Skip to content
Publicly Available Published by De Gruyter April 12, 2016

The insulin autoimmune syndrome (IAS) as a cause of hypoglycaemia: an update on the pathophysiology, biochemical investigations and diagnosis

  • Adel A.A. Ismail EMAIL logo

Abstract

Insulin autoimmune syndrome (IAS) is considered to be very rare in Caucasians. Understanding its pathophysiology is paramount in (a) appreciating its potential impact on analyses of pancreatic hormones and (b) explaining its highly variable clinical manifestations in non-diabetic, non-acutely ill patients with indeterminate hypoglycaemia. The underlying aetiology of IAS is the presence of variable affinity/avidity endogenous insulin antibodies in significant amounts. The two types of insulin antibodies namely antibodies which bind insulin and/or proinsulin(s) and receptor antibodies (insulin mimetic) will be discussed. Their biochemical and immunological roles in causing hypoglycaemia will be highlighted. Clinical manifestations of IAS can vary from mild and transient to spontaneous, severe and protracted hypoglycaemia necessitating in extreme cases plasmapheresis for glycaemic control. Antibodies of IAS can interfere in pancreatic immunoassay tests causing erroneous and potentially misleading results. Thorough testing for endogenous insulin antibodies must be considered in the investigations of non-diabetic, non-acutely ill patients with indeterminate and/or unexplained hypoglycaemia.

Introduction

Autoimmunity in general is beneficial; however, impairment of the immune system in recognising “self-antigens”, e.g. immunosenescence can trigger autoaggression which in some cases progresses to autoimmune disease [1], [2], [3]. Autoimmune diseases in some cases occur when the very low titre of endogenous antibodies rise causing transformation from benign autoimmunity to pathogenic autoimmunity. This progression is determined by both genetic influences and environmental triggers such as spontaneous genetic instability, acquired genetic mutations or antigenic modifications [3].

The very first case of insulin autoimmune syndrome (IAS) was described in the Tohoku Journal of Experimental Medicine in 1972 [4] by Yukimasa Hirata in a 47-year-old Japanese man with severe spontaneous hypoglycaemia. IAS is now the third most common cause of hypoglycaemia in Japan after insulinoma and extrapancreatic neoplasms. The first case in a Caucasian was independently reported in a 40-year-old Norwegian with hyperglycaemia and profound reactive hypoglycaemia after meals by Følling and Norman, and appeared in Diabetes in 1972 [5]. The very different clinical manifestations in these two very initial cases of IAS are noteworthy. Since then, some 70 cases of IAS in Caucasians with a wider range of clinical presentations have been fully described from almost all western European countries, North and South America and an abstract of one case in 2010 from the UK [6]. Newly found cases of IAS with already well-documented aetiologies and manifestations but lacking new or additional information may be considered unsuitable for further publication by most academic reviewers.

IAS is also known as Hirata’s disease or insulin autoimmune hypoglycaemia (IAH). It is essentially a rare autoimmune disorder caused by the presence of variable and significant increase in the titre and affinity/avidity of endogenous antibodies which bind insulin/proinsulin and/or insulin receptor (insulin mimetic). Antibodies of the IAS have a unique double-whammy impact namely a pathological role in causing hypoglycaemia as well as the potential to interfere in some immunoassays of pancreatic hormone measurements. IAS has been shown to clinically and biochemically mimic other causes of hypoglycaemia such as insulinoma, insulin administration (self-harm or malicious). Appreciating this dual clinical-analytical impact has made its consideration in the differential diagnosis of non-diabetic, non-critically ill patients with indeterminate cause of hypoglycaemia judicious and prudent.

In this short review, the pathophysiology of IAS will be briefly described and the heterogeneity of endogenous antibodies and their clinical and analytical repercussions in mimicking other causes of hypoglycaemia will be summarised. This would help clinical and laboratory practitioners to plan appropriate biochemical investigations, perform and correctly interpret relevant immunoassay results. Occasionally, a paraprotein in some patients may also exhibit an ability to bind insulin, thus causing hypoglycaemia by a mechanism similar to IAS [7], [8], [9].

