You are on page 1of 7

J. Smooth Muscle Res.

(2011) 47 (2): 47–53 47

Invited Review for 2010 Hirosi Kuriyama Award


The functional role of intramuscular interstitial cells of
Cajal in the stomach
Yoshihiko KITO
Department of Physiology, Nagoya City University Medical School, Japan

Received January 27, 2011; Accepted February 2, 2011

Abstract
Intramuscular interstitial cells of Cajal (ICC-IM) are found within the smooth muscle
layers of the stomach. ICC-IM are mainly spindle shaped cells with bipolar processes
orientated along the long axis of surrounding smooth muscle cells. ICC-IM make close
contacts with nerve varicosities and form gap junctions with neighbouring smooth muscle
cells, indicating that ICC-IM mediate enteric motor neurotransmission. These
morphological properties of ICC-IM are similar throughout the stomach. However, the
electrical properties of these cells differ from region to region. In the fundus, ICC-IM
generate spontaneous transient depolarizations (STDs), resulting in an ongoing discharge of
unitary potentials in the smooth muscle cells. ICC-IM in the corpus generate slow waves
and as they fire at the highest frequency they serve as the dominant pacemaker cells in the
stomach. On the other hand, ICC-IM in the antrum generate the secondary component of
slow waves triggered by the initial component that propagates passively from myenteric
ICC (ICC-MY). Thus, the different electrical properties of ICC-IM play a critical role in
creating the distinct functions of the proximal and distal regions of the stomach such that
the fundus acts as a reservoir of food, the corpus as a dominant pacemaker region, while the
antrum acts as a region for mixing and propulsion of food.

Key words: intramuscular interstitial cells of Cajal, neurotransmission, slow waves,


spontaneous transient depolarization, unitary potential

Introduction

The stomach has three types of motor functions; (1) storage of the food in the fundus and body
of the stomach, (2) mixing of food by peristaltic contractions which occurs in the body and moves
toward to the antrum, (3) emptying of the chyme (the resulting mixture of food and stomach
secretions) from the stomach into the duodenum. These stomach functions are regulated by
distinct electrical activity recorded in the smooth muscle cells in various regions of the stomach.
In the fundus, smooth muscle cells are electrically quiescent or, in some occasions, generate the
discharge of membrane noises (unitary potentials). Smooth muscle cells in the corpus generate
small amplitude (up to 15mV) slow waves at a high frequency. Slow waves in the antrum are 20–

Correspondence to: Dr. Y. Kito, Department of Physiology, Nagoya City University Medical School, Mizuho-ku,
Nagoya 467-8601, Japan
Phone: +81-52-853-8131 Fax: +81-52-842-1538 e-mail: y.kito@med.nagoya-cu.ac.jp
48 Y. KITO

30 mV in amplitude and consist of two components together with spike potentials on top of the
secondary component. The resting membrane potential of smooth muscle cells in the antrum is
more negative than that in the fundus (Komori and Suzuki, 1986, Xue et al., 1996). These
electrical properties are common in mouse (Hirst et al., 2002; Beckett et al., 2004), rat (Xue et al.,
1996), guinea pig (Komori and Suzuki, 1986), dog ( Szurszewski, 1981) and originate in interstitial
cells of Cajal (ICC) rather than the smooth muscle cells (Sanders, 1996). ICC are mesenchymal in
origin and express Kit, a receptor tyrosine kinase. Signalling via Kit is necessary for ICC
development and maintenance of phenotype (Sanders et al., 1999). Neither smooth muscle cells
nor enteric neurons express Kit. Thus ICC can be identified by Kit antibodies (Sanders et al.,
1999) and found in the all regions of the gastrointestinal (GI) tract from the oesophagus to the
inner sphincter region of the anus (Ward and Sanders, 2001, Komuro, 2006). In the stomach, at
least two types of ICC exist: highly branching, network-forming ICC located near the myenteric
plexus (ICC-MY or ICC-MP) and long, thin, spindle-shaped ICC scattered amongst the smooth
muscle cells (intramuscular ICC or ICC-IM). This review will focus on the functional roles of ICC-
IM in gastric motility.

