Toward Connecting Metabolism To The Exocytotic Si - 2017 - Trends in Cell Biolog

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Special Issue: Membrane Biology

Opinion
Toward Connecting
Metabolism to the Exocytotic
Site
Mourad Ferdaoussi1 and Patrick E. MacDonald1,*,@
Within cells the regulated exocytosis of secretory granules controls multiple
Trends
physiological functions, including endocrine hormone secretion. Release of the
Multiple receptor-mediated and meta-
glucose-regulating hormone insulin from pancreatic islet b cells is critical for bolic inputs converge at the exocytotic
whole-body metabolic homeostasis. Impaired insulin secretion appears early in site to amplify secretion in response to
a Ca2+-raising stimuli. In pancreatic b
the progression to type 2 diabetes (T2D). Key mechanisms that control the b-cell cells, these control the magnitude of an
exocytotic response, mediating the long-known but little understood metabolic insulin secretory response.
amplification of insulin secretion, are becoming clearer. Recent insights indicate
Glucose not only elicits electrical firing
a convergence of metabolism-driven signals, such as lipid-derived messengers and Ca2+ responses in b cells, but also
and redox-dependent deSUMOylation, at the plasma membrane to augment enhances the exocytotic response to
that Ca2+.
Ca2+-dependent insulin exocytosis. These pathways have important implica-
tions for the metabolic control of hormone secretion, for the functional com- Two key pathways, the glycerolipid/
pensation that occurs in obesity, and for impaired insulin secretion in diabetes. free fatty acid cycle-driven generation
of monoacylglycerol and the mitochon-
drial export of reducing equivalents to
Insulin Secretion, Functional Compensation, and Diabetes generate cytosolic NADPH, are pro-
The regulated secretion of pancreatic endocrine hormones is critical to the physiologic control of posed to underlie a metabolic amplifi-
metabolic homeostasis. Insulin released from b cells of the pancreatic islets of Langerhans is cation of insulin secretion.

required to promote the uptake and storage of circulating glucose following a meal. To achieve
SUMOylation at the exocytotic site is a
the balancing act required to maintain glucose levels within a relatively narrow range, an exquisite key regulator of secretory function in
control of hormone release must be exerted that balances stimulatory and inhibitory signals. multiple cell types, and in b cells is
Moreover, insulin secretory capacity must adjust to prevailing insulin demand. It is widely thought controlled by cytosolic redox signals
downstream of NADPH.
that an increased capacity of islets to secrete insulin represents a compensatory response to
insulin resistance, which occurs in concert with metabolic stress and obesity [1]. Such adaptive
changes take the form of increased islet mass, observed both in models of insulin resistance in
rodents (i.e., [2]) and in obese nondiabetic humans [3–5]. However, in humans this increase may
be marginal [6] and of unclear ontology [7] and recent work suggests a parallel adaptive increase
in insulin secretory capacity that likely occurs on a cell-by-cell basis [8–10]. This latter effect,
which can be termed a ‘functional’ compensation, has received relatively less attention but
recent work suggests that a failure of such adaptive responses likely precedes the establishment
of type 2 diabetes (T2D) [9], the tipping point for which occurs when insulin secretion from
genetically predisposed b cells [11] fails to compensate for insulin resistance (Figure 1A) [12]. The
recognition of a causative role for b-cell dysfunction in T2D has prompted a recent call for a
1
Department of Pharmacology and
Alberta Diabetes Institute, University of
‘b-cell centric’ definition of diabetes [13]. Alberta, Edmonton, AB, Canada T6G
2E1
Insulin secretion from pancreatic islet b cells is triggered by the initiation of electrical activity and
subsequent Ca2+-regulated exocytosis (Figure 1B); a process which shares much in common *Correspondence:
pmacdonald@ualberta.ca
with other regulated secretory systems [14]. Briefly, stimulus-secretion coupling in the b cell (P.E. MacDonald).
involves the sensing of extracellular glucose concentrations and the transduction of this into an @
Twitter: @bcellorg

Trends in Cell Biology, March 2017, Vol. 27, No. 3 http://dx.doi.org/10.1016/j.tcb.2016.10.003 163
© 2016 Elsevier Ltd. All rights reserved.
(A) Obesity (B)
Glucose

Hyperinsulinemia Insulin resistance Mito.


VDCC
Insulin
β-cell granules
compensaon
Insulin
β-cell dysfuncon and
decompensaon

Hormones,
Hyperglycemia cytokines,
type 2 diabetes (T2D) transmiers

Figure 1. Type 2 Diabetes and the Pancreatic b Cell. (A) Enhanced function and mass of b cells occur in concert with
hyperinsulinemia, obesity, and insulin resistance. This is generally considered to be a compensatory response that
maintains normoglycemia. Type 2 diabetes results when these adaptations fail. (B) A consensus model for glucose-
stimulated insulin secretion from pancreatic b cells. Glucose triggers a metabolism-dependent depolarization through the
ATP-dependent closure of ATP-sensitive K+ (KATP) channels. This elicits action potential firing and entry of Ca2+ through
voltage-dependent Ca2+ channels (VDCCs). Multiple receptor-mediated and metabolic coupling factor-mediated pathways
(green broken lines) act to amplify the exocytotic response of insulin granules to Ca2+. Abbreviation: Mito, mitochondrion.

electrical signal. This is accomplished through the regulation of ATP-sensitive K+ (KATP) channels
by a glucose-dependent control of the cytoplasmic ATP-to-ADP ratio. Closure of these channels
when glucose increases results in action potential firing and Ca2+ entry that triggers exocytosis
mediated by the soluble NSF attachment protein receptor (SNARE) proteins. However, like many
peptide-secreting cells, pancreatic b cells store as much as 100-fold more insulin than is
released in response to physiologic, or even supraphysiologic, stimuli (see [15,16] for recent
comparison of insulin release vs. content from human islets). One might ask: ‘What determines
whether secretory granules are stored or released in response to any given stimulus?’, ‘Can cells
control these processes to adapt to prevailing insulin requirements?’, and ‘What role does this
have in determining the success or failure of compensatory responses?’