Brief note on the pathophysiology of hypoglycaemia of IAS

IAS is a rare cause of hypoglycaemia with higher incidence in genetically predisposed individuals such as those with HLA-DR4 [10], [11], [12]. There are two major triggers, namely infection/reinfection and drugs particularly in patients with a history of other autoimmune disorders or autoimmune polyendocrine syndromes [10], [11], [12], [13], [14], [15]. A range of common viral infections/reinfections such as coxsackie B influenza, mumps and rubella are known to cause islet cell antibodies and diabetes type 1. However, these viruses have also been shown to increase the concentration of endogenous insulin antibodies in 1986 [16] and hepatitis C virus was added to the list in 2002 [17]. A higher concentration of endogenous antibodies in asymptomatic non-diabetic individuals is considered a risk factor for potential overt disease [16], [18].

Until 2003, drugs containing sulfur/sulfydryl groups (e.g. thiols with -SH) such as methimazole, captopril, D-penicillamine, hydralazine, glutathione, methionine, mercaptans, clopidogrel, aurothioglucose, imipenem, penicillin G and diltiazem [19], [20], [21] were deemed to have the potential to trigger IAS. An additional drug namely α-lipoic acid (ALA) which also contains sulfur (cyclic disulfide) was also reported to cause IAS in Japanese [19]; this was quickly followed by numerous reports in patients all of whom were also Japanese [21], [22], [23], [24], [25], [26]. The first case in a Caucasian (Italian) appeared in Sicily in 2011 [27]; however, within 22 months, six more cases were also identified by the same medical centre in Sicily [28]. Lipoic-acid is sold over-the-counter (OTC) and on the Internet as a universal antioxidant drug/free radical scavenger because of its unique solubility in both water and fat; it is also prescribed for conditions such as peripheral polyneuropathy [26]. It appears that the incidence of drug-induced IAS can occur in significant number in genetically predisposed Caucasians too. Clinical manifestations of IAS varied, irrespective of the underlying cause, from mild, intermittent and transient to spontaneous, protracted and very severe, necessitating in extreme cases plasmapheresis for glycaemic control [11], [29], [30], [31], [32], [33].

A suggested mechanism of action of endogenous antibodies which bind insulin and proinsulin(s) is that the initial pancreatic response to a rise in blood sugar would be “ineffective” in lowering blood glucose, because insulin/proinsulin(s) bind to these endogenous antibodies causing postprandial hyperglycaemia to persist for longer. This in turn results in prolonged pancreatic secretion of insulin and C-peptide in equimolar amounts as well as ~3% of proinsulin which has 1/10th of the insulin bioactivity, until the endogenous antibodies binding capacity is exceeded. Only then the unbound/free active insulin rises and reduces blood glucose, and at a concentration of ~4.5 mmol/L (i.e. the glycaemic threshold), pancreatic secretion decreases/ceases [31].

In patients with IAS, the t1/2 of insulin/proinsulin(s) increase manyfold, from minutes to potentially hour(s) [29], [31], [34], [35], [36], [37], but that of C-peptide which may not bind to these antibodies remains unaffected at ~25 min [31]. Insulin/proinsulin(s) bound to these endogenous antibodies in blood represent a “reservoir”; its subsequent dissociation from these antibodies would be dependent on the intrinsic dissociation rate constant (i.e. K−1) [38], [39]. A significant K−1 would continuously release free/unbound bioactive insulin/proinsulin(s) enough to cause hypoglycaemia of varying severity usually within 2–6 h or longer. Hyperinsulinaemia also inhibits counter-regulatory mechanisms of glucose homeostasis (suppresses glycogenolysis, gluconeogenesis and ketogenesis). The severity, duration and swing from hyper- to hypoglycaemia and the intensity and duration of the hypoglycaemic phase itself have been attributed to (a) the endogenous antibody characteristics, namely its intrinsic association and dissociation rate constants to insulin/proinsulin(s), (b) its titre/concentration and (c) its capacity, i.e. antibodies’ valency. More than one class of endogenous antibodies may also coexist in the same patient, e.g. high affinity/low capacity and high capacity/low affinity [31], [40]. The presence of such endogenous binding antibodies has distorted glucose metabolism [31], [35] with phases of hyper- and hypoglycaemia, which led some to refer to these swings as diabetes and hypoglycaemia [41], [42]. Three classes of immunoglobulin antibodies namely IgG, IgA and IgM are so far been identified with the most common being IgG [7], [19], [29], [31], [33], [43].

