You are on page 1of 11

C H A P T E R

26
Chloride Transport in Glioma Growth
and Cell Invasion
Harald Sontheimer

o u t l i n e

I. Introduction 519 VII. Clinical Use of Chlorotoxin 525


II. Gliomas and their Lineage 520 VIII. Cell Volume Changes Associated with
Cell Proliferation 526
III. Glioma Migration and Invasion 521
IX. Conclusions 528
IV. Cl Transport and Cell Volume Regulation
in Glioma Cells 521 Acknowledgements 528
V. Changes in Cell Volume of Invading Cell References 529
Require Cl Efflux via Clc Channels 523
VI. Mechanism of Chlorotoxin Action on
Glioma Invasion 524

I.  Introduction properties of the organ it originated from. Hence, they


present as liver or lung cells growing within brain and
Brain tumors fall into two principal categories, pri- tumors typically grow as confined solid masses. This
mary and secondary. Primary tumors are often called is not the case for primary brain tumors which often
gliomas and originate in the brain. Secondary or meta- lack clear boundaries between normal and malignant
static brain tumors are peripheral cancers that invade brain tissue. A representative example is illustrated
the brain. Together they account for well over 100,000 in Fig. 26.1A, which shows a cerebral glioma with
new cancer cases diagnosed each year in the USA, characteristic diffuse margins. An important differ-
of which approximately 40,000 are primary tumors ence between these two cancer types relates to how
(according to data from the Central Brain Tumor the tumors spread and form metastasis. Metastatic
Registry of the United States, CBTRUS). In addition brain tumors disseminate hematogenously throughout
to their dissimilar origin, primary and secondary the body and enter the brain through the vasculature.
brain tumors differ in many aspects of their etiology By comparison, primary brain tumors rarely metasta-
and biology. For example, metastatic cancers are eas- size into the periphery but instead spread within the
ily distinguishable from normal brain tissue as they brain often reaching distant sites such as the contralat-
represent the new growth of cancerous tissue with the eral brain hemispheres or the spinal cord, as illustrated

Physiology and Pathology of Chloride Transporters and Channels in the Nervous System 519 © 2009, Elsevier Inc.
520 26.  Chloride Transport in Glioma Growth and Cell Invasion

may not always be of defined lineage. Among the dis-


tinguishing features are immuno-positivity for cer-
tain glial associated antigens (Kleihues et al., 1995),
for example to glial fibrillary acidic protein (GFAP),
myelin associated glycoprotein (MAG), myelin basic
protein (MBP), S100 beta or vimentin. GFAP and/or
S100 positive cells are frequently termed astrocytomas,
MBP- or MAG-positive cells; Oligodendrogliomas
and cells that stain for both sets of markers are mixed
gliomas. While these names imply a known and well-
defined lineage relationship of these tumors with
normal glial cell type or their progenitor cells, such a
relationship has not actually been demonstrated and
the cell types of origin remain controversial. In stud-
ies addressing this question, investigators have trans-
fected glial progenitor cells with known mutations
in oncogenes and tumor suppressor genes and have
been able to induce a malignant transformation that
yielded tumor growth in mice, suggesting that com-
mitted glial progenitor cells may indeed be the most
likely cell type of origin (Dai et al., 2001).
Gliomas exhibit many of the characteristic features
of systemic cancers which include mutations in the
tumor suppressor genes P16 and P53, and amplification
and overexpression of certain oncogenic growth factor
receptors including EGF-R or PDGF-R (Von Deimling
Figure 26.1  Primary glioma at autopsy. A. Poorly defined et al., 1995). As with other cancers, angiogenesis or
margins are characteristic of cerebral gliomas, like the one shown in the induction of new blood vessels in response to the
this example (arrows). B. Although gliomas rarely metastasize out- release of vascular endothelial growth factor is com-
side the brain, they often present secondary tumors in other parts mon (Plate and Risau, 1995). Furthermore, the release
of the brain, often distant from the site of the primary tumor. These of matrix degrading enzymes that facilitate the remod-
secondary tumors are highlighted in B by white ovals. Copyrighted
images: University of Alabama at Birmingham, Department of
eling of the tumor associated extracellular space is com-
Pathology. mon and facilitates cell invasion (Giese et al., 1994).
A glioma diagnosis is almost always fatal as current
treatment options are limited (Butowski et al., 2006).
By the time a tumor is detectable, it has frequently
in the example shown in Fig. 26.1B. These cancers seeded tumor cells throughout the nervous system,
spread by active cell migration without extravasating and upon surgery these cells can quickly give rise to
into the vasculature. This is reminiscent of neuronal recurrent malignancies. The diffuse pattern of cellular
and glial cell migration during brain development or invasion illustrated in Fig. 26.1 not only makes com-
stem cells and microglial cells in the adult brain sug- plete surgical resection impossible, but also limits focal
gesting that some of the underlying mechanisms of treatments such as exogenous beam irradiation as cells
migration may be shared between these cells. remote from the tumor will escape the radiation. Upon
recurrence, many gliomas become even more malig-
nant. Recurrence is believed to result from cells that
have invaded surrounding brain areas. Surprisingly,
II.  Gliomas and their lineage little is known about the underlying mechanisms.
For example, pathways of cell migration are poorly
Although primary brain tumors can originate from understood as are molecules involved in chemotaxis
a number of growth competent cells in the brain or and path finding. These aspects of glioma biology are
spinal cord, the majority of them appear to derive promising areas for future research as they may yield
from glial cells or their precursors. Reflecting this pre- more effective therapeutic tools. An important aspect
sumed relationship these tumors are collectively called of tumor biology that has been well studied in recent
gliomas. These include a diverse group of cancers that years and which will be ­discussed in greater detail in
IV.  Cl Transport and cell volume regulation in glioma cells 521

