6
Breakthroughs and Views Molecular effects of proinsulin C-peptide q Jan Johansson, a, * Karin Ekberg, b Jawed Shafqat, a Mikael Henriksson, a Alexander Chibalin, b John Wahren, b and Hans Jornvall a a Department of Medical Biochemistry and Biophysics, Karolinska Institutet, SE-171 77 Stockholm, Sweden b Department of Surgical Sciences, Karolinska Hospital, SE-171 76 Stockholm, Sweden Received 10 June 2002 Abstract The proinsulin C-peptide has been held to be merely a by-product in insulin biosynthesis, but recent reports show that it elicits both molecular and physiological effects, suggesting that it is a hormonally active peptide. Specific binding of C-peptide to the plasma membranes of intact cells and to detergent-solubilised cells has been shown, indicating the existence of a cell surface receptor for C-peptide. C-peptide elicits a number of cellular responses, including Ca 2þ influx, activation of mitogen-activated protein (MAP) kinases, of Na þ ,K þ -ATPase, and of endothelial NO synthase. The pentapeptide EGSLQ, corresponding to the C-terminal five residues of human C-peptide, mimics several of the effects of the full-length peptide. The pentapeptide displaces cell membrane- bound C-peptide, elicits transient increase in intracellular Ca 2þ concentration and stimulates MAP kinase signalling pathways and Na þ ,K þ -ATPase. The Glu residue of the pentapeptide is essential for displacement of the full-length C-peptide, and free Glu can partly displace bound C-peptide, suggesting that charge interactions are important for receptor binding. Many C-peptide effects, such as phosphorylation of MAP-kinases ERK 1 and 2, stimulation of Na þ ,K þ -ATPase and increases in intracellular calcium concentrations are inhibited by pertussis toxin, supporting interaction of C-peptide with a G-protein-coupled receptor. However, all C-peptide effects cannot be explained in this manner, and it is possible that additional interactions are involved. Combined, the available observations show that C-peptide is biologically active and suggest a molecular model for its physiological effects. Ó 2002 Elsevier Science (USA). All rights reserved. The proinsulin C-peptide has long been considered merely a by-product of insulin biosynthesis. Thus, it is of structural importance in the folding of proinsulin by providing a spacer sequence that can be removed once folding is completed [1,2]. Early on, C-peptide was ex- amined for possible metabolic effects, but none was then established, see [2,3]. The observation that C-peptide varies considerably between species with regard to its amino acid sequence and chain-length has been taken as an additional indication for lack of C-peptide biological activity. As a consequence, C-peptide has generally been held to lack an active, hormone-like role. However, new findings in diabetic patients and animal models of dia- betes, correlated with defined molecular effects, have changed this view [4–13]. Here, we review the current understanding of the molecular mechanisms involved in C-peptide action and the structural properties of C-peptide in relation to cell binding and activity. We also briefly summarise the in vivo effects of C-peptide. C-peptide primary and secondary structures Human C-peptide has 31 residues; it lacks basic residues and contains five acidic residues (Fig. 1). In water, C-peptide is devoid of detectable stable sec- ondary structure, but the N-terminal 11 residues form an a-helical structure in 95% aqueous trifluoroethanol [14]. The region of positions 13–25, containing five Gly residues, is in parts lacking in some mammals, and has not been reported to have a stable secondary structure. The C-terminal five residue segment (encompassing positions 27–31), which contains two highly conserved residues, has been found to possess biological activity of its own, and is also unstructured in trifluoroethanol [14]. Biochemical and Biophysical Research Communications 295 (2002) 1035–1040 www.academicpress.com BBRC q Abbreviations: CP, proinsulin C-peptide; eNOS, endothelial NO synthase; MAP, mitogen-activated protein; NMDA, N-methyl-D- aspartate; NO, nitric oxide; PI3-K, phosphoinositide 3-kinase. * Corresponding author. Fax: +46-8-337462. E-mail address: [email protected] (J. Johansson). 0006-291X/02/$ - see front matter Ó 2002 Elsevier Science (USA). All rights reserved. PII:S0006-291X(02)00721-0

Molecular effects of proinsulin C-peptide

Embed Size (px)

Citation preview

Page 1: Molecular effects of proinsulin C-peptide

Breakthroughs and Views

Molecular effects of proinsulin C-peptideq

Jan Johansson,a,* Karin Ekberg,b Jawed Shafqat,a Mikael Henriksson,a

Alexander Chibalin,b John Wahren,b and Hans J€oornvalla

a Department of Medical Biochemistry and Biophysics, Karolinska Institutet, SE-171 77 Stockholm, Swedenb Department of Surgical Sciences, Karolinska Hospital, SE-171 76 Stockholm, Sweden