Another form of endogenous antibodies namely insulin receptor antibodies was the sole cause of hypoglycaemia in some cases. However, both insulin binding and insulin receptor antibodies have occurred in the same patient, sequentially or simultaneously [43], [44], [45], [46], [47], [48], [49]. A suggested mechanism is an idiotypic immune response (see reference [49] for more details).

Summary

The described biochemical and clinical manifestations of IAS were variable, occurring when endogenous antibodies capable of binding insulin, proinsulin(s) and/or insulin receptor antibodies were produced in significant amount. The presence of these endogenous antibodies is the hallmark IAS underpinning its aetiology. Their identification was essential in differential diagnosis in non-diabetic, non-acutely ill patients with indeterminate hypoglycaemia and in confirming the diagnosis of IAS.

Immunoassays of pancreatic hormones and potential interference from endogenous antibodies of IAS

Screening tests for hypoglycaemic drugs (sulfonylureas and biguanides), non-islet cell tumour-induced hypoglycaemia, e.g. testing for IGF(s), β-hydroxybutyrate, pituitary and adrenocortical functions will not be considered being outside the scope of this short review and are well-described elsewhere [50], [51] and in text books of endocrinology.

Current pancreatic immunoassays are almost exclusively commercial, optimised “with justification” to measure these moieties in serum essentially free from endogenous insulin and/or proinsulin antibodies. Proinsulin is proteolytically processed within the pancreas by several endopeptidases to generate and predominantly secrete insulin and C-peptide in equimolar amounts. Some 3% of intact proinsulin is also secreted together with four partially processed split proinsulin intermediates. These split conversions are biologically more potent than proinsulin itself, which has 1/10th of the bioactivity of insulin [52], [53], [54], present in higher concentration in serum than proinsulin [52], [53], [54] and similarly retain both insulin and C-peptide components (see reference [54] for more details).

The inbuilt “insulin component” in “proinsulin and the four proinsulin split intermediates” collectively referred to as proinsulin(s) in this review have the potential to cross-react (albeit at dissimilar rates), measured and may be “improperly” reported as “free” insulin in some immunoassays. The “C-peptide component” may similarly cross-react, measured and also may be “incorrectly” reported as “free” C-peptide by some immunoassays [54], [55], [56], [57]. The presence of insulin, proinsulin(s) binding and/or receptor antibodies has shown to increase the concentration of one or more of these moieties thus causing unpredictable inaccuracy in immunoassay results [53], [54], [55], [56], [57]. For this reason, testing for endogenous insulin antibodies is considered prudent or even necessary in non-diabetic, non-critically ill patients with indeterminate cause of hypoglycaemia. Tests for “functional” insulin/proinsulin binding antibodies are initially carried out (being relatively more common and simple to perform) and if negative, tests for receptor antibodies are then considered if all forms of IAS are to be excluded. Multiple blood samples during phases of hypoglycaemia as well as few hours after a meal or an extended OGTT (~5 h) were also useful tools in diagnosis. Pancreatic hormones which bind to endogenous antibodies progressively rise with time and remain elevated for long while others which do not bind declined with their normal t1/2 [31], [32], a distorted pattern consistent with the presence of endogenous antibodies.

