that octreotide, which mainly<br />

binds to the SSTR-2 subtype, would not be<br />

the ideal SST analogue in the treatment of PC.<br />

In other words, octreotide will have no effect<br />

on PC (or other tumors), which express the<br />

SSTR-1, SSTR-3 <strong>and</strong> SSTR-4 subtypes, because<br />

it has little or no binding affinities <strong>for</strong><br />

these receptors. Thus, especially SSTR-2 represents<br />

an SSTR subtype target <strong>for</strong> the development<br />

of more specific SST analogues with<br />

higher affinities to these receptor subtypes<br />

[55]. In contrast, Prevost et al. [58] demonstrated<br />

a 57-kD SSTR-2 subtype in all prostatic<br />

normal <strong>and</strong> tumoral tissues; thus, the<br />

SSTR-2 subtype probably represents the<br />

SSTR subtype target <strong>for</strong> the specific SST analogs.<br />

Sinisi et al. [59] also demonstrated with<br />

the PCR technique that the SSTR-1 subtype<br />

was expressed only in the epithelial cells of<br />

PC, the SSTR-2 only in the epithelial cells of<br />

normal prostate, while the SSTR-3 subtype<br />

was undetectable in normal <strong>and</strong> tumor epithelial<br />

cells, but the SSTR-4 <strong>and</strong> SSTR-5 subtypes<br />

were expressed in the epithelial as well<br />

as in the stromal cells of PC. Thus, there are<br />

differences between normal <strong>and</strong> tumoral samples<br />

in the SSTR expression in the human<br />

prostate epithelial <strong>and</strong> stromal cells in vitro<br />

[57, 59]. Because some SSTR-2-selective SST<br />

analogs (e.g. S<strong>and</strong>ostatin) are probably ineffective<br />

in the treatment of PC, this observation<br />

would suggest that the absence of SSTR-2<br />

could inhibit the growth <strong>and</strong> development of<br />

PC by analogs with a high binding <strong>for</strong> this<br />

receptor subtype. The three SSTRsmRNAs<br />

(1, 2 <strong>and</strong> 3) are expressed simultaneously not<br />

only in tumoral tissue but also in the host peritumoral<br />

vascular system; thus, some prostate<br />

tumor tissues bind octreotide due to a direct<br />

action on tumor cell SSTRs or through action<br />

on peritumoral vessels (due to SSTRs or<br />

growth factors), by altering the hemodynamics<br />

of the tumoral blood circulation [55].<br />

The SSTR can be generally measured [60]<br />

using (1) biochemical techniques, i.e. binding<br />

assays on tissue homogenates in vitro,<br />

(2) autoradiography <strong>for</strong> visualization in tissue<br />

sections, (3) systemic injection of a radio-<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 115


Fig. 2. Mechanisms of antiproliferative actions of SST-14 <strong>and</strong> SST analogs in advanced PC.<br />

chemical-coupled SST analog into normal<br />

subjects, which specifically labels SST target<br />

tissue <strong>and</strong> localizes it by a subsequent in vitro<br />

autoradiography <strong>and</strong> (4) in vivo visualization<br />

of SST receptor-positive tumors, with the injection<br />

of [ 123 I or 131 I]-coupled Tyr3- or 111 In-<br />

[DTPA]-coupled octreotide. Radiolabeled<br />

with 123 I, 131 I or 111 In SST analogs are synthesized<br />

<strong>for</strong> localization of prostate tumor <strong>and</strong> its<br />

metastases containing SST receptors, using<br />

scanning techniques [49]. In 8 of 31 biopsied<br />

patients with metastatic hormone-refractory<br />

PC, SST receptors were expressed both in<br />

vitro <strong>and</strong> in vivo, <strong>and</strong> visualized with the<br />

octreoscan technique [56]. SST analog RC-<br />

160 labeled with 99m Tc was shown to have a<br />

400% higher 24-hour tumor uptake compared<br />

with 111 In-DTPA-octreotide due to specific<br />

tumor binding sites in nude mice bearing<br />

experimental human PCs [61]. In contrast,<br />

111 In-DTPA-D-Phe1-octreotide scintigraphy<br />

identified only 3 of 7 patients with hormoneresistant<br />

PC who had the highest tumor to<br />

background ratio <strong>and</strong> responded to octreotide<br />

116 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

therapy [62]. The specific binding of a ß-emitting<br />

isotope-coupled SST analog to high-affinity<br />

membrane SSTR subtypes offers the opportunity<br />

to give brachytherapy to those tumors,<br />

which have the specific SSTR subtypes.<br />

Mechanisms of Antiproliferative Actions<br />

of SST-14 <strong>and</strong> SST <strong>Analogs</strong> in<br />

Advanced PC<br />

Five possible main antitumor actions of<br />

SST-14 <strong>and</strong> SST analogs have been discussed<br />

(fig. 2, 3). First, they inhibit secretory <strong>and</strong><br />

hormonal effects on tumor growth. Prolactin<br />

acts as a cofactor on the growth of PC, alone<br />

as well as synergistically with GH [31, 32].<br />

GH is of great importance <strong>for</strong> the suppression<br />

of various tumors, among them PC, in<br />

which GFs, such as IGF-1, are involved. Due<br />

to local production of IGF-1, GH directly<br />

promotes cell differentiation as well as indirectly<br />

clonal expansion [31]. Second, somatostatin<br />

<strong>and</strong> its analogs act as antimitogens<br />

synergistically with endogenous GFs (IGF-1


Fig. 3. Molecular basis of the antiproliferative actions of octreotide in advanced PC.<br />

<strong>and</strong> IGF-2, PDGF, FGF, TGF-· <strong>and</strong> TGF-ß),<br />

which are involved in proliferation <strong>and</strong> phenotypic<br />

trans<strong>for</strong>mation, or (directly) inhibit<br />

DNA synthesis <strong>and</strong> proliferation [63, 64].<br />

Third, S<strong>and</strong>ostatin <strong>and</strong> its analogs have direct<br />

antiproliferative actions due to inhibition of<br />

SSTR, reversing the stimulatory effect of EGF<br />

(1) on phosphorylation of the tyrosine kinase<br />

portion of EGF receptors (EGF-R) <strong>and</strong> of<br />

tyrosine acceptor proteins, <strong>and</strong> (2) on cell<br />

growth [28]. EGF promotes phosphorylation<br />

of cell membrane proteins of 170 kD (as EGF-<br />

R) <strong>and</strong> of 60 kD (as does LHRH-R). SST<br />

analogs (RC-160, RC-121) combined with<br />

LHRH analogs can increase the activity of<br />

tyrosine phosphatases, thus nullifying the<br />

EGF effects on the activation of dephosphorylation<br />

of such membrane peptides [65, 66].<br />

Fourth, somatostatin <strong>and</strong> its analogs inhibit<br />

(fig. 3) the production of oncogene products<br />

or the overexpression of protooncogenes,<br />

which act in the tyrosine kinase stimulatory<br />

pathway [61]. Oncogene products are similar<br />

to GFs or are aberrant GF receptors, which<br />

promote tumor cell growth. The protooncogene<br />

c-sis codes <strong>for</strong> the B chain of PDGF,<br />

while the viral oncogene erbB produces aberrant<br />

EGF-R [65]. Also, the EGF-R shares<br />

homology with erbB <strong>and</strong> c-erbB-2/new oncogenes<br />

[63, 67, 68]. Fifth, finally, SST-14 <strong>and</strong><br />

some of its analogs have actions via specific<br />

SSTR on the peritumoral vessels [69], <strong>and</strong> are<br />

able to inhibit (due to FGF-related molecules)<br />

angiogenesis. Thus, another potentially therapeutic<br />

effect of somatostatin, octreotide <strong>and</strong><br />

other analogs in vivo may partially depend on<br />

its suppressive action on tumor angiogenesis<br />

[70].<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 117


In vitro <strong>and</strong> in vivo Studies of the<br />

Antitumor Actions of SST-14 <strong>and</strong><br />

SST <strong>Analogs</strong> in PC:<br />

Animal <strong>and</strong> Human Models<br />

Dunning R-3327H rats (with <strong>and</strong>rogensensitive<br />

PC) were treated with modern SST<br />

analogs (S<strong>and</strong>ostatin, RC-160, RC-121) alone<br />

or combined with D-Trp6-LHRH analog or<br />

castration [71–74]. Final tumor volume, percentage<br />

change from initial tumor <strong>and</strong> final<br />

tumor weight were decreased due to the inhibitory<br />

effect of LHRH analog (or castration),<br />

effects which were potentiated by the<br />

superactive SST analogs, i.e. RC-98-I, RC-<br />

160, RC-121 [71–73]. Schally’s group [50]<br />

demonstrated with the so-called superactive<br />

SST analogs higher <strong>and</strong> more specific binding<br />

to human <strong>and</strong> rat prostatic adenocarcinomas<br />

as well as to human ovarian <strong>and</strong> several human<br />

pancreatic cancers. Thus, different SST<br />

analogs had to be developed <strong>for</strong> diagnostic<br />

<strong>and</strong> clinical use other than octreotide, since<br />

octreotide is mainly effective in hormonehypersecreting<br />

endocrine tumors, or in those<br />

with the specific (<strong>for</strong> octreotide) SSTR-2 subtype.<br />

The combination of [D-Trp6]-LHRH <strong>and</strong><br />

RC-160 in rats or in nude mice bearing xenografts<br />

of the hormone-dependent human<br />

prostate tumor PC-82 has shown a greater<br />

inhibition of tumor growth than [D-Trp6]-<br />

LHRH or RC-160 alone [31, 75]. Sustained<br />

release <strong>for</strong>mulations of SST analogs, such as<br />

Somatulin, also inhibited tumor growth after<br />

castration of male Copenhagen rats bearing<br />

Dunning R-3327H prostate tumors, but<br />

not in combination with LHRH analogs<br />

[73]. Male nude mice with PC-82 tumors,<br />

treated with slow-acting [D-Trp6]-LHRH agonist<br />

or LHRH antagonist (SB-75) or RC-<br />

160, showed greater tumor inhibition after<br />

treatment with SB-75, than with LHRH agonist,<br />

but no significant inhibition with<br />

118 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

RC-160 [76]. Combination therapy with SB-<br />

75 plus RC-160 achieved the best results,<br />

when it was started soon after the diagnosis<br />

of PC [77]. A decrease of IGF-1 <strong>and</strong> GH or<br />

a downregulation of EGF-R by RC-160 may<br />

improve the hormonal treatment of PC [75].<br />

Androgen-independent human PC cell line<br />

(PC-3, DU-45, R3327-AT-1 PC) xenografts<br />

in nude mice treated with RC-160 plus a<br />

GRP antagonist (RC-3095) showed a significant<br />

growth inhibition, when the therapy<br />

was started at an early stage of tumor development<br />

[78–80]. SST-14 inhibited cell proliferation<br />

<strong>and</strong> protein secretion of the relatively<br />

indolent tumor LNCaP, probably mediated<br />

by the activation of phosphotyrosyl<br />

protein phosphatases [81]. Lanreotide, a<br />

slow-acting SST analog, applied topically to<br />

the surgical site on xenografts of human<br />

prostate tumors (PC-3, DU-145, H-1579) in<br />

athymic male rats, inhibited tumor proliferation<br />

[82].<br />

Long-term exposure of cultured tumor<br />

cells (e.g. transplantable rat pituitary tumor<br />

7315b or PRL-secreting cell line 7315c) to<br />

octreotide led to a loss of sensitivity with<br />

respect to its inhibitory effects on tumor cell<br />

growth, as well as to its hormone-inhibitory<br />

effects [60]. These desensitized tumor cells<br />

have lost their SST receptors, which reappeared<br />

after the withdrawal of octreotide.<br />

Thus, the gradual decrease of the growthinhibitory<br />

action of octreotide during the<br />

long-term treatment may be caused by the<br />

desensitization of SSTR, <strong>and</strong> is not due to<br />

selection of receptor-negative cell clones.<br />

50% of all PCs have neuroendocrine (NE)<br />

cell populations with the presence of neurohormonal<br />

peptides (NSE, chromogranin, serotonin,<br />

·-HCG, SST-14, calcitonin, ACTH,<br />

ß-endorphin, <strong>and</strong> others). The incidence of<br />

NE differentiation in PC cells is higher in<br />

fresh than in <strong>for</strong>malin-fixed biopsy specimens<br />

[83]. Certain NE peptides such as bombesin


<strong>and</strong> VIP can increase the invasive potential of<br />

PC cells <strong>and</strong> may contribute to the rapid progression<br />

of PCs containing NE cell populations,<br />

<strong>and</strong> RC-160 did not alter the invasion<br />

of these PC-3 cells [84]. Thus, PCs appear to<br />

be more complex <strong>and</strong> heterogeneous than previously<br />

thought, exhibiting endocrine differentiation.<br />

PC cases with carcinoid tumor <strong>and</strong><br />

hypercalcemia due to PTHrP [85, 86] <strong>and</strong> a<br />

case with Cushing’s syndrome due to CRHproducing<br />

PC [87] showed that these unusual<br />

PC cases with overexpression of NE differentiation<br />

may have major life-threatening effects,<br />

which are not significantly affected by<br />

octreotide treatment [86]. A review of the<br />

world’s literature on this topic suggested that<br />

PC cells with NE differentiation (1) allow<br />

screening <strong>for</strong> PC <strong>and</strong>/or monitoring <strong>for</strong> recurrence<br />

of PC, (2) are resistant to hormone therapy,<br />

<strong>and</strong> thus (3) show poorer prognosis <strong>and</strong><br />

correlate directly with the tumor grade [75].<br />

NE cells have a regulatory role in PC growth,<br />

<strong>and</strong> are prognostically useful in prostatic<br />

high-grade adenocarcinoma [1, 88, 89]. Only<br />

a partial correlation was observed between<br />

NE serum markers <strong>and</strong> immunohistochemical<br />

findings in 22 patients with PC [90]. Only<br />

chromogranin A showed a correlation (but<br />

not SST, NSE, chromogranin B, pancreastatin),<br />

being a useful serum marker in predicting<br />

the extent of NE differentiation in prostate<br />

tumors [91].<br />

Certain PCs, which secrete somatostatin,<br />

are difficult to localize with the in vivo visualization<br />

technique with an isotope-labeled SST<br />

analog, because of a competition at their<br />

SST receptors. Two observations might in<br />

part explain this competition at the receptor<br />

levels with locally produced somatostatin:<br />

(1) the in vitro detection of SSTR could be<br />

improved after repeated additional wash procedures<br />

of the tumor samples, <strong>and</strong> (2) higher<br />

doses of octreotide are necessary in patients<br />

with such SST-producing PCs (with NE dif-<br />

ferentiation) in order to suppress tumor<br />

growth [60].<br />

In summary, many experimental findings<br />

suggest that both indirect <strong>and</strong> direct effects<br />

may play a role in the tumor growth-inhibitory<br />

action of SST analogs. In particular, the<br />

early tumor growth-inhibitory effects of SST<br />

analogs are probably due to the inhibition of<br />

the tumor angiogenesis. Finally, the decrease<br />

of the tumor growth-inhibitory action over<br />

time is mainly due to the desensitization <strong>and</strong><br />

downregulation of the tumor SSTR.<br />

Clinical Trials of the <strong>Treatment</strong> of<br />

Advanced Hormone-Refractory PC<br />

Introduction<br />

PC is often (60–80%) locally advanced or<br />

metastatic at diagnosis, making surgical removal<br />

<strong>and</strong> radiotherapy ineffective. Metastatic<br />

PC responds to normal hormonal manipulations,<br />

but once progression occurs new<br />

treatment modalities are required; at that<br />

time, specific <strong>and</strong> systemic antitumor therapy<br />

is better than local treatment, as ER. Alternative<br />

hormonal therapy involves <strong>and</strong>rogen deprivation<br />

with surgical or chemical castration<br />

estrogens, anti<strong>and</strong>rogens, inhibitors of <strong>and</strong>rogen<br />

metabolism or adrenal suppression. Complete<br />

<strong>and</strong>rogen blockade may be a more superior<br />

hormonal treatment. PCs have initially<br />

<strong>and</strong>rogen-dependent or <strong>and</strong>rogen-sensitive<br />

<strong>and</strong> <strong>and</strong>rogen-independent cell populations.<br />

Thus, in the later stages of PC a significant<br />

amount of tumor is <strong>and</strong>rogen-independent,<br />

<strong>and</strong> needs alternative antiproliferative treatment.<br />

Additionally, advanced PCs have been<br />

noted to represent chemotherapeutically relatively<br />

nonresponsive tumors, with a median<br />

response rate of only 8.7% among 26 new<br />

drug trials during 1987–1991 [92]. There<strong>for</strong>e,<br />

because of cytotoxic drug resistance, there is a<br />

need to develop new methods to overcome<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 119


such resistance with new classes of antimitogen<br />

agents. Many SST analogs have been<br />

found to inhibit the growth of animal <strong>and</strong><br />

human PC cells.<br />

Controlled Clinical Trials<br />

Sporadic clinical studies in patients with<br />

stage D2 PC (relapsed during treatment with<br />

flutamide plus castration) treated with SST<br />

analogs alone or combined with bromocriptine<br />

showed some advantage in tumor growth<br />

inhibition [93, 94]. Schally et al. [95] reviewed<br />

the hormone treatment of advanced<br />

PC with agonistic <strong>and</strong> antagonistic LHRH<br />

analogs or LHRH analogs linked to cytotoxic<br />

radicals, with very active SST analogs (SB-75,<br />

RC-160) or with bombesin/GRP antagonists.<br />

Thus, they tried to delay or prevent the relapse,<br />

<strong>and</strong> to improve the therapy of PC. Parmar<br />

et al. [96] treated 16 of 25 poor-risk<br />

patients with hormone refractory metastatic<br />

PC (failing to respond to total <strong>and</strong>rogen<br />

blockade) with an SST analog BIM-23014. 2<br />

of these 16 patients showed partial response<br />

<strong>and</strong> 3 no change of the disease, with only a<br />

few side effects, such as mild diarrhea <strong>and</strong><br />

abdominal cramps. In contrast, S<strong>and</strong>ostatin<br />

treatment of 24 patients with hormone-resistant<br />

PC with a dose of 100 Ìg ! 3 times/day<br />

<strong>for</strong> 6 weeks did not give any objective evidence<br />

of tumor regression, but only of tumor<br />

progression [97]. The author concluded that<br />

S<strong>and</strong>ostatin eventually stimulates PC growth,<br />

but may have sensitized tumor cells to subsequent<br />

chemotherapy, because salvage chemotherapy<br />

resulted in an objective tumor regression<br />

in 5 of 6 patients treated after progression.<br />

In a small group of 5 PC patients relapsing<br />

during hormonal treatment with rapidly<br />

increasing PSA levels, octreotide administered<br />

as a subcutaneous infusion, at a dose of<br />

400–1,000 Ìg/day <strong>for</strong> a period of 2–6 weeks<br />

offered only a moderate <strong>and</strong> temporary inhibition<br />

of PC growth [97, 98]. The author con-<br />

120 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

Table 4. Percentage <strong>and</strong> quality of response in 8<br />

responding stage D2 PC patients after the combined<br />

treatment of CAB plus octreotide<br />

Previous<br />

treatment<br />

Total<br />

patients<br />

Number<br />

responding<br />

Quality<br />

of response<br />

No 6 3 (50) 3 (oPR + sCR)<br />

Yes 8 3 (37.5)<br />

CAB 6 1 (16.5) 1 (oNC + sPR)<br />

iHT 2 2 (100) 1 (oPR + sPR)<br />

Total 14 6 (33)<br />

Figures in parentheses represent percentage. iHT =<br />

Incomplete hormone treatment; CR = complete response;<br />

PR = partial response; NC = no change; o =<br />

objective; s = subjective.<br />

cluded that more specific SST analogs may<br />

give better results possibly in higher doses,<br />

alone or in combination with LHRH agonists<br />

or antagonists. Slow-acting SST analog (Somatulin)<br />

in continuous intravenous infusion<br />

<strong>and</strong> in a dose escalation trial, administered to<br />

patients with advanced metastatic hormonerefractory<br />

PC, was well tolerated but no clinical<br />

responses were noted, even with the highest<br />

doses of 24 mg/day [99]. Somatulin treatment<br />

should be evaluated in less advanced<br />

PCs, or in combination with other antiproliferative<br />

agents.<br />

In the Theagenion <strong>Cancer</strong> Center my collaborators<br />

<strong>and</strong> I evaluated (in total) 18 carefully<br />

selected patients with advanced stage D2<br />

PC <strong>and</strong> treated 14 of them with a combined<br />

treatment of CAB (castration or triptorelin<br />

plus flutamide) plus octreotide (S<strong>and</strong>ostatin),<br />

in doses of 0.2 mg ! 2 times/day, s.c. <strong>for</strong> at<br />

least 12 months [24]. 3 of 6 patients without<br />

previous hormone treatment (50%) showed<br />

the best response (objective partial plus subjective<br />

complete response). Among the 8 patients<br />

previously treated with hormone only 3


Table 5. Natural history <strong>and</strong> follow-up of all 18 stage D2 PC patients after the treatment with CAB plus octreotide<br />

or CAB alone<br />

n Age<br />

years<br />

Histology<br />

Diff. Inv.<br />

Duration<br />

of R<br />

months<br />

Survival, months<br />

DFS total<br />

Deaths Cause of death<br />

AHA HF Dis.<br />

CAB<br />

4<br />

R 2 79 2 (high) 012 12 18 1 – – 1<br />

NR 2 71 2 (low) 0– 02 2 1 1 –<br />

Octreotide + CAB 14<br />

R 6 68 5 (low) 1 17 17 18.5 5 1 – 4<br />

NR 8 68 8 (low) 3 – – 4.5 8 2 2 4<br />

R = Response; NR = no response; Diff. = differentiation; Inv. = invasive; DFS = disease-free survival;<br />

n = number; AHA = acute heart attack; HF = heart failure; Dis. = disease.<br />

(37.5%) responded poorly to this treatment<br />

regimen (2 with no change <strong>and</strong> 1 with a partial<br />

objective response plus a partial subjective<br />

response) (table 4). The total response rate<br />

was 33% with a mean disease-free survival of<br />

17 months compared to 12 months when<br />

CAB was given alone (table 5). Total mean<br />

survival was similar <strong>for</strong> the two therapy<br />

groups. IGF-1 <strong>and</strong> EGF serum levels during<br />

the combined treatment (with the addition of<br />

S<strong>and</strong>ostatin) decreased significantly in all responding<br />

patients, <strong>and</strong> corresponded to the<br />

drop in PSA levels, or eventually in prostatic<br />

acid phosphatase serum levels. These observations<br />

suggest that PSA <strong>and</strong> prostatic acid<br />

phosphatase levels as well as IGF-1 <strong>and</strong> EGF<br />

serum levels may be useful as a marker of SST<br />

analog therapy in advanced PC.<br />

Octreotide has a consistent <strong>and</strong> persistent<br />

suppressive effect on hormone release (<strong>for</strong><br />

example in acromegaly), without the occurrence<br />

of therapy escapes. However, there are<br />

some escapes from the antineoplastic therapy<br />

of tumors. Thus, an escape from therapy occurs<br />

through the development of SSTR-negative<br />

tumor cell clones, or due to an early<br />

insensitivity of hormonal secretion, related to<br />

downregulation of the SST receptors on these<br />

tumors. This downregulation can be restored<br />

after interruption of octreotide treatment. Additionally,<br />

a continuing sensitivity of hormone<br />

secretion to octreotide with a simultaneously<br />

progressive tumor growth may need<br />

an increase in the dose of octreotide, particularly<br />

if the somatostatin scintigraphy in vivo<br />

is positive [60]. IGF-1 serum levels do probably<br />

not remain consistently lowered during<br />

long-term octreotide treatment of patients<br />

with PC, because its suppressive effect is<br />

mostly transient; additionally, one would take<br />

into consideration that IGF-1 serum levels in<br />

cancer patients tend to decrease during tumor<br />

progression.<br />

Criticism of the Design of Clinical Trials<br />

Clinical trials with SST analog treatment<br />

of patients with metastatic hormone-refractory<br />

PC are based on multiple in vivo <strong>and</strong> in<br />

vitro studies in animal models <strong>and</strong> in patients<br />

with human PC. Un<strong>for</strong>tunately, at present the<br />

studies have major disadvantages, such as<br />

inclusion of only a small number of patients,<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 121


inadequate trial design, in particular patient<br />

r<strong>and</strong>omization in groups according to hormone<br />

resistance, <strong>and</strong> finally, the use of widely<br />

varying dosage regimens <strong>and</strong> routes of administration.<br />

Additionally, none of the trials hitherto<br />

reported has been based on SSTR determination<br />

be<strong>for</strong>e starting the SST analog therapy.<br />

In our clinical study octreotide was combined<br />

with CAB from the start [24], while in<br />

most other trials only the LHRH analog was<br />

added. We believe such a clinical trial should<br />

be planned at specific medical centers with<br />

extensive experience in this field.<br />

In summary, in planning clinical trials<br />

with SST analogs in patients with metastatic<br />

or advanced PC, it has to be decided<br />

(1) whether SSTR (based on scintigraphic results)<br />

should be included, (2) whether IGF-1<br />

serum levels should be regularly measured<br />

during therapy, <strong>and</strong> (3) whether the dose of<br />

the SST analog used will be constant, increased,<br />

or even decreased according to the<br />

results of the in vivo scintigraphy measurements.<br />

Previous <strong>and</strong> recent clinical trials with<br />

octreotide in patients with PC leave uncertainty<br />

with regard to the optimal daily dose,<br />

the optimal route of administration, or the<br />

optimal duration of treatment.<br />

Conclusions<br />

<strong>Somatostatin</strong> seems to be an endogenous<br />

inhibitory GF in several organ systems, while<br />

exogenous SST analogs also have inhibitory<br />

effects on the growth of a variety of tumors;<br />

thus, one could speculate that the expression<br />

of SSTR on several PC cells might represent<br />

a general inhibitory control mechanism<br />

through which well-differentiated PC is inhibited<br />

in its growth. Thus SST analogs may<br />

have their place among the therapeutic<br />

antimitogen maneuvers in patients with metastatic<br />

<strong>and</strong> almost hormone-refractory PC.<br />

122 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

Radiolabeled SST analogs with 123 I, 131 I or<br />

111 In may detect the specific binding sites of<br />

these drugs, thus predicting their antiproliferative<br />

actions. PC could be partially visualized<br />

in vivo via an isotope-coupled SST analog<br />

<strong>and</strong> could be chronically treated with specific<br />

SST analogs according to a positive<br />

scan. Un<strong>for</strong>tunately, the value of the localization<br />

procedure with SSTR scintigraphy<br />

was lower in PCs than in other tumors. The<br />

detection of more than 2 SSTR subpopulations,<br />

of which 5 have been cloned, increases<br />

the possibility of using different SST analogs<br />

as selective antitumor agents <strong>for</strong> the treatment<br />

of disseminated hormone-resistant PC.<br />

Immunohistochemical NE differentiation of<br />

PC cells <strong>and</strong> serum NE hormonal markers<br />

may also predict the prognosis of such tumors<br />

or the usefulness of SST analog treatment<br />

in advanced PC.<br />

Projections <strong>for</strong> the Future<br />

Multicenter, prospective, clinical, controlled<br />

trials are needed with a great number<br />

of patients r<strong>and</strong>omized according to their status:<br />

with or without previous incomplete hormone<br />

treatment or refractory to hormone. Patients<br />

should be r<strong>and</strong>omized to: (1) CAB only,<br />

(2) CAB plus appropriate SST analog, (3) SST<br />

analog only, (4) SST analog with systemic<br />

cytotoxic therapy, (5) SST analog with cytotoxic<br />

radicals, (6) SST analog with LHRH<br />

agonists/antagonists <strong>and</strong> finally (7) SST analog<br />

with GRP/bombesin-antagonists.<br />

Specific antimitogen-acting SST analogs<br />

should be controlled on the basis of SSTR<br />

subtype determination, thus predicting the<br />

specific analogs with the highest binding capacity<br />

<strong>and</strong> having the best antiproliferative<br />

actions on hormone-resistant PCs. New specific<br />

radiolabeled (with 123 I, 131 I, 111 In) SST<br />

analogs, using scanning techniques, could


demonstrate SSTR-positive PCs. Thus, they<br />

might be used to monitor the efficacy of therapy,<br />

<strong>and</strong> with the coupling of ß-emitting radionuclide<br />

should make a kind of brachytherapy<br />

possible. The development of slow-release<br />

SST analog preparations (subcutaneous, intramuscular<br />

or intranasal) may increase patient<br />

compliance with this long-term antitumor<br />

treatment.<br />

References<br />

1 Bostwick DG, Myers RP, Oesterling<br />

JE: Staging of prostate cancer. Semin<br />

Surg Oncol 1994;10:60–72.<br />

2 Boring CC, Squires TS, Tong T:<br />

<strong>Cancer</strong> statistics, 1991. CA <strong>Cancer</strong> J<br />

Clin 1991;41:19–36.<br />

3 Labrie F, Dupont A, Belanger A:<br />

Complete <strong>and</strong>rogen blockade <strong>for</strong><br />

treatment of prostate cancer; in Important<br />

Advances in Oncology. Philadelphia,<br />

Lippincott, 1985, pp 193–<br />

217.<br />

4 Paulson DF: The natural history of<br />

prostate cancer. Adv Oncol 1988;4:<br />

10–17.<br />

5 Koslowski JM, Grayhack JT: Carcinoma<br />

of the prostate; in Gillenwater<br />

JY, Grayhack JT, Howard SS, et al.<br />

(eds): Adult <strong>and</strong> Pediatric Urology,<br />

ed 2. Chicago, Year Book Medical<br />

Publishers, 1991.<br />

6 Grayhack JT, Kozlowski JM: Endocrine<br />

therapy in the management of<br />

advanced prostate cancer. The case<br />

<strong>for</strong> early initiation of treatment.<br />

Urol Clin North Am 1980;7:639–<br />

642.<br />

7 Grayhack JT, Keeler IC, Kozlowski<br />

JM: Carcinoma of the prostate: Hormonal<br />

therapy. <strong>Cancer</strong> 1987;60<br />

(suppl 3):589.<br />

8 Schroeder FH: Early versus delayed<br />

endocrine treatment in metastatic<br />

prostatic cancer; in Murphy GP,<br />

Khoury S (eds): Therapeutic Progress<br />

in Urological <strong>Cancer</strong>s. Proceedings<br />

of an International Symposium<br />

held in Paris, France, June 29–<br />

July 1, 1988. New York, Liss, 1989,<br />

p 253.<br />

9 Zincke H: Extended experience with<br />

surgical treatment of stage D1 adenocarcinoma<br />

of prostate: Significant<br />

influences of immediate adjuvant<br />

hormonal treatment (orchidectomy)<br />

on outcome. Urology 1989;23(suppl<br />

5):27.<br />

10Huggins C, Hodges CV: Studies on<br />

prostate cancer. I. The effect of castration<br />

of estrogen <strong>and</strong> of <strong>and</strong>rogen<br />

injection on serum phosphatases in<br />

metastatic carcinoma of the prostate.<br />

<strong>Cancer</strong> Res 1941;1:293–297.<br />

11 Vainas I: Prostate cancer: Current<br />

strategy of treatment; in Boutis L,<br />

Dimitriadis K (ed): Advances in<br />

Medical Oncology. Proceedings of<br />

the Seminar of European School of<br />

Oncology, Thessaloniki, 1995, pp<br />

220–231.<br />

12 Meyer FJ, Craw<strong>for</strong>d ED: The role of<br />

endocrine therapy in the management<br />

of local <strong>and</strong> distant recurrence<br />

of prostate cancer following radical<br />

prostatectomy or radiation therapy.<br />

Urol Clin North Am 1994;4:707–<br />

715.<br />

13 Walsh PC, Madden JD, Harrod MJ:<br />

Familial incomplete male pseudohermaphroditism,<br />

type 2: Decreased<br />

dihydrotestosterone <strong>for</strong>mation<br />

in pseudovaginal perineoscrotal<br />

hypospadias. N Engl J Med 1974;<br />

291:944–949.<br />

14 Catalona WJ: Radical surgery <strong>for</strong><br />

advanced prostate cancer <strong>and</strong> <strong>for</strong> radiation<br />

failure. J Urol 1992;147:<br />

916–917.<br />

Acknowledgments<br />

I am indebted to Dr. Frances Gillepsie, consultant<br />

anesthesiologist, <strong>for</strong> her skillful ef<strong>for</strong>ts in correcting the<br />

manuscript.<br />

15 Santen RJ: Endocrine aspects of<br />

prostate cancer; in Becker KL (ed):<br />

Principles <strong>and</strong> Practice of Endocrinology<br />

<strong>and</strong> Metabolism. Philadelphia,<br />

Lippincott, 1990, pp 1656–<br />

1663.<br />

16 Labrie F, Dupont A, Belanger A:<br />

New approach in the treatment of<br />

cancer: Complete instead of only<br />

partial removal of <strong>and</strong>rogens. Prostate<br />

1983;4:579–594.<br />

17 Rohner TJ Jr: Choosing hormonal<br />

therapy treatment <strong>for</strong> patients with<br />

metastatic prostatic cancer; in Wise<br />

HA, Klein LA (eds): Practical Points<br />

in Urologic Practice. Philadelphia,<br />

Lea & Febiger, 1990, pp 1–4.<br />

18 Isaacs JT, Coffey DS: Adaptation<br />

versus selection as the mechanism<br />

responsible <strong>for</strong> the relapse of prostatic<br />

cancer to <strong>and</strong>rogen ablation<br />

therapy as studied in the Dunning<br />

R-3327-H adenocarcinoma. <strong>Cancer</strong><br />

Res 1981;41:5070–5074.<br />

19 Murphy GP, Beckley S, Brady MF:<br />

<strong>Treatment</strong> of newly diagnosed prostate<br />

cancer patients with chemotherapy<br />

agents in combination with hormones<br />

versus hormones alone. <strong>Cancer</strong><br />

1983;51:1264–1266.<br />

20Suarez AJ, Lamm DL, Radwin HM:<br />

Androgen priming <strong>and</strong> cytotoxic<br />

therapy in advanced prostatic cancer.<br />

<strong>Cancer</strong> Chemother Pharmacol<br />

1982 ;8:261–269.<br />

21 Linehan WM, La Rossa R, Stein C:<br />

Use of suramin in treatment of patients<br />

with advanced prostate carcinoma<br />

(abstract 131). J Urol 1990;<br />

143/2:221A.<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 123


22 Fong CJ, Kozlowski JM, Lee C: Effects<br />

of retinoid acid on the proliferation<br />

<strong>and</strong> differentiation of <strong>and</strong>rogen<br />

responsive prostatic cancer cell<br />

lines. LNCaP (abstract 753). J Urol<br />

1991;145/2:401A.<br />

23 Motta MM, Moretti RM, Marelli<br />

MM: LHRH <strong>and</strong> somatostatin: Examples<br />

of peptidergic control of<br />

prostate cancer growth; in Jonat W,<br />

Kaufmann M, Munk K (eds): Hormone-Dependent<br />

Tumors. Basel,<br />

Karger, 1995, pp 332–344.<br />

24 Vainas I, Pasaitou K, Galaktidou G,<br />

Maris K, Christodoulou K, Const<strong>and</strong>inidis<br />

C, Kortsaris AH: The<br />

role of somatostatin analogues in<br />

complete anti<strong>and</strong>rogen treatment. J<br />

Exp Clin <strong>Cancer</strong> Res 1997;16/1:<br />

119–126.<br />

25 Schally AV, Coy DH, Meyers CA:<br />

Hypothalamic regulatory hormones.<br />

Annu Rev Biochem 1978;47:89–<br />

128.<br />

26 Reichlin D: <strong>Somatostatin</strong>. N Engl<br />

J Med 1983;309:1495–1501, 1556–<br />

1563.<br />

27 Scarpignato C: Octreotide, the synthetic<br />

long-acting somatostatin analogue:<br />

Pharmacological profile; in<br />

Scarpignato C (ed): Octreotide:<br />

From Basic Science to Clinical Medicine.<br />

Basel, Karger, 1996, vol 10, pp<br />

54–72.<br />

28 Mascardo RN, Sherline P: <strong>Somatostatin</strong><br />

inhibits rapid centrosomal<br />

separation <strong>and</strong> cell proliferation induced<br />

by epidermal growth factor.<br />

Endocrinology 1982;111:1394–<br />

1396.<br />

29 Schally AV, Comaru-Schally AM,<br />

Redding T: Antitumor effects of<br />

analogs of hypothalamic hormones<br />

in endocrine-dependent cancers.<br />

Proc Soc Exp Biol Med 1984;175:<br />

259–281.<br />

30Schally AV, Cai R-Z, Torres-Alleman<br />

I: Endocrine, gastrointestinal<br />

<strong>and</strong> antitumor activity of somatostatin<br />

analogs; in Moody TW (ed):<br />

Neural <strong>and</strong> Endocrine Peptides <strong>and</strong><br />

Receptors. New York, Plenum<br />

Press, 1986, pp 73–88.<br />

31 Schally AV: Oncological application<br />

of somatostatin analogs. <strong>Cancer</strong> Res<br />

1988;48:6977–6985.<br />

32 Schally AV, Redding TW, Paz-Boula<br />

JI, Comaru-Schally AM, Mathe<br />

G: Current concept <strong>for</strong> improving<br />

treatment of prostate cancer based<br />

on combination of LHRH agonists<br />

with other agents; in Murphy GP,<br />

Khoury S, Kuss R, Chatelein D, Denis<br />

L (eds): Prostate <strong>Cancer</strong>. Part A.<br />

Research, Endocrine <strong>Treatment</strong> <strong>and</strong><br />

Histopathology. New York, Liss,<br />

1987, pp 173–197.<br />

33 Veber DF, Freidinger RM,<br />

Schwenk-Perlow D: A potent cyclic<br />

hexapeptide analog of somatostatin.<br />

Nature 1981;292:55–58.<br />

34 Bauer W, Briner U, Boepfner W:<br />

SMS-201-995: A very potent <strong>and</strong> selective<br />

octapeptide analogue of somatostatin<br />

with prolonged action.<br />

Life Sci 1982;31:1133–1140.<br />

35 Yamada Y, Post SR, Wang K, Tager<br />

HS, Bell GI, Seino S: Cloning <strong>and</strong><br />

functional characterization of a family<br />

of human <strong>and</strong> mouse somatostatin<br />

receptors expressed in brain,<br />

gastrointestinal tract, <strong>and</strong> kidney.<br />

Proc Natl Acad Sci USA 1992;89:<br />

251–255.<br />

36 Yamada Y, Reisine T, Law SF, Ihara<br />

Y, Kubota A, Kagimoto S, Seino<br />

M, Seino Y, Bell GI, Seino S: <strong>Somatostatin</strong><br />

receptors, an exp<strong>and</strong>ing<br />

gene family: Cloning <strong>and</strong> functional<br />

characterization of human SSTR3, a<br />

protein coupled to adenylcyclase.<br />

Mol Endocrinol 1992;6:2136–2142.<br />

37 Rohrer L, Raulf F, Bruns C, Buettner<br />

R, Hofstaedter F, Schuele R:<br />

Cloning <strong>and</strong> characterization of a<br />

fourth human somatostatin receptor.<br />

Proc Natl Acad Sci USA 1993;<br />

90:4196–4200.<br />

38 Yamada Y, Kagimoto S, Kubota A,<br />

Yasuda K, Masuda K, Someya Y,<br />

Ihara Y, Li Q, Imura H, Seino S, Seino<br />

Y: Cloning, functional expression<br />

<strong>and</strong> pharmacological characterization<br />

of a fourth (hSSTR4) <strong>and</strong> a<br />

fifth (hSSTR) human somatostatin<br />

receptor subtype. Biochem Biophys<br />

Res Commun 1993;195:844–852.<br />

39 O’Caroll AM, Raynor K, Lolait<br />

SJ, Reisine T: Characterization of<br />

cloned human somatostatin receptor<br />

SSTR5. Mol Pharmacol 1994;<br />

46:291–298.<br />

124 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

40Lewin MJ-M, LeRomance M: <strong>Somatostatin</strong><br />

receptors; in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine. Prog<br />

Basic Clin Pharmacol. Basel, Karger,<br />

1996, vol 10, pp 23–24.<br />

41 Cai RZ, Szoke B, Lu E, Fu D, Redding<br />

TW, Schally AV: Synthesis <strong>and</strong><br />

biological activity of highly potent<br />

octapetide analogs of somatostatin.<br />

Proc Natl Acad Sci USA 1986;83:<br />

1896–1900.<br />

42 Cai RZ, Karashima T, Guoth J,<br />

Szoke B, Olsen D, Shally AV: Superactive<br />

octapeptide somatostatin<br />

analogs containing tryptophan residue<br />

in position 1. Proc Natl Acad<br />

Sci USA 1987;84:2502–2506.<br />

43 Scarpignato C, Camboni MG: Safety<br />

profile of octreotide; in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine.<br />

Prog Basic Clin Pharmacol. Basel,<br />

Karger, 1996, vol 10, pp 296–309.<br />

44 Fredenrich A, Sosset C, Beranard<br />

JL, Sadoul J-L, Freychet P: Acute<br />

pancreatitis after short term octreotide.<br />

Lancet 1991;338:52–53.<br />

45 Sadoul J-L, Benchimol D, Thyss A,<br />

Freychet P: Side-effects of octreotide<br />

withdrawal. Lancet 1992;339:<br />

376.<br />

46 Vidal J, Sacanella E, Munoz E, Miro<br />

JM, Navarro S: Acute pancreatitis<br />

related to octreotide in a patient<br />

with acquired immunodeficiency<br />

syndrome. Pancreas 1994;9:395–<br />

397.<br />

47 Reubi JC, Maurer R, von Werder K,<br />

Torhorst J, Klijn JGM, Lamberts<br />

SWJ: <strong>Somatostatin</strong> receptors in human<br />

endocrine tumors. <strong>Cancer</strong> Res<br />

1987;47:551–558.<br />

48 Reubi JC, Maurer R, Klijn JGM:<br />

High incidence of somatostatin receptors<br />

in human meningiomas:<br />

Biochemical characterization. J Clin<br />

Endocrinol Metab 1986;63:433–<br />

438.<br />

49 Lamberts SWJ, Hofl<strong>and</strong> LJ, van<br />

Koetsveld PM: Parallel in vivo <strong>and</strong><br />

in vitro detection of functional somatostatin<br />

receptors in human endocrine<br />

pancreatic tumors: Consequences<br />

with regard to diagnosis, localization<br />

<strong>and</strong> therapy. J Clin Endocrinol<br />

Metab 1990;71:566–574.


50Skarlovic G, Cai RZ, Schally AV:<br />

Evaluation of receptors <strong>for</strong> somatostatin<br />

in various tumors using different<br />

analogs. J Clin Endocr Metab<br />

1990;70:661–669.<br />

51 Radulovic S, Cai R-Z, Milovanovic<br />

S, Schally AV: The binding of bombesin<br />

<strong>and</strong> somatostatin <strong>and</strong> their<br />

analogs to human colon cancers.<br />

Proc Soc Exp Biol Med 1992;200:<br />

394–401.<br />

52 Kadar T, Redding TW, Ben-Danid<br />

M, Schally AV: Receptors <strong>for</strong> prolactin,<br />

somatostatin <strong>and</strong> luteinizing<br />

hormone-releasing hormone in experimental<br />

prostate cancer after<br />

treatment with analogs of luteinizing<br />

hormone-releasing hormone <strong>and</strong><br />

somatostatin. Proc Natl Acad Sci<br />

USA 1988;85:890–894.<br />

53 Fekete M, Redding TW, Comaru-<br />

Schally AM, Pontes JE, Connelly<br />

RW, Srkalovic G, Schally AV: Receptors<br />

<strong>for</strong> luteinizing hormone-releasing<br />

hormone, somatostatin, prolactin<br />

<strong>and</strong> epidermal growth factor<br />

in rat <strong>and</strong> human prostate cancers<br />

<strong>and</strong> in benign prostate hyperplasia.<br />

Prostate 1989;14:191–208.<br />

54 Reubi JC, Waser B, Schaer JC,<br />

Markwalder R: <strong>Somatostatin</strong> receptors<br />

in human prostate <strong>and</strong> prostate<br />

cancer. J Clin Endocrinol Metab<br />

1995;80:2806–2814.<br />

55 Reubi JC, Schaer JC, Laissue JA,<br />

Waser B: <strong>Somatostatin</strong> receptors<br />

<strong>and</strong> their subtypes in human tumors<br />

<strong>and</strong> in peritumoral vessels. Metabolism<br />

1996;45/8(suppl 1):39–41.<br />

56 Nilsson S, Reubi JC, Kalkner KM,<br />

Laissure JA, Horisberger U, Olerud<br />

C, Westlin JE: Metastatic hormone<br />

refractory prostatic adenocarcinoma<br />

expresses somatostain receptors<br />

<strong>and</strong> is visualized in vivo by [ 111 In]labeled<br />

DTPA-D-[Phe-]-octreotide<br />

scintigraphy. <strong>Cancer</strong> Res 1995;<br />

55(suppl 23):5805s–5810s.<br />

57 Tatoud R, Degeorges A, Prevost G,<br />

Hoepffner JL, Gauville C, Millot G,<br />

Thomas F, Calvo F: <strong>Somatostatin</strong><br />

receptors in prostate tissues <strong>and</strong> derived<br />

cell cultures, <strong>and</strong> the in vitro<br />

growth inhibitory effect of BIM-<br />

23014 analog. Mol Cell Endocrinol<br />

1995;113/2:195–204.<br />

58 Prevost G, Benamouzig R, Veber N,<br />

Fajac A, Tatoud R, Defeorges A,<br />

Eden P: The somatostatin receptor<br />

subtype 2 is expressed in normal<br />

<strong>and</strong> tumoral human tissues 70. <strong>Cancer</strong><br />

Detect Prev 1997;21/1:62–70.<br />

59 Sinisi AA, Bellastella A, Prezioso D,<br />

Nicchio MR, Lotti T, Salvatore M,<br />

Pasquali D: Different expression<br />

patterns of somatostatin receptor<br />

subtypes in cultured epithelial cells<br />

from human normal prostate cancer.<br />

J Clin Endocrinol Metab 1997;<br />

82:2566–2569.<br />

60Lamberts SWJ, Krenning EP, Reubi<br />

J-C: The role of somatostatin <strong>and</strong> its<br />

analogs in the diagnosis <strong>and</strong> treatment<br />

of tumors. Endocr Rev 1991;<br />

12:450–482.<br />

61 Thakur ML, Kolan H, Li J, Wiaderkiewicz<br />

R, Parallela VR, Daggaraju<br />

R, Schally AV: Radiolabelled somatostatin<br />

analogs in prostate cancer.<br />

Nucl Med Biol 1997;24/1:105–<br />

113.<br />

62 Kalkner KM, Nisslon S, Westlin JE:<br />

[ 111 In-DTPA-D-Phe 1]-octreotide<br />

scintigraphy in patients with hormone-refractory<br />

prostatic adenocarcinoma<br />

can predict therapy outcome<br />

with octreotide treatment: A<br />

pilot study. Anticancer Res 1998;<br />

18/1B:513–516.<br />

63 Heldin CH, Westermark B: Growth<br />

factors: Mechanism of action <strong>and</strong> relation<br />

to oncogenes. Cell 1984;37:9–<br />

20.<br />

64 Todaro GJ, Fryling C, DeLarco JE:<br />

Trans<strong>for</strong>ming growth factors produced<br />

by certain human tumor cells:<br />

Polypeptides that interact with epidermal<br />

growth factor receptors. Proc<br />

Natl Acad Sci USA 1980;77:5258–<br />

5262.<br />

65 Liebow C, Lee MT, Schally A: Antitumor<br />

effects of somatostatin mediated<br />

by the stimulation of tyrosine<br />

phosphatase. Metabolism 1990;39:<br />

163–166.<br />

66 Lee MT, Liebow C, Kamer AR,<br />

Schally AV: Effects of epidermal<br />

growth factor <strong>and</strong> analogs of luteinizing<br />

hormone-releasing hormone<br />

<strong>and</strong> somatostatin on phosphorylation<br />

<strong>and</strong> dephosphorylation of tyrosine<br />

residues of specific protein substrates<br />

in various tumors. Proc Natl<br />

Acad Sci USA 1991;88:1656–1660.<br />

67 Downward J, Yarden Y, Maye<br />

E: Close similarity or epidermal<br />

growth factor receptor <strong>and</strong> v-erb-B<br />

oncogene protein sequences. Nature<br />

1984;307:521–527.<br />

68 Doolittle RF, Hunkapillar MW,<br />

Hood LE: Simian sarcoma virus oncogene,<br />

v-sis, is derived from the<br />

gene (or genes) encoding a plateletderived<br />

growth factor. Science 1983;<br />

221:275–277.<br />

69 Reubi JC: Octreotide <strong>and</strong> neuroendocrine<br />

tumors: Basic knowledge<br />

<strong>and</strong> therapeutic potential; in Scarpignato<br />

C (ed): Octreotide: From<br />

Basic Science to Clinical Medicine.<br />

Prog Basic Clin Pharmacol. Basel,<br />

Karger, 1996, vol 10, pp 246–269.<br />

70Danesi R, DelTacca M: Effects of<br />

octreotide on angiogenesis; in Scarpignato<br />

C (ed): Octreotide: From<br />

Basic Science to Clinical Medicine.<br />

Prog Basic Clin Pharmacol. Basel,<br />

Karger, 1996, vol 10, pp 234–245.<br />

71 Murphy WA, Lance VA, Moreau S,<br />

Moreau JP, Coy DH: Inhibition of<br />

rat prostate tumor growth by an octapeptide<br />

analog of somatostatin.<br />

Life Sci 1987;40:2515–2522.<br />

72 Schally AV, Redding TW: <strong>Somatostatin</strong><br />

analogs as adjuncts to agonists<br />

of luteinizing hormone-releasing<br />

hormone in treatment of experimental<br />

prostate cancer. Proc Natl Acad<br />

Sci USA 1987;84:7275–7279.<br />

73 Siegel RA, Tolcsvai L, Rudin M:<br />

Partial inhibition of the growth of<br />

transplanted Dunning rat prostate<br />

tumors with the long-acting somatostatin<br />

analogue s<strong>and</strong>ostatin<br />

(SMS 201-955). <strong>Cancer</strong> Res 1988;<br />

48:4651–4655.<br />

74 Zalatnai A, Paz-Bouza JI, Reddding<br />

TW, Schally AV: Histologic changes<br />

in the rat prostate cancer model after<br />

treatment with somatostatin analogue<br />

<strong>and</strong> D-Trp-6-LHRH. Prostate<br />

1988;12/1:85–98.<br />

75 Milovanovic SR, Radulovic S,<br />

Groot K, Schally AV: Inhibition of<br />

growth of PC-82 human prostate<br />

cancer line xenografts in nude mice<br />

by bombesin antagonist RC-3095 or<br />

in combination of agonist [D-Trp6]luteinizing<br />

hormone-releasing hormone<br />

<strong>and</strong> somatostatin analog RC-<br />

160. Prostate 1992;20/4:269–280.<br />

Octreotide <strong>and</strong> Prostatic Carcinoma Chemotherapy 2001;47(suppl 2):109–126 125