Role of ICC-IM in enteric neurotransmission

The physiological role of ICC-IM was first identified as mediators of neurotransmission (Burns
et al., 1996). In control mice stomach, nerve stimulation evoked cholinergic excitatory, nitrergic
inhibitory and purinergic inhibitory junction potentials in the circular muscle layers from the
fundus to antrum. However, in the stomach of Kit receptor kcockout W/Wv mice, which lacks ICC-
IM, the cholinergic and nitrergic components were greatly reduced in spite of the presence of
normal numbers of cholinergic and nitrergic nerve terminals (Ward and Sanders, 2001), suggesting
that ICC-IM rather than smooth muscle cells are the main target for nerve-released acetylcholine
(ACh) and nitric oxide (NO).
Although recent studies are against the involvement of ICC-IM in neurotransmission
(Huizinga et al., 2008; Zhang et al., 2009), a role for ICC-IM in neurotransmission has been
extensively confirmed by morphological studies (Thuneberg, 1982, Ward and Sanders, 2001,
Komuro, 2006). Double labeling of excitatory motor neurons with vesicular ACh transporter
(vAChT) and ICC with Kit showed that excitatory motor neurons are closely associated with ICC-
IM (Ward and Sanders, 2006). The same structural arrangement was also observed for nitrergic
inhibitory motor neurons. Morphological studies with electron microscopy have revealed very
close contacts (< 20 nm) between the varicosities of enteric motor neurons and ICC-IM (Wang et
al., 1999). Ultrastructural studies have also shown that synapse-like specializations exist between
enteric nerve terminals and ICC-IM (Komuro, 2006). In contrast, enteric nerves do not form close
apposition with the smooth muscle cells or form synaptic specializations. Further support comes
from the fact that ICC-IM are directly innervated by nerves containing immunohistochemical
staining for synaptic specific molecules; e.g. soluble N-ethylmaleimide sensitive factor or fusion
protein attachment protein receptor (SNARE) proteins involved in the neurovesicle docking to
postsynaptic membranes in nerve fibers and postsynaptic density (PSD) proteins (Beckett et al.,
2005). Furthermore, ICC-IM express several receptors for neurotransmitters, including M2 and
ICC-IM in stomach 49

M3 receptors, NK1 and NK3 receptors and VIP-1 receptors, as well as non-selective cation
channels such as TRP4 and TRP6 (Epperson et al., 2000), suggesting that electrical signals induced
in ICC-IM by neuronal inputs can conduct to surrounding smooth muscle cells since ICC-IM form
gap junctions with smooth muscle cells (Seki and Komuro, 2002).
Unlike ICC-IM, ICC-MY are not directly innervated by enteric motor neurons in the stomach.
In the gastric antrum of guinea pig, both cholinergic and nitrergic responses are elicited by nerve
stimulation in the circular muscle layer, while nerve stimulation has no effect on ICC-MY (Dickens
et al., 2000). However this is not the case for all regions of the GI tract. It was shown recently
that excitatory motor neurons which release ACh and tackykinins (TKs) innervate both ICC-MY
and smooth muscle cells in the mouse colon (Bayguinov et al., 2010). The activity in ICC-MY
evoked by enteric motor neurons paces the longitudinal and circular muscle layers, which are
synchronized during the colonic migrating motor complexes (CMMC).