Initial tests for antibodies were aimed at establishing or excluding the presence of “excessive functional binding” of labelled insulin/proinsulin in patient’s serum vs. normal sera free from endogenous antibodies. Initial and critical screening tests used were able to detect excessive “functional” binding irrespective of antibodies or paraprotein class such as IgM, IgG and IgA [7], [9]. Pragmatic, albeit not the only tests used were polyethylene glycol (PEG 6000 at 12.5% final concentration) or ammonium sulfate (50% final concentration) precipitation of immunoglobulins [34]. The finding of a higher amount of labelled insulin and/or proinsulin in the precipitate in patient’s serum compared with those found in normal sera (controls) was indicative of “endogenous functional binding” consistent with IAS. The limitation of such a simple test has been generally recognised because the amount of labelled insulin bindings to antibodies could vary from one sample to another taken from the same patient, being dependent on the number of “endogenously unoccupied” binding sites on the antibodies in the sample used (potentially being reduced postprandially but increase after prolonged fast). An alternative test used was to repeat the analyses of insulin, proinsulin and C-peptide before and after immunoglobulin/antibody precipitation in a patient’s serum and controls which allow assessment of the ratio of bound/free moieties. Currently, some laboratories routinely report both “free insulin” following PEG precipitation as well as total insulin. An additional test used was the Scatchard plot to identify the nature of the binding’s macromolecule, its class and binding characteristics. Some kit tests detect the physical presence of one specific class of insulin antibodies such as the common IgG and may be considered informative only in positive cases; however, if the outcome is negative, searching for other insulin/proinsulin antibodies of different class such as IgM or IgA as well as other binding macromolecules (e.g. paraprotein) was prudent and diagnostically necessary [7], [8]. Furthermore, kits which detect the immunological presence of a particular class of antibodies with a high degree of specificity may not detect others as well as proinsulin(s) antibodies if present. Investigating the presence of receptor antibodies was also considered in cases of unexplained hypoglycaemia in which exclusion of all forms of IAS per se was pursued and necessary.

Summary

Direct and thorough testing for excessive endogenous “functional” binding of insulin, proinsulin and/or receptor antibodies in non-diabetic, non-acutely ill patients with indeterminate cause of hypoglycaemia was necessary in all the IAS cases described. A systematic and cautious approach would however be prudent. Initial tests for insulin/proinsulin antibodies must be able to detect all classes/subclasses of immunoglobulins antibodies which if negative should be followed by tests for insulin receptor antibodies. Endogenous insulin/proinsulin and/or receptor antibodies when present have confirmed the diagnosis of IAS with certainty and avoided misinterpretation of pancreatic immunoassay results.

On the incongruity of IASs caused by endogenous insulin antibodies

Immunogenicity of a substance is generally influenced by numerous factors such as its molecular size, chemical complexity/heterogeneity and individual genotype. Proinsulin has 86 amino acids with a MW of 9390 Da and is considered more immunogenic [58] than insulin (51 amino acids) with MW of 5808 Da. No specific autoantibody against low MW C-peptide in IAS has been documented (this connecting-peptide with “unknown” function has 31 amino acids and MW of 3021 Da). The t1/2 of proinsulin and C-peptide are similar at ~25 min and 6-fold greater than the insulin t1/2 of ~4 min.

Endogenous antibodies are known to bind insulin and/or proinsulin(s) in patients with IAS. However, binding of C-peptide to endogenous antibodies in IAS though unreported remained a possibility, e.g. C-peptide was known to bind to antibodies in some insulin-dependent diabetic patients [58]. Currently, the assumption is that C-peptide does not bind to specific C-peptide antibodies in patients with IAS and its t1/2 remains unaltered. Elevated concentrations of C-peptide may be attributed to cross-reactivities with the C-peptide in proinsulin(s) which may be immunologically recognised, measured and “inappositely” reported as “free C-peptide” in some immunoassays [54], [55], [56], [57]. The above complexities and the heterogeneity of C-peptide measurements by immunoassays have therefore been a source of confusion, dependent on the nature of endogenous antibodies and immunoassay’s characteristics. The following are examples (without being prescriptive) of the reported adverse clinical and analytical manifestations of endogenous antibodies in patients with IAS.

