Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Experimental Physiology (1999).

84, 131-136
Printed in Great Britain

PHYSIOLOGICAL SOCIETY SYMPOSIUM:


THE PHYSIOLOGY AND PATHOPHYSIOLOGY O F THE
LOWER URINARY TRACT

THE ACTIVATION O F BLADDER WALL AFFERENT


NERVES
JOHN MORRISON*
Department of Physiology, United Arab Emirates University, PO Box 17666, A1 Ain, UAE

AFFERENT MECHANISMS

The functional state of the lower urinary tract is monitored more or less continuously by
afferent nerves, which provide a sensory input in the control of the bladder and the external
urinary sphincter. The functional state of the former is regulated by the autonomic nerves to
the bladder, principally the parasympathetic (pelvic) nerves, which cause the detrusor muscle
to contract, and possibly also by the sympathetic (mainly hypogastric nerve) innervation,
which may reduce the resting tone of the smooth muscle. Continence is normally maintained
by the resistance of the urethral sphincters; the tone generated by the skeletal muscle external
sphincter is under the control of the somatic a-motoneurones running in the pudendal nerve,
and possibly some somatic efferents in the pelvic nerve of some species.
The afferents monitor the volume of the bladder and the amplitude of bladder contraction;
these two factors are important in the regulation of bladder function, the former of relevance
particularly to the physiology of the filling phase, and the latter to maintenance of voiding
contractions of the detrusor muscle. In the pelvic and hypogastric nerves of cats, various
authors have described ‘in-series’ tension receptors in the detrusor muscle with AS afferent
fibres (Iggo, 1955; Floyd et al. 1976; Habler, 1993) that respond to distension or contraction
of the bladder, but cannot distinguish between them; these afferents, at least those in the pelvic
nerve, are believed to play a part in the regulation of the amplitude of micturition contractions
by providing a sensory input that is used as positive feedback to maintain the contraction of
the viscus (Morrison, 1995). But the absence of volume receptors in these descriptions has
always been an intellectual problem because there are some aspects of neural control that
cannot be performed just by in-series receptors. For example, the ability to remain continent
at high bladder volumes depends on sensing a large volume in the bladder, and suppressing the
micturition reflex. Current thinking states that the ‘in-series’ tension receptors are involved in
a positive feedback that facilitates and coordinates the micturition contraction, and it seems
unlikely that a single type of sensory receptor can undertake both contrasting functions
(Morrison, 1995). Indeed there is evidence that the sympathetic activity may be switched on at
high bladder volumes (Gjone, 1966; Edvardsen, 1967), and may oppose the contraction
induced by the micturition reflex.
*Email: John.Morrison@mail.uaeu.ac.ae

Presented at the Guy’s Hospital, London Meeting of the Society in January 1998.

1736
1469445x, 1999, 1, Downloaded from https://physoc.onlinelibrary.wiley.com/doi/10.1111/j.1469-445X.1999.tb00078.x by Nat Prov Indonesia, Wiley Online Library on [24/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
132 J. MORRISON