76 Redding TW, Schally AV, Radulovics<br />

S, Milovanovic SF, Szepeshazik<br />

K, Isaacs JT: Sustained release <strong>for</strong>mulations<br />

of luteinizing hormonereleasing<br />

hormone antagonist, SB-<br />

75, inhibit proliferation <strong>and</strong> enhance<br />

apoptotic cell death of human<br />

prostate carcinoma (PC-82) in male<br />

nude mice. <strong>Cancer</strong> Res 1992;52:<br />

2538–2544.<br />

77 Yano T, Pinski J, Szepeshazik K,<br />

Milovanovic SF, Groot K, Schally<br />

AV: Effect of microcapsules of luteinizing<br />

hormone-releasing hormone<br />

antagonist SB-75 <strong>and</strong> somatostatin<br />

analog RC-160 on endocrine<br />

status <strong>and</strong> tumor-growth in the<br />

Dunning R-3327 H rat prostate cancer<br />

model. Prostate 1992;20/4:297–<br />

310.<br />

78 Pinski J, Halmos G, Schally AV: <strong>Somatostatin</strong><br />

analog RC-160 <strong>and</strong> bombesin/gastrin-releasing<br />

peptide antagonist<br />

RC-3095 inhibit the growth<br />

of <strong>and</strong>rogen-independent DU-145<br />

human prostate cancer in nude<br />

mice. <strong>Cancer</strong> Lett 1993;71/1–3:<br />

189–196.<br />

79 Pinski J, Schally AV, Halmos G,<br />

Szepeshazi K: Effect of somatostatin<br />

analog RC-160 <strong>and</strong> bombesin/gastrin-releasing<br />

peptide antagonist<br />

RC-3095 on growth of PC-3 human<br />

prostate-cancer xenografts in nude<br />

mice. Int J <strong>Cancer</strong> 1993;55:963–<br />

967.<br />

80Pinski J, Reile H, Halmos G, Groot<br />

K, Schally AV: Inhibitory effects of<br />

somatostatin analogue RC-160 <strong>and</strong><br />

bombesin/gastrin-releasing peptide<br />

antagonist RC-3095 on the growth<br />

of the <strong>and</strong>rogen-independent Dunning<br />

R-3327-AT-1 rat prostate cancer.<br />

<strong>Cancer</strong> Res 1994;54/1:169–<br />

174.<br />

81 Brevini TA, Bianchi R, Motta M:<br />

Direct inhibitory effect of somatostatin<br />

on growth of the human prostate<br />

cancer cell line LNCaP: Possible<br />

mechanism of action. J Clin Endocrinol<br />

Metab 1993;77:626–631.<br />

82 Bodgem AE, LePage D, Zwicker S,<br />

Grant W, Silver M: Proliferative response<br />

of human prostate tumor xenografts<br />

to surgical trauma <strong>and</strong> the<br />

transurethral resection of the prostate<br />

controversy. Br J <strong>Cancer</strong> 1996;<br />

73/1:73–78.<br />

83 Abrahamsson PA, Wadstrom LB,<br />

Alumets J, Falkmer SS, Grimelius<br />

L: Peptide-hormone <strong>and</strong> serotonin<br />

immunoreactive tumor cells in carcinoma<br />

of the prostate. Pathol Res<br />

Pract 1987;182/3:298–307.<br />

84 Hoosein NM, Logothetis CJ, Chung<br />

LW: Differential effects of peptide<br />

hormones bombesin, vasoactive intestinal<br />

polypeptide <strong>and</strong> somatostatin<br />

analog RC-160 on the invasive<br />

capacity of hormone prostatic carcinoma<br />

cells. J Urol 1993;149:1209–<br />

1213.<br />

85 Rojas-Corona RR, Chen LZ, Mahadenia<br />

PS: Prostatic carcinoma<br />

with endocrine features. A report of<br />

a neoplasm containing multiple immunoreactive<br />

hormonal substances.<br />

Am J Clin Pathol 1987;88:759–762.<br />

86 Shulkes A, Fletcher DR, Rubinstein<br />

C, Ebeling PR, Martin TJ: Production<br />

of calcitonin gene related peptide,<br />

calcitonin <strong>and</strong> PTH related<br />

protein by a prostatic adenocarcinoma.<br />

Clin Endocrinol (Oxf) 1991;34:<br />

387–393.<br />

87 Fjellestad-Paulsen A, Abrahamsson<br />

PA, Bjartell A, Grino A, Grimelius<br />

L, Hedel<strong>and</strong> H, Falkmer S: Carcinoma<br />

of the prostate with Cushing’s<br />

syndrome. A case report with histochemical<br />

<strong>and</strong> chemical demonstration<br />

of immunoreactive corticotropin-releasing<br />

hormone in plasma<br />

<strong>and</strong> tumoral tissue. Acta Endocrinol<br />

(Copenh) 1988;119:506–516.<br />

88 di Sant’Agnese PA: Neuroendocrine<br />

differentiation in human prostatic<br />

carcinoma. Hum Pathol 1992;23:<br />

287–296.<br />

89 Abrahamsson PA: Neuroendocrine<br />

differentiation <strong>and</strong> hormone-refractory<br />

prostate cancer. Prostate Suppl<br />

1996;6:3–8.<br />

90Angelsen A, Syversen V, Stridsberg<br />

M, Haugen CA, Mjolnerod OK,<br />

Waldum HL: Use of neuroendocrine<br />

serum markers in the followup<br />

of patients with cancer of the<br />

prostate. Prostate 1997;31/2:110–<br />

117.<br />

126 Chemotherapy 2001;47(suppl 2):109–126 Vainas<br />

91 Angelsen A, Syversen V, Haugen<br />

CA, Stridsberg U, Mjolnerod OK,<br />

Waldum HL: Neuroendocrine differentiation<br />

in carcinomas of the<br />

prostate: Do neuroendocrine serum<br />

markers reflect immunohistochemical<br />

findings? Prostate 1997;30/1:1–<br />

6.<br />

92 Yagoda A, Petrylak D: Cytotoxic<br />

chemotherapy <strong>for</strong> advanced hormone-resistant<br />

prostate cancer.<br />

<strong>Cancer</strong> 1993;71(suppl 3):1098–<br />

1109.<br />

93 Bono AV, Pozzi E, Robustelli della<br />

Cuna G, Preti P: Prostatic cancer:<br />

Survey of hormonal treatment in<br />

Europe. J Int Med Res 1990;18<br />

(suppl 1):11–25.<br />

94 Dupont A, Boucher H, Cusan L, Lacourciere<br />

V, Emond J, Labrie F: Octreotide<br />

<strong>and</strong> bromocriptine in patients<br />

with stage D2 prostate cancer<br />

who relapsed during treatment with<br />

flutamide <strong>and</strong> castration (letter).<br />

Eur J <strong>Cancer</strong> 1990:26:770–771.<br />

95 Schally AV, Radulovic S, Comaru-<br />

Schally AM: Experimental <strong>and</strong> clinical<br />

studies in hormone-dependent<br />

cancers; in Mazzaferri EL, Samaan<br />

NA (eds): Endocrine Tumors. Ox<strong>for</strong>d,<br />

Blackwell Scientific Publications,<br />

1993, chap 6, pp 49–73.<br />

96 Parmar H, Charlton CK, Phillips<br />

RH, Edwards I, Bejot JL, Thomas F,<br />

Lightman SL: Therapeutic response<br />

to somatostatin analogue, BIM<br />

23014, in metastatic prostatic cancer.<br />

Clin Exp Metastasis 1992;10/1:<br />

3–11.<br />

97 Logothetis CJ, Hossan EA, Smith<br />

TL: SMS 201-995 in the treatment<br />

of refractory prostatic carcinoma.<br />

Anticancer Res 1994;14/6B:2731–<br />

2734.<br />

98 Verhelst J, De Longueville M, Ongena<br />

P, Denis L, Mahler C: Octreotide<br />

in advanced prostatic cancer relapsing<br />

under hormonal treatment.<br />

Acta Urol Belg 1994;62/1:83–88.<br />

99 Figg WD, Thibault A, Cooper MR,<br />

Reid R, Headlee D, Dawson N,<br />

Kohler DR, Reed E, Sartor O: A<br />

phase I study of the somatostatin<br />

analogue somatuline in patients<br />

with metastatic hormone-refractory<br />

prostate cancer. <strong>Cancer</strong> 1995;75:<br />

2159–2164.


Chemotherapy 2001;47(suppl 2):127–133<br />

Gastrointestinal <strong>Cancer</strong><br />

Refractory to Chemotherapy:<br />

A Role <strong>for</strong> Octreotide?<br />

Stefano Cascinu Vincenzo Catalano Paolo Giordani<br />

Anna Maria Baldelli Romina Agostinelli Giuseppina Catalano<br />

Section of Experimental Oncology, Division of Medical Oncology,<br />

Azienda Ospedaliera S. Salvatore, Pesaro, Italy<br />

Key Words<br />

Octreotide W Gastrointestinal cancers W<br />

Chemotherapy<br />

Abstract<br />

Although octreotide has been shown to inhibit<br />

the growth of gastrointestinal (GI) tumors<br />

in vitro <strong>and</strong> in vivo, preliminary clinical<br />

trials have reported disappointing results <strong>for</strong><br />

this somatostatin analog in patients with GI<br />

cancers. The results of these trials probably<br />

reflect the difficulty in assessing the therapeutic<br />

potential of an agent such as octreotide<br />

in GI cancers. Thus, it is possible that<br />

treatment with octreotide could be useful in<br />

the stabilization of disease if it is associated<br />

with an improvement in survival. On the basis<br />

of these considerations five r<strong>and</strong>omized<br />

trials were carried out to evaluate the therapeutic<br />

potential in patients with GI cancers.<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0127$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Four trials (one in patients with colorectal<br />

carcinoma <strong>and</strong> three in patients with carcinoma<br />

of the pancreas) did not show any advantage<br />

of octreotide in untreated patients; in<br />

contrast, one trial reported that octreotide<br />

prolonged survival in patients with GI cancers<br />

refractory to chemotherapy. Some clinical<br />

features of the latter study (treatment<br />

with chemotherapy, different schedules)<br />

may explain these conflicting results. Although<br />

data from r<strong>and</strong>omized trials suggest<br />

that octreotide is not effective in untreated<br />

asymptomatic advanced GI cancer patients,<br />

further studies are warranted to assess the<br />

efficacy of octreotide in chemotherapy refractory<br />

patients in order to clarify the impact<br />

of octreotide in terms of not only survival but<br />

also on the patients’ quality of life.<br />

Stefano Cascinu, MD<br />

Division of Medical Oncology<br />

Maggiore University Hospital<br />

I–43100 Parma (Italy)<br />

Fax +39 521 995448<br />

Copyright © 2001 S. Karger AG, Basel


Gastrointestinal (GI) cancer accounts <strong>for</strong> a<br />

large proportion of all human tumors [1].<br />

Although gastric cancer is declining in incidence,<br />

it remains the second most common<br />

malignancy in the world. Surgery is the only<br />

treatment which offers a chance of a complete<br />

cure <strong>for</strong> localized gastric cacarcinoma. However,<br />

at diagnosis 75% of all patients with gastric<br />

cancer have disseminated disease. Even<br />

among the subgroups of patients who undergo<br />

a potentially curative resection, relapse is<br />

common, consequently the 5-year survival<br />

ranges from 10 to 15% of all patients with<br />

newly diagnosed gastric cancer. Because of<br />

this poor outcome, the use of systemic treatment<br />

has been a subject of great interest. The<br />

currently available data indicate that with<br />

new combination cytotoxic regimens, approximately<br />

half of the patients with metastatized<br />

gastric cancer may benefit from chemotherapy<br />

by reduction of tumor-related symptoms<br />

<strong>and</strong>/or prolongation of survival [2].<br />

In the western countries, colorectal carcinoma<br />

represents, after lung cancer, the second<br />

leading cause of deaths due to neoplasms.<br />

During the past decades, knowledge about<br />

this carcinoma has considerably increased,<br />

but in the advanced disease, little progress has<br />

been made in improving patient survival. In<br />

advanced colorectal cancer in fact (at least<br />

40% of patients will have metastases sometime<br />

during the course of their illness) a st<strong>and</strong>ard<br />

treatment has not yet been established.<br />

For more than 30 years, fluorouracil has been<br />

the drug of choice even if tumor response<br />

rates are not more than 10–15%, with a median<br />

survival of about 1 year. Further innovative<br />

compounds (irinotecan, oxaliplatin) are<br />

now being evaluated in clinical trials producing<br />

promising results [3].<br />

Carcinoma of the pancreas is the fourth<br />

cause of cancer-related death in the United<br />

States, with approximately 28,000 deaths recorded<br />

in 1997. Although there are worldwide<br />

variations, similar figures have been reported<br />

in other western countries, <strong>and</strong> about 30,000<br />

deaths are reported in the European countries<br />

each year. Pancreatic cancer is an aggressive<br />

disease <strong>and</strong> only 5–22% of patients are eligible<br />

<strong>for</strong> a potentially curative resection at the<br />

time of diagnosis. Furthermore, the prognosis<br />

is unfavorable <strong>for</strong> this selected group of patients<br />

with only 10–30% 5-year survival rates.<br />

In patients with locally advanced or metastatic<br />

pancreatic cancer, conventional methods of<br />

treatment, including radiotherapy <strong>and</strong> chemotherapy,<br />

offer little benefit <strong>and</strong> their role in<br />

prolonging survival, ameliorating symptoms<br />

or improving quality of life still remains a<br />

matter of debate [4]. The outcome is even<br />

more dismal <strong>for</strong> patients with advanced gastrointestinal<br />

cancer nonresponsive to chemotherapy.<br />

Median survival <strong>for</strong> these patients is<br />

about 3–6 months, <strong>and</strong> symptoms are frequently<br />

difficult to control [5]. Consequently,<br />

there has been much interest in novel <strong>for</strong>ms of<br />

therapy which may be more effective in patients<br />

with GI malignancies nonresponsive to<br />

chemotherapy. There is some evidence suggesting<br />

that GI tumors may be partly hormone-dependent<br />

<strong>and</strong> that hormonal manipulation<br />

may have a role to play in the management<br />

of these cancers [6]. In fact, previous<br />

studies have shown that the tumor growth <strong>and</strong><br />

cellular proliferation are controlled by GI hormones<br />

<strong>and</strong> growth factors such as EGF <strong>and</strong><br />

IGF-1 [7, 8]. One of the most important naturally<br />

occurring antiproliferative hormones is<br />

somatostatin. <strong>Somatostatin</strong> has been shown<br />

to inhibit proliferation of GI cancers both in<br />

vitro <strong>and</strong> in vivo [9]. However, the short halflife<br />

of native somatostatin (1–3 min) <strong>and</strong><br />

the need <strong>for</strong> its intravenous administration<br />

makes long-term somatostatin therapy impractical<br />

[10]. Octreotide, a synthetic somatostatin<br />

analog, differs substantially from<br />

natural somatostatin in that it has a much<br />

longer half-life [11]. Octreotide inhibits the<br />

128 Chemotherapy 2001;47(suppl 2):127–133 Cascinu/Catalano/Giordani/Baldelli/<br />

Agostinelli/Catalano


growth <strong>and</strong> development of tumor via one or<br />

more of the following mechanisms: a direct<br />

inhibitory effect on the tumor via specific<br />

somatostatin receptors, direct or indirect inhibition<br />

of the production of IGF-1 <strong>and</strong>/or other<br />

growth factors, inhibition of angiogenesis,<br />

reduction in tumor blood flow <strong>and</strong> stimulation<br />

of apoptosis.<br />

The possibility of a direct effect of octreotide<br />

on the tumor tissue implies that the<br />

tumor cells bear somatostatin receptors<br />

(SSTR). In recent years, it has been documented<br />

that many nonendocrine tumors bear<br />

somatostatin receptor subtypes. At present 5<br />

somatostatin receptor subtypes have been<br />

cloned <strong>and</strong> characterized [12]. It was found<br />

that the antiproliferative effects of octreotide<br />

in vitro are mediated via the activation of the<br />

SSTR-2 subtype which is functionally a phosphotyrosine<br />

phosphatase [13]. The loss of<br />

SSTR-2 gene expression may provide a<br />

growth advantage <strong>for</strong> tumors which do not<br />

express this subtype <strong>and</strong> may explain, at least<br />

in part, the failure of somatostatin analogs to<br />

inhibit the growth <strong>and</strong> development of cancers.<br />

Furthermore, the restoration of SSTR-2<br />

expression to human pancreatic cells resulted<br />

in a significant reduction in the growth of<br />

these cells [14]. Octreotide can also exert an<br />

antiproliferative effect by binding to SSTR-5<br />

receptors <strong>and</strong> stimulate apoptosis by binding<br />

the SSTR-3 subtype. However, octreotide has<br />

little or no binding affinity to the somatostatin<br />

receptor subtypes expressed in gastric,<br />

pancreatic <strong>and</strong> colorectal cancers [12]. Nevertheless,<br />

octreotide has been demonstrated<br />

to inhibit their growth <strong>and</strong> development.<br />

These observations would suggest that octreotide<br />

can indirectly inhibit the growth <strong>and</strong><br />

development in these tumors. Several studies<br />

have shown that EGF <strong>and</strong> IGF-1 may be<br />

implicated in the autocrine control of the<br />

growth <strong>and</strong> development of gastric <strong>and</strong> colorectal<br />

tumors. In vitro both EGF <strong>and</strong> IGF-1<br />

promoted cell proliferation of gastric <strong>and</strong><br />

colorectal cancer cell lines [15–17]. In addition,<br />

both growth factors enhanced the response<br />

of these cells to gastrin. Thus, the inhibition<br />

of the secretion of these trophic factors<br />

may have therapeutic potential in the management<br />

of GI cancers [18]. Recently, we have<br />

demonstrated that octreotide resulted in a significant<br />

decrease in serum IGF-1 levels in<br />

colorectal cancer patients which was associated<br />

with a decrease in the proliferative activity<br />

of the tumor cells [19]. Furthermore, we have<br />

demonstrated that there is no correlation between<br />

the dose of octreotide <strong>and</strong> the magnitude<br />

of serum IGF-1 levels [20]. Stewart et al.<br />

[21] <strong>and</strong> Iftikhar et al. [22] reported a similar<br />

interference in tumor cell kinetics after treatment<br />

with octreotide in patients with colonic<br />

<strong>and</strong> rectal cancer.<br />

According to data from Klijn et al. [23] no<br />

significant effects were observed in our patients<br />

on serum levels of HGH <strong>and</strong> EGF.<br />

The antiproliferative effects of octreotide<br />

may also be mediated by inhibition of angiogenesis.<br />

The inhibition of angiogenesis by octreotide<br />

in vivo could result indirectly from<br />

an inhibition of IGF-1 [24]. However, octreotide<br />

may also inhibit angiogenesis by suppression<br />

of paracrine angiogenic factors such as<br />

vascular endothelial growth factor. This suggestion<br />

is supported by the observation that<br />

octreotide significantly reduced serum <strong>and</strong><br />

tissue vascular endothelial growth factor concentrations<br />

in patients with colorectal cancers<br />

[25].<br />

Octreotide has been shown to inhibit the<br />

growth of GI tumors both in vitro <strong>and</strong> in<br />

experimental animal studies [15, 26, 27]. On<br />

the basis of these experimental data, pilot<br />

clinical trials with octreotide were carried out<br />

in GI tumors. In spite of the promising preclinical<br />

results, these initial studies in man<br />

were disappointing (table 1). Klijn et al. [23]<br />

treated 34 patients with gastric <strong>and</strong> colonic<br />

Octreotide in GI <strong>Cancer</strong>s Chemotherapy 2001;47(suppl 2):127–133 129


Table 1. Octreotide in GI cancers: pilot clinical trials<br />

Authors, year Reference<br />

No.<br />

Tumor Number<br />

of<br />

patients<br />

Clinical response<br />

partial stable<br />

Survival<br />

months<br />

Savage et al., 1987 28 Colon 1004 n.r.<br />

Klijn et al., 199023 Colon 16 04 8<br />

Pancreas 14 03 2<br />

Stomach 4 01 n.r.<br />

Friess et al., 1993 31 Pancreas 22 03 5<br />

Rosenberg et al., 1995 33 Pancreas 12 03 12<br />

n.r. = Not reported.<br />

cancers with octreotide. Although octreotide<br />

treatment stabilized the disease in 27% of<br />

these patients there was no benefit in terms of<br />

survival. However, interestingly most patients<br />

expressed a subjective improvement in<br />

the absence of serious side effects.<br />

Savage et al. [28] treated 10 patients without<br />

finding any indication that octreotide can<br />

alter the rate of growth of advanced GI tumors.<br />

In contrast, Smith et al. [29] found a<br />

modest increase in survival in 12 colorectal<br />

cancer patients treated with octreotide but no<br />

objective responses were observed. Similar results<br />

have been reported <strong>for</strong> patients with<br />

pancreatic cancer treated with octreotide, observing<br />

no benefit [30–32]. Conversely, Rosenberg<br />

et al. [33] reported an improvement<br />

in survival with a concomitant treatment of<br />

octreotide <strong>and</strong> tamoxifen compared to historical<br />

controls (12 vs. 3 months).<br />

Overall, these preliminary results in patients<br />

are disappointing but may reflect the<br />

difficulty in assessing the activity of an agent<br />

such as octreotide considering only the objective<br />

responses of tumors. It is possible in fact<br />

that treatment with octreotide could be useful<br />

even obtaining only a stabilization of disease<br />

if this is associated with an improvement in<br />

survival. These considerations were used to<br />

justify five r<strong>and</strong>omized trials in which the<br />

major end points were survival <strong>and</strong> time to<br />

progression of disease. Three studies tested<br />

octreotide in untreated advanced pancreatic<br />

cancer <strong>and</strong> one in untreated advanced colorectal<br />

cancer [34–37]. However, none of the<br />

trials showed any benefit <strong>for</strong> octreotide treatment<br />

in terms of disease progression <strong>and</strong> survival<br />

(table 2).<br />

In 1995, we published the results of a r<strong>and</strong>omized<br />

trial comparing octreotide with the<br />

best available supportive care in patients with<br />

advanced GI cancers refractory to conventional<br />

chemotherapy [38]. The primary outcome<br />

measure was duration of survival. Fiftyfive<br />

patients (15 stomach, 16 pancreas, 24<br />

colorectal) received octreotide (200 Ìg 3 times<br />

a day <strong>for</strong> 5 days a week), while 52 (14 stomach,<br />

16 pancreas, 22 colorectal) received the<br />

best supportive care only. The survival of<br />

patients treated with octreotide was significantly<br />

longer (median 20 weeks) than the 11<br />

weeks of median survival observed in the control<br />

group (p ! 0.001) (table 3). This survival<br />

advantage of patients treated with octreotide<br />

was present in all three types of tumors. Furthermore,<br />

in 22 patients (40%) octreotide relieved<br />

pain <strong>and</strong> allowed discontinuation of<br />

analgesics.<br />

130 Chemotherapy 2001;47(suppl 2):127–133 Cascinu/Catalano/Giordani/Baldelli/<br />

Agostinelli/Catalano


Table 2. Octreotide in GI cancers: r<strong>and</strong>omized trials in untreated patients<br />

Authors, year Reference<br />

No.<br />

Goldberg et al., 1995<br />

34 Octreotide<br />

Placebo<br />

Burch et al., 1995 35 Octreotide<br />

Placebo<br />

Roy et al., 1998 36 Octreotide/5FU<br />

FU<br />

Pederzoli et al., 1998 37 Octreotide<br />

Placebo<br />

n.r. = Not reported.<br />

<strong>Treatment</strong> Tumor Number<br />

of patients<br />

Colon 131<br />

129<br />

Pancreas 42<br />

43<br />

Octreotide 15 24 16 15 22 16<br />

Placebo 14 22 16 8 12 8<br />

Survival<br />

weeks<br />

Octreotide in GI <strong>Cancer</strong>s Chemotherapy 2001;47(suppl 2):127–133 131<br />

17<br />

16.8<br />

n.r.<br />

n.r.<br />

Pancreas 284 22.6<br />

21.6<br />

Pancreas 93<br />

92<br />

16<br />

16.9<br />

Table 3. Octreotide in GI cancers: a r<strong>and</strong>omized trial in refractory chemotherapy patients<br />

<strong>Treatment</strong> Number of patients<br />

It is difficult to explain the disparity of our<br />

results <strong>and</strong> those of the other r<strong>and</strong>omized<br />

trials. However, some clinical <strong>and</strong> methodological<br />

aspects could help to explain these<br />

different results. In our study, only patients<br />

with chemotherapy-refractory disease were<br />

entered, whereas other trials enrolled asymptomatic<br />

patients who had not been treated<br />

with chemotherapy. It is well known that chemotherapy<br />

can improve survival in patients<br />

with advanced GI cancer, so that cytotoxic<br />

drug administrations in some patients could<br />

have determined differences in survival. Un<strong>for</strong>tunately,<br />

in most of these trials the further<br />

treatments of patients failing octreotide therapy<br />

were not reported. Furthermore, there are<br />

stomach colorectal pancreas<br />

Survival, weeks<br />

stomach colorectal pancreas<br />

data suggesting that in the end stages of GI<br />

cancers, when bowel obstruction is present,<br />

octreotide may provide some symptomatic<br />

benefit, <strong>and</strong> may also be a determinant of survival<br />

[39].<br />

Another problem arising from the studies<br />

cited above could be the continuous administration<br />

of octreotide <strong>for</strong> a prolonged period<br />

without interruptions. Preclinical data has<br />

shown that continuous administration of octreotide<br />

results in desensitization or tachyphylaxis<br />

of its inhibitory effects on somatostatin<br />

receptors <strong>and</strong> concomitant increase of<br />

plasma IGF-1 concentration within 6–10<br />

days following initiation of therapy. Possibly,<br />

desensitization may be prevented or delayed


if octreotide is administered intermittently as<br />

in our case [9]. Thus, although advanced GI<br />

cancer patients were included in these trials<br />

assessing the role of octreotide, there is a<br />

potential <strong>for</strong> the difference in activity due to<br />

the different subset of patients (untreated vs.<br />

refractory to chemotherapy, <strong>and</strong> so with<br />

‘more advanced disease’) <strong>and</strong> to the variations<br />

in dose <strong>and</strong> schedule of octreotide.<br />

References<br />

1 Ahlgren JD, Mac Donald JS: Gastrointestinal<br />

Oncology, ed 3. Philadelphia,<br />

Lippincott, 1992.<br />

2 Schipper DL, Wagener DJT: Chemotherapy<br />

of gastric cancer. Anticancer<br />

Drugs 1996;7:137–149.<br />

3 Labianca R, Pessi MA, Zamparelli<br />

G: <strong>Treatment</strong> of colorectal cancer.<br />

Drugs 1997;53:593–607.<br />

4 Graziano F, Catalano G, Cascinu S:<br />

Chemotherapy <strong>for</strong> advanced pancreatic<br />

cancer: The history is changing.<br />

Tumori 1998;84:308–311.<br />

5 Cascinu S, Fedeli A, Luzi Fedeli S,<br />

Catalano G: Salvage chemotherapy<br />

in colorectal cancer patients with<br />

good per<strong>for</strong>mance status <strong>and</strong> young<br />

age after failure of 5fluorouracil/leucovorin<br />

combination. J Chemother<br />

1992;4:46–49<br />

6 Townsend CM, Singh P, Thompson<br />

JC: Effects of gastrointestinal peptides<br />

on gastrointestinal cancer<br />

growth. Gastroenterol Clin 1989;18:<br />

777–791.<br />

7 Lahm H, Svardet L, Laurent PL,<br />

Fisher JR, Leyhan A, Givel JC,<br />

Odartchenko N: Growth regulation<br />

<strong>and</strong> co-stimulation of human colorectal<br />

cancer cell lines by insulin-like<br />

growth factors I, II <strong>and</strong> trans<strong>for</strong>ming<br />

growth factor ·. Br J <strong>Cancer</strong> 1992;<br />

65:341–346.<br />

8 Durrant LG, Watson SA, Hall A,<br />

Morris DL: Co-stimulation of gastrointestinal<br />

tumor growth by gastrin,<br />

trans<strong>for</strong>ming cell growth factor<br />

alpha <strong>and</strong> insulin like growth factors<br />

<strong>and</strong> cancer. Br J <strong>Cancer</strong> 1991;63:<br />

67–70.<br />

9 Lamberts SWJ: Potential role of somatostatin<br />

analogues in the treatment<br />

of cancer. Eur J Clin Invest<br />

1987;17:281–287.<br />

10Sheppard MC, Shapiro B, Pimstone<br />

B, Kronhein S, Berelowitz M, Gregory<br />

M: Metabolic clearance <strong>and</strong><br />

plasma half disappearance time of<br />

exogenous somatostatin in man. J<br />

Clin Endocrinol Metab 1979;48:50–<br />

53.<br />

11 Lamberts SWJ, van der Lely AJ, de<br />

Herder WW, Hofl<strong>and</strong> LJ: Octreotide.<br />

N Engl J Med 1996;334:246–<br />

254.<br />

12 Reubi JC: Octreotide <strong>and</strong> nonendocrine<br />

tumours: Basic knowledge <strong>and</strong><br />

therapeutic potential, in Scarpignato<br />

C (ed): Octreotide: From Basic<br />

Science to Clinical Medicine. Prog<br />

Basic Clin Pharmacol. Basel, Karger,<br />

1996, vol 10, pp 246–269.<br />

13 Buscail L, Esteve JP, Saint-Laurent<br />

N: Inhibition of cell proliferation by<br />

the somatostatin analogue RC-160<br />

is mediated by somatostatin receptor<br />

subtype SSTR2 <strong>and</strong> SSTR5<br />

through different mechanisms. Proc<br />

Natl Acad Sci USA 1995;92:1580–<br />

1584.<br />

14 Delesque N, Buscail L, Esteve JP,<br />

Saint-Laurent N, Muller C, Weckbecker<br />

G, Bruns C, Vaysse N, Susini<br />

C: SST2 somatostatin receptor expression<br />

reverses tumorigenicity of<br />

human pancreatic cancer cells. <strong>Cancer</strong><br />

Res 1997;57:956–962.<br />

15 Dy DY, Whitehead RH, Morris DL:<br />

SMS 201.995 inhibits in vitro <strong>and</strong> in<br />

vivo growth of human colon cancer.<br />

<strong>Cancer</strong> Res 1992;52:917–923.<br />

In conclusion, the present data suggest that<br />

octreotide is not effective in the management<br />

of untreated asymptomatic GI cancer patients.<br />

Nevertheless, although we are aware<br />

that the interpretation of our results requires<br />

caution in view of other conflicting results in<br />

untreated patients, we believe that additional<br />

studies of octreotide are warranted to assess<br />

its efficacy in chemotherapy refractory GI<br />

cancer patients.<br />

16 Pollak MN, Polychronakos C, Guyda<br />

H: Somatomedin analogue SMS<br />

201–995 reduces serum IGF-I levels<br />

in patients with neoplasms potentially<br />

dependent on IGF-I. Anticancer<br />

Res 1989;9:889–892.<br />

17 Baghdiguian S, Verrier B, Gerard C,<br />

Fantini J: Insulin like growth factor<br />

is an autocrine regulator of human<br />

colon cancer cell differentiation <strong>and</strong><br />

growth. <strong>Cancer</strong> Lett 1992;62:23–<br />

33.<br />

18 Baserga R: The insulin-like growth<br />

factor I receptor: A key to tumor<br />

growth? <strong>Cancer</strong> Res 1995;55:249–<br />

252.<br />

19 Cascinu S, Del Ferro E, Grianti C,<br />

Ligi M, Ghiselli M, Foglietti G, Saba<br />

V, Lungarotti F, Catalano G: Inhibition<br />

of tumor cell kinetics <strong>and</strong> serum<br />

insulin growth factor I levels by octreotide<br />

in colorectal cancer patients.<br />

Gastroenterology 1997;113:<br />

767–772.<br />

20Cascinu S, Del Ferro E, Ligi M, Rocchi<br />

MBL, Castellani A, Graziano F,<br />

Ghi<strong>and</strong>oni G, Catalano G: Lack of<br />

correlation between octreotide dose<br />

<strong>and</strong> decrease of serum insulin<br />

growth factor I in advanced colorectal<br />

cancer patients. GI <strong>Cancer</strong> 1997;<br />

2:139–142.<br />

21 Stewart GJ, Connor JL, Lawson JA,<br />

Preketes A, King J, Morris DL: Octreotide<br />

reduces the kinetic index,<br />

proliferating cell nuclear antigenmaximum<br />

proliferative index, in patients<br />

with colorectal cancer. <strong>Cancer</strong><br />

1995;76:572–578.<br />

132 Chemotherapy 2001;47(suppl 2):127–133 Cascinu/Catalano/Giordani/Baldelli/<br />

Agostinelli/Catalano


22 Iftikhar SY, Watson SA, Morris DL:<br />

The effect of long acting somatostatin<br />

analogue SMS 201.995 therapy<br />

on tumor kinetic measurements <strong>and</strong><br />

serum tumor marker concentrations<br />

in primary rectal cancer. Br J <strong>Cancer</strong><br />

1991;63:971–974.<br />

23 Klijn JGM, Hoff AM, Planting<br />

ASTh, Verweij J, Kok T, Lamberts<br />

SWJ, Portengen H, Foekens JA:<br />

<strong>Treatment</strong> of patients with metastatic<br />

pancreatic <strong>and</strong> gastrointestinal<br />

tumors with the somatostatin<br />

analogue S<strong>and</strong>ostatin: A phase II<br />

study including endocrine effects.<br />

Br J <strong>Cancer</strong> 1990;62:627–630.<br />

24 Mallet B, Vialettes B, Haroche S,<br />

Ecoffer P, Gastaut P, Taubert JP,<br />

Vague P: Stabilization of severe proliferative<br />

diabetic retinopathy by<br />

long-term treatment with SMS 201–<br />

995. Diabète Métab 1992;18:438–<br />

444.<br />

25 Cascinu S, Del Ferro E, Ligi M,<br />

Staccioli MP, Giordani P, Catalano<br />

V, Agostinelli R, Muretto P, Catalano<br />

G: Inhibition of vascular endothelial<br />

growth factor by octreotide in<br />

colorectal cancer patients. <strong>Cancer</strong><br />

Invest, in press.<br />

26 Manni A: <strong>Somatostatin</strong> <strong>and</strong> growth<br />

hormone regulation in cancer. Biotherapy<br />

1992;4:31–36.<br />

27 Schally AV: Oncological applications<br />

of somatostatin analogues.<br />

<strong>Cancer</strong> Res 1988;48:6977–6985.<br />

28 Savage AP, Calam J, Wood CB,<br />

Bloom SR: SM 201–995 treatment<br />

<strong>and</strong> advanced intestinal cancer: A<br />

pilot study. Aliment Pharmacol<br />

Ther 1987;1:133–139.<br />

29 Smith JP, Croitorou R, Townsend<br />

CM, Thompson JC: Effects of octreotide,<br />

a long-acting somatostatin<br />

analog, on advanced colon cancer.<br />

Gastroenterology 1992;102:399.<br />

30Ebert M, Friess H, Beger HG,<br />

Büchler MW: Role of octreotide in<br />

the treatment of pancreatic cancer.<br />

Digestion 1994;55:48–51.<br />

31 Friess H, Büchler MW, Beglinger C,<br />

Weber A, Kunz J, Fritsch K, Dennler<br />

HJ, Beger HG: Low-dose octreotide<br />

treatment is not effective in patients<br />

with advanced pancreatic cancer.<br />

Pancreas 1993;8:540–555.<br />

32 Friess H, Büchler MW, Ebert M,<br />

Malfertheiner P, Dennler HJ, Beger<br />

HG: <strong>Treatment</strong> of advanced pancreatic<br />

cancer with high-dose octreotide.<br />

Int J Pancreatol 1993;14:290–<br />

291.<br />

33 Rosenberg L, Barkun AN, Denis<br />

MH, Pollak M: Low-dose octreotide<br />

<strong>and</strong> tamoxifen in the treatment of<br />

adenocarcinoma of the pancreas.<br />

<strong>Cancer</strong> 1995;75:23–28.<br />

34 Goldberg RM, Moertel CG, Wie<strong>and</strong><br />

HJS, Krook JE, Schutt AJ, Veeder<br />

MH, Maillard JA, Dalton RJ: A<br />

phase III evaluation of a somatostatin<br />

analogue (octreotide) in the treatment<br />

of patients with asymptomatic<br />

advanced colon carcinoma. <strong>Cancer</strong><br />

1995;76:961–966.<br />

35 Burch PA, Block M, Wie<strong>and</strong> HS,<br />

Veeder MH, Michalak JC, Hatfield<br />

AK, Wright K: A phase III evaluation<br />

of octreotide versus chemotherapy<br />

with 5FU or 5FU/leucovorin in<br />

advanced exocrine pancreatic cancer.<br />

Proc Am Soc Clin Oncol 1995;<br />

14:488.<br />

36 Roy A, Jacobs A, Bukowsky R, Cunningham<br />

D, Hamm J, Schlag PM,<br />

Rosen P, Francois E, Finley G, Lipton<br />

A, Bruckner H, Haller D, Conroy<br />

T, Goel R, Price P, Smith G,<br />

Mietlowski W, Linnart R, Russo D,<br />

Kay A: A phase III trial of SMS 201–<br />

995 LAR <strong>and</strong> continuous infusion<br />

5FU in unresectable stage II, III, IV<br />

pancreatic cancer. Proc Am Soc Clin<br />

Oncol 1998;17:987.<br />

37 Pederzoli P, Maurer U, Vollmer K,<br />

Büchler MW, Kjaeve J, Van Cutsem<br />

E, Di Carlo V, Stauder H, Bergan A,<br />

Ebert M, Kiese B, Raymond MC,<br />

Kay A: Phase III trial of SMS 201–<br />

995 LAR vs placebo in unresectable<br />

stage II, III, IV pancreatic cancer.<br />

Proc Am Soc Clin Oncol 1998;17:<br />

988.<br />

38 Cascinu S, Del Ferro E, Catalano G:<br />

A r<strong>and</strong>omized trial of octreotide vs<br />

best supportive care only in advanced<br />

gastrointestinal cancer patients<br />

refractory to chemotherapy.<br />

Br J <strong>Cancer</strong> 1995;71:97–101.<br />

39 Mercadante S, Spoldi E, Caraceni A,<br />

Maddaloni S, Simonetti MT: Octreotide<br />

in relieving gastrointestinal<br />

symptoms due to bowel obstruction.<br />

Palliat Med 1993;7:295–299.<br />

Octreotide in GI <strong>Cancer</strong>s Chemotherapy 2001;47(suppl 2):127–133 133


Chemotherapy 2001;47(suppl 2):134–149<br />

Pancreatic <strong>Cancer</strong>:<br />

Does Octreotide Offer Any Promise?<br />

Lawrence Rosenberg<br />

The Pancreatic Diseases Centre, McGill University Health Centre <strong>and</strong> Department of Surgery,<br />

McGill University, Montreal, Canada<br />

Key Words<br />

Pancreatic cancer W <strong>Somatostatin</strong> W<br />

<strong>Somatostatin</strong> receptors W <strong>Somatostatin</strong><br />

analogs W Octreotide W RC-160<br />

Abstract<br />

The incidence of adenocarcinoma of the pancreas<br />

has risen steadily over the past 4 decades.<br />

Since pancreatic cancer is diagnosed<br />

at an advanced stage, <strong>and</strong> because of the<br />

lack of effective therapies the prognosis of<br />

such patients is extremely poor. Despite advances<br />

in our underst<strong>and</strong>ing of the molecular<br />

biology of pancreatic cancer, the systemic<br />

treatment of this disease remains unsatisfactory.<br />

Systemic chemotherapy <strong>and</strong> the administration<br />

of biologically active molecules<br />

such as tumor necrosis factor or interferons<br />

have not resulted in significant improvements<br />

in response rates or patient survival.<br />

New treatment strategies are obviously<br />

needed. This paper will discuss current advances<br />

in the use of somatostatin analogs in<br />

the management of pancreatic cancer.<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

Copyright © 2001 S. Karger AG, Basel<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0134$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Introduction<br />

The incidence of adenocarcinoma of the<br />

pancreas has risen steadily over the past 4<br />

decades. It currently st<strong>and</strong>s at approximately<br />

29,000 new cases per year in North America<br />

[1], making it the second most common gastrointestinal<br />

malignancy <strong>and</strong> the fifth leading<br />

cause of adult deaths from cancer [2]. The disease<br />

is characterized by its aggressive nature.<br />

The diagnosis of pancreatic cancer is usually<br />

established at an advanced stage, <strong>and</strong> a lack of<br />

effective therapies leads to an extremely poor<br />

prognosis. A meta-analysis of 144 reported<br />

series including approximately 37,000 patients<br />

found the median survival time to be 3<br />

months [3]. According to these findings, 65%<br />

of patients with pancreatic cancer will die<br />

within 6 months from the time of diagnosis,<br />

<strong>and</strong> about 90% within 1 year. Surgical resection,<br />

if per<strong>for</strong>med early enough, is presently<br />

the only effective <strong>for</strong>m of curative therapy [4].<br />

However, fewer than 15% of patients with<br />

pancreatic cancer are potential c<strong>and</strong>idates <strong>for</strong><br />

a curative resection [5] due to spread of the<br />

cancer to adjacent tissues or beyond [6]. Only<br />

Dr. Lawrence Rosenberg<br />

Montreal General Hospital, 1650 Cedar Ave., L9-424<br />

Montreal, Que. H3G 1A4 (Canada)<br />

Tel. +1 514 937 6011, ext. 4346, Fax +1 514 934 8210<br />

E-Mail cxlw@musica.mcgill.ca


1–4% of patients with adenocarcinoma of the<br />

pancreas will survive 5 years after diagnosis<br />

[7, 8]. Thus, the incidence rates are virtually<br />

identical to mortality rates.<br />

Approximately half of all patients with<br />

pancreatic cancer have metastatic disease at<br />

the time of diagnosis [9, 10], while most of the<br />

rest have locally advanced, unresectable disease<br />

[11, 12]. Metastatic pancreatic cancer is<br />

one of the most chemotherapy-resistant tumors,<br />

as evidenced by the fact that pancreatic<br />

cancer has the lowest 5-year survival rate (3%)<br />

of any cancer listed in the Surveillance, Epidemiology<br />

<strong>and</strong> End Results (SEER) data base of<br />

the NCI [13].<br />

Computed tomography (CT) <strong>and</strong> magnetic<br />

resonance imaging have made it easier to<br />

determine the diagnosis <strong>and</strong> to define the<br />

stage of the disease. CT-guided needle biopsies<br />

<strong>and</strong> laparoscopy have resulted in fewer<br />

unnecessary laparotomies, while biliary stenting<br />

can reduce the need <strong>for</strong> an invasive operative<br />

procedure in patients with advanced tumors.<br />

Un<strong>for</strong>tunately, these advances have not<br />

resulted in disease detection at an earlier stage<br />

[14].<br />

Reports of 5-year survival among patients<br />

managed with nonsurgical therapies remain<br />

anecdotal. Thus, of 150 patients who have<br />

survived <strong>for</strong> more than 10 years after their<br />

diagnosis of pancreatic cancer, only 12 have<br />

been cured by nonsurgical therapies [3].<br />

Clearly more effective therapies need to be<br />

developed.<br />

Epidemiology <strong>and</strong> Etiologic Factors<br />

A number of factors that may contribute to<br />

the pathogenesis of pancreatic cancer have<br />

recently been identified. These have been<br />

classified as environmental factors, pathological<br />

factors (e.g. chronic pancreatitis), genetic<br />

factors (e.g. familial pancreatic cancer), <strong>and</strong><br />

occupational exposure [15, 16]. Currently,<br />

cigarette smoking is the most firmly established<br />

risk factor associated with pancreatic<br />

cancer. Pancreatic malignancies can be induced<br />

in animals through long-term administration<br />

of tobacco-specific N-nitrosamines or<br />

by parenteral administration of other N-nitroso<br />

compounds [17–19]. Induction of pancreatic<br />

cancer in these experimental models<br />

can be influenced by additional factors, including<br />

changes in bile acid composition, cholecystokinin<br />

levels, diet <strong>and</strong> pancreatic duct<br />

obstruction [20–23].<br />

Clinically, numerous case-control <strong>and</strong> cohort<br />

studies have reported an increased risk of<br />

pancreatic cancer <strong>for</strong> smokers in both the<br />

United States <strong>and</strong> Europe, <strong>and</strong> current estimates<br />

suggest that approximately 30% of pancreatic<br />

cancer cases may be attributed to cigarette<br />

smoking [24, 25].<br />

Pathology<br />

Most malignant pancreatic tumors (95%)<br />

are believed to arise from the exocrine portion<br />

of the gl<strong>and</strong> <strong>and</strong> have light-microscopic features<br />

consistent with those of adenocarcinomas.<br />

Much more infrequent are tumors that<br />

arise from acinar cells or islet cells. Primary<br />

nonepithelial tumors of the pancreas (e.g. lymphomas<br />

or sarcomas) are exceedingly rare.<br />

Natural History of Pancreatic <strong>Cancer</strong><br />

Adenocarcinoma of the pancreas metastasizes<br />

to regional lymph nodes at an early stage<br />

of the disease, <strong>and</strong> subclinical liver metastases<br />

are present in the majority of patients at the<br />

time of diagnosis, even though findings from<br />

imaging studies may be otherwise normal. Patients<br />

who undergo surgical resection <strong>for</strong> localized<br />

nonmetastatic cancer of the head of<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 135


the pancreas have a long-term survival rate of<br />

approximately 20% <strong>and</strong> a median survival of<br />

15–19 months. However, disease recurrence<br />

following a potentially curative Whipple resection<br />

is the norm.<br />

Local recurrence occurs in up to 85% of<br />

patients who undergo surgery alone, <strong>and</strong> local-regional<br />

tumor control may be improved<br />

by combined modality therapy involving both<br />

chemoradiation <strong>and</strong> surgery. Liver metastases<br />

then become the dominant <strong>for</strong>m of tumor<br />

recurrence <strong>and</strong> occur in 50–70% of patients<br />

following potential curative combined modality<br />

treatment. Patients with locally advanced,<br />

nonmetastatic disease have a median survival<br />

of 6–10 months, while those with metastatic<br />

disease have a short survival (3–6 months),<br />

the length of which depends on the extent of<br />

disease <strong>and</strong> per<strong>for</strong>mance status.<br />

Because of the prognosis <strong>and</strong> the patterns<br />

of treatment failure associated with adenocarcinoma<br />

of the pancreas, any proposed treatment<br />

must not be worse than the disease. The<br />

low cure rate <strong>and</strong> modest median survival following<br />

Whipple’s resection m<strong>and</strong>ate that treament-related<br />

morbidity be low <strong>and</strong> treatmentrelated<br />

death be rare. A recent report of the<br />

experience from Johns Hopkins [26] demonstrates<br />

that this can be achieved by careful<br />

selection of patients who undergo therapy. In<br />

addition, however, the development of innovative<br />

treatment strategies directed at the<br />

known sites of tumor recurrence should be<br />

directed towards improvements in patient<br />

survival <strong>and</strong> quality of life.<br />

A Brief History of Pancreatic <strong>Cancer</strong><br />

Therapies<br />

Phase II Trials<br />

Single-agent phase II trials in patients with<br />

advanced pancreatic cancer reported large<br />

variations in response rates [27, 28]. In evalu-<br />

ating these studies, it is important to recognize<br />

that the clinical trial methodology <strong>and</strong><br />

the criteria <strong>for</strong> judging objective response<br />

have changed over time [27]. Phase II trials in<br />

the 1970s frequently included patients with a<br />

variety of different tumors in a single trial,<br />

<strong>and</strong> there<strong>for</strong>e the published response rates<br />