IAS with endogenous antibodies which exclusively/predominantly bind insulin

The increase in the insulin t1/2 from minutes to potentially hour(s) caused a significant elevation in serum insulin concentration. Generally, insulin concentrations tended to be very high in IAS and rose further after meals/glucose administration to even higher concentrations [32], [57], [59]. The dynamics of C-peptide and/or proinsulin secretion and clearance remained largely unaltered, i.e. secreted in response to meals and/or glucose administration and declining thereafter with clearance at their normal rates to potentially low or very low concentrations, as shown in seven of the 17 serial samples taken from a Dutch case of IAS over a period of 48 h, i.e. 40% [32]. Ephemeral increase in serum concentrations of proinsulin(s) and/or C-peptide is therefore likely to occur shortly after meals or glucose administration. Needless to say that a biochemical profile of high serum insulin, low serum proinsulin and C-peptide with negative endogenous insulin/proinsulin(s) binding and receptor antibodies would be consistent with exogenous insulin administration (self-harm or malicious).

IAS with endogenous antibodies which predominantly bind proinsulin(s)

Serum proinsulin(s) concentration is normally ~10%–15% that of insulin. Predominantly proinsulin antibodies have caused a significant increase in proinsulin concentration and its t1/2. Presumably other proinsulin split intermediates (not normally considered among pancreatic hormone immunoassay measurements) can potentially bind to these endogenous proinsulin antibodies. The serum concentration of insulin and C-peptide may also be elevated in some immunoassays with significant cross-reactivities [54], [55], [56], [57], [60]. The clinical manifestation of hypoglycaemia varied [61]. Concomitant increase in serum proinsulin, insulin and C-peptide concentrations was interpreted as a prima facie evidence of “phantom” insulinoma [62]. Of interest too are two recent studies by Vezzosi et al. (2007; on 33 patients) [63] and Guettier et al. (2013; on 56 patients) [64] who suggested that end-of-fast hypoglycaemia associated with serum proinsulin >22 pmol/L best detect cases of insulinomas. However, these well-executed and informative studies on patients free from endogenous antibodies did not highlight the possibility of misdiagnosis if insulin/proinsulin antibodies are present but not excluded a priori.

IAS with mixed binding endogenous antibody(s) against insulin and proinsulin(s)

Both serum insulin and proinsulin were elevated with variable C-peptide concentrations dependent on immunoassays’ cross-reactivities [3], [31], [62], [65]. Elevated serum insulin concentrations associated with high proinsulin and C-peptide concentrations may be taken as prima facie evidence of an insulinoma if endogenous antibodies are not excluded a priori. Distinguishing between a single antibody which binds both moieties vs. two separate clones of antibodies has required more work a posteriori and was of academic rather than clinical interest.

IAS with endogenous insulin receptor antibodies (type-B insulin resistance)

Type-B insulin resistance is complex, common in women, usually associated with hyperglycaemia and extreme resistance to insulin. However, in ~24% of patients [66], the insulin receptor antibody can be agonistic thus causing hypoglycaemic manifestations. The postulate is that the binding of agonistic antibodies to receptors inhibits endogenous insulin binding, thus reducing its removal by endocytosis and subsequent intracellular metabolic degradation [67] resulting in hyperinsulinaemia and concomitant hypoglycaemia with inverse C-peptide/insulin ratio (normally C-peptide concentration in blood is ~5 times higher than insulin) [68]. The first case of insulin receptor antibodies was reported in 1976 [68] with new variants still being described [69], [70], [71], [72]. Reported hypoglycaemic manifestations vary occurring as fasting, postprandial or spontaneous and in some can be severe and protracted. Episodes of hyperglycaemia may also precede hypoglycaemia [73], [74], [75], [76], [77], [78], [79], attributed by some to the simultaneous co-existent of agonistic and antagonistic insulin receptor antibodies, the titre, nature and ratio of which may also change with time [51].