There are two examples of inhibitory influences on the micturition reflex. In the rat, Lumb
(1 986) and Chandler et al. (1994) described excitation of raphe-spinal neurones by bladder
distension. These neurones are believed to inhibit bladder motility. The second example of an
inhibitory pathway that would be activated by vesical distension is the nucleus reticularis pontis
oralis, where Kimura et al. (1995) identified a urine storage centre. It would be surprising if
the very receptors that initiated or facilitated the micturition reflex were also responsible for
inhibitory phenomena such as suppression of the desire to void or facilitation of urine storage.
This separation into volume receptors and tension receptors has not been a feature of the
scientific literature until recently, because most authors have used the cat as an experimental
model, and others have been particularly keen to differentiate between receptors that subserve
innocuous and noxious sensations.
Sengupta & Gebhart (1994) have described bladder nociceptors in an experiment where, on
average, ten afferents per animal were studied whilst undergoing repeated distensions of the
bladder to 80 mmHg; a proportion of these afferents became desensitized. However, it would
be desirable not to subject the bladder to repeated distensions of these pathophysiological
pressures. In studies from this laboratory that will be described below, bladder pressures never
exceeded 40 mmHg. The bladder of each animal was cannulated and kept empty until the start
of the recording. One unit per animal was studied. Distension was effected with saline at 37 "C.
Potentially sensitizing solutions were introduced into the bladder once only per experiment,
after a series of control measurements during saline distension. Great care was taken not to
overdistend the bladder at any stage, and this was confirmed at the end of the experiment by
the lack of trauma to the viscus, i.e. no evidence of haematuria, thus ensuring that the properties
of the units could not have been altered by previous manipulations of the viscus.
The sampling procedure for isolation of the units depends on finding a filament of the dorsal
roots containing afferents that respond to stimulation of the pelvic nerve at the bladder base; this
procedure eliminates any sampling bias associated with searching for units using mechanical
stimuli.
In the rat, conduction velocities greater than 1.3 m s-l are classified as A6-units (Waddell et
al. 1989). Sengupta & Gebhart (1994) described a population of C fibres in the rat whose mean
conduction velocity was 1.7 m s-'. Clearly there is some incompatibility between the definitions.
In the cat, C-afferents are generally accepted as having conduction velocities of less than 2 m s-'.
Some authors (e.g. Habler et al. 1990) have described cat C fibre afferents which are normally
insensitive to distension and have been called 'silent' afferents; some of these afferents may be
sensitized by various forms of chemical inflammation of the bladder mucosa (Habler et al.
1990).
In the rat there is now evidence that many C-afferents with conduction velocities of less than
1.3 m sC1 do respond to slow distension of the viscus with physiological volumes; some of
these are 'volume' receptors that do not respond to contractions of the bladder. Twenty-three
percent of A6 fibres and 64 YOof C fibres studied during bladder contractions induced by ventral
root stimulation behaved in this way; so the volume receptor afferents are mainly C fibres that
discharge during a normal cycle of distension of the bladder, but with higher thresholds than
those of the A6 'in-series' tension receptors.
Many C fibres which can be seen in the mucosa of the bladder contain peptides, and it may
be that the receptors that do not respond to detrusor contractions may have endings in the
mucosa. If these peptides were released into the interstitial fluid around these mechanosensitive
endings, they would be bathed in a cocktail of endogenously active agents arising from the
nerve itself and the surrounding tissue. The ability to transduce mechanical signals depends on
1469445x, 1999, 1, Downloaded from https://physoc.onlinelibrary.wiley.com/doi/10.1111/j.1469-445X.1999.tb00078.x by Nat Prov Indonesia, Wiley Online Library on [24/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ACTIVATION O F BLADDER WALL AFFERENT NERVES 133

Neurotubules
Protein Movement

Smooth Muscle
Inflammatory
Mediators

Fig. 1. This diagram shows a sensory ending associated with smooth muscle and monitoring forces generated in the bladder
wall. The membrane proteins in the ending include receptors, and those shown are receptors for neurokinin A (the NK-2
receptor), inflammatory mediators (such as bradykinin), and trophic agents (such as nerve growth factor).

membrane proteins such as the mechanogated channel, which is a membrane protein trans-
ported from the nucleus of the cell along a intracellular network of neurotubules. These proteins
impart specific sensitivities to the nerve ending and Fig. 1 shows just some of the types of
membrane protein on which the nerve ending depends to perform its function. These include
not only the mechanogated channel, but also a series of pharmacological binding sites, and
some neuropeptides that can be released as neurotransmitters or neuromodulators. The latter
include substance P, neurokinin A (NKA) and calcitonin gene-related peptide (CGRP), while
the membrane sensitivity depends on receptors for inflammatory mediators, neurochemical
mediators such as NKA and nerve growth factor (NGF) (Su et al. 1986; Persson et al. 1995;
Steers et al. 1991, 1996). This review will consider briefly the effects of neurokinin A and
NGF.
SENSITIZATION