Spontaneous electrical activities of ICC-IM

ICC-IM in the antrum


Single bundles of antral circular muscles generate unitary potentials, which sum together to
trigger regenerative potentials (Edwards et al., 1999). In intact tissue preparations isolated from
the antrum, those regenerative potentials (secondary component of slow waves) are superimposed
on passive waves of pacemaker depolarization initiated by ICC-MY (Hirst and Edwards, 2006).
The secondary component of the slow wave is not detected in the antrum of W/Wv mice which lack
ICC-IM (Dickens et al., 2001), indicating that ICC-IM generate these unitary potentials in the
antrum. Thus, antral ICC-IM play a key role in the generation of gastric slow waves. Discharge
of unitary potentials depends on Ca2+ release from internal stores in ICC-IM, since BAPTA-AM,
an intracellular Ca2+ chelator, and cyclopiazonic acid, an inhibitor of internal store uptake of Ca2+,
inhibit the generation of unitary potentials (Suzuki and Hirst, 1999). When regenerative potentials
are initiated by membrane depolarization in isolated circular muscle bundles, the onset of the
initiation of regenerative potentials occurs with a minimum latency of about 1s (Suzuki and Hirst,
1999). It has been suggested that membrane depolarization evokes the production of a second
messenger like inositol 1, 4, 5 trisphosphate (IP3) (Ganitkevich and Isenberg, 1993). In fact, slow
waves are absent in tissues devoid of type-1 IP3 receptors (Suzuki et al., 2000).
As described above, neurally released ACh and NO also selectively target ICC-IM in the
antrum. Cholinergic responses initiate regenerative potentials by resulting from a synchronous
discharge and summation of unitary potentials, while nitrergic responses prevent the discharge of
unitary potentials, giving rise to membrane hyperpolarization (Suzuki et al., 2003).

ICC-IM in the fundus


A single bundle of circular muscle cells from the gastric fundus also displays a discharge of
unitary potentials, which are thought to be generated in ICC-IM because this discharge is not
detected in fundic circular muscle layers isolated from W/Wv mice (Beckett et al., 2004). Recently,
it was demonstrated that ICC-IM generate spontaneous transient depolarizations (STDs) in the rat
50 Y. KITO

gastric fundus (Kito et al., 2009) in two forms: bust-type STDs and continuous-type STDs. Both
types of STDs are generated by the summation of single-unit STDs. It seems likely that ICC-IM
trigger attenuated STDs as unitary potentials in the circular muscle layer through the electrical
connections between ICC-IM and circular muscle layer, as confirmed by the demonstration of the
movement of dye between these cells (Kito et al., 2009).
In the antrum, the frequency of unitary potentials is increased when a single bundle of circular
muscle tissues is depolarized, resulting in the generation of the regenerative potentials (Suzuki
and Hirst, 1999). In other words, antral ICC-IM, which are sensitive to voltage, contribute to slow
wave activity by amplifying the pacemaker depolarizations initiated in ICC-MY. In contrast, in the
fundus, the frequency of unitary potentials is not modulated by changes in membrane potentials
(Beckett et al., 2004). Thus, unitary potentials do not sum to develop regenerative potentials in
the fundus. The simple explanation of those observations is that fundic ICC-IM lack voltage
sensitivity. Alternatively, the electrical properties of fundic muscles and the relatively few gap
junctions in the fundus compared with the antrum (Seki and Komuro, 2002) combined to create an
organ with relatively short syncytial units which do not display slow wave activity.
Sodium nitroprusside (SNP), a NO donor, hyperpolarized the membrane of ICC-IM, and
decreased the frequency of STDs. SNP also hyperpolarized the membrane of circular muscles, and
abolished unitary potentials (Kito et al., 2009). Thus, the role of ICC-IM in the fundus seems to
maintain the membrane at depolarized levels (–40 to –50 mV) to allow the generation of
appropriate responses to inhibitory neurotransmitters in cases of receptive or adaptive relaxations
(Arakawa et al., 1997).