Following their biogenesis, insulin granules must be trafficked to the cell periphery, where they
are biochemically ‘primed’ to attain release competence. These key aspects of secretory
function have been reviewed recently in detail [14]. Thus, while insulin secretion rises in response
to a glucose-dependent increase in intracellular Ca2+, the magnitude of the secretory response
can also be regulated by processes that mediate the conversion of mature secretory granules
from a storage to a releasable pool by controlling granule trafficking and priming. These latter
pathways contribute as much as half of the glucose-regulated secretory response [17]. In short,
in pancreatic endocrine b cells, glucose-stimulated electrical signals determine when insulin
secretion occurs and other pathways are critical in determining how much insulin is released.
Recent findings have illuminated key signals controlling insulin exocytotic competence to
modulate secretory capacity.

Triggering and Amplifying Insulin Secretion


In response to a step-wise increase in glucose, insulin secretion from isolated islets occurs in a
biphasic manner [16]: a rapid first-phase peak is followed by a prolonged, and often rising,
second phase of insulin secretion. In its simplest form, the consensus model for glucose-
stimulated insulin secretion is shown in Figure 1B, where the major metabolic signal controlling
insulin secretion is a glucose-dependent rise in the intracellular ATP-to-ADP ratio and inhibition of
KATP channels to depolarize the cell, elicit action potential firing, and allow Ca2+ entry largely
through L-type Ca2+ channels. Because the resulting rise in intracellular Ca2+ initiates the

164 Trends in Cell Biology, March 2017, Vol. 27, No. 3


SNARE-mediated fusion of insulin-containing secretory granules with the plasma membrane,
and is absolutely required for insulin secretion, this is often referred to as a ‘triggering’ pathway
that is essentially similar to well-conserved excitation–secretion coupling processes in many
excitable cells. However, a rise in intracellular Ca2+ is only one facet controlling the physiological
secretory response, and while ATP-dependent processes are required for secretory granule
exocytosis [18–20], there are pathways in addition to glucose-dependent ATP:ADP signaling
and Ca2+ increases that act as ‘amplifying’ mechanisms to modulate the magnitude of the
secretory response. Such pathways may be of key importance since they effectively control the
‘secretory capacity’ of b cells, determining how much of the cellular insulin content is available for
release at any given moment. While the insulin content of a cell may be an important determinant
of secretion (many groups, including ours, often present secretory data as a percentage of total
content), the exact correlation between content and secretion remains less than clear, particu-
larly in humans. For instance, unpublished data of 101 nondiabetic human islet preparations
from our human islet program [15] suggest that, while absolute insulin secretion positively
correlates with insulin content (p < 0.001), this association is weak (R2 = 0.13) and cannot
account for most variation in the magnitude of human islet in vitro insulin responses. Additional
mechanisms, perhaps with genetic underpinnings [11], contribute to the variation in human
insulin secretory capacity.

Generally speaking, the moment-to-moment amplitude of a secretory response can be impacted


by several potential mechanisms (Figure 1B). The most prominent example of receptor-mediated
amplification is the enhancement of insulin secretion by G-protein-coupled incretin receptors,
which forms the basis for the action of newer classes of diabetes therapeutics such as the incretin
mimetics. Activation of the glucagon-like peptide-1 receptor, for example, augments the secre-
tory response to glucose through cAMP-dependent and cAMP-independent mechanisms [21].
Having recently commented on this [22], it will not be discussed in further detail here. However,
the mechanisms underlying another key pathway controlling b-cell secretory capacity, the
metabolic amplification of insulin secretion, are only now becoming clear.

Metabolic Amplification of Insulin Exocytosis


More than 20 years ago a classic series of experiments [23,24] demonstrated that glucose
metabolism itself exerts a control of insulin secretion that is separate from its ability to elicit
electrical activity and a rise in Ca2+. Originally referred to as a KATP-independent pathway
regulating insulin secretion, this is now more appropriately referred to as the metabolic amplifi-
cation of insulin secretion [25]. In short, one or more metabolic coupling factors contribute to
increasing exocytotic competence separately from the induction of electrical activity. Much of
this effect likely occurs at or near the exocytotic site, after the regulated trafficking of secretory
granules to the plasma membrane [26]. The nature of the coupling factors responsible for the
metabolic amplification of exocytosis has been debated, although one of the earliest proposed
was glutamate [27], further support for which is found in recent studies of islet-specific
mitochondrial glutamate dehydrogenase knockout mice [28].