Hyperinsulinaemia in these patients has been attributed to (a) increased pancreatic hormone secretion of insulin to compensate for the peripheral insulin resistance and (b) concomitant reduction in insulin clearance. Elevated serum insulin associated with low proinsulin/C-peptide may be misdiagnosed as injectable insulin (self-harm or malicious) if receptor antibodies are not excluded a priori. Several methods have been described exploiting the pleiotropic effect of insulin-receptor binding; examples are glucose oxidation and tyrosine kinase stimulation; others are based on inhibition of insulin binding to RBC receptors (radioreceptor assay) and immunoprecipitation of the insulin receptors [78], [79]. Radioreceptor inhibition assays are commonly used in clinical laboratories and assess the magnitude of reduction in the binding of labelled-insulin to insulin receptors in vitro. The presence of insulin binding antibodies can interfere in radioreceptor assays and its removal/separation prior to radioreceptor analysis was considered necessary.

IAS with both insulin/proinsulin binding and receptor antibodies

This occurred sequentially [43] or in combination [44], [45], [46], [47], [48] as part of the idiotypic immune response (for more details, see [48]). The clinical and biochemical manifestations were dependent on the preponderance and characteristics of each antibody, which have the potential to make some biochemical tests, in some cases inconsistent and even confusing [43], [44], [45], [46], [47], [48]. A positive test for insulin and/or proinsulin(s) binding antibodies establishes the diagnosis of IAS; additional tests for receptor antibodies were necessary to confirm this entity of IAS [74], [78].

Comments and conclusions

It may be important to comment further on some aspects of the endogenous antibodies which underpin the aetiology of IAS to help elucidate its highly variable biochemical and clinical manifestations, i.e. onset, spontaneousness, duration and severity of hypoglycaemia. As mentioned before, some viruses can increase insulin antibodies. An important feature of the immune response is that it remembers if it has seen such a pathogen (or a closely related antigen) before, reacts to a second and further exposure in a manner different than after a primary exposure. The reaction to a second exposure, e.g. reinfection(s) or sustained exposure reflects the immune memory response and “affinity maturation”. In such case, antibodies production are of the IgG, IgA, IgD or IgE class rather than the IgM, commonly produced after an initial exposure, i.e. class switching [80]. Furthermore, the immune response to a second and further exposure can be fast and vigorous with progressive increase in the titre of high affinity/avidity endogenous antibodies which in some cases can double every few hours [81], [82]. It is therefore reasonable to expect that the titre, capacity and affinity/avidity of antibody production in response to pathogens and/or antigenic modifications to be variable and unpredictable [2], [3], [16], [17], [18].

Of the many features of antibodies of IAS, probably the most crucial is their affinities/avidities. The binding between an antigen (Ag) and antibody (Ab) is not a reaction but a reversible interaction. The rate law equations (zero, first and second rate kinetic reactions) for this reversible interaction are describable by the forward/association and backward/dissociation rate constants (i.e. K1 and K−1, respectively) and are applicable to any Ag-Ab interaction, both in vivo and in vitro and follow a reversible bimolecular thermodynamic principle (for further details, see [38], [39], [49], [81], [82]). The association binding reaction (K1) is usually dominant and fast; the reverse dissociation rate reaction (K−1) determines the length of time the antigen remains bound to their antibodies and is influenced by the number and types of non-covalent bonds that are formed between the antigenic-determinant, i.e. epitope(s) on the antigen and the corresponding antigen-combining site on the paratope(s) on the antibody [81], [82]. The association reaction ranges over several orders of magnitude; the reverse dissociation rate reaction (K−1) has similarly wide range [81], [82]). The equilibrium constant (Keq) is the ratio between these reactions (reactants) at equilibrium and the greater the strength of the bonds, the higher its Keq which is typically between 108 and 1010 (a 100-fold difference). A dominant K1 is necessary in order to develop an endogenous insulin/proinsulin(s) pool in the first place to cause IAS. The titre and capacity of antibodies would define the size of this pool of bound insulin and/or proinsulin(s) and K−1 would determine the rate of insulin/proinsulin(s) released into the blood, affecting the onset, duration and severity of hypoglycaemia [31], [40]. The clinical course of most autoimmune diseases may not be straightforward, and tend to wax and wane by obscure and poorly understood mechanism(s). Of interest is that such a phenomenon has also been demonstrated in patients with IAS as shown by the epidemiological study of Uchigata et al. on 197 Japanese patients in whom many had spontaneous remission without any “positive” treatment and in others after the discontinuation of drugs containing sulfur. In few patients however, recurrent attacks have subsequently ensued without exposure to any drug prior to their recurrent hypoglycaemic episodes [83].