The relationship between volume and spike rate is not constant in many small fibre afferents,
and the threshold or slope of the relationship or gain can be adjusted or modulated in various
ways. Figure 2 shows the process called sensitization, in which at a given bladder volume the
frequency of action potential in the afferent fibres is increased. This can be achieved either by
reducing the volume at which afferent firing is initiated or by increasing the slope of the
volume-frequency relationship. Consider how the brain would interpret a particular spike rate,
such as that shown by the horizontal dashed line in Fig. 2. In the normal bladder that amount
of neural activity might be associated with a full bladder. However, after sensitization, the
subject would think that the bladder was fuller than it actually is. This will give rise to the
clinical symptom of frequency.
It has been known for some time that some of these endings are sensitive to inflammatory
mediators such as bradykinin (Floyd et al. 1977), and what is becoming clearer is that some
membrane proteins present in the sensory endings, such as the NK-2 receptor, which respond
1469445x, 1999, 1, Downloaded from https://physoc.onlinelibrary.wiley.com/doi/10.1111/j.1469-445X.1999.tb00078.x by Nat Prov Indonesia, Wiley Online Library on [24/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
134 J. MORRISON

Spike rate Spike rate

I Normal Sensitize
Reduced Threshold

Volume
-c Volume

L Reduced Threshold and


Spike rate Increased Gain
(Slope of S-R curve) and
Resting Discharge
Volume

Fig. 2. The changes occurring in the spike frequency- volume relationship during sensitization include a reduction in
threshold (the volume at which spike discharge increases), an increase in thc slope of gain of the relationship, and the
occurrence o f resting discharge.

to the presence of chemicals such as neurokinin A in the local micro-environment of the nerve
ending, can alter the transduction process, possibly through interference with second messenger
systems such as inositol 1,4,5-trisphosphate (lP3). Another aspect of sensitization is that the
afferent neurone may become spontaneously active at rest, even when the bladder is empty.
Changes in urinary composition have been shown to influence the sensitivity of bladder
afferent endings (Jiang & Morrison, 1995, 1996), possibly by acting on intraepithelial or
submucosal nerve endings. The stimuli are alterations to the composition of normal constituents
of urine found in extreme physiological states in man, e.g. a pH of 4.5 (the lowest urinary pH,
such as might occur during a diabetic ketoacidosis), a high potassium concentration (about
300 mM) or high osmolality (2000 mosmol kg-', consistent with the urinary concentration
during dehydration, or diabetic ketoacidosis).
Lowering urinary pH to 4.5 sensitizes pelvic afferent endings in the bladder, increasing the
slope of the pressure-response curve, and inducing a resting discharge (Jiang et al. 1994).
Sensitization would predispose to lowering the sensory threshold to distension and sensory
urgency, and to reflex hyperactivity of the detrusor.
Raising the potassium concentration of the fluid in the bladder also sensitizes primary
afferents in the pelvic nerve; about one-third of those tested were sensitized by 300 mmol K' ;
about two-thirds were sensitized by 400 mmol K' and all were sensitized by 500 mmol K'
(Jiang & Morrison, 1995). The sensitization consisted of (a) the development of a resting
discharge significantly greater than that present when the bladder was empty in the control part
of the experiment, (b) a reduction in the threshold of the afferents, (c) an increase in the slope
of the pressure-response curve, and (d) the development of mechanosensitivity in 'silent' C-
afferents. Similarly hyperosmolality (2000 mosmol kg-') achieved by increasing NaCl or glucose
concentrations in the intravesical fluid also caused sensitization.
1469445x, 1999, 1, Downloaded from https://physoc.onlinelibrary.wiley.com/doi/10.1111/j.1469-445X.1999.tb00078.x by Nat Prov Indonesia, Wiley Online Library on [24/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
A C T I V A T I O N OF B L A D D E R W A L L A F F E R E N T N E R V E S 135

It is known that neurogenic extravasation occurs as a result of stimulation of afferents