ICC-IM in the corpus


Gastric slow waves are initiated in the corpus which lacks ICC-MY but contains ICC-IM. The
shape of corporal slow waves is similar to the regenerative secondary component of antral slow
waves (Hashitani et al., 2005). Thus, it seems likely that corporal slow waves are generated by
corporal ICC-IM. In isolated antral preparations, slow waves occur irregularly at ~3 cycles min–1
(Dickens et al., 2000). Single bundles of antral circular muscles, with no ICC-MY attached,
generate regenerative potentials at ~1 cycle min–1 (Edwards et al., 1999). On the other hand,
curiously, the corpus generates slow waves at the highest frequency of discharge (5 cycles min–1)
without a network of ICC-MY (Hashitani et al., 2005).
Corporal slow waves trigger large amplitude, square-shaped potential transients in ICC-MY.
These potential changes depolarize antral ICC-IM to initiate the secondary component of antral
slow waves. Rapid circumferential conduction of slow waves by antral ICC-IM triggers a ring of
contraction. In this way, all regions of the stomach generate slow waves at the same frequency as
the corpus, indicating that corporal ICC-IM act as the dominant pacemaker cells (Hirst and
Edwards, 2006).
Taken together, it is concluded that ICC-IM in different regions of the stomach have quite
distinct electrical properties.
ICC-IM in stomach 51

Pathology of ICC in the stomach

As W/Wv mice develop postsurgical ileus like symptoms, loss of ICC may well be involved in
several GI motility disorders such as achalasia, gastroparesis, diabetic gastroenteropathy, chronic
idiopathetic intestinal pseudoobstraction, chronic constipation and inflammatory bowel disease
(Sanders et al., 1999). However, it is currently unknown whether loss of function is the cause of
motor dysfunction or a consequence of the development of motor symptoms, because it usually
takes a long time for patients to develop their syndromes before they are diagnosed as motility
disorders. Further studies are required to elucidate these important questions.
On the other hand, gain-of-function mutations of c-kit genes leads to the development of
gastrointestinal stromal tumors (GISTs) (Hirota et al., 1998). GISTs are the most common
mesenchymal tumors of the human gut. GISTs are considered to originate from ICC, since most
GISTs express mutated Kit, which is constitutively active, resulting in ICC hyperplasia (Hirota and
Isozaki, 2006). GISTs occur mostly in the stomach (60–70%), followed by the small intestine (20–
30%), colon (5%) and the esophagus (<1%) (Agaimy et al., 2007). Interestingly, GISTs are
commonly seen in the gastric cardia, fundus and proximal body in the stomach. As above, fundic
and corporal muscles contain ICC-IM but lack network of ICC-MY. Thus, in the proximal stomach,
ICC-IM rather than ICC-MY seem to be the main subtype of ICC that give rise to GISTs.

Concluding remark

This review summarizes current evidence describing the multiple roles of ICC-IM in the
stomach. Briefly, it has been reported that: 1) ICC-IM act as mediators of enteric motor
neurotransmission in all regions of the stomach. 2) ICC-IM in the fundus generate STDs to
maintain the membrane potentials of smooth muscles at positive levels. 3) ICC-IM in the corpus
provide the dominant pacemaker frequency in the stomach. 4) ICC-IM in the antrum generate the
secondary component of the slow wave. As it has been revealed that ICC play important roles in
the physiological function of smooth muscle tissues in GI tract, we cannot carry out smooth muscle
research in GI tract without considering interstitial cells of Cajal.

Acknowledgements

The author is grateful to Dr. R. J. Lang for critical reading of the manuscript. This project was
supported by research grant from Japan Society for the Promotion of Science to Y. K. (No.
19790460).