Numerous additional cytosolic factors influenced by metabolism have since been suggested to
play key coupling roles, and many have been discussed in detail elsewhere [29,30]. Among
these, at least two pathways (Figure 2A) have emerged as important contributors to the
metabolic amplification of insulin release. One is defined by the malonyl-CoA-dependent
inhibition of carnitine palmitoyltransferase-1, which increases the activity of the glycerolipid/free
fatty acid (GL/FFA) cycle. This results in the cytosolic generation of monoacylglycerol (MAG)
species [31,32], which activate Munc13-1 translocation to the plasma membrane where it
promotes granule priming [33] to enhance Ca2+-induced exocytosis, presumably by inducing
structural changes in syntaxin-1A [34]. A second pathway involves the generation of NADPH
[35–38] via cytosolic isocitrate dehydrogenase [39], malic enzyme [40], or the pentose

Trends in Cell Biology, March 2017, Vol. 27, No. 3 165


(A) Glucose
Malate Pentose
phosphate
pathway
ME G6PDH
IMP
6PGD
ADSS S-AMP

Citrate
NADPH
CIC ACO1 ?
Mitochondrial matrix

Isocitrate SENP1 S2
NADPH 2 GSH
ICDc GRX1
GSR
NADP+ SENP1 SH2
GSSG

Glutathione
biosynthesis
OGC α-KG GOT1
+
Glutamate

brane
cAMP
GC1

a mem
MPC Acetoacetate
? ATGL
DAG

Plasm
DAGL
Acetyl-CoA
TG HSL
Malonyl-CoA MAG
ABHD6

CPT-1 Fay acids


GL/FFA
Glucose cycle

(B)
Glucose
Glucose/FFA

NADPH
HSL
DAGL
GSH
s
Tomosyn

s s
s
MAG
SENP1
SENP1
Ca2+
Munc13-1 s
Synaptotagmin
Tomosyn

s
Syntaxin1a

Syntaxin1a

VDCC

Figure 2. Hypothesis for the Metabolic Amplification of Insulin Exocytosis. (A) The mitochondrial export of reducing
equivalents, and possibly activity of the pentose phosphate pathway, results in cytosolic generation of NADPH, which acts
via glutaredoxin1 (GRX1) to enhance SUMO-protease activity at the exocytotic site (green). Additional signals, such as

(Figure legend continued on the bottom of the next page.)

166 Trends in Cell Biology, March 2017, Vol. 27, No. 3


phosphate pathway [41]. Although a positive role for cytosolic NADPH in insulin secretion has
been questioned [42] in favor of a mechanism involving the (iso)citrate-dependent generation of
acetyl-CoA via ATP citrate lyase (ACLY) and downstream protein acylation [43], clear evidence
for this remains lacking. ACLY knockdown or inhibition is reported both to impair insulin
secretion [40,44] and to have no effect [45,46]. In addition, a detailed new study [47] shows
that neither knockdown nor pharmacological inhibition of ACLY affects acetyl-CoA generation or
insulin secretion, implicating an alternative mechanism for glucose-dependent acetyl-CoA
increases via mitochondrial export of acetoacetate, which could regulate the GL/FFA cycle
by producing malonyl-CoA (Figure 2A).

Thus, the cytosolic generation of NADPH likely contributes to the metabolic amplification of insulin
secretion, and current thinking suggests that this occurs by the regeneration of reduced
glutathione (GSH) and signaling via glutaredoxin1 (GRX1) [36,37]. Clearly, much remains to
be elucidated regarding these upstream metabolic amplifying pathways, including how they
may interact. For example, the mitochondrial production of glutamate, in addition to a direct role in
exocytosis that appears cAMP dependent [48], may contribute to NADPH-dependent signaling
by maintaining the overall cytosolic pool of GSH [35] (Figure 2A). In addition, cooperation between
the GL/FFA cycle and this pathway seems possible when one considers that Munc13-1-
dependent granule priming could facilitate secretion in concert with redox-dependent post-
translational modifications downstream of the NADPH–GSH–GRX1 pathway shown in
Figure 2B and discussed in further detail in the following section.

Redox, Small Ubiquitin-like Proteins, and Insulin Secretion


The post-translational modification of proteins with small ubiquitin-like modifier (SUMO) peptides
(called SUMOylation), of which there are several isoforms (a fifth being just recently identified
[49]), occurs at target lysine residues. Unlike ubiquitylation however, SUMO attachment does not
generally signal for proteasomal degradation, but instead regulates target function and protein–
protein interactions [50]. SUMOylation has a diverse range of physiological functions, including
classical roles in transcription factor targeting and cell cycle regulation along with emerging roles
in mitochondrial function, metabolism, and metabolic sensing. It is these latter roles which may
play key functions in coupling nutrient sensing to hormone secretion.

SUMOylation via SUMO1 blocks glucose and Ca2+-dependent insulin exocytosis at a point
downstream of insulin granule targeting to the plasma membrane [51]. This inhibitory effect on
insulin secretion may act as a key ‘brake’ to limit insulin secretion. SUMOylation is reversed by
the family of sentrin/sumo-specific proteases (SENPs; [52]). Although there are six SENP
isoforms, beyond SENP1 we know little about the role that most of these may play in exocytosis.
However, consistent with an inhibitory effect of SUMOylation on insulin exocytosis, loss of

