05.08.2013 Views

Voiding dysfunction.pdf - E-Lib FK UWKS

Voiding dysfunction.pdf - E-Lib FK UWKS

Voiding dysfunction.pdf - E-Lib FK UWKS

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.

VOIDING DYSFUNCTION


URRENT CLINICAL UROLOGV<br />

Eric A. Klein, SERIES EDITOR<br />

<strong>Voiding</strong> Dysfunction: Diagnosis arzd Treatment, edited by Rodney A.<br />

Appell, 2000<br />

~ana~e~ent of Prostate Cancer, edited by Eric A. Klein, 2000


VOIDING<br />

DYSFUNCTION<br />

DUGNOSIS<br />

Edited by<br />

RODNEY A. APPELL,<br />

Chehnd Clinic Foundntion,<br />

Chehd OH<br />

HUMANA PRESS<br />

TOTOWA, NEW<br />

JERSEY<br />

MD


0 2000 Huniana Press Inc.<br />

999 Riverview Drive, Suite 208<br />

Totowa, New Jersey 075 12<br />

For additional copies, pricing for bulk purchases, and/or information about other Huinana titles,<br />

contact Huinana at the above address or at any of the following numbers: Tel: 973-256-1 699.<br />

Fax. 973-256-834 I : E-mail: humana~liiinianapr.com or visit our Website at http.//humanapress.com<br />

All rights reserved No part of this book may be reproduced. stored in a retrieval system, or transmitted in<br />

any forin or by any means. electronic, mechanical, photocopying. microfilming. recording. or otherwise<br />

without written perinission from the Publisher.<br />

All articles, comments, opinions, conclusions. or recoininendations are those of the author(s), and do not neces-<br />

sarily reflect the views of the publisher.<br />

Due diligence has been taken by the publishers. editors, and authors of this book to ensure the accuracy of<br />

the information published and to describe generally accepted practices. The contributors herein have care-<br />

fully checked to ensure that the drug selections and dosages set forth in this text are accurate in accord with<br />

the standards accepted at the time of publication. Notwithstanding, as new research, changes in government<br />

regulations, and knowledge froin clinical experience relating to drug therapy and drug reactions constantly<br />

occurs, the reader is advised to check the product information provided by the manufacturer of each drug for<br />

any change in dosages or for additional warnings and contraindications. This is of utmost importance when<br />

the recommended drug herein is a new or infrequently used drug. It is the responsibility of the health care<br />

provider to ascertain the Food and Drug Administration status of each drug or device used in their clinical<br />

practice. The publisher, editors, and authors are not responsible for errors or omissions or for any conse-<br />

quences from the application of the information presented in this book and make no warranty, express or<br />

implied. with respect to the contents in this publication.<br />

This publication is printed on acid-free paper.=<br />

ANSI 239.48- 1984 (American National Standards Institute)<br />

Permanence of Paper for Printed <strong>Lib</strong>rary Materials.<br />

Cover design by Patricia F. Cleary.<br />

Photocopy Authorization Policy:<br />

Authorization to photocopy items for internal or personal use, or the internal or personal use of specific<br />

clients, is granted by Huniana Press Inc., provided that the base fee of US $8.00 per copy, plus US $00.25<br />

per page, is paid directly to the Copyright Clearance Center at 233 Rosewood Drive. Danvers. MA 0 1923.<br />

For those organizations that have been granted a photocopy license from the CCC, a separate system of<br />

payment has been arranged and is acceptable to Huniana Press Inc. The fee code for users ofthe Transactional<br />

Reporting Service is [O-89603-659-6100 $8.00 + $00.251.<br />

Printed in the United States of America. 10 9 8 7 6 5 4 3 2 1<br />

<strong>Voiding</strong> <strong>dysfunction</strong> . diagnosis and treatmededited by Rodney A. Appell<br />

p. : cm.--(Current clinical urology)<br />

Includes bibliographical references and index.<br />

ISBN 0-89603-659-6 (alk. paper)<br />

1. Urination disorders. I. Appell, Rodney A. 11. Series.<br />

[DNLM: 1. Urination Disorders4iagnosis. 2. Urination Disorders-therapy. WJ 146<br />

V889 20001<br />

RC901.75.V65 3000<br />

6 16.6'24~2 1<br />

DNLM/DLC 00-027599<br />

for <strong>Lib</strong>rary of Congress CIP


PREFACE<br />

The purpose of <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment is to bring the<br />

reader up to date on all clinical aspects of voiding <strong>dysfunction</strong>, not just<br />

urodynamic and other evaluative techniques, but varying nuances in presentation<br />

of the individual problems that may occur during voiding, sometimes as a<br />

manifestation of an underlying disease or disorder, and at other times because<br />

of individual patient circumstances, including the patient’s behavior or the effects<br />

of treatment for an unrelated disorder. <strong>Voiding</strong> <strong>dysfunction</strong> includes disorders<br />

of urinary storage as well as the emptying of the lower urinary tract. Although<br />

urinary incontinence is a huge problem worldwide as exemplified by the absolute<br />

need of the World Health Organization (WHO) to sponsor the mammoth<br />

undertaking of an international consultation on the topic in mid-1998, other<br />

voiding disorders also adversely affect the quality of the lives of the individuals<br />

afflicted. These problems are often placed on the “back burner” with respect to<br />

basic and clinical research because they do not result in death; however, it is<br />

important to recognize that the quality of one’s life is as important as mere<br />

existence itself, as demonstrated by the extent to which people will go to try<br />

therapies (many unproven) at tremendous expense to make small improvements<br />

in the quality of their respective lives.<br />

Fortunately, knowledge of lower urinary tract function and <strong>dysfunction</strong> has<br />

advanced rapidly over the last half of the 20th century from the early phase of<br />

urodynamics (a term first used in 1953 by D. M. Davis [I]) to the current neurological<br />

era (2).<br />

I want to express my appreciation to an excellent group of contributors to<br />

<strong>Voiding</strong> Dysfunction: Diagnosis and Treatment for their timely response to my<br />

request to put down on paper, in an explicable manner, the current states of the<br />

art and science of voiding <strong>dysfunction</strong>. First, there is a presentation on the<br />

current background information necessary, to understand the physiology of the<br />

normal lower urinary tract and how to use that information to classify the various<br />

voiding <strong>dysfunction</strong>s. This is followed by an overview the of manner in which<br />

voiding <strong>dysfunction</strong> is diagnosed, stressing the importance of individualizing<br />

the evaluation to demonstrate the voiding dysfuction. Following this, there is a<br />

comprehensive discussion of major neurological problems and their adverse effects<br />

on voiding fbnction. This section includes discussions of specific diseases<br />

V


vi Preface<br />

and disorders that make up the so-called “neurogenic bladder,” which is, quite<br />

obviously, not a single entity. Individual discussions include the effects of stroke,<br />

multiple sclerosis, spinal cord problems, diabetes mellitus, and lumbar disc<br />

disorders on lower urinary tract function. The next section involves <strong>dysfunction</strong><br />

in individual patients only, with comprehensive reviews of urinary incontinence<br />

and urinary retention. Attention in the following section revolves around voiding<br />

<strong>dysfunction</strong>s unique to males: bladder outlet obstruction and postprostatectomy<br />

incontinence. The final segment covers general topics involved in newer modali-<br />

ties of treatment for voiding <strong>dysfunction</strong> and includes phannacologic therapy,<br />

electrical stimulation, and surgery for intractable instability-augmentation<br />

cytoplasty .<br />

I am certain readers will find <strong>Voiding</strong> Dysfiuictiorz: Diagnosis and Treatment<br />

informative, practical, and clinically relevant.<br />

References<br />

Rodney A. Appell, MD<br />

1. Davis, D. M. (1953) The mechanisms of urologic disease. WB Saunders,<br />

Philadelphia.<br />

2. Hinman, F., Jr. (1996) Urodynamics I: Foreword. Ui-01. Clin. N. i4mer-. 23, xi-xii.


CONTENTS<br />

Preface.. ....................................................................................... v<br />

List of Contributors ................................................................... .ix<br />

Part I. Introduction<br />

1 Pathophysiology and Classification<br />

of <strong>Voiding</strong> Dysfunction ..................................................... 3<br />

Alan J. Wein and Eric S. Rovner<br />

2 <strong>Voiding</strong> Dysfunction: Diagnostic Evaluation ....................... 25<br />

Victor W. ~ittj and ~ichael Ficazzola<br />

Part II, Neurogenic Vesico-Urethral Dysfunction<br />

3 Cerebrovascular Accidents ................................................ 63<br />

Serge Peter ~arinkovjc and Gopal H. Badlani<br />

4 Multiple Sclerosis ............................................................... 83<br />

Jerry G. Blaivas<br />

5 Diagnosis and Treatment of Spinal Cord injuries<br />

and Myeloneuropathie ................................................. 1 15<br />

Steven W. Sukin and Timothy B. Boone<br />

6 Diabetic Bladder Dysfunction ............................................ 139<br />

Howard B. Goldman and Rodney A. Appell<br />

7 Lumbar Disc Disease ....................................................... 149<br />

Howard B. Goldman and Rodney A. Appell<br />

Part 111. Female <strong>Voiding</strong> Dysfunction<br />

8 Treatment of Stress Urinary Incontinence ........................ 163<br />

Raymond Rackley<br />

9 Urinary Retention in Women ............................................ 185<br />

Roger R. Dmochowski<br />

Part N . Male <strong>Voiding</strong> Dysfunction<br />

10 Bladder Outlet Obstruction in Males ................................. 225<br />

Edward F. Ikeguchi, Alexis E. Te, James Choi,<br />

and Steven A. Kaplan<br />

Vii


...<br />

v111 Contents<br />

11 Post-Prostatectomy Incontinence ..................................... 247<br />

Harriette M. Scarper0 and J. Christian Winters<br />

Part V. Treatment Modalities<br />

12 Clinical Pharmacology ..................................................... 275<br />

Scott R. Serels and Rodney A, Appell<br />

13 Treatment of Detrusor Instability<br />

with Electrical Stimulation ............................................ 297<br />

Steven W. Siege1<br />

14 Treatment of Detrusor instability<br />

with Augmentation Cystoplasty ........................................ 31 5<br />

Joseph M. Khoury<br />

Index.. ................................................................................... ,326


CONTRIBUTORS<br />

RODNEY A. APPELL, MD * Section of <strong>Voiding</strong> Dysfunction and Female Urology,<br />

Department of Urology, Cleveland Clinic Foundation, Cleveland, OH<br />

GOPAL H. BADLANI, MD * Department of Urology, Long Island Jewish Medical<br />

Center, New Hyde Park, NY<br />

JERRY G. BLAwAs, MD * Urologic Oncology, New York Hospital-Cornell<br />

Medical Center, New York, NY<br />

TIMOTHY B. BOONE, MD, PHD * Scott Department of Urology, Baylor College<br />

of Medicine, Houston, TX<br />

JAMES CHOI, MD * Department of Urology, Columbia University Medical Center,<br />

New York, NY<br />

ROGER R. DMOCHOWSKI, MD, FACS North Texas Center,jor Urinary Control,<br />

Fort Worth, TX<br />

MICHAEL FICAZZOLA, MD * Department of Urology, New York University Medical<br />

Center, New York, NY<br />

HOWARD B. GOLDMAN, MD Department of Urology, Case Western Reserve<br />

University, Cleveland, OH<br />

EDWARD F. IKEGUCHI, MD Department of Urology, Columbia University<br />

Medical Center, New York, NY<br />

STEVEN A. KAPLAN, MD Department of Urology, Columbia University<br />

Medical Center, New York, NY<br />

hsEPn M. KHOURY, MD, SCD 9 Division of urology, University ofNorth Carolina<br />

Medical Center, Chapel Hill, NC<br />

SERGE PETER MARINKOVIC,<br />

Medical Center, New Hyde Park, NY<br />

VICTOR W. NITTI,<br />

Center, New York, NY<br />

RAYMOND RACKLEY, MD 0 Section of <strong>Voiding</strong> Dysfunction, Department<br />

of Urology, Cleveland Clinic Foundation, Cleveland, OH<br />

ERIC S. ROVNER, MD Division of Urology, Hospital of the University<br />

of Pennsylvania, University of Pennsylvania School of Medicine,<br />

Philadelphia, PA<br />

HARRIETTE M. SCARPERO, MD Department of Urology, Ochsner Clinic,<br />

Louisiana State University Health Sciences Center, New Orleans, LA<br />

MD 0 Department of Urology, Long Island Jewish<br />

MD * Department of Urology, New York University Medical<br />

ix


X Contributol<br />

SCOTT R. SERELS, MD 0 Bladder Control Center, Urology Association of Norwalk,<br />

Nomvalk, CT<br />

STEVEN W. SIEGEL, MD Metropolitan Neurologic Specialists, St. Paul, MN<br />

STEVEN W. SUKIN, MD Scott Department of Urology, Baylor College<br />

of Medicine, Houston, TX<br />

ALEXIS E. ?"E, MD * Department of Urology, Columbia University Medical<br />

Center, New York, NY<br />

ALAN J. WEIN, PHD 0 Division of Urology, Hospital of the University<br />

of Pennsylvania, University uf Pennsylvania School ofMedicine,<br />

Philadelphia, PA<br />

J. CHRISTIAN WINTERS, MD * Department of Urology, Ochsner Clinic, Louisiana<br />

State University Health Sciences Center, New Orleans, LA


I INTRODUCTION


This Page Intentionally Left Blank


l Pathophysiology and Classification<br />

of <strong>Voiding</strong> Dysfunction<br />

Alan J. Wein, MD,<br />

and Eric S. Rovner? MD<br />

CONTENTS<br />

NORMAL LOWER URINARY TRACT FUNCTION<br />

ABNORMALITIES OF FILLING/STORAGE<br />

AND EMPTYING: OVERVIEW<br />

CLASSIFICATION OF VOIDING DYSFUNCTION<br />

REFERENCES<br />

The lower urinary tract functions as a group of inter-related structures<br />

whose purpose is to bring about efficient and low-pressure bladder<br />

filling, low-pressure urine storage with perfect continence (in the adult),<br />

and periodic voluntary urine expulsion (in the adult), again at low<br />

pressure. A simple way of looking at the pathophysiology of all types<br />

of voiding <strong>dysfunction</strong> will be presented, followed by a discussion of<br />

various systems of classification. Consistent with the author’s philosophy<br />

and prior attempts to make the understanding, evaluation, and<br />

management of voiding <strong>dysfunction</strong> as logical as possible (l), a functional<br />

and practical approach will be favored.<br />

NORMAL LOWER URINARY TRACT FUNCTION<br />

Whatever disagreements exist regarding anatomic, morphologic,<br />

physiologic, pharmacologic, and mechanical details involved in both<br />

the storage and expulsion of urine by the lower urinary tract, the<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

3


4 Wein and Rovner<br />

“experts” would agree on certain points (1,2,). The first is that the<br />

micturition cycle involves two relatively discrete processes: (1) bladder<br />

filling and urine storage and (2) bladder emptying. The second is that,<br />

whatever the details involved, one can succinctly summarize these<br />

processes from a conceptual point of view as follows.<br />

Bladder filling and urine storage require:<br />

1. Accommodation of increasing volumes of urine at a low intravesical<br />

pressure and with appropriate sensation.<br />

2. A bladder outlet that is closed at rest and remains so during increases<br />

in intra-abdominal pressure.<br />

3. Absence of involuntary bladder contractions.<br />

Bladder emptying requires:<br />

1. A coordinated contraction of the bladder smooth musculature of adequate<br />

magnitude.<br />

2. A concomitant lowering of resistance at the level of the smooth and<br />

striated sphincter.<br />

3. Absence of anatomic (as opposed to functional) obstruction.<br />

The term smooth sphincter refers to a physiologic but not an anatomic<br />

sphincter, and one that is not under voluntary control (this includes<br />

the smooth musculature of the bladder neck and proximal urethra). The<br />

striated sphincter refers to the striated musculature, which is a part of<br />

the outer wall of the urethra in both the male and female (intrinsic or<br />

intramural) and the bulky skeletal muscle group that surrounds the<br />

urethra at the level of the membranous portion of the male and the<br />

middle segment in the female (extrinsic or extramural). The extramural<br />

portion, the classically described “external urethral sphincter’’ is under<br />

voluntary control (see refs. 3 and 4 for a detailed discussion).<br />

Any type of voiding <strong>dysfunction</strong> must result from an abnormality<br />

of one or more of the factors previously listed. This two-phase concept<br />

of micturition, with the three components of each related either to the<br />

bladder or the outlet, provides a logical framework for a functional<br />

categorization of all types of voiding <strong>dysfunction</strong> and disorders as<br />

related primarily to fillinglstorage or to emptying (see tables 1 and 2).<br />

There are indeed some types of voiding <strong>dysfunction</strong> that represent<br />

combinations of fillinglstorage and emptying abnormalities. Within<br />

this scheme, however, these become readily understandable, and their<br />

detection and treatment can be logically described. All aspects of urodynamic<br />

and video urodynamic evaluation can be conceptualized as to<br />

exactly what they evaluate in terms of either bladder or outlet activity


Chapter 1 / Pathophysiology and Classification 5<br />

Table 1<br />

The Functional Classification of <strong>Voiding</strong> Dyshnction<br />

Failure to store<br />

Because of the bladder<br />

Because of the outlet<br />

Failure to empty<br />

Because of the bladder<br />

Because of the outlet<br />

during filling/storage or emptying (see Table 3). One can easily classify<br />

all known treatments for voiding <strong>dysfunction</strong> under the broad categories<br />

of whether they facilitate fillinghtorage or emptying and whether they<br />

do so by an action primarily on the bladder or on one or more of the<br />

components of the bladder outlet (see Tables 4 and 5). Finally, the<br />

individual disorders produced by various neuromuscular <strong>dysfunction</strong>s<br />

can be thought of in terms of whether they produce primarily storage<br />

or emptying abnormalities, or a combination of the two.<br />

A ~ N O ~ I T I OF E S FILLING/STORGE<br />

AND EMPTYING: OVERVIEW<br />

The pathophysiology of failure of the lower urinary tract to fill with<br />

or store urine adequately may be secondary to reasons related to the<br />

bladder, the outlet, or both. Hyperactivity of the bladder during filling<br />

can be expressed as phasic involuntary contractions, as low compliance,<br />

or as a combination. Involuntary contractions are most commonly seen<br />

in association with neurologic disease or following neurologic injury;<br />

however, they may also be associated with aging, inflammation or<br />

irritation of the bladder wall, bladder outlet obstruction, or they may<br />

be idiopathic. Decreased compliance during filling may be secondary<br />

to neurologic disease, usually at a sacral or infrasacral level, but may<br />

also result from any process that destroys the viscoelastic or elastic<br />

properties of the bladder wall. Storage failure may also occur in the<br />

absence of hyperactivity secondary to hypersensitivity or pain during<br />

filling. Irritation and inflammation can be responsible, as well as neurologic,<br />

psychologic, or idiopathic causes. The classic clinical example<br />

is interstitial cystitis.<br />

Decreased outlet resistance may result from any process that damages<br />

the innervation, structural elements, or support of the smooth or striated


6 Wein and Rovner<br />

Table 2<br />

The kpanded Functional Classification<br />

Failure to store<br />

Because of the bladder<br />

Detrusor hyperactivity<br />

Involuntary contractions<br />

Suprasacral neurologic disease<br />

Bladder outlet obstruction<br />

Idiopathic<br />

Decreased compliance<br />

Neurologic disease<br />

Fibrosis<br />

Idiopathic<br />

Detrusor hypersensitivity<br />

Inflammatory<br />

Infectious<br />

Neurologic<br />

Psychologic<br />

Idiopathic<br />

Because of the outlet<br />

Stress incontinence<br />

Nonfunctional bladder necldproximal urethra<br />

Failure to empty<br />

Because of the<br />

Neurologic<br />

Myogenic<br />

Psychogenic<br />

Idiopathic<br />

Because of the<br />

Anatomic<br />

bladder<br />

outlet<br />

Prostatic obstruction<br />

Bladder neck contracture<br />

Urethral stricture<br />

Urethral compression<br />

Functional<br />

Smooth sphincter dyssynergia<br />

Striated sphincter dyssynergia<br />

sphincter. This may occur with, neurologic disease or injury, surgical<br />

or other mechanical trauma, or aging. Assuming the bladder neck and<br />

proximal urethra are competent at rest, lack of a stable suburethral<br />

supportive layer (see ref. 5) seems a plausible explanation of the primary<br />

factor responsible for genuine stress urinary incontinence in the female.


Chapter 1 / Pathophysiology and Classification 7<br />

Table 3<br />

Urodynamics Simplified"<br />

Bladder<br />

FILLING/STORAGE Pvesb Pde,"(FCMGd) UPPf<br />

PHASE DLPPe<br />

FLUOROh<br />

EMPTYING<br />

PHASE<br />

Pve; (VC?V~G)~ Pde/<br />

MUPPl<br />

FLUORO"<br />

EM@<br />

(<br />

FLOW"<br />

1<br />

"This functional conceptualization of urodynamics categorizes each study as to<br />

whether it examines bladder or outlet activity during the fillinglstorage or emptying<br />

phase of micturition. In this scheme, uroflow and residual urine integrate the activity<br />

of the bladder and the outlet during the emptying phase.<br />

bTcTotal bladder (P,,,) and detrusor (Pdet) pressures during a filling cystometrogram<br />

(FCMG).<br />

dFilling cystometrogram.<br />

eDetrusor leak point pressure.<br />

fUrethral pressure profilometry.<br />

g Valsalva leak point pressure.<br />

h Fluoroscopy of outlet during fillinglstorage.<br />

"jTotal bladder and detrusor pressures during a voiding cystometrogram (VCMG).<br />

k<strong>Voiding</strong> cystometrogram.<br />

'Micturitional urethral pressure profilometry.<br />

"Fluoroscopy of outlet during emptying.<br />

"Electromyography of periurethral striated musculature.<br />

"Flowmetry.<br />

PResidual urine.<br />

Such failure can occur because<br />

of laxity or hypermobility, each resulting<br />

in a failure of the normal transmission of intra-abdominal pressure<br />

increases to the bladder outlet. The primary etiologic factors may be<br />

any of the causes of pelvic floor relaxation or weakness.<br />

The treatment of fillinghtorage abnormalities is directed toward<br />

inhibiting bladder contractility, decreasing sensory input, or mechani-<br />

cally increasing bladder capacity; or, toward increasing outlet resis-<br />

tance, either continuously or just during increases in intra-abdominal<br />

pressure.<br />

Absolute or relative failure to empty results from decreased bladder<br />

contractility (a decrease in magnitude or duration), increased outlet<br />

resistance, or both. Absolute or relative failure of bladder contractility


8 Wein and Rovner<br />

Table 4<br />

Therapy to Facilitate Bladder Filling and Urine Storage<br />

Inhibiting Bladder Contractility, Decreasing Sensory Input and/or<br />

Increasing Bladder Capacity<br />

Behavioral therapy<br />

Timed bladder emptying<br />

Bladder training; biofeedback<br />

Pharmacologic therapy<br />

Anticholinergic agents<br />

Musculotropic relaxants<br />

Calcium Antagonists<br />

Potassium channel openers<br />

Prostaglandin inhibitors<br />

Beta-adrenergic agonists<br />

Alpha-adrenergic antagonists<br />

Tricyclic antidepressants<br />

Dimethyl sulfoxide (DMSO)<br />

Polysynaptic inhibitors<br />

Therapy decreasing sensory input<br />

Bladder overdistention<br />

Electrical stimulation (Reflex inhibition); Neuromodulation<br />

Acupuncture<br />

Interruption of innervation<br />

Central (subarachnoid block)<br />

Sacral rhizotomy, selective sacral rhizotomy<br />

Perivesical (peripheral bladder denervation)<br />

Augmentation cystoplasty<br />

Increasing outlet resistance<br />

Physiotherapy and biofeedback<br />

Electrical stimulation<br />

Pharmacologic therapy<br />

Alpha-adrenergic agonists<br />

Tricyclic antidepressants<br />

Beta-adrenergic antagonists, agonists<br />

Estrogens<br />

Vesicourethral suspension (Stress urinary incontinence)<br />

Nonsurgical mechanical compression<br />

Periurethral polytef injection<br />

Periurethral collagen injection<br />

Occlusive and supportive devices; urethral plugs<br />

Surgical mechanical cornpression<br />

Sling procedures<br />

Closure of the bladder outlet<br />

Artificial urinary sphincter<br />

Bladder outlet reconstruction<br />

Circumventing the problem<br />

Antidiuretic hormone-like agents<br />

Diuretics<br />

Intermittent catheterization<br />

Continuous catheterization<br />

Urinary diversion<br />

External collecting devices<br />

Absorbent products


Chapter 1 / Pathophysiology and Classification. 9<br />

Table 5<br />

Therapy to Facilitate Bladder Emptying<br />

Increasing intravesical pressure and/or bladder contractility<br />

External Compression, Valsalva maneuver<br />

Promotion or initiation of reflex contractions<br />

Trigger zones or maneuvers<br />

Bladder training, tidal drainage<br />

Pharmacologic therapy<br />

Parasympathomimetic agents<br />

Prostaglandins<br />

Blockers of inhibition<br />

Alpha-adrenergic antagonists<br />

Opioid antagonists<br />

Reduction cystoplasty<br />

Electric stimulation<br />

Directly to the bladder or spinal cord<br />

To the nerve roots<br />

Transurethral intravesical electrotherapy<br />

Decreasing outlet resistance<br />

At a site of anatomic obstruction<br />

At the level of the prostate<br />

Pharmacologic<br />

Decrease prostatic size<br />

Decrease prostatic tone<br />

Balloon dilatation<br />

Intraurethral stent<br />

Urethral stricture repair or dilatation<br />

At the level of the smooth sphincter<br />

Pharmacologic therapy<br />

Transurethral resection or incision of the bladder neck<br />

Y-V plasty of the bladder neck<br />

Alpha-adrenergic antagonists<br />

Beta-adrenergic agonists<br />

At the level of the striated sphincter<br />

Biofeedback and psychotherapy<br />

Pharmacologic therapy<br />

Skeletal muscle relaxants<br />

Benzodiazepines<br />

Baclofen<br />

Dantrolene<br />

Surgical sphincterotomy, botulinum A toxin<br />

Urethral overdilation<br />

Urethral stent<br />

Pudendal nerve interruption<br />

Alpha-adrenergic antagonists<br />

Circumventing the problem<br />

Intermittent catheterization<br />

Continuous catheterization<br />

Urinary diversion


10 Wein and Rovner<br />

may result from temporary or permanent alteration in one or more of<br />

the neuromuscular mechanisms necessary for initiating and maintaining<br />

a normal detrusor contraction. Inhibition of the voiding reflex in a<br />

neurologically normal individual may also occur by a reflex mechanism<br />

secondary to painful stimuli, especially from the pelvic and perineal<br />

areas-or such an inhibition may be psychogenic. Non-neurogenic<br />

causes can also include impairment of bladder smooth muscle function,<br />

which may result from overdistention, severe infection, or fibrosis.<br />

Pathologically increased outlet resistance is generally seen in the<br />

male and is most often secondary to anatomic obstruction, but it may<br />

be secondary to a failure of coordination (relaxation) of the striated or<br />

smooth sphincter during bladder contraction. Striated sphincter dyssynergia<br />

is a common cause of functional (as opposed to fixed anatomic)<br />

obstruction in patients with neurologic disease or injury.<br />

The treatment of emptying failure generally consists of attempts to<br />

increase intravesical pressure or facilitate or stimulate the micturition<br />

reflex, to decrease outlet resistance, or both. If all else fails or the<br />

attempt is impractical, intermittent catheterization is an effective way<br />

of circumventing emptying failure.<br />

CLASSIFICATION OF VOIDING DYSFUNCTION<br />

The purpose of any classification system should be to facilitate<br />

understanding and management. A good classification should serve as<br />

intellectual shorthand and should convey, in a few key words or phrases,<br />

the essence of a clinical situation. An ideal system for all types of<br />

voiding <strong>dysfunction</strong> would include or imply a number of factors:<br />

I. the conclusions reached from urodynamic testing;<br />

2. expected clinical symptoms<br />

3. the approximate site and type of a neurologic lesion, or lack of one.<br />

If the various categories accurately portray pathophysiology, treatment<br />

options should then be obvious, and a treatment “menu” should be<br />

evident. Most systems of classification for voiding <strong>dysfunction</strong> were<br />

formulated to describe <strong>dysfunction</strong>s secondary to neurologic disease<br />

or injury. The ideal system should be applicable to all types of voiding<br />

<strong>dysfunction</strong>. Based upon the data obtained from the neurourologic<br />

evaluation, a given voiding <strong>dysfunction</strong> can be categorized in a number<br />

of descriptive systems. None are perfect.


Chapter 1 / Pathophysiology and Classification 11<br />

Sensory neuron lesion<br />

Incomplete, balanced<br />

Complete, unbalanced<br />

Motor neuron lesion<br />

Balanced<br />

Imbalanced<br />

Sensory-motor neuron lesion.<br />

Upper motor neuron lesion<br />

Complete, balanced<br />

Complete, unbalanced<br />

Incomplete, balanced<br />

Incomplete, unbalanced<br />

Lower motor neuron lesion<br />

Complete, balanced<br />

Complete, unbalanced<br />

Incomplete, balanced<br />

Incomplete, unbalanced<br />

Mixed lesion<br />

Table 6<br />

The Bors-Comarr Classification<br />

Upper somatomotor neuron, lower visceromotor neuron<br />

Lower somatomotor neuron, upper visceromotor neuron<br />

Normal somatomotor neuron, lower visceromotor neuron<br />

Bors-Comarr ClasszJicdtion (see Table 6 and re$ 6)<br />

This classification system was deduced from clinical observation of<br />

patients with traumatic spinal cord injury. This system applies only to<br />

patients with neurologic <strong>dysfunction</strong> and considers three factors:<br />

1. The anatomic localization of the lesion;<br />

2. The neurologic Completeness of the lesion, and<br />

3. Whether lower urinary tract function is “balanced” or “unbalanced.”<br />

The latter terms are based solely on the percentage of residual urine<br />

relative to bladder capacity. “Unbalanced” implies greater than 20%<br />

residual urine in a patient with an upper motor neuron (UMN) lesion<br />

or 10% in a patient with a lower motor neuron (LMN) lesion. This<br />

relative residual urine volume was ideally meant to imply coordination<br />

(synergy) or dyssynergia of the smooth and striated sphincters during<br />

bladder contraction or attempted micturition by abdominal straining or<br />

Crede. The determination of the completeness of the lesion is made on<br />

the basis of a thorough neurologic examination. The system erroneously


12 Wein and Rovner<br />

assumed that the sacral spinal cord is the primary reflex center for<br />

micturition. “LMN” implies collectively the preganglionic and postganglionic<br />

parasympathetic autonomic fibers that innervate the bladder and<br />

outlet and originate as preganglionic fibers in the sacral spinal cord.<br />

The term is used in an analogy to efferent somatic nerve fibers, such<br />

as those of the pudendal nerve, which originate in the same sacral cord<br />

segment but terminate directly on pelvic floor striated musculature<br />

without the interposition of ganglia. “UMN” is used in a similar analogy<br />

to the somatic nervous system to describe those descending autonomic<br />

pathways above the sacral spinal cord (the origin of the motor efferent<br />

supply to the bladder).<br />

In this system, “upper motor neuron bladder” refers to the pattern<br />

of micturition that results from an injury to the suprasacral spinal cord<br />

after the period of spinal shock has passed, assuming that the sacral<br />

spinal cord and the sacral nerve roots are intact and that the pelvic and<br />

pudendal nerve reflexes are intact. Lower motor neuron bladder refers<br />

to the pattern resulting if the sacral spinal cord or sacral roots are<br />

damaged and the reflex pattern through the autonomic and somatic<br />

nerves that emanate from these segments is absent. This system implies<br />

that if skeletal muscle spasticity exists below the level of the lesion,<br />

the lesion is above the sacral spinal cord and is by definition an UMN<br />

lesion. This type of lesion is characterized by detrusor hyperreflexia<br />

during filling. If flaccidity of the skeletal musculature below the level<br />

of a lesion exists, an LMN lesion is assumed to exist, implying detrusor<br />

areflexia. Exceptions occur and are classified in a “mixed lesion” group<br />

characterized either by detrusor hyperreflexia with a flaccid paralysis<br />

below the level of the lesion or by detrusor areflexia with spasticity<br />

or normal skeletal muscle tone neurologically below the lesion level.<br />

The use of this system is illustrated as follows. A complete, unbalanced,<br />

UMN lesion, implies a neurologically complete lesion above<br />

the level of the sacral spinal cord that results in skeletal muscle spasticity<br />

below the level of the injury. Detrusor hyperreflexia exists during filling,<br />

but a residual urine volume of greater than 20% of the bladder capacity<br />

is left after bladder contraction, implying obstruction in the area of<br />

the bladder outlet during the hyperreflexic detrusor contraction. This<br />

obstruction is generally due to striated sphincter dyssynergia, typically<br />

occurring in patients who are paraplegic and quadriplegic with lesions<br />

between the cervical and the sacral spinal cord. Smooth sphincter<br />

dyssynergia may be seen as well in patients with spinal cord lesions<br />

above the level of T6, usually in association with autonomic hyperreflexia.<br />

An LMN lesion, complete, unbalanced, implies a neurologi-


Chapter 1 I Pathophysiology and Classification 13<br />

Table 7<br />

The Hald-Bradley Classification<br />

Suprasacral lesion<br />

Suprasacral spinal lesion<br />

Infrasacral lesion<br />

Peripheral autonomic neuropathy<br />

Muscular lesion<br />

cally complete lesion at the level of the sacral spinal cord or of the<br />

sacral roots, resulting in skeletal muscle flaccidity below that level.<br />

Detrusor areflexia results, and whatever measures the patient may use<br />

to increase intravesical pressure during attempted voiding are not suffi-<br />

cient to decrease residual urine to less than 10% of bladder capacity.<br />

This classification system applies best to spinal cord injury patients<br />

with complete neurologic lesions after spinal shock has passed. It is<br />

difficult to apply to patients with multicentric neurologic disease and<br />

cannot be used at all for patients with non-neurologic disease. The<br />

system fails to reconcile the clinical and urodynamic variability exhib-<br />

ited by patients who, by neurological exam alone, seem to have similar<br />

lesions. The period of spinal shock that immediately follows severe<br />

cord injury is generally associated with bladder areflexia, whatever the<br />

status of the sacral somatic reflexes. Temporary or permanent changes<br />

in bladder or outlet activity during fillinghtorage or emptying may<br />

occur secondary to a number of factors such as chronic overdistention,<br />

infection, and reinnervation or reorganization of neural pathways fol-<br />

lowing injury or disease; such changes make it impossible to always<br />

accurately predict lower urinary tract activity solely on the basis of the<br />

level of the neurologic lesion. Finally, although the terms “balanced”<br />

and “unbalanced” are helpful, in that they describe the presence or<br />

absence of a certain relative percentage of residual urine, they do not<br />

necessarily imply the true functional significance of a lesion, which<br />

depends on the potential for damage to the lower or upper urinary<br />

tracts, and also on the social and vocational disability that results.<br />

Huld-Bradley Classzfication (see ruble 7 and ref: 7)<br />

This is described as a simple neurotopographic classification. A<br />

supraspinal lesion implies synergy between detrusor contraction and<br />

smooth and striated sphincters, but defective inhibition of the voiding<br />

reflex. Detrusor hyperreflexia generally occurs and sensation is usually


14 Wein and Rovner<br />

preserved. However, depending on the site of the lesion, detrusor areflexia<br />

and defective sensation may be seen. A suprasacral spinal lesion<br />

is roughly equivalent to what is described as a UMN lesion in the<br />

Bors-Comarr classification. An infrasacral lesion is roughly equivalent<br />

to an LMN lesion. Peripheral autonomic neuropathy is most frequently<br />

encountered in the diabetic and is characterized by deficient bladder<br />

sensation, gradually increasing residual urine and ultimate decompensation,<br />

with loss of detrusor contractility. A muscular lesion can involve<br />

the detrusor itself, the smooth sphincter, or any portion, or all, of the<br />

striated sphincter. The resultant <strong>dysfunction</strong> is dependent on which<br />

structure is affected. Detrusor <strong>dysfunction</strong> is the most common and<br />

generally results from decompensation, following long-standing bladder<br />

outlet obstruction.<br />

This is a primarily neurologic system based upon a conceptualization<br />

of central nervous system control of the lower urinary tract as including<br />

four neurologic “loops.” Dysfunctions are classified according to the<br />

loop affected.<br />

Loop 1 consists of neuronal connections between the cerebral cortex<br />

and the pontine-mesencephalic micturition center; this coordinates voluntary<br />

control of the detrusor reflex. Loop 1 lesions are seen in conditions<br />

such as brain tumor, cerebrovascular accident or disease, and<br />

cerebral trophy with dementia. The final result is characteristically<br />

detrusor hyperreflexia.<br />

Loop 2 includes the intraspinal pathway of detrusor muscle afferents<br />

to the brain-stem micturition center and the motor impulses from this<br />

center to the sacral spinal cord. Loop 2 is thought to coordinate and<br />

provide for a detrusor reflex of adequate temporal duration to allow<br />

complete voiding. Partial interruption by spinal cord injury results in<br />

a detrusor reflex of low threshold and in poor emptying with residual<br />

urine. Spinal cord transection of loop 2 acutely produces detrusor<br />

areflexia and urinary retention-spinal shock. After this has passed,<br />

detrusor hyperreflexia results.<br />

Loop 3 consists of the peripheral detrusor afferent axons and their<br />

pathway in the spinal cord; these terminate by synapsing on pudendal<br />

motor neurons that ultimately innervate periurethral striated muscle.<br />

Loop 3 was thought to provide a neurologic substrate for coordinated


Chapter l / Pathophysiology and Classification 15<br />

reciprocal action of the bladder and striated sphincter. Loop 3 <strong>dysfunction</strong><br />

could be responsible for detrusor-striated sphincter dyssynergia or<br />

involuntary sphincter relaxation.<br />

Loop 4 consists of two components. Loop 4A is the suprasacral<br />

afferent and efferent innervation of the pudendal motor neurons to the<br />

periurethrla striated musculature which synapse on pudendal motor<br />

neurons in Onuf's nucleus-the segmental innervation of the periurethral<br />

striated muscle. In contrast to the stimulation of detrusor afferent<br />

fibers, which produce inhibitory postsynaptic potentials in pudendal<br />

motor neurons through loop 3, pudendal nerve afferents produce excitatory<br />

postsynaptic potentials in those motor neurons through loop 4B.<br />

These provide for contraction of the periurethral striated muscle during<br />

bladder filling and urine storage. The related sensory impulses arise<br />

from muscle spindles and tendon organs in the pelvic floor musculature.<br />

Loop 4 provides for volitional control of the striated sphincter. Abnormalities<br />

of the suprasacral portion result in abnormal responses of the<br />

pudendal motor neurons to bladder filling and emptying, manifested<br />

as detrusor-striated sphincter dyssynergia, and/or loss of the ability to<br />

voluntarily contract the striated sphincter.<br />

This system is sophisticated and reflects the ingenuity and neurophysiologic<br />

expertise of its originator, Dr. William Bradley. For some neurologists,<br />

this method may be an excellent way to conceptualize the<br />

neurophysiology involved, assuming that they agree on the existence<br />

and significance of all four loops. Most urologists find this system<br />

difficult to use for many types of neurogenic voiding <strong>dysfunction</strong> and<br />

not at all applicable to non-neurogenic voiding <strong>dysfunction</strong>. Urodynamically,<br />

it may be extremely difficult to test the intactness of each loop<br />

system, and multicentric and partial lesions are difficult to describe.<br />

The Lapides Class$kation (see Table 8 and re$ S)<br />

This is a modification of a system originally proposed by McLellan<br />

(a neurologist) in 1939 (9). This remains one of the most familiar<br />

systems to urologists and non-urologists because it describes in recog-<br />

nizable shorthand the clinical and cystometric conditions of many types<br />

of neurogenic voiding <strong>dysfunction</strong>.<br />

A sensory neurogenic bladder results from intemption of the sensory<br />

fibers between the bladder and spinal cord or the afferent tracts to the<br />

brain. Diabetes mellitus, tabes dorsalis, and pernicious anemia are most<br />

commonly responsible. The first clinical changes are described as those


16 Wein Rovner<br />

Table 8<br />

The Lapides Classification.<br />

Sensory neurogenic bladder<br />

Motor paralytic bladder<br />

Uninhibited neurogenic bladder<br />

Reflex neurogenic bladder<br />

Autonomous neurogenic bladder<br />

of impaired sensation of bladder distention. Unless voiding is initiated<br />

on a timed basis, varying degrees of bladder overdistention can result<br />

with resultant hypotonicity. With bladder decompensation, significant<br />

amounts of residual urine are found and, at this time, the cystometric<br />

curve generally demonstrates a large capacity bladder with a flat high<br />

compliance low pressure filling curve.<br />

A motor paralytic bladder results from disease processes that destroy<br />

the parasympathetic motor innervation of the bladder. Extensive pelvic<br />

surgery or trauma may produce this. Herpes zoster has been listed as<br />

a cause as well, but recent evidence suggests that the voiding <strong>dysfunction</strong><br />

seen with herpes is more related to a problem with afferent input.<br />

The early symptoms may vary from painful urinary retention to only<br />

a relative inability to initiate and maintain normal micturition. Early<br />

cystornetric filling is normal but without a voluntary bladder contraction<br />

at capacity. Chronic overdistention and decompensation may occur in<br />

a large capacity bladder with a flat, low-pressure filling curve; a large<br />

residual urine may result.<br />

The uninhibited neurogenic bladder was described originally as<br />

resulting from injury or disease to the “corticoregulatory tract.” The<br />

sacral spinal cord was presumed to be the micturition reflex center,<br />

and this “corticoregulatory tract” was believed to exert an inhibitory<br />

influence on the sacral micturition reflex center. A destructive lesion<br />

in this tract would then result in overfacilitation or lack of inhibition<br />

of the micturition reflex. Cerebrovascular accident, brain or spinal cord<br />

tumor, Parkinson’s disease, and demyelinating disease are the most<br />

common causes in this category. The voiding <strong>dysfunction</strong> is most<br />

often characterized symptomatically by frequency, urgency, and urge<br />

incontinence. Urodynamically, one sees normal sensation with an involuntary<br />

bladder contraction at low filling volumes. Residual urine is<br />

characteristically low unless anatomic outlet obstruction or true smooth<br />

or striated sphincter dyssynergia occurs. The patient generally can


Chapter 1 I Pathophysiology and Classification 17<br />

initiate a bladder contraction voluntarily, but is often unable to do so<br />

during cystometry because sufficient urine storage cannot occur before<br />

detrusor hyperreflexia is stimulated.<br />

Reflex neurogenic bladder describes the postspinal shock condition<br />

that exists after complete interruption of the sensory and motor pathways<br />

between the sacral spinal cord and the brain stem. Most commonly,<br />

this occurs in traumatic spinal cord injury and transverse myelitis, but<br />

may occur with extensive demyelinating disease or any process that<br />

produces significant spinal cord destruction as well. Typically, there<br />

is no bladder sensation and there is inability to initiate voluntary micturition.<br />

Incontinence without sensation generally occurs because of low<br />

volume involuntary bladder contraction. Striated sphincter dyssynergia<br />

is the rule. This type of lesion is essentially equivalent to a complete<br />

UMN lesion in the Bors-Comarr system.<br />

An autonomous neurogenic bladder results from complete motor and<br />

sensory separation of the bladder from the sacral spinal cord. This may<br />

be caused by any disease that destroys the sacral cord or causes extensive<br />

damage to the sacral roots or pelvic nerves. There is inability to voluntarily<br />

initiate micturition, no bladder reflex activity, and no specific bladder<br />

sensation. This type of bladder is equivalent to a complete LMN lesion<br />

in the Bors-Comarr system and is also the type of <strong>dysfunction</strong> seen in<br />

patients with spinal shock. This characteristic cystometric pattern is initially<br />

similar to the late stages of the motor or sensory paralytic bladder,<br />

with a marked shift to the right of the cystometric filling curve and a<br />

large bladder capacity at low intravesical pressure. However, decreased<br />

compliance may develop, secondary either to chronic inflammatory<br />

change or to the effects of denervationlcentralization with secondary neuromorphologic<br />

and neuropharmacologic reorganizational changes.<br />

Emptying capacity may vary widely, depending on the ability of the<br />

patient to increase intravesical pressure and on the resistance offered<br />

during this increase by the smooth and striated sphincters.<br />

These classic categories in their usual settings are usually easily<br />

understood and remembered, and this is why this system provides an<br />

excellent framework for teaching some fundamentals of neurogenic<br />

voiding <strong>dysfunction</strong> to students and non-urologists. Unfortunately,<br />

many patients do not exactly “fit” into one or another category. Gradations<br />

of sensory, motor, and mixed lesions occur, and the patterns<br />

produced after different types of peripheral dene~ation/defunctionalization<br />

may vary widely from those which are classically described.<br />

The system is applicable only to neuropathic <strong>dysfunction</strong>.


18 Wein Rovner<br />

Table 9<br />

A Urodynamic Classification“<br />

Detrusor hyperreflexia (or normoreflexia)<br />

Coordinated sphincters<br />

Striated sphincter dyssynergia<br />

Smooth sphincter dyssynergia<br />

Nonrelaxing smooth sphincter<br />

Detrusor areflexia<br />

Coordinated sphincters<br />

Nonrelaxing striated sphincter<br />

Denervated striated sphincter<br />

Nonrelaxing smooth sphincter<br />

“Adapted with permission from ref. 10.<br />

Urodynamic Clas$cation (see Table 9 and re$ 10)<br />

Systems of classification have evolved based solely on objective<br />

urodynamic data. When exact urodynamic classification is possible,<br />

this sort of system can provide an exact description of the voiding<br />

<strong>dysfunction</strong> that occurs. If a normal or hyperreflexic detrusor exists<br />

with coordinated smooth and striated sphincter function and without<br />

anatomic obstruction, normal bladder emptying should occur.<br />

Detrusor hyperreflexia is most commonly associated with neurologic<br />

lesions above the sacral spinal cord. Striated sphincter dyssynergia is<br />

most commonly seen after complete suprasacral spinal cord injury,<br />

following the period of spinal shock. Smooth sphincter dyssynergia is<br />

seen most classically in autonomic hyperreflexia when it is characteristi-<br />

cally associated with detrusor hyperreflexia and striated sphincter dys-<br />

synergia. Detrusor areflexia may be secondary to bladder muscle<br />

decompensation or to various other conditions that produce inhibition<br />

at the level of the brain-stem micturition center, the sacral spinal cord,<br />

bladder ganglia, or bladder smooth muscle.<br />

This classification system is easiest to use when detrusor hyper-<br />

reflexia or normoreflexia exists. Thus, a typical Tl0 level paraplegic<br />

exhibits detrusor hyperreflexia, smooth-sphincter synergia, and striated-<br />

sphincter dyssynergia. When a voluntary or hyperreflexic contraction<br />

cannot be elicited, the system is more difficult to use, because it is not<br />

appropriate to speak of true sphincter dyssynergia in the absence of<br />

an opposing bladder contraction. There are obviously many variations<br />

and extensions of such a system. Such systems work only when total<br />

urodynamic agreement exists among classifiers. Unfortunately, there


Chapter 1 / Pathophysiology and Classification 19<br />

Table 10<br />

The International Continence Society Classification“<br />

Storage Phase <strong>Voiding</strong> Phase<br />

Bladder Function<br />

Detrusor activity<br />

Normal or stable<br />

Overactive<br />

Unstable<br />

Hyperreflexic<br />

Bladder sensation<br />

Normal<br />

Increased or hypersensitive<br />

Reduced or hyposensitive<br />

Absent<br />

Bladder capacity<br />

Normal<br />

High<br />

Low<br />

Compliance<br />

Normal<br />

High<br />

LOW<br />

Urethral Function<br />

Normal<br />

Incompetent<br />

“Adapted with permission from ref. 11.<br />

Bladder Function<br />

Detrusor activity<br />

Normal<br />

Underactive<br />

Acontractile<br />

Urethral Function<br />

Normal<br />

Obstructive<br />

Overactive<br />

Mechanical<br />

are many voiding <strong>dysfunction</strong>s that do not fit neatly into a urodynarnic<br />

classification system that is agreed upon by all “experts.” As sophisti-<br />

cated urodynarnic technology and understanding improve, this type of<br />

classification system may supplant some others in general use.<br />

International Continence Society Classz;ficdtion<br />

(see Table IO and ref: I I)<br />

This is in many ways an extension of a urodynamic classification<br />

system. The storage and voiding phases of micturition are described<br />

separately, and, within each, various designations are applied to describe<br />

bladder and urethral function (11).


20 Wein and Rovner<br />

Normal bladder function during fillingktorage implies no significant<br />

rises in detrusor pressure (stability). Overactive detrusor function indi-<br />

cates the presence of involuntary contractions. If owing to neurologic<br />

disease, the term detrusor hyperreflexia is used; if not, the phenomenon<br />

is known as detrusor instability. Bladder sensation can be categorized<br />

only in qualitative terms as indicated. Bladder capacity and compliance<br />

(A) volume/A pressure) are cystometric measurements. Normal urethral<br />

function during filling/storage indicates a positive urethral closure pres-<br />

sure (urethral pressure minus bladder pressure) even with increases in<br />

intra-abdominal pressure. Incompetent urethral function during filling/<br />

storage implies urine leakage in the absence of a detrusor contraction.<br />

This may be secondary to genuine stress incontinence, intrinsic sphinc-<br />

ter <strong>dysfunction</strong>, or an involuntary fall in urethral pressure in the absence<br />

of a detrusor contraction.<br />

During the voiding/emptying phase of micturition, normal detrusor<br />

activity implies voiding by a voluntarily initiated, sustained contraction<br />

that also can be suppressed voluntarily. An underactive detrusor defines<br />

a contraction of inadequate magnitude or/and duration to empty the<br />

bladder with a normal time span. An acontractile detrusor is one that<br />

cannot be demonstrated to contract during urodynamic testing. Areflexia<br />

is defined as acontractility owing to an abnormality of neural control,<br />

implying the complete absence of centrally coordinated contraction.<br />

Normal urethral function during voiding indicates opening prior to<br />

micturition to allow bladder emptying. An obstructed urethra is one<br />

which contracts against a detrusor contraction or fails to open (nonrelax-<br />

ation) with attempted micturition. Contraction may be owing to smooth<br />

or striated sphincter dyssynergia. Striated sphincter dyssynergia is a<br />

term that should be applied only when neurologic disease is present. A<br />

similar syndrome but without neurologic disease is called <strong>dysfunction</strong>al<br />

voiding. Mechanical obstruction is generally anatomical and caused by<br />

BPH, urethral or bladder neck stricture, scarring or compression, or,<br />

rarely, kinking of a portion of the urethra during straining.<br />

<strong>Voiding</strong> <strong>dysfunction</strong> in a classic T10 level paraplegic after spinal<br />

shock has passed would be classified as follows:<br />

1. Storage phase: overactive hyperreflexic detrusor, absent sensation, low<br />

capacity, normal compliance, normal urethral closure function.<br />

2. '<strong>Voiding</strong> phase: overactive obstructive urethral function, ? normal detrusor<br />

activity (actually, hyperreflexic).<br />

The voiding <strong>dysfunction</strong> of a stroke patient with urgency inconti-<br />

nence would most likely be classified during storage as overactive


Chapter 1 I Pathophysiology and Classification 21<br />

hyperreflexic detrusor, normal sensation, low capacity, normal compliance,<br />

and normal urethral closure function. During voiding, the <strong>dysfunction</strong><br />

would be classified as normal detrusor activity and normal urethral<br />

function, assuming that no anatomic obstruction existed.<br />

The Functional System (see Tables I and 2)<br />

Classification of voiding <strong>dysfunction</strong> can also be formulated on a<br />

simple functional basis, describing the <strong>dysfunction</strong> in terms of whether<br />

the deficit produced is primarily one of the fillinghtorage or emptying<br />

phase of micturition (see Table 1) (1,1.2)* This type of system was<br />

proposed initially by Quesada et al. (13), and is an excellent alternative<br />

when a particular <strong>dysfunction</strong> does not readily lend itself to a generally<br />

agreed upon classification elsewhere, This simple-minded scheme<br />

assumes only that, whatever their differences, all “experts” would agree<br />

on the two-phase concept of micturition and upon the simple overall<br />

mechanisms underlying the normality of each phase (see section on<br />

normal lower urinary tract function).<br />

Storage failure results either because of bladder or outlet abnormalities<br />

or a combination. Bladder abnormalities include involuntary bladder<br />

contractions, low compliance, and hypersensitivity. The outlet abnormalities<br />

include only an intermittent or continuous decrease in outlet<br />

resistance.<br />

Similiarly, emptying failure can occur because of bladder or outlet<br />

abnormalities or a combination of the two. The bladder side includes<br />

inadequate or unsustained bladder contractility, and the outlet side<br />

includes anatomic obstruction and sphincter(s) dyssynergia.<br />

Failure in either category generally is not absolute, but more frequently<br />

relative. Such a functional system can easily be “expanded”<br />

and made more complicated to include etiologic or specific urodynamic<br />

connotations (see Table 2). However, the simplified system is perfectly<br />

workable and avoids argument in those complex situations in which<br />

the exact etiology or urodynamic mechanism for a voiding <strong>dysfunction</strong><br />

cannot be agreed upon.<br />

Proper use of this system for a given voiding <strong>dysfunction</strong> obviously<br />

requires a reasonably accurate notion of what the urodynamic data<br />

show. However, an exact diagnosis is not required for treatment. It<br />

should be recognized that some patients do not have only a discrete<br />

storage or emptying failure, and the existence of combination deficits<br />

must be recognized to properly utilize this system of classification. The


22 Wein and Rovner<br />

classic T10 paraplegic after spinal shock generally exhibits a relative<br />

failure to store because of detrusor hyperreflexia and a relative failure<br />

to empty because of striated sphincter dyssynergia. With such a combi-<br />

nation deficit, to utilize this classification system as a guide to treatment<br />

one must assume that one of the deficits is primary and that significant<br />

improvement will result from its treatment alone, or, that the voiding<br />

<strong>dysfunction</strong> can be converted primarily to a disorder either of storage<br />

or emptying by means of nonsurgical or surgical therapy. The resultant<br />

deficit can then be treated or circumvented. Using the same example,<br />

the combined deficit in a T10 paraplegic can be converted primarily<br />

to a storage failure by procedures directed at the dyssynergic striated<br />

sphincter; the resultant incontinence (secondary to detrusor hyper-<br />

reflexia) can be circumvented (in a male) with an external collecting<br />

device. Alternatively, the deficit can be converted primarily to an empty-<br />

ing failure by pharmacologic or surgical measures designed to abolish<br />

or reduce the detrusor hyperreflexia, and the resultant emptying failure<br />

can then be circumvented with clean intermittent catheterization. Other<br />

examples of combination deficits include impaired bladder contractility<br />

with sphincter <strong>dysfunction</strong>, bladder outlet obstruction with detrusor<br />

hyperactivity, bladder outlet obstruction with sphincter malfunction,<br />

and detrusor hyperactivity with impaired contractility.<br />

One of the advantages of this functional classification is that it allows<br />

the individual the liberty of “playing” with the system to suit his or<br />

her preferences without an alteration in the basic concept of “keep it<br />

simple but accurate and informative.” For instance, one could easily<br />

substitute “overactive or oversensitive bladder” and “outlet insuffi-<br />

ciency” for “because of the bladder” and “because of the outlet” under<br />

“failure to store” in Table 1. One could choose to subcategorize the<br />

bladder reasons for overactivity (see Table 2) in terms of neurogenic,<br />

myogenic, or anatomic etiologies and further subcategorize neurogenic<br />

in terms of increased afferent activity, decreased inhibitory control,<br />

increased sensitivity to efferent activity, and so forth. The system<br />

is flexible.<br />

An additional advantage to the functional system is that the underly-<br />

ing concepts can be used repeatedly to simplify many areas in neurourol-<br />

ogy. One logical extension makes urodynamics become more readily<br />

understandable (see Table 3), whereas a different type of adaptation<br />

functions especially well as a “menu” for the categorization of all the<br />

types of treatment for voiding <strong>dysfunction</strong> (see Tables 4 and 5).<br />

The major problem with the functional system is that not every<br />

voiding <strong>dysfunction</strong> can be reduced or converted primarily to a failure


Chapter 1 / Pathophysiology and Classification 23<br />

of storage or emptying. Additionally, although the functional classifica-<br />

tion of therapy that is a correlate of this scheme is entirely logical and<br />

complete, there is a danger of accepting an easy therapeutic solution<br />

and of thereby overlooking an etiology for a voiding <strong>dysfunction</strong> that<br />

is reversible at the primary level of causation. Non-neurogenic voiding<br />

<strong>dysfunction</strong>s, however, can be classified within this system, including<br />

those involving only the sensory aspect of micturition.<br />

It is obvious that no type of system is perfect. Each offers something<br />

to every clinician; although his or her level and type of training, interests,<br />

experience, and prejudices regarding the accuracy and interpretation<br />

of urodynamic data will determine a system’s usefulness. The ideal<br />

approach to a patient with voiding <strong>dysfunction</strong> continues to be a thor-<br />

ough neurourologic evaluation. If the clinician can classify a given<br />

patient’s voiding <strong>dysfunction</strong> in each system or can understand why<br />

this cannot be done, there is enough working knowledge to proceed<br />

with treatment. However, if the clinician is familiar with only one or<br />

two of these systems and a given patient does not fit these, and the<br />

reason is uncertain, the patient should certainly be studied further and<br />

the voiding <strong>dysfunction</strong> better characterized, at least before irreversible<br />

therapy is undertaken.<br />

1.<br />

2.<br />

3.<br />

4.<br />

5.<br />

6.<br />

7.<br />

8.<br />

REFERENCES<br />

Wein AJ. Barrett DM (1988) <strong>Voiding</strong> Firriction and Dysfimctioii: A Logical and<br />

Prnctical Approach. Chicago: Year Book Medical Publishers, Tnc.<br />

Wein AJ ( 1998) Pathophysiology and categorization of voiding <strong>dysfunction</strong>. In:<br />

Walsh PC, Retik AB, Vaughan ED. Jr, Wein AJ, eds. Curnpbell’s Urology (8th<br />

ed.) Philadelphia: WB Saunders Co., pp. 917-926.<br />

Zderic SA, Levin RM, Wein AJ (1996) <strong>Voiding</strong> function: relevant anatomy,<br />

physiology, pharmacology and molecular aspects. In: Gillenwater JY, Grayhack<br />

JT, Howards SS, Duckett JW, Jr, eds., Adult orid Pediatric Urology, Chicago:<br />

Year Book Medical Publishers, pp. 1159-1219.<br />

Steers WD (1 998) Physiology and pharmacology of the bladder and urethra. In:<br />

Walsh PC, Retik AB, Vaughan ED. Jr. Wein AJ, eds., Ccinzpbell’s Uro-ology, (8th<br />

ed.), Philadelphia: WB Saunders Co., pp. 870-916.<br />

DeLancey JOL (1994) Structural support of the urethra as it relates to stress<br />

urinary incontinence: the hammock hypothesis. Ail7 J Obstet Gyizecol 170: 17 13.<br />

Bors E. Cornarr AE (197 1) Neurological Urology. Baltimore: University Park<br />

Press.<br />

Hald T, Bradley WE (1982) The Urirtai? Bladder: Neurology and Dynamics,<br />

Baltimore: Williams and Wilkins.<br />

Lapides J (1 970) Neur~muscular, vesical and ureteral <strong>dysfunction</strong>. In: Campbell<br />

MF and Harrison JH, eds., Urology. Philadelphia: WB Saunders Co, pp. 1343-<br />

1379.


24 Wein and Rovner<br />

9. McLellan FC (1939) The Neurogenic Bladder. Springfield: Charles Thomas Co..<br />

pp. 57-70, 116-185.<br />

10. Krane RJ, Siroky MB (1984) Classification of voiding <strong>dysfunction</strong>: Value of<br />

classification systems. In: Barrett DM. Wein AJ, eds., Controversies in Nertro-<br />

Urology. New York: Churchill Livingstone, pp. 223-238.<br />

11. Abrams P. Blaivas JG. Stanton SL, Andersen JT (1990) The standardization of<br />

terminology of lower urinary tract function recommended by the International<br />

Continence Society. lid Urogynecol J 1 :45.<br />

12. Wein AJ (1981) Classification of neurogenic voiding <strong>dysfunction</strong>. J Urol 125:605.<br />

13. Quesada EM, Scott FB, Cardus D (1968) Functional classification of neurogenic<br />

bladder <strong>dysfunction</strong>. Arch Phys Med Relzabil 49:692.


Victor M7 Nitti, MD<br />

and Michael Ficdnnola, MD<br />

CONTENTS<br />

INTRODUCTION<br />

CLASSIFICATION OF VOIDING DYSFUNCTION<br />

HISTORY<br />

PHYSICAL EXAMINATION<br />

LABORATORY TESTING<br />

SIMPLE TESTS<br />

FOR EVALUATING VOIDING DYSFUNCTION<br />

URODYNAMICS<br />

ENDOSCOPY<br />

URINARY TRACT IMAGING<br />

REFERENCES<br />

INTRODUCTION<br />

<strong>Voiding</strong> <strong>dysfunction</strong> usually presents in one of two ways. The first<br />

is in the form of symptoms. Symptoms related to voiding <strong>dysfunction</strong><br />

are broadly referred to as lower urinary tract symptoms (LUTS). LUTS<br />

have classically been divided into obstructive symptoms such as diffi-<br />

culty initiating a stream, decreased force of urinary stream, need to<br />

push and strain to void (stranguria), hesitancy or intermittent urine<br />

flow, and irritative symptoms such as urinary frequency, urgency, and<br />

nocturia. In addition, symptoms of incontinence and lower abdominal<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Hurnana Press Inc., Totowa, NJ


26 Nitti and Ficazzola<br />

or pelvic pain may exist. The second way in which voiding <strong>dysfunction</strong><br />

presents is in the form of urinary tract decompensation such as incomplete<br />

bladder emptying or urinary retention, renal insufficiency, and<br />

recurrent urinary tract infections. It is possible for patients who present<br />

with urinary tract decompensation to have little or no symptoms. In<br />

the case of symptoms, evaluation and treatment are often driven by<br />

the degree of bother to the patient. In many cases, patients with mild<br />

LUTS of a minimal bother will not even bring these to the attention<br />

of their physician. However, when urinary tract decompensation is<br />

diagnosed, a more aggressive diagnostic and treatment plan must be<br />

implemented. There are also patients who have diseases known to effect<br />

the lower urinary tract and cause voiding <strong>dysfunction</strong>, yet do not have<br />

significant symptoms or obvious signs of decompensation. These<br />

include patients with a variety of neurological conditions such as spinal<br />

cord injuries or multiple sclerosis, or non-neurological conditions such<br />

as prior pelvic irradiation or extensive pelvic surgery. In many cases<br />

careful evaluation of the urinary tract will uncover underlying voiding<br />

<strong>dysfunction</strong>. Thus the diagnostic evaluation of voiding <strong>dysfunction</strong> will<br />

be influenced by the type and degree of bother of symptoms, the<br />

presence of urinary tract decompensation, and coexisting medical conditions<br />

that might affect the lower urinary tract or its treatment.<br />

In this chapter we will discuss the diagnostic evaluation of voiding<br />

<strong>dysfunction</strong>. It is important to realize that this evaluation is not the<br />

same for all patients and will be influenced by many factors, including<br />

those previously mentioned. Before discussing the evaluation in detail,<br />

we will first present a classification system for voiding <strong>dysfunction</strong>,<br />

which should help the clinician plan a proper diagnostic evaluation<br />

and treatment plan.<br />

CLASSIFICATION OF VOIDING DYSFUNCTION<br />

In order to formulate a plan for the diagnostic evaluation of voiding<br />

<strong>dysfunction</strong>, an understanding of the possible causes of symptoms<br />

or urinary tract decompensation and the possible manifestations of a<br />

coexisting condition is necessary. In order to accomplish this, a practical<br />

classification of voiding <strong>dysfunction</strong> is invaluable. The functional clas-<br />

sification system proposed and popularized by Wein (1) is simple and<br />

practical and allows treatment options to be formulated according to<br />

classification. In simple terms, voiding <strong>dysfunction</strong> can be divided into<br />

three categories:


Chapter 2 I Diagnostic Evaluation 27<br />

1. Failure to store urine.<br />

2. Failure to empty urine.<br />

3. Failure to store and empty.<br />

For example, the symptom of urinary frequency or incontinence is<br />

usually associated with <strong>dysfunction</strong> of the storage phase of micturition,<br />

whereas decreased force of stream or elevated postvoid residual are<br />

associated with <strong>dysfunction</strong> of the emptying phase. In addition, we can<br />

view voiding <strong>dysfunction</strong> in simple anatomical terms:<br />

l. Bladder <strong>dysfunction</strong> (overactive, underactive).<br />

2. Bladder outlet <strong>dysfunction</strong> (overactive, underactive).<br />

3. Combined bladder and outlet <strong>dysfunction</strong>.<br />

These two concepts can be combined so that one can imagine<br />

that a patient could present with urinary incontinence (failure to<br />

store) secondary to bladder overactivity or bladder outlet underactivity.<br />

Similarly a patient with urinary retention (failure to empty) might have<br />

an underactive-or hypocontractile-bladder or an overactive-or<br />

obstructing-outlet. Failure to empty and failure to store as well as<br />

bladder and outlet <strong>dysfunction</strong> are not mutually exclusive conditions<br />

and can exist in multiple combinations. These very simple concepts<br />

can be applied to all types of voiding <strong>dysfunction</strong>. Therefore when<br />

evaluating voiding <strong>dysfunction</strong>, from history and physical examination<br />

to simple and comprehensive testing, keeping these concepts in mind<br />

can greatly facilitate the process.<br />

HISTORY<br />

The patient’s history is the first step in directing the clinician toward<br />

the appropriate evaluation and treatment. It should provide the clinician<br />

with a detailed account of the precise nature of the patient’ S symptoms.<br />

It is important to remember that the history is only as accurate as the<br />

patient’s ability to describe their symptoms, therefore, some skill is<br />

required by the physician to obtain this infomation. This is especially<br />

true for patients who have difficulty communicating or those who are<br />

anxious or embarrassed about their condition. A classic illustration of<br />

this is in the study by Diokno et al. (2) in noninstitutional elderly<br />

patients. They found that only 37% of incontinent men and 41% of<br />

incontinent women told a physician about their condition (2).<br />

The history begins with an assessment of a patient’s symptoms and<br />

their onset. Each symptom should be characterized as to its onset,


28 Nitti and Ficazzola<br />

frequency, duration, severity, and bother; and exacerbating or relieving<br />

factors. It is important to note whether the onset of the symptom<br />

occurred after a specific event such as surgery, childbirth, menopause,<br />

or with the use of a new medication. Any prior treatments by other<br />

physicians for their symptoms and the resultant outcome should also<br />

be noted. Specific questions about childhood and adolescent voiding<br />

troubles or problems with toilet training should be asked.<br />

Patients will often present with one or more voiding symptoms that<br />

have been traditionally separated into irritative or obstructive in nature.<br />

Irritative voiding symptoms are common presenting complaints that<br />

may herald a number of different types of voiding <strong>dysfunction</strong>. Urgency<br />

is defined as an intense desire to void secondary to an abrupt sensation<br />

of bladder discomfort or as a conditional response from the fear of<br />

urine leakage. Frequency is defined as more than seven diurnal voids<br />

and may reflect excessive fluid intake, diuretic use, or excessive caffeine<br />

consumption. Nocturia is nighttime frequency and may be secondary<br />

to detrusor overactivity, reduced bladder capacity, or excessive fluid<br />

caffeine intake prior to bedtime. Daytime frequency without nocturia<br />

may be suggestive of timing of diuretic medications or a psychogenic<br />

component to the voiding <strong>dysfunction</strong>. Dysuria refers to the burning<br />

sensation that occurs during micturition and implies bladder, urethral,<br />

or prostatic inflammation. Obstructive voiding symptoms include<br />

decreased force of urinary stream, straining to void, hesitancy (the<br />

prolonged interval necessary to voluntarily initiate the urinary stream),<br />

and interruption of urinary stream. They may be present in men with<br />

bladder outlet obstruction secondary to benign prostatic enlargement<br />

or urethral stricture, or in women with pelvic organ prolapse. Abrams<br />

(3) has suggested replacing the terns obstructive and irritative symptoms<br />

with symptoms of storage (e.g., frequency, urgency, incontinence)<br />

and symptoms of voiding (hesitancy, decreased force of stream, incomplete<br />

emptying) (3). This is consistent with the functional classification<br />

presented above and may be more useful when evaluating patients<br />

initially by history.<br />

One of the most distressing of all urinary symptoms is incontinence.<br />

When evaluating the symptom of incontinence, it is essential to query<br />

extensively about the nature and severity of the incontinence. From a<br />

conceptual standpoint, urinary incontinence can occur as a result of<br />

bladder overactivity or inadequate urethral sphincter function (or a<br />

combination of both) and these in turn can result from a variety of<br />

conditions (4). Urinary incontinence is simply defined as the involuntary


Chapter 2 / Diagnostic Evaluation 29<br />

loss of urine, however, this can be further characterized according to<br />

the information relayed by the patient:<br />

1.<br />

2.<br />

3.<br />

4.<br />

Urge incontinence: The symptom is incontinence is associated with a<br />

sudden uncontrollable desire to void. This condition is usually due to<br />

involuntary detrusor contractions.<br />

Stress incontinence: The symptom is incontinence that occurs during<br />

coughing, sneezing, physical exertion, changes in body position, or<br />

other action that causes an increase in abdominal pressure. This condi<br />

tion may be caused by sphincter abnormalities or bladder overactivity<br />

provoked by physical activity.<br />

Unconscious incontinence: The symptom of incontinence is uncon-<br />

scious and occurs without patient awareness of urges or stress or<br />

increases in abdominal pressure. This condition may be caused by<br />

bladder overactivity, sphincter abnormalities, overflow, or extraurethr<br />

causes such as a fistula or ectopic ureter.<br />

Continuous leakage: The symptom is a complaint of continuous loss<br />

of urine. This may be caused by sphincter abnormalities or extraure-<br />

thral causes.<br />

It is not always possible to determine the etiology of incontinence<br />

based on history alone. However, careful history taking as to the nature<br />

of the incontinence and when it occurs is critical in directing the<br />

diagnostic evaluation and treatment options.<br />

As previously mentioned, it is essential to determine the duration of<br />

symptoms. When symptoms are acute, or subacute, history may reveal<br />

an obvious cause of transient voiding <strong>dysfunction</strong> as amedication or acute<br />

nonurologic illness. This is demonstrated by the mnemonic DIAPPERS<br />

developed by Resnick (5) to describe causes of transient incontinence:<br />

delerium, infection, atrophic vaginitislurethritis, pharmaceuticals, psychological,<br />

endocrine, restricted mobility, and stool impaction.<br />

Questionnaires and symptoms scores can be helpful in assessing the<br />

type of symptoms a patient has, the degree of 'bother of symptoms,<br />

and/or effect on quality of life that symptoms produce. Two such<br />

example are the American Urologic Association Symptom Index (International<br />

Prostate Symptom Score), originally designed to assess men<br />

with LUTS secondary to benign prostatic hyperplasia (6), and the I-<br />

QOL, designed to assess the impact of incontinence on quality of life<br />

(7). In cases where voiding <strong>dysfunction</strong> does not present danger to a<br />

patient, it is often the degree of bother or effect on an individual<br />

patient' S quality of life that drives evaluation and treatment decisions.<br />

Symptom scores and quality of life assessments are very useful in


30 Nitti and Ficazzola<br />

monitoring response to treatment for an individual or in clinical<br />

trials.<br />

After a thorough assessment of a patient’s presenting symptoms, the<br />

history should then focus on related areas. There are several aspects<br />

of a patient’ S history that may be intimately related to voiding function.<br />

Sexual and bowel <strong>dysfunction</strong> are often associated with voiding<br />

<strong>dysfunction</strong>. Therefore the review of symptoms should focus on these<br />

areas including defecation (constipation, diarrhea, fecal incontinence,<br />

changes in bowel movements), sexual function, dysparunia, and pelvic<br />

pain. As neurological problems are frequently associated with voiding<br />

<strong>dysfunction</strong>, a thorough neurological history is critical, including known<br />

neurologic disease as well as symptoms that could be related to occult<br />

neurological disease (back pain, radiculopathy, extremity numbness,<br />

tingling, or weakness, headaches, changes in eyesight, and so on). In<br />

addition to a focused history regarding LUTS and voiding <strong>dysfunction</strong>,<br />

a thorough urological history is important. This includes a history<br />

of hematuria, urinary tract infections, sexually transmitted diseases,<br />

urolilthiasis, and urological malignancy and their treatment.<br />

The past medical history should provide information about concurrent<br />

medical diseases, obstetric and gynecologic history, past surgical history,<br />

and medication use. Many medications have profound effects on the<br />

lower urinary tract or can effect fluid mobilization and urine production<br />

and thus contribute to LUTS. Examples of medications that may be associated<br />

with voiding <strong>dysfunction</strong> include alpha-adrenergic agonists, such<br />

as pseudoephedrine, diuretics, antidepressants, and anticholinergics.<br />

A detailed history of known neurological diseases (e.g., stroke,<br />

Parkinson’s disease, spinal cord injury, multiple sclerosis, myelodysplasia,<br />

and so on) is important because these diseases have the potential<br />

to affect bladder and sphincteric function. A history of medical diseases<br />

such as diabetes or congestive heart failure can cause LUTS by their<br />

effects on the lower urinary tract or fluid mobilization.<br />

For women with voiding <strong>dysfunction</strong>, obstetrical and gynecological<br />

history is extremely important. Pregnancy and childbirth, particularly<br />

vaginal delivery, are associated with voiding <strong>dysfunction</strong>, especially<br />

incontinence and pelvic prolapse. Thus, number of pregnancies, deliveries<br />

(including method, i.e., vaginal vs cesarean), and the onset of<br />

the symptoms in relation to these events is important. Symptoms can<br />

sometimes be related to a woman’s menstrual cycle and careful questioning<br />

in this regard should be done. A women’s hormone status (pre-,<br />

peri-, or postmenopausal) and the onset of symptoms with changes in<br />

status should be noted. Estrogen deficiency may cause or contribute


Chapter 2 / Diagnostic Evaluation 31<br />

to LUTS, including irritative symptoms of frequency and urgency, as<br />

well as incontinence. If the patient is postmenopausal, it is important<br />

to note whether the patient is being treated with hormone replacement<br />

therapy. Careful questioning regarding a history of endometriosis and<br />

gynecological malignancy should be performed.<br />

Prior surgery may have effects on lower urinary tract function. This<br />

includes surgery on the lower urinary tract (e.g., prostate surgery in<br />

men or incontinence surgery in women). Other pelvic surgery such as<br />

gynecological surgery or lower intestinal surgery also may affect the<br />

bladder directly or indirectly through damage to the nerve supply to<br />

the bladder or sphincter. For example, new onset total incontinence<br />

after hysterectomy raise the suspicion of a vesicovaginal fistula while<br />

urinary retention after an abdominal perineal resection of the rectum<br />

may be indicative of injury to the neural innervation of the bladder.<br />

History of pelvic radiation for treatment of pelvic malignancy (urologi-<br />

cal, gynecological, or rectal) is important as this can have a marked<br />

effect on lower urinary tract function and LUTS.<br />

PHYSICAL FXAMINATION<br />

A complete physical exam is important; however, certain aspects of<br />

the exam need to be emphasized. A focused physical examination<br />

should be performed to:<br />

1.<br />

2.<br />

3.<br />

4.<br />

5.<br />

6.<br />

Assess the bladder for masses and fullness;<br />

Assess the external genitalia;<br />

Assess the pelvic floor, including anal sphincter tone, and thoroughly<br />

examine for support defects, prolapse, and other pelvic conditions<br />

in women;<br />

Assess the prostate in men;<br />

Demonstrate incontinence in patients with that symptom;<br />

Detect neurologic abnor~alities that may contribute to voiding <strong>dysfunction</strong>.<br />

The abdominal exam, which includes examination of the flanks,<br />

begins with inspection for scars, masses, or hernias. Examination of<br />

the back should be performed to check for scars and scoliosis which may<br />

be an indication of potential spine abnormalities that may contribute to<br />

voiding <strong>dysfunction</strong>. Suprapubic palpation is performed to determine<br />

if the patient has a distended bladder or pelvic mass.<br />

In women, a systematic examination of the vagina and pelvis is<br />

important. This is first done in lithotomy position and may be repeated


32 Nitti and Ficazzola<br />

with the patient standing. The external genitalia are first inspected<br />

followed by evaluation of the vaginal mucosa for signs of atrophic<br />

vaginitis, indicating estrogen deficiency, previous surgery, and vaginal<br />

discharge. The urethral meatus should be observed and the urethra<br />

palpated for any abnormalities. The anterior vaginal compartment is<br />

examined next. This can be aided by applying slight pressure wall with<br />

the posterior blade of a small vaginal speculum. The position of the<br />

urethra, bladder neck, and bladder can be observed at rest and with<br />

straining to evaluate support of these structures and determine the<br />

presence of urethral hypermobility and cystocele. Also with coughing<br />

and straining, the urethra should be observed for urine loss and whether<br />

that loss occurs with hypermobility . In cases of stress incontinence,<br />

a Q-tip test may be performed to determine the degree of urethral<br />

hypermobility. This is done by placing a Q-tip inside the urethra with<br />

its tip at the urethrovesical junction and observing the degree of rotation<br />

with straining. Hypermobility is defined as a resting or straining angle<br />

greater than 30" from the horizontal. The central vaginal compartment<br />

is examined next. The uterus and cervix should be evaluated at rest<br />

and with straining to determine prolapse. Bimanual examination is<br />

done to evaluate the presence of uterine, adnexal, or other pelvic masses.<br />

If the patient has had a hysterectomy, the vaginal cuff should be assessed<br />

for enterocele. This is often best accomplished by first retracting the<br />

anterior vaginal wall and then the posterior wall. Finally the posterior<br />

vaginal compartment is examined by retracting the anterior vaginal<br />

wall with the speculum blade. A large rectocele is easily identifiable.<br />

Examination for rectocele can be aided by a simultaneous rectal exam,<br />

which will also assess perineal integrity, anal sphincter tone, bulbocav-<br />

ernosus reflex, and rectal prolapse. If it is difficult to distinguish between<br />

a high rectocele and enterocele, simultaneous rectal and vaginal exam<br />

with the index finger and thumb may palpate the enterocele sliding<br />

above the anterior rectal wall. The pelvic examination may be repeated<br />

with the patient standing with one foot on the floor and the other<br />

on a step. This may give a more realistic assessment of the degree<br />

of prolapse.<br />

Genital examination in men should include inspection of the urethral<br />

meatus and palpation of the penile and bulbar urethra. During rectal<br />

exam, anal sphincter tone and bulbocavernosus reflex are assessed, as<br />

is the size and consistency of the prostate.<br />

A focused neurologic exam may yield important information regard-<br />

ing voiding <strong>dysfunction</strong>. Symptoms related to voiding <strong>dysfunction</strong> may


Chapter 2 I Diagnostic Evaluation 33<br />

be among the earliest manifestation of nervous system disease (8). A<br />

neurologic exam should begin with an observation of the patient’s<br />

general appearance and gait. Lack of coordination, tremor, or facial<br />

asymmetry may be signs of neurologic disease. The upper and lower<br />

extremities are evaluated for gross motor coordination, strength, and<br />

sensation along dermatome distributions. The deep tendon reflexes at<br />

the knee and ankle should be assessed, as well as bulbocavernosus and<br />

perianal reflexes, including an evaluation of the muscle tone of the<br />

anal sphincter. A patient with a neurologic lesion at the sacral level<br />

may demonstrate diminished anal sphincter tone, perianal sensation,<br />

or absent bulbocavernosus reflex. Patients with a suprasacral spinal<br />

cord lesion will demonstrate spastic paralysis and hyperactive reflexes<br />

below the level of the lesion.<br />

LABOMTORY TESTING<br />

Urine analysis is part of the standard evaluation of the patient with<br />

LUTS and voiding <strong>dysfunction</strong>. Urinalysis can screen for pyuria, bacturia<br />

hematuria, and the presence of glucosuria or proteinuria. <strong>Voiding</strong><br />

<strong>dysfunction</strong> and LUTS can be associated with infection, malignancy,<br />

or medical illness such as diabetes, which can be discovered as a result<br />

of an abnormal urine analysis. When abnormalities are found on urine<br />

analysis, further testing may be warranted such as urine culture in cases<br />

of suspected infection or urine cytology, endoscopic, and radiographic<br />

studies when microscopic hematuria is present.<br />

Blood tests are useful in select cases of voiding <strong>dysfunction</strong>. The<br />

most common tests are those that evaluate renal function, e.g., serum<br />

blood urea nitrogen and creatinine, in cases where renal insufficiency<br />

is known or suspected.<br />

In select cases, more specific blood and urine testing may be performed,<br />

but these are usually dependent on patient history and physical<br />

as well as the results of simple tests.<br />

SIMPLE TESTS<br />

FOR EVALUATING VOIDING DYSFUNCTION<br />

When history and physical exam alone are insufficient to make a<br />

diagnosis or institute treatment, or when more objective information


34 Nitti and Ficazzola<br />

is desired, the clinician may start with simple tests to evaluate lower<br />

urinary tract function. These are noninvasive or minimally invasive<br />

(placement of a urethral catheter) tests that can provide information<br />

that may influence treatment or further diagnostic evaluation. The most<br />

basic of these include a voiding and intake diary, measurement of<br />

postvoid residual volume, uroflowmetry, and pad testing. Bedside or<br />

eyeball urodynamics is also a simple test and is described at the end<br />

of the section on urodynamics.<br />

<strong>Voiding</strong> and Intake Diary<br />

A diary is an extremely useful way to describe the nature and quantify<br />

the severity of symptoms such as frequency, nocturia, and incontinence.<br />

It also provides a baseline to assess future treatment. A voiding and<br />

intake diary should include the time, type and amount of fluid intake,<br />

the time and amount of each void, and any associated symptoms such<br />

as incontinence, extreme urgency or pain. The diary should be done<br />

for a period of 24 h, and several days are preferred. The patient should<br />

also note how representative a particular 24-h period was of his or her<br />

normal symptoms.<br />

Postvoid Residual<br />

The postvoid residual volume (PVR) is defined as the volume of urine<br />

remaining in the bladder immediately following voiding. It provides<br />

information on the ability of the bladder to empty as well as its functional<br />

capacity (voided volume plus PVR). Normal lower urinary tract function<br />

is usually associated with a negligible PVR; however, there is no<br />

agreement as to what an “abnormal” or clinically significant value is.<br />

Thus, elevated PVR is somewhat arbitrary. This being the case, most<br />

would agree that a PVR of greater than l00 mL is elevated, although<br />

perhaps not significant. Elevated PVR may be an indication of detrusor<br />

hypocontractility or bladder outlet obstruction and may prompt further<br />

evaluation depending on the patient and the symptoms or condition<br />

being evaluated. PVR can be measured directly by in and out urethral<br />

catherization or determined noninvasively by ultrasonography. Portable<br />

bladder scanners based on ultrasound technology are now available to<br />

determine bladder volume.


Chapter 2 / Diagnostic Evaluation 35<br />

'I<br />

I J<br />

i time to maximum tiow<br />

I<br />

Flow time<br />

maximum<br />

time I<br />

I<br />

I<br />

Fig. 1. Normal flow curve and pattern depicting the terminology of the International<br />

Continence Society relating to urodynarnic the description of urinary flow. Adapted<br />

with permission from ref. 8a.<br />

Uroflowmetry<br />

The determination of urinary flow rate over time, or uroflowmetry,<br />

is a simple way to measure bladder emptying. In and of itself, a uroflow<br />

is rarely able to determine the cause of voiding <strong>dysfunction</strong>; however, in<br />

conjunction with a careful history and physical, it can provide valuable<br />

information. In addition, it is extremely useful in selecting patients for<br />

more complex urodynmic testing. Uroflow is measured by a device<br />

called a uroflowmeter. Modern uroflowmeters consist of electronic<br />

collection equipment with graphic expression of the flow rate as a<br />

function of time.<br />

Common parameters determined by uroflowmetry include (see<br />

Fig. l):<br />

l. Voided volume: Actual volume of urine voided.<br />

2. Flow time: Time during which measurable flow occurs.<br />

3. Total voiding time: The total time of void taking into account periods<br />

of no flow in the patient with an int~~ittent pattern.<br />

4. Maximum flow rate (Qmax): The highest flow rate achieved during the<br />

voiding episode (i.e. the highest point on the curve).


36 Nitti and Ficazzola<br />

5. Time to maximal flow: The elapsed time from the beginning of voiding<br />

to the point of maximal flow. It is generally about one third of the<br />

total voided time.<br />

6. Mean flow rate (Qave): Voided volume divided by flow time. Only<br />

interpretable if flow is continuous and uninterrupted.<br />

There is considerable overlap in flow rates between normal and<br />

abnormal patients. Qmax seems to have a greater specificity than Qave in<br />

determining abnormal voiding (9) and has become the most commonly<br />

used uroflow parameter. Urinary flow rate and Qnlax varies as a function<br />

of patient age, sex, anxiety, and voided volume. Several nomograms<br />

have been established to classify Qmax as normal or abnormal based on<br />

Qmax and voided volume (10-12). Since Drach et al. (13) described a<br />

linear relationship between Qmax and voided volume above a voided<br />

volume of 150 mL, and a hyperbolic relationship between the two<br />

below this volume, voided volume of 150 mL has been commonly<br />

used as a minimum volume to make an accurate assessment of the<br />

uroflow. This is a major limitation in patients who do not routinely<br />

void with volumes 150 mL or greater.<br />

In addition to the parameters previously mentioned, the overall flow<br />

pattern is also important. Visual inspection of the uroflow curve can<br />

give valuable insight into the patients voiding <strong>dysfunction</strong>. A normal<br />

flow curve is a continuous, almost bell-shaped, smooth curve (Fig. 1).<br />

Interruptions or spikes and valleys in the flow may be indicative of<br />

voiding <strong>dysfunction</strong>. Intermittent flow pattern is characterized by one or<br />

more episodes of flow increasing and then decreasing and is commonly<br />

owing to abdominal straining (Fig. 2). A single uroflow measurement<br />

may not be representative for a given patient (14). Therefore, measuring<br />

multiple voiding events is a much more reliable indicator of the patient’s<br />

voiding pattern. In addition, to have a reliable uroflow measurement,<br />

the flow event must represent the patient’s usual voiding pattern. This<br />

can be simply assessed by asking the patient after completing the<br />

uroflow test if the void was representative of their typical pattern with<br />

respect to its volume and force.<br />

Abnormal flow rates and patterns are suggestive of voiding dysfunc-<br />

tion. They may be caused by a variety of conditions, including bladder<br />

outlet obstruction, impaired detrusor contractility, or learned voiding<br />

behaviors. Although an abnormal uroflow suggests a problem, it cannot<br />

differentiate between causes or make a definitive diagnosis (15). More<br />

comprehensive urodynamic testing is needed to make a precise diagnosis.<br />

Uroflow is more cominonly utilized in men as opposed to women,<br />

probably because of the relatively high incidence of bladder outlet


Chapter 2 I Diagnostic Evaluation 37<br />

..................................................<br />

125 ml& Flow Rate :<br />

...................................................................<br />

...................................................................<br />

....................................................................<br />

Results of UROFL<br />

..............................<br />

<strong>Voiding</strong> Time T100 65 S<br />

Flow Time TQ 42 S<br />

Time to max Flow TQnax 17 S<br />

Max Flow Rate Qmax 5.5 ml/s<br />

Ruerage Flow Rate Qave 2.8 ml/s<br />

Vo i ded Vo 1 ume Vcomp 118 m1<br />

Fig. 2. Abnormal flow pattern, showing interrupted stream and low Qmax<br />

(5 mL/s). Adapted with permission from ref. 9.<br />

obstruction and decreased flow in elderly men (16,17’). Again it must<br />

be emphasized that uroflow alone cannot make a diagnosis of obstruction.<br />

In males, Qtnm and voided volume normally decrease with age.<br />

Giman et al. (18) found that in a group of asymptomatic men of ages<br />

40-79 yr, Qmax dropped from a median of 20.3 mL/s (40-44 yr) to 11.5<br />

mL/s (75-79 yr), and voided volume decreased from a median of 356<br />

to 223 mL over this same time period. Uroflow is not as frequently<br />

used for women with voiding <strong>dysfunction</strong>. Uroflow tends to be higher<br />

in normal women than it is in age-matched normal men (19). This is<br />

owing to the fact that the female urethra is short and only the voluntary<br />

part of the external urethral sphincter seems to provide any sort of<br />

outlet resistance during voiding. In women, noma1 Qmax ranges between<br />

20-36 mL/s with a bell-shaped curve and short flow time (20).<br />

Pad Test<br />

For patients who have the symptom of incontinence, the pad test is<br />

a method that can be used to quantify urine loss based on the measurement<br />

of weight gain of absorbent pads over a period of time. Several


38 Nitti and Ficazzola<br />

different pad tests have been described, using variable periods of time<br />

(20 min to 48 h) and conditions (exercises, daily activity) (21-24).<br />

Shorter tests are usually performed with a standard bladder volume. The<br />

pad test provides a quantification of urine loss, but cannot distinguish<br />

between causes of incontinence. Pad testing is commonly used in<br />

research and clinical trials on incontinence treatments. It may also be<br />

useful in clinical practice to assess severity of incontinence and response<br />

to treatment.<br />

Urodynamics is the study of the transport, storage, and evacuation<br />

of urine by the urinary tract. It is comprised of a number of tests that<br />

individually or collectively can be used to gain information about lower<br />

urinary tract function and can provide a precise diagnosis of the etiology<br />

of voiding <strong>dysfunction</strong>. However, in order to use urodynamics in a<br />

practical and effective way, it is important for the clinician to how<br />

when and why a urodynamic investigation should be performed. There<br />

are situations in which the information gained from history and physical<br />

exam, and simple tests like uroflow and postvoid residual determination,<br />

is sufficient to make a clinical decision. In other cases, it is necessary<br />

to monitor physiological parameters in a more comprehensive and<br />

precise way. For example, sometimes a limited evaluation cannot adequately<br />

or accurately explain a patient’s symptoms, or a patient may<br />

have a medical condition known to affect the urinary tract in potentially<br />

serious and/or unpredictable ways (e.g., neurological disease).<br />

Once the decision has been made to perform urodynamics on a<br />

particular patient, it is important to consider what information is<br />

expected from the test (25). The simple fact that a patient has symptoms<br />

or a disorder that may affect the lower urinary tract is not sufficient<br />

to start the urodynamic evaluation. A list of problems or questions that<br />

should be solved or answered by urodynamics should be made before<br />

any testing is performed. All patients are not alike and therefore each<br />

urodynamic evaluation may be different depending on the information<br />

needed to answer the questions relevant to a particular patient. We follow<br />

three important rules before starting a urodynamic evaluation (25):<br />

1. Decide on questions to be answered before starting a study.<br />

2. Design the study to answer these questions.<br />

3. Customize the study as necessary.


Chapter 2 / Diagnostic Evaluation 39<br />

By following these simple rules, one can maximize the chance of<br />

obtaining useful information from a study. It is also critical to have<br />

an understanding of the possible causes of a patient’s symptoms and the<br />

urodynamic manifestations of a pre-existing condition. The functional<br />

classification previously described is particularly useful in this regard,<br />

because urodynamic testing easily classifies problems into failure to<br />

empty or failure to store and further into bladder <strong>dysfunction</strong> and/or<br />

bladder outlet <strong>dysfunction</strong>.<br />

Midtichannel Urodynamics<br />

The various components of the urodynamic evaluation include moni-<br />

toring of bladder pressure during filling (cystometrogram), monitoring<br />

of bladder pressure and simultaneous urinary flow rate during voiding<br />

(voiding pressure Bow study), and monitoring of pelvic Boor and exter-<br />

nal sphincter activity (electromyography). Usually abdominal pressure<br />

is also monitored during filling and voiding so that subtracted detrusor<br />

pressure can be determined (see Cystometrogram). In select cases,<br />

the urethral pressure can also be assessed during storage and voiding<br />

(urethral pressure profilometry). It is when these components are com-<br />

bined together as “multichannel urodynamics” that a most sophisticated<br />

study of the lower urinary tract is obtained.<br />

CYSTOMETROGRAM<br />

The cystometrogram (CMG) is a measure of the bladder’s response<br />

to being filled. Normally the bladder should store increasing volumes<br />

of urine at low pressure and without involuntary contractions. CMG<br />

determines the pressure-volume relationship within the bladder and<br />

also provides a subjective measure of bladder sensation with the cooper-<br />

ation of the patient. Ideally, the CMG should mimic noma1 bladder<br />

filling and gives an accurate assessment of true bladder function.<br />

CMG is performed by filling the bladder at a constant rate (usually<br />

10-100 mL/s) with fluid (normal saline or contrast media) or gas (such<br />

as carbon dioxide). Filling occurs via a catheter, which is inserted<br />

transurethrally or suprapubically. Usually there are two lumens on the<br />

catheter, one to measure pressure and one to fill the bladder. Most<br />

urodynamisists have abandoned gas cystometry because fluid is more<br />

physiologic and allows the determination more parameters. The reader<br />

is referred elsewhere for a more detailed description of the technical<br />

aspects of cystometry (26).