Received 10 June 2002

Abstract

The proinsulin C-peptide has been held to be merely a by-product in insulin biosynthesis, but recent reports show that it elicits

both molecular and physiological effects, suggesting that it is a hormonally active peptide. Specific binding of C-peptide to the

plasma membranes of intact cells and to detergent-solubilised cells has been shown, indicating the existence of a cell surface receptor

for C-peptide. C-peptide elicits a number of cellular responses, including Ca2þ influx, activation of mitogen-activated protein (MAP)

kinases, of Naþ,Kþ-ATPase, and of endothelial NO synthase. The pentapeptide EGSLQ, corresponding to the C-terminal five

residues of human C-peptide, mimics several of the effects of the full-length peptide. The pentapeptide displaces cell membrane-

bound C-peptide, elicits transient increase in intracellular Ca2þ concentration and stimulates MAP kinase signalling pathways and

Naþ,Kþ-ATPase. The Glu residue of the pentapeptide is essential for displacement of the full-length C-peptide, and free Glu can

partly displace bound C-peptide, suggesting that charge interactions are important for receptor binding. Many C-peptide effects,

such as phosphorylation of MAP-kinases ERK 1 and 2, stimulation of Naþ,Kþ-ATPase and increases in intracellular calcium

concentrations are inhibited by pertussis toxin, supporting interaction of C-peptide with a G-protein-coupled receptor. However, all

C-peptide effects cannot be explained in this manner, and it is possible that additional interactions are involved. Combined, the

available observations show that C-peptide is biologically active and suggest a molecular model for its physiological effects. � 2002Elsevier Science (USA). All rights reserved.

The proinsulin C-peptide has long been consideredmerely a by-product of insulin biosynthesis. Thus, it isof structural importance in the folding of proinsulin byproviding a spacer sequence that can be removed oncefolding is completed [1,2]. Early on, C-peptide was ex-amined for possible metabolic effects, but none was thenestablished, see [2,3]. The observation that C-peptidevaries considerably between species with regard to itsamino acid sequence and chain-length has been taken asan additional indication for lack of C-peptide biologicalactivity. As a consequence, C-peptide has generally beenheld to lack an active, hormone-like role. However, newfindings in diabetic patients and animal models of dia-betes, correlated with defined molecular effects, havechanged this view [4–13]. Here, we review the current

understanding of the molecular mechanisms involved inC-peptide action and the structural properties ofC-peptide in relation to cell binding and activity. Wealso briefly summarise the in vivo effects of C-peptide.

C-peptide primary and secondary structures

Human C-peptide has 31 residues; it lacks basicresidues and contains five acidic residues (Fig. 1). Inwater, C-peptide is devoid of detectable stable sec-ondary structure, but the N-terminal 11 residues forman a-helical structure in 95% aqueous trifluoroethanol[14]. The region of positions 13–25, containing five Glyresidues, is in parts lacking in some mammals, and hasnot been reported to have a stable secondary structure.The C-terminal five residue segment (encompassingpositions 27–31), which contains two highly conservedresidues, has been found to possess biological activityof its own, and is also unstructured in trifluoroethanol[14].

Biochemical and Biophysical Research Communications 295 (2002) 1035–1040

www.academicpress.com

BBRC

qAbbreviations: CP, proinsulin C-peptide; eNOS, endothelial NO

synthase; MAP, mitogen-activated protein; NMDA, N-methyl-D-

aspartate; NO, nitric oxide; PI3-K, phosphoinositide 3-kinase.* Corresponding author. Fax: +46-8-337462.

E-mail address: [email protected] (J. Johansson).

0006-291X/02/$ - see front matter � 2002 Elsevier Science (USA). All rights reserved.PII: S0006 -291X(02 )00721 -0

Page 2: Molecular effects of proinsulin C-peptide

C-peptide binding to cells

Binding of rat C-peptide 1 to cultured rat islet tumorcells, predominantly composed of insulin-secretingb-cells, was demonstrated using a radioligand bindingtechnique [15]. In contrast, human C-peptide was re-ported not to bind to crude cell membranes preparedfrom skeletal muscle, using the same technique [16].Proinsulin binding to cells has been reported, and humanC-peptide was able to compete with proinsulin, but notwith insulin, for binding to IM-9 lymphoblasts [17–19].Moreover, human C-peptide has been recently shown tobind to cell membranes from human renal tubular cells,skin fibroblasts, and saphenous vein endothelial cellsusing fluorescence correlation spectroscopy [20–22].Saturation of the C-peptide binding to the cell surfacewas then reported at low nanomolar concentrations.Addition of an excess of unlabelled C-peptide displacesthe bound C-peptide, but all-D-amino acid C-pep-tide, scrambled C-peptide, insulin, proinsulin, IGF-I, orIGF-II do not compete for binding, demonstrating thespecificity and the chiral nature of C-peptide binding tocell membranes.The detergent CHAPS can release C-peptide-binding

macromolecules from cells, which suggests that bindingcomponents can be purified and characterised [21].However, detection of binding required strict control ofconditions and short handling times, which complicatedinterpretations and further purification attempts.