adenylosuccinate (S-AMP) generated downstream of the pentose phosphate pathway (purple) and mitochondrial glutamate
export (blue), may feed into this (although the latter may also affect secretory granules directly, in a cAMP-dependent manner).
In concert, insulin granule priming is increased by the glucose (and free fatty acid)-dependent generation of monoacylglycerol
(MAG) species via the glycerolipid/free fatty acid (GL/FFA) cycle (yellow), which is upregulated upon carnitine palmitoyl-
transferase 1 (CPT-1) inhibition by malonyl-CoA produced from acetoacetate exported to the cytosol, possibly via the
mitochondrial pyruvate carrier (MPC). (B) In b cells, glucose-dependent deSUMOylation and MAG-dependent signaling may
work together to increase the pool of releasable insulin granules. A deSUMOylation-dependent release of syntaxin-1A from
tomosyn could provide substrate on which MAG-activated Munc13-1 can act to enhance granule priming. Many other
regulatory events, such as deSUMOylation of synaptotagmin VII, may also occur to facilitate exocytosis. Abbreviations: 6PGD,
6-phosphogluconate dehydrogenase; ABHD6, alpha/beta-Hydrolase domain containing 6; ACO1, aconitase 1; ADSS,
adenylosuccinate synthase; ATGL, adipose triglyceride lipase; CIC, citrate carrier; DAG, diacylglycerol; DAGL, diacylglycerol
lipase; G6PDH, glucose-6 phosphate dehydrogenase; GC1, glutamate carrier 1; GOT1, cytosolic aspartate aminotransfer-
ase; GSH, glutathione; GSR, glutathione reductase; GSSG, oxidized glutathione; HSL, hormone sensitive lipase; ICDc,
cytosolic isocitrate dehydrogenase; IMP, inosine monophosphate; ME, malic enzyme; MPC, mitochondrial pyruvate carrier;
OGC, 2-oxoglutarate carrier; S, SUMO1; SENP1, sentrin/SUMO-specific protease 1; SUMO, small ubiquitin-like modifier; TG,
triglyceride; VDCCs, voltage-dependent Ca2+ channels; /-KG, /-ketoglutaric acid.

Trends in Cell Biology, March 2017, Vol. 27, No. 3 167


SENP1 in islets results in impaired insulin exocytosis, insulin secretion, and glucose homeostasis
[35]. Similar to deubiquitinating enzymes [53], this SENP isoform is sensitive to the cellular redox
state [54], which may underlie its sensitivity to hypoxia [55]. Indeed, a GSH-dependent increase
in SENP1 activity at the exocytotic site facilitates insulin exocytosis [35] although some key
questions, such as why SENP1 localizes to secretory granules in insulin-producing cells as
opposed to the nucleus, where it resides in most other cells, remain to be studied. Both nuclear
localization [56] and export [57] sequences within the enzyme have been reported.

SUMOylation is regulated by redox, and in particular has been implicated in responses to


oxidative stress. Underlying mechanisms include the reactive oxygen species-dependent con-
trol of SUMO-ligase activity [58], and inactivation [54] or stabilization [59] of SENPs. Considered
mostly in respect to pathophysiological oxidative stress, oxidative signals may increase or
decrease SUMOylation depending on the strength and timing of the redox change and the
target(s) in question, a topic that has been recently reviewed [60]. The pancreatic b cell is unique
in its cytosolic redox control. Insulin secreting cells express low levels of some antioxidant
proteins, leaving them vulnerable to oxidative damage [61], while key redox regulating proteins
(such as GRX1 [36,37]) appear to have been co-opted into a signaling rather than protective role.
Indeed, glucose stimulates an increase in NADPH, which has now been demonstrated through
metabolomic [41], biochemical [35,36], and imaging [38] approaches, the latter confirming a rise
in cytosolic NADPH. This signals through GRX1 to amplify insulin exocytosis [36,37] and our
recent work connects this to deSUMOylation at the exocytotic site [51] through a direct
facilitation of SENP1 activity by GSH and GRX1 [35], presumably by modulating key thiol
groups within the enzyme [54]. Additional signals, such as pentose phosphate pathway-derived
adenylosuccinate (Figure 2A), may also feed into the control of insulin exocytosis by SENP1 [62].
Effectively, we hypothesize that the glucose-dependent export of reducing equivalents from the
mitochondria signals through NADPH–GSH–GRX1 to promote deSUMOylation via SENP1 at
the exocytotic site. This is supported by the significant impairment in the metabolic amplification
of insulin secretion and subsequent reduction in glucose tolerance in islet-specific SENP1
knockout mice [35].

One key question that remains is the identity of the SUMO-dependent targets at the exocytotic
site. We showed that the exocytotic Ca2+ sensor synaptotagmin VII appears deSUMOylated
upon glucose stimulation [51]. Subsequently, several other exocytotic and granule-trafficking
proteins have been shown to be SUMOylated. These include syntaxin-1A [63], synapsin1a [64],
RIM1/ [65], Kv2.1 [66], and tomosyn [67,68]. With the exception of synapsin [69], all these play
known roles in the regulation of insulin exocytosis [70–74]. Of the aforementioned targets,
tomosyn (of which there are two major isoforms) may be particularly interesting given its
structural role in binding of syntaxin-1A [75,76]. Syntaxin-1A can itself be SUMOylated (although
the authors of this work propose a role in endocytosis, rather than a direct regulation of syntaxin-
dependent exocytosis by SUMOylation [63]) but also contains at least two putative cytoplasmic
SUMO-interacting motifs (based on amino acid sequence analysis [77]). Thus, a glucose-
dependent release of syntaxin-1A from tomosyn could be mediated by a change in SUMOylation
status and/or degradation of tomosyn [78]. It may be interesting to consider whether we can
connect a glucose-dependent release of syntaxin-1A from this inhibitory interaction to the MAG-
dependent activation of syntaxin-1A via Munc13-1 (Figure 2B).