this chapter pertains to biophysical and biomechanical


adaptations that support the migration and invasion
of gliomas into normal brain tissue. Some of these
findings may pertain to other migratory cells in the
brain and even to other cancers.

III.  Glioma migration and


invasion

As illustrated in Fig. 26.1, the boundaries between


a primary glioma and normal brain tissue are often
difficult to delineate at a macroscopic level. At a
microscopic level, thousands of glioma cells will have
diffusely invaded the surrounding areas of brain tissue,
and, over time, they will have spread to very distant
sites. Wherever possible, invading glioma cells appear
to take advantage of other structures in the brain to
migrate. For example, they frequently migrate along
nerve fiber bundles or, as illustrated in Fig. 26.2A,
along blood vessels. Whether the spaces along these Figure 26.2  Invading glioma cells in situ. A. Confocal images
structures are more favorable for migration, or whether of invading D54MG cells stably expressing EGFP (green). The
invading glioma cells adhere to blood vessels (red). Cells often
there are other guidance cues, or a more slippery extra-
exhibit an elongated wedge-shaped appearance as shown in the
cellular matrix, is not entirely clear. Without question, lower panel. B. The elongated shape is quite apparent in this elec-
however, the narrowness of the extracellular space pro- tron micrograph that captured an invading glioma cell (*, recog-
vides a significant impediment to cell migration. At the nized because of the abundance of ribosomes and other organelles
electron microscopic level, invading cells appear elon- that incorporate lead citrate and give a darker appearance) extend-
ing between normal brain cells, likely astrocytes (based on their
gated, wedge shaped, and with an overall shrunken
large nuclei and abundance of electron dense, glycogen deposits
appearance (Fig. 26.2B). This has led to the hypothesis throughout the cytoplasm). C. Cell shrinkage requires water efflux
that glioma cells may dynamically adjust their cell vol- which is driven by the concomitant efflux of Cl and K through
ume as they invade. As illustrated in cartoon form in their respective ion channels. Glutamate is shown as a possible
Fig. 26.2C, and further discussed below, recent findings motogenic stimulus acting via AMPA receptors that raise intracel-
lular [Ca2] which may in turn activate BK channels. (Panel B is
support this hypothesis and suggest an important role
reproduced with permission from Soroceanu et al., 1999.)
for Cl channels and transporters in this context.

block the K-Cl cotransporters (Ernest et al., 2005).


IV.  Cl Transport and cell Furthermore, volume regulation is supported when Cl
volume regulation in is replaced by halide ions such as I or Br with known
permeability to Cl channels, but not when gluconate
glioma cells substitutes for Cl (Ernest, 2007). These studies suggest
that RVD in glioma cells utilizes Cl as osmolyte, which
As extensively discussed in Chapter 15 in this book, is released from the cell through Cl channels.
all eukaryotic cells have developed powerful mecha- An important question is whether Cl may simi-
nisms to maintain a constant cell volume even when larly act as an osmolyte during cell volume changes
extracellular osmotic conditions change. Glioma cells are associated with cell invasion, a process that is very
not an exception; when exposed to a 50% hyposmotic different from cell volume changes elicited by osmotic
challenge they regulate their volume back to baseline challenges. Under hyposmotic conditions, a gradi-
within just a few minutes. As illustrated in Fig. 26.3A, ent for Cl efflux is favored by the dilution of extra-
this regulatory volume decrease (RVD) is inhibited by cellular ions with water, whereas under isosmotic
drugs known to block Cl channels including NPPB, conditions, this is not the case, unless the cell has a
Cd2 and DIDS with almost complete inhibition by sufficiently high [Cl]i. Hence, the hypothesized cell
NPPB and Cd2 (Ernest et al., 2005). The remaining shrinkage of invading cells requires that intracellu-
volume regulatory response is inhibited by drugs that lar Cl be accumulated so that an outward directed
522 26.  Chloride Transport in Glioma Growth and Cell Invasion