Cellular effects of C-peptide

Effects on intracellular Ca2þ levels. C-peptide in-creases the intracellular concentration of calcium [23–25]. This is also true for the C-terminal pentapeptide ofC-peptide [25]. Addition of the calcium chelator EGTAto the medium abolishes the effects of C-peptide on in-tracellular calcium levels, suggesting that the increase ismediated by influx of extracellular calcium [23]. Like-wise, addition of pertussis toxin prevents the effects onintracellular calcium levels of both full-length C-peptideand of the C-terminal pentapeptide [25], which suggest

that G-protein-dependent effects are upstream of theinflux of calcium. It appears worthwhile to further ex-plore the effects of specific inhibitors on the C-peptide-elicited calcium increase to find out what signallingpathways are involved.Effects on mitogen-activated protein (MAP) kinase.

C-peptide has been found to induce phosphorylation ofthe MAP-kinases ERK 1 and 2 of a mouse embryonicfibroblast cell line (Swiss 3T3), while reverse-sequenceC-peptide and all-D-amino acid C-peptide did not stim-ulateMAP-kinase phosphorylation [26].Moreover, ERK1 and 2 phosphorylations were also observed in humanrenal tubular cells upon incubation with homologousC-peptide [27]. Preincubation of the cells with pertussistoxin or with PD98059, aMAP-kinase kinase 1 inhibitor,abolished the stimulatory effect on ERK phosphoryla-tion. Combined, these data suggest that C-peptide acti-vates a MAP-kinase-dependent signalling pathway.Effects on Naþ,Kþ-ATPase activity. A stimulatory ef-

fect of C-peptide on Naþ,Kþ-ATPase activity has beendemonstrated in both in vitro and in vivo studies. Thus,incubation of rat renal tubular segments with homolo-gous C-peptide in the concentration range 10–100 nMelicited a rise in Naþ,Kþ-ATPase activity as measured byhydrolysis of [32P]ATP [23,28]. The increase in activitywas concentration-dependent and could be blocked bypretreatment of the tubular segments with pertussis toxinor addition of by FK 506, a specific inhibitor of the Ca2þ-calmodulin-dependent protein phosphatase 2B. The lat-ter enzyme has an important function in rat renal tubularcells by converting the phosphorylated, low-activityform of Naþ,Kþ-ATPase to its dephosphorylated, activeform. In addition, incubation of the tubular segments ina Ca2þ free medium was found to completely abolish theC-peptide signal [23]. Human renal tubular cells in pri-mary culture have also been studied with regard toNaþ,Kþ-ATPase activity after exposure to humanC-peptide [27]. Oubain-sensitive uptake of 86Rbþ wasthen used as a measure of Naþ,Kþ-ATPase activity. Aconcentration-dependent stimulation was seen in therange 1–10 nM of C-peptide, and the effect was abolishedby pretreatment with pertussis toxin.Red blood cells frompatientswith type 1diabetes show

reduced Naþ,Kþ-ATPase activity [29] and have impaireddeformability. The latter phenomenon is reversible byexposure of the red blood cells to C-peptide [30]. TheC-peptide effect appears to be mediated via Naþ,Kþ-ATPase, because pretreatment of the cells with ouabain, aspecific inhibitor of Naþ,Kþ-ATPase, abolished the effectof C-peptide on the red cell deformability. Rat C-peptide2 given in replacement doses to diabetic BB/Wor rats, andhuman C-peptide administered in supra-physiologicaldoses to streptozotocin diabetic rats, have been found tostimulate Naþ,Kþ-ATPase activity in the sciatic nerveand to partially correct the diabetes-induced reduction inNaþ,Kþ-ATPase activity [6,8].

Fig. 1. Proinsulin C-peptide amino acid sequence. The upper line

shows the amino acid sequence of human C-peptide and the bars

correspond to the number of alternative amino acids found at each

position in 20 mammalian C-peptide amino acid sequences (J€oornvall et

al., unpublished).