Most likely, a concerted control of the SUMOylation of numerous targets at the exocytotic site
underlies the regulation of exocytosis. Perhaps this can explain apparent discrepant findings
from us and others that suggest either a positive role [64,65,79] or a negative role [35,51] for
SUMOylation in exocytosis that may be cell-type dependent, even between islet cell types
[51,79]. In glucagon-secreting pancreatic /-cells, for example, increased SUMOylation enhan-
ces depolarization-induced exocytosis [79]. While this might be consistent with an increase in

168 Trends in Cell Biology, March 2017, Vol. 27, No. 3


Outstanding Questions
No diabetes Type 2 diabetes
What effects do metabolic coupling
Stored Genec risk; factors that augment secretory
Stored
olic insulin stress; inflammaon;
responses have on secretory dynam-
tab on reacve oxygen species insulin
Me lifica
am
p ics? For example, what roles do these
Increased β-cell Releasable Increased β-cell play in the classical biphasic nature of
Releasable
blood glucose metabolism Ele insulin blood glucose metabolism insulin insulin secretion? Do these signals
c
sig trica
na l
ls control both first- and second-phase
Insulin Insulin
secretion?
secreon secreon

Can we reconcile multiple putative met-


abolic coupling factors controlling the
amplification of insulin secretion, par-
Figure 3. Contribution of Metabolic Amplification to the Control of Secretory Capacity, and Potential Role in Type 2
ticularly the two pathways described
Diabetes. Metabolic amplifying factors control key steps in the conversion of stored insulin to a ‘releasable’ form at the
herein? Do these pathways cooperate
plasma membrane. Glucose-stimulated electrical activity and Ca2+ entry then trigger insulin secretion from that ‘releasable
to augment secretion? If so, how?
pool’. Some studies, including our own, suggest that the glucose-dependent amplification pathways operate at glucose
levels lower than those required to stimulate electrical activity. As such, these pathways could serve to maintain or control
the size of the releasable granule pool before secretion occurs. Early in type 2 diabetes these pathways may be impaired, What controls SUMOylation at the
resulting in reduced maintenance of the ‘releasable insulin’ pool and manifesting as an impaired first-phase secretory exocytotic site, and to what extent
response. Abbreviation: EC50, half maximal effective concentration. does nuclear versus cytoplasmic local-
ization of SUMO-proteases (particularly
SENP1) differ between cell types and
glucagon exocytosis under hypoglycemic conditions, it remains unclear whether the glucose- under different conditions?
dependent ‘amplifying’ pathways discussed here impact glucagon secretion, either positively or
negatively. Regardless, the exact milieu of SUMO targets at the membrane in any given cell may What are the ultimate SUMOylated tar-
get(s) that regulate secretion at the
determine the physiological response to upregulating or downregulating SUMOylation. Answer- exocytotic site? Can one target explain
ing this question will require not just additional insight into potential SUMOylation sites within the SUMO-dependent control of exo-
exocytotic and related proteins, but detailed information on the cell-type-specific impact of cytosis, or do we need to consider the
SUMOylation status of multiple pro-
SUMOylation on protein–protein interactions at the exocytotic site.
teins in complex? Can differences in
the exocytotic machinery complex
Amplification of Exocytosis, Secretory Compensation, and Type 2 Diabetes between cells explain apparent dis-
The metabolic amplifying pathway appears to operate at glucose concentrations below that crepancies of whether SUMO blocks
required to initiate electrical activity and action potential firing [25] (Figure 3). The relevance of this or enhances regulated exocytosis?

finding (which is consistent with our observations on the glucose-dependent amplification of


What roles do these pathways have in
insulin exocytosis [35]) has been largely underappreciated. Although additional confirmation of early compensatory increases in secre-
this is needed, the implication is that metabolic amplifying pathways are active even at ‘basal’ tion that occur in the context of insulin
glucose levels, and can act to maintain a pool of releasable granules and thus control the size of resistance? And conversely, is failure
of this metabolic-exocytotic signaling
both the first- and second-phase insulin responses. Subtle changes in the efficiency at which
causative of impaired insulin secretory
secretory granules are primed (or otherwise attain release competence) can have important responses in type 2 diabetes?
implications to the secretory capacity of an islet – in compensation for insulin resistance for
example. Recent findings consistent with this suggest an enhanced ‘efficacy’ of Ca2+ to
stimulate insulin release early in b-cell adaptation to insulin resistance [10], which may be related
to enhanced metabolic signaling [80]. Presumably, an ability to maintain such functional
adaptations would prevent diabetes [12], and indeed new data provide some evidence for
impaired regulation of exocytotic competence early in the progression to diabetes in animals [9]
and in b cells from established human type 2 diabetes [35]. Is it possible that the metabolic
control of secretory capacity (i.e., operation of a metabolic amplification pathway at subthreshold
glucose) is impaired early in diabetes, reducing the releasable granule pool and manifesting as
the well-known loss-of-first-phase secretion seen in type 2 diabetes?