Figure 26.3  Volume regulation in glioma cells. A. On exposure to a 50% hypotonic challenge, glioma cells swell and regulate their vol-
ume back to baseline or even below the baseline level. This regulatory volume decrease is partially inhibited by NPPB or DIDS and is almost
completely inhibited by NPPB and Cd2. B. Glioma cells maintain an elevated intracellular [Cl] which is accumulated via the bumetanide-
sensitive Na-K-Cl cotransporter, NKCC1. Pharmacological inhibition of the cotransporter by 20 μM bumetanide causes a decrease in [Cl]i.
Exposure to 40 μM DIOA causes an increase in [Cl]i above control levels, presumably by inhibition of KCC mediated Cl efflux. C. Western
blot of lysates from two glioma cell-lines D54-MG and U251, and from samples obtained from one patient (GBM50), show prominent expres-
sion of NKCC1 but absence of NKCC2. Rat kidney lysates were used as control for NKCC2. D. Immunostaining also shows prominent mem-
brane associated labeling for NKCC1 in representative U251 glioma cells. Antibodies directed against NKCC1 and NKCC2 were from Alpha
Diagnostics, and were used at a dilution of 1:500. (Panels A and B are reproduced with permission from Ernest et al. (2005), and panel C is
reproduced with permission from Ernest and Sontheimer (2007).)

electrochemical gradient for Cl is maintained. Using [Cl]i was estimated from EGABA (Habela et al., 2009).
the Cl sensitive fluorescence indicator SPQ, intracel- These studies indicated an intracellular [Cl] of
lular [Cl] was measured in glioma cells and found to 105 mM in glioma cells, a value close to that deter-
be around 100 mM (Ernest, 2007), a value far greater mined using SPQ. Hence, glioma cells maintain a
than that typically observed in mature central neurons steep outward directed gradient for Cl. In most cells,
(7–10 mM) , glial cells (30–40 mM) or mature primary Cl is actively accumulated via the Na-K-2Cl
afferent neurons (45 mM) (see Chapters 7, 19 and cotransporter (NKCC1), which is a widely expressed
22 in this volume). These findings were recently con- Cl importer (Chapters 2, 16 and 19 in this volume).
firmed by patch-clamp studies at the single cell level Western blot and immunostaining analyses of sev-
in which the reversal potential of Cl currents was eral glioma cell-lines, including those obtained from
used as an indirect indicator of [Cl]i. Since glioma acute patient biopsies, demonstrated prominent
cells in culture lack ligand-gated Cl channels, such expression of NKCC1 and absence of NKCC2 (see
as the GABAA receptor-channel widely expressed in Fig. 26.3C and 26.3D as well as Ernest and Sontheimer,
neurons, recombinant ligand-gated non-inactivating 2007). Gliomas also express KCC1 and KCC3 (Ernest
Cl channels (GABA-rho) were introduced into glial et al., 2005). Consistent with NKCC1 being princi-
cells, and stable cell-lines expressing GABA-gated pally responsible for the accumulation of intracellu-
Cl channels were created. Gramicidin-perforated lar Cl above electrochemical equilibrium in gliomas,
patch recordings allowed determination of the rever- pharmacological inhibition of the cotransporter with
sal potential of the GABA-induced currents (EGABA). bumetanide causes a significant drop in intracellular
V.  Changes in cell volume of invading cell require Cl efflux via clc channels 523