1036 J. Johansson et al. / Biochemical and Biophysical Research Communications 295 (2002) 1035–1040

Page 3: Molecular effects of proinsulin C-peptide

Effects on endothelial nitric oxide synthase (eNOS)activity. In a study using bovine aortic cells and a re-porter cell assay it could be demonstrated that exposureof cells to C-peptide resulted in stimulation of eNOSactivity and release of NO in a concentration- and time-dependent manner [24]. Addition of an eNOS blockingsubstance (L-NAME) or a calcium-binding agent in-hibited the C-peptide effect, suggesting that the effect ofC-peptide on eNOS is mediated via an increase in in-tracellular Ca2þ [24]. In keeping with these observations,C-peptide is reported to elicit a dilation of skeletalmuscle arterioles via a NO-mediated mechanism [31].Moreover, leukocyte–endothelium interactions inducedin rat mesenteric venules by both thrombin andL-NAME are diminished by C-peptide [32]. C-peptideadministration is associated with an increase in eNOSmRNA levels and a marked increase in basal NO releasefrom aortic endothelial cells. C-peptide may thus act asan inhibitor of leukocyte/endothelium interactions byinhibiting endothelial cell adhesion molecules viamaintenance of NO release from endothelial cells [32].In support of this hypothesis, C-peptide exerts cardio-protective effects against leukocyte-mediated reperfusioninjury in the isolated rat heart, most likely due to re-duced leukocyte adherence to vascular endotheliumsecondary to stimulation of NO production by the en-dothelium [33].Replacement administration of C-peptide to patients

with type 1 diabetes results in a concentration-depen-dent rise in muscle blood flow [34]. The C-peptide in-duced rise in muscle blood flow is seen both at rest andduring physical exercise and can be blocked by additionof L-NAME [7,35]. C-peptide administration is alsoreported to result in increases in blood flow to the kid-neys, the skin, and the nerves in the diabetic state[5,9,36].

Activity of C-peptide fragments and regions

Some of the conserved C-peptide residues (Fig. 1) arethe consequence of the requirements for folding andprocessing of proinsulin. The biological effects and cellbinding of C-peptide are likely to be mediated through aspecific region(s) of the peptide, and its sequence con-servation may partly be a reflection of this. Attemptshave been made to locate the activities and properties ofC-peptide with respect to its sequence. The rat C-termi-nal pentapeptide (EVARQ) was found to possess 100%of the Naþ,Kþ-ATPase stimulatory activity of the intactmolecule, while the remaining segment, des-(27–31)C-peptide, failed to elicit detectable activity [28]. Fur-thermore, human C-peptide bound to renal tubular cellmembranes was effectively displaced by a 1000-foldmolar excess of the human pentapeptide EGSLQ [20],indicating that the C-terminal fragment is involved in a

specific binding process. The pentapeptide EGSLQ, butnot des-(27–31)C-peptide, elicits an increase in intracel-lular calcium [25] and causes phosphorylation of MAP-kinases in human renal tubular cells [27]. Thus, fourdifferent approaches demonstrate a molecular effect ofthe intact C-peptide and a similar effect of its C-terminalpentapeptide. This supports the view that the C-terminalsegment constitutes an active site of C-peptide.An internal fragment corresponding to positions

11–19 of the C-peptide also exhibits some stimulatoryeffects on Naþ,Kþ-ATPase activity [28], and it is re-ported that a C-peptide mid-portion is important fornormalisation of glucose-induced vascular dysfunctionin a rat model [6]. The mid-fragment sequence is highlyvariable between species, and a peptide comprisingamino acid residues 11–19 of C-peptide does not dis-place cell membrane-bound human C-peptide [22]. Thissuggests that the mechanisms of the mid-portion seg-ment are different from those of the C-terminal segment.The possible importance of Glu per se and of Glu-

terminal peptides has been considered [22,37]. C-peptideand other Glu-terminal peptides have been shown toselectively suppress N-methyl-D-aspartate (NMDA)glutamate receptor-mediated secretion of gonadotropin-releasing hormone in rat hypothalamic explants [37].LL-Glutamate stimulates insulin secretion in the ratpancreas via stimulation of NMDA receptors. Thus,C-peptide could be involved in an auto-feedbackmechanism of insulin secretion.The current view is that the C-terminal pentapeptide

segment of C-peptide is important for specific binding tomolecular targets and for evoking biological effects, andthat a Glu residue is involved in these processes.