40 Nitti and Ficazzola<br />

Filling/storage phase<br />

Volume<br />

Fig. 3. Idealized normal adult cystometrogram. Phases I, 11, and I11 occur during<br />

filling and phase IV during voiding. Phase I reflects the bladder’s initial response<br />

to filling. Phase 11 is the tonus limb and reflects bladder pressure during the<br />

majority of the filling phase. As the vesicoelastic properties of the bladder reach<br />

their limit phase, IT1 is entered where pressures begin to increase just prior to<br />

phase IV, the voluntary contraction phase. (Clinically the pressure change from<br />

phases T to I1 is often imperceptible and-depending on how much the bladder<br />

is allowed to fill-phase 111 may not be appreciated as a significant pressure<br />

change.) Adapted with permission from ref. 8a.<br />

Several parameters may be evaluated by CMG:<br />

1. Sensation.<br />

2. Filling pressure.<br />

3. Presence of involuntary or unstable contractions.<br />

4.. Compliance (filling pressure in relation to volume).<br />

5. Capacity.<br />

6. Control over micturition.<br />

Sensation is the part of cystometry that is truly subjective and there-<br />

fore requires an alert and attentive patient and clinician. It is difficult<br />

to quantify sensation, which can vary greatly under different conditions.<br />

It may be affected by factors such as infusion rate and temperature of<br />

the filling solution as well as the level of distraction of the patient.<br />

The true value of the CMG with respect to sensation is when a symptom<br />

is mimicked and sensation can be correlated with changes in vesical<br />

pressure.<br />

Normally as the bladder fills, it maintains a relatively constant and<br />

low pressure, usually not exceeding 5-10 cm H20 above starting or<br />

baseline pressure (Fig. 3). It will store increasing volumes of urine<br />

without a significant rise in pressure. The only time the bladder should


Chapter 2 / Diagnostic Evaluation. 41<br />

contract is during the voluntary act of voiding. However, under certain<br />

circumstances, a bladder may contract involuntarily. Involuntary con-<br />

tractions may be associated with the symptoms of frequency, urgency,<br />

urge incontinence, pain, or the perception of a normal desire to void.<br />

The International Continence Society (ICs) has classified involuntary<br />

detrusor contractions into two broad categories (27):<br />

1. Detrusor instability: involuntary contractions not associated with an<br />

underlying neurologic lesion. Detrusor instability may be caused by<br />

inflammation, infection, bladder outlet obstruction, aging, so and forth,<br />

or in many cases is “idiopathic.” Activity may be spontaneous or<br />

provoked by certain movements, position, etc. ICs originally required<br />

a contraction of at least 15 cm H20 for the classification of unstable<br />

bladder. However, we feel Comfortable classifying any involuntary<br />

rise in detrusor pressure that is accompanied by an urge as detrusor<br />

instability.<br />

2. Detrusor hyperreflexia: involuntary contractions associated with a<br />

known neurologic lesion. Typically, these are upper motor neuron<br />

lesions such as supra sacral spinal cord lesions, multiple sclerosis, or<br />

cerebrovascular accident.<br />

Detrusor instability and detrusor hyperreflexia may look identical<br />

on CMG (Fig. 4). However, instability is more likely to be phasic<br />

and of lower pressure and can more often be inhibited. Hyperreflexic<br />

contractions tend to be more numerous and of higher pressure. These<br />

are only generalizations, however, and the terms instability and hyperreflexia<br />

are strictly defined by the patient’s neurological status and not<br />

the CMG appearance of the involuntary contraction.<br />

Compliance refers to the change in volume/change in pressure and<br />

is expressed in mL/cm H20. Since the bladder is normally able to hold<br />

large volumes at low pressures it is said to be highly compliant. The<br />

spherical shape of the bladder as well as the viscoelastic properties of<br />

its components contribute to its excellent compliance allowing storage<br />

of progressive volumes of urine at low pressure. When the bladder is<br />

unable to maintain low pressure with increasing volume, compliance<br />

is decreased or impaired. This may occur when there is prolonged<br />

bladder outlet obstruction as the “overworked bladder” becomes scarred<br />

as smooth muscle is replaced by collagen. It can also result from<br />

radiation, surgery on the bladder, tuberculosis, and other chronic infectious<br />

diseases, as well as other conditions that affect the size, shape,<br />

and vesicoelastic properties of the normal bladder. The measurement<br />

of compliance is important because high prolonged elevated storage<br />

pressures have been shown to have a deleterious effect on the kidneys.


42 Nitti and Ficazzola<br />

pressure<br />

cm water<br />

Involuntary contractions<br />

PHASIC INSTABILITY<br />

Detrusor instability<br />

or<br />

Destrusor hyperreflexia<br />

Volume<br />

ml<br />

Fig. 4. Involuntary detrusor contractions. Note the multiple involuntary rises in<br />

bladder pressure during filling. This could be an example of detrusor instability<br />

or hyperreflexia and that distinction would be determined by the absence or<br />

presence of a neurological lesion.<br />

McGuire and associates have shown that in neurogenic bladder<br />

sustained pressures of 40 cmH20 or greater during storage can lead to<br />

upper tract damage (28). In practical terms, the concept of compliance<br />

is actually more useful than its absolute value. The calculated value<br />

of compliance (in cmH20) is probably less important than the absolute<br />

bladder pressure during filling (Fig. 5).<br />

When bladder pressure is monitored through a transurethral or suprapubic<br />

catheter, the pressure that is recorded is actually the sum of intraabdominal<br />

pressure and the pressure generated by the detrusor itself,<br />

either through a contraction or wall tension with bladder filling, i.e.,<br />

compliance. Cystometry can be performed as a single channel study<br />

where the bladder pressure (Pves) is measured and recorded during<br />

filling and storage or as a multi-channel study where abdominal pressure<br />

(Pabd) is subtracted from Pves to give the detrusor pressure (Pdet).<br />

Abdominal pressure can be recorded via a small balloon catheter that


Chapter 2 / Diagnostic Evaluation 43<br />

Fig. 5. Impaired compliance. This CMG shows a gradual loss of compliance<br />

between 40 and 280 m1 with filling pressures of 17 cm H20 at 280 mL. Between<br />

280 m1 and 330 mL there is a rapid rise in pressure to 41 cm H20 which continues<br />

to rise to 51 crnH20 until filling is stopped at 400 mL. The storage pressures<br />

occurring from 300 m1 on clearly present potential danger to the kidneys, Adapted<br />

with permission from ref. 26.<br />

is place~d in the rectum or vagina. The ability to calculate subtracted<br />

detrusor pressure allows one to distinguish between a true rise in<br />

detrusor pressure and the effect of increased abdominal pressure (e.g.,<br />

straining, Valsalva) (Fig. 6).<br />

VOIDING PRESSURE FLOW STUDY<br />

Cystometry assesses the bladder's response to filling, however, by<br />

itself tells nothing about the bladder's ability to empty. This can be<br />

determined by allowing a patient to void (voluntarily or involuntarily)<br />

during bladder pressure monitoring. When the simultaneous measure-<br />

ment of uroflow is added, i.e., voiding pressure-flow study, detrusor<br />

contractility as well as the resistance of the bladder outlet can be<br />

determined (Fig. 7). In fact, detrusor pressure during voiding is actually<br />

determined by the amount of outlet resistance. The more resistance,


44 Nitti and Ficazzola<br />

a<br />

B<br />

. .<br />

- loo<br />

- 100<br />

- rm 0<br />

-100 0<br />

0<br />

WeS<br />

cm HZO<br />

-5<br />

.. .<br />

cm H20<br />

P-bd<br />

ern H20<br />

Wet<br />

em H20<br />

Fig. 6. Adding intra-abdominal pressure monitoring gives a better representation<br />

of the true detrusor pressure. (A) Single channel study showing multiple spikes<br />

and rises in vesical pressure (Pves). Without having simultaneous monitoring of<br />

intra-abdominal pressure, it is impossible to tell if these pressure spikes are due<br />

to a rise in detrusor or abdominal pressure. (B) Same tracing with intra-abdominal<br />

pressure monitoring added. It can now be determined that the changes in Pves<br />

are due to the changes in abdominal pressure (Pabd). When one looks at the<br />

subtracted detrusor pressure curve (Pdet), it is flat without any significant rises.<br />

Adapted with permission from ref. 26.<br />

the more forceful the bladder will contract (29). Since any given bladder<br />

has a set bladder outlet relation, the higher the pressure, the lower the<br />

flow. Thus obstruction can be defined as relatively high pressure and<br />

low flow. In cases where contractility is impaired, the same detrusor<br />

pressure will result in a lower flow.<br />

<strong>Voiding</strong> pressure flow studies are extremely important in evaluating<br />

lower urinary tract symptoms and voiding <strong>dysfunction</strong>. For example,<br />

bladder outlet obstruction and/or impaired detrusor contractility may be<br />

directly responsible for symptoms (e.g., decreased uroflow, incomplete<br />

emptying); abnormal voiding can affect the bladder’s ability to store<br />

and result in an abnormal cystometrogram (e.g., detrusor instability or<br />

decreased compliance caused by bladder outlet obstruction).<br />

In general, one can diagnose bladder outlet obstruction when high<br />

pressure and low flow are seen on a voiding pressure flow study.


Chapter 2 / Diagnostic Evaluation 45<br />

I l l I I I I I i l l 1 1 I I I I I I I I I I<br />

. . . . . . . . . . . . . . . . . . . . . . . . . .<br />

"W<br />

. , . : , . . , . : . * , . . : * . * * . : , . . . . . flow<br />

. . . . . . . . . . . . . . . . . . . . . . . . . ' . * E.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . *t<br />

Fig. 7. Normal voiding pressure-flow study in a male. During voluntary voiding<br />

on the right portion of the tracing, there is a rise in Pdet associated with flow.<br />

There is a slight decrease in Pabd during the second half of voiding, with Pabd<br />

returning to just below baseline. The uroflow curve is relatively normal with<br />

Qmax = 22 mL/s. Detrusor pressure during voiding was about 30 cm H20.<br />

Sometimes this is obvious (Fig. S), but other times a more sophisticated<br />

analysis of the pressure flow relationship is needed. In men, several such<br />

analyses or nomograms have been described to diagnose obstruction and<br />

impaired contractility (Figs. 9 and 10) (16,30).<br />

LEAK-POINT PRESSXJRE<br />

Leak-point pressures are used to determine the pressure at which<br />

involuntary loss of urine occurs. There are two types of leak-point<br />

pressures that can be determined and each represents the measurement<br />

of a completely different parameter and concept (31):<br />

1. Abdominal leak-point pressure (ALPP). This pressure, also known as<br />

the Valsalva leak-point pressure, is the amount of abdominal pressure<br />

required to cause leakage of urine, in the absence of a rise in Pdet. It<br />

is a measure of the resistance of the sphincter to rises in Pabd and is


46 Nitti and Ficazzoia<br />

6:40 loa0 UDS-120<br />

425min<br />

t i l l I I I I I I I I I I I I I I I I I I 1 1 i 1 1<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . ’1% . flaw<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . - 0<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . - 8-<br />

. . . .<br />

Fig. 8. Classic bladder outlet obstruction seen as high pressure-low flow voiding.<br />

Pves, intravesical pressure; Pabd, abdominal pressure; Pdet, detrusor pressure.<br />

Adapted with permission from ref. 29a.<br />

commonly used in evaluating patients with stress incontinence (Fig.<br />

l l). There is no ‘“nomal” ALPP, because there should not be leakage<br />

when any physiologic abdominal pressure is generated. Very low leak<br />

point pressures are suggestive of intrinsic urethral <strong>dysfunction</strong>.<br />

2. Bladder leak-point pressure (BLPP). This pressure, also known as<br />

detrusor leak-point pressure, is the amount of detrusor pressure require<br />

to cause leakage in the absence of a rise in Pabd (32). It is a measure<br />

of the resistance of the sphincter to rises in Pdet. BLPP is useful in<br />

cases of detrusor <strong>dysfunction</strong>, where storage pressures are elevated,<br />

usually in cases of impaired compliance (Fig. 12). The higher the<br />

BLPP, the more resistance is provided by the sphincter and the more<br />

likely is upper tract damage. BLPP that is 40 cm W20 or greater during<br />

storage can lead to upper tract damage (28).<br />

ELECTROMYOGRAPHY<br />

The storage and emptying phases of the micturition cycle are affected<br />

by the perineal musculature including the striated external urethral


Chapter 2 I Diagnostic Evaluation 47<br />

200<br />

Pdet<br />

Unobstructed<br />

0 t I<br />

0 Flow 25<br />

Fig. 9. The Abrams-Griffiths Nomogram. Maximum flow rate and the conesponding<br />

detrusor pressure at maximum flow are plotted against each other, with detrusor<br />

pressure in cm H20 on the y-axis and uroflow rate in mL/s on the x-axis. Data<br />

from 117 males (>55 years) evaluated for possible prostatic obstruction were used<br />

to create zones on their graph representing obstructed, unobstructed and equivoc<br />

micturition. The location of the plotted maximum flow point on the graph determines<br />

the presence or absence of obstruction or an equivocal state. (In this case<br />

the patient would be obstructed.)<br />

sphincter. Sphincter activity can be measured<br />

during urodynamic testing<br />

either by surface electrodes (similar to those used for electrocardiogram)<br />

or by inserting needle electrodes directly into the sphincter muscle<br />

(33). When surface electrodes are used, it is actually the activity of<br />

the anal sphincter that is measured, but it is extremely rare for anal<br />

sphincter activity to be dissociated from urethral sphincter activity.<br />

Normally, external sphincter activity will increase as bladder pressure<br />

rises (e.g., with detrusor instability) as a way to protect against inconti-<br />

nence. Conversely, the external sphincter should relax during voiding<br />

(Fig. 13). In fact, normally, the external sphincter relaxes before a<br />

bladder contraction is initiated. Failure of the external sphincter to<br />

relax during voiding can result in lower urinary tract symptoms and<br />

<strong>dysfunction</strong>. This can be a learned behavior (<strong>dysfunction</strong>al voiding) or<br />

can occur involuntarily as a result of neurological disease (detrusor-<br />

external sphincter-dyssynergia). Either can result in significant voiding


48 Nitti and Ficazzola<br />

20 30 40 50<br />

pre 17/18 TURP 19/20 post<br />

Grade: IllN -> OIN<br />

Type: corn pro3 I v e<br />

Fig. 10. Linear PURR nomogram. This nomogram developed by Schafer allows<br />

for the determination of obstruction independent of contractility. The degree of<br />

obstruction is determined by 7 zones on the pressure-flow diagram labeled 0 to VI<br />

corresponding to increasing grades of obstruction: grade 0 and I, no obstruction;<br />

grade 11, equivocal or mild obstruction; grade I11 to VI, increasing severity of obstruc-<br />

tion. The boundary between grades 2 and 3 corresponds to the boundary between<br />

equivocal and obstructed in Abranis-Griffiths nomogram. Contractility zones are<br />

also seen: ST, strong; N, normal; W, weak. Adapted with permission from ref. 30.<br />

<strong>dysfunction</strong> as outlet resistance during voiding is increased creating a<br />

functional obstruction.<br />

VIDEOURODYNAMICS<br />

In certain cases, multichannel urodynamic testing is unable to provide<br />

a precise diagnosis. In such cases videourodynamics may be necessary.<br />

Videourodynamics refers to the simultaneous measurement and display<br />

of urodynamic parameters with radiographic visualization of the lower<br />

urinary tract (34,35). In these cases, the bladder is filled with radio-<br />

graphic contrast filling during urodynamics. Because all urodynamic<br />

parameters previously mentioned are visualized simultaneously with<br />

the radiographic appearance of the lower urinary tract, the clinician<br />

can better appreciate their interrelationships and recognize artifacts.<br />

Videourodynamics is the most precise way to evaluate lower urinary


Chapter 2 / Diagnostic Evaluation 49<br />

................................<br />

UDS-l20<br />

J I I I I I I I ~ I I I I I I I I I 1 I I I I ~ I I I I I I I ~ ~<br />

*;til ' 1Iw<br />

l<br />

~ 1 I<br />

................................<br />

....................................<br />

*)it ' ' '190<br />

.............................. M<br />

.......................<br />

I<br />

0<br />

an H20<br />

Fig. 11. Urodynarnic study of a 60-year-old female with stress incontinence. The<br />

multiple spikes in Pves and Pabd represent coughing and straining. The point<br />

where leakage was demonstrated, marked by the vertical line, was at a Pves (and<br />

Pabd) of 11 1 cm HzO. This is the abdominal leak point pressure (ALPP) which<br />

is usually measured on the Pabd curve.<br />

tract function and disturbances in micturition. In cases of complex<br />

voiding <strong>dysfunction</strong>, videourodynamics can be invaluable (Fig. 14).<br />

BEDSIDE CYSTOMETRICS<br />

In cases where very basic urodynamic information such as sensation,<br />

capacity, and presence of detrusor contractions (voluntary or involuntary)<br />

is desired, a very simple CMG, known as bedside or eyeball<br />

urodynamics, can be performed without any special urodynamic equipment.<br />

The urodynamic information provided, although limited, is sometimes<br />

enough to aid in malung a differential diagnosis (36). Bedside<br />

cystometrics are useful when sophisticated equipment is unavailable<br />

or when standard urodynamic testing would produce undue discomfort<br />

for a particular patient. Simple CMG can be done with a urethral<br />

catheter, catheter tip syringe, and filling solution (saline or water). Prior<br />

to starting the patient is asked to void, if possible. A standard straight


50 Nitti and Ficazzola<br />

j : I<br />

1:<br />

l I l t l l l l l l I I I I l I l I I l l l l l l l Y l l l l l l l l t<br />

. 45 '<br />

............<br />

....................<br />

....................... ....<br />

BL<br />

t<br />

*<br />

* '<br />

I<br />

I "<br />

I<br />

' * " " '<br />

* I<br />

" ' " " ' '<br />

.......................... .....<br />

.................... ...'.!...'.....<br />

. f<br />

. I<br />

.......................... I . . . . . . . . .<br />

l * . p"'.<br />

UDS-120<br />

PIbd<br />

Fig. 12, Urodynamic study of a child with myelomeningocele and incontinence.<br />

The patient bas poorly a compliant bladder with Pdet rising during filling. Leakage<br />

was demonstrated at a Pdet of 38 cm H20, marked by the vertical line. This is<br />

the bladder leak-point pressure (BLPP).<br />

catheter 14-18 French is introduced into the bladder and postvoid<br />

residual is measured. A Foley catheter may be used if occlusion of the<br />

bladder neck is necessary or if one is already indwelling. A catheter<br />

tip syringe (usually 60 mL) is connected to the end of the catheter<br />

after its plunger is removed. The catheter and attached syringe are held<br />

directly upright and the top of the syringe is usually located about<br />

15-20 cm above the patient's bladder (Fig. 15). The bladder is then<br />

slowly filled by gravity at 50-m.L increments as water or saline is<br />

poured into the syringe. The catheter and syringe should be kept as<br />

low as possible to allow for a steady infusion. Sensations and the<br />

volumes at which they occur can be noted. If it is necessary to raise<br />

the syringe to maintain flow rate, this usually means that intravesical<br />

pressure is increasing (contraction, poor compliance, or abdominal<br />

straining). When a sudden rise in pressure occurs, fluid may be seen<br />

backing up into the syringe. This is usually indicative of an involuntary<br />

contraction and may be accompanied by an urge; however, an abrupt<br />

decrease in compliance or abdominal straining can also do this. It is


Chapter 2 / Diagnostic Evaluation 51<br />

Fig. 13. Normal electromyography of the pelvic floor during a pressure flow study.<br />

Note the voluntary increase in ENIG activity during the involuntary detrusor<br />

contraction and the decrease in activity during voluntary voiding. Adapted with<br />

permission from ref. 33.<br />

helpful to gently place a hand on, the patient’s abdomen and careful<br />

observe for such movement. When it is important to assess voluntary<br />

contractility, the catheter can be raised to its maximum height and with<br />

the syringe empty, the patient is asked to void. If the top of the syringe<br />

is 25 cm above the bladder and fluid is raised to that level, then at<br />

least 25 cmH20 pressure was generated. If fluid is not seen backing<br />

up into the syringe, the catheter is slowly lowered to see if there is<br />

any increase in pressure. Again, a hand should be kept on the abdomen<br />

to rule out, as best as is possible, abdominal straining. Obviously, when<br />

a detrusor contraction coexists with abdominal straining, it is extremely<br />

difficult to sort out the two.<br />

ENDOSCOPY<br />

Cystourethroscopy has been used in the evaluation of patients with<br />

lower urinary tract symptoms and voiding <strong>dysfunction</strong> in order to assess<br />

the bladder, the urethra, and prostate and look for extraurethral causes<br />

of incontinence.


52 Nitti<br />

Ficazzola<br />

Fig. 14. Videourodynamic evaluation of a woman in retention with primary bladder<br />

neck obstruction. Figure shows urodynamic tracing (top line is vesical pressure,<br />

second is abdominal pressure, third is detrusor pressure and forth flow) is as well<br />

as fluoroscopic image during voiding. There is a sustained detrusor contraction<br />

of >40 cm H20 with no opening of the bladder neck and flow. no Without imaging,<br />

the point of obstruction could not be localized. Adapted with permission from<br />

ref. 35.<br />

Evaluation of the Bhdder<br />

Cystourethroscopy to assess the bladder has been used as a routine<br />

part of the evaluation of males and females with LUTS and voiding<br />

<strong>dysfunction</strong>) to rule out intravesical pathology as a potential source of<br />

symptoms (37). The bladder is examined for an intravesical lesion or<br />

process that may be causing symptoms of instability, irritability, or<br />

incontinence. However a thorough review of the literature would sug-<br />

gest that, in routine cases, cystourethroscopy is optional and has its<br />

greatest value in select cases (38,39). There has been particular concern<br />

that irritative symptoms may be a warning of intravesical pathology<br />

such as carcinoma of the bladder. However there is no direct diagnostic<br />

value of cystourethroscopy in a patient with an unstable bladder unless<br />

microscopic hematuria is present (40,41). This would support the rou-<br />

tine use of cystourethroscopy for patients with irritative symptoms<br />

when clinical suspicion of an intravesical lesion is high, such as when<br />

hematuria is present or when other tests such as radiography suggest


Chapter 2 / Diagnostic Evaluation 53<br />

Fig. 15. Bedside or eyeball urodymanics. A urethral catheter is placed in the<br />

bladder and raised vertically. A syringe (without its plunger) is connected to the<br />

end of the catheter and the bladder filled. is Adapted with permission from ref. 26.<br />

intravesical pathology. However, cystourethroscopy is extremely help-<br />

ful and recomended in cases where symptoms cannot be explained<br />

by other more routine diagnostic testing or in patients who have failed<br />

to respond to seemingly appropriate treatment. Cystourethroscopy<br />

should be part of routine follow-up for patients whose voiding dysfunc-<br />

tion is managed by a chronic indwelling catheter or bladder augmen-<br />

tation or substitution.<br />

Evaluation of the Fmale Bhdder Outlet<br />

Endoscopic evaluation of the female bladder outlet in cases of stress<br />

incontinence has been advocated by several authors (433). However,<br />

others have found it to be inferior to other tests such as urodynamics<br />

(44-46). Most experts would now agree that cystourethroscopy is inadequate<br />

to judge the functional integrity of the bladder outlet and will<br />

underestimate the presence of intrinsic sphincter deficiency. Endoscopic<br />

evaluation of the female bladder outlet is also of little value in the<br />

uncomplicated patient with urge incontinence (47). In select women,<br />

cystourethroscopy to evaluate the bladder outlet can be extremely helpful;<br />

for example, in cases of urethral trauma or foreshortening, previous


54 Nitti and Ficazzola<br />

anti-incontinence surgery, suspected intraurethral pathology (e.g., ure-<br />

thral diverticulum), or bladder outlet obstruction. When evaluating the<br />

female urethra, it is extremely important to use an endoscope, which<br />

allows for complete visualization of the urethra. A rigid scope with no<br />

beak (or a short beak) to allow distension of the urethra, and a narrow<br />

lens (0-30”) or a flexible scope are optimal.<br />

Evaluation of the Male Bhdder Outlet<br />

Lower urinary tract symptoms in men were previously thought to<br />

be caused by benign prostatic hyperplasia. However, it is now known<br />

that a variety of conditions can cause such symptoms, including bladder<br />

outlet obstruction, detrusor instability, and impaired detrusor contractility.<br />

Cystoscopic evaluation of men with LUTS used to be routine;<br />

however, now its indications are more limited. Based on the available<br />

evidence and world literature, The World Health Organization Third<br />

International Consultation on BPH made the following recommendation:<br />

“Diagnostic endoscopy of the lower urinary tract is an optional test<br />

in the standard patient with symptoms of prostatism because<br />

-the outcomes of intervention are unknown,<br />

-the benefits do not outweigh the harms of the invasive study, and<br />

-the patients’ preferences are expected to be divided.<br />

However, endoscopy is recommended as a guideline at the time of<br />

surgical treatment to rule out other pathology and to assess the<br />

shape and size of the prostate which may have an impact on the<br />

treatment modality chosen. ” (48)<br />

Cystourethroscopy has a more definitive role in men who have<br />

undergone surgical treatment of the prostatic for benign or malignant<br />

disease when anatomic causes of postoperative voiding <strong>dysfunction</strong><br />

are suspected (e.g., bladder neck contracture or anastomotic stricture).<br />

Extmurethrul Incontinence<br />

Endoscopy can be an<br />

invaluable tool in the diagnosis and treatment<br />

of extraurethral incontinence due to vesicovaginal fistula and ectopic<br />

ureter. Cystourethroscopy can precisely localize a fistula site in the<br />

bladder and help plan surgical correction. Occasionally, a small fistula


Chapter 2 / Diagnostic Evaluation 55<br />

that is not seen on physical exam or by radiographic studies can be<br />

diagnosed only by cystoscopy. Incontinence owing to ectopic ureter<br />

in the female is usually diagnosed by radiographic studies. However,<br />

the exact location of the ureteral orifice in the urethra or vagina can<br />

be identified by cystourethroscopy and/or vaginoscopy. This can be<br />

extremely helpful in the planning of corrective surgery.<br />

URINARY TRACT IMAGING<br />

In certain cases of voiding <strong>dysfunction</strong>, imaging studies, including<br />

radiography, ultrasonography, magnetic resonance, and nuclear scan-<br />

ning, are an important part of the evaluation. Specifically, when detri-<br />

mental effects on the upper urinary tract or anatomical abnormalities<br />

of the upper and lower urinary tract are suspected, such studies can<br />

be useful. We will limit our discussion to imaging of the upper and<br />

lower urinary tract; however, there are cases where a urologic work-<br />

up of voiding <strong>dysfunction</strong> may prompt radiographic investigation of<br />

the nervous system or spine (e.g., in cases of suspected neurogenic<br />

voiding <strong>dysfunction</strong>).<br />

Upper Urinary Tract Imaging<br />

As mentioned previously, voiding <strong>dysfunction</strong> can be<br />

a cause of renal<br />

damage or deterioration of function. In 1981, McCuire and colleagues<br />

popularized the now universally accepted concept of the relationship<br />

between high bladder storage pressure and renal deterioration (28).<br />

Thus, in cases of known or suspected high storage pressures (e.g.,<br />

neurogenic voiding <strong>dysfunction</strong>) or elevated postvoid residual upper<br />

urinary tract, imaging is an important and necessary part of the evalua-<br />

tion. Also in certain patients with incontinence, an extra-urethral cause<br />

may be suspected (e.g., ectopic ureter or urinary tract fistula).<br />

The upper tract imaging modalities most commonly used are intra-<br />

venous urography (IVIJ), ultrasonography, computerized tomography<br />

(CT scan), magnetic resonance imaging (MRI), and isotope scanning.<br />

The usefulness of each individual test often depends on local exper-<br />

tise.<br />

Intravenous urography (IVU) is a standard radiographic examination<br />

of the upper urinary tract. Successful examination is dependent upon<br />

adequate renal function. Renal <strong>dysfunction</strong>, obstruction, congenital


56 Nitti and Ficazzola<br />

anomalies, fistula, stones, and tumors may be detected. IVU is the<br />

appropriate first study when ureteral ectopia is suspected. It is also an<br />

appropriate first imaging study when ureterovaginal fistula is suspected<br />

(49,50). Confirmation of the fistula, its size, and its exact location<br />

are often obtained with retrograde ureteropyelography. IVU can also<br />

be used to evaluate hydroureteronephrosis; however, ultrasound has<br />

become the procedure of choice for this because it is more cost-effective<br />

and does not expose the patient to intravenous contrast (see below).<br />

Ultrasonography is an excellent tool for imaging of the upper urinary<br />

tracts. It is totally noninvasive, and successful imaging of the kidneys<br />

is independent of renal function. Ultrasound can be used to assess many<br />

features of renal anatomy including renal size and growth, hydronephrosis,<br />

segmental anomalies, stones, and tumors. In the evaluation of the<br />

patient with lower urinary tract <strong>dysfunction</strong>, the detection of hydronephrosis<br />

is extremely important and may be an indication of vesicoureteral<br />

reflux or obstruction. However, no correlation exists between the<br />

degree of dilatation and the severity of obstruction. Also renal blood<br />

flow can be detected by the Doppler technique. Ultrasound is an excellent<br />

tool to follow the degree of hydronephrosis over time or in response<br />

to treatment.<br />

CT scanning can also provide much useful information about the<br />

anatomy of the upper urinary tract. Information can be independent of<br />

renal function; however, the addition of intravenous contrast can highlight<br />

specific anatomic characteristics (dependent on renal function).<br />

In most cases of voiding <strong>dysfunction</strong>, adequate information about renal<br />

anatomy can be obtained with ultrasonography; however, there may<br />

be select cases where CT is beneficial. MRI offers some of the same<br />

benefits as CT in the evaluation of the upper urinary tracts. It has the<br />

advantage over CT in that all planes of imaging are possible. As<br />

technology advances, MRI may play an increasing role in the evaluation<br />

of hydronephrosis and urinary tract anomalies in the future.<br />

When it is necessary to investigate functional characteristics of the<br />

upper urinary tract, nuclear isotope scanning is useful. Renography is<br />

used to examine the differential function of the two kidneys as well<br />

as how they drain. There are many physiological factors and technical<br />

pitfalls that can influence the outcome, including the choice of radionucleotide,<br />

timing of diuretic injection, state of hydration and diuresis,<br />

fullness or back pressure from the bladder, variable renal function, and<br />

compliance of the collecting system (51,52). Diuresis renography with<br />

bladder drainage may be performed when obstructive uropathy is suspected<br />

(53).


Chapter 2 / Diagnostic Evaluation 57<br />

Lower Urinary Truct Imuging<br />

When anatomical factors are considered as a possible cause of void-<br />

ing <strong>dysfunction</strong>, lower urinary tract imaging can be useful. A plain<br />

radiograph of the abdomen and pelvis (KUB) can determine the pres-<br />

ence of stones or foreign bodies, which can be contributing to LUTS<br />

and voiding <strong>dysfunction</strong>. The lower spine can also be evaluated (spina<br />

bifida occulta or sacral agenesis). A cystogram, with stress and oblique<br />

views, can be useful in select cases of pelvic prolapse (especially when<br />

the degree of cystocele is difficult to determine on a physical exam)<br />

and stress incontinence in women to determine support defects (54).<br />

<strong>Voiding</strong> cystourethrography (VCUG) may detect abnormalities such<br />

as posterior urethral valves, urethral diverticulum, urethral stricture, or<br />

other anatomic abnormalities of the urethra. In cases where female<br />

bladder outlet obstruction is suspected the VCUC in addition to urody-<br />

namic studies is important. We prefer videourodynamics in such cases<br />

(55); however, when not available the VCUG can be done separately.<br />

The VCUG can also determine the presence of vesicoureteral reflux.<br />

Ultrasound of the bladder has become a convenient, noninvasive way<br />

to determine postvoid residual, especially with the introduction of small<br />

portable devices made specifically for determination of bladder volume.<br />

REFERENCES<br />

1. Wein AJ (1981) Classification of neurogenic voiding <strong>dysfunction</strong>. J Urol125:605-<br />

607.<br />

2. Dioho AC, Brock BM, Brown MB, et al. (1986) Prevalence of urinary and other<br />

urologic symptoms in the non-institutionalized elderly. J Urol 136: 1022-1025.<br />

3. Abrams P (1 994) New words for old: lower urinary tract symptoms for prostatism.<br />

Br Med J 308:929.<br />

4. Blaivas JG, Heritz DM (1996) Classification, diagnostic evaluation and treatment<br />

overview. In: Topics in Clinical Urology-Evaluation and Treatment of Urinary<br />

Incontinence, (Blaivas JG, ed.), New York: Igaku-Shoin, pp. 22-45.<br />

5. Resnick NM (1994) Urinary incontinence in the elderly. Medical Grand<br />

Rounds 3:281-290.<br />

6. Barry MJ, Fowler FJ, O’Leary MP, et al. (1992) The American Urological Association<br />

symptom index for benign prostatic hyperplasia. J Urol 148:1549-1557.<br />

7. Wagner TH, Partick DL, Bavendam TG, et al. (1995) Quality of life in persons<br />

with urinary incontinence: development of a new measure. Urology 47:67-72.<br />

8. DeJong F (1967) The Neurologic Examination. Philadelphia: W.B. Saunders.<br />

8a. Kim YH, Boone TB (1998) <strong>Voiding</strong> pressure-flow studies. In Practical Urodynamics<br />

Nitti VW, ed., Philadelphia: WB Saunders, pp. 52-64.<br />

9. Boone TB, Kim YH (1998) Uroflowmetry. In: Practical Urodynamics, Nitti, VW,<br />

ed., Philadelphia: W.B. Saunders, pp. 28-37.


58 Nitti and Ficazzola<br />

10. Siroky MB, Olsson CA, Krane RJ (1979) The flow rate nomogram: I. Development.<br />

J. Urd 122:665-668.<br />

11. Haylen BT, Ashby D, Sutherst JR, et al. (l 989) Maximum and average urine flow<br />

rates in normal male and female populations-the Liverpool nomograms. Brit J<br />

Ur01 64~30-38.<br />

12. Haylen BT, Parys BT, Anyaegbunam WI, et al. (1990) Urine flow rates in male<br />

and female urodynamic patients compared with the Liverpool nomograms. Br J<br />

Urol 65:483-487.<br />

13. Drach GW, Layton TN, Binard WJ (1979) Male peak urinary flow rate: relationships<br />

to volume voided and age. J Urol 122:210-214.<br />

14. Barry MJ, Giman CJ, O’Leary MP, et al. (1995) Using repeated measures of<br />

symptom score, uroflowmetry and prostatic specific antigen in clinical the management<br />

of prostate disease. J Urol 153:99-103.<br />

15. Chancellor MB, Blaivas JG, Kaplan SE, et al. (1991) Bladder outlet obstruction<br />

versus impaired detrusor contractility: the role of uroflow. J Urd 145:810-<br />

812.<br />

16. Abrams PH, Griffths DJ (1979) The assessment of prostatic obstruction from<br />

urodynamic measurements and from residual urine. Br J Urd 5 I : 129-134.<br />

17. McLoughlin J, Gill KP, Abel PD, et al. (1990) Syrnptorns versus flow rates versus<br />

urodynamics in the selection of patients for prostatectomy. Br J Urol 66:303-305.<br />

18. Girman CJ, Panser LA, Chute CG, et al. (1993) Natura1 history of prostatism:<br />

urinary flow rates in a community-base study. J Urol 150:887-892.<br />

19. Drach GW, Ignatoff J, Layton T (1979) Peak urinary flow rate: observations in<br />

female subjects and comparison to male subjects. J Urd 122:215-219.<br />

20. Jorgensen JB, Jensen KM-E (1996) Uroflowmetry. Urol Clin N Am. 23:237-242.<br />

21. Hahn I, Fall M (1991) Objective quantification of stress urinary incontinence: a<br />

short reproducible, provocative pad-test. Neurourol Urodyn 10:475-48 1.<br />

22. Jorgensen L, Lose G, Andersen JT (1987) One hour pad-weighing for test objective<br />

assessment of female incontinence. Obstet Gynecol 69:39-41.<br />

23. Jorgensen L, Lose G, Thunedborg P (1987) Diagnosis of mild stress incontinence<br />

in females: 24-hour home pad test versus the 1-hour ward test. Neurourol Urodyn<br />

6: 165-166.<br />

24. Versi E, Anand D, Smith P, et al. (1996) Evaluation of the home pad test in the<br />

investigation of female urinary incontinence. Br J Obstet Gynecol103(2): 162-167.<br />

25. Nitti VW, Combs AJ (1998) Urodynarnics when why and how. In: Practical<br />

Urodynamics, Nitti VW, ed., Philadelphia: WB Saunders, pp. 15-26.<br />

26. Nitti VW (1998) Cystometry and abdominal pressure monitoring. In: Practical<br />

Urodynarnics, Nitti VW, ed., Philadelphia: W.B. Saunders, pp. 38-51.<br />

27. Abrams P, Blaivas JG, Stanton SA, and Andersen JT (1988) The standardization of<br />

terminology of lower urinary tract function. Scand J Urol NephroZ (Suppl.) 1145.<br />

28. McGuire EM, Woodside JR, Borden TA (1981) Prognostic value of urodynamic<br />

testing in myelodysplasic children. J Urd 126:205.<br />

29. Griffiths DJ (1973) The mechanics of the urethra and of micturition. Br J Urd<br />

451497-507.<br />

30. Schafer W (1990) Principles and clinical application of advanced urodynarnic<br />

analysis of voiding function. Urol Clin NArn 17553-566.<br />

31. Cespedes RD, McGuire EJ (1998) Leak point pressures. Practical In: Urodynamics,<br />

Nitti VW, ed., Philadelphia: WB Saunders, pp. 94-107.<br />

32. McGuire EJ, Fitzpatrick CC, Wan J, et al. (1993) Clinical assessment of urethral<br />

sphincter function. J Urd 150: 1452-1454.


Chapter 2 / Diagnostic Evaluation 59<br />

33. O’Donnell PD. Electromyography. In: Practical Urodynamics, Nitti WV, ed.,<br />

Philadelphia: WB Saunders, 1988 pp. 65-71.<br />

34. Blaivas JG, Fischer DM (1981) Combined radiographic and urodynamic monitoring:<br />

advances in techniques. J Urol 125:693-694.<br />

35. Nitti NW, Raz S (1996) Urinary retention. In: Female Urology, 2nd ed, Raz S,<br />

ed., Philadelphia: WB Saunders, pp. 197-213.<br />

36. Ouslander J, Leach G, Abelson S, et al. (1988) Simple versus multichannel<br />

cystometry in the evaluation of bladder function in an incontinent geriatric population.<br />

J Urol 140:1482-1486.<br />

37. Langmade CF, Oliver JA (1984) Simplifying the management of stress incontinence.<br />

Am J Obstet Gynecol 149:24.<br />

38. Fischer-Rasmussen W, Hansen RI, Stage P (1986) Predicative value of diagnostic<br />

tests in the evaluation of female urinary stress incontinence. Acta Obstet Gynecol<br />

Scand 65:291.<br />

39. Cardoza LD, Stanton SL (1980) Genuine stress incontinence and detrusor instability-a<br />

review of 200 patients. Br J Obstet Gynecol 87:184.<br />

40. Mundy AR (1985 j The unstable bladder. Urol Clin N Am 12:317-328.<br />

41. Duldulao IUi, Diokno AC, Mitchell B (1996): Value of urinary cytology in<br />

women presenting with urge incontinence and/or irritative voiding symptoms.<br />

J Urol 157:113-116.<br />

42. Robertson JR (1976) Urethroscopy-the neglected gynecological procedure. Clin<br />

Obstet Gynecol 19:3 15-340.<br />

43. Aldridge CW Jr, Beaton JH, Nanzig RP (1978) A review of office cystoscopy<br />

and cystornetry. Am J Obstet Gynecol 13 1:432-437.<br />

44. Scotti RJ, Ostergard DR, Guillaume AA, Kohatsu KE (1990) Predicative value<br />

of urethroscopy compared to urodynamics in the diagnosis of genuine stress<br />

incontinence. J Reprod Med 35:772-776.<br />

45. Horbach NS, Ostergard DR (1994) Predicting intrinsic sphincter <strong>dysfunction</strong> in<br />

women with stress urinary incontinence. Obstet Gynecol 84: 188-192.<br />

46. Govier FE, Pritchett TR, Kornman JD (1994) Correlation of the cystoscopic<br />

appearance and functional integrity of the female urethral sphincteric mechanism.<br />

Urology 44:250-253.<br />

47. Sand PR, Hill RC, Ostergard DA (1987) Supine urethroscopic and standing<br />

cystometry as screening methods for the detection of detrusor instability. Obstet<br />

Gynecol 7057-60.<br />

48. World Health Organization Proceedings of the 3rd Consultation on Benign Prostatic<br />

Hyperplasia (BPH), Monaco, June 26-28, 1995, pp. 191-193.<br />

49. Mandal AK, Shwma SK, Vaidyanathan S, Goswami AK (1990) Ureterovaginal<br />

fistula: summary of 18 years’ experience. Br J Urol 65:453.<br />

50. Murphy DM, Grace PA, O’Flynn JD (1982) Ureterovaginal fistula: a report of<br />

12 cases and review of the literature. J Urol 128:924.<br />

51. Conway JJ (1992) “Well-tempered” diuresis renography: it’s historical development,<br />

physiological and technical pitfalls, and standardized technique protocol.<br />

Seminars Nuclear Med 22174.<br />

52. Hvistendahl JJ, Pedersen TS, Schmidt F, Hansen W, Djurhuus JC, Frokier J (1987)<br />

The vesico-renal reflex mechanism modulates urine output during elevated bladd<br />

pressure. Scand J Urol Nephrol 186(31 suppl.): 24.<br />

53. O’Reilly PH (1992) Diuresis renography. Recent advances and recommended<br />

protocols. Br J Urol 69: 113.


GO Nitri and Ficazzola<br />

54. Shapiro R, Raz S (1983) Clinical applications of the radiographic evaluation of<br />

female incontinence. In FernaEe Urology, Raz S, ed., Philadelphia: WB Saunders,<br />

pp. 123-136.<br />

55. Nitti VW, Tu LM, Gitlin J Diagnosing bladder outlet obstruction in women.<br />

J UPUZ 161: 1535-1540.


11 NEUROGENIC VESICO-URETH<br />

DYSFUNCTION


This Page Intentionally Left Blank


3 Cerebrovascular Accidents<br />

serge Peter Mdrinkovic, MD<br />

md Gopd H. Badhi, MD<br />

CONTENTS<br />

INTRODUCTION<br />

CENTRAL INNERVATION<br />

OF THE LOWER URINARY TRACT<br />

AND THE PONTINE MICTURITION<br />

CENTER (PMC)<br />

INCONTINENCE<br />

INCIDENCE OF URINARY<br />

PATHOPHYSIOLOGY OF INCONTINENCE<br />

EARLY PRESENTATION<br />

LATE PRESENTATION<br />

URODYNAMICS STUDIES<br />

HEMISPHERIC DOMINANCE<br />

AND VOIDING DYSFUNCTION<br />

UROLOGICAL EVALUATION<br />

HISTORY<br />

TECHNICAL DIFFICULTIES<br />

INTERPRETATION OF THE STUDY<br />

SPECIFIC SITUATIONS<br />

MEDICAL AND SURGICAL<br />

URINARY RETENTION IN WOMEN<br />

URINARY INCONTINENCE IN MEN<br />

AND WOMEN<br />

ALTERNATIVE THERAPIES<br />

SUMMARY<br />

REFERENCES<br />

MANAGEMENT<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfinction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Hurnana Press Inc., Totowa, NJ<br />

63


64 Marinkovic<br />

INTRODUCTION<br />

Cerebrovascular accidents (CVA) or stroke are a common and serious<br />

neurologic event in the elderly. The residual effects can be permanent<br />

or temporary and vary in their morbidity. These impairments may<br />

include the loss of memory, vision, speech, motor function, and control<br />

of micturition. A CVA is defined as the acute onset of a focal neurologic<br />

deficit caused by an occlusive event (i.e., cerebral emboli or atherosclerotic<br />

thrombis) or a hemorrhagic focus. Cerebrovascular accidents are<br />

the third leading cause of death in the United States (I) and although<br />

their incidence is still one tenth that of myocardial infarction, they<br />

result in over 30 billion dollars in annual health care costs (2). Risk<br />

factors for CVA’s include hypertension, diabetes mellitus, smoking,<br />

increased serum cholesterol level, alcohol consumption, obesity, stress,<br />

and sedentary lifestyle (3-6). Greater than 500,000 CVAs occur annually<br />

in the United States, one-third of which are fatal, one-third requiring<br />

long-term nursing care, and the remaining one-third returning home,<br />

many at their previous level of function (7). Recently, the incidence<br />

of cerebrovascular accidents in individuals over 70 years of age and<br />

their mortality have been shown to be declining (8,9).<br />

CVA can have a profound effect on genitourinary function. The<br />

voiding <strong>dysfunction</strong> can range from urinary retention to total urinary<br />

incontinence (10-14). The CVA can also result in significant sexual<br />

<strong>dysfunction</strong>, Since CVA predominantly occurs in the elderly population,<br />

the issue is further influenced by coexisting genitourinary <strong>dysfunction</strong><br />

(e.g., bladder outlet obstruction in males, stress urinary incontinence<br />

(SUI) in females). Cornorbid medical conditions can effect the genitourinary<br />

system (e.g., Diabetes mellitus, vascular <strong>dysfunction</strong>, coronary<br />

artery disease). Furthermore, changes secondary to aging (e.g., bladder<br />

instability, change in circadian rhythm, influence of hormonal withdrawal<br />

in females, cognitive impairment), can impair function of the<br />

genitourinary tract. It is generally recognized that genitourinary <strong>dysfunction</strong><br />

improves as the time from the onset of stroke increases. It is<br />

not precisely clear if this is owing to improvement in genitourinary<br />

function or declining disability of the patient.<br />

Urinary incontinence immediately after a stroke has also been determined<br />

to infer a poor overall prognosis (1.5,16). Wade and Hewer (16)<br />

followed 532 patients shortly after the onset of a CVA. Patients with<br />

urinary incontinence within 1 wk of CVA fared poorly, with 50% dying<br />

within 6 mo of presentation. Incontinence at 3 wk also predicted a<br />

higher risk of dying and a decreased chance of regaining mobility.


Chapter 3 / Cerebrovascular Accidents 65<br />

They also found a 44% incidence of urinary incontinence and 63% of<br />

these patients showed an alteration in their level of consciousness, as<br />

compared to 1% of 300 continent patients demonstrating a similar<br />

change in level of consciousness. Taub et al. (15) reviewed 639 regis-<br />

tered CVA patients for disability at 3 and 12 mo. Initial incontinence<br />

was found to be the best single indicator of future disability with a<br />

sensitivity and specificity of 60% and 78%. Using urinary incontinence<br />

as a good predictor of morbidity and mortality, these patients can be<br />

stratified into a high-risk group requiring more detailed and intensive<br />

medical surveillance for a potentially better outcome.<br />

CENTRAL INNERVATION<br />

OF THE LOWER URINARY TRACT<br />

AND THE PONTINE MICTURITION CENTER (PMC)<br />

The central nervous system (CNS) has an important role in regulating<br />

the ability of the bladder to facilitate the storage and emptying of<br />

urine. These CNS functions may be divided into three specific areas:<br />

suprapontine, pontine, and spinal centers.<br />

<strong>Voiding</strong> is coordinated by the neurons of the pontine-mesencephalic<br />

gray matter or the Pontine Micturition Center (PMC) (17,18). <strong>Voiding</strong><br />

depends on the spinobulbospinal reflex through the PMC after receiving<br />

input from the hypothalmus, thalmus, basal ganglia cerebellum, and<br />

cerebral cortex. The input from the suprapontine center is predominantly<br />

inhibitory but may have partial facilitory action. The major inhibitory<br />

areas appear to be the cerebellum, basal ganglia, and cerebral cortex,<br />

while facilitation may be regulated by the posterior hypothalmus and<br />

anterior pons (I 7-23). Additionally, most of the suprapontine input into<br />

the PMC is inhibitory. Interruption of this input by a cerebrovascular<br />

accident, Parkinson’s disease, or brain tumor may result in detrusor<br />

overactivity or detrusor hyperreflexia (DH). This detrusor hyperreflexia<br />

may manifest itself with symptomatic frequency, urgency, and urge<br />

incontinence.<br />

The PMC regulates efferent stimuli to the bladder and external<br />

sphincter by its two regions, medial and lateral. The medial region<br />

regulates motor stimuli to the detrusor muscle by overseeing the communication<br />

between the reticulospinal tracts and the sacral intermediolateral<br />

cells carrying the preganglionic parasympathetic neurons that<br />

innervate the bladder detrusor muscle, Electrical stimulation of the


66 Marinkovic<br />

Badlani<br />

median region elicits a decrease in pelvic floor electromyographic<br />

(EMG) activity and urethral pressure with a concomitant increase in<br />

intravesical pressure. The lateral region regulates the corticospinal stimuli<br />

to Onuf‘s nucleus in the sacral spinal cord, which innervates the<br />

levator ani, and urethral and anal sphincters. Electrical stimulation of<br />

this area results in an increase of urethral pressure and levator ani (EMG)<br />

activity but with only no or minimal elevation in intravesical pressure.<br />

INCIDENCE OF URINARY INCONTINENCE<br />

The incidence of early poststroke urinary incontinence varies from<br />

57-83% (11,26,31,32). These studies suggest that incontinence may be<br />

transitory and related to the patient’ S immobility and altered mental<br />

status. In a study of 151 patients, Borrie et al. (26) reported an initial<br />

incontinence rate of 60%. After 1 mo, this incidence was reduced to 29%<br />

at 1 mo. Additionally, 66% of patients with mild incontinence at 1 mo<br />

regained continence at 3 mo. Brocklehurst et al. (32) found an initial<br />

incontinence rate of 39%. They demonstrated that by 2 mo 55% were<br />

continent and at 6 mo this increased to 80%. However, the 2 and 3 yr<br />

follow-up data demonstrate an incidence higher than the general population.<br />

Improvement with time is impressive and should be relayed to<br />

the patient and family. Communicating this information could improve<br />

the patient’s self-image and be an encouraging factor for the future.<br />

PATHOPHYSIOLOGY OF INCONTINENCE<br />

Frontal cortical lesions from a CVA may affect a patient’s higher<br />

cognitive functions. This may be reflected by the patient’ S inability to<br />

suppress a reflex detrusor contraction resulting in urinary incontinence.<br />

This incontinence can be owing to<br />

1. Detrusor hyperreflexia;<br />

2. CVA related language and cognitive impairments with normal bladder<br />

function; or<br />

3. Overflow incontinence secondary to detrusor hyporeflexia (mediated<br />

by either neuropathies or medications) (13).<br />

The concomitant diagnosis of dementia and benign prostatic hyperplasia<br />

(BPH) in males and urethral incontinence in females may also contribute


Chapter 3 / Cerebrovascular Accidents 67<br />

to urinary incontinence. However, it must be emphasized that inconti-<br />

nence in the CVA patient merits evaluation to determine its etiology.<br />

EARLY PRESENTATION<br />

Acute urinary retention commonly occurs after a CVA. The neurophysiologic<br />

explanation for this detrusor areflexia is unknown but is<br />

referred to as cerebral shock (25). The size, location, and extent of a<br />

CVA may have a profound affect on micturition. However, acute urinary<br />

retention may not necessarily be a result of CVA but may be related<br />

to the patient’s inability to communicate their need to void, impaired<br />

consciousness, temporary overdistension of the bladder, and restricted<br />

mobility (26). Pre-existing comorbidities such as bladder outlet obstruction,<br />

diabetic cystopathy, and various medications (with anticholinergic<br />

side effects) may also be responsible for urinary retention. Prospective<br />

urodynamic studies on patients soon after unilateral CVAs demonstrated<br />

a 21 % incidence of overflow incontinence because of detrusor hyporeflexia.<br />

However, a significant number of patients in this study were<br />

diabetics or patients on anticholinergic medications (13). Our urodynamic<br />

evaluation of patients soon after cerebellar infarcts have detected<br />

a high incidence of urinary retention. In contrast, cerebellectomy in an<br />

animal model (27) results in hyperreflexia rather than areflexia; thus,<br />

the role of the cerebellum in micturition is still speculative and will<br />

need more prospective clinical and animal studies for clarification.<br />

LATE PRESENTATION<br />

The symptoms of bladder <strong>dysfunction</strong> after established CVA include<br />

frequency, urgency, and urge incontinence (11,30). These symptoms<br />

are generally a result of detrusor hyperreflexia (14,28,29,30). Tsuchida<br />

et al. (30) evaluated 39 patients urodynamically after a CVA. The mean<br />

time from CVA to urodynamic evaluation was 19 m0 (range 1 1 da to<br />

13 yr). They found that 66% of the patients complained of frequency<br />

or urge incontinence and the remaining 33% had either dysuria or<br />

urinary retention. Tsuchida determined that the symptoms of urgency<br />

or frequency are related to detrusor hyperreflexia, and this may also<br />

contribute to urge incontinence.


68 Marinkovic and Badlani<br />

Table 1<br />

Patient Urodynamic Findings on CMG and EMG<br />

Group CMG EMG<br />

Normal Normal<br />

Hyperre flexic Normal<br />

H y perreflexc AVCS<br />

Hyperreflexic DSD<br />

Areflexic Normal<br />

Used with permission from ref. 33.<br />

URODYNAMICS STUDIES<br />

Published studies assessing a CVA’s affect on the bladder have<br />

largely been performed in a retrospective fashion with evaluations<br />

occurring from days to years after their initial event. Detrusor hyper-<br />

reflexia has been demonstrated to be the main finding in their voiding<br />

<strong>dysfunction</strong> (12,14,28,30). Other studies have attempted to correlate<br />

the site of ischemic or hemorrhagic infarct with urodynamic findings<br />

but have been inconclusive (28).<br />

Burney et al. (33) performed a prospective study of 60 patients,<br />

correlating the site of ischemic or hemorrhagic CVA by either CaT Scan<br />

or magnetic resonance imaging (MRI) with the results from urodynamic<br />

testing. All patients were imaged within 72 hr of presentation. If a<br />

patient had a diagnostic infarct, the patient underwent a filling and<br />

voiding multichannel cystometrogram (CMG) study with a simultane-<br />

ous electromyogram (EMG) recording with a needle electrode. Patients<br />

were then separated into five groups based on the results of their<br />

urodynamic testing (see Table 1). This study demonstrated that the<br />

majority of frontalparietal and internal capsular lesions caused detrusor<br />

hyperreflexia and absent volitional control of the external sphincter<br />

(AVCS). Those patients with temporo-occipital lesions had normal<br />

urodynamic studies (see Table 2).<br />

Burney et al. (33) also found that 47% of their patients were in<br />

urinary retention and overflow incontinence. Various lesions were<br />

responsible for bladder areflexia, including the frontalparietal area,<br />

internal capsule, basal ganglia, thalamus, pons, and cerebellum (see<br />

Table 2). Eighty-five percent of the hemorrhagic infarcts (17/20) exhib-<br />

ited an areflexic bladder, whereas only 10% (n = 4) of the ischemic<br />

CVA’s demonstrated detrusor areflexia (see Table 3). Will the type of


Chapter 3 / Cerebrovascular Accidents 69<br />

Table 2<br />

Site of CVA Correlated with Urodynamic Findings<br />

Site of lesion Group I Group 2 Group 3 Group 4 Group 5<br />

Frontalparietal 0 0 6 0 2<br />

Occipital 1 0 0 0 0<br />

Temporal 1 0 0 0 0<br />

Multifocal 0 0 2 0<br />

Internal capsule 2 0 12 0 4<br />

Basal ganglia 2 2 3 0 4<br />

Combined 0 0 1 3 0<br />

Thalamus 2 0 0 0 2<br />

Pons 0 0 2 0 3<br />

Cerebellum 0 0 0 0 6<br />

Totals 8 2 24 5 21<br />

Used with permission from ref. 33.<br />

infarct determine the type of voiding <strong>dysfunction</strong>? This has yet to be<br />

confirmed, but areflexia has been seen after large infarcts of either<br />

varieties (28,34).<br />

Electromyograms performed in the majority of poststroke patients<br />

reveal uninhibited relaxation of the external sphincter during or preced-<br />

ing detrusor contractions with resultant urinary incontinence (15). Inter-<br />

estingly, detrusor sphincter dyssynergia (DSD) is uncommon in the<br />

poststroke period (13,35). However, pseudodyssynergia, demonstrated<br />

by the voluntary contraction of the external sphincter during an involun-<br />

tary detrusor contraction, may be more commonly observed in the<br />

recovery phase and should not be misinterpreted as DSD (36,37).<br />

Because of these confounding clinical possibilities, we recommend that<br />

there be close supervision of all urodynamic data and clinical symptoms<br />

by the urologist during a urodynamic study.<br />

Table 3<br />

Type of CVA and Urodynamic Findings<br />

Type of CVA Group I Group 2 Group 3 Group 4 Group 5<br />

Hemorrhagic (n = 20) 1 0 1 1 17<br />

Ischemic (n = 40) 7 2 23 4 4<br />

Totals (n = 60) 8 2 24 5 21<br />

Used with permission from ref. 33.


70 Marinkovic<br />

No of patients<br />

Average age<br />

Capacity<br />

No hyperreflexia<br />

No areflexia<br />

No normal<br />

Table 4<br />

Urodynamic Results with Right Hemiplegia<br />

Used with permission from ref. 12.<br />

13 males<br />

77 yr<br />

288 mL<br />

6<br />

2<br />

5<br />

HEMISPHERIC DOMINANCE<br />

AND VOIDING DYSFUNCTJON<br />

11 females<br />

84 yr<br />

317 mL<br />

6<br />

3<br />

2<br />

Hemispheric dominance has been well-established for language,<br />

whereas the temporal lobe has been linked to musical aptitude. Patients<br />

who have experienced a right hemispheric CVA have been known to<br />

incur more sexual <strong>dysfunction</strong> (38). In 1980, Khan et al. (14) postulated<br />

that patients after nondominant CVAs were less likely to have urinary<br />

incontinence. If so, what effects would a dominant hemispheric stroke<br />

have on voiding?<br />

We retrospectively evaluated 44 symptomatic patients admitted to<br />

a rehabilitation unit after a stroke (12,33). The mean patient age was<br />

81.2 yr and the time from CVA to urodynamic evaluation ranged from<br />

1-12 mo. Tables 4 and 5 demonstrate their urodynamic findings with<br />

right hemiplegia or left hemiplegia. Both results were similar and we<br />

concluded that, as with many other studies (14,28) there was no signifi-<br />

cant difference in bladder <strong>dysfunction</strong> with dominant and nondominant<br />

hemispheric WAS.<br />

Table 5<br />

Urodynamic Results with Lek Hemiplegia<br />

No of patients 14 males 6 females<br />

Average age 78 yr 86 yr<br />

Capacity 210 mL 416 mL<br />

No hyperreflexia 6 2<br />

No areflexia 0 4<br />

No normal 8 0<br />

Used with permission from ref. 12.


Chapter 3 / Cerebrovascular Accidents 71<br />

UROLOGICAL EVALUATION<br />

The 1990s have seen the implementation of a multidisciplinary<br />

approach to the evaluation of the strokes (39). Neurologist, internist,<br />

and psychiatrist play key roles, whereas the urologist is consulted as<br />

needed. Our knowledge of voiding <strong>dysfunction</strong> in stroke patients largely<br />

comes from the evaluation of only symptomatic patients.<br />

In most instances, urological evaluation begins when the patient is<br />

transferred from an acute-care facility to a rehabilitation or a stroke<br />

unit. Like any other disease, process assessment begins with a history<br />

predating the stroke and proceeds with a laboratory and specialized<br />

tests such as a urodynamic evaluation and endoscopy. There are often<br />

problems related to the evaluation of the elderly. These can be broadly<br />

categorized into four areas:<br />

1. Obtaining an adequate history;<br />

2. Technical difficulty with doing studies in the elderly;<br />

3. Interpreting a urodynarnic study (What is incidental due to aging?<br />

What is abnormal?); and<br />

4. Multiple coexisting etiologies in a single patient.<br />

When urodynamics are indicated, the best approach is to begin with<br />

simple screening urodynamics (uroflow and simple cystometry) then<br />

proceed to complex urodynarnics (multichannel urodynamics or video<br />

urodynamics), as needed. However, keep in mind that the evaluation<br />

and the extent to which it is carried out depend not only on the likely<br />

diagnosis, but also on the patient’s comorbid conditions, level of func-<br />

tioning, and the wishes of the patient and/or his or her family.<br />

HISTORY<br />

Diagnosis largely depends on the history of symptoms. The elderly<br />

have been known to underreport their problems, especially when it is<br />

related to urinary incontinence. Interviewing the patient may not provide<br />

you with sufficient information on which to base a diagnosis. This is<br />

often compounded in those patients with dominant hemispheric CVAs<br />

with aphasia and those with dementia. Interviewing the caregiver<br />

(spouse, relative, or nurse’ S aide) may be necessary to provide additional<br />

information. A questionnaire given to the family, along with a voiding<br />

chart that is brought back on the second visit, is very helpful to assess<br />

ingested fluid and functional bladder capacity. The first visit is used


72 Marinkovic<br />

Badlani<br />

to rule out reversible causes of incontinence such as infection, stool<br />

impaction, medications, and so forth. Targeted questions related to<br />

symptoms are helpful in relation to symptoms such as urinary frequency.<br />

It is prudent to ask if the frequency is associated with high or low<br />

volume, whether it is day or night, and if it is associated with any<br />

symptoms of dysuria. Selective nocturia in the elderly can be caused<br />

by increased urinary output during the night due to loss of circadian<br />

rhythm, and/or peripheral pooling of fluid, which is mobilized with<br />

recumbency. More importantly, nocturia can be owing simply to a poor<br />

sleep pattern. A patient’s prescription and often-overlooked over-thecounter<br />

medications may have urological implications. In the physical<br />

examination, specific attention should be detailed to the mobility, visual<br />

acuity, hand coordination, and mental status of the patient in addition<br />

to the focused urological exam, including a digital rectal exam pelvic<br />

floor assessment in females, and prostatic evaluation in males. A<br />

regional neurological examination allows assessment of the ability to<br />

isolate the pelvic floor and voluntary contraction of the same. A voided<br />

urine analysis for infection and hematuria follows. If hematura is present<br />

appropriate assessment with cytology, intravenous pyelogram and<br />

endoscopy should be considered.<br />

TECHNICAL DIFFICULTIES<br />

There are several technical difficulties in doing urodynamics in the<br />

elderly, and the studies are tailored to individual patients rather than<br />

a set study being done on every individual. Often the elderly have<br />

decreased mobility and therefore may not be able to stand up or use<br />

the uroflow machine. In such cases, only a supine study may be possible.<br />

Every attempt is made to have them void prior to the study as a postvoid<br />

residual is an important parameter in the elderly. Use of a rectal pressure<br />

balloon is often mandatory in the elderly owing to the straining and<br />

Valsalva during the cystometry. The most useful study is cystometry<br />

and should be used as a screening urodynarnics study along with a<br />

postvoid residual. If complex urodynamics is then required, it should<br />

be performed with fluoroscopy along with the measurement of several<br />

other parameters (i.e., Valsalva leak-point pressure). Electromyogram<br />

(EMG) is done in select cases as DSD is an uncommon finding in<br />

CVA. Fluoroscopy is helpful in preselecting patients for EMG.


Chapter 3 I Cerebrovascular Accidents 73<br />

INTERP~TATION OF THE STUDY<br />

Since uninhibited bladder contractions can occur in continent elderly<br />

patients, the significance of bladder instability alone in the presence<br />

of incontinence is unclear. The uninhibited contraction can be present<br />

independently of associated stress incontinence and/or outlet obstruction.<br />

The reported incidence of age-related uninhibited contractions is<br />

10% in women and 25-35% in men. Other age-related changes include,<br />

as mentioned earlier, increased urine output at night, prostatic enlargement<br />

in men, and changes in the urethra in the female related to loss<br />

of hormonal support. As such, the urethral pressure profile may change<br />

in the elderly as a normal aging process rather than being an abnormal<br />

finding. Ultimately, it is better not to rely on a single parameter, but<br />

rather review the whole picture in context of the associated conditions.<br />

SPECIFIC SITUATIONS<br />

Benign prostatic hyperplasia (BPH) and stroke are common clinical<br />

situations involving a male patient who has some degree of incontinence<br />

following CVA. In such cases, one is left to wonder if detrusor overacti-<br />

vity is secondary to cerebral vascular accident and/or bladder outlet<br />

obstruction. It is helpful to get a history of voiding disturbances prior<br />

to the stroke since results following prostatectomy in CVA patients<br />

report persistence of voiding disturbances and unsatisfactory bladder<br />

control. It is prudent to evaluate these patients by sophisticated urody-<br />

namic techniques. One of the parameters that helps in differentiating<br />

the two is if a flow can be obtained from the patients (patients with<br />

BPH will generally have a low flow rate). Other techniques used are<br />

measurements of micturating urethral pressure profile during the void-<br />

ing contraction. This test described by Yalla et al. (40) allows one to<br />

detect the point of obstruction or to ascertain if the posterior urethra<br />

is isobaric with the bladder during voiding. This test requires the use<br />

of fluoroscopy and a special catheter. It is also helpful clinically to<br />

know the effect of anticholinergics on this hyperreflexia preoperatively<br />

as 30% of the patients with hyperreflexia may not respond to anticholin-<br />

ergics and postprostatectomy. These patients will be significantly incon-<br />

tinent. The drawback of using anticholinergics preoperatively is that<br />

the patient may end up in urinary retention. The patient and family are<br />

prepared for the same. The use of minimally invasive techniques sach


74 Marinkovic and Badlani<br />

as temporary or permanent prostatic stents (Conticath [41] and Urolume<br />

[42,43]) may also be useful in these scenarios.<br />

Diagnostic dilemma in females includes coexisting bladder hyper-<br />

reflexia secondary to CVA and urethral incontinence due to sphincteric<br />

malposition and or intrinsic deficiency. Assessment of SUI can be<br />

difficult if there is low-volume bladder hyperreflexia; conversely, detru-<br />

sor hypeneflexia may be missed if the leak-point pressure is low and<br />

the bladder is not filled to an adequate volume. The latter is easy to<br />

overcome by occluding the outlet with a Foley balloon and filling the<br />

bladder to assess its filling function. A low-volume detrusor hyper-<br />

reflexia with a coexisting open bladder neck at rest on fluoroscopy can<br />

be a difficult problem.<br />

MEDICAL AND SURGICAL MANAGEMENT<br />

Retention and Obstruction in Men<br />

In the poststroke period, with the onset of cerebral shock and urinary<br />

retention, the need for clean intermittent catheterization (CIC) every<br />

4-6 h may be a useful minimally invasive therapy. As spontaneous<br />

voiding occurs and postvoid residuals consistently return to less than<br />

100 mL, CIC can be discontinued. If CIC is not feasible owing to<br />

patient or caretaker factors, an indwelling Foley catheter can be used<br />

and changed once a month. For those patients with documented bladder<br />

outlet obstruction, new medical and surgical therapies are becoming<br />

avail able.<br />

For some patients, long-acting oral medications called alphal blockers<br />

such as terazosin (44) (5 or 10 mg, qd), or doxazosin (45) (4 or 8<br />

mg, qd) can be utilized if the patient is medically stable. They selectively<br />

antagonize the alpha, receptors, which mediate prostate smooth muscle<br />

tension, improve American Urological Association Symptom Scores,<br />

increase Qmax (peak flow rate mL/s), and reduce postvoid residuals. Their<br />

efficacy is dose-dependent and dosing needs to be titrated. However,<br />

potential side effects including dizziness, fatigue, headache, postural<br />

hypotension, and may preclude its use in the elderly. Tamsulosin46 (0.4<br />

or 0.8 mg, qd) is a potent and selective alphal, antagonist and does not<br />

require dose titration. It does not have significant effect on blood<br />

pressure and may be better tolerated.<br />

On the horizon are investigational, temporary prostatic stents or<br />

catheters. The Conticath (41); (Fig. 1) bridges the space from the


Chapter 3 I Cerebrovascular Accidents 75<br />

Fig. 1. Conticath-a temporary prostatic catheter.<br />

bladder neck through the prostate and ends proximal to the external<br />

sphincter and can be used up to 30 da. The Conticath allows for voiding<br />

with much less discomfort and there are few contraindications for its<br />

use. Additionally, for high-risk patients, a permanent prostatic stent<br />

can be utilized. The Urolume (42,43) functions similar to the temporary<br />

stent, and can be placed under local anesthetic with minimal patient<br />

discomfort. It has proven to be an effective, low-risk therapy.<br />

Less invasive endoscopic procedures such as transurethral microwave<br />

thermotherapy (477 transurethral needle ablation of the prostate<br />

(48) Holmium (49) laser, and Indigo (50,51) laser prostatectomy may<br />

prove useful in the post-CVA patient. These procedures can all be<br />

performed under local anesthetic and reduce the operative risk to the<br />

patient. However, the urologist should refrain from performing a traditional<br />

transurethral resection of the prostate in the poststroke period<br />

for 6-12 m0 because incontinence and morbidity may be increased.<br />

Lum and Marshall (34) identified several variables that can affect outcome<br />

in the postprostatectomy CVA patient. These factors can lead to<br />

an unsatisfactory outcome in upwards of 50% of the patients. The most<br />

closely correlated with poor clinical outcome (higher incidence of<br />

urinary incontinence) were age greater than 70, worsening neurological<br />

symptoms, site of CVA (bilateral or nondominant lesions), lack of<br />

associated urinary symptoms, and prostatectomy within 1 yr of a CVA.


76 Marinkovic<br />

Fig. 2. Inflow device-24F intraurethral valved catheter with intraluminal pump.<br />

URINARY MTENTION IN WOMEN<br />

Until recently women in urinary retention secondary to detrusor<br />

hyporeflexia have the following available: clean intermittent catheterization<br />

(CIC), or chronic indwelling Foley catheter. Although CIC<br />

offers excellent long-range management of urinary retention, its use<br />

in CVA is restricted by the inability of the patient to perform this<br />

procedure owing to hemiparesis or hemiplegia, unless a family member<br />

or caregiver is able to do this. Chronic indwelling Foley catheters have<br />

been the alternative. A new investigational therapy is the In-Flow device<br />

(Influence Medical Technologies, San Francisco, CA) (Fig. 2), a 24F<br />

intraurethral valved catheter with intraluminal pump for a female atonic<br />

bladder. The catheter is fixed at the bladder neck by flexible silicone<br />

fins, which open like petals of a flower inside the bladder, and by an


Chapter 3 / Cerebrovascular Accidents 77<br />

external tab at the urethral meatus. To urinate, the patient or caregiver<br />

holds a small hand-held battery operated magnetic unit near the lower<br />

abdomen, close to the pubic area. This activates a valve in the catheter<br />

and drives an intraluminal pump to draw urine from the bladder at a<br />

flow rate similar to that seen in normal urination. This device may<br />

prove to be the device of choice if clinical trials comparing it to CIC<br />

are successful. 1 "<br />

URINARY INCONTINENCE IN MEN AND WOMEN<br />

The urologist goal with the incontinent patient is to restore a socially<br />

acceptable level of urinary continence while minimizing the risks of<br />

infection. Patients with detrusor hyperreflexia can be treated with timed<br />

voiding and concomitant fluid restriction and anticholinergic medications.<br />

Outlet obstruction in males and detrusor hyperactivity with<br />

impaired contraction (DHIC) in males and females can result in urinary<br />

retention with the injudicious use of anticholinergic medications (i.e.,<br />

oxybutynin, dicyclomine hydrochloride, or hyoscyamine sulfate USP).<br />

In male patients, outflow obstruction should be treated prior to addressing<br />

detrusor hyperreflexia. We recommend the periodic monitoring of<br />

postvoid residuals in DHIC patients after the initiation of anticholinergic<br />

therapy, and educating these patients about the risks of urinary retention<br />

so that the risk can be minimized. Tolterodine (52,53) is a new anticholinergic<br />

that has been recently developed and given FDA approval for<br />

1998. It has less binding to cholinergic receptors in the salivary glands<br />

and leads to less dry mouth and discontinuation of therapy secondary<br />

to adverse affects. For women with detrusor hyperreflexia and SUI,<br />

several oral medications are available that cause increased bladder outlet<br />

resistance, such as imipramine or phenylpropanolamine hydrochloride/<br />

guaifenesin. There are many new minimally invasive surgical procedures<br />

that can also address concomitant SUI in women.<br />

If medical and/or behavioral therapy fail in the female CVA patient<br />

with urethral incontinence, a few minimally invasive procedures are<br />

available that provide fair to excellent improvement in SUI:<br />

1. Periurethral or transurethral collagen injection (54);<br />

2. Urosurge, (Urosurge Inc, Coralville, CA) (Fig. 3), an investigational<br />

self-detachable balloon system that is placed at the bladder neck und<br />

endoscopic guidance similar to collagen therapy; and


78 Marinkovic<br />

Fig. 3. Urosurge device. A self detachable balloon system for stress urinary<br />

incontinence<br />

Badlani<br />

3. Pubovaginal sling witwwithout bone anchors and with either autologous<br />

tissue (56) (rectus fascia or fascia lata), allogenic (57) (donor<br />

cadaver fascia), or synthetic materials (polypropylene (581 or polytetrafluoroethylene<br />

1.591).<br />

Allogenic and synthetic materials lead to decreased operating room<br />

time and cost without sacrificing patient outcome or morbidity as com-<br />

pared to autologous tissue. Our success rate and complication rate using<br />

polypropylene with bone anchors (58) has been outstanding in this<br />

patient population. It is imperative that detrusor <strong>dysfunction</strong> be<br />

addressed first before treating urethral incontinence.<br />

ALTERNATIVE THERAPIES<br />

Sometimes all that may be necessary for incontinence therapy in<br />

the stroke patient will be simple behavioral modifications such as fluid<br />

restriction and initiating a timed voiding schedule both determined<br />

from a voiding diary. Many demented patients who no longer have<br />

socially appropriate behavior may also benefit substantially from a<br />

prompted voiding schedule. Celber et al. (13) reported that 37% of


Chapter 3 / Cerebrovascular Accidents 79<br />

severely handicapped (aphasia, dementia, or immobility) incontinent<br />

patients had normal functioning bladders. Interesting, even these debilitated<br />

patients benefited from prompted voiding schedules and fluid<br />

restriction. This method of patient management is very labor intensive<br />

and works best in a home environment with one-to-one patient attention.<br />

For patients who fail medicalhehavioral therapy and are poor surgical<br />

candidates, there are less-than-ideal alternatives. The use of a carefully<br />

supervised chronic indwelling Foley catheter may be considered,<br />

with monthly catheter changes. However, complications can occur with<br />

a Foley catheter, such as increased detrusor hyperactivity, urinary tract<br />

infections, bladder calculi, squamous cell carcinoma of the bladder,<br />

adenocarcinoma of the bladder, urethral erosion, and fistulas. A suprapubic<br />

catheter may result in a similar complication profile to the chronic<br />

indwelling Foley and thus should also be approached with caution.<br />

Condom catheters are available but disturbing problems such as skin<br />

excoriation, maintenance of the device, and urinary tract infections<br />

may also limit their usefulness.<br />

CIC is another viable alternative therapy for those patients with<br />

overflow incontinence secondary to bladder outlet obstruction and/<br />

or detrusor hyporeflexia. The patient or caretaker will need proper<br />

instruction, manual dexterity, and motivation to perform this minimally<br />

invasive procedure with success (60). While treatment attempts at curing<br />

a condition, management can aim at lessening the impact of voiding<br />

<strong>dysfunction</strong> and improving the quality of life.<br />

SUMIVURY<br />

The urologist now plays an important role in the multidisciplinary<br />

evaluation of poststroke voiding <strong>dysfunction</strong>. An accurate assessment<br />

of voiding <strong>dysfunction</strong> is feasible in most patients for implementation<br />

of a logical therapeutic plan in conjunction with other caregivers. His<br />

or her role can help improve patient self-esteem, and quality of life,<br />

and reduce morbidity through an accurate diagnosis and implementation<br />

of noninvasive or minimally invasive therapies.<br />

REFERENCES<br />

1. Kurtzke JF (1985) Epidemiology of cerebrovascular disease. In: Cerebrovascular<br />

Survey Report for the National Institute of Neurological and Communicative


80 Marinkovic and Badlani<br />

2.<br />

3.<br />

4.<br />

Disorders and Stroke. McDowell FM, Caplan LR, eds., Bethesda, MD: National<br />

Institute of Neurological and Communicative Disorders and Stroke, pp. 1-34.<br />

Yatsu FM, Grotta JC, Villar-Cordova C (1997) In: Cerebrovascular Disorders in<br />

Scientijc American Medicine, v01 4, Dale DC, Fedeman DD, eds., New York,<br />

NY: Scientific American Inc, pp. 1-12.<br />

Gorelick PB (1995) Stroke prevention. Arch Neurol 52:347-52.<br />

Love BB, Biller J, Jones MP, et al. (1990) Cigarette smoking: a risk factor for<br />

cerebral infarction in young adults. Arch Neurol 47:693-697.<br />

SHEP Cooperative Research Group (1993) Prevention of stroke by antihyperten-<br />

5.<br />

sive drug treatment in older persons with isolated systolic hypertension. JAMA<br />

265:3255-3258.<br />

6. Toni D, De Michele M, Fiorelli M, et al. (1994) Influence of hyperglycemia on<br />

infarct size and clinical outcome of acute ischemic stroke patients with intracranial<br />

arterial occlusion. J Neurol Sci 123: 129-13 1.<br />

7. Heart and Stroke Facts Statistics (1993) American Heart Association, Dallas, TX.<br />

8. May DS, Kittner SJ (1994) Use of Medicare clairns data to estimate national<br />

trends in stroke incidence, 1985-1991. Stroke 25:2343-2346.<br />

9. Klag MJ, Whelton PK, Seider AJ (1989) Decline in US stroke mortality, demographic<br />

trends and antihypertensive treatment. Stroke 20: 14-21.<br />

10. Gross JC (1990) Bladder <strong>dysfunction</strong> after a stroke: it’s not always inevitable.<br />

J Gerontologic Nursing 16:20-23.<br />

11. Arunabh MB, Badlani GH (1993) Urologic problems cerebrovascular in<br />

accidents.<br />

Problems Urol 7:41-53.<br />

12. Badlani GH, Vohara S, Motola JA (1991) Detrusor behavior in-patients with<br />

dominant hemispheric strokes. Neurourol Urodyn 10: 1 19-121.<br />

13. Gelber DA, Good DC, Laven LJ, et al. (1993) Causes of urinary incontinence<br />

after acute hemispheric stroke. Stroke 24:378-382.<br />

14. Khan Z, Hertanu J, Yang WC, et al. (1980) Predictive correlation of urodynamic<br />

<strong>dysfunction</strong> and brain injury after cerebrovascular accident. J Urol 126:86-88.<br />

15. Taub NA, Wolfe CD, Richardson E, et al. (1994) Predicting the disability of firsttime<br />

stroke sufferers at one year. 12-month follow-up of a population-based cohort<br />

in Southeast England. Stroke 25:352-357.<br />

16. Wade DT, Hewer RL (1978) Outlook after an acute stroke: Urinary incontinence<br />

and loss of consciousness compared in 532 patients. Quart J Med 56:601-<br />

608.<br />

17. Carlsson CA (1978) The supraspinal control of the urinary bladder. Acta Phamzacol<br />

Toxicol 43(Supp 2):8-12.<br />

18. DeGroat WC, Steers WD (1990) Autonomic regulation of the urinary bladder<br />

and sexual organs. In: Loewy AD, Spyer KM, eds., Central Regulation of the<br />

Autonomic Functions. London: Oxford University Press, pp. 3 13-333.<br />

19. Bradley WE, Tim GW, Scott FB (1974) Innervation of the detrusor muscle and<br />

urethra. UroZ Clin N Am 1:3-27.<br />

20. Bradley WE, Sundin T (1982) The physiology and pharmacology of the urinary<br />

tract <strong>dysfunction</strong>. Clin Neuropharmacol 5:131-158.<br />

21 a Bhatia NN, Bradley WE (1983) Neuroanatomy and physiology: innervation of<br />

the urinary tract. In: Raz S, ed., Female Urology, Philadelphia: WB Saunders,<br />

pp. 12-32.<br />

22. Tang PC (1955) Levels of the brainstem and diencephalon controlling micturition<br />

reflex. J Neurophysiol 18:583-595.


Chapter 3 I Cerebrovascular Accidents 81<br />

23. Tang PC, Ruch TC (1956) Localization of brainstem and diencephalic areas<br />

controlling the micturition reflex. J Comp Neurol 106:213-231.<br />

24. Nathan PW (1976) The central nervous connections of the bladder. In: Williams<br />

DI, Chisholm CD, eds., Scientijic Foundations of Urology, v01 2, Chicago, IL:<br />

Year Book Medical Publishers, pp. 51-58.<br />

25. Hald T, Bradley WE (1982) The nervous control of the urinary bladder. In:<br />

Hald T, Bradley WE, eds., The Urinary Bladder: Neurology and Urodynamics,<br />

Baltimore, MD: Williams and Wilkens, pp. 48-51.<br />

26. Borrie M, Campbell A, Caradoc-Davies TH, et al. (1986) Urinary incontinence<br />

after stroke: a prospective study. Age Aging 15: 177-18 1.<br />

27. Nishizawa 0, Ebina K, Sugaya K, et al. (1989) Effect of cerebellectomy on reflex<br />

micturition in the decerebrate dog as determined by urodynamic evaluation. Urol<br />

IntE 44:152-154.<br />

28. Khan 2, Starer P, Yang WC, et al. (1990) Analysis of voiding disorders in patients<br />

with cerebrovascular accidents. Urology 35:265-270.<br />

29. Motola JA, Mascarenas B, Badlani GH (1988) Cerebrovascular accidents: urodynamic<br />

and neuroanatomical findings. J Urol 139:512A.<br />

30. Tsuchida S, Noto H, Yamaguchi 0, et al. (1983) Urodynamic studies on hemiplegic<br />

patients after cerebrovascular accident. Urology 2 1 :3 15-3 18.<br />

31. Adams M, Baron M, Caston MA (1966) Urinary incontinence in the acute phase<br />

of cerebrovascular accident. Nurse Res 15:lOO-105.<br />

32. Brocklehurst JC, Andrews K, Richards B, et al. (1985) Incidence and correlates<br />

of incontinence in strokes patients. J Am Geriatr Soe 33:540-542.<br />

33. Burney T, Senapati M, Desai S, et al. (1996) Effects of cerebrovascular accident<br />

on micturition. Uro Clinics N Am 23:483-486.<br />

34. Lum SK, Marshall VR (1982) Results of prostatectomy in patients following a<br />

cerebrovascular accident. Br J Urol 54:186-189.<br />

35. Motola JA, Badlani GH (1990) Cerebrovascular accidents, urological effects and<br />

management. Clin Geriatr Med 6:55-58.<br />

36. Blaivas JG (1982) The neurophysiology of micturition: a clinical study of 550<br />

patients. J Urol 127:958-963.<br />

37. Wein A, Barrett DM (1982) Etiologic possibilities of increased pelvic floor electromyography<br />

activity during cystometry. J Urol 127:949-952.<br />

38. Coslett HB, Heilman KM (1986) Male sexual function: impairment after right<br />

hemisphere stroke. Arch Neurol 43: 1036-1039.<br />

39. Coletta EM, Murphy JB (1994) Physical and functional assessment of the elderly<br />

stroke patient. Am Fam Physician 49:1777-1785.<br />

40. Yalla S, Blute R, Waters W, et al. (1980) Urodynamic evaluation of prostatic<br />

enlargements with micturitional vesicourethral static pressure profiles. J Urol<br />

125~685-689.<br />

41. Lightner DJ, Barrett DM, Schmidt R, et al. (1998) Conticath: a simple new<br />

catheter designed for continence and volitional voiding past the obstructed urethra.<br />

J Urol 159:303.<br />

42. Oesterling JE, Epstein H. (1993) North America Urolume Study Group. J Urol<br />

149: 10.<br />

43. Corujo M, Badlani GH (1998) Uncommon complications of permanent stents.<br />

J Endo 12:385-388.<br />

44, Lepor H, Auerbach S, Puvas-Baez A, et al. (1992) A randomized multicenter<br />

placebo controlled study of the efficacy and safety of terazosin in the treatment<br />

of benign prostatic hyperplasia. J Urol 148:1467-70.


82 Marinkovic and Badlani<br />

45. Gillenwater JY, Conn RL, Chrysant SG, et al. (1995) Doxazosin for the treatment<br />

of benign prostatic hyperplasia in-patients with mild to moderate essential hypertension:<br />

A double blind, placebo controlled dose response multicenter study.<br />

J Urol 154: 110-1 14.<br />

46. Lepor H, The Tamsulosin Investigator Group (1995) Clinical evaluation of tamsulosin,<br />

a prostate selective alpha,, antagonist. J Urol 153:274A.<br />

47. Sravodimosk D, Goldfischer E, Klima W, et al. (1998) Transurethral microwave<br />

thermotherapy for management of benign prostatic hyperplasia: A single-institution<br />

experience. Urology 51:1008-1011.<br />

48. Beduschi M, Oesterling J (1998) Transurethral needle ablation of the prostate: A<br />

minimally invasive treatment for symptomatic benign prostatic hyperplasia. Mayo<br />

Clin Proc 73:696-701.<br />

49. Gilling P, Cass C, Cresswell M (3996) The use of the Holmium laser in the<br />

treatment of benign prostatic hyperplasia. J Endo 10:459-461.<br />

50. Issa MM, Townsend M, Jimine VK, Miller LE, Anastasia K (1998) A new<br />

technique of intraprostatic fiber placement to minimize thermal injury to prostatic<br />

urothelium during indigo interstitial laser thermal therapy. Urology 5 l( 1): 105-1 10.<br />

51. Martov A, Kilchukov 2, Gushchin B (1998) Long-term follow-up in interstitial<br />

laser coagulation in BPH treatment. J Endo 12(Supp 1):195.<br />

52. Hills CJ, Winter SA, Balfour JA (1998) Tolterodine. Drugs 55(6):813-820.<br />

53. Abrams P, Freeman R, Anderstrom C, Mattiasson A (1998) Tolterodine, a new<br />

antimuscarinic agent: as effective but better tolerated than oxybutynin in-patients<br />

with an overactive bladder. Br J Urol 81(6):801-810.<br />

54. Stanton S, Monga AK (1997) Incontinence in elderly women: is periurethral<br />

collagen an advance? Br J Obstet Gynaecol 104(2):154-157.<br />

55. Appell R (1998) Techniques and results in the implementation of the artificial<br />

urinary sphincter in women with type I11 stress urinary incontinence by a vaginal<br />

approach. Neurourol Urodynamics 7:613-617.<br />

56. Marinkovic S, Mian H, Evankovich M, et al. (1998) Burch versus the pubovaginal<br />

sling. Inti Urogynecol J 8:260-265.<br />

57. Wright EJ, Iselin CE, Webster G (1998) Pubovaginal sling using cadaveric allograft<br />

fascia lata for the treatment of intrinsic sphincter deficiency. J Urol 159(Supp<br />

1):215.<br />

58. Hom D, Desautel MG, Lumerman 5, et al. (1998) Pubovaginal sling using polypropylene<br />

mesh and Vesica bone anchors. Urology 5 1:708-7 13.<br />

59. Barbailias GA, Liatsikos EN, Athanasopoulos A (1997) Goretex sling urethral<br />

suspension in Type 3 female urinary incontinence: clinical results and urodynamic<br />

changes. Intl Urogynecol J 8:344--50.<br />

60. Rosen N, Ravalli R (1990) Intermittent self-catherization. Clinic Geriatr Med<br />

61101-107.


CONTENTS<br />

INTRODUCTION<br />

NEUROPHYSIOLOGY OF MICTURITION<br />

VIDEOURODYNAMICS<br />

TREATMENT OVERVIEW<br />

TRANSVAGINAL DENERVATION<br />

(INGELMANN-SUNDBERG PROCEDURE)<br />

SUBTRIGONAL PHENOL AND ALCOHOL INJECTIONS<br />

DETRUSOR MYOTOMY, BLADDER TRANSECTION<br />

AND CYSTOLYSIS<br />

DETRUSOR MYECTOMY<br />

(BLADDER AUTO-AUGMENTATION)<br />

AUGMENTATION CYSTOPLASTY & (WITH WITHOUT<br />

CONTINENT CATHETERIZABLE ABDOMINAL<br />

STOMA) AND CONTINENT URINARY DIVERSION<br />

SUMMARY<br />

REFERENCES<br />

INTRODUCTION<br />

Multiple sclerosis is a chronic demyelinating disease with a variable<br />

clinical course. It is most common in the third to fifth decades of life<br />

and has a female to male ratio of about 3: 1. In North America and<br />

Europe, the incidence is approximately 10 new cases per 100,000<br />

population between the ages of 20 to 50. The majority of patients have<br />

mild neurologic abnormalities, which tend to exacerbate and remit over<br />

time. The most severe cases are characterized by chronic progression,<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

83


84 Blaivas<br />

which may ultimately lead to quadriplegia. The clinical course is classified<br />

as acute, progressive, chronic, and benign (I).<br />

The etiology of multiple sclerosis is unknown, but the prevailing<br />

theory suggests that it is an autoimmune disorder (1,2). From a pathologic<br />

standpoint it is characterized by focal inflammatory and demyelinating<br />

lesions scattered throughout the nervous system (3). The demyelinating<br />

plaques have a predilection for the white matter of the cervical<br />

spinal cord, particularly the posterior and lateral columns (4,5). Since<br />

these are the pathways that subserve vesical and urethral function (64,<br />

bladder symptoms are prominent in the majority of patients (9-14).<br />

Bladder symptoms and urodynamic findings are subject to the same<br />

exacerbations and remissions as the underlying disease process.<br />

Some studies have reported that there is no correlation between<br />

bladder symptoms and either the duration of illness or age of the patient.<br />

However, others have shown a correlation between urinary symptoms<br />

and duration of disease and that as the disease progresses, urologic<br />

symptoms become common, eventually affecting at least 50% of men<br />

and 80% of women. In a series of 212 multiple sclerosis (MS) patients<br />

screened at a university clinic, Koldewijn et al. found that urologic<br />

symptoms correlated with the degree of disability, but not duration of<br />

disease (15). Over half of the patients with a Kurtzke score of greater<br />

than four suffered from urinary storage symptoms.<br />

Previously, it had been thought that urinary tract complications led<br />

to a high incidence of deaths in multiple sclerosis, ranging from 21-55%<br />

(16,17). At autopsy, Samellas et al. found that at autopsy, 55% of<br />

patients had hydronephrosis, or pyelonephritis that had directly contributed<br />

to their death (17). Leibowitz et al. on the other hand, reported<br />

that only 5% of deaths in MS patients were attributed to urologic<br />

complications (18). With development of sophisticated urodynamic<br />

testing, modern therapeutic interventions and more careful monitoring,<br />

urologic causes of death has decreased dramatically. However, the<br />

morbidity of lower urinary tract disorders, particularly incontinence,<br />

is still great. In a prospective study, Blaivas found that those who<br />

develop urologic complications were either women with severe incontinence<br />

who were managed with indwelling catheters or men who had<br />

DESD. Women who developed DESD did not seem to be at nearly so<br />

high a risk (19).<br />

Lower urinary tract symptoms in patients afflicted with multiple<br />

sclerosis are usually attributed to the underlying condition, yet this is<br />

not necessarily the case. Nonneurogenic urologic disorders such as<br />

cystitis, urolithiasis, bladder and prostate cancer, and benign prostatic


Chapter 4 I Multiple Sclerosis 85<br />

hypertrophy are probably as common in these patients as in the general<br />

population and require the same End of diagnostic evaluation. Once<br />

these urologic conditions are excluded, the diagnosis is best accomplished<br />

when there is a clear understanding of the underlying pathophysiology.<br />

This, in turn, requires knowledge about the neurophysiology<br />

of micturition and the effect of specific neurologic lesions on normal<br />

physiology.<br />

NEUROPHYSIOLOGY OF MICTURITION<br />

Normal voiding is accomplished by activation of the “micturition<br />

reflex.” This is a coordinated event characterized by relaxation of the<br />

striated urethral sphincter, contraction of the detrusor, opening of the<br />

vesical neck and urethra and the onset of urine flow (20). The micturition<br />

reflex is integrated in the pontine micturition center, which is located<br />

in the rostral brain stem (20,21,22). Interruption of the neural pathways<br />

connecting the “pontine micturition center” to the “sacral micturition<br />

center” usually results in detrusor external sphincter dyssynergia<br />

(23,20,24). Detrusor external sphincter dyssynergia is characterized by<br />

simultaneous involuntary contractions of the detrusor and the external<br />

sphincter. The involuntary detrusor contractions cause incontinence; the<br />

involuntary sphincter contractions result in bladder outlet obstruction.<br />

Detrusorexternal sphincter dyssynergia is commonly seen in patients<br />

with spinal cord involvement because of the demyelinating plaques,<br />

which are the hallmark of multiple sclerosis. Patients with untreated<br />

detrusorexternal sphincter dyssynergia are at high risk for developing<br />

urologic complications such as vesicoureteral reflux, hydronephrosis,<br />

urolithiasis, and urosepsis (19). When the neurologic lesion is above<br />

the pons, the “micturition reflex” is intact and when micturition is<br />

affected it is usually characterized by loss of voluntary control. This<br />

is usually manifest as detrusor hyperreflexia, but losing awareness of<br />

bladder events or concern about the consequences of incontinence are<br />

also common. Although these patients are usually incontinent, they are<br />

at little risk for developing urologic complications unless they are<br />

treated with an indwelling vesical catheter, or have other urologic<br />

abnormalities. Neurologic lesions that interfere with the sacral reflex<br />

arcs are less common. These lesions result in various combinations of<br />

detrusor areflexia, intrinsic sphincter deficiency, and paralysis of the<br />

striated urethral sphincter.