Unlike pancreatic hormones which have t1/2 of minutes extended to hours in IAS, antibodies per se (irrespective of class or subclass) have a much longer t1/2 of 3 weeks or more thus allowing belated detection. Testing for “insulin antibodies” is now included among the first-line tests in the recently published clinical practice guidelines by the American Endocrine Society in 2009 [84] and reiterated in a short note in 2012 [85]. Consideration and searching for various forms of functional binding antibodies a priori is important because of the vulnerabilities of many immunoassays to produce results which if interpreted in isolation could be highly misleading. Antibodies which underpin IAS have the ability to bind insulin and/or proinsulin(s) and the potential to distort not only their own measurements [54], [55], [56], [57] but also that of C-peptide immunoassays. This may also help to explain the wide range of concentrations reported in the literature [31], [34], [54], [55], [56], [57], [81], [85], [86], [87]. The specificities of immunoassays of insulin, proinsulin and C-peptide are known to vary from one reagent antibody to another leading to poor “between-methods” interchangeability [54], [55], [56], [57], [88], [89], [90]. However, despite these inaccuracies, immunoassay results remained on the whole useful and informative when interpreted in the context of the prevailing antibodies and with appropriately/timely taken samples during fasting and after meals and/or an extended OGTT [31], [32]. Direct testing for all forms of endogenous insulin/proinsulin antibodies remains however, the sine qua non practices in establishing or excluding the diagnosis of IAS in non-diabetic, non-critically ill individuals [85].

In conclusion, new informative cases of IAS in Caucasians continued to appear in the world literature [27], [28], [69], [70], [75], [90], [91], [92], [93], [94], [95], [96], [97], [98], [99]. The very wide range of their clinical presentations and biochemical findings re-emphasised that IAS is a highly heterogeneous syndrome. Furthermore, it confirms that IAS per se has no specific clinical features of its own but has the potential to mimic other causes of hypoglycaemia clinically and biochemically. In this paper, the term IAS was confined to hypoglycaemia in non-diabetic, non-acutely ill patients in whom thorough testing for insulin/proinsulin binding and receptor antibodies was essential for confirming diagnosis and rational interpretation of pancreatic hormone data. Potential misdiagnosis is therefore likely to occur if these two interdependent entities namely the presence and nature of endogenous antibodies and immunoassays cross-reactivities are not considered and taken into account. Finally, the term IAS must not be confused with the effect of endogenous insulin antibodies in patients with brittle diabetes triggered by exogenous administration of insulin or insulin analogues causing poor glycaemic control and hypoglycaemia by a mechanism similar to IAS [100], [101], [102], [103].

  1. Author contributions: The author has accepted responsibility for the entire content of this submitted article and approved submission.

  2. Research funding: None declared.

  3. Employment or leadership: None declared.

  4. Honorarium: None declared.

  5. Competing interests: The funding organization(s) played no role in the study design; in the collection, analysis, and interpretation of data; in the writing of the report; or in the decision to submit the report for publication.