Molecular mechanisms of C-peptide action

As described above several actions of C-peptide areapparently mediated via G-protein-coupled pathways,but it is possible that further interactions are involved(Fig. 2).(i) Pertussis toxin treatment inhibits C-peptide stim-

ulation of Naþ,Kþ-ATPase activity [23,27], calcium in-flux [25], and activation of MAP kinases [26,27],indicative of receptors coupled to adenylate cyclase,specifically transmembrane G-protein-coupled receptorswith the a-subunit of the Gi=G0 subtype. Interestingly,pertussis toxin not only blocks C-peptide effects but alsointerferes with C-peptide binding to cell membranes [20].Moreover, activation of protein kinase C and phos-phoinositide 3-kinase (PI3-K) is apparently involved inC-peptide induced phosphorylation of MAP kinases[26].(ii) Interactions between C-peptide and receptors with

catalytic activity are indicated by results showing that C-peptide attenuates protein tyrosine phosphatase activity

J. Johansson et al. / Biochemical and Biophysical Research Communications 295 (2002) 1035–1040 1037

Page 4: Molecular effects of proinsulin C-peptide

[38]. Protein tyrosine phosphatases inactivate the insulinsignalling pathway by dephosphorylation of the insulinreceptor, insulin receptor substrates, and MAP kinases.Hence, C-peptide and insulin might have a synergisticeffect on the insulin signalling pathway at the level of theinsulin receptor. This is further corroborated by therecent finding that C-peptide at physiological concen-trations mimics insulin effects in myoblasts; it activatesinsulin receptor tyrosine kinase, insulin receptor sub-strate-1 tyrosine phosphorylation, PI3-K activity, andMAP kinase phosphorylation [39]. If C-peptide is addedin the presence of high insulin concentrations, no furthereffects are observed, indicating that C-peptide and in-sulin may use the same signalling pathway [39]. Theseauthors suggested that low C-peptide levels enhanceinsulin effects, while at supra-physiological concentra-tions C-peptide blunts insulin effects. However, C-pep-tide, unlike insulin, does not activate Akt (protein kinaseB), suggesting that C-peptide also works via mecha-nisms distinct from those of insulin. C-peptide-inducedstimulation of glycogen synthesis in the myoblasts wasblocked by Wortmannin, an inhibitor of PI3-K activity,but not by pertussis toxin [39]. In contrast to thesefindings, Zierath et al. [16] found that C-peptide stimu-lates glucose transport in human muscle strips, and thatthese effects were not mediated via the insulin receptoror tyrosine kinase activation.(iii) C-peptide interactions with ligand-gated ion

channels coupled to glutamate receptors have beenconsidered because of the capacity of free glutamic acidto partly displace C-peptide from cell membranes, andby the importance of N-terminal Glu in intact C-peptideand its C-terminal pentapeptide in binding and biolog-ical activity [22,37]. The observed effects of C-peptide on

NMDA receptors [37] would also fit with an involve-ment of a particular class of glutamate receptors, but itremains possible that C-peptide Glu residues are crucialin any type of receptor interaction. However, the first 26amino acid residues of rat C-peptide, including anN-terminal Glu, lack the stimulatory activity of intactC-peptide on Naþ,Kþ-ATPase activity in rat renaltubular segments [28], and human C-peptide (1–26) hasa poor capacity to displace C-peptide bound to humancell membranes [22].It has also been suggested that the effects of C-peptide

may be mediated by direct membranotropic mechanismsinstead of by classical receptor–ligand interactions [6].However, the hallmarks of traditional pore-formingpeptides are lacking in C-peptide; it shows no tendencyto self-associate and it is overall hydrophilic and nega-tively charged [10]. In line with this, C-peptide was laterfound not to associate in a stable manner with lipidmembranes or micelles [14].The above observations are of interest regarding

possible diversity of C-peptide action, but it should bepointed out that C-peptide concentrations differ widelybetween the experimental systems used. Saturation ofbinding to cells is apparently reached already within thephysiological concentration range [20]. As a conse-quence, receptor-mediated physiological effects arelikely to take place at or below the physiologicalC-peptide concentration (0.5–1:5� 10�9 M), whereassupra-physiological concentrations might induce non-specific effects.In summary, the current view on C-peptide action

(Fig. 2) is that G-protein-coupled receptor(s) areinvolved, since addition of pertussis toxin reducesC-peptide binding to cells and blunts C-peptide effectson signal mediators, as well as enzyme activity. Exper-imental data in support of C-peptide binding to otherreceptor types indicate that additional signal pathway(s)may also be affected, which may converge with aG-protein-coupled receptor pathway at different levels.Notably, relaxin, a member of the insulin-like peptidehormone family, was recently found to work viaG-protein-coupled receptors and adenylate cyclaseactivation, but also possible to elicit effects via a tyrosinekinase pathway [40,41].

In vivo effects of C-peptide replacement therapy

Replacement of C-peptide in rats with streptozotocin-induced diabetes is accompanied by correction ofglomerular hyperfiltration, diminished levels of microal-buminuria, and regression of glomerular hypertrophy[11,12]. Likewise, when C-peptide is administered inreplacement doses to patientswith type 1 diabetes, there issignificant reduction of both glomerular hyperfiltrationand urinary albumin excretion [4,5,13].