were often based on a small number of patients<br />

with a particular cancer. As a result,<br />

these studies were difficult to interpret from a<br />

statistical point of view.<br />

Prior to 1985, trials relied primarily on an<br />

estimation of tumor size by physical examination,<br />

<strong>and</strong> responses were defined as shrinkage<br />

of a palpable abdominal mass by 50% or<br />

more, or a reduction in the palpable liver span<br />

by 30% or more [28]. The inherent inaccuracy<br />

of these techniques, intraobserver <strong>and</strong> interobserver<br />

variability, <strong>and</strong> the influence of confounding<br />

factors on the size of the measured<br />

lesions all contributed to the initial reports of<br />

high response rates <strong>for</strong> drugs such as 5-fluorouracil<br />

(5-FU), chlorambucil, <strong>and</strong> mitomycin,<br />

as well as the failure to confirm these<br />

promising response rates in subsequent trials,<br />

especially when CT scans were used to determine<br />

tumor response [27].<br />

Phase III Trials<br />

Two kinds of comparative studies have<br />

been carried out in patients with advanced<br />

pancreatic cancer: (1) those that compared<br />

active treatment to best supportive care (to<br />

determine whether chemotherapy made any<br />

difference in the outcome of these patients),<br />

<strong>and</strong> (2) those that compared multiple agent<br />

regimens to single-agent chemotherapy (to determine<br />

whether combinations of drugs with<br />

distinct mechanisms of action could improve<br />

outcome compared to that achieved by single<br />

agents) [27]. Of the three trials that compared<br />

active treatment to best supportive care, two<br />

demonstrated no significant difference [29,<br />

30], <strong>and</strong> the third suggested a substantial sur-<br />

136 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg


vival advantage in favor of a five-drug regimen<br />

[31]. However, subsequent trials of the<br />

regimen failed to duplicate the promising results<br />

of the original study [32].<br />

Despite a number of potentially interesting<br />

preclinical findings <strong>and</strong> promising phase II<br />

studies, virtually no progress has been made<br />

in the chemotherapy <strong>for</strong> advanced pancreatic<br />

cancer during the past 30 years.<br />

Current Approach to Single-Agent<br />

Chemotherapy<br />

The thymidylate synthase inhibitor 5-FU<br />

remains the most extensively evaluated chemotherapeutic<br />

agent <strong>for</strong> pancreatic cancer<br />

[33–35]. Despite numerous trials, however,<br />

its efficacy remains questionable. Between<br />

1991 <strong>and</strong> 1994, 25 investigational new drugs<br />

were evaluated in phase II trials <strong>for</strong> the treatment<br />

of pancreatic cancer. The median response<br />

rate in these trials was 0% (range 0–<br />

14%) <strong>and</strong> the median survival was 3 months<br />

[28]. Inactive drugs that have undergone evaluation<br />

over the past 5 years include iproplatin<br />

[36], trimetrexate [37], edatrexate [38], fazarabine<br />

[39], diaziquone [40], mitoguazone<br />

[40], <strong>and</strong> amonafide [41]. One trial conducted<br />

during this period focused on gemcitabine<br />

(2),2)-difluoro-2)-deoxycytidine).<br />

Gemcitabine<br />

Gemcitabine is a deoxycytidine analog<br />

with structural similarities to cytarabine. As a<br />

prodrug, gemicitabine must be phosphorylated<br />

to its active metabolites – gemcitabine<br />

diphosphate <strong>and</strong> gemcitabine triphosphate.<br />

In both preclinical <strong>and</strong> clinical testing, gemcitabine<br />

demonstrated greater activity against<br />

solid tumors than did cytarabine [42]. These<br />

observations have been explained by the following<br />

properties of gemcitabine: (1) it is 3–4<br />

times more lipophilic than cytarabine, resulting<br />

in greater membrane permeability <strong>and</strong><br />

cellular uptake; (2) it has higher affinity <strong>for</strong><br />

deoxycytidine kinase, <strong>and</strong> (3) the intracellular<br />

retention of gemcitabine triphosphate, an active<br />

metabolite, is prolonged [43].<br />

Following a phase I study [44], gemcitabine<br />

was evaluated in a multicenter trial of 44<br />

patients with advanced pancreatic cancer<br />

[45]. Although the objective response rate to<br />

this drug was only 11% <strong>and</strong> median survival<br />

was 5.6 months, a number of potentially important<br />

observations were made in this trial<br />

[45]. The 1-year survival rate was a remarkably<br />

high 23%, <strong>and</strong> the responses observed<br />

appeared to be somewhat prolonged, i.e. from<br />

4 to more than 20 months. Perhaps the most<br />

unexpected outcome of this study, however,<br />

was the impact of gemcitabine on tumorrelated<br />

symptoms. Of the 5 patients who had<br />

an objective response to gemcitabine, 4 were<br />

able to resume normal daily activities. Three<br />

of the 5 patients were also able to reduce their<br />

daily requirement <strong>for</strong> analgesics. An additional<br />

14 patients who did not meet the radiological<br />

criteria <strong>for</strong> an objective response experienced<br />

disease stabilization <strong>for</strong> 4 months or<br />

more, <strong>and</strong>, of these, 9 had an improvement in<br />

per<strong>for</strong>mance status.<br />

One hundred sixty patients with newly<br />

diagnosed, unresectable pancreatic cancer<br />

were recruited in a phase III trial [46]. A total<br />

of 126 patients completed a period during<br />

which pain was stabilized <strong>and</strong> then they were<br />

r<strong>and</strong>omized to treatment with gemcitabine or<br />

5-FU. Fifteen patients (23.8%) treated with<br />

gemcitabine achieved a clinical beneficial response<br />

compared with only 3 patients (4.8%)<br />

treated with 5-FU (p = 0.002). The median<br />

duration of the clinical beneficial response <strong>for</strong><br />

gemcitabine-treated patients was 18 weeks<br />

compared to 13 weeks <strong>for</strong> the 5-FU-treated<br />

patients. Gemcitabine also proved superior to<br />

5-FU in terms of the trial’s secondary endpoints.<br />

Median survival <strong>for</strong> gemcitabinetreated<br />

patients was 5.6 months compared to<br />

4.4 months <strong>for</strong> 5-FU-treated patients. In addi-<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 137


tion, the probability of survival at 1 year was<br />

18% in the gemcitabine group, significantly<br />

greater than the 2% in the 5-FU group. Few<br />

objective responses, however, were observed<br />

in either treatment arm.<br />

A subsequent phase II study enrolled 63<br />

patients with pancreatic cancer that had progressed<br />

despite treatment with 5-FU [47]. To<br />

be eligible <strong>for</strong> the trial, patients had to have a<br />

significant degree of tumor-related symptoms.<br />

In this study, 17 patients (27%) experienced<br />

a clinically beneficial response to gemcitabine,<br />

the median duration of which was 14<br />

weeks. Median survival of all patients treated<br />

in this trial was 3.8 months. Objective responses<br />

were seen in 6 (10.5%) of the 57<br />

patients with measurable disease.<br />

While these results could suggest that gemcitabine<br />

should become the accepted firstline<br />

therapy <strong>for</strong> patients with advanced pancreatic<br />

adenocarcinoma, the median survival<br />

<strong>for</strong> patients with metastatic disease was still<br />

less than 6 months, with few patients achieving<br />

long-term disease stabilization. Furthermore,<br />

some of the effects attributed to chemotherapy<br />

may not be substantially different<br />

from what can be achieved with aggressive<br />

supportive care alone. In fact the use of clinical<br />

benefit response as a valid means to determine<br />

the efficacy of gemcitabine has itself<br />

been questioned [48]. Thus the median survival<br />

was less than 4 months, <strong>and</strong> 85% did not<br />

survive 7 months. Eight patients had extension<br />

to regional organs or lymph nodes without<br />

distant metastases. One patient survived<br />

4.4 years following the completion of a trial of<br />

5-FU <strong>and</strong> the start of gemcitabine. Thus,<br />

there is no evidence from this study that gemcitabine<br />

improved the survival of patients<br />

with pancreatic cancer.<br />

The quality of life one seeks to evaluate by<br />

defining net patient benefit must take into<br />

account the duration of remaining survival<br />

available to the patient. Gelber [48] has sug-<br />

gested that per<strong>for</strong>ming treatment comparisons<br />

based on the amount of time patients<br />

spend in clinical health states characterized<br />

by relatively good quality of life might be a<br />

better indicator of net patient benefit than<br />

defining a percentage of patients who achieve<br />

some criteria of response. The fact that treatments<br />

which produce higher response rates do<br />

not always yield better survival also argues<br />

against putting too much emphasis in estimating<br />

response rates as a guide to net benefit.<br />

Seventeen patients had a clinically beneficial<br />

response in the phase II study, <strong>and</strong> the<br />

investigators claimed that the treatment was<br />

generally well tolerated [47]. However, although<br />

the toxicities were reported as moderate,<br />

more patients had some noticeable adverse<br />

experiences than achieved a clinical<br />

benefit response. The evidence <strong>for</strong> substantial<br />

benefit <strong>for</strong> gemcitabine is not overwhelming,<br />

<strong>and</strong> additional studies will be required to<br />

more fully define its role in the treatment of<br />

pancreatic cancer.<br />

Other Approaches to Pancreatic<br />

<strong>Cancer</strong><br />

Despite advances in our underst<strong>and</strong>ing of<br />

the molecular biology of pancreatic cancer,<br />

the systemic treatment of metastatic disease<br />

remains unsatisfactory. Systemic chemotherapy<br />

<strong>and</strong> the administration of biologically active<br />

molecules such as tumor necrosis factor<br />

or interferons [49, 50] have not resulted in significant<br />

improvements in response rates or<br />

patient survival. New treatment strategies are<br />

obviously needed. A number of more general<br />

areas of investigation may yield more promising<br />

results. One of these involves interruption<br />

or modulation of growth factors <strong>and</strong> signal<br />

transduction pathways. One example is the<br />

successful treatment of carcinoma of the<br />

breast that has been achieved by endocrine<br />

138 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg


manipulation. The presence of estrogen receptors<br />

on neoplastic breast tissue is correlated<br />

with response to ovarian ablation <strong>and</strong>/or<br />

antiestrogen treatment. A similar approach to<br />

the treatment of pancreatic cancer seems justified<br />

because of the presence of estrogen<br />

receptors in pancreatic carcinoma [51–54] as<br />

well as in normal pancreatic tissue [55, 56]. In<br />

fact, the use of tamoxifen in 80 patients with<br />

ductal adenocarcinoma of pancreas has been<br />

reported in a case-control study to increase<br />

the median survival time from 3 to 7 months<br />

[57, 58]. However, steroid hormones may not<br />

be the most important regulator of pancreatic<br />

cell proliferation. Other potential influences<br />

include the growth factor IGF-1 <strong>and</strong> the<br />

growth inhibitor somatostatin.<br />

Scientific Background <strong>for</strong> the Use of<br />

<strong>Somatostatin</strong>-Based Therapies<br />

<strong>Somatostatin</strong> is a tetradecapeptide that<br />

elicits a variety of biological processes including<br />

inhibition of hormonal secretion <strong>and</strong> cell<br />

proliferation [59]. In some patients, analog<br />

therapy leads to an inhibition of tumor<br />

growth [60–62]. However, the use of native<br />

somatostatin is limited because of its very<br />

short plasma half-life <strong>and</strong> the need <strong>for</strong> continuous<br />

infusion. The recent development of<br />

long-acting somatostatin analogs, such as RC-<br />

160 <strong>and</strong> octreotide, however, has made clinical<br />

trials possible.<br />

These properties of somatostatin <strong>for</strong>m the<br />

basis <strong>for</strong> the treatment of hormone-producing<br />

pituitary or gastroenteropancreatic tumors by<br />

long-acting analogs of the native hormone<br />

[60]. Thus, hormonal suppression is produced<br />

in patients with acromegaly or with neuroendocrine<br />

tumors such as insulinoma, glucagonoma,<br />

gastrinoma, vipoma, or carcinoid syndrome<br />

by somatostatin analogs resulting in<br />

symptomatic relief [60].<br />

<strong>Somatostatin</strong> can exert an antiproliferative<br />

activity by either indirectly inhibiting angiogenesis<br />

or hormone <strong>and</strong> growth factor release,<br />

or by acting directly on neoplastic cells [59,<br />

60, 62]. For example, a number of gastrointestinal<br />

hormones, including gastrin <strong>and</strong> cholecystokinin,<br />

have trophic effects on pancreatic<br />

tissue [63] <strong>and</strong> can stimulate the growth of<br />

pancreatic tumors. <strong>Somatostatin</strong> suppresses<br />

the secretion <strong>and</strong> action of these peptides, <strong>and</strong><br />

this may also contribute to its antiproliferative<br />

activity [64, 65]. In addition, somatostatin<br />

<strong>and</strong> its analogs may also act by reducing<br />

levels of growth factors such as EGF <strong>and</strong> IGF-<br />

1 that are thought to be important in neoplastic<br />

processes [66–70]. This latter possibility is<br />

of considerable interest because both tamoxifen<br />

[70] <strong>and</strong> octreotide [71] have been shown<br />

to lower circulating levels of IGF-1 <strong>and</strong> the<br />

combination has recently been reported to<br />

lower IGF-1 levels more substantially than<br />

either agent alone [72]. This observation<br />

raises the possibility of therapeutic synergy if<br />

the two agents were used together, a suggestion<br />

which is supported by a report that<br />

tamoxifen <strong>and</strong> octrotide are an effective treatment<br />

<strong>for</strong> human pancreatic cancers growing<br />

in nude mice [73]. However, as appealing as<br />

this suggestion is, Klijn et al. [74] measured<br />

insulin, IGF-1, <strong>and</strong> EGF levels in a clinical<br />

study of the use of octreotide <strong>for</strong> pancreatic<br />

cancer. Chronic treatment with a octroetide<br />

had no effect on EGF levels. However, these<br />

investigators observed early significant decreases<br />

in insulin <strong>and</strong> IGF-1; the levels of both<br />

growth factors had returned to pretreatment<br />

values by 5 days <strong>and</strong> 4 weeks, respectively.<br />

This may have been due to the downregulation<br />

of the receptors responsible <strong>for</strong> inhibiting<br />

the release of these trophic factors. This is a<br />

well-recognized effect of even short-term<br />

treatment with octroetide, e.g. loss of initial<br />

potent inhibitory effects, which may be manifested<br />

clinically as tachyphylaxis. This study<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 139


does not support suppression of trophic peptides<br />

as an important mechanism by which<br />

somatostatin inhibits pancreatic cancer, but it<br />

does not eliminate the possibility that these<br />

hormones may influence pancreatic tumor<br />

growth.<br />

The direct actions of somatostatin are mediated<br />

by specific receptors that have been<br />

detected using binding assay or autoradiography<br />

in various human normal <strong>and</strong> tumor tissues<br />

[59, 60]. Recently, five somatostatin receptor<br />

subtypes (SSTR-1–5) <strong>and</strong> one splice<br />

variant have been cloned from humans, mice<br />

<strong>and</strong> rats [75–80].<br />

Buscail et al. [81, 82] have demonstrated a<br />

distinct profile <strong>for</strong> binding of clinically used<br />

somatostatin analogs, SMS 201-995 (octreotide),<br />

BIM 23014 (lanreotide), <strong>and</strong> RC-160<br />

(vapreotide). These analogs bind with high<br />

affinity to SSTR-2, SSTR-3, <strong>and</strong> SSTR-5 <strong>and</strong><br />

with low affinity to SSTR-1 <strong>and</strong> SSTR-4 [75–<br />

85]. The biological functions mediated by the<br />

five SSTR(s) have not yet been completely<br />

established. Recently, after stable expression<br />

of SSTR(s), it was demonstrated that only<br />

SSTR-2 <strong>and</strong> SSTR-5 mediated the antiproliferative<br />

effect of the somatostatin analogs octreotide<br />

<strong>and</strong> vapreotide [82, 83]. The inhibition<br />

of tumor growth by somatostatin <strong>and</strong> its<br />

analogs is rather more complex than these<br />

studies would imply, <strong>and</strong> many gaps remain<br />

in our knowledge. For instance, it is difficult<br />

to measure the binding affinities of the somatostatin<br />

analogs <strong>and</strong> their selectivity, since<br />

most competitive studies use different SSTR<br />

receptor clones from different species expressed<br />

in different cell lines <strong>and</strong> tested with<br />

different radiolig<strong>and</strong>s [86–88].<br />

The antiproliferative action of somatostatin<br />

<strong>and</strong> its analogs on pancreatic cancer has<br />

been demonstrated in vitro <strong>and</strong> in vivo.<br />

Schally [62] reported that the somatostatin<br />

analog RC-160 reduced the weight <strong>and</strong> volume<br />

of the tumor <strong>and</strong> prolonged host survival<br />

in nitrosoamine-induced pancreatic carcinoma<br />

in Syrian golden hamsters. Upp et al. [89]<br />

demonstrated that octreotide inhibited the<br />

growth <strong>and</strong> development of two human pancreatic<br />

cancers, SKI <strong>and</strong> CAV, in nude mouse<br />

xenografts.<br />

In keeping with these findings, recent experimental<br />

evidence has indicated that somatostatin<br />

analogs may act directly on neoplastic<br />

cells by upregulating activity of phosphotyrosine<br />

phosphatases [81, 90], an activity<br />

which would be expected to reduce proliferation<br />

stimulated by growth factors that act via<br />

tyrosine kinase signal transduction pathways.<br />

In fact, Fisher et al. [91, 92] have reported<br />

that when somatostatin receptors were<br />

present on pancreatic tumor cell lines, somatostatin<br />

analogs inhibited their proliferation<br />

both in vivo <strong>and</strong> in vitro. Eleven specimens<br />

of pancreatic adenocarcinoma tissue were<br />

harvested from patients undergoing surgery<br />

<strong>and</strong> nine human pancreatic carcinoma cell<br />

lines were maintained in vitro. All tumors<br />

were assayed <strong>for</strong> expression of mRNA of all<br />

five somatostatin receptors using RT-PCR.<br />

Their findings with the MIA PaCa-2 cell line<br />

are consistent with the results of previous<br />

studies [73, 90, 93–95] <strong>and</strong> support the hypothesis<br />

that high numbers of high affinity<br />

somatostatin receptors on the surface of pancreatic<br />

tumors render the tumor susceptible<br />

to inhibition by somatostatin. Their study<br />

indicated that seven of nine pancreatic cell<br />

lines showed expression of at least one subtype<br />

of somatostatin receptor mRNA, but<br />

only one cell line showed functional somatostatin<br />

receptors on the surface. In this regard,<br />

these findings are consistent with previous<br />

studies showing that somatostatin receptors<br />

are rarely detected on the surface of pancreatic<br />

adenocarcinomas [96]. Data also suggests<br />

that the presence of mRNA <strong>for</strong> the somatostatin<br />

receptor in the clinically acquired specimens<br />

cannot be taken as evidence of the pres-<br />

140 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg


ence of cell surface somatostatin receptors in<br />

those tumors.<br />

Among the recently cloned somatostatin<br />

receptor subtypes, SSTR-2 displays the highest<br />

affinity <strong>for</strong> the somatostatin analogs [75,<br />

79–83, 85]. It has been demonstrated that this<br />

subtype mediates the antiproliferative effect<br />

of the stable analogs octreotide <strong>and</strong> vapreotide<br />

through the stimulation of a tyrosine<br />

phosphatase activity, both in rat pancreatic<br />

cells that endogenously expressed SSTR-2<br />

<strong>and</strong> in cells transfected with SSTR-2 cDNA<br />

[81, 82, 97, 98]. In contrast to that observed in<br />

normal tissue or benign lesions, there is a loss<br />

of gene expression of SSTR-2 in pancreatic<br />

adenocarcinoma, <strong>and</strong> metastases derived<br />

from these primaries.<br />

In addition, SSTR-2 was not expressed in<br />

the pancreatic cancer cell lines, except in the<br />

MIA PaCa-2 cells. Interestingly, RC-l60 inhibited<br />

growth of this latter cell line both in<br />

vitro <strong>and</strong> in vivo through binding of highaffinity<br />

sites <strong>and</strong> stimulation of a tyrosine<br />

phosphatase activity [90]. The molecular<br />

mechanism responsible <strong>for</strong> downregulation of<br />

SSR2 gene expression could occur at the transcriptional<br />

or translational level. Because expression<br />

of the SSR2 gene appears to predict<br />

the response of pancreatic cancer to octreotide,<br />

strategies to increase SSR2 gene expression<br />

in tumors lacking such receptors may<br />

prove beneficial.<br />

All of these observations argue in favor of<br />

the role of SSTR-2 in the negative regulation<br />

of pancreatic cell growth. The loss of expression<br />

of SSTR-2 in neoplastic pancreatic cells<br />

may provide a growth advantage <strong>for</strong> tumors<br />

<strong>and</strong> their metastases. Moreover, a direct antiproliferative<br />

effect of somatostatin <strong>and</strong> its<br />

analogs may thus be excluded in the absence<br />

of SSTR-2 receptors.<br />

The mRNAs of SSTR-1, SSTR-4 <strong>and</strong><br />

SSTR-5 subtypes are present both in normal<br />

<strong>and</strong> cancerous tissues from exocrine pancreas.<br />

We found previously that among these three<br />

subtypes, SSTR-5 also mediated the antiproliferative<br />

effect of somatostatin analogs [82].<br />

However, the mechanism implicated in the<br />

mediation of cell growth inhibition does not<br />

involve activation of tyrosine phosphatase activity<br />

[82]. In these conditions, <strong>and</strong> contrary<br />

to SSTR-2, activation of SSTR-5 cannot<br />

counteract the effect of growth factors acting<br />

via tyrosine kinase-dependent receptors [82]<br />

such as epidermal growth factor, insulin-like<br />

growth factor, <strong>and</strong> insulin that are implicated<br />

in the growth of pancreatic <strong>and</strong> colon cancers<br />

[60, 90, 98, 99].<br />

Only a few studies have so far addressed<br />

the potential benefit of combined treatment<br />

with octreotide <strong>and</strong> various chemotherapeutic<br />

agents. Lamberts et al. [100] combined<br />

octreotide with vincristine, methotrexate, 5-<br />

FU or suramin <strong>and</strong> found an additive interaction.<br />

More recently, Weckbecker et al. [101]<br />

demonstrated the inhibitory effect of octreotide<br />

in combination with the chemotherapeutic<br />

agents Taxol, 5-FU, doxorubicin <strong>and</strong><br />

mitomycin C on the growth of AR42J pancreatic<br />

cancer cells in vitro. The dose-dependent<br />

antiproliferative effects of mitomycin C,<br />

doxorubicin <strong>and</strong> Taxol were synergistically<br />

enhanced by octreotide. Combinations of octreotide<br />

<strong>and</strong> 5-FU resulted either in additive<br />

or, at high concentrations of the chemotherapeutic<br />

agent, in synergistic interactions. These<br />

experiments suggest a modulatory role <strong>for</strong> octreotide<br />

in combination with widely used anticancer<br />

drugs. The additive to synergistic interaction<br />

of octreotide with these chemotherapeutic<br />

agents in in vitro <strong>and</strong> in vivo models<br />

warrants clinical studies to explore the potential<br />

of such combinations in the treatment of<br />

pancreatic cancer.<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 141


Clinical Studies:<br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong><br />

Preliminary results of somatostatin analog<br />

therapy in patients with tumors other than<br />

pancreatic have been encouraging [102–106].<br />

The rationale <strong>for</strong> the use of somatostatin <strong>and</strong><br />

its analogs can be briefly summarized as follows.<br />

<strong>Somatostatin</strong> decreases growth of the<br />

normal pancreas [107], <strong>and</strong> at least one somatostatin<br />

analog, SMS 201-995, inhibits the<br />

growth of human pancreatic adenocarcinoma<br />

in nude mice [89]. The demonstration of somatostatin<br />

receptors in exocrine pancreatic<br />

adenocarcinomas [107] <strong>and</strong> the large body of<br />

evidence [73, 108–114] demonstrating antiproliferative<br />

activity of somatostatin <strong>and</strong> somatostatin<br />

analogs on experimental pancreatic<br />

neoplasms in vitro <strong>and</strong> in vivo justify clinical<br />

studies on the potential therapeutic role of<br />

these drugs in pancreatic cancer.<br />

The first clinical study on somatostatin<br />

analog treatment was published by Klijn et al.<br />

[115]. This group [74] went on to treat 14<br />

patients who had metastatic pancreatic cancer<br />

with three daily subcutaneous injections<br />

of octreotide (100–200 pg/injection) <strong>for</strong> an<br />

average of 7 weeks <strong>and</strong> observed no antitumor<br />

effect.<br />

Friess et al. [116, 117] <strong>and</strong> Ebert et al.<br />

[118], using octreotide at a low dose level<br />

[3 ! 100–200 Ìg/day) <strong>and</strong> a high dose level<br />

(3 ! 2000 Ìg/day), suggested that the effects<br />

of this somatostatin analog were dose-dependent.<br />

These observations are in accord with<br />

the dependent relationship of somatostatin<br />

analogs on the proliferation in breast cancer<br />

cell lines [119]. In the high-dose study of<br />

Friess et al. [117], the median survival increased<br />

from 4 to 6 months with symptomatic<br />

<strong>and</strong> clinical improvement. A positive impact<br />

on the course of disease was confirmed by a<br />

r<strong>and</strong>omized trial of octreotide versus best<br />

supportive care [120] based on a low-dose<br />

therapy given 5 days/week. In this trial a significant<br />

advantage in duration of survival <strong>and</strong><br />

in percentage of stable disease was observed<br />

<strong>for</strong> the octreotide-treated patients although no<br />

objective response was reached.<br />

The only objective response reported <strong>for</strong><br />

somatostatin analogs in pancreatic cancer was<br />

observed by Canobbio et al. [121], who administered<br />

BIM 23014 in dosages between<br />

250 <strong>and</strong> 1,000 Ìg/day to 18 evaluable patients.<br />

However, only one partial response<br />

was observed at the highest dose level. Huguier<br />

et al. [122] in a r<strong>and</strong>omized prospective<br />

study of 86 patients who were given a similar<br />

treatment regimen demonstrated no significant<br />

increase in median survival rates using<br />

life table analysis.<br />

Weckbecker et al. [123] have demonstrated<br />

potentiation of the anticancer effect of<br />

tamoxifen by concomitant infusion of highdose<br />

octreotide in the rat DMBA mammary<br />

cancer model. In order to further define a possible<br />

beneficial role <strong>for</strong> combination hormonal<br />

therapy with octreotide <strong>and</strong> tamoxifen, Rosenberg<br />

et al. [124] studied the effect of<br />

chronic administration of these two inhibitory<br />

agents on survival of a prospective series of<br />

12 consecutive patients with biopsy-proven<br />

ductal adenocarcinoma of the pancreas followed<br />

up from 1990 to 1993. Five of these<br />

patients had resectable disease <strong>and</strong> the remaining<br />

7 were deemed unresectable. <strong>Treatment</strong><br />

consisted of 100 Ìg of octreotide subcutaneously<br />

3 times daily <strong>and</strong> 10 mg of tamoxifen<br />

orally twice daily. Patients had regular follow-up<br />

examinations at 4- to 6-week intervals<br />

until the time of death. The major outcome<br />

was the median duration of patient survival in<br />

months from the time of diagnosis. The outcome<br />

of this group of patients was compared<br />

to a cohort of 68 patients with a biopsy-documented<br />

diagnosis of ductal adenocarcinoma<br />

of the pancreas treated between 1985 <strong>and</strong><br />

1990. Data was collected on age, gender, stage<br />

142 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg


at diagnosis (based on the TNM classification<br />

of the International Union against <strong>Cancer</strong>),<br />

type of treatment <strong>and</strong> duration of survival<br />

from the time of diagnosis.<br />

The median survival times of the octreotide/tamoxifen-treated<br />

group <strong>and</strong> the historical<br />

cohort were 12 <strong>and</strong> 3 months, respectively,<br />

<strong>and</strong> the 1-year actuarial survival was 59 <strong>and</strong><br />

16%, respectively.<br />

The median survival times of the resected<br />

(n = 5) <strong>and</strong> the unresected patients (n = 7)<br />

were 20 <strong>and</strong> 12 months, respectively, <strong>and</strong> the<br />

1-year actuarial survival was 80 <strong>and</strong> 31%,<br />

respectively. The median survival times of the<br />

resected octreotide/tamoxifen-treated group<br />

<strong>and</strong> the resected historical cohort were 20 <strong>and</strong><br />

12 months, respectively, <strong>and</strong> the 1-year actuarial<br />

survival was 80 <strong>and</strong> 44%, respectively.<br />

The median survival times of the 7 unresected<br />

octreotide/tamoxifen-treated patients <strong>and</strong> the<br />

59 unresected historical patients were 12 <strong>and</strong><br />

2.5 months, respectively, <strong>and</strong> the 1-year actuarial<br />

survival was 31 <strong>and</strong> 11%, respectively.<br />

Cox’s proportional hazard analysis confirmed<br />

that treatment <strong>and</strong> resection both independently<br />

predicted a longer survival (p ! 0.01).<br />

The significance of a possible interaction between<br />

both treatment <strong>and</strong> resection could not<br />

be fully determined due to the small sample<br />

size.<br />

Although an assessment of quality of life<br />

<strong>and</strong> pain control was not prospectively assessed<br />

in this study, a post hoc analysis demonstrated<br />

that morphine sulfate was not required<br />

until the final 4–6 weeks of illness in 9<br />

of 12 patients that had already succumbed to<br />

their disease.<br />

CT scanning examinations of the twelve<br />

cases were per<strong>for</strong>med at intervals of 6–8<br />

weeks. No objective responses were seen in<br />

terms of tumor regression; rather the data was<br />

more consistent with a slowing of tumor progression.<br />

The 3 patients surviving at the time<br />

of publication had stable disease.<br />

Possible explanations <strong>for</strong> the apparent<br />

benefit of octreotide/tamoxifen therapy in<br />

this study include (1) diagnosis of the historical<br />

cohort later in the course of the disease,<br />

(2) diagnosis of the cases earlier in the course<br />

of the disease, or (3) a change in the biological<br />

aggressiveness of the disease as a consequence<br />

of treatment. The first two explanations are<br />

unlikely as the data indicated that the stage at<br />

which the disease was diagnosed in cases <strong>and</strong><br />

the cohort was similar. While it might appear<br />

that more patients in the octreotide/tamoxifen-treated<br />

group underwent resection, <strong>and</strong><br />

were there<strong>for</strong>e ‘more curable’, the TNM stage<br />

at the time of diagnosis was similar in both<br />

groups.<br />

The most recent report is the phase II study<br />

by Fazeny et al. [125], which was designed to<br />

investigate the efficacy <strong>and</strong> toxicity of octreotide<br />

combined with goserelin in patients with<br />

advanced pancreatic cancer. Octreotide was<br />

injected subcutaneously in dosages increasing<br />

weekly, starting with 50 Ìg twice daily, until<br />

the level of maintenance therapy of 500 Ìg<br />

3 times a day was reached. In addition, 3.8 mg<br />

goserelin acetate was administered subcutaneously<br />

at monthly intervals. A median of 7<br />

cycles (range 1–27 cycles) was applied. In<br />

comparison to the 40% of patients who had<br />

no change in their disease while on high-dose<br />

octreotide [117], Fazeny et al. [125] demonstrated<br />

that 500 Ìg octreotide 3 times per day<br />

in combination with goserelin resulted in one<br />

partial response <strong>and</strong> no disease progression in<br />

70% of participants. The observations suggest<br />

that combining octreotide with an LHRH<br />

analog might be of therapeutic benefit in patients<br />

with pancreatic cancer <strong>and</strong> could compensate<br />

<strong>for</strong> the potential advantage of a higher<br />

dose of octreotide. Overall, however, the regimen<br />

under investigation did not meet the criteria<br />

<strong>for</strong> sufficient antitumor effectiveness.<br />

Nevertheless, this study rein<strong>for</strong>ces the concept<br />

that pancreatic cancer is principally re-<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 143


sponsive to endocrine therapy <strong>and</strong> there<strong>for</strong>e<br />

the further investigation of hormonal manipulation<br />

seems worth while in the future.<br />

Conclusions<br />

In summary, while it is clear from the<br />

numerous studies conducted on experimental<br />

neoplasms in vitro <strong>and</strong> in vivo that somatostatin<br />

analogs inhibit growth of exocrine pancreatic<br />

cancers, clinical studies have demonstrated<br />

that somatostatin analog therapy<br />

probably does not produce an adequate clinical<br />

response in patients with advanced pancreatic<br />

cancer.<br />

An explanation <strong>for</strong> the discrepancy between<br />

the animal <strong>and</strong> human trials has not<br />

been elucidated. It is possible that a higher<br />

dose of somatostatin <strong>and</strong> its analogs should<br />

have been used in the human trials. Many of<br />

the animal studies used doses at least an order<br />

of magnitude greater than those used in the<br />

clinical trials. However, it is possible that the<br />

main reason <strong>for</strong> the failure of adjuvant treatment<br />

with somatostatin analogs is that most<br />

human pancreatic cancers lack receptors <strong>for</strong><br />

somatostatin. <strong>Somatostatin</strong> binding sites were<br />

poorly expressed or not detected in many<br />

tumors from nonresponsive patients [60, 108,<br />

126], <strong>and</strong> this could be one of the explanations<br />

<strong>for</strong> a lack of direct antiproliferative<br />

effect of analog therapy. Moreover, considering<br />

the differences in the pharmacological<br />

profile <strong>and</strong> the biological function mediated<br />

by somatostatin receptor subtypes, the therapeutic<br />

response may depend on the subtype<br />

expression <strong>for</strong> a given tumor. In contrast to<br />

that observed in normal tissue or benign lesions,<br />

there is a loss of gene expression of<br />

SSTR-2 in pancreatic adenocarcinomas, <strong>and</strong><br />

metastases derived from these primaries<br />

[127]. The loss of expression of SSTR-2 in<br />

neoplastic pancreatic cells may provide a<br />

growth advantage <strong>for</strong> tumors <strong>and</strong> their metastases.<br />

There<strong>for</strong>e, a direct antiproliferative effect<br />

of somatostatin <strong>and</strong> its analogs may thus<br />

partly be excluded in the absence of SSTR-2.<br />

Alternatively, the different results of animal<br />

<strong>and</strong> human studies with different analogs of<br />

somatostatin may reside in the fact that there<br />

are different binding affinities of the various<br />

somatostatin analogs to somatostatin receptors<br />

[107, 126, 128].<br />

Future Directions<br />

The management of pancreatic cancer has<br />

improved relatively little in the past few decades;<br />

the major advances have been better<br />

preoperative staging <strong>and</strong> a decrease in surgical<br />

morbidity <strong>and</strong> mortality. New cytotoxic<br />

drugs, novel biological agents or biological<br />

response modifiers, new surgical <strong>and</strong> radiotherapeutic<br />

techniques, or a combination of<br />

these modalities have all failed to have any<br />

appreciable impact on the final outcome of<br />

this disease. As local-regional treatment becomes<br />

more effective, the dominant site of<br />

failure has shifted to hepatic metastases.<br />

There<strong>for</strong>e, future improvements in survival<br />

duration will result either from effective systemic<br />

or regional therapy directed at subclinical<br />

liver metastases or from strategies <strong>for</strong><br />

screening <strong>and</strong> early diagnosis directed at increasing<br />

the number of patients eligible <strong>for</strong><br />

potentially curative surgery. Further improvements<br />

in the quality of patient survival will<br />

result from the application of innovative multimodality<br />

therapy to carefully selected patients<br />

<strong>and</strong> the avoidance of unnecessary patient<br />

morbidity due to the inappropriate use<br />

of surgery, radiation, <strong>and</strong>/or chemotherapy in<br />

poorly selected patients with advanced disease.<br />

Although the indication <strong>for</strong> the use of somatostatin<br />

analogs in the treatment of pan-<br />

144 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg


creatic cancer has yet to be decisively established,<br />

data from numerous in vitro <strong>and</strong> animal<br />

studies, as well as the results of several<br />

clinical trials rein<strong>for</strong>ce the concept that pancreatic<br />

cancer is principally responsive to endocrine<br />

therapy.<br />

It has been suggested that improved methodology<br />

to assess the interaction of somatostatin<br />

(<strong>and</strong> its analogs) with its target at a cellular<br />

level may allow further development of this<br />

agent. A novel variation of this approach used<br />

an octapeptide analog of somatostatin that<br />

contained methotrexate attached to the ·amino<br />

group of D-phenylalanine in position<br />

1. Subsequent experiments with the MIA<br />

PaCa-2 human pancreatic cell lines in nude<br />

mice demonstrated significant inhibition of<br />

tumor growth [102].<br />

In addition, the attractive features of this<br />

therapeutic concept are the absence of severe<br />

References<br />

1 L<strong>and</strong>is S, Taylor M, Bolden S, et al:<br />

<strong>Cancer</strong> statistics, 1998. CA <strong>Cancer</strong> J<br />

Clin 1998;48:6–30.<br />

2 Wingo PA, Tong T, Bolden S: <strong>Cancer</strong><br />

statistics, 1995. CA <strong>Cancer</strong> J<br />

Clin 1995;45:8–30.<br />

3 Gudjonsson B: <strong>Cancer</strong> of the pancreas.<br />

<strong>Cancer</strong> 1987;60:2284–2303.<br />

4 Tsuchiya R, Noda T, Harada N,<br />

Miyamoto T, Tomioka T, Yamamoto<br />

K, et al: Collective review of small<br />

carcinomas of the pancreas. Ann<br />

Surg 1986;203:77–81.<br />

5 Friess H, Büchler MW, Beglinger C,<br />

Weber A, Kunz J, Fritsch K, et al:<br />

Low-dose octreotide treatment is<br />

not effective in patients with advanced<br />

pancreatic cancer. Pancreas<br />

1993;8:540–545.<br />

6 Kelly DM, Benjamin IS: Pancreatic<br />

carcinoma. Ann Oncol 1995;6:19–<br />

28.<br />

7 Williamson RCN: Pancreatic cancer:<br />

The greatest oncologic challenge.<br />

Br Med J 1988;296:445–446.<br />

side effects, the frequently observed improvement<br />

of patients’ per<strong>for</strong>mance status <strong>and</strong> the<br />

advantage of treatment administered completely<br />

outside the hospital during the entire<br />

period of therapy. Furthermore, the definitive<br />

identification of different receptor subtype(s)<br />

mediating the antiproliferative effects, the expression<br />

of these receptor subtype(s) in this<br />

tumor <strong>and</strong> the development of subtype-specific<br />

analogs should lead to controlled trials of<br />

hormonal manipulation in this malignancy<br />

which is virtually untouched by any systemic<br />

therapy at present.<br />

Acknowledgment<br />

8 Gordis L, Gold EB: Epidemiology of<br />

pancreatic cancer. World J Surg<br />

1984;8:808–821.<br />

9Douglass H: Adjuvant therapy <strong>for</strong><br />

pancreatic cancer. World J Surg<br />

1995;19:170–174.<br />

10 American <strong>Cancer</strong> Society: <strong>Cancer</strong><br />

Facts <strong>and</strong> Figures – 1991. New<br />

York, American <strong>Cancer</strong> Society,<br />

1991.<br />

11 Connolly M, Dawson P, Michelassi<br />

F, et al: Survival in 1001 patients<br />

with carcinoma of the pancreas.<br />

Ann Surg 1987;206:366–373.<br />

12 Singh S, Longmire W, Reber H: Surgical<br />

palliation <strong>for</strong> pancreatic cancer:<br />

The UCLA experience. Ann<br />

Surg 1990;212:132–139.<br />

13 Ries LAG, Miller BA, Hankey BF,<br />

et al: SEER <strong>Cancer</strong> Statistics Review,<br />

1973–1991: Tables <strong>and</strong><br />

Graphs. Bethesda, National <strong>Cancer</strong><br />

Institute, 1994, IH No 94-2789, pp<br />

356–368.<br />

L. Rosenberg is a Senior Clinician-Scientist supported<br />

by the Fonds de la Recherche en Santé du Québec<br />

(FRSQ).<br />

14 Blackstock AW, Cox AD, Tepper<br />

JE: <strong>Treatment</strong> of pancreatic cancer:<br />

Current limitations, future possibilities.<br />

Oncology 1996;10:301–330.<br />

15 Fl<strong>and</strong>ers TY, Foulkes WD: Pancreatic<br />

adenocarcinoma: Epidemiology<br />

<strong>and</strong> genetics. J Med Genet 1996;<br />

33:889–898.<br />

16 Abruzzese JL: Pancreatic cancer:<br />

Overview of current <strong>and</strong> future therapeutic<br />

approaches. Educational<br />

Book of the American Society of<br />

Clinical Oncology, 33rd Annual<br />

Meeting, 1997, pp 65–70.<br />

17 Rivenson A, Hoffman D, Prokopczyk<br />

B, et al: Induction of lung <strong>and</strong><br />

exocrine pancreatic tumors in F344<br />

rats by tobacco-specific <strong>and</strong> Arecaderived<br />

N-nitrosamines. <strong>Cancer</strong> Res<br />

1988;48:6912–6917.<br />

18 Pour PM, Rivenson A: Induction of<br />

a mixed ductal-squamous-islet cell<br />

carcinoma in a rat treated with a<br />

tobacco-specific carcinogen. Am J<br />

Pathol 1989;134:627–631.<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 145


19Hoffmann D, Rivenson A, Chung<br />

FL, et al: Nicotine-derived N-nitrosamines<br />

(TSNA) <strong>and</strong> their relevance<br />

in tobacco carcinogenesis. Crit Rev<br />

Toxicol 1991;21:305–311.<br />

20 Ogawa T, Makino T, Mizumoto K,<br />

et al: Promoting effect of truncal vagotomy<br />

on pancreatic carcinogenesis<br />

initiated with N-nitrosobis-(2-oxopropyl)<br />

amine in Syrian golden<br />

hamsters. Carcinogenesis 1991;12:<br />

1227–1230.<br />

21 Corra S, Kazakoff K, Lawson TA, et<br />

al: Cholecystokinin inhibits DNA<br />

alkylation induced by N-nitrosobis<br />

(2-oxopropyl)amine (BOP) in hamster<br />

pancreas. <strong>Cancer</strong> Lett 1992;62:<br />

251–256.<br />

22 Hoffmann D, Rivenson A, Abbi R,<br />

et al: A study of tobacco carcinogenesis:<br />

Effect of the fat content of<br />

the diet on the carcinogenic activity<br />

of 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone<br />

in F344 rats. <strong>Cancer</strong><br />

Res 1993;53:2758–2761.<br />

23 Rosenberg L, Brown RA, Duguid<br />

WP: Development of experimental<br />

cancer in the head of the pancreas by<br />

surgical induction of tissue injury.<br />

Am J Surg 1984;147:146–151.<br />

24 Silverman DT, Dunn JA, Hoover<br />

RN, et al: Cigarette smoking <strong>and</strong><br />

pancreas cancer: A case-control<br />

study based on direct interviews. J<br />

Natl <strong>Cancer</strong> Inst 1994;86:1510–<br />

1516.<br />

25 La Vecchia C, Boyle P, Francesschi<br />

S, et al: Smoking <strong>and</strong> cancer with<br />

emphasis on Europe. Eur J <strong>Cancer</strong><br />

1991;27:94–104.<br />

26 Yeo C, Cameron J, Lillemoe K, et al:<br />

Pancreaticoduodenectomy <strong>for</strong> cancer<br />

of the head of the pancreas. Ann<br />

Surg 1995;221:721–733.<br />

27 Rothenberg ML: New developments<br />

in chemotherapy <strong>for</strong> patients with<br />

advanced pancreatic cancer. Oncology<br />

1996;10:18–22.<br />

28 Rothenberg ML, Abbruzzese JL,<br />

Moore M, Portenoy RK, Robertson<br />

JM, Wanebo HJ: A rationale <strong>for</strong> exp<strong>and</strong>ing<br />

the endpoints <strong>for</strong> clinical<br />

trials in advanced pancreatic carcinoma.<br />

<strong>Cancer</strong> 1996;78(3 suppl):<br />

627–632.<br />

29Andersen JR, Friss-Mollek A,<br />

Hancke S, et al: A controlled trial of<br />

combination chemotherapy with 5-<br />

FU <strong>and</strong> BCNU in pancreatic adenocarcinoma.<br />

Sc<strong>and</strong> J Gastroenterol<br />

1981;16:973.<br />

30 Frey C, Twomey P, Keehn R, et al:<br />

R<strong>and</strong>omized study of 5-FU <strong>and</strong><br />

CCNU in pancreatic cancer: Report<br />

of the Veteran Administration Surgical<br />

Adjuvant <strong>Cancer</strong> Chemotherapy<br />

Study Group. <strong>Cancer</strong> 1981;47:<br />

27–32.<br />

31 Mallinson CN, Rake MO, Cocking<br />

JB, et al: Chemotherapy in pancreatic<br />

cancer: Results of a controlled,<br />

prospective, r<strong>and</strong>omised, multicentre<br />

trial. Br Med J 1980;281:<br />

1589–1591.<br />

32 Cullinan S, Moertel CG, Wie<strong>and</strong><br />

HS, et al: A phase III trial on the<br />

therapy of advanced pancreatic cancer:<br />

Evaluations of the Mallinson<br />

regimen <strong>and</strong> combined 5-fluorouracil,<br />

doxorubicin, <strong>and</strong> cisplatin. <strong>Cancer</strong><br />

1990;65:2207–2212.<br />

33 Kelsen D: The use of chemotherapy<br />

in the treament of advanced gastric<br />

<strong>and</strong> pancreatic cancer. Semin Oncol<br />

1994;21:58–66.<br />

34 Bukowski RM: Role of chemotherapy<br />

in patients with adenocarcinoma<br />

of the pancreas. Adv Oncol 1995;11:<br />

25.<br />

35 Carter SK: The integration of chemotherapy<br />

into a combined modality<br />

approach <strong>for</strong> cancer treatment.<br />

VI. Pancreatic adenocarcinoma.<br />

<strong>Cancer</strong> Treat Rev 1975;3:193.<br />

36 Hubbard KP, Pazdur R, Ajani JA, et<br />

al: Phase II evaluation of iproplatin<br />

in patients with advanced gastric<br />

<strong>and</strong> pancreatic cancer. Am J Clin<br />

Oncol 1992;15:524–527.<br />

37 Carlson RW, Doroshow JH, Odujinrin<br />

OO, et al: Trimetrexate in locally<br />

advanced or metastatic adenocarcinoma<br />

of the pancreas: A phase<br />

II study of the Northern Cali<strong>for</strong>nia<br />

Oncology Group. Invest New Drugs<br />

1990;8:387–389.<br />

38 Casper ES, Schwartz GK, Johnson<br />

B, et al: Phase II trial of edatrexate<br />

in patients with advanced pancreatic<br />

cancer. Invest New Drugs 1992;<br />

10:313–316.<br />

39Casper ES, Schwartz GK, Kelsen<br />

DP: Phase II trial of fazarabine (arabinofuranosyl-5-azacytidine)<br />

in patients<br />

with advanced pancreatic adenocarcinoma.<br />

Invest New Drugs<br />

1992;10:205–209.<br />

146 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg<br />

40 Bukowski RM, Fleming TR, Mac-<br />

Donald JS, et al: Evaluation of combination<br />

chemotherapy <strong>and</strong> phase II<br />

agents in pancreatic adenocarcinoma:<br />

A Southwest Oncology Group<br />

study. <strong>Cancer</strong> 1993;71:322–325.<br />

41 Linke K, Pazdur R, Abbruzzese J, et<br />

al: Phase II study of amonafide in<br />

advanced pancreatic adenocarcinoma.<br />

Invest New Drugs 1991;9:353–<br />

356.<br />

42 Hertel LW, Boder GB, Kroin JS, et<br />

al: Evaluation of the antitumor activity<br />

of gemcitabine (2),2)-difluoro-<br />

2)-deoxycytidine). <strong>Cancer</strong> Res 1990;<br />

50:4417–4422.<br />

43 Heinemann V, Hertel LW, Grindley<br />

GB, et al: Comparison of the cellular<br />

pharmacokinetics <strong>and</strong> toxicity of<br />

2),2)-difluoro-2)-deoxycytidine <strong>and</strong><br />

1-ß-D arabinofuranosylcytosine.<br />

<strong>Cancer</strong> 1988;48:4024–4031.<br />

44 Abbruzzese JL, Grunewald R,<br />

Weeks EA, et al: A phase I clinical,<br />

plasma <strong>and</strong> cellular pharmacology<br />

study of gemcitabine. J Clin Oncol<br />

1991;9:491–498.<br />

45 Casper ES, Green MR, Kelsen DP,<br />

et al: Phase II trial of gemcitabine<br />

(2),2)-difluoro-2)-deoxycytidine) in<br />

patients with adenocarcinoma of the<br />

pancreas. Invest New Drugs 1994;<br />

12:29–34.<br />

46 Burris HA III, Moore MJ, Andersen<br />

J, et al: Improvements in survival<br />

<strong>and</strong> clinical benefit with gemcitabine<br />

as first-line therapy <strong>for</strong> patients<br />

with advanced pancreas cancer: A<br />

r<strong>and</strong>omized trial. J Clin Oncol<br />

1997;15:2403–2413.<br />

47 Rothenberg ML, Moore MJ, Cripps<br />

MC, et al: A phase II trial of gemcitabine<br />

in patients with 5-FU-refractory<br />

pancreas cancer. Ann Oncol<br />

1996;7:347–353.<br />

48 Gelber RD: Gemcitabine <strong>for</strong> pancreatic<br />

cancer: How hard to look <strong>for</strong><br />

clinical benefit? An American perspective.<br />

Ann Oncol 1996;7:335–<br />

337.<br />

49Abbruzzese JL, Levin B, Ajani JA,<br />

et al: A phase I trial of recombinant<br />

human interferon-gamma <strong>and</strong> recombinant<br />

human tumor necrosis<br />

factor in patients with gastrointestinal<br />

cancer. <strong>Cancer</strong> Res 1989;49:<br />

4057–4061.