86 Blaivas<br />

Symptoms<br />

Lower urinary tract symptoms may be classified as storage symptoms,<br />

emptying symptoms or mixed storage, and emptying (25). Storage<br />

symptoms include urinary frequency, urgency, incontinence, nocturia,<br />

and pain. Emptying symptoms include hesitancy, difficulty starting,<br />

weak stream, and a feeling of incomplete bladder emptying. About<br />

three fourths of patients with multiple sclerosis complain primarily of<br />

storage symptoms. The remainder has emptying or mixed symptoms.<br />

Many studies have found that there is no correlation between urologic<br />

symptoms, neurologic findings, and urodynamic findings.<br />

Blaivas performed prospective urodynamic study in 67 consecutive<br />

patients with MS and urinary bladder symptoms (19). Treatment was<br />

individualized on the basis of the underlying pathophysiology and consisted<br />

of intermittent catheterization (21 %), observation (27%), surgical<br />

(12%), drugs (g%), voiding training (6%), and external condom drainage<br />

(6%). In 18 patients (27%), lesser forms of treatment were unsuccessful,<br />

and indwelling vesical catheters were required. Symptoms correlated<br />

poorly with urodynamic findings, and treatment based on symptoms<br />

alone would have been ineffective in over half the patients. It is interesting<br />

to note that both voiding symptoms as well as urodynamic findings often<br />

changes over the course of time. Wheeler and coworkers noted that 55%<br />

of patients had different urodynamic findings after repeat examination<br />

(26). Blaivas and associates found 15% of patients had markedly different<br />

urodynamic findings upon repeat studies (11). <strong>Voiding</strong> <strong>dysfunction</strong> has<br />

also been documented in asymptomatic patients with MS. Bemelmans<br />

et al. evaluated 27 patients with no urologic complaints and a definite<br />

diagnosis of multiple sclerosis (mean duration of diagnosis-5 yr) (27).<br />

Fifty-two percent (14/27) had urodynamic abnormalities and 50%<br />

required urologic follow-up and therapeutic intervention.<br />

Diagnostic Evaluation<br />

As always, evaluation begins with a history and physical examination.<br />

Urinary tract infection (UTI) is common is MS, and for this reason,<br />

UTI should be excluded by urinalysis before further evaluation commences.<br />

Urologic History<br />

The history begins with a detailed account of the precise nature of<br />

the patient’s symptoms. Each symptom should be characterized and


Chapter 4 / Multiple Sclerosis 87<br />

quantified as accurately as possible. When more than one symptom is<br />

present, the patient’s assessment of the relative severity of each should<br />

be noted. The patient should be asked how often he/she urinates during<br />

the day and night and how long helshe can comfortably wait between<br />

urinations. It should be determined why he/she voids as often as he/<br />

she does. Is it because of a severe urge or is it merely out of convenience<br />

or an attempt to prevent incontinence? The severity of incontinence<br />

should be graded. Does the patient lose a few drops or saturate his/<br />

her outer clothing? Are protective pads worn? Do they become saturated?<br />

How often are they changed? Is the patient aware of the act of<br />

incontinence or does he/she just find hidherself wet? Is there a sense<br />

of urgency first? If so, how long can micturition be postponed? Does<br />

urge incontinence occur? Does stress incontinence occur during coughing,<br />

sneezing, rising from a sitting to standing position, or only during<br />

heavy physical exercise? If the incontinence is associated with stress,<br />

is urine lost only for an instant during the stress or is there uncontrollable<br />

voiding? Is the incontinence positional? Does it ever occur in the lying<br />

or sitting positions? Is there difficulty initiating the stream requiring<br />

pushing or straining to start? Is the stream weak or interrupted? Is there<br />

postvoid dribbling? Has the patient ever been in urinary retention?<br />

In order to document the nature and severity of urinary incontinence,<br />

a micturition diary and pad test are most useful. Conceptually, a pad<br />

test will provide a semi-objective measurement of urine loss over a<br />

given period of time. A number of pad tests have been described<br />

(28,29), but none has met with widespread approval, mainly because<br />

of poor test-retest validation (30,31). The simplest pad test can be<br />

done by having the patient change his or her pads every 6 h for one<br />

representative 24 h period while he or she is taking Pyridium. The<br />

amount of staining on the pads is a rough estimate of the severity of<br />

the incontinence. Alternatively, the pads can be weighed and the total<br />

weight, minus the weight of an unused pad recorded in the patient’ S<br />

record as an estimate of the volume of urine loss (1 gm approximately<br />

equals 1 mL, of urine). We believe the pad test is very useful and<br />

recommend that, once completed, the patient simply be asked to state<br />

how representative it was. To this end, the patient is asked to rate the<br />

day, on a scale of 1-10, with respect to how bad the incontinence was<br />

on the day that the pad test was recorded. The main purpose of these<br />

tests is to grossly quantitate the relative severity of the incontinence,<br />

so that it can be determined whether the symptoms are being accurately<br />

reproduced during subsequent examinations. Although there may be<br />

great variability in the actual data accumulated by these instruments,


88 Blaivas<br />

simply asking the patient whether or not the diary and pad test were<br />

representative of a “good” or “bad” day will usually suffice for clini-<br />

cal purposes.<br />

Physical Examination<br />

The physical examination should focus on detecting anatomic and<br />

neurologic abnormalities that contribute to the patient’s symptoms. The<br />

nature of the incontinence should be determined by examining the<br />

patient with a full bladder as discussed in the following section entitled<br />

“eyeball urodynamics.” The neurourologic examination begins by<br />

observing the patient’s gait and demeanor as he first enters the office.<br />

A slight limp or lack of coordination, an abnormal speech pattern,<br />

facial asymmetry, or other abnormalities may be subtle signs of a<br />

neurologic condition. The abdomen and flanks should be examined<br />

for masses, hernias, and a distended bladder. Rectal examination will<br />

disclose the size and consistency of the prostate. The sacral dermatomes<br />

are evaluated by assessing anal sphincter tone and control, perianal<br />

sensation, and the bulbocavernosus reflex. With a finger in the rectum,<br />

the patient is asked to squeeze as if he were in the middle of urinating<br />

and trying to stop. A lax or weakened anal sphincter or the inability<br />

to voluntarily contract and relax are signs of neurologic damage, but<br />

some patients simply don’t know or don’t understand how to the contract<br />

these muscles, while others may be too embarrassed to comply with<br />

the instructions. The bulbocavernosus reflex is checked by suddenly<br />

squeezing the glans penis or clitoris and feeling (or seeing) the anal<br />

sphincter and perineal muscles contract. Alternatively, the reflex may<br />

be initiated by suddenly pulling the balloon of the Foley catheter against<br />

the vesical neck. The absence of this reflex in men is almost always<br />

associated with a neurologic lesion, but the reflex is not detectable in<br />

up to 30% of otherwise noma1 women (23).<br />

In men, the size and consistency of the prostate should be assessed.<br />

In women a vaginal examination should be performed with both an<br />

empty bladder (to check the pelvic organs) and a full bladder (to check<br />

for incontinence and prolapse) (33 not cited). With the bladder semifull<br />

in the lithotomy position, the patient is asked to cough or strain<br />

in an attempt to reproduce the incontinence. The degree of urethral<br />

hypermobility is assessed by the Q-tip test (34-37). The Q-tip test was<br />

devised as a method to differentiate type I stress incontinence from<br />

type I1 stress incontinence at a time when bead chain cystourethrograms<br />

were routinely used for this purpose. It is performed by inserting well


Chapter 4 / Multiple Sclerosis 89<br />

lubricated sterile cotton tipped swab gently through the urethra into<br />

the bladder. Once in the bladder, the Q-tip is withdrawn to the point<br />

of resistance, which is at the level of the vesical neck. The resting<br />

angle from the horizontal is recorded. The patient is then asked to<br />

strain and the degree of rotation is assessed. Hypemobility is defined<br />

as a resting or straining angle of greater than 30 degrees from the<br />

horizontal and is equivalent to stress incontinence type 11, provided<br />

that incontinence is demonstrated.<br />

Upper Urinu y l’rdct Evuluation<br />

Upper urinary tract evaluation includes measurement of serum BUN<br />

& creatinine and upper tract imaging (intravenous pyelogram (IVP),<br />

renal ultrasound, CAT scan or MRI). Renal ultrasound is recommended<br />

for screening because this is the easiest, least expensive and devoid of<br />

potential complication.<br />

Urodyndmic Evuluation<br />

There are two basic reasons for performing urodynamic investigation<br />

in patients with MS: to determine the precise etiology of the patient’s<br />

symptoms and to identify urodynamic risk factors for the development<br />

of upper urinary tract deterioration. Urodynamic risk factors include<br />

detrusorexternal sphincter dyssynergia, low bladder wall compliance,<br />

bladder outlet obstruction, and vesicoureteral reflux (1 9,3840).<br />

Urodynamic technique range from simple “eyeball urodynamics”<br />

to sophisticated multichannel synchronous video/pressure/~ow~M~<br />

studies. We believe that synchronous multichannel videourodynamics<br />

offers the most comprehensive, artifact free means of arriving at a<br />

precise diagnosis and we perfom them routinely in patients with multiple<br />

sclerosis.<br />

Before starting the urodynamic study the following information<br />

should be known:<br />

1. What symptoms are you trying to reproduce?<br />

2. What is the functional bladder capacity (the usual voided volume based<br />

on the voiding diary)?<br />

3. Does the patient usually empty his or her bladder (based on examinations<br />

& measurement of postvoid residual urine)?<br />

4. Is the uroflow normal, i.e., is there a likelihood of either urethral<br />

obstruction or impaired detrusor contractility?<br />

5. Is urinary incontinence a complaint? If so, the nature of the incontinence


90 Blaivas<br />

and the severity should be determined from a diary, pad test, and exam<br />

with a full bladder.<br />

6. Is there a neurologic lesion that could cause detrusorexternal sphincter<br />

dyssynergia, detrusor hyperreflexia, or detrusor areflexia?<br />

VIDEOURODYNAMICS<br />

Synchronous measurement and display of urodynamic parameters<br />

with radiographic visualization of the lower urinary tract videourody-<br />

namics is the most precise diagnostic tool for evaluating disturbances<br />

of micturition (41). In these studies, radiographic contrast is used as<br />

the infusant for cystometry. Depending on the level of sophistication<br />

required, other urodynamic parameters such as abdominal pressure,<br />

urethral pressure, uroflow, and sphincter electromyography may be<br />

recorded as well. There are important advantages to synchronous video/<br />

pressure flow studies compared to conventional single channel urody-<br />

namics and to conventional cystography and voiding cystourethrogra-<br />

phy . By simultaneously measuring multiple urodynamic variables one<br />

gains a better insight into the underlying pathophysiology. Moreover,<br />

since all variables are visualized simultaneously one can better appreci-<br />

ate their interrelationships and identify artifacts with ease.<br />

Common Urodynamic Patterns in MS<br />

Detrusor hyperreflexia is seen in over 50% of patients with MS<br />

(9,II, 12,42,43,44,45,46). Detrusor-external sphincter dyssynergia is<br />

found in about half or the patients with detrusor hyperreflexia<br />

(9,11,12,42,43,45,46,47). Detrusor areflexia is much less common.<br />

Most MS patients generally one of these distinct urodynamic pat-<br />

terns:<br />

1. Detrusor hyperreflexia without bladder outlet obstruction (Fig. l).<br />

2. Detrusor hyperreflexia and detrusor-external sphincter dyssynergia<br />

(Figs. 2 and 3).<br />

3. Detrusor hyperreflexia with vesical necldprostatic urethral obstruction<br />

(Fig. 4).<br />

4. Detrusor hyperreflexia and detrusor-external sphincter dyssynergia<br />

with vesical necldprostatic urethral obstruction (Fig. 5).<br />

5. Detrusor areflexia and normal bladder compliance (Fig. 6).<br />

6. Low bladder compliance.


Chapter 4 / Multiple Sclerosis 91<br />

Fig. 1. Detrusor hyperreflexia in a 54 year old woman with exacerbating &<br />

remitting multiple sclerosis. (A) Urodynamic tracing. At a bladder volume of 165<br />

m1 there was an involuntary detrusor contraction (arrow) which reached a maximum<br />

of 100 cm H20 (during sphincter contraction). Qmax = 13 ml/S and = 25 cm<br />

H20. Throughout the detrusor contraction, there was increased EMG activity as<br />

she tried to abort the detrusor contraction. When she relaxed, (at Qmax) there<br />

was complete ENIG silence. Unintubated uroflow (far right) was normal (19 ml/<br />

S). Flow = uroflow; Pves = vesical pressure; Pabd = abdominal pressure; Pdet =<br />

detrusor pressure (Pves-Pabd); EMG = EMG of the pelvic floor (surface electrodes).<br />

(B) X-ray obtained at Qmax show a normal urethra. Courtesy of Jerry<br />

G. Blaivas, MD, with permission


92 Blaivas<br />

r<br />

...................................<br />

. . . . . . . . . . . . . . . .<br />

, , 1 1 1 , 4 . , 1 , .<br />

..............<br />

.......... I IUW<br />

...........<br />

........... i f 0<br />

Fig. 2. Detrusor hyperreflexia and detrusor-external sphincter dyssynergia in a 56<br />

year old woman with multiple sclerosis. (C) Urodynamic tracing. At a bladder<br />

volume of 135 cni H20 there was an involuntary detrusor contraction (arrow)<br />

which reached a maxinium of 107 cm H20 with a Qmax = 5 ml/S. Throughout<br />

the detrusor contraction, there was increased EMG activity. Flow = uroflow;<br />

Pves = vesical pressure; Pabd = abdominal pressure; Pdet = detrusor pressure<br />

(Pves-Pabd); EMG = EMG of the pelvic floor (surface electrodes). (D) X-ray<br />

obtained at the height of the detrusor contraction. The proximal urethra is dilated.<br />

The point of obstruction is the external urethral sphincter, marked by the arrow.<br />

Courtesy of Jerp G. Blaivns, MD, with permission


Chapter 4 / Multiple Sclerosis 93<br />

I . . I , . . . t . l . . l . ~ . l . " . . . . . . , . . . * . . *<br />

* . . * . . l * . . . . . 4 .<br />

Fig. 3. Detrusor hyperreflexia and detrusor-external sphincter dyssynergia a in 53<br />

year old paraplegic man. (A) Urodynamic tracing. At a bladder volume of 165 ml<br />

there was an involuntary detrusor contraction (arrow) which reached a maximum of<br />

l00 cm H20 with a Qmax = 5 ml/S. Throughout the detrusor contraction, there<br />

was increased EMG activity. Flow = uroflow; Pves = vesical pressure; Pabd =<br />

abdominal pressure; Pdet = detrusor pressure (Pves-Pabd); EMG = EMG of the<br />

pelvic floor (surface electrodes). (E) X-ray obtained at the height of the detrusor<br />

contraction. The proximal urethra is dilated. The of point obstruction is the external<br />

urethral sphincter, marked by the arrow. Courtesy of Jerry G. BEaivas, MD,<br />

with permission


94 Blaivas<br />

Fig. 4. Detrusor hyperreflexia and vesical necWprostatic urethral obstruction in a<br />

50 year old quadriplegic man with multiple sclerosis. (A) Urodynamic tracing.<br />

At a bladder volume of 82 m1 there was an involuntary detrusor contraction<br />

(arrow) which reached a maximum of 88 cm H20and Qmax = 6 ml/S. Throughout<br />

the detrusor contraction, there was no change in EMG activity. Flow = uroflow;<br />

Pves = vesical pressure; Pabd = abdominal pressure; Pdet = detrusor pressure<br />

(Pves-Pabd); EMG = EMG of the pelvic floor (surface electrodes). (B) X-ray<br />

obtained at Qmax shows a diffusely narrowed prostatic urethra. Courtesy of Jerry<br />

G. Blaivas, MD, with permission


Chapter 4 I Multiple Sclerosis 95<br />

Fig. 5. Detrusor hyperreflexia and detrusor-external sphincter dyssynergia with vesical ne<br />

prostatic urethral obstruction in a 37 year old man with multiple sclerosis. (C) Urodynamic<br />

tracing. At a bladder volume of 128 cm H20 there was an involuntary detrusor contraction<br />

(arrow) which reached a maximum of 187 cm H20 with no flow at all. Throughout the det<br />

contraction, there was increased EMG activity. At pdetmax, was there no contrast was visualized<br />

in the urethra, indicating both vesical neck obstruction and DESD. The fall in detrusor<br />

pressure in the middle of the detrusor contraction occurred as the external sphincter relaxed a<br />

bit and the urethra opened a bit, but the uroflow was too low to be measured by the flowm<br />

Flow = uroflow; Pves = vesical pressure; Pabd = abdominal pressure; Pdet = detrusor pressure<br />

(Pves-Pabd); EMG = EMG of the pelvic floor (surface electrodes). (D) X-ray obtained at the<br />

height of the detrusor contraction. There is no visualization of the urethra at all, indicating<br />

obstruction of the prostatic urethra as well as the external sphincter. Courtesy of Jerry G.<br />

Blaivas, MD, with permission


96 Blaivas<br />

Fig. 6. Detrusor areflexia and normal bladder compliance in a 61 year old women<br />

with occult spina bifida. (A) Urodynarnic tracing. The bladder was filled to capacity<br />

(650 rnl) at which point she felt uncomfortably full. Pdetmax was 2 cm H20.<br />

When asked to void, she strained and the EMG activity increased. There was no<br />

detrusor contraction and no flow. Flow = uroflow; Pves = vesical pressure; Pabd =<br />

abdominal pressure; Pdet = detrusor pressure (Pves-Pabd); EMG = EMG of the<br />

pelvic floor (surface electrodes). (B) X-ray obtained at the height of straining.<br />

There is no visualization of the urethra at all and no flow. Courtesy of Jerry G.<br />

Blaivas, MD, with permission


Chapter 4 I Multiple Sclerosis 97<br />

Fig. '7. (DS j Low bladder compliance and vesicoureteral reflux a 48 in yearold neurologically<br />

normal man. (A) Urodynamic tracing. The bladder was filled to capacity (850 mlj<br />

at which point he felt uncomfortably full. Pdetmax during filling 50 was cm H20 and there<br />

was a steep rise in detrusor pressure. When asked to void, he had a detrusor contraction to<br />

75 cm H20, without any flow all at and no contrast appeared in the urethra. EMG activity<br />

remained unchnaged Flow = uroflow; Pves = vesical pressure; Pabd = abdominal pressure;<br />

Pdet = detrusor pressure (Pves-Pabdj; EMG = EMG of the pelvic floor (surface<br />

electrodes). (B) X-ray obtained the at end of bladder filling (prior to detrusor the contraction)<br />

shows grade 3 vesicoureteral reflux and no contrast in the urethra. Courtesy ofJerry<br />

G. Bluivus, MD, with permission


98 Blaivas<br />

TREATMENT OVERVIEW<br />

Although the ultimate goal of therapy is the restoration of normal<br />

voiding, this is usually not possible unless the neurologic abnormalities<br />

remit, or if the symptoms are caused by a coincidental urologic condition<br />

such as urinary tract infection or benign prostatic hypertrophy. The<br />

social and emotional consequences of urinary bladder symptoms, particularly<br />

incontinence, must always be weighed against the risks to the<br />

patients’ general health imposed by the proposed therapy. For example,<br />

a woman with mild, yet uncontrollable incontinence resulting from<br />

refractory detrusor hyperreflexia may be better managed with incontinence<br />

pads (which pose threat no to her health) than with an indwelling<br />

vesical catheter (which predisposes her to urinary infection).<br />

The unpredictable course of MS makes it difficult to know what to<br />

expect or how to plan for the future. If the general neurologic condition<br />

of the patient can be improved, the voiding <strong>dysfunction</strong> will generally<br />

also improve. The primary goal in managing a patient with MS is to<br />

control the symptoms and minimizing the disability (48). Treatment<br />

of each patient requires an open mind, a broad array of therapeutic<br />

strategies and the compassion and tenacity to tailor the therapeutic<br />

approach to the individual needs of the patient. Further, the potential<br />

for urologic and neurologic progression makes long-term urologic follow-up<br />

essential. We recommend that patients be followed biannually<br />

with renal and bladder ultrasound alternating with videourodynamics.<br />

The most effective treatment is predicated on a clear understanding<br />

several factors:<br />

l. The patients symptoms and how they impact on his or her daily life,<br />

2. The physical, psychosocial and environmental milieu that the patient<br />

experiences and, finally<br />

3. The underlying pathophysiology causing the symptoms (as determined<br />

by urodynamic studies).<br />

From a scientific standpoint, it is the underlying pathophysiology<br />

that directs treatment, but there are a number of adjunctive therapies<br />

and considerations that for many patients are equally important.<br />

These include:<br />

l.<br />

2.<br />

3.<br />

Identifying and treating remediable causes of MS exacerbations (see<br />

above),<br />

Techniques to maximize the patient’s independence and activities of<br />

daily living,<br />

Behavior modification to treat and deal with symptoms,


Chapter 4 / Multiple Sclerosis 99<br />

4. Biofeedback and pelvic floor exercises, and<br />

5. Absorbent pads and incontinence appliances.<br />

Remediuble Cuwes of MS Exucerbutions<br />

There are a number of remediable factors that can precipitate an<br />

exacerbation of MS; these should be sought and treated whenever<br />

possible. They include urinary tract infection, fatigue, increased body<br />

temperature, skeletal spasticity, psychic or physical trauma, and emotional<br />

stress (48,49).<br />

Mmimizing Activities of Dui4 Living<br />

For patients with disabilities, there are many environmental and<br />

behavioral interventions that can increase their independence and functionality.<br />

Some things require only common sense; others require the<br />

expertise of occupational and physical therapists, physiatrists, orthopedic<br />

surgeons, and makers of prosthetic devices. The consultation of<br />

these professionals should be sought whenever the need arises. For<br />

example, a patient with muscular weakness may benefit from sitting<br />

in a high, rather than a low chair to make it easier for him or her to<br />

rise from the sitting to standing position. The judicious use of canes<br />

and walkers should be obvious, but some patients have aversions to<br />

such aids and may need counseling. Depression, an ever present threat,<br />

should be identified and treated. Other assistive devices such as a Velcro<br />

pants zipper, a. tenodesis hand splint, a catheter with extension which<br />

will reach the toilet without having to transfer out of the wheelchair<br />

may allow a person to maintain independence and permit intermittent<br />

self catheterization which might otherwise be impractical.<br />

Behuvior Modzjicution<br />

Behavior modification is an effective technique for ameliorating<br />

lower urinary tract symptoms, which is useful in a number of circum-<br />

stances. Simple behavioral techniques, such as diary keeping, adjust-<br />

ment of oral intake of fluids and voiding by the clock, by decreasing<br />

the amount of urine that is in the bladder at any point in time, minimizes<br />

the likelihood of incontinence. Pelvic floor exercises, with or without<br />

biofeedback, not only has the potential to strengthen the sphincter, but<br />

also heightens the patient’s awareness about bladder and sphincter<br />

events. Contraction of the pelvic floor muscles and urinary sphincter


100 Blaivas<br />

can, in some patients, prevent momentary incontinence owing to stress<br />

and prevent or abort involuntary detrusor contractions.<br />

DDAVP, a synthetic analog of antidiuretic hormone has proven (in<br />

our unpublished experience) to be very useful in selected patients with<br />

nocturia and/or enuresis. As a general rule, by decreasing the amount<br />

of overnight urine production, the number of nocturnal voids and enure-<br />

sis can be reduced. Patients are asked to keep a baseline 24 h voiding<br />

diary. On the night before their first treatment, the patient takes 0.1<br />

mg DDAVP at bedtime and is scheduled for an office visit the next<br />

afternoon. This office visit is extremely important because complica-<br />

tions owing to water toxicity or overhydration, when they occur, usually<br />

do so about 12 h or so after the first dose. On that visit, the patient<br />

is specifically queried about the warning signs of toxicity-headache,<br />

dizziness, nausea, vomiting, shortness of breath, mental confusion etc.<br />

Vital signs are taken and serum electrolytes are drawn. If there no<br />

signs of toxicity, the patient can be titrated up to a maximum dose of<br />

about 0.4 mg HS.<br />

Kinn treated thirteen patients with advanced MS and urge urinary<br />

incontinence with DDAVP in a double-blind crossover study (50). They<br />

demonstrated that total voided volume during the 6 h test period was<br />

reduced from 440 mL to 325 mL during treatment with DDAVP<br />

(p


Chapter 4 / Multiple Sclerosis 101<br />

until it is absolutely clear that the voiding picture and urodynamic<br />

findings have stabilized over time, generally, a period of about 2 yr.<br />

For long term management, an indwelling catheter should only be<br />

considered as a last resort in those patients who are unable, or unwilling.<br />

to undergo continent urinary diversion or cutaneous ileo-cystostomy .<br />

In our judgement, the long term hazards of an indwelling catheter do<br />

not justify its use except in rare instances.<br />

Detrusor Hyperreflexia<br />

The most common treatment for detrusor hyperreflexia is anticholin-<br />

ergic medication with or without behavior modification and intermittent<br />

catheterization as the need arises (52). Many patients on these regimens<br />

can be stable for years with a reasonable quality of life. We generally<br />

begin with a short acting medication, such as oxybutinin, in order to<br />

test the effects of a single dose. The patient is given 2.5-5 mg oxybu-<br />

tynin, asked to drink and evaluated by measurement of uroflow and<br />

postvoid residual urine 2-4 hours later. If he or she voids with an<br />

acceptable flow and residual urine, an anticholinergic medication is<br />

prescribed, If residual urine is unacceptable, the patient is scheduled<br />

for another session to learn intermittent self catheterization. Whether<br />

or not intermittent catheterization is used, the medication is titrated to<br />

an appropriate dosage based on the patient’s symptoms, voiding diaries<br />

and pad tests.<br />

The two most coinmon anticholinergics that we use are oxybutynin<br />

chloride (dosage is 2.5-10 ing TID or QID) and tolterodine (2 mg<br />

BID). A long acting (once a day) formulation of oxybutynin has just<br />

been approved by the FDA, but there is no clinical experience with it<br />

as of the writing of this chapter. For some patients, anticholinergics<br />

have proven most useful when taken on a PRN basis. Tricyclic antide-<br />

pressants are useful as primary or adjunctive treatment. The prototype<br />

is imipramine. The usual starting dose is 10-25 mg HS which can be<br />

titrated up to a maximum of about 150 mg by increasing the dose<br />

every third day or so. In some patients the effects of tricyclics and<br />

anticholinergics are additive, and they may be used in combination.<br />

It is thought that anticholinergics and tricyclics work primarily by<br />

decreasing the frequency and amplitude of involuntary detrusor contrac-<br />

tions. However, the warning time (from the onset of the urge to void<br />

until the iiicontinent episode) remains unchanged (52). Thus, the useful-<br />

ness of these medications may be enhanced by a behavior modification


102 Blaivas<br />

program of timed voiding (prior to the anticipated onset of the unstable<br />

detrusor contraction).<br />

Denervution<br />

Although denervation procedures have been used for decades in an<br />

attempt to prevent detrusor overactivity, with a few exceptions, we do<br />

not believe the long or short term success, when weighed against<br />

potential complications, warrants their use. Even when initially success-<br />

ful, over the long term, many patients develop low bladder compliance<br />

that proves a greater threat to their health and well being than the<br />

original condition. A great variety of techniques have been described<br />

including sacral blockade (53), dorsal rhizotomy (54), transvaginal<br />

denervation (55,561, subtrigonal injection of phenol or alcohol (57-601,<br />

cystolysis, bladder tnyotomy, and transection (61). An intrathecal baclo-<br />

fen infusion pump has been recently used to reduce severe spasticity<br />

resulting from neurological disease. In addition to reducing spasticity,<br />

improvement in bladder capacity and a decrease in urgency and urge<br />

incontinence has also been reported (62). Longer trials of this technique<br />

will determine if the system is reliable long-term and if tolerance<br />

develops. The use of other intrathecal pharmacological agents to inhibit<br />

the unstable bladder is a possibility for the future.<br />

TRANSVAGINAL DENERVATI ON<br />

(INGELMANN-SUNDBERG PROCEDURE)<br />

In 1959, Ingelman-Sundberg described a transvaginal technique<br />

intended to accomplish partial denervation of the subtrigonal nerve<br />

supply to the bladder. He reported an 88% success rate in 34 women<br />

with urinary frequency, urgency, urge incontinence, and bladder pain.<br />

A subsequent series corroborated these results and cited a 6-15%<br />

recurrence rate in the two series (55). Several other authors cited success<br />

rates of about 50% (56). Wan and McGuire, in an abstract, reported<br />

that 72% of 62 patients were “either cured or significantly improved”<br />

at least 1 yr following surgery (63).<br />

Patient selection is of critical importance. Only those with refractory<br />

detrusor instability who respond favorably to a temporary nerve block<br />

should be considered candidates for the procedure. The nerve block is


Chapter 4 / Multiple Sclerosis 103<br />

performed by injecting a long acting local anesthetic (1% Marcaine)<br />

beneath the trigone through a transvaginal approach. Under digital<br />

guidance, a 22 gage spinal needle is inserted into the anterior vaginal<br />

wall at the bladder neck. The needle is directed laterally toward the<br />

vaginal fornix 1 cm lateral to the cervix to a depth of about 3 cm and<br />

5-10 mL of 1% Marcaine is injected. If the patient reports an overt<br />

clinical improvement in the 6-8 h after injection, surgery may be<br />

considered. If there is no beneficial response other options should be<br />

considered. If the patient develops significant residual urine or is unable<br />

to void after the local anesthesia, she should be warned that she has<br />

a higher than normal risk of requiring intermittent self catheterization<br />

postoperatively.<br />

The operative procedure is performed with the patient in the dorsal<br />

lithotomy position. Some surgeons perform cystourethroscopy and<br />

insert bilateral ureteral catheters to aid in the recognition of the transmural<br />

ureters and to avoid their injury during the dissection. A Foley<br />

catheter is placed transurethrally to decompress the bladder and to<br />

assist in identifying the vesical neck. A vertical midline or inverted<br />

“U” anterior vaginal incision is made from the midurethra to approx<br />

2-4 cm proximal to the bladder neck. The anterior vaginal wall is<br />

dissected free from the urethra and bladder and a long scissors is used<br />

to dissect the soft tissue between the bladder, beneath the trigone, and<br />

the anterior vaginal wall. Care must taken to avoid injury to the ureters<br />

at this point. In addition, it is important to confine the limits of the<br />

dissection to the lateral limits of the trigone and to be careful not to<br />

perforate the endopelvic fascia and enter the retropubic space lest stress<br />

incontinence develop postoperatively. The vaginal wall is closed with<br />

a running 20 chromic suture. A vaginal pack and Foley or percutaneous<br />

suprapubic cystotomy are placed and the patient is given a voiding<br />

trial the following day.<br />

Despite the encouraging results reported by the original investigators<br />

and by McGuire, transvaginal partial bladder denervation has never<br />

achieved a significant level of clinical utility and there have not been<br />

enough studies to determine its efficacy.<br />

SUBTRIGONAL PHENOL AND ALCOHOL INJECTIONS<br />

Subtrigonal phenol injection has been employed in the treatment of<br />

patients with refractory detrusor instability and sensory urgency, but


104 Blaivas<br />

lack of overall efficacy and the possibility of bladder and ureteral<br />

necrosis have all but eliminated its clinical use (64,65).<br />

Most initial reports suggested that the technique was effective<br />

(66,58,64,59,65). Harris reported that 60% of patients developed an<br />

acontractile detrusor, but 30% were complicated by vesicovaginal fistula<br />

after extravesical subtrigonal injection of 50% ethanol for detrusor<br />

instability in 10 patients (67). Subtrigonal injection has fallen out of<br />

favor because of recent disappointing outcomes. McInerney in a long<br />

term follow-up of 97 patients, had only 19% long-term success, 24%<br />

short term but unsustained benefit and 57% failure (60). In addition<br />

there was a 17% complication rate: urinary retention (1 transient, 7<br />

permanent requiring intermittent catheterization, 4 nerve palsies, erectile<br />

impotence in 1 of 9 men, and 2 bladder mucosal necrosis). Many<br />

of the phenol failures subsequently underwent augmentation cystoplasty<br />

as the definite treatment with success.<br />

DETRUSOR MYOTOMY, BLADDER TRANSECTION<br />

AND CYSTOLYSIS<br />

Bladder transection (cystocystoplasty) was introduced for the treatment<br />

of patients with refractory detrusor instability by Turner-Warwick<br />

and Ashken in 1967 (68). The procedure is essentially a circumferential<br />

bladder transection with immediate reconstruction. An incision was<br />

made through the full thickness of the bladder wall and perivesical<br />

tissue between a point 1 to 2 cm. lateral to each ureteral orifice. The<br />

incision through the bladder wall is then closed in 1 layer with absorbable<br />

suture (69,70). Various modifications were made involving partial<br />

versus complete bladder transection and cystolysis. In 1972 Mahony<br />

and Laferte reported on the use of multiple detrusor myotomies in<br />

the treatment of detrusor hyperreflexia refractory to pharmacological<br />

therapy (71). In cystolysis the bladder is circumferentially dissected as<br />

if a cystectomy were to be performed, but the bladder is left in place<br />

(72). Essenhigh and others have used this procedure in over 100 patients<br />

and reported satisfactory results in approx 75% (69,70,73,74).<br />

In all of these procedures, the initial reports were quite encouraging.<br />

Nevertheless, none of them ever achieved widespread acceptance and<br />

no long term results have been reported. Further, even the original<br />

investigators have largely abandoned them. In our opinion, there is no<br />

role for them in current clinical practice.


Chapter 4 / Multiple Sclerosis 105<br />

DETRUSOR MYECTOMY<br />

(BLADDER AUTO-AUGMENTATION)<br />

Detrusor myectomy was first reported by Cartwright and Snow in<br />

1989 in pediatric patients with neurogenic bladder (75). In this procedure,<br />

the bladder is exposed through an extraperitoneal approach and a<br />

plane developed between the detrusor and underlying mucosa. Detrusor<br />

muscle fibers on the anterior surface and dome of the bladder are<br />

excised being careful to leave the mucosa intact. If successful, detrusor<br />

myectomy obviates the need for enterocystoplasty with its attendant<br />

morbidity. Although initial reports, with short term follow-up have<br />

shown excellent success rates, much more clinical data is needed before<br />

the results of this procedure can be evaluated (7576).<br />

AUGMENTATION CYSTOPLASTY (WITH & WITHOUT<br />

CONTINENT CATHETERIZAl3LE ABDOMINAL STOMA)<br />

AND CONTINENT URINARY DIVERSION<br />

We believe that augmentation cystoplasty is the treatment of choice<br />

for MS patients with refractory detrusor hyperreflexia and/or low bladder<br />

compliance (77,78,79,80). However, it should only be considered<br />

when all conservative measures have failed in patients who are able<br />

and willing to accept permanent intermittent self catheterization should<br />

that prove necessary. A continent urinary diversion or augmentation<br />

cystoplasty with a continent abdominal stoma is especially useful for<br />

female paraplegics and quadriplegics and other patients who are physically<br />

unable to perform intermittent self catheterize through the urethra<br />

because of physical limitations.<br />

These procedures are far preferable to either urinary conduit diversions<br />

or an indwelling vesical catheter for both medical and psychosocial reasons.<br />

From a psychosocial viewpoint, a continent abdominal stoma<br />

relieves the patient of the burden of an external urinary drainage bag.<br />

From a medical standpoint, the creation of a large capacity, low-pressure<br />

internal urinary reservoir greatly reduces the chances of urosepsis, urolithiasis,<br />

hydronephrosis, and ultimately, renal failure. Further, these procedures<br />

result in improved self-image and sexual experiences (81).<br />

The purpose of augmentation cystoplasty is to create a urinary reservoir<br />

that permits the bladder to painlessly accept large volumes of<br />

urine at low filling pressures. If sphincter function is intact, no further


106 Blaivas<br />

operative procedure is necessary. However, when sphincteric incontinence<br />

is present a concomitant pubovaginal sling or a sphincter prosthesis<br />

may be required. If intermittent catheterization is not possible, a<br />

cutaneous ileo-cystostomy is a most reasonable alternative. In this<br />

procedure, a small section of terminal ileum is isolated to act as a<br />

conduit between the urinary bladder and the anterior abdominal wall;<br />

an ileal “bladder chimney.” This procedure was reported in 1957 by<br />

Cordonnier (82), but subsequently attention focused on augmentation<br />

cystoplasty and formal supravesical urinary diversion in treating those<br />

who might otherwise benefit from a bladder chimney.<br />

The creation of a bladder chimney is a major abdominal surgical<br />

endeavor, but is usually well tolerated by the patient. A twenty centimeter<br />

segment of distal ileum is isolated in standard fashion. The proximal<br />

end is spatulated widely and anastomosed to a similarly spatulated flap<br />

developed from the dome of the bladder. This wide anastomosis is<br />

required to avoid an hourglass deformity, which could impede egress<br />

of urine from the bladder. The distal end of the ileal segment is delivered<br />

to the anterior abdominal wall as the stoma of a Bricker ileal urinary<br />

conduit would be fashioned.<br />

The advantage of this procedure over standard cystectomy and ileal<br />

diversion is the avoidance of manipulation of the ureters, thereby eliminating<br />

the problems associated with uretero-intestinal anastomoses such<br />

as reflux, stenosis, devascularization, urinary extravasation, or complete<br />

anastomotic disruption. Urinary collection is achieved with a standard<br />

urostomy drainage bag, without the need for an indwelling urethral or<br />

suprapubic catheter, which would predispose to chronic infection, tissue<br />

erosion, neoplasia, or calculus formation. Bladder capacity is preserved,<br />

and the gradual destruction of the bladder associated with indwelling<br />

catheters is obviated by this effective drainage procedure. Furthermore,<br />

the procedure is potentially reversible, as the bladder remains in place,<br />

with the ureters in normal anatomic position (83).<br />

If there is sphincteric incontinence, concomitant in women pubovaginal<br />

sling usually suffices; in rare instances, when there has been urethral<br />

destruction, surgical bladder neck closure may be necessary in either<br />

sex (8485).<br />

Detrzlsor Hyperreflexia with Vesical lVech/Prostatic<br />

Urethral Obstruction<br />

In men with detrusor overactivity and prostatic obstruction, it is<br />

usually not possible with any degree of certainty to distinguish detrusor


Chapter 4 / Multiple Sclerosis 10’7<br />

hyperreflexia (owing to multiple sclerosis) from detrusor instability<br />

(owing to prostatic obstruction). Logic dictates that detrusor overacti-<br />

vity is more likely to be because of multiple sclerosis if the patient is<br />

unaware of the involuntary detrusor contractions, cannot abort them,<br />

and is incontinent. Our experience defies this logic, and we do not<br />

believe that we can reliably make the necessary. The practical conse-<br />

quence of this is that we are unable to predict whether or not a man<br />

will be incontinent after transurethral incision or resection. Accordingly,<br />

if this form of therapy is chosen, it must be done so with informed<br />

consent and contingency plans for managing postoperative urinary<br />

incontinence because of persistent detrusor overactivity.<br />

Detrusor Hyperreflexia With Detrusor-External<br />

Sphincter Dyssynergia<br />

Patients with detrusor hyperreflexia and DESD are at higher risk for<br />

upper urinary tract damage, stones, and urosepsis than the other two<br />

patterns (19,86). Further, when treated with anticholinergics they are<br />

more likely to develop urinary retention and require intermittent catheterization.<br />

For practical purposes, though, treatment is pretty much the<br />

same as outlined above. The most common treatment is anticholinergics<br />

and intermittent catheterization. In men, if intermittent catheterization<br />

is not an option, external sphincterotomy or an urethral stent may<br />

be considered. However, such destructive procedures should only be<br />

considered when there is no reasonable hope for neurologic regression<br />

and after more conservative therapies have been unsuccessful or are<br />

not possible. Different surgical techniques have been used for external<br />

sphincterotomy, using electrocautery, laser or a special urethrotome to<br />

incise one or more positions through the sphincter (87-91).<br />

Anatomically, the bulk of the striated sphincter is anteromedial,<br />

while the neurovascular bundles are lateral to the membranous urethra.<br />

Incisions at the three and nine o’clock endoscopic positions are associated<br />

with injury to the neurovascular bundles of the corporal bodies<br />

and reduced potency. The 12 o’clock sphincterotomy, described by<br />

Madersbacher et al. (87), and Yalla et al. (91) is the method of choice,<br />

because incision at this site decreases the risk of significant arterial<br />

hemorrhage and erectile <strong>dysfunction</strong>. Hemorrhage associated with the<br />

12 o’clock incision usually emanates from venous structures, and abates<br />

spontaneously with catheter placement.<br />

Complications of conventional external sphincterotomy include a<br />

reoperation rate of from 12-26%, hemorrhage requiring blood trans-


108 Blaivas<br />

fusion in 5-23%, and erectile <strong>dysfunction</strong> in from 2.8 to 64% of patients<br />

(88,90,92,93). Several minimally invasive surgical alternatives to traditional<br />

sphincterotomy have been recently developed - laser sphincterotomy<br />

and sphincter stent placement.<br />

Laser Sphincterotomy<br />

External sphincter ablation using the contact Nd:YAG (Neodymium:<br />

Yittrium Aluminum Garnet) laser permits carbonization as well as<br />

vaporization of tissue. Because of the near infra-red wavelengths used,<br />

minimal laser energy produces approx 4 mm of penetration. per pass.<br />

Encouraging laser sphincter results, with over 1 yr of follow-up, have<br />

recently become available (89).<br />

The Sphincter Stent<br />

Endoluminal urethral stent prostheses have recently been used as an<br />

alternative treatment for DESD. Several intraurethral stents are currently<br />

available, but only the UroLumeTM (American Medical Systems,<br />

Minnetonka, MN) has been systematically tested in the membranous<br />

urethral for the treatment of DESD. The Multicenter North American<br />

Trial data of 153 patients treated with the sphincter stent has shown<br />

promising results (94). The efficacy of the stent prosthesis approximates<br />

that of external sphincter destruction, with much less morbidity.<br />

Demor Hyperreflexia with Detrusor-External Sphincter<br />

Dyssynergia and Vesical NeckIProstatic Urethral Obstruction<br />

These patients are treated essentially the same as those with DESD<br />

except that, if external sphincterotomy or a urethral stent is used, the<br />

prostatic obstruction must be addressed at the sarne time. This can be<br />

accomplished by transurethral resection or incision of the prostate or<br />

by placing the stent through both the proximal urethral obstruction and<br />

the membranous urethra.<br />

Detrusor Areflexia<br />

Intermittent self catheterization is the simplest and most effective<br />

means of managing patients with detrusor areflexia. Over time, though,<br />

some of these patients develop low bladder compliance or detrusor<br />

hyperreflexia that requires a change in management (95,15). There is


Chapter 4 I Multiple Sclerosis 109<br />

no proven pharmacologic therapy for detrusor areflexia. Bethanechol<br />

chloride is, in our experience, ineffective for the treatment of detrusor<br />

areflexia. Although bethanechol can raise intravesical pressure, this<br />

does not imply that coordinated voiding occurs. In fact, urethral pressure<br />

also increases with bethanechol medications. Therefore the use of bethanechol<br />

in detrusor hyperreflexia does not make any sense, and its use<br />

in detrusor hyperreflexia with DESD is dangerous and is contraindicated.<br />

In patients who cannot be managed by intermittent catheterization,<br />

the only alternatives to an indwelling catheter are cutaneous urinary<br />

conduit diversion and cutaneous ileocystotomy.<br />

Concarrent Stress Incontinence<br />

In some patients with MS, particularly those with conus medullaris<br />

lesions, there is weakness and denervation of the pelvic floor striated<br />

sphincter muscles leading to sphincteric incontinence (95,96). Others<br />

have sphincteric incontinence because of anatomic abnormalities unre-<br />

lated to MS. In either case, there are many effective surgical treatments<br />

for sphincteric incontinence. However, because of the unpredictable<br />

nature of MS, there is probably a higher incidence of urinary retention<br />

and incontinence owing to detrusor overactivity in these patients.<br />

Because of this, it seems prudent to offer periurethral collagen injections<br />

as primary treatment. This serves a dual purpose-it offers the possibility<br />

of effective treatment, and it may serve as a proxy for the effects of<br />

surgical treatment.<br />

Lower urinary tract symptoms in patients afflicted with MS are<br />

very common. In general, treatment should be predicated on a clear<br />

understanding of the underlying pathophysiology, and for most patients,<br />

this means that urodynamic studies should be part of the evaluation.<br />

Because of the unpredictable course of MS, therapy should commence<br />

with the least invasive. Ablative or irreversible surgery should be<br />

reserved for patients with stable or progressive disease who have no<br />

reasonable hope for recovery in whom conservative therapy has failed.<br />

Most patients with exacerbating and remitting forms of MS will con-<br />

tinue in that pattern for long periods of time; these patients are best<br />

managed conservatively for as long as possible.


110 Blaivas<br />

1.<br />

2.<br />

3.<br />

4.<br />

5.<br />

6.<br />

7.<br />

REFERENCES<br />

McFarlin DE, McFarland HF (1982) Multiple Sclerosis, NEJM, 307:1183-1188.<br />

Poser, CM (1981) Multiple Sclerosis, A Critical Update, Eur Neurol 20:394<br />

MacDonald W1 (l 974) Pathophysiology in multiple sclerosis. Brain 97: 179-196.<br />

Fog T (1950) Topographic distribution of plaques in the spinal cord in multiple<br />

sclerosis. Arch Neurol Psychiatry 63:382.<br />

Oppenheimer DR (1978) The cervical cord in multiple sclerosis. Neuropathol<br />

App Neurobiol4: 15 1.<br />

Barrington FJF (1933) The localization of the paths serving micturition in the<br />

spinal cord of the cat. Brain 56126.<br />

Nathan PW, Smith NC (195 1) The centripetal pathway from the bladder and<br />

urethra within the spinal cord. J Neurol Neurosurg Psychiatry 14:262.<br />

Nathan PW, Smith NC (1958) The centrifugal pathway for micturition with the<br />

8.<br />

spinal cord. J Neurol Neurosurg Psychiatry 21 : 177.<br />

9. Anderson JT, Bradley WE (1976) Abnormalities of detrusor and sphincter function<br />

in multiple sclerosis. Br J UroE 48:193.<br />

10. Awad SA, Gajewski JB, Sogbein SK, Murray TJ, Field CA (1 984) Relationship<br />

between neurological and urological status in patients with multiple sclerosis.<br />

J Urol 132:499-502.<br />

11. Blaivas JG, Bhimani G, Labib KEl (1979) Vesicourethral <strong>dysfunction</strong> in multiple<br />

sclerosis. J Urol 122:342-347.<br />

12. Blaivas JG (1980) Management of bladder <strong>dysfunction</strong> in multiple sclerosis.<br />

Neurology 30: 12- 18.<br />

13. Miller H, Simpson CA, Yeates WE( (1965) Bladder <strong>dysfunction</strong> in multiple sclerosis.<br />

Br Med J 1:1265.<br />

14. Piazza DH, Diokno AC (1979) Review of neurogenic bladder in multiple sclerosis.<br />

Urology 14: 3 3.<br />

15. Koldewijn EL, Hommes OR, Lemmens WAJG, Debruyne MJ, Van Kerrebroeck<br />

PEV (1995) Relationship between lower urinarytract abnormalities and disease<br />

related parameters in multiple sclerosis. J Urol 154:169--173.<br />

16. Rheinhold M (1950) Prognosis in disseminated sclerosis. Br Med J 1 : 160.<br />

17. Samellas W, Rubin B (1965) Management of upper urinary tract complications in<br />

multiple sclerosis by means of urinary diversion to an ileal conduit. J Urol 93:548.<br />

18. Leibowitz U, Kahana E, Jacobsen SG, Alter M (1972) The cause of death in<br />

multiple sclerosis. In: Leibowitz U, ed. Progress in Multiple Sclerosis: Research<br />

and Treatment. New York, NY: Academic Press, 196.<br />

19. Blaivas JG, Barbalias CA (1984) Detrusor-external sphincter dyssynergia in men<br />

with multiple sclerosis: an ominous urologic condition. J Urd 131:91-4.<br />

20. Blaivas JG (1982) The neurophysiology of micturition: a clinical study of 550<br />

patients. J Urol 127:958.<br />

21. Bradley WE, Conway CJ (1966) Bladder representation in the pontine mesencephalic<br />

reticular formation. Exp Neurol 16:237.<br />

22. DeGroat WC, Booth AM, Krier J, Milner RJ, Morgan C, Nadelhaft I (1979)<br />

Neural control of the urinary bladder and large intestine. In: McBroods C, Koizumi<br />

K, Sato A, eds. Integrative Functions of the Autonomic Nervous System, Amsterdam,<br />

the Netherlands, Elsevier, Chap. 4.<br />

23. Blaivas JG, Sinha HP, Zayed AAH, Labib KB (1981) Detrusorexternal sphincter<br />

dyssynergia, J Urol 125:541.


Chapter 4 / Multiple Sclerosis 111<br />

24.<br />

25.<br />

26.<br />

McGuire EJ, Brady S (1979) Detrusor-sphincter dyssynergia. J Urol 121:774.<br />

Wein AJ (1981) Classification of neurogenic voiding <strong>dysfunction</strong>. J Urol 125:<br />

605.<br />

Wheeler JS, Siroky MB, Pavlakis AJ, et al. (1983) The changing neurourologic<br />

pattern of multiple sclerosis. J Urol 130: 123.<br />

27. Bemelmans BLH, Homes OR, Van Kenebroeck PEV, Lemmens WAJG, Doesburg<br />

WH, Debruyne FMJ (1991) Evidence for early lower urinary tract <strong>dysfunction</strong><br />

in clinically silent multiple sclerosis. J Urol 145: 1219-1224.<br />

28. Abrams PH, Blaivas JG, Stanton SL, Andersen JT (1988) Standardization of lower<br />

urinary tract function. Neurourol Urodynam 29:458.<br />

29. Hahn I, Fall M (199 1) Objective quantification of stress urinary incontinence: A<br />

short, reproducible, provocative pad-test. Neurourol Urodynam 10:475.<br />

30. Christensen SJ, Colstrup H, Hertz JB, et al. (1986) Inter- and intra-departmental<br />

variations of the perineal pad weighing test. Neurourol Urodynam 5:23.<br />

31. Lose G, Gammelgaard J, Jorgensen TJ (1986) The one-hour pad weighing test:<br />

Reproducibility and the correlation between the test result, start volume in the<br />

bladder and the diuresis. Neurourol Urodynam 5:17.<br />

32. Blaivas JG, Sinha HP, Zayed AAH, Labib KB (1981) Detrusor external sphincter<br />

dyssynergia: a detailed EMG study. J Urol 545.<br />

33. Blaivas JG, Hertitz DM, Romanzi LJ (1995) Early late versus repair of vesicovaginal<br />

fistulas: Vaginal and abdominal approaches. J Urol 153: 1 110.<br />

34. Bergman A, Bhatia NN (1987) Urodynamic appraisal of the Marshall-Marchetti<br />

test in women with stress urinary incontinence. Urology 29:458.<br />

35. Crystle C, Charme L, Copeland W (1971) Q-tip test in stress urinary incontinence.<br />

Obstet Gynecol 38:313.<br />

36. Montz FJ, Stanton SL (1986) Q-tip test in female urinary incontinence. Obstet<br />

Gynecol 67:259.<br />

37. Walters MD, Diaz K (1987) A study of continent and incontinent women. Obstet<br />

Gynecol 70(2):208.<br />

38. Ghoneim GM, Bloom DA, McGuire EJ, Stewar KL, (1989) Bladder compliance<br />

in meningomyelocele children. J Urol 141:1404.<br />

39. McGuire EJ, Woodside JR, Borden TA, Weiss RM (1981) The prognostic significance<br />

of urodynamic testing in myelodysplastic patients. J Urol 125:205.<br />

40. Zoubek J, McGuire EJ, Noff F, Delancey JOL (1989) Late occurence of urinary<br />

tract damage in patients successfully treated with radiation for cervical carcinoma.<br />

J Urol 141:1347.<br />

41. Blaivas JH, Chancellor M (1996) Atlas of Urodynamics, Williams & Wilkins, Baltimore.<br />

42. Bradley WE, Logothetis JL, Timm GW (1973) Cystometric and sphincter abnormalities<br />

in multiple sclerosis. Neurology 23: l 13 1.<br />

43. Goldstein I, Siroky MB, Sax DS, Krane RJ (1982) Neurourologic abnormalities<br />

in multiple sclerosis. J Urol 128:541-545.<br />

44. Philip T, Read DJ, Higson RH (1981) The urodynamic characteristics of multiple<br />

sclerosis. Br J Urol 53:672.<br />

45. Piazza DH, Diokno AC (1979) Review of neurogenic bladder in multiple sclerosis.<br />

Urology 14:33.<br />

46. Summers JL (1978) Neurogenic bladder in the woman with multiple sclerosis.<br />

J Urol 120:555.<br />

47. Ketelaer, 1977.


112 Blaivas<br />

48.<br />

49.<br />

50.<br />

51.<br />

52.<br />

53.<br />

54.<br />

55.<br />

56.<br />

57.<br />

58.<br />

59.<br />

60.<br />

61.<br />

62.<br />

63.<br />

64.<br />

65.<br />

66.<br />

67.<br />

68.<br />

Schneitzer L (1978) Rehabilitation of patients with multiple sclerosis. Arch Phv<br />

Men Rehnbil 59:430.<br />

Poser CM (1 980) Exacerbation, activity. and progression in multiple sclerosis.<br />

Arch Neiirol 37:47 I.<br />

Kinn AC, Larsson PO (1990) Desmopressin: A new principle for symptomatic<br />

treatment of urgency and incontinence in patients with multiple sclerosis. Scnizd<br />

J Urol Nephrol 24: 109-1 12.<br />

Hilton P, Hertogs K, Stanton SL (1983) The use of Desmopressin (DDAVP) for<br />

nocturia in women with multiple sclerosis. J Nertro Nezrroszrrg Psvdz 46:854-<br />

855.<br />

Wein, AJ, Nigro. DA (1996) Pharinacologic Therapy of Urinary Incontinence,<br />

Ch 3, in Evaluation and Treatment of Urinary Incontinence, Jerry G. Blaivas, ed..<br />

Igaku-Shoin, New York, pp. 46-68.<br />

Alloussi S. Loew F, Mast GJ, Wolf D (1984) Treatment of detrusor instability<br />

of the urinary bladder by selective sacral blockade. Br J Urol 56:464.<br />

Rockswold GL, Bradley WE (1978) The use of sacral nerve block in the evaluation<br />

and treatment of neurologic bladder disease. J Urol 18:415-417.<br />

Ingelman-Sundberg A ( 1978) Partial bladder denervation for detrusor dyssynergia.<br />

Clin Obstet Gyriecol 21 :797-805.<br />

Hodgkinson CP, Drukker BH (1 977) Intravesical nerve resection for detrusor<br />

dyssynergia. Act0 Obstet Gynecol Scnricl 56:401.<br />

Blackford HN, Murrary K, Stephenson TP. Mundy AR (1982) Results of transvesi-<br />

cal infiltration of the pelvic plexuses with phenol in 1 16 patients. Br J Urol56:647-<br />

649.<br />

Blackford HN, Murray K, Stephenson TP, et al. (1984) Results of transvesical<br />

infiltration of the pelvic plexuses with phenol in 116 patients. Br J Urol56:647-<br />

649.<br />

Cameron-Strange A, Millard RJ ( 1988) Management of refractory motor urge<br />

incontinence by transvesical phenol injection. Br J Urol 62:323.<br />

McInerney PD, Vanner TF. Matenhelia S, Stephenson TP (1 99 1) Assessment of<br />

the long-term results of subtrigonal phenolisation. Brit J Urology 67:586-587.<br />

Freiha FS, Stamey TA (1980) Cystolysis: A procedure for the selective denervation<br />

of the bladder. J Urol 123:360.<br />

Nanninga JB, Frost F and Penn R (1989) Effect of intrathecal baclofen on bladder<br />

and sphincter function. J UroZ 142101-105.<br />

Wan J, McGuire EJ. Wang SC, Cerny JC, Hodgkinson CP (1991) Ingelman-<br />

Sundberg denervation for detrusor instability. J Urol 145:Abstract #8 1.<br />

Parkhouse HF, Gilpin SA, Gosling JA, et al. (1987) Quantitative study of phenol<br />

as a neurolytic agent in the urinary bladder. Br J Urol 60:410-412.<br />

Wall LL, Stanton SL (1989) Transvesical phenol injection of pelvic plexuses in<br />

females with refractory urge incontinence. Br J Urol 63:465.<br />

Ewing R, Bultitude MT. Shuttleworth KED (1982) Subtrigonal phenol injection<br />

for urge incontinence secondary to detrusor instability in female. Br J Urol54:689-<br />

692.<br />

Harris RG. Constantinou CE, Stanley TA (1 988) Extravesical subtrigonal injection<br />

of 50 per cent ethanol for detrusor instability. J Urol 140: 11 1-1 16.<br />

Turner-Warwick RT, Ashen MH (1967) The functional results of partial. subtotal<br />

and total cystoplasty with special reference to ureterocaecocystoplasty. selective<br />

sphincterotomy and cystocystoplasty. Br J Urol 39:3.


Chapter 4 / Multiple Sclerosis 113<br />

69. Mundy AR (1980) Bladder transection for urge incontinence associated with<br />

detrusor instability. Br J Urd 52:480-483.<br />

70. Mundy AR (1982) The surgical treatment of urge incontinence of urine. J Urd<br />

128:481-483.<br />

71. Mahony DT, Laferte RD (1972) Studies of enuresis. IV. Multiple detrusor myotomy-a<br />

new operation for the rehabilitation of severe detrusor hypertrophy and<br />

hypercontractility. J Urol 107: 1064.<br />

72. Hindmarsh JR, Essenhigh SM, Yeates WK (1977) Bladder transection for adult<br />

enuresis. Br J Urol 49:5 15.<br />

73. Essenhigh DM, Yeates WK (1973) Transection of the bladder with particular<br />

reference to enuresis. Brit J Urol 45:299.<br />

74. Janknegt RA, Moonen WA and Schrinemachers LMH (1979) Transection of the<br />

bladder as a method of treatment in adult enuresis nocturna. Brit J Urol 51:275.<br />

75. Cartwright PC, Snow BW (1989) Bladder autoaugmentation: early clinical experience.<br />

J Urol 142:505.<br />

76. Cartwright PC, Snow BW (1995) Bladder autoaugmentation. Adv in Urol 8:273.<br />

77. Sidi AA, Reinberg Y, Gonzalez R (1986) Influency of segment of configuration<br />

of the outcome of augmentation enterocystoplasty. J Urol 136:1201-1204.<br />

78. Luangkhot R, Peng BCH, Blaivas JG (1991) Ileocystoplasty for the management<br />

of the refractory neurogenic bladder: Surgical technique and urodynamic findings.<br />

J Urol 145:1340.<br />

79. Linder A, Leach GE, Raz S (1983) Augmentation cystoplasty in the treatment of<br />

neurogenic bladder <strong>dysfunction</strong>. J Urd 129: 1007.<br />

80. Smith RB, Van Cangh P, Skinner DG, Kaufmann JJ, Goodwin WE (1977) Augmentation<br />

enterocystoplasty: A critical review. J Urd 118:35.<br />

81. Moreno et al. (1995).<br />

82. Cordonnier JJ (1957) Ileocystostomy for neurogenic bladder. J Urd 78:605-610.<br />

83. Rivas DA, Karasick S, Chancellor MB (1995) Cutaneous ileocystostomy (bladder<br />

chimney) for the treatment of severe neurogenic vesical <strong>dysfunction</strong>. Paraplegia,<br />

in press.<br />

84. Wan J, McGuire EJ (1990) Augmentation cystoplasty and closure of the urethra<br />

for the destroyed lower urinary tract. J Am Paraplegia Soc 13:40-45.<br />

85. Zimmern PE, Hadley HR, Leach GE, Raz S (1985) Transvaginal closure of the<br />

bladder neck and placement of a suprapubic catheter for destroyed urethra after<br />

long-term indwelling catheterization. J Urol 134:554-557.<br />

86. McGuire EJ, Savastano JA (1984) Urodynamic findings and clinical status following<br />

vesical denervation procedures for control of incontinence. J Urol132:87-88.<br />

87. Madersbacher H, Scott FB (1976) The twelve o’clock sphincterotomy: Technique,<br />

indications, results, Paraplegia 13:261-267.<br />

88. Perkash I (1976) Modified approach to sphincterotomy in spinal cord injury<br />

patients: indications, technique, and results in 32 patients. Paraplegia 13:247.<br />

89. Rivas DA, Chancellor MB, Staas WE, Gomella LG (1995) Contact Nd:YAG laser<br />

ablation of the external sphincter in spinal cord injured men with detrusor sphincter<br />

dyssynergia. Urology, in press.<br />

90. Schellhamer PT;, Hackler RH, Bunts RC (1973) External sphincterotomy: an<br />

evaluation of 150 patients with neurogenic bladder. J Urol 110:199-202.<br />

91. Yalla SV, Fam BA, Gabilondo FB, Jacobs S, DiBenedetto M, Rossier AB, Gittes<br />

RF (1977) Anteromedian external sphincterotomy: Technique, rationale and cornplications.<br />

J Urol 117:489-493.


114 Blaivas<br />

92.<br />

93.<br />

94.<br />

95.<br />

96.<br />

Lockhart JL, Vorstman B, Weinstein D, Politano VA (1986) Sphincterotomy<br />

failure in neurogenic bladder disease. J UroZ 135:86-89.<br />

Whitmore WF, Fam BA, Yalla SV (1978) Experience with anteromedian (12<br />

O’clock) external urethral sphincterotomy in 100 male subjects with neuropathic<br />

bladders. J UroE 50:99.<br />

Chancellor MB, Rivas DA(1994) Current Management of Detrusor Sphincter<br />

Dyssynergia in Advances. In McGuire EJ, ed. UroZogy Mosby-Year Book, Chi-<br />

cago, IL.<br />

Swash M, Snooks SJ, Chalmers DHK (1987) Parity as a factors in incontinence<br />

in multiple sclerosis. Arch NeuroZ44:504-508.<br />

Mathers SE, Ingram DA, Swash M (1990) Electrophysiology of motor pathways<br />

for sphincter control in multiple sclerosis. J Neuro Neurosurg Psych 53:955--960.


5 Diagnosis and Treatment<br />

of Spinal Cord Injuries<br />

and M~eloneuropat~ies<br />

Steven W: Sukin, MD<br />

dnd Timothy B. Boone, MD, PHD<br />

CONTENTS<br />

INTRODUCTION<br />

SPINAL-CORD INJURIES<br />

MYELODYSPLASIA<br />

OTHER DISEASES OF THE SPINAL CORD<br />

LONG-TERM UROLOGIC MANAGEMENT<br />

FOLLOW-UP PREVENTION OF COMPLICATIONS<br />

REFERENCES<br />

INTRODUCTION<br />

This chapter will cover the diagnosis and management of spinal-<br />

cord injuries and myeloneuropathies. In order to understand the basis<br />

of management, we will focus initially on the historical background,<br />

the neurobiology, and the general patterns of voiding <strong>dysfunction</strong> in<br />

spinal-cord injury and myelodysplasia. The general evaluation of void-<br />

ing <strong>dysfunction</strong> as it pertains to each particular lesion of the spinal<br />

cord will be discussed and we will review their respective management.<br />

Historically, the urologic complications of spinal-cord injury have<br />

been a primary source of morbidity and mortality. During World War<br />

I, the mortality after spinal-cord injury was approximately 80% owing<br />

to urinary tract infection and lack of antibiotics (I). In the past, urologists<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfinction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

115


116 Sukin and Boone<br />

had advocated the theory of a “Balanced Bladder” (2). Patients were<br />

felt to be clinically stable if they were able to void with minimal<br />

residual urine. This theory was not based on objective data and did<br />

not take into account the intravesical pressure elevations during urine<br />

storage and elimination. Improvements in urological care involving<br />

antibiotics, management of nephrolithiasis, the use of intermittant catheterization,<br />

the use of an objective means of evaluating the “Balanced<br />

Bladder” through urodynamics, and close life-long urologic follow-up<br />

have led to near-normal life expectancy for spinal-cord injury patients<br />

(I). Currently, major causes of mortality after spinal-cord injury are<br />

no longer owing to renal failure, but owing to pulmonary and cardiovascular<br />

complications. Despite these advancements in urologic care, urological<br />

complications continue to be a source of patient morbidity.<br />

The normal pattern of micturition requires an intact neural axis for<br />

controlled and coordinated lower urinary tract function. The Two-Phase<br />

Concept of Micturition (3) is a functional method of understanding<br />

this process. Initially, normal voiding function requires bladder filling<br />

and urine storage with a bladder able to accommodate increasing volumes<br />

of urine at low pressures and a bladder outlet that remains closed.<br />

Bladder emptying requires a coordinated and sustained contraction of<br />

the bladder muscle with concomitant relaxation of the external urethral<br />

sphincter. This micturition reflex is organized in the rostral brain stem.<br />

Parasympathetic and somatic components of the sacral spinal cord<br />

and sympathetic innervation from the thoracolumbar cord are highly<br />

integrated to regulate normal micturition (4).<br />

<strong>Voiding</strong> <strong>dysfunction</strong> occurs when there are abnormalities of filling,<br />

storage, and emptying (3). This may be a failure to store secondary to<br />

the bladder, the outlet, or both. Bladder filling can be affected by<br />

hyperactivity from phasic involuntary contractions, low detrusor cornpliance,<br />

or both. Storage failure can occur secondary to pain or hypersensitivity<br />

during filling. Outlet resistance may be decreased from any<br />

process that damages the innervation or structural elements of the<br />

smooth or striated urethral sphincter. A failure to empty may result from<br />

decreased bladder contractility, increased outlet resistance, or both.<br />

Animal studies have been used to uncover the micturition reflex<br />

pathways in neurally intact animals and compare normal function<br />

to the <strong>dysfunction</strong> following spinal-cord injury. These studies have<br />

shown that the afferent limb of the uninjured micturition reflex,<br />

travelling via myelinated A-delta fibers, switches to “silent” C-fiber<br />

afferents following spinal-cord injury (Figs. 1 and 2) (5-7). These


Chapter 5 / Spinal Cord Injuries 117<br />

CMG - Pressure Flow Study<br />

Fig. 1. Normal micturition reflex pathways with A-delta afferents mediating sen-<br />

sory input to the Pontine Micturition Center.<br />

11.11<br />

f Spinal Transection<br />

Pontine<br />

Micturition<br />

Center<br />

Fig. 2. Following Spinal Cord Injury C-Unmyelinated<br />

fibers take over to mediate<br />

the afferent limb of the neurogenic bladder.


118 Sukin and Boone<br />

findings also have been verified through electrophysiologic studies<br />

using capsaicin, a neurotoxin that disrupts C-fiber afferents (8). The<br />

rhythmic bladder contractions induced by bladder distention after<br />

spinal-cord injury are blocked by the intravesical instillation of<br />

capsaicin (5,9).<br />

After an injury to the spinal cord, the processes of edema, hemorrhage,<br />

and cord necrosis ensue. Wemmorrhagic necrosis primarily<br />

occurs in the gray matter of the cord and may extend both cephalad<br />

and caudad (10). The injury can be extensive or may be confined to<br />

several cord segments. Focal trauma induces rapid opening of the bloodspinal<br />

cord barrier (11) and the cord is exposed to a new environment<br />

including serum, platelets, neutrophils, monocytes, and macrophages<br />

with the subsequent release of cytokines. A vasogenic type of edema<br />

is present that includes transmitters from injured cells. Anoxia may<br />

occur after cord injury owing to decreased blood flow. A cascade of<br />

events develops by the accumulation of excitatory amino acids, of<br />

which glutamate is believed to be the primary one in the spinal cord (12),<br />

activation of various glutamine receptors, and increasing intracellular<br />

calcium (13). These processes lead to further cell injury, with neuronal<br />

damage and cell death.<br />

Regardless of the cause of a spinal-cord lesion, the general patterns<br />

of voiding <strong>dysfunction</strong> can be deciphered depending on the level of<br />

injury. An injury above the brainstem usually causes detrusor hyperreflexia.<br />

Injuries above the level of S2 most often cause detrusor hyperreflexia<br />

with detrusor external sphincter dyssynergia (DESD). Generally,<br />

complete lesions above this area, but below the area of the<br />

sympathetic outflow, result in detrusor hyperreflexia, smooth sphincter<br />

synergia, and striated sphincter dyssynergia. Lesions above the spinal<br />

cord level of T7 and T8 may result in smooth sphincter dyssynergia<br />

as well, Finally, lesions below the level of S2 most commonly cause<br />

detrusor areflexia with fixed external urethral sphincter tone.<br />

Many classification systems of voiding <strong>dysfunction</strong> have been<br />

used in the past. It is desirable to choose a system that is applicable<br />

for all types of voiding <strong>dysfunction</strong> and makes the treatment options<br />

fairly obvious by the category of <strong>dysfunction</strong> that is used. We feel<br />

that urinary bladder symptoms are best classified as fillinglstorage<br />

abnormalities, emptying abnormalities or combinations of the two<br />

(14). Using this functional classification, treatment can be directed<br />

at the underlying physiologic abnormality that is responsible for<br />

the symptom.


Chapter 5 / Spinal Cord Injuries 119<br />

SPINAL-CORD INJURIES<br />

The major causes of spinal-cord injury during peacetime are motor<br />

vehicle accidents, diving accidents, and falls. Other causes include disc<br />

prolapse, acute myelitis, surgery of thoracic aortic aneurysms, and<br />

occasionally aortography (15,161. Over 12,000 new cases of traumatic,<br />

spinal-cord injury occur each year in the United States, with an incidence<br />

estimated at 32 new injuries per million annually and a prevalence of<br />

906 cases per million (17). Eighty-five percent of all cases are in men.<br />

The majority of these injuries occur at or above the T-12 vertebral<br />

level. Approximately half of the injuries result in quadriplegia and the<br />

other half in paraplegia. Multiple-level injuries may occur, and even<br />

with a single-level injury, cord damage may not be confined to a single<br />

cord segment. Approximately 53.8% of injuries are incomplete and<br />

46.2% are complete.<br />

After a spinal-cord injury occurs, spinal shock is seen, in which<br />

there is a period of decreased excitability of the spinal-cord segments<br />

at and below the level of the lesion. Both flaccid paralysis of the skeletal<br />

muscle mass and an absence of visceral reflexes occur below the level<br />

of the lesion. This includes a suppression of autonomic and somatic<br />

activity. The bladder is areflexic and clinically develops urinary retention.<br />

Usually, reflex bladder activity returns within 2-12 wk, but may<br />

not return for as long as 6-12 mo. Most peripheral somatic reflexes<br />

of the sacral-cord segment, including the anal and bulbocavernosus,<br />

may never disappear, or if they do, may return within minutes or hours<br />

after the injury (18). Radiologically, the bladder has a smooth contour<br />

with no evidence of trabeculation. The bladder neck is usually closed<br />

and competent unless there has been prior surgery or in some cases of<br />

thoracolumbar injury where it remains open (19). By EMG, the maximum<br />

urethral closing pressure is lower than normal but is still<br />

maintained at the level of the external sphincter (20). A loss of voluntary<br />

control is, however, seen at the external sphincter. Overflow incontinence<br />

may develop, but because sphincter tone does continue, generally<br />

incontinence does not occur unless the bladder is grossly overdistended.<br />

The optimal initial management is intermittent catheterization.<br />

After a period of spinal shock, a recovery phase occurs in which<br />

there is a return of reflex activity. This stage of recovery usually lasts<br />

6-12 wk in complete suprasacral spinal-cord lesions, but may last up<br />

to a year or two (21). Generally, there is a return of detrusor contractility,<br />

if the distal spinal cord is intact but is simply isolated from higher


120 Sukin and Boone<br />

centers. Initially, the return of reflex activity is not strongly sustained<br />

and only produces low-pressure changes. The strength and duration of<br />

these involuntary contractions usually increases, producing involuntary<br />

voiding, usually with incomplete bladder emptying. This period is often<br />

manifested by involuntary voiding between catheterizations. Once there<br />

is no longer any sign of neurologic recovery, a stable phase is seen,<br />

in which the urodynamic pattern remains fairly stable over time. This<br />

stability in bladder function cannot be assumed, however. In many<br />

patients with detrusor areflexia, a loss of bladder compliance may<br />

develop over time with the development of high storage pressures.<br />

Both radiologic and urodynamic evaluations are essential in the care<br />

of patients with spinal-cord injury. Renal ultrasound has the advantage<br />

over intravenous pyelograms in following spinal-cord injury patients<br />

because it does not require a bowel prep, has reduced radiation exposure,<br />

and has no risk of anaphylaxis. Renal scintigraphy is useful in calculating<br />

the glomerular filtration rate, creatinine clearance, and differential<br />

renal function.<br />

The urodynamic evaluation is considered the best method of detecting<br />

and categorizing bladder <strong>dysfunction</strong> (22). Many sophisticated urodynamic<br />

modalities exist and the physician must decide which test is<br />

best based on availablity and experience. Urodynamics helps identify<br />

those patients at risk of developing urological sequelae and those<br />

patients that will require early intervention. One reason urodynamics<br />

is felt to be essential is that the type of voiding <strong>dysfunction</strong> a patient<br />

has after spinal-cord injury does not always correlate with the level of<br />

injury (22). Many patients suffer from multiple injuries, multiple levels<br />

of injury, or occasional occult brain injury. In 80% of patients, the<br />

clinical exam can give information on the type of neurogenic <strong>dysfunction</strong>,<br />

completeness of the injury, activity of the external sphincter, but<br />

urodynamics is required to accurately assess bladder pressures (23).<br />

The use of the ASIA (American Spinal Injury Association) Score and<br />

the SSEP (Somatosensory Evoked Potentials), which have been helpful<br />

in predicting overall neurologic recovery after spinal-cord injury, have<br />

shown to only correlate with the recovery of somatic nerve function<br />

(external sphincter) and not with the recovery of bladder function (24).<br />

The urodynamic findings of 489 consecutive patients with spinalcord<br />

lesions of varying causes have been retrospectively reviewed and<br />

have revealed three general patterns of voiding <strong>dysfunction</strong> depending<br />

on the level of injury (22). Detrusor hyperreflexia with synergistic<br />

external sphincter function is seen primarily after incomplete spinalcord<br />

injury; detrusor hyperreflexia with DESD is seen primarily after


Chapter 5 I Spinal Cord Injuries 121<br />

complete thoracic and cervical level lesions; and detrusor areflexia is<br />

seen after sacral and many lumbar injuries. Although there is a general<br />

correlation between the neurological level of injury and the expected<br />

vesicourethral function, this analysis verifies that the correlation is not<br />

absolute, and clinical neurological examination alone is not adequate<br />

to predict neurourological <strong>dysfunction</strong>. A urodynamic evaluation provides<br />

a more accurate diagnosis.<br />

Detrusor hyperreflexia may evolve into forceful, prolonged detrusor<br />

contractions from brief, low-level contractions as reflex activity returns<br />

(25). This is felt to be an exaggerated reflex response to bladder filling<br />

after suprasacral trauma from the collateral formation of new neural<br />

pathways, the loss of inhibitory impulse transmission, or the emergence<br />

of more primitive alternative pathways (26).<br />

Detrusor external sphincter dyssynergia is caused by a lesion that<br />

disrupts the coordination of signals between the pontine mesencephalic<br />

reticular formation center and the sacral reflex center of the spinal cord.<br />

Most patients with suprasacral spinal-cord injuries have evidence of<br />

DESD (22). With DESD, there is an inappropriate elevation in activity<br />

of the external urethral sphincter during an involuntary detrusor contraction<br />

(Fig. 3). This causes a functional obstruction with poor emptying<br />

and elevated detrusor pressures (27,28). Ten to twenty percent of men<br />

with. spinal-cord injury and DESD also demonstrate dysynergia of the<br />

bladder neck internal sphincter when the injury is above the level of<br />

T-6 (29). Patients with injuries to the sacral cord usually develop<br />

detrusor areflexia, with low pressures at volumes up to 5OOccs. This<br />

initial bladder compliance does not always remain unchanged. subset A<br />

of patients will develop low compliance and may represent a complex<br />

response to neurologic decentralization (20,30). This increased tone<br />

has been shown in animal studies with lower motor neuron injuries,<br />

in which there is an adrenergic overgrowth in the bladder (31). The<br />

classic outlet finding after sacral spinal-cord injury is a competent but<br />

nonrelaxing smooth muscle sphincter and a striated sphincter that retains<br />

some fixed tone but is not under voluntary control (18,19). There is<br />

little consensus on the function of the bladder neck or smooth sphincter<br />

after a sacral injury, however.<br />

The goal of management of patients with spinal-cord injury is to<br />

avoid high-pressure voiding, maintain detrusor compliance, and prevent<br />

acute sepsis, autonomic dysreflexia, and the deterioration of renal function.<br />

Urologic care should always begin in the acute period after injury.<br />

Initial management usually involves the use of an indwelling catheter,<br />

but the method chosen has been shown to have no significant adverse


122 Sukin and Boone<br />

r"urethrr;r<br />

EM<br />

1<br />

t l It.<br />

Fig. 3. Videourodynamic image of Detrusor Hyperreflexia with Detrusor External<br />

Sphincter Dyssynergia (DESD).


Chapter 5 / Spinal Cord Injuries 123<br />

effect on eventual outcome (32). Clean intermittent catheterization<br />

(CIC) is most desirable once the patient is stable. CIC leaves the bladder<br />

without foreign material and provides regular complete emptying of<br />

the bladder (33). Irreversible surgery is usually not recommended in<br />

the first year after injury before the patient’s bladder function has<br />

stabilized. Once the patient recovers from spinal shock, a detailed<br />

evaluation of the lower urinary tract is carried out. Further discussion<br />

of the management of spinal-cord injury will occur later in this chapter.<br />

MYELODYSPLrzSL4<br />

Neural tube defects are the most common cause of neurogenic bladder<br />

<strong>dysfunction</strong> in children. The incidence is approximately 1 in 1000<br />

in the United States but varies by geography (34). Over the last decade,<br />

this incidence has declined (35). The development of the spinal cord<br />

and vertebrae begin approximately on the eighteenth day of gestation<br />

and the closure of the canal is completed by day 35. Spina bifida<br />

represents a spectrum of abnormalities in the development of the neural<br />

tube and spinal cord. Spina bifida occulta occurs when only a bony<br />

defect is seen. A meningocele is defined as a meningeal sac with intact<br />

neural elements. Myelomeningocele or spina bifida cystica contains<br />

neural elements within a skin covered intact sac, whereas the sac is<br />

open in spina bifida aperta. Myelomeningocele account for over 90%<br />

of all open dysraphic states (36). Most spinal defects occur at the level<br />

of the lumbar vertebrae, followed by sacral, thoracic, and cervical (35).<br />

The underlying etiology of myelodyplasia is thought to be multifactorial.<br />

Epidemiological studies have shown that an environmental component<br />

is likely to be involved as the incidence of neural dysraphisms<br />

varies with geography and different periods of time. Evidence for<br />

genetic sources are many with an increase in female preponderance,<br />

ethnic variability, and familial tendencies. Myelodysplasia is also associated<br />

with Mecklel’s Syndrome, trisomies 13 and 18, and cloacal<br />

exstrophy (37). In addition, there is a 2-5% chance that a second sibling<br />

will be born with the same condition when spina bifida is already<br />

present in a family member (38). A periconceptual folic acid deficiency<br />

has been implicated as a cause of spina bifida. One multicenter study<br />

reported a 72% reduction in the recurrence of newborn neural tube<br />

defects with women who received a 4 mg supplement of folic acid<br />

during the periconceptual period (39). Other studies have shown that


124 Sukin and Boone<br />

the risk of occurrence of first-time neural tube defects decreased by<br />

60% by the use of folic acid (40).<br />

The diagnosis of myelomeningocele is most often made at the time<br />

of delivery. A prenatal diagnosis can be made by ultrasound (41).<br />

Elevated alpha-fetoprotein (AFP) by amniocentesis may also make<br />

the diagnosis more suspect (42), although the use of AFP has been<br />

controversial since both false positives and false negative results have<br />

been reported (43). While the role of elective cesarean section in fetuses<br />

with myelodysplasia may reduce the risk of intrauterine trauma to the<br />

neural elements (44), this remains controversial (45).<br />

A full urological evaluation is required in all newborns with spinalcord<br />

dysraphisms. Initially, the neonatal assessment begins with a<br />

detailed physical examination. It is important to document neurologic<br />

status with attention to abdominal muscle tone, sphincter tone, function<br />

of the sacral reflex arc, and lower extremity function. A urinalysis,<br />

urine culture, and serum creatinine should be obtained. Documentation<br />

of the renal anatomy can provide baseline information about the radiologic<br />

appearance of the upper and lower urinary tracts as well as the<br />

condition of the sacral spinal cord and the central nervous system<br />

(CNS). These initial studies are important for many reasons. They help<br />

identify babies at risk of urinary tract deterioration, help the physician<br />

counsel parents about the child's future bladder and sexual function,<br />

and can be used to compare with later studies (46).<br />

Renal ultrasound should be done as soon as possible to document<br />

the state of the upper tracts and thickness of the bladder wall. The<br />

postvoid residual urine volume can be helpful in assessing the completeness<br />

of spontaneous voiding by the neonate. The ultrasound is also an<br />

important means of examining the entire spine to rule out other neural<br />

tube defects (43). Studies have shown that less than 10% of newborns<br />

with spina bifida have hydronephrosis. A voiding cystourethrogram<br />

(VCUG) is also mandatory after a negative urine culture has been<br />

documented. This examination should be done promptly if hydronephrosis<br />

is seen on ultrasound. Sixteen percent of newborns with myelodysplasia<br />

have reflux while '23.5% have a bladder diverticulum.<br />

Urodynamics may be delayed in the newborn until it is safe to<br />

transport the child to the urodynamic suite and place the baby in the<br />

supine position for the test. This evaluation is important as the level<br />

of the bony defect will not necessarily determine the functional cord<br />

level. The region of involvement may be affected by the degree of<br />

involvement of the spinal cord and nerve roots, CNS anomalies (hydrocephalus,<br />

Arnold Chiari malformation), and surgical trauma during sac


Chapter 5 / Spinal Cord Injuries 125<br />

Fig. 4. Areflexic bladder with loss of compliance and an open bladder neck<br />

throughout filling.<br />

closure (47’). The typical patient with myelodysplasia shows<br />

an areflexic<br />

bladder with an open bladder neck (Fig. 4). The bladder generally fills<br />

until the resting fixed external sphincter pressure is reached, and then<br />

leaking ensues (detrusor leak-point pressure) (48). Urodynamic testing<br />

during the neonatal period may also have prognostic value in regard<br />

to the upper tracts (46). Patients with vesical filling pressure below 4.0


126 Sukin and Boone<br />

mmHg do not usually have vesicoureteral reflex, whereas 8 l% of<br />

patients with pressures greater than 40 mmHg have been noted to have<br />

ureteral dilation and 68% to have vesicoureteral reflux.<br />

Management of the neonate is based on the urologic evaluation<br />

including urodynamics. Diapers are acceptable and optimal if the neo-<br />

nate voids spontaneously. The Crede maneuver generally should be<br />

avoided, for reasons which will be discussed later in this chapter, but<br />

can be used initially if it is effective at emptying the bladder until the<br />

lower urinary tract can be evaluated. If the infant cannot empty his<br />

bladder by spontaneous voiding or Crede, then intermittant catheteriza-<br />

tion is started even before urodynamics is done. The neonate should<br />

be placed on antibiotics until the absence of reflux has been documented<br />

by VCUC.<br />

The goal of management in children with neurogenic bladder dys-<br />

function is to preserve renal function, prevent urinary tract infections,<br />

and allow socially acceptable continence. After puberty, many patients<br />

with myelodysplasia will note an improvement in their continence.<br />

These patients are less tolerant of any degree of incontinence, however<br />

(48). In the adult population, the management can be more complex<br />

because of prior surgery, loss of bladder compliance, and upper tract<br />

<strong>dysfunction</strong>.<br />

OTHER DISEASES OF THE SPINAL CORD<br />

Any lesion of the cord can potentially cause voiding <strong>dysfunction</strong>.<br />

There are many rare and uncommon causes such as tabes dorsalis,<br />

pernicious anemia, transverse myelitis, central cord syndrome, and<br />

polio. Spinal-cord tumors, tuberculosis, and arteriovenous malforma-<br />

tions involving the cord are more common sources of spinal-cord<br />

pathology.<br />

LONG-TERM UROLOGIC MANAGEMENT<br />

In order to best manage a patient with voiding <strong>dysfunction</strong> caused<br />

by a spinal-cord lesion or injury, it is important to consider the patient' s<br />

age, sex, level of lesion, degree of ambulation, manual dexterity, and


Chapter 5 / Spinal Cord Injuries 127<br />

independence. On the basis of both the clinical and urodynamic evaluation,<br />

as mentioned previously, urinary bladder symptoms may be classified<br />

as fillinghtorage abnormalities, emptying abnormalities or a combination<br />

of these (14,19), and treatment can be directed at the underlying<br />

urodynamic abnormality. The goal of management is to preserve renal<br />

function, avoid infection, provide freedom from catheters, and maintain<br />

urinary continence. Urologic management is directed towards achieving<br />

complete emptying of the bladder with low pressures. If bladder storage<br />

pressures are suitably low or can be made suitably low by nonsurgical<br />

means, the problem can be treated primarily as an emptying failure,<br />

and intermittent catheterization can be used, when practical, as a safe<br />

and effective way of satisfying many of the goals of treatment.<br />

For patients with a lower motor neuron lesion and an areflexic<br />

bladder, clean intermittent catheterization is the treatment of choice.<br />

Self-intermittent catheterization has been associated with few complications<br />

and provides periodic complete emptying of the bladder (50).<br />

Reports have shown that 87.5% of patients on CIC have stable upper<br />

tracts (Cass). These same reports have shown an associated decrease<br />

in vesicoureteral reflux in 75%, with 90% of patients maintaining sterile<br />

urine. Moreover, approx 34-49% of patients remain dry on this regimen.<br />

Still, there are complications seen in male patients with greater than<br />

five years of intermittent catheterization (23), including vesicoureteral<br />

reflux, stone disease, pyelonephritis, and even loss of kidney function.<br />

The use of a chronic indwelling catheter is never desirable because<br />

of its complications including epididymitis, urethrocutaneous fistula,<br />

traumatic hypospadias, and squamous cell carcinoma. Yet, chronic<br />

catheterization remains the most common form of management in<br />

patients who are tetraplegic and bedridden. Many female patients,<br />

unable to use an external collecting device, are managed with catheter<br />

drainage because they fail pharmacologic therapy and/or have limited<br />

hand function. McGuire followed 35 women managed with either an<br />

indwelling catheter or CIC for 2-12 yr following spinal-cord injury and<br />

found a significant reduction in the incidence of autonomic dysreflexia,<br />

febrile UTIs, pyelonephritic scarring by IVP, and bladder stones in<br />

patients managed with intermittent catheterization. This same study<br />

showed 92% of women with long-term indwelling catheters eventually<br />

had incontinence around the catheter and 54% had urethral erosion,<br />

whereas none on CIC had these complications (51). Catheterization or<br />

adult diapering with proper skin care in the individual unwilling or<br />

unable to catheterize is a better alternative to chronic catheterization.


128 Sukin and Boone<br />

The Crede maneuver and “trigger voiding” (a reflex contraction<br />

initiated by manual stimulation over a sacral or lumbar dermatonie)<br />

continue to be popular methods of bladder management but are contrain-<br />

dicated in patients with high outlet resistance, decreased compliance,<br />

DESD, and severe vesicoureteral reflux. These methods of bladder<br />

emptying can cause vesicoureteral reflux and upper tract deterioration.<br />

External catheters are another alternative but contraindicated in patients<br />

with high detrusor leak-point pressures. Finally, a cystostomy tube is<br />

the simplest of all bladder diversions but carries many of the same<br />

risks as an indwelling urethral catheter (e.g., infection, stones, cancer).<br />

Pharmacologic therapy in the management of voiding <strong>dysfunction</strong><br />

after spinal-cord injury is based on our knowledge of the innervation<br />

of the bladder and urethra as well as receptor localization for drug action<br />

in the lower urinary tract. Urodynamic studies define the condition as<br />

a failure to store or empty and pharinacologic therapy is designed to<br />

fit the overall bladder management. In the patient with a hyperreflexic<br />

bladder, a variety of antispasmodic and anticholinergic medications are<br />

available and effective. In patients with suprasacral injury causing<br />

hyperreflexia with DESD, the goal of pharmacotherapy is to control<br />

reflex bladder activity so that intermittent catheterization can be per-<br />

formed with a low pressure system.<br />

Alpha blockers may reduce outlet resistance and improve bladder<br />

emptying, although there is no evidence that this therapy is effective<br />

in helping with DESD. The effect of alpha blockers on bladder function<br />

in the spinal-cord patient has been investigated prospectively and has<br />

been shown to improve detrusor compliance by as much as 73% and<br />

improve episodes of dy sreflexia and incontinence (52). This therapy<br />

would therefore be optimal in patients that cannot tolerate anticholiner-<br />

gic agents or have persistently high bladder pressures. This suggests<br />

that in the spinal-cord-injured patient, alpha blockers may have an<br />

effect either on the receptors in the detrusor muscle or a central effect.<br />

Sundin et al. have shown that the alpha adrenergic receptors that are<br />

normally found in the bladder base become prominent in the detrusor<br />

body after decentralization (53).<br />

The use of cholinergic agonists such as bethanechol to facilitate blad-<br />

der emptying in patients with areflexic bladders has not proven beneficial.<br />

While cholinergics have been shown to raise baseline bladder pressure<br />

and increase maximal detrusor pressure, they do not completely empty<br />

the bladder. Cholinergic agonists cause a simultaneous contraction of the<br />

bladder, bladder neck, and urethra, preventing coordinated micturition


Chapter 5 I Spinal Cord Injuries 129<br />

(54). Poor intestinal absorption of bethanechol is another reason that cholinergic<br />

agonists lack efficacy. Cholinergic agonists also have been<br />

shown to actually inhibit the release of acetylcholine from parasympathetic<br />

postganglionics via a presynaptic feedback mechanism (55-57).<br />

The potential use of capsaicin for the treatment of neurogenic detrusor<br />

hyperreflexia has been evaluated. A prospective, randomized study<br />

showed significant improvements in urinary continence, reduced voiding<br />

frequency, and general patient satisfaction (58). Urodynamic evaluation<br />

in patients who received capsaicin also had significant improvements<br />

in bladder capacity and maximum detrusor pressure vs those<br />

who received a placebo. Capsaicin is highly neurotoxic to unmyelinated<br />

C-fibers and has been used in the past for its analgesic properties. In<br />

this study, 100 mL of capsaicin in 30% ethanol was placed intravesically<br />

one time for 2-5 min, depending on the patients’ tolerance of the<br />

medication. Side effects are common and almost systematic with suprapubic<br />

pain, urgency, flushing, hematuria, and autonomic dysreflexia,<br />

but usually resolve within 2 wk after installation.<br />

If conservative methods fail, surgical alternatives may be required.<br />

Patients are selected for surgical management on the basis of a careful<br />

neurourological evaluation, patient compliance, and hand function. A<br />

sphincterotomy is indicated in patients with DESD and elevated<br />

intravesical storage pressure that can be associated with deterioration<br />

of the upper urinary tracts (46). Sphincterotomy should lower the detrusor<br />

leak-point pressure to an acceptable level, thus treating the <strong>dysfunction</strong><br />

primarily as one of emptying. The resultant storage failure can be<br />

handled with an external collecting device. For patients who are unable<br />

to self-cath, this procedure allows urinary drainage with low pressure,<br />

often significantly reducing postvoid residuals (59). In addition, there<br />

is evidence that 70-90% of patients show substantial improvement<br />

after sphincterotomy with a reduction in vesicoureteral reflux and upper<br />

tract deterioration (60). Use of this procedure has waned recently<br />

because of both the significant complications associated with this procedure<br />

and the long-term failure rates. Sphincterotomy, which involves<br />

the transurethral resection of the striated external urethral sphincter, is<br />

plagued with significant bleeding requiring blood transfusion in 5-23%<br />

(61-63). In addition, recurrent obstruction is reported in 12-26%<br />

(59,64) and erectile <strong>dysfunction</strong> in 2.8-64% (61,63,65,66). Less than<br />

50% of patients, moreover, have been reported to be successfully managed<br />

with a condom catheter after sphincterotomy, and eventually<br />

require a suprapubic tube (67).