Fig. 2. Signal pathways of interest in relation to reported C-peptide

effects. The central (blue) pathway represents C-peptide effects medi-

ated by G-protein-coupled receptor(s) (GPCR), i.e., are blocked by

pertussis toxin [23,25–27]. The right (yellow) pathway represents effects

which appear to be mediated via the insulin receptor and are not in-

hibited by pertussis toxin [39]. The left (orange) pathway represents the

finding that C-peptide can bind to an ion-channel-coupled glutamate

receptor [37], which possibly could result in Ca2þ influx. See text for

further details. PKC, protein kinase C; PI3-K, phosphoinositide

3-kinase; MAPK, mitogen-activated protein kinase; eNOS, endothelial

NO synthase.

1038 J. Johansson et al. / Biochemical and Biophysical Research Communications 295 (2002) 1035–1040

Page 5: Molecular effects of proinsulin C-peptide

Effects of C-peptide replacement therapy on functionaland structural changes in peripheral nerves have beenstudied in rats that spontaneously develop diabetes.C-peptide administration for two months was found toprevent a major proportion of the nerve conductionvelocity defect and the paranodal swelling that otherwiseoccurred [8]. In patients with autonomic nerve dysfunc-tion, increased heart rate variability during deep breath-ing has been seen following C-peptide administration[13,42]. In addition, preliminary evidence from a studyinvolving C-peptide replacement therapy in type 1diabetes patients without overt symptoms of neuropathyindicate that three months of treatment results in signifi-cant improvement of sensory nerve conduction velocityand vibration thresholds (Ekberg et al., unpublishedobservation). C-peptide replacement therapy in rats isknown to be accompanied by significant improvement innerve Naþ,Kþ-ATPase activity [6,8] and in nerve bloodflow [9]. The relevance of the observed C-peptide effects issupported by the fact that both deficient NO formationand reduced levels of Naþ,Kþ-ATPase are factors ofpathogenetic importance for diabetic neuropathy [43].The present data suggest that C-peptide replacementtherapy together with insulin therapy in type 1 diabetespatients may be beneficial in preventing or retarding thedevelopment of long-term complications.

Acknowledgment

This work was supported by the Juvenile Diabetes Foundation

(project JDFI 1-1999-647).

References

[1] R. Mackin, Proinsulin: recent observations and controversies,

Cell. Mol. Life Sci. 54 (1998) 696–702.

[2] A.E. Kitabchi, Proinsulin and C-peptide: a review, Metab. Clin.

Exp. 26 (1977) 547–587.

[3] J. Wahren, K. Ekberg, J. Johansson, M. Henriksson, A. Prama-

nik, B.-L. Johansson, R. Rigler, H. J€oornvall, Role of C-peptide in

human physiology, Am. J. Physiol. 278 (2000) E759–E768.

[4] B. Johansson, S. Sj€ooberg, J. Wahren, The influence of human

C-peptide on renal function and glucose utilization in type 1

(insulin-dependent) diabetic patients, Diabetologia 35 (1992) 121–

128.

[5] B. Johansson, A. Kernell, S. Sj€ooberg, J. Wahren, Influence of

combined C-peptide and insulin administration on renal function

and metabolic control in diabetes type 1, J. Clin. Endocrinol.

Metab. 77 (1993) 976–981.

[6] Y. Ido, A. Vindigni, K. Chang, L. Stramm, R. Chance, W. Heath,

R. DiMarchi, E. Di Cera, J. Williamson, Prevention of vascular

and neural dysfunction in diabetic rats by C-peptide, Science 277

(1997) 563–566.

[7] B. Johansson, B. Linde, J. Wahren, Effects of C-peptide on blood

flow, capillary diffusion capacity and glucose utilization in the

exercising forearm of type 1 (insulin-dependent) diabetic patients,

Diabetologia 35 (1992) 1151–1158.

[8] A. Sima, W. Zhang, K. Sugimoto, D. Henry, Z. Li, J. Wahren, G.

Grunberger, C-peptide prevents and improves chronic type 1

diabetic neuropathy in the BB/Wor-rat, Diabetologia 44 (2001)

889–897.

[9] M. Cotter, N. Cameron, C-peptide effects on nerve conduction

and blood flow in streptozotocin-induced diabetic rats: modula-

tion by nitric oxide synthase inhibition, Diabetes 50 (2001) A184.

[10] D.F. Steiner, A.H. Rubenstein, Proinsulin C-peptide-biological

activity?, Science 277 (1997) 531–532.