50 Abbruzzese JL, Levin B, Ajani JA,<br />

et al: A pilot phase II trial of recombinant<br />

human interferon-gamma<br />

<strong>and</strong> recombinant human tumor necrosis<br />

factor in patients with gastrointestinal<br />

malignancies: Results<br />

of a trial terminated by excessive<br />

toxicity. J Biol Response Modif<br />

1992;9:522–527.<br />

51 Andren-S<strong>and</strong>erg A, Borg S, Dawiskiba<br />

I, Ferno M: Estrogen receptors<br />

<strong>and</strong> estrogen binding protein in pancreatic<br />

cancer. Digestion 1982;25:<br />

12.<br />

52 Berz C, Holl<strong>and</strong>er C, Miller B: Endocrine<br />

responsive pancreatic carcinoma<br />

steroid binding <strong>and</strong> cytotoxicity<br />

studies in human tumor cell<br />

lines. <strong>Cancer</strong> Res 1986;46:2276–<br />

2281.<br />

53 Greenway B, Iqbal MJ, Johnson PJ,<br />

Williams R: Oestrogen receptor proteins<br />

in malignant <strong>and</strong> fetal pancreas.<br />

Br Med J 1981;283:751–753.<br />

54 Satake K, Yoshimoto T, Mukai R,<br />

Umeyama K: Estrogen receptors in<br />

7, 12-dimethylbenz (a) anthracene<br />

(DMBA) induced pancreatic carcinoma<br />

in rats <strong>and</strong> in human pancreatic<br />

carcinoma. Clin Oncol 1982;<br />

8:49–54.<br />

55 S<strong>and</strong>berg AA, Rosenthal HE: Steroid<br />

receptors in exocrine gl<strong>and</strong>s:<br />

The pancreas <strong>and</strong> prostate. J Steroid<br />

Biochem 1979;11:293–299.<br />

56 Pousette A, Carlstrom K, Skolde<strong>for</strong>s<br />

H, Wilking N, Theve NO: Purification<br />

<strong>and</strong> partial characterization of a<br />

17-estradiol-binding macromolecule<br />

in the human pancreas. <strong>Cancer</strong><br />

Res 1982;42:633–637.<br />

57 Theve NO, Pousette A, Carlstrom<br />

K: Adenocarcinoma of the pancreas<br />

– a hormone sensitive tumor? A preliminary<br />

report on Nolvadex treatment.<br />

Clin Oncol 1983;9:193–197.<br />

58 Wong A, Chan A: Survival benefit of<br />

tamoxifen therapy in adenocarcinoma<br />

of pancreas. A case-control<br />

study. <strong>Cancer</strong> 1993;71:2200–2203.<br />

59Lewin MJM: The somatostatin receptors<br />

in the GI tract. Annu Rev<br />

Physiol 1992;54:455–469.<br />

60 Lamberts SWJ, Krenning EP, Reubi<br />

JC: The role of somatostatin <strong>and</strong> its<br />

analogs in the diagnosis <strong>and</strong> treatment<br />

of tumors. Endocr Rev 1991;<br />

12:450–458.<br />

61 Arnold R, Benning R, Neuhaus R,<br />

Rolwage M, Trautmann ME: Gastroenteropancreatic<br />

endocrine tumors:<br />

Effect of s<strong>and</strong>ostatin on tumor<br />

growth. Digestion 1993;<br />

54(suppl 1):72–75.<br />

62 Schally AV: Oncological applications<br />

of somatostatin analogues.<br />

<strong>Cancer</strong> Res 1988;48:6977–6985.<br />

63 Johnson LR: Effects of gastrointestinal<br />

hormones on pancreatic growth.<br />

<strong>Cancer</strong> 1981;47:1640–1645.<br />

64 Comaru-Schally M, Schally AV:<br />

LH-RH agonists as adjuncts to somatostatin<br />

analogs in the treatment<br />

of pancreatic cancer; in Lunefield B,<br />

Vickery B (eds): International Symposium<br />

on Gn-RH Analogues in<br />

<strong>Cancer</strong> <strong>and</strong> Human Reproduction.<br />

Boston, Kluwer Academic Publishers,<br />

1990, pp 203–210.<br />

65 Konturek SJ, Bilski J, Jaworek J,<br />

Tasler J, Schally AV: Comparison of<br />

somatostatin <strong>and</strong> its highly potent<br />

hexa- <strong>and</strong> octapeptide analogs on<br />

exocrine <strong>and</strong> endocrine pancreatic<br />

secretion. Proc Soc Exp Biol Med<br />

1988;187:241–249.<br />

66 Stoscheck CM, King LE Jr: Role of<br />

epidermal growth factor in carcinogenesis.<br />

<strong>Cancer</strong> Res 1986;46:1030–<br />

1037.<br />

67 Goustin AS, Leof EB, Shipley GS,<br />

Moses HL: Growth factors <strong>and</strong> cancer.<br />

<strong>Cancer</strong> Res 1986;46:1015–<br />

1029.<br />

68 Korc M, Magnum BE: Recycling of<br />

epidermal growth factor in a human<br />

pancreatic carcinoma cell line. Proc<br />

Natl Acad Sci USA 1985;82:6172–<br />

6175.<br />

69Lamberts SWJ, Koper JW, Reubi<br />

JC: Potential role of somatostatin<br />

analogues in the treatment of cancer.<br />

Eur J Clin Invest 1987;17:281–<br />

287.<br />

70 Pollak M, Constantino J, Polychronakos<br />

C, Blauer S, Guyda H, Redmond<br />

C, Fisher B, Margolese R: Effect<br />

of tamoxifen on serum insulinlike<br />

growth factor I levels in stage 1<br />

breast cancer patients. J Natl <strong>Cancer</strong><br />

Inst 1990;82:1693–1697.<br />

71 Pollak M, Polychronakos C, Guyda<br />

H: <strong>Somatostatin</strong> analogue SMS<br />

201–995 reduces serum IGF levels<br />

in patients with neoplasms potentially<br />

dependent on IGF-1. Anticancer<br />

Res 1989;9:889–891.<br />

72 Huynh H, Pollak M: Enhancement<br />

of tamoxifen-induced suppression<br />

of insulin-like growth factor I gene<br />

expression <strong>and</strong> serum level by a somatostatin<br />

analogue. Biochem Biophys<br />

Res Commun 1994;203/1:<br />

253–259.<br />

73 Poston GJ, Townsend CM Jr, Rajaraman<br />

S, Thompson JC, Singh P:<br />

Effect of somatostatin <strong>and</strong> tamoxifen<br />

on the growth of human pancreatic<br />

cancers in nude mice. Pancreas<br />

1990;5:151–157.<br />

74 Klijn JGM, Hoff AM, Planting AS,<br />

et al: <strong>Treatment</strong> of patients with<br />

metastatic pancreatic <strong>and</strong> gastrointestinal<br />

tumors with the somatostatin<br />

analogue S<strong>and</strong>ostatin: A phase II<br />

study including endocrine effects.<br />

Br J <strong>Cancer</strong> 1990;62:627–630.<br />

75 Yamada Y, Post SR, Wang K, Tager<br />

HS, Bell GI, Seino S: Cloning <strong>and</strong><br />

functional characterization of a<br />

family of human <strong>and</strong> mouse somatostatin<br />

receptors expressed in<br />

brain, gastrointestinal tract, <strong>and</strong> kidney.<br />

Proc Natl Acad Sci USA 1992;<br />

89:251–255.<br />

76 Yamada Y, Reisine T, Law S, Ihara<br />

Y, Kubota A, Kagimoto S, Seino M,<br />

Seino Y, Bell GI, Seino S: <strong>Somatostatin</strong><br />

receptors, an exp<strong>and</strong>ing gene<br />

family: Cloning <strong>and</strong> functional characterization<br />

of human SSTR3, a<br />

protein coupled to adenylyl cyclase.<br />

Mol Endocrinol 1992;6:2136–2142.<br />

77 Xu Y, Song J, Bruno JF, Berelowitz<br />

M: Molecular cloning <strong>and</strong> sequencing<br />

of a human somatostatin receptor,<br />

hSSTR4. Biochem Biophys Res<br />

Commun 1993;193:648–652.<br />

78 Yamada Y, Kagimoto S, Kubota A,<br />

Yasuda K, Masuda K, Someya Y,<br />

Ihara Y, Li Q, Imura H, Seino S, Seino<br />

Y: Cloning, functional expression<br />

<strong>and</strong> pharmacological characterization<br />

of a fourth (hSSTR4) <strong>and</strong><br />

fifth (hSSThS) human somatostatin<br />

receptor subtype. Biochem Biophys<br />

Res Commun 1993;195:844–852.<br />

79Bell GI, Reisine T: Molecular biology<br />

of somatostatin receptors. Trends<br />

Neurosci 1993;16:34–38.<br />

80 Hoyar D, Bell GI, Berelowitz M,<br />

Epelbaum J, Feniuk W, Humphrey<br />

PPA, O’Carroll A-M, Patel YC,<br />

Schonbrunn A, Taylor JE, Reisine<br />

T: Classification <strong>and</strong> nomenclature<br />

of somatostatin receptors. Trends<br />

Pharmacol Sci 1995;16:86–88.<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 147


81 Buscail L, Delesque N, Esteve J-P,<br />

Saint-Laurent N, Prats H, et al:<br />

Stimulation of tyrosine phosphatase<br />

<strong>and</strong> inhibition of cell proliferation<br />

by somatostatin analogues: Mediation<br />

by human receptor subtypes<br />

SSTR1 <strong>and</strong> SSTR2. Proc Natl Acad<br />

Sci USA 1994:91:2315–2319.<br />

82 Buscail L, Esteve J-P, Saint-Laurent<br />

N, Bertr<strong>and</strong> V, Resine T, O’Carroll<br />

AM, Bell GI, et al: Inhibition of cell<br />

proliferation by the somatostain analogue<br />

RC-160 is mediated by somatostatin<br />

receptor subtypes SSTR2<br />

<strong>and</strong> SSSTR5 through different<br />

mechanisms. Proc Natl Acad Sci<br />

USA 1995;92:1580–1584.<br />

83 Raynor K, Murphy WA, Coy DH,<br />

Taylor JE, Moreau J-P, Yasuda K,<br />

Bell GI, Reisine T: Cloned somatostatin<br />

receptors: Identification of<br />

subtype selective peptides <strong>and</strong> demonstration<br />

of high affinity linear<br />

peptides. Mol Pharmacol 1993;43:<br />

838–844.<br />

84 O’Carroll A-M, Raynor K, Lolait SJ,<br />

Reisine T: Characterization of<br />

cloned human somatostatin receptor<br />

SSTR5. Mol Pharmacol 1994;<br />

46:291–298.<br />

85 Patel YC, Srikant CB: Subtype selectivity<br />

of peptide analogs <strong>for</strong> all<br />

five cloned human somatostatin receptors<br />

(hsstr 1–5). Endocrinology<br />

1994;135:2814–2817.<br />

86 Miller GV, Preston SR, Woodhouse<br />

LF, Farmery SM: <strong>Somatostatin</strong><br />

binding in human gastrointestinal<br />

tissues: Effect of cations <strong>and</strong> somatostatin<br />

analogues. Gut 1993;34:<br />

1351–1356.<br />

87 Liebow C, Reilly C, Serrano M,<br />

Schally AV: <strong>Somatostatin</strong> analogues<br />

inhibit growth of pancreatic cancer<br />

by stimulating tyrosine phosphatase.<br />

Proc Natl Acad Sci USA 1989;<br />

86:2003–2007.<br />

88 Patel YC, Panetta R, Escher E,<br />

Greenwood M, Srikant CB: Expression<br />

of multiple somatostatin receptor<br />

genes in AtT–20 cells. Evidence<br />

<strong>for</strong> a novel somatostatin–28 selective<br />

receptor subtype. J Biol<br />

Chem 1994;269:1506–1509.<br />

89Upp JR, Olson D, Polson FJ, et al:<br />

Inhibition of growth of two human<br />

pancreatic adenocarcinomaa in vivo<br />

by somatostatin analog SMS 201-<br />

995. Am J Surg 1988;155:29–35.<br />

90 Liebow C, Reilly C, Serrano M,<br />

Schally AV: <strong>Somatostatin</strong> analoges<br />

inhibit growth of pancreatic cancer<br />

by stimulating tyrosine phosphatase.<br />

Proc Natl Acad Sci USA 1989;<br />

86:2003–2007.<br />

91 Fisher WE, Boros LG, O’Dorisio<br />

MS, O’Dorisio TM, Schirmer WJ:<br />

<strong>Somatostatin</strong> receptor status of pancreatic<br />

adenocarcinoma predicts response<br />

to somatostatin therapy in<br />

vitro <strong>and</strong> in vivo. Surg Forum 1995;<br />

46:137–140.<br />

92 Fisher WE, Muscarella P, O’Dorisio<br />

TP, O’Dorisio MS, Kim JA, Doran<br />

TA, et al: Expression of the somatostatin<br />

receptor subtype-2 gene predicts<br />

response of human pancreatic<br />

cancer to octreotide. Surgery 1996;<br />

120:234–240.<br />

93 Liebow C, Hierowski M, DuSapin<br />

K: Hormonal control of pancreatic<br />

cancer growth. Pancreas 1986;l:44–<br />

48.<br />

94 Weckbecker G, Liu R, Tolcsvai L:<br />

Inhibitory effect of the somatostatin<br />

analog octreotide (SMS-201-995) on<br />

the growth of human <strong>and</strong> animal<br />

tumors in rodent models. Contrib<br />

Oncol 1992;42:362–369.<br />

95 Radulovic S, Comaru-Schally AM,<br />

Milovanovic S, Schally AV: <strong>Somatostatin</strong><br />

analogue RC-l60 <strong>and</strong> LH-<br />

RH antagonist SB-75 inhibit growth<br />

of MIA PaCa-2 human pancreatic<br />

cancer xenografts in nude mice.<br />

Pancreas 1993;8:88–97.<br />

96 Singh P, Townsend CM Jr, Poston<br />

GJ, Reubi JC: Specific binding of<br />

cholecystokinin, estradiol <strong>and</strong> somatostatin<br />

to human pancreatic xenografts.<br />

J Steroid Biochem Mol<br />

Biol 1991;39:759–767.<br />

97 Vidal C, Rauly I, Zeggari M, Delesque<br />

N, Esteve J-P, Saint-Laurent N,<br />

Vaysse N, Susini C: Up-regulation<br />

of somatostatin receptors by epidermal<br />

growth factor <strong>and</strong> gastrin in<br />

pancreatic cells. Mol Pharmacol<br />

1994;45:97–104.<br />

98 Tahiri-Jouiti N, Cambillau C, Viguerie<br />

N, Vidal C, Buscail L, Saint-<br />

Laurent N, Vaysse N, Susini C:<br />

Characterization of a membrane tyrosine<br />

phosphatase activity in AR4-<br />

2J cells: Regulation by somatostatin.<br />

Am J Physiol 1992;262:G1007–<br />

G1014.<br />

148 Chemotherapy 2001;47(suppl 2):134–149 Rosenberg<br />

99 Viguerie N, Tahiri-Jouti N, Ayral<br />

A-M, Cambillau C, Scemama J-L,<br />

et al: Direct inhibitory effects of a<br />

somatostatin analog, SMS 201-<br />

995, on AR4-2J cell proliferation<br />

via pertussis toxin-sensitive guanosine<br />

triphosphate-binding protein-independent<br />

mechanism. Endocrinology<br />

1989;124:1017–1025.<br />

100 Lamberts SWJ, von Koetsveld P,<br />

Hofl<strong>and</strong> LJ: The interrelationship<br />

between the antimitotic action of<br />

the somatostatin analog octroetide<br />

<strong>and</strong> that of cytostatic drugs <strong>and</strong><br />

suramin. Int J <strong>Cancer</strong> 1991;48:<br />

938–941.<br />

101 Weckbecker G, Raulf F, Tolcsvai<br />

L, Bruns C: Potentiation of the<br />

anti-proliferative effects of anticancer<br />

drugs by octreotide in vitro<br />

<strong>and</strong> in vivo. Digestion 1996;<br />

57(suppl 1):22–28.<br />

102 Radulovic S, Nagy A, Szoke B, et<br />

al: Cytotoxic analog of somatostatin<br />

containing methotrexate inhibits<br />

growth of MIA PaCa-2 human<br />

pancreatic xenografts in nude<br />

mice. <strong>Cancer</strong> Lett 1992;62:263–<br />

271.<br />

103 Vennin PH, Peyret JP, Bonneterre<br />

J: Effect of the long-acting somatostatin<br />

analogue SMS 201–995 in<br />

advanced breast cancer. Anticancer<br />

Res 1989;9:153- 156.<br />

104 Guliana JM, Guillausseau PJ, Caron<br />

J, Siame-Mourot C, Calmettes<br />

C, Modigliant E: Effects of shortterm<br />

subcutaneous administration<br />

of SMS 201-995 on calcitonin plasma<br />

levels in patients suffering<br />

from medullary thyroid carcinoma.<br />

Horm Metab Res 1989;21:<br />

584–586.<br />

105 Kraenzlin ME, Ch’ng JC, Wood<br />

SM, Carr DH, Bloom SR: Longterm<br />

treatment of a VIPoma with<br />

somatostatin analogue resulting in<br />

remission of symptoms <strong>and</strong> possible<br />

shrinkage of metastases. Gastroenterology<br />

1985;88:185–187.<br />

106 Parmar H, Bogden A, Mollard M,<br />

de Rouge B, Phillips RH, Lightman<br />

SL: <strong>Somatostatin</strong> <strong>and</strong> somatostatin<br />

analogues in oncology.<br />

<strong>Cancer</strong> Treat Rev 1989;16:95–<br />

115.


107 Morisset J, Genik P, Lord A, Solomon<br />

TE: Effects of chronic administration<br />

of somatostatin on rat exocrine<br />

pancreas. Regul Pept 1982;<br />

4:49–58.<br />

108 Fekete M, Zalatnai A, Comura-<br />

Schally AM, Schally AV: Membrane<br />

receptors <strong>for</strong> peptides in experimental<br />

<strong>and</strong> human pancreatic<br />

cancers. Pancreas 1989;4:521–<br />

528.<br />

109Davies NM, Kapur P, Gillespie J,<br />

Schally AV, Guillou PJ, Poston<br />

GJ: Inhibitory effect of somatostatin<br />

analog RC-160 on EGF- <strong>and</strong><br />

trans<strong>for</strong>ming growth factor alpha<br />

(TGF-·)-stimulated pancreatic<br />

cancer growth in vivo. Br J <strong>Cancer</strong><br />

1991;64(suppl 15):4.<br />

110 Schally AV, Srkalovic G, Szende<br />

B, Redding TW, Janaky T, Juhasz<br />

A, et al: Antitumor effects of analogs<br />

of LH-RH <strong>and</strong> somatostatin:<br />

Experimental <strong>and</strong> clinical studies.<br />

J Steroid Biochem Mol Biol 1990;<br />

37:1061–1067.<br />

111 Szende B, Srkalovic G, Schally<br />

AV, Lapis K, Groot K: Inhibitory<br />

effects of analogs of luteinizing<br />

hormone-releasing hormone <strong>and</strong><br />

somatostatin on pancreatic cancers<br />

in hamsters. <strong>Cancer</strong> 1990;65:<br />

2279–2290.<br />

112 Paz-Bouza JR, Redding TW,<br />

Schally AV: <strong>Treatment</strong> of nitrosamine-induced<br />

pancreatic tumors<br />

in hamsters with analogues of somatostatin<br />

<strong>and</strong> luteinizing hormone-releasing<br />

hormone. Proc<br />

Natl Acad Sci USA 1987;84:1112–<br />

1116.<br />

113 Poston GJ, Gillespie J, Guillou PJ:<br />

Biology of pancreatic cancer. Gut<br />

1991;32:800- 812.<br />

114 Szende B, Zalatnai A, Schally AV:<br />

Programmed cell death (apoptosis)<br />

in pancreatic cancer of hamsters<br />

after administration of analogs of<br />

luteinizing hormone releasing hormone<br />

<strong>and</strong> somatostatin. Proc Natl<br />

Acad Sci USA 1989;86:1643–<br />

1647.<br />

115 Klijn JGM, Setyono-Han B, Bakker<br />

GH, Henkelman MS, Portengen<br />

H, Foekens JA: Effects of somatostatin<br />

analog (S<strong>and</strong>ostatin)<br />

treatment in experimental <strong>and</strong> human<br />

cancer; in Klijn JGM, Paridaens<br />

R, Foekens JA (eds): Hormonal<br />

Manipulation of <strong>Cancer</strong>:<br />

Peptides, Growth Factors <strong>and</strong><br />

New (Anti)steroidal agents.<br />

EORTC Monograph Series. New<br />

York, Raven Press, 1987, vol 19,<br />

pp 459–468.<br />

116 Friess H, Büchler MW, Beglinger<br />

Ch, Weber A, Kunz J, Fritsch K,<br />

Beger HG: Low dose octreotide<br />

treatment is not effective in patients<br />

with advanced pancreatic<br />

cancer. Pancreas 1993;8:540–545.<br />

117 Friess H, Büchler MW, Ebert M,<br />

Malfertheiner P, Dennler HJ, Beger<br />

HG: <strong>Treatment</strong> of advanced<br />

pancreatic cancer with high dose<br />

octreotide. Int J Pancreatol 1993;<br />

14:290–291.<br />

118 Ebert M, Friess H, Beger H, et al:<br />

Role of octreotide in the treatment<br />

of pancreatic cancer. Digestion<br />

1994;55(suppl 1):48–51.<br />

119Scambia G, Benedetti BP, Baiocchi<br />

G, Perrone L, Iacobelli S, Mancuso<br />

S: Antiproliferative effects of<br />

somatostatin analog SMS 201-995<br />

on three human breast cancer cell<br />

lines. J <strong>Cancer</strong> Res Clin Oncol<br />

1988;144:106–108.<br />

120 Cascinu S, Del Ferro E, Catalano<br />

G: A r<strong>and</strong>omized trial of octreotide<br />

vs best supportive care only in<br />

advanced cancer patients refractory<br />

to chemotherapy. Br J <strong>Cancer</strong><br />

1995;71:97–101.<br />

121 Canobbio L, Boccardo F, Cannata<br />

D, Gallotti P, Epis R: <strong>Treatment</strong> of<br />

advanced pancreatic cancer with<br />

the somatostatin analogue BIM<br />

23014. <strong>Cancer</strong> 1992;69:648–650.<br />

122 Huguier M, Samama G, Testart J,<br />

et al: <strong>Treatment</strong> of adenocarcinoma<br />

of the pancreas with somatostatin<br />

<strong>and</strong> gonadoliberin (luteinizing<br />

hormone-releasing hormone).<br />

Am J Surg 1992;164:348–353.<br />

123 Weckbecker G, Tolcsvai L, Stolz<br />

B, Pollak M, Brun C: <strong>Somatostatin</strong><br />

analogue octreotide enhances the<br />

antineoplastic effects of tamoxifen<br />

<strong>and</strong> ovariectomy on 7,12-dimethylbenz(a)anthracene-induced<br />

rat<br />

mammary carcinomas. <strong>Cancer</strong><br />

Res 1994;54:6334–6337.<br />

124 Rosenberg L, Barkun AN, Denis<br />

MH, Pollak M: Low dose octreotide<br />

<strong>and</strong> tamoxifen in the treatment<br />

of adenocarcinoma of pancreas.<br />

<strong>Cancer</strong> 1995;75:23–28.<br />

125 Fazeny B, Baur M, Prohaska M,<br />

Hudec M, Kremnitzer M, Meryn<br />

S, Huber H, Grunt T, Tuchmann<br />

A, Dittrich C: Octreotide combined<br />

with goserelin in the therapy<br />

of advanced pancreatic cancer –<br />

Results of a pilot study <strong>and</strong> review<br />

of the literature. J <strong>Cancer</strong> Res Clin<br />

Oncol 1997;123:45–52.<br />

126 Reubi JC, Horisberger U, Essed<br />

CE, Jeekel J, Klijn JHG, Lamberts<br />

SWJ: Absence of somatostatin receptors<br />

in human exocrine pancreatic<br />

cancer. Gastroenterology<br />

1988;95:760–763.<br />

127 Buscail L, Saint-Laurent N,<br />

Chastre E, Vaillant J-C, Gespach<br />

C, Capella G, Kalthoff H, Lluis F,<br />

Vayesse N, Susini C: Loss of sst2<br />

somatostatin receptor gene expression<br />

in human pancreatic <strong>and</strong> colorectal<br />

cancer. <strong>Cancer</strong> Res 1996;56:<br />

1823–1827.<br />

128 Pinski J, Milovanovic S, Yano T,<br />

Hamaoui A, Radulovic S, Cai R-Z,<br />

et al: Biological activity <strong>and</strong> receptor<br />

binding characteristics of various<br />

human tumors of acetylated<br />

somatostatin analogs. Proc Soc<br />

Exp Biol Med 1992;200:49–56.<br />

Pancreatic <strong>Cancer</strong> <strong>and</strong> Octreotide Chemotherapy 2001;47(suppl 2):134–149 149


Chemotherapy 2001;47(suppl 2):150–161<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

Elias A. Kouroumalis<br />

Department of Gastroenterology, University Hospital, Heraklion, Greece<br />

Key Words<br />

Octreotide W Hepatocellular carcinoma W<br />

<strong>Somatostatin</strong> receptors W Liver tumors<br />

Abstract<br />

Inoperable liver tumors have an unfavorable<br />

natural course despite various therapeutic<br />

modalities. Octreotide, a somatostatin<br />

analog, has shown considerable antitumor<br />

activity on animal models of various hepatic<br />

tumors <strong>and</strong> on isolated cell culture lines. In<br />

this paper, a review of the experimental evidence<br />

is presented. Moreover clinical papers<br />

of case reports of uncontrolled studies of<br />

patients are also reviewed. The majority of<br />

clinical studies provide evidence of a clinical<br />

<strong>and</strong> biochemical response of liver endocrine<br />

tumors while regression of tumor size is a<br />

rare event. A r<strong>and</strong>omized controlled trial of<br />

octreotide in the treatment of advanced hepatocellular<br />

carcinoma has shown a significant<br />

survival benefit in the treated patients.<br />

Literature reports indicate a stimulatory ef-<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0150$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

fect of octreotide on Kupffer cells as a possible<br />

antitumor mechanism, but other antiproliferative<br />

actions of octreotide have been suggested<br />

but not proved. Finally the question of<br />

the presence <strong>and</strong> affinity of somatostatin receptors<br />

on liver tumor tissue is discussed. In<br />

conclusion, according to our experience, octreotide<br />

administration is the best available<br />

treatment <strong>for</strong> advanced inoperable hepatocellular<br />

carcinoma <strong>and</strong> future better patient<br />

selection, based on receptor subtypes, might<br />

further improve the results.<br />

Introduction<br />

Copyright © 2001 S. Karger AG, Basel<br />

Although octreotide appears to have considerable<br />

antimitotic activity in various nonendocrine<br />

tumors, its use in treating liver cancers<br />

is mostly limited to endocrine metastatic<br />

or primary liver neoplasms, with only limited<br />

experience in other hepatic primaries.<br />

E.A. Kouroumalis, MD, PhD<br />

Associate Professor of Gastroenterology<br />

Head, Department of Gastroenterology<br />

University Hospital, PO Box 1352, Heraklion Crete (Greece)<br />

Fax +30 81 542085, E-Mail Kouroum@danae.med.uoc.gr


Epidemiology<br />

Liver neoplasms are among the most common<br />

tumors in many parts of the world. Hepatocellular<br />

carcinomas <strong>and</strong> cholangiocarcinomas<br />

comprise most of the liver tumors,<br />

while endocrine tumors, hepatoblastomas,<br />

bile duct cystadenocarcinomas, angiosarcomas,<br />

malignant hemangioendotheliomas,<br />

rhabdomyosarcomas <strong>and</strong> primary lymphomas<br />

are much rarer <strong>and</strong> have been reported<br />

mostly as case reports. An extensive review<br />

of liver tumors has recently been published<br />

[1].<br />

Hepatocellular carcinomas have an annual<br />

occurrence of at least 1,000,000 new cases [2].<br />

The geographical distribution is highly uneven<br />

with three main groups of countries being<br />

identified in terms of incidence rates. The<br />

highest rates are found in southeast Asia <strong>and</strong><br />

tropical Africa with an incidence of as high as<br />

150,000 cases/year. The lower rates are found<br />

in the western countries, South America <strong>and</strong><br />

India, while Japan, Middle East <strong>and</strong> the Mediterranean<br />

countries belong to the intermediate<br />

group [3, 4]. Overall hepatocellular carcinoma<br />

is estimated to be the seventh most<br />

common cancer in men <strong>and</strong> the ninth in<br />

women [5]. However, in Japan hepatocellular<br />

carcinoma is the third most frequent cause of<br />

male cancer deaths [6]. Furthermore, in Japan<br />

the number of deaths due to hepatocellular<br />

carcinoma is increasing constantly [7].<br />

The reported increase in the hepatocellular<br />

carcinoma incidence in western countries is<br />

probably due to better diagnostic modalities<br />

<strong>and</strong> detection programs in patients with cirrhosis<br />

[8].<br />

It seems that there is no racial predisposition<br />

to liver cancer. Differences among races<br />

can be explained on the basis of differences in<br />

exposure to well-established risk factors, like<br />

HBV <strong>and</strong> HCV infection <strong>and</strong> aflatoxins or<br />

other chemicals [3]. A similar explanation can<br />

be given <strong>for</strong> the observed differences among<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

immigrants when compared to the population<br />

in their native country. It has been established<br />

that the incidence of hepatocellular carcinoma<br />

in immigrants from countries where the<br />

prevalence of the disease is high falls to that of<br />

their country of adoption. The mean age of<br />

patients is lower in high incidence areas, <strong>and</strong><br />

Africans seem to develop the tumor earlier in<br />

life than Asians.<br />

The incidence rates <strong>for</strong> cholangiocarcinoma<br />

are highest in southeast Asia, particularly<br />

Hong Kong <strong>and</strong> Thail<strong>and</strong> [9,10]. In these<br />

areas the high incidence of the disease is<br />

linked to infestations with liver flukes, like<br />

Clonorchis sinensis <strong>and</strong> Opisthorchis viverrini.<br />

In western countries, cholangiocarcinomas<br />

are much less common, but a significant incidence<br />

occurs in patients with inflammatory<br />

bowel disease <strong>and</strong> primary sclerosing cholangitis<br />

[11, 12].<br />

Among the rarer tumors of the liver, epidemiological<br />

data exists <strong>for</strong> hepatoblastomas.<br />

The majority of hepatoblastomas (90%) occur<br />

under 5 years of age <strong>and</strong> account <strong>for</strong> almost<br />

40% of all liver tumors in childhood but are<br />

responsible <strong>for</strong> only 0.2–5.8% of all infant<br />

malignancies [13, 14].<br />

Epidemiology of liver endocrine tumors<br />

has not been well established.<br />

Natural History<br />

Little is known about the natural history of<br />

untreated primary liver tumors, since some<br />

<strong>for</strong>m of treatment is usually attempted. Survival<br />

in cases of hepatocellular carcinoma is<br />

strongly related to the Okuda classification<br />

[15]. In a recent study in Crete, a median survival<br />

of 16, 7 <strong>and</strong> 2 months <strong>for</strong> Okuda stage I,<br />

II <strong>and</strong> III liver tumors, respectively, was observed<br />

<strong>for</strong> patients who did not receive any<br />

therapy [16]. These findings are in agreement<br />

with earlier studies [17, 18]. Interestingly in<br />

this study, HCV positivity was not a risk factor<br />

<strong>for</strong> decreased survival, while HBeAg or<br />

Chemotherapy 2001;47(suppl 2):150–161 151


anti-c positivity carried a high relative risk of<br />

3.8 <strong>and</strong> 3.4, respectively. This observation is<br />

in accord with the Japanese experience where<br />

HBV-related cancer has a reduced 3-year survival<br />

rate compared to HCV-related disease<br />

[19]. In our study albumin concentration was<br />

inversely related to mortality with an 11%<br />

decrease in the hazard rate <strong>for</strong> each unit<br />

increase of this protein [16]. The natural history<br />

of liver metastatic endocrine tumors is<br />

less well known. In a recent study of patients<br />

with histologically confirmed liver metastases<br />

of endocrine tumors (gastrinomas, carcinoids,<br />

nonfunctioning pancreatic tumors <strong>and</strong> calcitonin-secreting<br />

tumors), a 90% progression<br />

was observed within a mean of 11.5 months<br />

[20]. The natural history of cholangiocarcinomas<br />

is dismal. Most patients deteriorate<br />

steadily <strong>and</strong> die within a few months after<br />

diagnosis [21].<br />

Current Therapeutic Approach<br />

Undoubtedly, the best available treatment<br />

<strong>for</strong> all liver tumors is complete surgical resection.<br />

In a recent report, 532 cirrhotics<br />

with hepatocellular carcinoma were evaluated.<br />

Some <strong>for</strong>m of liver resection was feasible<br />

in 44.7% of them (238 patients). Hospital<br />

mortality was 4.6% <strong>and</strong> 5-year actuarial survival<br />

rate was 41.3% [22]. Forty-one patients<br />

in this series were transplanted with a 5-year<br />

actuarial survival rate of 58.1%. Moreover, in<br />

most countries a differential diagnosis of hepatocellular<br />

carcinoma is made when the tumor<br />

is already inoperable.<br />

Other therapeutic modalities <strong>for</strong> liver tumors<br />

include selective arterial embolization<br />

with either iodine-131 Lipiodol or Lipiodolepirubicin<br />

[23]. In a recent report [24], the<br />

median survival rates at 6 <strong>and</strong> 12 months following<br />

Lipiodol-epirubicin were less than 65<br />

<strong>and</strong> 50%, respectively. Only when embolization<br />

is per<strong>for</strong>med in stage I Okuda patients,<br />

the 12-month survival rates are greater than<br />

50% [25]. Moreover, following several embolizations,<br />

arterial obstruction may give rise to<br />

the development of small collateral arteries<br />

that make repeated attempts to embolize the<br />

tumor difficult. Development of these collaterals<br />

may render the tumors resistant to emboli<br />

<strong>and</strong> their growth continues despite repeated<br />

treatments [26]. Arterial chemoembolization<br />

has been used in the treatment of metastatic<br />

endocrine liver tumors as well as primary<br />

hepatocellular carcinoma [20]. In a large<br />

series of carcinoid tumors, 40 patients with<br />

bipolar hepatic disease underwent embolization<br />

as primary treatment, followed by octreotide<br />

administration [27]. The 5-year survival<br />

rate was 56% <strong>and</strong> was accompanied by markedly<br />

decreased 5-hydroxyindole-acetic acid<br />

(5-HIAA) levels.<br />

Percutaneous ethanol injection has also<br />

been used extensively to treat hepatocellular<br />

carcinomas, especially tumors not exceeding<br />

2–3 cm in diameter [28]. Histopathologic examination<br />

after therapy revealed that in many<br />

cases ethanol injection can completely destroy<br />

the tumor [29, 30]. In addition ethanol injection<br />

has achieved considerably high long-term<br />

survival rates. In a study of 162 hepatocellular<br />

carcinoma patients with a single nodule the<br />

1-, 2- <strong>and</strong> 3-year survival rates were 90, 80<br />

<strong>and</strong> 63%, respectively [30]. In another Italian<br />

study of ethanol injection of small tumors,<br />

survival rates were clearly associated with the<br />

Child-Pugh classification. Five-year survival<br />

rates of 47, 29 <strong>and</strong> 0% were reported <strong>for</strong> Child<br />

A, B <strong>and</strong> C, respectively [31]. Similar findings<br />

were found in 105 western patients with hepatocellular<br />

carcinoma smaller than 5 cm in<br />

diameter where the 5-year overall survival<br />

rate was 32% on Child A <strong>and</strong> B patients [32].<br />

However, local recurrences are frequent, particularly<br />

when the tumor size exceeds 3 cm in<br />

diameter [33]. Moreover, in our experience,<br />

ethanol injection of larger tumors is of limited<br />

value, mainly because the alcohol diffuses<br />

152 Chemotherapy 2001;47(suppl 2):150–161 Kouroumalis


apidly into the surrounding liver parenchyma,<br />

because these large liver cancers are well<br />

vascularized (unpubl. observations). Fewer reports<br />

exist on the use of ethanol injection in<br />

metastatic liver tumors [34, 35] <strong>and</strong> cholangiocarcinomas<br />

[36]. Long-term survivals have not<br />

been reported so far <strong>for</strong> these tumors, but surgical<br />

excision of some tumors after ethanol<br />

injection showed an extensive necrosis ranging<br />

from 50 to 90% of the tumor.<br />

Hepatocellular carcinomas are assumed to<br />

be radioresistant. However radiation therapy<br />

has been recently reported in 22 patients with<br />

hepatocellular cancer receiving a total dose of<br />

58–68 Gy [37]. In these patients the tumor<br />

size never exceeded that prior to radiotherapy<br />

except in 1 patient. The 1-year survival rate<br />

was 68%, but definitive conclusions on this<br />

<strong>for</strong>m of treatment cannot be reached, without<br />

further studies since the treated tumors in this<br />

group were very heterogeneous <strong>and</strong> additional<br />

modalities were also used. Three of the patients<br />

had multiple ethanol injections <strong>and</strong> 17<br />

others had undergone transarterial embolization<br />

be<strong>for</strong>e the radiation therapy. Only 2 patients<br />

had radiation as the only therapeutic<br />

modality.<br />

Hormonal manipulations have been used<br />

<strong>for</strong> the treatment of hepatocellular carcinomas.<br />

In a recent r<strong>and</strong>omized controlled study,<br />

80 patients with Okuda II <strong>and</strong> Okuda III<br />

hepatocellular carcinoma were r<strong>and</strong>omized to<br />

either tamoxifen treatment or no treatment.<br />

Tamoxifen-treated patients were reported to<br />

have a 22% 1-year survival compared to almost<br />

10% of the nontreated patients. Only<br />

minor side effects were attributable to tamoxifen<br />

[38].<br />

The anti<strong>and</strong>rogen flutamide has been<br />

shown to have no therapeutic benefit in advanced<br />

hepatocellular carcinoma. The median<br />

survival time of patients treated with the<br />

drug was 2–5 months [39]. In a large anti<strong>and</strong>rogen<br />

trial of 244 patients with hepatocel-<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

lular carcinoma there was no survival advantage<br />

<strong>and</strong> indeed survival rates seemed even<br />

better <strong>for</strong> the placebo group [40]. In cholangiocarcinomas<br />

no effective drug treatment<br />

has been reported.<br />

Chemotherapy has been unsuccessful in<br />

the treatment of hepatocellular carcinoma as<br />

emphasized by two recent reports using combination<br />

chemotherapy. In a pilot study, intrahepatic<br />

arterial chemotherapy with methotrexate,<br />

5-fluorouracil, cisplatin <strong>and</strong> subcutaneous<br />

interferon-·2b were used in 16 patients<br />

with advanced hepatocellular carcinoma with<br />

portal thrombosis [41]. Median survival was 7<br />

months only. In another phase II Italian multicenter<br />

study, 5-fluorouracil plus high-dose<br />

levofolinic acid <strong>and</strong> hydroxyurea were used in<br />

50 patients with hepatocellular carcinoma.<br />

Median survival was again a disappointing<br />

5.8 months <strong>and</strong> many patients developed<br />

mild leukopenia <strong>and</strong> thrombocytopenia [42].<br />

Scientific Background<br />

Presence of <strong>Somatostatin</strong> Receptors in<br />

Neoplastic Tissue<br />

The diverse biological effects of somatostatin<br />

are mediated through somatostatin receptors<br />

that are coupled to a variety of signal<br />

transduction pathways. These include adenylate<br />

cyclase, ionic conductance channels <strong>and</strong><br />

protein phosphatase [43]. Recently, five such<br />

receptors have been cloned. They belong to<br />

the family of G-protein-coupled receptors [44,<br />

45]. The SSTR-2 receptor has so far been best<br />

characterized, <strong>and</strong> its negative effect on cell<br />

growth has been established [45]. According<br />

to these recent findings, the tyrosine phosphatase<br />

SHP1 is an essential component of<br />

the inhibitory growth signalling mediated<br />

through this receptor [46]. In vitro studies<br />

have shown that octreotide demonstrates<br />

high affinity binding properties to SSTR-2,<br />

Chemotherapy 2001;47(suppl 2):150–161 153


SSTR-3 <strong>and</strong> SSTR-5 receptors, while no binding<br />

affinity is found towards receptors<br />

SSTR-1 <strong>and</strong> SSTR-4 [47–49]. There<strong>for</strong>e when<br />

octreotide receptor studies are per<strong>for</strong>med using<br />

labelled octreotide only three out of the<br />

five known receptors are identified.<br />

The presence of somatostatin receptors has<br />

been verified in many endocrine tumors by<br />

autoradiographic techniques. Moreover, the<br />

presence of somatostatin receptors has been<br />

extensively investigated in vivo by using<br />

111 In-labelled octreotide scintigraphy [50, 51].<br />

In a recent study of metastatic liver colorectal<br />

carcinomas, somatostatin receptors were<br />

identified in only 1 out of 10 patients included<br />

in that study. This study would suggest<br />

that either octreotide scintigraphy is not a sensitive<br />

method <strong>for</strong> detection of somatostatin<br />

receptors in these patients or that the tumors<br />

do not express such receptors [52].<br />

Interestingly, in a small series, octreotide<br />

scintigraphy revealed SSTR receptors in 57%<br />

of patients with metastatic liver neuroblastomas,<br />

<strong>and</strong> patients with receptors had a better<br />

outcome than those who were negative [53].<br />

Similar findings were found in a single case of<br />

neuroblastoma [54].<br />

In a large series of patients with the Zollinger-Ellison<br />

syndrome, octreotide scintigraphy<br />

proved to be very accurate in detecting extrahepatic<br />

disease but the specificity of liver hot<br />

spots was only 60% [55]. Similar findings<br />

were reported <strong>for</strong> carcinoid tumor, where extrahepatic<br />

abdominal metastases were easily<br />

detected by scintigraphy but the detection of<br />

liver tumors was not superior to CT or ultrasonography<br />

[56–58].<br />

In another study of 18 patients with hepatic<br />

neuroendocrine tumors, octreotide scintigraphy<br />

showed a sensitivity of 94%, indicating<br />

that at least in these tumors, there is a<br />

large number of receptors easily detected by<br />

111 In-octreotide [59]. In contrast, it seems that<br />

medullary thyroid carcinomas contain very<br />

few somatostatin receptors as identified by<br />

scintigraphy [60]. Since octreotide mostly<br />

binds to SSTR-2 receptors <strong>and</strong> to a lesser<br />

extent to SSTR-3 <strong>and</strong> SSTR-5 receptors, octreotide<br />

scintigraphy will demonstrate only<br />

the presence of these subtypes. Receptors in<br />

hepatocellular carcinomas were detected by a<br />

radiolig<strong>and</strong> method in our own study, but<br />

identification of the receptor subtypes was not<br />

done [61].<br />

Inhibition of Tumor Growth:<br />

In vitro Studies<br />

There are very few data from in vitro studies.<br />

In pancreatic tumor cells, somatostatin<br />

<strong>and</strong> analogs antagonize the mitogenic effect of<br />

growth factors, acting on tyrosine kinase receptors<br />

such as epidermal growth factor [62].<br />

In human hepatocellular cell cultures, a 4-fold<br />

increase in the expression of IGF-1 binding<br />

protein has been reported after incubating the<br />

cells with octreotide [63]. Since IGF-1 is considered<br />

to be a trophic factor in hepatocellular<br />

carcinomas, this upregulation suggests that<br />

any antimitotic effect of octreotide on hepatocarcinomas<br />

is not likely to be mediated via<br />

inhibition of trophic hormones. However,<br />

further studies need to be done.<br />

Inhibition of Tumor Growth with<br />

Octreotide: Animal Models<br />

Octreotide has been found effective in inhibiting<br />

tumor growth in a variety of experimental<br />

models. Early studies demonstrated<br />

that experimentally derived (fibrosarcoma<br />

<strong>and</strong> adenocarcinoma) liver metastases are inhibited<br />

by octreotide [64, 65]. In an interesting<br />

study, nude mice were xenografted with a<br />

neuroendocrine cell line <strong>and</strong> received treatment<br />

with either high-dose octreotide or interferon-·<br />

or a combination of both. A 3-fold<br />

increase of apoptotic cells was observed in the<br />

octreotide group while the interferon-· group<br />

showed no evidence of increased apoptosis.<br />

154 Chemotherapy 2001;47(suppl 2):150–161 Kouroumalis


However, tumor growth inhibition was more<br />

pronounced in the combination group [66].<br />

Hepatocarcinogenesis induced by nitrosomorpholine<br />

in rats was significantly less in animals<br />

treated with somatostatin [67]. Morris hepatoma<br />

3924A cells were implanted in the liver <strong>and</strong><br />

subcutaneous tissues of rats which were partially<br />

hepatectomized <strong>and</strong> treated with either<br />

octreotide or placebo. Partial hepatectomy significantly<br />

increased tumor growth but treatment<br />

with octreotide resulted in a 10-fold<br />

decrease in liver tumor growth <strong>and</strong> a 3-fold<br />

reduction in subcutaneous tumor growth compared<br />

to controls [68].<br />

Inhibition of liver regeneration after partial<br />

hepatectomy by octreotide has been reported<br />

in another study. Octreotide did not<br />

influence hepatic or portal blood flow. However,<br />

octreotide significantly increased reticuloendothelial<br />

system activity [69]. In the same<br />

study the growth of adenocarcinoma <strong>and</strong> fibrosarcoma<br />

cells implanted into the partially<br />

hepatoctomized rat livers was significantly<br />

decreased by octreotide. These results confirm<br />

earlier observations by the same group<br />

who demonstrated that blockade of the reticuloendothelial<br />

system resulted in an increased<br />

tumor growth while octreotide increased the<br />

activity of the reticuloendothelial system [70,<br />

71]. However, octreotide only partially reversed<br />

the inhibition of growth <strong>and</strong> development<br />

of hepatic metastases in rats with gadolinium<br />

chloride-induced blockade of the<br />

Kupffer cells. This observation suggests that<br />

apart from stimulation of reticuloendothelial<br />

system, octreotide might have other mechanisms<br />

of action in inhibiting the growth<br />

of hepatic tumors. Two other mechanisms<br />

might be implicated. First, a direct antiproliferative<br />

effect through receptor-mediated<br />

growth inhibition has been postulated but<br />

never proved. Second, a reduction in tumor<br />

blood flow is a further possibility whereby<br />

octreotide may inhibit the growth of hepatic<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

metastases. Interestingly, the results from the<br />

above mentioned group do not confirm earlier<br />

studies in experimental hepatic metastases<br />

attributing the arrest of tumor growth to<br />

a reduction of hepatic arterial flow [72]. It<br />

should be noted, however, that reduction of<br />

tumor blood flow by octreotide may be dependent<br />

on the vascularity of the tumor. Fibrosarcomas<br />

are well vascularized while adenocarcinomas<br />

are relatively avascular <strong>and</strong> this difference<br />

might account <strong>for</strong> the divergent results<br />

reported in these two studies. Moreover, since<br />

tumor blood vessels lack receptors <strong>for</strong> vasoactive<br />

substances, any change of blood flow<br />

could be indirect through an effect on blood<br />

flow of the normal liver vessels. The distribution<br />

of octreotide receptors in normal liver<br />

blood vessels has not been studied. Finally,<br />

another possibility <strong>for</strong> the antitumor effect of<br />

octreotide could be an indirect action through<br />

inhibition of hormonal trophic factors. Although<br />

this possibility has not been adequately<br />

explored, there is evidence that this might<br />

not be an important factor, as previously<br />

mentioned.<br />

Rats were subcutaneously inoculated with<br />

mammary adenocarcinoma cells. A combination<br />

treatment with somatostatin plus insulin,<br />

which greatly increased the insulin/glucagon<br />

ratio, resulted in a reversed cachexia of the<br />

host, an effect not found with somatostatin<br />

alone. No appreciable effect on tumor growth<br />

was reported [73].<br />

Clinical Studies<br />

Results from Case Reports<br />

Most, if not all case reports refer to metastatic<br />

liver disease either from primary endocrine<br />

malignancies or from adenocarcinoma<br />

of the large bowel. An extremely rare case of a<br />

28-year-old woman with a metastatic ACTHsecreting<br />

pituitary carcinoma has been re-<br />

Chemotherapy 2001;47(suppl 2):150–161 155


ported. Liver metastases from this ACTHsecreting<br />

pituitary carcinoma have been<br />

treated with octreotide <strong>for</strong> almost 2 years.<br />

Paradoxically ACTH production was stimulated<br />

by octreotide treatment [74].<br />

Despite this unexplained secretion of<br />

ACTH, the development of the disease was<br />

very slow. Un<strong>for</strong>tuantely, the influence of octreotide<br />

cannot be fully assessed since the<br />

patient had a complementary hepatic chemoembolization<br />

treatment.<br />

Most case reports deal with either metastatic<br />

or the rare primary liver carcinoids.<br />

Only 18 cases of primary liver carcinoids have<br />

been reported so far. One such case of a<br />

patient with a malignant carcinoid of the liver<br />

with 18 years of follow-up has recently been<br />

reported [75]. Artery chemoembolization, hepatic<br />

lobectomy <strong>and</strong> octreotide treatment<br />

have all been used <strong>for</strong> the treatment of carcinoid<br />

in this patient who continues to be<br />

asymptomatic at the time of publication. A<br />

5-year survival has also been reported in a<br />

patient with the carcinoid syndrome after octreotide<br />

administration [76]. Two cases of primary<br />

liver VIPomas treated with octreotide<br />

<strong>and</strong> followed by liver surgery also resulted<br />

in amelioration of symptoms. One of the<br />

patients experienced diarrhea, hypokalemia<br />

<strong>and</strong> hypercalcemia. All these manifestations<br />

were reversible after treatment. Again arterial<br />

liver embolization was also used limiting the<br />

assessment of octreotide benefit [77]. This<br />

study confirms earlier case reports of longterm<br />

successful treatment of hepatic VIPomas<br />

with octreotide [78–80]. Patients with metastatic<br />

medullary thyroid carcinomas in the<br />

liver treated with octreotide have been reported<br />

to have a good symptomatic <strong>and</strong> biochemical<br />

response [81] despite the fact that<br />

results from octreotide scintigraphy indicated<br />

a relative lack of receptors in this type of<br />

tumor [60]. However, caution should be exercised<br />

in interpreting this finding. As men-<br />

tioned be<strong>for</strong>e scintigraphy will only detect<br />

SSTR-2, SSTR-3 <strong>and</strong> SSTR-5 receptors to<br />

which octreotide mostly binds. Moreover,<br />

since this is not a sensitive method, relatively<br />

large doses of octreotide might act on low<br />

receptor levels, not detected by scintigraphy.<br />

From case reports the only conclusion that<br />

can be drawn is that octreotide provides an<br />

effective symptomatic control. There is only<br />

one case report of a 68-year-old male with<br />

advanced hepatocellular carcinoma, who was<br />

successfully treated with the long-acting somatostatin<br />

analog lanreotide [82]. No case<br />

reports of hepatocellular carcinoma treated<br />

with octreotide have been published.<br />

Clinical Trials<br />

Clinical trials of octreotide in liver tumors<br />

are limited <strong>and</strong> mostly uncontrolled. In an<br />

early study of octreotide effectiveness on liver<br />

metastases derived from endocrine tumors,<br />

a significant reduction in tumor blood<br />

flow was caused by octreotide in 8 patients<br />

[83]. This finding was associated with a slight<br />

temporary reduction of the size of the tumors.<br />

Octreotide has been used as a palliative<br />

treatment of metastatic carcinoids of the liver.<br />

In an uncontrolled trial, 40 patients with<br />

bipolar liver disease were treated this way.<br />

Urinary 5-HIAA levels were still reduced by<br />

55% after 71 B 11 months of follow-up. The<br />

5-year survival was 56%. Symptomatic relief<br />

was found in 85% of patients. Most deaths<br />

were related to cardiovascular incidents. For<br />

comparison, 14 patients were treated by surgical<br />

resection alone, which achieved anatomical<br />

<strong>and</strong> biochemical cure. 5-HIAA levels were<br />

normal after 69 B 6.2 months of follow-up<br />

<strong>and</strong> these patients were asymptomatic. Two<br />

out of 14 patiens died of causes unrelated to<br />

the tumor. Despite the fact that there was no<br />

control group <strong>for</strong> the octreotide group, it is<br />

concluded that administration of octreotide<br />

156 Chemotherapy 2001;47(suppl 2):150–161 Kouroumalis


offers a symptomatic <strong>and</strong> survival benefit<br />

[84]. This report confirms earlier trials. In the<br />

largest report from the Mayo Clinic, 66 patients<br />

with metastic carcinoid of the liver were<br />

treated with octreotide. Complete or nearcomplete<br />

control of symptoms of diarrhea<br />

was obtained in 77%, of flushing in 87% <strong>and</strong> a<br />

50% reduction of urinary 5-HIAA was obtained<br />

in 70% of patients. The mean survival<br />

was more than 3 years, which is longer than in<br />

previous reports with combination chemotherapy<br />

[85].<br />

In another smaller uncontrolled clinical<br />

trial, 10 patients with metastatic liver carcinoids<br />

were treated with artery embolization,<br />

intra-arterial 5-fluorouracil <strong>and</strong> octreotide. In<br />

60% of the patients the tumor size was reduced<br />

by 50%, while the mean survival was<br />

58 months [86].<br />

In a phase II clinical trial, patients with<br />

metastatic carcinoid liver tumors were treated<br />

with either 500 or 1,000 Ìg t.i.d. of octreotide.<br />

Some patients with metastatic medullary thyroid<br />

carcinoma, pancreatic islet cell tumors<br />

<strong>and</strong> Merkel cell carcinomas were also included.<br />

The carcinoid syndrome was documented<br />

in 16 patients <strong>and</strong> abnormal urinary<br />

5-HIAA excretion in 15. Median treatment<br />

duration was 5 months. Tumor regression,<br />

symptom response <strong>and</strong> biochemical response<br />

were evaluated. Symptomatic <strong>and</strong> biochemical<br />

responses were reported as satisfactory<br />

with an overall response of 73 <strong>and</strong> 77%,<br />

respectively. However, tumor regression was<br />

reported in only 3% of patients [87]. In 27<br />

patients, the disease was stabilized <strong>for</strong> at least<br />

6 months (range 6–32 months). The median<br />

survival <strong>for</strong> all patients was 22 months (range<br />

1–32 months).<br />

In the prospective German octreotide multicenter<br />

phase II trial, 200 Ìg octreotide t.i.d.<br />

were given daily <strong>for</strong> 1 year to 103 patients<br />

with advanced nonoperable endocrine tumors.<br />

Sixty-four patients had a functional<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