130 Sukin and Boone<br />

Sphincter prostheses and balloon dilation of the external sphincter are<br />

more recent options associated with reduced morbidity and acceptable<br />

patient satisfaction. The use of an artificial sphincter must be carefully<br />

planned to avoid infection and erosion in patients with loss of sensation.<br />

Balloon dilation of the sphincter works but may not be a durable<br />

treatment. Long-term studies evaluating urethral stents are underway.<br />

Initial results have revealed similar effectiveness with sphincterotomy<br />

but reduced morbidity. A prospective comparison of the UroLume<br />

endourethral Wallstent prosthesis (American Medical Systems, MN)<br />

vs sphincterotomy in patients with DESD and voiding pressures greater<br />

than 60 mm H20 showed that sphincter prosthesis has similar efficacy<br />

but is easier technically, less morbid, and less expensive (68). Complications<br />

of sphincter prosthesis included migration of the prosthesis and<br />

bladder neck obstruction. Twenty-seven percent of patients in this study<br />

required a second prosthesis to bridge the entire external urethral<br />

sphincter.<br />

Bladder augmentation cystoplasty is designed to create a bladder<br />

with increased compliance and capacity. It is primarily used in patients<br />

with a hyperreffexic, small-capacity, poorly compliant bladder, in which<br />

there is evidence of deteriorating renal function or incontinence between<br />

catheterization, which is refractory to medical therapy. This operation<br />

is ideal in patients who can perform intermittent catheterization but<br />

are incontinent or have upper tract deterioration. Renal insufficiency<br />

is a contraindication to augmentation enterocystoplasty.<br />

Urinary diversion is used only as a last resort but may be a patient’s<br />

only viable alternative once the problems associated with chronic catheterization<br />

develop. An array of diversions such as the ileal conduit,<br />

ileovesicostomy, or continent diversion are available to the patient based<br />

on the patient’s underlying renal function, motivation, compliance,<br />

and hand function. Supravesical diversion is indicated when there is<br />

progressive hydronephrosis and intractable upper urinary tract dilation,<br />

recurrent upper urinary tract infections, and considerable difficulty with<br />

the patient performing inte~ittent catheterization. With any diversion,<br />

there is a high rate of long-term complications (69). The initial upper<br />

tract improvements seen in patients after ileal conduit diversion have<br />

been shown to give way to deterioration. Therefore, ileal loop diversion<br />

is a last resort treatment for neurogenic bladder <strong>dysfunction</strong>.<br />

Neurostimulation is a promising method currently under investigation<br />

using electrical stimulation to control micturition and urinary incontinence.<br />

This technique was initially introduced in the 1960s using<br />

implantable electrodes in the detrusor muscle (70). Both intradural and


Chapter 5 / Spinal Cord Injuries 131<br />

extradural techniques of sacral neurostimulation are available (71,72).<br />

<strong>Voiding</strong> <strong>dysfunction</strong> after suprasacral spinal-cord injury responds well<br />

to sacral-root stimulation but has not been effective in lesions at the<br />

level of the cauda equina and below. A major difficulty in this treatment<br />

has been the inability to simultaneously relax the striated external<br />

sphincter and make the detrusor contract. Dorsal rhizotomy with both<br />

an extradural or intradural approach has become an important component<br />

of neurostimulation as it reduces detrusor hyperreflexia, improves<br />

bladder capacity, improves detrusor responses to ventral root stimulation<br />

and contraction, and diminishes the spasticity of the pelvic floor<br />

and external sphincter. A review of a multicenter experience using<br />

the Finetech-Brindley sacral anterior root stimulator showed that the<br />

technique of complete intradural posterior sacral rhizotomies, which<br />

was much more common than an extradural approach, and the implantation<br />

of a ~ine~ech- rind ley sacral anterior root stimulator improved<br />

continence, evacuation of urine, and bladder capacity (73). A reduction<br />

in vesicoureteral reflux, urinary tract infections, and autonomic<br />

dysreflexia was also noted. The stimulator can also assist in defecation<br />

and help men achieve penile erections, although only one-third of these<br />

patients actually use the stimulated erections for coitus. Long-term<br />

evaluation and comparison of sacral anterior root stimulation vs intermittent<br />

catheterization must be made in terms of safety, efficacy, and<br />

cost. Complications of these procedures can arise from potential infection<br />

of these implanted medical devices or nerve electrodes, nerve<br />

destruction with rhizotomies, possible erectile <strong>dysfunction</strong>, and bladdersphincter<br />

dyssynergia (74).<br />

FOLLOW-UP P"ENTION OF COMPLICATIONS<br />

Annual urologic follow-up is necessary in all patients with spinal-<br />

cord injury or myeloneuropathies, regardless of the nature or ease of<br />

bladder management. In fact, approx 40% of all SCI patients would<br />

die of renal insufficiency if they were left completely untreated (7.5,76).<br />

The evaluation should rule out any chronic, symptomatic infection of<br />

the urinary tract or urolithiasis and include an upper urinary tract<br />

assessment with either an intravenous pyelogram or renal ultrasound.<br />

There is continuing debate on which of these radiologic evaluations<br />

is superior.<br />

Urinary-tract infections are the most common urologic complication<br />

in patients with spinal-cord injuries (77). All urinary infections in these


132 Sukin and Boone<br />

patients are considered complicated because an uncomplicated infection<br />

is defined as community acquired without structural or neurologic<br />

abnormalities. IJTIs are responsible for many uroseptic episodes in the<br />

SCI population. Many risks exist including instrumentation, intermittent<br />

catheterization, stones, urethral strictures, benign prostatic hyperplasia,<br />

and detrusor external sphincter dyssynergia. Other causes include bladder<br />

overdistention, vesicoureteral reflux, large postvoid residuals, and<br />

high voiding pressures. The aim in management is to prevent, not to<br />

treat. This is best done by ensuring the well-being of the patient,<br />

providing adequate urinary drainage, and preserving healthy bacterial<br />

commenseral environment. Antibiotics should only be used during<br />

episodes of clinical infection, not asymptomatic bacteriuria. Asymptomatic<br />

bacteriuria does not usually need to be treated unless the patient<br />

has either a urease-producing organism, vesicoureteral reflux, or both.<br />

Approximately 23% of patients have vesicoureteral reflux after spinal-cord<br />

injury, as reported by Bors (77a). Causes include a Hutch<br />

para-ureteral diverticula, urinary infection, and high intravesicle storage<br />

pressure compromising the ureterovesical junction (78). Reflux can<br />

lead to reflux nephropathy with subsequent renal impairment and<br />

chronic urolithiasis.<br />

Stone disease develops in 8-15% of patients with spinal-cord injury,<br />

often caused by urease-producing bacterial infections, forming primarily<br />

struvite stones. The highest incidence occurs in patients with indwelling<br />

catheters (79). Other causes include chronic infections, immobilization<br />

with hypercalcemia (80), increased age, and complete neurological<br />

lesions. Prompt recognition and treatment of stone disease is necessary<br />

to adequately eliminate infection from the urologic tract in these<br />

patients.<br />

Carcinoma of the bladder has an incidence of 2-10% in this population.<br />

Squamous-cell carcinoma is the most common type of malignancy.<br />

The primary risk is long-term use of an indwelling catheter and therefore<br />

should be avoided. Annual bladder surveillance in these patients has<br />

not influenced survival as most of these patients present with advanced<br />

disease. It is generally recommended that for patients with a 10-yr<br />

history of an indwelling catheter or greater, annual cystoscopy with<br />

bladder biopsy be performed along with evaluation of the upper tracts.<br />

Autonomic dysreflexia is a potentially life-threatening complication<br />

in these patients. This disorder represents an autonomic response, which<br />

is primarily sympathetic, to specific visceral stimuli in patients with<br />

spinal-cord injury above the level of T6 (81). An incomplete compensatory<br />

parasympathetic outflow will occur above the level of injury. This


Chapter 5 / Spinal Cord Injuries 133<br />

phenomena is more common in patients with cervical injuries, and<br />

common triggers include bowel and bladder distention. Symptoms may<br />

involve piloerection, diaphoresis, pounding headache, flushing above<br />

the level of the injury, and may be associated with sudden and severe<br />

hypertension accompanied by reflex bradycardia. Although bradycardia<br />

is most common, tachycardia and arrythmias may be present. Hypertension<br />

may be of varying severity from causing a mild headache to a<br />

seizure or life-threatening cerebral hemorrhage. Therefore, it is necessary<br />

to monitor the blood pressure in these patients with any visceral<br />

stimulation. Immediate management when dysreflexia occurs is to discontinue<br />

the procedure. A life-threatening episode can be terminated<br />

by Procardia l0 mg sublingual (82), which has also been shown to be<br />

capable of preventing this syndrome when given orally 30 min prior<br />

to the procedure (82). The use of general or spinal anesthesia may be<br />

used in refractory cases. Prophylaxis against dysreflexia includes the<br />

elimination of visceral stimulation and chronic alpha-adrenergic blockade<br />

(83). In addition, a number of ablative procedures have been used for<br />

intractable dysreflexia including sympathectomy, sacral neurectomy,<br />

rhizotomy, cordectomy, and dorsal-root ganglionectomy (81).<br />

REFERENCES<br />

1. Graham SD (1 981) Present urological treatment of spinal cord injury patients.<br />

J Urol 126:l-4.<br />

2. Bors E (1957) Neurogenic Bladder, Urol Sur 7:177-250.<br />

3. Wein A, Barrett D (1988) <strong>Voiding</strong> function and <strong>dysfunction</strong>: a logical andpractical<br />

approach. Chicago: Year Book Medical Publishers.<br />

4. Blaivias JG (1982) The neurophysiology of micturition: a clinical study of SS0<br />

patients. J Urol 127:958-963.<br />

S. de Groat WC, Booth AM, Yoshiyarnd, M (1993) Neurophysiology of micturition<br />

and its modification in animal models of human disease. In: Maggi CA, ed. The<br />

Autonomic Nervous System, vol. 3. London, Harwood Academic Publishers, 1993,<br />

pp. 227-290.<br />

6. de Groat WC, Nadelhaft I, Milne RJ, Booth AM, Morgan C, Thor K (1981)<br />

Organization of the sacral parasympathetic reflex pathways to the urinary bladder<br />

and large intestine. JAuton New Syst 3:13S-160.<br />

7. de Groat WC, Kawatani M, Hisarnitsu T, Cheng CL, Ma CP, Thor K, Steers W,<br />

Roppolo JR (1990) Mechanisms underlying the recovery of urinary bladder function<br />

following spinal cord injury. J Auton New Syst 30 (Suppl):S71”77.<br />

8. Janig W, Koltzenburg M (1990) On the function of spinal primary afferent fibres<br />

supplying colon and urinary bladder. J Auton New Syst 30 (Suppl):S89-96.<br />

9. McMahon S, Morrison J (1982) Spinal neurones with long projections activated<br />

from the abdominal viscera of the cat. J Physiol (Lond) 322: 1-20.<br />

10. Hughes J (1978) Puthology of the Spinal Cord. London: Saunders.


134 Sukin and Boone<br />

11. Schall T, Bacon K (1994) Chemokines, leucocyte trafficking, and inflammation.<br />

Curr Opin Immunol6:865-873.<br />

12. Fagg GE, Foster AC (1983) Amino acid neurotransmitters and their pathways in<br />

the mammalian central nervous system. Neuroscience 9:701-709.<br />

13. Rothman S, Olney J (1987) Excitotoxicity and the NMP receptor. Trends Neurosc<br />

10:299-302.<br />

14. Wein AJ (1981) Classification of neurogenic voiding <strong>dysfunction</strong>. J UroZ 125:605-<br />

609.<br />

15. Bors E, Comm AE (1971) Disturbances of micturition: acute onset. Neurolog<br />

Urol University Park, Baltimore, MD. 1971. pp. 181-184.<br />

16. Roaf R (1960) A study of the mechanics of spinal injuries. J Bone Joint Surg<br />

42B:810-816.<br />

17. DeVivo M, Rutt R, Black K (1982) Trends in spinal cord injury demographics and<br />

treatment outcomes between 1973 and 1986. Arch Phys Med Rehabil73:424-430.<br />

18. Thomas D, O’Flynn K (1994) Urodynarnics: Principles, Practice, Application. In:<br />

Mundy AR, Stephenson TP, Wein AJ, Spinal Cord Injury. London: Churchill<br />

Livingstone, pp. 345.<br />

19. Sullivan M, Yalla S (1992) Spinal cord injury and other forms of myeloneuropathies.<br />

Problem Urol 6:643.<br />

20. Fam B, Yalla SV (1988) Vesicourethral <strong>dysfunction</strong> in spinal cord injury and its<br />

management. Semin Neurol 8: 150-155.<br />

21 * Wein A, Rovner E (1999) Adult voiding <strong>dysfunction</strong> secondary to neurologic<br />

disease or injury. In: Ball T, Jr., ed. AUA Update Series Vol. XVIII. AUA Office<br />

of Education, Houston, TX. pp. 42-48.<br />

22. Kaplan SA, Chancellor MB, Blaivas JG (1991) Bladder and sphincter behavior<br />

in patients with spinal cord lesions. J Urol 146: 1 13-1 17.<br />

23. Wyndaele JJ (1992) Neurourology in spinal cord injured patients. Paraplegia<br />

30:50-53.<br />

24. Curt A, Rodic B, Schurch B, Dietz V (1997) Recovery of bladder function in<br />

patients with acute spinal cord injury: significance of ASIA scores and somatosensory<br />

evoked potentials. Spinal Cord 35:368-373.<br />

25. Rudy D, Awad S, Downie J (1988) External sphincter dyssynergia: an abnormal<br />

continence reflux. J. Urol 140: 105-1 10.<br />

26. de Groat W, Karvatni M (1985) Neural control of the urinary bladder: Possible<br />

relationship between peptidergic inhibitory mechanisms and detrusor instability.<br />

Neurourol Urodyn 4:285-300.<br />

27. Anderson RU (1983) Urodynamic patterns after acute spinal cord injury: association<br />

with bladder trabeculation in male patients. J Urol 129:777-779.<br />

28. Ruutu M (1985) Cystometrographic patterns in predicting bladder function after<br />

spinal cord injury. Paraplegia 23:243-252.<br />

29. Chancellor MB, Rivas DA, Linsenmeyer T, Abdill CA, Ackman CF, Appell RA,<br />

et al. (1994) Multicenter trial in North America of UroLume urinary sphincter<br />

prosthesis. J Urol 152:924-930.<br />

30. McGuire E (1984) Clinical evaluation and treatment of neurogenic vesical <strong>dysfunction</strong>.<br />

In: <strong>Lib</strong>ertino J, ed., International Perspectives in Urology. Baltimore: Williams<br />

and Wilkins.<br />

31. Sundin T, Dahlstrom A (1973) The sympathetic innervation of the urinary bladder<br />

and urethra in the normal state and after parasympathetic denervation at the spinal<br />

root level. An experimental study in cats. Scand J Urol Nephrol 7:131-149.


Chapter 5 I Spinal Cord Injuries 135<br />

32. Lloyd L (1986) New trends in urologic management in spinal cord injured patients.<br />

J Cent New Syst Trauma 3:3-12.<br />

33. Guttman L, Frankel H (1966) The value of intermittent catheterisation in the early<br />

management of traumatic paraplegia and tetraplegia. Paraplegia 4:63-84.<br />

34. Stein SC, Feldman JG, Friedlander M, Klein RJ (1982) Is myelomeningocele a<br />

disappearing disease? Pediatrics 695 11-514.<br />

35. Laurence K (1989) A declining incidence of neural tube defects in UK. Z Kinderchir<br />

44(Suppl 1):5 1.<br />

36. Stark G (1977) Spina BiJida: Problems and Management. Oxford: Blackwell<br />

Scientific Publications.<br />

37. Toriello H (1982) Periconceptual vitamin supplementation for the prevention of<br />

neural tube defects: a review. Spinal BiJda Therapy 459.<br />

38. Scarff T, Fronczak S (1981) Myleomeningocele: a review and update. Rehab<br />

Lit 42:143-146.<br />

39. MRC Vitamin Study Research Group (1991) Prevention of neural tube defects:<br />

results of the Medical Research Council Vitamin Study. Lancet 338:131-137.<br />

40. Werler MM, Shapiro S, Mitchell AA (1993) Periconceptional folic acid exposure<br />

and risk of occurrent neural tube defects. JAMA 269: 1257-126 1.<br />

41. Hobbins J, Grannum P, Berkowitz RL, et al. (1979) Ultrasound in the diagnosis<br />

of congenital anomalies. Am J Obstet Gynecol 134:331-345.<br />

42. Brock DJH, Sutcliffe RC (1 972) Alpha Fetoprotein in the antenatal diagnosis of<br />

anencephaly and spina bifida. Lancet 2:197-199.<br />

43. Watson WJ, Chescheir NC, Katz VL, Seeds JW (1991) The role of ultrasound<br />

in evaluation of patients with elevated maternal serum alpha-fetoprotein: a review.<br />

Ohstet Gynecol 78:123-128.<br />

44. Chervenak FA, Duncan C, Ment LR, et al. (1984) Perinatal management of<br />

myelomeningocele. Obstet Gynecol 63:376-380.<br />

45. Bensen JT, Dillard RG, Burton BK (1988) Open spina bifida: Does caesarian<br />

section improve prognosis. Obstet Gynecol 71 532-534.<br />

46. McGuire E, Woodside J, Ta B, et al. (1981) Prognostic value of urodynamic<br />

testing in myelodysplastic patients. J Urol 126:205-209.<br />

47. Kroovand RL, Bell W, Hart LJ, Benfield KY (1990) The effect of back closure<br />

on detmsor function in neonates with myelomeningocele. J Urd 144:423-425.<br />

48. McGuire E, Denil J Adult myelodysplasia. In: AUA Update Series Vol. X. AUA<br />

Office of Education. Houston, TX. 1991. pp. 297.<br />

49. Blaivas JG (1980) Management of bladder <strong>dysfunction</strong> in multiple sclerosis.<br />

Neurology 30: 12-18.<br />

50. Lapides J, Diokno A, Silber S, Lowe B (1972) Clean intermittent self-catheterization<br />

in the treatment of urinary tract disease. J Urol 107:7458-7461.<br />

51. McGuire E, Savastano J (1986) Comparative urological outcome in women with<br />

spinal cord injury. J Urol 135:730-731.<br />

52. Swierzewski SJ, 3rd, Gormley EA, Belville WD, Sweetser PM, Wan J, McGuire<br />

EJ (1 994) The effect of terazosin on bladder function in the spinal cord injured<br />

patient. J UroZ 151:951-954.<br />

53. Sundin T, Dahlstrom A, Norlen L, Svedmyr N (19’77) The sympathetic innervation<br />

and adrenoreceptor function of the human lower urinary tract in the norma1 state<br />

and after parasympathetic denervation. Invest Urol 14:322-328.<br />

54. Sogbein SK, Downie JW, Awad SA (1984) Urethral response during bladder<br />

contraction induced by subcutaneous bethanechol chloride: elicitation of a sympathetic<br />

reflex urethral constriction. J Urd 13 1 :791-795.


136 Sukin Boone<br />

55. D’ Agostino G, Kilbinger H, Chiari MC, Grana E (1 986) Presynaptic inhibitory<br />

muscarinic receptors modulating (3H) acetylcholine release in the rat urinary<br />

bladder. J Pharm Exper Ther 239522-528.<br />

56. Somogyi G, de Groat W (1990) Modulation of the release of (3H) norepinephrine<br />

from the base and body of the rat urinary bladder by endogenous adrenergic and<br />

cholinergic mechanisms. J Pharmacol Exp Ther 255:204-210.<br />

57. Alberts P (l 995) Classification of the presynaptic muscarinic receptor subtype<br />

that refulates 3H-acetylcholine secretion in the guinia pig urinary bladder in vitro.<br />

J Pharmacol Exp Ther 274:458-468.<br />

58. Seze M, Wiart L, Joseph P (1998) Capsaicin and neurogenic detrusor hyperreflexia:<br />

A double-blind placebo controlled study in 20 patients with spinal cord lesions.<br />

Neurourol Urodynamics 17513-523.<br />

59. Perkash I (1976) Modified approach to sphincterotomy in spinal cord injury<br />

patients: Indications, technique, and results in 32 patients. Paraplegia 13:247-260.<br />

60. Wein A, Raezer D, Benson G (1976) Management of neurogenic bladder <strong>dysfunction</strong><br />

in the adult. Urology 83432-434.<br />

61. Carrion H, Braun B, Politano V (1979) External sphincterotomy at the 12 o’clock<br />

position. J Urol 121:462.<br />

62. Madersbacher H (1976) The twelve o’clock sphincterotomy: Technique, indications,<br />

results. Paraplegia 13:261-267.<br />

63. KIiviat M (1975) Transurethral sphincterotomy: Relationship of site of incision to<br />

postoperative potency and delayed hemorrhage. J Urol 114:399-401.<br />

64. Schellhammer P, Hackler R, Bunts R (1973) External sphincterotomy: An evaluation<br />

of 150 patients with neurogenic bladder. J Urol 110: 199-202.<br />

65. Yalla S, Fam B, Gabilondo F, Jacobs J, DiBenedetto M, et al. (l 977) Anteromedian<br />

external sphincterotomy: technique, rationale, and complications. J Urol l 17:489-<br />

493.<br />

66. Thomas D (l 976) The effect of trans-urethral surgery on penile erection in spinal<br />

injury patients. Paraplegia 13:286-289.<br />

67. Vapnek JM, Couillard DR, Stone AR (1994) Is sphincterotomy the best management<br />

of the spinal cord injured bladder? J Urol 151:961-964.<br />

68. Rivas D, Chancellor M, Bagley, D (1994) Prospective comparison of external<br />

sphincter prosthesis placement and external sphincterotomy in men with spinal<br />

cord injury. J Endourol 8:89-93.<br />

69. Moeller B (1977) Some observations on 31 spinal cord injury patients on whom<br />

the Bricker procedure was pedormed. Paraplegia 15:230.<br />

70. Boyce WH, Lathem JE, Hunt LD (1964) Research related to the development<br />

of an artificial electrical stimulator for the paralyzed human bladder: a review.<br />

J Urol 91:41-51.<br />

71. Brindley GS, Polkey CE, Rushton DN, Cardoza L (1986) Sacral anterior root<br />

stimulators for bladder control in paraplegia: The first 50 cases. J Neurol Neurosurg,<br />

Psychiat, 49: 1104-1 1 14.<br />

72. Tanagho EA, Schmidt RA, Orvis BR (1989) Neural stimulation for control of<br />

voiding <strong>dysfunction</strong>: a preliminary report in 22 patients with serious neuropathic<br />

voiding disorders. J Urol 142:340-345.<br />

73. Kerrebroech P, Koldewijn E, Debruyne M (1993) Worldwide experience with the<br />

Finetech-Brindley Sacral Anterior Root stimulator. Neurol Urodynamics 12:497-<br />

503.<br />

74. Merrill DC (1979) The treatment detrusor of incontinence by electrical stimulation.<br />

J Urol 122:5 15-5 17.


Chapter 5 / Spinal Cord Injuries I37<br />

75. Hackler RH (1977) A 25-year prospective mortality study in the spinal cord injured<br />

patient: comparison with the long-term living paraplegic. J Urol 117:486-488.<br />

76. Jonas U, Jones LW, Tanagho EA (1975) Recovery of bladder function after spinal<br />

cord transection. J Urol. 113:626-628.<br />

77. Young J. Spinal cord injury statistics: Experience of regional spinal injury systems,<br />

Good Samaritan Medical Center Phoenix, AZ. 1982 p. 152.<br />

77a. Bors E (1957) Neurogenic bladder. Urologic Survey 7:177.<br />

78. Duckett JW, Jr (1982) Ureterovesical junction and acquired vesicoureteral reflux.<br />

J Urol 127:249.<br />

79. Hall MK, Hackler RH, Zampieri TA, Zampieri JB (1989) Renal calculi in spinal<br />

cord-injured patient: association with reflux, bladder stones, and foley catheter<br />

drainage. Urology 34: 126- 128.<br />

80. Claus-Walker J, Campos R, Carter R, Vallbona C, Liscomb H (1972) Calcium<br />

excretion in quadraplegia. Arch Phys Med Rehabil 53:14-20.<br />

81. Trop CS, Bennett CJ (1991) Autonomic dysreflexia and its urological implications:<br />

a review. J Urol 146:1461-1469.<br />

82. Dykstra DD, Sidi AA, Anderson LC (1987) The effect of nifedipine on cystoscopyinduced<br />

autonomic hyperreflexia in patients with high spinal cord injuries.<br />

J Urol 138:1155-1157.<br />

83. Chancellor MB, Erhard MJ, Hirsch IH, Stass WE Jr (1994) Prospective evaluation<br />

of terazosin for the treatment of autonomic dysreflexia. J UroZ 15 1 : 1 1 1-1 13.


This Page Intentionally Left Blank


6 Diabetic Bladder Dysfunction<br />

CONTENTS<br />

INTRODUCTION<br />

EPIDEMIOLOGY<br />

PRESENTATION<br />

DIAGNOSIS<br />

URODYNAMIC FINDINGS<br />

PATHOPHYSIOLOGY<br />

TREATMENT<br />

CONCLUSION<br />

REFERENCES<br />

INTRODUCTION<br />

The prevalence of diabetes mellitus (DM) has reached epidemic<br />

proportions among the population of the United States. The American<br />

Diabetes Association recently reported that the prevalence of diagnosed<br />

and undiagnosed DM in the United States is currently 6% and rising<br />

(I). This increase is felt to be secondary to a rising incidence of obesity<br />

as well as a change in the criteria for the diagnosis of DM (from a<br />

fasting blood glucose of 140 mgldL to 126 mgldL). Besides impaired<br />

blood glucose regulation, many direct and indirect sequelae of DM can<br />

occur. Lower urinary tract <strong>dysfunction</strong> is relatively common in this<br />

group of patients and the etiology of these symptoms is commonly<br />

attributed to DM. Owing to the high incidence of DM and its frequent<br />

involvement of the lower urinary tract, a complete understanding of<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

139


140 Goldman Appell<br />

its epidemiology, physiology, presentation, diagnosis, and treatment is<br />

essential for clinicians caring for these patients.<br />

EPIDEMIOLOGY<br />

It is difficult to estimate exactly what percentage of diabetic patients<br />

suffer from bladder <strong>dysfunction</strong>. Though many patients are referred<br />

for evaluation of urinary symptoms, the total pool from which they<br />

are extracted is frequently hard to determine. In addition, many patients<br />

may have other co-existing problems which can either mimic or mask<br />

the findings of a diabetic neurogenic bladder. For example, women<br />

may have stress incontinence, prolapse, or pelvic floor <strong>dysfunction</strong>,<br />

whereas men may have prostatic obstruction, urethral stricture disease,<br />

or a host of other obstructive processes, all of which can be confused<br />

with or coexist with a diabetic neurogenic bladder.<br />

In some series, as many as 52% of randomly evaluated diabetic<br />

patients were found to have urologic symptoms (2). Even among<br />

“asymptomatic” patients, many have unrecognized symptoms. Ueda et<br />

al. (3) examined 53 “asymptomatic” diabetic patients and found that<br />

upon careful questioning 21 (40%) complained of voiding symptoms.<br />

They also noted no correlation of bladder <strong>dysfunction</strong> with type or<br />

duration of DM. In addition, the presence or absence of diabetic retinopathy<br />

did not correlate with the presence or absence of bladder <strong>dysfunction</strong>.<br />

Thus, a considerable number of patients with DM, even when<br />

asymptomatic, will be found to have bladder <strong>dysfunction</strong> upon careful<br />

evaluation.<br />

PESENTATION<br />

Generally the onset of diabetic neurogenic bladder <strong>dysfunction</strong> is<br />

insidious and may not be recognized until it has reached an advanced<br />

stage. Early signs and symptoms may be overlooked by the patient,<br />

but impaired bladder sensation is usually the first manifestation of<br />

lower urinary tract involvement. Symptoms may begin as infrequent<br />

voiding, decreased stream, hesitancy in initiating the urinary flow,<br />

dribbling due to overflow incontinence, and the sensation of incomplete<br />

emptying. Upon questioning the primary symptom may be only the<br />

loss of desire to void with a resultant pattern of voiding by the clock


Chapter 6 / Diabetic Bladder Dysfunction 141<br />

or from habit only. This constellation of symptoms is classically referred<br />

to as “diabetic cystopathy” (4).<br />

It is becoming more evident, however, that a significant number of<br />

diabetic patients may present with symptoms of urgency and frequency.<br />

A recent study by Kaplan et al. (5) noted that the most common<br />

symptoms in a group of diabetic patients referred for evaluation of<br />

urologic complaints were nocturia greater than two times (87%) and<br />

urinary frequency greater than every 2 h (78%). In contrast, the more<br />

classic diabetic cystopathy symptoms of hesitancy, decreased stream,<br />

and the sensation of incomplete bladder emptying were elicited in 62,<br />

52, and 45% of the patients. Because the symptoms associated with<br />

the classic diabetic bladder are insidious in nature, perhaps those<br />

patients are not well-represented in studies such as Kaplan’s, which<br />

rely on patients referred for specific urologic symptoms. In any event,<br />

it is clear that patients with a diabetic bladder can present with any of<br />

a full spectrum of voiding complaints from urgency and frequency to<br />

difficulty emptying.<br />

DIAGNOSIS<br />

When evaluating the diabetic patient for signs of bladder <strong>dysfunction</strong>,<br />

a complete history and physical exam as well as urine culture is essential.<br />

Many other conditions that may cause or aggravate vesico-urethral<br />

<strong>dysfunction</strong> must be considered since cerebrovascular or lumbar disc<br />

disease, as well as a host of other neurological disorders, can cause<br />

voiding symptoms that mimic those found in DM. Similarly, gross<br />

physical findings such as a cystocele or other forrns of pelvic prolapse<br />

in women or an enlarged prostate or indurated portion of the urethra<br />

(possible stricture) in men can contribute to significant voiding <strong>dysfunction</strong>.<br />

Testing of perineal sensation, sphincter tone, and the bulbocavernosus<br />

reflex can help identify a peripheral neuropathy consistent with<br />

DM. Urine culture may identify an infection that could be the source<br />

of the patient’s symptoms. In some patients, it may be difficult to<br />

separate the bladder <strong>dysfunction</strong> caused by diabetes from that of a coexisting<br />

disease.<br />

The diagnosis of diabetic neurogenic bladder and the exact type of<br />

<strong>dysfunction</strong> present is made most readily with urodynamic testing. This<br />

may include a simple cystometrogram, uroflow, simultaneous pressure/<br />

flow studies, sphincter electromyography (EMG), and urethral pressure<br />

profilometry or evaluation of leak-point pressures. Obtaining a postvoid


142 Goldman and Appell<br />

residual (PVR) may also help determine the efficiency of bladder empty-<br />

ing. The main differential diagnosis is stress incontinence (primarily<br />

in women) or outflow obstruction. The use of simultaneous pressure/<br />

flow studies with determination of leak-point pressures or urethral<br />

pressures will help differentiate between these two entities and the<br />

diabetic neurogenic bladder, though in some cases with co-existing<br />

problems, the determination of the role played by each entity may<br />

be difficult. In addition, electrophysiologic testing may be useful for<br />

assessment of peripheral neuropathy.<br />

URODYNMIC FINDINGS<br />

Classic urodynamic findings (6) include impaired bladder sensation,<br />

increased cystometric capacity, decreased bladder contractility,<br />

impaired uroflow, and an increased PVR. Cystometric examination<br />

may show detrusor areflexia but detrusor instability is also frequently<br />

found (see Table l). Urethral closure pressure profiles or Valsalva leak-<br />

point pressures are generally normal, but as noted previously, serve to<br />

exclude other forms of pathology.<br />

Table 1<br />

Urodpamic Features in Diabetes Mellitus<br />

Cystometrogram<br />

Decreased bladder sensation<br />

Increased bladder capacity<br />

Bladder contractility<br />

Areflexic<br />

Hyperreflexic<br />

Uroflowmetry<br />

Normal or decreased<br />

Urethral pressure profile<br />

Normal<br />

Leak point pressure<br />

Absent (infinity)<br />

Sphincter EMG<br />

Normal or abnormal<br />

Electrophysiologically evoked response<br />

Normal or abnormal


Chapter 6 / Diabetic Bladder Dysfunction 143<br />

In Kaplan’s study (5), 52% of the diabetic patients had urodynamically<br />

documented detrusor instability, 23% had impaired detrusor contractility,<br />

24% had poor compliance, 1 l% had indeterminate findings,<br />

and 24% had detrusor areflexia (5). Another study (7’) of elderly diabetic<br />

nursing home patients (mean age 80 yr; 19 women, 4 men) who presented<br />

with symptoms of urinary <strong>dysfunction</strong> showed that 61% had<br />

detrusor instability, 13% normal contractions, 17% voluntary contractions<br />

of low magnitude, and 9% no contractions at all. The authors<br />

concluded by emphasizing that one cannot assume that elderly diabetic<br />

patients with urologic symptoms have poorly contracting bladders (classic<br />

diabetic bladder), and that urodynamic studies can be helpful in<br />

determining the etiology and choosing the appropriate therapy in<br />

these patients.<br />

In contrast, earlier reports (4) and the more recent study by Ueda<br />

et al. (3) have noted urodynamic changes more consistent with those<br />

described with classic diabetic cystopathy. Ueda et al. (3) noted detrusor<br />

instability in 25% of patients (study involved 28 women and 25 men),<br />

but noted that all of the patients (with detrusor instability) had a history<br />

of cerebrovascular disease. No patient without cerebrovascular disease<br />

had detrusor instability. In addition, they noted significant increases in<br />

bladder volume at first desire to void and maximal capacity, decreases<br />

in detrusor contractility, and increases of PVR in their cohort of asymptomatic<br />

diabetic patients. Similarly, a study by Goldman et al. (8) noted<br />

that diabetics who also had gastroparesis had delayed first Sensation,<br />

increased capacity, and increased PVR. Perhaps the difference between<br />

these sets of findings is related to a selection bias regarding whether<br />

or not symptoms requiring evaluation were volunteered by the patients.<br />

In Kaplan’s study (5), patients had been referred because of voiding<br />

symptoms, whereas in Ueda’s study (3) the patients were unselected.<br />

It is clear though that there is no one set of urodynamic findings that<br />

applies to all cases of diabetic neurogenic bladder. Rather, it is imperative<br />

to test accurately for and diagnose the specific type of bladder<br />

<strong>dysfunction</strong> occurring in order to appropriately treat the patient.<br />

PATHOPHYSIOLOGY<br />

The cause of the bladder <strong>dysfunction</strong> seen in DM is primarily the<br />

result of peripheral and autonomic neuropathy. The classic diabetic<br />

cystopathy has been attributed to diminished sensation leading to a<br />

chronically overstretched bladder, which results in myogenic failure


144 Goldman Appell<br />

and the inability to mount a satisfactory detrusor contraction. The<br />

findings of detrusor instability in a significant number of patients implies<br />

that the cortical or spinal regulatory tracts can be affected too.<br />

In diabetes, somatic and autonomic neuropathy generally exist concurrently.<br />

This has been corroborated by studies demonstrating that<br />

the electrically stimulated bulbocavernosus reflex latency times are<br />

significantly prolonged (9). However, this is seen primarily in advanced<br />

peripheral pudendal neuropathy. More subtle changes in nerve function<br />

may be ascertained by detrusor and urethral electromyelography as<br />

described by Bradley and colleagues (IO), where there can be an<br />

increased latency of the evoked response in the perineal musculature<br />

as well as anal and urethral sphincters. It should be remembered that<br />

when simultaneous <strong>dysfunction</strong> of bowel, bladder, and genitals occurs<br />

early, it may be owing to cauda equina infarction from the vascular<br />

complications of diabetes (I I ).<br />

Currently, there are several theories regarding the pathophysiology<br />

of diabetic neuropathy. Symmetric peripheral neuropathy (polyneuropathy)<br />

may be the most common complication of DM (12). The autonomic<br />

neuropathy associated with DM is "generally thought to occur later.<br />

Several forms of abnormal impulse conduction patterns have been<br />

described, but in general, electrophysiologic studies show slowing of<br />

evoked responses (11). The impedance of conduction has been demonstrated<br />

in demyelinated nerve fibers, and when more severe conduction<br />

delay occurs this may result in not only temporal delay, but complete<br />

nerve blockade (13). Although fibers such as those in the autonomic<br />

nervous system may be affected first owing to their smaller diameter,<br />

shorter length, and smaller sudace-to-volume ratio, clinical manifestations<br />

generally appear later than peripheral symmetric polyneuropathy.<br />

Disturbance of the peripheral nerve metabolism is thought to play<br />

an important role in the neuropathy of diabetes as well. The metabolic<br />

derangement causes <strong>dysfunction</strong> of the Schwann cell with resultant<br />

segmental demyelination as well as axonal degeneration (14). It is<br />

hypothesized that increased tissue levels of sorbitol may account for<br />

the tissue destruction and peripheral neuropathy (15). Another postulated<br />

cause of peripheral nerve damage is the reduction of myo-inositol<br />

in peripheral nerves (15). Interplay may occur between the sorbitol<br />

and myo-inositol pathways, but it has not been confirmed (16). Finally,<br />

the increased glycation of peripheral nerve cytoskeleton proteins is felt<br />

to contribute to the pathogenesis of diabetic neuropathy (17').<br />

Whatever the exact metabolic cause, recent studies have clearly<br />

demonstrated neural <strong>dysfunction</strong> involving the bladder. Steers et al.


Chapter 6 / Diabetic Bladder Dysfunction 145<br />

(18) demonstrated changes in both afferent and efferent pathways innervating<br />

the bladder in an animal model. Itoh et al. (19) demonstrated<br />

significant decreases of acetylcholinesterace staining and activity in<br />

detrusor muscle, suggesting the presence of an autonomic neuropathy.<br />

These and other findings indicate that the alterations in bladder function<br />

found in diabetic patients are the result of derangements in both the<br />

peripheral and autonomic nervous systems.<br />

Treatment involves symptomatic relief, prevention of infection,<br />

maintenance of renal function, and achievement of continence while<br />

providing adequate bladder emptying (20). Often treatment decisions<br />

must be based on the results of urodynamic testing since effective<br />

therapy requires an understanding of the cause of the individual patient’s<br />

symptoms. Patients with detrusor instability are best managed with<br />

anticholinergic or smooth muscle relaxants to diminish the frequency<br />

and amplitude of involuntary contractions. Oxybutynin hydrochloride,<br />

imipramine, and dicyclomine are among the medications that are commonly<br />

used, although side effects may limit their application. The<br />

introduction of tolteradine and extended-release oxybutynin hydrochloride,<br />

which reportedly have fewer side effects, may increase the applicability<br />

of medical therapy.<br />

For those with impaired detrusor contractility or detrusor areflexia<br />

no cure is available, as yet. Timed voidings on a consistent and regular<br />

schedule is effective in those with diminished sensation but who still<br />

can empty their bladder. Significant detrusor decompensation will<br />

necessitate intermittent catheterization. Clean intermittent catheterization<br />

(CIC) is usually easily learned and accepted by most patients.<br />

By keeping the catheterized volumes under 400 mL (or less if poor<br />

compliance or elevated bladder pressures are noted on urodynamics<br />

studies the risk of renal damage is minimized. Occasionally, if recurrent<br />

urinary tract infections become a problem, antimicrobial prophylaxis<br />

is necessary. Metoclopromide, with its cholinergic effects, has been<br />

used to treat diabetic neurogenic bladder <strong>dysfunction</strong>, as improvement<br />

has been seen in patients with delayed gastric emptying (21). On the<br />

other hand, bethanechol chloride, a purely parasympathomimetic agent,<br />

is rarely effective (22). In short, the success of bladder rehabilitation<br />

in this subgroup of patients depends upon the degree of detrusor<br />

decompensation.


146 Goldman Appell<br />

CONCLUSION<br />

Bladder <strong>dysfunction</strong> in those with DM is relatively common. Given<br />

the high incidence of DM in the general population, the clinician must<br />

be prepared to evaluate and treat these patients effectively. Though the<br />

classic diabetic cystopathy refers to the bladder with poor sensation and<br />

detrusor failure, many patients will have primarily detrusor instability.<br />

Thus, urodynamic testing is frequently crucial to determine the actual<br />

type of bladder <strong>dysfunction</strong> in the individual and to allow for the<br />

institution of appropriate therapy. Hopefully, future research will shed<br />

more light onto the exact pathophysiology of this disorder and allow<br />

for more effective bladder rehabilitation.<br />

REFERENCES<br />

1. American Diabetes Association (1998) Screening for type I1 diabetes. Diabetes<br />

Care 21520-22.<br />

2. Ioanid CP, Noica N, Pop T (198 Incidence 1) and diagnostic aspects of the bladder<br />

disorders in diabetics. Eur Urol 7:211-214.<br />

3. Ueda T, Yoshimura N, Yoshida 0 (1997) Diabetic cystopathy: relationship to<br />

autonomic neuropathy detected by sympathetic skin response. J Urol 157:580-<br />

584.<br />

4. Frimoldt-Moller A (1980) Diabetic cystopathy: epidemiology and related disorders.<br />

Ann Intern Med 92:318-321.<br />

5. Kaplan SA, Te AE, Blavis JG (1995) Urodynamic findings patients in with diabetic<br />

cystopathy. J Urol 152:342-344.<br />

6. Appell RA, Whiteside HV (1991) Diabetes and other peripheral neuropathies<br />

affecting lower urinary tract function. In: Gane RJ, Siroky MB, eds., Clinical<br />

Neurourology. Boston: Little, Brown, pp. 365-375.<br />

7. Starer P, <strong>Lib</strong>ow L (1990) Cystometric evaluation of bladder <strong>dysfunction</strong> in elderly<br />

diabetic patients. Arch Int Med 150:810-813.<br />

8. Goldman HB, Dmochowski RR (1997) Lower urinary tract <strong>dysfunction</strong> in patients<br />

with gastroparesis. J Urol 157: 1823-1825.<br />

9. Goldstein I, Siroky MB, mane RJ (1983) Impotence in diabetes mellitus. In:<br />

Krane RJ, Siroky MB, Goldstein I, eds., Male Sexual Dysfinction, Boston: Little,<br />

Brown, pp. 77-86<br />

10. Bradley WE, Timm GW, Rockswold GL, Scott FB (1975) Detrusor and urethral<br />

electromyelography. J Urol 114:891-894.<br />

11. Bradley WE (1979) Neurologic disorders affecting the urinary bladder. In: Krane<br />

RJ, Siroky MB, eds., Clinical Neuro-Urology, Boston: Little, Brown, pp. 245-255.<br />

12. Ellenberg M (1976) Diabetic neuropathy; clinical aspects. Metabolism 25:1627-<br />

1632.<br />

13. Waxman SG (1980) Pathophysiology of nerve conduction; relation to diabetic<br />

neuropathy. Ann Intern Med 92:297-300.


Chapter G / Diabetic Bladder Dysfunction 147<br />

14. Spritz N, Singh H, Marinan B (1975) Decrease in myelin content of rabbit sciatic<br />

nerve with aging and diabetes. Diabetes 24:680-684.<br />

15. Clements RS Jr (1979) Diabetic neuropathy: new concepts of its etiology. Diabetes<br />

28:604-611.<br />

16. Finegold D, Lattirner SA, Nolle S, Bernstein M, Green DA, et al. (1983) Polyol<br />

pathway activity and myo-inositol metabolism. A suggested relationship in the<br />

pathogenesis of diabetic neuropathy. Diabetes 32:988-992.<br />

17. Ryle C, Leow CK, Donaghy M (1997) Nonenzymatic glycation of peripheral<br />

and central nervous system proteins in experimental diabetes mellitus. Muscle<br />

Nerve 20:577-584.<br />

18. Steers WD, Mackway-Gerardi AM, Ciambotti J, deGroat WC (1994) Alterations<br />

in neural pathways to the urinary bladder of the rat in response to streptozotocininduced<br />

diabetes. J Auton New Syst 47:83-94.<br />

19. Itoh H. Morikawa A, Maliino I (1994) Urinary bladder <strong>dysfunction</strong> in spontaneously<br />

diabetic Chinese hamsters. Diabet Res Clin Pract 22: 163-170.<br />

20. Appell RA, Baum NH (1990) Neurogenic bladder in patients with diabetes mellitus.<br />

Practical Diabetology 9: 1-6.<br />

21. Nestler JE, Stratton MA, Hakim CA (1983) Effect of metaclopromide on diabetic<br />

neurogenic bladder. Clin Pharmacol 283-85.<br />

22. Mundy AR, Blaivis JG (1984) Nontraumatic neurological disorders. In: Mundy<br />

AR, Stephenson TP, and Wein AJ, eds., Urodynamic Principles, Practices and<br />

Application. Edinburgh: Churchill Livingstone, pp. 278-287.


This Page Intentionally Left Blank


7<br />

/ Lumbar Disc Disease<br />

Howard B. Goldman, MD<br />

and Rodney A. Appell, MD, FACS<br />

INTRODUCTION<br />

NEUROPATHOPHYSIOLOGY<br />

CLINICAL FEATURES<br />

EVALUATION<br />

TREATMENT<br />

CONCLUSION<br />

REFERENCES<br />

INTRODUCTION<br />

A significant proportion of individuals with lumbar disc disease<br />

experience voiding <strong>dysfunction</strong>, Specifically, 27 (1) to 92% (2) of these<br />

patients are noted to have such <strong>dysfunction</strong> and in most cases detrusor<br />

areflexia is found. This chapter reviews the pertinent neuropathophysi-<br />

ology, clinical features, evaluation, and treatment for patients with<br />

voiding <strong>dysfunction</strong> resulting from lumbar disc disease.<br />

NEIJROPATHOPWSIOLOGY<br />

A review of the relevant neuroanatomy is necessary before discussing<br />

the neuropathology resulting from lumbar disc disease.<br />

Innervation of the lower urinary tract is derived from both the autonomic<br />

and somatic nervous systems. The parasympathetic pelvic nerves,<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfinction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

149


150 Goldrnan Appell<br />

EHerent<br />

Autonomic<br />

Parasympathetic<br />

(Pelvic nerve)<br />

Sympathetic<br />

(Hypogastric nerve)<br />

Somatic<br />

(Pudendal nerve)<br />

Table l<br />

Innervation of the Lower Urinary Tract<br />

Spinal Cord Segment Function<br />

S2-S4<br />

T11-L2<br />

s3-s4<br />

Excitatory to bladder<br />

Inhibitory to bladder<br />

Excitatory to urethra<br />

Control over<br />

external sphincter<br />

Aferent<br />

Tension receptors in bladder wall<br />

Nociceptors S2-S4 Micturition ini<br />

Sensory S2-S3 Perineal sensa<br />

which emanate from the second through fourth sacral segments of the<br />

spinal cord, provide the principle excitatory input to the bladder. The<br />

somatic nerves originate from the third and fourth sacral segments and<br />

provide innervation to the external sphincter and other pelvic-floor<br />

musculature. Finally, the sympathetic pathways of the hypogastric<br />

nerves arise from the lower thoracic and upper lumbar segments and<br />

provide inhibitory input to the bladder body as well as excitatory input<br />

to the urethra and bladder base. Afferent activity may travel from the<br />

bladder to the spinal cord along both sets of autonomic nerves. However,<br />

those that convey information from tension receptors and nociceptors<br />

in the bladder wall and are the most important for initiating micturition<br />

travel via the parasympathetic nerves to the sacral segments of the<br />

cord. Sensory nerves from the vagina and clitoris travel in the somatic<br />

nerves to the sacral cord as well (see Table l) (3).<br />

In the adult, the sacral segments of the spinal cord are at the level<br />

of the first and second lumbar vertebral bodies. This distal end of the<br />

spinal cord is commonly called the conus medullaris. The spinal-cord<br />

segments are named for the vertebral body at which its nerve roots<br />

exit the spinal canal. Thus, though the first sacral segment of the spinal<br />

cord is located at Ll, its nerve roots run in the subarachnoid space<br />

posterior to the L2 to L5 vertebral bodies till reaching the first sacral<br />

~~


Chapter 7 / Lumbar Disc Disease 151<br />

vertebral body, at which point they exit the canal. Thus, all of the<br />

sacral nerves that originate at the L1 and L2 levels run posterior to the<br />

lumbar vertebral bodies until reaching their appropriate site of exit<br />

from the spinal canal. This group of nerves running at the distal end<br />

of the spinal cord is more commonly referred to as the cauda equina.<br />

The most frequent sites of lumbar disc prolapse are the L4/5 and L51<br />

S1 intervertebral spaces (4-6). Usually, prolapse is in a posteriolateral<br />

direction not affecting the majority of the cauda equina. However, in<br />

1-15% of cases, central-disc prolapse occurs and compression of the<br />

cauda eyuina may result (4) (Fig. 1). In fact, in some instances a large<br />

posteriolateral disc prolapse may migrate medially and cause cauda<br />

equina compression as well (4). Thus, fusing this neuroanatomy with<br />

the known neural innervation of the lower urinary tract, one can see<br />

how prolapse anywhere along the lumbar spine could interfere with<br />

the parasympathetic (S2-S4) and somatic (S3-S4) innervation, whereas<br />

only prolapse at the upper lumbar spine, which is relatively rare, would<br />

affect the sympathetic (T11-L2) innervation. Along with interfering<br />

with parasympathetic and somatic innervation to the bladder and urethra,<br />

afferent stimuli from the bladder and perineal sensation, both<br />

served via sacral segments, would face interference as well.<br />

Thus, lumbar disc prolapse interferes with stimuli from the bladder<br />

needed to initiate micturition, excitatory input to the bladder required<br />

for detrusor contraction, and somatic innervation to the external sphincter<br />

and pelvic-floor musculature. The resultant urologic finding is one<br />

of detrusor areflexia frequently with an impaired sensation of filling.<br />

This urologic manifestation of compression of the sacral nerves is<br />

usually only one aspect of the cauda equina syndrome (compression<br />

of the lumbosacral nerves). Classically, saddle anesthesia, bilateral<br />

sciatica, lower back pain-and in men, impotence-is noted (7).<br />

Though detrusor areflexia is the most common finding in these patients,<br />

detrusor hyperreflexia has been reported as well (8). It has been suggested<br />

that the mechanisms of injury in this group of patients is progressive<br />

disc herniation causing irritation and excitation of the sacral<br />

nerve roots.<br />

Besides direct compression, a prolapsing disc can affect the sacral<br />

nerves by interfering with blood flow to and from the cauda equina<br />

(9). Experimental investigations have shown changes in the intraneural<br />

venous blood flow with compression of these veins leading to congestion<br />

and ischemia of the nerve roots. Delmarter et al. (9) constricted<br />

the cauda equina to varying degrees in an animal model and evaluated<br />

changes in bladder function. Fifty percent constriction resulted in no


152 Goldman and Appell<br />

Fig. 1. (A) MRI with sagital section showing prolapse of L5-S 1 disc into subarach-<br />

noid space.<br />

significant cystometric changes, but did cause venous congestion of<br />

the nerve roots and ganglia. Seventy-five percent constriction caused<br />

detrusor areflexia, increased bladder capacity, and overflow inconti-<br />

nence, as well as arterial narrowing and venous congestion of the nerve<br />

roots and ganglia.


Chapter 7 / Lumbar Disc Disease 153<br />

Fig. 1. (B) Same patient, coronal view showing central disc prolapse with compression<br />

of sacral nerve roots.<br />

Complete damage to the conus or sacral roots produces paralysis of<br />

the detrusor such that it is acontractile and must be emptied using<br />

external pressure (straining or Crede) or a catheter. Although detrusor<br />

contractility is absent, there may be poor compliance and a steady rise<br />

of detrusor pressure during the filling phase, which indicates that the<br />

injury to the cauda equina is incomplete. Fortunately, most of these<br />

lesions are partial and have a more profound effect on the autonomically<br />

innervated bladder than on the somatically innervated external sphincter.<br />

Therefore, if the conus medullaris is partially damaged even without<br />

sacral-root damage, activity in the pelvic floor and external sphincter<br />

remains intact despite the loss of detrusor contractility. The efficiency<br />

with which the bladder empties in the absence of detrusor contractility<br />

depends on sufficient external force and the absence of any outflow


154 Goldrnan Appell<br />

obstruction. Even if detrusor function returns, it is usually inadequate<br />

to result in efficient emptying of the bladder and does not indicate<br />

detrusor-striated external sphincter dyssynergia, as it is a form of func-<br />

tional obstruction by the external sphincter because there is no detru-<br />

sor contractility.<br />

CLINICAL FMTIJRES<br />

Recent studies suggest that a minority of patients presenting with<br />

lumbar disc protrusion will have voiding <strong>dysfunction</strong>. Bartolin et al.<br />

(1) prospectively studied 114 patients (37 women, 77 men) who complained<br />

of low back pain and were found to have lumbar disc protrusion<br />

requiring surgical treatment. Of this group, 31 (27.2%) were found to<br />

have detrusor areflexia, whereas detrusor activity was normal in the<br />

remaining 83. Specifically, 3 of 8 with L3, 10 of 54 with L4, and 18<br />

of 52 with L5 disc protrusion had detrusor areflexia. All of these 31<br />

patients reported difficulty voiding requiring straining. Clearly, this is<br />

a select group as many patients do not require surgery and likely have<br />

less severe disc prolapse. Those with less severe prolapse probably<br />

have a lower rate of voiding <strong>dysfunction</strong> as well. In contrast to this<br />

report, Rosomoff et al. (2) reported rates of voiding <strong>dysfunction</strong> as<br />

high as 92%; however, his study had far less stringent diagnostic criteria<br />

for areflexia than that of Bartolin.<br />

As noted earlier, patients will frequently present with a constellation<br />

of symptoms representative of the cauda equina syndrome. O’Flynn et<br />

al. (4) reported on 30 patients with lumbar disc prolapse and urinary<br />

<strong>dysfunction</strong>. In 23 patients, a long history of lower back pain existed<br />

prior to the development of urinary symptoms. However, on occasion<br />

voiding <strong>dysfunction</strong> may be the only or first symptom of disc prolapse.<br />

Sylvester et al. (8) reported on two women who presented with painless<br />

urinary retention, but no other neurological findings were noted. In<br />

both cases, an MRI revealed lumbar central disc prolapse and, in<br />

retrospect, both patients noted the experience of minor lower back pain<br />

prior to presentation.<br />

A prolapsed intervertebral disc may produce a cauda equina syndrome<br />

with obstructive voiding symptoms on a permanent or intermittent<br />

basis. Therefore, the clinical presentation depends on the extent<br />

of injury to the autonomic parasympathetic nervous input to the lower<br />

urinary tract. Clinically, the patient describes pain in the lower back<br />

radiating in a girdle-like fashion along the lumbar dermatome involved;


Chapter 7 / Lumbar Disc Disease 155<br />

physical examination may reveal reflex and sensory changes consistent<br />

with nerve-root compression (1 0). <strong>Voiding</strong> symptoms, when present,<br />

are obstructive in nature (11) and include compromised urinary flow<br />

rate, interrupted stream owing to abnormal straining to void, residual<br />

urine, and incontinence. The obstructive Symptoms are secondary to<br />

the degree of detrusor denervation. The incontinence, however, may<br />

be owing to lack of resistance at the level of the external sphincter<br />

owing to pelvic-floor denervation or overflow. Denervation is more<br />

commonly seen in patients with recurrent or repetitive intervertebral<br />

disc problems and spinal stenosis (11).<br />

The most characteristic findings on physical examination are sensory<br />

loss in the perineum or perianal area (associated with the S2-4 dermatomes),<br />

sensory loss on the lateral foot (S 1-2 dematomes), or both (12).<br />

A unilateral or mild sensory disturbance indicates a better prognosis, as<br />

prolonged sensory deficits imply that the bladder will not recover<br />

because its normal function requires an intact visceral reflex arc of the<br />

sacral roots (6,13). The bulbocavernosus reflex (BCR) should be<br />

checked because reflects it pudendal (somatic) nerve function. Although<br />

this reflex is absent in all patients with complete lower motor neuron<br />

lesions of the sacral cord, care must be taken not to assume this is the<br />

case since 19% of healthy female subjects do not have a detectable<br />

reflex on physical examination alone (14). In one series, which included<br />

patients with injuries to the conus medullaris and cauda equina having<br />

various causes, the BCR reflex was absent or significantly diminished<br />

in 84% of the patients and perineal sensation and muscle-stretch reflexes<br />

were compromised in 77% of patients (IS). These findings correlated<br />

well with a diagnosis of pelvic-floor denervation.<br />

EVALUATION<br />

When other features of the cauda equina syndrome are present it is<br />

easy to suspect lumbar disc prolapse as the cause of voiding <strong>dysfunction</strong>.<br />

It is a more difficult diagnostic dilemma when other signs, in particular<br />

pain, are absent. One’ S suspicion should be aroused when faced with<br />

a woman who is straining to void or having difficulty emptying her<br />

bladder as outlet obstruction is unusual in females. A thorough history<br />

and physical examination with a focused neurological evaluation is<br />

mandatory. In the male, prostatic obstruction or urethral stricture disease<br />

may cause similar findings and should be excluded. Other disease states<br />

such as diabetes mellitus (DM), multiple sclerosis (MS), vitamin B12


156 Goldman ADD^<br />

deficiency, or infectious neuropathies such as herpes simplex should<br />

be noted. Careful questioning concerning lower back pain, gait disturbances,<br />

or any other symptoms that were present when the voiding<br />

<strong>dysfunction</strong> began can be helpful. For example, the presence of flu<br />

symptoms and gait disturbance in the past may be related to an infectious<br />

neuropathy. Sensory examination of the perineal area and lateral foot<br />

(both associated with sacral segments) may give clues to a neurologic<br />

origin. The BCR and anal sphincter tone as well can suggest a lesion<br />

involving sacral segments or nerve roots. Assessment of the patients<br />

postvoid residual using either a straight catheter or an ultrasound bladder-scan<br />

device can also give information on the efficiency of bladder<br />

emptying.<br />

Further urodynamic testing is frequently required to document the<br />

voiding <strong>dysfunction</strong>. ~rodyn~ically, detrusor motor <strong>dysfunction</strong> and<br />

weakened external sphincter activity are the primary findings even in<br />

urologically asymptomatic patients (11). A cystometrogram (CMG)<br />

with simultaneous pelvic-floor electromyography (EMG) is sufficient<br />

for urodynamic diagnosis and the predominant finding is usually detrusor<br />

areflexia associated with sphincter neuropathy (5,10,11,15). However,<br />

simultaneous pressure flow studies are frequently needed if there<br />

is a question of outlet obstruction. As discussed previously, in cases<br />

in which the somatic innervation remains intact, the external sphincter<br />

may appear to be incapable of relaxing when the patient strains to<br />

void and actually constitutes an obstructive factor that may require<br />

pharmacologic manipulation or intermittent catheterization. Thus, confirmation<br />

that intravesical pressure elevation on the CMG is owing to<br />

abdominal straining and not a true detrusor contraction (indicating<br />

normal function or obstruction) can be obtained by using subtracted<br />

pressure measurements or by synchronous perineal floor and rectus<br />

abdominis EMG (16). Sacral-evoked potentials, specifically the bulbocavernosus<br />

reflex latency time, can be used to study the integrity of<br />

the sacral reflex arc further; in this set of patients the response is<br />

usually absent or prolonged (1 7). Table 2 summarizes the neurourologic<br />

features of patients with cauda equina syndrome (12).<br />

When lumbar disc disease is suspected, magnetic resonance imaging<br />

(MRI) of the lumbar spine should be performed. MRI defines<br />

the anatomical detail much more precisely than myelography (which<br />

was formerly the diagnostic test of choice). If MRI reveals lumbar<br />

disc prolapse, neurosurgical or orthopaedic consultation should be obtained.


Chapter 7 I Lumbar Disc Disease I57<br />

Table 2<br />

Major Neurourologic Features of Lumbar Disc Disorders<br />

Symptoms Obstructive voiding symptomatology<br />

Signs Perineal or perianal sensory loss<br />

Absent or diminished bulbocavernosus reflex<br />

Tests CMG: detrusor areflexia<br />

EMG: neuropathic changes<br />

CMG, cystometrogram; EMG, electromyography.<br />

Treatment consists of correcting the underlying cause of the problem,<br />

usually requiring laminectomy. However, in some instances laminectomy<br />

may not improve the voiding symptoms. Shapiro (7) reviewed<br />

the cases of 14 patients who presented with cauda equina syndrome<br />

secondary to lumbar disc herniation, of whom 13 had incontinence.<br />

When evaluating variables to predict resolution of incontinence, he<br />

noted that the time between onset of Symptoms and laminectomy was<br />

important. Seven of 7 patients operated on within 48 h regained continence,<br />

whereas only 2 of 6 operated on at lengths greater than 48 h<br />

from presentation regained continence. Other studies though, such as<br />

that by O’Flynn et al. (4) report a much poorer rate of detrusor recovery.<br />

They noted only 1 of 26 patients with prelaminectomy incontinence<br />

regained completely normal bladder function. This disparity between<br />

the reports of Shapiro (7) and O’Flynn et al. (4) may be owing to<br />

the use of different outcome measures-Shapiro used “continence”<br />

whereas O’Flynn used “completely normal bladder function.’’ Thus,<br />

though it is likely that many patients become continent and can void<br />

by abdominal straining, few may actually regain “normal” bladder<br />

function. The sequelae of laminectomy such as spinal stenosis or arachnoiditis<br />

can affect bladder function as well. O’Flynn et al. (4) reported<br />

that four patients who had no urinary symptoms prior to surgery had<br />

voiding <strong>dysfunction</strong> postoperatively.<br />

The challenge for the clinician is to attain symptomatic relief for<br />

the patient by allowing adequate bladder emptying without incontinence<br />

while preserving upper urinary-tract function. Patients with lower motor<br />

neuron lesions do not develop reflex vesical activity, and storage function<br />

may or may not be normal. If bladder compliance is adequate, a<br />

trial of cholinergic stimulation with bethanechol chloride alone (50 mg


158 Goldman Appell<br />

up to four times daily) or in combination with metoclopramide (5-10<br />

mg up to four times daily) may be given, but should not be continued<br />

if no significant response is attained by the time medication has been<br />

prescribed for a month (1 8). In most cases cholinergic therapy is unsatis-<br />

factory and clean intermittent catheterization (CIC) is required.<br />

Hypertonicity or decreased compliance with a concomitant fixed-<br />

outlet resistance can impair ureteral function and place the upper urinary<br />

tracts at risk. In these patients, a regimen of anticholinergic medications<br />

usually combined with self-catheterization should be utilized. When<br />

this is unsuccessful, bladder augmentation may be necessary to lower<br />

detrusor pressure and protect the upper urinary tracts. Again, emptying<br />

is handled by self-catheterization. When incontinence results from poor<br />

sphincter function, the options are similar to those of any patient with<br />

intrinsic sphincter <strong>dysfunction</strong>-an artificial urethral sphincter, colla-<br />

gen injections, or-in women-a pubovaginal sling. A sling is preferred<br />

whenever the patient may require self-catheterization postoperatively,<br />

as repeated catheterization may displace the collagen that is injected<br />

and render the patient incontinent again.<br />

CONCLUSION<br />

<strong>Voiding</strong> <strong>dysfunction</strong> may present in conjunction with other neuro-<br />

logic symptoms or as an isolated event in patients with a prolapsed<br />

intervertebral disc. Damage to the parasympathetic, somatic, and sen-<br />

sory nervous innervation of the lower urinary tract is the usual mecha-<br />

nism of injury. In most cases detrusor areflexia results. A careful history<br />

and examination in conjunction with urodynamic testing and MRI<br />

imaging will establish the diagnosis. Although surgical intervention<br />

(laminectomy) may not always give complete bladder recovery, it can<br />

frequently allow for improvement. Various bladder management tech-<br />

niques (of which CIC is the mainstay) will help achieve good bladder<br />

emptying and preservation of the upper urinary tracts.<br />

REFERENCES<br />

1. Bartolin Z, Gilja I, Bedalov G, Savic I (1998) Bladder function in patients with<br />

lumbar intervertebral disc protrusion. J Urol 159:969-971.<br />

2. Rosomoff HL, Johnston JD, Gallo AE, et al. (1963) Cystometry in the evaluation of<br />

nerve compression in lumbar spine disorders. Surg Gynecol Obstet 117:263-270.


Chapter 7 / Lumbar Disc Disease 159<br />

3. deGroat WC (1 996) Neuroanatomy and neurophysiology: Innervation of the lower<br />

urinary tract. In: Raz S, ed. Female Urology, Philadelphia: WB Saunders, pp.<br />

28-30.<br />

4. O’Flynn KJ, Murphy R, Thomas DG (1992) Neurogenic bladder <strong>dysfunction</strong> in<br />

lumbar intervertebral disc prolapse. Br J Urol 69:38-40.<br />

5. Andersen JT, Bradley WE (1976) Neurogenic bladder <strong>dysfunction</strong> in protruded<br />

lumbar disc and after laminectomy. Urology 8:94-96.<br />

6. Scott PJ (1965) Bladder paralysis in cauda equina lesions from disc prolapse.<br />

JBone Joint Surg 47:224-227.<br />

7. Shapiro S (1993) Cauda equina syndrome secondary to lumbar disc herniation.<br />

Neurosurgery 32:743-747.<br />

8. Sylvester PA, McLoughlin J, Sibley GN, Dorman PJ, Kabala J, Ormerod IEC<br />

(1 995) Neuropathic urinary retention in the absence of neurological signs. Postgrad<br />

Med J 7 1 :747-748<br />

9. Delamarter RB, Bohlrnan HH, Bodner D, Biro C (1990) Urologic function after<br />

experimental cauda equina compression-cystornetrograrns versus corticalevoked<br />

potentials. Spine 15:864-870.<br />

10. Bradley WE (1978) Neurologic disorders affecting the urinary bladder. In: mane<br />

RJ, Siroky MB, eds., Clinical Neuro-Urology, 1st ed. Boston: Little, Brown,<br />

pp. 245-255.<br />

11. Appell RA, Levine RL (l 98 l) Assessment of vesico-urethral <strong>dysfunction</strong> in lumbar<br />

spine disorders. In: Proceedings of the Eleventh Annual Meeting of the InternationaE<br />

Continence Society. Edited by T. Sundin, A Mattiasson, Sweden: Skogs<br />

Trelleborg, 186-7.<br />

12. Appell RA (1993) <strong>Voiding</strong> <strong>dysfunction</strong> and lumbar disc disorders. Problems<br />

Urol 7( 1):35-40.<br />

13. Susset JG, Peters ND, Cohen SI, Ghoneim GM (1982) Early detection of neurogenic<br />

bladder <strong>dysfunction</strong> caused by protruded lumbar disc. UroEogy 20:461--463<br />

14. Blaivas JG, Zayed AAH, Labib KB (1981) The bulbocavernosus reflex in urology:<br />

a prospective study of 299 patients. J Urol 126197-199.<br />

15. Pavlakis AJ, Siroky MB, Goldstein I, Krane RJ (1983) Neurourologic findings<br />

in conus medullaris and cauda equina injury. Arch Neurol 40:570-573.<br />

16. Koff SA, Kass EJ (1982) Abdominal wall electromyography: a non-invasive<br />

technique to improve pediatric urodynamic accuracy. J Urol 127:736-739<br />

17. Krane RJ, Siroky MB (1 980) Studies in sacral evoked potentials. J Urol 124:872-<br />

876.<br />

18. Appell RA (1992) Clinical pharmacology in neurourology. Problems Urol 6622-<br />

642.


This Page Intentionally Left Blank


I I I FEMALE VOIDING DYSFUNCTION


This Page Intentionally Left Blank


8 Treatment of Stress<br />

Urinary Incontinence<br />

CONTENTS<br />

INTRODUCTION<br />

BEHAVIORAL INTERVENTIONS<br />

PHARMACOLOGICAL MANAGEMENT<br />

SURGICAL MANAGEMENT<br />

SUMMARY<br />

REFERENCES<br />

INTRODUCTION<br />

Stress urinary incontinence (SUI) is defined as urinary leakage sec-<br />

ondary to an increase in abdominal pressure (Valsalva maneuvers, which<br />

place “stress” on the bladder and bladder support mechanisms) and is<br />

classified for treatment purposes by urologist into anatomic hypermobil-<br />

ity of the bladder neck and proximal urethra, intrinsic sphincteric defi-<br />

ciency, or a combination of both. Anatomic hypemobility-related SUI<br />

(accounting for 85% of SUI) develops most commonly with aging,<br />

hormonal changes, traumatic or prolonged vaginal delivery, and pelvic<br />

surgery. Some patients with a well-supported bladder neck and urethra<br />

will still have SUI due to intrinsic sphinteric deficiency (15% of SUI).<br />

In these cases, intrinsic sphinteric deficiency may be owing to damage<br />

of the urethral closure mechanism after prior pelvic or anti-incontinence<br />

procedures, pelvic radiation, trauma, or neurogenic and medical disor-<br />

ders resulting in denervation injury. A combination of anatomic hyper-<br />

mobility and intrinsic sphinteric deficiency often exist together, and<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

163


164 RacMey<br />

Table 1<br />

Outcomes of Treatment for Urinary Incontinence(')<br />

Improved (%) 54-95<br />

Cured (%) 12-16<br />

Treatment<br />

Behavioral Pharmocolagic<br />

77 -<br />

44 78-92<br />

contemporary theories suggest that all cases of SUI have a component<br />

of intrinsic sphinteric deficiency, whether dynamically induced by the<br />

anatomic movement of a hypermobile bladder and urethra or inherently<br />

acquired by intrinsic <strong>dysfunction</strong> of the urethral closure mechanism,<br />

regardless of its anatomic position. The consolidation of the anatomic<br />

hypermobility and intrinsic sphincteric deficiency as etiologies of SUI<br />

provides a fundamental basis for expanding individual patient treatm<br />

options that incorporate various aspects of each management and treatment<br />

approach available for this condition.<br />

After the diagnosis and classification of SUI are established, the<br />

three major categories of intervention are broadly divided for discussion<br />

purposes into behavioral, pharmacologic, and surgical options. Successful<br />

intervention depends upon the understanding of treatment goals<br />

defined during the relationship established between the patient and<br />

physician. Ideally, the therapy should eliminate or improve symptoms,<br />

restore <strong>dysfunction</strong>al anatomy, and result in an improved quality of life<br />

for the patient. In choosing from available management and treatment<br />

options, informed patients are those who have received information on<br />

the expected outcomes, risk and benefits of each of these options for<br />

their particular problems (see Table 1). Management and treatment for<br />

all forms of urinary incontinence are traditionally staged with the least<br />

invasive forms of behavioral interventions offered first, then the introduction<br />

of pharmacologic agents when indicated and tolerated, and<br />

finally, after optimization of the pelvic floor and bladder physiology,<br />

the use of reconstructive surgery that strives to restore a normal bladder<br />

state of continence and voiding function.<br />

The long-term efficacy and cost analyses of individual treatment<br />

options clearly favor the options of all forms of intervention over the<br />

benign neglect of this condition (2). The use of surgical interventions<br />

are clearly the most cost-effective approach in the long-term, yet the<br />

increased risk associated with these treatments and the limited number<br />

of specialist trained to perform these procedures have traditionally


Chapter 8 I Stress Urinary Incontinence 165<br />

made nonsurgical options a priority. The concern of increased risk<br />

associated with these procedures are lessened with contemporary reports<br />

defining better ways to perform traditional procedures using minimally<br />

invasive approaches (3-5), improvements in anesthesia and overall<br />

reduction in their use (6,7'), and proper patient selection for procedures<br />

that continue to expand in number and indications (8). The reduction<br />

in risk associated with invasive procedures has shifted the focus of this<br />

risk management onto pharmaceutical interventions, which by necessity<br />

are used long-term and associated with significant side-effects (l),<br />

especially in the setting of increased polypharmacy use in the elderly<br />

(9). The goal of treatment intervention offered by the physician to the<br />

patient with SUI seeking an improved quality of life is the integration of<br />

the three main treatment options defined in this review. When instituted<br />

properly and analyzed as a composite, all forms of outcome measures<br />

will reveal a rewarding endeavor established between the patient and<br />

physician.<br />

BEHAVIORAL INTER~NTIO~S<br />

Behavioral therapy may be easily initiated because it includes a<br />

range of behavioral modifications that are dependent on the type of<br />

voiding <strong>dysfunction</strong> diagnosed. For patients with incontinence, the<br />

understanding of normal voiding physiology, the awareness of the<br />

quantity of fluid intake, timing of fluid intake, and avoidance of bladder<br />

stimulants can result in modest improvement of their urinary symptoms.<br />

Simple adjustment in the timing of medications such as the avoidance of<br />

diuretics before recumbency and emptying the bladder before reaching<br />

maximum functional capacity (timed 'voiding) are often overlooked<br />

changes a patient with SUI may be able to institute. Patients with<br />

congestive heart failure andlor fluid retention in the lower extremity<br />

will benefit from promoting a diuretic phase before periods of recumbency<br />

by elevation of the lower extremities when sitting, by initiating<br />

light walking programs for muscle pumping of congested veins in the<br />

leg, by using pressure gradient stockings, and by the judicious use of<br />

diuretics and an avoidance of high-salt diets. Changing access to the<br />

toilet (support bars and lighting), toilet substitutes (bedside commode),<br />

adjustment of garments for easier undressing, and providing mobility<br />

aids will prevent many episodes of mixed stress and urge incontinence


166 Racklev<br />

in patients with congenitive <strong>dysfunction</strong>, mobility problems, and physi-<br />

cal handicaps.<br />

Teaching methods to inhibit the urge sensation to void in order to<br />

gradually expand the voiding interval is a form of behavioral therapy<br />

often referred to as bladder retraining. Bladder retraining has its greatest<br />

benefit in patients with urge or mixed urge and stress incontinence.<br />

The excellent outcome results achieved with these methods should not<br />

be mistaken for the poor results associated with the treatment of pure<br />

stress incontinence or the stress component of mixed incontinence.<br />

These methods for treating urge incontinence rely on the sympathetic-<br />

mediated negative feedback inhibition augmented by the somatic-medi-<br />

ated contraction of the external urinary sphincter on bladder contrac-<br />

tions. By voluntary contraction of the external urinary sphincter, most<br />

patients are able to extinguish low-level bladder contractions (urgency).<br />

Retraining or improving a patient’s ability to increase her external<br />

urinary sphincter tone with a voluntary effort may aid in the treatment<br />

of urge and urge incontinence, lead to less urinary frequency, and<br />

improve functional bladder capacity. If the patient can predict the timing<br />

of a stress maneuver such as a cough, sneeze, or physical activity that<br />

will lead to incontinence, then bladder-retraining methods may help to<br />

augment the external urethral sphincteric resistance to leakage when<br />

employed prior to the occurrence of the stress maneuver. Other methods<br />

to inhibit the urge sensation include distraction techniques, and biofeed-<br />

back using deep breathing and relaxation exercises. All these techniques<br />

of bladder retraining appear to work best in patients who have forms<br />

of urge or mixed incontinence but not stress incontinence. Furthermore<br />

these patients must be cognitively intact, properly motivated, and have<br />

the capacity to understand and follow instructions. This select group<br />

of patients has the highest rate of improvement scores owing to the<br />

treatment of urge incontinence and are some of the few patients who<br />

actually have a documented cure of their incontinence by behavioral<br />

intervention.<br />

Biofeed&uck<br />

While biofeedback may aid in teaching patients behavioral interven-<br />

tions as previously mentioned for the treatment of urge incontinence,<br />

the merits of this therapy for treating stress incontinence are due to<br />

pelvic-floor muscle rehabilitation, training, and control. The goal of<br />

this form of behavioral therapy is to teach the patient to increase


Chapter 8 I Stress Urinary Incontinence 167<br />

intraurethral pressure by contracting the periurethral muscles prior to<br />

activities that cause stress incontinence. Distinct from the concept of<br />

negative feedback inhibition on the bladder by external urethral sphincter<br />

contraction for treating urge incontinence, exercising the pelvicfloor<br />

muscles (the levator ani muscles) may result in decreasing urethral<br />

hypemobility and increasing urethral resistance for the prevention of<br />

SUI. Kegel exercises were originally proposed in 194 1 for strengthening<br />

the pelvic-floor musculature (lo), but since that time, these exercise<br />

have been improperly taught and improperly used by patients (11).<br />

Although many patients can be instructed on the identification of the<br />

proper muscle group for pelvic-floor muscle exercises in the office,<br />

the best results are achieved by repetitive training, use of biofeedback,<br />

and consultation or initiation of therapy by a physical therapist or a<br />

nurse with specialized training (12). Manometry and electromyography<br />

can be used to measure external and sphincter activity or vaginal muscle<br />

activity when teaching pelvic muscle contraction and relaxation. The<br />

most common error in pelvic muscle training involves the contraction<br />

of other muscle groups (abdominal, back, gluteal, or thigh muscles).<br />

Once the pelvic muscles are located and control obtained, the next<br />

goal of therapy is to increase the patient’s strength. Pelvic muscle<br />

rehabilitation alone can increase resting muscle tone, but active maximal<br />

contraction of the periurethral muscles through control and strengthening<br />

is needed to prevent incontinence through the urethra due to transient<br />

rises in abdominal pressures.<br />

Other forms of therapy for pelvic-floor muscle rehabilitation include<br />

intravaginal weighted cones for exercise, the use of electrical stimulation,<br />

and the introduction of electromagnetic energy. Vaginal cones<br />

introduced in 1985 for strengthening pelvic-floor muscles provide an<br />

inexpensive, effective, nonlabor intensive or time-consuming treatment<br />

for mild stress incontinence (13). Electrical stimulation involves the<br />

application of electrical current to the pelvic-floor muscle via rectal or<br />

vaginal electrodes. Although the side effects are minimal, the results<br />

appear to last only as long as the device is used, and may not show<br />

any added advantage over properly performed exercises except in those<br />

patients with a denervated pelvic-floor musculature (14,15). In addition<br />

to rehabilitation of the levator ani muscles, the rehabilitation of the<br />

obturator internus muscle may be of importance, as this muscle with<br />

its attendant fascia shares a common fascial insertion with the levator<br />

ani muscles called the arcus tendinous fascia. Whereas pelvic floor<br />

rehabilitation may appear to be useful for all patients, the etiology of


168 Rackley<br />

SUI may result from a traumatic peripheral neuropathy ‘with resultant<br />

muscle atrophy and <strong>dysfunction</strong> typical of intrinsic sphincteric defi-<br />

ciency, rather than fascial breaks or lengthening associated with ana-<br />

tomic hypermobility. Therefore, standard rehabilitation techniques<br />

require at least partial innervation to the muscle groups of interest<br />

unless the nerves are bypassed by an electrical stimulator or electromag-<br />

netic energy.<br />

Currently, there is no reproducible neurologic test that can determine<br />

which patients are likely to benefit from pelvic floor rehabilitation. The<br />

low risk of this noninvasive therapy, not the cost (office visits, labor<br />

intensity), has made pelvic-floor rehabilitation an acceptable first-line<br />

therapy for patients with mild SUI owing to anatomic hyperrnobility,<br />

not intrinsic sphinteric deficiency.<br />

Containment Devices and Support Prosthesis<br />

Conservative treatments such as urethral meatal occlusive devices,<br />

urethral inserts, plugs, and catheters, as well as vaginal pessaries (16-<br />

18), tampons, and contraceptive devices (19) for bladder neck support<br />

provide successful management options for selected patients with stress<br />

incontinence. The success is dependent on multiple factors including<br />

patient selection, patient motivation, and training of the patient and<br />

instructor. Extensive patient education by a health care professional<br />

requires significant time and expensive allocations that are resulting in<br />

the continued trend to develop nonsurgical alternatives that require<br />

minimal patient education with maximal acceptance. While use of these<br />

devices are not curative of the patient’ S underlying voiding <strong>dysfunction</strong>,<br />

they do provide management options for their symptoms that may serve<br />

as an adjuvant to other interventions listed below or may be the only<br />

choice of treatment available to the patient for a variety of medical<br />

and social reasons.<br />

Estrogens<br />

The ability of sex hormones to induce physiological changes of<br />

the lower urinary tract during the menstrual cycle, pregnancy, and<br />

menopause is the rational for using these agents pharmacologically to


Chapter 8 I Stress Urinary Incontinence 169<br />

maintain and augment urinary continence. In particular, estrogens are<br />

known to have a direct effect on lower urinary-tract organs and to<br />

modulate the autonomic nervous system activity of the bladder and<br />

urethra (20,21).<br />

Since the majority of women with stress incontinence are perimenopausal<br />

or postmenopausal, restoring the tissue integrity of the urethral,<br />

bladder neck, and vaginal epithelium with the use of estrogen therapy<br />

serves as a fundamental adjuvant to the other management options<br />

available for urinary incontinence. Clinical studies have revealed its<br />

subjective usefulness especially in conjunction with behavioral interventions<br />

and pharmacologic agents (22). By promoting upregulation<br />

of neurotransmitter receptor function, estrogen therapy may augment<br />

the effect of anticholinergics and alpha-adrenergic agonist (23). Therefore,<br />

it is recommended as baseline therapy for all forms of urinary<br />

incontinence and is usually initiated as a first-line intervention following<br />

the detection of estrogen deficiency on the pelvic examination performed<br />

on the initial physician visit. For patients with long-standing<br />

estrogen deficiency, symptoms of stress incontinence may not change<br />

with replacement therapy, but the restoration of tissue integrity through<br />

the increase in blood supply will aid in tissue manipulation associated<br />

with surgical interventions, urethral occlusive devices, or vaginal pessary<br />

use.<br />

Estrogens are used either systemically or topically. Side effects<br />

may include an increased risk for gynecological malignancies, fluid<br />

retention, depression, nausea, vomiting, elevated blood pressure, gallstones,<br />

and cardiovascular effects such as stroke and myocardial infarction.<br />

However, in addition to the aforementioned benefits of improved<br />

voiding function, hormone replacement therapy confers an overall benefit<br />

in mortality as compared to nontreated patients (24). Topical vaginal<br />

application through direct placement or vaginal implants are preferable<br />

because of the decreased systemic absorption and preferential local<br />

uptake (25). When mild atrophic changes of the vaginal epithelium are<br />

noted on pelvic examination despite systemic estrogen replacement,<br />

additional topical vaginal therapy may be useful in order to improve<br />

voiding function without incurring side effects from increasing the<br />

systemic administration. Yearly PAP smears and breast examinations<br />

should be performed when placing or following patients on estrogen<br />

therapy. In general, the use of estrogens is restricted in patients with.<br />

a history of deep venous thrombosis, embolization, and gynecologic<br />

malignancies.