Cl (Fig. 26.3B). As we will be discussing below, high staining was predominant in invading processes of
intracellular [Cl] is possibly required for immature isolated ­glioma cells. In an effort to further identify the
neurons and glioma cells to migrate. High Cl might Cl channels functioning in gliomas, cells were treated
facilitate water extrusion and cell shrinkage, processes with antisense oligonucleotides to known members
necessary for migrating cells to navigate through con- of the ClC Cl channels’ super family. These stud-
fined extracellular spaces. ies showed prominent expression of currents attrib-
utable to ClC-2 and ClC-3, respectively (Olsen et al.,
2003). ClC-2 currents, known to be sensitive to Cd2,
V.  Changes in cell volume of inwardly rectifying, and potentiated by a negative pre-
pulse to 120 mV, were selectively eliminated in
invading cell require Cl efflux
glioma cells treated with ClC-2 antisense oligonucle-
via clc channels otides (Fig. 26.4A). As expected, these currents were
unaltered by ClC-3 antisense oligonucleotides.
As illustrated in Fig. 26.2C migratory glioma cells ClC-3 channels giving rise to outwardly rectifying
appear to undergo profound changes in cell volume as currents that show time-dependent inactivation and
they invade surrounding tissues. We hypothesize that are sensitive to NPPB were greatly reduced in glioma
these spontaneously occurring changes in cell volume cells treated with ClC-3 antisense oligonucleotides
are driven by efflux of Cl and obligatory water. The (Fig. 26.4B). These data are consistent with both ClC-2
notion that a favorable outward Cl gradient is estab- and ClC-3 channels being functionally expressed in
lished by NKCC1 was experimentally tested in a recent gliomas. However, functional expression of ClC-3
study (Habela et al., 2009). Glioma cells were stably channels in the plasma membrane is controversial, as
transfected with GABA-rho channels, and subjected discussed in detail in Chapter 12 of this volume. ClC-
to volume measurements using a Coulter Counter. 3 knockout mice primarily show CNS pathology asso-
Application of GABA induced opening of Cl chan- ciated with loss of synaptic vesicles in hippocampal
nels which caused progressive cell volume decrease. neurons (Stobrawa et al., 2001). Thus, whether ClC-3
This cell shrinkage was not observed in untransfected protein is able to generate functional channels in the
cells, or in the absence of GABA. This suggests that plasma membrane has been questioned. Immuno-
opening of Cl channels causes efflux of Cl associated gold electron microscopy of human gliomas, however,
with obligatory water loss, and cell shrinkage. These shows immunoreactivity associated with both plasma
experiments also suggest that Cl efflux is sufficient to membrane and intracellular vesicles (Fig. 26.4C).
induce a volume decrease in glioma cells. Interestingly, Further, ClC-3 in cultured glioma cells colocalizes to
these studies were made by inserting a ligand-gated the -subunit of cholera-toxin, which binds to lipid
Cl channel which could be activated on demand, but raft domains, arguing for membrane localization of
glioma cells express a significant resting Cl conduc- the Cl channel (see merged signals in Fig. 26.4D).
tance (Ransom et al., 2001). Indeed, when recorded Figure 26.5A shows that currents with the biophysi-
using perforated patch-clamp technique to avoid dis- cal signature of ClC-3 are inhibited in a dose-dependent
turbing cytosolic Cl, glioma cells exhibit a resting Cl fashion by chlorotoxin (Cltx), a peptide isolated from
conductance sensitive to NPPB and DIDS. These cur- the venom of the scorpion Leiurus quinquestriatus
rents are outwardly rectifying, show time-dependent (DeBin et al., 1993). This toxin might inhibit Cl chan-
inactivation at positive potentials and exhibit the fol- nels with some specificity (McFerrin and Sontheimer,
lowing anion permeability sequence: IBrCl. 2005). Importantly, as shown in Fig. 26.5B–C, when
However, although the currents could be potentiated cells were challenged to cross a transwell barrier that
by cell swelling, this was not required for current acti- mimics the spatial constraints of the extracellular
vation. Because Cl channel inhibitors still lack speci- space in the brain, cell migration across the barrier was
ficity, attributing the inhibitory effect of NPPB and inhibited when the Cl conductance was blocked with
DIDS to a specific ion channel is not yet possible. As NPPB (Ransom et al., 2001), Cd2 or Cltx (Soroceanu
a first step towards the molecular identification of the et al., 1999). Of all these drugs, NPPB and Cltx were
Cl channels expressed in glioma cells, Western blots the most effective inhibitors of cell migration in the
using lysates obtained from gliomas isolated from transwell assays. Cltx in both its native and recom-
patients were probed with antibodies directed against binant form inhibited transwell migration in a dose-
epitopes of cloned Cl channels. These studies dem- dependent fashion (Deshane et al., 2003). Further, a
onstrated the presence of ClC-2, ClC-3 and ClC-5 fluorescently labeled Cltx showed binding to the cell
proteins in all gliomas examined (Olsen et al., 2003). surface of human malignant glioma cells in patient
Further, immunolabeling studies showed that ClC-3 biopsies. Based on these data we proposed ClC-3 as a
524 26.  Chloride Transport in Glioma Growth and Cell Invasion