tumor. Most of the patients had either carcinoids<br />

or islet cell tumors. Fifty-two patients<br />

had a CT-confirmed tumor progression be<strong>for</strong>e<br />

treatment. Thirty-seven percent of patients<br />

from this group experienced stabilization<br />

of tumor growth <strong>for</strong> at least 3 months.<br />

Median duration of stable disease was 18<br />

months. Tumor regression has not been seen<br />

in any patient [88]. Different results have<br />

been reported in an uncontrolled trial of somatostatin<br />

in liver metastatic endocrine tumors,<br />

where no objective response was found<br />

[20]. In conclusion, octreotide seems to have a<br />

beneficial effect on symptomatology <strong>and</strong> biochemical<br />

response of endocrine tumors. It<br />

also seems to stabilize the tumor growth in<br />

some patients while tumor regression is not<br />

observed.<br />

There is only one r<strong>and</strong>omized controlled<br />

trial of octreotide in the management of hepatocellular<br />

carcinoma [61]. <strong>Somatostatin</strong> receptors<br />

were identified in liver biopsy tissue<br />

from various diseases, including hepatocellular<br />

carcinomas, using a radiolig<strong>and</strong> method.<br />

Fifty-eight patients, the majority of whom<br />

had Okuda stage II <strong>and</strong> III tumors, were r<strong>and</strong>omized<br />

to receive either no treatment or<br />

subcutaneous octreotide 250 Ìg twice daily.<br />

Treated patients had a significantly increased<br />

median survival (13 months) compared to<br />

controls (4 months) <strong>and</strong> a significantly increased<br />

cumulative survival rate at 6 <strong>and</strong> 12<br />

months (75 vs. 37% <strong>and</strong> 56 vs. 13%, respectively).<br />

Furthermore, the quality of life was<br />

improved in the treatment group. In 5 patients,<br />

small satellite tumors, below 3 cm in<br />

diameter, disappeared after 6–12 months of<br />

treatment. In 4 additional patients, the size of<br />

the tumor remained unchanged. In view of<br />

the results from this trial carried out in our<br />

unit [61], we currently believe that octreotide<br />

is the treatment of choice <strong>for</strong> inoperable hepatocellular<br />

carcinoma. Hitherto no trial has<br />

been done on cholangiocarcinomas.<br />

Chemotherapy 2001;47(suppl 2):150–161 157


Conclusions<br />

There is substantial evidence from experimental<br />

studies <strong>and</strong> from a controlled clinical<br />

trial in man that octreotide is a useful drug <strong>for</strong><br />

the treatment of not only metastatic liver<br />

endocrine tumors, but also of primary hepatocellular<br />

carcinoma, either as an adjuvant therapy<br />

after surgical resection, or a sole treatment.<br />

The relative lack of side effects of<br />

octreotide is particularly promising in this<br />

respect. Caution should be excercised in patients<br />

with diabetes since hyperglycemia can<br />

be a problem.<br />

<strong>Treatment</strong> of hepatic metastases derived<br />

from colon carcinoma also seems to be another<br />

area in which octreotide treatment is justified,<br />

although controlled trials in patients are<br />

necessary to assess the efficacy of this somatostatin<br />

analog in this indication.<br />

Evidence that the antiproliferative effect of<br />

octreotide on hepatic tumors is immune-mediated<br />

via the stimulation of the reticuloendothelial<br />

system of the liver has been carefully<br />

documented [70, 71]. However, other modes<br />

of action of octreotide on hepatic tumors<br />

require further investigation.<br />

Projections <strong>for</strong> the Future<br />

In a recent study, gallbladder visualization<br />

was obtained with octreotide scintigraphy, indicating<br />

that somatostatin receptors are also<br />

found in the gallbladder mucosa [89]. If this<br />

finding can be verified by more specific studies,<br />

octreotide may prove to be useful <strong>for</strong> the<br />

treatment of cholangiocarcinomas.<br />

Careful examination of the receptor subtypes<br />

in liver <strong>and</strong> bile duct tissue in health <strong>and</strong><br />

disease using more sophisticated techniques<br />

than octreotide scintigraphy is urgently required.<br />

Moreover, the mode of antiproliferative<br />

action of octreotide should be delineated.<br />

Besides the possible immune modulation<br />

through stimulatory effects on Kupffer cells,<br />

there are certain other possibilities that have<br />

not been adequately explored. There is also<br />

evidence <strong>for</strong> a possible effect of octreotide<br />

on hepatic tumor angiogenesis, but further<br />

studies are required. Similarly a direct antiproliferative<br />

effect through somatostatin receptors,<br />

or through inhibition of tumor trophic<br />

hormones should be examined. Moreover,<br />

a recent report indicates that a somatostatin<br />

analog activates hepatoma cell apoptosis<br />

in vitro in both drug-sensitive <strong>and</strong> drugresistant<br />

cell lines [90]. This is a possible<br />

mechanism of octreotide antitumor activity<br />

that should be further explored. Another area<br />

<strong>for</strong> future research is the development of longacting<br />

octreotide or somatostatin analogs,<br />

making protracted treatment with octreotide<br />

more acceptable to patients. In this respect<br />

S<strong>and</strong>ostatin-LAR <strong>and</strong> lanreotide development<br />

might prove of particular significance.<br />

Note added in proof<br />

Recent studies [91, 92] have explored the clinical<br />

potential of lanreotide (30 mg i.m. every 14 days) in<br />

the treatment of advanced hepatocellular carcinoma.<br />

Besides a successful case report [91], 38% of patients<br />

had stable disease while the remaining ones progressed<br />

during treatment. Since in vitro studies [92] clearly<br />

demonstrated the ability of lanreotide to decrease the<br />

S phase fraction along with induction of apoptosis in<br />

Hep G2 cells in a dose-dependent fashion, the disappointing<br />

clinical results could be ascribed to the use of<br />

suboptimal doses of the peptide.<br />

158 Chemotherapy 2001;47(suppl 2):150–161 Kouroumalis


References<br />

1 Anthony PP: Tumours <strong>and</strong> tumourlike<br />

lesions of the liver <strong>and</strong> biliary<br />

tract; in McSween RNM, Anthony<br />

PP, Scheuer PJ, Burt AD, Portmann<br />

BC (eds): Pathology of the Liver,<br />

ed 3. Edinburgh, Churchill Livingstone,<br />

1994, pp 635–711.<br />

2 Rustgi VK: Epidemiology of hepatocellular<br />

carcinoma. Gastroenterol<br />

Clin North Am 1987;16:545–551.<br />

3 Linsel A: Primary liver cancer: Epidemiology<br />

<strong>and</strong> etiology; in Waneb<br />

HJ (ed): Hepatic <strong>and</strong> Biliary <strong>Cancer</strong>.<br />

New York, Dekker, 1986, pp 3–15.<br />

4 Simonetti RG, Camma C, Fiorello<br />

P, Politi F, D’Amico G, Pagliaro L:<br />

Hepatocellular carcinoma. A worldwide<br />

problem <strong>and</strong> the major risk<br />

factors. Dig Dis Sci 1991;36:962–<br />

972.<br />

5 EI Refaie A, Savage K, Bhattacharya<br />

S, Khakoo S, Harrison TJ, El-<br />

Batanony M, Soliman ES, Nasr S,<br />

Mokhtar N, Amer K, Scheuer PJ,<br />

Dhillon AP: HCV associated hepatocellular<br />

carcinoma without cirrhosis.<br />

J Hepatol 1996;24:277–285.<br />

6 Trends in the Health of Nation: J<br />

Health Welfare Stat 1997;44:53–55.<br />

7 The Ministry of Health <strong>and</strong> Welfare<br />

in Japan: Vital Statistics of Japan<br />

1985, 1986, 1987. Tokyo, Ministry<br />

of Health <strong>and</strong> Welfare, 1987.<br />

8 Andersen IB, Sorensen TIA, Prener<br />

A: Increase in incidence of disease<br />

due to diagnostic drift: Primary liver<br />

cancer in Denmark 1943–85. Br<br />

Med J 1991;302:437–440.<br />

9 Belamaric J: Intrahepatic bile duct<br />

carcinoma <strong>and</strong> C. sinensis infection<br />

in Hong Kong. <strong>Cancer</strong> 1973;31:<br />

468–473.<br />

10 Srivatanakul P, Sontipong S, Chotiwan<br />

P, Parkin DM: Liver cancer in<br />

Thail<strong>and</strong>. Temporal <strong>and</strong> geographic<br />

variations. J Gastroenterol Hepatol<br />

1988;3:413–420.<br />

11 Rosen CB, Nagorney DM, Wiesner<br />

RH, Coffey RJ, La Russo NF:<br />

Cholangiocarcinoma complicating<br />

primary sclerosing cholangitis. Ann<br />

Surg 1991;213:21–25.<br />

12 D’Haens GR, Lashner BA, Hanauer<br />

SB: Pericholangitis <strong>and</strong> sclerosing<br />

cholangitis are risk factors <strong>for</strong> dysplasia<br />

<strong>and</strong> cancer in ulcerative colitis.<br />

Am J Gastroenterol 1993;88:<br />

1174–1178.<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

13 Lack EE, Neave C, Vawter GF: Hepatoblastoma.<br />

Am J Surg Pathol<br />

1982;6:599–612.<br />

14 Schmidt D, Harms D, Lang W: Primary<br />

hepatic tumours in childhood.<br />

Virchows Arch 1985;407:387–405.<br />

15 Okuda K, Obata H, Nakajima Y:<br />

Prognosis of primary hepatocellular<br />

carcinoma. Hepatolgy 1984;4:36–65.<br />

16 Kouroumalis EA, Skordilis PG,<br />

Mosch<strong>and</strong>rea J, Alex<strong>and</strong>rakis G,<br />

Charoulakis N, Tzardi M, Manousos<br />

ON: Natural history of advanced<br />

hepatocellular carcinoma in Crete.<br />

Association with hepatitis C virus.<br />

Eur J Gastroenterol Hepatol 1997;9:<br />

981–989.<br />

17 Okuda K, Liver <strong>Cancer</strong> Study<br />

Group of Japan: Primary liver cancers<br />

in Japan. <strong>Cancer</strong> 1980;45:<br />

2663–2669.<br />

18 Okuda K, Obata H, Nakajima Y:<br />

Prognosis of primary hepatocellular<br />

carcinoma. Hepatology 1984;4:36–<br />

65.<br />

19 Eto H, Toriyama K, Itakura H: A<br />

clinicopathological study of hepatocellular<br />

carcinoma in Nagasaki,<br />

South Western Japan: The association<br />

of hepatitis B <strong>and</strong> C viruses.<br />

Southeast Asian J Trop Med Public<br />

Health 1994;25:88–92.<br />

20 Skinazi F, Zins M, Menu Y, Bernades<br />

P, Ruszniewski P: Liver metastases<br />

of digestive endocrine tumours.<br />

Natural history <strong>and</strong> response<br />

to medical treatment. Eur J Gastroenterol<br />

Hepatol 1996;8:673–678.<br />

21 Roberts JW: Carcinoma of the extrahepatic<br />

bile ducts. Surg Clin<br />

North Am 1986;66:751–756.<br />

22 Mazziotti A, Grazi GL, Cavallari A:<br />

Surgical treatment of hepatocellular<br />

carcinoma on cirrhosis: A Western<br />

Experience. Hepatogastroenterology<br />

1998;45:1281–1287.<br />

23 Madden MV, Krige JEJ, Bailey S,<br />

Beningfield SJ, Geddes C, Werner<br />

ID, Ternblanche J: R<strong>and</strong>omized<br />

trial of targeted chemotherapy with<br />

lipiodol <strong>and</strong> 5-epidoxorubicin compared<br />

with symptomatic treatment<br />

<strong>for</strong> hepatoma. Gut 1993;34:1598–<br />

1600.<br />

24 Bhattacharya S, Norell JR, Dusheiko<br />

GM, Hilson AJ, Dick R, Hobbs<br />

KE: Epirubicin-Lipiodol chemotherapy<br />

in the treatment of unresectable<br />

hepatocellular carcinoma. <strong>Cancer</strong><br />

1995;76:2202–2210.<br />

25 Bronowicki JP, Boudjema K, Chone<br />

L, Nis<strong>and</strong> G, Basin C, Pflumio F, Uhl<br />

G, Wenger JJ, Jaeck D, Boissel P,<br />

Bigard MA, Gaucher P, Vetter D,<br />

Doffoel M: Comparison of resection,<br />

liver transplantation <strong>and</strong> transcatheter<br />

oily chemoembolization in the<br />

treatment of hepatocellular carcinoma.<br />

J Hepatol 1996;24:293–300.<br />

26 Ikeda K: Prediction, <strong>Diagnosis</strong> <strong>and</strong><br />

<strong>Treatment</strong> of Hepatocellular Carcinoma.<br />

Tokyo, Medical Review,<br />

1995, pp 68–83 <strong>and</strong> 98–115.<br />

27 Wangberg B, Westberg G, Tylen U,<br />

Tisell L, Jansson S, Nilsson O, Johansson<br />

V, Schersten T, Ahlman H:<br />

Survival of patients with disseminated<br />

midgut carcinoid tumors after<br />

aggressive tumor reduction. World J<br />

Surg 1996;20:892–899.<br />

28 Seki T, Nonaka T, Kubota Y, Mizuno<br />

T, Sameshima Y: Ultrasonically<br />

guided percutaneous ethanol injection<br />

therapy <strong>for</strong> hepatocellular carcinoma.<br />

Am J Gastroenterol 1989;84:<br />

1400–1407.<br />

29 Livraghi T, Bolondi L, Lazzaroni S:<br />

Percutaneous ethanol injection in<br />

the treatment of hepatocellular carcinoma<br />

in cirrhosis: A study on 207<br />

patients. <strong>Cancer</strong> 1992;69:925–929.<br />

30 Shiina S, Tagawa K, Niwa Y, et al:<br />

Percutaneous ethanol injection therapy<br />

<strong>for</strong> hepatocellular carcinoma:<br />

Results in 146 patients. Am J Radiol<br />

1993;160:1023–1028.<br />

31 Livraghi T, Giorgio A, Marin G, Bolondi<br />

L, Lazzaroni S: Hepatocellular<br />

carcinoma <strong>and</strong> cirrhosis in 746 patients:<br />

Long term results of percutaneous<br />

ethanol injection. Radiology,<br />

1995;197:101–108.<br />

32 Lencioni R, Bartolozzi C, Caramella<br />

D, Paolicchi A, Carrai M, Maltinti G,<br />

Capria A, Tafi A, Conte PF, Bevilacqua<br />

G: <strong>Treatment</strong> of small hepatocellular<br />

carcinoma with percutaneous<br />

ethanol injection: Analysis of prognostic<br />

factors in 105 western patients.<br />

<strong>Cancer</strong> 1995;76:1737–1746.<br />

33 Ishii H, Okada S, Nose H, Okusaka<br />

T, Yoshimori M, Takayama T, Kosuge<br />

T, Yamasaki S, Sakamoto M,<br />

Hiroyashi S: Local recurrence of hepatocellular<br />

carcinoma after percutaneous<br />

ethanol injection <strong>Cancer</strong><br />

1996;77:1792–1796.<br />

34 Livraghi T, Festi D, Monti F, Salmi<br />

A, Vettori C: US-guided percutaneous<br />

alcohol injection of small he-<br />

Chemotherapy 2001;47(suppl 2):150–161 159


patic <strong>and</strong> abdominal tumours. Radiology<br />

1986;161:309–312.<br />

35 Livraghi T, Vettori C, Lazzaroni S:<br />

Liver metastases: Results of percutaneous<br />

ethanol injection in 14 patients.<br />

Radiology 1991;179:709–<br />

712.<br />

36 Suyama Y, Horishi M, Shizumi Y,<br />

Ebisni S, Maekawa T, Miyoshi M:<br />

Clinicopathological effects of USguided<br />

intratumoral ethanol injection<br />

therapy to small liver cancers<br />

with special references to its adequate<br />

injected volume. Nippon Gan<br />

Chiryo Gakkai Shi 1988;23:1727–<br />

1731.<br />

37 Matsuura M, Nakayima N, Arai K,<br />

Ito K: The usefulness of radiation<br />

therapy <strong>for</strong> hepatocellular carcinoma.<br />

Hepatogastroenterology 1998;<br />

45:791–796.<br />

38 Mannesis EK, Giannoulis G, Zoumpoulis<br />

P, Vafiadou I, Hadjiyannis S:<br />

<strong>Treatment</strong> of hepatocellular carcinoma<br />

with combined suppression<br />

<strong>and</strong> inhibition of sex hormones: A<br />

r<strong>and</strong>omized controlled trial. Hepatology<br />

1995;21:1535–1542.<br />

39 Chao Y, Chan WK, Huang YS, Teng<br />

HC, Wang SS, et al: Phase II of flutamide<br />

in the treatment of hepatocellular<br />

carcinoma. <strong>Cancer</strong> 1996;77:<br />

635–639.<br />

40 Grimaldi C, Bleiberg H, Gay F,<br />

Messner M, Rougier P, Kok TC, Cirera<br />

L, Cervantes A, De Greve J,<br />

Paillot B, Buset M, Nitti D, Sahmound<br />

D, Duez N, Wils J: Evaluation<br />

of anti<strong>and</strong>rogen therapy in unrespectable<br />

hepatocellular carcinoma:<br />

Results of a European Organization<br />

<strong>for</strong> Research <strong>and</strong> <strong>Treatment</strong><br />

multicentric double-blind trial. J<br />

Clin Oncol 1998;16:411–417.<br />

41 Urabe T, Kaneko S, Matsushita E,<br />

Unoura M, Kobayashi K: Clinical<br />

pilot study of intrahepatic arterial<br />

chemotherapy with methotrexate, 5fluorouracil,<br />

cisplatin <strong>and</strong> subcutaneous<br />

interferon-alpha-2b <strong>for</strong> patients<br />

with locally advanced hepatocellular<br />

carcinoma. Oncology 1998;<br />

55, 1:39–47.<br />

42 Gebbia V, Maiello E, Serravezza G,<br />

Giotta F, Testa A, Borsellino N,<br />

Pezzella G, Colucci G: 5-Fluorouracil<br />

plus high dose levofolinic acid<br />

<strong>and</strong> oral hydroxyurea <strong>for</strong> the treatment<br />

of primary hepatocellular carcinomas:<br />

Results of a phase II multicenter<br />

study of the Southern Italy<br />

Oncology Group (GOIM). Anticancer<br />

Res 1999;19, 2B:1407–1410.<br />

43 Reisine T, Bell GI: Molecular biology<br />

of somatostatin receptors. Endocr<br />

Rev 1995;16:427–442.<br />

44 Liapakis G, Tallent M, Reisine T:<br />

Molecular <strong>and</strong> functional properties<br />

of somatostatin receptor subtypes.<br />

Metabolism 1996;45(suppl 1):12–<br />

13.<br />

45 Patel YC, Greenwood M, Panetta R,<br />

Hukovic N, Grigorakis S, Robertson<br />

LA, Srikant CB: Molecular biology of<br />

somatostatin receptor subtypes. Metabolism<br />

1996;45(suppl 1):31–38.<br />

46 Lopez F, Estere JP, Buscail L, Delesque<br />

N, Saint-Laurent N, Thereniau<br />

M, Nahmias C, Vaysse N, Susini C:<br />

The tyrosine phosphatase SHP-1 associates<br />

with the sst2 somatostatin<br />

receptor <strong>and</strong> is an essential component<br />

of sst2-mediated inhibitory<br />

growth signalling. J Biol Chem<br />

1997;272:24448–24454.<br />

47 Raynor K, Murphy WA, Coy DH,<br />

Taylor JE, Moreau JP, Yasuka K,<br />

Bell G I, Reisine T: Cloned somatostatin<br />

receptors: Identification of<br />

subtype-selective peptides <strong>and</strong> demonstration<br />

of high affinity binding<br />

of linear peptides. Mol Pharmacol<br />

1993;43:838–844.<br />

48 Patel YC, Srikant CB: Subtype specificity<br />

of peptide analogs <strong>for</strong> all<br />

five cloned human somatostatin<br />

receptors. Endocrinology 1994;135:<br />

2814–2817.<br />

49 Bruns C, Raulf F, Hoyer D, Schloos<br />

J, Lübbert H, Weckebecker G: Binding<br />

properties of <strong>Somatostatin</strong> receptor<br />

subtypes. Metabolism 1996;<br />

45:17–20.<br />

50 Kurtanam A, Raderer M, Muller C,<br />

Prokesch R, Kaserer K, Eibenberger<br />

K, Koperna K, Niederle B, Virgolini<br />

I: Vasoactive intestinal peptide <strong>and</strong><br />

somatostatin receptor scintigraphy<br />

<strong>for</strong> differential diagnosis of hepatic<br />

carcinoma metastasis. J Nucl Med<br />

1997;38:880–881.<br />

51 Lombrano MB, McCarthy K, Adams<br />

L, Neitzschaman H: Metastatic<br />

carcinoid tumor imaged with CT<br />

<strong>and</strong> a radiolabeled analog: A case<br />

report. Am J Gastroenterol 1997;92:<br />

513–515.<br />

52 Seifert JK, Gorges R, Bockisch A,<br />

Junginger T: 111-indium DTPA octreotide<br />

scintigraphy in colorectal<br />

liver metastases. Langenbecks Arch<br />

Chir 1997;382:332–336.<br />

53 Shalaby-Rana E, Majd M, Andrich<br />

MP, Morassaghi N: In-111 pentetreotide<br />

scintigraphy in patients<br />

with neuroblastoma. Comparison<br />

160 Chemotherapy 2001;47(suppl 2):150–161 Kouroumalis<br />

with I-131 MIBG, N-Myconcogene<br />

amplification <strong>and</strong> patient outcome.<br />

Clin Nucl Med 1997;22:315–319.<br />

54 Manil L, Edeline V, Michon J,<br />

Neuenschw<strong>and</strong>er S, Lequen H, Lavocat<br />

C, Zucker JM: Could somatostatin<br />

scintigraphy be superior to<br />

MIBG scan in the staging of stage IV<br />

neuroblastome (Pepper’s syndrome)?<br />

Clin Nucl Med 1996;21:530–533.<br />

55 Cadiot G, Bonnard G, Lebtahi R,<br />

Sarda L, Ruszniewski P, Le Guludec<br />

D, Mignon M: Usefulness of somatostatin<br />

receptor scintigraphy in<br />

the management of patients with<br />

Zollinger-Ellison syndrome. Gut<br />

1997;41:107–114.<br />

56 Kisker O, Weinel RJ, Geks J, Zacara<br />

F, Joseph K, Rothmund M: Value of<br />

somatostatin receptor scintigraphy<br />

<strong>for</strong> preoperative localization of carcinoids.<br />

World J Surg 1996;20:162–<br />

167.<br />

57 Schillaci O, Scopinaro F, Angeletti<br />

S, Talaro R, Danieli R, Annibale B,<br />

Gualdi G, Delle-Fare G: SPECT improves<br />

accuracy of somatostatin receptor<br />

scintigraphy in abdominal<br />

carcinoid tumours. J Nucl Med<br />

1996;37:1452–1456.<br />

58 Gibril F, Reynolds JC, Doppman<br />

JL, Chen CC, Venzon DJ, Termanini<br />

B, Weber HC, Stewart CA,<br />

Jensen RT: <strong>Somatostatin</strong> receptor<br />

scintigraphy: Its sensitivity compared<br />

with that of other imaging<br />

methods in detecting primary <strong>and</strong><br />

metastatic gastrinomas. A prospective<br />

study. Ann Intern Med 1996;<br />

125:26–34.<br />

59 Ramage JK, Williams R, Buxton-<br />

Thomas M: Imaging secondary neuroendocrine<br />

tumours of the liver:<br />

Comparison of I-123 metaiodobenzylguanidine<br />

(MIBG) an In-111-labelled<br />

octreotide (Octreoscan). Q J<br />

Med 1996;89:539–542.<br />

60 Baudin E, Lumbroso J, Schlumberger<br />

M, Lechere J, Giammarile C, Gardet<br />

P, Roche A, Travegli JP, Parmentier<br />

C: Comparison of octreotide<br />

scintigraphy <strong>and</strong> conventional imaging<br />

in medullary thyroid carcinoma.<br />

J Nucl Med 1996;37:912–916.<br />

61 Kouroumalis E, Skordilis P, Thermos<br />

K, Vasilaki A, Mosch<strong>and</strong>rea J,<br />

Manousos ON: <strong>Treatment</strong> of hepatocellular<br />

carcinoma with octreotide:<br />

A r<strong>and</strong>omized controlled<br />

study. Gut, 1998;42:442–447.<br />

62 Liebow C, Reilly C, Serrano M,<br />

Schally A: <strong>Somatostatin</strong> analogues


inhibit the growth of pancreatic cancer<br />

by stimulating tyrosine phosphatase.<br />

Proc Natl Acad Sci USA 1989;<br />

86:2003–2007.<br />

63 Ren SG, Ezzat S, Melmed S, Brannstein<br />

GD: <strong>Somatostatin</strong> analogue induces<br />

insulin-like growth factor<br />

binding protein-1 expression in human<br />

hepatoma cells. Endocrinology<br />

1992;131:2479–2481.<br />

64 Frizelle FA: Octreotide inhibits the<br />

growth <strong>and</strong> development of three<br />

types of experimental liver metastasis.<br />

Br J Surg 1995;82:1577.<br />

65 Davies N, Kynaston H, Yates J,<br />

Nott DM, Nash J, Taylor BA, Jenkins<br />

SA: Octroetide inhibits the<br />

growth <strong>and</strong> development of three<br />

types of experimental liver metastasis.<br />

Br J Surg 1995;82:840–843.<br />

66 Imam H, Eriksson B, Lumkinius A,<br />

Janson ET, Lindgren PG, Wil<strong>and</strong>er<br />

E, Oberg K: Induction of apoptosis<br />

in neuroendocrine tumours of the<br />

digestive system during treatment<br />

with somatostatin analogs. Acta Oncol<br />

1997;36:607–614.<br />

67 Nakaizumi A, Uehara H, Bara M,<br />

Irishi H, Tatsuta M: Inhibition by<br />

somatostatin of hepatocarcinogenesis<br />

induced by N-nitrosomorpholine<br />

in Sprague-Dawley rats. Carcinogenesis<br />

1993;14:2601–2604.<br />

68 Schinded DT, Grosfeld JL: Hepatic<br />

resection enhances growth of residual<br />

intrahepatic <strong>and</strong> subcutaneous<br />

hepatoma which is inhibited by octreotide.<br />

J Pediatr Surg 1997;32:<br />

995–998.<br />

69 Davies N, Yates J, Kynaston H,<br />

Taylor BA, Jenkins SA: Effects of<br />

octreotide on liver regeneration <strong>and</strong><br />

tumor growth in the regenerating<br />

liver. J Gastroenterol Hepatol 1997;<br />

12:47–53.<br />

70 Davies N, Kynaston H, Yates J,<br />

Nott DM, Jenkins SA, Taylor BA:<br />

Reticuloendothelial stimulation: Levamisole<br />

compared. Dis Colon Rectum<br />

1993;36:1054–1058.<br />

71 Davies N, Kynaston H, Yates J,<br />

Taylor BA, Jenkins SA: Octreotide,<br />

the reticuloendothelial system <strong>and</strong><br />

experimental liver tumours. Gut<br />

1995;36:610–614.<br />

72 Hemingway DM, Jenkins SA, Cook<br />

TG: The effects of s<strong>and</strong>ostatin (octreotide<br />

SMS 201-995) infusion on<br />

splanchnic <strong>and</strong> hepatic blood flow<br />

in an experimental model of hepatic<br />

metastases. Br J <strong>Cancer</strong> 1992;65:<br />

396–398.<br />

Octreotide <strong>for</strong> <strong>Cancer</strong> of the Liver <strong>and</strong><br />

Biliary Tree<br />

73 Bartlett DL, Charl<strong>and</strong> SL, Torosian<br />

MH: Reversal of tumor-associated<br />

hyperglucagonemia as treatment <strong>for</strong><br />

cancer cachexia. Surgery 1995;118:<br />

87–97.<br />

74 Lormean B, Miossec P, Sibony M,<br />

Valensi P, Attali JR: Adrenocorticotropin<br />

producing pituitary carcinoma<br />

with liver metastasis. J Endocrinol<br />

Invest 1997;20:230–236.<br />

75 Mehta DC, Warner RR, Parnes I,<br />

Weiss M: An 18-year follow-up of<br />

primary hepatic carcinoid with carcinoid<br />

syndrome. J Clin Gastroenterol<br />

1996;23:60–62.<br />

76 Degushi H, Deghuci K, Tsukada T,<br />

Murashima S, Iwasaki E, Tsuda M,<br />

Kobayashi T, Shirakawa S: Longterm<br />

survival in a patient with malignant<br />

carcinoid treated with high<br />

dose octreotide. Intern Med 1994;<br />

33:100–102.<br />

77 Lundstedt C, Linjawi T, Amin T:<br />

Liver vipoma: Report of two cases<br />

<strong>and</strong> literature review. Abdom Imaging<br />

1994;19:433–437.<br />

78 Fluckinger A, Schlup P: Long-term<br />

therapy of a metastasizing pancreatic<br />

vipoma using the somatostatin<br />

derivative octreotide. Schweiz Med<br />

Wochenschr 1993;122:1221–1223.<br />

79 Procaccioante F, Piccozzi P, Fantini<br />

A, Pacifici M, Di Nardo A, Ribotta<br />

G, Delle-Fare G, Catani M, Ruggeri<br />

S, Romeo F: Vipoma: Surgical treatment.<br />

Minerva Chir 1992;43:135–<br />

142.<br />

80 Bradley C, Haddock G, Pickard<br />

RG, Rankin EM: Metastatic vipoma,<br />

arising from a colonic primary<br />

tumor. Eur J Surg Oncol 1989;15:<br />

386–389.<br />

81 Mahler C, Verhelst J, de Longueville<br />

M, Harris A: Long-term treatment<br />

of metastatic medullary thyroid carcinoma<br />

with the somatostatin analogue<br />

octreotide. Clin Endocrinol<br />

(Oxf) 1990;33:261–269.<br />

82 Raderer M, Heina MH, Kurtaran A,<br />

Kornek GV, Valenck JB, Oberhuber<br />

G, Vorbech F, Virgolini I, Scheithauer<br />

W: Successful treatment of an<br />

advanced hepatocellular carcinoma<br />

with the long-acting somatostatin<br />

analog lanreotide. Am J Gastroenterol<br />

1999;94:278–279.<br />

83 Cho KJ, Vinik AI: Effect of somatostatin<br />

analogue (octreotide) on<br />

blood flow to endocrine tumours<br />

metastatic to the liver: Angiographic<br />

evaluation. Radiology 1990;177:<br />

549–553.<br />

84 Ahlman H, Westberg G, Wangberg<br />

B, Nilsson O, Tylen U, Schersten T,<br />

Tisill LE: <strong>Treatment</strong> of liver metastases<br />

of carcinoid tumors. World J<br />

Surg 1996;20:196–202.<br />

85 Krols LK: Therapy of the malignant<br />

carcinoid syndrome. Endocrinol<br />

Metab Clin North Am 1989;18:<br />

557–568.<br />

86 Diaco DS, Hajarizadeh H, Mueller<br />

CR, Fletcher WS, Pommier RF, Woltering<br />

EA: <strong>Treatment</strong> of metastatic<br />

carcinoid tumours using multimodality<br />

therapy of octreotide acetate,<br />

intra-arterial chemotherapy <strong>and</strong> hepatic<br />

arterial chemoembolization.<br />

Am J Surg 1995;169:523–528.<br />

87 di Bartolomeo M, Bajetta E, Buzzoni<br />

R, Mariani L, Carnaghi C,<br />

Somma L, Zilembo N, di Leo A:<br />

Clinical efficacy of octreotide in the<br />

treatment of metastatic neuroendocrine<br />

tumours. A study by the Italian<br />

Trials in Medical Oncology Group.<br />

<strong>Cancer</strong> 1996;77:402–408.<br />

88 Arnold R, Trautmann MF, Creutzfeldt<br />

W, Benning R, Benning M,<br />

Neuhaus C, Jurgensen R, Stein<br />

K, Schafer H, Bruns C, Dennler<br />

HJ, S<strong>and</strong>ostatin Multicentre Study<br />

Group: <strong>Somatostatin</strong> analogue octreotide<br />

<strong>and</strong> inhibition of tumour<br />

growth in metastatic endocrine gastroenterol<br />

pancreatic tumours. Gut<br />

1996;38:430–438.<br />

89 Kipper MS, Krohn LD: Gallbladder<br />

visualization at twenty-four hours<br />

with octreotide scintigraphy. Potential<br />

false-positive finding. Clin Nucl<br />

Med 1997;22:253–254.<br />

90 Diaconu CC, Szathmari M, Keri G,<br />

Venetianer A: Apoptosis is induced<br />

in both drug-sensitive <strong>and</strong> multidrug-resistant<br />

hepatoma cells by somatostatin<br />

analogue TT-232. Br J<br />

<strong>Cancer</strong> 1999;80:1197–1203.<br />

91 Raderer M, Hejna MH, Kurtaran A,<br />

Kornek GV, Valencak JB, Oberhuber<br />

G, Vorbeck F, Virgolini I,<br />

Scheithauer W: Successful treatment<br />

of an advanced hepatocellular carcinoma<br />

with the long-acting somatostatin<br />

analog lanreotide. Am J Gastroenterol<br />

1999;94:278–279.<br />

92 Raderer M, Hejna MH, Muller C,<br />

Kornek GV, Kurtaran A, Virgolini<br />

I, Fiebieger W, Hamilton G, Scheithauer<br />

W: <strong>Treatment</strong> of hepatocellular<br />

cancer with the long acting somatostatin<br />

analog lanreotide in vitro<br />

<strong>and</strong> in vivo. Int J Oncol 2000;16:<br />

1197−1201.<br />

Chemotherapy 2001;47(suppl 2):150–161 161


Chemotherapy 2001;47(suppl 2):162–196<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology:<br />

A Look to the Future<br />

Spencer A. Jenkins a Howard G. Kynaston b Nick Davies c<br />

John N. Baxter a David M. Nott d<br />

a Academic Department of Surgery, Postgraduate Medical School, Morriston Hospital,<br />

Swansea, b Department of Urology, University Hospital of Wales, Cardiff, c Departments of<br />

Surgery, Royal Bournemouth Hospital, <strong>and</strong> d Chelsea <strong>and</strong> Westminster Hospital, London, UK<br />

Key Words<br />

<strong>Somatostatin</strong> analogs W Neoplasms W<br />

Therapy W Review<br />

Abstract<br />

In the past 15 years considerable advances<br />

have been made in our underst<strong>and</strong>ing of the<br />

molecular pharmacology of the mechanisms<br />

whereby somatostatin <strong>and</strong> its analogs mediate<br />

their direct <strong>and</strong> indirect antineoplastic<br />

effects. However, some important issues remain<br />

to be resolved, in particular the functional<br />

roles of the individual somatostatin<br />

receptors (SSTR-1–5) in tumor tissue <strong>and</strong><br />

up- or downregulation of the hSSTRs with<br />

prolonged administration of somatostatin<br />

analogs. Answers to these questions are essential<br />

be<strong>for</strong>e we can maximize the therapeutic<br />

efficacy of somatostatin analogs in<br />

cancer. For example, is continuous administration<br />

more or less effective than intermittent<br />

therapy? The role of somatostatin<br />

analogs in the management of acromegaly<br />

ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

0009–3157/01/0478–0162$17.50/0<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

<strong>and</strong> to a lesser extent neuroendocrine tumors<br />

is firmly established. The development<br />

of depot preparations of all 3 somatostatin<br />

analogs currently available <strong>for</strong> clinical use<br />

will undoubtedly improve both patient compliance<br />

<strong>and</strong> quality of life in patients with<br />

these conditions. There are only likely to be<br />

minor differences in the therapeutic efficacy<br />

of octreotide, lanreotide <strong>and</strong> vapreotide<br />

since all three analogs exert the majority of<br />

their antineoplastic effects via hSSTR-2 <strong>and</strong><br />

hSSTR-5 <strong>and</strong> at the end of the day, price<br />

may well dictate which of these drugs oncologists<br />

use to provide symptomatic palliation<br />

of acromegaly <strong>and</strong> neuroendocrine tumors.<br />

Apart from some notable exceptions,<br />

somatostatin analog therapy has proven to<br />

be very disappointing in the management of<br />

advanced malignancy. Improvements in the<br />

management of solid tumors are likely to<br />

come only from combination therapy of somatostatin<br />

analogs with cytotoxic agents or<br />

other hormones in both advanced malignancy<br />

<strong>and</strong> in the adjuvant setting. Clinical<br />

S.A. Jenkins, MD<br />

Departments of General Surgery & Urology<br />

Ward 5-A, University Hospital of Wales<br />

Health Park, Cardiff CF 14 4XW (UK)<br />

Tel. +44 2920 744971, Fax +44 2920 744179


trials with clear-cut objective outcome measures<br />

<strong>and</strong> health-related quality of life assessment<br />

are needed to evaluate the therapeutic<br />

efficacy of combination treatment in<br />

advanced malignancy <strong>and</strong> as an adjuvant to<br />

surgery. Particular attention needs to be paid<br />

to possible adverse effects of somatostatin<br />

analog therapy on the immune response to<br />

cancer. Further studies are required to establish<br />

whether the adverse effects of somatostatin<br />

analog therapy alone or in combination<br />

with cytotoxics or other hormones can be<br />

reversed with appropriate immunomodulatory<br />

treatment. Targeted somatostatin analog<br />

radiotherapy <strong>and</strong> chemotherapy are currently<br />

being investigated <strong>and</strong> the results of<br />

these studies are awaited with interest. Novel<br />

approaches using combinations of somatostatin<br />

analogs with antiangiogenic drugs<br />

or gene therapy are of particular interest <strong>and</strong><br />

may provide important advances in the management<br />

of cancer in the not too distant<br />

future.<br />

Introduction<br />

Copyright © 2001 S. Karger AG, Basel<br />

Hypothalamic factors which regulate the<br />

secretion of growth hormone were first identified<br />

in 1968 by Krulich et al. [1]. Subsequently,<br />

in 1973, Brazeau et al. [2] identified a 14amino<br />

acid peptide inhibitor of growth hormone<br />

release-inhibiting factor (SRIF), the<br />

name subsequently being changed to somatostatin.<br />

Following production of a synthetic<br />

replicate of the native hormone [3] it soon<br />

became apparent that somatostatin had a<br />

wide variety of effects on the endocrine, gastrointestinal,<br />

immune, circulatory <strong>and</strong> central<br />

nervous systems.<br />

Although somatostatin is widely distributed<br />

throughout the body, immunochemical<br />

studies indicated that the peptide is confined<br />

to three cell types. <strong>Somatostatin</strong> is found in<br />

classic ‘open type’ endocrine cells from which<br />

it is directly excreted into the blood [4], paracrine<br />

cells with long cytoplasmic extensions<br />

which terminate on putative effector cells [5]<br />

<strong>and</strong> in neurones where it may function as a<br />

neurotransmitter [6–8] or is released from the<br />

nerve endings into the blood [4]. Thus, depending<br />

on its site of elaboration, somatostatin<br />

may function as a hormone, a neurohormone,<br />

a neurotransmitter or a parahormone.<br />

Consequently, the sites of elaboration<br />

<strong>and</strong> release of somatostatin in the body, together<br />

with a short half-life of approximately<br />

2 min [9], make it the almost ideal regulatory<br />

peptide.<br />

The plethora of effects mediated by somatostatin<br />

led to the suggestion that the hormone<br />

could have a therapeutically beneficial<br />

effect in a wide variety of indications. However,<br />

the short half-life of somatostatin restricts<br />

its application as a drug since it has to<br />

be administered by a continuous infusion<br />

to maintain a sustained biological effect <strong>and</strong><br />

hence a therapeutic response. Although intravenous<br />

administration of somatostatin does<br />

not present a problem in the management of<br />

patients hospitalized <strong>for</strong> acute indications<br />

such as variceal <strong>and</strong> nonvariceal upper gastrointestinal<br />

bleeding or acute pancreatitis, its<br />

short plasma half-life is a serious disadvantage<br />

to the exploitation of its full therapeutic<br />

potential in those conditions requiring longterm<br />

treatment. Not surprisingly there<strong>for</strong>e,<br />

following determination of the structure of<br />

somatostatin, peptide chemists began to try<br />

<strong>and</strong> synthesize more stable agonists with a<br />

longer duration of action. Furthermore, since<br />

somatostatin inhibits the release of such a vast<br />

array of hormones, the other <strong>and</strong> perhaps the<br />

most <strong>for</strong>midable challenge <strong>for</strong> the peptide<br />

chemists was to design analogs which were<br />

not only stable, but more selective in their<br />

mode of action.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 163


Mechanism of Action of<br />

<strong>Somatostatin</strong> <strong>and</strong> Its <strong>Analogs</strong><br />