innervating the bladder (Koltzenburg & McMahon, 1986; McMahon & Abel, 1987), and it is
probable that neuropeptides present in the afferent innervation are responsible for these changes.
The tachykinins substance P and neurokinin A, and CGRP appear to be released from afferent
nerve terminals during stimulation of the unmyelinated afferent fibres. We have previously found
that neurokinin-2 binding sites, which have a high affinity for neurokinin A, are present on
afferent nerves in the bladder mucosa (Nimmo et al. 1992, 1993); we also tested the hypothesis
that neurokinin A may play a part in the sensitization of mechanoreceptive afferents in the rat
bladder. It is suggested that NK-2 binding sites are autoreceptors on sensory endings: Kibble &
Morrison (19964 found that the NK-2 agonist P-Alas-NKA(4- 10) sensitized many mechano-
sensitive afferents in the rat bladder. Furthermore the NK-2 receptor antagonist SR48968 was
able to block the action of the agonist, suggesting that the sensitizing effect was mediated by
an NK-2 receptor mechanism (Kibble & Morrison, 1996b).
Finally, the possibility exists that part of the normal response of the afferent to distension
may be mediated by the NK-2 receptor. One might hypothesize that NKA release begins once
a threshold level of afferent discharge is achieved, and that sensitization is a physiological
process thereafter. The hypothesis was tested by comparing the responses of afferents to
distension before and after treatment with the NK-2 antagonist SR 48968: about 30 % of the
normal mechanoreceptor activity seen at 30 mmHg intravesical pressure can be blocked by
the NK-2 antagonist (Kibble & Morrison, 1996b). This finding suggests that release of NKA
and sensitization of afferents may be a physiological process; furthermore it demonstrates a
potential target for modulating afferent sensitivity.
As stated previously, nerve growth factor is released from smooth muscle cells and appears to
influence the sprouting of afferents in the hypertrophied bladder. It is also postulated that the
hyperactivity of the hypertrophied bladder is due to an increase in afferent sensitivity that may
be due to NGF. Dmitreva & McMahon (1996) have indeed demonstrated that NGF can sensitize
afferents in the bladder wall. Thus NGF and NKA may both be involved in sensitization
phenomena in the bladder.

We would like to thank The Wellcome Trust and MRC for funding this research.

REFERENCES

CHANDLER,
M. J., OH,U. T., HOBBS,S . F. & FOREMAN,
R. D. (1994). Responses of feline raphe-spinal
neurons to urinary bladder distension. Journal of the Autonomic Nervous System 47, 21 3-224.
DMITREVA,
N. & MCMAHON,S. B. (1996). Sensitisation of visceral afferents by nerve growth factor in the
adult rat. Pain 66, 87-97.
EDVARDSEN,
P. (1967). Nervous control of urinary bladder in cats. Acta Neurologica Scandinavica 43,
543-563.
FLOYD,
K., HICK,V. E., KOLEY,J. & MORRISON,
J. F. B. (1977). The effects of bradykinin on afferent
units in intra-abdominal sympathetic nerve trunks. Quarterly Journal of Experimental Physiology 62,
19-25.
GJONE,R. (1966). Peripheral autonomic influence on the motility of the urinary bladder in the cat. Acta
Physiologica Scandinavica 66,72-80.
HABLER,H.-J., JANE, W. & KOLTZENBURG,M. (1990). Activation of unmyelinated afferent fibres by
mechanical stimuli and inflammation of the urinary bladder in the cat. Journal of Physiology 425,
545-562.
HABLER,H.-J., JXNIG, W. & KOLTZENBURG, M. (1993). Myelinated primary afferents of the sacral spinal
cord responding to slow filling and distension of the cat urinary bladder. Journal of Physiology 463,
449-460.
1469445x, 1999, 1, Downloaded from https://physoc.onlinelibrary.wiley.com/doi/10.1111/j.1469-445X.1999.tb00078.x by Nat Prov Indonesia, Wiley Online Library on [24/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
136 J . MORRISON