Concluding Remarks
Multiple metabolic signaling pathways converge to control Ca2+-regulated exocytosis. In pan-
creatic b cells, the mitochondrial generation of ATP triggers electrical activity, Ca2+ influx, and
contributes to energy-requiring processes in secretory granule priming. Additional metabolism-
driven signals amplify the exocytotic response by increasing the pool of releasable insulin
granules. Key emerging pathways include the GL/FFA cycle-dependent production of MAG

Trends in Cell Biology, March 2017, Vol. 27, No. 3 169


and the generation of cytosolic NADPH, which ultimately signals through SENP1 to promote
deSUMOylation at the exocytotic site. Differences in SUMO targets may contribute to cell-type-
specific differences in the role of SUMOylation in the control of exocytosis. Several key questions
remain (see Outstanding Questions), including the contribution of metabolic amplification to both
first- and second-phase insulin secretion, the potential for these multiple pathways to interact in
the control of exocytosis, their contribution to functional adaptation to insulin resistance, and their
role in impaired insulin secretion that precipitates T2D. Understanding the pathways that amplify
insulin secretion provides hope that early intervention to maintain appropriate secretory capacity,
for example, by targeting the rescue of metabolic amplification as recently suggested [81], can
contribute to improved diabetes treatment. More broadly, convergence of multiple signaling
pathways connecting metabolism to the exocytotic site may have important implications for many
endocrine and neuronal functions, such as gut hormone release and central nutrient sensing.

Acknowledgments
Research on pancreatic islet biology in the MacDonald laboratory is funded by the Canadian Institutes of Health Research,
the Canadian Diabetes Association, and the Alberta Diabetes Foundation. P.E.M. is supported by a Killam Annual
Professorship.

References
1. Solomon, T.P.J. et al. (2014) Determining pancreatic b-cell com- 18. Eliasson, L. et al. (1997) Rapid ATP-dependent priming of secre-
pensation for changing insulin sensitivity using an oral glucose tory granules precedes Ca2+-induced exocytosis in mouse pan-
tolerance test. Am. J. Physiol. Endocrinol. Metab. 307, E822–E829 creatic B-cells. J. Physiol. (Lond.) 503, 399–412
2. El Ouaamari, A. et al. (2016) SerpinB1 promotes pancreatic b cell 19. Takahashi, N. et al. (1999) Post-priming actions of ATP on Ca2
+
proliferation. Cell Metab. 23, 194–205 -dependent exocytosis in pancreatic beta cells. Proc. Natl. Acad.
3. Butler, A.E. et al. (2003) Beta-cell deficit and increased beta-cell Sci. U.S.A. 96, 760–765
apoptosis in humans with type 2 diabetes. Diabetes 52, 102–110 20. Barg, S. et al. (2001) Priming of insulin granules for exocytosis by
4. Saisho, Y. et al. (2013) b-Cell mass and turnover in humans: granular Cl–uptake and acidification. J. Cell. Sci. 114, 2145–2154
effects of obesity and aging. Diabetes Care 36, 111–117 21. Shigeto, M. et al. (2015) GLP-1 stimulates insulin secretion by
5. Klöppel, G. et al. (1985) Islet pathology and the pathogenesis of PKC-dependent TRPM4 and TRPM5 activation. J. Clin. Invest.
type 1 and type 2 diabetes mellitus revisited. Surv. Synth. Pathol. 125, 4714–4728
Res. 4, 110–125 22. Kolic, J. and Macdonald, P.E. (2015) cAMP-independent effects of
6. Rahier, J. et al. (2008) Pancreatic beta-cell mass in European GLP-1 on b cells. J. Clin. Invest. 125, 4327–4330
subjects with type 2 diabetes. Diabetes Obes. Metab. 10, 32–42 23. Gembal, M. et al. (1992) Evidence that glucose can control insulin
7. Linnemann, A.K. et al. (2014) Pancreatic b-cell proliferation in release independently from its action on ATP-sensitive K+ chan-
obesity. Adv. Nutr. 5, 278–288 nels in mouse B cells. J. Clin. Invest. 89, 1288–1295

8. Gonzalez, A. et al. (2013) Insulin hypersecretion in islets from diet- 24. Sato, Y. et al. (1992) Dual functional role of membrane depolari-
induced hyperinsulinemic obese female mice is associated with zation/Ca2+ influx in rat pancreatic B-cell. Diabetes 41, 438–443
several functional adaptations in individual b-cells. Endocrinology 25. Henquin, J.C. (2000) Triggering and amplifying pathways of regu-
154, 3515–3524 lation of insulin secretion by glucose. Diabetes 49, 1751–1760
9. Do, O.H. et al. (2016) Changes in beta cell function occur in 26. Mourad, N.I. et al. (2010) Metabolic amplifying pathway increases
prediabetes and early disease in the Lepr (db) mouse model of both phases of insulin secretion independently of beta-cell actin
diabetes. Diabetologia 59, 1222–1230 microfilaments. Am. J. Physiol. Cell Physiol. 299, C389–C398
10. Chen, C. et al. (2016) Alterations in b-cell calcium dynamics and 27. Maechler, P. and Wollheim, C.B. (1999) Mitochondrial glutamate
efficacy outweigh islet mass adaptation in compensation of insulin acts as a messenger in glucose-induced insulin exocytosis. Nature
resistance and prediabetes onset. Diabetes 65, 2676–2685 402, 685–689
11. Rosengren, A.H. et al. (2012) Reduced insulin exocytosis in human 28. Vetterli, L. et al. (2012) Delineation of glutamate pathways
pancreatic b-cells with gene variants linked to type 2 diabetes. and secretory responses in pancreatic islets with b-cell-specific
Diabetes 61, 1726–1733 abrogation of the glutamate dehydrogenase. Mol. Biol. Cell 23,
12. Ha, J. et al. (2015) A mathematical model of the pathogenesis, 3851–3862
prevention and reversal of type 2 diabetes. Endocrinology 157, 29. Maechler, P. (2013) Mitochondrial function and insulin secretion.
624–635 Mol. Cell Endocrinol. 379, 12–18
13. Schwartz, S.S. et al. (2016) The time is right for a new classification 30. Prentki, M. et al. (2013) Metabolic signaling in fuel-induced insulin
system for diabetes: rationale and implications of the b-cell-centric secretion. Cell Metab. 18, 162–185
classification schema. Diabetes Care 39, 179–186 31. Zhao, S. et al. (2014) //b-Hydrolase domain-6-accessible mono-
14. Thorn, P. et al. (2016) Exocytosis in non-neuronal cells. J. Neuro- acylglycerol controls glucose-stimulated insulin secretion. Cell
chem. 137, 849–859 Metab. 19, 993–1007
15. Lyon, J. et al. (2015) Research-focused isolation of human islets 32. Zhao, S. et al. (2015) //b-Hydrolase domain-6 and saturated long
from donors with and without diabetes at the Alberta Diabetes chain monoacylglycerol regulate insulin secretion promoted by
Institute IsletCore. Endocrinology 157, 560–569 both fuel and non-fuel stimuli. Mol. Metab. 4, 940–950
16. Henquin, J-C. et al. (2015) Dynamics of glucose-induced insulin 33. Kang, L. et al. (2006) Munc13-1 is required for the sustained
secretion in normal human islets. Am. J. Physiol. Endocrinol. release of insulin from pancreatic beta cells. Cell Metab. 3,
Metab. 309, E640–E650 463–468
17. Henquin, J.C. (2009) Regulation of insulin secretion: a matter of 34. Yang, X. et al. (2015) Syntaxin opening by the MUN domain
phase control and amplitude modulation. Diabetologia 52, underlies the function of Munc13 in synaptic-vesicle priming.
739–751 Nat. Struct. Mol. Biol. 22, 547–554