In the past 15 years substantial advances<br />

have been made in elucidating the mechanisms<br />

whereby somatostatin <strong>and</strong> its analogs<br />

exert their antineoplastic effects [<strong>for</strong> comprehensive<br />

reviews see 10–15]. In brief, somatostatin<br />

<strong>and</strong> its analogs elicit direct antineoplastic<br />

effects via a complex variety of intracellular<br />

transduction pathways leading to inhibition<br />

of cell proliferation or by stimulation of<br />

apoptosis. Clearly, the direct antineoplastic<br />

effects of somatostatin <strong>and</strong> its analogs are<br />

confined to cells expressing hSSTRs. However,<br />

somatostatin <strong>and</strong> its analogs can exert<br />

antineoplastic effects via a number of indirect<br />

mechanisms including inhibition of the release<br />

<strong>and</strong> end-organ effects of trophic growth<br />

factors <strong>and</strong> hormones, angiogenesis <strong>and</strong> a reduction<br />

in tumor blood flow. Stimulation of<br />

the reticuloendothelial system (RES) appears<br />

to be a very important mechanism whereby<br />

somatostatin <strong>and</strong> its analogs inhibit the<br />

growth <strong>and</strong> development of liver tumors [11].<br />

The indirect antineoplastic effects of somatostatin<br />

<strong>and</strong> its analogs are independent of the<br />

presence of hSSTRs on the tumor cells. There<strong>for</strong>e,<br />

a possible beneficial effect of somatostatin<br />

analog therapy is not restricted to tumors<br />

expressing hSSTRs.<br />

Future Prospects<br />

Be<strong>for</strong>e discussing how it may be possible to<br />

improve somatostatin analog therapy of neoplasia<br />

it is important at the outset to emphasize<br />

the limitations of such treatment. Thus,<br />

apart from the stimulation of apoptosis by<br />

somatostatin <strong>and</strong> its analogs which could be<br />

considered cytotoxic, the remainder of their<br />

antineoplastic effects are cytotoxic. There<strong>for</strong>e,<br />

somatostatin analog therapy of neoplasia<br />

in symptomatic patients cannot be expected<br />

to be curative but only to provide remission<br />

<strong>and</strong> palliation of unpleasant <strong>and</strong> sometimes<br />

life-threatening side effects, thereby increasing<br />

survival <strong>and</strong> improving the quality of life.<br />

This is well illustrated by the experience of<br />

octreotide in the management of hypersecretory<br />

GEP tumors. Control of hormone hypersecretion<br />

by octreotide is observed in approximately<br />

55% of patients with vasoactive intestinal<br />

polypeptide (VIP)-omas <strong>and</strong> a reduction<br />

in the size of liver metastases has been reported<br />

in approximately 10%. Control of tumor<br />

growth occurs in approximately 50% of<br />

VIP-oma patients treated with octreotide.<br />

However, the reduction in the size of liver<br />

metastases secondary to VIP-omas is temporary<br />

<strong>and</strong> transient <strong>and</strong> control of tumor<br />

growth in response to octreotide therapy is<br />

lost after 8–16 months of treatment. Eventually,<br />

all the patients escape from the inhibitory<br />

effects of octreotide therapy on tumor progression<br />

<strong>and</strong> hormonal hypersecretion <strong>and</strong> become<br />

symptomatic. The major challenge <strong>for</strong><br />

somatostatin analog therapy in oncological indications<br />

is, there<strong>for</strong>e, to determine how we<br />

can take full advantage of their antineoplastic<br />

effects which, of course, will vary with the<br />

type of cancer being treated. For most solid<br />

tumors, surgery offers the only chance of cure<br />

<strong>and</strong> should be offered to all patients who are<br />

fit enough to undergo surgical intervention.<br />

However, <strong>for</strong> the majority of patients, a truly<br />

curative resection of their tumor(s) may not be<br />

feasible. In such patients debulking or neoadjuvant<br />

therapy to reduce the tumor burden<br />

may improve the response to adjuvant therapies<br />

<strong>and</strong>, in some cancers, increase the resectability<br />

rate. Finally, in patients with advanced<br />

metastatic malignancy, surgery may be contraindicated<br />

<strong>and</strong> in this situation effective palliation<br />

to maintain quality of life <strong>for</strong> as long<br />

as possible is the primary challenge <strong>for</strong> the<br />

oncologist. It seems likely, there<strong>for</strong>e, that the<br />

164 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


major role of somatostatin analog therapy is an<br />

adjuvant to surgery <strong>and</strong> in providing effective<br />

palliation <strong>for</strong> patients with advanced malignancies<br />

<strong>for</strong> most of the major cancers. There<br />

are, however, some exceptions to this rule. For<br />

example, elderly acromegalics with small tumors<br />

are particularly sensitive to the antineoplastic<br />

effects of octreotide. Furthermore, no<br />

escape from the inhibitory effects of octreotide<br />

on the hypersecretion of hormones by pituitary<br />

tumors occurs in the vast majority of acromegalics<br />

even after 10 years of therapy. There<strong>for</strong>e,<br />

in elderly acromegalics, somatostatin analog<br />

therapy is the treatment of choice since the<br />

life expectancy of such patients is unlikely to be<br />

influenced by their pituitary tumor.<br />

In order to maximize the therapeutic potential<br />

of somatostatin analog therapy in neoplasia<br />

there are a number of areas which<br />

require further investigation, in particular,<br />

the significance of heterogeneous expression<br />

of individual hSSTRs within <strong>and</strong> between different<br />

tumors, <strong>and</strong> their functional role <strong>and</strong><br />

their desensitization <strong>and</strong> downregulation.<br />

Expression <strong>and</strong> Functional Role of<br />

<strong>Somatostatin</strong> Receptors in Human<br />

Tumors<br />

Early competitive binding assays using radiolabelled<br />

somatostatin <strong>and</strong> its analogs <strong>and</strong><br />

autoradiography indicated that somatostatin<br />

receptors were expressed in a wide variety of<br />

normal tissue <strong>and</strong> tumors such as those of the<br />

nervous system, pituitary, gastrointestinal<br />

tract, pancreas, endocrine gl<strong>and</strong>s <strong>and</strong> lymphoid<br />

tissue [15–17]. The early binding studies<br />

indicated that the majority of human tumors<br />

expressed somatostatin receptors with<br />

a high binding affinity <strong>for</strong> somatostatin-14,<br />

somatostatin-28 <strong>and</strong> octreotide. However, a<br />

number of tumors appeared to express somatostatin<br />

receptors with a high binding affinity<br />

<strong>for</strong> somatostatin but a low binding affinity <strong>for</strong><br />

octreotide, e.g. pituitary adenomas, glial tumors,<br />

meningiomas, some GEP tumors, medullary<br />

thyroid carcinomas, ovarian cancers<br />

<strong>and</strong> breast tumors. Furthermore, normal <strong>and</strong><br />

cancerous gastric <strong>and</strong> colonic tissue were<br />

demonstrated to express somatostatin receptors<br />

with a high binding affinity <strong>for</strong> the native<br />

hormone, low binding affinities <strong>for</strong> lanreotide<br />

<strong>and</strong> somatuline but with little or no binding<br />

affinity <strong>for</strong> octreotide [18]. Interestingly,<br />

these early binding receptor studies suggested<br />

that somatostatin receptors were preferentially<br />

expressed in well-differentiated compared<br />

to less differentiated tumors [19, 20].<br />

These observations suggest somatostatin receptors<br />

may represent markers of differentiation<br />

in some cancers <strong>and</strong> that loss of functional<br />

hSSTRs may be of clinical importance in<br />

the progression of neoplasia. Thus, since somatostatin<br />

receptors play an important role in<br />

the physiological control of cell proliferation,<br />

loss of hSSTR expression in neoplastic cells<br />

would confer a proliferative advantage to<br />

those cells <strong>and</strong> their progeny. The net result<br />

would be the emergence of a dominant rapidly<br />

proliferating clone of hSSTR-negative cells<br />

<strong>and</strong> the progression of the neoplasm to a more<br />

aggressive, less differentiated tumor. Although<br />

it has not yet been established which<br />

of the 5 hSSTR are responsible <strong>for</strong> the progression<br />

of a tumor to a less differentiated<br />

more aggressive phenotype, the genes which<br />

encode <strong>for</strong> these receptors <strong>and</strong> their proteins<br />

which mediate their intracellular transduction<br />

pathways should be regarded as tumor<br />

suppressor genes. This suggestion is supported<br />

by the observation that a point mutation in<br />

the gene encoding <strong>for</strong> hSSTR-2 results in a<br />

proliferative advantage in small cell lung cancer<br />

cells in vitro [21]. These observations have<br />

two important clinical correlates. Firstly, it<br />

is well established that long-term somatostatin<br />

analog therapy results in downregula-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 165


tion of hSSTRs. Clearly, it will be important<br />

to establish whether or not long-term therapy<br />

with somatostatin analogs results in well-differentiated<br />

tumors becoming phenotypically<br />

more aggressive, thereby adversely affecting<br />

survival. Secondly, restoration or modulation<br />

of hSSTR expression in anaplastic tumors by<br />

gene transfer may revolutionize somatostatin<br />

analog therapy of neoplasia.<br />

More recently, cloning of 5 hSSTR has<br />

enabled the expression of individual somatostatin<br />

receptor expressions to be determined<br />

using techniques such as in situ hybridization,<br />

RNA-ase protection <strong>and</strong> reverse transcriptase<br />

polymerase chain reactions (RT-PCR) [22–<br />

37]. Studies using these techniques indicate<br />

that there is considerable heterogeneity between<br />

<strong>and</strong> within individual tumors with respect<br />

to the density of the individual hSSTRs<br />

expressed. Thus, in most tumors, hSSTR-2<br />

which is thought to be the receptor whereby<br />

somatostatin analogs exert the majority of<br />

their antiproliferative effects was expressed in<br />

the highest density. The other hSSTRs were<br />

also expressed in tumors expressing hSSTR-2,<br />

but in lower densities <strong>and</strong> in differing patterns<br />

in individual neoplasms. Perhaps more importantly<br />

with respect to the present generation<br />

of somatostatin analogs available <strong>for</strong> clinical<br />

use, some of the most common malignancies<br />

such as cancers of the pancreas, stomach<br />

<strong>and</strong> prostate do not express hSSTR-2 [18, 34].<br />

In addition there is some controversy on<br />

whether or not colorectal adenocarcinomas<br />

express hSSTR-2. In an early report using<br />

competitive binding assays, primary cultures<br />

of both normal <strong>and</strong> cancerous colonic tissue<br />

were reported not to bind to octreotide suggesting<br />

that these cells do not express hSSTR-<br />

2 [18]. However, in a subsequent study using<br />

RT-PCR the same group reported that the<br />

mucosa of colonic cancer <strong>and</strong> adjacent normal<br />

tissue expressed all 5 hSSTRs [37]. One<br />

possible explanation <strong>for</strong> these divergent re-<br />

sults is that high affinity somatostatin receptors<br />

have been reported on vascular <strong>and</strong> stromal<br />

cells [34, 38, 39]. Thus, in the second<br />

study by the Southampton group [37], the<br />

highly sensitive PT-PCR technique used to<br />

determine the hSSTR expression in colonic<br />

biopsies may have detected hSSTR-2 in contaminating<br />

peritumoral blood vessels <strong>and</strong> stroma<br />

rather than on the tumor cells per se. Finally,<br />

a proportion of tumors which normally<br />

express a high density of hSSTR-2 do not<br />

express this subtype [15, 16]. The variable<br />

expression of the hSSTRs between various<br />

tumors emphasizes the need to define the precise<br />

characteristics of the receptor subtypes in<br />

tumor tissue prior to somatostatin analog therapy<br />

if the direct antineoplastic effects of these<br />

drugs are to be maximized. Clearly such studies<br />

should be carried out on tumor cells per se<br />

<strong>and</strong> not on whole tissue homogenates, because<br />

of possible confounding contamination with<br />

peritumoral blood vessels <strong>and</strong>/or stroma. In<br />

addition, there is a need to define the proportion<br />

of binding attributable to each of the<br />

hSSTRs on the neoplastic cell to further optimize<br />

somatostatin analoge therapy <strong>and</strong> maximize<br />

the direct antineoplastic effects of these<br />

drugs. However, determination of hSSTR expression<br />

may not in itself be sufficient to fully<br />

exploit the direct antineoplastic effects of somatostatin<br />

analog therapy since it has yet to be<br />

determined whether or not detection of<br />

hSSTR mRNA implies the presence of functional<br />

proteins on the cell surface <strong>and</strong> intact<br />

intracellular pathways. Clearly, the relationship<br />

between hSSTR expression <strong>and</strong> functional<br />

receptor-proteins requires the development<br />

of specific antibodies to the latter. Not surprisingly<br />

there<strong>for</strong>e, the biological roles of each of<br />

the hSSTRs remains to be elucidated <strong>and</strong> defining<br />

the functional role of each of the subtypes<br />

requires the development of specific<br />

analogs, antagonists <strong>and</strong> antibodies. The complexity<br />

of the problems provides a major chal-<br />

166 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


lenge to peptide chemists <strong>and</strong> molecular biologists<br />

but such studies are essential to fully<br />

exploit the therapeutic potential of somatostatin<br />

analog therapy in neoplasia.<br />

Desensitization <strong>and</strong> Downregulation<br />

Cells can regulate their sensitivity to lig<strong>and</strong>s<br />

such as drugs or hormones by modifying<br />

their response (desensitization) or altering<br />

receptor density (downregulation or<br />

upregulation). Desensitization or tachyphylaxis<br />

may be achieved rapidly through functional<br />

uncoupling of the receptor proteins on<br />

the cell membrane that activate intracellular<br />

transduction pathways <strong>and</strong> is thought to be<br />

the initial response whereby cells regulate<br />

their response to continuous application of<br />

agonists. Cells can also modulate their response<br />

to continuous exposure of an agonist<br />

via a decrease in receptor density achieved by<br />

internalization of receptors (downregulation).<br />

In the case of persistent presence of an agonist,<br />

the agonist-receptor complex is internalized<br />

the structurally intact receptors stored<br />

within the cell <strong>and</strong> at this stage could be recycled<br />

if the stimulus was withdrawn. If stimulation<br />

persists the receptors are eventually broken<br />

down but can be resynthesized when the<br />

stimulus is eventually removed. Through the<br />

uncoupling <strong>and</strong> downregulation of receptors,<br />

cells are capable of decreasing the magnitude<br />

of their response to a constant level of agonist<br />

stimulation. This phenomenon is well illustrated<br />

by the effects of somatostatin on the<br />

modulation of potassium channels in neocortical<br />

cells [40]. Thus, some cells rapidly desensitize<br />

to continuous somatostatin application<br />

(less than 30 s) whereas other neurones are<br />

only partially desensitized or are refractory<br />

to more prolonged exposure of the hormone<br />

(5 min). However, with prolonged exposure<br />

(30 min) to somatostatin all neocortical cells<br />

are completely desensitized [40]. It seems<br />

likely that rapid desensitization of some neocortical<br />

cells following that exposure to somatostatin<br />

(less than 30 s) results from uncoupling<br />

of the G proteins from the somatostatin<br />

receptors whereas desensitization of neurones<br />

initially refractory to continuous exposure of<br />

hormone may require downregulation of the<br />

receptor.<br />

Desensitization <strong>and</strong> downregulation have<br />

important consequences <strong>for</strong> the clinical use of<br />

somatostatin analogs. As long ago as 1989<br />

Londong et al. [41] demonstrated that the initial<br />

potent inhibition of gastric secretion by<br />

relatively low doses of subcutaneous administration<br />

of octreotide was lost after 7 days of<br />

continuous administration. Subsequently, the<br />

initial potent inhibitory effects of octreotide<br />

on pancreatic enzyme excretion were also<br />

demonstrated to be lost after 7 days’ continuous<br />

subcutaneous administration [42]. The<br />

important question with respect to oncology<br />

is whether or not long-term somatostatin analog<br />

therapy results in downregulation of the<br />

receptors responsible <strong>for</strong> mediating their antineoplastic<br />

effects <strong>and</strong> which of the hSSTR are<br />

affected. There is good evidence that somatostatin<br />

<strong>and</strong> its analogs can upregulate or downregulate<br />

SSTR expression in a variety of cell<br />

lines <strong>and</strong> in tumor-bearing experimental animals<br />

[43–52]. It is now clear that regulation of<br />

individual SSTR expression by somatostatin<br />

<strong>and</strong> its analogs is cell subtype- <strong>and</strong> agonistspecific<br />

[43–52]. Furthermore, there is considerable<br />

evidence to suggest that SSTR expression<br />

is sensitive to the regulatory effects<br />

of glucocorticoids, thyroid hormones, estrogens,<br />

insulin <strong>and</strong> growth factors [53–57],<br />

which, at least in part, may explain why somatostatin<br />

analogs exhibit both a proliferative<br />

<strong>and</strong> antiproliferative effect on the same<br />

cell line depending on the culture conditions<br />

[58–62]. Although much progress has been<br />

made in our underst<strong>and</strong>ing of SSTR expres-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 167


sion at the cellular level the molecular events<br />

responsible <strong>for</strong> these effects are poorly under<br />

stood. In addition, the functional significance<br />

of either upregulation or downregulation of<br />

the individual SSTRs has yet to be elucidated,<br />

<strong>and</strong> in particular, with respect to the therapeutic<br />

<strong>and</strong> diagnostic use of somatostatin analogues,<br />

whether or not the expression of<br />

mRNAs <strong>for</strong> the individual hSSTRs by neoplastic<br />

cells implies a functional receptor protein<br />

<strong>and</strong> intact signal transduction mechanism(s).<br />

It is also important to distinguish<br />

between internalization <strong>and</strong> downregulation.<br />

Thus, internalization of the hSSTRs after<br />

brief exposure to somatostatin or its analogs is<br />

cell-, agonist- <strong>and</strong> receptor-specific [43–52]<br />

<strong>and</strong> involves endocytosis of the receptorlig<strong>and</strong><br />

complex which may be accompanied<br />

by recruitment of cellular SSTRs to the cell<br />

membrane [46] <strong>and</strong> may represent an important<br />

upregulatory mechanism whereby cells<br />

can rapidly modulate their response to agonist<br />

stimulation. Upregulation, whereby neoplastic<br />

cells increase the density of cell membrane<br />

SSTRs, has obvious implications in both the<br />

visualization (somatostatin scintigraphy) <strong>and</strong><br />

therapy (targeted chemotherapy or radiotherapy)<br />

of tumors using somatostatin analogs.<br />

However, as described earlier in this section,<br />

internalization is also the first step involved<br />

in downregulation in response to a persistent<br />

stimulus. It should be possible to distinguish<br />

between internalization resulting in upregulation<br />

from internalization heralding the onset<br />

of downregulation by repeat somatostatin receptor<br />

scintigraphy, if these were the only<br />

mechanisms involved in determining the response<br />

to escape from the antineoplastic effects<br />

of somatostatin analog therapy. Moreover,<br />

clinically, the situation is more complex<br />

since there are other mechanisms which may<br />

be responsible <strong>for</strong> escape from the inhibitory<br />

effects of somatostatin analog therapy on<br />

tumor growth <strong>and</strong> hypersecretory states.<br />

For example, the antineoplastic effects of<br />

somatostatin analogs on tumor growth would<br />

be limited to those cells bearing hSSTRs or<br />

whose growth is inhibited by their indirect<br />

antiproliferative effects. Thus, a proportion of<br />

cells in each tumor will not respond to either<br />

the direct or indirect antineoplastic effects of<br />

somatostatin analog therapy, the net result<br />

being the growth <strong>and</strong> development of clones<br />

of differentiated cells refractory to analog<br />

treatment. It seems likely that escape is a combination<br />

of these phenomena since in a small<br />

proportion of VIP-oma patients the initial<br />

inhibitory effects octreotide on hormone hypersecretion<br />

is lost after as little as 4 days’<br />

therapy [63]. Downregulation or desensitization<br />

of hSSTRs responsible <strong>for</strong> the inhibitory<br />

effects of octreotide would appear to be the<br />

only plausible explanation <strong>for</strong> escape in<br />

these patients. However, <strong>for</strong> the majority<br />

of VIP-oma patients escape appears after a<br />

much longer period of octreotide therapy (6–<br />

16 months) suggesting that downregulation<br />

<strong>and</strong> possibly the growth of rapidly proliferating<br />

clones of cells may both be responsible <strong>for</strong><br />

loss of efficacy of analog therapy. Possibly,<br />

somatostatin scintigraphy be<strong>for</strong>e <strong>and</strong> after somatostatin<br />

analog therapy may, at least in<br />

part, be able to differentiate between the relative<br />

roles of downregulation <strong>and</strong> the growth of<br />

hSSTR-negative clones in the phenomenon of<br />

escape, but is unlikely to provide a definitive<br />

answer to this problem. Irrespective of the<br />

mechanisms responsible <strong>for</strong> escape from somatostatin<br />

analog therapy, this phenomenon<br />

represents a considerable problem to the longterm<br />

use of these drugs in oncology. However,<br />

the problem of developing resistance to drug<br />

therapy is not new <strong>and</strong> applies to a wide variety<br />

of therapeutic agents. For example, it is<br />

well established that chronic antibiotic therapy<br />

is the best way to induce antibiotic resistance.<br />

Similarly with respect to oncology, it<br />

has long been recognized that prolonged che-<br />

168 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


motherapy leads to the development of cytotoxic<br />

resistant clones of tumor cells. More<br />

recently, it has been generally accepted that<br />

<strong>and</strong>rogen blockade with luteinizing hormonereleasing<br />

hormone (LHRH) or anti<strong>and</strong>rogens<br />

needed to be permanent to normalize surrogate<br />

end points such as prostatic-specific antigen<br />

(PSA) <strong>and</strong> improve survival in patients<br />

with prostatic cancer. This hypothesis was<br />

based on the assumption that total <strong>and</strong>rogen<br />

deprivation resulted in a rapid sustained response<br />

which outweighed the development of<br />

hormone resistance. More recently, intermittent<br />

<strong>and</strong>rogen blockade in a relatively small<br />

number of patients was demonstrated to be<br />

without risk <strong>and</strong> to significantly improve<br />

quality of life after initial normalization of<br />

PSA in patients with prostate cancer [64, 65].<br />

It remains to be established in r<strong>and</strong>omized<br />

controlled trials whether intermittent <strong>and</strong>rogen<br />

blockade, by delaying development of<br />

hormone resistance, improves survival compared<br />

to continuous therapy in patients with<br />

carcinoma of the prostate. However, there are<br />

no good data on whether or not intermittent<br />

somatostatin analog therapy prevents or reduces<br />

the rate of downregulation or the development<br />

of hormone resistant clones of differentiated<br />

cells. Un<strong>for</strong>tunately, the three analogs<br />

currently available <strong>for</strong> clinical use, octreotide,<br />

vapreotide <strong>and</strong> lanreotide, have very<br />

similar antineoplastic effects, the majority of<br />

which are mediated predominantly via the<br />

same hSSTR. There<strong>for</strong>e, a treatment option<br />

alternating octreotide, vapreotide <strong>and</strong> lanreotide<br />

is unlikely to delay escape. A more realistic<br />

approach may be intermittent administration<br />

of long-acting depot preparations of octreotide,<br />

lanreotide <strong>and</strong> vapreotide, but such<br />

clinical studies will require careful monitoring<br />

until the reliability <strong>and</strong> safety of such an<br />

approach can be justified.<br />

Future Clinical Studies<br />

In the past 15 years by far <strong>and</strong> away the<br />

most successful use of octreotide has been in<br />

the treatment of acromegaly. As discussed<br />

earlier, although octreotide <strong>and</strong> other somatostatin<br />

analogs effectively control hypersecretion<br />

of growth hormone or insulin-like growth<br />

factor-1 (IGF-1) in acromegalics, these drugs<br />

are cytotaxic <strong>and</strong> not cytotoxic. Consequently,<br />

in younger patients surgery is the treatment<br />

of choice since it offers the change of a<br />

complete cure. If surgery is not curative, radiotherapy<br />

is a second option but normalization<br />

of growth hormone <strong>and</strong> IGF-1 can take<br />

up to 2 years following such treatment. Octreotide<br />

is very effective in providing good<br />

long-term (112 years) control of hormone hypersecretion<br />

in acromegalics who fail surgery<br />

<strong>and</strong>/or radiotherapy. Since no escape occurs<br />

from octreotide inhibition of hypersecretion<br />

even during very prolonged periods of treatment<br />

in acromegalics, somatostatin analog<br />

therapy is the treatment of choice in elderly<br />

patients with this pituitary tumor. A number<br />

of reports suggest that lanreotide <strong>and</strong> vapreotide<br />

are as effective as octreotide in the management<br />

of acromegaly, but since these analogs<br />

have only fairly recently become available<br />

<strong>for</strong> clinical use, there is obviously no<br />

in<strong>for</strong>mation on very long-term therapy in<br />

acromegalics. An important advance in the<br />

management of acromegaly with somatostatin<br />

analog therapy has been the development<br />

of long-acting depot preparations of octreotide,<br />

lanteotide <strong>and</strong> vapreotide which will<br />

improve patient compliance <strong>and</strong> more importantly<br />

quality of life.<br />

The role of somatostatin analog therapy in<br />

GEP tumors is limited because relapse occurs<br />

in all patients usually within 2 years following<br />

commencement of treatment. Consequently,<br />

somatostatin analog therapy is now<br />

considered to be palliative <strong>and</strong> should be<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 169


eserved <strong>for</strong> when other treatment options<br />

have failed. Surgery is the only treatment<br />

which offers the chance of a complete cure in<br />

patients with GEP tumors even when liver<br />

metastases are present. There is little doubt<br />

that surgery <strong>for</strong> GEP tumors has been greatly<br />

enhanced by somatostatin scintigraphy [66]<br />

or positron emission tomography <strong>for</strong> in vivo<br />

studies of 5-hydroxytryptophan metabolism<br />

[67]. Both techniques are superior to conventional<br />

radiological techniques in localizing<br />

small lesions preoperatively <strong>and</strong> enable the<br />

surgeon to increase the chance of achieving a<br />

curative resection. There is some evidence<br />

that surgery can alter the natural history of<br />

gastrinomas [68] but there is an urgent need<br />

<strong>for</strong> good r<strong>and</strong>omized controlled trials to establish<br />

the benefits of radical surgical intervention<br />

in patients with GEP tumors. Furthermore,<br />

even if radical curative surgery<br />

cannot be per<strong>for</strong>med, debulking procedures<br />

should be considered to reduce the tumor load<br />

which may enhance the efficacy of medical<br />

treatments such as chemotherapy with or<br />

without liver embolization, tumor-targeted<br />

radiotherapy, immunotherapy <strong>and</strong> somatostatin<br />

analog treatment. The precise role <strong>and</strong><br />

timing of the adjuvant therapies has yet to be<br />

fully determined <strong>and</strong> a multimodality approach,<br />

using different therapies synchronously<br />

or metachronously, is required to maximize<br />

the benefits of medical treatment of<br />

GEP tumors which are not resectable. Such<br />

studies should take the <strong>for</strong>m of controlled<br />

trials, survival <strong>and</strong> quality of life being the<br />

primary outcome measures. Interestingly,<br />

neuroendocrine tumors of the gastrointestinal<br />

tract <strong>and</strong> pancreas are the only cancers where<br />

hepatic metastases are not a contraindication<br />

to liver transplantation. However, extrahepatic<br />

disease is a contraindication to liver transplantation.<br />

Thus, several reports indicate that<br />

liver transplantation provides good symptomatic<br />

relief <strong>and</strong> is occasionally curative in<br />

patients with neuroendocrine tumors in<br />

whom resection is not feasible [69–72]. Disease<br />

recurrence is common but good palliation<br />

can be achieved with medical therapy<br />

<strong>and</strong> survival is longer in patients with carcinoid<br />

tumors than other neuroendocrine tumors<br />

following liver transplantation [72].<br />

However, the indications <strong>and</strong> timing of transplantation<br />

still needs to be evaluated <strong>for</strong><br />

neuroendocrine tumors. Furthermore, liver<br />

transplantation is not a realistic treatment option<br />

<strong>for</strong> the majority of patients with metastatic<br />

GEP tumors because of the shortage of<br />

donor livers.<br />

With respect to somatostatin analog therapy<br />

of GEP tumors, it is generally accepted<br />

that it should be reserved <strong>for</strong> patients in<br />

whom all treatment options have failed. However,<br />

somatostatin analog therapy can potentiate<br />

cytotoxic therapy in vitro <strong>and</strong> in experimental<br />

animals [73–75]. Furthermore, there<br />

is compelling experimental evidence that octreotide<br />

is a very potent inhibitor of the<br />

growth <strong>and</strong> development of hepatic metastases<br />

in experimental animals [11]. Consequently,<br />

it is feasible that short-term somatostatin<br />

analog therapy of GEP tumors in combination<br />

with cytotoxic therapy, liver embolization<br />

<strong>and</strong> hepatic resection may potentiate<br />

the palliative effects of these treatment modalities.<br />

Such an approach may prolong the<br />

time during which these adjuvant therapies<br />

provide symptomatic control without precluding<br />

the use of somatostatin analogs as a<br />

palliative therapy when all other treatment<br />

options have been exhausted. However, further<br />

controlled trials are required to substantiate<br />

this hypothesis.<br />

Clinical experience with somatostatin analogs<br />

over the past 15 years has clearly demonstrated<br />

that these compounds are very successful<br />

in the management of acromegaly <strong>and</strong><br />

are a valuable palliative therapy <strong>for</strong> neuroendocrine<br />

tumors. In contrast, although there<br />

170 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


is considerable evidence that somatostatin<br />

analogs have a potent antineoplastic effect<br />

on neoplasms of the breast, colon, pancreas,<br />

prostate, lung <strong>and</strong> other common malignancies<br />

in experimental animals <strong>and</strong> in vitro, the<br />

results of clinical trials of these drugs have<br />

been very disappointing (see other papers in<br />

this volume <strong>for</strong> detailed reviews on the results<br />

of somatostatin analog therapies in individual<br />

malignancies). However, the vast majority of<br />

early clinical trials were uncontrolled <strong>and</strong> carried<br />

out in patients with advanced metastatic<br />

disease <strong>and</strong> the methodology used to assess<br />

health-related quality of life (HRQL) inadequate.<br />

It is important to recognize that we<br />

should not be deterred by the lack of a therapeutic<br />

benefit of somatostatin analogs in advanced<br />

malignancy since the majority of antineoplastic<br />

treatments are also ineffective in<br />

this clinical setting but are substantially more<br />

effective in patients with a relatively low tumor<br />

burden. There is no reason to suspect that<br />

somatostatin analog therapy will not be more<br />

effective in patients with a low tumor burden<br />

compared to those with advanced disease<br />

<strong>and</strong>, indeed, there is a wealth of preclinical<br />

evidence to suggest that this is the case. However,<br />

there is a paucity of good studies in<br />

man to support this hypothesis, <strong>and</strong> as yet, we<br />

must accept that the clinical efficacy of somatostatin<br />

analog therapy in patients with a<br />

low tumor burden is unproven. This immediately<br />

raises ethical problems of conducting<br />

controlled trials of somatostatin analog therapy<br />

in patients with a relatively low tumor burden<br />

if other treatments have been previously<br />

demonstrated to exert a clinically beneficial<br />

effect. This problem can be overcome by<br />

carrying out r<strong>and</strong>omized controlled trials<br />

comparing a therapy with established clinical<br />

efficacy alone or combined with somatostatin<br />

analogs. Such an approach has an advantage<br />

in patient recruitment since more will be<br />

amenable to participate in a trial in which<br />

they are r<strong>and</strong>omized to receive a treatment<br />

with an established therapeutic benefit alone<br />

or in combination with a novel drug with as<br />

yet unproven clinical efficacy. On the other<br />

h<strong>and</strong>, such an approach requires very large<br />

numbers of patients to be recruited into the<br />

trial to identify significant differences in outcome<br />

attributable to somatostatin analog<br />

therapy over <strong>and</strong> above that observed with<br />

the established therapy with adequate power.<br />

There can be little argument that future<br />

trials of somatostatin analog therapy of neoplasia<br />

should be r<strong>and</strong>omized controlled studies<br />

wherever possible. In general, the methodology<br />

<strong>for</strong> the design of clinical trials is well<br />

documented <strong>and</strong> investigators should take<br />

care to include all the data required by medical<br />

journals <strong>for</strong> the reporting <strong>and</strong> publication<br />

of such studies in preparing the protocol [76].<br />

With respect to somatostatin analog therapy<br />

of neoplasia, additional problems need to be<br />

addressed, in particular HRQL <strong>and</strong> dosing<br />

regimes. HRQL includes both specific <strong>and</strong><br />

generic instruments <strong>and</strong> care must be taken in<br />

analyzing the data since there are methodological<br />

difficulties which have not been completely<br />

resolved <strong>for</strong> r<strong>and</strong>omized controlled<br />

trials. Thus, generic <strong>and</strong> specific questionnaires<br />

used to assess the HRQL have multiple<br />

questions covering several domains which<br />

give rise to analytical problems. Analysis of<br />

individual questions or subscales would increase<br />

the likelihood of obtaining a considerable<br />

number of significant difference by<br />

chance whereas combining a large number of<br />

question into a single HRQL index may not<br />

be advisable as in<strong>for</strong>mation may be lost. A<br />

reasonable compromise is to limit the number<br />

of individual questions analyzed to those relevant<br />

to the disease <strong>and</strong> expected treatment<br />

effects <strong>and</strong> to combine only those that are<br />

related. However, advice should be sought<br />

from an expert in statistics <strong>and</strong> health services<br />

research, particularly experienced with<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 171


HRQL, prior to commencing any study, not<br />

only to select the best generic <strong>and</strong> specific<br />

instruments <strong>and</strong> how to analyze the data, but<br />

also to check that the study has sufficient power<br />

to detect important changes in HRQL where<br />

they exist with 95% confidence intervals.<br />

The dosing regime <strong>for</strong> somatostatin analog<br />

therapy in controlled studies of their efficacy<br />

in solid tumors is difficult to define. It may<br />

well be that higher doses of somatostatin analog<br />

are required to control tumor growth than<br />

to control hormone hypersecretion characteristic<br />

of acromegaly <strong>and</strong> neuroendocrine tumors,<br />

indications in which the greater part of<br />

the evidence of clinical efficacy of these drugs<br />

in neoplasia has been accumulated. Unlike<br />

chemotherapy, somatostatin analogs are well<br />

tolerated even at very high doses <strong>and</strong> hence<br />

the dosage cannot be based on maximal tolerable<br />

side effects. It is tempting to select a highdose<br />

regime <strong>for</strong> somatostatin analog therapy<br />

but this may not be advisable since as will be<br />

discussed in a subsequent section, these drugs<br />

may have adverse effects on the immune system.<br />

There<strong>for</strong>e, care must be taken to ensure<br />

that any increased antineoplastic effects of<br />

high-dose somatostatin analog therapy are<br />

not negated by increasing suppression of the<br />

immune system, the body’s major defense<br />

against neoplasia. In addition, high-dose somatostatin<br />

analog therapy may result in more<br />

rapid downregulation of the hSSTRs responsible<br />

<strong>for</strong> mediating the antineoplastic effects<br />

of somatostatin analogs <strong>and</strong>/or the development<br />

of somatostatin-resistant clones of tumor<br />

cells. At present, there<strong>for</strong>e, the dosing<br />

regime <strong>for</strong> somatostatin analog therapy of solid<br />

tumors has to be somewhat empirical <strong>and</strong><br />

based on in<strong>for</strong>mation from the few studies<br />

which have demonstrated a clinical benefit in<br />

patients with advanced malignancy or extrapolation<br />

of doses that result in plasma levels<br />

of these drugs in experimental animals in<br />

which antineoplastic effects have been ob-<br />

served. The latter approach is fraught with<br />

difficulties, however, since extrapolation of<br />

results obtained in experimental animals can<br />

be very misleading when applied to man. At<br />

present, there is perhaps arguably, only one<br />

really well-designed prospective r<strong>and</strong>omized<br />

control of somatostatin analog therapy in advanced<br />

malignancy which has clearly demonstrated<br />

the therapeutic potential of these<br />

drugs as antineoplastic agents [77]. In this<br />

study, patients with advanced inoperable somatostatin<br />

receptor-positive hepatocellular<br />

carcinoma were r<strong>and</strong>omized to 500 Ìg octreotide<br />

s.c. b.d. or no treatment. The results of<br />

this study clearly indicated that the survival<br />

of octreotide treated patients was significantly<br />

better than that of controls. Multivariate analysis<br />

confirmed that octreotide was a major<br />

factor influencing survival. Tumor size was<br />

also investigated at regular intervals in both<br />

groups of patients. In patients r<strong>and</strong>omized to<br />

receive no treatment all tumors continued to<br />

grow until the death of the patient. Similarly<br />

in the octreotide-treated patients, all large tumors<br />

continued to grow. However, small satellite<br />

tumors (!3 cm in diameter) either disappeared<br />

or remained the same size after prolonged<br />

octreotide therapy [77]. There<strong>for</strong>e, <strong>for</strong><br />

the present, it would be reasonable to use a<br />

dose of 500 Ìg somatostatin analog twice a<br />

day in initial <strong>for</strong>mal clinical trials to evaluate<br />

the therapeutic potential of these drugs in<br />

patients with advanced malignancy <strong>and</strong> possibly<br />

a reduced dose in patients with a relatively<br />

low tumor load. The response to somatostatin<br />

analog therapy may vary with both the hSSTR<br />

expression of individual tumors <strong>and</strong> any therapy<br />

used in combination with these drugs.<br />

There is no in<strong>for</strong>mation available on how<br />

these confounding factors may influence the<br />

response to somatostatin analog therapy, <strong>and</strong><br />

until such data becomes available, it would be<br />

unwise to speculate on varying dosing regimes.<br />

172 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


With respect to clinical trial methodology<br />

of the efficacy of somatostatin analog therapy<br />

it is important to specify in the eligibility criteria<br />

that only patients with somatostatin receptor-positive<br />

tumors are included. Fortunately,<br />

most human solid tumors express<br />

hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong> can be detected<br />

by somatostatin scintigraphy using stable radioisotopes<br />

of octreotide <strong>and</strong> vapreotide or by<br />

positron emission tomography (PET). However,<br />

there are exceptions such as cancer of<br />

the prostate which do not express hSSTR-2 or<br />

hSSTR-5 <strong>and</strong> in these tumors the presence of<br />

somatostatin receptors needs to be based on<br />

tissue obtained by biopsy or at operation.<br />

Indeed, ideally tissue should be obtained from<br />

all tumors prior to therapy with somatostatin<br />

analog therapy to establish the hSSTR profile<br />

in order to determine whether or not this may<br />

influence the response to therapy with these<br />

drugs.<br />

Combination Therapy<br />

Chemotherapy<br />

In many ways, the role of cytotoxic therapy<br />

in malignancy resembles that of somatostatin<br />

analogs. Thus, early dramatic responses<br />

achieved by chemotherapy in fairly uncommon<br />

malignancies such as Hodgkin’s disease,<br />

histiocytic lymphoma, germ cell tumors, choriocarcinoma<br />

<strong>and</strong> various solid tumors of<br />

childhood led to a bewildering number of<br />

clinical studies of cytotoxic therapy in all cancers.<br />

However, it is now greatly accepted that<br />

chemotherapy alone is rarely curative <strong>and</strong> in<br />

the majority of patients is used to provide palliation.<br />

In this situation, a decision to use<br />

cytotoxic drugs inevitably entails balancing<br />

the potential benefits against the toxic side<br />

effects of chemotherapy. Consequently, the<br />

use of toxic <strong>and</strong> hence potentially fatal cytotoxics<br />

should be questioned unless they are<br />

likely to lead to a high incidence of durable<br />

complete remissions with the occasional cure<br />

or regression of the tumor to such an extent<br />

that symptomatic relief offers a prolonged improvement<br />

in HRQL. However, <strong>for</strong> most advanced<br />

metastatic cancers the effects of chemotherapy<br />

have been largely disappointing,<br />

cytotoxics providing only modest improvements<br />

in response rates <strong>and</strong> survival. A good<br />

example is provided by a critical meta-analysis<br />

of the large number of r<strong>and</strong>omized controlled<br />

trials comparing single agent versus<br />

polychemotherapy in the management of recurrent<br />

early breast cancer [78]. In general,<br />

this meta-analysis indicated that polychemotherapy<br />

resulted in a modest increase in response<br />

<strong>and</strong> survival compared to single agent<br />

chemotherapy but at the expense of increased<br />

toxicity. However, the clinical relevance of<br />

this large meta-analysis of chemotherapy of<br />

early breast cancer recurrence is limited because<br />

of the paucity of data on HRQL [78].<br />

Fortunately, in the case of breast cancer, these<br />

tumors respond to less toxic hormonal treatment.<br />

Thus, in recurrent early breast cancer,<br />

hormonal treatment, single agent or combination<br />

of tamoxifen, megestrol <strong>and</strong> aromatase<br />

inhibitors compared favorably in terms of<br />

response rate <strong>and</strong> survival with the majority<br />

of chemotherapeutic regimes [78]. However,<br />

there was no improvement in response rates<br />

or survival in trials comparing chemotherapy<br />

alone or combined with hormone treatment<br />

[79]. For the medical oncologist, the results of<br />

these r<strong>and</strong>omized controlled trials of chemotherapy<br />

<strong>and</strong> hormonal treatment allow <strong>for</strong><br />

some limited clinical decision making in the<br />

management of advanced metastatic breast<br />

cancer. Hence, most oncologists would prefer<br />

single agent or combination hormonal therapy<br />

to chemotherapy because of the relatively<br />

reduced toxicity. For many solid tumors,<br />

however, cytotoxic therapy may be the only<br />

therapeutic modality available <strong>for</strong> the man-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 173


agement of advanced metastatic malignancy.<br />

In these situations it is much more difficult to<br />

decide whether or not cytotoxic therapy has<br />

any overall beneficial effects particularly because<br />

of the lack of good HRQL data.<br />

<strong>Somatostatin</strong> analogs potentiate the cytotoxic<br />

effects <strong>and</strong> decrease the toxicity of cytotoxic<br />

agents [73–75], suggesting they may improve<br />

the response rate, survival <strong>and</strong> HRQL<br />

of chemotherapy of advanced metastatic disease.<br />

This hypothesis merits further investigation<br />

in the <strong>for</strong>m of large r<strong>and</strong>omized controlled<br />

trials <strong>and</strong> particular emphasis should<br />

be placed on the HRQL component of the<br />

studies which hitherto has largely been neglected.<br />

Targeted Chemotherapy<br />

An interesting <strong>and</strong> challenging concept is<br />

the targeting of chemotherapy to somatostatin<br />

receptor-positive tumors by synthesizing<br />

conjugates of somatostatin analogs <strong>and</strong> cytotoxics<br />

[79, 80]. These preliminary studies indicated<br />

that the antineoplastic effects of the<br />

cytotoxic radicals in these conjugates on<br />

growth of a variety of human cancer effects<br />

was preserved in vitro. Although conjugation<br />

of the cytotoxic somatostatin analogs reduced<br />

their binding affinities to somatostatin receptors,<br />

the conjugates were still capable of inhibiting<br />

the release of growth hormone in the<br />

nanomolar range. Finally, conjugates of somatostatin<br />

analogs <strong>and</strong> cytotoxics were less<br />

toxic <strong>and</strong> more potent in inhibiting the<br />

growth of breast <strong>and</strong> pancreatic tumors in<br />

experimental animals than the cytotoxic<br />

agents alone following intravenous administration<br />

[79, 80]. Although these preliminary<br />

results are a potentially exciting development<br />

in the use of somatostatin analogs in oncology<br />

there are a number of problems which require<br />

further evaluation. Firstly, somatostatin statin<br />

receptor-positive cells are widely distributed<br />

throughout the body <strong>and</strong> hSSTR-2 is<br />

the most commonly expressed subtype in<br />

both normal <strong>and</strong> neoplastic tissue. Un<strong>for</strong>tunately,<br />

the present generation of somatostatin<br />

analogs all bind to hSSTR-2 with high affinity.<br />

There<strong>for</strong>e, synthesis of conjugates of octreotide,<br />

lanreotide <strong>and</strong> vapreotide cytotoxics<br />

would not really target chemotherapy to neoplastic<br />

tissue per se, but to all cells expressing<br />

hSSTR-2 or hSSTR-5. In the preliminary<br />

studies of conjugates or somatostatin analogs<br />

<strong>and</strong> cytotoxics tissues distribution was not<br />

determined <strong>and</strong> toxicity was defined only in<br />

terms of death during a very short treatment<br />

period [79, 80]. Clearly much work is required<br />

to address these issues to establish the relative<br />

toxicities of the conjugates of cytotoxics <strong>and</strong><br />

somatostatin analogs compared to the chemotherapeutic<br />

agent alone. In addition, targeting<br />

of chemotherapy to tumors after systemic administration<br />

of somatostatin analog <strong>and</strong> cytotoxics<br />

will require the development of newer<br />

hSSTR-selective analogs to tumors not expressing<br />

hSSTRs to which octreotide, lanreotide<br />

<strong>and</strong> vapreotide do not bind. If, however,<br />

these problems can be resolved somatostatin<br />

analog targeting of cytotoxics to neoplastic<br />

cells may prove to be a valuable treatment<br />

option <strong>for</strong> advanced metastatic cancer.<br />

Hepatic Metastases Derived from<br />

Colorectal Primaries<br />

At present, the most promising role of chemotherapy<br />

in cancer is as an adjuvant to surgery<br />

<strong>and</strong>/or radiotherapy. Of particularly interest<br />

with respect to somatostatin analog<br />

therapy is their potential benefit when used in<br />

combination with cytotoxics as an adjuvant to<br />

surgery in common abdominal malignancies.<br />

Space does not allow <strong>for</strong> a comprehensive discussion<br />

of the natural history of all types of<br />

gastrointestinal malignancies <strong>and</strong> the rationale<br />

<strong>for</strong> adjuvant cytotoxic therapy in these<br />

174 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


Table 1. Colorectal cancer stage,<br />

definitions, frequency at diagnosis<br />

<strong>and</strong> survival<br />

Dukes<br />

stage<br />

indications following surgery, but perhaps is<br />

best illustrated by colorectal cancer. The prognosis<br />

<strong>for</strong> patients with colorectal cancer is<br />

summarized in table 1 <strong>and</strong> is associated with<br />

the staging of the disease, which in turn, is<br />

predominantly influenced by the development<br />

of liver metastases which have traditionally<br />

been equated with the imminent<br />

demise of the patient. For patients who undergo<br />

a theoretically curative resection of<br />

their colorectal primary <strong>and</strong> in whom no<br />

overt hepatic metastases are present at the<br />

time of operation, only 35–40% will be truly<br />

cured. A small proportion (12–15%) will develop<br />

extrahepatic hepatic disease (including<br />

local recurrence) <strong>and</strong> the remainder will develop<br />

liver metastases within 2–5 years of<br />

their theoretically curative resection. These<br />

observations would suggest that in those patients<br />

who develop hepatic metastases, occult<br />

micrometastases are present in the liver at the<br />

time of their theoretically curative resection<br />

of their colorectal primary or that tumorigenic<br />

cells are shed into the portal circulation during<br />

surgery <strong>and</strong> seed in the hepatic parenchyma.<br />

Not surprisingly, there<strong>for</strong>e, there has<br />

been considerable interest in the use of adjuvant<br />

chemotherapy, systemic or regional, to<br />

prevent or delay the development of hepatic<br />

Definition Approximate<br />

frequency<br />

at diagnosis<br />

%<br />

5-year<br />

survival<br />

A <strong>Cancer</strong> confined to bowel wall 10 80–85<br />

B1 <strong>Cancer</strong> which penetrates the bowel<br />

wall without spread through serosa 10–15 70–75<br />

B2 <strong>Cancer</strong> which penetrates wall of<br />

bowel beyond serosa 20–30 60–70<br />

C Lymph node involvement 25 35–40<br />

D Liver metastases 30 3–6<br />

metastases <strong>and</strong> hence improve survival after a<br />

theoretically curative resection of a colorectal<br />

primary. In brief, as long ago as 1988 a metaanalysis<br />

of r<strong>and</strong>omized adjuvant trials of 5fluorouracil<br />

(5-FU) <strong>for</strong> up to 1 year demonstrated<br />

a marginal overall reduction in the relative<br />

risk of death by approximately 17% <strong>and</strong><br />

an absolute 5-year survival benefit of 3–4%<br />

[81, 82]. More recently, however, much more<br />

impressive results have been reported using<br />

6–12 months’ adjuvant combination therapy<br />

with 5-FU <strong>and</strong> the immunomodulatory drug<br />

levamisole or by biomodulating 5-FU with<br />

leucovorin [81]. However, the statistical significant<br />

improvement in survival of these systemic<br />

therapies was confined to patients with<br />

Dukes C tumors <strong>and</strong> correlated with a similar<br />

reduction in the incidence of hepatic metastases<br />

[81]. Similar improvements in survival<br />

<strong>and</strong> the incidence of liver metastases is associated<br />

with a 7-day perioperative portal vein<br />

infusion of 5-FU following resection of a colorectal<br />

primary [81]. As would be expected, the<br />

toxicity of 7 days of 5-FU was markedly less<br />

than 6–12 months’ treatment with this cytotoxic.<br />

The major risk reduction of developing<br />

overt hepatic metastases <strong>and</strong> overall survival<br />

improvement, with portal vein infusion of 5-<br />

FU, was confined to patients with Dukes C<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 175<br />