170 RacMey<br />

Adpba-Adrenergic Agents<br />

Because sympathetic-nerve stimulation of the alpha-adrenergic<br />

receptors located in the intrinsic sphincter or bladder neck region produces<br />

smooth muscle contractions, cases of mild stress incontinence<br />

have responded to the use of oral alpha-adrenergic agents (26). In the case<br />

of stress incontinence, the most commonly used agents are ephedrine,<br />

pseudoephedine, and phenylpropanolamine; all are the active components<br />

of commonly used over-the-counter decongestants. The usual adult<br />

dosage of phenylpropanolamine is 25-100 mg in sustained-release form,<br />

given orally twice a day; that of pseudoephedrine is l 5-30 mg three times<br />

a day. Side effects are numerous and are associated with the drug’s lack<br />

of bladder neck selectivity. Typical manifestations of the nonspecific<br />

action of this medication include an increase in blood pressure, stomach<br />

cramping, and central nervous system (CNS) symptoms of excitation<br />

and drowsiness. These drugs must be used with caution in patients with<br />

hypertension, arrhythmias, angina, respiratory difficulties, hyperthyroidism,<br />

and diabetes. Imipramine, a tricyclic antidepressant, has adual action<br />

on the lower urinary tract by its effect on bladder contractions through<br />

its anticholinergic effects while increasing bladder neck resistance<br />

through its alpha-adrenergic agonist properties. Although only approved<br />

by the FDA for enuresis in children, the usual adult dosage is 10-25 mg<br />

once to three times daily. Side effects may include postural hypotension<br />

and cardiac conduction disturbances in the elderly. Reported outcomes<br />

measures for using a pharmacologic agent such as an alpha-adrenergic<br />

medication are clinically significant, both from an efficacy as well as a<br />

side-effect profile.<br />

SURGICAL MANAGEMENT<br />

Introduction<br />

As stated earlier, SUI may be owing to anatomic hypermobility<br />

or intrinsic sphincteric deficiency of the urethral closure mechanism.<br />

Although no standardized methodology exist to clearly define these<br />

two forms of stress incontinence, most experts would agree that the<br />

available methods of urethral evaluation can determine the contribution<br />

of these two factors in an individual patient. Thus, the classification<br />

of stress incontinence exists between the spectrum of anatomic displacement,<br />

which induces a transient incompetence of the urethral closure<br />

mechanism to a well-supported urethral that has intrinsic incompetence


Chapter 8 / Stress Urinary Incontinence 171<br />

of its closure mechanism. Therefore, by definition, any women with<br />

evidence of SUI must have some component of intrinsic sphincteric<br />

deficiency, whether induced transiently by the dynamic hypermobility<br />

movement of the urethra during a stress maneuver or caused by a rise<br />

in intra-abdominal pressure that when transmitted through the stable<br />

bladder is greater than the closure pressure of the incompetent but<br />

supported urethra.<br />

Clinical experience with objective urodynamic evaluations (27’) sup-<br />

port this classification of stress incontinence. The efficacy of any treat-<br />

ment option is based on how well the recommendation will address these<br />

two components of stress incontinence and is why surgical procedures<br />

produce the highest long-term efficacy rates of all available treatments.<br />

The surgical treatments outlined below have evolved in design over<br />

the years to better address the following:<br />

1. Restoration of the urethra to its proper resting position and to stabilize<br />

this position during increases in intra-abdominal pressure;<br />

2. Augmentation of intraurethral pressures for restoration of intrinsic<br />

urethral closure; or<br />

3. A combination of both restoration of support and augmentation of<br />

urethral closure.<br />

Implants<br />

Periurethral and transurethral injection of bulking agents at the level<br />

of the proximal urethra have been used extensively for years to increase<br />

the outflow resistance of the urethra for the treatment of SUI due to<br />

intrinsic sphincteric deficiency. Injectables are able to increase urethral<br />

closure function without significantly resulting in increases of urethral<br />

closure pressure, which would lead to a rise in voiding pressure. The<br />

effect is to correct the incompetent urethral closure mechanism without<br />

clinically disturbing voiding function.<br />

Although few are currently in use, many choices for injectable sub-<br />

stances have been tried; they include sclerosing solutions (28), polyte-<br />

trafluroethylene (PTFE) paste (29), glutaraldehyde cross-linked bovine<br />

collagen (GAX-collagen) (30), autologous fat (31), silicone particles<br />

(32), and many more. Currently, the only injectable materials acceptable<br />

for use in the United States are autologous material such as fat, and<br />

xenogenic collagen, Contingen TM (C.R. Bard Inc., Covington, CA).<br />

Autologous fat provides the advantages of easy accessibility, availabil-<br />

ity, affordability, and biocompatibility; however, the efficacy is abysmal<br />

owing to poor graft neovascularity and long-term viability (1O--ZO%)


172 Racklet.<br />

(33). The currently used GAX-collagen previously mentioned is a suspension<br />

of 35% highly purified bovine collagen in a phosphate buffer.<br />

The cross-linking and hydrolysis of the bovine collagen that is 95%<br />

type I and 1-5% type I1 collagen reduces its antigenicity and increases<br />

its resistance to collagenase degradation. Allergic reactions are still<br />

possible and skin testing is mandatory and may be positive in 4% of<br />

females (30). Adverse events occurring with this injectable are rare<br />

and self-limiting; they include transient retention (6%), urinary tract<br />

infection (6%), de novo urgency (12%), and hematuria (3%) as reported<br />

in the North American Contigen Study Group (34).<br />

Initial studies such as the North American Contigen Study Group<br />

suggested that the best candidates for successful Contigen injectable<br />

therapy were patients with good anatomic urethral support and intrinsic<br />

sphincteric deficiency. Additional studies which have followed the<br />

initial experiences with Contigen demonstrate equal efficacy in the<br />

treatment of patients with anatomic hypermobility and/or intrinsic<br />

sphincteric deficiency (35-37). These reports have demonstrated overall<br />

success rates of 80% (40% cures, 40% improved) with the probability<br />

of remaining dry without additional collagen at 72% at 1 yr, 57% at<br />

2 yr, and 45% at 3 yr. If deterioration in continence status occurs after<br />

achieving a successful outcome, repeat collagen injections will restore<br />

a successful condition.<br />

Based on the aforementioned reports, all patients with SUI are potential<br />

candidates for Contigen injectable therapy as long as they do not<br />

have a known hypersensitivity to bovine collagen, an untreated urinary<br />

tract infection, or unmanaged detrusor instability, which may obviate<br />

the patient’s subjective awareness of achieving a successful treatment<br />

of stress incontinence. Techniques and results of transurethral and<br />

periurethral injection have been reviewed in depth in several reports<br />

(38). For patients with intrinsic sphincteric deficiency, a larger amount<br />

of injectable material appears to be needed. For patients with transient<br />

induction of intrinsic sphincteric deficiency owing to anatomic hypermobility<br />

in which the posterior urethra moves away from the anterior<br />

urethra during rotational descent of the proximal urethra, technical<br />

consideration of placing the bulking agent between the 5:OO and 7:OO<br />

o’clock position is warranted in order to preserve a coapted urethral<br />

closure mechanism despite anatomic movement or location.<br />

While injectable therapy for stress incontinence has a learning curve<br />

and requires cystoscopic equipment, all patients may receive this therapy<br />

in the office or outpatient surgical setting under a local block and<br />

no postprocedure restrictions. The reliable, cost-effective and well-


Chapter 8 I Stress Urinary Incontinence 173<br />

controlled method of administering this therapy places this surgical<br />

option at the forefront of all treatment algorithms for treating stress<br />

urinary incontinence in the absence of pelvic organ prolapse. While<br />

the safety and efficacy profile of Contigen injectable therapy is excel-<br />

lent, the ideal material is still being sought and should combine the<br />

ease of administration with minimal tissue reaction, no material migra-<br />

tion, and persistence over time. Combined with estrogen replacement<br />

therapy and behavioral modification, injectable therapy will continue<br />

to find a high acceptance among patients and physicians.<br />

Bhdder Neck Suspension Procedures<br />

RETROPUBIC APPROACHES<br />

FOR BLADDER NECK SUSPENSIONS<br />

In unoperated patients with stress incontinence predominantly owing<br />

to anatomic incontinence defined as hypermobility of the bladder neck<br />

and urethra with high abdominal leak-point or urethral pressure profiles,<br />

bladder neck suspensions procedures give adequate results in this highly<br />

selective group. The etiology and important implication that all stress<br />

incontinence is owing to various degrees of intrinsic sphincteric deficiency<br />

has only recently been considered since the original urodynamic<br />

description of this forrn of urethral <strong>dysfunction</strong> was introduced by<br />

McGuire in 1981 (39). At the time bladder neck suspensions were<br />

introduced in 1949 for correction. of stress incontinence, hypersuspensions,<br />

and overcorrection of anatomic changes of the bladder neck<br />

support were the prevailing theories. The first publication of an attempt<br />

to suspend the bladder neck describes an anterior urethropexy technique,<br />

Marshall-Marchetti-Krantz procedure, which has mostly fallen out of<br />

favor owing to the complications from sutures placed close or into<br />

the anterior urethra with subsequent high retropubic fixation to the<br />

symphysis pubis (40). Interestingly, the authors of this landmark paper<br />

suggested that stress incontinence was not entirely owing to lack of<br />

anatomic support since correction of the hypermobility failed to successfully<br />

treat stress incontinence in some patients. Modifications of retropubic<br />

bladder neck suspension procedures, which include stable suture<br />

fixation into Cooper’s ligament and lateral placement of the suspending<br />

sutures into the periurethral tissue at the level of the bladder neck, have<br />

popularized the use of the Burch bladder neck suspension (Fig. 1).<br />

With satisfactory long-term outcomes, the open abdominal Bur& procedure<br />

is considered one of the gold standards in the surgical treatment


174 Racklev<br />

Fig. 1. Burch bladder neck suspension.<br />

of anatomic stress incontinence when an abdominal approach is indicated.<br />

In two contemporary reviews of surgical procedures for the<br />

treatment of stress incontinence, retropubic bladder neck suspensions<br />

achieve an 83435% long-term objective success rate when used as a<br />

first-time procedure (41,42). The disadvantages of open abdominal<br />

approaches are well documented and the historical trend towards less<br />

invasive procedures popularized the development of transvaginal needle<br />

suspensions from 1959 to the 3990s.<br />

TRANSVAGINAL APPROACHES<br />

LADDER NECK SUSPENSIONS<br />

Pereyra introduced the first description of a transvaginal approach<br />

to bladder neck suspensions that was later modified into its contemporary<br />

form by Raz and is now referred to as the Pereyra-Raz procedure<br />

(Fig. 2) (43). Despite the realization of poor long-term success of this<br />

procedure in recent reviews (43,44), several fundamental principles<br />

popularized through the use of this technique for surgical correction<br />

of stress incontinence warrant recognition. This transvaginal approach<br />

uses intraoperative endoscopy to insure procedural safety. Owing to<br />

the breakdown of the native supporting endopelvic fascia for entering<br />

the retropubic space, permanent suture to repair the iatrogenic paravaginal<br />

defect and to provide suspension of the bladder neck and urethra .<br />

is required for long-term success (45). By 1973, Stamey introduced<br />

the concept of preserving the native endopelvic fascia support by using<br />

a ligature carrier, which is negotiated through the retropubic space<br />

under endoscopic control (46). To prevent pull-through failures, pledget


Chapter 8 / Stress Urinary Incontinence 175<br />

a<br />

Fig. 2. Pereyra-Raz bladder neck suspension.<br />

support was added to the sutures at the level of the endopelvic fascia<br />

and surgeons were advised to stabilize, not suspend, the urethra and<br />

bladder neck competency. When analyzed collectively, the long-term<br />

outcome success of 67% for these transvaginal procedures appears to<br />

be only slightly better than the 61 % success rate for the use of the<br />

obsolete anterior colporrhaphy procedure (42). The real merits of trans-<br />

vaginal approaches for bladder neck suspensions is based on incorporat-<br />

ing the knowledge gained from the clinical experience of performing<br />

these procedures into other retropubic and transvaginal approaches for<br />

the surgical correction of stress incontinence as demonstrated in the<br />

percutaneous bladder suspension procedure (Fig. 3) (2,4,46-50). These<br />

merits include the following:<br />

S.<br />

2.<br />

3.<br />

4.<br />

Retropubic dissection, which allows for scarification that supplements<br />

the native support of the endopelvic fascia;<br />

Use of permanent suture material for repair of an iatrogenic paravaginal<br />

defect;<br />

Stable suture fixation which allows for unstressed scarification and<br />

potentially prevents suture pull-through failures from the vaginal wall<br />

and endopelvic fascia; and<br />

Adjusting the tension of the supporting sutures to ensure stabilization,<br />

not suspension or overcorrection of the bladder neck and urethra.<br />

LAPAROSCOPIC APPROACHES<br />

FOR BLADDER NECK SUSPENSIONS<br />

Interest in developing laparoscopic approaches to treat SUI has<br />

increased due to reduced patient morbidity and improved convalescence


176 Rackley<br />

Arcus<br />

lev<br />

Symphysis pubis<br />

~rethrope~vi~ complex<br />

Fig. 3. Percutaneous bladder neck suspension.<br />

over traditional open, retropubic bladder neck suspensions (Fig. 3).<br />

While the early results of laparoscopic bladder neck suspensions are<br />

promising (51), the approach of this minimally invasive procedure has<br />

unfortunately led to changes in the method of a performing the Burch<br />

procedure to the point where few direct outcome comparisons can be<br />

made between the open vs laparoscopic approach (52). Changing the<br />

surgical access should not change the surgical method of performing a<br />

retropubic suspension; however, the fundamental techniques responsible<br />

for the success of the traditional open procedures have been abandoned in<br />

the rush to improve the operative time of this minimally invasive surgery<br />

using suture fixation and synthetic mesh devices (53,5#).<br />

Sling Procedures<br />

Unlike bladder neck suspensions, sling procedures are specifically<br />

designed to address both anatomic hypermobility and intrinsic sphinc-<br />

teric deficiency components of SUI. Sling procedures have generally<br />

been reserved for the treatment of complex or previously failed cases<br />

of SUI, which has obviated their direct comparison to bladder neck<br />

suspensions in patients undergoing primary treatment. Despite the his-<br />

torical use of sling procedures in complex cases of intrinsic sphinteric<br />

deficiency, the long-term success rate of 83% combined with the report<br />

of general medical and surgical complication rates equal to transvaginal<br />

or retropubic bladder neck suspension techniques suggest that the sling<br />

procedure should be considered as the primary surgical treatment of<br />

choice over bladder neck suspensions for stress incontinence (42). As


Chapter 8 / Stress Urinary Incontinence 177<br />

Fig. 4. Modified pubovaginal sling.<br />

outlined previously, incorporation of minimally invasive techniques of<br />

bladder neck suspensions has led to Contemporary modifications of sling<br />

procedures, which are now considered technically easier to perform than<br />

bladder neck suspensions.<br />

Although conceptually introduced in the early 1900s in various<br />

complex surgical forms, the contemporary reemergence of sling procedures<br />

is credited to the introduction of the rectus fascia pubovaginal<br />

sling by McGuire and Lyton in 19'78 (55). Modifications of the original<br />

procedure include shorter sling length, less traumatic disruption of the<br />

native endopelvic fascia support, and securing the sutures to a nonmobile<br />

point of fixation (Fig. 4).<br />

In addition to autologous fascia, which has the advantages of durability<br />

and lower removal rate secondary to infection and erosion (42),<br />

allograft (56) and synthetic (57) materials are often considered as a<br />

substitute sling choice based on ease of acquisition and surgeons preference<br />

for perf'oming quicker operative procedures. While a variety of<br />

other autologous tissues have been introduced, the use of the vaginal<br />

wall as a sling material by Raz (58) has undergone minor changes that<br />

have resulted in its acceptance as a viable alternative for use in patients<br />

with acceptable tissue quality and vaginal capacity, especially those<br />

with concomitant poor detrusor function (59). The in situ vaginal wall<br />

sling (Fig. 5) with preservation of the urethropelvic complex (endopelvic<br />

fascia and vaginal wall attachments to the arcus tendineous)<br />

(60) provides durable results in patients with an abdominal leak point


178 RacMey<br />

Fig. 5. In situ vaginal wall sling.<br />

pressure greater than 50 cm of water (61). While many proponents of<br />

bladder neck suspensions such as the Burch procedure contend that<br />

these procedures do to some degree support the urethra on a sling of<br />

vaginal wall tissue owing to the location of the suppoting sutures, it<br />

is interesting to note in our recent report that patients undergoing an<br />

in situ vaginal wall sling with a Valsalva leak-point pressure less than<br />

50 cm of water pressure (equivalent to a urethral closure pressure of<br />

20 cm or less) have a failure rate similar to that of a Burch procedure<br />

(62). The similarity for risk of failure for these two nearly identical<br />

procedures that differ mainly in surgical approach, suggest that the<br />

elastic properties of the vaginal wall makes this sling material a poor<br />

choice in patients with overtly poor urethral incompetence who will<br />

need more suburethral support than can be afforded by the vaginal wall<br />

tissue. Thus, in patients undergoing urodynamic evaluation for poor<br />

detrusor function with concomitant stress incontinence, the degree of<br />

intrinsic incompetence must be ascertained when a choice of using the<br />

vaginal wall as a sling material is being considered.<br />

Sling procedures have traditionally been performed from an abdominal-perineal<br />

approach and more recently by laparoscopic methods<br />

(63,64). Two procedures incorporating a transvaginal approach are now<br />

under consideration. With the use of transvaginal bone anchors into the<br />

undersurface of the inferior pubic rami or areas of the pubic symphysis,<br />

fixation of the supporting sling sutures can be placed from a transvaginal<br />

approach (65). Ulmsten has introduced a technique of synthetic sling


Chapter 8 / Stress Urinary Incontinence 179<br />

placement via a transvaginal, retrograde, retropubic passage of the<br />

supporting sling to the level of the rectus fascia (5). Owing to the<br />

coefficient of friction between the sling material and all the intervening<br />

tissue, no effort is made to tie down the sling material to the rectus<br />

fascia in an attempt to reduce urethral obstruction from hypersuspension<br />

of the sling.<br />

Traditional sling procedures have a long history of producing successful<br />

long-term outcomes for patient with stress incontinence due to<br />

intrinsic sphincteric deficiency. With recent reports revealing that the<br />

general medical and surgical complications (42), as well as the technical<br />

skills of these procedures are equal to those of bladder neck suspensions,<br />

the indications for slings have been extended to treat stress incontinence<br />

owing to anatomic hypermobility (66). Urodynamic evaluation of urethral<br />

competency, not detrusor <strong>dysfunction</strong>, may be obviated by choosing<br />

to use fascial and synthetic slings since these procedures can safely<br />

and successfully treat the entire spectrum of intrinsic sphincteric deficiency.<br />

Patients with stress incontinence and a urethral competence<br />

characterized or clinically judged as greater than 50 cm of water will<br />

do well with a vaginal wall sling procedure (61). Furthermore, based<br />

on comparative clinical outcome studies (59), the vaginal wall sling<br />

results in less mechanical obstruction of the urethra and should be<br />

considered as the sling of choice in patients with documented poor<br />

detrusor function and concomitant stress incontinence. There are currently<br />

no absolute indications for allograft or synthetic materials for<br />

sling formation. As such, concern over early failure rates of allograft<br />

fascia (67) and higher removal rates of synthetic materials (57) have<br />

tempered the enthusiasm for their use on a routine basis. With the<br />

introduction of techniques and technology that provide placement of<br />

slings entirely from a transvaginal approach that obviates counter incisions<br />

for harvesting autologous sling material, the need for obtaining<br />

material other than the vaginal wall will continue to drive the search<br />

for improvements in allograft, heterologous, and synthetic material. As<br />

improvements in behavioral, medical, and other surgical interventions<br />

for treating stress incontinence continue, surgical operations for stress<br />

incontinence will be reserved for patients with the concomitant need<br />

of additional abdominal or transvaginal surgery (tubal ligation, cholecystectomy,<br />

pelvic organ prolapse, and so on). Thus, the approach<br />

required for the additional concomitant surgery and knowledge of the<br />

degree of urethral competency will greatly influence the decision of<br />

the surgical technique chosen to treat the stress incontinence.


180 Raclcley<br />

Artificial Urinary Sphincter<br />

If the goal of surgical treatment is to establish a normal voiding<br />

pattern that allows the patient to remain dry between voiding, then the<br />

artificial urinary sphincter is one of the best treatment options capable<br />

of achieving this result. The artificial urinary sphincter is intended to<br />

permit intermittent urethral compression for maintaining continence<br />

with voluntary reduction of urethral resistance during voiding, as<br />

opposed to other surgical procedures in which the compressive nature<br />

of static urethral resistance (i.e., injectable material, sling, suspension<br />

of periurethral tissue) to incontinence also results in voiding against a<br />

relative obstruction. However, not all females are candidates for the<br />

artificial urinary sphincter; indications for use include adequate manual<br />

dexterity, mental capacity, motivation to manipulate the device each<br />

time for voiding, normal detrusor function, and normal bladder compliance.<br />

Widespread use of the artificial urinary sphincter is limited by<br />

its technical difficulty in placement whether transvaginal or abdominal,<br />

and is usually reserved for patients with complex etiologies of SUI<br />

(68-70). With the infection rate for primary artificial urinary sphincter<br />

implants at 14% and the nonmechanical failure rate at 13%, the overall<br />

5-yr expected product survival is 72% (71). The dry continence rates<br />

of 67% limits the use of this device as a first line surgical treatment<br />

for SUI despite overall patient satisfaction near rates 92%. When considered<br />

in the context of implantation for complex cases of stress incontinence<br />

who have failed multiple prior procedures, the overall success<br />

and patient satisfaction rates are extremely impressive for these subset<br />

of patients with limited options for achieving continence.<br />

SUMMARY<br />

The overall emphasis on successful management of female urinary<br />

incontinence begins with the combination of a thorough evaluation,<br />

selection of individualized behavior modifications, and possibly pharmacologic<br />

supplementation, Only for patients who have failed to<br />

achieve an acceptable continence status with nonsurgical management<br />

and who have maximized their potential for overall pelvic-floor reha<br />

tation and voiding function should reconstructive surgery then be considered.<br />

A recent report (1) of a 10-yr expected cost per elderly patient


Chapter 8 I Stress Urinary Incontinence 181<br />

with chronic stress incontinence in 1994 dollars reveals that untreated<br />

incontinence is the most expensive health care choice ($86,726). Com-<br />

parative costs are the lowest for surgical therapies and the highest for<br />

behavioral therapy (bladder neck suspensions, $25,388; pharmacologic<br />

therapy, $62,021 ; and behavioral therapy, $68,924). Through continued<br />

advances in prevention, patient evaluation, and optimization of the<br />

pelvic floor and bladder function, the current and future potential for<br />

increased long-term effectiveness of all management options for<br />

patients who seek an acceptable continence status to improve their<br />

quality of life will continue to have a favorable impact on the over-<br />

all cost.<br />

REFERENCES<br />

1. Ramsey SD, Wagner TH, Bavendam TG (1996) Estimated costs of treating stress<br />

urinary incontinence in elderly women according to the AHCPR clinical practice<br />

guidelines. Am J Man Cure 2:147-150.<br />

2. Appell RA (1997) The use of bone anchoring in the surgical management of<br />

female stress urinary incontinence. World J Urol 15:300-305.<br />

3. Vasavada SP, Rackley RR, Appell RA (1998) In situ anterior vaginal wall sling<br />

formation with preservation of the endopelvic fascia for treatment of stress urinary<br />

incontinence. Intl Urogynecol J 9:379-384.<br />

4. Levin S, Bennet AE, Levin D, Danielli L, Levin R, Sidi A (1998) Minimally<br />

invasive surgical treatment of female stress urinary incontinence. Intl Urogynecol<br />

J 9:405-408.<br />

5. Ulmsten U, Henriksson L, Johnson P, Varhos G (1996) An ambulatory surgical<br />

procedure under local anesthesia for treatment of female urinary incontinence.<br />

Intl Urogynecol J 7:81-85.<br />

6. Haab F, Leach G (1997) Feasibility of outpatient percutaneous bladder neck<br />

suspension under local anesthesia. Urology 50(4):585-581.<br />

7. Carr LK, Walsh PJ, Abraham VE, Webster GD (1997) Favorable outcome of<br />

pubovaginal slings for geriatric women with stress incontinence. J Urol<br />

157(1):125-128.<br />

8. Urinary Incontinence Guideline Panel (1992) Urinary Incontinence in Adults: A<br />

Patient’s Guide. Rockville, MD: US Dept of Health and Human Services, PHs,<br />

AHCPR; AHCPR pub. no. 92-0040.<br />

9. Ghoniem GM, Hassouna M (1997) Alternatives for the pharmacologic management<br />

of urge and stress urinary incontinence in the elderly. J Wound Ostomy<br />

Contin Nurs 24(6):3 11-3 18.<br />

10. Kegel A (1949) Progressive resisting exercise in the functional restoration of the<br />

perineal muscles. Am J Obstet Gynecol 56:238-240.<br />

11. Bo K, Larger S, David S, et al. (1988) Knowledge about and ability to correct<br />

pelvic floor muscle exercises in women with urinary stress incontinence. Neurourol<br />

Urodyn 71:261-263.


Rackley<br />

12. Burgio KL, Robinson JC, Engel BT (1986) The role of biofeedback in Kegel<br />

exercise training for stress urinary incontinence. Am J Ostet Gynecol 154:58-64.<br />

13. Wilson DP, Borland M (1995) Vaginal cones for the treatment of genuine stress<br />

incontinence. Aust N Z J Obstet Gynaecol 95: 1049-105 1.<br />

14. Jonasson A, Larsson, Pschera H, Nylund L (1990) Shortterm maximal electrical<br />

stimulation: a conservative treatment of urinary incontinence. Gynecol Obstet<br />

Invest 30:120-124.<br />

15. Richardson DA, Miller KL, Siege1 SW, Karram MM, Blackwood NB, Staskin<br />

DR. (1996) Pelvic floor electrical Stimulation: A comparison of daily and everyother-day<br />

therapy for genuine stress incontinence. Urology 48:110-118.<br />

16. Nygaard I. (1995) Prevention of exercise incontinence with mechanical devices.<br />

J Reprod Med 40:89-94.<br />

17. Hahn I, Milson I (1996) Treatment of female stress urinary incontinence with<br />

a new anatomically shaped vaginal device (Conveen Continence Guard). Br J<br />

Urol 77:711-715.<br />

18. Davilla WD, Ostermann KN (I 996) The bladder neck support prosthesis: a nonsurgical<br />

approach to stress incontinence. Design, efficacy and short term safety. Acta<br />

Obstet Gyn Scand 75:170-173.<br />

19. Suarez GM, Baum NH, Jacobs J (1991) Use of standard contraceptive diaphragm<br />

in management of stress urinary incontinence. Urology 37: 119-122.<br />

20. Rud T (1980) Urethral pressure profile in continent women from childhood to<br />

old age. Acta Obstet Gynecol Scand 59:331--335.<br />

21. Batra SC, Losif CS (1986) Female urethra: A target for estrogen action. J Urol<br />

291418-420,<br />

22. Fantl JA, Cardozo L, McClish DK, et al. (1994) Estrogen therapy in the management<br />

of urinary incontinence in postmenopausal women: a meta-analysis. First<br />

report of the hormones and urogenital therapy committee. Obstet Gynecol<br />

83:12-18.<br />

23. Walter S, Kjaergaard B, Lose G, et al. (1990) Stress urinary incontinence in<br />

postmenopausal women treated with oral estrogen (estriol) and an alpha adrenoceptor<br />

stimulating agent (phenylpropanolamine): A randomized double blind placebo<br />

controlled study. Intl Urogynecol J l :74-78.<br />

24. Goldstein F, Stampfer MJ, Colditz GA, et al. (1997) Postmenopausal hormone<br />

therapy and mortality. New Engl J Med 336(25):1769-1775.<br />

25. Johnston A (1996) Estrogens: pharmacokinetics and pharmacodynamics spe- with<br />

cial reference to vaginal administration and the new estradiol formulation-Estring.<br />

Acta Obstet Gynecol Scan 165: 16-25.<br />

26. Wein AJ, Van Arsdale KN (1991) Pharmacologic therapy. In: Krane RJ, Siroky<br />

MB, eds., Clinical-Neuro-Urology. Boston: Little Brown, pp. 1-33.<br />

27. English SF, Amundsen CL, McGuire EJ (1999) Bladder neck competency at rest<br />

in women with incontinence. J UroE 161:578-580.<br />

28. Murless BC (1938) The injection treatment of stress incontinence. J Obstet Gynaecol<br />

Br Emp 45:67-71.<br />

29. Cole ED (ed.) (1990) Diagnostic and therapeutic technology assessment (DATTA)<br />

JAMA 269:263.<br />

30. Appell RA (1994) Collagen injection therapy for urinary incontinence. Urol Clin<br />

NAm 21:177-182.<br />

31. Santarosa RP, Blaivas JG (1994) Periurethral injection of autologous fat for the<br />

treatment of sphincteric incontinence. J Urol 15 1 :607-611.


Chapter 8 / Stress Urinary Incontinence 183<br />

32. Buckley JF, Scott R, Meddings R, Lingham V, Buck AC, Aitchison M, Deane<br />

RF, Kirk D, and Kyle KF (1992) Injectable silicone microparticles: a new treatment<br />

for female stress incontinence. J Urol 1473266A.<br />

33. Bartynski J, Marion MS, Wang TD (1990) Histopathologic evaluation of adipose<br />

autografts in a rabbit ear model. Otolaryngol Head Neck Surg 102:314-321.<br />

34. McGuire EJ, Appell RA (1994) Transurethral collagen injection for urinary incontinence.<br />

Urology 43:413-415.<br />

35. Faerber GJ (1 996) Endoscopic collagen injection therapy in elderly women with<br />

type l stress urinary incontinence. J Urol 155(2):512-514.<br />

36. Smith DN, Appell RA, Winters JC, Rackley RR (1997) Collagen injection therapy<br />

for female intrinsic sphincteric deficiency? J Urol 157: 1275.<br />

37. Herschorn S, Radomski SB (1997) Collagen injections for genuine stress urinary<br />

incontinence: patient selection and durability. Zntl Urogynec J 8: 18-24.<br />

38. Winter JC, Appell RA (1996) Bioinjectables in the treatment of urinary incontinence.<br />

In: Smith AD, Badlani DH, Clayman RV, Jorda GH, Kavoussi LR, Lingeman<br />

JE, et al., eds., Smith’s Textbook of EndouroZogy. St. Louis: Quality Medical<br />

Publishing Inc. 1255-1265.<br />

39. McGuire EJ, Woodside JR, Borden TA, Weiss RM (1981) Prognostic value of<br />

urodynamic testing in myelodysplastic patients. J Urol 126:205-209.<br />

40. Marshall VF, Marchetti AA, Krantz KE (1949) The correction of stress incontinence<br />

by simple vesicourethral suspension. Surg Gynecol Obstet 88:509-518.<br />

41. Jarvis GJ (1994) Surgery for genuine stress incontinence. Br J Obstet Gynaecol<br />

101(2):371-374.<br />

42. Leach GE, Dmochowski RR, Appell RA, et al. (1997) Female Stress Urinary<br />

Incontinence Clinical Guidelines Panel. Summary report on the surgical management<br />

of Female Stress Urinary Incontinence. J Urol 158:875-880.<br />

43. Pererya AJ (1959) A simplified surgical procedure for the correction of stress<br />

incontinence in women. West J Surg 67:223-226.<br />

44. Trockman BA, Leach GE, Hamilton J, Sakamoto M, Santiago L, Zirnmern PE<br />

(1995) Modified Pereyra bladder neck suspension: 10-year mean followup using<br />

outcomes analysis in 125 patients. J Urd 154:1841-1847.<br />

45. Korn A (1994) Does the use of permanent suture material affect outcome of the<br />

modified Pereyra procedure? Obstet Gyn 83(1): 104-10’7.<br />

46. Stamey TA (1973) Endoscopic suspension of the vesical neck for urinary incontinence.<br />

Surg Gynecol Obstet 136:547-549.<br />

47. Gittes W, Laughlin KR (1987) No-incision pubovaginal suspension for stress<br />

incontinence. J Urol 138:568-570.<br />

48. Benderev TV (1992) Anchor fixation and other modifications of endoscopic bladder<br />

neck suspension. Urology 40:409-418.<br />

49. Nativ 0, Levine S, Madjar S, Tssaq E, Moskovitz B, Beyar M (1 997) Incisionaless<br />

per vaginal bone anchor cystourethropexy for the treatment of female stress urinary<br />

incontinence. Experience with the first 50 patients. J Urol 158: 1742- 1744.<br />

50. Leach GE (1988) Bone fixation technique for transvaginal needle suspensions.<br />

Urology 31:388-390.<br />

51. Lobel RW, Davis GD (1997) Long results term of laparascopic Burch urethropexy.<br />

J Am Assoc Gynecol hpro 4(3):341-345.<br />

52. Polascik TJ, Moore RG, Rosenberg MT, Kavoussi LR (1995) Comparison of<br />

laparascopic and open retropubic urethropexy for treatment of stress urinary incontinence.<br />

Urology 45(4):647-652.


184 Rackley<br />

53. ROSS (1996) J Two techniques of laparascopic Burch repair for stress incontinence:<br />

a prospective, randomized study. J Am Ass Gynecol Laparo 3(3):35-354.<br />

54. Chau-Su 0, Presthus J, Beadle E (1993) Laparoscopic bladder neck re-suspension<br />

using hernia mesh and surgical staples. J hparoendoscop Surg 3(6):563-566.<br />

55. McGuire EJ, Lyton B (1978) Pubovaginal sling procedures for stress incontinence.<br />

J Urol 119232-85.<br />

56. Wright JE, Iselin CE, Cm LIS, Webster GD (1998) Pubovaginal sling using<br />

cadaveric allograft fascia for the treatment of intrinsic sphincter deficiency.<br />

J Urol 160:759-762.<br />

57. Iglesia CB, Fenner DE, Brubaker L (1997) The use of mesh in gynecologic<br />

surgery. Intl Urogynecol J 8:105-115.<br />

58. Raz S, Siege1 AL, Short JL, Snyder JA (1989) Vaginal wall sling. J Urol 141:43.<br />

59. Kaplan SA, Santarosa RP, Re AE (1996) Comparison of fascial and vaginal wall<br />

slings in the management of intrinsic sphincteric deficiency. Urology 47:885.<br />

60. Haas CA, Rackley RR, Vasavada SP, Winters JC, Appell RA (1997) Comparison<br />

of in situ vaginal wall sling procedures with and without preservation of the<br />

endopelvic fascia and use of bone anchor suture fixation. J Urol 175(4):1034A.<br />

61. Appell RA, Goldman HB, and Rackley RR (1998) Efficacy and predictors of<br />

success for the in-situ anterior vaginal wall sling with bone anchoring. J Urol<br />

159(5):161A.<br />

62. Sand PK, Bowen LW, Ostergard DR, Nakanishi AM (1988) Hysterectomy and<br />

prior incontinence surgery as risk factors for failed retropubic cystourethropexy.<br />

J Reprod Med 33: 171-174.<br />

63. Gilling PD, Fraundorfer MR, Sealey C, et al. (1994) Laparoscopic extraperitoneal<br />

approaches to female urinary incontinence: the colposuspension and pubovaginal<br />

sling. J Urol 151:334A.<br />

64. Narepalem N, Kieder KJ, Winfeld HN, et al. (1995) Laparoscopic urethral sling<br />

for the treatment of intrinsic urethral weakness (type I11 stress urinary incontinence).<br />

Tech Urol 1(2):102-105.<br />

65. Madjar S, Beyar M, Nativ 0 (1998) Pubic bone anchoring in the treatment of<br />

women with stress urinary incontinence: New applications to an old concept. Intl<br />

Urogynecol J 9:416-418.<br />

66. Zaragoza MR (1996) Expanded indications for the pubovaginal sling: treatment<br />

of type 2 or 3 stress incontinence. J Urol 156: 1620-1622.<br />

67. Chaikin DC, Blavias JG (1998) Weakened cadaveric fascial sling: an unexpected<br />

cause of failure. J Urol 160:2151.<br />

68. Appell RA (1988) Techniques and results in the implantation of the artificial<br />

sphincter in women with type I11 stress urinary incontinence. Neurourol Urodyn<br />

7~613-619.<br />

69. Long RL, Barrett DM (1996) Artificial sphincter: abdominal approach. In: Raz S,<br />

ed. Female Urology, 2nd ed., Philadelphia: WB Saunders Co., 419-427.<br />

70. Wang Y, Hadley HR (1996) Artificial sphincter: transvaginal approach. In: Raz S,<br />

ed. Female Urology, 2nd ed. Philadelphia: WB Saunders Co., 428-434.<br />

7 1. Elliot DS, Barrett DM (1998) The artificial urinary sphincter in the female: Indications<br />

for use, surgical approach, and results. Intl Urology J 9:409-415.


9 Urinary Retention in Women<br />

Roger R. Dmochowski, MD, FACS<br />

CONTENTS<br />

INTRODUCTION<br />

PATHOPHYSIOLOGY<br />

EVALUATION<br />

ETIOLOGY<br />

TREATMENT<br />

REFERENCES<br />

INTRODUCTION<br />

By definition, the woman afflicted with urinary retention is unable<br />

to completely empty her bladder. However, it is difficult to characterize<br />

the condition of urinary retention on the basis of the magnitude of<br />

residual volume alone, and more important to evaluate and render<br />

treatment for the presenting symptomatic complex and any associated<br />

urinary tract pathophysiology. Retention may be symptomatic or<br />

asymptomatic. Symptoms commonly associated with retention may<br />

include pure irritative (urgency, frequency), or obstructive (incomplete,<br />

bladder emptying, hesitancy) components, with or without abdominal or<br />

suprapubic discomfort, and urinary incontinence (either due to overflow<br />

incontinence or detrusor overactivity). Urinary tract sequellae associated<br />

with retention include: alterations in bladder storage characteristics<br />

such as compliance changes, which result in abnormal bladder filling<br />

pressures, detrusor overactivity, abnormalities in bladder sensation, and<br />

effects on renal function owing to abnormal vesical storage pressures<br />

and/or recurrent urinary tract infection.<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfinction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ


186 Drnochowski<br />

Retention in women tends to be an overlooked diagnosis, insidious<br />

in onset, owing to the chronic time frame over which the condition<br />

evolves. In part, diagnostic delay results from the perceived lack of<br />

frequency of this condition in the general population. However, transient<br />

causes of urinary retention also can cause acute symptomatology, such<br />

as inability to void or spontaneous incontinence owing to overflow<br />

of urine.<br />

This chapter will provide an overview of the pathophysiology and<br />

etiologies of incontinence, with an emphasis placed on diagnostic evalu-<br />

ation to identify the aforementioned sequellae of retention and potential<br />

underlying etiologies. Therapeutic approaches pertinent to particular<br />

etiologies will also be discussed.<br />

PATHOPHYSIOLOGY<br />

Normal bladder storage and emptying function depends on integration<br />

between detrusor and sphincter function. During filling, sphincteric<br />

closure provides continence. Bladder filling should occur at low pressures<br />

and these pressures should not substantially change (5 cm of<br />

water) during the storage cycle. No detrusor contractile activity should<br />

occur during this phase. At vesical capacity, detrusor contraction and<br />

sphincteric relaxation occur as essentially simultaneous events. Detrusor<br />

contraction should be of sufficient magnitude and adequately facilitated<br />

to empty the bladder.<br />

This intricate mechanism is dependent on integrity of all the components<br />

of this process. Dysfunction of the urethral sphincter may result<br />

in obstruction, which is commonly associated with neurologic etiologies<br />

of retention. Detrusor <strong>dysfunction</strong> may arise from ultrastructural<br />

changes within the smooth muscle, bladder interstitium, or nerves intrinsic<br />

to bladder function. Although detrusor changes seen with aging<br />

(trebeculation, fibrosis, atrophy or hypertrophy, and decreased nerve<br />

receptor density) have long been recognized (1-3), recent work by<br />

Elbadawi and colleagues has further elucidated the relationship between<br />

ultrastructural changes and concurrent symptoms (4-7).<br />

EVALUATION<br />

As previously noted, knowledge of the time course of retention<br />

or urinary tract obstruction is necessary to determine the level and


Chapter 9 / Female Urinary Retention 187<br />

appropriateness of diagnostic work-up. Although urodynamics are useful<br />

for assessing vesical compliance and presence or absence of detrusor<br />

instability, no urodynamic parameter has been identified as being singularly<br />

diagnostic for obstruction. Pressure/flow nomograms are standardized<br />

for males with outlet obstruction, but do not apply readily to<br />

females owing to variability in voiding dynamics in women. <strong>Voiding</strong><br />

<strong>dysfunction</strong> in women is very poorly reported owing to the fact that<br />

women tend to underreport their voiding symptoms. Additionally, the<br />

diagnosis of obstruction and symptomatic correlation of symptoms have<br />

not been well delineated in patients.<br />

The most appropriate method with which to diagnose obstruction is<br />

to utilize a three-component analysis: history and physical findings<br />

(with cystoscopy considered as a component of physical examination),<br />

symptomatic appraisal (inclusive of subjective and objective methods),<br />

and urodynamic results. Findings consistent with obstruction in any of<br />

the three categories should alert the diagnostician to the possibility of<br />

retention or obstruction in that particular patient.<br />

Thorough historical information provides the basis for accurate diagnosis<br />

and subsequent intervention. In addition to standard complete<br />

review of all prior surgical and medical interventions, some concept<br />

of the time focus of the patient’s symptoms should be delineated.<br />

Chronic onset of symptoms is more likely consistent with a degenerative<br />

process such as pelvic prolapse and/or the possibility of a neurogenic<br />

component. Retention or voiding obstructive symptoms that have an<br />

immediate time correlation to surgical intervention indicate iatrogenic<br />

cause as the main etiology of the obstruction. Retention occurring in<br />

the acute time frame may be indicative more of medication and/or<br />

surgical insult such as hemorrhoidectomy.<br />

Evaluation for retention in the acute time frame can often be delayed<br />

until the presumed inciting etiology has been removed and/or modified.<br />

Initially, during this period, intervention should include clean intermittent<br />

catheterization (CIC). Persistent or continued retention prompts<br />

evaluation. It is best, if possible, to await ambulatory status and return<br />

to normal activities before pursuing neurourologic evaluation. If the<br />

patient is unable to continue intermittent catheterization or perform<br />

intermittent catheterization, chronic catheter drainage does remain an<br />

option.<br />

In addition to history and physical examination, vaginal examination<br />

remains a crucial component of the evaluation of the female with.<br />

chronic obstructive symptomatology. Definitive vaginal examination<br />

should include evaluation of not only the anterior vaginal wall to exclude


188 Dmochowski<br />

cystocele formation, but also examination of the other components of<br />

vaginal vault support including the apeduterus and also the posterior<br />

wall to exclude rectocele.<br />

The resting position of the urethra and bladder neck should be<br />

evaluated in the patient who is status postsuspension, as hyperangulation<br />

of this area may be coexistent with obstruction. The presence of significant<br />

periurethral fibrosis indicates potential iatrogenic cause for obs<br />

tion. Additionally the presence or absence of incontinence at the time<br />

of the examination especially with stress maneuvers should be elucidated.<br />

Also during this time an assessment of the vaginal mucosa,<br />

specifically evaluating for atrophy, should be undertaken.<br />

Focused neurologic examination specifically evaluates the sensory<br />

and motor functions of the sacral dermatomes. This should include not<br />

only assessment of perianal sensation but also the status of voluntary<br />

and resting anal sphincter tone and of the bulbocavernosus reflux. Longtrack<br />

evaluation of the lower extremities includes sensory, fine touch,<br />

pinprick, and reflex testing of the appropriate dermatomes.<br />

The urodynamic evaluation is a crucial component of the evaluation<br />

of the female with chronic voiding <strong>dysfunction</strong> and obstructive symptomatology.<br />

Given the reported lack of correlation between symptomatology,<br />

urodynamic testing does allow some baseline determination<br />

of underlying bladder function prior to intervention for obstructive<br />

symptomatology. However, it should be kept in mind that urodynamic<br />

testing does not provide completely irrefutable data. Specifically, if the<br />

female patient is unable to void in the urodynamic lab, this finding<br />

may not be reflective of true retention. Therefore postvoid residual<br />

volume and simple noninvasive uroflowometry should be used in conjunction<br />

with this analysis.<br />

Noninvasive uroflowometry does provide a simple screen when utilized<br />

with postvoid residual testing. The identification of high postvoid<br />

residual volume associated with a diminished flow may be indicative<br />

of outlet obstruction; however, the possibility of detrusor myogenic<br />

failure must also be considered (Fig. 1).<br />

~ultichannel subtracted pressure flow testing allows a more complete<br />

assessment of bladder contraction when identified. Not only does<br />

multichannel study assess the ability of the bladder to effectively store<br />

urine, it also assesses for bladder filling pressures associated with<br />

compliance characteristics and the presence or absence of detrusor<br />

instability. The assessment of total bladder capacity also is crucial<br />

in evaluating bladder storage capabilities. Electromyography (EMG)<br />

activity at rest may give some indication of possible denervation.


Chapter 9 / Female Urinary Retention 189<br />

Fig. 1. Screening uroflow may indicate possible obstructive phenomena. Curve<br />

at left is from stress-incontinent patient with no clinical signs or symptoms of<br />

obstruction. Curve at right is markedly prolonged and of low amplitude, possibly<br />

indicating obstruction, but also consistent with poor detrusor contractility.<br />

When pressure flow testing is performed, high-pressure, low-flow<br />

evaluation may indicate obstruction; "however, these criteria are not<br />

completely normalized in females at this time. Electromyographic activity<br />

during voiding may be indicative of dyssynergic activity that may<br />

be linked to an underlying neurologic etiology for the obstructive<br />

symptomatology. Evaluation of abdominal pressure during the pressure<br />

flow evaluation is crucial to assess for the presence or absence of<br />

abdominal straining (8). Debate regarding criteria for outlet obstruction<br />

continues on the basis of pressure flow observations. Some investigators<br />

have attempted to validate diagnostic parameters of peak flow less than<br />

15 cc/s, associated with peak voiding pressure of greater than or equal<br />

to 20 cm H20. However, other investigators feel that these criteria are<br />

too arbitrary in the absence of symptomatic assessment and correlation<br />

with parameters obtained from normal controls (9,lO). Massey and<br />

Abrams (l l) utilize four parameters including: significant residual volume,<br />

diminished uroflow rate (less than 20 mL/s), peak detrusor voiding<br />

pressures of greater than 50 cm of water, and high urethral resistance<br />

indices. Other authors have utilized only flow rates and utilized cutoff<br />

values of 15 mL/s with voiding volumes of 200 d.., or more. Bass<br />

used 15 mL/s and 100 ccs of voiding volume (1 l a). Blaivas has utilized<br />

the previously mentioned criteria of peak pressures of 20 cm of water<br />

with peak flows of less than 12 (1 lb).<br />

Authors with extensive experience in this field, however, have noted<br />

that the diagnosis of bladder outlet obstruction in women solely on the<br />

basis of pure urodynamics may be very difficult to define because no<br />

urodynamic parameter has been reported to be predictive of outcome<br />

after urethrolysis in patients felt to be obstructed after incontinence<br />

procedures. Specifically patients with low-pressure voiding may have<br />

outcomes similar to those patients with high-pressure voiding. In


190 Dmochowski<br />

- """"_ 4-<br />

. "-"-."-. ". "."."-I " "-"." t<br />

""-<br />

" ""._""._". """""""."""- "_." """."<br />

I<br />

""-.""_."-.-"""""_"."_"""".""""" +<br />

""""""" i


Chapter 9 / Female Urinary Retention 19 1<br />

Fig. 2. Detrusor curves from three different patients. All were diagnosed as<br />

obstructed on the basis of symptoms, residual urine volume, and clinical examination<br />

findings. (A) Low-pressure voiding in elderly woman with high residual<br />

and periurethral fibrosis from needle suspension. (B) Higher pressure contraction<br />

accompanied by detrusor instability during filling. Irritative symptoms were<br />

marked, in this patient. (C) Pressures approximating 100 cm H20 in woman with<br />

urethral obstruction from sling (complete and acute), note magnitude and quality<br />

of detrusor contraction.<br />

women who void by perineal relaxation, even a small degree of outlet<br />

obstruction may cause significant voiding <strong>dysfunction</strong>. Therefore, stan-<br />

dardized voiding pressures may not apply to this sub-group of women<br />

(12) (Fig. 2).<br />

Simultaneous pressure flow analysis coupled with radiographic (flu-<br />

oroscopic) monitoring provides crucial data regarding simultaneous<br />

structural and functional activity within the lower urinary tract. Video<br />

urodynamics not only allows the assessment of exact point of obstruc-<br />

tion, especially in the patient who is postsurgical, it also allows the<br />

determination of any functional element of obstruction in those patients<br />

who may have learned patterns of voiding <strong>dysfunction</strong>. This study does


192 Dmochowski<br />

provide the gold standard to which all other testing modalities should<br />

be obtained (Fig. 3).<br />

In the absence of videourodynamics, sequential urodynamics supple-<br />

mented by standing voiding cystourethrography (VCUG) will provide<br />

complimentary data. VCUG is performed in lateral, anterior-posterior,


Chapter 9 I Female Urinary Retention 193<br />

Fig. 3. Fluoroscopy and urethral obstruction. (A) Normal, patent, unobstructed<br />

urethra. Urethral lumen has no angulation. (B) Bladder diverticulum in woman<br />

with obstructive symptoms after sling. Volume of diverticulum was approximately<br />

50 cc. (C) Oversuspended bladder neck with no mobility of funneling during<br />

voiding attempt. Patient had significant periurethral fibrosis on pelvic examinati<br />

(D) Prolapse and voiding <strong>dysfunction</strong>. Grade 2 cystocele, in patient with persistent<br />

irritative complaints after suspension. (E) Prolapse and voiding <strong>dysfunction</strong>. Grade<br />

3 cystocele. Patient had residual volumes of 150 cc. (F) Prolapse and voiding<br />

<strong>dysfunction</strong>. Grade 4, central defect. Patient was unable to void except by manual<br />

relocation of vaginal contents. (G) Bladder neck angulation and torquing effect<br />

from suspension suture in close apposition to urethral wall of at bladder level neck.


194 Dmochowski<br />

Fig. 3. Continued<br />

and slightly oblique positions with and without straining and allows<br />

assessment of functional integrity of the pelvic floor in relationship to<br />

urethral and bladder prolapse. VCUG is a very useful test in the patient<br />

who may have occult prolapse not recognized at the time of physi-<br />

cal examination.<br />

This author believes that cystourethroscopy provides a crucial<br />

component of evaluation utilizing a 0 and 30" lens with a nonbeaked<br />

cystoscope. This modality provides for urethral mucosal evaluation<br />

and some determination of any obstructive phenomena caused by<br />

periurethral lesions such as urethral diverticula. Also, hyper-angulation<br />

of the urethra can be identified with this approach. This is especially<br />

important in a post surgical patient who may have had sling<br />

rotation, distortion, or translocation. Additionally, cystourethroscopic<br />

evaluation is crucial to help identify any mucosal lesions within the<br />

urethra or bladder that may provide some component of obstruction<br />

or retention.


Chapter 9 / Fernale Urinary Retention 195<br />

Fig. 3. Continued<br />

In summary, in the woman with obstructive symptoms, subjective<br />

and objective assessment of the patient’s presenting symptoms with<br />

voiding diary, symptom score, and assessment of symptomatic bother<br />

index combined with complex videourodynamic and endoscopic inves-<br />

tigation will provide the most complete estimation of presence and


196 Dmochowski<br />

Fig. 3. Continued<br />

magnitude of definable urinary tract obstruction. Videourodynamics<br />

will localize the site of obstruction. Cystoscopy will identify the point<br />

of obstruction as well as to determine eo-existent lesions.<br />

No one study is felt (by this author) to be conclusive in terms of<br />

diagnostic potential.


Chapter 9 / Female Urinary Retention 197<br />

Fig. 3. Continued<br />

ETIOLOGY<br />

Urinary retention in women can arise from numerous etiologies. As<br />

was previously mentioned, the time course of onset of retention should<br />

be determined (if possible, by symptomatic appraisal). Distinguishing


198 Dmochovvski<br />

recent initiation of symptoms may allow the examiner to determine a<br />

transient etiology such as: endocrine <strong>dysfunction</strong>, urinary tract infection,<br />

immobility, bowel <strong>dysfunction</strong> (including fecal impaction), or psychogenic<br />

cause. Any of these etiologies, alone or in combination, can<br />

produce acute loss of bladder contractile capabilities. Medication dose<br />

adjustments or addition of new medications to a therapeutic regimen<br />

is a particularly important cause of voiding <strong>dysfunction</strong> in patients with<br />

chronic neurologic conditions such as Parkinson’s disease, treatment<br />

of which is dependent on several classes of drugs with significant<br />

anticholinergic side effects (13).<br />

Chronic causes of urinary retention can be broadly classitied as<br />

anatomic/obstructive, neurogenic/functional, and physiologiclage related<br />

(see Table 1) (141.5).<br />

SPECIFIC ETIOLOGIES<br />

Anutomic/Obstructive<br />

GYNECOLOGIC CAUSES OF<br />

BLADDER OUTLET OBSTRUCTION<br />

Gynecologic causes of voiding <strong>dysfunction</strong> arise from angulation or<br />

distortion of the proximal urethral and bladder neck caused by periure-<br />

thral or perivesical gynecologic lesions. The most common gynecologic<br />

cause of outlet obstruction is that resulting from large cystocele forma-<br />

tion. Cystocele formation results in acute kinking or angulation of the<br />

proximal urethra and bladder neck in relationship to the bladder base.<br />

The patient will often complain of positional voiding <strong>dysfunction</strong> (lean-<br />

ing forward, backward, or having to stand to completely evacuate).<br />

She may also report digital manipulation and/or manual reduction of<br />

the cystocele in order to void. These patients may also have co-existing<br />

recurrent urinary-tract infection and stress incontinence.<br />

In Gardy’s study, greater than 40% of women had a significant<br />

postvoid residual, however only 3% developed urinary retention owing<br />

to the large cystocele (54). These authors noted that cystoceles of high<br />

grades (Grades 111-IV) tend to have more significant postvoid residuals<br />

than do smaller grade cystoceles. Ureteral obstruction associated with<br />

prolapse of the anterior vaginal wall may result in oliguria, anuria, and<br />

hydronephrosis.<br />

Large enteroceles and/or rectoceles may also compress the proximal<br />

urethra and bladder neck when dislocated. Constipation and/or other


Chapter 9 / Female Urinary Retention 199<br />

Table l<br />

Etiologies of Chronic Urinary Retention<br />

~nato~ic/Obstructive Neurogenic/~<br />

Gynecologic<br />

Prolapse<br />

Cystocele<br />

VaultlUterine prolapse<br />

Uterine<br />

Myoma<br />

Uterine retroversion<br />

Cervical carcinoma<br />

Endometrial carcinoma<br />

Urethral<br />

Carcinoma<br />

Congenital<br />

Ectopic ureterocele<br />

Female urethral valves<br />

Anterior vaginal wall cysts<br />

Inflammatory<br />

Urethral diverticulum<br />

Intrinsic stricture<br />

S kene’ S duct<br />

Meatal stenosis<br />

Caruncle<br />

Bladder Neck<br />

Primary bladder neck <strong>dysfunction</strong><br />

Iatrogenic<br />

Postsurgical<br />

Incontinence procedures<br />

Urethral instrumentation<br />

Inadequate urethral reconstruction<br />

Upper motor neuron lesion<br />

Pseudodyssynergia<br />

Parkinsons’s disease<br />

Movement disorders<br />

Autonomic neuropathy<br />

True detrusor sphincter<br />

Dyssynergia<br />

Spinal-cord injury<br />

Demyelinating disease<br />

Lower motor neuron lesion<br />

Cauda equina injury<br />

Pelvic nerve injury<br />

Demyelinating disease<br />

Peripheral neuropathy<br />

Tabes dorsalis<br />

Diabetes mellitus<br />

Herpes zoster<br />

Spinal dysraphisrn<br />

Functional<br />

Dysfunctional voiding<br />

Psychogenic retention<br />

Hinman syndrome<br />

Physio~ogic/Age-Related<br />

Detrusor hyperactivity impaired<br />

contractility syndrome<br />

(DHIC)<br />

Decreased tonus<br />

Age-related<br />

Prolonged obstruction<br />

bowel <strong>dysfunction</strong> may exacerbate the voiding <strong>dysfunction</strong>. Impaired<br />

detrusor contractility may further complicate large prolapse (55).<br />

Treatment of vaginal prolapse is predominantly surgical by restoring<br />

normal vaginal access and depth as well as anterior vaginal wall support<br />

(56-63). Temporary therapy may be initiated with pessary support. In<br />

selected patients with severe intercurrent disease, this option may pro-<br />

vide the best chronic management option.


200 Drnochowski<br />

Less common causes of gynecologic obstruction include large uterine<br />

fibroids, especially those located in the anterior uterine segment or<br />

lower uterine segment (64-66). Benign or malignant uterine lesions as<br />

well as ovarian cysts or tumors may also lead to obstruction by compression<br />

or angulation of the proximal urethra and bladder neck (67).<br />

Vaginal, vulvar, and cervical lesions when large rnay also cause<br />

retention owing to impingement on the bladder floor, trigone, or urethra.<br />

Uterine retroversion usually does not result in urinary voiding <strong>dysfunction</strong><br />

unless associated with pregnancy (68,69). During the first trimester<br />

of pregnancy, retention has been reported because the lower uterine<br />

segment distorts, in a fulcrum-type fashion, the proximal urethra and<br />

bladder neck. This usually occurs in the first trimester before 16 wk<br />

(70,71). Therapy usually consists of manipulation of the uterus from<br />

the retroverted position. Usually, with enlargement of the uterus after<br />

the first trimester, retention resolves.<br />

Urethral lesions may also cause obstruction. This is most commonly<br />

seen with urethral diverticula although this author has noted the same<br />

finding with anterior vaginal wall cysts and Skene’s lesions. Additionally<br />

anterior urethral stricture disease, although uncommon in women<br />

may also result in obstruction (Fig. 4). Other rare causes of obstruction<br />

including urethral valves (72,73) and ectopic ureteroceles have been<br />

reported to cause urinary retention, especially in young girls. Urethral<br />

caruncles may also cause obstruction especially when inflamed and<br />

enlarged. In elderly patients, complete urethral prolapse rnay also result<br />

in urinary retention. Urethral carcinoma, although rare in women, can<br />

obstruct the urethral lumen and result in urinary retention also (74).<br />

Primary Bladder Neck Obstruction. This entity initially was<br />

thought to be very common etiology for voiding <strong>dysfunction</strong> in women<br />

and has been termed Marion’ S disease (44). It is defined as incomplete<br />

relaxation of the bladder neck in association with adequate magnitude<br />

and amplitude detrusor contraction. Several large series now have<br />

reported the very rare nature of this disease. Both Farrar and Massey<br />

have shown that in large groups of unscreened patients reporting for<br />

urodynamic evaluation, this is a very unlikely cause of obstructive<br />

symptomatology (45,46). However it can be well-delineated in some<br />

patients and should be excluded with fluoroscopic evaluation. The<br />

underlying etiology may be representative of a learned voiding <strong>dysfunction</strong>,<br />

and/or a true organic disease as some authors have reported<br />

increased fibrosis and collagenous deposition within the bladder neck<br />

in these patients (47). Whether or not these histologic findings are<br />

cause or result of any component of bladder neck rigidity is unknown.


Chapter 9 / Female Urinary Retention 20 1<br />

Fig. 4. Two different cases of intrinsic, mucosal structure disease in women. (A)<br />

Demonstrates a pinpoint mid-urethral orifice in a woman who had undergone<br />

multiple dilatations. In (B) Definitive pelvic radiation had been administered for<br />

uterine malignancy. Both patients had high residuals and predominantly obstructive<br />

complaints.


202 Drnochowski<br />

Some authors have also attributed the abnormality to an increased<br />

number of sympathetic receptors in the bladder neck increasing sensitiv-<br />

ity to normal neurogenic stimulation.<br />

Irritative symptoms (frequency, urgency, and nocturia) often pre-<br />

dominate early, but may subsequently progress to a high postvoid<br />

residual and occasional retention in these patients. Urinary tract infec-<br />

tions may also complicate this disease (48-52).<br />

Blaivas has defined the entity in urodynarnic terms as being associ-<br />

ated with detrusor contraction of greater than 20 cm H20 associated<br />

with urinary flow rates of less than 12 mL/s and with evidence of a<br />

nonrelaxing bladder neck (M). Other authors have identified patients<br />

with voiding pressures of greater than 60 cm of H20 with these same<br />

fluoroscopic components (52). The fluoroscopic evidence of primary<br />

bladder neck obstruction is the lack of funneling or poor opening of<br />

the proximal urethra and bladder neck in association with adequate<br />

amplitude and magnitude detrusor contraction. This results in poor or<br />

absent urinary flow.<br />

Therapy for this disease has rested on surgical ablation of the bladder<br />

neck using transurethral and/or open surgical technique. Transurethral<br />

technique has utilized two incisions vs a single midline incision in<br />

order to prevent sphincteric injury. In the largest series to date, 38<br />

wornen with this disease were described mainly on the basis of symp-<br />

tomatology and cystoscopic evaluation. In this Scandinavian series, a<br />

98% success rate at 1 m0 and a less successful 75% rate at greater<br />

than 4 yr follow-up was reported using transurethral incision (49).<br />

Other authors have described, making a 12-o’clock incision instead<br />

of low lateral incisions at the bladder neck as part of the transurethral<br />

treatment of this disease (53). Open correction of this entity has relied<br />

on WV plasty of the bladder neck.<br />

Other authors have described subjective improvement with alpha-<br />

adrenergic blocking agents with well-reported urodynamic pre and post<br />

operative variables.<br />

DIABETIC CYSTOPATHY<br />

Chronic manifestations of diabetes mellitus (DM) affect the lower<br />

urinary tract in both sexes. Often in diabetics, cystopathy is associated<br />

with demonstrable peripheral neuropathy in the patient who has insulin-<br />

dependent diabetes mellitus (IDDM).


Chapter 9 / Female Urinary Retention 203<br />

Diabetes is associated with segmental neuronal demyelination, axonal<br />

degeneration, and peripheral nervous system degeneration of the<br />

autonomic and somatic nervous systems in the affected population<br />

(16,17). It has been previously shown that reflex-evoked potentials in<br />

patients suffering from diabetes showed diminished conduction velocities<br />

that may be subclinical even in the face of normal neurourologic<br />

examinations. This finding implies a chronic, progressive disease course<br />

with potentially insidious consequences for the patient, even if subtle<br />

physical examination signs are present.<br />

It is most often assumed that the diabetic presents with decreased<br />

bladder sensation resulting in urinary retention owing to chronic bladder<br />

overdistention and resultant poor bladder contractility. Often it is also<br />

presupposed that the patient has increased bladder capacity. Recent<br />

data would suggest that detrusor instability may be the most common<br />

manifestation of diabetic cystopathy and that impaired contractility or<br />

loss of bladder contraction may be less common except as a late<br />

manifestation of the disease in this patient population (18). The scenario<br />

of latent loss of sensation accompanied by increases in bladder capacity<br />

and high postvoid residual volumes however remains the most profound<br />

manifestation of diabetes in the lower urinary tract (19-22).<br />

Therapy in this population depends at what stage the disease is<br />

identified. The goal is to improve bladder emptying and treat underlying<br />

bladder instability if identified on urodynamic evaluation.<br />

Because of the delay in reflex sensation activity (23,24), patients<br />

may do well with a timed voiding schedule with or without Crede or<br />

abdominal straining and/or double and triple voiding. With the presence<br />

of chronic over distention injury, catheter drainage may be necessary,<br />

with clean intermittent catheterization (CIC) providing the best management<br />

for this eventuation. Utilization of parasympathetic agonist agents<br />

has been recommended to facilitate emptying, but most recent data<br />

suggest that these agents are not useful (22,25). Resection or incision<br />

of the bladder neck has been reported to be useful in some patients<br />

but the possibility of causing iatrogenic sphincteric or mucosal injury<br />

and subsequent stricture disease must be considered (22).<br />

PELVIC NERVE INJURY<br />

Extensive pelvic surgery is commonly the cause of injury to sympa-<br />

thetic and parasympathetic ganglia and roots in the true pelvis. In<br />

women, this is most commonly seen in colon surgery (abdomino-<br />

perineal) resection as well as low anterior resection. Also, radical


204 Dmochowski<br />

hysterectomy (Wertheim hysterectomy) is responsible for significant<br />

disruption of pelvic afferent and efferent nervous function because of<br />

extensive lateral pelvic and uterine cervical dissection (26-30). The<br />

incidence of voiding <strong>dysfunction</strong> following these procedures is widely<br />

disparate and has been reported in up to 80% of patients. Injury may<br />

occur to all three branches of the nervous system (parasympathetic,<br />

sympathetic, and somatic-pudendal distributions) and the resultant functional<br />

status may be representative of any combination of complete or<br />

incomplete injury of these three distributions (31). Recurrent pelvic<br />

malignancy may also affect all three branches of the nervous system<br />

within the pelvis.<br />

Often these patients present with an acontractile bladder; however,<br />

partial lesions may result in very poorly compliant and highly unstable<br />

bladders that may have a component of poor emptying as well (32-35).<br />

Bladder contractility may be retained in some of these patients unless<br />

total parasympathic denervation has occurred. Isolated sympathetic<br />

injury may result in sphincteric <strong>dysfunction</strong> and stress incontinence,<br />

which may be severe. Approximately 50% of patients with parasympathetic<br />

nerve denervation also experience sympathetic nerve injury. Radical<br />

dissection associated with the Wertheim procedure occurs distal to<br />

the level of the cervix and accounts for the combined <strong>dysfunction</strong>s<br />

noted in these patients.<br />

Patients often retain a sensation of fullness with poor contractility<br />

and in the early postoperative period may have a component of poor<br />

filling pressures. Urinary retention is usually short-lived, but may be<br />

chronic in up to 30-50% of patients (29,33).<br />

NEUROGENIC DYSSYNERGIA<br />

Interruption of the descending tracks between the pontine mesence-<br />

phalic reticular formation and the sacral cord results in disruption<br />

in communication between the detrusor and pudendal nucleus. This<br />

disruption results in a loss of coordination between bladder and sphinc-<br />

ter. Subsequently, simultaneous bladder contraction and sphincteric<br />

activity occur resulting in high-pressure intravesical storage and empty-<br />

ing (75). Three patterns of dyssynergia have been described, with Type-<br />

I being the most severe because of its presence throughout the voiding<br />

cycle. The most deleterious results of dyssynergia include upper-tract<br />

decompensation and bladder <strong>dysfunction</strong>,<br />

Spinal-cord injury is the most common cause of destrusor sphincter<br />

dyssynergia, but other neurologic diseases such as multiple sclerosis


Chapter 9 I Female Urinary Retention 205<br />

(MS), transverse myelitis associated with acquired immune deficiency<br />

(AIDS), and other demyelinating diseases are associated with this condition.<br />

Depending on the completeness of the neurologic injury, patients<br />

with a specific diagnosis may or may not demonstrate sphincteric<br />

dyssynergia. In a recent study of MS patients, only approximately<br />

1/3 had detrusor sphincter dyssynergia presumably resulting from the<br />

incomplete nature of their disease (76).<br />

LEARNED VOIDING DYSFUNCTION<br />

Non-organic Urinary Retention. Retention resulting from no overt<br />

organic basis is considered psychogenic until proven otherwise. This<br />

finding may be a phenomena that has contributing components both<br />

from central and peripheral nervous system stimulation. Retention has<br />

been reported in patients who are severely traumatized or otherwise<br />

recipients of psychologic stress.<br />

In a series evaluating a group of women presenting with acute<br />

retention, Barrett (36) found that the condition was usually temporary<br />

and responded to noninterventive managements and CIC. Urodynamics<br />

in these patients were normal aside from delayed appreciation of<br />

bladder-filling events. Several of those patients were noted to have<br />

large-capacity bladders. Despite the acute nature of the presentation,<br />

one-fifth of the patients reported by Barrett experienced detrusor myogenic<br />

decompensation and were dependent on chronic catheterization<br />

for bladder drainage.<br />

Dysfunctional <strong>Voiding</strong>. A less overt cause of urinary retention in<br />

women is that associated with <strong>dysfunction</strong>al voiding. The pathophysiologic<br />

event behind <strong>dysfunction</strong>al voiding is loss of coordination or<br />

incomplete coordination between detrusor and sphincteric activity during<br />

voiding in the absence of discrete neurologic lesion. This finding<br />

was first reported in children by Hinrnan (37). This presentation is now<br />

described as the Hinman syndrome and is also described as nonneurogenic<br />

neurogenic voiding <strong>dysfunction</strong> (38-40). The patients initially<br />

reported on by Hinman had dramatic changes in both upper and lower<br />

urinary tracts. This syndrome represents an end-stage variant of the<br />

<strong>dysfunction</strong>al voiding syndrome, which actually manifests in less severe<br />

forms in a significant number of patients in general urologic practice<br />

(41,42). The patients with severe <strong>dysfunction</strong>al voiding habit will have<br />

associated encopresis and bowel <strong>dysfunction</strong> as well as nocturnal voiding<br />

<strong>dysfunction</strong> and enuresis. They will also show dramatic upper-tract


206 Drnochowski<br />

urinary-tract changes such as hydronephrosis and even renal func-<br />

tional compromise.<br />

Urodynamics may reveal incomplete bladder emptying and high-<br />

pressure bladder storage. Fluoroscopic evaluation may show significant<br />

bladder trebeculation with vesicoureteral reflux and high voiding pres-<br />

sures (41). The patient’s voiding pattern will often be staccato or<br />

intermittent and associated with spontaneous sphincteric activity that<br />

occurs during the voiding event, indicating an attempt at volitional<br />

overriding of the voiding reflex arch (40). The electromyographic<br />

(EMG) burst activity results in decrease in urinary flow and marked<br />

increase in voiding pressure. Furthermore, there may be fluoroscopic<br />

evidence of urethral dilitation proximal to the sphincter owing to relative<br />

obstruction caused by pelvic-floor muscular contraction. This will be<br />

associated with interruption in the urinary flow pattern.<br />

McGuire (42) has postulated that the abnormality in these patients<br />

is detrusor overactivity (instability) resulting in volitional sphincteric<br />

activity to overcome sudden unanticipated detrusor activity. This repre-<br />

sents an exaggeration of the normal reflex control (continence reflex)<br />

to prevent incontinence in patients who experience urgency-type symp-<br />

tomatology. However, once this becomes a chronic voiding habit, the<br />

habituation results in chronic voiding <strong>dysfunction</strong>.<br />

In light of this, McGuire has recommended anticholinergic medica-<br />

tion as a primary modality to treat the hyperactive detrusor (42). Other<br />

components of pelvic-floor rehabilitation, including timed voiding and<br />

biofeedback have been reported to be useful in these patients. Some<br />

authors have used smooth and skeletal muscle-relaxing agents such as<br />

Baclofen and Diazepam (43), because these agents tend to be more<br />

successful with sphincteric <strong>dysfunction</strong> resulting from neurologic phe-<br />

nomenon. Strict behavior modification has also been reported to be<br />

useful in many of these patients by altering voiding habits and fluid<br />

intake.<br />

OTHER CAUSES OF SPHINCTER DYSFUNCTION<br />

Other causes of sphincteric <strong>dysfunction</strong> arise from spasticity of the<br />

proximal urethra. Spasticity may result from irritation or urethritis as<br />

well as from periurethral inflammatory phlegmon such as adnexal dis-<br />

ease, vaginal infections, and abscesses of the glandular components in<br />

the periurethral area. Retention may result owing to high urethral closure<br />

pressures (77,78).


Chapter 9 I Female Urinary Retention 207<br />

External sphincter spasticity has also been clinically reported in<br />

patients who are status postanal and -rectal surgery. Pudendal nerve<br />

block has been reported as being salubrious in these patients by decreas-<br />

ing the tone of the surrounding pelvic-floor muscles (78). The etiology<br />

of this spasticity is thought to be owing to local factors such as pain<br />

resulting from infectious, inflammatory, or traumatic causes. Ameliora-<br />

tion of the inflammatory or other insulting lesion with muscle relaxants<br />

and, when appropriate, antibiotics may be beneficial. Alpha-blockade<br />

has also been reported as being beneficial in these patients (77) (Fig. 5).<br />

Physiologic/Age Related<br />

IMPAIRED DETRUSOR CONTRACTILITY<br />

Recently, aging related affects on the detrusor function had been<br />

reported to result in poor detrusor contraction and emptying. In their<br />

seminal work, Resnick and Yalla described detrusor hyperactivity associated<br />

with impaired contractility (DHIC) (4,5). These authors identified<br />

the syndrome in approximately 30% of elderly patients who demonstrated<br />

components of bladder overactivity (instability) associated with<br />

poor bladder emptying. This syndrome was manifested in the absence of<br />

overt outlet obstruction. During urodynamic evaluation, these patients<br />

demonstrated detrusor contractions of adequate magnitude but of poor<br />

duration. The authors noted that approximately one-half of all of those<br />

studied who had detrusor overactivity (instability) had functional<br />

impairment of the detrusor. Many of these patients demonstrated<br />

abdominal augmentation (straining) of voiding. Approximately onethird<br />

of the patients’ bladder volume was voided, the remainder being<br />

retained as residual volume, No patients were in retention as part of<br />

the entry criteria for this study. The authors felt that there could be a<br />

progressive nature to this disease and in fact several of the patients<br />

did progress over the study time frame (79).<br />

The exact etiology for this deterioration in detrusor function may<br />

result from collagenous ingrowth in the detrusor, disrupting the smooth<br />

muscle syncytium. There may also be impairment in neuromuscular<br />

transmission that is related to the myopathic degeneration. These<br />

authors stressed the absolutely crucial nature of urodynamic evaluation<br />

to exclude outlet obstruction as a component of the presenting scenario<br />

(80). Also, it is necessary to exclude components such as fecal <strong>dysfunction</strong>,<br />

constipation, and medication effect as being causative to the<br />

<strong>dysfunction</strong>.


208 Drnochowski<br />

Fig. 5. Fluoroscopy in two patients with sphincteric <strong>dysfunction</strong>. (A) T6 incomplete<br />

paraplegic with high pressure bladder filling and dyssynergia. Note significant<br />

trebeculation. (B) Small-capacity bladder with early trebeculation in woman with<br />

urgency/frequency syndrome and sphincteric overactivity with voiding (Pseudodyssynergia).


Chapter 9 / Female Urinary Retention 209<br />

A<br />

"""""<br />

Fig. 6. Detrusor hyperactivity, impaired contractility ("DHIC). 80-year-old woman<br />

with mixed voiding complaints. Urodynamics demonstrates a poorly facilitated<br />

bladder contraction augmented by significant abdominal straining (arrow). Also<br />

note the low amplitude instability occurring earlier in the filling cycle.<br />

Other patients manifest less overt detrusor hyperactivity and poor<br />

contractility and indeed may present only with impaired contractility<br />

that may lead to chronic detrusor decompensation and retention. Most<br />

common causes for this include medication affect, immobility, and<br />

fecal impaction. The underlying contribution of aging to this entire<br />

syndrome cannot be under estimated (Fig. 6).<br />

TREATMENT<br />

Urinary outflow obstructionhetention after surgical procedures to<br />

correct incontinence can occur due to several technical factors. Suture<br />

tension and placement represent the most important variables. Over-<br />

suspension or hyper-suspension of the proximal urethra and bladder<br />

neck due to suture tension results in these structures being positioned


210 Dmochowski<br />

high in the retropubic space. This high position produces a physiologic<br />

kinking effect at the urethra and bladder neck, effectively compressing<br />

the urethra during bladder contraction. Distal to mid-urethral or juxta-<br />

urethral placement of sutures under tension can also result in inefficient<br />

bladder emptying from angulation of that section of the urethra in<br />

relation to more proximal areas of the urethra and bladder neck. Zim-<br />

mern et al. reported acute angulation of the urethra producing obstruc-<br />

tion after Marshall Marchetti Kranz suspension, incorporating distal<br />

suture placement (81). This outcome has been commonly reported in<br />

the Marshall-Marchetti and some laparoscopic approaches. Because of<br />

this finding, Marshall actually modified his procedure to avoid suture<br />

placement in close proximity to the wall of the urethra. Obstruction<br />

may also result from periurethral fibrosis that may be caused by either<br />

extensive dissection and hemorrhage and/or extravasation of urine at<br />

the time of the procedure (82-88).<br />

Even well-performed procedures can cause increases in urethral<br />

resistance in the chronic time frame after suspension. This is manifested<br />

by increased voiding pressures (89). Patients with postoperative voiding<br />

<strong>dysfunction</strong> are most clearly delineated by the fact that they had normal<br />

emptying prior to the procedure, which postoperatively has been sub-<br />

stantially altered.<br />

<strong>Voiding</strong> by abdominal straining or pelvic-floor relaxation when cou-<br />

pled with an outlet obstructive procedure may result in relative obstruc-<br />

tion producing significant voiding <strong>dysfunction</strong>. Juma reported 20 of 77<br />

patients with a noncontractile bladder preoperatively. Twelve of those<br />

women had significant postvoid residuals at greater than 90 d postopera-<br />

tively. Overall patients with poorly contractile bladders accounted for<br />

approximately 90% of his patients with high postvoid residuals. When<br />

evaluating all patients the actual incidence of obstruction in this patient<br />

group was only 2.5% (90).<br />

Obstruction or retention resulting after sling procedures arises not<br />

only from excessive tension placed on the sling, but also from buckling<br />

or proximal migration of the sling, which produces a narrow bow-<br />

string effect under the bladder neck or urethra (Fig. 7). Additionally,<br />

sling displacement distally or uneven sling tension at the bladder neck<br />

producing a torque or rotational deformity at this location that will also<br />

result in disturbed voiding dynamics. The exact incidence of outlet<br />

obstruction after suspension procedures is probably underreported,<br />

given the variability in diagnostic criteria for obstruction (91-95).<br />

Due to the incomplete nature of complication reporting, the American<br />

Urologic Association’s Female Stress Incontinence Guidelines Panel,


Chapter 9 / Female Urinary Retention 211<br />

Fig. 7. Cystoscopic imaged in two women obstructed after sling. (A) Sling irnpression<br />

on ventrum of urethra at bladder neck. (B) Impression on urethra of sling<br />

that had translocated distally and caused obstruction owing to location at distal<br />

urethra. Cystoscope is just inside meatus. (C) Fluoroscopic image of same patient.<br />

Note urethral narrowing distally, not related to sphincteric activity or to <strong>dysfunction</strong>al<br />

voiding.


212 Dmochowski<br />

Fig. 9.7 Continued<br />

by panel consensus, estimated the risk of obstruction after suspension<br />

procedures to be low. However, obstruction (as defined by need for<br />

catheter drainage greater than 4 wk after surgery) occurring after slingtype<br />

procedures was estimated to occur in 1 out of 20 operations (96).<br />

Explicit review of all collated data, indicated a tendency for slings and<br />

retropubic procedures to have more associated postoperative obstructive<br />

phenomena than transvaginal or anterior repair-type procedures.<br />

Urethrolysis is the fundamental therapy for retention caused by<br />

incontinence procedure. Results with urethrolysis vary; however, most<br />

authors report success rates (relief of obstruction and continence) in<br />

'70-80% of patients (see Table 2). Techniques to accomplish satisfactory<br />

urethrolysis include retropubic or transvaginal approaches. Semilunar,<br />

suprameatal dissection has been described as an isolated procedure, or<br />

has also been combined with more rigorous bladder neck urethrolysis<br />

performed from an introital approach. The goal of all methods is complete<br />

lysis of adhesions and mobility of proximal urethra and bladder<br />

neck. This author performs urethrolysis from below when the inciting<br />

procedure was accomplished by the same route. For patients who have


213<br />

@<br />

ri<br />

P-<br />

5 v;,<br />

n<br />

@<br />

(v?<br />

0<br />

S<br />

GI<br />

P-<br />

W<br />

S<br />

n<br />

Q z<br />

a)<br />

!3<br />

z<br />

Q z


214 Drnochowski<br />

had a laparoscopic suspension or open retropubic, the procedure is<br />

performed through a suprapubic approach. Slings may be incised in<br />

the midline as an initial stage, and this will often decrease resistance<br />

sufficiently so as to allow spontaneous voiding. Other authors use only<br />

one route preferentially. Adjunctive procedures include Martius labial<br />

interposition into the retropubic space to prevent re-fibrosis, and/or<br />

vaginal-wall interposition for the same purpose.<br />

Two current points of contention center on timing of urethrolysis<br />

and the need for re-suspension at the time of procedure. If the woman<br />

is not voiding by 4-6 wk postoperatively, most likely spontaneous<br />

voiding will not resume without intervention. Longer time delays result<br />

in more significant scarring, and add to the difficulty in performing<br />

the procedure. Early release of tension can often be performed by suture<br />

incision alone. Re-suspension has long been advocated as adjunctive<br />

at the time of urethrolysis to prevent recurrent stress incontinence.<br />

Several authors, including Goldman et al. (97) have recommended no<br />

attempt at re-suspension. Results data supports this approach, as the<br />

incidence of recurrent stress incontinence is no greater than after those<br />

procedures during which re-suspension is performed. In general, 70%<br />

of women will void satisfactorily and be continent after urethrolysis<br />

(98-101).<br />

The best solution for postoperative outlet obstruction is satisfactory<br />

intraoperative technique. Care in suture or sling placement, complete<br />

lack of tension (urethra supported in neutral position, i.e., parallel to<br />

floor), and sling fixation in planar configuration will minimize the<br />

possibility of retention. No fool-proof tension application exists, but<br />

laxity of suspending sutures is the most important preventative step<br />

that can be performed (l 02,103).<br />

Catheterization and Other Modzalities<br />

Treatment for these retentive and obstructive conditions hinges on<br />

adequate bladder drainage. The optimal form of bladder drainage at<br />

this time is clean intermittent catheterization (CIC). Chronic catheteriza-<br />

tion is to be decried and avoided if at all possible. CIC was first<br />

recommended by Goodman (I03b); however, Lapides (l 04) popular-<br />

ized the utilization of this therapy in the lower urinary tract. It is a<br />

reasonable therapy for patients of all ages and with all conditions. It<br />

can also be utilized by caregivers and family members for those patients<br />

who lack manual dexterity and/or motivation to perform the procedure.


II Sterility<br />

Chapter 9 / Female Urinary Retention 215<br />

is not crucial; however, adherence to frequency intervals is<br />

necessary to keep bladder volumes under 400-mL thresholds. A high<br />

risk of infection is associated with overdistension and this may be<br />

owing to ischemic injury as indicated by the work of Lapides and<br />

colleagues (I 05-1 08).<br />

Pharmacologic manipulation with parasympathetic agonist agents<br />

has been utilized extensively in the past; however, meta-analysis of<br />

these agents has shown no significant effect on voiding or decrease in<br />

urinary residuals with administration of these medications through any<br />

route. These agents do have some effect on the lower urinary tract,<br />

but there is also simultaneous effect on other organ systems that may<br />

cause significant side effects. No controlled studies substantiate the<br />

utilization of these medications in patients with voiding <strong>dysfunction</strong>.<br />

However, there is some indication that these agents may be useful for<br />

patients with partial lower motor neuron lesions. Oral decomposition<br />

of these drugs indicates that the subcutaneous administration is probably<br />

the most adequate method with which to administer them. Administration<br />

by any route is associated with significant side effects, which tend<br />

to be lessened although not completely ameliorated by the subcutaneous<br />

administration route (1 09-1 15).<br />

A variety of valvular mechanisms are in clinical trails. These devices<br />

attempt to recreate intrinsic urethral function by providing an on demand<br />

control to activate voiding by decreasing internal valvular resistance.<br />

Current mechanisms employ pressure driven and magnetic controlled<br />

avenues. Clinical data is immature at this time.<br />

The newest intervention providing an option in the management<br />

of women with urinary retention is that of sacral neuromodulation.<br />

Previously demonstrated as efficacious in refractory urinary urgency<br />

and frequency syndromes, this modality has recently been reported as<br />

successful in women with urinary retention. Fifty-eight percent of<br />

women dependent on intermittent catheterization were able to void<br />

spontaneously and not require catheter drainage in one Dutch study<br />

(116). The pathophysiologic explanation for this is as yet unresolved;<br />

however, the concept of stochastic resonance has been proposed to<br />

explain the seemingly divergent salutary effect of neuromodulation on<br />

both urgency and frequency symptoms as well as urinary retention.<br />

Stochastic resonance refers to acoustic wave modulation, which occurs<br />

during harmonic activity (R. Schmidt, personal communication). Analogy<br />

between this acoustic activity and neural signal transmission implies<br />

the possibility of nerve transmission manipulation in a similar fashion.<br />

Clinical data to support this thesis is not yet available.


216 Dmochowski<br />

1. Susset JG, Servot-Viguier D, Lamy F,<br />

bladders. Invest Urol 16:204-206.<br />

et al. (1978) Collagen in 155 human<br />

2. Levy BJ, Wight TN (1990) Structural changes in the aging submucosa: new<br />

3.<br />

morphologic criteria for the evaluation of the unstable human bladder. J Urol<br />

144:1044-1055.<br />

Gilpin SA, Gilpin CJ, Dixon JS, et al. (1986) The effect of age on the autonomic<br />

innervation of the urinary bladder. J Urol 150: 168 1.<br />

4. Elbadawi A, Yalla SV, Resnick NM (1993) Structural basis of geriatric voiding<br />

<strong>dysfunction</strong>. I. Methods of a prospective ultrastructural/urodynamic study and an<br />

overview of the findings. J Urol 150:1650-1656.<br />

5. Elbadawi A, Yalla SV, Resnick NM (1993) Structural basis of geriatric voiding<br />

<strong>dysfunction</strong>. 11. Aging detrusor: normal versus impaired<br />

150: 1657-1 667.<br />

contractility. J Urol<br />

6. Elbadawi A, Yalla SV, Resnick NM (1993) Structural basis of geriatric voiding<br />

<strong>dysfunction</strong>. 111. Detrusor overactivity. J Urol 150: 1668-1680.<br />

7. Elbadawi A, Yalla SV, Resnick NM (1993) Structural basis of geriatric voiding<br />

<strong>dysfunction</strong>. IV. Bladder outlet obstruction. J Urol 150: 1681-1695.<br />

8. Barrett DM, Wein AJ (1995) <strong>Voiding</strong> <strong>dysfunction</strong>. Diagnosis, classification and<br />

management. In: Gillenwater JY, Grayhack JT, Howards SS, Duckett JW, eds.,<br />

Adult and Pediatric Urology, 2nd ed., vol. 1, St. Louis: Mosby Year Book, p. 1001.<br />

9. Krane RJ, Siroky<br />

Brown.<br />

MB (1991) Clinical Neuro-Urology, 2nd ed. Boston: Little<br />

10. Webster GI), Kreder KJ (1990). <strong>Voiding</strong> <strong>dysfunction</strong> following cystourethropexy:<br />

its evaluation and management. J Urol 144:670-673.<br />

S 1. Massey J, Abrams PA (1988) Obstructed voiding in the<br />

61:36-39.<br />

female. Br J Urd<br />

S 1A. Bass JS, Leach GE (1991) Bladder outlet obstruction in women. Prob Urd 5:141-<br />

146.<br />

11B. Chancellor MB, Blaivas JG (1992) Diagnostic<br />

ogy 6(4) ~604-62 1.<br />

evaluation. Problems in Urd-<br />

12. Lemack G, Zimern P, Foster B (l 999) Relationship between Urogenital Distress<br />

Inventory-6 questionnaire and urodynamic findings. J Urd 161 (4) 292(A):77.<br />

13. Dmochowski RR (l 999) Female voiding <strong>dysfunction</strong> associated with Parkinson’s<br />

disease. Intl J Urogynecol (In press).<br />

14. Blaivas JG, Labib KB (1 978) Acute urinary retention in women: complete urodynamic<br />

evaluation. Urology 4:383.<br />

15. Wein AJ, Levin RM, Barrett DM (1995) <strong>Voiding</strong> function and <strong>dysfunction</strong>.<br />

<strong>Voiding</strong> function relevant to anatomy, physiology and pharmacology. In: Gillenwater<br />

JY, Grayhack JT, Howards SS, Duckett JW, eds., Adult and Pediatric<br />

Urology, 2nd ed., vol. l., St. Louis: Mosby Year Book, p. 933.<br />

16. Ewing DJ, Clarke BF (1986) Diabetic autonomic neuropathy: present<br />

and future prospects. Diabetes Care 9:648-665.<br />

insights<br />

17. Van Poppel H, Stessens R, Van Darnme B, et al. (1988) Diabetic cystopathy.<br />

Neuropathological examination of urinary bladder biopsies. Eur Urd 15: 128- 13 1.<br />

18. Kaplan SA, Te AE, Blaivas JG (1995) Urodynamic findings in patients with<br />

diabetic cystopathy. J Urol 153:342-344.


Chapter 9 / Female Urinary Retention 217<br />

19. Frimodt-Moller C (1980) Diabetic cystopathy: epidemiology and related disorders.<br />

Ann Intern Med 92:318-321.<br />

20. Ellenberg M (1980) Development of urinary bladder <strong>dysfunction</strong> in<br />

mellitus. Ann Intern Med 92:321-322.<br />

diabetes<br />

21. Bradley (1980) Diagnosis of urinary bladder <strong>dysfunction</strong> in diabetes mellitus.<br />

Ann Intern Med 92:323-326.<br />

22. Frimodt-Moller C (1980) Treatment of diabetic cystopathy. Ann Intern Med<br />

921327-328.<br />

23. Bradley WE, Timm GW, Scott FB (1975) Cystometry V:<br />

6:654.<br />

sensation. Urology<br />

24. Andersen ST, Bradley WE (1976) Early detection of diabetic visceral neuropathy.<br />

An electrophysiologic study of bladder and urethral<br />

25:1100-1105.<br />

innervation. Diabetes<br />

25. Mundy AR, Blaivas JG (1984) Nontraumatic neurological disorders. In: Mundy<br />

AR, Stephenson TP, Wein AJ, eds., Urodynamic Principles, Practices and Application.<br />

Edinburgh: Churchill Livingstone, p. 278.<br />

26. Kirkergaad P, Hjortrup A, Sanders S (198 1) Bladder <strong>dysfunction</strong> after low anterior<br />

resection for mid-rectal cancer. Am J Surg 141:266-268.<br />

27. Forney JP (1980) The effect of radical hysterectomy on bladder function. Am J<br />

Obstet Gynecol 138:374-382.<br />

28. Fowler JW, Brenner DN, Moffatt LEF (1978) The incidence and consequences of<br />

damage to the parasympathetic nerve supply the to bladder after abdominoperineal<br />

resection of the rectum for carcinoma. Br J Urol 50:95-98.<br />

29. Seski JC, Diokno AC (1977) Bladder <strong>dysfunction</strong> after radical abdominal hysterectomy.<br />

Am J Obstet Gynecol 128:643-651.<br />

30. Eickenberg HC, et al. (1976) Urologic complications following abdominoperineal<br />

resection. J Urol 115: 180.<br />

31. Smith PH, Ballantyne B (1968) The neuroanatomical basis for denervation of<br />

the urinary bladder following major pelvic surgery. Br J Surg 55:929-933.<br />

32. Marshall VF, Pollack RS, Miller C (l 946) Observations on urinary <strong>dysfunction</strong><br />

after excision of the rectum. J UroE 55:40.<br />

33. Woodside JR, Crawford ED (1980) Urodynamic features in pelvic plexus injury.<br />

J Urol 124:657-658.<br />

34. Yalla SV, Andriole G (1984) Vesicourethral <strong>dysfunction</strong> following pelvic visceral<br />

ablative surgery. J Urol 132:503-509.<br />

35. Pavlakis AJ (1991) Cauda equina and pelvic plexus injury. In: Krane RJ, Siroky<br />

MB, eds., Clinical Neuro-UroEogy, 2nd ed. Boston: Little, Brown, p. 333.<br />

36. Barrett DM (l 978) Evaluation of psychogenic urinary retention. J Urd 120: 19 1-<br />

192.<br />

37. Hinman F, Raumann FW (1973) Vesical and ureteral damage from voiding<br />

38.<br />

<strong>dysfunction</strong> in boys without neurologic or obstructive disease. J Urol 109:727-<br />

732.<br />

Allen TD (1977) The non-neurogenic neurogenic bladder. J Urol 117:232-238.<br />

39. Allen TD, Bright I11 (1978) Urodynamic patterns in children with <strong>dysfunction</strong>al<br />

voiding problems. J Urol 119:247-249.<br />

40. Hinman F Jr (1986) Non-neurogenic neurogenic bladder (the Hinrnan syndrorne)-15<br />

years later. J Urd 136:769-777.<br />

41. Jorgensen TM, Djurhuus JC, Schroder HD (1982) Idiopathic detrusor sphincter<br />

dyssynergia in neurourologically normal patients with voiding abnormalities. Eur<br />

Urol 8: 107-1 10.


218 Drnochowski<br />

42.<br />

43 I)<br />

44.<br />

45.<br />

46.<br />

47.<br />

48.<br />

49.<br />

McGuire EJ, Savastano JA (1984) Urodynarnic studies in enuresis and the nonneurogenic<br />

neurogenic bladder. J Urol 132:299-302.<br />

Kaplan WE, Firlit CR, Schoenberg HW (1980) The female urethral syndrome.<br />

External sphincter spasticity as an etiology. J Urol 124:48-49.<br />

Marion G (1933) Surgery of the neck of the bladder. Br J Urd 5:351.<br />

Massey JA, Abrams PA (1988) Obstructed voiding in the female. Br J Urol<br />

61~36-39.<br />

Farrar DJ, Osborne JL, Stephenson TP, et al. (1976) A urodynamic view of<br />

bladder outflow obstruction in the female: Factors influencing the results of<br />

treatment. Br J Urol 47:815.<br />

Mayo ME (1982) Primary bladder neck obstruction. Surg Rounds 5:66.<br />

Gronbaek K, Struckmann JR, Frimodt-Moiler C (1992) The treatment of female<br />

bladder neck <strong>dysfunction</strong>. Scand J Urol Nephrol 26: 113-1 18.<br />

Norgaard JP, Swartz-Sorenson S, Djurhuus JC (1984) Functional bladder neck<br />

obstruction in women. Urology 39:207.<br />

50. Diokno AC (1986) Bladder neck obstruction in women. Neurourol Urodyn 5:325.<br />

51. Diokno AC, Hollander JB, Bennett CJ (1984) Bladder neck obstruction in wome<br />

a real entity. J Urol 132:294-298.<br />

52. Axelrod SL, Blaivas JG (1987) Bladder neck<br />

137~497-499.<br />

obstruction in women. J Urol<br />

53. Delaere KPJ, Debruyne FMJ, Moonen WA (1983) Bladder neck incision in the<br />

female: a hazardous procedure? Br J Urol 55:283-286.<br />

54. Gardy M, Kozminski M, DeLancy J, et al. (1991) Stress incontinence and cystoceles.<br />

J Urol 145:1211-1216.<br />

55. Nichols DH (1985) Vaginal prolapse<br />

Am 12:325.<br />

affecting bladder function. Urol Clin N<br />

56. Nichols DH, Milley PS, Randall CL (1970) Significance of restoration of normal<br />

vaginal depth and axis. Obstet Gynecol 36:251-256.<br />

57 * Raz S, Nitti VW, Bregg KJ (1993) Transvaginal repair of enterocele. J Urd<br />

149~724-730.<br />

58. Raz S, Klutke CG, Golub J (1989) Four-corner bladder and urethral suspension<br />

for moderate cystocele. J Urol 142:712-715.<br />

59. Raz S, Little NA, Juma S, et al. (1991) Repair of severe anterior vaginal wall<br />

prolapse (grade IV cystourethrocele). J Urol 146:988-992.<br />

60. Nitti VW (1994) Transvaginal repair of enterocele, with variations. Conternp<br />

Urol 6(4):50.<br />

61. Nichols DH (1982) Sacrospinous fixation for massive eversion of the vagina. Am<br />

J Obstet Gynecol 142:901-904.<br />

62. Moschcowitz AV (1912) The pathogenesis, anatomy and cure of prolapse of the<br />

rectum. Surg Gynecol Obstet 14:7.<br />

63. Raz S (1992) Vaginal prolapse. In: Raz S, ed., Atlas of Transvaginal Surgery.<br />

Philadelphia: WB Saunders, p. 103.<br />

64. Ward JN, Lavengood RW Jr, Draper JW (1968) Pseudo Bladder neck syndrome<br />

in women. J UroE 99:64.<br />

65. Carve11 JF, Stubbs SM (1977) Acute retention of urine in a woman due to an<br />

exoenteric leiomyoma. Br J Urol4950.<br />

66. Kingsworth AN (1984) Urinary retention due to ovarian cyst. Br J Urol56:439-<br />

440.<br />

67. Doran J, Roberts M (1978) Acute urinary retention in the female. Br J Urol47:793.


Chapter 9 / Female Urinary Retention 219<br />

68. Silva PD, Rerberich W (1986) Retroverted impacted gravid uterus with acute<br />

urinary retention: report on two cases and review of the literature. Obstet Gynecol<br />

68:121-123.<br />

69. Ilansen JH, Asmussen M (1985) Acute urinary retention in the first trimester of<br />

pregnancy. Acta Obstet Gynecol Scand 64:279-280.<br />

70. Burton S, Maurer JE, Lich R (1951) Acute retention in pregnancy. J Urol64:578.<br />

71. Goldbert KA, Kwart ANI (1981) Intermittent urinary retention in first trimester<br />

of pregnancy. Urology 17:270-27 1.<br />

72. Cherrie RJ, Leach GE, Raz S (1983) Obstructing urethral valve in a woman: a<br />

case report. J Urd 129:1051-1052.<br />

73. Nesbit RM, McDonald HP Jr, Busby S (1964) Obstructing valves in the female<br />

urethra. J Urol 91 :79.<br />

74. Sarosdy MF (1987) Urethral carcinoma. AUA Update<br />

Houston: American Urological Association.<br />

Series, vol. 6, no. 13,<br />

75. McGuire E (1979) Electromyographic evaluation<br />

<strong>dysfunction</strong>. Urol Clin N Am 6:21.<br />

of sphincter function and<br />

76. Sirls LT, Zimmern PE, Leach GE (1994) Role of limited ovulation and aggressive<br />

medical management in multiple sclerosis.<br />

151:946-950.<br />

A review of 113 patients. J Urol<br />

77. Raz S, Smith RB (1976) External sphincter spasticity syndrome in female patients.<br />

J Urol 1 15 :443-446.<br />

78. Webster GR (1975) Combined video/pressure/flow cystourethrography in female<br />

patients with voiding disturbances. Urology 5:209-215.<br />

79. Resnick NM, Yalla SV (198’7) Detrusor hyperactivity with impaired contractile<br />

function-an unrecognized but common cause of incontinence in elderly patients.<br />

JAMA 257:3076-3081.<br />

80. Resnick NM, Yalla SV, Laurino E (1989) The pathophysiology of urinary incontinence<br />

among institutionalized elderly persons. N Engl J Med 320: 1-7.<br />

81. Zimern PE, Hadley HR, Leach GE, Raz S (1987) Female urethral obstruction<br />

after Marshall Marchetti Kranz operation. J Urol 138:517-520.<br />

82. Nitti VW, Raz S (1994) Obstruction following anti-incontinence procedures:<br />

diagnosis and treatment with transvaginal urethrolysis. J Urol 152:93-98.<br />

83. Webster GD, Kreder KJ (1990) <strong>Voiding</strong> <strong>dysfunction</strong> following cystourethropexy:<br />

its evaluation and management. J Urol 144:670-673.<br />

84. Rost A, Fiedler U, Fester C (1979) Comparative analysis of the results of<br />

suspension-urethroplasty according to Marshall-Marchetti-~antz and of ure-<br />

85.<br />

throvesicopexy with adhesive. Urol Intl 34: 167.<br />

Mundy AR (1983) A trial comparing the Stamey bladder neck suspension with<br />

colposuspension for the treatment of stress incontinence. Br J Urol 55:687-690.<br />

86. Cardoza LD, Stanton SL, Williams JE (1979) Detrusor instability following<br />

surgery for genuine stress incontinence. Br J UroE 51:204-207.<br />

87. Marchetti AA, Marshall VM, Shultis LD (1957) Simple vesicourethral suspension.<br />

A survey. Am J Obstet Gynecol 7457.<br />

88. Leach GE, Raz S (1984) Modified Pereya bladder neck suspension after previously<br />

failed anti-incontinence surgery. Urology 23:359-362.<br />

89. Pope AJ, Shaw PJR, Coptcoat et MJ, al. (1990) Changes in bladder function following<br />

a surgical alteration in outflow resistance. Neurourol Urodyn 9:503-508.<br />

90. Juma S, Sdrales L (1993) Etiology of urinary retention after bladder neck suspension<br />

(abstract) J Urol 149:400(A).