170 Trends in Cell Biology, March 2017, Vol. 27, No. 3


35. Ferdaoussi, M. et al. (2015) Isocitrate-to-SENP1 signaling ampli- 58. Bossis, G. and Melchior, F. (2006) Regulation of SUMOylation by
fies insulin secretion and rescues dysfunctional b cells. J. Clin. reversible oxidation of SUMO conjugating enzymes. Mol. Cell 21,
Invest. 125, 3847–3860 349–357
36. Ivarsson, R. et al. (2005) Redox control of exocytosis: regulatory 59. Yan, S. et al. (2010) Redox regulation of the stability of the SUMO
role of NADPH, thioredoxin, and glutaredoxin. Diabetes 54, protease SENP3 via interactions with CHIP and Hsp90. EMBO J.
2132–2142 29, 3773–3786
37. Reinbothe, T.M. et al. (2009) Glutaredoxin-1 mediates NADPH- 60. Feligioni, M. and Nisticò, R. (2013) SUMO: a (oxidative) stressed
dependent stimulation of calcium-dependent insulin secretion. protein. Neuromolecular Med. 15, 707–719
Mol. Endocrinol. 23, 893–900 61. Tiedge, M. et al. (1997) Relation between antioxidant enzyme gene
38. Cameron, W.D. et al. (2016) Apollo-NADP+: a spectrally tunable expression and antioxidative defense status of insulin-producing
family of genetically encoded sensors for NADP+. Nat. Methods cells. Diabetes 46, 1733–1742
13, 352–358 62. Gooding, J.R. et al. (2015) Adenylosuccinate is an insulin secre-
39. Ronnebaum, S.M. et al. (2006) A pyruvate cycling pathway involv- tagogue derived from glucose-induced purine metabolism. Cell
ing cytosolic NADP-dependent isocitrate dehydrogenase regu- Rep. 13, 157–167
lates glucose-stimulated insulin secretion. J. Biol. Chem. 281, 63. Craig, T.J. et al. (2015) SUMOylation of syntaxin1A regulates
30593–30602 presynaptic endocytosis. Sci. Rep. 5, 17669
40. Guay, C. et al. (2007) A role for ATP-citrate lyase, malic enzyme, 64. Tang, L.T-H. et al. (2015) SUMOylation of synapsin Ia maintains
and pyruvate/citrate cycling in glucose-induced insulin secretion. synaptic vesicle availability and is reduced in an autism mutation.
J. Biol. Chem. 282, 35657–35665 Nat. Commun. 6, 7728
41. Spégel, P. et al. (2013) Time-resolved metabolomics analysis of b- 65. Girach, F. et al. (2013) RIM1/ SUMOylation is required for fast
cells implicates the pentose phosphate pathway in the control of synaptic vesicle exocytosis. Cell Rep. 5, 1294–1301
insulin release. Biochem. J. 450, 595–605
66. Plant, L.D. et al. (2011) SUMO modification of cell surface Kv2.1
42. Panten, U. and Rustenbeck, I. (2008) Fuel-induced amplification of potassium channels regulates the activity of rat hippocampal
insulin secretion in mouse pancreatic islets exposed to a high neurons. J. Gen. Physiol. 137, 441–454
sulfonylurea concentration: role of the NADPH/NADP+ ratio. Dia-
67. Williams, A.L. et al. (2011) Structural and functional analysis of
betologia 51, 101–109
tomosyn identifies domains important in exocytotic regulation.
43. Panten, U. et al. (2016) Acute metabolic amplification of insulin J. Biol. Chem. 286, 14542–14553
secretion in mouse islets: role of cytosolic acetyl-CoA. Metab. Clin.
68. Geerts, C.J. et al. (2014) Tomosyn interacts with the SUMO E3
Exp. 65, 1225–1229
ligase PIASg. PLoS ONE 9, e91697
44. Flamez, D. et al. (2002) Critical role for cataplerosis via citrate in
69. Wendt, A. et al. (2012) Synapsins I and II are not required for insulin
glucose-regulated insulin release. Diabetes 51, 2018–2024
secretion from mouse pancreatic b-cells. Endocrinology 153,
45. MacDonald, M.J. et al. (2007) Feasibility of pathways for transfer 2112–2119
of acyl groups from mitochondria to the cytosol to form short
70. Cheviet, S. et al. (2006) Tomosyn-1 is involved in a post-docking
chain acyl-CoAs in the pancreatic beta cell. J. Biol. Chem. 282,
event required for pancreatic beta-cell exocytosis. J. Cell. Sci. 119,
30596–30606
2912–2920