%


Fig. 1. Percentage of hepatic replacement<br />

by K12Tr tumor in octreotide-treated<br />

rats with or without<br />

blockade of RES with gadolinium<br />

chloride (GAD). Octreotide<br />

almost completely inhibited the<br />

growth of hepatic tumor in rats<br />

with a functional hepatic RES.<br />

Blockade of the hepatic RES with<br />

GAD significantly increased tumor<br />

growth compared to controls. Octreotide<br />

inhibited tumor growth in<br />

rats with hepatic RES blockade but<br />

not to the same extent as it did in<br />

animals with a functional hepatic<br />

RES [after 90].<br />

disease [81]. Furthermore, the benefits of 6–<br />

12 months of treatment with 5-FU <strong>and</strong> levamisole<br />

or folinic acid are equivalent to<br />

those of 7 days’ perioperative portal vein with<br />

5-FU, although the combination of the two<br />

could be additive. Consequently, there is now<br />

a substantial body of evidence to suggest that<br />

all patients with Dukes C colorectal primaries<br />

should be treated with adjuvant systemic cytotoxic<br />

therapy using 5-FU combined with<br />

levamisole or folinic acid or a perioperative<br />

intraportal infusion of 5-FU. The situation in<br />

patients with Dukes B lesions is less certain<br />

since they are less likely to develop hepatic<br />

metastases than those with Dukes C disease,<br />

<strong>and</strong> hence, could be entered into a trial comparing<br />

surgery alone or combined with cytotoxic<br />

therapy. We have clearly demonstrated<br />

that octreotide inhibits the growth <strong>and</strong> development<br />

of hepatic tumors derived by intraportal<br />

injection of a variety of tumorigenic<br />

cells [11]. The mechanisms whereby octreotide<br />

inhibits the growth <strong>and</strong> development of<br />

hepatic tumors is mediated primarily via<br />

stimulation of the hepatic RES system (fig. 1)<br />

although direct <strong>and</strong> indirect antiproliferative<br />

effects of the somatostatin analogs are also<br />

contributory [11]. Levamisole stimulates hepatic<br />

RES activity, <strong>and</strong> this may partly explain<br />

its beneficial effect when used with 5-<br />

FU as an adjuvant to surgical resection of<br />

colorectal tumors [81]. However, octreotide<br />

is significantly more potent than levamisole<br />

in stimulating hepatic RES activity [11].<br />

Furthermore, since octreotide has additional<br />

direct <strong>and</strong> indirect antineoplastic effects<br />

apart from stimulation of hepatic RES activity,<br />

it may be superior to levamisole when<br />

combined with 5-FU in inhibiting the development<br />

of liver metastases after surgical resection<br />

of a primary colorectal tumor. Finally,<br />

since somatostatin <strong>and</strong> its analogs are cytoprotective<br />

with respect to the liver [11],<br />

concomitant administration of cytotoxics<br />

with somatostatin analogs may improve the<br />

response rates [73–75] <strong>and</strong> at the same time,<br />

protect the liver from hepatotoxic effects of<br />

the chemotherapeutic agents. We believe<br />

that the available evidence, although experimental,<br />

suggests that somatostatin analogs<br />

may have a beneficial effect on the development<br />

of liver metastases when used as an<br />

176 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


adjuvant to surgery in colorectal cancer <strong>and</strong><br />

that this area warrants urgent clinical investigation.<br />

Pancreatic <strong>Cancer</strong><br />

The possible therapeutic potential of somatostatin<br />

analogs as an adjuvant to surgery<br />

is not confined to the development of hepatic<br />

metastases derived from colorectal primaries.<br />

Thus, other gastrointestinal malignancies<br />

such as those of the stomach, pancreas, small<br />

intestine, gallbladder <strong>and</strong> extrahepatic bile<br />

ducts frequently metastasize to the liver.<br />

Some of these cancers such as pancreatic adenocarcinoma<br />

have low resectability rates <strong>and</strong><br />

curative surgery is only possible in 6–8% of<br />

patients [see 82]. Furthermore, even following<br />

a theoretically extensive resection of adenocarcinoma<br />

of the pancreas the 5-year survival<br />

is only approximately 25%. Following surgical<br />

resection <strong>for</strong> adenocarcinoma of the pancreas<br />

there are three major patterns of recurrence<br />

which occur with approximately equal<br />

frequency, namely local recurrence, hepatic<br />

metastases undetected at the time of operation<br />

<strong>and</strong> local recurrence with metastatic disease.<br />

There is now compelling evidence that<br />

the best means of preventing local recurrence<br />

after resection <strong>for</strong> pancreatic adenocarcinoma<br />

is radiotherapy whereas <strong>for</strong> metastatic spread<br />

chemotherapy is the treatment of choice.<br />

However, since the pattern of recurrence after<br />

pancreatic resection is both local <strong>and</strong> systemic,<br />

an improvement in survival is only<br />

likely to be observed with combinations of<br />

radiotherapy <strong>and</strong> chemotherapy. In view of<br />

the poor prognosis, even after a theoretically<br />

curative resection <strong>and</strong> because there is some<br />

limited evidence from phase 2 trials that adjuvant<br />

chemotherapy <strong>and</strong> radiotherapy may<br />

improve survival of patients with pancreatic<br />

adenocarcinoma, it is both surprising <strong>and</strong> dis-<br />

appointing that there are a paucity of r<strong>and</strong>omized<br />

controlled trials to evaluate the efficacy<br />

of adjuvant therapies in this indication.<br />

Clearly, there is an urgent need <strong>for</strong> confirmation<br />

of these putative benefits of adjuvant<br />

radiotherapy <strong>and</strong> chemotherapy in patients<br />

with resectable pancreatic tumors in the <strong>for</strong>m<br />

of large controlled r<strong>and</strong>omized trials. <strong>Somatostatin</strong><br />

analog therapy could be a valuable<br />

option <strong>for</strong> such phase III studies since there is<br />

good evidence that these compounds potentiate<br />

the cytotoxic effects <strong>and</strong> reduce the side<br />

effects of chemotherapy [73–75]. In addition,<br />

as discussed above, stimulation of hepatic<br />

RES activity by somatostatin analogs may<br />

be particularly important in inhibiting the<br />

growth <strong>and</strong> development of liver metastases<br />

[11]. All patients with pancreatic cancer in<br />

which surgery is possible should be entered<br />

into such r<strong>and</strong>omized controlled trials irrespective<br />

of the size of the tumor <strong>and</strong> such<br />

studies should be sufficiently large to detect<br />

differences between the therapies under investigation<br />

<strong>and</strong> between predefined subgroups<br />

with sufficient power.<br />

Gastric <strong>Cancer</strong><br />

Like pancreatic adenocarcinoma, gastric<br />

cancer has a very dismal prognosis <strong>and</strong> a high<br />

proportion of patients develop liver metastases.<br />

As discussed by Cascinu et al. [83] in<br />

another paper in this volume, there is an<br />

urgent need <strong>for</strong> large controlled trials to evaluate<br />

the efficacy of adjuvant therapies after<br />

surgery. <strong>Somatostatin</strong> analog therapy may, as<br />

discussed earlier <strong>for</strong> colorectal <strong>and</strong> pancreatic<br />

cancer, be a valuable treatment <strong>for</strong> gastric<br />

cancer to potentiate the effects of chemotherapy<br />

<strong>and</strong> via their stimulatory effect on hepatic<br />

RES activity inhibit the growth <strong>and</strong> development<br />

of liver metastases. Since colorectal,<br />

pancreatic <strong>and</strong> gastric cancers do not express<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 177


hSSTR-2, the receptor subtype whereby octreotide,<br />

vapreotide <strong>and</strong> lanreotide exert their<br />

direct <strong>and</strong> antineoplastic effects, the development<br />

of specific analogs with high binding<br />

affinities <strong>for</strong> individual hSSTRs would be required<br />

to maximize somatostatin analog therapy<br />

of any of these tumor cells remaining after<br />

resection. It is unlikely, however, that such<br />

analogs will be available <strong>for</strong> clinical use in the<br />

near future. An alternative strategy could be<br />

to use native somatostatin rather than one<br />

of its analogs perioperatively in conjunction<br />

with intraportal 5-FU in those cancers which<br />

do not express hSSTR-2 to maximize the potential<br />

antineoplastic effects of the hormone.<br />

Subsequently, 6–12 months of systemically<br />

administered 5-FU which may provide additional<br />

benefits to 7 days of intraportal administration<br />

of this cytotoxic [81] could be again<br />

given concomitantly with somatostatin analogs.<br />

Such an approach would, however, be<br />

very expensive <strong>and</strong> the therapeutic advantage<br />

of such therapy would have to be demonstrated<br />

to be very beneficial to justify the<br />

costs. Furthermore, in order to demonstrate a<br />

marked therapeutic advantage of somatostatin<br />

<strong>and</strong> its analogs currently available <strong>for</strong> clinical<br />

use, over <strong>and</strong> above adjuvant cytotoxic<br />

<strong>and</strong> radiotherapy in patients with gastric,<br />

pancreatic <strong>and</strong> colorectal cancer would require<br />

very large r<strong>and</strong>omized controlled trials<br />

which would be both difficult to organize <strong>and</strong><br />

fund.<br />

Liver Tumors<br />

The presence of overt liver metastases is<br />

not uncommon in gastrointestinal malignancies<br />

<strong>and</strong> is associated with a very poor prognosis.<br />

Colonic cancer is unique among solid<br />

tumors in that surgical resection of hepatic<br />

<strong>and</strong> pulmonary metastases can prolong longterm<br />

survival <strong>and</strong> offers the only real possibil-<br />

ity of cure [84–86]. Nevertheless, even after a<br />

technically curative resection of colorectal hepatic<br />

metastases, recurrence occurs in approximately<br />

60% of patients [86] suggesting<br />

that there is an urgent need <strong>for</strong> postoperative<br />

adjuvant therapy. The role of chemotherapy,<br />

immunotherapy, hepatic artery or portal vein<br />

embolization, brachytherapy, cryotherapy, interstitial<br />

laser hypothermia <strong>and</strong> alcohol injection<br />

as adjuvant <strong>and</strong> neoadjuvant therapies in<br />

the management of colorectal hepatic metastases<br />

has been critically examined in a recent<br />

review [87]. A strong case can be made <strong>for</strong> a<br />

multimodality approach using combinations<br />

of techniques to optimize the management of<br />

colonic liver metastases <strong>and</strong> hence prolong<br />

survival <strong>and</strong> increase the chance of cure.<br />

However, the efficacy of new adjuvant <strong>and</strong><br />

neoadjuvant therapies have yet to be rigorously<br />

evaluated in r<strong>and</strong>omized controlled trials<br />

be<strong>for</strong>e they are incorporated into <strong>for</strong>mal<br />

structured management strategies <strong>for</strong> the<br />

management of colonic metastases. Theoretically,<br />

somatostatin analog therapy may have a<br />

potentially beneficial effect when used in conjunction<br />

with these adjuvant or neoadjuvant<br />

treatment modalities <strong>for</strong> the management of<br />

colonic hepatic metastases. Thus, somatostatin<br />

analogs may potentiate the effects of adjuvant<br />

or neoadjuvant chemotherapy <strong>and</strong> reduce<br />

the hepatoxicity of cytotoxic agents [11,<br />

73–75]. Perhaps more importantly, however,<br />

the growth <strong>and</strong> development of hepatic tumors<br />

is significantly increased following partial<br />

hepatectomy [88, 89]. Octreotide inhibits<br />

the growth of hepatic tumors after partial<br />

hepatectomy, possibly via stimulation of hepatic<br />

(RES) activity which is significantly reduced<br />

after liver resection [89]. However, the<br />

direct <strong>and</strong> indirect antineoplastic effects of<br />

octreotide may also be contributory to the<br />

inhibition of tumor growth by this somatostatin<br />

analog after partial hepatectomy [11]. On<br />

the negative side, in addition to inhibiting the<br />

178 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


growth of hepatic tumors, octreotide also suppresses<br />

liver regeneration following partial hepatectomy.<br />

The potential <strong>for</strong> inhibition of<br />

regeneration in the human liver by somatostatin<br />

analogs may cause some concern. However,<br />

anecdotal evidence suggests that octreotide<br />

has no clinically detectable effect on liver<br />

regeneration in patients undergoing hepatic<br />

resection [Wu, pers. commun.]. Nevertheless,<br />

the potential detrimental effect of octreotide<br />

on liver resection after partial hepatectomy<br />

needs to be addressed more <strong>for</strong>mally prior to<br />

advocating the use of somatostatin analogs as<br />

an adjuvant to surgery in patients with liver<br />

metastases. There is an urgent need <strong>for</strong> further<br />

investigation of the effects of somatostatin<br />

analogs in this area since the benefits of<br />

such treatment combined with other adjuvant<br />

<strong>and</strong> neoadjuvant therapies to surgery could be<br />

considerable in the management of hepatic<br />

metastases.<br />

Hormonal Therapy<br />

There is compelling experimental evidence<br />

to suggest that somatostatin analog therapy<br />

may be of value when used in combination<br />

with other hormonal treatments in the management<br />

of estrogen-positive breast cancer<br />

<strong>and</strong> prostate cancer which is discussed in<br />

detail in another paper in this volume [90–<br />

92]. However, the results of clinical studies of<br />

somatostatin analogs in combination with<br />

other hormonal therapies in the management<br />

of breast <strong>and</strong> prostate cancer have been very<br />

disappointing [91]. It is generally accepted<br />

that tamoxifen is the hormonal treatment of<br />

choice <strong>for</strong> estrogen-positive metastatic breast<br />

cancer. In spite of compelling evidence that<br />

octreotide in combination with tamoxifen<br />

may be more effective than tamoxifen alone<br />

in the management of metastatic breast cancer<br />

there is now good evidence from two ran-<br />

domized controlled trials that this is not the<br />

case [92, 93]. Thus, in both trials the time to<br />

disease progression, objective response rates<br />

<strong>and</strong> survival were not significantly different<br />

between patients treated with tamoxifen<br />

alone or combined with octreotide [92–94].<br />

Moreover, in both trials the incidence of side<br />

effects was higher in patients treated with<br />

combination therapy than with tamoxifen<br />

alone [92, 93]. One of these studies initially<br />

r<strong>and</strong>omized patients to tamoxifen or octreotide<br />

alone or combination therapy [92]. The<br />

octreotide only arm of the trial was dropped<br />

when the somatostatin analog was found to be<br />

associated with a rapid time to disease progression<br />

<strong>and</strong> to produce no objective response<br />

[92]. Finally, in one of these trials, combined<br />

tamoxifen <strong>and</strong> octreotide treatment resulted<br />

in a significantly greater reduction in serum<br />

IGF-1 than tamoxifen alone in a limited cohort<br />

of patients [92]. Both these clinical trials<br />

can be criticized on the grounds that they<br />

were underpowered, particularly since multiple<br />

outcome measures were assessed, to completely<br />

evaluate any possible benefit of octreotide<br />

in the management of metastatic breast<br />

cancer. However, both studies are strongly<br />

suggestive that combined tamoxifen <strong>and</strong> octreotide<br />

therapy is unlikely to confer any<br />

worthwhile clinical benefit over tamoxifen<br />

alone, particularly since combination therapy<br />

was associated with an increased number of<br />

side effects which could adversely affect<br />

HRQL. It is unlikely that vapreotide or lanreotide<br />

are more effective than octreotide<br />

when used in combination with tamoxifen in<br />

the management of metastatic breast cancer<br />

since all three analogs exhibit similar biological<br />

effects. There<strong>for</strong>e, any benefit of these two<br />

newer analogs over octreotide is likely to be<br />

only marginal.<br />

The mechanisms whereby octreotide <strong>and</strong><br />

tamoxifen in combination do not confer any<br />

therapeutic benefit over <strong>and</strong> above that of<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 179


antiestrogen therapy alone is not clear. In the<br />

trial carried out by Ingle et al. [92], combined<br />

octreotide <strong>and</strong> tamoxifen therapy resulted in<br />

significantly greater reductions in serum IGF-<br />

1 than tamoxifen alone in patients with metastatic<br />

breast cancer without conferring any<br />

clinical benefits over <strong>and</strong> above that obtained<br />

with antiestrogen alone. However, in this<br />

study, serum IGF-1 levels were only measured<br />

in a cohort of patients be<strong>for</strong>e <strong>and</strong> 6 weeks after<br />

commencing combined tamoxifen <strong>and</strong> octreotide<br />

therapy or tamoxifen alone [92]. It has<br />

previously been reported that the initial suppression<br />

of serum IGF-1 levels by octreotide<br />

in patients with breast cancer is gradually lost<br />

approximately 8–14 weeks after commencing<br />

therapy [95]. Possibly, there<strong>for</strong>e, the failure of<br />

combined octreotide <strong>and</strong> tamoxifen to confer<br />

any additional clinical benefits compared to<br />

antiestrogen alone in patients with breast cancer<br />

may be due, at least in part, to the gradual<br />

loss of efficacy in the control of IGF-1 <strong>and</strong> possibly<br />

other mitogenic factors, with prolonged<br />

administration of the somatostatin analog.<br />

This hypothesis requires further investigation<br />

in patients with breast cancer since it could<br />

suggest alternative dosing regimes which may<br />

provide improvements in clinical outcome.<br />

For example, withdrawal of octreotide therapy<br />

as soon as IGF-1 levels begin to increase <strong>and</strong><br />

recommencement when the serum levels of<br />

this mitogen are within the range of those<br />

observed in patients with breast carcinoma<br />

treated with tamoxifen alone may prolong the<br />

period whereby combination therapy potentiates<br />

suppression of this growth factor. Intermittent<br />

octreotide administration combined<br />

with continuous tamoxifen therapy may delay<br />

complete loss of efficacy of the somatostatin<br />

analog in the control of IGF-1 secretions thereby<br />

possibly conferring a clinical benefit in<br />

patients with breast cancer.<br />

The molecular pharmacology of the relationship<br />

between the effects of estrogen, an-<br />

tiestrogens <strong>and</strong> somatostatin analogs on the<br />

expression of hSSTRs is complex <strong>and</strong> not fully<br />

understood. Firstly, estrogen has been reported<br />

to upregulate hSSTR-2 <strong>and</strong> hSSTR-3<br />

expression in prolactin-secreting rat pituitary<br />

tumor 7315b cells both in vivo <strong>and</strong> in vitro<br />

[56]. Secondly, tamoxifen either decreases<br />

or increases estrogen-induced expression of<br />

hSSTRs in different human breast cancer cell<br />

lines in vitro [96]. Thirdly, in vitro studies<br />

have demonstrated that the antineoplastic effects<br />

of somatostatin on MCF-7 estrogenreceptor-positive<br />

human breast cancer cells<br />

are attenuated by estradiol [97]. Fourthly, tamoxifen<br />

potentiates the suppression of both<br />

IGF-1 expression <strong>and</strong> serum levels of IGF-1<br />

[98, 99]. Finally, octreotide potentiates the<br />

antineoplastic effects of tamoxifen in oophorectomized<br />

rats with 7,12-dimethylbenz(a)anthracene-induced<br />

mammary tumors [100].<br />

The above observations clearly indicate that<br />

the relationships between somatostatin analogs,<br />

estrogens <strong>and</strong> antiestrogens are complex<br />

<strong>and</strong> require further investigation. Of particular<br />

importance is the as yet unknown effects of<br />

tamoxifen on individual hSSTR expression in<br />

patients with breast cancer. Studies in vitro<br />

indicate that this antiestrogen can upregulate<br />

or downregulate somatostatin receptor expression<br />

in different human breast cancer cell<br />

lines [96]. Clearly, depending on its overall<br />

effect on in vivo hSSTR expression in patients<br />

with breast cancer, tamoxifen could potentiate<br />

or attenuate the direct antineoplastic<br />

effects of somatostatin analog therapy. However,<br />

it should be pointed out that we do not<br />

as yet know the functional roles of hSSTRs in<br />

breast cancer <strong>and</strong>, hence, the possible therapeutic<br />

benefits of somatostatin analogs in<br />

terms of their direct antineoplastic effects.<br />

There is some preliminary evidence to suggest<br />

that hSSTR expression may play a functional<br />

role in the growth <strong>and</strong> development of breast<br />

cancer in man. Thus, preliminary observa-<br />

180 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


tions indicate that the presence of hSSTRs in<br />

the tumor of patients with breast cancer increases<br />

the probability of a longer disease-free<br />

survival compared to those in whom the tumors<br />

are hSSTR-negative [20]. It should be<br />

borne in mind, however, that hSSTRs are not<br />

homogeneously distributed in breast cancer<br />

[101]. Furthermore, the expression of individual<br />

hSSTRs in human breast cancer is variable<br />

[27]. The majority of human breast cancers<br />

do express hSSTR-2, the subtype whereby<br />

the current generation of somatostatin analogs<br />

mediate most of their antineoplastic effects.<br />

Some breast tumors do not, however,<br />

express hSSTR-2 [27]. These observations<br />

suggest that hSSTR-positive areas of the tumor<br />

may respond to the direct antineoplastic<br />

effects of somatostatin analog therapy thereby<br />

conferring a growth advantage compared to<br />

those neoplastic cells which do not express<br />

somatostatin receptors. However, any beneficial<br />

effects that somatostatin analog therapy<br />

may have in terms of their direct antineoplastic<br />

cells may be lost during long-term administration<br />

due to the development of somatostatin-resistant<br />

clones of breast tumor cells. At<br />

present, there is no clear indication of the<br />

temporal relationships between the commencement<br />

of somatostatin analog therapy<br />

<strong>and</strong> downregulation of the hSSTRs which mediate<br />

their direct antiproliferative effect in<br />

breast cancer. Furthermore, it is unknown<br />

whether tamoxifen, by upregulating hSSTR<br />

expression, potentiates the direct antineoplastic<br />

effects of somatostatin analogs. Alternatively,<br />

if tamoxifen downregulates hSSTR expression<br />

in breast cancer it could attenuate<br />

the direct antineoplastic effects of somatostatin<br />

analog therapy. Finally, tamoxifen may<br />

not have any effect on hSSTR expression in<br />

patients with breast cancer. Clearly, the issues<br />

discussed above are complex <strong>and</strong> will not be<br />

easy to resolve in patients with breast cancer.<br />

Nevertheless, there is an urgent need to an-<br />

swer at least some of these questions in order<br />

to fully rationalize the therapeutic potential<br />

of combination tamoxifen <strong>and</strong> somatostatin<br />

analog therapy in breast cancer. For example,<br />

intermittent somatostatin analog administration<br />

may be more beneficial than continuous<br />

treatment to delay the downregulation of<br />

hSSTR expression <strong>and</strong> the loss of efficacy in<br />

suppressing IGF-1 secretion during combination<br />

therapy with tamoxifen in breast cancer.<br />

There are a considerable number of permutations<br />

of how to maximize somatostatin analog<br />

therapy <strong>for</strong> breast carcinoma in combination<br />

with tamoxifen alone, as an adjuvant to chemotherapy<br />

prior to commencing tamoxifen<br />

treatment or intermittent administration, <strong>and</strong><br />

all eventualities need to be explored in good<br />

phase 2 trials be<strong>for</strong>e proceeding to phase 3<br />

r<strong>and</strong>omized controlled trials. All we can conclude<br />

at present is that somatostatin analogs<br />

in combination with tamoxifen confer no advantage<br />

over antiestrogen treatment alone in<br />

the management of metastatic breast cancer.<br />

Indeed in both r<strong>and</strong>omized controlled trials<br />

the complication rate was higher with combination<br />

therapy than with tamoxifen alone<br />

which may adversely affect the HRQL [91,<br />

92].<br />

In a third very small r<strong>and</strong>omized controlled<br />

trial tamoxifen was compared with triple<br />

therapy comprising the antiestrogen in<br />

combination with octreotide <strong>and</strong> a prolactin<br />

inhibitor in patients with breast cancer [102].<br />

Interestingly, unlike previous controlled trials<br />

[91, 92], serum IGF-1 levels were not significantly<br />

different between the two groups of<br />

patients [102]. The objective response rates<br />

were better in patients receiving triple therapy<br />

compared to tamoxifen alone but survival<br />

was not improved [102]. However, it is very<br />

difficult to make any definitive conclusions<br />

on the efficacy of triple therapy compared<br />

with tamoxifen alone since the study was considerably<br />

underpowered. Furthermore, in our<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 181


opinion it would be unwise to proceed to controlled<br />

trials comparing tamoxifen with polyhormonal<br />

therapies until the issues relating to<br />

combining the antiestrogen with somatostatin<br />

analogs have been fully resolved.<br />

It has been suggested that tamoxifen plus<br />

somatostatin analog therapy may be more<br />

effective than tamoxifen alone in the adjuvant<br />

setting than in metastatic breast cancer. This<br />

hypothesis is based on phase 2 studies suggesting<br />

that somatostatin analogs potentiate the<br />

effects of tamoxifen in suppressing serum<br />

IGF-1 levels [99] <strong>and</strong> from experimental studies<br />

which indicate that combination therapy<br />

is more effective than monotherapy in rats<br />

with a low tumor burden from those with<br />

more advanced neoplasia [100]. Indeed, clinical<br />

trials are currently in progress to examine<br />

this hypothesis. However, in view of the very<br />

disappointing results obtained with combined<br />

tamoxifen <strong>and</strong> somatostatin analog therapy in<br />

patients with metastatic breast disease [91,<br />

92] adjuvant trials of such combination therapy<br />

should be monitored carefully by independent<br />

Safety Monitoring <strong>and</strong> Advisory Committees<br />

(SMAC). Furthermore, adjuvant tamoxifen<br />

therapy <strong>for</strong> patients with early breast<br />

cancer is established as being both safe <strong>and</strong><br />

effective [103]. The only controversy over<br />

tamoxifen as an adjuvant therapy <strong>for</strong> early<br />

breast cancer is <strong>for</strong> how long treatment should<br />

be continued after surgery <strong>for</strong> early breast<br />

cancer, i.e. 1, 2 or 5 years [103]. Given that we<br />

know that prolonged somatostatin analog<br />

therapy can lead to downregulation of hSSTR<br />

expression <strong>and</strong> loss of efficacy in controlling<br />

IGF-1 secretion, it is very difficult to see a role<br />

<strong>for</strong> these drugs in very long-term adjuvant<br />

management programs <strong>for</strong> preventing recurrent<br />

early breast cancer. Furthermore, in view<br />

of the increased risk of side effects of combined<br />

tamoxifen <strong>and</strong> somatostatin analog<br />

therapy compared with tamoxifen alone in<br />

patients with metastatic breast cancer [91,<br />

92], particularly attention should be paid to<br />

HRQL by SMAC’s monitoring adjuvant trials<br />

of this combination therapy. When assessing<br />

HRQL account must also be taken of the pattern<br />

of the method of delivery of the somatostatin<br />

analogs, since although facilitated by the<br />

availability of long-acting depot preparations<br />

of octreotide, vapreotide <strong>and</strong> lanreotide, this<br />

still involves injections. Finally, although not<br />

always the ideal setting <strong>for</strong> carrying out complex<br />

economic analyses, this does need to be<br />

taken into account when considering prolonged<br />

expensive adjuvant therapies. Perhaps<br />

somatostatin analog therapy in combination<br />

with tamoxifen may prove to be superior to<br />

antiestrogen alone in the prevention of recurrent<br />

early breast cancer but we will not be able<br />

to answer this until the results of such trials<br />

become available.<br />

The rationale <strong>for</strong> the use of somatostatin<br />

analogs in the management of advanced prostatic<br />

cancer <strong>and</strong> the results of preliminary<br />

phase 1 <strong>and</strong> phase 2 clinical trials are discussed<br />

in detail by Vainas [91] in a separate<br />

paper in this volume. In brief, hSSTR expression<br />

is different in normal <strong>and</strong> neoplastic cells<br />

<strong>and</strong> between stromal <strong>and</strong> epithelial cells [34].<br />

Of particular importance is that apart from<br />

hSSTR-5, the other hSSTRs expressed in primary<br />

cultures of epithelial <strong>and</strong> stromal cells<br />

(table 2) do not bind with high affinity to the<br />

somatostatin analogs currently available <strong>for</strong><br />

clinical use [34]. These observations would<br />

suggest that any beneficial effects of octreotide,<br />

vapreotide or lanreotide in advanced<br />

prostatic cancer are likely to be mediated by<br />

their indirect rather than direct antineoplastic<br />

mechanisms of action. Exploitation of the full<br />

direct antineoplastic effects of somatostatin<br />

analog therapy in prostate cancer patients<br />

awaits (1) the development of novel analogs<br />

specific <strong>for</strong> hSSTR-1 or hSSTR-5 subtypes<br />

or (2) effective methods of delivering the<br />

hSSTR-2 gene to neoplastic cells. Thus, at<br />

182 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


Table 2. Expression of somatostatin<br />

receptor subtypes (hSSTR) in<br />

primary cultures of epithelial<br />

<strong>and</strong> stromal cells on normal human<br />

prostates <strong>and</strong> prostate<br />

cancers [after 34]<br />

present, any benefit of somatostatin analog<br />

therapy in prostatic cancer is likely to be<br />

mediated largely by their indirect antineoplastic<br />

effects.<br />

With respect to the clinical trials of somatostatin<br />

analog therapy in advanced prostate<br />

cancer it is difficult to make any conclusions<br />

on their therapeutic efficacy [91] since such<br />

studies in general included (1) patients with<br />

hormone-refractory or non-hormone-refractory<br />

disease, (2) relatively small numbers of<br />

patients, (3) used somatostatin analog therapy<br />

alone or combined with bromocriptine or<br />

complete <strong>and</strong>rogen blockade, (4) used wide<br />

variations in the dosing regimes <strong>and</strong><br />

(5) lacked objective outcome measures including<br />

HRQL. Consequently, at present it<br />

would be unwise to embark on large r<strong>and</strong>omized<br />

controlled trials of somatostatin analog<br />

therapy in cancer of the prostate but rather to<br />

carry out good quality phase 2 trials to determine<br />

the therapeutic efficacy <strong>and</strong> tolerability<br />

of octreotide, lanreotide <strong>and</strong> vapreotide in<br />

this indication.<br />

An obvious clinical setting to initially examine<br />

the therapeutic efficacy <strong>and</strong> tolerability<br />

of somatostatin analogs is in patients with<br />

hormone-refractory prostate cancer since at<br />

present there are no clear-cut treatment algorithms<br />

<strong>for</strong> second-line management. Hormone-refractory<br />

prostate cancer usually manifests<br />

itself after complete <strong>and</strong>rogen blockade<br />

<strong>and</strong> is thought to be the result of selection<br />

hSSTR-1 hSSTR-2 hSSTR-3 hSSTR-4 hSSTR-5<br />

Normal epithelial cells – + – + +<br />

Normal stromal cells – – – – –<br />

Neoplastic epithelial cells + – – + +<br />

Neoplastic stromal cells – – – – –<br />

– = Not detectable; + = detectable.<br />

<strong>and</strong>/or cloning of preexisting or de novo hormone-independent<br />

cell lines. Hormonal escape<br />

of prostate cancer cells during complete<br />

<strong>and</strong>rogen blockade is a gradual event <strong>and</strong><br />

monitoring of PSA has become accepted as<br />

the diagnostic test of choice to measure the<br />

progression of disease. Early diagnosis of hormonal<br />

escape based on a rise in PSA levels in<br />

patients with advanced prostate cancer who<br />

are still asymptomatic provides an opportunity<br />

<strong>for</strong> early institution of adjuvant therapies,<br />

<strong>and</strong> hence gives a reasonable chance <strong>for</strong> investigating<br />

both the efficacy <strong>and</strong> tolerability of<br />

the additional treatment being investigated.<br />

The principal outcome measure should be<br />

survival <strong>and</strong> the main secondary end point<br />

HRQL, the latter being assessed by both specific<br />

<strong>and</strong> generic instruments. With respect to<br />

somatostatin analog therapy as adjuvant to<br />

complete <strong>and</strong>rogen blockade some consideration<br />

should be given to whether treatment<br />

should be intermittent or continuous. It has<br />

been demonstrated that octreotide in combination<br />

with complete <strong>and</strong>rogen blockade in<br />

patients with advanced prostate cancer resulted<br />

in significant reductions in the serum<br />

levels of IGF-1 <strong>and</strong> epidermal growth factor<br />

(EGF) which paralleled the reduction in PSA<br />

[104]. Possibly there<strong>for</strong>e, serum IGF-1 <strong>and</strong><br />

EGF levels could be used as serum markers of<br />

early escape from somatostatin analog therapy.<br />

<strong>Somatostatin</strong> analog therapy could then<br />

be stopped <strong>and</strong> re-instituted when the serum<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 183


levels of IGF-1 <strong>and</strong> EGF return to those<br />

observed prior to commencing treatment. A<br />

reduction of serum IGF-1 <strong>and</strong> EGF levels<br />

after recommencing somatostatin analog therapy<br />

to those observed when these drugs were<br />

initially administered would be strongly suggestive<br />

that intermittent treatment dosing regimes<br />

may be preferable to continuous administration.<br />

This hypothesis clearly needs to<br />

be initially substantiated in phase 2 trials. It is<br />

m<strong>and</strong>atory that serum PSA levels are monitored<br />

at regular intervals in patients with advanced<br />

prostate cancer being treated with a<br />

somatostatin analog as an adjuvant in the<br />

event that combination therapy results in rapid<br />

progression of disease.<br />

In patients with advanced prostate cancer<br />

in whom the disease is completely refractory<br />

to complete <strong>and</strong>rogen blockade, somatostatin<br />

analog therapy appears to be of very limited<br />

value [91]. Only in this setting should<br />

phase 2 trials of somatostatin analog therapy<br />

combined with other treatment modalities<br />

such as bombesin antagonists, LHRH antagonists<br />

or chemotherapy be considered. The<br />

principal primary <strong>and</strong> secondary outcome<br />

measures should again be survival <strong>and</strong><br />

HRQL. As discussed above, consideration<br />

should be given as to whether somatostatin<br />

analog therapy should be intermittent or<br />

continuous.<br />

In recent years, there has been increasing<br />

interest on whether or not intermittent complete<br />

<strong>and</strong>rogen blockade may delay the onset<br />

of hormone-refractory prostate cancer compared<br />

to continuous treatment <strong>and</strong> hence improve<br />

survival <strong>and</strong> increase the tolerability of<br />

chemical castration [64, 65]. Although the<br />

results of preliminary trials suggest that intermittent<br />

complete <strong>and</strong>rogen blockade may be<br />

superior to continuous treatment [64, 65], a<br />

comparison of these two <strong>for</strong>ms of treatment<br />

urgently requires confirmation in a r<strong>and</strong>omized<br />

controlled trial. It is tempting to specu-<br />

late that somatostatin analog therapy may<br />

confer an additional benefit, if used <strong>for</strong> limited<br />

periods between cycles of <strong>and</strong>rogen suppression<br />

in patients with advanced prostate<br />

cancer. Increases in the serum levels of IGF-1<br />

or EGF could be used as indicators of escape<br />

from hormone suppression by the somatostatin<br />

analogs <strong>and</strong> to discontinue treatment.<br />

Serum PSA could be used to both detect any<br />

deleterious effect of somatostatin analog therapy<br />

on tumor progression <strong>and</strong> <strong>for</strong> recommencing<br />

<strong>and</strong>rogen suppression. Intermittent<br />

<strong>and</strong>rogen suppression <strong>and</strong> somatostatin analog<br />

therapy may be a tempting therapeutic<br />

modality <strong>for</strong> the management of advanced<br />

prostate cancer, particularly in the younger<br />

male but it has to be borne in mind that intermittent<br />

<strong>and</strong>rogen blockade has yet to be demonstrated<br />

to be superior to continuous therapy<br />

in prolonging survival. Furthermore, the<br />

potential <strong>for</strong> harm exists <strong>for</strong> both intermittent<br />

<strong>and</strong>rogen suppression <strong>and</strong> <strong>for</strong> somatostatin<br />

analog therapy. Any trials evaluating alternating<br />

<strong>and</strong>rogen suppression with somatostatin<br />

analog therapy should be very carefully monitored.<br />

Pancreatic cancer is associated with an extremely<br />

poor prognosis even following a theoretically<br />

curative resection. The etiology,<br />

prognosis <strong>and</strong> therapeutic modalities <strong>for</strong> adenocarcinoma<br />

of the pancreas are discussed in<br />

detail by Rosenberg [82] in another paper in<br />

this volume. In brief, pancreatic adenocarcinoma<br />

cells express receptors <strong>for</strong> estrogens,<br />

progesterone <strong>and</strong> <strong>and</strong>rogens suggesting that<br />

this neoplasm may be responsive to hormonal<br />

treatment. Indeed, there is some evidence to<br />

suggest that somatostatin analog therapy combined<br />

with either tamoxifen [105] or LHRH<br />

analogs [106] may have some clinically beneficial<br />

effects in adenocarcinoma of the pancreas<br />

<strong>and</strong> the results of these studies are discussed<br />

in detail by Rosenberg [82]. Suffice it<br />

to say that both studies suggest that pancreatic<br />

184 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


cancer may be responsive to hormonal manipulation<br />

with somatostatin analog therapy<br />

combined with either tamoxifen [105] or<br />

LHRH analogs [106] <strong>and</strong> warrants further<br />

investigation particularly in patients with<br />

small tumors who undergo a theoretically curative<br />

resection.<br />

Immunotherapy<br />

<strong>Cancer</strong> cells expressing tumor-specific antigens<br />

initiate a complex immune response<br />

involving a complex cascade of cellular <strong>and</strong><br />

molecular interactions to destroy the tumor<br />

cell [107]. There is accumulating evidence to<br />

suggest that the failure of the immune system<br />

to eliminate a tumor results from the ability of<br />

tumor antigens to stimulate an effective immune<br />

response [107]. Possibilities <strong>for</strong> such<br />

failure include loss of major histocompatibility<br />

class 1 or its costimulator B7 expression,<br />

aberrant antigen presentation <strong>and</strong> the inhibition<br />

of the cellular <strong>and</strong> molecular components<br />

of the immune system by factors secreted<br />

from or expressed on tumor cells [107]. Further,<br />

surgical or anesthetic trauma compromise<br />

immune function [108–114] <strong>and</strong> can<br />

stimulate the release of growth factors [115].<br />

Thus, in the immediate postoperative period<br />

in the cancer patient, these changes in immune<br />

function may accelerate the growth of<br />

microscopic local tumor cells <strong>and</strong> distant micrometastatic<br />

deposits <strong>and</strong> increase the risk of<br />

tumor cells that shed into the circulation during<br />

resection of the primary cancer establishing<br />

themselves as distant micrometastases.<br />

Not surprisingly, in the past 20 years<br />

there has been increasing interest in developing<br />

therapies to stimulate the immune response<br />

against tumors in patients with advanced<br />

metastatic disease to stabilize disease<br />

progression <strong>and</strong> as an adjuvant to surgery to<br />

increase the chances of achieving a curative<br />

resection. A variety of immunotherapies have<br />

been suggested to stimulate the immune system<br />

in cancer patients including (1) intralesional<br />

administration of nonspecific immunostimulants<br />

such as BCG or Cryptosporidium<br />

parvum, (2) administration of cytokines,<br />

(3) adoptive immunotherapy with lymphokine-activated<br />

killer cells administered<br />

either systematically or regionally, alone or in<br />

combination with cytokines or chemotherapy<br />

<strong>and</strong> (4) combination therapy with cytokines<br />

<strong>and</strong> immunomodulating hormones [116–<br />

120]. It is beyond the scope of this paper to<br />

discuss the clinical efficacy of all these therapeutic<br />

modalities except to state that none of<br />

these immunotherapies <strong>for</strong> neoplasia have<br />

fulfilled their potential <strong>and</strong> that future strategies<br />

are likely to focus on genetic immunotherapy.<br />

Of particular relevance to this paper<br />

are the consequences of somatostatin analog<br />

therapy on the immune response to cancer<br />

<strong>and</strong> whether or not the clinical efficacy of<br />

these drugs could be improved by concomitant<br />

immunotherapy.<br />

Initial interest in the use of combination<br />

somatostatin analog <strong>and</strong> immunotherapy focused<br />

on the possible synergistic effects of<br />

octreotide <strong>and</strong> interferon-· in inhibiting the<br />

proliferation of endocrine tumor cells [120].<br />

This approach appeared to be justified by early<br />

anecdotal reports suggesting that octreotide<br />

<strong>and</strong> interferon-· in combination appeared to<br />

be a valuable immunoendocrine therapy <strong>for</strong><br />

metastatic carcinoid tumors [121, 122]. However,<br />

interferon-· elicits a number of other<br />

immune responses apart from its antiproliferative<br />

effect on endocrine cells including stimulation<br />

of natural killer cell activity [118] suggesting<br />

that it may be a valuable adjuvant to<br />

somatostatin analog treatment of nonendocrine<br />

tumors. Un<strong>for</strong>tunately, the therapeutic<br />

potential of somatostatin analog therapy combined<br />

with immunotherapies has been relatively<br />

ignored as an option <strong>for</strong> the manage-<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 185


ment of cancer. This may represent a considerable<br />

oversight since hSSTRs are expressed<br />

on a variety of normal <strong>and</strong> neoplastic lymphoid<br />

cells [38]. Furthermore, somatostatin<br />

analogs inhibit the proliferation of a variety of<br />

lymphoid cells, immunoglobulin production<br />

<strong>and</strong> natural killer cell activity [38, 123–126].<br />

These inhibitory effects of somatostatin analog<br />

therapy on the immune system may have<br />

profound implications on their therapeutic efficacy<br />

in neoplasia. Apart from suggesting<br />

that somatostatin analog therapy may play a<br />

role in the management of lymphomas, these<br />

drugs may attenuate or completely negate any<br />

potential therapeutic benefit in terms of the<br />

direct <strong>and</strong> indirect antineoplastic effects on<br />

tumor growth. It is even possible that the<br />

immunosuppressive effects of somatostatin<br />

analogs outweigh their therapeutic beneficial<br />

effects resulting in a more rapid tumor progression.<br />

Although purely speculative at<br />

present the immunosuppressant effects of somatostatin<br />

analog may account, at least in<br />

part, <strong>for</strong> the very disappointing results of<br />

these drugs in advanced malignancy. Furthermore,<br />

careful consideration should be given<br />

to the use of adjuvant somatostatin analog<br />

therapy as an adjuvant to surgery in the immediate<br />

postoperative period. As discussed<br />

above, surgery <strong>and</strong> anesthesia compromise<br />

the immune system [108–114] <strong>and</strong> administration<br />

of somatostatin analogs intra- or postoperatively<br />

may, by further immunocompromising<br />

the patient, accelerate the growth<br />

of tumor deposits remaining after radical resection.<br />

There is an abundance of evidence to<br />

indicate that somatostatin analog therapy<br />

used as <strong>and</strong> adjuvant to surgery may prevent<br />

the growth <strong>and</strong> development of hepatic metastases<br />

via stimulation of hepatic RES activity<br />

[11] but this may be at the expense of early<br />

local recurrence. Similarly, somatostatin analogs,<br />

because of their inhibitory effects on other<br />

functions of the immune system, may ac-<br />

celerate the growth of established micrometastases<br />

in locations other than the liver.<br />

Again these hypotheses require further investigation.<br />

Finally, <strong>and</strong> perhaps more importantly,<br />

investigations are required to determine<br />

whether or not it is possible to potentiate<br />

the individual benefits of somatostatin<br />

analog <strong>and</strong> concomitant therapies such as the<br />

administration of cytotoxics, by stimulating<br />

immune function by adoptive immunotherapy<br />

or genetic immunotherapy by combination<br />

treatments. Resolving all the problems discussed<br />

above represent a considerable challenge<br />

to oncologists <strong>and</strong> immunologists <strong>and</strong><br />

require carefully designed studies in experimental<br />

animals <strong>and</strong> in man. Nevertheless,<br />

such studies are required to substantiate the<br />

above hypotheses since they are critical to<br />

optimizing somatostatin analog therapy of<br />

neoplasia.<br />

Novel Therapeutic Approaches to the<br />

Management of Neoplasia<br />

Targeted Radiotherapy<br />

<strong>Somatostatin</strong> scintigraphy using stable radiolabelled<br />

compounds of octreotide <strong>and</strong><br />

more recently vapreotide is a well-established<br />

technique <strong>for</strong> the localization of primary <strong>and</strong><br />

metastatic tumors expressing hSSTR-2 <strong>and</strong><br />

hSSTR-5. <strong>Somatostatin</strong> receptor scintigraphy<br />

can, <strong>and</strong> indeed should, be used <strong>for</strong> the selection<br />

of patients with neuroendocrine <strong>and</strong> other<br />

solid tumors who are likely to benefit from<br />

somatostatin therapy <strong>and</strong> <strong>for</strong> monitoring<br />

treatment. The major limitation of somatostatin<br />

receptor scintigraphy using radiolabelled<br />

lig<strong>and</strong>s of octreotide is that the technique<br />

will only allow detection of those tumors<br />

expressing hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong><br />

possibly those neoplasms expressing hSSTR-3<br />

in sufficient density to allow visualization.<br />

Radiolig<strong>and</strong>s of vapreotide may be more use-<br />

186 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


ful than radiolabelled lig<strong>and</strong>s of octreotide in<br />

visualizing those tumors that express hSSTR-<br />

4 in a sufficient density, but not hSSTR-2 <strong>and</strong><br />

hSSTR-5. Visualization of tumors by vapreotide<br />

scintigraphy but not by octreotide scintigraphy<br />

may provide a rationale <strong>for</strong> the selection<br />

of patients that are likely to benefit from<br />

therapy with vapreotide or lanreotide rather<br />

than octreotide, but this hypothesis requires<br />

confirmation in prospective controlled trials.<br />

Targeted radiotherapy based on the binding<br />

of somatostatin analog radiolig<strong>and</strong>s to tumors<br />

expressing hSSTRs is an obvious extension<br />

of somatostatin scintigraphy. This is not<br />

a new concept <strong>and</strong> as long ago as 1995 Wiseman<br />

<strong>and</strong> Kvols [127] reviewed the results<br />

of targeted radiotherapy using 131 I-metaiodobenzylguanidine<br />

( 131 I-MIBG) <strong>and</strong> 111 In-<br />

DTPA-D-Phe1octreotide ( 111 In-pentetreotide)<br />

in patients with pheochromocytoma,<br />

neuroblastoma, carcinoid tumors, medullary<br />

thyroid carcinoma <strong>and</strong> paragangliomas. Targeted<br />

radiotherapy with 131 I-MIBG resulted<br />

in variable tumor responses with the most<br />

encouraging results being observed in patients<br />

with pheochromocytomas [127]. 111 In-pentetreotide<br />

has also been used <strong>for</strong> targeting radiotherapy<br />

in a very small number of patients<br />

with neuroendocrine tumors <strong>and</strong> resulted in<br />

objective tumor responses [127]. These authors<br />

concluded that targeted radiotherapy<br />

with somatostatin analogs coupled to ß- or ·emitting<br />

radioisotopes would be necessary to<br />

obtain a significant <strong>and</strong> desirable tumor response<br />

[127]. This point has been emphasized<br />

repeatedly by medical oncologists <strong>and</strong> indeed<br />

there are a number of experimental studies<br />

evaluating ß- or ·-emitting somatostatin radiolig<strong>and</strong>s.<br />

For example, rhenium-188 (t 1 /2<br />

16.9 h; beta-max 2.5 units) coupled to vapreotide<br />

administered intralesionally (prostate) or<br />

intracavitarily (mammary <strong>and</strong> small cell lung<br />

cancer) significantly reduced or eliminated<br />

the tumor burden in nude mice [128]. Local<br />

or regional administration of radiolabelled<br />

high energy ß- or Á-radiolig<strong>and</strong>s of somatostatin<br />

analogs may have some advantages over<br />

intravenous administration. Firstly, local or<br />

regional administration of such radiolig<strong>and</strong>s<br />

may reduce systemic toxicity. Secondly, somatostatin<br />

receptors are widely distributed in<br />

normal tissues throughout the body <strong>and</strong> intravenous<br />

administration of high energy ß- or Áradiolig<strong>and</strong>s<br />

of somatostatin analogs may produce<br />

side effects over <strong>and</strong> above those observed<br />

with regional conventional or somatostatin<br />

analog-targeted radiotherapy. Finally,<br />

recent studies have suggested that there may<br />

be a blood-tumor barrier, thereby possibly<br />

preventing diffusion of radiolabelled somatostatin<br />

analogs from the capillary tumor to<br />

neoplastic cells [129].<br />

The concept of targeted radiotherapy of<br />

tumors using radiolig<strong>and</strong>s of somatostatin<br />

analogs remains a very attractive approach<br />

<strong>for</strong> the treatment of neoplasia. The fact that it<br />

is over 12 years since the concept of octreotide-targeted<br />

radiotherapy of neoplasia was<br />

first proposed <strong>and</strong> we are still awaiting any<br />

good phase 2 clinical trials of the efficacy <strong>and</strong><br />

tolerability of this technique highlights the<br />

practical difficulties involved in developing<br />

this therapy. Nevertheless, a number of<br />

groups are actively investigating targeted radiotherapy<br />

of tumors using high energy ß- <strong>and</strong><br />

Á-emitters of somatostatin analog radiolig<strong>and</strong>s<br />

<strong>and</strong> the results of early studies in patients<br />

are awaited with interest.<br />

Angiogenesis<br />

Angiogenesis is defined as the development<br />

of de novo capillaries from preexisting<br />

blood vessels <strong>and</strong> is an important process in<br />

the growth of solid cancers. The pathophysiology<br />

of angiogenesis in solid tumors <strong>and</strong> the<br />

potential role of somatostatin analogs as antiangiogenic<br />

drugs in neoplasia has been comprehensively<br />

reviewed by Danesi <strong>and</strong> Del<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 187


Tacca [130] <strong>and</strong> Woltering et al. [131]. In<br />

brief, using in vitro models of angiogenesis,<br />

octreotide inhibits the development but not<br />

the growth of new capillaries [130, 131]. Furthermore,<br />

octreotide has been demonstrated<br />

to inhibit angiogenesis after chemical-induced<br />

corneal injury in rats <strong>and</strong> possibly in<br />

patients with diabetic retinopathy [130]. Although<br />

it is almost impossible to quantify the<br />

antiangiogenic effects of somatostatin analog<br />

therapy on tumor progression directly, inhibition<br />

of angiogenesis is theoretically an important<br />

mechanism whereby these drugs exert<br />

their antineoplastic effects. Thus, tumor cells<br />

are genetically diverse <strong>and</strong> unstable <strong>and</strong> hence<br />

the direct <strong>and</strong> indirect antineoplastic effects<br />

of somatostatin analog therapy may not result<br />

in complete elimination of all cancer cells. In<br />

contrast, endothelial cells in tumor vessels are<br />

genetically more stable <strong>and</strong> less likely to develop<br />

resistance to therapeutic procedures.<br />

For example, endogenous antagonists to angiogenesis<br />

such as angiostatin <strong>and</strong> endostatin<br />

which selectively inhibit endothelial cells to<br />

respond to antigenic signals elicit marked regression<br />

of tumors in experimental animals<br />

[132, 133]. Furthermore, gene transfer of angiostatin<br />

significantly inhibits tumor growth<br />

in experimental animals [134]. Alternatively,<br />

although a large number of angiogenic factors<br />

are implicated in tumor vascularization, strategies<br />

targeted at blocking these factors can<br />

reduce tumor growth [131]. A promising approach<br />

to inhibition of tumor growth by inhibiting<br />

angiogenesis is the development of<br />

gene therapies. For example, transfection of<br />

tumor cells with antisense oligonucleotides<br />

against vascular endothelial growth factor<br />

mRNA has been reported to reduce the<br />

growth rate of gliomas [135]. Similarly, genetic<br />

transfer of a soluble Tie 2 receptor (endothelium-specific<br />

receptor tyrosine kinase)<br />

blocks activation of Tie 2 receptors on tumor<br />

cells [135]. Furthermore, these investigators<br />

demonstrated that treatment of tumor-bearing<br />

animals with the soluble Tie 2 receptor<br />

using recombinant adenovirus as a vector almost<br />

completely inhibited angiogenesis but<br />

significantly inhibited the growth of both the<br />

primary tumor <strong>and</strong> metastases [135]. Although<br />

the efficacy <strong>and</strong> tolerability of the specific<br />

antagonists <strong>and</strong> genetic therapy of angiogenesis<br />

has not yet been reported in man, clinical<br />

studies are ongoing. It appeared logical to<br />

us that the direct <strong>and</strong> indirect antineoplastic<br />

effects of somatostatin analog therapy combined<br />

with its antigenic action could potentiate<br />

the effects of specific antagonists <strong>and</strong><br />

gene therapy of angiogenesis. We are currently<br />

evaluating this hypothesis intensively <strong>and</strong> if<br />

the results of our studies suggest that somatostatin<br />

analogs potentiate that of specific antiangiogenic<br />

therapies, this may prove to be a<br />

very valuable approach to the management of<br />

neoplasia.<br />

Gene Therapy<br />

Gene therapy has immense potential <strong>for</strong><br />

the treatment of many <strong>for</strong>ms of diseases including<br />

cancer [136]. Although there are a<br />

number of problems which remain to be resolved<br />

be<strong>for</strong>e gene therapy can be routinely<br />

applied to patients with malignancy, initial<br />

studies in man are very promising [136]. With<br />

respect to somatostatin analog therapy there<br />

are a number of areas in which effective gene<br />

therapy may be used to potentiate the antineoplastic<br />

effects of these drugs. Perhaps the<br />

most obvious application of gene therapy to<br />

somatostatin analog treatment of neoplasia is<br />

the delivery of hSSTR-2 <strong>and</strong> hSSTR-5 genes<br />

together with the genes that encode their<br />

membrane proteins to cancers such as pancreatic,<br />

gastric <strong>and</strong> colorectal cancers that do<br />

not express these receptor subtypes. The somatostatin<br />

analogs currently available <strong>for</strong><br />

clinical uses, octreotide, vapreotide <strong>and</strong> lanreotide,<br />

exert the majority of their antineo-<br />

188 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


plastic effects via hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong><br />

it follows there<strong>for</strong>e that effective transfer of<br />

genes encoding <strong>for</strong> hSSTR-2 <strong>and</strong> hSSTR-5 as<br />

well as their membrane proteins to cancers<br />

which do not express these receptors subtypes<br />

may render them responsive to the direct antineoplastic<br />

effects of the current generation of<br />

somatostatin analogs. In some tumors such as<br />

breast cancer, hSSTRs are not uni<strong>for</strong>mly distributed<br />

throughout the neoplastic tissue<br />

[100]. Using genetic transfer of genes encoding<br />

<strong>for</strong> hSSTR-2 <strong>and</strong> hSSTR-5 <strong>and</strong> their<br />

membrane proteins, it may be possible to<br />

increase the number of cells expressing<br />

hSSTRs, thereby enhancing the efficacy of<br />

somatostatin analog therapy. Tissue targeting<br />

is crucial to effective gene therapy <strong>and</strong> this<br />

problem has still to be overcome be<strong>for</strong>e such<br />

treatment can become part of routine clinical<br />

practice <strong>for</strong> common malignancies. The use of<br />

retroviral vectors which preferentially target<br />

cancer cells by their specific integration into<br />

proliferating cells may be one option. However,<br />

perhaps a more practical approach to<br />

ensure tumor-restricted expression of the<br />

transgenes is the use of tumor-specific promoters<br />

or high affinity monoclonal antibodies.<br />

Currently, there is an explosion of<br />

activity in researching effective, specific <strong>and</strong><br />

less toxic delivery of genes to target tissue <strong>and</strong><br />

it is not overoptimistic to speculate that in the<br />

not too distant future, genetic transfer of<br />

hSSTRs <strong>and</strong> their membrane proteins will potentiate<br />

the direct antineoplastic effects of<br />

somatostatin analogs in tumors which do not<br />

express hSSTR-2 or hSSTR-5. It is also not<br />

inconceivable that genetic transfer of hSSTRs<br />

to hormone-resistant tumor cells in tumors<br />

resulting from prolonged somatostatin analog<br />

administration may prevent or delay escape<br />

from the antineoplastic effects of these drugs.<br />

<strong>Somatostatin</strong> analog therapy of cancer<br />

may also be enhanced by genetic therapy of<br />

oncogenes <strong>and</strong> replacement of tumor suppres-<br />

sor genes. For example, antisense inhibition<br />

of IGF-1 has been demonstrated to induce<br />

apoptosis in rat hepatocellular carcinoma<br />

cells [137]. A number of tumors express IGF-<br />

1 receptors <strong>and</strong> suppression of IGF-1 levels by<br />

somatostatin analogs, as discussed previously,<br />

provides a rationale <strong>for</strong> the indirect antineoplastic<br />

effects of these drugs in a variety of<br />

human neoplasms. A combination of somatostatin<br />

analog therapy alone or combined<br />

with other hormonal treatments or chemotherapy<br />

on tumor cells transfected with antisense<br />

inhibitors of IGF-1 may well result in<br />

increased tumor responses.<br />

A high proportion of human cancers contain<br />

mutations of the p53 suppressor gene<br />

[138]. There is considerable evidence to suggest<br />

that p53-dependent apoptosis plays an<br />

important role in enhancing the therapeutic<br />

efficacy of chemotherapy <strong>and</strong> radiotherapy.<br />

Conversely, tumors that contain p53 mutations<br />

generally, but not always, are more resistant<br />

to chemotherapy <strong>and</strong> radiotherapy. Not<br />

surprisingly, there<strong>for</strong>e, there has been considerable<br />

interest in restoration or modulation of<br />

p53 function by gene therapy to increase<br />

apoptosis <strong>and</strong> enhance tumor sensitivity to<br />

chemotherapy <strong>and</strong> radiotherapy. Studies on<br />

human cancer cell lines <strong>and</strong> experimental animals<br />

suggest that restoration of p53 function<br />

using retroviral or adenoviral vectors inhibits<br />

cell proliferation in vitro, the growth of tumors<br />

in vivo <strong>and</strong> sensitizes tumor cells to chemotherapy<br />

[139–141]. Further, as long ago as<br />

1996, Roth et al. [139] reported that retroviral<br />

wild-type p53 gene transfer in 9 patients with<br />

non-small-cell lung cancer, in whom conventional<br />

therapy had failed, resulted in increased<br />

apoptosis <strong>and</strong> tumor regression in 3<br />

patients. These data suggest that gene replacement<br />

with p53 is a promising therapeutic<br />

approach to the management of malignancy.<br />

However, it should be pointed out that as yet,<br />

available gene vectors <strong>for</strong> p53 are not highly<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 189