220 Dmochowski<br />

91. Bwch JC (1961) Urethrovaginal fixation to Cooper’s ligament for correction of<br />

stress incontinence, cystocele and prolapse. Am J Obstet Gynecoj 81 281.<br />

92. Shull BL, Baden WF (1989) A six-year experience with paravaginal defect<br />

93.<br />

repair for stress urinary incontinence. Am J Obstet Gynecol 160: 1432-1439.<br />

McGuire EJ, Letson W, Wang S (1989) Transvaginal urethrolysis after obstructive<br />

urethral suspension procedures. J Urd 142:1037.<br />

94. Stamey TA (1 980) Endoscopic suspension of the vesical neck for urinary incontinence<br />

in females: Report on 203 consecutive patients. Ann Surg 192:465-471.<br />

95. Spencer JR, O’Conor VJ Jr., Schaeffer AJ (1987) Comparison of endoscopic<br />

suspension of the vesical neck with suprapubic vesicourethropexy for treatment<br />

of stress urinary incontinence. J Urd 137:411-415.<br />

96. Leach GE, Dmochowski RR, Appell RA, Blaivas JG, Hadley HR, Luber KM, et<br />

al. (1997) Female stress Urinary Incontinence Clinical Guidelines Panel summary<br />

report on surgical management of female stress urinary incontinence. The American<br />

Urological Association. J Urol 158:875--880.<br />

97. Goldman HB, Rackley RR, Appell RA (1999) The efficacy of urethrolysis<br />

98.<br />

without re-suspension for iatrogenic urethral obstruction. J UroE 161: 196-199.<br />

Can LK, Webster GD (199’7) <strong>Voiding</strong> <strong>dysfunction</strong> following incontinence surgery:<br />

diagnosis and treatment with retropubic or transvaginal urethrolysis. J<br />

Urol 157:821--823.<br />

99. Foster HE, McGuire EJ (1993) Management of urethral obstruction with transvaginal<br />

urethrolysis. J Urol 150: 1448-145 l.<br />

100. Dmochowski RR, Leach GE, Zimmern PE, Roskamp DA, Ganabathi K (1994)<br />

Urethrolysis to relieve outlet obstruction after prior incontinence surgery. J Urol<br />

(part 2) 15 l :420A.<br />

101. Wang AC (1996) Burch colposuspension vs Stamey bladder redo suspension.<br />

A comparison of complications with special emphasis on detrusor instability<br />

and voiding <strong>dysfunction</strong>. J Reprod Med 41529-533.<br />

102. Rovner ES, Ginsberg DA, Raz S (199’7) A method for intraoperative adjustment<br />

of sling tension: prevention of outlet obstruction during vaginal wall<br />

Urology 50:273-276.<br />

sling.<br />

103. McLennon Mt, Bent AE (199’7) Sling incision with associated vaginal wall<br />

interposition for obstructed voiding secondary to suburethral sling procedure.<br />

Zntl Urogyn J 8:168.<br />

103B. Guttman L (1954) Initial treatment of traumatic paraplegia. Proc R Soc Med<br />

47: 1103-1 105.<br />

104. Lapides J, Costello RT, Zierdl DK, et al. (1968) Primary causes and treatment of<br />

recurrent urinary tractinfectioninwornen: preliminaryreport. J Urol100:552-555.<br />

105. Lapides J, Diokno AC, Silber SJ, et al. (l 972) Clean, intermittent self-catheterization<br />

in the treatment of urinary tract disease. J Urol 107:458-461.<br />

106. Lapides J, Diokno AC, Could FR, et al. (1976) Further observation on selfcatheterization.<br />

J Urol 1 16: 169-1 71.<br />

107. Diokno AC, Sands LP, Hollander JB, et al. (I 983) Fate of patients started on clean<br />

intermittent self-catheterization therapy l0 years ago. J Urd 129: l 120-1 122.<br />

108. Bu&e A, Vollset SE (1993) Risk factors for bacteremia and clinical urinary<br />

tract infection in patients treated with clean intermittent catheterization.<br />

J Urol 14952’143 l.<br />

109. Barrett DM (1981) The effect of oral bethanechol chloride on voiding in female<br />

patients with excessive residual urine. A randomized double-blind study.<br />

J Urol 126:640-642.


Chapter 9 / Female Urinary Retention 221<br />

110. Wein AJ, Malloy TR, Shofer F, Raezer DM (1 980) The effects of bethanechol<br />

chloride on urodynamic parameters in normal women and in women with significant<br />

residual urine volumes. J Urol 124:397-399.<br />

111. Wein AJ, Raezer DM, Malloy TR (1980) Failure of the bethanechol supersensitivity<br />

test to predict improved voiding after a subcutaneous bethanechol administration.<br />

J Urol 123:202-203.<br />

112. Wein AJ, Hanno PM, Dixon DO, et al. (1978) The effect of oral bethanechol<br />

chloride on the cystometrogram of the normal male adult. J Urol 120:330-331.<br />

113. Blaivas JG (1984) If you currently prescribe bethanechol chloride for urinary<br />

retention, please raise your hand. Neurourol Urodyn 3:209.<br />

114. Awad SA, McGinnis RHJ, Downie JW (1984) The effectiveness of bethanechol<br />

chloride in lower motor neuron lesions: The importance of mode of administration.<br />

Neurourol Urodyn 3:173.<br />

115. Downie JW (1984) Bethanechol chloride in urology-a discussion of issues.<br />

Neurourol Urodyn 321.<br />

116. Koldewijn E, Meulman E, Bermelmans B, van Kerrebroek, Debruyne F (1999)<br />

Neuromodulation effective in voiding <strong>dysfunction</strong> despite high reoperation rate.<br />

J Urol 161(4), 984(A):255.


This Page Intentionally Left Blank


N ~ LVOIDING E DYSFUNCTION


This Page Intentionally Left Blank


es<br />

Edward F, Ikeguchi, ALexis F. Te,<br />

James Cboi, and Steven A. KapLan<br />

INTRODUCTION<br />

PRESENTATION<br />

EVALUATION<br />

ETIOLOGIES AND TREATMENT<br />

REFERENCES<br />

INTRODUCTION<br />

Bladder outlet obstruction in males may be a complex syndrome of<br />

either dynamic functional or fixed anatomic forces resulting in resis-<br />

tance to the flow of urine. Particularly in males, bladder outlet obstruc-<br />

tion is a frequently encountered finding. From a diagnostic standpoint,<br />

the evaluation of bladder outlet obstruction relies heavily upon tradi-<br />

tional methods. While the most common etiology of bladder outlet<br />

obstruction in males relates to prostatic obstruction, an accurate history<br />

and physical exam may alert the clinician that other causes may be at<br />

hand. Newer technologies have, however, proven to be particularly<br />

helpful to the clinician in ascertaining the correct diagnosis and etiology<br />

of bladder outlet obstruction prior to the institution of therapy. Urolo-<br />

gists have also seen the treatment of bladder outlet obstruction change<br />

dramatically over the past decade. While transurethral prostate resection<br />

was the most commonly performed procedure by urologists of yester-<br />

year, it now competes with multiple alternative surgical and medical<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfinction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

225


226 Ikeguchi et al.<br />

therapies. Finally, in addressing bladder outlet obstruction in males,<br />

more attention has been given to functional voiding disorders commonly<br />

seen in younger male populations. Herein, an overview of the evaluation<br />

and treatment of bladder outlet obstruction in males will be presented,<br />

addressing the most commonly utilized therapies for the most commonly<br />

seen etiologies.<br />

PRESENTATION<br />

The way a patient with bladder outlet obstruction presents depends<br />

upon the etiology and the chronicity of the problem. Unless symptoms<br />

are acute, few patients experience pain. Thus, what prompts a patient<br />

to seek medical attention is dependent on subjective symptoms, which<br />

are tolerated by individuals to varying degrees. While the majority of<br />

patients will admit to obstructive symptoms for some time prior to<br />

presentation, on occasion a patient will develop urinary retention, uri-<br />

nary sepsis, or incontinence as their first sign of bladder outlet<br />

obstruction.<br />

EVALUATION<br />

Histo y<br />

Lower urinary tract symptoms (LUTS) may be divided into either<br />

obstructive or irritative symptoms. Patients with bladder outlet obstruc-<br />

tion generally complain of obstructive symptoms. However, as the<br />

bladder reacts to the presence of outlet obstruction, the response may<br />

result in the onset of detrusor instability. Thus irritative voiding symp-<br />

toms may also be present. The obstructive vs irritative voiding symp-<br />

toms are listed below:<br />

Obstructive voiding symptoms:<br />

1. Urinary hesitancy.<br />

2. Dribbling/weak urinary stream.<br />

3. Sensation of incomplete bladder emptying.<br />

4. Prolonged urination.<br />

Irritative voiding symptoms:<br />

1. Urinary frequency.<br />

2. Nocturia.


Chapter 10 / Male Bladder Outlet Obstruction 227<br />

3. Urinary urgency.<br />

4. Urge incontinence.<br />

In addition, the clinician should elicit a medical history pertinent to<br />

the various etiologies for bladder outlet obstruction. These might<br />

include the presence of a split or narrow stream, the presence of concomitant<br />

medical/neurologic history, medication lists, and so forth.<br />

The urologist should make an assessment of the patient’s subjective<br />

complaints. This is best accomplished with the use of the American<br />

Urological Association (AUA) Symptom Index. The AUA index is<br />

based on a score ranging from 0 to 35, and classifies the patient’s<br />

symptoms as mild (0 to 7), moderate (8 to 19), or severe (20 to 35).<br />

Not only is the AUA symptom index helpful in determining the severity<br />

of symptoms, it may be used to monitor the progression of the disease.<br />

In addition, the AUA symptom index may be the single most important<br />

parameter followed to determine the efficacy of treatment, since the<br />

patient’ S motivation in seeking medical attention usually relates to<br />

subjective complaints and quality of life.<br />

The physical examination should be complete, though a focused<br />

urinary tract exam is the most pertinent in diagnosing bladder outlet<br />

obstruction. The abdominal examination should take note of suprapubic<br />

fullness or pain. The penis and scrotum should be inspected, as should<br />

the prostate via digital rectal examination to ascertain size, fluctuance<br />

or bogginess, and the presence of prostate cancer signs.<br />

URINE<br />

Lab Tests<br />

Routine urine analysis and culture should be sent in order to rule<br />

out the presence of any infectious process. Frequently, in the face of<br />

true bladder outlet obstruction, urinary infection may be present on the<br />

basis of urinary stasis.<br />

BLOOD<br />

Although there are no absolute guidelines for blood work in the case<br />

of bladder outlet obstruction, the clinician should bear several points<br />

in mind. Patients with bladder outlet obstruction may be subject to


228 Ikeguchi et al.<br />

significant renal injury reflected in an elevated blood urea nitrogen and<br />

serum creatinine. Although we currently do not obtain these tests in<br />

all patients with LUTS, it should be considered, particularly for patients<br />

presenting with large postvoid residual volumes or patients presenting<br />

to an emergency department acutely with retention.<br />

In terns of analysis of serum prostate specific antigen (PSA), there<br />

currently is no established guideline for its testing. Although patients<br />

with bladder outlet obstruction are frequently in the age range for<br />

prostate screening, routine PSA screening is still a point of controversy.<br />

Furthermore, bladder outlet obstruction and urinary retention may result<br />

in falsely elevated PSA levels, as can the diagnosis of benign prostatic<br />

hyperplasia (BPH).<br />

URINARY FLOW RATE<br />

Diagnostic Procedures<br />

In general, urinary flow rates correlate poorly with patient symptoms.<br />

Nonetheless, urinary flow assessment is frequently used as an objective<br />

parameter to follow the progress of treatment. The measurement of<br />

urinary flow may also be used in conjunction with postvoid residual<br />

assessment in identifying patients with attenuated detrusor function.<br />

Patients with moderate to severe symptoms but without diminished<br />

peak urinary flow rates (>l5 mL/s) may benefit from more in-depth<br />

urodynamic testing, i.e., pressure/flow studies.<br />

Another aspect of urinary flow rate assessment is the interpretation<br />

of flow patterns. In general, a normal flow pattern should rise to a peak<br />

rate and return to baseline in the form of a sine-wave curve. Certain<br />

patterns may be indicative of abdominal straining. Although this is<br />

not pathognomonic for either bladder outlet obstruction or impaired<br />

detrusor contractility, it may be another finding that prompts more<br />

sophisticated urodynamic evaluation.<br />

Nonetheless, from a practical standpoint it is important to remember<br />

that urinary flow assessment is neither sensitive nor specific for bladder<br />

outlet obstruction. Significant bladder outlet obstruction may be present<br />

in the face of normal flow rates, and a poor urinary flow rate does not<br />

distinguish between bladder outlet obstruction and an impaired detrusor.<br />

OSTVOID RESIDUAL<br />

As in the case of urinary flow rate, the postvoid residual has poor<br />

correlation with the degree of subjective complaints. The postvoid


Chapter 10 / Male Bladder Outlet Obstruction 229<br />

residual may, however, be helpful as a comparative parameter when<br />

assessments can be made before and after treatment. It may also serve<br />

as a helpful monitoring tool for patients with the diagnosis of bladder<br />

outlet obstruction being treated with medications.<br />

Nonetheless, as a single modality of assessment for the bladder outlet<br />

obstruction, postvoid residual measurement lacks both sensitivity and<br />

specificity. Patents with larger postvoid residual determinations should<br />

be considered for urodynamic evaluation. Recent data suggests that<br />

men with elevated residual urine at baseline may be at greater risk for<br />

urinary retention. We routinely obtain postvoid residual measurements<br />

as part of the initial evaluation of LUTS.<br />

CYSTOSCOPY<br />

Cystoscopy in the male patient with symptoms of bladder outlet<br />

obstruction is an important part of its diagnosis. When the diagnosis<br />

of bladder outlet obstruction is clear, though the exact etiology is not<br />

absolutely certain (i.e., prostatic obstruction vs urethral stricture), the<br />

performance of a flexible cystoscopic examination under local anesthesia<br />

is an invaluable part of the work-up. Not only is there minimal<br />

morbidity to the patient, it may help further elucidate the patient’s<br />

underlying disease. Particularly in the case of prostatic obstruction, the<br />

clinician can determine certain variables such as the presence or absence<br />

of obstruction at the level of the bladder neck, median lobe, lateral lobes,<br />

and measure prostatic urethral length. Ultimately, this information may<br />

be used by the urologist to cater therapy to each individual.<br />

RODYNAMICS<br />

When the exact diagnosis or etiology of bladder outlet obstruction<br />

is not entirely clear, urodynamic evaluation coupled with cystoscopy<br />

may be an invaluable tool. In first making the diagnosis of bladder<br />

outlet obstruction, the patient requires a fully functional detrusor from<br />

the standpoint that a detrusor contraction should be able to be generated.<br />

If the patient is unable to generate a detrusor contraction greater than<br />

15 cm H20 during cystometry, the patient should be treated for impaired<br />

detrusor contractility or acontractile detrusor. In such situations, the<br />

concomitant presence of bladder outlet obstruction is unable to be ascer-<br />

tained.<br />

Provided the patient can generate an adequate detrusor contraction,<br />

the presence of bladder outlet obstruction is based on the combination


230 Ikeguchi et al.<br />

of low urinary flow in the face of elevated detrusor pressure. Often<br />

this can be represented graphically with the use of the Abrams-Griffith<br />

nomogram. Furthermore, provided the patient is able to initiate a urinary<br />

stream during the study, the exact location of the obstruction may be<br />

determined with the simultaneous use of fluoroscopy.<br />

In general, conventional urodynamic parameters have not been found<br />

to correlate well with the AUA symptom score. At our institution, we<br />

have found that the two urodynamic parameters that most accurately<br />

correlate with worsening LUTS in males is the presence of involuntary<br />

detrusor contractions and the presence of prolonged detrusor contraction<br />

duration (I).<br />

The degree to which urodynamic evaluation is utilized varies widely<br />

according to the medical center and the treating urologists’ practices.<br />

Nonetheless, urodynamic testing may in some cases be the only sure<br />

way to make an accurate diagnosis, short of the empiric institution<br />

of therapy.<br />

ETIOLOGIES AND TREATMENT<br />

Prostatic Obstruction-Benign Prostatic Hyperplmia<br />

Although many different prostatic conditions may result in symptoms<br />

of bladder outlet obstruction, the most common cause is benign prostatic<br />

hyperplasia (BPH). Although BPH will be the focus of this section,<br />

the reader should bear in mind that other common conditions of the<br />

prostate resulting in bladder outlet obstruction are prostatic carcinoma<br />

and acute prostatitis. In most cases, the patient’s history and digital<br />

rectal examination will be suggestive of these diagnoses.<br />

BPH is a process by which there is an increased number of cells in<br />

both the epithelial and stromal compartments of the gland. BPH is<br />

highly prevalent in the aging population and has been reported in 90%<br />

of glands of men aged 85 in autopsy series. In the United States,<br />

approximately 250,000 transurethral prostate resections were performed<br />

in 1993 with a national expenditure in the billions of dollars. Even<br />

with the advent of better medical therapies, in 1996 approximately<br />

l 16,000 transurethral procedures were performed, comprising 94% of<br />

all prostate procedures performed for BPH (2).<br />

BPH also takes its toll upon the patient’s quality of life and sense<br />

of well being. It has been shown that over half of men between the


Chapter 10 / Male Bladder Outlet Obstruction 23 1<br />

ages of 40 and 79 without a prior diagnosis of BPH will curtail at least<br />

one activity of daily life owing to their urinary difficulties (3).<br />

The natural history of benign prostatic hyperplasia has been<br />

addressed in several studies. Most notably, in 1997, Jacobsen et al.<br />

reported their findings on the natural history of BPH in a cohort of<br />

over 2,000 randomly selected men aged 40-79 in Olmstead County,<br />

Minnesota. They found that in men with urinary symptoms and physical<br />

findings consistent with BPH, the incidence of urinary retention was<br />

dramatically increased with age and symptoms. The incidence of acute<br />

urinary retention among men in the general population in the studied<br />

age range was 7/1,000 person-years of follow-up. This was in contrast<br />

to men in their 70s that had a 1 in 10 chance of acute urinary retention<br />

within 5 years follow-up and even greater risk if they had urinary<br />

symptoms. Jacobsen al. et identified several risk factors for acute urinary<br />

retention as being age greater than 70 yr, AUA symptom index of 8<br />

or greater, and a prostatic size determined by digital rectal examination<br />

of greater than 30 g in size (4).<br />

MEDICAL THERAPY<br />

Medical therapy has revolutionized the way in which urologists<br />

approach the patient with bladder outlet obstruction owing to BPH.<br />

Once the diagnosis of BPH is made in a patient with LIJTS, they should<br />

be offered a trial of medical therapy. Only under certain circumstances<br />

should a patient be considered better served with first-line surgical<br />

therapy. These situations would include the presence of bladder calculi,<br />

upper tract deterioration diagnosed by ultrasound, or acute urinary<br />

retention.<br />

The two main theories behind the pathophysiology prostatic obstruction<br />

are based on the either the concept of a fixed obstruction caused<br />

by prostatic enlargement or the contraction of the prostatic urethra<br />

owing to a degree of sympathetic tone stimulating the alpha,-adrenoceptors<br />

of the prostatic smooth muscle (5). These theories may be correlated<br />

with the currently available classes of medications available for the<br />

treatment of BPH. While the 5-alpha reductase inhibitors act upon<br />

the prostate by minimizing the hormonal stimulus to prostatic growth<br />

resulting in shrinkage, the alpha-blockers cause smooth muscle relaxation<br />

to ease the proposed tension around the urethra resulting in<br />

obstruction. Both forms of medical therapy will be discussed below.<br />

Alpha-blockade. Alpha-blockers represent the mainstay of medical<br />

therapy currently used for the treatment of bladder outlet obstruction


232 Ikeguchi et al.<br />

owing to BPH. The rationale behind the mechanism of alpha-blocker<br />

therapy is the interaction of the smooth muscle (stromal) and glandular<br />

(epithelial) components of the prostate, resulting in a situation of<br />

dynamic obstruction from a sympathetic stimulation of smooth muscle<br />

receptors in the prostate. Alpha-blockers act on the alpha,-adrenoceptors<br />

that are found mainly in the prostatic stroma and capsule and the bladder<br />

neck (6). Given the mechanism of its action, alpha-blocker therapy<br />

manifests the expected adverse reactions on an individual basis. The<br />

most common adverse effects include weakness, asthenia, hypotension<br />

(particularly postural) and dizziness, and retrograde ejaculation. These<br />

complications are reversible with the discontinuation of therapy. In<br />

general, alpha-blocker therapy can be expected to result in an increase<br />

in peak urinary flow rate approximating 3.8 mL/s and 74% likelihood<br />

of improved subjective urinary symptoms.<br />

Many varieties of alpha-blockers have been shown to be both safe and<br />

effective. Elhilali et al. reported on this profile as seen with terazosin, as<br />

studied in 224 patients in a double-blind, placebo-controlled fashion.<br />

They found that compared to placebo, terazosin was safe and effective<br />

in improving both obstructive and irritative voiding symptoms as well<br />

as improved urinary flow rates (7). Terazosin was also found to have<br />

a beneficial anti-hypertensive effect on patients with baseline hypertension<br />

while having minimal blood pressure alteration in normotensive<br />

patients (8).<br />

In a pooled analysis of three double-blind, placebo-controlled studies,<br />

the alpha-blocker doxazosin was assessed with regard to safety and<br />

efficacy. Doxazosin was found to have a statistically significant<br />

improvement in peak urinary flow rate, symptom severity, and bothersomeness<br />

when compared to placebo. Furthermore, doxazosin was<br />

found to have a greater impact when utilized for patients with more<br />

advanced symptoms when compared with patients with mild symp-<br />

toms (9).<br />

Alpha-blockers as a class of BPW medical therapy have also been<br />

compared to the other major therapeutic class of drug, the 5-alpha<br />

reductase inhibitors, in a double-blind, placebo-controlled fashion.<br />

Safety and efficacy were compared in 1229 men with BPH randomized<br />

to groups consisting of placebo, terazosin, finasteride, and finasteride<br />

plus terazosin. At 1 -yr follow-up, the changes in AUA symptom score<br />

and flow rate was statistically better in the terazosin and terazosin plus<br />

finasteride groups compared to the placebo and finasteride groups.<br />

~inasteride was found to be no efficacious than placebo (10).


Chapter 10 I Male Bladder Outlet Obstruction 233<br />

Selective vs Nonselective Alphal, Blockade. Recently another subclass<br />

of alpha-blockers has come into widespread clinical use. Pharmacologic<br />

analysis has revealed that there are two subtypes of the alpha,adenoceptor<br />

designated 1 a and lb (1 I). While the alpha,,-adrenoceptor<br />

is found in great quantities in the prostate, the alphalb-adrenoceptor is<br />

more responsible for the contraction of peri-arterial smooth muscle<br />

(12,13). It has thus been postulated that a medical therapeutic targeted<br />

to the alpha,,-receptor would be more specific to prostatic symptoms<br />

and result in fewer side effects.<br />

Tamsulosin is a selective alpha,,-adrenoceptor antagonist which, in<br />

doses ranging from 0.2-0.4 mg daily, has been found to have a beneficial<br />

effect on symptoms of BPH while minimizing complaints of dry mouth<br />

and dizziness (14).<br />

While it is difficult to exactly equate a dosage of tamsulosin to<br />

other nonselective alphal, blockers, one of the attractive features of the<br />

selective alphal, blockers is the fact that no titration is necessary, thus<br />

simplifying the patient’ S initial regimen.<br />

A comparison between tamsulosin at a dosage of 0.2 mg/d and<br />

terazosin at a tiltrated dosage of l-S mg/d has been conducted in a<br />

group of Korean patients followed with maximal urinary flow rate<br />

and International Prostate Symptom Score (IPSS) in a single-blind,<br />

nonplacebo controlled fashion. While both medications resulted in<br />

improved maximal flow rates and Symptoms, tamsulosin resulted in<br />

a statistically significant improvement in the occurrence of adverse<br />

events (15).<br />

While there are currently no specific guidelines regarding the usage<br />

of one form of alpha-blocker therapy over another, clinicians are encouraged<br />

to tailor therapy to each individual patient.<br />

eduetase Inhibitors. The S-alpha reductase inhibitors<br />

are a class of medications, which acts on the crucial conversion of<br />

testosterone to its most active form dihydro-testosterone. The major<br />

medication in the class has been finasteride. Finasteride has been found<br />

to induce prostatic tissue changes via a suppression of the prostatic<br />

epithelium. In human studies performed on men with BPH, finasteride<br />

has been found to result in an approx SO% decrease in serum prostate<br />

specific antigen levels as well as a 21 % decrease in prostate volume (16).<br />

Furthermore, in clinical studies finasteride has been found to have<br />

beneficial effects on the signs and symptoms of BPH, as well as patients’<br />

overall health-related quality of life. While the toxicity profile of finasteride<br />

is minimal, some studies have reported a slight reduction in sexual


234 Ikeguchi et al.<br />

function (17). In general, finasteride therapy has an approximately 3%<br />

risk of erectile <strong>dysfunction</strong> 3% risk of <strong>dysfunction</strong>al ejaculation, and<br />

3-5% chance of diminished libido. Although finasteride therapy has<br />

not been found to definitively result in retrograde ejaculation, a small<br />

percentage of patients may complain of their own subjective sense of<br />

abnormal ejaculation.<br />

Nonetheless, many of the positive reports of finasteride effects on<br />

signs and symptoms of BPH have not held-up in direct comparison<br />

studies with the alpha-blocker medications (10). However, in one of<br />

the largest clinical studies addressing finasteride, the long-term effects<br />

of finasteride therapy was addressed in a cohort of over 3000 men with<br />

moderate to severe symptoms relating to BPH. Over a 4-yr period in<br />

a randomized, double-blind, placebo-controlled fashion, finasteride was<br />

found to reduce the incidence of prostate surgery and acute urinary<br />

retention by approximately one-half, while improving urinary symptom<br />

score (18).<br />

Thus, in spite of conflicting reports, the 5-alpha reductase inhibitors<br />

may still have a significant role in the scheme of medical therapy for<br />

BPH. From this standpoint, a new concept in medical therapy of BPH<br />

may emerge from a prophylactic standpoint, As demonstrated by<br />

Jacobsen et al., a subset of patients at greatest risk for the development<br />

of complications related to BPH may be able to be identified (4). In<br />

such case, clinicians should be alerted to those patients at higher risk<br />

of the ultimate development of BPH complications. The institution of<br />

prophylactic measures may prove to be beneficial to the patient and<br />

cost-effective for the economy in the long run.<br />

SURGICAL THERAPY<br />

Transur~thral Resection and TVP. The surgical therapy for BPH<br />

has a long and colorful history. While transurethral resection of the<br />

prostate (TURP) has remained the standard to which other therapies<br />

are compared, many newer techniques have found their own niche in<br />

the general scheme of surgical management of BPH. This is evidenced<br />

by the fact that the incidence of TURP in the United States reached<br />

its peak in 1987 and subsequently declined in use, most likely owing<br />

to improvements in medical therapy and the advent of several alternative<br />

surgical techniques (19).<br />

While the consensus remains that patients with LUTS should first<br />

be given a trial of medications, once medications have failed, the best<br />

remaining therapy appears to be surgical. In a multicenter randomized


Chapter 10 / Male Bladder Outlet Obstruction. 235<br />

trial, the institution of TURP was compared to watchful waiting in<br />

over 550 men with moderate LUTS owing to BPH. Over an average<br />

follow-up period of 2.8 yr, immediate TURP was found to be superior<br />

to watchful waiting in regards to treatment failures and in eradicating<br />

moderate to severe LUTS (20).<br />

Conventional TURP is generally considered to be a well-tolerated<br />

procedure with a perioperative mortality rate of approximately 2%. The<br />

kinds of immediate postoperative complications vary widely and in-<br />

clude: urinary tract infection, urinary retention, bleeding requiring blood<br />

transfusion, epididymitis, TURP syndrome, urinary incontinence, and<br />

sexual <strong>dysfunction</strong>. Regarding the risk of erectile <strong>dysfunction</strong>, although<br />

historical retrospective studies have reported risks ranging between<br />

5-35%, more recent prospective studies have not found a greater risk<br />

of erectile <strong>dysfunction</strong> amongst men undergoing TURP when compared<br />

to men with BPH under watchful waiting protocols (20). Clearly, the<br />

most commonly reported form of sexual <strong>dysfunction</strong> after TURP is<br />

retrograde ejaculation, which may occur in as many as 75% of patients.<br />

The risks of retrograde ejaculation should be discussed with the patient<br />

prior to operation particularly in those men who are younger and<br />

sexually active. Long-term complication from TURP may include the<br />

occurrence of bladder neck contracture and urethral structure.<br />

Benefits of TURP on urinary parameters have generally been reported<br />

as an increase in peak urinary flow rate of about 9.8 mL/s and an 88%<br />

likelihood of an improvement in subjective symptoms.<br />

In addition to the standard techniques of transurethral resection,<br />

recent advancements in the area of transurethral electrovaporization of<br />

the prostate (TVP) have made this option increasingly popular. TVP<br />

utilizes conventional TURP equipment while using a combination of<br />

electrosurgical vaporization and coagulation energy delivered through<br />

a modified form of rolling TURP “loop” or “roller” in order to desiccate<br />

and vaporize obstructing prostatic tissue. The objective behind TVP is<br />

the minimize the risk of bleeding and TURP syndrome, thus allowing<br />

the clinician a longer operative interval during which resection may<br />

be carried out for larger sized prostates. The draw-backs of TVP have<br />

been sited as being the increased time required to complete the resection<br />

and the inability to obtain prostatic tissue for pathologic review. We<br />

have found that both of these difficulties may be overcome in any given<br />

patient with the use of a combination of conventional loop followed by<br />

electrovaporization.<br />

The safety and efficacy of TVP has been investigated. In a series<br />

of 93 consecutive patients with LUTS undergoing TVP, 96% of patients


236 Ikeguchi et al.<br />

were able to undergo voiding trial within 24 h of operation and be<br />

discharged home. Both peak urinary flow rate and AUA symptom<br />

scores were found to improve in quantities comparable to TURP. The<br />

mean decrease in hematocrit was 1.4 mL/dL and TURP syndrome was<br />

not seen. No patients complained of erectile <strong>dysfunction</strong> although the<br />

incidence of retrograde ejaculation was 92%. An 8% incidence of<br />

irritative voiding symptoms was noted. TVP was concluded to be both<br />

safe and effective and minimized the postoperative hospital stay (21).<br />

TUP. Owing to the increasing body of evidence to suggest a significant<br />

role of the prostatic capsule in the pathophysiology of bladder<br />

outlet obstruction, the transurethral incision of the prostate (TUIP) has<br />

become popularized as a less extensive operation, which may offer<br />

symptomatic and objective relief. Rather than a full resection of the<br />

inner adenoma of the gland, either one or two incisions are made into<br />

the prostatic urethra between the limits of the bladder neck or trigone<br />

to the verumontanum. The incisions are generally made on either or<br />

both sides of the 6 o’clock position. While descriptions of the procedure<br />

have included use of any cutting device ranging from a cold knife to<br />

a laser, we currently utilize a resectoscope loaded with an electrovaporization<br />

roller. The extent of the incision should extend at least to the<br />

level of the prostatic capsule, with many reports advocating the usage<br />

of a through-capsule incision resulting in the visualization of extracapsular<br />

fat.<br />

While treatment outcomes are not superior to TURP, TUP is still<br />

reported to improve flow rates and symptom scores better than with<br />

medical therapy. The most notable aspect of TUP is the markedly lower<br />

incidence of retrograde ejaculation, which has varied between 0-37%.<br />

In a prospective trial comparing TURP, TUIP, and prostatic balloon<br />

dilation, in 60 men with LUTS, TUIP was found to be the method of<br />

choice for young patients with smaller prostatic adenoma and/or bladder<br />

neck obstruction (22).<br />

Minimally Invasive Surgical Therapies. As an effort to further<br />

lessen the invasive nature of TURP and TIJIP, several attempts have<br />

been made at developing technologies amenable to minimal (i.e. local)<br />

anesthesia. Although each treatment is less extensive; however, the<br />

minimally invasive nature often comes at the expense of efficacy. Given<br />

the risk of marginal efficacy for all of these therapies, patients should<br />

be well-informed of the risks. Clinicians should also try to appropriately<br />

select patients with mild to moderate symptoms, which generally have<br />

a better chance at improvement than those with more severe problems.


Chapter 10 I Male Bladder Outlet Obstruction 237<br />

Furthermore, as in any algorithm for the treatment of BPH symptoms,<br />

patients should be offered a trial of medical therapy prior to attempts<br />

at surgical procedures.<br />

Inter~stitia~ Theories: ~rans~rethr~l Needle Ablation ( ~ ~ and ~ A<br />

I~terstitiaZ Diode Laser. Interstitial therapies of the prostate are attractive<br />

from the standpoint that there is no violation and resection of the<br />

prostatic urethral mucosa. It would seem intuitive that this would result<br />

in much less bleeding, urethral irritation, and intra-operative discomfort.<br />

Interstitial therapies are based upon the principal of delivering some<br />

form of destructive energy to the prostatic adenoma. Ultimately, the<br />

destroyed prostatic tissue is passively resorbed, resulting in a "defect"<br />

in the prostatic urethra. In general, these therapies work best for patients<br />

with smaller prostatic gland size and median lobe involvement.<br />

TUNA therapy utilizes radio-frequency energy delivered through a<br />

shielded needle device to ablate prostatic tissue. In a prospective study<br />

of patients with bladder outlet obstruction and LUTS, TUNA was<br />

performed under conditions of intra-urethral anesthesia in an outpatient<br />

setting. Patients starting with urinary symptoms measured by the international<br />

prostate symptom score (IPSS) at average 20.8 t- 4.5 improved<br />

to 6.2 t- 2.9 at l-yr follow-up. Likewise, urinary flow improved from<br />

an average baseline of 8.2 t- 3.4 mL/s to l 5.9 t- 2.1 at l-yr postoperatively<br />

(23).<br />

Interstitial laser coagulation of the prostate has also been studied<br />

recently. In a group of l12 men with LUTS secondary to BPH, interstitial<br />

laser therapy was undertaken. AUA symptoms score was found to<br />

decrease from a baseline average of 20.9-7.9 at 6-m0 follow-up. Peak<br />

urinary flow rate was also found to improve with therapy, from a<br />

baseline of 8.0 mL/s to 14.2 mL/s at 6 mo. Reported complications<br />

were minimal although a 2.7% of patients required re-treatment or<br />

underwent TURP (24).<br />

Prostatic ~icrowave Therapy. Prostatic microwave therapy relies<br />

on hyperthermia produced by the emission of microwave energy from<br />

a specially engineered catheter. The energy is delivered to the lateral<br />

lobes of the prostate, while there is simultaneous cooling of the urethral<br />

mucosa. Intraprostatic temperatures attained by the currently available<br />

devices may reach 45"-75" C and patients generally undergo one or<br />

more hour-long sessions.<br />

Reports of treatment outcomes have been conflicting. In a metaanalysis<br />

of recent clinical studies, microwave therapy was found to<br />

result in significant sy~ptomatic improvement as well as increased


238 Ikeguchi et al.<br />

urinary flow rates of 35% above baseline (25). However, in another<br />

study comparing various microwave techniques to TURP in a randomized<br />

controlled fashion, microwave therapy did not result in any<br />

improvement in the objective parameters of bladder outlet obstruction<br />

as measured by urodynamics. Nonetheless, microwave therapies did<br />

result in an improvement in subjective symptoms with a mean AUA<br />

symptom score decrease from 18.4-5.2. Complications following prostatic<br />

microwave therapy have been found to be low but do include<br />

retrograde ejaculation in as many as 22% in some series (26).<br />

Urethral Stent. Urethral stents were first introduced in 1987 as an<br />

alternative means of treating urethral strictures. Its technology has since<br />

been applied to the treatment of bladder outlet obstruction secondary<br />

to BPH for patients who are at high surgical risk. From a technical<br />

standpoint, the most challenging aspect of obtaining a favorable outcome<br />

is in the proper initial positioning of the device. The device is<br />

a titanium mesh that is loaded into a delivery mechanism resembling<br />

a rigid cystoscopy apparatus. Under direct vision, the stent should be<br />

deployed from the bladder neck without extending beyond the verumontanum.<br />

The procedure is very fast to perform and can be performed<br />

under spinal or local anesthesia depending on the individual patient.<br />

Over a period of several weeks, the normal urethral epithelium will<br />

grow into the interstices of the stent and ultimately should result in an<br />

open urethra covered with epithelium. Complications from urethral<br />

stent placement include stent migration, stone formation, colonization,<br />

and persistent urinary infection and intractable detrusor instability.<br />

In an effort to study the long term efficacy and safety of urethral<br />

stent placement for the treatment of bladder outlet obstruction secondary<br />

to BPH, 144 patients underwent implantation and were followed for<br />

up to 24 mo. Amongst patients in urinary retention, at 24 mo, average<br />

AUA symptom score was 5.21 2 0.81 with an average peak urinary<br />

flow rate of 11.34 t- 1.12. Among patients treated for symptoms of<br />

bladder outlet obstruction, at 24 mo follow-up, AUA symptom score<br />

decreased from 15.89 J. 0.47 to 9.33 2 0.86 while peak urinary flow<br />

rate improved from 8.59 t- 0.41 mL/s to 11.43 2 1.12 mWs. Although<br />

the incidence of stent removal for migration was as high as 19%, the<br />

authors concluded that given proper placement, urethral stent placement<br />

was a safe and effective therapy for high-risk patients with prostatic<br />

obstruction (27).<br />

Primary Bladder Neck Obstruction. Primary bladder neck obstruction<br />

is a condition usually seen in men of a slightly younger age group<br />

(20-50 yr) than those with BPH are. The exact pathophysiology of the


Chapter 10 / Male Bladder Outlet Obstruction 239<br />

disease is poorly understood, and patients are commonly treated for<br />

years with therapies geared towards other diagnoses such as chronic<br />

prostatitis. Typically, patients present with complaints of moderate to<br />

severe LUTS with diminished urinary flow rates and an overall clinical<br />

scenario consistent with bladder outlet obstruction. We currently<br />

employ video-urodynamics in making the diagnosis, with which voiding<br />

cystourethrogram will reveal a “cut-off” demarcation of contrast at the<br />

bladder neck (28).<br />

MEDICAL THERAPY<br />

Alpha-blockade. While previously primary bladder neck obstruction<br />

was thought to be a result of the deposition of fibrous tissue around<br />

the bladder neck, histologic analysis has found this to be untrue (29).<br />

Other theories of increased sympathetic nervous input to the bladder<br />

neck have also been proposed. This has lead to investigations mapping<br />

neuropeptides to the bladder neck, which have confined the presence<br />

of a possible neuropeptide-Y mediated sympathetic <strong>dysfunction</strong> as the<br />

etiology of primary bladder outlet obstruction (30).<br />

Thus the rationale behind the usage of alpha-blocking agents as the<br />

treatment for primary bladder outlet obstruction extends from this the-<br />

ory. Nonetheless, alpha-blockers of all types have been found to be<br />

notoriously ineffective in treating this group of patients. While many<br />

patients presented with their options will choose an initial trial of<br />

medical therapy, most will ultimately come to operation. In a retrospec-<br />

tive review of 36 men with primary bladder neck obstruction, Trockman<br />

et al. studied outcomes after treatment with alpha-blockers, TUIP, and<br />

watchful waiting. They found that of patients begun on alpha-blocker<br />

therapy, only 30% continued this treatment owing to a lack of effi-<br />

cacy (31).<br />

SURGICAL THERAPY<br />

Endoscopic Bladder Neck Incision and TUIP. As discussed pre-<br />

viously, the mainstay of therapy for primary bladder neck obstruction<br />

is transurethral incision of the bladder neck, and in some cases TUIP.<br />

Formal TURP is usually not indicated since most patients with primary<br />

bladder neck obstruction will have small prostates on cystoscopic exam-<br />

ination. Furthermore, since the majority of patients diagnosed with<br />

primary bladder neck obstruction are young, risks of retrograde ejacula-<br />

tion usually make more aggressive surgical therapies unacceptable.


240 Ikeguchi et al.<br />

Amongst men undergoing transurethral incision of the bladder neck,<br />

'reported success rates have been excellent. In our own experience,<br />

of 33 patients undergoing bladder neck incision, 30 had a marked<br />

improvement in symptoms and an average peak urinary Bow rate of<br />

15.7 mL/s at 6 mo follow-up, with no patients developing retrograde<br />

ejaculation (32).<br />

ral Stricture Disease. Urethral stricture disease has plagued<br />

men for centuries, with descriptions of the problem dating back to the<br />

ancient civilizations of India, Egypt and Greece. To this day, urethral<br />

stricture disease remains a major etiology of bladder outlet obstruction.<br />

Common etiologies of urethral strictures include sexually transmitted<br />

diseases (STDs), trauma either incidental or iatrogenic, and least com-<br />

monly carcinoma. Furthermore, as ever before, a single effective therapy<br />

remains elusive. Even amongst urologists, there is relatively little con-<br />

sensus. Part of the problem arises from that fact that few cases of<br />

urethral strictures are predictably categorized to correspond to progno-<br />

sis. Although it is generally true that traumatic strictures are short while<br />

urethritis results in longer strictures, the behavior and disposition of a<br />

urethral stricture in any given patient rnay behave differently from<br />

one individual to the next. Owing to this unpredictability, the natural<br />

tendency of both clinician and patient is to favor less invasive tech-<br />

niques, which frequently prove to be suboptimal in efficacy.<br />

The diagnosis of urethral stricture disease is frequently suggested<br />

in the patient's history. Particularly in younger men, the history of<br />

pertinent risk factors and the onset of LUTS should prompt the clinician<br />

to be suspicious of an underlying stricture. Definitive diagnosis may<br />

be confirmed radiographically with antegrade and retrograde urethral<br />

contrast studies. In some cases, for patients presenting in urinary reten-<br />

tion, an indwelling Foley catheter or suprapubic catheter rnay be<br />

required in order to temporize before definitive therapy may be insti-<br />

tuted.<br />

ON, ENDOSCOPIC U~T~ROTO~Y,<br />

Endoscopic urethrotomy or urethral dilation is frequently the first<br />

line of therapy offered to patients when the diagnosis of urethral stricture<br />

is first made. Given the minimally invasive nature of the procedure, there<br />

are fewer operative risks. However, the problem with these therapies is<br />

their lack of long-term efficacy. Studies of long-term urethrotomy follow-up<br />

reveal a success rate of approx 70% for single short (4 cm)


Chapter 10 / Male Bladder Outlet Obstruction 241<br />

strictures) and 18% for longer strictures. These percentages appear to<br />

be consistent regardless of the location or etiology of the stricture<br />

disease. Furthermore, with each recurrence, the chance of stricture<br />

resolution with urethrotomy or dilation dwindles in an exponential<br />

fashion. Ultimately, the degree of urethral and spongiosa1 fibrosis<br />

extends to a point that re-incisions of failed urethrotomies are nearly<br />

useless. Clearly it behooves the clinician to move ahead with more<br />

aggressive open repair when an initial conservative attempt has failed.<br />

In addressing the issue of endoscopic urethrotomy versus urethral<br />

dilation, a recent prospective randomized trial was performed compar-<br />

ing the two. Among 210 men with proven urethral stricture disease,<br />

106 underwent filiform dilation 104 underwent internal urethrotomy.<br />

For all patients, there was a 12-1310 recurrence rate of 40% for strictures<br />

less than 2 cm and 80% for those greater than 4 cm. The authors found<br />

that both treatment modalities were equivalent in efficacy for the initial<br />

treatment of strictures and that both methods had a higher failure rate<br />

with increasing stricture length. They concluded that either dilation or<br />

urethrotomy could be used as first line therapy for strictures less than<br />

2 cm, open urethroplasty for strictures greater than 4 cm, and a trial<br />

of dilation or urethroplasty for strictures measuring between 2 cm and<br />

4 cm (33).<br />

Urethral stent technology was initially developed for the therapy of<br />

urethral stricture disease and has since been extrapolated to the therapy<br />

of BPH (see above). In the treatment of urethral strictures, some studies<br />

have now revealed the efficacy of these devices. In general, urethral<br />

stents have been used for patients with short strictures located in the<br />

bulbar urethra. As in the case with its BPH indication, the procedure<br />

may be accomplished with minimal anesthesia and stress to the patient,<br />

and may be ideal for patients who are poor surgical candidates. Deploy-<br />

ment of the device is the most crucial factor and usually necessitates<br />

the performance of a urethral dilation or incision of the stricture. Once<br />

the delivery device is able to pass the stricture, the stent is disengaged<br />

such that the length of the stent overlaps the stricture.<br />

Although some studies have reported success rates greater than 80971,<br />

most lack long-term follow-up. In 1995, Badlani et al. reported on the<br />

reports of a prospective multicenter study undertaken to evaluate the<br />

long-term efficacy of the Urolume stent used for patients with recurrent<br />

strictures of the bulbar urethra. Out of 175 patients enrolled, at I and


242 Ikeguchi et al.<br />

2 yr follow-up, 139 and 8 1 patients were available for evaluation,<br />

respectively. The authors found that the re-treatment rate was 14.3%<br />

at l-yr postimplantation compared to 75.2% preimplantation, and con-<br />

cluded that the treatment of an excellent therapeutic option for recurrent<br />

bulbar urethral strictures (34).<br />

3ehavioraZ <strong>Voiding</strong> Dysfinction (Pseudodyssynergia)<br />

Bladder outlet obstruction may also occur in the setting of an anatomically<br />

normal male patient owing to a dynamic obstruction caused by<br />

a lack of coordination between the detrusor and bladder neck smooth<br />

muscle (bladder neck dyssynergia) or striated muscle external sphincter<br />

(pseudodyssynergia). While psydodyssynergia generally implies a condition<br />

in a neurologically intact patient, similar conditions may be the<br />

result of a neurologic lesion (detrusor external sphincter dyssynergia,<br />

DESD) or in neurologically intact children (Binman’s bladder).<br />

Although the exact etiology of this condition is unknown, it is generally<br />

considered to be a learned behavior of failed pelvic relaxation. It has<br />

been postulated that the condition may arise as reaction to a negative<br />

response to micturition such as pain, and frequently is seen in younger<br />

male patients who have been treated for long periods for chronic prostatitis.<br />

Although not an absolute part of the syndrome, there is a preponderance<br />

of patients who are generally apprehensive and nervous young<br />

men who could be described as type A personalities.<br />

Typically, patients may complain of either obstructive or irritative<br />

symptoms with otherwise unremarkable prostatic examination and postvoid<br />

residual determination. The definitive diagnosis of pseudodyssynergia<br />

requires the use of pressure-flow-EMG studies. The presence<br />

of increased striated sphincter activity corresponding with detrusor<br />

contraction in the absence of abdominal straining is consistent with<br />

the diagnosis. On voiding cystourethrography, there may be a demonstration<br />

of narrowing of the urethra at the level of the membranous<br />

urethra. As in any condition of bladder outlet obstruction, secondary<br />

bladder responses may be manifest in the form of detrusor hypertrophy,<br />

diminished capacity and compliance, and instability, or bladder<br />

decompensation with larger capacity and residual volumes.<br />

Therapy in these cases may frequently be started with a trial of<br />

alpha-blockers though this usually meets with minimal success. Preferred<br />

therapy for such patients include behavioral modification. and<br />

biofeedback, with symptomatic improvement seen in over 80% of<br />

patients over a 6-m0 period (35).


Chapter 10 I Male Bladder Outlet Obstruction 243<br />

REFERENCES<br />

1. Kaplan SA, Reis RB (1996) Significant correlation of the American Urological<br />

Association symptom score and a novel urodynamic parameter: detrusor contraction<br />

duration. J Urol Nov; 156(5):1668-1672.<br />

2. Graves EJ (1995) Detailed Diagnoses and Procedures, National Discharge Survey,<br />

1993. National Center for Health Statistics. Vital Health Stat 13:122.<br />

3. Garraway WM, Russell EBAW, Lee RJ, Collins GN, McKelvie GB, Hehir M,<br />

Rogers ACN, Simpson RJ (1993) Impact of previously unrecognized benign<br />

prostatic hyperplasia on the daily activities of middle-aged and elderly men. Br<br />

J Gen Pract 43:318-321.<br />

4. Jacobsen SJ, Jacobson DJ, Girman CJ, Roberts RO, Rhodes T, Guess HA, Lieber<br />

MM (1997) Natural history of prostatism: risk factors for acute urinary retention.<br />

J Urol Aug 158(2):481-487.<br />

5. Chapple CR (1995) a-Adrenergic blocking drugs in bladder outflow obstruction;<br />

what potential has al-adrenoceptor selectivity? Br J Urol 76 (Suppl 1):47-55.<br />

6. Yamada S, Tanaka C, Kimura R, Kawabe K (1994) Alpha-l-adrenoceptors in<br />

human prostate: characterization and binding characteristics of alpha- 1-antagonists.<br />

Life Sci 54: 1845-1854.<br />

7. Elhilali MM, Rarnsey EW, Barkin J, Casey RW, Boake RC, Beland G, Fradet<br />

Y, Trachtenberg J, Orovan WL, Schick E, Klotz LH (1996) A multicenter, randomized,<br />

double-blind, placebo-controlled study to evaluate the safety and efficacy<br />

of terazosin in the treatment of benign prostatic hyperplasia. Urology Mar<br />

47(3):335-342.<br />

8. Debruyne FM, Witjes WP, Fitzpatrick J, Kirby R, Kirk D, Prezioso D (1996)<br />

The international terazosin trial: a multicentre study of the long-term efficacy and<br />

safety of terazosin in the treatment of benign prostatic hyperplasia. The ITT Group.<br />

Eur Urol 30(3):369-376.<br />

9. Roehrborn CG, Siege1 RL (1996) Safety and efficacy of doxazosin in benign<br />

prostatic hyperplasia: a pooled analysis of three double-blind, placebo-controlled<br />

studies. Urology Sep; 48(3):406-415.<br />

10. Lepor H, Williford WO, Barry MJ, Brawer MK, Dixon CM, Gormley G, Haakenson<br />

C, Machi M, Narayan P, Padley RJ (1996) The efficacy of terazosin, finasteride,<br />

or both in benign prostatic hyperplasia. Veterans Affairs Cooperative Studies<br />

Benign Prostatic Hyperplasia Study Group. NEngE JMed Aug 22;335(8):533-539,<br />

11. Hieble JP, Bylund DB, Clarke DE, et al. (1995) International Union of Pharmacology.<br />

X Recommendation for nomenclature of al-adrenoreceptor: Consensus<br />

Update. Pharmacol Rev 47:267-270.<br />

12. Foglar R, Shibata K, Horie K, et al. (1995) Use of recombinant al-adrenoreceptors<br />

to characterize subtype selectivity of drugs for the treatment of prostatic hypertrophy.<br />

Eur J Pharmacol Mol Pharmaco1288:201-207.<br />

13. Hatano A, Takahashi H, Tamalci M, et al. (1994) Pharmacological evidence of<br />

distinct al-adrenoreceptor subtypes mediating the contraction of human prostatic<br />

urethra and peripheral artery. Br J Pharmacol 113:723-728.<br />

14. Chapple CR (1996) Selective al-adrenoreceptor antagonists in benign prostatic<br />

hyperplasia: rationale and clinical experience. Eur Urol 29: 129-144.<br />

15. Lee E, Lee C (1997) Clinical comparison of selective and non-selective alpha 1Aadrenoreceptor<br />

antagonists in benign prostatic hyperplasia: studies on tamdosin in<br />

a fixed dose and terazosin in increasing doses. Br J Urd Oct;80(4):606-611.


244 Ikeguchi et al.<br />

16. Marks LS, Partin AW, Gormley GJ, Dorsey FJ, Shery ED, Garris JB, Subong<br />

EN, Stoner E, deKernion JB (1997) Prostate tissue composition and response to<br />

Gnasteride in rnen with symptomatic benign prostatic hyperplasia. J Urol<br />

Jun;157(6):2171-2178.<br />

17. Ginnan CJ, Kolman C, Liss CL, Bolognese JA, Binkowitz BS, Stoner E (1996)<br />

Effects of finasteride on health-related quality of life in rnen with symptomatic<br />

benign prostatic hyperplasia. Finasteride Study Group. Prostate Aug; 29(2):83-90.<br />

18. McConnell JD, Bruskewitz R, Walsh P, Andriole G, Lieber M, Holtgrewe HL,<br />

Albertsen P, Roehrborn CC, Nickel JC, Wang DZ, Taylor AM, Waldstreicher J<br />

(1998) The effect of finasteride on the risk of acute urinary retention and the need<br />

for surgical treatment among rnen with benign prostatic hyperplasia. Finasteride<br />

Long-Term Efficacy and Safety Study Group. N Engl J Med Feb 26;338(9):557-<br />

563.<br />

19. Lu-Yao CL, Barry ML, Chang CH, Wasson JH, Wennberg JE (1994) Transurethral<br />

resection of the prostate among Medicare beneficiaries in the United States: time<br />

trends and outcomes. Prostate Patient Outcomes Research Team (PORT). Urology<br />

N0~;44(5):692-698.<br />

20. Wasson JH, Reda DJ, Bruskewitz RC, Elinson J, Keller AM, Henderson WG<br />

(1995) A comparison of transurethral surgery with watchful waiting for moderate<br />

symptoms of benign prostatic hyperplasia. The Veterans Affairs Cooperative<br />

Study Group on Transurethral Resection of the Prostate. N EngE J Med Jan<br />

12; 332(2):75-79.<br />

21. Te AE, Santarosa R, Eaplan SA (I 997) Electrovaporization of the prostate: electrosurgical<br />

modification of standard transurethral resection in 93 patients with benign<br />

hyperplasia. J EndouroE Feb; 1 I ( 1):7 1-75.<br />

22. Saporta L, Aridogan IA, Erlich N, Yachia D (1 996) Objective and subjective<br />

comparison of transurethral resection, transurethral incision and balloon dilatation<br />

of the prostate. A prospective study. Eur Urol 29(4):439-45.<br />

23. Campo B, Bergamaschi F, Corrada P, Ordesi G (1997) Transurethral needle<br />

ablation (TUNA) of the prostate: a clinical and urodynarnic evaluation. Urology<br />

Jun;49(6):847-850,<br />

24. Muschter R, de la Rosette JJ, Whitfield H, Pellerin JP, Madersbacher S, Gillatt<br />

D (1 996) Initial human clinical experience with diode laser interstitial treatment<br />

of benign prostatic hyperplasia. Urology Aug;48(2):223-228.<br />

25. de la Rosette JJ, D’ Ancona FC, Debruyne FM (1997) Current status of thermotherapy<br />

of the prostate. J Urol Feb;157(2):430-438.<br />

26. Ahrned M, Bell T, Lawrence WT, Ward JP, Watson GM (1997) Transurethral<br />

microwave thermotherapy (Prostatron version 2.5) compared with transurethral<br />

resection of the prostate for the treatment of benign prostatic hyperplasia: a<br />

randomized, controlled, parallel study. Br J Urol Feb 79(2): 181-185.<br />

27. Kaplan SA, Chiou RK, Morton WJ, Katz PG (1995) Long-term experience utilizing<br />

a new balloon expandable prostatic endoprosthesis: the Titan stent. North American<br />

Titan Stent Study Group. Urology Feb;45(2):234-240.<br />

28. Kaplan SA, Ikeguchi EF, Santarosa RP, D’ Alisera PM, Hendricks J, Te AE, Miller<br />

MI (1996) Etiology of voiding <strong>dysfunction</strong> in men less than 50 years of age.<br />

Urology Jun;47(6):836-839.<br />

29. Norlen LJ, Blaivas JG (1986) Unsuspected proximal urethral obstruction in young<br />

and middle-aged men. J Urol 135:972.<br />

30. Crowe R, Noble J, Robson T, Soediono P, Milroy EJG and Burnstock G (1995)<br />

An increase of neuropeptide Y but not nitric oxide synthase-immunoreactive


Chapter 10 / Male Bladder Outlet Obstruction 245<br />

nerves in the bladder neck from male patients with bladder neck<br />

J Urd 154: 123 1.<br />

dyssynergia.<br />

31. Trockman BA, Gerspach J, Dmochowski R, Haab F, Zimmern PE, Leach GE<br />

(1996) Primary bladder neck obstruction: urodynamic findings and treatment<br />

results in 36 men. J Urd Oct;l56(4): 1418-1420.<br />

32. Kaplan SA, Te AE, Jacobs BZ (1994) Urodynamic evidence of vesical neck<br />

obstruction in men with misdiagnosed chronic nonbacterial prostatitis and the<br />

therapeutic role of endoscopic incision of the bladder neck. J UroE Dec;152(6<br />

Pt 1):2063-2065.<br />

33. Steenkamp JW, Heyns CF, de Kock ML (1997) Internal urethrotomy versus<br />

dilation as treatment for male urethral strictures: a prospective, randomized comparison.<br />

J Urol Jan 357(1):98-101.<br />

34. Badlani GH, Press SM, Defalco A, Oesterling JE, Smith AD (1995) Urolume<br />

endourethral prosthesis for the treatment of urethral stricture disease: long-term<br />

results of the North American Multicenter UroLume Trial. Urology May<br />

45(5):846-856.<br />

35. Kaplan SA, Santarosa RP, D’Alisera PM, Fay BJ, Ikeguchi EF, Hendricks J,<br />

Klein L, Te AE (1997) Pseudodyssynergia (contraction of the external sphincter<br />

during voiding) misdiagnosed as chronic nonbacterial prostatitis and the role of<br />

biofeedback as a therapeutic option. J Urd Jun 157(6):2234-2237.


This Page Intentionally Left Blank


1 l post-Prostatectomy Incontinence<br />

CONTENTS<br />

INTRODUCTION<br />

INCIDENCE OF INCONTINENCE<br />

AFTER PROSTATECTOMY<br />

ANATOMY OF THE PROSTATE GLAND<br />

AND URINARY SPHINCTER<br />

PROSTATECTOMY: TECHNICAL POINTS<br />

REGARDING PRESERVATION OF CONTINENCE<br />

ETIOLOGY OF INCONTINENCE<br />

FOLLOWING PROSTATECTOMY<br />

EVALUATION OF POSTPROSTATECTOMY<br />

INCONTINENCE<br />

TREATMENT OF POSTPROSTATECTOMY INCONTINEN<br />

TREATMENT OF POSTPROSTATECTOMY<br />

INCONTINENCE: A COST COMPARISON<br />

SUMMARY<br />

REFERENCES<br />

INTRODUCTION<br />

Adenocarcinoma of the prostate is the most frequently occurring<br />

cancer in men (I). Recently, the American Urologic Association<br />

released guidelines for the management of prostate cancer. These guide-<br />

lines confirmed that radical prostatectomy is the optimal treatment for<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ


248 Scarper0 and<br />

localized prostate cancer in healthy men (2). Incontinence following<br />

prostatectomy is a condition with significant adverse effects on quality<br />

of life that all physicians caring for patients with prostate disorders<br />

will encounter. Complete knowledge of the etiology of incontinence<br />

following prostatectomy and the options of treatment will facilitate<br />

treatment of these often distressed patients.<br />

INCIDENCE OF INCONTINENCE<br />

AFTER ~ROSTA~ECTO~~<br />

The incidence of urinary incontinence is approximately l-3% in<br />

patients undergoing transurethral resection of the prostate or open prostatectomy<br />

for benign disease (3,4). The incidence of incontinence following<br />

radical prostatectomy has been reported to range from 2.5 (5)<br />

to 87% (6). This wide discrepancy is a result of several factors. The<br />

definition of incontinence varies widely among series, ranging from<br />

any degree of wetting or restricted to total incontinence. In addition,<br />

the method of data acquisition has a significant effect on reported rates<br />

of incontinence. Studies that involve patient questionnaires andlor direct<br />

patient input generally have higher rates of incontinence than data<br />

obtained by chart review or physician interview. In a sample of Medicare<br />

patients undergoing radical prostatectomy, 47% had leakage of<br />

urine daily, and 6% needed surgical intervention (7). Incontinence after<br />

prostatectomy has a significant negative impact on a patient’s quality<br />

of life. Herr discovered in a questionnaire-based study that in 26% of<br />

patients surveyed, incontinence seriously affected the overall quality<br />

of life, and that 53% of patients 5 years after surgery would not undergo<br />

surgery again (8). In a survey of quality of life issues in men treated<br />

for localized prostate cancer compared to a normal age-matched control,<br />

patients following radical prostatectomy scored significantly worse on<br />

a scale evaluating urinary function (9). Postoperative interval also may<br />

effect reported rates of continence, as urinary control progressively<br />

improves in the year following surgery (10). Almost all patients have<br />

some degree of incontinence immediately after catheter removal, but<br />

several series document a progressive reduction in incontinence rates<br />

up to one year after prostatectomy (11,12). Thus, it may take up to<br />

one year to determine the final continence status of a patient following<br />

prostatectomy.


Chapter 11 / Post-Prostatectomy Incontinence 249<br />

ANATOMY OF THE PROSTATE GLAND<br />

AND URINARY SPHINCTER<br />

The prostate gland is comprised of three glandular zones and one<br />

nonglandular region, the anterior fibromuscular stroma. The three glandular<br />

regions arise from different segments of the prostatic urethra<br />

(13). The transition zone constitutes about 5% of prostatic glandular<br />

tissue and is located on both sides of the prostatic urethra. It is in this<br />

zone that benign prostatic hyperplasia develops (14). The transition<br />

zone is separated from the central and peripheral zones of the prostate<br />

by the surgical capsule, which marks adequate depth of resection of<br />

adenomatous prostatic tissue when performing transurethral prostatectomy.<br />

The peripheral zone comprises about ’70% of the mass of the<br />

glandular prostate, and it is in this zone that most prostate cancers arise<br />

(15). It is located at the posterior and lateral aspects of the prostate<br />

gland with its ducts exiting from the base of the verumontanum to the<br />

prostatic apex. The central zone constitutes about 25% of the glandular<br />

tissue around the base of the prostate, and the ejaculatory ducts travel<br />

through the central zone to the verumontanum.<br />

The male urethral sphincter consists of two anatomic zones responsible<br />

for continence (16) (Fig. l). The proximal ure~hral sphincter is<br />

centered at the bladder neck and includes smooth musculature of the<br />

bladder base, bladder neck, and prostatic urethra to the verumontanum.<br />

These smooth muscle fibers are arranged largely in a circular fashion<br />

at the bladder neck and proximal urethra, and this area has abundant<br />

alpha adrenergic innervation (1 7). The distal urethral sphincter extends<br />

from the verumontanum, across the apex of the prostate and into the<br />

perineal membrane. The distal urethral sphincter is comprised of three<br />

parts: smooth periurethral muscle along the prostatic and membranous<br />

urethra, the rhabdosphincter of specialized slow-twitch striated muscle<br />

fibers contributing to increased urethral tone, and extrinsic striated<br />

muscle, “the urogenital diaphragm,” containing fast-twitch striated<br />

fibers that are recruited under periods of increased intraabdominal<br />

pressure (18). The striated muscle fibers of the rhabdosphincter have<br />

a concentric, cylindrical design around the prostatomembranous urethra<br />

and are not contained within a “flat” urogenital diaphragm (19).<br />

Surrounding musculofascial and skeletal structures provide a framework<br />

critical for the male urethral sphincter to function properly (20).<br />

The pubourethral ligaments comprise a median suspensory mechanism<br />

for the urethra beneath the subpubic arch (21). Dorsal anchoring of


250 Scarper0 and Winters<br />

Fig. 1. Male urethral sphincter anatomy.<br />

the rhabdosphincter is carried out by the connection of the sphincter<br />

to Denonvillier’s fascia by the midline dorsal raphe (22). It has been<br />

demonstrated that this raphe is continuous with a broad musculofascial<br />

plate providing a broad base of support, suspending and stabilizing the<br />

rhabdosphincter (20). These supporting structures facilitate urethral<br />

elevation and closure promoting continence.<br />

Variations of the shape of the prostatic apex have been shown to<br />

clinically important (23). Myers et al. noted that in “doughnut shape”<br />

prostates with large lateral lobes and prostatic hypertrophy anterior to<br />

the urethra, the urethra exits at the apex. In “croissant shape” prostates<br />

with little anterior hypertrophy and a prostatic notch, the urethra exits<br />

proximal to the apex of the prostate. The authors concluded that the<br />

variation in shape of the prostatic apex should be considered at radical<br />

prostatectomy so that dissection can spare urethral length and periure-<br />

thral supporting structures (23).<br />

“W ”


Chapter l1 I Post-Prostatectomy Incontinence 25 1<br />

Prostatic surgery destroys the proximal urethral sphincter, and conti-<br />

nence totally relies on the distal urethral sphincter (24). Therefore, a<br />

thorough understanding of the anatomy of the prostate and surrounding<br />

structures is essential to avoid destruction of the distal urethral sphinc-<br />

teric mechanism.<br />

PROSTATECTOMY: TECHNICAL POINTS<br />

REGARDING PRESERVATION OF CONTINENCE<br />

A knowledge of male pelvic and sphincteric anatomy will facilitate<br />

surgeons to insure better outcomes and improved continence rates<br />

postoperatively. Perhaps the best treatment of incontinence following<br />

prostatectomy is to prevent it by preserving sphincteric anatomy and<br />

pelvic-floor supporting structures without compromising cancer control.<br />

Transurethral Resection of the Prostate<br />

and Open Prostatectomy<br />

In many cases, prostatic hypertrophy extends beyond the verumonta-<br />

num (25,26) (Fig. 2). Thus, the verumontanum remains a critical land-<br />

mark when performing a transurethral prostatectomy. If the resection<br />

is carried past the veru, violation of the distal urethral sphincter will<br />

occur, resulting in postoperative incontinence. The distal limit of resec-<br />

tion should be at the verumontanum, even if a small rim of adenoma is<br />

left behind. This small amount of adenoma may provide some protection<br />

against stress urinary incontinence (27). Care should be taken to mini-<br />

mize bleeding to insure adequate visualization, and cautery should be<br />

employed judiciously in this area to avoid thermal muscle damage.<br />

Most frequently damage to the distal urethral sphincter occurs anteriorly<br />

between the 10 and 2 o’clock positions (16).<br />

When performing open prostatectomy for benign disease, great care<br />

must be taken to preserve the distal sphincteric mechanism. Most of<br />

these glands are quite large with adenomatous tissue extending beyond<br />

the veru. Meticulous technique should be employed when separating<br />

the urethra during enucleation. If one places the adenoma on too much<br />

upward traction, this may also pull upward a significant portion of the<br />

distal sphincter resulting in transection of the sphincter. One should<br />

avoid upward pulling of the distal adenoma during enucleation, and<br />

thin out the urethra as much as possible before transection. One may


252 Scarpero and Winters<br />

Fig. 2. Schematic illustratioii of benign prostate adenoma extending into distal<br />

urethral sphincter.<br />

consider the enucleation of one lateral lobe at a time, or open the<br />

capsule and prostatic urethra to visualize the veru prior to enucleation<br />

of the distal adenoma.<br />

Radical Retrop ubic Prosta tectomy<br />

Radical retropubic prostatectomy has become the most predomi-<br />

nately performed procedure to treat localized prostate cancer following<br />

observations by Walsh and colleagues that have allowed for a more<br />

anatomic dissection, resulting in decreased blood loss and preservation<br />

of erectile function in a significant number of patients (28). By preserv-<br />

ing the cavernous nerves, O’Donnell and Finan reported that improved<br />

continence rates are achieved (29). However, Steiner et al. (30) in a<br />

larger series reported no difference in continence rates regardless of<br />

whether a nerve-sparing approach was used or not. They suggested


Chapter 11 / Post-Prostatectomy Incontinence 253<br />

“anatomic factors,” such as improved visualization and preservation of<br />

the urethra at the prostatic apex are responsible for improved continence.<br />

Bladder neck preservation has been reported to promote continence by<br />

sparing circular fibers of the bladder neck and proximal urethra (31,32).<br />

Gomez et al. added that bladder neck preservation can be performed<br />

with no increased risk of isolated positive margin at the bladder neck<br />

(33). To date, there are no controlled urodynamic studies to verify that<br />

bladder neck preservation promotes continence. Licht et al. (34) noted<br />

that bladder neck preservation had no effect on continence postoperatively,<br />

and it appears that meticulous closure of the bladder neck around<br />

the catheter has no influence on postoperative continence (30). Bladder<br />

neck preservation does appear to reduce the incidence of anastomotic<br />

strictures.<br />

At present, considerable interest lies in the preservation of urethral<br />

length and integrity as the major mechanism promoting urinary continence<br />

after radical prostatectomy. Urodynamic data has revealed that<br />

continent patients have significantly longer functional urethral length<br />

than incontinent patients (35). Myers (22) described the use of a<br />

vein retractor to elevate the prostatic apex in order to achieve the<br />

highest possible transection of the membranous urethra to preserve<br />

urethral length. Walsh et al. (36) described incorporating tissue of the<br />

dorsal venous complex in the vesicourethral anastomosis, promoting an<br />

earlier return to continence. Klein (37) and Stamey (38) have described<br />

modified apical dissections incorporating the tissue posterior to the<br />

urethra into the vesicourethral anastornosis. Klein (39) describes a<br />

modification of technique allowing the urethra to remain attached to<br />

its posterior fascia (Fig. 3). He reported improved continence rates and<br />

an earlier return to continence. These modifications not only stress<br />

meticulous preservation of urethral length, but stabilize the urethra to<br />

the periurethral supporting structures which assist in maintaining continence.<br />

Rudicul Perineul Prostutectomy<br />

The perineal approach to radical prostatectomy has been described<br />

as advantageous, because the urethra is better visualized facilitating<br />

the vesicourethral anastomosis (40). Harris and Thompson (41) noted<br />

that an “anatomic” approach to perineal prostatectomy achieved conti-<br />

nence rates of 97.5% after 18 MO. Several important points were stressed


254 Scarper0 and<br />

Fig. 3. (A-E) Steps in apical dissection of urethra. (Note: Preservation of maximal<br />

urethral length by blunt mobilization of prostatic apex off urethra [A] division<br />

of Denonvillier’s fascia with the urethra attached [D], and incorporation of Denonvillier’s<br />

fascia into anastomotic sutures [E].) (Adapted with permission from<br />

ref. 39.)<br />

to achieve these excellent results: careful dissection of the striated<br />

urethral sphincter, transection of the urethra at the prostatic apex, and<br />

bladder neck preservation. Isolated positive margins at the bladder neck<br />

were encountered, and the authors stressed proper patient selection for<br />

bladder neck preservation.


Chapter 11 I Post-Prostatectomy Incontinence 255<br />

Table 1<br />

Etiology of Postprostatectomy Incontinence Based on Urodynamic Findings<br />

~-<br />

No. of % Bladder % Sphincteric<br />

patients <strong>dysfunction</strong> incompetence % Combined<br />

Khan (42) et 63 al. 49.2 22.2 25.4<br />

Goluboff et (43) al. 56 61 5 34<br />

Chao and (44) 74 Mayo 4 57 39<br />

Winters (45) et 65 al. 1.5 70.8 27.7<br />

Fitispatrick (46) et 68 al. 16.2 26.5 38.2<br />

Yalla et al. (48) 21 38 62 l0<br />

ETIOLOGY OF INCONTINENCE<br />

FOLLOWING PROSTATECTOMY<br />

When considering the etiology of urinary incontinence, it is essential<br />

to identify that leakage may occur as a result of an abnormality of<br />

bladder and/or sphincteric function. Urodynamic studies provide information<br />

concerning the relative contributions of bladder andor sphincteric<br />

<strong>dysfunction</strong> in patients with incontinence after prostatectomy.<br />

The presence of isolated bladder <strong>dysfunction</strong> as the predominant<br />

cause of incontinence following prostatectomy has been identified in<br />

reports utilizing urodynarnic studies (42,43). These reports concluded<br />

that sphincteric weakness is not the major factor contributing to incontinence<br />

following prostatectomy. In contrast, sphincteric weakness alone<br />

has been implicated as the predominant cause of postprostatectomy<br />

incontinence (444.5) (see Table 1 >. Several factors have been proposed<br />

to explain for this discrepancy (45). Many studies documenting a high<br />

incidence of bladder <strong>dysfunction</strong> alone contain a large percentage of<br />

patients following transurethral resection of the prostate (43,46,47).<br />

Bladder <strong>dysfunction</strong> has been found preoperatively in 45% of patients<br />

with symptomatic bladder outlet obstruction, and 38% of patients have<br />

persistent detrusor instability following prostatectomy (48). A high<br />

incidence of preoperative bladder <strong>dysfunction</strong> likely effected the results<br />

in many series identifying bladder <strong>dysfunction</strong> as the predominant cause<br />

of postprostatectomy incontinence. These findings were confirmed as<br />

Winters et al. noted a higher incidence of combined bladder and<br />

sphincter <strong>dysfunction</strong> in patients following TUN? compared to patients


256 Scarper0 and<br />

following radical prostatectomy in whom isolated sphincter <strong>dysfunction</strong><br />

was the predominant cause (45). Detrusor instability owing to decreased<br />

compliance after prostatectomy has been documented to improve within<br />

1 year in a prospective evaluation in men (49). Thus, in groups of<br />

patients studied early after prostatectomy, the findings of bladder<br />

<strong>dysfunction</strong> may improve with time. This may explain the high incidence<br />

of bladder <strong>dysfunction</strong> in a recent review in which the mean<br />

time between prostatectomy and urodynamic study was only 19 mo<br />

(50). The principal importance of competent sphincteric function was<br />

identified by Hammerer and Huland, who noted statistically significant<br />

differences in maximal urethral closure pressure and functional urethral<br />

length between continent and incontinent patients without changes in<br />

bladder function (51). Valsalva leak-point pressures (VLPP) have not<br />

been identified to correlate with incontinence severity (45,52). Thus<br />

there appears to be no role utilizing the absolute value of VLPP in<br />

application of therapy for incontinence following prostatectomy.<br />

In summary, in patients following radical prostatectomy undergoing<br />

urodynamic studies 2 l year postoperatively, isolated sphincteric<br />

<strong>dysfunction</strong> is the most likely cause. A significant number of these<br />

patients will also have concomitant bladder <strong>dysfunction</strong>. In patients<br />

following TURP, a higher incidence of bladder <strong>dysfunction</strong> is present,<br />

and is likely to be the predominant cause if urodynamics are done less<br />

than 1 year postoperatively. It has been documented that symptoms do<br />

not accurately predict the diagnosis (49); therefore, urodynarnic studies<br />

are impo~ant-particularly if considering surgical treatment.<br />

Anastomotic strictures are a possible factor in the cause of incontinence<br />

following prostatectomy. Strictures may result in incomplete<br />

bladder emptying with resultant overflow incontinence or contribute<br />

to sphincteric weakness. Scarring of the anastomotic region may extend<br />

down into the urethral sphincteric mechanism and impair urethral closure.<br />

Dilatation andlor incision of urethral strictures are often followed<br />

by sphincteric weakness.<br />

As with all clinical situations, the evaluation of postprostatectomy<br />

incontinence begins with a complete history and physical examination.<br />

The history should include detailed information about the urinary leak-<br />

age. When did the leakage occur after surgery, and was any leakage


Chapter 11 / Post-Prostatectomy Incontinence 257<br />

present prior to surgery? The events occurring during leakage (activity,<br />

urgency or no sensation) as well as associated urinary symptoms (force<br />

of stream, complete emptying) are important factors. Day and nighttime<br />

pad usage should be documented to measure the severity of leakage.<br />

A pad-weight test may also more accurately quantify urine leakage. Any<br />

associated neurologic symptoms or history of previous incontinence<br />

therapy should be obtained. The physical examination should include<br />

focus on the neurologic status of the perineum and lower extremities.<br />

Rectal examination should be performed to rule out recurrence in cancer<br />

cases, and to assess the size of the prostate in patients following TURP.<br />

A urinalysis and residual urine should be obtained in all patients, and<br />

PSA testing should be considered in all patients with a history of<br />

prostate cancer to rule out recurrence.<br />

Cystourethroscopy should be performed to inspect for urethral strictures<br />

or bladder neck contractures. Gross inspection of the distal urethral<br />

sphincter may reveal abnormalities, but cystoscopic examination does not<br />

determine sphincter function. In cases of incontinence following TURP,<br />

distortion or absence of the verumontanum suggests sphincteric injury<br />

and the presence of residual adenoma can be confirmed. Bladder pathology<br />

(tumors, stones, or diverticuli) can be detected. An important feature<br />

in performing cystoscopy is to assess the length of the urethra between<br />

the external sphincter and the bladder neck. If the urethrovesical anastomosis<br />

is at the level of the external sphincter without sufficient urethral<br />

length, a transurethral route for injection therapy may be difficult (53).<br />

Radiologic evaluation may consist of retrograde urethrography<br />

(RUG) or voiding cystourethrography (VCUG). These studies are capable<br />

of identifying anatomic causes of incontinence (stricture or bladder<br />

neck contracture). The bladder neck is visualized during VCUG and<br />

should remain closed during filling. Straining films may detect leakage<br />

of urine across the bladder neck. The VCUG also permits the identification<br />

of bladder trabeculation, diverticuli, vesicoureteral reflux, or residual<br />

urine. As with cystoscopy, these studies offer anatomic information<br />

alone and do not assess bladder or sphincteric function. Perhaps the<br />

greatest utility of VCUG is in combination with urodynamic studies,<br />

videourodynamics.<br />

Urodynamic evaluation is essential to determine the etiology of<br />

incontinence following prostatectomy, and should be performed in all<br />

patients for whom invasive therapy is considered or in patients who<br />

fail conservative treatment. The appropriate urodynamic investigation<br />

should allow determination of bladder and urethral function during<br />

filling as well as assess bladder contractility and the presence of


258 Scarper0 and Winters<br />

obstruction during voiding. To delineate these functions, a multichannel<br />

pressure flow study is the most appropriate test. This study requires<br />

the presence of a rectal catheter to allow determination of true detrusor<br />

pressure, and the presence of simuultaneous uroflow to measure flow<br />

velocity during voiding. The use of electromyography of the pelvic floor<br />

in indicated in cases complicated by potential neurologic <strong>dysfunction</strong>. In<br />

routine cases it will add little to the diagnosis of postprostatectomy<br />

incontinence. Urethral pressure profilometry will provide infomation<br />

concerning urethral closing pressure and functional urethral length.<br />

There are no established normal values in male patients and the shape<br />

and magnitude of the urethral pressure profile varies greatly with indi-<br />

vidual techniques (54). Videourodynamics provides the most complete<br />

evaluation, combining the anatomic detail of VCUG with the functional<br />

assessment of pressure-flow studies.<br />

During the filling phase of the pressure-flow study, information<br />

concerning detrusor compliance or instability as well as cystometric<br />

capacity is obtained. At 200, 250, and 300 mL volume, the patient is<br />

asked to strain. If urinary leakage occurs during straining in the absence<br />

of a rise in true detrusor pressure, sphincteric weakness is documented.<br />

As stated previously, the exact value of the VLPP is less important in<br />

males as is making the diagnosis of stress incontinence. The Jilling<br />

phase of the urodynamic study is critical, as one should determine the<br />

whether the bladder is stable with normal storage pressures and that<br />

the sphincter is competent during straining maneuvers. During the<br />

voiding phase, information regarding bladder contractility or the pres-<br />

ence of bladder outlet obstruction is present. As in cases of outlet<br />

obstruction related to benign prostatic hyperplasia (BP€€), the voiding<br />

phase of the pressure flow study will distinguish between obstruction<br />

(high detrusor pressure, low urinary flow) and impaired contractility<br />

(low detrusor pressure, low urinary flow). The authors routinely perform<br />

pressure-flow studies followed by flexible cystoscopy in the evaluation<br />

of incontinent patients following prostatectomy.<br />

T~TMENT OF<br />

POSTPROSTATECTOMY INCONTINENCE<br />

After performance of urodynamic testing to determine the exact<br />

etiology of postprostatectomy incontinence, appropriate treatment decisions<br />

can be made. The following treatment recommendations are based<br />

on these urodynamic findings.


Chapter 11 / Post-Prostatectomy Incontinence 259<br />

Sphincteric Incompetence<br />

CONSERVATIVE TREATMENT<br />

Pelvic-floor muscle exercises called Kegel exercises can speed the<br />

recovery of continence following prostatectomy; therefore, patients<br />

should be instructed and perform the exercises before prostatectomy<br />

(55). There are several studies which suggest the utility of electrical<br />

stimulation (56), behavior training (57), and biofeedback (58) in the<br />

management of incontinence after prostate surgery. Unfortunately,<br />

many of these studies contain only a small number of patients, and<br />

many patients still experience wetting requiring pad use even though<br />

considered “improved.” These conservative measures offer little prom-<br />

ise in the correction of severe or total urinary incontinence.<br />

PHARMACOLOGIC TREATMENT<br />

Alpha-adrenergic agents (ephedrine, pseudoephedrine, and phenylpropranolamine)<br />

may improve leakage by increasing outlet resistance.<br />

These drugs have a very limited effectiveness, and should be utilized<br />

mainly in mild cases in combination with timed voiding and behavioral<br />

therapy. Imipramine (10-25 mg tid) possesses anticholinergic activity<br />

and direct smooth muscle inhibition of the bladder. Imipramine also<br />

blocks reuptake of norepinephrine thus theoretically increasing outlet<br />

resistance. The efficacy of this drug is probably as a result of its effect<br />

on the bladder, but its use should be considered in cases of sphincteric<br />

incompetence (59).<br />

INJECTION THERAPY<br />

The ideal material for periurethral injection is one that is easily<br />

injected, biocompatible, and causes little or no inflammatory reaction.<br />

There should be no migration of the injected material, and it should<br />

maintain its bulking effect for a long period of time. Polytetrafluoroeth-<br />

ylene (Teflon) paste was one of the first agents utilized in the treatment<br />

of male incontinence. Politano reported on ”70 males treated with<br />

Teflon injections, and found that 88% of patients post-TURP and 67%<br />

of patients following radical prostatectomy were cured or improved<br />

(60). However, other authors have not duplicated these results, and<br />

documented concerns of granuloma formation and particle migration<br />

have precluded the use of Teflon as an injectable agent (61).


260 Scarper0 and<br />

Table 2<br />

Efficacy of Collagen in the Treatment of Postprostatectomy Incontinence<br />

Shortliffe et al. (63) 16 3<br />

McGuire and Appell (64) 134 22<br />

Aboseif et al. (65) 88 42<br />

Herschorn et al. (66) 10 2<br />

Cumrnings et al. (67) 19 4<br />

Kageyama et al. (68) 10 1<br />

Griebling et al. (69) 25 2<br />

Smith et al. (70) 57 2<br />

Improvedl M<br />

No. of Social follow-up<br />

Cases Dry continence Failed (mo)<br />

5<br />

70<br />

33<br />

5<br />

7<br />

5<br />

8<br />

20<br />

7<br />

42<br />

13<br />

3<br />

8<br />

4<br />

15<br />

35<br />

9-23<br />

> 12<br />

10<br />

5 .7<br />

10.4<br />

NA<br />

13.3<br />

9.4<br />

To date, the most commonly utilized injectable agent is collagen.<br />

Glutaraldehyde cross-linked bovine collagen (Contigen, CR Bard Co.,<br />

Covington, GA) is both biocompatible and biodegradable. No inflam-<br />

matory reaction or granuloma formation is elicited, and no cases of<br />

particle migration are identified (62). Shortliffe et al. were the first to<br />

utilize collagen for the treatment of urinary incontinence (63). In 16<br />

male patients, 50% were either dry or improved. Results of initial<br />

studies utilizing transurethral injections of collagen in the treatment of<br />

postprostatectomy incontinence were encouraging. In a multi-center<br />

trial of 134 patients, 52% had significantly improved and 16% were<br />

dry at l-yr follow-up, and of 60 patients, 47% were significantly<br />

improved with 25% dry at 2-yr follow-up (64). Aboseif et al. reported<br />

on 88 patients with a 10-1110 follow-up. Forty-eight percent were dry<br />

and another 38% were improved (65). In this study it is important to<br />

note that patients required up to five injections, and some patients did<br />

not improve until after the fourth injection. Herschorn et al. reported<br />

less encouraging results with a 20% dry rate at 5.7 mo follow-up (66),<br />

and many reports that followed were not able to reproduce the earlier<br />

promising reports (67-70) (see Table 2). Smith et al. achieved social<br />

continence in only 38.7% of patients and documented poor results in<br />

patients with continuous supine leakage, those wearing penile clamps,<br />

patients with a bladder neck defect greater than l cm and patients<br />

following radiation therapy. The authors recommended these patients<br />

not undergo collagen-injection therapy (70). In an effort to improve<br />

the results of collagen implantation in men, Appell et al. (71) and


Chapter l1 I Post-Prostatectomy Incontinence 261<br />

Klutke et al. (72) introduced “antegrade” placement of collagen through<br />

a suprapubic tract. The rationale is to inject collagen into supple, well-<br />

vascularized tissue above the scar of the vesicourethral anastomosis to<br />

improve results. Utilizing this technique, Klutke et al. achieved dryness<br />

in 25% of patients and 45% were significantly improved (73). Vasavada<br />

and Appell (74) presented long-term follow-up in 29 men treated with<br />

the antegrade collagen injection technique. The initial (2 wk) dry rate<br />

was 76%. The dry rate decreased to 66% at 6 m0 and 24% at 1 yr.<br />

These authors concluded that early success is not predictive of a long-<br />

term outcome. From these data, it is apparent that collagen injection<br />

therapy is not the ideal treatment of incontinence following prostatec-<br />

tomy. Patients selected should be those with mild to moderate stress<br />

incontinence, and they should be counseled that multiple injections<br />

will be needed. It is important to stress that social continence is a more<br />

realistic goal than total dryness (70).<br />

~RTI~ICI~L ~ R I ~ SPHINCTER<br />

~ R Y<br />

To date, the AMS 800 artificial urinary sphincter (American Medical<br />

Systems, Minnetonka, MN) is the definitive procedure of choice in<br />

patients with moderate to severe sphincteric incompetence after prostatectomy.<br />

It is imperative that patients undergo urodynamic studies<br />

preoperatively to insure that the bladder compliance is normal and<br />

that there are no involuntary bladder contractions. In males following<br />

prostatectomy, the sphincter cuff is placed around the proximal bulbous<br />

urethra via a midline perineal incision. The bulbospongiosus muscle<br />

is largely dissected off of the urethra (Fig. 4), and the measuring device<br />

is placed around the urethra to determine the appropriate cuff size (Fig.<br />

5). Most frequently a 4.5-cm cuff is employed. Rarely are 4.0-cm cuffs<br />

utilized except in cases of tissue atrophy. Once the cuff is placed, the<br />

pressure reservoir is placed by incising the rectus fascia through a small<br />

suprapubic incision. The rectus muscle is then split bluntly, and the<br />

reservoir is placed in a subrectus pouch. Routinely a 61-70-cm reservoir<br />

is used. If the patient has poor-quality tissues or a history of radiation<br />

therapy, a 5 I -60-cm reservoir is used. The pump mechanism is placed<br />

in a dependent position in the scrotum through the suprapubic incision<br />

(Fig. 6). The sphincter is filled with isotonic contrast medium to allow<br />

radiologic identification of possible leaks. The mechanism is left deacti-<br />

vated for 6 wk. to prevent erosion.<br />

Long-term success rates and patient satisfaction are well-documented<br />

following i~plantation of artificial sphincters. Each patient should be


262 Scarper0 and<br />

Fig. 4. Perineal approach to bulbar urethra for placement of artificial urinary<br />

sphincter.<br />

counseled that some patients will experience “minor leakage” (social<br />

continence), and that there is a significant risk of reoperation. In a<br />

questionnaire-based study, Litwiller et al. revealed that 90% of patients<br />

reported satisfaction with the artificial sphincter and would have the<br />

procedure done again after a mean follow-up of 2 yr (7.5). An interesting<br />

finding was that the need for sphincter revision did not effect the rate<br />

of patient satisfaction adversely. Montague reviewed the results of 156<br />

men implanted with a mean follow-up of 41.6 mo (76). Total continence<br />

was achieved in 75% of patients, and 15% were improved. The revision<br />

rate in this series was 19.3%. Haab et al. (77) confirmed excellent<br />

patient satisfaction and results, as 54/68 men were socially continent<br />

with significant reduction in pad usage after 3.5 yr follow-up. These<br />

authors noted an overall revision rate of 25%. They cited a reduction<br />

in the revision rate from 44.4 to 12.4% after the introduction of sphincter<br />

modifications in 1987 (77). In perhaps the largest series, Elliot and<br />

Barrett (78) reviewed the revision rate of 313 men following artificial<br />

sphincter implantation with a mean follow-up of 68.8 rno. Forty-two<br />

percent of patients implanted before modifications in the artificial<br />

sphincter required revision surgery, but after sphincter modification,


Chapter 11 / Post-Prostatectomy Incontinence 263<br />

Fig. 5. The use of measuring “tape” to determine appropriate cuff size. The most<br />

common size cuff implanted is 4.5 cm. A 4.0-cm cuff is employed in this particular<br />

case of tissue atrophy.<br />

only 17% needed revision. It is very well-documented that following<br />

modifications in sphincter design, the revision rate is approx 20%.<br />

The most common causes of device failure are fluid leak or urethral<br />

tissue atrophy. Both problems create a decrease in closing pressure of<br />

the cuff. Alteration in pumping mechanics may signify device malfunc-<br />

tion. When a patient experiences a leak in the device, the number of<br />

pumps may decrease, or the patient may not be able to pump the device<br />

at all. If contrast was placed in the balloon reservoir, a X-ray may<br />

confirm loss of fluid in the reservoir. It is important to remember that<br />

small leaks may be present with a normal appearing reservoir on X-<br />

ray. The cuff is the most common site of a leak followed by the balloon<br />

(79). The site of the leak is determined at surgical exploration, and the<br />

device should be replaced. Occasionally the patient may describe an<br />

increasing number of pumps to empty the cuff. This may signify urethral<br />

tissue atrophy as increased fluid is allowed into the cuff. Atrophy should<br />

be suspected in any patient who has a normally functioning device that<br />

starts leaking 2 6 mo after implantation. When this problem is con-<br />

firmed, the simplest solution is to decrease the cuff size to a smaller


264 Scarper0 and<br />

Fig. 6. Schematic representation of artificial urinary sphincter implanted in a male.<br />

(Courtesy of American Medical Systems.)<br />

size. Other options include proximal repositioning of a new cuff (80)<br />

or a double-cuff technique, which requires implanting a second cuff<br />

around the bulbar urethra (81). Cuff erosion is fortunately much less<br />

of a problem now that delayed activation is the rule (82).<br />

The most precise way to diagnose erosion is with urethroscopy.<br />

Patients may present urinary tract infection (UTI), perineal pain, irritative<br />

voiding symptoms, or pain and induration at the pump. These<br />

findings wmant consideration of possible erosion. Almost all cases<br />

of erosion should be managed by removal of the entire prosthesis.<br />

Occasionally, if identified early and associated with sterile urine, only<br />

the cuff may need to be removed. The authors recommend this practice<br />

with great caution, as in our experience we have never been able to<br />

successfully salvage a sphincter in these circumstances. A repeat sphincter<br />

may be implanted after 4-6 mo. It is imperative that patients be<br />

instructed to carry medical identification cards andlor bracelets as one<br />

of the most common causes of erosion is instr~mentation of the urethra


Chapter 11 / Post-Prostatectomy Incontinence 265<br />

with the sphincter activated. Many patients assume that all hospital<br />

staff are aware of artificial sphincter function. In reality, we have all<br />

encountered patients who develop problems after instrumentation that<br />

probably could have been avoided.<br />

A difficult clinical problem is the management of simultaneous<br />

urethral strictures or bladder neck contractures and stress urinary incon-<br />

tinence. Mark et al. (83) reported 26 patients with stress incontinence<br />

and urethrovesical stricture no longer than 2.0 cm. They found synchro-<br />

nous incision of the stricture and insertion of an artificial urinary sphinc-<br />

ter to be safe and provide excellent results. Haab et al. recommend<br />

incision of the stricture followed by 6 wk of patient self-catheterization.<br />

The sphincter is implanted 6 wk after stricture incision, and the patient<br />

may periodically continue self-catheterization after sphincter implanta-<br />

tion (24). The implantation of an artificial sphincter following radiation<br />

therapy is controversial. With the use of a low-pressure (51-60 cm)<br />

balloon reservoir, acceptable success rates can be achieved with a<br />

higher revision rate (84,85). In these patients it is advisable to leave<br />

the sphincter deactivated for 12 wk after implantation.<br />

MALE SUBURETHRAL SLING<br />

The use of pubovaginal slings in the management of female stress<br />

urinary incontinence (SUI) is well-established. The use of slings in male<br />

patients has been previously described (86). More recently, Schaeffer et<br />

al. (87) introduced a bulbourethral sling procedure utilizing Teflon<br />

bolsters achieving a 75% success rate-some patients required “retight-<br />

ening” procedures. In this series the revision rate was 27% with a 9%<br />

incidence of erosion or infection. These authors concluded that this<br />

procedure is effective in the management of postprostatectomy inconti-<br />

nence. At present we recommend this procedure only in experimental<br />

trials. More experience and follow-up with this procedure is needed,<br />

as the potential to create significant obstruction and/or irritative voiding<br />

symptoms exists.<br />

Bludder Dysfinction<br />

CONSERVATIVE TREATMENT<br />

As in all clinical presentations of detrusor overactivity, the mainstay<br />

of therapy is timed voiding techniques, avoidance of dietary irritants<br />

and fluid restriction when appropriate.