[11] M. Sj€ooquist, W. Huang, B.L. Johansson, Effects of C-peptide on

renal function at the early stage of experimental diabetes, Kidney

Int. 54 (1998) 758–764.

[12] B. Samneg�aard, S. Jacobson, G. Jaremko, B.-L. Johansson, M.

Sj€ooquist, Effects of C-peptide on renal function, albuminuria andglomerular volume in diabetic rats, Kidney Int. 60 (2001) 1258–

1265.

[13] B.L. Johansson, K. Borg, E. Fernqvist-Forbes, A. Kernell, T.

Odergren, J. Wahren, Beneficial effects of C-peptide on incipient

nephropathy and neuropathy in patients with Type 1 diabetes

mellitus, Diabet. Med. 17 (2000) 181–189.

[14] M. Henriksson, J. Shafqat, E. Liepinsh, M. Tally, J. Wahren, H.

J€oornvall, J. Johansson, Unordered structure of proinsulinC-peptide in aqueous solution and in the presence of lipid vesicles,

Cell. Mol. Life Sci. 57 (2000) 337–342.

[15] P. Flatt, S. Swanston-Flatt, S. Hampton, C. Bailey, V. Marks,

Specific binding of the C-peptide of proinsulin to cultured B-cells

from a transplantable rat islet cell tumor, Biosci. Rep. 6 (1986)

193–199.

[16] J. Zierath, A. Handberg, M. Tally, H. Wallberg-Henriksson,

C-peptide stimulates glucose transport in isolated human skeletal

muscle independent of insulin receptor and tyrosine kinase

activation, Diabetologia 39 (1996) 306–313.

[17] P. Jehle, M. Lutz, R. Fussgaenger, High affinity binding sites for

proinsulin in human IM-9 lymphoblasts, Diabetologia 39 (1996)

421–432.

[18] P. Jehle, R. Fussgaenger, N. Angelus, R. Jungwirth, B. Saile, M.

Lutz, Proinsulin stimulates growthof small intestinal crypt-like cells

acting via specific receptors, Am. J. Physiol. 276 (1999) E262–E268.

[19] M. Faehling, R. Fussgaenger, P. Jehle, High affinity binding sites

for proinsulin on human umbilical vein endothelial cells (HU-

VEC), Diabetologia 42 (1999) 259–260.

[20] R. Rigler, A. Pramanik, P. Jonasson, G. Kratz, O. Jansson, P.

Nygren, S. St�aahl, K. Ekberg, B. Johansson, M. Uhl�een, H. J€oornvall,

J. Wahren, Specific binding of proinsulin C-peptide to human cell

membranes, Proc. Natl. Acad. Sci. USA 96 (1999) 13318–13323.

[21] M. Henriksson, A. Pramanik, J. Shafqat, Z. Zhong, M. Tally, K.

Ekberg, J. Wahren, R. Rigler, J. Johansson, H. J€oornvall, Specific

binding of proinsulin C-peptide to intact and detergent-solubilized

human skin fibroblasts, Biochem. Biophys. Res. Commun. 280

(2001) 423–427.

[22] A. Pramanik, K. Ekberg, Z. Zhong, J. Shafqat, M. Henriksson, O.

Jansson, A. Tibell, M. Tally, J. Wahren, H. J€oornvall, R. Rigler, J.

Johansson, C-peptide binding to human cell membranes: impor-

tance of Glu27, Biochem. Biophys. Res. Commun. 284 (2001) 94–

98.

[23] Y. Ohtomo, A. Aperia, B. Sahlgren, B. Johansson, J. Wahren,

C-peptide stimulates rat renal tubular Naþ,Kþ-ATPase activity in

synergism with neuropeptide Y, Diabetologia 39 (1996) 199–205.

[24] T. Kunt, T. Forst, R. Lehmann, A. Pfuetzner, M. L€oobig, O.Harzer, M. Engelbach, J. Beyer, Human C-peptide increases

calcium influx into endothelial cells, Diabetes 47 (1998) A30.

[25] J. Shafqat, L. Juntti-Berggren, Z. Zhong, K. Ekberg, M. K€oohler,

P.-O. Berggren, J. Johansson, J. Wahren, H. J€oornvall, ProinsulinC-peptide and its analogues induce intracellular Ca2þ increases in

human renal tubular cells, Cell. Mol. Life Sci., 2002 (in press).

[26] T. Kitamura, K. Kimura, B.D. Jung, K. Makondo, S. Okamoto,

X. Canas, N. Sakane, T. Yoshida, M. Saito, Proinsulin C-peptide

rapidly stimulates mitogen-activated protein kinases in Swiss3T3

fibroblasts: requirement of protein kinase C, phosphoinositide3-

J. Johansson et al. / Biochemical and Biophysical Research Communications 295 (2002) 1035–1040 1039

Page 6: Molecular effects of proinsulin C-peptide

kinase and pertussis toxin-sensitive G-protein, Biochem. J. 355

(2001) 123–129.