Figure 26.4  Glioma cells express functional ClC-2 and ClC-3 channels. Using specific antisense oligonucleotides to ClC-2 (A) and ClC-
3 (B), it was possible to identify currents attributable to these channels, respectively. The Western blots illustrate effective reduction in cor-
responding protein expression following antisense treatment, demonstrating specificity of the effects observed in the membrane currents.
C. Immuno-gold EM with antibodies to ClC-3 show clusters of channels at the cell surface (thin white arrow) as well as in intracellular vesicles
(thick white arrow). D. Merged image of triple immunolabeling of cultured D54-MG (human glioma cell line). ClC-3 antibody (labeled with
alexa 546) shows that this protein colocalizes with lipid rafts which are identified by immunolabeling of the beta subunit of cholera-toxin
(fluorescein-conjugated cholera-toxin  subunit). Nuclei were counterstained blue with DAPI. (Panels A and B were modified from Olsen et al.
(2003); C is an unpublished image; and D is reproduced with permission from McFerrin and Sontheimer (2006).)

prime candidate for mediating the Cl fluxes required on its mechanism of action; Cltx has a potential clini-
to accomplish the cell shrinkage needed for glioma cal use as an anti-invasive drug. While biophysical
cell migration. The data further suggest that ClC-3 studies suggested that Cltx inhibits Cl channels in
may be a biological target of Cltx and that the latter glioma cells, it took up to 15 minutes to achieve its
might be a potent inhibitor of glioma cell migration. maximal effect. This long delay questioned a direct
action on the channels; channel-specific toxins typi-
cally act almost instantaneously. Using a His-tagged
recombinant Cltx, Deshane and collaborators were
VI.  Mechanism of chlorotoxin
able to isolate a protein complex and analyzed it by
action on glioma invasion mass-spectroscopy (Deshane et al., 2003). These stud-
ies showed that matrix-metalloproteinase-2 (MMP-2),
The finding that Cltx inhibited Cl channels (Fig. a 72 kD protein that is highly expressed on the surface
26.5A) and was effective in preventing cell invasion of invading glioma cells, could be the primary binding
in transwell assays (Fig. 26.5B) prompted further studies site for Cltx. However, the isolated protein complex
VII.  Clinical use of chlorotoxin 525

Figure 26.5  Inhibition of Cl channels with chlorotoxin retards glioma cell migration. A. Representative currents in response to volt-
age steps, recorded before (control) and 15 minutes after application of 1 M chlorotoxin. Outwardly rectifying, inactivating Cl− currents were
recorded by whole-cell patch-clamp in D54-MG glioma cells using 20 mV voltage steps ranging from −120 to 160 mV. B. To show that a chlo-
rotoxin-sensitive Cl− conductance is required for migration across a spatial barrier, D54-MG glioma cells were plated on the upper surface of a
Transwell insert with 8 μm pores and allowed to migrate for 4 hours towards vitronectin coated on the bottom of the filter insert (top left). Under
control conditions, most cells migrated successfully, indicated by crystal violet staining of cells (control). In the presence of 5 μM chlorotoxin only
a few cells migrated through the filter. C. Chlorotoxin inhibits glioma cell migration. Dose–response curve of D54-MG cells treated with His-Cltx
or commercial Cltx peptide (Alomone) and analyzed by matrigel invasion assay at 24 h post-treatment. Half maximal inhibition (IC50) for Cltx was
184 nM. Percent inhibition was calculated as the decrease in the number of migrated cells normalized to control. (Panel A reproduced with per-
mission from McFerrin and Sontheimer (2006); panels B and C are reproduced with permission from Deshane et al. (2003).)

also contained ClC-3 channels and several regulators


of MMP-2. To further investigate how Cltx may have
VII.  Clinical use of chlorotoxin
decreased Cl channel function, cell surface expres-
sion was examined by biotinylation. These studies In light of the specific binding of Cltx to cultured
showed that upon application of Cltx, membrane glioma cells, it was logical to explore the biological
associated ClC-3 channels gradually disappeared and activity of this molecule in animal models of malig-
were almost undetectable at the surface after 30 min- nant glioma. Using a radiolabeled peptide we dem-
utes (McFerrin and Sontheimer, 2005). Further, it was onstrated its specific binding to human gliomas
observed that the plasma membrane channels ended xenografted into the cerebrum of immunocompro-
up in intracellular caveolar vesicles. In the presence of mised mice (Soroceanu et al., 1998). These studies were
filipin, a sterol-binding drug that prevents the forma- followed by screening of human tissues searching
tion of caveolar vesicles, Cltx lost its effect on ClC-3 for specific binding of Cltx (Lyons et al., 2002). These
channel internalization. This suggested that binding of studies, which examined over 100 samples, revealed
the toxin induces the internalization of ClC-3 channels binding of Cltx to gliomas of all malignancy grades,
together with Cltx into caveolar raft vesicles. These as well as to tumors that share an embryological rela-
findings explain the intracellular trapping of Cltx tionship with them. The latter includes primarily can-
observed in other studies, including those in humans, cers originating from neuroectodermally derived
as discussed below. tissues such as melanoma or small lung cell carcinomas.
526 26.  Chloride Transport in Glioma Growth and Cell Invasion