266 Scarper0 and<br />

PHARMACOLOGIC THERAPY<br />

Anticholinergic therapy may be initiated with oxybutinin chloride<br />

(5 mg tid) or probantheline bromide (15 mg tid). Hyoscyamine may<br />

be given in as a sustained release tablet or in a sublingual immediate<br />

release preparation. For refractory cases of incontinence or in patients<br />

with a marked reduction in bladder capacity, imipramine (25 mg tid)<br />

may be added to either of these anticholinergic medications. Patients<br />

should be warned of anticholinergic side effects, which can be mini-<br />

mized by starting these medications at a low dose and increasing the<br />

dosage periodically. The addition of new anticholinergic formulations<br />

(tolterodine, ditropan XL) may minimize these side effects, which limit<br />

therapy in many patients.<br />

SURGICAL THERAPY<br />

When aggressive therapy does not control incontinence, repeat uro-<br />

dynamic investigation should be undertaken to document persistent<br />

bladder <strong>dysfunction</strong> and to evaluate sphincter function. When aggres-<br />

sive anticholinergic therapy does not control bladder overactivity, blad-<br />

der augmentation should be considered. The various procedures for<br />

bladder augmentation are discussed in detail in chapter 14.<br />

Combined Bladder Dysfinction and Sphincteric Incompetence<br />

The treatment of this patient group with postprostatectomy incontinence<br />

is most challenging. The basic approach to these patients involves<br />

initial behavioral and pharmacological attempts to control the detrusor<br />

<strong>dysfunction</strong>. Once the bladder storage function normalizes, the sphincteric<br />

incompetence can be corrected. This usually involves repeating<br />

the urodynamic study while the patient remains on anticholinergic<br />

therapy. This differentiates the cause of persistent wetting-either<br />

sphincteric incompetence alone or refractory bladder <strong>dysfunction</strong>. If<br />

isolated sphincteric incompetence is identified, the patient proceeds to<br />

correction of the sphincteric incompetence and remains on anticholinergic<br />

therapy.<br />

The treatment of the patient with refractory bladder <strong>dysfunction</strong> and<br />

sphincteric incompetence is controversial. Leach et al. (50) concluded<br />

that it is essential to control bladder <strong>dysfunction</strong> prior to insertion of<br />

an artificial sphincter, as this will minimize persistent leakage, damage<br />

to the upper tracts or possible worsening of bladder <strong>dysfunction</strong>. In


Chapter 1 I / Post-Prostatectomy Incontinence 267<br />

this review of 215 patients, they did not encounter patients whose<br />

bladder <strong>dysfunction</strong> was refractory to medical therapy. The safety of<br />

simultaneous implantation of an artificial sphincter and entero-<br />

cystoplasty (SS) indicates that this form of management may be applica-<br />

ble in this group of patients. These patients should be forewarned about<br />

the possible need for self-catheterization following surgery and the risk<br />

of infection. Perez and Webster (89) noted that there has never been<br />

a case report of upper tract deterioration in postprostatectorny men with<br />

a hyperactive detrusor following implantation of an artificial urinary<br />

sphincter. In patients with detrusor hyperactivity, acceptable rates of<br />

continence were achieved when compared to men with stable detrusor<br />

function after sphincter implantation (89). The authors concluded that<br />

men with postprostatectomy urinary incontinence should not be<br />

excluded from consideration from artificial urinary sphincter implanta-<br />

tion based on the failure to meet ideal implantation characteristics.<br />

Only a controlled, prospective comparison of these patient groups will<br />

resolve this issue.<br />

TREATMENT OF POSTPROSTATECTOMY<br />

INCONTINENCE: A COST COMPARISON<br />

Stothers et al. (90) reported on the results of 20 patients treated<br />

surgically for incontinence following prostatectomy. Ten patients<br />

underwent transurethral collagen injections and 10 patients underwent<br />

sphincter implantation. Only one collagen patient achieved complete<br />

dryness, and three collagen patients ultimately had a sphincter<br />

implanted. These authors concluded that an artificial sphincter is a more<br />

effective solution than collagen with a higher incidence of patient<br />

dryness and long-term success. Brown et al. (91) after reviewing the<br />

charges of hospital procedures and undergarments concluded that the<br />

cost of sphincter implantation compared favorably to the cost of colla-<br />

gen injections (general anesthesia) in patients requiring three injections.<br />

In patients with severe leakage (>9 undergarmentdd) the artificial<br />

sphincter is more cost-effective than continued pad usage.<br />

SUMMARY<br />

Postprostatectomy incontinence can adversely effect one’ S quality<br />

of life. Isolated sphincteric incompetence is the most common etiology


268 Scarper0 and<br />

following radical prostatectomy and combined bladder <strong>dysfunction</strong><br />

and sphincteric incompetence is the most common etiology following<br />

TURP. Urodynamic studies are important in obtaining a precise diagnosis.<br />

Transurethral injection of collagen is moderately effective and can<br />

be utilized in cases of mild to moderate stress incontinence. The artificial<br />

urinary sphincter remains the most definitive treatment for sphincteric<br />

incompetence. In cases of combined bladder <strong>dysfunction</strong> and sphincteric<br />

incompetence, anticholinergic therapy should be given in an<br />

attempt to normalize bladder pressure prior to correction of sphincter<br />

deficiency. A thorough discussion of all the treatment options with<br />

patients is essential and allows the patient to actively participate in the<br />

treatment process.<br />

REiFEREiNCES<br />

1. Boring C, Squires T, Tong T (1993) Cancer Statistics. CA Cuncer J Clin 43:7-26.<br />

2. Middleton R, Thompson I, Austenfeld M, et al. (1995) Prostate cancer clinical<br />

guidelines Panel summary report on the management of clinically localized prostate<br />

cancer. J Urol 154:2144-2148.<br />

3. Marks J, Light JK (1989) Management of urinary incontinence after prostatectomy<br />

with the artificial urinary sphincter. J Urol 142:302-304.<br />

4. Mebust W, Holtgrewe H (1989) Current status of transurethral prostatectomy: a<br />

review of the AUA National Cooperative Study. World J Urol 6: 194-1 99.<br />

5. Walsh P, Jewett A (1980) Radical Surgery for prostatic cancer. Cuncer 45:1906.<br />

6. Rudy D, Woodside J, Jeffry R, et al. (l 984) Urodynamic evaluation of incontinence<br />

in patients undergoing modified Campbell radical retropubic prostatectomy: a<br />

prospective study. J Urol 132:708-712.<br />

7. Fowler F, Barry M, Lu-yao G, et al. (1995) Effect of radical prostatectomy<br />

for prostate cancer on patient quality of life: Results from a Medicare survey.<br />

Urology 45 : 1007.<br />

8. Herr H (1994) Quality of life of incontinent men after radical prostatectomy.<br />

J Urol 151:652-654.<br />

9. Litwin M, Hays R, Fink A, Ganz P, Leake B, Leach G, Brook R (1995) Qualityof-life<br />

outcomes in men treated for localized prostate cancer. JAMA 273: 129-135.<br />

10. O’Donnell P, Finan B, Barnett T, Brookover T (1990) Continence recovery following<br />

radical prostatectomy. Neurol Urodyn 9:251-256.<br />

1 1. Hammerer P, Huland H (1997) Urodynamic evaluation of changes in urinary<br />

control after radical retropubic prostatectomy. J Urol 157:233-236.<br />

12. Geary E, Dendinger T, Freiha F, Stamey T (1 995) Incontinence and vesical neck<br />

strictures following nerve sparing radical retropubic prostatectomy. Urology<br />

45 : 1000.<br />

13. McNeal JE (1 968) Regional morphology and pathology of the prostate. Am J<br />

Surg Pathol49:347-357.<br />

14. McNeal JE (1978) Origin and evolution of benign prostatic enlargement. Invest<br />

Urol 15:340-345.


Chapter 11 / Post-Prostatectomy Incontinence 269<br />

15. McNeal J, Redwine E, Freiha F, Stamey T (1 988) Zonal distribution of the prostatic<br />

adenocarcinoma: correlation with histologic patterns and direction of spread. Am<br />

J Surg Pathol 12:897-906.<br />

16. Turner Warwick R (1983) The sphincter mechanisms: their relationship to prostatic<br />

enlargement and its treatment. In: Hinman F Jr, ed., Benign Prostatic Hypertrophy.<br />

New York: Springer-Verlag, p. 809.<br />

17. Caine N, Raz S, Ziegler M (1975) Adrenergic and cholinergic receptors in the<br />

human prostate, prostatic capsule and bladder neck. Br J Urol 47:193-202.<br />

18. Eilbadawi A (1996) Functional anatomy of the organs of micturition. Urol Clin<br />

N Am 23:177-210.<br />

19. Oelrich TM (1 980) The urethral sphincter muscle in the male. Am JAnat 158:229.<br />

20. Burnett A, Mostwin J (1998) In situ anatomical study of the male urethral sphincteric<br />

complex: relevance to continence preservation following major pelvic surgery.<br />

J Urol 160:1301-1306.<br />

21. Steiner MS (1994) The puboprostatic ligament and the male urethral suspensory<br />

mechanism: an anatomic study. Urology 44:530-534.<br />

22. Myers RP (1991) Male urethral sphincteric anatomy and radical prostatectomy.<br />

Urol Clin N Am m21 1-227.<br />

23. Myers R, Goellner J, Cahill D (1987) Prostate shape, external striated urethral<br />

sphincter and radical prostatectomy. J Urol 138543-550.<br />

24. Haab F, Yamaguchi R, Leach G (1996) Postprostatectomy incontinence. Urol<br />

Clin N Am 23:447-457.<br />

25. McNeal J (1972) The prostate and prostatic urethra-a morphologic synthesis.<br />

J Urol 107:1008-1016.<br />

26. Shah P Jr, Abrarns P, Fineley R, et al. (1979) The influence of prostatic anatomy<br />

and the different results of prostatectomy according to the surgical approach. Br<br />

J Urol 5 1549-55 1.<br />

27. Levy J, Seay T, Wein A (1996) Postprostatectomy incontinence. AUA Update<br />

Series, Lesson 8, Vol XV, 58-66.<br />

28. Walsh P, Donker P (1982) Impotence following radical prostatectomy: insight<br />

into etiology and prevention. J Urol 128:492-497.<br />

29. O’Donnell P, Finan B (1989) Continence following nerve-sparing radical prostatectomy.<br />

J Urd 142:1227-1228.<br />

30. Steiner M, Morton R, Walsh P (1991) Impact of anatomical radical prostatectomy<br />

on urinary continence. J Urol 145:s 12-5 15.<br />

31. Donat SM, Fair WR (1996) Radical Retropubic Prostatectomy: bladder neck<br />

preservation. AUA Update Series, Lesson 7, Vol XV, 50-57.<br />

32. Malizia A, Banks D, Newton E, et al. (1989) Modified radical retropubic prostatectomy:<br />

double continence technique. J UroE 141:316A4.<br />

33. Gomez C, Soloway M, Civantos F, et al. (1993) Bladder neck preservation and<br />

its impact on positive surgical margins during radical prostatectomy. Adult<br />

Urol 42:689-694.<br />

34. Licht M, Klein E, Tuason I, Levin H (1994) Impact of bladder neck preservation<br />

during radical prostatectomy on continence and cancer control. Urology 44:883-<br />

887.<br />

35. Presti J Jr, Schmidt R, Narayan P, Carroll P, Tanagho E (1990) Pathophysiology<br />

of urinary incontinence after radical prostatectomy. J UroE 143:975-978.<br />

36. Walsh P, Quinlan D, Morton R, Steiner M (1990) Radical Retropubic prostatectomy.<br />

Improved anastomosis and urinary continence. Urol Clin NAm 17:679-684.


270 Scarper0 and Winters<br />

37. Klein EA (1992) Early continence after radical prostatectomy. J Urol 148:92-94.<br />

38. Stamey TA (1991) The Stanford radical retropubic prostatectomy. J Urol 145:49A<br />

(V-42).<br />

39. Klein EA (1993) Modified apical dissection for early continence after radical<br />

prostatectomy. Prostate 22:217-220.<br />

40. Carlin B, Resnick M (1995) Anatomic approach radical to perineal prostatectomy.<br />

Urol Clin N Am 22:461-473.<br />

41. Harris M, Thompson I (1996) The anatomic radical perineal prostatectomy: a<br />

contemporary and anatomic approach. Urology 48:762-768.<br />

42. Khan Z, Mieza M, Starer P, Singh V (1991) Post-prostatectomy incontinence: a<br />

urodynamic and fluoroscopic point of view. Urology 38:483-488.<br />

43. Goluboff E, Chang D, Olsson C, Kaplan S (1995) Urodynamics and the etiology<br />

of post-prostatectomy urinary incontinence: the initial Columbia experience.<br />

J Urol 153:1034-1037.<br />

44. Chao R, Mayo M (1995) Incontinence after radical prostatectomy: Detrusor or<br />

sphincter causes. J Urol 154:16-18.<br />

45. Winters J, Appell R, RacMey R (1998) Urodynamic findings in postprostatectomy<br />

incontinence. Neurol Urodynam 17:493-498.<br />

46. Fitzpatrick J, Gardiner R, Worth P (1 979) The evaluation of 68 patients with<br />

postprostatectomy incontinence. Br J Urol 5 1:552-555.<br />

47. Yalla S, Karsh L, Kearney G, et al. (1982) Post-prostatectomy incontinence:<br />

urodynamic assessment. Neurol Urodynam 1 :77-8 1.<br />

48. Andersen JT (1976) Detrusor hyperreflexia in benign infravesical obstruction, a<br />

cystometric study. J Urol 115532-534.<br />

49. Foote J, Yun S, Leach G (1991) Post-prostatectomy incontinence: Pathophysiology,<br />

evaluation and management. Urol Clin N Am 18:229-241.<br />

50. Leach G, Trockman B, Wong A, et al. (l 996) Post-prostatectomy incontinence:<br />

urodynamic findings and treatment outcomes. J Urd 155:1256-1259.<br />

51. Hamerer P, Huland H (1997) Urodynannic evaluation of changes in urinary<br />

control after radical retropubic prostatectomy. J Urol 157:233-236.<br />

52. Walker C, Mason D, Joseph A, Juma S (1995) Does the leak point pressure predict<br />

the severity of urinary incontinence in male patients. J Urol 153:277A.<br />

53. Appell R (1 996) Personal communication.<br />

54. Winters J, Appell R (1997) Practical urodynamics. In: O’Donnell P, ed., Urinary<br />

Incontinence. St. Louis, MO: Mosby-Year Book, pp. 181-189.<br />

55. Moul J (1994) For incontinence after prostatectomy, tap a diversity of treatments.<br />

Contemp Urol 6:78-88.<br />

56. Sotiropoulos A, Yeaw S, Lattimer J (1976) Management of urinary incontinence<br />

with stimulation: observations and results. J Urol 116:747-750.<br />

57. Meaglia J, Joseph A, Chang M, et al. (1990) Postprostatectomy urinary incontinence:<br />

response to behavioral training. J Urd 144:674-676.<br />

58. Burgio K, Stutzman R, Engel B (1989) Behavioral training for postprostatectomy<br />

urinary incontinence. J Urol 141:303-306.<br />

59. Worth P (1984) Postprostatectomy incontinence. In: Mundy A, Stephenson T,<br />

Wein A, eds., Urodynamics. New York, NY: Churchill Livingstone, pp. 205-21 1.<br />

60. Politano V (1992) Transurethral Polytef for post-prostatectomy urinary incontinence.<br />

Br J Urol 69:26-28.<br />

61. Malizia A, Reiman H, Myers R, et al. (1984) Migration and granulomatous reaction<br />

after periurethral injection of polytef (teflon). JAMA 251:3277-3281.


Chapter 11 / Post-Prostatectomy Incontinence 271<br />

62. Remacle M, Marbaix E (1988) Collagen implants in the human larynx: pathologic<br />

dissemination of two cases. Arch Otolaryngol 245:203-209.<br />

63. Shortliffe L, Freiha F, Kessler R, Stamey T, Constantinou C (1989) Treatment<br />

of urinary incontinence by the periurethral implantation of glutaraldehyde crosslinked<br />

collagen. J Urol 141:538-541.<br />

64. Appell R, McGuire E, DeRidder P, Bennett A, Webster G, Badlani et G, al. (1994)<br />

Summary of effectiveness in the prospective, open, multicenter investigation of<br />

Contigen implant for incontinence due to intrinsic sphincteric deficiency in males.<br />

J Urd, part 2, 151:217A (abstract 174).<br />

65. Aboseif S, O’Connell H, Usui A, McGuire E (1996) Collagen injection for intrinsic<br />

sphincteric deficiency in men. J Urol 155:lO-13.<br />

66. Herschorn S, Radomslci S, Steele D (1992) Early experience with intraurethral<br />

collagen injections for urinary incontinence. J Urol 148: 1797-1 800.<br />

67. Cumings J, Boullier J, Parra R (1996) Transurethral collagen injections in the<br />

therapy of post-radical prostatectomy stress incontinence. J Urd 155:1011-1013.<br />

68. Kageyama S, Kawabe K, Suzuki K, Ushiyama T, Suzuki T, Aso Y (1 994) Collagen<br />

Implantation for post-prostatectomy incontinence: early experience with a transrectal<br />

ultrasonographically guided method. J Urol 152: 1473-1475.<br />

69. Griebling T, Kreder K, Williams R (1997) Transurethral collagen injection for<br />

treatment of postprostatectomy urinary incontinence in men. Urology 49:907-912.<br />

70. Smith D, Appell R, Rackley R, Winters J (1998) Collagen injection therapy in<br />

postprostatectomy incontinence. J Urol 160:364-367.<br />

71. Appell R, Vasavada S, Rackley R, Winters J (1996) Percutaneous antegrade<br />

collagen injection therapy for urinary incontinence following radical prostatectomy.<br />

Urology 48:769-772.<br />

72. Klutke C, Nadler R, Andriole G (1995) Antegrade collagen injection: New<br />

technique for postprostatectomy stress incontinence (Surgeon’ S Workshop)<br />

J Endourol 9:513-515.<br />

73. Klutke C, Nadler R, Tiemann D, Andriole G (1996) Early results with antegrade<br />

collagen injection for post-radical prostatectomy stress urinary incontinence.<br />

J Urol 156:1703-1706.<br />

74. Vasavada S, Appell R, Rackley R, Winters J (1997) Long term follow-up in men<br />

treated with antegrade collagen for post-prostatectomy incontinence. J Urol<br />

1573394A.<br />

75. Litwiller S, Kim K, Fone P, White R, Stone A (1996) Post-prostatectomy incontinence<br />

and the artificial urinary sphincter: a long term study of patient satisfaction<br />

and criteria for success. J Urol 156:1975-1980.<br />

76. Montague DK (1992) The artificial urinary sphincter (AS 800): experience in 166<br />

consecutive cases. J Urol 147:380-382.<br />

77 * Haab F, Trockman B, Zimern P, Leach G (1997) Quality of life and continence<br />

assessment of the artificial urinary sphincter in men with minimum 3.5 years<br />

follow-up. J Urd 158:435-439.<br />

78. Elliot D, Barrett D (1998) Mayo Clinic long-term analysis of the functional<br />

durability of the AMS 800 artificial urinary sphincter: a review of 323 cases.<br />

J Urol 159:1206-1208.<br />

79. Light JK (1990) Complications of surgery for male urinary incontinence including<br />

the artificial urinary sphincter. In: Marshall F, ed., Urologic Complications, 2nd<br />

ed. St. Louis, MO: Mosby Year Book, pp. 328-337.<br />

80. Couillard D, Vapnek J, Stone A (1995) Proximal artificial sphincter cuff repositioning<br />

for urethral atrophy incontinence. Urology 45:653-656.


272 Scarper0 and Winters<br />

81. Brito, Mulcahy J, Mitchell M, Adam M (1993) Use of a double cuff AMS 800<br />

urinary sphincter for severe stress incontinence. J Urol 149283-285.<br />

82. Goldwasser B (1990) Avoiding the problems of artificial urinary sphincter implantation.<br />

Probl Urol 4: 187-1 99.<br />

83. Mark S, Perez L, Webster G (1994) Synchronous management of anastomotic<br />

contracture and stress urinary incontinence following radical prostatectomy.<br />

J Ur~l 151:1202-1204.<br />

84. Martins F, Boyd S (1995) Artificial urinary sphincter in patients following major<br />

pelvic surgery andor radiotherapy: are they less favorable candidates? J Urol<br />

153~1188-1193.<br />

85. "rang Y, Hadley R (1 992) Experiences with the artificial urinary sphincter in the<br />

irradiated patient. J Urol 147:612-613.<br />

86. Raz S, McGuire E, Erlich R (1988) Fascial sling to correct male neurogenic<br />

sphincter incompetence: the McGuire/Raz procedure approach. J Urd 139:528--<br />

531.<br />

87. Schaeffer A, Clemens J, Ferrari M, Stamey T (1988) The male bulbourethral sling<br />

procedure for post-radical prostatectomy incontinence. J Urol 159:1510-1515.<br />

88. Gonzalez R, Nguyen D, Koleilat N, Sidi A (1989) Compatibility of enterocystoplasty<br />

and the artificial urinary sphincter. J Urol 144(2 pt 2): 502-504.<br />

89. Perez L, Webster G (1992) Successful outcome of artificial urinary sphincters in<br />

men with post-prostatectomy incontinence despite adverse implantation features.<br />

J Urol 148: 1166-1 170.<br />

90. Stothers L, Chopra A, Raz S (1995) A cost-effectiveness and utility analysis of<br />

the artificial urinary sphincter and collagen injections in the treatment of postprostatectomy<br />

incontinence. J Urol 1533278A.<br />

91. Brown J, Elliot D, Barrett D (1998) Postprostatectomy urinary incontinence: A<br />

comparison of the cost of conservative versus surgical management. Urology<br />

51:715-720.


EN TIES


This Page Intentionally Left Blank


1 2 Clinical Pharmacology<br />

Scott R. SereLs and Rodney A. AppeLL<br />

CONTENTS<br />

INTRODUCTION<br />

FAILURE TO EMPTY<br />

FAILURE TO STORE<br />

CONCLUSION<br />

REFERENCES<br />

INTRODUCTION<br />

There are many types of voiding <strong>dysfunction</strong> that can be treated with<br />

pharmacologic therapy. These voiding problems are often described in<br />

one of two broad categories. There are those problems related to storage<br />

(i.e., failure to store) and those related to emptying (i.e., failure to<br />

empty). Difficulty with storage can result from detrusor hyperactivity,<br />

stress urinary incontinence (SUI), or problems with permeability (i.e.,<br />

interstitial cystitis).<br />

Failure to empty usually is caused by either a hypocontractile bladder<br />

or perhaps a bladder outlet obstruction such as that seen with benign<br />

prostatic hypertrophy (BPH) or the failure to relax the bladder neck<br />

and/or external sphincter.<br />

The treatments for the various types of voiding <strong>dysfunction</strong> are best<br />

understood once the neuroanatomy/anatomy is first delineated. The<br />

major anatomic areas that are involved in micturition are the detrusor,<br />

the smooth muscle of the posterior urethra, the prostate, and the striated<br />

muscle of the external sphincter. The neuroanatomy involves the pelvic,<br />

hypogastric, and pudendal nerves.<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Hurnana Press Inc., Totowa, NJ<br />

275


276 Serels and Appell<br />

It is the afferent nerves from the S2-S4 levels that once Stimulated<br />

immediately act on the efferent nerves. This then leads to stimulation<br />

of the pelvic nerves resulting in relesae of acetylcholine (Ach). The<br />

Ach acts on the parasympathetic receptors on the detrusor muscle to<br />

cause a contraction. Furthermore, several other substances have been<br />

noted to be related after pelvic nerve stimulation. These substances<br />

include histamine, prostaglandin, and serotonin. These noncholinergic<br />

induced contractions are believed to be owing to the liberation of<br />

adenosine triphosphate.<br />

In addition, the afferents to the urethra that travel through the pelvic<br />

and pudendal nerves to the spinal cord can also result in Sympathetic<br />

outflow from the thoraco-lumbar region (i.e., epinephrine and norepinephrine).<br />

The adrenergic receptors of the urethra and bladder neck respond<br />

to norepinephrine released by the sympathetic nervous system, and this<br />

facilitates bladder storage and urinary continence. This is accomplished<br />

by relaxing the bladder body where the beta receptors are most prevalent<br />

as well as by contracting the bladder outlet via the alpha receptors.<br />

Beta receptors (i.e., beta-2) appear along with the acetylcholine (Ach)<br />

receptors in the bladder body and result in smooth muscle relaxation.<br />

On the other hand, the alpha receptors (i.e., alpha-l ) predominate in<br />

the bladder neck and proximal urethra where they cause contraction<br />

when stimulated. In addition, the striated muscle of the pelvic floor<br />

serves as the external sphincter and is innervated by the pudendal nerve<br />

which originates from the sacral cord and uses acetylcholine as its<br />

neurotransmitter. Figure 1 summarizes the neuroanatomy previously<br />

described.<br />

Micturition is a complex interaction of both the central and autonomic<br />

nervous system. During filling, the bladder suppresses reflex contractions<br />

by the help of the beta adrenergic system and the cortico-regulatory<br />

tract. The voiding phase is initiated by relaxation of the external sphincter<br />

as well as the smooth muscle of the bladder outlet. This relaxation<br />

is then followed by a parasympathetic-induced contraction of the detrusor<br />

muscle.<br />

Rased on the previous discussion, it is clear that voiding <strong>dysfunction</strong><br />

can be affected by manipulating the interaction of the neurotransmitters.<br />

Medications are available that can alter transport, synthesis, storage,<br />

release, binding to receptors, degradation, and/or reuptake of these<br />

neurotransmitters. However, most agents that interact with the nervous<br />

system lack specificity for the lower urinary tract and therefore side<br />

effects are common.


Chapter 12 / Clinical Pharmacology 277<br />

Pelvic nerve<br />

(parasympathetics)<br />

A<br />

Ach<br />

L Ac<br />

Hypogastric<br />

Ach<br />

N.<br />

(sympathetic)<br />

Fig. 1. Diagram showing the innervation to the bladder, bladder neck, and exter-<br />

nal sphincter<br />

FAILURE TO EMPTY<br />

Medications to Fucilitute Bhdder Emptying<br />

Normal detrusor contractions can be inhibited by any alteration in<br />

the neuromuscular mechanism that is responsible for initiating and<br />

maintaining micturition. These can result from spine injuries, neurologic<br />

diseases, or events such as pelvidperineal pain, psychogenic problems,<br />

and myogenic impairment that influence the nociceptor reflex mechanisms.<br />

Ach results in bladder contractions via the parasympathetic<br />

nervous system, but it cannot be used as a medication because it is<br />

rapidly broken down by acetylholinesterase. Bethanechol chloride (BC)<br />

is a cholinergic agent that is selective for the bladder and gastrointestinal<br />

tract without being affected by cholinesterase. Bethanecol has been<br />

used for many years to treat hypotonic detrusor activity (I). It is used<br />

in the form of 5-10 mg subcutaneously if the patient is awake with<br />

no known outlet obstruction (2). For partial bladder emptying, oral<br />

bethanecol at a dose of 25-100 mg four times daily may be used.


278 Serels and Appell<br />

Bethanecol has also been used to stimulate reflex bladder contractions<br />

in patients who have had suprasacral spinal-cord injuries (3). The<br />

contraindications include peptic ulcer disease, cardiac arrhythmias,<br />

bladder or bowel obstruction, bronchial asthma, and hyperthyroidism.<br />

In addition, acute circulatory arrest may be caused by intramuscular<br />

or intravenous injection. Other side effects include flushing, nausea,<br />

vomiting, diarrhea, bronchospasm, headache, salivation, sweating, and<br />

visual changes. Overall in several studies, BC has not been demonstrated<br />

to cause sustained physiologic bladder contractions in individuals with<br />

voiding <strong>dysfunction</strong> (4-7).<br />

Metoclopramide (Reglan) and cisapride are two drugs that may be<br />

useful in facilitating bladder emptying. Metaclopramide is a dopamine<br />

antagonist with cholinergic properties. On the other hand, cisapride is<br />

a synthetic benzamide that promotes release of ACH. Both of these<br />

drugs have been shown to be effective in the gastrointestinal tract.<br />

However, their efficacy in the lower urinary tract has yet to be proven<br />

conclusively (8).<br />

Prostaglandins (PG) have been used to facilitate emptying. It is<br />

thought that these substances contribute to the bladder tone and bladder<br />

contractile activity (9,lO). PGE2 and PGF2a, specifically, have been<br />

shown to cause bladder contractility. These medications, when administered<br />

intravesically, improved bladder emptying and reduced the frequency/duration<br />

of postoperative urinary retention (1 0-1 6). Nevertheless,<br />

the use of PG to facilitate bladder emptying has not been completely<br />

determined. The side effects include hypotension, hypertension,<br />

pyrexia, diarrhea, and vomiting.<br />

Alpha adrenergic blockers may facilitate transmission through the<br />

parasympathetic ganglia and in turn cause bladder contractility in addition<br />

to decreasing outlet resistance. There have been several studies that<br />

promoted the use of alpha blockade to prevent postoperative retention<br />

(17,18) or in the treatment of nonobstructive urinary retention (19). It<br />

is possible that the alpha adrenergic blockers may not just relax the<br />

bladder outlet, but rather facilitate the detrusor reflex by a direct or<br />

indirect effect on the Parasympathetic ganglia. This phenomenon was<br />

demonstrated by DeGroat through his studies on cats (20,21).<br />

Furthermore, opiod antagonists are thought to perhaps help stimulate<br />

reflex bladder activity, but the findings to support this hypothesis<br />

are not completely convincing (22-24). Their use is based on the<br />

belief that opiods inhibit the micturition reflex, though the data<br />

is inconclusive.


Chapter 12 I Clinical Pharmacology 279<br />

Medications to Decreuse Outlet Resistance<br />

As mentioned previously, the smooth muscle of the bladder base<br />

and proximal urethra contain predominately alpha-adrenergic receptors.<br />

Therefore, alpha-blockers primarily affect the smooth muscle of the<br />

bladder neck and proximal urethra as well as the striated sphincter<br />

tone. Phenoxybenzamine was the original alpha-adrenolytic agent used<br />

to treat voiding <strong>dysfunction</strong>. The side effects of this medication are<br />

orthostatic hypotension, reflex tachycardia, nasal congestion, diarrhea,<br />

meiosis, sedation, nausea, and vomiting. It has also been shown to<br />

have mutagenic activity, and in animals it induced peritoneal sarcomas<br />

and lung tumors (25). Terazosin (Hytrin), doxazosin (Cardura), and<br />

prazosin (Minipress) are selective alpha-blocking drugs. These drugs<br />

are selective for the alphal receptor. Prazosin differs because it has a<br />

short duration of action and therefore must be administered every 4-6<br />

h. The other drugs can be given once daily (26,27). The use of these<br />

drugs has been promoted for the treatment of voiding <strong>dysfunction</strong><br />

related to benign prostatic hypertrophy (BPH). In this scenario, they<br />

are believed to affect the alpha receptors in the prostatic stroma and<br />

capsule. Another agent, Alfuzosin, is a more selective alpha, antagonist<br />

with similar efficacy to prazosin (28). Tarnulosin is an alphal blocker<br />

that is selective for the alpha,, (majority of receptors on prostate stroma<br />

and likely bladder neck) and alphald receptor subtypes over the alphalb<br />

(29,301. This medication therefore causes less side effects (31-32).<br />

There are several other medications that have been tried to decrease<br />

bladder outlet resistance. Beta-adrenergic agonists have been shown<br />

to decrease urethral pressure but have not been proven to be clinically<br />

useful (33,341. Nitric oxide (NO) has been described as a possible<br />

relaxant of the smooth muscle of the bladder outlet (35). However, its<br />

role in the treatment of voiding <strong>dysfunction</strong> has yet to be completely<br />

elucidated. Furthermore, owing to its ubiquity, selectivity for the lower<br />

urinary tract seems difficult.<br />

Benzodiazepines (diazepam), dantrolene, and baclofen have been<br />

tried to treat voiding <strong>dysfunction</strong> owing to the striated sphincter. All<br />

of these dmgs are considered antispasmotics. Dantrolene works directly<br />

on the skeletal muscle, and the other two medications affect the central<br />

nervous system through their interactions with inhibitory neurotransmitters.<br />

Benzodiazepines influence the neurotransmitter gamma-aminobutyric<br />

acid (GABA) at both the presynaptic and postsynaptic sites in<br />

the brain and spinal cord. The side effects are central nervous system


280 Serels and Appell<br />

(CNS) depression, which results in ataxia, lethargy, and drowsiness<br />

(36). These agents may be worthwhile in patients who cannot relax<br />

their pelvic floor, such as that seen in Hinman’s syndrome. However,<br />

it has not been proven to be effective in detrusor sphincter dysynergia<br />

that is seen with true neurologic disorders. Baclofen is thought to<br />

activate the GABA receptors which causes a decrease in the release<br />

of excitatory transmitters onto motor neurons by increasing potassium<br />

conductance or by inhibiting calcium influx. This in turn results in<br />

normalizing interneuron activity and decreasing motor neuron activity,<br />

which reduces spasticity (37). It is primarily used to treat skeletal<br />

muscle spasticity in a variety of neurologic conditions. The usual dose<br />

is 5 mg twice daily (bid), which can be increased to three times daily<br />

(tid) until a maximum of 20 mg four times daily (qid). Side effects<br />

include drowsiness, insomnia, weakness, dizziness, rash, and purititis.<br />

Withdrawal suddenly of baclofen has provoked hallucinations, tachy-<br />

cardia, and anxiety and should as a result be performed gradually.<br />

Administration of baclofen into the subarachnoid space by an infusion<br />

pump has resulted in decreased skeletal spasticity as well as decreased<br />

striated sphincter dysynergia and bladder hyperactivity (38-40).<br />

Dantrolene is believed to inhibit the release of calcium ions from<br />

the sarcoplasmic reticulum of the striated muscle fibers which inhibits<br />

muscle contraction. This inhibition, however, is not complete and con-<br />

traction is not totally abolished. It has nevertheless been reported to<br />

improve voiding <strong>dysfunction</strong> caused by detrusor-striated sphincter<br />

dysynergia (41). Hepatotoxicity, dizziness, diarrhea, and euphoria have<br />

all been reported side effects. Furthermore, the risk of hepatic injury<br />

is reported to be greater in women (42).<br />

Medication to Treat Outlet Obstruction<br />

Cawed by Benign Prostatic Hypertrophy<br />

As part of the normal aging process, prostatic tissue continues to grow<br />

and can potentially cause a bladder outlet obstruction. Histologically,<br />

50430% of the prostate volume is composed of stromal tissue, with<br />

the glandular tissue comprising the remaining 2040%. The alpha<br />

adrenergic blockers described previously act on the stromal tissue of<br />

the prostate.<br />

Finasteride (Proscar), a selective inhibitor of 5-alpha-reductase,<br />

decreases the conversion of testosterone to dihydrotestosterone. Therefore,<br />

it prevents the growth of the glandular component of the prostate.<br />

In several studies, Proscar has been shown to improve urinary symptoms


Chapter 12 / Clinical Pharmacology 281<br />

Table 1<br />

Mechanisms of Actions of Phytotherapeutic Agents<br />

~ ~ _ _ _ - _ _<br />

Possible mechanisms of actions for phytotherapeutic agents<br />

Salpha-reductase inhibition<br />

Anti-inflammatory effects<br />

Reduction in sex hormone-binding globulin<br />

Influences cholesterol metabolism<br />

Growth factor inhibition<br />

Improvement in detrusor function<br />

Anti-androgen effects<br />

Anti-estrogenic effects<br />

Arornatase inhibition<br />

and reduce the volume of the prostate in men with benign prostatic<br />

hyperplasia and enlarged glands (43-46). In a more recent study by<br />

NIcConnell et al. Proscar was noted to reduce symptoms and prostate<br />

volume, increase the urinary flow rate, and reduce the probability of<br />

surgery for acute retention (47). The usual dose is 5 mg once daily.<br />

Another group of medications, known as phytotherapeutic agents<br />

because they are extracted from plants, is currently being investigated<br />

as a treatment for BPH. These extracts include: Saw palmetto, pygeurn<br />

africanum, beta-sitosterol, pollen extract, pumpkin seeds, and south<br />

african star grass. The exact mechanism of action is also uncertain.<br />

The proposed mechanisms are shown in Table l.<br />

There have been several studies performed looking at phytotherapy<br />

but at this point have not shown convincing evidence that it is effective<br />

in relieving the symptoms of BPH (48,491. The studies thus far have<br />

been uncontrolled or of limited follow-up. Phytotherapy does not result<br />

in any serious side effects and does not seem to influence PSA values.<br />

Nevertheless, the exact role for this therapy has yet to be determined.<br />

FAILURE TO STORE<br />

Medications to Decreuse Bhdder Contrdctility<br />

Acetycholine induces the postganglionic parasympathetic muscarinic<br />

receptor sites of the bladder smooth muscle. Therefore, anticholinergic<br />

agents have been used to depress bladder contractions. There are five<br />

genes that code for muscarinic receptors, represented by ml-m5 (SO).


282 Serels and Appell<br />

M1-M4 represent the protein products of these genes (51,52). M2<br />

receptors are the most prevalent type of receptor in the bladder, whereas<br />

M3 receptors are believed to be most responsible for the contractile<br />

response (53,54).<br />

Propantheline bromide (Pro-Banthine) is an antimuscarinic drug that<br />

is used at an oral dose of 15-30 mg every 4-6 h. It is best absorbed<br />

if taken before meals.<br />

Atropine is also an agent that can be used to decrease bladder<br />

contractility at a dose of 0.5 mg. Scopolamine is another belladonna<br />

alkaloid that is used in a “patch” form that delivers 0.5 mg daily. The<br />

results of this therapy have been mixed and its efficacy is uncertain<br />

(55-57). Furthermore, skin irritation has been a problem with the patch,<br />

Hyoscyamine (Cystospaz) and hyoscyamine sulfate (Levsin) are<br />

other belladonna alkaloids that can be used. All antimuscarinic agents<br />

can produce the following side effects: inhibition of salivary secretion,<br />

blockade of ciliary muscle of the lens to cholinergic stimulation (blurred<br />

vision for near objects), tachycardia, drowsiness, and inhibition of<br />

intestinal motility. These agents are contraindicated in narrow-angle<br />

glaucoma and should be used with great caution in those with bladder<br />

outlet obstruction. The side effects of the aforenoted medications tend<br />

to limit their use. Tolterodine is a new antimuscarinic agent that has<br />

been shown to be selective for bladder tissue over salivary tissue in<br />

both in vitro and in vivo cat studies (58,59). In fact, tolterodine has<br />

an eight times lower affinity for the parotid gland than oxybutynin<br />

(60). These tissue selective effects seem to be as a result of differential<br />

affinities of tolterodine for the receptors of the bladder vs the salivary<br />

glands. The clinical studies thus far have shown a decrease in the<br />

number of micturitions, a decrease in the number of episodes of incontinence,<br />

and an increase in the voided volume (61-63). The usual dose<br />

is 2 mg bid.<br />

Musculotropic agents have been separated from pure anticholinergics<br />

because these agents have anesthetic properties in addition to their<br />

anticholinergic properties. Oxybutynin chloride (Ditropan) is the most<br />

commonly prescribed medication of this type. The recommended dose<br />

is 5 mg three to four times daily. The side effects appear to be doserelated.<br />

The summary results of six randomized controlled studies show<br />

a cure rate of 28-44% and a reduction in urge incontinence of 9-56%,<br />

with side effects occurring in 2-66% (64). Intravesical Ditropan has<br />

also been investigated and shown to be effective with less side<br />

effects (65-67).


Chapter 12 I Clinical Pharmacology 283<br />

In addition, dicyclomine hydrochloride (Bentyl) has been shown to<br />

be effective for detrusor hyperreflexia (68,69). The usual dose is 10-20<br />

mg three times a day.<br />

Flavoxate hydrochloride (Urispas) has been shown to have a direct<br />

inhibitory effect on smooth muscle but very minimal anticholinergic<br />

properties (70,71). It has been used successfully in treating detrusor<br />

hyperactivity (72,73), however, there are no studies from North America<br />

indicating any efficacy with this agent. The dosage is 100-200 mg<br />

three to four times daily.<br />

The intracellular release of calcium is paramount to the contractile<br />

response seen in all muscle, and bladder smooth muscle is no exception.<br />

Nifedipine, a calcium channel blocker, has been shown to inhibit bladder<br />

contractions (70,74). Palpitations, weakness, rash, nausea, constipation,<br />

dizziness, headache, facial flushing, hypotension, and abdominal discomfort<br />

are all potential side effects that may occur with calcium<br />

channel blockers.<br />

Potassium channel openers are a potentially useful therapy for detrusor<br />

overactivity. These medications relax smooth muscle by increasing<br />

potassium efflux which results in membrane hyperpolarization (70).<br />

However, there have been several clinical trials that have failed to<br />

demonstrate an improvement in detrusor hyperreflexia (75-77).<br />

Prostaglandin inhibitors have been used to facilitate bladder storage.<br />

The proposed mechanism of action may be related to the modulation<br />

of the inflammatory response to local irritation or to a modification of<br />

the afferent sensory inervation (70,78). Unfortunately, clinical studies<br />

have failed to demonstrate a significant benefit (79,80). The side effects<br />

commonly seen are rash, constipation, nausea, vomiting, headaches,<br />

and indigestion.<br />

The bladder also contains beta adrenergic receptors, and thus their<br />

stimulation has been attempted to increase bladder relaxation. A beta<br />

agonist, terbutaline, has been noted to have some benefit in the overactive<br />

bladder (81-83). These findings were not universal (W), and the<br />

exact role of beta agonists has yet to be determined.<br />

Tricyclic antidepressants (TCA), such as imipramine hydrochloride,<br />

decrease bladder contractility and increase outlet resistance. These<br />

agents have essentially three major pharmacological actions: they block<br />

the reuptake of amine neurotransmitters (i.e., norepinephrine and serotonin),<br />

they act centrally as sedatives and possibly via antihistaminic<br />

properties, and they have both central and peripheral anticholinergic<br />

effects (85-88).


284 Serels and Appell<br />

Interestingly, the main anticholinergic effect of imipramine is systemic<br />

with only a weak antimuscarinic effect (89,90), but it also causes<br />

a direct nonadrenergic and nonanticholinergic inhibitory effect (90-92).<br />

It has been noted that imipramine has an additive effect when used<br />

with antimuscarinic agents (93). The usual dose of imipramine is 25<br />

mg four times a day with a reduced dose given to elderly patients. The<br />

dosage used to treat depression is much higher. This medication can<br />

also be used to treat nocturnal enuresis in children at a dose of 10-50<br />

mg daily. The side effects are the same as those seen with other systemic<br />

anticholinergics. One can see hypotension and CNS side effects such as<br />

weakness, Parkinsonian effects, fatigue, and tremors. Most importantly,<br />

TCAs can produce arrhythmias and should be used with caution in<br />

those with cardiac disease (86). There have also been several reports<br />

in children of severe side effects, such as vomiting, headaches, lethargy,<br />

abdominal distress, and irritability, following abrupt cessation, and<br />

therefore it should be discontinued gradually. Furthermore, its use is<br />

contraindicated in those taking monoamine oxidase inhibitors because<br />

serious CNS toxicity can be induced.<br />

Another potential mediator of smooth muscle relaxation is nitric<br />

oxide (NO) (94,95). NO is believed to be a nonadrenergic noncholinergic<br />

neurotransmitter but its exact role in treating bladder hyperactivity<br />

is still under investigation.<br />

Capsaicin is a new compound that is obtained from hot peppers and<br />

is highly selective for the sensory neurons in mammals (96,97). Thus,<br />

repeated administration of capsaicin, either systemically or topically,<br />

induces desensitization and inactivation of the sensory nerves by cresting<br />

reversible antinociceptive and anti-inflammatory action. The action<br />

of topical or local capsaicin is owing to the blockade of the C-fiber<br />

conduction and inactivation of the neuropeptides released from the<br />

peripheral nerve endings. By using systemic capsaicin, antinociception<br />

is produced by activating specific receptors on the afferent nerve terminals<br />

in the spinal cord, which results in blockade of spinal neurotransmission<br />

by the prolonged inactivation of sensory neurotransmitter<br />

release. Topical use prevents systemic side effects and acts primarily<br />

on the small diameter nociceptor receptors, which, in turn, prevents<br />

loss of sensation to touch and pressure as well as loss of motor function,<br />

which are owing to larger diameter nociceptor receptors. There have<br />

been several studies supporting the potential use of capsaicin for the<br />

hyperactive bladder (98-102). However, it is not commercially available.<br />

Local drug instillation has been associated with gross hematuria,


Chapter 12 / Clinical Pharmacology 285<br />

severe pains upon instillation, and precipitation of autonomic<br />

dysreflexia in some.<br />

Medications to Increme Outlet Resistance<br />

Alpha adrenergic agonists have been used to increase bladder outlet<br />

resistance. Their use is based on studies that show an abundance of<br />

alpha-adrenergic receptors at the bladder neck and proximal urethra<br />

(103,104). Ephedrine is a treatment used for stress urinary incontinence<br />

(SUI). It enhances the release of norepinephrine from sympathetic<br />

neurons and stimulates directly the alpha and beta-adrenergic receptors<br />

(105). The usual dose is 25-50 mg QID and tachyphylaxis has been<br />

reported. Pseudophedrine, a steriosomer of ephedrine, can also be used<br />

at a dose of 30-60 mg QID. The use of these drugs for severe SUI is<br />

limited and may only be of benefit for minimal wetting (106,107).<br />

Phenylpropanoamine (PPA) has the same pharmacologic properties as<br />

ephedrine and has been used for SUI at a dosage of 50 mg BID. The<br />

side effects of these medications include anxiety, headaches, tremor,<br />

weakness, palpitations, cardiac arrhythmias, hypertension, and respira-<br />

tory difficulties. Thus, these medications should be used with caution<br />

in all patients with cardiac disease, hypertension, increased age, and/<br />

or hyperthyroidism.<br />

Estrogen therapy for SUI has, as with other pharmacologic agents,<br />

been noted to have mixed results (108-118). It is hypothesized that<br />

estrogens may work by changing the autonomic innervation, the recep-<br />

tor content, the function of the smooth muscle, the estrogen binding<br />

sites, the supporting tissues, or the mucosal seal mechanism (119-123).<br />

There have been others that endorsed the use of estrogens in combina-<br />

tion with alpha agonists (I 11,112,116,124,125). Estrogens, however,<br />

can increase the risk of endometria1 cancer if they, are used unopposed<br />

in those with an intact uterus. Progestins exert a protective affect in<br />

these situations. There is also an association between breast cancer and<br />

estrogen replacement in those receiving such therapy for more than 15<br />

years. The beneficial effects of estrogen include osteoporosis prevention<br />

and decreased cardiovascular disease. Estrogen therapy can be adminis-<br />

tered orally, transdemally, subcutaneously implanted, and topically.<br />

Furthermore, the topical approach has been shown to have an added<br />

psychologic benefit (126). A new silicone ring impregnated with estra-<br />

diol is being introduced for vaginal use. This new device offers a


286 Sereis and Appeli<br />

continuous delivery over a 90-d period and has shown to be very<br />

acceptable in certain women (126,127).<br />

Medications for InJEammutory Conditions<br />

of the Bladder Cuzlsing Storuge Failure<br />

Interstitial cystitis (IC) is a syndrome whose pathogenesis and etiology<br />

remain a mystery. Even the diagnosis of this entity is difficult<br />

and considered one of exclusion. There are four currently proposed<br />

etiologies for this syndrome: inflammation, vascular insufficiency, epithelial<br />

leak, and deficiency of proteoglycans (i.e., glycosaminoglycans-<br />

GAG-). Dimethysulfoxide (DMSO), after being approved for use in<br />

1977, has been shown to induce remission in 35-40% of the patients<br />

(128) and is a mainstay of treatment (129). However, no controlled<br />

clinical studies have been performed. DMSO is a derivative of lignin,<br />

which is a product of the wood-pulp industry. Its therapeutic properties<br />

include: anti-inflammatory properties, analgesic properties, collagen<br />

dissolution, muscle relaxation, and mast-cell histamine release. The<br />

usual dose is a 50% solution instilled for 5-10 min. This therapy can<br />

be used as a one time dose, repeated weekly for 6-8 wk, or continued<br />

weekly for 4-6 mo. Some even advocate using DMSO indefinitely. If<br />

the patient has pain with administration, she should receive 2% viscous<br />

lidocaine. A flare of symptoms has also been noted and usually disappears<br />

over 24 h and diminishes with subsequent treatments. DMSO is<br />

generally well-tolerated; however, reports of cataracts in animals have<br />

been noted with long-term use. Thus, it is recommended that patients<br />

receiving chronic therapy undergo slit-lamp evaluation at 3-6 mo<br />

intervals.<br />

Intravesical silver nitrate has been tried, but no good controlled<br />

studies have supported its implementation (130,131). This therapy<br />

should not be used after bladder biopsy because tissue damage may<br />

result if the silver nitrate is introduced outside the bladder.<br />

Heparin has been shown to have significant activity in approximately<br />

50% of patients (133). It is believed to work by restoring the GAG<br />

layer of the urothelium. 10,000 units (U) of heparin is instilled daily<br />

for 3-4 mo and then reduced to 3-4 timedwk. If there is no effect<br />

after 3 mo, it can be increased to 20,000 U. It may take 2-4 mo to<br />

work and should be continued for at least 6 mo.<br />

Pentosanpolysulfate (Elmiron) is a sulfated polysaccharide, which is<br />

thought to augment the bladder surface defense mechanism or possibly<br />

detoxify urine agents that can irritate the bladder surface. In a controlled


Chapter 12 / Clinical Pharmacology 287<br />

study, Elmiron improved the symptoms of 42% of the patients in the<br />

treatment group as compared to 20% in the placebo group (133). These<br />

findings were further corroborated in several other studies (134-136).<br />

The usual dose is 100 mg tid. Response to this medication is first seen<br />

after 6-10 wk, and patients generally do better after 6-12 mo.<br />

Other medications, including corticosteroids, have been tried but the<br />

data to support their use is sketchy. The assumption is that if inflammation<br />

contributes to IC, steroids may help in the treatment. Thus, some<br />

physicians use an intravesical "cocktail," which includes heparin and<br />

steroids weekly for 6 wk.<br />

Antidepressant therapy have been used in patients with chronic pain<br />

who lose sleep owing to depression. The usual medication is amitriptyline<br />

25 mg at bed time. Fluxetine (Prozak) can be used at a dose of<br />

20 mg and increasing to 40 mg as needed. Zoloft (sertraline) may also<br />

be used at a dose of 50-100 mg.<br />

Antihistamines have also been used to inhibit mast-cell release (137-<br />

140). Some patients benefit from this therapy although no controlled<br />

studies have been performed. The beneficial effects are thought to occur<br />

approximately 2-3 MO after treatment.<br />

Medications to Treat Nocturnal Enuresis<br />

as a Cause of Failure to Store<br />

A synthetic antidiuretic hormone peptide analog known as DDAVP<br />

(l -deaminino-%" arginine vasopressin) is used to treat noctural enure-<br />

sis in children and adults (141,142). The medication is administered<br />

by nasal spray at a dose of 10-40 pg at bedtime. Hyponatremia is a<br />

rare complication that must be watched for in all patients, especially<br />

the elderly. There is also an oral form available that may be useful in<br />

treating this entity.<br />

CONCLUSION<br />

There are a plethora of pharmacologic treatment options for voiding<br />

<strong>dysfunction</strong>. These medications can affect both the filling and storage<br />

phase of micturition. As a general principal, therapy should be started<br />

at a low dose and titrated upward as necessary. Thus, this will keep<br />

the side effects to a minimum while maximizing the therapeutic effects.<br />

Many advances in our understanding of neurophysiology and pharmacology<br />

are ongoing. There are many new drugs being developed as


288 Serels and Appell<br />

well as innovative new ways to administer them. It therefore is important<br />

that testing of these medications be conducted in properly developed<br />

and implemented trials so that the most useful information can be<br />

obtained. Effectively developed studies will, in turn, allow the advance-<br />

ment of pharmacologic treatment of voiding <strong>dysfunction</strong> and provide<br />

the best future care to our patients.<br />

1. Lee L (1949) The clinial use of Urocholine in <strong>dysfunction</strong>s of the bladder.<br />

J Urol 62:300.<br />

2. Wein AJ (1986) Specific methods of pharmacologic treatment. In: Stanton SL,<br />

Tanagho EA, eds., Surgery of Female Incontinence New York, NY: Springer-<br />

Verlag, 229-250.<br />

3. Parkash L (1975) Intermittent catheterization and bladder rehabilitation in spinal<br />

cord injury patients. J Urol 114:230.<br />

4. Ban-ett DM, Wein AJ (1991) <strong>Voiding</strong> <strong>dysfunction</strong>: Diagnosis, classification and<br />

management. In: Gillenwater JY, Grayhack ST, Howards ST, Duckett JW, eds.<br />

Adult and Pediatric Urology, 2nd ed. St. Louis, MO: Mosby-Year Book, p. 1001.<br />

5. Finkbeiner AK (1985) Is bethanechol chloride clinically effective in promoting<br />

bladder emptying? J Urol 134:443.<br />

6. Wein AJ, Malloy T, Shofar FE (1980) The effects of bethanechol chloride on<br />

urodynamic parameters in normal women and in women with significant residual<br />

urine volumes. J Urd 124397.<br />

7. Wein AJ, Raezer DO, Malloy T (1980) Failure of the bethanechol supersensitivity<br />

test to predict improved voiding after subcutaneous bethanechol administration.<br />

J Urol 123:202.<br />

8. Wyndaele JJ, Van Kerrebroeck R (1995) The effects of 4 weeks treatment with<br />

cisapride on cystometric parameters in spinal cord injury patients. A double blind,<br />

placebo controlled study. Paraplegia 33:625.<br />

9. Wein AJ, Levin M, Barrett DM (1991) <strong>Voiding</strong> function: relevant anatomy,<br />

physiology, and pharmacology. In: Duckett JW, Howards ST, Grayhack JT, Gillenwater<br />

JY, eds., Adult and Pediatric Urology, 2nd ed. St. Louis, MO: Mosby-<br />

Year Book, p. 933.<br />

10. Jaschevatsky OE, AnderTnan S, Shalit A (1985) Prostaglandin F2a for prevention<br />

of urinary retention after vaginal hysterectomy. Obstet Gynecol 66:244.<br />

11. Bultitude M, Hills N, Shuttleworth K (1976) Clinical and experimental studies<br />

on the action of Prostaglandins and their synthesis inhibitors on detrusor muscle<br />

in vitro and in vivo. Br J Urol 48:631.<br />

12. Despond A, Bultitude M, Hills N, et al. (1980) Clinical experience with intravesical<br />

prostaglandin E2: a prospective study of 36 patients. Br J Urol 53:357.<br />

13. Vaidyanathan S, Rao M, Mapa M (1981) Study of instillation of 15(S)-15-methyl<br />

prostaglandin F2a in patients with neurogenic bladder <strong>dysfunction</strong>. J Urol 126381.<br />

14. Tammela TL, Kontturi M, Lukkarinen 0 (1987) Intravesical prostaglandin F2 for<br />

promoting bladder emptying after surgery for female stress incontinence. Br J<br />

Urol 60:43.


Chapter 12 / Clinical Pharmacology 289<br />

15. Koonings P, Bergman A, Ballard CA (1990) Prostaglandins for enhancing detrusor<br />

function after surgery for stress incontinence in women. J Reprod Med 35: l.<br />

16. Stanton DL, Cardozo LD, Ken-Wilson R (1974) Treatment of delayed onset of<br />

spontaneous voiding after surgery for incontinence. Urology 13:494.<br />

17. Tammela TL (1986) Prevention of prolonged voiding problems after unexpected<br />

postoperative urinary retention: comparison of phenoxybenzamine and carbachol.<br />

J Urol 136: 1254.<br />

18. Goldman G, Levian A, Mazor A (1988) Alpha-adrenergic blocker for posthernioplasty<br />

urinary retention. Arch Surg 123(1):35--36.<br />

19. Raz S, Smith RB (1976) External sphincter spasticity syndrome in female patients.<br />

J Urol 115:443.<br />

20. de Groat WC (1993) Anatomy and physiology of the lower urinary tract. Urol<br />

Clin NAm 20(3):383.<br />

21. de Groat WC, Booth AM (l 984) Autonomic systems to the urinary bladder and<br />

sexual organs. In: Dyck PJ, Thomas PJS, Lambert EH, eds. Periph.eraZ Neuropathy.<br />

Philadelphia, PA: WB Saunders, p. 285.<br />

22. Murray KHA, Fencley RC1 (1982) Endorphins: A role in urinary tract function?<br />

The effect of opioid blockade on the detrusor and urethral sphincter mechanisms.<br />

Br J Urol 54:638.<br />

23. Galeano C, Jubelin B, Biron L, et al. (1986) Effect of naloxone on detrusorsphincter<br />

dyssynergia in chronic spinal cat. NeurouroE Urodyn. 5:203.<br />

24. Wheeler JS, Jr. Robinson CJ, Culkin DJ, et al. (1987) Naloxone efficacy in bladder<br />

rehabilitation of spinal cord injury patients. J Urd 137:1202.<br />

25. Hoffman BB, Lefkowitz RJ (1990) Adrenergic receptor antagonists. In: Gilman,<br />

AG, Rail TW, Nies AS, Taylor P, eds., Goodman and Gilrnan’s The Phamacological<br />

Basis of Therupeutics. 8th ed. New York, NY: Pergamon Press, 221.<br />

26. Physician’s Desk Reference. (1992) Montraler, NJ: Medical Economics Co.<br />

27. Lepor H (1990) Role of long acting selective alpha-I blockers in the treatment<br />

of benign prostatic hyperplasia. Urol Clin N Am 17:651.<br />

28. Buzelin JM, Herbert M, Biondin P ( 1993) Alpha-blocking treatment with afuzosin<br />

in symptomatic benign prostatic hyperplasia: comparative study with prazosin.<br />

The PRAZALF Group. Br J Urol 72(6):922.<br />

29. Lepor H, Tang R, Kobayashi S, et al. (1995) Localization of the alpha-IA-adrenoreceptor<br />

in the human prostate. J Urol 154:2096.<br />

30. Noble AJ, Williams-Chess R, Couldwell C, et al. (1997) The effects of tamulosin,<br />

a high affinity antagonist at functional aIA- and at D- adrenoreceptor subtypes.<br />

Br J Pharrn 120:23 l.<br />

31. Cbapple C, Wyndaele JJ, Nordling J, et al. (1996) Tamulosin, the first prostateselective<br />

alpha I adrenoreceptor antagonist. A meta-analysis of two randomized<br />

placebo-controlled, multicentre studies in patients with benign prostatic obstruction<br />

(symptomatic BPH). European Tamsulosin Study Group. Euro UroE 29(2):155.<br />

32. Wilde M, McTavish D (1996) Tamsulosin. A review of its pharmacological<br />

properties and therapeutic potential in the management of symptomatic benign<br />

prostatic hyperplasia. Drugs 52(6):883.<br />

33. Raz S, Caine M (1971) Adrenergic receptors in the female canine urethra. hest Urol 9:319.<br />

34. Vaidyanathan S, Rao M, Bapna B (1980) Beta adrenergic activity in human<br />

proximal urethra. J Urol 124:869.<br />

35. Anderson IQ3 (1993) Pharmacology of lower urinary tract smooth muscles and<br />

penile erectile tissues. Pharmacol Rei1 45:253.


290 Serels and Appell<br />

36. Shader RI, Greenblatt DJ (1993) Use of benzodiazepines in anxiety disorders.<br />

N Engl J Med 328: 1398.<br />

37. Milanov IG (1991) Mechanisms of baclofen action on spasticity. Acta Neutral<br />

Scand 85:304.<br />

38. Kurns JJM, Delhaas EM (1991) Intrathecal baclofen infusion in patients with<br />

spasticity and neurogenic bladder disease. World J Urol 9:99.<br />

39. Nannings J, Frost F, Penn R. Effect of intrathecal baclofen on bladder and sphincter<br />

function. J Urol 142:101.<br />

40. Loubser PG, Narayan RK, Sadin KJ (1991) Continuous infusion of intrathecal<br />

baclofen: Long term effects on spasticity in spinal cord. Paraplegia 29:48.<br />

41. Murdock M, Sax D, Krane R (1976) Use of dantrolene sodium in external sphincter<br />

spasm. Urology 8: 133.<br />

42. Ward AK, Chaffman MO, Sorkin EM (1986) A review of its p~~macodynamic<br />

and pharacokinetic properties and therapeutic use in malignant hyperthermia, the<br />

neurologic syndrome and an update of its use in muscle spasticity. Drugs 32: 130.<br />

43. Gomley GJ, Stoner E, Bruskewitz RC, et al. (1992) The effect of finasteride in<br />

men with benign prostatic hyperplasia. NEJM 327: l 3 85-91.<br />

44. McConnell JD, Wilson JD, George FW, Geller J, Pappas F, Stoner E (1992)<br />

Finasteride, an inhibitor of Sa-reductase, suppresses prostatic dihydrotestosterone<br />

in men with benign prostatic hyperplasia. J Clin Endocrinol Metab 74:505-8.<br />

45. Boyle P, Could AL, Roehrborn CG (1996) Prostate volume predicts outcome<br />

of treatment of benign prostatic hyperplasia with finasteride: meta-analysis of<br />

randomized clinical trials. Urology 48:398-405.<br />

46. Andersen JT, Nickel JC, Marshall VR, Schulman CC, Boyle P (1997) Finasteride<br />

significantly reduces acute urinary retention and need for surgery in patients with<br />

symptomatic benign prostatic hyperplasia. Urology 49:839-45.<br />

47. McConnell JD, Bruskewitz R, Walsch P, et al. (1998) The effect of finasteride<br />

on the risk of acute urinary retention and the need for surgical treatment among<br />

men with benign prostatic hypertrophy. NEJM 338557-563.<br />

48. Lowe FC, Ku JC (1996) Phytotherapy in treatment of benign prostatic hyperplasia:<br />

a critical review. Urology 48( 1): 12-20.<br />

49. Carraro JC, Raynaud JP, Koch G, et al. (1996) Comparison of phytotherapy<br />

(Permixon) with finasteride in the treatment of benign prostate hyperplasia: a<br />

randomized international study of 1,098 patients. Prostate 29(4):23 1-40.<br />

SO. Bonner TI (1989) The molecular basis of muscarinic receptor diversity. Trends<br />

Neurosci 12: 148.<br />

51. Poli E, Monica B, Zappia L (1992) Antimuscarinic activity of telemyepine on<br />

isolated human urinary bladder: role no for M-l receptors. Gen PharmacoZ23:659.<br />

52. Eglen RM, Watson N (1996) Selective muscarinic receptor agonists and antagonists.<br />

Plzarmacol Toxicol 78:59,<br />

53. Nilvebrant L, Anderson ICE, Giliberg PG, et al. (1997) Tolterodine-a new<br />

bladder-selective antimuscarinic agent. Eur J Pharm 327: 195.<br />

54. Tobin G, Sjogren C (1995) In vivo and in vitro effects of muscarinic receptor<br />

antagonists on contractions and release of E3H] acetylcholine release in the rabbit<br />

urinary bladder. Eur J Pharm 28: 1.<br />

55. Weiner LB, Baum NH, Suarez GM (1986) New method for management of<br />

detrusor instability: transdermal scopolamine. Urology 28:208.<br />

56. Cornelia JL, Bent AE, Ostergard DR, et al. (1990) Prospective study utilizing<br />

transdermal scopolamine in detrusor instability. Urology 35:96.


Chapter 12 I Clinical Pharmacology 29 1<br />

57. Greenstein A, Chen J, Matzkin HE (1992) Transdermal scopolamine in prevention<br />

of post open prostatectomy bladder contractions. Urology 39:215.<br />

58. Nilvebrant L, Hallen B, Larsson G (1997) A new bladder selective muscarinic<br />

receptor antagonist: preclinical pharmacological clinical and data. Life Sci 60: 1 129.<br />

59. Nilvebrant L, Sundquist S, Giliberg PG (1996) Tolterodine is not subtype (ml-m5)<br />

selective but exhibits functional bladder selectivity in vivo. Neurol Urodyn 15:310.<br />

60. Nilvebrant L, Glas G, Jonsson A, et al. (1994) The in vitro pharmacological<br />

profile of tolterodine-a new agent for the treatment of urinary urge incontinence.<br />

Neurourol Urodyn 13:433.<br />

61. Appell RA (1997) Clinical efficacy and safety of tolterodine in the treatment of<br />

overactive bladder: a pooled analysis. Urology 50:90-96.<br />

62. Abrams P, Freeman R, Anderstrom C, et al. (1998) Tolterodine, a new antimuscarinic<br />

agent: as effective but better tolerated than oxybutynin in patients with an<br />

overactive bladder. Br J Urol 81(6):801-810.<br />

63. Jonas U, Hofner K, Madersbacher H, Holmdahl TH (1997) Efficacy and safety<br />

of two doses of tolterodine versus placebo in patients with detrusor overactivity<br />

and symptoms of frequency, urge incontinence, and urgency: urodynamic evaluation.<br />

The International Study Group. World J Urol 15(2):144-1.51.<br />

64. Urinary Incontinence Guideline Panel (1992) Urinary Incontinence in Adults:<br />

Clinical Practice Guidelines. AHCPR Pub. No. 92-0038. Rockville, MD: Agency<br />

for Health Care Policy and Research, Public Health Service, US Department of<br />

Health and Human Services.<br />

65. Kat0 K, Kitada S, Chun A, et al. (1989) In vitro intravesical instillation of<br />

anticholinergic, antispasmodic and calcium blocking agents (rabbit whole bladder<br />

model.) J Urd 141:1471.<br />

66. Mizunaga M, Nfiyata M, Kaneko S, et al. (1994) Intravesical instillation of<br />

oxybutynin hydrochloride therapy for patients with a neuropathic bladder. Pamplegta<br />

32(1):25.<br />

67. Szollar SM, Lee SM (1 996) Intravesical oxybutynin for spinal cord injury patients.<br />

Spinal Cord 34:284.<br />

68. Downie J, Twiddy D, Awad SA (1997) Antimuscarinic and non-competitive<br />

antagonist properties of dicyclomine hydrochloride in isolated human and rabbit<br />

bladder muscle. J Pharmacol Exp Ther 201:662.<br />

69. Fischer C, Diokno A, Lapides J (1978) The anticholinergic effects of dicyclomine<br />

hydrochloride in inhibited neurogenic bladder <strong>dysfunction</strong>. J Urol 120:328.<br />

70. Andersson lE (1988) Current concepts in the treatment of disorders of micturition.<br />

Drugs 35:477.<br />

71. Briggs RS, Castieden CM, Asher (1980) MJ The effect of flavoxate on ~lninhibited<br />

detrusor contractions and urinary incontinence in the elderly. J Urol 123:656.<br />

72. Delaere KJP, Michiels HGE, Debruyne FMJ, et al. (I 977) Flavoxate hydrochloride<br />

in the treatment of detrusor instability. Urol Intl 32:377.<br />

73. Jonas U, Petri E, Kissal J (1979) The effect of flavoxate on hyperactive detrusor<br />

muscle. Eur Urol 5:106.<br />

74. Forman A, Andersson K, Henriksson, et al. (1978) Effects of nifedipine on the<br />

smooth muscle of the human urinary tract in vitro and in vivo. Acta Pharmacol<br />

Toxicol 43: 1 l 1.<br />

75. Hedlund H, Mattiasson A, Andersson ICE (1990) Lack of effect of pinacidil on<br />

detrusor instability in men with bladder outlet obstruction. J Urd 143336914.<br />

76. Nurse DE, Restorick JM, Mundy AR (1991) The effect of cromakalim on the<br />

normal and hyperreflexic human detrusor muscle. Br J Urol 68:27.


292 Serels and Appell<br />

77. Komersova K, Rogerson IW, Conway EL, et al. (1995) The effect of Levcromokalim<br />

(BRL 38227) on bladder function in patients with high spinal cord lesions.<br />

Br J Clin Pharm 39:207.<br />

78. Downie JW, Karmazyn M (1984) Mechanical trauma to bladder epithelium liberates<br />

prostanoids which modulate neurotransmission in rabbit detrusor muscle.<br />

J Pharmacol Exp Ther 230:445.<br />

79. Cardozo LD, Stanton SL, Robinson H, et al. (1980) Evaluation of Aurbiprofen in<br />

detrusor instability. Br Med J 2:281.<br />

80. Cardozo L, Stanton SL (1979) An object comparison of the effects of parenterally<br />

administered drugs in patients suffering from detrusor instability. J Urol 122:58.<br />

81. Norien L, Sundin T, Waagstein F (1978) Beta-adrenoceptor stimulation of the<br />

human urinary bladder in vivo. Acta Pharmacol Toxicol 43:5.<br />

82. Lindholm P, Lose G (1986) Terbutaline (Bricanyl) in the treatment of female<br />

urge incontinence. Urol Lnt 41:158.<br />

83. Gruneberger A (1984) Treatment of motor urge incontinence with clenbuterol<br />

and flavoxate hydrochloride. Br J Obstet Gynecol 91:275.<br />

84. Castieden CM, Morgan B (1980) The effect of beta adrenoceptor agonists on<br />

urinary incontinence in the elderly. Br J Clin Pharmacol l0:619.<br />

85. Barrett DM, Wein AJ (1991) <strong>Voiding</strong> <strong>dysfunction</strong>: diagnosis, classification and<br />

management. In: Gillenwater, JY, Grayhack JT, Howards ST, Duckett W, eds.,<br />

Adult and Pediatric Urology, 2nd ed. St. Louis, MO: Mosby-Year Book, p. 1001.<br />

86. Baldesarini RJ (1990) Drugs and the treatment of psychiatric disorders. In: Gilman,<br />

AG, Rail TW, Nies AS, Taylor P, eds., Goodman and Gilman 'S The Pharmacological<br />

Basis of Therapeutics, 8th ed. New York, NY: Pergamon Press, p. 383.<br />

87. Hollister LE (1986) Current antidepressants. Ann Rev Pharmacol Toxicol 26:23.<br />

88. Richelson E (1990) Antidepressants and brain neurochemistry. Mayo Clin Proc<br />

65: 1227.<br />

89. Levin RM, Staskin D, Wein AJ (1983) Analysis of the anticholinergic and musculotropic<br />

effects of desmethylirnipramine on the rabbit urinary bladder. Urol Res<br />

11:259.<br />

90. Levin RM, Wein AJ (1984) Comparative effects of five tricyclic compounds on<br />

the rabbit urinary bladder. Neurourol Urodyn 3:127.<br />

91. Benson GS, Sarshik SA, Raczer DM, et al. (1977) Bladder muscle contractility:<br />

Comparative effects and mechanisms of action of atropine, propantheline, flavoxate<br />

and irniprarmine. Urology 9:31.<br />

92. Olubadewo J (1980) The effect of imipramine on rat detrusor muscle contractility.<br />

Arch Int Pharmacodyn Ther 145:84.<br />

93. Raezer DM, Benson GS, Wein AJ (1977) The functional approach to the management<br />

of the pediatric neuropathic bladder: a clinical study. J Urol l I7:649.<br />

94. James MJ, Birmingham AT, Hill SJ (1993) Partial medication by nitric oxide of<br />

the relaxation of human isolated detrusor strips in response to electrical field<br />

stimulation. Br J Clin Pharmarcol 35:366.<br />

95 * Andersson W3 (1993) Pharmacology of lower urinary tract smooth muscles and<br />

penile erectile tissues. Pharmacol Rev 45:253.<br />

96. Maggi CA (1992) Therapeutic potential of capsaicin-like molecules-Studies in<br />

animals and humans. Life Sci 5 l : 1777.<br />

97. Dray A (1993) Mechanism of action of capsaicin-like molecules on sensory<br />

neurons. Life Sci 5 l: 1759.<br />

98. Maggi CA (1991) Capsaicin and primary afferent neurons: from basic science to<br />

human therapy? J Auton New Syst 33:1-14.


Chapter 12 / Clinical Pharmacology 293<br />

99. Maggi CA, Barbanti G, Santicioli P, et al. (1989) Cystornetric evidence that<br />

capsaicin-sensitive nerves modulate the afferent branch of micturition reflex in<br />

humans. J Urol 142:150.<br />

1 00. Fowler CJ, Beck RO, Gerrard S, et al. (1996) Intravesical capsaicin for treatment<br />

of detrusor hyperreflexia. J Spinal Cord Med 19:190.<br />

101. Das A, Chancellor M, Watdnabe T, et al. (1996) Intravesical capsaicin in neurologic<br />

impaired patients with detrusor hyperreflexia. J Spinal Cord Med 19: 190.<br />

102. Chandiramani VA, Peterson T, Duthie CS, et al. (1996) Urodynamic changes<br />

during therapeutic intravesical instillations of capsaicin. Br J Urd 77(6):792.<br />

103. Nilvebrant L, Sundquist S, Giliberg PG (1996) Tolterodine is not subtype (mlm5)<br />

selective but exhibits functional bladder selectivity in vivo. Neurol Urodyn<br />

15:310.<br />

104. Wein AJ, Van Arsdalen W, Levin RM (1991) Pharmacologic therapy. In: Krane<br />

RJ, Siroky MB, eds. Clinical-Neuro-Urology Boston: Little Brown, p. 523.<br />

l 05. Hoffman BB, Lefkowtiz RJ (1990) Andrenergic receptor agonists. In: Gilman<br />

AG, Rail TW, Nies AS, Taylor P, eds., Goodman and Gilman’s The Pharmacological<br />

Basis oj’l’herapeutics, 8th ed., New York, NY: Pergamon Press, p. 221.<br />

106. Diokno A, Taub M (1 975) Ephedrine in treatment of urinary incontinence.<br />

Urology 5:624.<br />

107. O’brink A, Bunne G (1 978) The effect of alpha adrenergic stimulation in stress<br />

incontinence. Urd Intl 12:205.<br />

108. Salmon UJ, Walter RI, Geist SH (1941) The use of estrogens in the treatment<br />

of dysuria and incontinence in postmenopausal women. Am J Obstet Gynecol<br />

42:845.<br />

109. Rud T (1980) The effects of estrogens and gestagens on the urethral pressure<br />

profile of urinary continent and stress incontinent women. Acta Obstet Gynecol<br />

Scand 59:265.<br />

110. Raz S, Ziegler M, Caine M (1973) The role of female hormones in stress<br />

incontinence. Proc 16th Congr Society International d’urologic, vol. I, Paris,<br />

p. 397.<br />

111. Schreiter F, Fuchs P, Stockamp K (1976) Estrogenic sensitivity of a-receptors<br />

in the urethra musculature. Urol Intl 3 1 : 13.<br />

1 12. Beisland HO, Fossberg E, Moer A, et al. (1984) Urethral sphincteric insufficiency<br />

in postmenopausal females: treatment with phenylpropanolamine and estriol<br />

separately and in combination. Urol Intl 39:211.<br />

113. Bhatia NN, Bergman A, Karrain MM (1989) Effects of estrogen on urethral<br />

function in women with urinary incontinence. Am J Obstet Gynecol 160: 176.<br />

114. Karram MM, Yeko TR, Sauer MV, et al. (1989) Urodynamic changes following<br />

hormonal replacement therapy in women with premature ovarian failure. Obstet<br />

Gyneco174:208.<br />

115. Walker S, Wolf H, Barleto H, et al. (I 978) Urinary incontinence in post menopausal<br />

women treated with estrogens. Urol Intl 33:135.<br />

116. Hilton P, Stanton SL (1983) The use of intravaginal estrogen cream in genuine<br />

stress incontinence. Br J Obstet Gyneco190:940.<br />

1 17. Samsioe G, Jansson L, Mellstrom D, et al. (1985) Occurrence, nature and treatment<br />

of urinary incontinence in a 70-year-old fernale population. Mut~ritas.<br />

7:335.<br />

118, Cardozo L (1990) Role of estrogens in the treatment of female urinary incontinence.<br />

J Am Geriatric Soc 38:326.


294 Serels and Appell<br />

119. Bump RC, Friedman CL (1986) Intraluminal urethral pressure measurements in<br />

the female baboon: Effects of hormonal manipulation. J Urol 136:508.<br />

120. Batra SC, Losif CS (1983) Female urethra: a target for estrogen action. J Urol<br />

129:418.<br />

121. Batra S, Byellin L, Sjogren C (l 986) Increases in blood flow of the female rabbit<br />

urethra following low dose estrogens. J Urol 136:1360.<br />

122. Rud T (1980) Urethral pressure profile in continent women from childhood to<br />

old age. Acta Obstet Gynecol Scand 59:331.<br />

123. Versi E, Cardozo L, Buncat L, et al. (1 988) Correlation of urethral physiology<br />

and skin collagen in post menopausal women. Br J Urol Gynecol95:147.<br />

124. Kinn AC, Lindskog M (1988) Estrogens and phenylpropanolamine in combination<br />

for stress urinary incontinence in postmenopausal women. Urology 32:273.<br />

125. Walter S, Kjaergaard B, Lose G, et al. (1990) Stress urinary incontinence in<br />

postmenopausal women treated with oral estrogen (estriol) and an alpha-adrenoceptor<br />

stimulating agent (phenylpropanolamine): a randomized double-blind placebo-controlled<br />

study. Intl Urogynecol J 1:74.<br />

126. Ayton RA, Darling GM, Murkies AL, et al. (1996) A comparative study of<br />

safety and efficacy of continuous low dose estradiol released from a vaginal ring<br />

compared with conjugated equine estrogen vaginal cream in the treatment of<br />

postmenopausal urogenital atrophy. Br J Obstet Gynecol 103:351.<br />

127. Johnston A (1996) Estrogens: pharmacokinetics and pharmacodynamics with<br />

special reference to vaginal administration and the new estradiol formulation-<br />

Estring. Acta Obstet Gynecol Scan 165: 16.<br />

128. Parsons CL and Parsons KJ (1996) Interstitial cystitits. In: Raz S, eds., Female<br />

Urology Philadelphia, PA: WB Saunders, pp. 167-182.<br />

129. Sant GR (1987) Intravesical 50% dimethyl sulfoxide (RIMSO-50) in treatment<br />

of interstitial cystitis. Urology 29: 17-26.<br />

130. Dodson AI (1926) Hunner’s ulcer of the bladder: a report of 10 cases. Virginia<br />

Med Monthly 53:305.<br />

131. Pool TL (1967) Interstitial cystitis: clinical considerations and treatment. CEin<br />

Obstet Gynecol 10: 185.<br />

132. Parsons CL, Housley T, Schmidt JD, Lebow D (1994) Treatment of interstitial<br />

cystitis with intravesical heparin. Br J Urol 73504-7.<br />

133. Parsons CL, Mulholland SG (1987) Successful therapy of interstitial cystitis with<br />

pentosanpolysulfate. J Urol 13851 3-516.<br />

134. Mulholland SG, Hanno P, Parsons CL, Sant GR, Staskin DR (1990) Pentosan<br />

polysulfate sodium for therapy of interstitia1 cystitis: a double-blind placebocontrolled<br />

clinical study. Urology 35(6):552-58.<br />

135. Parsons CL, Benson G, Childs SJ, Hanno P, Sant GR, Webster G (1993) A<br />

quantitatively controlled method to prospectively study interstitial cystitis and<br />

which demonstrates the efficacy of pentosan polysulfate. J Urol 150:845-48.<br />

136. Hanno PM (1997) Analysis of long-term Elmiron therapy for interstitial Cystitis.<br />

Urology 49(5A S~ppl):93-99.<br />

137. Larsen S, et al. (1982) Mast cells in interstitial cystitis. Br J Urol 54:283.<br />

138. Smith B, Dehner LP (1972) Chronic ulcerating interstitial cystitis. A study of<br />

28 cases. Arch Pathol 93:76.<br />

139. Bohne AW, Hodson JNI, Rebuck JW, Reinhard RE (1962) An abnormal leukocyte<br />

response in interstitial cystitis. J Urd 88:387.


Chapter 12 / Clinical Pharmacology 295<br />

140. Simons JL (1961) Interstitial cystitis: an explanation for the beneficial effect<br />

of an antihistamine. J UroE 85:149.<br />

14 l. Norgaard JP, Rillig S, Djurhuus JC (1989) Nocturnal enuresis: an approach to<br />

treatment based on pathogenesis. J Pediatr 114:705.<br />

142. Rew DA, Fundle JSH (1989) Assessment of the safety of regular DDAVP therapy<br />

on primary nocturnal enuresis. Br J UroE 63:352.


This Page Intentionally Left Blank


Treatment of Detrusor Instability<br />

pvith Electrical Stimulation<br />

Steven K Siegel MD<br />

CONTENTS<br />

NORMAL URINE STORAGE AND EVACUATION<br />

DISRUPTED MICTURITION BALANCE<br />

MECHANISM OF ACTION<br />

OF ELECTRICAL STIMULATION<br />

ELECTRICAL STIMULATION<br />

AS A TREATMENT MODALITY<br />

CONCLUSIONS<br />

REFERENCES<br />

There are an estimated 13 million Americans who suffer from incon-<br />

tinence. Urge incontinence is conservatively estimated to account for<br />

40% of all urinary incontinence patients (I). Of this population, two-<br />

thirds suffer from chronic or established incontinence. Yet patients<br />

diagnosed with urinary incontinence owing to detrusor instability have<br />

had limited treatment options. Nonsurgical interventions, including diet<br />

modification, behavioral techniques (pelvic muscle exercises, biofeed-<br />

back, timed voiding), drug therapies, and containment devices are com-<br />

monly used to treat the condition. If these therapies are unsuccessful<br />

or unsatisfactory to the patient, surgical interventions such as bladder<br />

denervation procedures, augmentation cystoplasty, or urinary diversion<br />

may be considered. These alternatives have their own set of risks and<br />

consequences, making them unattractive to the majority of patients.<br />

According to the 1996 National Association for Continence (NAFC)<br />

survey of 2,000 incontinent persons in the US, although more treatments<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfunction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Hurnana Press Inc., Totowa, NJ<br />

297


298 Siege1<br />

are available to urge incontinent patients, 63% of these patients reported<br />

they were "not satisfied" with their treatment outcomes. The lack of<br />

effective treatments for urge incontinence is particularly disturbing<br />

given the debilitating nature of this condition. Incontinent patients<br />

commonly experience loss of self-esteem, shame, depressive symptoms,<br />

embarrassment, anger, and a significant loss of quality of life (I). Urge<br />

incontinence is especially difficult given the severity and unpredictable<br />

nature of leaking episodes. Consequently, patients restrict or avoid<br />

social interactions, and have difficulties meeting daily responsibilities.<br />

Over the past 30 years, advances have been made in the treatment<br />

of urinary incontinence via electrical stimulation of neural pathways<br />

that control bladder function. New therapies have emerged offering<br />

treatment that is reversible and does not preclude other complimentary<br />

alternatives for management. They include use of transvaginal/transanal<br />

Stimulation, needle stimulation, and sacral nerve stimulation (SNS)<br />

devices. Electrical stimulation therapies involving these devices may<br />

differ in their mechanism of action, patient compliance and acceptability,<br />

efficacy, and suitability for a given underlying condition. The<br />

various forms of electrical stimulation have been used for symptoms<br />

of stress, urge and mixed incontinence, urgency-frequency syndromes,<br />

urinary retention, and painful bladder disorders. Regardless of clinical<br />

diagnosis, the goal of therapy is to alter lower urinary tract function<br />

by stimulation of pudendal or sacral nerves and modulate pelvic visceral<br />

and striated muscle behavior.<br />

NORMAL URINE STORAGE AND EVACUATION<br />

Normal micturition relies on urine storage and release as reciprocal<br />

functions in which there is precise coordination between the detrusor,<br />

striated muscles of the pelvic floor (levator anihphincter), and the<br />

external urinary sphincter. Storage of urine during bladder filling<br />

requires the bladder to be compliant in order to distend without increased<br />

pressure, and stable, so that the detrusor does not contract causing<br />

sudden increased pressure and possible incontinence. Coordination of<br />

these muscles systems is controlled by nervous system components<br />

located in the brain, spinal cord, bladder, and urethra via reflex mechanisms.<br />

Tension (afferent) receptors in the bladder wall respond to<br />

distention, transmitting signals through the A-delta fibers when transvesical<br />

pressure approaches 5-10 cm H20 (2,3). As the bladder fills,<br />

the detrusor remains relaxed and the pelvic floor tightens (guarding


Chapter 13 / Treatment of Detrusor Instability 299<br />

reflex). With continued filling, the afferent signals to the sacral cord<br />

and brain stem become stronger, and the guarding reflex increases<br />

accordingly. Upon initiation of voluntary voiding, the pelvic floor<br />

relaxes and the detrusor contracts (4). A positive feedback loop between<br />

bladder afferent and pelvic efferent neurons allows for efficient bladder<br />

emptying with minimal residual. This action is facilitated by supraspinal<br />

input through the pontine micturition center.<br />

DISRUPTED MICTURITION BALANCE<br />

Conscious control allows voiding to occur at convenient times. Inhibitory<br />

reflexes (autonomic and somatic) coordinated by the pons are<br />

needed to keep the sacral micturition reflex in balance. If these reflexes<br />

are overly inhibited, the balance is tipped towards urgency and urge<br />

incontinence. If they are overly facilitated, the balance is shifted towards<br />

urinary retention. Viewed in this way, detrusor instability and urge<br />

incontinence simply represent the flip side of urinary retention. Both<br />

syndromes represent a central nervous system (CNS) <strong>dysfunction</strong>,<br />

which secondarily affects pelvic visceral function. It also follows that<br />

symptoms of bowel <strong>dysfunction</strong> (irritable bowel or chronic constipation)<br />

are likely to be owing to the same imbalance of reflexes. It becomes<br />

clear that any therapy directed at the end organ (i.e., anticholinergics,<br />

denervation procedures, augmentation) or any therapy that does not<br />

positively affect both bladder and bowel function (i.e., anticholinergics)<br />

is unlikely to be completely successful.<br />

MECHANISM OF ACTION<br />

OF ELECTRICAL STIMULATION<br />

The way in which electrical stimulation improves voiding dysfunc-<br />

tion is unknown, but most experts agree that the therapies work by<br />

producing a modulatory effect on sacral nerve reflexes (5-40). Stimula-<br />

tion restores a balance between sacral reflexes that are either overly<br />

inhibited or facilitated. Depending on where the stimulation is delivered,<br />

the mechanism and potential benefit may be different. For example,<br />

transvaginal or transanal devices first stimulate skin receptors and<br />

myelinated A-delta fibers, causing the striated muscles of the pelvic<br />

floor to contract, and secondarily turning off inappropriate detrusor<br />

contractions by augmenting the guarding reflex. These actions may


300 Siege1<br />

be duplicated by conscious efforts and practiced behaviors (Kegel<br />

exercises), but they must be incorporated into everyday behavior in<br />

order to have long term success. If effective over the long term, there<br />

is an implication that the abnormal reflex activity can be modified by<br />

consistent corrective behavior, or that the <strong>dysfunction</strong>s in the nervous<br />

system can be altered permanently.<br />

Sacral nerve stimulation may work in a similar fashion, but also<br />

affects the nervous system through direct stimulation of unmyelinated<br />

C-fiber sacral nerve afferents, which have a much higher threshold of<br />

stimulation than the myelinated A-delta fibers. It is certain that these<br />

nerves play an important role in abnormal bladder activity and in<br />

syndromes with pelvic pain (10). Because SNS devices are implanted,<br />

they can continue to modulate abnormal reflex activity even if the<br />

patient is incapable of conscious behavioral change, or if the changes<br />

in the CNS are otherwise permanent.<br />

ELECTRICAL STIMULATION<br />

AS A TREATMENT MODALITY<br />

Devices for electrical stimulation therapy are the result of observations<br />

that spontaneous or artificially evoked bladder hyperactivity can<br />

be inhibited by electrical stimulation of the pelvic floor and sacral<br />

nerves (5-7). Animal experiments and electrophysiologic studies in<br />

humans have confirmed that spinal inhibitory systems capable of interrupting<br />

a detrusor contraction can be activated by electrical stimulation<br />

(11,12). The afferent anorectal branches of the pelvic nerve, afferent<br />

nonmuscular somatic (sensory) fibers in the pudendal nerve, and muscle<br />

afferents from the limbs have all been found sensitive to low-frequency<br />

stimulation (6,8,13-19). Various devices are currently on the market<br />

utilizing these nerve paths.<br />

There are two main types of electrical stimulation in use, long-term,<br />

or chronic, and acute maximal functional electrical stimulation. Chronic<br />

stimulation is delivered below the sensory threshold usually for more<br />

than 6 h a day over a number of months, or-with the newer implantable<br />

devices-for years. Maximal functional Stimulation is used for short<br />

periods of time (15-30 min) at varying intervals from daily to weekly,<br />

for lengths of time ranging for 10-20 wk. In both types of stimulation<br />

electrical current is pulsed at frequencies based on clinical diagnosis.<br />

In detrusor instability, the goal of therapy is to increase inhibitory<br />

impulses to the bladder using afferent nerve pathways (see Fig. 1).