References

Agaimy, A., Wunsch,P.H., Hofstaedter, F., Blaszyk, H., Rummele, P., Gaumann, A., Dietmaier, W. and
Hartmann, A. (2007). Minute gastric sclerosing stromal tumors (GIST tumorlets) are common in
adults and frequently show c-KIT mutations. Am. J. Surg. Pathol. 31: 113–120.
Arakawa, T., Uno, H., Fukuda, T., Higuchi, K., Kobayashi, K. and Kuroki, T. (1997). New aspects of gastric
52 Y. KITO

adaptive relaxation, reflex after food intake for more food: involvement of capsaicin-sensitive
sensory nerves and nitric oxide. J. Smooth Muscle Res. 33: 81–88.
Bayguinov, P.O., Hennig, G.W. and Smith, T.K. (2010). Ca2+ imaging of activity in ICC-MY during local
mucosal reflexes and the colonic migrating motor complex in the murine large intestine. J. Physiol.
(Lond.). 588: 4453–4474.
Beckett, E.A., Bayguinov, Y.R., Sanders, K.M., Ward, S.M. and Hirst, G.D. (2004). Properties of unitary
potentials generated by intramuscular interstitial cells of Cajal in the murine and guinea-pig
gastric fundus. J. Physiol. (Lond.). 559: 259–269.
Beckett, E.A., Takeda, Y., Yanase, H., Sanders, K.M. and Ward, S.M. (2005). Synaptic specializations exist
between enteric motor nerves and interstitial cells of Cajal in the murine stomach. J. Comp.
Neurol. 493: 193–206.
Burns, A.J., Lomax, A.E., Torihashi, S., Sanders, K.M. and Ward, S.M. (1996). Interstitial cells of Cajal
mediate inhibitory neurotransmission in the stomach. Proc. Natl. Acad. Sci. USA 93: 12008–
12013.
Dickens, E.J., Edwards, F.R. and Hirst, G.D. (2000). Vagal inhibition in the antral region of guinea pig
stomach. Am. J. Physiol. 279: G388–G399.
Dickens, E.J., Edwards, F.R. and Hirst, G.D. (2001). Selective knockout of intramuscular interstitial cells
reveals their role in the generation of slow waves in mouse stomach. J. Physiol. (Lond.). 531: 827–
833.
Edwards, F.R., Hirst, G.D. and Suzuki, H. (1999). Unitary nature of regenerative potentials recorded from
circular smooth muscle of guinea-pig antrum. J. Physiol. (Lond.). 519: 235–250.
Epperson, A., Halton, W.J., Callaghan, B., Doherty, P., Walker, R.L., Sanders, K.M., Ward, S.M. and
Horowitz, B. (2000). Molecular markers expressed in cultured and freshly isolated interstitial cells
of Cajal. Am. J. Physiol. 279: C529–C539.
Ganitkevich, Vya. and Isenberg, G. (1993). Membrane potential modulates inositol 1, 4, 5-trisphosphate-
mediated Ca2+ transients in guinea pig coronary artery. J. Physiol. (Lond.). 504: 47–55.
Hashitani, H., Garcia-Londono, A.P., Hirst, G.D. and Edwards, F.R. (2005). Atypical slow waves generated
in gastric corpus provide dominant pacemaker activity in guinea pig stomach. J. Physiol. (Lond.).
569: 459–465.
Hirota, S. and Isozaki, K. (2006). Pathology of gastrointestinal stromal tumors. Pathol. Int. 56: 1-9.
Hirota, S., Isozaki, K., Moriyama, Y., Hashimoto, K., Nishida, T., Ishiguro, S., Kawano, K., Hanada, M.,
Kurata, A., Takeda, M., Muhammad Tunio, G., Matsuzawa, Y., Kanakura, Y., Shinomura, Y. and
Kitamura, Y. (1998).Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors.
Science 279: 577–580.
Hirst, G.D., Beckett, E.A., Sanders, K.M. and Ward, S.M. (2002). Regional variation in contribution of
myenteric and intramuscular interstitial cells of Cajal to generation of slow waves in mouse gastric
antrum. J. Physiol. (Lond.). 540: 1003–1012.
Hirst, G.D. and Edwards, F.R. (2006). Electrical events underlying organized myogenic contractions of the
guinea pig stomach. J. Physiol. (Lond.). 576: 659–665.
Huizinga, J.D., Liu, L.W., Fitzpatrick, A., White, E., Gill, S., Wang, X.Y., Zarate, N., Krebs, I., Choi, C.,
Starret, T., Dixit, D. and Ye, J. (2008). Deficiency of intramuscular ICC increases fundic muscle
excitability but does not impede nitrergic innervation. Am. J. Physiol. 294: G589–G594.
Kito, Y., Sanders, K.M., Ward, S.M. and Suzuki, H. (2009). Interstitial cells of Cajal generate spontaneous
transient depolarizations in the rat gastric fundus. Am. J. Physiol. 297: G814–G824.
Komori, K. and Suzuki, H. (1986). Distribution and properties of excitatory and inhibitory junction
potentials in circular muscle of the guinea-pig stomach. J. Physiol. (Lond.). 370: 339–355.
Komuro, T. (2006). Structure and organization of interstitial cells of Cajal in the gastrointestinal tract. J.
Physiol. (Lond.). 576: 653–658.
Sanders, K.M. (1996). A case for interstitial cells of Cajal as pacemakers and mediators of
ICC-IM in stomach 53