efficient but work is ongoing to increase p53<br />

transduction efficiency. Furthermore, malfunctions<br />

of other components of the p53<br />

pathway may be important in the development<br />

of malignancy. For example, overexpression<br />

of the mdm-2 oncogene (which encodes<br />

<strong>for</strong> a phosphoprotein which blocks the<br />

activity of p53) in human sarcomas may be<br />

responsible <strong>for</strong> the development towards malignancy<br />

although p53 per se is normal [142].<br />

Since mdm-2 interacts with p53 via a small<br />

molecular interface to inhibit its function, it is<br />

possible that a small molecule could be designed<br />

to disrupt this interaction. Nevertheless,<br />

in spite of the complexity of p53 function<br />

in carcinogenesis, the observation that p53<br />

can be restored by gene therapy in some<br />

tumors raises the possibility of combining<br />

such treatment with somatostatin analog administration.<br />

This suggestion is based on observations<br />

which suggest somatostatin analogs<br />

upregulate p53 function <strong>and</strong> hence apoptosis<br />

by activation of hSSTR-3 [143]. Possibly,<br />

there<strong>for</strong>e, the beneficial effects of genetic<br />

transfer of p53 to tumors that contain mutations<br />

of this suppressed gene may be enhanced<br />

by combination therapy with somatostatin<br />

analogs. Upregulation of p53 function<br />

by combination with somatostatin analogs<br />

may also increase the sensitivity of the tumors<br />

to chemotherapy <strong>and</strong> radiation therapy.<br />

Clearly, further studies are required to substantiate<br />

these hypotheses.<br />

The examples of a possibility of potentiating<br />

gene therapy by concomitant somatostatin<br />

analog administration discussed above <strong>and</strong> in<br />

the sections on Angiogenesis <strong>and</strong> Immunotherapy<br />

only represent a small number of the<br />

large number of possibilities <strong>for</strong> maximizing<br />

the therapeutic efficacy of these drugs by taking<br />

advantage of the impact that the ‘genetic<br />

revolution’ is likely to make on the management<br />

of neoplasia in the not too distant future.<br />

Summary <strong>and</strong> Conclusion<br />

The role of somatostatin therapy in the<br />

management of acromegaly is well established.<br />

Similarly, somatostatin analog therapy<br />

of neuroendocrine tumors is accepted as a<br />

valuable treatment option <strong>for</strong> providing excellent<br />

palliation when all else fails, although<br />

there may be room <strong>for</strong> improvement. In contrast<br />

the therapeutic efficacy of somatostatin<br />

analog monotherapy in the management of<br />

nonendocrine solid tumors has proven to be<br />

largely very disappointing. Combination therapy<br />

of somatostatin analogs with cytotoxics<br />

or other hormonal treatments targeted somatostatin<br />

analog chemotherapy or radiotherapy<br />

in both advanced malignancy <strong>and</strong> in<br />

the adjuvant setting may prove to be very<br />

much more effective than somatostatin analog<br />

monotherapy. Carefully controlled clinical<br />

studies with objective outcome measures <strong>and</strong><br />

HRQL assessments are required to evaluate<br />

combination therapies. Particular attention<br />

should be given to the possible adverse effects<br />

of somatostatin analog therapy on the immune<br />

response to neoplasia <strong>and</strong> consideration<br />

given to the possibility of concomitant<br />

immunostimulatory therapy. The question of<br />

whether or not intermittent somatostatin<br />

analog therapy is preferable to continuous<br />

treatment needs to be addressed <strong>for</strong>mally.<br />

This is now possible with slow-release preparations<br />

of octreotide, vapreotide <strong>and</strong> lanreotide.<br />

Other possible novel therapeutic approaches<br />

such as combination treatment with<br />

antiangiogenic drugs require investigation. Of<br />

particular importance in the future is the need<br />

to exploit the major advances in gene therapy<br />

to optimize the full potential of somatostatin<br />

analogs in the management of neoplasia. Although<br />

somatostatin analogs have not turned<br />

out to be the ‘magic bullet’ <strong>for</strong> the management<br />

of cancer as many oncologists thought<br />

they would when octreotide was first intro-<br />

190 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


duced into clinical medicine all is not doom<br />

<strong>and</strong> gloom. Indeed, if we heed the lessons<br />

learned from previous experience with somatostatin<br />

analog therapy in the management of<br />

neoplasia <strong>and</strong> logically <strong>and</strong> <strong>for</strong>mally explore<br />

possible combinational treatments to maximize<br />

the therapeutic potential of these drugs in<br />

cancer the rewards could be significant. To<br />

quote Winston Churchill at the end of the<br />

References<br />

1 Krulich L, Dharwalh A, McCann S:<br />

Stimulated <strong>and</strong> inhibitory effects of<br />

purified hypothalmic extracts on<br />

growth hormone release from the<br />

rat pituitary in vivo. Endocrinology<br />

1968;83:783–790.<br />

2 Brazeau P, Vale W, Burger R, Ling<br />

N, Butcher M, Rivier J, Guilleman<br />

R: Hypothalmic peptide that inhibits<br />

the secretion of immunoreactive<br />

pituitary growth hormone. Science<br />

1973;179:77–79.<br />

3 Rivier J: <strong>Somatostatin</strong>: Total solid<br />

phase synthesis. J Am Chem Soc<br />

1974;96:2986–2992.<br />

4 Larson LI: Gastrointestinal cells<br />

producing endocrine, neueuroendocrine<br />

<strong>and</strong> paracrine messengers.<br />

Clin Gastroenterol 1980;9:485–<br />

515.<br />

5 Brown M, Vale W: Central nervous<br />

system effects of hypothalmic peptides.<br />

Endocrinology 1975;96:1333–<br />

1336.<br />

6 Cohen M, Rosing E, Wiley K, Slater<br />

I: <strong>Somatostatin</strong> inhibits adrenergic<br />

<strong>and</strong> cholinergic neurotransmission<br />

in smooth muscle. Life Sci 1978;23:<br />

1659–1664.<br />

7 Rioux F, Kerovac R, St Pierre S:<br />

<strong>Somatostatin</strong>: Interaction with the<br />

sympathetic nervous system in guinea<br />

pigs. Neuropeptides 1981;1:319–<br />

327.<br />

8 Belanger A, Labrie F, Borgeat M:<br />

Inhibition of growth hormone <strong>and</strong><br />

thyrotropin release by growth release<br />

inhibiting hormone. Mol Cell<br />

Endocrinol 1974;1:329–338.<br />

9 Sheppard M, Shapiro B, Berelowitz<br />

M, Pimstone B: Metabolic clearance<br />

<strong>and</strong> plasma half-life disappearance<br />

time of endogenous somatostatin in<br />

man. J Clin Endocrinol Metab 1979;<br />

48:50–53.<br />

10 Hofl<strong>and</strong> LJ, Visser-Wisselaar HA,<br />

Lamberts SW: <strong>Somatostatin</strong> analogues:<br />

Clinical application in relation<br />

to human somatostatin receptor<br />

subtypes. Biochem Pharmacol 1995;<br />

50:287–297.<br />

11 Davies N, Cooke TG, Jenkins SA:<br />

Therapeutic potential of octreotide<br />

in the treatment of liver metastases.<br />

Anticancer Drugs 1996;7(suppl 1):<br />

23–31.<br />

12 Meyerhof W: The elucidation of somatostatin<br />

receptor functions: A<br />

current view. Rev Physiol Biochem<br />

Pharmacol 1998;183:55–108.<br />

13 Polak MN, Schally AV: Mechanisms<br />

of antineoplastic action of somatostatin<br />

analogues. Proc Soc Exp<br />

Biol Med 1998;217:143–152.<br />

14 Bousquet C, Puente E, Buscail L,<br />

Vaysse N, Suisini C: Antiproliferative<br />

effect of somatostatin <strong>and</strong><br />

analogs. Chemotherapy 2001;47<br />

(suppl 2):30–39.<br />

15 Patel YC: Molecular pharmacology<br />

of somatostatin receptor subtypes.<br />

J Endocrinol Invest 1997;20:348–<br />

367.<br />

16 Patel YC, Greenwood MT, Panetta<br />

R, Demchyslyn L, Niznik H, Srikant<br />

CB: The somatostatin receptor<br />

family. Life Sci 1995;57:1249–<br />

1265.<br />

North African Campaign in the Second<br />

World War: ‘This is not the end, nor is it the<br />

beginning of the end, it is only the end of the<br />

beginning’. He proved to be correct <strong>and</strong> this<br />

may very well turn out to be the case <strong>for</strong> the<br />

role of somatostatin analogs in the management<br />

of cancer.<br />

17 Reubi JC: Octreotide <strong>and</strong> nonendocrine<br />

tumours. Basic knowledge <strong>and</strong><br />

therapeutic potential; in Lomax P,<br />

Scarpignato C, Vessel E (eds): Octreotide<br />

from Basic Science to Clinical<br />

Medicine. Prog Basic Clin Pharmacol.<br />

Basel, Karger, 1996, vol 10,<br />

pp 246–269.<br />

18 Miller GV, Preston SR, Woodhouse<br />

LF, Farmery SM, Primrose JN: <strong>Somatostatin</strong><br />

binding in human gastrointestinal<br />

tissues: Effects of cations<br />

<strong>and</strong> somatostatin analogues.<br />

Gut 1993;34:1351–1356.<br />

19 Reubi JC, Laissue J, Kenniag E,<br />

Lamberts SW: <strong>Somatostatin</strong> receptors<br />

in human cancer: Incidence,<br />

characteristics, functional correlates<br />

<strong>and</strong> clinical implications. J Steroid<br />

Biochem Mol Biol 1992;43:27–35.<br />

20 Foekens JA, Portengen H, Van Putten<br />

WL, Trapman AM, Reubi JC,<br />

Alexieva-Figush J, Klijn JG: Prognostic<br />

value of receptors <strong>for</strong> the insulin<br />

like growth factor 1, somatostatin<br />

<strong>and</strong> epidermal growth factor<br />

in human breast cancer. <strong>Cancer</strong> Res<br />

1989;49:7002–7009.<br />

21 Zhang CY, Yokogoshi Y, Yoshimoto<br />

K, Fujinaka Y, Matsumoto T,<br />

Saito S: Point mutation of the somatostatin<br />

receptor 2 gene in human<br />

small cell lung cancer cell line<br />

COR-L103. Biochem Biophys Res<br />

Commun 1995;211:805–815.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 191


22 Greenman Y, Melmed S: Heterogenous<br />

expression of two somatostatin<br />

receptor subtypes in pituitary tumours.<br />

J Clin Endocrinol Metab<br />

1994;78:398–403.<br />

23 Greenman Y, Melmed S: Expression<br />

of three somatostatin receptor<br />

analogues on pituitary adenomas.<br />

Evidence <strong>for</strong> preferential sstr5 expression<br />

in the mammosomatograph<br />

lineage. J Clin Endocrinol<br />

Metab 1994;79:398–403.<br />

24 Kubota F, Yamada Y, Kagimoto S,<br />

Shimatsu IM, Tsuda K, Imura H,<br />

Seino S, Seino Y: Identification of<br />

somatostatin receptors subtypes <strong>and</strong><br />

an implication of the efficacy of somatostatin<br />

analogue SMS 201–995<br />

in treatment of human endocrine tumours.<br />

J Clin Invest 1994;93:1321–<br />

1325.<br />

25 Panetta R, Patel YC: Expression of<br />

mRNA <strong>for</strong> all five human somatostatin<br />

receptors (hsstr1–5) in pituitary<br />

tumours. Life Sci 1995;56:<br />

333–342.<br />

26 Miller GM, Alex<strong>and</strong>er JM, Bikhal<br />

HA, Katznelson LN, Zervas A: <strong>Somatostatin</strong><br />

receptor subtype gene<br />

expression in pituitary adenomas. J<br />

Clin Endocrinol Metab 1995;80:<br />

1386–1392.<br />

27 Vikic-Topic S, Raisch KP, Kvols<br />

LK, Vuk-Pavlovic S: Expression of<br />

somatostatin receptor subtypes in<br />

breast carcinoma, carcinoid tumours<br />

<strong>and</strong> renal cell carcinoma. J<br />

Clin Endocrinol Metab 1995;80:<br />

2974–2979.<br />

28 Buscail L, Delesque N, Esteve J-P,<br />

Saint Laurent N, Prats H: Stimulation<br />

of tyrosine phosphatase <strong>and</strong> inhibition<br />

of cell proliferation by somatostatin<br />

analogues by human receptors<br />

subtypes sstr1 <strong>and</strong> sstr2.<br />

Proc Natl Acad Sci USA 1994;19:<br />

2315–2319.<br />

29 Buscail L, Esteve J-P, St Laurent N,<br />

Bertr<strong>and</strong> B, Reisine T, O’Carroll<br />

AM, Bell GI, et al: Inhibition of cell<br />

proliferation by the somatostatin<br />

analogue RC-160 mediated by somatostatin<br />

receptor subtypes SSTR<br />

<strong>and</strong> SSTR5 through different mechanisms.<br />

Proc Natl Acad Sci USA<br />

1995;92:1580–1584.<br />

30 Srikant CB, Shew SH: Octapeptide<br />

somatostatin analogue SMS 201–<br />

995 induces translocation of intracellular<br />

PTPIC to membranes in<br />

MCF-7 human breast adenocarcinoma<br />

cells. Endocrinology 1996;137:<br />

3461–3468.<br />

31 Ain KB, Taylor KD, Tofiq S, Venkataraman<br />

G: <strong>Somatostatin</strong> receptor<br />

subtype expression in human<br />

thyroid <strong>and</strong> thyroid carcinoma cells.<br />

J Clin Endocrinol 1997;82:1857–<br />

1862.<br />

32 Shimon I, Yan X, Taylor JE, Weiss<br />

MH, Culler MD, Melmed S: <strong>Somatostatin</strong><br />

receptor (sstr) subtypeselective<br />

analogues differentially<br />

suppress in vitro growth hormone<br />

<strong>and</strong> prolactin in human pituitary adenomas.<br />

Novel potential therapy <strong>for</strong><br />

functional pituitary tumours. J Clin<br />

Invest 1997;100:2386–2392.<br />

33 Reubi JC, Horiberger U, Studer UE,<br />

Waser B, Laissue JA: Human kidney<br />

as a target <strong>for</strong> somatostatin:<br />

High affinity receptors in tubules<br />

<strong>and</strong> vasa recta. J Clin Endocrinol<br />

Metab. 1993;77:1323–1328.<br />

34 Sinisi AA, Bellastella A, Prezioso D,<br />

Nicchio MR, Lotti T, Salvatore M,<br />

Pasquali D: Different expression<br />

patterns of somatostatin receptor<br />

subtypes in cultured cells from human<br />

normal <strong>and</strong> prostate cancer. J<br />

Clin Endocrinol Metab 1987;82:<br />

2566–2569.<br />

35 Hoffl<strong>and</strong> LJ, De Herder WW, Visser-Wisselaar<br />

HA, Van Ufflen C,<br />

Waaijers M, Zuyderwijk J, et al:<br />

Dissociation between the effects of<br />

somatostatin (SS) <strong>and</strong> octapeptide<br />

SS-analogues on hormone release in<br />

a small group of pituitary <strong>and</strong> islet<br />

cell tumours. J Clin Endocrinol<br />

1997;82:3011–3018.<br />

36 Mato E, Matias-Guiu X, Chico A,<br />

Webb SM, Cabezas R, Berna L, Deleiva<br />

A: <strong>Somatostatin</strong> <strong>and</strong> somatostatin<br />

receptor subtype expression in<br />

medullary thyroid carcinoma. J Clin<br />

Endocrinol Metab 1998;83:2417–<br />

2420.<br />

37 Laws SAM, Gough AC, Primrose<br />

JN: Expression of somatostatin receptor<br />

subtype messenger RNA in<br />

human colonic cancers <strong>and</strong> normal<br />

mucosa (abstract). Br J Surg 1985;<br />

82:1544.<br />

38 Reubi JC, Horisberger U, Laissue J:<br />

High density of somatostatin receptors<br />

in veins surrounding human<br />

cancer tissue: Role in tumour host<br />

interaction. Int J <strong>Cancer</strong> 1994;56:<br />

681–688.<br />

39 Reubi JC, Mazzuchelli L, Hennig I,<br />

Laissue JA: Local up-regulation of<br />

neuropeptide receptors in host<br />

blood vessels around human colorectal<br />

cancers. Gastroenterology<br />

1996;110:1719–1726.<br />

40 Wang H, Li M, Dichter M, Reisine<br />

T: Lack of cross-desensitisation of<br />

somatostatin-14 <strong>and</strong> somatostatin-<br />

28 receptors coupled to potassium<br />

channels in rat neocortical neurones.<br />

Mol Pharmacol 1990;38:<br />

357–361.<br />

41 Londong W, Angerer MA, Kutz K,<br />

L<strong>and</strong>graf R, Londong V: Diminishing<br />

efficacy of octreotide (SMS<br />

201–995) on gastric functions of<br />

healthy subjects during one week administration.<br />

Gastroenterology<br />

1989;96:713–722.<br />

42 Friess H, Bordihn K, Ebert M, Malfertheiner<br />

P, Kenner T, Dennier HJ,<br />

Bulcher MW: Inhibition of pancreatic<br />

secretion under long-term<br />

octreotide treatment in humans. Digestion<br />

1994;55(suppl 1):10–15.<br />

43 Srikant B, Heisler S: Relationship<br />

between receptor binding <strong>and</strong> biopotency<br />

of somatostatin-14 <strong>and</strong> somatostatin-28<br />

in mouse pituitary tumour<br />

cells. Endocrinology 1985;<br />

117:271–285.<br />

44 Presky DH, Schonbrunn A: <strong>Somatostatin</strong><br />

pre-treatment increases the<br />

number of somatostatin receptors<br />

on GH4C1 pituitary cells <strong>and</strong> does<br />

not reduce cellular responsiveness to<br />

somatostatin. J Biol Chem 1988;<br />

263:714–721.<br />

45 Pinsky J, Halmos G, Yano T, Szeperhazi<br />

K, Quin Y, Ertl T, Scally<br />

AV: Inhibition of growth of MKN45<br />

human gastric-carcinoma xenografts<br />

in nude mice by treatment<br />

with bombesin/gastrin-releasingpeptide<br />

antagonist (RC-3095) <strong>and</strong><br />

somatostatin analogue RC-160. Int<br />

J <strong>Cancer</strong> 1994;57:574–580.<br />

46 Bruno JF, Xu Y, Berelowitz M: <strong>Somatostatin</strong><br />

regulates somatostatin<br />

receptor subtype mRNA expression<br />

in GH3 cells. Biochem Biophys Res<br />

Commun 1995;202:1738–1743.<br />

192 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


47 Hofl<strong>and</strong> LJ, Van Koetsveld PM,<br />

Waaijers M, Zuyderwijk Breeman<br />

WAP, Lamberts SWJ: Internalisation<br />

of the radioiodinated somatostatin<br />

analogue [125I-Tyr3]OC-<br />

TREOTIDE by mouse <strong>and</strong> human<br />

pituitary tumour cells. Increase by<br />

unlabelled octreotide. Endocrinology<br />

1995;136:3698–3706.<br />

48 Parmer H, Phillips RH, Lightman<br />

SL: <strong>Somatostatin</strong> analogues: Mechanism<br />

of action. Recent Results <strong>Cancer</strong><br />

Res 1993;912:1–24.<br />

49 Reisine T, Bell CI: Molecular properties<br />

of somatostatin receptors.<br />

Neuroscience 1995;674:777–790.<br />

50 Patel YC, Greenwood M, Panetta R,<br />

Hukovic N, Grigorakis S, Robertson<br />

LA, Srikant CB: Molecular biology<br />

of somatostatin receptor subtypes.<br />

Metabolism 1996;45:31–38.<br />

51 Patel YC: Molecular pharmacology<br />

of somatostatin receptors. J Endocrinol<br />

Invest 1997;20:348–367.<br />

52 Meyerhof W: The elucidation of somatostatin<br />

receptor functions; a current<br />

view. Rev Physiol Biochem<br />

Pharmacol 1998;133:55–108.<br />

53 Schonbrunn A: Glucocorticoids<br />

down-regulate somatostatin receptors<br />

on pituitary cells in culture. Endocrinology<br />

1982;110:1147–1154.<br />

54 Hinkle PM, Peronne MH, Schonbrunn<br />

A: Mechanism of thyroid hormone<br />

stimulation of thyrotropin-releasing<br />

hormone action. Endocrinology<br />

1981;108:199–205.<br />

55 Krimura N, Hayafuji C, Kimura N:<br />

Characterisation of 17-beta-estradiol-dependant<br />

<strong>and</strong> -independent<br />

somatostatin receptor subtypes in<br />

rat anterior pituitary. J Biol Chem<br />

1989;264:7033–7044.<br />

56 Visser-Wisselaar HA, Hofl<strong>and</strong> LJ,<br />

Van Uffelen CJ, Van Koetsveld PM,<br />

Lambert SW: <strong>Somatostatin</strong> receptor<br />

manipulation. Digestion 1996;57<br />

(suppl 1):7–10.<br />

57 Chou CK, Ho LT, Ting LP, Hu CP,<br />

Su TS, Chang WC, et al: Selective<br />

suppression of insulin-induced proliferation<br />

of cultured human hepatoma<br />

cells by somatostatin. J Clin Invest<br />

1987;99:175–178.<br />

58 Septono-Han B, Henkleman MS,<br />

Foekens JA, Klijin JGM: Direct inhibitory<br />

effects of somatostatin analogues<br />

on the growth of human<br />

breast cancer cells. <strong>Cancer</strong> Res<br />

1987;47:1566–1570.<br />

59 Ruiz-Torres P, Lucio FJ, Gonzalez-<br />

Rubio M, Rodriguez-Puyol M, Rodriguez-Puyol<br />

D: A dual effect of<br />

somatostatin on the proliferation of<br />

cultured rat mesangial cells. Biochem<br />

Biophys Res Commun 1993;<br />

195:1057–1062.<br />

60 Ishizuka J, Beauchamp RD, Evers<br />

BM, Townsend CM, Thompson JC:<br />

Unexpected growth-stimulatory effect<br />

of somatostatin analogue on cultured<br />

human pancreatic carcinoid<br />

cells. Biochem Biophys Res Commun<br />

1992;185:577–581.<br />

61 Koper JW, Markstein R, Kohler C,<br />

Kwekkeboom DJ, Avezaat CJJ,<br />

Lamberts SWJ, Reubi JC: <strong>Somatostatin</strong><br />

inhibits the activity of adenylate<br />

cyclase in cultured human meningioma<br />

cells <strong>and</strong> stimulates their<br />

growth. J Clin Endocrinol Metab<br />

1992;74:543–547.<br />

62 Ain KB, Taylor KD: <strong>Somatostatin</strong><br />

analogues affect proliferation of human<br />

thyroid carcinoma cell lines<br />

in vitro. J Clin Endocrinol Metab<br />

1994;78:1097–1102.<br />

63 Koelz A, Kraenzlin M, Gyr M, Meir<br />

V, Bloom S, Heitz P, Stalder H:<br />

Escape of the response to a longacting<br />

somatostatin analogue (SMS<br />

201–995) in patients with VIPoma.<br />

Gastroenterology 1987;82:527–531.<br />

64 Goldenberg SL, Brochovsky N,<br />

Gleave M, Sullivan LD, Akakura K:<br />

Intermittent <strong>and</strong>rogen suppression<br />

in the treatment of prostate cancer:<br />

A preliminary report. Urology 1995;<br />

45:839–845.<br />

65 Oliver RTD, Williams G, Paris<br />

AMI, Bl<strong>and</strong>y JP: Intermittent <strong>and</strong>rogen<br />

deprivation after PSA –<br />

Complete response as a strategy to<br />

reduce induction of hormone-resistant<br />

prostate cancer. Urology 1997;<br />

49:79–82.<br />

66 Alex<strong>and</strong>er HR, Fraker DL, Norton<br />

JA, Bartlett DL, Tio L, Benjamin<br />

SB, Doppman JL, Goebel F, Serrano<br />

J, Gibril F, Jensen RT: Prospective<br />

study of somatostatin receptor<br />

scintigraphy <strong>and</strong> its operative outcome<br />

in patients with Zollinger-Ellison<br />

syndrome. Ann Surg 1998;228:<br />

121–129.<br />

67 Orle<strong>for</strong>s H, Sundrin A, Helstrom H,<br />

Bjurling P, Bergstrom M, Lilja A:<br />

Positon emission tomography with<br />

5 hydroxytryptophan in neuroendocrine<br />

tumours. J Clin Oncol 1998;7:<br />

2534–2541.<br />

68 Franker DL, Norton JA, Alex<strong>and</strong>er<br />

HR, Venzon DJ, Jensen RT: Surgery<br />

in Zollinger-Ellison syndrome<br />

alters the natural history of gastrinoma.<br />

Ann Surg 1994;220:320–<br />

330.<br />

69 Routley D, Ramage J, McPearke J:<br />

Orthotopic liver transplantation in<br />

the treatment of metastatic neuroendocrine<br />

tumours of the liver. Liver<br />

Transplant Surg 1995;1:118–121.<br />

70 Dousset B, Hussin D, Soubrane O:<br />

Metastatic endocrine tumours: Is<br />

there a place <strong>for</strong> liver transplantation?<br />

Liver Transplant Surg 1995;1:<br />

111–117.<br />

71 Lang H, Oldhafer KJ, Weimann A:<br />

Liver transplantation <strong>for</strong> metastatic<br />

neuroendocrine tumours. Ann Surg<br />

1997;225:347–354.<br />

72 Le-Treut YP, Delpero JR, Dousset<br />

A: Results of liver transplantation in<br />

the treatment of metastatic neuroendocrine<br />

tumours. A 31 case French<br />

multicentre study. Ann Surg 1997;<br />

225:355–364.<br />

73 Lee JM, Erlich R, Bruchner HW,<br />

Roboz J, Beasley J, Ohnuma T: Octreotide<br />

acetate (S<strong>and</strong>ostatin) increases<br />

in vivo accumulation of doxirubicin.<br />

Proc Am Assoc <strong>Cancer</strong> Res<br />

1993;34:285.<br />

74 Stewart G, Romani R, Lawson J,<br />

Morris DL: SMS 201–995 inhibits<br />

the growth of liver metastases from<br />

colorectal cancer <strong>and</strong> increase in vitro<br />

toxicity of fluorouracil (abstract).<br />

Proc Am Assoc <strong>Cancer</strong> Res 1993;34:<br />

296.<br />

75 Weckbecker G, Raulf F, Tolcsva I,<br />

Bruno L: Potentiation of the antiproliferative<br />

effects of anti cancer<br />

drugs by octreotide in vivo <strong>and</strong> in<br />

vitro. Digestion 1996;57(suppl 1):<br />

27–30.<br />

76 Begg C, Cho M, Eastwood S, Horton<br />

R, Moher D, Oklin I: Improving the<br />

quality of reporting of r<strong>and</strong>omised<br />

controlled trials: The consort statement.<br />

JAMA 1996;76:637–639.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 193


77 Kouroumalis E, Skordilis P, Thermas<br />

K, Vasilaki A, Mosch<strong>and</strong>rua J,<br />

Manousatous ON: <strong>Treatment</strong> of<br />

hepatocellular carcinoma with octreotide:<br />

A r<strong>and</strong>omised controlled<br />

study. Gut 1998;42:442–447.<br />

78 Fossati R, Confalonieri C, Torri V,<br />

Ghisl<strong>and</strong>i E, Penna A, Pistotti V,<br />

Tinazzi A, Liberati A: Cytotoxic <strong>and</strong><br />

hormonal treatment <strong>for</strong> metastatic<br />

breast cancer: A systemic review of<br />

published r<strong>and</strong>omised trials involving<br />

31,500 women. J Clin Oncol<br />

1998;16:3439–3460.<br />

79 Radulovic S, Nagy A, Szoke B,<br />

Schally A: Cytotoxic analogue of somatostatin<br />

containing methotrexate<br />

inhibits growth of MIA PaCA-2 human<br />

pancreatic xenografts in nude<br />

mice. <strong>Cancer</strong> Lett 1992;62:263–<br />

271.<br />

80 Nagy A, Schally AV, Halmos G, Armatis<br />

P, Cai RZ, Csernus V: Synthesis<br />

<strong>and</strong> biological evaluation of cytotoxic<br />

analogues of somatostatin containing<br />

doxorubicin <strong>and</strong> its potent<br />

derivative 2-pyrrolinodoxorubicin.<br />

Br J Surg 1986;83:456–460.<br />

81 Taylor I: Liver metastases from<br />

colorectal cancer: Lessons from past<br />

<strong>and</strong> present clinical studies. Br J<br />

Surg 1996;83:456–460.<br />

82 Rosenberg L: Pancreatic cancer:<br />

Does octreotide offer any promise?<br />

Chemotherapy 2001;47(suppl 2):<br />

134–149.<br />

83 Cascinu S, Catalo V, Giordani P,<br />

Baldelli AM, Agostinelli R, Catalanoo<br />

G: GI cancer refractory to chemotherapy:<br />

A room <strong>for</strong> octreotide.<br />

Chemotherapy 2001;47(suppl 2):<br />

127–133.<br />

84 Hughes KS, Simon RM, Songhorabod<br />

IS, Adson MA, Ilstrup DM: Resection<br />

of the liver <strong>for</strong> colorectal metastases:<br />

A multi-institutional study<br />

of patterns of recurrence. Surgery<br />

1986;100:279–284.<br />

85 Scheele J, Stangl R, Altendorf-Hofmann<br />

A, Paul M: Resection of colorectal<br />

liver metastases. World J Surg<br />

1995;19:59–71.<br />

86 Scheele J, Stangl R, Altendorf-Hofmann<br />

A, Gall FP: Indicators of prognosis<br />

after hepatic resection <strong>for</strong> colorectal<br />

secondaries. Surgery 1991;<br />

110:13–29.<br />

87 Geoghegan JC, Scheele J: <strong>Treatment</strong><br />

of colorectal liver metastases. Br J<br />

Surg 1999;86:158–169.<br />

88 Asaga T, Suzuki K, Umeda M, Sugimasa<br />

Y, Takemiya S, Okamoto T:<br />

The enhancement of tumour growth<br />

after partial hepatectomy <strong>and</strong> the effect<br />

of sera obtained from hepatectomised<br />

rats on tumour cell growth.<br />

Jpn J Surg 1991;21:669–675.<br />

89 Loizidou MC, Lawrence RJ, Holt S:<br />

Facilitation by partial hepatectomy<br />

of tumour growth within the rat liver<br />

following intraportal injection of syngeneic<br />

tumour cells. Clin Exp Metastasis<br />

1991;9:335–349.<br />

90 Davies N, Yates J, Kynaston H,<br />

Taylor BA, Jenkins SA: Effects of<br />

octreotide on liver regeneration <strong>and</strong><br />

tumour growth in the regenerating<br />

liver. J Gastroenterol Hepatol 1997;<br />

12:47–53.<br />

91 Vainas IG: Octreotide in the management<br />

of hormone-refractory<br />

prostate cancer. Chemotherapy<br />

2001;47(suppl 2):109–126.<br />

92 Ingle JN, Suman VJ, Kardinal CG:<br />

A r<strong>and</strong>omised trial of tamoxifen<br />

alone or combined with octreotide<br />

in the treatment of metastastic<br />

breast carcinoma. <strong>Cancer</strong> 1999;85:<br />

1284–1292.<br />

93 Ottestad M, Boni C, Bryce C, Klijo<br />

J, Kiese B, Mietlouski U, et al: A<br />

phase 3 trial of octreotide pamoate<br />

(op lar) <strong>and</strong> tamoxifen versus placebo<br />

<strong>and</strong> tamoxifen in metastatic<br />

breast cancer (abstract). Proc Assoc<br />

Clin Oncol 1999;18:110.<br />

94 Ingle JN, Kardnal CG, Suman VJ,<br />

Krook JE, Hatfield AK: Octreotide<br />

as a first-line treatment <strong>for</strong> women<br />

with metastatic breast cancer. Invest<br />

New Drugs 1996;14:235–237.<br />

95 Vennin P, Peyratt JP, Bonneterre J,<br />

Louchez MM, Harris AG, Demaille<br />

A: The effect of a long-acting somatostatin<br />

analogue SMS 201–995<br />

(S<strong>and</strong>ostatin) in advanced breast<br />

cancer. Anticancer Res 1989;9:153–<br />

156.<br />

96 Xu U, Song J, Berelowitz M, Bruno<br />

JF: Estrogen regulates somatostatin<br />

receptor subtype 2 messenger ribonucleic<br />

acid expression in human<br />

breast cancer cells. Endocrinology<br />

1996;137:5634–5640.<br />

97 Setyono-Han B, Heinkelman MS,<br />

Foekens JA, Klijn GM: Direct inhibitory<br />

effects of somatostatin<br />

(analogues) on the growth of human<br />

breast cancer cells. <strong>Cancer</strong><br />

Res 1987;47:1556–1570.<br />

98 Huynh HT, Pollait M: Enhancement<br />

of tamoxifen-induced suppression<br />

of insulin-like growth factor,<br />

gene expression <strong>and</strong> serum level<br />

by a somatostatin analogue. Biochem<br />

Biophys Res Commun 1984;<br />

203:253–259.<br />

99 Pollak M, Constantino J, Polychronakos<br />

C, Blauer SA, Guyda V,<br />

Redmond C, et al: Effect of tamoxifen<br />

on serum insulin-like growth<br />

factor 1 levels in stage 1 breast cancer<br />

patients. J Natl <strong>Cancer</strong> Inst<br />

1990;82:1643–1697.<br />

100 Weckbecker G, Tolcsvai L, Stolz<br />

B, Pollak M, Bruns C: <strong>Somatostatin</strong><br />

analogue octreotide enhances<br />

the antineoplastic effects of<br />

tamoxifen <strong>and</strong> ovariectomy on<br />

7, 12 - dimethylbenz(·)anthraceneinduced<br />

rat mammary carcinomas.<br />

<strong>Cancer</strong> Res 1994;54:6134–<br />

6137.<br />

101 Reubi JC, Torhorst J: The relationship<br />

between somatostatin, epidermal<br />

growth factor <strong>and</strong> steroid hormone<br />

receptors in breast cancer.<br />

<strong>Cancer</strong> 1989;64:1254–1260.<br />

102 Bontebal M, Foekens JA, Lamberts<br />

SWJ, Jong FH, Van Putten<br />

WLJ, Braun HJ, et al: Feasibility,<br />

endocrine <strong>and</strong> antitumour effects<br />

of a triple therapy with tamoxifen,<br />

a somatostatin analogue <strong>and</strong> an<br />

antiprolactin in postmenopausal<br />

metastatic breast cancer: A r<strong>and</strong>omised<br />

study with long-term follow-up.<br />

Br J <strong>Cancer</strong> 1998;77:115–<br />

122.<br />

103 Early Breast <strong>Cancer</strong> Trialists Collaborative<br />

Group: Tamoxifen <strong>for</strong><br />

early breast cancer: An overview of<br />

the r<strong>and</strong>omised trials. Lancet<br />

1998;351:1451–1467.<br />

104 Vainas C, Pasaitou G, Galakitdou<br />

K, Maris K, Christodoulou C,<br />

Constantidins C, Korsakis AH:<br />

The role of somatostatin analogues<br />

in complete anti<strong>and</strong>rogen treatment<br />

in patients with prostate cancer.<br />

J Exp Clin Res 1997;16:119–<br />

126.<br />

194 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


105 Rosenberg L, Barkun AN, Denis<br />

MH, Pollak M: Low dose of octreotide<br />

<strong>and</strong> tamoxifen in the<br />

treatment of adenocarcinoma of<br />

the pancreas. <strong>Cancer</strong> 1985;735:<br />

23–28.<br />

106 Fazenzy B, Baur M, Prohaska M,<br />

Hudec M, Kremnitzer M, Meryn<br />

S, et al: Octreotide combined with<br />

gorserelin in the therapy of advanced<br />

pancreatic cancer – Results<br />

of pilot study <strong>and</strong> review of the literature.<br />

J <strong>Cancer</strong> Res Clin Oncol<br />

1997;123:45–52.<br />

107 Chouabib S, Asselin-Paturel C,<br />

Mami-Chouabib F, Caignard A,<br />

Blay JY: The host-tumour immune<br />

conflict from immunosuppression<br />

to resistance <strong>and</strong> destruction. Immunol<br />

Today 1997;18:493–497.<br />

108 Park S, Brody JL, Wallace HC,<br />

Blakemore WS: Immunosuppressive<br />

effect of surgery. Lancet 1971;<br />

i:53–55.<br />

109 Roth JA, Golub SH, Grimm EA,<br />

Eiber FR, Morton DL: Effects of<br />

operation on immune response in<br />

cancer patients: Sequential evaluation<br />

of in vitro lymphocyte function.<br />

Surgery 1976;79:46–51.<br />

110 Tarpley JL, Twomey PL, Catalona<br />

WJ, Chreiten PB: Suppression of<br />

cellular immunity by anaesthesia,<br />

by surgery <strong>and</strong> operation. J Surg<br />

Res 1977;22:195–201.<br />

111 Guillou PJ, Hegarty J, Ramsden<br />

C: Changes in human natural killer<br />

cell activity early <strong>and</strong> late after renal<br />

transplantation using conventional<br />

immunosuppression. Transplantation<br />

1982;33:414–420.<br />

112 Lennard TWJ, Shenton BK, Borzotta<br />

A: The influence of surgical<br />

operations on components of the<br />

human immune system. Br J Surg<br />

1985;129:771–776.<br />

113 Redmond HP, Watson RW,<br />

Houghton T, Condron C, Watson<br />

RGK, Bouchier-Hayes D: Immune<br />

function in patients undergoing<br />

open versus laparoscopic<br />

cholecystectomy. Arch Surg 1994;<br />

129:1240–1246.<br />

114 Klava A, Windsor A, Boylston<br />

AW, Reynolds JV, Ramsden CW,<br />

Guillou PJ: Monocyte activation<br />

after open <strong>and</strong> laparoscopic surgery.<br />

Br J Surg 1997;84:1152–<br />

1156.<br />

115 Takeda K, Fuijii N, Nitta Y, Sakihara<br />

H, Nakatama K, Rishiki H, et<br />

al: Murine tumour cells metastasising<br />

selectively in the liver: Ability<br />

to produce hepatocyte-activating<br />

cytokines interleukin-1 <strong>and</strong> -6.<br />

Jpn J <strong>Cancer</strong> Res 1991;82:1299–<br />

1308.<br />

116 Terry WD, Roseberg SA: Immunotherapy<br />

of Human <strong>Cancer</strong>. New<br />

York, Elsevier, 1982.<br />

117 Rosenberg SA: The development<br />

of new immunotherapies <strong>for</strong> the<br />

treatment of cancer using interleukin-2.<br />

Ann Surg 1988;208:121–<br />

135.<br />

118 Muisani P, Modesti A, Giovarelli<br />

M, Cavallo F, Colombo MP: Cytokine<br />

tumour cell death <strong>and</strong> immunogenicity:<br />

A question of choice.<br />

Immunol Today 1997;13:32–36.<br />

119 Kobari M, Egawa S, Shibuya T,<br />

Sunamura K, Saitoh M, Matsuno<br />

P: Effect of intraportal adoptive<br />

immunotherapy on liver metastases<br />

after resection of pancreatic<br />

cancer. Br J Surg 2000;87:43–47.<br />

120 Lisson P: Immunoneuroendocrine<br />

therapy <strong>for</strong> endocrine tumours:<br />

Recent developments. Exp Opin<br />

Invest Drugs 1995;4:1267–1272.<br />

121 Janson E, Allstrom M, Andersson<br />

T: Octreotide <strong>and</strong> interferon-alpha:<br />

A new combination <strong>for</strong> the<br />

treatment of malignant carcinoid<br />

tumours. Eur J <strong>Cancer</strong> 1992;28:<br />

1647–1650.<br />

122 Joensuu H, Katra K, Kujari H:<br />

Dramatic response of metastatic<br />

carcinoid to a combination of interferon<br />

<strong>and</strong> octreotide. Acta Endocrinol<br />

1992;126:184–185.<br />

123 Van Hagen PM, Krenning EP,<br />

Kwekkerboom DJ, Reubi JC,<br />

Van de Anker-Lugtenburgp PJ: <strong>Somatostatin</strong><br />

<strong>and</strong> the immune <strong>and</strong><br />

haematopoietic system: A review.<br />

Eur J Clin Invest 1994;24:91–99.<br />

124 Balibrea JL, Ariaz-Diaz J, Gruz G,<br />

Vara E: Effect of pentoxifylline<br />

<strong>and</strong> somatostatin on tumour necrosis<br />

factor production by human<br />

pulmonary macrophages. Circ<br />

Shock 1994;43:51–56.<br />

125 Muscettola M, Grasso G: <strong>Somatostatin</strong><br />

<strong>and</strong> VIP reduce interferon<br />

production in human peripheral<br />

mononuclear cells. Immunobiology<br />

1990;80:419–430.<br />

126 Sirriani MC, Annigale B, Fais S,<br />

Delle-Faue G: Inhibitory effects of<br />

somatostatin <strong>and</strong> some analogues<br />

on human natural killer cell activity.<br />

Peptides 1994;15:1033–1036.<br />

127 Wiseman GA, Kvols K: Therapy<br />

of neurometastatic tumours with<br />

radiolabelled MIBG <strong>and</strong> somatostatin<br />

analogues. Semin Nucl Med<br />

1995;25:272–278.<br />

128 Zamora PO, Bender H, Gulhke S,<br />

Marek MJ, Knapp FF, Rhodes BA,<br />

Biersack HJ: Pre-clinical experience<br />

with Rb-188-RC-160, a radiolabelled<br />

somatostatin analogue<br />

<strong>for</strong> use in peptide-targeted radiotherapy.<br />

Anticancer Res 1997;17:<br />

1803–1808.<br />

129 Jain RK: The next frontier of molecular<br />

medicine: Delivery of therapeutics.<br />

Nat Med 1998;4:656–<br />

657.<br />

130 Danesi R, Del Tacca M: Effects of<br />

octreotide on angiogenesis; in<br />

Scarpignato C (ed): Octreotide:<br />

From Basic Science to Clinical<br />

Medicine. Prog Basic Clin Pharmacol.<br />

Basel, Karger, 1996, vol 10,<br />

pp 234–245.<br />

131 Woltering EA, Watson JC, Alperin-Lea<br />

C, Sharma C, Keenan E,<br />

Kurdzawa DL, Barrie R: <strong>Somatostatin</strong><br />

analogues: Angiogenesis inhibitors<br />

with novel mechanisms of<br />

action. Invest New Drugs 1997;15:<br />

77–86.<br />

132 O’Reilly M, Boehm T, Shing Y,<br />

Fukai N, Vasios G, Lane W, et al:<br />

Endostatin: An endogenous inhibitor<br />

of angiogenesis <strong>and</strong> tumour<br />

growth. Cell 1997;88:277–285.<br />

133 O’Reilly M, Holmgren L, Chen C,<br />

Folkman J: Angiostatin induces<br />

<strong>and</strong> sustains dormancy of human<br />

primary tumours in mice. Nat<br />

Med 1996;2:689–692.<br />

134 Griscelli LH, Bennaceur-Griscelli<br />

A, Soria J, Opolon P, Soria C: Angiostatin<br />

gene transfer: Inhibition<br />

of tumour growth in vivo by blockage<br />

of endothelial cell proliferation<br />

associated with a mitosis arrest.<br />

Proc Natl Acad Sci USA 1998;95:<br />

6367–6372.<br />

<strong>Somatostatin</strong> <strong>Analogs</strong> in Oncology Chemotherapy 2001;47(suppl 2):162–196 195


135 Lin P, Buxton JA, Acheson A,<br />

Radziejewski C, Maisonperre PC,<br />

Yancopoulos CD, et al: Antiangiogenic<br />

gene therapy targeting the<br />

endothelium-specific receptor tyrosine<br />

kinase Tie 2. Proc Natl<br />

Acad Sci USA 1998;95:8829–<br />

8834.<br />

136 Roth JA, Cristiano J: Gene therapy<br />

<strong>for</strong> cancer: What have we done<br />

<strong>and</strong> where are we going? J Natl<br />

Invest 1999;89:21–36.<br />

137 Ellouk-Achard S, Djenabi S, De<br />

Oliveira GA, Desauty G, Duc HT,<br />

Zohair M, et al: Induction of apoptosis<br />

in rat hepatocarcinoma cells<br />

by expression of 1GF-1 antisense<br />

c-DNA. J Hepatol 1998;29:807–<br />

818.<br />

138 Hollstein M, Rice K, Greenblatt<br />

MS, Soussi T, Fuchs R, Sorlie T, et<br />

al: Database of p53 gene somatic<br />

mutations in human tumours <strong>and</strong><br />

cell lines. Nucleic Acids Res 1994;<br />

2:3351.<br />

139 Fujiwara T, Cai DW, Georges RN,<br />

Mukhopadhyay T, Grimm EA,<br />

Roth JA: Therapeutic effect of a<br />

wild type p53-expression vector in<br />

an orthotopic lung cancer model. J<br />

Natl <strong>Cancer</strong> Inst 1994;86:1458–<br />

1462.<br />

140 Spitz FR, Nguyen D, Skibber JM,<br />

Cusack J, Roth JA, Cristiano RJ:<br />

In vivo adenovirus mediated p53<br />

tumour supressor gene therapy <strong>for</strong><br />

colorectal cancer. Anticancer Res<br />

1996;16:3415–3422.<br />

141 Ogawa N, Fujiwara T, Kagawa S,<br />

Nishizaki M, Morimoto Y, Tanida<br />

T, et al: Novel combination therapy<br />

<strong>for</strong> human colon cancer with<br />

adenovirus-mediated wild type<br />

p53 gene transfer <strong>and</strong> DNA-damaging<br />

chemotherapeutic agents. Int<br />

J <strong>Cancer</strong> 1997;73:367–370.<br />

142 Onliner JD, Kinzler KW, Meltzer<br />

PS, George DL, Vogelstein B: Amplification<br />

of a gene encoding a p53associated<br />

protein in human sarcomas.<br />

Nature 1992;358:80–83.<br />

143 Sharma K, Patel Y, Srikant C: Subtype<br />

specific induction of wild type<br />

p53 <strong>and</strong> apoptosis, but not cell cycle<br />

arrest by human somatostatin<br />

receptor 3. Mol Endocrinol 1966;<br />

10:1688–1696.<br />

196 Chemotherapy 2001;47(suppl 2):162–196 Jenkins/Kynaston/Davies/Baxter/Nott


ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Author Index Vol. 47, Suppl. 2, 2001<br />

Agostinelli, R. 127<br />

Amoroso, D. 62<br />

Baldelli, A.M. 127<br />

Baxter, J.N. 162<br />

Boccardo, F. 62<br />

Bousquet, C. 30<br />

Buscail, L. 30<br />

Carney, D.N. 78<br />

Cascinu, S. 127<br />

Catalano, G. 127<br />

Catalano, V. 127<br />

Davies, N. 162<br />

Dean, A. 54<br />

Giordani, P. 127<br />

Jenkins, S.A. 162<br />

Kouroumalis, E.A. 150<br />

Kynaston, H.G. 162<br />

Nott, D.M. 162<br />

Öberg, K. 40<br />

O’Byrne, K.J. 78<br />

Pelosini, I. 1<br />

Puente, E. 30<br />

Rosenberg, L. 134<br />

Scarpignato, C. 1<br />

Schally, A.V. 78<br />

Steward, W.P. 78<br />

Susini, C. 30<br />

Thomas, A. 78<br />

Vainas, I.G. 109<br />

Vaysse, N. 30<br />

197


ABC<br />

Fax + 41 61 306 12 34<br />

E-Mail karger@karger.ch<br />

www.karger.com<br />

© 2001 S. Karger AG, Basel<br />

Accessible online at:<br />

www.karger.com/journals/che<br />

Subject Index Vol. 47, Suppl. 2, 2001<br />

Anti<strong>and</strong>rogen blockade, complete<br />

109<br />

Breast cancer 62<br />

<strong>Cancer</strong> treatment 1<br />

Chemotherapy 78, 127<br />

Gastrointestinal cancers 127<br />

[ 111 In]pentetreotide 78<br />

Hepatocellular carcinoma 150<br />

Hormonal maneuvers, alternative<br />

109<br />

Lanreotide 1, 40<br />

Liver tumors 150<br />

Lung cancer 78<br />

Neoplasms 162<br />

Neuroendocrine tumors 40<br />

Octastatin 40<br />

Octreoscan 40<br />

Octreotide 1, 40, 109, 127, 134,<br />

150<br />

Opioids 54<br />

Palliative care 54<br />

Pancreatic cancer 134<br />

Proliferation 30<br />

Prostate cancer, hormonerefractory<br />

109<br />

RC-160 134<br />

Receptor 78<br />

– signaling 30<br />

Review 162<br />

<strong>Somatostatin</strong> 1, 54, 62, 78, 134<br />

– analogs 62, 134, 162<br />

– receptors 30, 134, 150<br />

– – scintigraphy 1<br />

– receptor-targeted radiotherapy<br />

1<br />

Therapy 162<br />

Vapreotide 1

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!