In contrast, non-malignant tissues were ­universally neg- s­ urvival, following administration of three or six doses
ative. In 2002, a synthetically manufactured Cltx was of Cltx. Importantly, imaging studies such as those
approved by the US Food and Drug Administration illustrated in Fig. 26.6 suggest that Cltx is retained at
(FDA) to be examined in 18 patients in a phase I study. the tumor for 5–8 days. This observation is in good
Like in the previous preclinical studies, this trial used agreement with the internalization of Cltx together
a radiolabeled form of the peptide which was intro- with ClC-3 and MMP-2 into caveolar vesicles. The
duced through an intrathecal catheter. The radiolabel therapeutic efficacy of Cltx is therefore, in all likeli-
could then be detected by whole-body gamma cam- hood, due to (1) the internalization of ClC-3 channels
era scans (Fig. 26.6A), or with greater resolution, by and decreased cell migration and (2) the trapping of
SPEC imaging (Fig. 26.6B). In this study, fluid sam- the radiolabel toxin which could have its own effect on
ples were collected to determine the pharmacokinet- cellular DNA.
ics of the molecule. The data from this clinical study
were published in 2006 (Mamelak et al., 2006). Sample
images like those shown in Fig. 26.6 were published in
2005 (Hockaday et al., 2005). The safety and localiza- VIII.  Cell volume changes
tion data gathered in the phase I trial justified further associated with cell
use of Cltx in 59 patients, in a phase II clinical study
which concluded recently. Preliminary data released
proliferation
from this trial showed a significant increase in mean
In addition to being highly invasive, primary brain
tumors also exhibit relentless growth, with mitotic
indices suggesting that over 30% of high-grade glio-
mas are in the active process of cell division. As cells
divide, they give rise to two daughter cells of approx-
imately half the volume of the parent cell. Yet, within
just a few hours, cell size and volume are restored
in both daughter cells. Surprisingly, little is known
about cell volume changes occurring in dividing cells
in general (see Chapter 27 in this volume). In a recent
study, we imaged complete cycles of cell division
using three-dimensional time-lapsed video micros-
copy following individual cells from birth through
to the next cell division giving rise to new daughter
cells (Fig. 26.7A, B). In this study, cell volume was
obtained from 200 to 400 serial sections at each time
point, allowing relatively accurate cell volume mea-
surements for the entire cell cycle (Fig. 26.7C). We
demonstrated a reduction in cell volume prior to the
M-phase of the cell cycle (Fig. 26.7D), a phenomenon
which we termed ‘‘pre-mitotic volume condensation’’
(Habela and Sontheimer, 2007). Regardless of the
cell volume that a cell maintains during interphase,
it condenses to a volume of approximately 6000-fL
prior to entering into M-phase, approximately 6 h
before giving rise to two daughter cells of approxi-
mately 3000-fL volume (Fig. 26.7D). The condensed
Figure 26.6  The Cl channel inhibitor chlorotoxin localizes to cells have already synthesized the cell membrane of
gliomas in vivo. A. A single dose of 131I-chlorotoxin given to a patient the two daughter cells, as this is readily visible by the
in a phase I clinical study shows tumor-specific localization in whole- thickened membrane (Fig. 26.8A). This finding was
body scans performed over a 5 day period (modified from Shen et al., entirely unexpected, as the common assumption has
2005). B. Overlay of MRI and SPECT images showing tumor-specific
retention of chlorotoxin, 8 days after administration of the drug.
been that cells grow in size continuously until divi-
Axial view of T1-Wc (left), coregistered (middle), and SPECT (right). sion occurs. A contraction of the cytoplasmic volume
(Reproduced with permission from Hockaday et al., 2005.) was not expected. Furthermore, the fact that the cell
VIII.  Cell volume changes associated with cell proliferation 527

Figure 26.7  3D time-lapse imaging of glioma cells division allows accurate determination of cell volume throughout the cell cycle pro-
cess. A. 3D projections created from image z-stacks computed from 200 optical sections such as those shown in B, and rendered in 3D using
ImagePro. This program also computed volumes in fL for each 3D rendered cell. B. Sections from the z-stack used to generate the correspond-
ing projections shown in A. Images are in chronological order from 1 to 5. C. Volume measurements (in fL) at specific time points relative to
division are shown for four cells including the cell in A (green triangle symbols). Time points 1 through 4 correspond to projections 1–4 in B.
For each cell, M-phase was set at t  0 minutes. Note the convergence in volumes immediately before M-phase, where volumes are tightly
clustered around 6000 fL. D. Cells assume a volume of 6000 fL as they reach M-phase, regardless of their volume during interphase (n  14
cells). (Reproduced with permission from Habela and Sontheimer, 2007.)