IGGO,A. (1955). Tension receptors in the stomach and the urinary bladder. Journal offhysiology 128,
593-607.
JIANG,W., KAWATANI, M. & MORRISON,J. F. B. (1994). The effects of urinary pH on pelvic nerve
mechanoreceptors and ‘silent’ afferents in the bladder of the anaesthetized rat. Journal of P hysiology
481.P, 23-24P.
JIANG,W. & MORRISON, J. F. B. ( 1 994). The effects of changing urinary composition on ‘silent’ afferents
in the rat bladder. Journal of Physiology 476.P, 50--51P.
JIANG,W. & MORRISON, J. F. B. (1995). The effects of high urinary potassium concentration on pelvic
nerve mechanoreceptors and ‘silent’ afferents from the rat bladder. Advances in Experimental
Medicine and Biology 385, 237-239.
JIANG,W. & MORRISON, J. F. B. (1996). Sensitization of pelvic afferent neurones from the rat bladder.
Journal of the Autonomic Nervous System 58, 187-1 88.
KIBBLE,A. & MORRISON, J. F. B. (1996a). In vivo pelvic afferent responses to a Neurokinin A agonist in
the anaesthetized rat. Journal offhysiology 491.P, 148-149P.
KIBBLE,A. & MORRISON, J. F. B. (19966). Changes in mechanosensitivity and sensitization of urinary
bladder afferents in the anaesthetized rat following administration of the selective NK-2 antagonist
SR48968. Journal of Physiology 497.P, 18-19P.
KIMURA, Y., UKAI,Y., KIMURA,K., SUGAYA, K. & NISHIZAWA, 0. (1995). Inhibitory influence from the
nucleus reticularis pontis oralis on the micturition reflex induced by electrical stimulation o f the pontine
micturition centre in cats. Neuroscience Letters 195, 214-216.
KOLTZENBURG, M. & MCMAHON,S. B. (1986). Plasma extravasation in the rat urinary bladder following
mechanical, electrical and chemical stimuli: evidence for a new population of chemosensitive primary
sensory afferents. Neuroscience Letters 72, 352-356.
LUMB,B. M. (1986). Brainstem control of visceral afferent pathways in the spinal cord. Progress in Brain
Research 67, 279-293.
MCMAHON, S. B. & ABEL,C. (1987). A model for the study of visceral pain states: chronic inflammation
o f the chronic decerebrate rat urinary bladder by irritant chemicals. Pain 28, 109-127.
MORRISON, J. F. B. (1995) The excitability of the micturition reflex. Scandinavian Journal of Urology
and Nephrology Supplement 175, 2 1-26.
NIMMO,A. J., ANDERSON,P. 0. & MORRISON, J. F. B. (1991). Reactive changes in neurokinin receptor
density following selective denervations and outlet obstruction of rat bladder. Journal of Physiology
446, S24P.
NIMMO,A. J., ANDERSON,P. 0. & MORRISON, J. F. B. (1992). NK-2 binding sites associated with afferent
ncrve endings in the rat bladder increase in number during urethral obstruction. X X X I I Congress ofthe
International Physiological Sciences 18.8/0.
PERSSON,K., SANDO,J. J., TUTTLE,J. B. & STEERS,W. D. (1995). Protein kinase C in cyclic stretch-
induced nerve growth factor production by urinary tract smooth muscle cells. American Journal of
Physiology 269, C1018-1024.
SASAKI, M., MORRISON, J. F. B., SATO,Y. & SATO,A. (1994). Effect of mechanical stimulation of the skin
on the external urethral sphincter muscles in anaesthetised cats. Japanese Journal of Physiology 44,
575--590.
SENGUPTA, J. N. & GEBHART,G. F. (1994). Mechanosensitive properties of pelvic nerve afferent fibres
innervating thc urinary bladder of thc rat. Journal of Neurophysiology 72, 2420-2430.
STEERS,W. D., CREEDON, D. J. & TU’ITLE, J. B. (1996). Immunity to nerve growth factor prevents affcrent
plasticity following urinary bladder hypertrophy. Journal of Urology 155, 379-385.
STEERS,W. D., KOLBECK,S., CREEDON, D. & TUTTLE,J. B. (1991). Nerve growth factor in the urinary
bladder of the adult regulates neuronal form and function. Journal of Clinical Investigation 88,
1709-17 15.
Su, H . C., WHARTON, J., POLAK,J. M., MULDERRY, P. K., GHATEI,M. A,, GIBSON,S. J., TERENGHI, G.,
MORRISON, J. F. B., BALLESTA, J. & BLOOM,S . R. (1986). Calcitonin gene-related peptide immuno-
reactivity in afferent neurons supplying the urinary tract: Combined retrograde tracing and immuno-
Iiistochemistry. Neuroscience 3, 727-747.
WADDELL,P. J., LAWSON,S. N. & MCCARTHY, P. W. (1989). Conduction velocity changes along the
processes of rat primary sensory neurones. Neuroscience 30, 577-584.

You might also like