[27] A. Chibalin, A. Dumitriescu, K. Ekberg, B.-L. Johansson, J.

Wahren, C-peptide stimulates Na,K-pump through MAP kinase-

dependent pathway in the human renal tubular cells, Acta

Physiol., 2002 (in press).

[28] Y. Ohtomo, T. Bergman, B. Johansson, H. J€oornvall, J. Wahren,

Differential effects of proinsulin C-peptide fragments on Naþ,Kþ-

ATPase activity of renal tubule segments, Diabetologia 41 (1998)

287–291.

[29] D. De La Tour, D. Raccah, M. Jannot, T. Coste, C. Rougerie, P.

Vague, Erythrocyte Na/K ATPase activity and diabetes: relation-

ship with C-peptide level, Diabetologia 41 (1998) 1080–1084.

[30] T. Kunt, S. Schneider, A. Pfutzner, K. Goitum, M. Engelbach, B.

Schauf, J. Beyer, T. Forst, The effect of human proinsulin C-

peptide on erythrocyte deformability in patients with type I

diabetes mellitus, Diabetologia 42 (1999) 465–471.

[31] M.E. Jensen, E.J. Messina, C-peptide induces a concentration-

dependent dilation of skeletal muscle arterioles only in presence of

insulin, Am. J. Physiol. 276 (1999) H1223–H1228.

[32] R. Scalia, K.M. Coyle, B.J. Levine, G. Booth, A.M. Lefer,

C-peptide inhibits leukocyte–endothelium interaction in the

microcirculation during acute endothelial dysfunction, FASEB

J. 14 (2000) 2357–2364.

[33] L.H. Young, Y. Ikeda, R. Scalia, A.M. Lefer, C-peptide exerts

cardio protective effects in myocardial ischemia-reperfusion, Am.

J. Physiol. Heart Circ. Physiol. 279 (2001) H1453–H1459.

[34] K. Ekberg, B-L. Johansson, J. Wahren, Stimulation of blood flow

by C-peptide in patients with type 1 diabetes, Diabetologia 44

(Suppl. 1) (2001) A323.

[35] B-L. Johansson, J. Pernow, J. Wahren, Muscle vasodilatation by

C-peptide is NO-mediated, Diabetologia 42 (Suppl. 1) (1999)

A324.

[36] T. Forst, T. Kunt, T. Pohlmann, K. Goitom, M. Engelbach, J.

Beyer, A. Pf€uutzner, Biological activity of C-peptide on the skinmicrocirculation in patients with insulin-dependent diabetes

mellitus, J. Clin. Invest. 101 (1998) 2036–2041.

[37] J.P. Bourguignon, M.L. Alvarez Gonzalez, A. Gerard, P. Fran-

chimont, Gonadotropin releasing hormone inhibitory autofeed-

back by subproducts antagonist at N-methyl-D-aspartate

receptors: a model of autocrine regulation of peptide secretion,

Endocrinology 134 (1994) 1589–1592.

[38] Z.G. Li, X. Qiang, A.A. Sima, G. Grunberger, C-peptide

attenuates protein tyrosine phosphatase activity and enhances

glycogen synthesis in L6 myoblasts, Biochem. Biophys. Res.

Commun. 280 (2001) 615–619.

[39] G. Grunberger, X. Qiang, Z. Li, S.T. Mathews, D. Sbrissa, A.

Shiseva, A.A.F. Sima, Molecular basis for the insulinomimetic

effects of C-peptide, Diabetologia 44 (2001) 1247–1257.

[40] S.Y. Hsu, K. Nakabayashi, S. Nishi, J. Kumagai, M. Kudo, O.D.

Sherwood, A.J.W. Hsueh, Activation of orphan receptors by the

hormone relaxin, Science 295 (2002) 671–674.

[41] R. Ivell, This hormone has been relaxin’ too long, Science 295

(2002) 637–638.

[42] B. Johansson, K. Borg, E. Fernqvist-Forbes, T. Odergren, S.

Remahl, J. Wahren, C-peptide improves autonomic nerve func-

tion in IDDM patients, Diabetologia 39 (1996) 687–695.

[43] N.E. Cameron, S.E.M. Eaton, M.A. Cotter, S. Tesfaye, Vascular

factors and metabolic interactions in the pathogenesis of diabetic

neuropathy, Diabetologia 44 (2001) 1973–1988.

1040 J. Johansson et al. / Biochemical and Biophysical Research Communications 295 (2002) 1035–1040