46. Joseph, J.W. et al. (2007) Normal flux through ATP-citrate lyase or
71. Gandasi, N.R. and Barg, S. (2014) Contact-induced clustering of
fatty acid synthase is not required for glucose-stimulated insulin
syntaxin and munc18 docks secretory granules at the exocytosis
secretion. J. Biol. Chem. 282, 31592–31600
site. Nat. Commun. 5, 3914
47. El Azzouny, M. et al. (2016) Knockdown of ATP citrate lyase in
72. Gandini, M.A. et al. (2011) Functional coupling of Rab3-interacting
pancreatic beta cells does not inhibit insulin secretion or glucose
molecule 1 (RIM1) and L-type Ca2+ channels in insulin release.
flux and implicates the acetoacetate pathway in insulin secretion.
J. Biol. Chem. 286, 15757–15765
Mol. Metab. 5, 980–987
73. Zhang, W. et al. (2006) Tomosyn is expressed in beta-cells and
48. Gheni, G. et al. (2014) Glutamate acts as a key signal linking
negatively regulates insulin exocytosis. Diabetes 55, 574–581
glucose metabolism to incretin/cAMP action to amplify insulin
secretion. Cell Rep. 9, 661–673 74. Dai, X.Q. et al. (2012) The voltage-dependent potassium channel
subunit Kv2.1 regulates insulin secretion from rodent and human
49. Liang, Y-C. et al. (2016) SUMO5, a novel poly-SUMO isoform,
islets independently of its electrical function. Diabetologia 55,
regulates PML nuclear bodies. Sci. Rep. 6, 26509
1709–1720
50. Yeh, E.T.H. (2009) SUMOylation and de-SUMOylation: wrestling
75. Fujita, Y. et al. (1998) Tomosyn: a syntaxin-1-binding protein that
with life's processes. J. Biol. Chem. 284, 8223–8227
forms a novel complex in the neurotransmitter release process.
51. Dai, X.Q. et al. (2011) SUMOylation regulates insulin exocytosis Neuron 20, 905–915
downstream of secretory granule docking in rodents and humans.
76. Yizhar, O. et al. (2004) Tomosyn inhibits priming of large dense-
Diabetes 60, 838–847
core vesicles in a calcium-dependent manner. Proc. Natl. Acad.
52. Hickey, C.M. et al. (2012) Function and regulation of SUMO Sci. U.S.A. 101, 2578–2583
proteases. Nat. Rev. Mol. Cell Biol. 13, 755–766
77. Zhao, Q. et al. (2014) GPS-SUMO: a tool for the prediction of
53. Cotto-Rios, X.M. et al. (2012) Deubiquitinases as a signaling target sumoylation sites and SUMO-interaction motifs. Nucleic Acids
of oxidative stress. Cell Rep. 2, 1475–1484 Res. 42, W325–W330
54. Xu, Z. et al. (2008) Molecular basis of the redox regulation of 78. Bhatnagar, S. et al. (2014) Phosphorylation and degradation of
SUMO proteases: a protective mechanism of intermolecular disul- tomosyn-2 de-represses insulin secretion. J. Biol. Chem. 289,
fide linkage against irreversible sulfhydryl oxidation. FASEB J. 22, 25276–25286
127–137
79. Dai, X.Q. et al. (2014) SUMO1 enhances cAMP-dependent exo-
55. Kunz, K. et al. (2016) SUMO signaling by hypoxic inactivation of cytosis and glucagon secretion from pancreatic /-cells. J. Physiol.
SUMO-specific isopeptidases. Cell Rep. 16, 3075–3086 (Lond.) 592, 3715–3726
56. Bailey, D. and O’Hare, P. (2004) Characterization of the localization 80. Irles, E. et al. (2015) Enhanced glucose-induced intracellular sig-
and proteolytic activity of the SUMO-specific protease, SENP1. naling promotes insulin hypersecretion: pancreatic beta-cell func-
J. Biol. Chem. 279, 692–703 tional adaptations in a model of genetic obesity and prediabetes.
57. Kim, Y.H. et al. (2005) Desumoylation of homeodomain-interacting Mol. Cell Endocrinol. 404, 46–55
protein kinase 2 (HIPK2) through the cytoplasmic-nuclear shuttling 81. Perry, R.J. et al. (2016) Imeglimin lowers glucose primarily by
of the SUMO-specific protease SENP1. FEBS Lett. 579, amplifying glucose-stimulated insulin secretion in high fat fed
6272–6278 rodents. Am. J. Physiol. Endocrinol. Metab. 311, E461–E470

Trends in Cell Biology, March 2017, Vol. 27, No. 3 171

You might also like