neurotransmission in the gastrointestinal tract. Gastroenterol. 111: 492–515.


Sanders, K.M., Ordog, T., Koh, S.D., Torihashi, S. and Ward, S.M. (1999). Development and plasticity of
interstitial cells of Cajal. Neurogastroenterol. Motil. 11: 311–338.
Seki, K. and Komuro, T. (2002). Distribution of interstitial cells of Cajal and gap junction protein, CX 43 in
the stomach of wild-type and W/WV mutant mice. Anat. Embryol. 206: 57–65.
Suzuki, H. and Hirst, G.D. (1999). Regenerative potentials evoked in circular smooth muscle of the antral
region of guinea-pig stomach. J. Physiol. (Lond.). 517: 563–573.
Suzuki, H., Takano, H., Yamamoto, Y., Komuro, T., Saito, M., Kato, K. and Mikoshiba, K. (2000).
Properties of gastric smooth muscles obtained from mice which lack inositol trisphosphate
receptor. J. Physiol. (Lond.). 525: 105–111.
Suzuki, H., Ward, S.M., Bayguinov, Y.R., Edwards, F.R. and Hirst, G.D. (2003). Involvement of
intramuscular interstitial cells in nitrergic inhibition in the mouse gastric antrum. J. Physiol.
(Lond.). 546: 751–763.
Szurszewski, J.H. (1981). Electrical basis for gastrointestinal motility. In Physiology of the Gastrointestinal
Tract, ed. Johnson R, pp. 1435–1465. Raven Press, New York.
Thuneberg, L. (1982). Interstitial cells of Cajal: intestinal pacemaker cells? Adv. Anat. Embryol. Cell. Biol.
71: 1–130.
Wang, X.Y., Sanders, K.M. and Ward, S.M. (1999). Intimate relationship between interstitial cells of Cajal
and enteric nerves in guinea-pig small intestine. Cell Tissue Res. 295: 247–256.
Ward, S.M. and Sanders, K.M. (2001). Interstitial cells of Cajal: primary targets of enteric motor
innervation. Anat. Rec. 262: 125–135.
Ward, S.M. and Sanders, K.M. (2006). Involvement of intramuscular interstitial cells of Cajal in
neuroeffector transmission in the gastrointestinal tract. J. Physiol. (Lond.). 576: 675–682.
Xue, L., Fukuta, H., Yamamoto, Y. and Suzuki, H. (1996). Properties of junction potentials in gastric smooth
muscle of the rat. Jpn. J. Physiol. 46: 123–130.
Zhang, Y., Carmichael, S.A., Wang, X.Y., Huizinga, J.D. and Paterson, W.G. (2009). Neurotransmission in
lower esophageal sphincter of W/Wv mutant mice. Am. J. Physiol. 298: G14–G24.

You might also like