membrane thickens as the cell volume condenses sug- condensation is accompanied by condensation of
gests that, at this stage, cells have membrane folds nuclear chromatin and indeed, the two processes
ready to be unfolded once a cell division and separa- appear to occur in close synchrony (Habela and
tion of two daughter cells has occurred. Upon divi- Sontheimer, 2007). The initial condensation of the
sion, to achieve normal volume, each daughter cell cytoplasm and hence the chromosomal condensa-
only needs to reaccumulate water through the uptake tion are mediated by the efflux of Cl through the
of Na and Cl, presumably via NKCC1. The vol- same ClC-3 channels that are involved in cell volume
ume changes that may occur through the cell cycle decreases associated with invading cells since shRNA
are illustrated in Fig. 26.8B. Importantly, studies that knock-down of ClC-3 impaired cell condensation
directly compared intracellular [Cl] in M-phase cells (Habela et al., 2008). While pharmacological studies
versus the bipolar interphase cells showed a 40% have long suggested a role for Cl channels in cell
reduction in [Cl]i in the condensed M-phase cells, division, these studies are the first to ascribe a mech-
suggesting that Cl efflux is mechanistically linked anistic role to these channels in cell division; they
to the cell volume reduction (Habela et al., 2009). mediate cytoplasmic condensation through water loss,
Closer examination also showed that cytoplasmic a necessary step for cells to enter the M-phase.
528 26.  Chloride Transport in Glioma Growth and Cell Invasion

Figure 26.8  A. Membrane thickening of M-phase cells following volume condensation are displayed by labeling cells expressing cytosolic
eGFP (green) with membrane bound DiI (red). A. In interphase, the cell membrane associated DiI is a thin membrane layer surrounding the
cytoplasm. In M-phase, this area is thickened suggesting a ruffled membrane. B. Changes in cell volume during the cell cycle illustrated in
cartoon form. As cells progress through G1/S, they increase their plasma membrane area and overall cell volume. As they progress to the M-
phase they condense their cytoplasmic volume but maintain their cell membrane area which becomes thickened and folded. Cell division into
two daughter cells divides membrane and cytoplasm equally between daughter cells. The acquisition of new membrane is accompanied by
uptake of Na, Cl and water through NKCC1, which establishes the normal cell size/volume. (Reproduced with permission from Habela
and Sontheimer, 2007.)

IX.  Conclusions developing neurons, stem cells, and other cell types
which migrate during development prior to settling
down, maturing and forming tissues. Furthermore,
Taken together, the data discussed in this chapter a similar role for Cl channels in the control of cell
suggest that Cl channels and transporters cooperate growth and proliferation may widen the therapeu-
to support dynamic changes in cell volume that gov- tic potential for Cl channel blockers as anti-cancer
ern cell proliferation and cell migration/invasion. The reagents.
outward electrochemical gradient for Cl is estab-
lished by the Na-K-2Cl cotransporter, NKCC1,
and this gradient permits rapid Cl efflux through
Acknowledgements
Cl channels and obligatory water movement
across the plasma membrane, ultimately leading to The author is grateful for the continued support
cell volume reduction. It is conceivable that similar by grants from the National Institutes of Health RO1
mechanisms operate in other migratory cells, e.g. NS-31234, RO1 NS-52634, NS-36692 and P50-CA97247.
REFERENCES 529