Chapter 13 I Treatment of Detrusor Instability 301<br />

Fig. 1. Position points and nerve pathways for electrical stimulation devices.<br />

Electrical stimulation also has an effect on the striated pelvic muscles<br />

responsible for inhibition of bladder contraction.<br />

Control of voiding mediated by electrical stimulation does not appear<br />

to be the result of a placebo effect, though this has not received extensive<br />

study. The nature of the treatment intervention generally precludes<br />

double-blind studies. There is evidence to suggest, however, that a<br />

significant placebo effect is not present. A study of vaginal stimulation<br />

included a period of active electrical stimulation followed by a period<br />

of mechanical stimulation of a vaginal electrode without connection<br />

to a pulse generator. There was no effect on incontinence in the absence<br />

of electrical stimulation (2). Clinical trials of the InterStimB device<br />

have also included a therapy evaluation period when electrical stimulation<br />

was turned off. Regardless of the clinical incontinence diagnosis,<br />

incontinence levels returned to the baseline values measured prior to<br />

n


302 Siege1<br />

Fig. 2. Electrode placement at the S3 nerve root. The sacral foramina are probed<br />

with an insulated needle using bony landmarks as a guide. There is little risk of<br />

direct nerve impingement with correct orientation of the needle.<br />

implantation of electrodes. The results suggest that a placebo effect is<br />

not apparent; electrical stimulation is necessary for inducing inconti-<br />

nence control in treated patients (20).<br />

Trunsvuginul and Trunsunal Electrical Stimuhtion Devices<br />

Noninvasive electrical stimulation of the pelvic floor using an exter-<br />

nally applied electrical source offers potential for retraining the basic<br />

physiologic responses of intact muscle tissue for patients with stress,<br />

urge, or mixed incontinence. Electrical stimulation induces a Kegel-<br />

type movement of the pelvic floor muscles and therefore may be useful


303


304 Siege1<br />

in patients who are unable to coi-rectly perform these exercises using<br />

verbal, written, or other visual cues. Devices of this type have been<br />

used transvaginally in females and transanally in males. The therapy<br />

involves vaginal or rectal probes connected to an external power source<br />

(see Table 1). Stiinulation causes the pelvic-floor muscles to contact<br />

involuntarily. The frequency of stimulation is modified depending on<br />

the type of incontinence. A frequency of 50Hz may be used in treatment<br />

of stress incontinence; lower frequencies, below 25Hz, are used for<br />

the treatment of urge incontinence (21 ). Typical treatment regimens<br />

involve use of the devices for 15-30 min twice per day, every other<br />

day. This regimen allows the muscles to recover from the refractory<br />

period induced by maximal functional stimulation.<br />

Studies have demonstrated improved continence in 54-7796 of<br />

patients at follow-up periods from 6 wk to 2 yr (22-26), although there<br />

has been some conflicting evidence (2 7-29). Differences in stimulation<br />

protocols make direct comparisons of the literature difficult (30). When<br />

pelvic-floor stimulation is combined with other therapies or pelvic floor<br />

muscle exercises are continued after treatment, curehmprovements rates<br />

have increased (31). This form of treatment is desirable because it may<br />

be effective for urge, stress, and mixed incontinence. Another obvious<br />

advantage of this type of therapy is its ease of use and relatively low<br />

cost. Patients can be taught easily, the cost is considerably less than a<br />

yearly prescription of tolterodine, and the therapy is completely revers-<br />

ible. Very few side effects were noted in clinical trials (22.25).<br />

Compliance with stimulation regimens is an important issue. A study<br />

of daily or every-other-day stiinulation for a treatment period of 20<br />

wk found that either regimen was effective in treating stress, urge, and<br />

mixed incontinence (22,25). Compliance was found to be significantly<br />

higher for the every-other-day protocol (up to 93% during the 20-wk<br />

evaluation period). The studies also demonstrated that a period of up<br />

to 12 wk was needed to determine if a patient would benefit from<br />

the stimulation. Compliance over longer periods of use has not been<br />

assessed. To maintain improvement, long-term attention must be paid<br />

to pelvic floor behavior, making the issue of compliance critical to the<br />

ultimate success of this therapy (22).<br />

Needle Stimulation Devices<br />

Needle stimulation for the treatment of urinaiy incontinence is an<br />

adaptation of transcutaneous electrical nerve stimulation (TENS). A<br />

stainless steel needle is inserted approx 5 cni cephalad from the medial


Chapter 13 I Treatment of Detrusor Instability 305<br />

malleolus just posterior to the margin of the tibia and then advanced<br />

to the medial edge of the fibula. Maximal functional electrical stimulation<br />

is applied using the posterior tibial nerve to inhibit the sacral<br />

micturition center.<br />

Stimulation via this route has been proposed as the optimal choice<br />

for patients with neurological lesions, in whom complete pelvic floor<br />

rehabilitation is not a clinical goal (32). Results from one study of 90<br />

patients using needle electrical stimulation has been reported (33).<br />

Patients with various voiding disturbances were treated once weekly<br />

for 20-30 min per session for 10 consecutive wk. Successful outcomes<br />

(at least 50% improvement in symptoms documented by voiding diaries)<br />

were reported for 81% of treated patients.<br />

The length of the trial period to assess efficacy of treatment is<br />

not known at this time. As with transvaginalltransanal stimulation,<br />

compliance is an issue of critical concern. In the case of the reported<br />

form of needle stimulation, patients must receive treatment during an<br />

outpatient office visit one to two times per week. To be desirable, it<br />

must be shown that a sustained benefit can be obtained without ongoing<br />

office-based treatment.<br />

Sacral Nerve Stimulation<br />

Sacral nerve stimulation (SNS) therapy acts on the neural reflexes<br />

of the bladder at the level of the S3 sacral nerves. The therapy is based<br />

on conclusions from animal experimentation and electrophysiologic<br />

studies that electrical stimulation of sacral nerves can modulate neural<br />

reflexes that influence bladder, sphincter, and pelvic-floor behavior<br />

(8,14,15,18,34) (see Fig. 2).<br />

The effects of SNS depend on the electrical stimulation of afferent<br />

axons in the spinal roots, which in turn modulate voiding and continence<br />

reflex pathways in the CNS (10). Electrical stimulation of the S3 nerve<br />

ramus is optimal because it contains the sensory fibers from the genitals<br />

and perineum, afferent and efferent fibers from the anterior part of the<br />

levator ani and urethral sphincter, and autonomic fibers from the detru-<br />

sor (15,35). Neuromodulation at this level involves unmyelinated fibers<br />

so that very low frequencies, around 10-25Hz, can be utilized.<br />

An SNS device consists of a lead implanted adjacent to a targeted<br />

sacral nerve, a pulse generator (PG) implanted in the lower abdomen<br />

or upper buttocks, and an extension that connects the lead to the IPG.<br />

Electrical pulses from the IPG are transmitted through the extension


306 Siege1<br />

and lead to the targeted sacral nerve via electrodes located at the distal<br />

end of the lead.<br />

TEST STIMULATION<br />

The first stage in the implantation of the device is test stimulation<br />

using a temporary electrode. ‘Unlike other electrical stimulation modal-<br />

ities, SNS provides the opportunity to test potential candidates, and to<br />

select the patients most suitable for long-term therapy (Fig. 3). During<br />

acute stimulation, a temporary electrode is used to determine the func-<br />

tional integrity of the sacral nerves. A successful response to subchronic<br />

test stimulation indicates that the patient’s symptoms are owing to<br />

CNS <strong>dysfunction</strong>, and are treatable with neuromodulation. Studies have<br />

reported constant reproducibility of the results of the subchronic test<br />

Stimulation compared to those of the surgical neuroprosthetic<br />

implant (36).<br />

TEMPORARY ELECTRODE IMPLANTATION<br />

Patients are placed in the prone position with pillows under the chest<br />

and abdomen to flatten the back and draped in a sterile manner. Bony<br />

landmarks are used to define the level of the sacral foramina on either<br />

side. The S3 is located at the level of the greater sciatic notch, one<br />

fingerbreath lateral to the midline spinous process. After infiltration of<br />

the skin and subcutaneous tissue with 1 % lidocaine, the posterior surface<br />

of the bony sacrum is probed with an insulated needle. Each plane of<br />

resistance, including the subcutaneous tissue, fascia, and posterior sacral<br />

surface is flooded with local anesthetic. The probing needle is oriented<br />

to follow the natural course of the sacral ramus through the bone table.<br />

When the foramen is localized, the spinal needle will drop off the<br />

posterior surface of the sacrum into the sacral foramen.<br />

Once the electrically insulated needle has been positioned for acute<br />

Stimulation, a graduated current amplitude is applied to the nerve to<br />

determine responses consistent with an S3 pattern. Because S3 is primarily<br />

responsible for levator function and has less contribution to the<br />

motor function of the lower extremity, stimulation at this level is<br />

preferable. The S3 responses are deepening of the buttocks groove<br />

(“bellows” response) and plantar flexion of the great toe only. When<br />

these responses have been discerned, a test stimulation lead is threaded<br />

through the needle lumen, and the needle is withdrawn, leaving the


Chapter 13 I Treatment of Detrusor Instability 307<br />

Fig. 3. Surgical Implantation of the SNS device. The lead is maintained by securing<br />

to the posterior sacral periosteum. The proximal portion of the electrode is routed<br />

toward the flank or upper buttock using an extension. The pulse generator is<br />

implanted in an abdominal pouch.<br />

lead in place. The portion of the test stimulation lead above the skin<br />

is secured using a breathable membrane dressing and position of the<br />

electrode confirmed by antero-posterior and lateral sacral radiograph.<br />

If the positioning of the temporary lead is appropriate and S3<br />

responses are obtained, a subchronic test is performed. The proximal<br />

end of the lead is connected to an external stimulator and a test period<br />

of 3-7 d follows. Patient responses to stimulation sensations aid in<br />

determining the electrical stimulus patterns that meet their individual<br />

needs and changes in incontinent symptoms are quantified in the voiding<br />

diary. Patients are contraindicated for implant if, during the test period,<br />

SNS is not found satisfactory in alleviating target symptoms, if the<br />

patient is unable to operate the device, or the treatment is not acceptable<br />

to the patient. The test stimulation lead is removed, and the patient is<br />

free to explore other treatment options. Patients who experience a 50%<br />

reduction of symptoms in at least one key symptom variable such as<br />

number of incontinence episodes, pads, or severity scores are considered<br />

candidates for long-term therapy.


308 Siege1<br />

SURGICAL IMPLANTATION<br />

The second stage, surgical implantation, is performed under general<br />

anesthesia. The patient receives prophylactic intravenous antibiotics,<br />

and is then placed in a prone position supported by a laminectomy<br />

frame. During surgery, the feet will remain exposed so that nerve<br />

responses may be assessed. A 6-12-cm midline incision is made over<br />

the spinous process to the level of the underlying lumbodorsal fascia,<br />

which is then cleaned and incised longitudinally. The underlying paraspinous<br />

muscle is split in the direction of its fibers, and with blunt<br />

dissection the posterior surface of the sacrum is exposed. An insulated<br />

needle is inserted into the foramen in the direction of the course of<br />

the nerve, and proper S3 responses are again identified as in the test<br />

stimulation. If necessary the position of the needle may be changed to<br />

obtain appropriate responses and, when determined, removed to allow<br />

insertion of a surgically implanted lead into the needle puncture site<br />

to a depth limited by a fixation cuff. The distal end of the lead contains<br />

four electrical contact points, which are tested intraoperatively to confirm<br />

the response obtained during prior test stimulation procedures.<br />

The perineum and foot are observed carefully for typical motor<br />

responses, and if necessary, the electrode may be removed and reinserted<br />

to improve the response. Ideally, the response at the perineum<br />

should be greater than that at the great toe. When satisfactory responses<br />

are obtained, the position of the lead is maintained by securing a preattached<br />

fixation cuff to the posterior sacral periosteum. The proximal<br />

portion of the electrode is routed toward the flank or upper buttock.<br />

A subcutaneous pouch for the IPG is created in the upper buttock,<br />

lateral to the edge of the sacrum, and below the posteror-superior iliac<br />

crest. The pouch is created in a position such that belts or clothing will<br />

not put pressure on the area. The IPG and the electrode are connected by<br />

an extension lead tunneled in the subcutaneous tissue between the<br />

midline and upper buttock incisions.<br />

CLINICAL EFFICACY<br />

A randomized, multicenter, clinical trial conducted in the US, Can-<br />

ada, and Europe evaluated the safety and efficacy of SNS therapy in<br />

three different patient populations: urge incontinence (155 patients),<br />

urgency-frequency (220 patients) and urinary retention (177 patients).<br />

Study inclusion criteria required that patients be refractory to conserva-<br />

tive forms of medical treatment. After a successful test stimulation


Chapter 13 I Treatment of Detrusor Instability 309<br />

procedure, qualified patients were randomized into one of two treatment<br />

groups: a stimulation group (treatment) and a delay group (control).<br />

The delay group served as the control arm of the study, and continued<br />

to use conservative treatments to manage their symptoms for a period<br />

of 6 mo. Control patients were then allowed to cross over to the<br />

treatment am of the study upon conclusion of the delay period. <strong>Voiding</strong><br />

diasies were used to collect efficacy outcome info~ation on accepted<br />

measures for each of the three patient populations. Health-related quality<br />

of life was also examined. Efficacy was evaluated by comparing<br />

outcomes of the treatment and control groups at 6 mo.<br />

Urge Incontine~ce. Of the 155 patients diagnosed with urge incontinence<br />

who underwent test stimulation, 98 qualified for implantation;<br />

58 patients were implanted and had data at 6 mo. Results of group<br />

sequential data analysis demonstrated that, compared to the control<br />

group, patients in the treatment group demonstrated clinically and statistically<br />

significant reductions in the frequency of urge incontinent episodes,<br />

the severity of leaking episodes, and the use of absorbent pads/<br />

diapers. At 6 mo, 74% of the treatment group patients reduced the<br />

frequency of incontinent episode by greater than 50%; 47% of these<br />

patients were completely dry. Of the treatment group patients who<br />

experienced heavy leaking at baseline, 77% had eliminated these types<br />

of leaks at 6 mo postimplant. The results were sustained at 12-1310 and<br />

18-1110 postimplant. Analysis of SF-36 results indicated statistically<br />

significant improvements in the domains of Physical Functioning, Vitality<br />

and General Health (20).<br />

Ur~ency~requen~y. Of the 220 patients with urgency-frequency,<br />

80 patients qualified for surgical implantation following a successful<br />

test stimulation procedure; 47 were randomized into the treatment<br />

group and 33 into the control group. At 6 mo postimplant, statistically<br />

significant reductions were documented in the treatment group with<br />

respect to the three primary diary variables: number voids/day, the<br />

volumehoid, and the degree of urgency ranking. 88% of the treatment<br />

group patients documented clinical success on these measures compared<br />

with 32% of patients in the control group. Patients in the treatment<br />

group demonstrated a significant reduction in pelviclbladder discomfort<br />

at 6 m0 and also documented clinically and statistically significant<br />

changes in the ability to store urine, the ability to empty urine, and<br />

a decrease in incontinent episodes per day; control group patients<br />

documented no statistically-significant changes in these parameters.<br />

Sustained clinical benefit was documented at 12 and at 18 m0<br />

post-implant.


310 Siege1<br />

Six months postimplant treatment group patients demonstrated sig-<br />

nificant improvements in 7 of the 10 health-related quality of life<br />

domains measured by the SF36 (Physical Functioning, Role Physical,<br />

Bodily Pain, General Health, Vitality, Social Functioning, and Mental<br />

Health). Implant patients also had more favorable perceptions of their<br />

general health status over time as compared to control patients (20).<br />

Retention. Of the 177 patients presenting with urinary retention, 68<br />

patients qualified for surgical implantation following a successful test<br />

stimulation procedure; 37 were randomized into the treatment group<br />

and 3 1 into the control group. Treated patients documented significant<br />

reductions in catheter volumes at 6-mo postimplant; 69% of the treated<br />

patients completely eliminated catheterization, with an additional 14%<br />

demonstrating 50% reduction in catheter volumes. Successful results<br />

were therefore achieved by 83% of the treated patients. Patients in the<br />

control group remained clinically unchanged with only 9% demonstra-<br />

ting a clinically-meaningful change in retention symptoms. There were<br />

also significant improvements in the ability to empty urine (decreased<br />

number of catheterizations per day, decreased total catheter volume<br />

per day, decreased maximum catheter volume); the ability to void urine<br />

(increased number of voids per day, increased total volume voided per<br />

day, increased average volume voided per void, increased maximum<br />

voided volume, improved urine stream force j; and increased patient<br />

comfort among implanted patients. Evidence of improved bladder func-<br />

tion was further demonstrated by the statistically significant reduction<br />

in urinary tract infections for the treatment group at 12 mo postimplant.<br />

Sustained clinical benefit was documented at 12 mo and at 18 mo post-<br />

implant.<br />

At the 6-mo follow-up analysis implant patients demonstrated sig-<br />

nificant improvements in Bodily Pain, as measured by the SF36 and<br />

had more favorable perceptions of their general health transition com-<br />

pared to control patients (20).<br />

The efficacy of SNS in the treatment of voiding <strong>dysfunction</strong> has<br />

been demonstrated and reported in earlier literature (13,34). Published<br />

results of clinical trials using the InterStimB device suggest that SNS<br />

in a safe and effective treatment alternative for a variety of urinary<br />

incontinence problems (8,15,18,36,37). The FDA approved the use of<br />

the Inters timB Continence Control System for the indication of refrac-<br />

tory urge incontinence in 1997. The Food and Drug Administration<br />

(FDA) is currently reviewing data for expanded indications of urgency-<br />

frequency and urinary retention. The device has been available for<br />

these indications in Canada, Europe, and Australia since 1994.


Chapter 13 / Treatment of Detrusor Instability 311<br />

Adverse events related to the therapy, the device, or implant proce-<br />

dure have been comprehensively recorded in the multicenter trial and<br />

include: pain at the lead implant site (21%) or at the IPG site (17%),<br />

which may require surgical revision but are resolvable. The probability<br />

of surgical revision. computed from survival analysis of clinical trial<br />

data, was 29% within the first 6 mo, and 12% in the second 6 nio,<br />

suggesting that the potential for surgical revision declines over time.<br />

It is likely that further refinements in technique, such as positioning<br />

of the IPG in the upper buttocks instead of the lower abdomen, and<br />

monitoring of sacral evoked response potentials during lead placement,<br />

will reduce the need for surgical revision in the future.<br />

Other adverse events associated with use include infectionhkin irri ta-<br />

tion (7%), technical problems (7%), and transient increases in electrical<br />

sensation (6%). Changes in bowel function (5%), nuinbness (1.3%),<br />

and suspected nerve injury tnay result during chronic treatment (


3 12 Siege1<br />

therapies and surgery. While it is reasonable to attempt other conservative<br />

treatments first, when these options fail or are otherwise not indicated,<br />

electrical stimulation should be considered before offering surgical<br />

treatment. Therapies involving electrical stimulation are based on<br />

the concept that urge incontinence owing to detrusor instability is<br />

not m end organ problem, but that the condition represents a CNS<br />

<strong>dysfunction</strong>. The problem, in turn, may be manifest as symptoms of<br />

urge incontinence, urgency-frequency syndromes, urinary retention, or<br />

bowel <strong>dysfunction</strong> (38). Stimulation therapies are able to modulate<br />

abnormal reflexes between the bladder, pelvic floor, and external sphincter.<br />

Unlike other electrical stimulation techniques with more limited<br />

applications, SNS holds great promise for a large number of patients<br />

who suffer a spectrum of lower urinary tract <strong>dysfunction</strong>. Implantation<br />

of an SNS device in properly selected patients is likely to provide<br />

significant and sustained relief from debilitating eurologic symptoms,<br />

is totally reversible, and does not preclude the use of other standard treatments.<br />

REFERENCES<br />

1. Fantl JA, Newman DK, Colling J, et al. (1996) Urinary Incontinence in Adults:<br />

Acute and Chronic Management, Clinical Practice Guideline No. 2 1996 Update.<br />

Rockville, Maryland: U.S. Department of Health and Human Services, Public<br />

Health Service, Agency for Health Care Policy and Research, AHCPR Publication<br />

No. 96-0682, March.<br />

2. Fall M (1985) Electrical pelvic floor stimulation for the control of detrusor instability.<br />

Neurourol Urodyn 4:329-335.<br />

3. Fall M, Lindstrom S (1991) Electrical stimulation: a physiologic approach to the<br />

treatment of urinary incontinence. Urd Clin NA 18(2):393-407.<br />

4. Wein AJ, Bmet DM (1988) <strong>Voiding</strong> Function and Dysfinction: A hgical and<br />

Practical Approach. Chicago: Yearbook Medical Publishers, Inc.<br />

5. Schmidt RA, Senn E, Tanagho EA (1990) Functional evaluation of sacral nerve<br />

root integrity: report of a technique. Urology 3.5(5):388-392.<br />

6. Schmidt RA (1988) Applications of neurostimulation. Neurourol Urodyn 7:585.<br />

7. Tanagho EA, Schmidt RA (1988) Electrical stimulation in the management of<br />

the neurogenic bladder. J Urd 140:1331.<br />

8. Thon W, Baskin L, Jonas U, et al. (1991) Neuromodulation of voiding <strong>dysfunction</strong><br />

and pelvic pain. World J Urd 9:138-141.<br />

9. Mersdorf A, Schmidt RA, Tanagho EA (1993) Topographic-anatomical basis of<br />

sacral neurostimulation: neuroanatomical variations. J Urol 149:345-349.<br />

10. Chancellor MB, deGroat WC (Submitted for publication) Hypotheses on how<br />

sacral nerve stimulation works for the treatment of detrusor overactivity and<br />

urinary retention.


Chapter 13 / Treatment of Detrusor Instability 313<br />

11. Blok BFM, van Maarseveen JTPW, Holstege G (1998) Electrical stimulation of<br />

the sacral dorsal gray commissure evokes relaxation of the external urethral<br />

sphincter in the cat. Neurosci Letters 249:68-70.<br />

12. Schultz-Lampel D, Jiang C, Lindstrom S, Thuroff JW (1998) Experimental results<br />

on mechanisms of action of electrical neuromodulation in chronic urinary retention.<br />

World J Urol 16:301-304.<br />

13. Vapnek JM, Schmidt RA (1991) Restoration of voiding in chronic urinary retention<br />

using the neuroprosthesis. World J Urol 9:142-144.<br />

14. Siegel SW (1992) Management of voiding <strong>dysfunction</strong> with an implantable neuroprosthesis.<br />

Urol Clin N Am 19(3):163-170.<br />

15. Diijkema H, Weil EHJ, Mijs P, Janknegt RA (1993) Neuromodulation of sacral<br />

nerves for incontinence and voiding <strong>dysfunction</strong>s. Eur Urol 24:72-7’7.<br />

16. Hassouna M (1994) Neural stimulation for chronic voiding <strong>dysfunction</strong>. J Urol<br />

153:2078-2080.<br />

17. Koldewijn EL, et al. (1994) Predictors of success with neuromodulation in lower<br />

urinary tract <strong>dysfunction</strong>: results of trial stimulation in 100 patients. J Urol<br />

15212071-2075.<br />

18. Bosch J, Groen J (1995) Sacral (S3) Segmental nerve stimulation as a treatment for<br />

urge incontinence in patients with detrusor instability: results of chronic electrical<br />

stimulation using an implantable neural prosthesis. J Urol 154:504-507.<br />

19. Bosch J, Groen J (1996) Treatment of refractory urge urinary incontinence with<br />

sacral spinal nerve stimulation in multiple sclerosis patients. Lancet 348:717-719.<br />

20. Medtronic data on file; MDT- 103.<br />

21. Empi (1994) The ~ndamentals of pelvic ftoor stimulation: innovative treatments<br />

for urinary incontinence. St. Paul, MN: Empi, Inc.<br />

22. Siegel SW, Richardson DA, Miller ILL, Karram MM, et al. (1997) Pelvic floor<br />

electrical stimulation for the treatment of urge and mixed urinary incontinence in<br />

women. Urology 50(6):934-940.<br />

23. Bent AE, Sand PK, Ostegard DR, Brubaker L (1993) Transvaginal electrical<br />

stimulation in the treatment of genuine stress incontinence and detrusor instability.<br />

IntE Urogynecol J 4:9-13.<br />

24. Sand PK, Richardson DA, Staskin DR, al. et (1 995) Pelvic floor electrical stimulation<br />

in the treatment of genuine stress incontinence: a multicenter, placebocontrolled<br />

trial. Am J Obstet Gynecol 183:72-79.<br />

25. Richardson DA, Miller ILL, Siegel SW, et al. (1996) Pelvic floor electrical stimulation:<br />

a comparison of daily and every-other-day therapy for genuine stress incontinence.<br />

Urology 48( 1):llO-118.<br />

26. Elgamasy AN, Lewis V, Hassouna ME, Ghoniern GM Effect (1996) of transvaginal<br />

stimulation in the treatment of detrusor instability. Urol Nurs 16(4):127-130.<br />

27. Brubaker L, Benson JT, Bent A, et al. (1997) Transvaginal electrical stimulation<br />

for female urinary incontinence. Am J Obstet Gynecol 177:536-540.<br />

28. Luber KM, Wolde-Tsadik G (1997) Efficacy of functional electrical stirnuIation<br />

in treating genuine stress incontinence: a randomized clinical trial. Neurourol<br />

Urodyn 16543-55 1.<br />

29. Kulseng-Hanssen S, Kristoffersen M, Larsen E (1998) Evaluation of the subjective<br />

and objective effect of maximal electrical stimulation in patients complaining of<br />

urge incontinence. Acta Obstet Gynecol Scand 168(Suppl):12-15.<br />

30. Bo K (1998) Effect of electrical stimulation on stress and urge urinary incontinence.<br />

Clinical outcome and practical recommendations based on randomized controlled<br />

trials. Acta Obstet Gynecol Scand 168 (Suppl):3-11.


3 14 Siege1<br />

31.<br />

32.<br />

33.<br />

34.<br />

35.<br />

36.<br />

37.<br />

38.<br />

Davila GW, Bernier F (1995) Multimodality pelvic physiotherapy treatment of<br />

urinary incontinence in adult women. Intl Urogynecol J 6:187-194.<br />

McGuire EJ, Shi-Chun Z, Horwinski ER, et al. (1983) Treatment of motor and<br />

sensory detrusor instability by electrical stimulation. J Urol 129:78-79.<br />

Stoller ML (1998) Afferent nerve stimulation for pelvic floor <strong>dysfunction</strong>.<br />

J Endourol 12( 1):S108 (Abstract F2-6).<br />

Elabbady AA, Hassouna MM, Elhilali MM (1994) Neural stirnulation for chronic<br />

voiding <strong>dysfunction</strong>s. J Urol 152:2076-2080.<br />

Appell RA (1998) Electrical stimulation for the treatment of urinary incontinence.<br />

Urology 51 (2A Suppl):24-26.<br />

Shaker HS, Hassouna M (1998) Sacral nerve root neuromodulation: an effective<br />

treatment for refractory urge incontinence. J Urol 159: 15 16-15 19.<br />

Shaker HS, Hassouna M (1998) Sacral root neuromodulation in idiopathic nonob-<br />

structive chronic urinary retention. J UroZ 159: 1476-1478.<br />

Stadelmaier MKE, Hohenfellner M, Gall FP (1995) Electrical stimulation of sacral<br />

spinal nerves for treatment of faecal incontinence. Lancet 346(8983):1124-1.127.


l 4 Treatment of Detrusor Instability<br />

With Augmentation Cystoplasty<br />

Joseph M. Khoury<br />

CONTENTS<br />

CLASSIFICATION OF CYSTOPLASTY<br />

INDICATIONS FOR AUGMENTATION CYSTOPLASTY<br />

CONTRAINDICATIONS TO AUGMENTATION<br />

CYSTOPLASTY<br />

SURGICAL CHOICES IN ACHIEVING THE COALS<br />

OF BLADDER RECONSTRUCTION<br />

COMPLICATIONS<br />

REFERENCES<br />

Augmentation cystoplasty can provide significant relief of refractory<br />

storage symptoms such as frequency, urgency, and urge incontinence<br />

in those patients who have failed conservative measures such as pharmacotherapy,<br />

behavioral techniques, and minimally invasive procedures<br />

such as neurornodulation. The concept of augmenting bladder capacity<br />

is not a new one. Von Mikulicz in 1899 was the first to use ileum to<br />

restore normal bladder capacity in humans (1). Other investigators at<br />

the turn of the century used different segments of bowel; however,<br />

enthusiasm waned until the early 1950s, when pioneers such as Kuss<br />

(21, Couvelaire (3), and Gil-Vernet (4) restored the impetus to pursue<br />

bladder reconstruction. The primary indication for augmentation enterocystoplasty<br />

at that time was tuberculous cystitis, which caused a stmcturally<br />

contracted bladder. Unfortunately, the postoperative mortality<br />

and morbidity was exceedingly high and other supravesical diversions<br />

such as the ileal conduit urostomy became the technique of choice in<br />

From: Current Clinical Urology: <strong>Voiding</strong> Dysfinction: Diagnosis and Treatment<br />

Edited by: R. A. Appell 0 Humana Press Inc., Totowa, NJ<br />

315


316 aoury<br />

the urologic community. The long-term sequelae of the Bricker loop<br />

became manifest after several years and included recurrent urinary<br />

infections, dissatisfaction with a wet urostomy, and the psychosocial<br />

stigma associated with an external collection device (5,6).<br />

In the 1960s a new surge of enthusiasm developed in reconstructive<br />

urology. Fluoroscopic multichannel urodynamics provided the scientific<br />

foundation for establishing the urophysiologic mechanisms through<br />

which the detrusor and outlet function in a coordinated manner. Likewise,<br />

Lapides’ hallmark contribution in 19’72 of clean intermittent selfcatheterization<br />

(clean intermittent catheterization; CIC) revolutionized<br />

the management of voiding <strong>dysfunction</strong>, particularly in patients with<br />

neuropathic bladder (7,8). In addition, the discovery of pharmacological<br />

agents to treat abnormal bladder storage and emptying characteristics<br />

established a new avenue for managing urinary incontinence in conjunction<br />

with intermittent self-catheterization (9). In spite of all the new<br />

technology and drugs to manage urinary incontinence, there are patients<br />

who remain refractory to such therapy and in whom surgery is necessary<br />

to correct their underlying disorder. Augmentation enterocystoplasty<br />

is a safe operation for this purpose.<br />

CLASSIFICATION OF CYSTOPLASTY<br />

Augmentation cystoplasty is a procedure used to increase bladder<br />

capacity by incorporating a segment of intestine or stomach into an<br />

unresected bladder. Augmentation cystoplasty is used commonly to<br />

manage intractable symptoms resulting from detrusor instability hyperflexia<br />

or impaired detrusor compliance.<br />

Substitution cystopZasty is an operation whereby a significant amount<br />

of bladder wall is resected prior to incorporating an intestinal segment<br />

with the bladder remnant. This procedure is used commonly in hypersensitive<br />

bladder states such as interstitial cystitis. The bladder is usually<br />

resected to its trigonal remnant and an isolated intestinal segment is<br />

then substituted for the resected bladder.<br />

Replacement or orthotopic cystoplasty is reserved for those cases<br />

in which a total cystectomy is necessary as in patients undergoing<br />

radical Cystectomy for bladder cancer. The urethra and distal sphincter<br />

mechanism are left intact so that the intestinal reservoir can be anastomosed<br />

to it.<br />

Autoaugmentation is a relatively new surgical procedure in which<br />

the detrusor muscle is dissected off from the bladder usually on its


Chapter 14 / Augmentation Cystoplasty 3 17<br />

anterior, lateral, and superior surfaces creating a large mucosal bulge<br />

or wide-mouth diverticulum. It obviates the potential complications of<br />

incorporating bowel segments into the urinary system and, therefore,<br />

is an alternative procedure in those patients who may be a candidate for<br />

augmentation enterocystoplasty. This procedure has also been termed<br />

detrusor myomectomy.<br />

INDICATIONS FOR AUGMENTATION CYSTOPLASTY<br />

The major indications to perform enterocystoplasty have changed<br />

considerably since the early 1950s. Antituberculous drugs have significantly<br />

decreased the end-stage manifestations of tuberculous cystitis,<br />

and, therefore, bladder reconstruction is rarely needed. In contemporary<br />

times, cystoplasty is now performed for detrusor hyperactivity resulting<br />

from such entities as poor bladder wall compliance, refractory detrusor<br />

instability or hyperreflexia, and a structurally reduced bladder capacity<br />

owing to inflammatory disease. Regardless of the etiology, these bladders<br />

are contracted and bladder capacity reduced, resulting in the symptoms<br />

of frequency, urgency, and urge incontinence. As noted earlier, the<br />

majority of patients with bladder overactivity and/or impaired detrusor<br />

compliance respond to anticholinergic agents; however, end-stage<br />

inflammatory bladders, particularly with thick-walled fibrosis do not,<br />

and surgical intervention is often necessary. In addition to the bladder<br />

other variables that may need to be addressed include the upper tracts<br />

and the bladder outlet. The presurgical evaluation will determine those<br />

variables that will need to be addressed and determine the magnitude<br />

of the urinary reconstruction.<br />

PREOPEMTIVE EVALUATION<br />

The preoperative surgical evaluation should include a thorough his-<br />

tory and physical examination to determine the underlying etiology of<br />

the bladder <strong>dysfunction</strong>, as well as laboratory, radiographic, endoscopic,<br />

and urodynmic studies. It is largely upon these results that the surgeon<br />

will decide whether the patient requires augmentation or substitution<br />

cystoplasty and what appropriate management of the ureters and outlet<br />

should be.


318 Khoury<br />

LABORATORY STUDIES<br />

The preoperative laboratory evaluation is necessary to assess renal<br />

function and acid-base status. Serum creatinine, a blood urea nitrogen<br />

(BUN), and 24" creatinine clearance are helpful measures to assess<br />

renal status, particularly in those individuals who may have renal insuf-<br />

ficiency or chronic renal failure. Urinalysis and culture should be per-<br />

formed to identify urinary infection, which should be eradicated prior<br />

to reconstruction. Urinary cytology may be helpful in those patients<br />

with sensory urgency in order to exclude carcinoma in situ.<br />

RADIOLOGIC STUDIES<br />

The entire urinary tract needs to be imaged prior to reconstruction.<br />

Upper tract evaluation will initially begin with intravenous urogram to<br />

assess the upper tracts, particularly if ureteral reconstruction is neces-<br />

sary. <strong>Voiding</strong> cystography will identify bladder abnormalities that may<br />

influence a decision regarding bladder substitution rather than simple<br />

augmentation and will identify the presence and severity of vesicoure-<br />

teric reflux. A cystogram may help to characterize the competence of<br />

the sphincter mechanisms and lower urinary tract patency.<br />

URODYNAMIC EVALUATION<br />

Urodynamic studies should evaluate bladder and outlet function<br />

during filling, storage, and voiding. This is best accomplished by using<br />

multichannel fluorourodynamics, which combine electronic urody-<br />

namic data collection and simultaneous cystourothrography. Filling<br />

cystometrogram will assess compliance, stability, storage, sensation,<br />

and capacity. Storage after filling will further define stability and compe-<br />

tence of the outlet. Pressure flow studies will document voiding dysfunc-<br />

tion and the Valsalva leak-point pressure can be used to evaluate the<br />

bladder outlet for intrinsic sphincter deficiency (10). Other urodynamic<br />

tools such as the urethral pressure profile, electromyography, and fluo-<br />

roscopy may further enhance information regarding the continence<br />

mechanisms.<br />

CYSTOURETHROSCOPY<br />

Endoscopic evaluation should precede all urinary tract reconstructive<br />

procedures to identify any structural bladder or urethral problems that


Chapter 14 / Augmentation Cystoplasty 319<br />

may alter the choice of the proposed reconstructive procedure. Signifi-<br />

cant outlet findings, urethral stricture disease, or trauma may help<br />

predict continence or catheterization difficulties, particularly in those<br />

individuals who have had prior outlet procedures performed. Finally,<br />

cystourethroscopy will identify any concomitant bladder pathology<br />

such as tumors, stones, or diverticuli that need to be identified prior<br />

to the reconstruction.<br />

CONTRAINDICATIONS TO<br />

AUGMENTATION CYSTOPLASTY<br />

There are relatively few contraindications to cystoplasty. Compro-<br />

mised renal function in the past was initially considered an absolute<br />

contraindication to bladder reconstruction as it could exacerbate meta-<br />

bolic problems owing to the absorptive surface of the bowel within<br />

the urinary tract. However, if the etiology of the underlying renal<br />

impairment is owing to hostile detrusor function, then augmentation<br />

cystoplasty may stabilize or improve renal function. It is of utmost<br />

importance to counsel the family that renal deterioration may continue<br />

even after bladder reconstruction and result in the need for renal trans-<br />

plantation.<br />

A well-motivated and compliant patient is imperative as bladder<br />

reconstruction surgery can be fraught with relatively severe postopera-<br />

tive problems should timely follow-up and review not be kept. From<br />

the outset the patient and family should be well aware that intermittent<br />

self-catheterization might be required on a routine basis and that devia-<br />

tion from this routine could precipitate severe injury to the urinary<br />

tract or metabolic problems such as uremia.<br />

SURGICAL CHOICES IN ACHIEVING THE GOALS<br />

OF BLADDER IXECONSTRUCTION<br />

The goals of bladder reconstruction are to create a low-pressure,<br />

capacious reservoir that is continent and guards the upper tract from<br />

the damaging effects of reflux, infection, and high intravesicle pressure.<br />

Therefore, the surgeon not only needs to consider reconstruction of<br />

the bladder reservoir but also whether ancillary procedures are necessary<br />

to correct reflux and outlet incompetence. These three variables will<br />

be addressed individually.


320 Khoury<br />

MANAGEMENT OF THE BLADDER<br />

Management of the bladder remnant should address the extent of<br />

bladder resection, choice of bowel segment, and reconfiguration of the<br />

bowel segment.<br />

Extent of Bhdder Resection and Preparation<br />

The extent of bladder resection, if at all, is determined by the pathology<br />

of the underlying disease for which cystoplasty is being performed.<br />

Patients with an overactive detrusor owing to neurologic or non-neurologic<br />

disease may benefit from simple augmentation without bladder<br />

resection. In this situation the bladder is incised, either sagittally or<br />

longitudinally, and a segment of detubularized bowel is incorporated<br />

into the valved bladder. Partial or supratrigonal cystectomy is recommended<br />

when the bladder is structurally diseased or symptomatic as<br />

in interstitial cystitis, or in those cases in which the bladder diverticulae<br />

are present or when the bladder is extremely thick-walled.<br />

When performing augmentation cystoplasty without bladder resection,<br />

the main objective is to achieve as long an anastomosis as possible<br />

with the intestinal segment, avoiding an hourglass configuration with<br />

the bowel patch acting as a diverticulum. This goal may be best accomplished<br />

by incising the bladder in the sagittal plane, which simplifies<br />

the procedure by avoiding dissection of the lateral pelvic walls and<br />

bladder incision close to the intramural ureters. This incision is begun<br />

approximately 2 cm cephalad to the bladder neck anteriorly and extends<br />

posteriorly 2 cm above to the interureteric ridge. This defect is then<br />

augmented using either a single segment of open bowel or commonly<br />

a cup patch (11).<br />

Choice of Bowel Segment<br />

Almost all segments of the gastrointestinal tract, including stomach,<br />

have been used for bladder augmentation. Each bowel segment has its<br />

own advantages and disadvantages, and its own strong proponents<br />

(12-14). Several factors influence the type of bowel employed in reconstruction.<br />

These include availability, anatomic considerations, impact<br />

on nutrition, bowel and renal function, the need to preserve the ileocecal<br />

valve, particularly in patients with myelodysplasia, and finally, the


Chapter 14 I Augmentation Cystoplasty 321<br />

surgeon’s preference. It is my preference to use ileum for simple augmentation;<br />

however, in some cases, particularly in patients with myelodysplasia,<br />

sigmoid colon lends itself to bladder reconstruction because<br />

of its redundancy and proximity to the bladder.<br />

Bowel Sepent Reconfiguration<br />

The ultimate outcome in cystoplasty reconstruction is a large-capac-<br />

ity, low-pressure system with normal compliance. This is best accom-<br />

plished by remodeling a tubular section of intestine into a sphere,<br />

thereby not only increasing capacity but also disrupting the normal<br />

propagation of peristaltic contractions so as to keep cystoplasty activity<br />

to a minimum (13,1543). The volume within a sphere is determined<br />

by its radius (V = 4/3R3); therefore, the surgical technique chosen<br />

should increase the overall radius of the system. In any event,<br />

cystoplasty size should achieve a total bladder volume of 500 cc to<br />

ensure a normal functional capacity of the reconstructed bladder.<br />

MANAGEMENT OF THE URETERS AND<br />

UPPER TRACTS<br />

Generally speaking, if reflux is present at the time of cystoplasty,<br />

then ureteral reimplantation is usually necessary. There are some reports<br />

indicating that vesicoureteral reflux is based solely on hostile detrusor<br />

factors and if a low-pressure reservoir is created through bladder recon-<br />

struction, the reflux will then resolve (19). Likewise, this has been the<br />

author’s experience for those patients with lower grades of reflux;<br />

however, those patients with either unilateral or bilateral Grade IV or<br />

V reflux may be better served with ureteral reimplantation. A high<br />

number of myelodysplastic patients undergoing cystoplasty will require<br />

simultaneous reimplantation (20). It is less common in the adult spinal<br />

cord-injured patient with neurogenic bladder <strong>dysfunction</strong>, and relatively<br />

uncommon in non-neurogenic patients with detrusor instability. Factors<br />

such as length of available ureter, peristaltic activity of the ureter, and<br />

nature of the bladder remnant will guide the most appropriate surgical<br />

technique. When the bladder remnant is not heavily trabeculated or<br />

fibrotic, the Cohen cross-trigonal reimplant may correct low-grade<br />

reflux (21). The ileal intesusception nipple technique is optimal for<br />

reimplantation of large, dilated poorly peristaltic ureters for the ileal<br />

intesusception has very low resistance and is unlikely to cause ureteral


322 aoury<br />

obstruction (22). Direct reimplantation into the augmenting bowel seg-<br />

ment has been favorable in those circumstances where the bladder<br />

remnant is either too small or pathology precludes standard reimplanta-<br />

tion. Examples of these include the split cuff nipple as described by<br />

Turner-Warwick (23), the Goodwin (24) and Carney-LeDuc (25) reim-<br />

plantations. All of these are appropriate techniques to implant the ureter<br />

into bowel.<br />

MANAGEMENT OF THE BLADDER OUTLET<br />

Many treatment options are available to manage the incompetent<br />

outlet and these are determined by the type and magnitude of the<br />

incontinence, dexterity of the patient and ambulatory status. Mild bladder<br />

neck incompetence can be treated pharmacologically by using drugs<br />

such as alpha agonists to increase outlet resistance (26). For more<br />

debilitating types of incontinence, a number of surgical procedures are<br />

available and include injection of periurethral bulking agents such as<br />

collagen or polytetrafluoroethylene (2 7), bladder neck reconstruction<br />

(28-30), sling cystourethropexy, or placement of an artificial urinary<br />

sphincter (31).<br />

COMPLICATIONS<br />

Metaliolic and Electrodyte Disorders<br />

Hyperchloremic metabolic acidosis is a potential problem whenever<br />

bowel is interposed within the urinary tract as a conduit or reservoir.<br />

The total surface area exposed to urine and the patients renal function<br />

are the two most important predisposing factors that will potentiate<br />

systemic acidosis. Nurse and Mundy reported on 48 patients who underwent<br />

assessment of acid-base and electrolyte balance following incorporation<br />

of different intestinal segments into the urinary tract (32). All<br />

patients had abnormal blood gases with most demonstrating metabolic<br />

acidosis with respiratory compensation. One-third of their patients had<br />

hyperchloremia. In a follow-up study of 28 patients, Mundy and Nurse<br />

investigated the consequences of metabolic acidosis with respect to<br />

calcium balance and skeletal mineralization (33). All of their patients<br />

had normal serum and 24 h calcium levels. Of the 28 patients, l2 were<br />

women between the ages of 45 and 50 and all had normal dual-photon<br />

absorptiometry bone scans, demonstrating no bone demineralization.


Chapter 14 / Augmentation Cystoplasty 323<br />

Of the 16 children in their series, 10 underwent ileocystoplasty and 6<br />

underwent colocystoplasty. All children with ileocystoplasties continued<br />

to grow within the same percentile as before their cystoplasty;<br />

however, 50% of the children with colocystoplasty demonstrated an<br />

approximately 20% diminution in their postcystoplasty growth rate.<br />

Although there were no dramatic changes in calcium balance or skeletal<br />

~in~ra~ization, the authors concluded that follow-up may be too short<br />

and that these changes may take several years to develop with chronic<br />

metabolic acidosis.<br />

<strong>Voiding</strong> DzficuZties<br />

It is not unusual for patients to experience voiding difficulty or<br />

incomplete bladder emptying after bladder reconstructive procedures<br />

particularly in those undergoing an outlet procedure. Preoperatively<br />

these patients should understand that clean intermittent self-catheterization<br />

may be necessary to allow for complete bladder emptying. Neurologically<br />

intact patients undergoing clam augmentation cystoplasty usually<br />

void efficiently following bladder reconstructive surgery. Those<br />

neurologically impaired patients, though, rarely void normally after<br />

cystoplasty and will require CIC. For those patients who do need<br />

to perform clean intermittent self-catheterization a large catheter (16<br />

French) should be used in order to evacuate mucous plugs that may<br />

obstruct the outlet, particularly in those children and young adults with<br />

myelodysplasia.<br />

Incontinence<br />

When urinary incontinence occurs following bladder reconstructive<br />

surgery urodynamic evaluation is necessary to identify whether the<br />

cystoplasty or the outlet is responsible. Should cystoplasty function be<br />

hyperactive initial therapy with anticholinergic agents may be helpful;<br />

if this fails, further augmentation using a patch of ileum can be successful.<br />

If the outlet is incompetent, then one of the previously discussed<br />

techniques may be needed for its correction.<br />

Urinary Infection<br />

Bacteriuria following augmentation cystoplasty is common and<br />

should be anticipated in those patients on CIC. Pyelonephritis, however,


324 laloury<br />

is uncommon, and if it does occur then imaging studies may be necessary<br />

to rule out ureteral reflux, urinary obstruction, or urinary calculi.<br />

Neoplasms of the bladder and/or bowel segment are relatively rare<br />

following bladder reconstructive surgery. Only 14 cases of neoplasms<br />

have been reported in patients undergoing cystoplasty, and over half<br />

of these operations were performed on tuberculous contracted bladders<br />

(34). The latency period from time of bladder reconstruction to diagnosis<br />

of tumor ranged from 3 to 24 years. The tumor histology was<br />

adenocarcinoma in 7 patients, transitional cell carcinoma in 5, small<br />

cell in l and sarcoma in 1. These and other anecdotal cases further<br />

emphasize the need for chronic surveillance in patients undergoing<br />

such procedures, particularly those in whom gross hematuria may occur.<br />

Filmer and Spencer recommend yearly cytologies and cystoscopy with<br />

biopsy beginning 10 years after enterocystoplasty (35).<br />

Spontuneozts Peforution<br />

Spontaneous perforation following cystoplasty is not an uncommon<br />

occurrence and is being reported more frequently in the literature. It<br />

is important to consider this diagnosis following cystoplasty in those<br />

patients who may have signs and symptoms of an acute abdomen as<br />

it is a potentially lethal complication of enterocystoplasty. The incidence<br />

is reported to be between 3-6.1% (36,37). The diagnosis should be<br />

prompted by a very high index of suspicion, and many times the<br />

diagnosis may be confirmed by retrograde cystogram or radio-<br />

nucleide cystography.<br />

Pregnancy A.er Bhdder Reconstruction<br />

Pregnancy following enterocystoplasty is becoming more common<br />

as women who undergo bladder reconstruction for congenital problems<br />

enter their childbearing years. These women during their pregnancy<br />

could potentially have significant morbidity including febrile urinary<br />

infections, premature labor, urinary tract obstruction, and compromised<br />

renal function. Renal function should be monitored vigilantly in these<br />

women with monthly serum creatinines and, when indicated, renal<br />

ultrasonography. Hill and Gamer (38) recommend that those women


Chapter 14 I Augmentation Cystoplasty 325<br />

who have undergone bladder reconstruction with the continence mechanism<br />

unaltered, be allowed to deliver vaginally. However, those who<br />

have had a bladder neck reconstruction, fascial sling, cystourethropexy,<br />

or artificial urinary sphincter, should undergo Cesarean section to avoid<br />

disruption of the underlying continence mechanism.<br />

1.<br />

2.<br />

3.<br />

4.<br />

5.<br />

REFERENCES<br />

Von Mikulicz J (1899 j Zur Operation der angeborenen Blasenspalte. Zentralbe<br />

Chir 20:64 1-643.<br />

Kuss R, Bitker M, Camey M, et al. (1970) Indications and early and late results<br />

of intestinocystoplasty: a review of 185 cases. J UroE 103:53.<br />

Couvelaire R (1950) La “petite vessie” des tuberculeux geniourinaires: essai<br />

de classification pacelore et variantesioute des cysto-intestine-plasties. J Urd<br />

(Paris) 56:38 1.<br />

Gil-Vernet JM (1965) The ileocolic segment in urologic surgery. J UroE 94:418.<br />

Pitts WR, Muecke EC (1979) A 20-year experience with ileal conduits: the fate<br />

of the kidneys. J Urol 122: 154.<br />

6. Schwarz GR, Jeffs RD (1975) Ilea1 conduit urinary diversion in children: computer<br />

analysis of followup from 2 to 16 years. J Urol 114:285.<br />

7. Diokno AC, Kass E, Lapides J (19’76) A new approach to myelodysplasia.<br />

J Urol 116:771-772.<br />

8. Lapides J, Diokno AC, Silber SJ, et al. Clean intermittent self-catheterization in<br />

the treatment of urinary tract disease. J Urol 107:458.<br />

9. Mulcahy JJ, James HE, McRoberts JW (1977) Oxybutynin chloride combined with<br />

intermittent clean catheterization in the treatment of myelomeningocele patients.<br />

J Urol 118:95.<br />

10. Wan J, McGuire EJ, Bloom DA, and Ritchey ML (1993 j Stress leak point pressure:<br />

a diagnostic tool for incontinent children. J Urol 150:700.<br />

11. Bramble FJ (1982) The treatment of adult enuresis and urge incontinence by<br />

enterocystoplasty. Br J Urol 54:693.<br />

12. Adams MC, Mitchell ME, Rink RC (1988) Gastrocystoplasty: an alternative<br />

solution in the severely compromised patient. J Urol 140:1152-1158.<br />

13. Sidi AA, Reinberg Y, Gonzalez R (1986) Influence of intestinal segment and<br />

configuration on the outcome of augmentation enterocystoplasty. J Urol 136: 1201-<br />

1204.<br />

14. Kuss R (1959) Colocystoplasty rather than ileocystoplasty. J Urol 82:587.<br />

15. Cheng C, Whitfield HN (1990) Cystoplasty: tubularization or detubularization?<br />

Br J Urd 66:30-34.<br />

16. Hinman F (1988) Selection of intestinal segments for bladder substitution: physical<br />

and physiological characteristics. J Urol 139:5 19.<br />

17. Koff S (1988) Guidelines to determine the size and shape of intestinal segments<br />

used for reconstruction. J Urol 140: 1150-15 1.<br />

18. Goldwasser BZ, Barrett DM, Webster GD, et al. (1987) Cystometric properties<br />

of ileum and right colon in patients following bladder augmentation, substitution<br />

and replacement. J UroE 138 (part 2): 1007-1008.


326 fioury<br />

19. Nasrallah PF, Aliabadi HA (1991 ) Bladder augmentation in patients with neurogenic<br />

bladder and vesicoureteral reflux. J Urol 146563.<br />

20. Gittes RE; (1977) Bladder augmentation procedures. In: <strong>Lib</strong>ertino JA, Zinman L,<br />

eds., Reconstructive Urologic Surgery: Pediatric and Adult, Baltimore: Williams<br />

& Wilkins, pp. 216-226.<br />

21. Cohen SJ (1975) Ureterozystoneostomie: eine neue antireflux Tecnik. Aktuel<br />

Urol 6:1.<br />

22. King LR (1987) Protection of the upper tracts in undiversion. In: King LK, Stone<br />

AR, Webster GD, eds. Bladder Reconstruction and Continent Urinary Diversion,<br />

Chicago: Year Book Medical Publishers, pp. 127-153.<br />

23. Kirby RS, Turner-Warwick R (1987) Substitution cystoplasty. In: King LK, Stone<br />

AR, Webster CD, eds. Bladder Reconstruction and Continent Urinary Diversion.<br />

Chicago: Year Book Medical Publishers, pp. 61.<br />

24. Goodwin WE, Harris AP, Kaufmann JJ, Beal JM (1953) Open transcolonic ureterointestinal<br />

anastomosis: a new approach. Surg Gynecol Obstet 97:295-300.<br />

25. LeDuc A, Camey M, Teillac P (1987) An original antireflux ureteroileal implantation<br />

technique: long term followup. J Urd 137:1156.<br />

26. Wein AJ (1987) Lower urinary tract function and pharmacologic management of<br />

lower urinary tract <strong>dysfunction</strong>. Urol Clin N Am 14:273.<br />

27. Lewis RI, Lockhart JL, Politano VA (1984) Periurethral polytetrafluoroethylene<br />

injections in incontinent female subjects with neurogenic urinary incontinence.<br />

J Urol 131:459.<br />

28. Young HH (1922) An operation for the cure of incontinence associated with<br />

epispadias. J Urol 7: l.<br />

29. Dees JE (1949) Congenital epispadias with incontinence. J Urol 62:5 13.<br />

30. Leadbetter GW (1964) Surgical correction of total urinary incontinence. J Urol<br />

91:261.<br />

31. Strawbridge LR, Kramer SA, Castillo OA, Barrett DM (1989) Augmentation<br />

cystoplasty and the artificial genitourinary sphincter. J UroE 142:297-301.<br />

32. Nurse DE, Mundy AR (1989) Metabolic complications of cystoplasty. Br J<br />

Urol 63:165-170.<br />

33. Mundy AR, Nurse DE (1992) Calcium balance growth skeletal and mineralization<br />

in patients with cystoplasties. Br J Urol 69:257-259.<br />

34. Golomb J, Klutke CG, Lewin KJ, et al. (1989) Bladder neoplasms associated with<br />

augmentation cystoplasty: report of 2 cases and literature review. J Urol 142:377-<br />

380.<br />

35. Filmer RB, Spencer JR (1990) Malignancies in bladder augmentations and intestinal<br />

conduits. J Urol 143:671-678.<br />

36. Hensle TW, Burbige KA (1990) Complications of urinary tract reconstruction.<br />

In: Marshall FF, ed., Urologic Complications, 2nd ed. St. Louis: CV Mosby<br />

Year Book.<br />

37. Braverman RW, Leboweitz RL (1991) Perforation of the augmented urinary<br />

bladder in nine children and adolescents: importance of cystography. Am J Roentgenol<br />

1575.<br />

38. Hill DE, Kramer SA (1990) Management of pregnancy after augmentation<br />

cystoplasty. J Urol 144 (part 2): 457-459.


INDEX<br />

A<br />

5-alpha reductase inhibitors, 233-234<br />

ALPP, 45-46<br />

Abdominal examination, 3 1<br />

American Spinal Injury Association<br />

Abdominal leak-point pressure (ALPP), (ASIA) Score, 120<br />

45-46<br />

American Urologic Association Symp-<br />

Abdominal stoma, 100<br />

tom Index, 29-30<br />

Abdominal straining<br />

Anatomic etiology<br />

voiding, 2 l0<br />

female urinary retention, 198-202<br />

Abrams-Griffiths Nomogram, 47f Anterior vaginal wall<br />

Absent volitional control of the external prolapse, 198<br />

sphincter (AVCS), 68, 69t Anticholinergics, 73, 101, 158, 266,<br />

Acetylcholine (Ach), 276, 28 1-282<br />

28 1-282<br />

Acontractile bladder<br />

Antidepressants, 287<br />

female urinary retention, 204 Antihistamines, 287<br />

Acquired immune deficiency syndrome Antimuscarinic agents, 284<br />

(AIDS)<br />

Anuria<br />

female urinary retention, 205<br />

female urinary retention, 198<br />

Activities of daily living (ADL) Areflexic bladder<br />

MS, 99<br />

LMN<br />

Acute urinary retention<br />

CIC, 127<br />

myelodysplasia, 125, 125f<br />

CVA, 67<br />

Artificial urinary sphincter, 26 1-265,<br />

ADL<br />

262f-264f<br />

MS, 99<br />

erosion, 264<br />

Age related etiology<br />

failure, 263-264<br />

female urinary retention, 207-209 SUI, 180<br />

AIDS<br />

ASIA Score, 120<br />

female urinary retention, 205 Assistive devices, 99<br />

Alcohol injection<br />

Atropine, 282<br />

MS, 103-1 04<br />

Augmentation cystoplasty, 3 16<br />

Alfuzosin, 279<br />

bladder management, 320-32 l<br />

Alpha adrenergic blockers, 168-1 69,<br />

bladder resection, 320<br />

23 l-233,239,242,259,278,285 bowel segment reconfiguration, l 32<br />

SCI, 128<br />

bowel segment selection, 320-32 1<br />

selective vs nonselective, 233<br />

bladder outlet, 322<br />

327


328 Index<br />

complications, 322-325<br />

electrolyte disorders, 322-323<br />

Behavioral voiding <strong>dysfunction</strong>, 242<br />

Benign prostatic hyperplasia (BPH)<br />

incontinence, 323<br />

bladder outlet obstruction, 230-242<br />

metabolic disorders, 322-323<br />

pharmacological therapy, 23 1-234,<br />

neoplasms, 324<br />

pregnancy, 324-325<br />

280-28 1<br />

surgery, 234-238<br />

spontaneous perforation, 324 incontinence, 73<br />

urinary infection, 323-324<br />

natural history, 23 l<br />

voiding difficulties, 323<br />

Bentyl, 283<br />

contraindications, 3 19<br />

Benzodiazepines, 279-280<br />

detrusor instability, 3 15-325 Beta adrenergic agonists, 279<br />

indications, 3 17<br />

Beta-sitosterol, 28 1<br />

MS, 105-1 09<br />

Bethanechol chloride, 128, 145, 157-158,<br />

preoperative evaluation, 3 17-3 19 277-278<br />

cystourethroscopy, 3 18-3 19 Biofeedback<br />

laboratory studies, 3 18<br />

SUI, 166-1 68<br />

radiologic studies, 3 18 Bladder<br />

urodynamic evaluation, 3 18 endoscopy, 52-53<br />

surgery, 3 19<br />

hyperactive<br />

ureters, 32 1-322<br />

filling, 5<br />

Autoaugmentation, 3 16-3 17<br />

innervation, 277f<br />

Autologous fat, 170-1 7 1<br />

Bladder augmentation<br />

Autonomic dysreflexia<br />

cystoplasty, 320-32 l<br />

SCI, 132-1 33<br />

SCI, 130<br />

Autonomous neurogenic bladder, 17 lumbar disc disease, 158<br />

AVCS, 68,69t<br />

Bladder autoaugmentation<br />

Axonal degeneration<br />

MS, 105<br />

female urinary retention, 203 Bladder cancer<br />

SCI, 132<br />

B<br />

Bladder chimney, 106<br />

Bladder <strong>dysfunction</strong><br />

Baclofen, 279-280<br />

Bacteriuria, 323-324<br />

Balanced Bladder, l 16<br />

Balloon dilation<br />

external sphincter<br />

SCI, 130<br />

BCR, 88<br />

lumbar disc disease, 155<br />

Bedside cystometrics, 49-5 l<br />

Behavioral modification<br />

post-prostatectomy, 265-267<br />

conservative treatment, 265<br />

pharmacologic therapy, 266<br />

with sphincteric incompetence,<br />

266-267<br />

surgery, 266-267<br />

Bladder emptying<br />

abnormalities, 5-1 0<br />

facilitation, 9t<br />

failure<br />

MS, 99-100<br />

treatment, 10<br />

poststroke, 78<br />

Bladder filling<br />

Behavioral therapy<br />

abnormalities, 5-1 0<br />

SUI, 165-166<br />

facilitation, 8t


Index 329<br />

requirements, 4<br />

Bladder leak-point pressure (BLPP), 46<br />

Bladder neck<br />

endoscopic incision, 239-240<br />

female urinary retention, 188<br />

Bladder neck obstruction<br />

female urinary retention, 200<br />

primary<br />

male, 238-242<br />

Bladder neck suspension<br />

Burch<br />

SUI, 173-174,174f<br />

percutaneous, 175, 176f<br />

Pereyra-Raz, 174, 175f<br />

SUI, 173-1 76<br />

Bladder neoplasms, 324<br />

Bladder outlet<br />

augmentation cystoplasty, 322<br />

endoscopy<br />

female, 53-54<br />

male, 54<br />

Bladder outlet obstruction<br />

diagnosis, 44--45,46f<br />

female urinary retention, 189-1 9 1<br />

gynecologic causes, 198-200<br />

male, 225-242<br />

behavioral voiding dysknction, 242<br />

BPH. See Benign prostatic<br />

hyperplasia<br />

cystoscopy, 229<br />

lab tests, 227-228<br />

patient history, 226-227<br />

physical, 227<br />

postvoid residual, 228-229<br />

presentation, 226<br />

MS, 104<br />

Blood tests, 33,227-228,3 18<br />

BLPP, 46<br />

Bors-Comarr classification<br />

voiding <strong>dysfunction</strong>, 1 1-1 3, l 1 t<br />

Bowel <strong>dysfunction</strong>, 30<br />

female urinary retention, 198<br />

Bowel neoplasms, 324<br />

BPH. See Benign prostatic hyperplasia<br />

(BPH)<br />

Bradley classification<br />

voiding <strong>dysfunction</strong>, 14-1 5<br />

Bulbocavernosus reflex (BCR), 88<br />

lumbar disc disease, 155<br />

Burch bladder neck suspension<br />

SUI, 173-1 74,174f<br />

Capsaicin, 129, 284285<br />

Cardura, 74,232,279<br />

Catheterization, 2 14-2 15<br />

Cauda equina syndrome, 154-1 55<br />

Cerebral shock, 67<br />

Cerebrovascular accidents (CVA), 6344<br />

bladder <strong>dysfunction</strong><br />

early presentation, 67<br />

late presentation, 67<br />

hemispheric dominance, 70<br />

patient history, 7 1-72<br />

urinary incontinence, 64-65<br />

alternative therapy, 78-79<br />

female, 77-78<br />

incidence, 66<br />

male, 77<br />

primary bladder neck obstruction, pathophysiology, 66-67<br />

23 8-242<br />

urinary retention<br />

urinary flow rate, 228<br />

female, 76-77, 76f<br />

urodynamics, 229-230<br />

male, 74-76<br />

Bladder retraining<br />

urodynamic studies, 68-70, 68t-70t<br />

SUI, 166<br />

urological evaluation, 7 l<br />

Bladder storage, 186<br />

Cervical lesions<br />

Bladder suspension<br />

urinary retention, 200<br />

female urinary retention, 209-2 l0 Cholinergic agonists<br />

Bladder transection<br />

SCI, 128-1 29<br />

c


330 Index<br />

Chronic continence<br />

Cystometrics<br />

incidence, 297<br />

bedside, 49-5 l<br />

Chronic indwelling catheter<br />

Cystometrogram (CMC), 38"43,40f, 42f<br />

LMN and areflexic bladder, 127 Cystometry<br />

Chronic urinary retention<br />

elderly, 72<br />

etiology, 199t<br />

Cystoplasty<br />

CIC. See Clean intermittent<br />

augmentation<br />

catheterization (CIC)<br />

MS, 105-1 09<br />

Cisparide, 278<br />

classification, 3 16-3 17<br />

Clean intermittent catheterization (CIC), Cystoscopic image<br />

74,76, 187,2 14,323<br />

DM, 145<br />

sling obstruction, 2 1 1 f<br />

cystoscopy<br />

LMN and areflexic bladder, 127 female urinary retention, 196<br />

lumbar disc disease, 158<br />

Cystospaz, 266,282<br />

poststroke, 79<br />

CMG, 38"43,40f, 42f<br />

Cystourethroscopy, 5 1-55, 194,3 18-3 19<br />

Cohen cross-trigonal reimplant, 32 1<br />

Collagen, 26O-26 1,260t<br />

I)<br />

injection, 77<br />

lumbar disc disease, 158<br />

Collagenous ingrowth<br />

detrusor, 207<br />

Colon surgery<br />

female urinary retention, 203-204<br />

Compliance, 4 1<br />

Condom catheter<br />

poststroke, 79<br />

SCI, 129<br />

Conticath, 74-75,75f<br />

Contigen, 172-1 73<br />

Continent urinary diversion<br />

MS, 105-1 09<br />

Continuous leakage, 29<br />

Corticoregulatory tract, 16<br />

Corticosteroids, 287<br />

Crede maneuver, 126, 128<br />

CT scans, 56<br />

CVA, see Cerebrovascular accidents<br />

(CV4<br />

Cystocele formation<br />

female urinary retention, 198<br />

Cystocystoplasty<br />

MS, 104<br />

Cystolysis<br />

Dantrolene, 279-280<br />

DDAVP ( 1 -deaminino-8-D arginine<br />

vasopressin), 100,287<br />

Decreased outlet resistance, 5-45<br />

Denervation<br />

MS<br />

treatment, 102<br />

Depression, 99<br />

DESD. See Detrusor external sphincter<br />

dyssynergia (DESD)<br />

Detrusor<br />

collagenous ingrowth, 207<br />

Detrusor areflexia<br />

DM, 145<br />

MS, 96f, 108-109<br />

SCI, 121<br />

Detrusor contractility<br />

impaired, 207-209<br />

lumbar disc disease, 153-154<br />

Detrusor contraction, 186<br />

Detrusor curves<br />

female urinary retention, 19Of-19 1 f<br />

Detrusor <strong>dysfunction</strong>, 186<br />

Detrusor external sphincter dyssynergia<br />

(DESD), 85<br />

MS, 92f-93f, 95f<br />

MS, 104<br />

detrusor hyperreflexia, 107, 108


Index 331<br />

SCI, 8, 121<br />

stents, 108<br />

Detrusor function, 186<br />

Detrusor hyperactivity with impaired<br />

contraction (DHIC), 77,207,209f<br />

Detrusor hyperreflexia, 12-14, 18,41<br />

CVA, 67,68<br />

DESD<br />

MS, 107-108,108<br />

SCI, 120-121, 122f<br />

MS, 9 1 f-95f<br />

treatment, 10 1-1 02<br />

neurogenic<br />

capsaicin, 129<br />

SCI, 120-1 l 2<br />

urethral obstruction<br />

MS, 106-107, 108<br />

Detrusor instability, 4 1<br />

augmentation cystoplasty, 3 15-325<br />

DM<br />

treatment, 145<br />

electrical stimulation, 297-3 12<br />

Detrusor myectomy<br />

MS, 105<br />

Detrusor myotomy<br />

MS, 104<br />

Detrusor sphincter dyssynergia (DSD)<br />

CVA, 69<br />

spinal-cord injury, 204-205<br />

DHIC, 77,207,209f<br />

Diabetes mellitus (DM), 139-146<br />

diagnosis, 140-141<br />

epidemiology, 1 3 9<br />

female urinary retention, 202-203<br />

pathophysiology, 143-145<br />

presentation, 139-140<br />

prevalence, 139<br />

treatment, 145<br />

urodynamic findings, 142-143, 142f<br />

Diabetic cystopathy, 14 I, 143-144<br />

female urinary retention, 202-203<br />

Diabetic neurogenic bladder<br />

diagnosis, 141-142<br />

Diazepam, 279<br />

Dicyclomine, 145<br />

Dicyclomine hydrochloride, 283<br />

Dimethysulfoxide (DMSO), 286<br />

Distal urethral sphincter, 249<br />

Ditropan, 101, 145,266,282<br />

Diuresis renography, 56<br />

DM, See Diabetes mellitus (DM)<br />

DMSO, 286<br />

Doxazosin, 74,232,279<br />

DSD<br />

CVA, 69<br />

spinal-cord injury, 204-205<br />

Dysfunctional voiding<br />

female urinary retention, 205-206<br />

Dyspareunia, 3 0<br />

Dysuria, 28<br />

E<br />

Elderly<br />

urodynamic studies<br />

technical difficulties, '72<br />

Electrical stimulation, 167<br />

action mechanism, 299-300<br />

acute maximal functional, 300<br />

chronic, 300<br />

detrusor instability, 297-3 12<br />

electrode placement, 302f<br />

needle stimulation devices, 304-305<br />

nerve pathways, 301f<br />

overview, 303t<br />

position points, 30 1 f<br />

SNS, 305-3 1 l<br />

transanal devices, 302-304<br />

transvaginal devices, 302-304<br />

Electrolyte disorders<br />

augmentation cystoplasty, 322-323<br />

Electromyography (EMG), 46-48<br />

CVA, 69<br />

elderly, 72<br />

female urinary retention, 188-1 89,206<br />

Elmiron, 286-287<br />

EMG. See Electromyography (EMG)<br />

Emptying function, 186. See aZso <strong>Voiding</strong><br />

Endocrine <strong>dysfunction</strong>


332 Index<br />

female urinary retention, 198<br />

Endoscopic urethrotomy, 240-24 1<br />

Endoscopy<br />

bladder, 52-53<br />

extraurethral incontinence, 54-55<br />

female bladder outlet, 53-54<br />

male bladder outlet, 54<br />

Enteroceles<br />

female urinary retention, 198-1 99<br />

Ephedrine, 170,259,285<br />

Estrogen, 168-1 69,285-286<br />

External sphincter<br />

innervation, 277f<br />

spasticity, 207<br />

Extraurethral incontinence<br />

endoscopy, 54-55<br />

F<br />

Failure to empty<br />

pharmacological therapy, 277-28 1<br />

bladder emptying facilitation, 277-<br />

278<br />

BPH outlet obstruction, 23 1-234,<br />

280-28 l<br />

decrease outlet resistance, 279-280<br />

Failure to store<br />

phamacological therapy, 28 1-287<br />

decrease bladder contractility,<br />

28 1-285<br />

increase outlet resistance,<br />

285-286<br />

inflammatory bladder conditions,<br />

286-287<br />

nocturnal enuresis, 287<br />

Fecal impaction<br />

female urinary retention, 198<br />

Female bladder outlet<br />

endoscopy, 53-54<br />

Female urethra<br />

endoscopy, 54<br />

Finasteride, 232,233-234,280-281<br />

Finetech-Brindley sacral anterior root<br />

stimulator, 13 1<br />

Flavoxate hydrochloride, 283<br />

Fluid restriction<br />

poststroke, 78<br />

Fluoroscopy<br />

elderly, 72<br />

pressure flow testing, 19 1-1 92<br />

sphincter <strong>dysfunction</strong>, 208f<br />

urethral obstruction, 192f-197f<br />

Fluxetine, 287<br />

Foley catheter<br />

poststroke, 79<br />

Frequency, 28<br />

SNS, 309-3 10<br />

Functional etiology<br />

female urinary retention, 202-207<br />

Functional system classification<br />

voiding <strong>dysfunction</strong>, 5t4t, 2 1-23,<br />

26-27<br />

G<br />

Gait, 88<br />

Genitals<br />

male<br />

examination, 32<br />

Glutaraldehyde cross-linked (GAX)<br />

bovine collagen, 172,260<br />

Gynecologic causes<br />

bladder outlet obstruction, 198-200<br />

H<br />

Hald Bradley classification<br />

voiding <strong>dysfunction</strong>, 13-14, 13t<br />

Hemiplegia<br />

urodynamic studies, 70t<br />

Heparin, 286<br />

Hinrnan’s syndrome, 205-206,280<br />

Holium laser, 75<br />

Hydronephrosis<br />

female urinary retention, 198<br />

Hyoscyamine, 266,282<br />

Hyoscyamine sulfate, 282<br />

Hyperactive bladder<br />

filling, 5<br />

Hyperchloremic metabolic acidosis, 322


Index 333<br />

Hysterectomy<br />

female urinary retention, 204<br />

Hytrin, 279<br />

IS, 5,286<br />

IVU, 55-56<br />

I<br />

K<br />

IDDM<br />

Kegel exercises, 167,259<br />

female urinary retention, 202-203<br />

Ilia1 intesusception nipple technique, 32 1<br />

L<br />

Imipramine, 145,259,283-284<br />

Irnrnobility<br />

female urinary retention, 198<br />

Impaired detrusor contractility, 207-209<br />

Implants<br />

SUI, 170-1 73<br />

Indigo laser prostatectomy, 75<br />

Indwelling Foley catheter<br />

poststroke, 79<br />

In-Flow device, 76, 76f<br />

Ingelmann-Sundberg procedure<br />

MS, 102-1 03<br />

Injection therapy, 259-26 1<br />

Insterstitial diode laser, 237<br />

Insulin-dependent diabetic mellitus<br />

(IDDM)<br />

female urinary retention, 202-203<br />

Intermittent catheterization<br />

myelodysplasia, 126<br />

Internittent flow pattern, 36, 37f<br />

Intermittent self catheterization, 108-1 09<br />

International Continence<br />

Society Classification<br />

voiding <strong>dysfunction</strong>, 19-2 l, 19t<br />

International Prostate Symptom Score,<br />

29-30<br />

InterStim Continence Control System, 10 3<br />

InterStirn device, 301<br />

Interstitial cystitis (IS), 5, 286<br />

Intervertebral disc<br />

prolapsed, 154-1 55<br />

Involuntary detrusor contractions<br />

classification, 4 1<br />

Irritative voiding<br />

symptoms, 28<br />

male bladder outlet obstruction,<br />

226-227<br />

Laminectomy, 157<br />

Lapides classification<br />

voiding <strong>dysfunction</strong>, 15-1 7, 16t<br />

Laser sphincterotomy<br />

MS, 108<br />

Leak-point pressure, 45-46<br />

Learned voiding <strong>dysfunction</strong>, 205-206<br />

Lesions<br />

lower motor neuron (LMN), 1-12<br />

urinary retention, 200<br />

vaginal<br />

urinary retention, 200<br />

Levsin, 282<br />

Linear PURR nomogram, 48f<br />

LMN<br />

areflexic bladder<br />

CIC, 127<br />

lesion, 1-1 2<br />

Low bladder compliance<br />

MS<br />

male, 97f<br />

Lower motor neuron (LMN)<br />

areflexic bladder<br />

CIC, 127<br />

lesion, 1-12<br />

Lower urinary tract<br />

central innervation, 65-66<br />

innervation, 149-1 54, 150t<br />

Lower urinary tract symptoms (LUTS), 25<br />

male bladder outlet obstruction, 226<br />

Lumbar disc disease, 149-1 58<br />

clinical features, 154-1 55<br />

evaluation, 155-1 56<br />

neurologic features, 157t<br />

neuropathophysiology, 149-1 54, 15Ot<br />

treatment, 157-1 58


334 Index<br />

urodynamics, 156 bladder transection, 10<br />

Lumbar disc prolapse, 15 1-1 52, 152f-153f comon urodynamic patterns, 90,<br />

LUTS, 25<br />

9 f-97f l<br />

male bladder outlet obstruction, 226 concurrent stress incontinence, 109<br />

continent urinary diversion, 105-1 09<br />

M<br />

cystolysis, 104<br />

denervation<br />

Male bladder outlet<br />

endoscopy, 54<br />

Male genitals<br />

examination, 32<br />

Male suburethral sling, 265<br />

Marion’s disease<br />

female urinary retention, 200<br />

Meningocele, 123<br />

Metabolic disorders<br />

augmentation cystoplasty, 322-323<br />

Metoclopramide, 145, 158,278<br />

Microwave therapy<br />

prostatic, 237-238<br />

Micturition, 1 16,276,298-299<br />

CVA, 67<br />

disrupted, 298-299<br />

neurophysiology, 85-90<br />

reflex pathways, 1 16-1 18, 1 17f<br />

two-phase concept, l 16<br />

Micturition cycle, 4,4647<br />

Micturition diary<br />

urinary incontinence, 87<br />

Micturition reflex, 85<br />

treatment, 102<br />

detrusor areflexia, 108-1 09<br />

detrusor hyperreflexia<br />

DESD, 107-1 08<br />

DESD and urethral obstruction, 108<br />

treatment, 10 1-1 02<br />

urethral obstruction, 106-1 07<br />

detrusor myectomy, 105<br />

detrusor myotomy, 104<br />

diagnostic evaluation, 86<br />

etiology, 84<br />

female urinary retention, 205<br />

incidence, 83<br />

laser sphincterotomy, I08<br />

physical examination, 88-89<br />

sphincter stent, 108<br />

subtrigonal phenol injection, 103-1 04<br />

transvaginal denervation, 102-1 03<br />

upper urinary tract evaluation, 89<br />

urinary tract symptoms, 86<br />

urodynamic evaluation, 89-90<br />

urologic history, 86-88<br />

videourodynamics, 90, 9 l f-97f<br />

Myelodysplasia, 123-1 26<br />

Minipress, 279<br />

Motor paralytic bladder, 16<br />

etiology, 123-124<br />

ureteral reimplantation, 32 1<br />

MRI, 56<br />

Myelomeningocele<br />

MS. See Multiple sclerosis (MS)<br />

child<br />

MS 800 artificial urinary sphincter, 261<br />

Mucosal structure disease<br />

female, 20 l f<br />

Multichannel subtracted pressure<br />

urodynamic study, 50f<br />

diagnosis, 124-1 25<br />

Myo-inositol<br />

peripheral nerves, 144<br />

flow testing<br />

N<br />

female urinary retention, 188<br />

Multiple sclerosis (MS), 83-1 09<br />

ADL, 99<br />

alcohol injection, 103-104<br />

augmentation cystoplasty, 105-109<br />

behavior modification, 99-1 00<br />

Nd:YAG laser<br />

MS, 108<br />

Needle electrical stimulation devices,<br />

304-305


Index 335<br />

Neoplasms<br />

Orthotopic cystoplasty, 3 16<br />

augmentation cystoplasty, 324<br />

Neural tube defects<br />

children<br />

Overworked bladder, 4 1<br />

Oxybutynin chloride, 101, 145,266,282<br />

incidence, 123<br />

Neurogenic bladder<br />

P<br />

diabetic<br />

diagnosis, 141-142<br />

ureteral reimplantation, 32 1<br />

Neurogenic detrusor hyperreflexia<br />

capsaicin, 129<br />

Neurogenic dyssynergia<br />

female urinary retention, 204-205<br />

Neurogenic etiology<br />

female urinary retention, 202-207<br />

Neurologic evaluation, 32-33<br />

female urinary retention, 188<br />

Neurostimulation<br />

SCI, 130-1 3 l<br />

Nifedipine, 283<br />

Pad test<br />

post-prostatectomy incontinence, 257<br />

urinary incontinence, 8’7<br />

Parasympathetic agonist agents<br />

pharmacologic manipulation<br />

female urinary retention, 2 15<br />

Parkinson’s disease<br />

female urinary retention, 198<br />

Pelvic floor muscle exercises, 99-1 00, 167,<br />

259<br />

Pelvic floor relaxation<br />

voiding, 2 IO<br />

Pelvic nerve injury<br />

Nitric oxide (NO), 279,284<br />

female urinary retention, 203-204<br />

Nocturia, 28<br />

Pelvic pain, 30<br />

selective<br />

elderly, 72<br />

Noninvasive uroflowmetry<br />

female urinary retention, 188<br />

Non-organic urinary retention<br />

fernale urinary retention, 205-206<br />

North American Contigen<br />

Study Group, 172<br />

Pelvis<br />

examination, 32<br />

Pentosanpolysulfate, 286-287<br />

Percutaneous bladder neck suspension,<br />

175, 176f<br />

Pereyra-Raz bladder neck suspension,<br />

174, 175f<br />

Peripheral nerves<br />

myo-inositol, 144<br />

0<br />

Peripheral nervous system degeneration<br />

female urinary retention, 203<br />

Obstruction diagnosis<br />

Pessaries<br />

method, 187<br />

vaginal, 168<br />

Obstructive etiology<br />

PG, 278<br />

female urinary retention, 198-202<br />

Pharmacologic manipulation<br />

parasympathetic agonist agents<br />

Obstructive voiding<br />

female urinary retention, 2 15<br />

symptoms, 28<br />

Phenoxybenzamine, 279<br />

male bladder outlet obstruction, 226 Phenylpropanolamine (PPA), 170,<br />

Oliguria<br />

259,285<br />

female urinary retention, 198 Physiologic etiology<br />

Open prostatectomy, 25 1-252<br />

female urinary retention, 207-209<br />

Opioid antagonists, 278<br />

Phytotherapeutics, 28 1,28 1 t


336 Index<br />

PMC, 65-66<br />

Pro-Banthine, 266, 282<br />

Pollen extract, 28 1<br />

Procardia, 133<br />

Polyneuropathy<br />

Prolapsed intervertebral disc, 154-1 55<br />

DM, 144<br />

Prompted voiding schedule<br />

Polytetrafluoroethylene (Teflon) paste, poststroke, 78<br />

259-260<br />

Propantheline, 266,282<br />

Pontine micturition center (PMC), 65-66 Propantheline bromide, 282<br />

Post-prostatectomy incontinence, 247-268 Proscar, 232-234,280-28 l<br />

etiology, 255-256,255t<br />

Prostaglandin inhibitors, 283<br />

evaluation, 256-258<br />

Prostaglandin (PG), 278<br />

cystourethroscopy, 257<br />

Prostate adenoma, 252f<br />

pad-weight test, 257<br />

Prostate gland<br />

patient history, 256-257<br />

anatomy, 249-25 1<br />

physical examination, 257<br />

examination, 88<br />

RUG, 257<br />

Prostatic microwave therapy, 23 7-23 8<br />

urinalysis, 257<br />

Prostatic stents, 74<br />

urodynamic, 257-258<br />

Proteoglycans, 286<br />

VCUG, 257-258<br />

Proximal urethra<br />

VLPP, 258<br />

spasticity, 206-207<br />

incidence, 248<br />

Proximal urethral sphincter, 249<br />

open prostatectomy, 25 1-252 Prozak, 287<br />

radical perineal prostatectomy, Pseudodyssnergia, 242<br />

253-254<br />

Pseudoephedrine, 170,259,285<br />

radical retropubic prostatectomy, Psychogenic cause<br />

252-253,254f<br />

female urinary retention, 198<br />

treatment, 258-267<br />

Pubovaginal sling, 78, 177f<br />

bladder <strong>dysfunction</strong>, 265-266 lumbar disc disease, 158<br />

bladder <strong>dysfunction</strong>/sphincteric Pumpkin seeds, 281<br />

incompetence, 266-267 PURR nomogram, 48f<br />

cost comparison, 267<br />

PVR, 34,228-229<br />

sphincteric incompetence,<br />

elderly, 72<br />

259-26 l<br />

Pyelonephritis, 323-324<br />

TUR, 25 1-252<br />

Pygeum africanum, 28 l<br />

Postvoid residual (PVR), 34,228-229<br />

elderly, 72<br />

Potassium channel openers, 283<br />

R<br />

PPA, 170,259,285<br />

Prazosin, 279<br />

Pregnancy<br />

augmentation cystoplasty, 324-325<br />

Pressure flow testing<br />

female urinary retention, 189<br />

radiographic monitoring, 19 1-1 92<br />

Primary bladder neck obstruction<br />

female urinary retention, 200<br />

Radical hysterectomy<br />

female urinary retention, 203-204<br />

Radical perineal prostatectomy, 253-254<br />

Radical retropubic prostatectomy,<br />

252-253,254f<br />

Radiographic monitoring<br />

pressure flow testing, 19 1-1 92<br />

Rectoceles<br />

female urinary retention, 198-1 99


Index 337<br />

Reflex neurogenic bladder, 17<br />

Regian, 145, 158,278<br />

Renography, 56<br />

diuresis, 56<br />

Replacement cystoplasty, 3 16<br />

Retention<br />

SNS, 310-311<br />

S<br />

Sacral dermatomes, 88<br />

Sacral nerve stimulation (SNS), 299-300,<br />

305-3 1 l<br />

adverse effects, 3 1 1<br />

clinical efficacy, 308-3 1 l<br />

retention, 3 10-3 1 1<br />

urge incontinence, 309<br />

urgency/frequency, 309-3 10<br />

surgical implantation, 308<br />

temporary electrode implantation, 306-<br />

307<br />

test stimulation, 306, 307f<br />

Sacral neuromodulation<br />

female urinary retention, 2 15<br />

Saw palmetto, 28 l<br />

SCI. See Spinal cord injury (SCI)<br />

Segmental neuronal demyelination<br />

female urinary retention, 203<br />

Selective nocturia<br />

elderly, 72<br />

Self catheterization<br />

lumbar disc disease, 158<br />

Self intermittent catheterization.<br />

LMN and areflexic bladder, 127<br />

Sensation, 40<br />

Sensory neurogenic bladder, 15-1 6<br />

Sertraline, 287<br />

Sexual dysfirnction, 30, 64<br />

Silver nitrate, 286<br />

Sling<br />

pubovaginal, 177f<br />

suburethral<br />

male, 265<br />

vaginal wall, 178f, 179<br />

Sling displacement<br />

female urinary retention, 2 10-2 12<br />

Sling obstruction<br />

cystoscopic image, 21 If<br />

Sling procedures<br />

female urinary retention, 2 10-2 12<br />

SUI, 176-1 80, 177f-178f<br />

Smooth muscle relaxants, 145<br />

Smooth sphincter, 4<br />

Smooth sphincter dyssynergia, 19<br />

SNS. See Sacral nerve stimulation (SNS)<br />

Somatosensory Evoked Potentials (SSEP),<br />

120<br />

South african star grass, 281<br />

Spasticity<br />

external sphincter, 207<br />

proximal urethra, 206-207<br />

Sphincter<br />

activity<br />

urodynamic testing, 47<br />

artificial urinary, 26 1-265,262f-264f<br />

erosion, 264<br />

failure, 263-264<br />

SUI, 180<br />

balloon dilation<br />

SCI, 130<br />

distal urethral, 249<br />

<strong>dysfunction</strong><br />

female urinary retention, 206-207<br />

fluoroscopy, 208f<br />

dyssynergia, 19<br />

external<br />

innervation, 277f<br />

spasticity, 207<br />

function, 186<br />

prostheses<br />

SCI, 130<br />

proximal urethral, 249<br />

smooth, 4<br />

stent<br />

MS, 108<br />

striated, 19<br />

urethral<br />

<strong>dysfunction</strong>, 186<br />

male, 249,250f<br />

urinary


338 Index<br />

anatomy, 249-25 l<br />

artificial, 180<br />

Sphincteric incompetence<br />

post-prostatectomy<br />

SCI, 132<br />

SSEP, 120<br />

Stents<br />

urethral, 24 1-242<br />

artificial urinary sphincter, 26 1-265, BPH, 238<br />

262f-264f<br />

Storage<br />

with bladder <strong>dysfunction</strong>, 266-267 symptoms, 28<br />

conservative treatment, 259 Stranguria, 25<br />

injection therapy, 259-26 l, 260t Stress incontinence, 29<br />

male suburethral sling, 265 female<br />

pharmacologic treatment, 259<br />

urodynamic study, 49f<br />

Sphincteric relaxation, 186<br />

management, 168<br />

Sphincterotomy<br />

MS, 109<br />

SCI, 129<br />

Stress urinary incontinence (SUI), 1631 8 l<br />

Spina bifida aperta, l23<br />

behavioral interventions, 165-1 68<br />

Spina bifida cystica, 123<br />

behavioral therapy, 165-1 66<br />

Spina bifida occulta, 123<br />

biofeedback, 166-1 68<br />

Spinal cord dysraphisms<br />

containment devices, 168<br />

neonatal<br />

support prosthesis, 168<br />

management, 124126<br />

bladder neck suspension, 173-1 76<br />

renal ultrasound, 124<br />

laparoscopic, 175-1 76<br />

urodynamics, 124-1 26<br />

retropubic, 173-1 74, 174f<br />

urological evaluation, 124-1 25 transvaginal, 174-1 75, 175f-176f<br />

VCUG, 124<br />

pharmacological management,<br />

Spinal cord injury (SCI), l 19-123,<br />

168-1 70<br />

121, 123<br />

alpha-adrenergic agents, 170<br />

cause, l 19<br />

estrogen, 168-1 69,286-287<br />

complication<br />

surgery, 170-1 80<br />

prevention, 13 1-1 33<br />

artificial urinary sphincter, 180<br />

detrusor sphincter dyssynergia,<br />

bladder neck suspension, 173-1 76<br />

204205<br />

implants, 170-1 73<br />

long-term urologic management, 126- sling procedures, 176-1 80,<br />

13 1<br />

177f-178f<br />

mortality, l 15-1 16<br />

treatment outcome, 163-1 65, 164t<br />

pharmacologic therapy, 128<br />

recovery phase, l 19-120<br />

spinal shock, 1 19<br />

surgery, 129<br />

ureteral reimplantation, 32 1<br />

urodynamic evaluation, 120-1 2 1<br />

Spinal shock, l 19<br />

Split cuff nipple, 322<br />

Spontaneous perforation<br />

augmentation cystoplasty, 324<br />

Squamous cell carcinoma<br />

Striated sphincter dyssynergia, 19<br />

Stroke. See Cerebrovascular accidents<br />

Substitution cystoplasty, 3 16<br />

Subtrigonal phenol injection<br />

MS, 103-104<br />

SUI. See Stress urinary incontinence (SUI)<br />

Suprapubic catheter<br />

poststroke, 79<br />

Surgery<br />

prio<br />

urinary tract function, 3 l


Index 339<br />

Surgical capsule, 249<br />

TUNA, 75,237<br />

Suture placement<br />

TUR, 236236,25 1-252<br />

female urinary retention, 209-2 l0 TVP, 235-236<br />

Suture tension<br />

female urinary retention, 209-2 10<br />

Two-Phase Concept of Micturition, l 16<br />

Symmetric peripheral neuropathy<br />

DM, 144<br />

U<br />

T<br />

Ultrasonography, 56<br />

Unconscious incontinence, 29<br />

Uninhibited neurogenic bladder, 16<br />

Tamsulosin, 233<br />

Upper motor neuron (UMN) lesion, 1-12<br />

TCA, 10 1,283-284<br />

Ureteral obstruction<br />

Teflon injections, 259-260<br />

female urinary retention, 198<br />

TENS, 306305<br />

Ureteral reimplantation<br />

Terazosin, 74,232,279<br />

augmentation cystoplasty, 32 1<br />

Timed voiding<br />

Ureters<br />

DM, 145<br />

augmentation cystoplasty, 32 1-322<br />

schedule<br />

Urethra<br />

poststroke, 78<br />

female<br />

SUI, 165-166<br />

endoscopy, 54<br />

Tolterodine, 77, 101 , 266, 282 Urethral dilation, 240-24 l<br />

Transanal electrical stimulation devices, Urethral obstruction<br />

3 02-3 04<br />

detrusor hyperreflexia<br />

Transcutaneous electrical nerve stimula- MS, 106-107, 108<br />

tion (TENS), 306305<br />

fluoroscopy, 192f-197f<br />

Transurethral electrovaporization of Urethral sphincter<br />

prostate (TVP), 235-236<br />

<strong>dysfunction</strong>, 186<br />

Transurethral incision of prostate male, 249,250f<br />

(TUIP), 236,239-240<br />

Transurethral microwave thermotherapy,<br />

75<br />

Transurethral needle ablation (TUNA), 75,<br />

237<br />

Transurethral resection (TUR), 236236,<br />

25 1-252<br />

Transvaginal denervation<br />

MS, 102-1 03<br />

Transvaginal electrical stimulation devices,<br />

302-304<br />

Transverse myelitis<br />

female urinary retention, 205<br />

Tricyclic antidepressants (TCA), 10 l, 283-<br />

284<br />

Urethral stent, 24 1-242<br />

BPH, 238<br />

Urethral stricture disease, 240<br />

Urethra position<br />

urinary retention, 188<br />

Urethra suspension<br />

female urinary retention, 209-2 l0<br />

Urethrolysis<br />

clinical experience, 2 13t<br />

female urinary retention, 12-2 2 14<br />

Urge incontinence, 29<br />

incidence, 297<br />

SNS, 309<br />

therapy, 166<br />

Urgency, 28, 166<br />

Trigger voiding, 128<br />

SNS, 309-3 10<br />

TUIP, 236,239-240<br />

Urinalysis, 33, 227, 318


340 Index<br />

Urinary diversion<br />

continent<br />

MS, 105-1 09<br />

SCI, 130<br />

Urinary incontinence, 28-29<br />

augmentation cystoplasty, 323<br />

child<br />

urodynamic study, 50f<br />

micturition diary and pad test, 8'7<br />

Urinary infection<br />

augmentation cystoplasty, 323-324<br />

Urinary retention<br />

female, 185-2 15<br />

etiology, 197-209<br />

evaluation, 186-1 98<br />

pathophysiology, 186<br />

treatment, 209-2 15<br />

Urinary sphincter<br />

anatomy, 249-25 1<br />

artificial<br />

SUI, 180<br />

Urinary tract<br />

lower<br />

function, 3-5<br />

innervation, 149-1 54, 150t<br />

Urinary tract imaging, 55-57<br />

lower, 57<br />

upper, 55-56<br />

Urinary tract infection<br />

female urinary retention, 198,202<br />

SCI, 13 1-1 32<br />

Urinary tract symptoms<br />

multiple sclerosis (MS), 84-85, 86<br />

Urine evacuation, 298-299<br />

Urine storage, 298-299<br />

abnormalities, 5-1 0<br />

facilitation, 8t<br />

failure, 2 1<br />

requirements, 4<br />

Urispas, 283<br />

Urodynamic<br />

classification<br />

voiding <strong>dysfunction</strong>, 18-19, 18t<br />

female urinary retention, 206<br />

simplification, 75<br />

Uroflowmetry, 35-37,35f<br />

female urinary retention, 188<br />

Urogenital diaphragm, 249<br />

Urolume, 75, 108<br />

Urosurge, 77,78f<br />

v<br />

Vagina<br />

examination, 3 1-32, 88-89<br />

urinary retention, 187-1 88<br />

Vaginal cones, 16'7<br />

Vaginal lesions<br />

urinary retention, 200<br />

Vaginal pessaries, 168<br />

Vaginal prolapse<br />

treatment, 199-200<br />

Vaginal stimulation, 30 1<br />

Vaginal wall<br />

prolapse, 198<br />

sling, 178f, 179<br />

Valsalva leak-point pressure, 45-46<br />

Valvular mechanisms<br />

female urinary retention, 2 15<br />

VCUG, 57<br />

female urinary retention, 192-1 94<br />

Vesical capacity, 186<br />

Vesicoureteral reflux<br />

SCI, 132<br />

Videourodynamics, 48"49,52f, 196<br />

<strong>Voiding</strong><br />

abdominal straining, 2 10<br />

pelvic floor relaxation, l0 2<br />

symptoms, 28<br />

<strong>Voiding</strong> and intake diary, 34<br />

<strong>Voiding</strong> cystourethrography (VCUG), 57<br />

female urinary retention, 192-1 94<br />

<strong>Voiding</strong> difficulties<br />

augmentation cystoplasty, 323<br />

<strong>Voiding</strong> <strong>dysfunction</strong>, 25-57<br />

classification, 1&23,26-27<br />

Bors-Comarr, l 1-1 3, l l t<br />

Bradley, 14-1 5<br />

functional system, 5t4, 2 1-23,<br />

26-27


Index 341<br />

Hald Bradley, 13-14, 13t<br />

International Continence Society<br />

Classification, 19-2 1, 19t<br />

Lapides, 15-17, 16t<br />

urodynamic, 18-1 9, 18t<br />

endoscopy, 5 1-55<br />

bladder evaluation, 52-53<br />

extraurethral incontinence, 54-55<br />

female bladder outlet, 53-54<br />

male bladder outlet, 54<br />

functional classification, 5t<br />

expanded, 6t<br />

laboratory testing, 33<br />

learned, 205-206<br />

patient history, 27-3 1<br />

physical examination, 3 1-33<br />

testing, 33-38<br />

pad test, 37-38<br />

postvoid residual, 34<br />

uroflowmetry, 35-37,35f<br />

voiding and intake diary, 34<br />

urinary tract imaging, 55-57<br />

lower, 57<br />

upper, 55-56<br />

urodynamics, 38-5 1<br />

bedside cystometrics, 49-5 1<br />

CMG, 38-43,40f, 42f<br />

electromyography, 46-48<br />

leak-point pressure, 45-46<br />

videourodynamics, 48-49,52f<br />

voiding pressure flow study, 43-45,<br />

45f, 51f<br />

<strong>Voiding</strong> function, 186<br />

<strong>Voiding</strong> pressure flow study, 43-45,45f,<br />

51f<br />

<strong>Voiding</strong> reflux<br />

inhibition, 10<br />

Vulvar lesions<br />

urinary retention, 200<br />

W<br />

Wertheim hysterectomy<br />

female urinary retention, 204<br />

z<br />

Zoloft, 287

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!