References Lyons, S.A., O’Neal, J., and Sontheimer, H. (2002). Chlorotoxin,


a scorpion-derived peptide, specifically binds to gliomas and
Butowski, N.A., Sneed, P.K., and Chang, S.M. (2006). Diagnosis and tumors of neuroectodermal origin. Glia 39, 162–173.
treatment of recurrent high-grade astrocytoma. J. Clin. Oncol. 24, Mamelak, A.N., Rosenfeld, S., Bucholz, R., Raubitschek, A., Nabors, L.B.,
1273–1280. Fiveash, J.B., Shen, S., Khazaeli, M.B., Colcher, D., Liu, A.,
Dai, C., Celestino, J.C., Okada, Y., Louis, D.N., Fuller, G.N., and Osman, M., Guthrie, B., Schade-Bijur, S., Hablitz, D.M., Alvarez, V.L.,
Holland, E.C. (2001). PDGF autocrine stimulation dedifferenti- and Gonda, M.A. (2006). Phase I single-dose study of intracavi-
ates cultured astrocytes and induces oligodendrogliomas and tary-administered iodine-131-TM-601 in adults with recurrent
oligoastrocytomas from neural progenitors and astrocytes in high-grade glioma. J. Clin. Oncol. 24, 3644–3650.
vivo. Genes Dev. 15, 1913–1925. McFerrin, M.B. and Sontheimer, H. (2006). A role for ion channels in
DeBin, J.A., Maggio, J.E., and Strichartz, G.R. (1993). Purification glioma cell invasion. Neuron Glia Biol. 2, 39–49.
and characterization of chlorotoxin, a chloride channel ligand Olsen, M.L., Schade, S., Lyons, S.A., Amarillo, M.D., and Sontheimer, H.
from the venom of the scorpion. Am. J. Physiol. 264, C361–C369. (2003). Expresssion of voltage-gated chloride channels in human
Deshane, J., Garner, C.C., and Sontheimer, H. (2003). Chlorotoxin glioma cells. J. Neurosci. 23, 5572–5582.
inhibits glioma cell invasion via matrix metalloproteinase-2. Plate, K.H. and Risau, W. (1995). Angiogenesis in malignant glio-
J. Biol. Chem. 278, 4135–4144. mas. Glia 15, 339–347.
Ernest, N.J. and Sontheimer, H. (2007). Extracellular glutamine is a Ransom, C.B., O’Neal, J.T., and Sontheimer, H. (2001). Volume-
critical modulator for regulatory volume increase in human gli- activated chloride currents contribute to the resting conductance
oma cells. Brain Res. 1144, 231–238. and invasive migration of human glioma cells. J. Neurosci. 21,
Ernest, N.J., Weaver, A.K., Van Duyn, L.B., and Sontheimer, H.W. 7674–7683.
(2005). Relative contribution of chloride channels and transport- Shen, S., Khazaeli, M.B., Gillespie, G.Y., and Alvarez, V.L. (2005).
ers to regulatory volume decrease in human glioma cells. Am. J. Radiation dosimetry of 131I-chlorotoxin for targeted radiother-
Physiol. Cell Physiol. 288, C1451–C1460. apy in glioma-bearing mice. J. Neurooncol. 71, 113–119.
Giese, A., Rief, M.D., Loo, M.A., and Berens, M.E. (1994). Determinants Soroceanu, L., Gillespie, Y., Khazaeli, M.B., and Sontheimer, H.
of human astrocytoma migration. Cancer Res. 54, 3897–3904. (1998). Use of chlorotoxin for targeting of primary brain tumors.
Habela, C.W., Ernest, N.J., Swindall, A.F., and Sontheimer, H. (2009). Cancer Res. 58, 4871–4879.
Chloride accumulation drives volume dynamics underlying cell Soroceanu, L., Manning, T.J., Jr, and Sontheimer, H. (1999).
proliferation and migration. J. Neurophysiol. 101, 750–757. Modulation of glioma cell migration and invasion using Cl– and
Habela, C.W., Olsen, M.L., and Sontheimer, H. (2008). ClC3 is a criti- K ion channel blockers. J. Neurosci. 19, 5942–5954.
cal regulator of the cell cycle in normal and malignant glial cells. Stobrawa, S.M., Breiderhoff, T., Takamori, S., Engel, D., Schweizer, M.,
J. Neurosci. 28, 9205–9217. Zdebik, A.A., Bosl, M.R., Ruether, K., Jahn, H., Draguhn, A.,
Habela, C.W. and Sontheimer, H. (2007). Cytoplasmic volume con- Jahn, R., and Jentsch, T.J. (2001). Disruption of ClC-3, a chloride
densation is an integral part of mitosis. Cell Cycle 6, 1613–1620. channel expressed on synaptic vesicles, leads to a loss of the hip-
Hockaday, D.C., Shen, S., Fiveash, J., Raubitschek, A., Colcher, D., pocampus. Neuron 29, 185–196.
Liu, A., Alvarez, V., and Mamelak, A.N. (2005). Imaging glioma Von Deimling, A., Louis, D.N., and Wiestler, O.D. (1995). Molecular
extent with 131I-TM-601. J. Nucl. Med. 46, 580–586. pathways in the formation of gliomas. Glia 15, 328–338.
Kleihues, P., Soylemezoglu, F., Schaueble, B., Scheithauer, B.W., and
Burger, P.C. (1995). Histopathology, classification and grading of
gliomas. Glia 15, 211–221.

You might also like