References

1. Fineberg SE, Biegel AA, Durr KL, Hufferd S, Fineberg NS, Anderson JH. Presence of insulin antibodies as regular feature of nondiabetic repertoire of immunity. Diabetes 1991;40:1187–93.10.2337/diab.40.9.1187Search in Google Scholar

2. Brickman CM, Shoenfeld Y. The mosaic of autoimmunity. Scand J Clin Lab Invest 2001;61(Suppl 235):3–15.Search in Google Scholar

3. Semsei I. On the role of aging in the etiology of autoimmunity. Gerentology 2002;48:179–84.10.1159/000052839Search in Google Scholar

4. Hirata Y, Ishizu H. Elevated insulin-binding capacity of serum proteins in a case with spontaneous hypoglycemia and mild diabetes not treated with insulin. Tohoku J Exp Med 1972:107:277–86.10.1620/tjem.107.277Search in Google Scholar

5. Følling I, Norman N. Hyperglycemia, hypoglycaemic attacks, and production of anti-insulin antibodies without previous known immunization. Immunological and functional studies in a patient. Diabetes 1972;21:814–26.10.2337/diab.21.7.814Search in Google Scholar

6. Sugunendran S, Malik M. Insulin autoimmune syndrome: a rare case of hypoglycaemia. Endocrine Abstracts 2010;21:70.Search in Google Scholar

7. Halsall DJ, Mangi M, Soos M, Fahie-Wilson MN, Wark G, Mainwaring-Burton R, et al. Hypoglycaemia due to an insulin binding antibody in a patient with IgA-k myeloma. J Clin Endocrinol Metab 2007;92:2013–16.10.1210/jc.2007-0075Search in Google Scholar

8. Redmon B, Pyzdrowski KL, Elson MK, Dalmasso AP, Nuttall FQ. Brief report: hypoglycemia due to a monoclonal insulin binding antibody in multiple myeloma. N Engl J Med 1992;26:994–8.10.1056/NEJM199204093261505Search in Google Scholar

9. Khant M, Florkowski CM, Livesey J. Insulin autoimmune syndrome due to IgG kappa paraprotein. Pathology 2004;36:86–7.10.1080/00313020310001643624Search in Google Scholar

10. Cervera R. The epidemiology and significance of autoimmune diseases in health care. Scand J Clin Lab Invest 2001;61(Suppl 235):27–30.10.1080/003655101753352013Search in Google Scholar

11. Redmon JB, Nuttall FQ. Autoimmune hypoglycaemia. Endocrinol Metab Clin North Am 1999;28:603–18.10.1016/S0889-8529(05)70090-6Search in Google Scholar

12. Uchigata Y, Kuwata S, Tokunaga K, Omen Y, Hirata Y, Kuwasta S, et al. Strong association of insulin autoimmune syndrome with HLA-DR 4. Lancet 1992;339:393–4.10.1016/0140-6736(92)90080-MSearch in Google Scholar

13. Eisenbarth GS, Gottlieb PA. Autoimmune polyendocrine syndromes. N Engl J Med 2004;350:2068–79.10.1056/NEJMra030158Search in Google Scholar

14. Li JK, Lee WK, Chan CH, Ng RS, Wong BK, Mo KK. Auto-insulin antibodies due to methimazole in a patient with Myasthenia Gravis and Graves disease. Ann Gerontol Geriatric Res 2015;2:1021–2.Search in Google Scholar

15. Rouabhia S, Ramanoelina J, Godmer P, Reach G, Dutel J-L, Guillevin L. Insulin autoimmune syndrome revealing systemic lupus erythematosus. Ann Med Interne (Paris) 2003;154:59–60.Search in Google Scholar

16. Bodansky HJ, Dean BM, Bottazzo GF, Grant PJ, McNally J, Hambling MH, et al. Islet cell antibodies and insulin antibodies in association with common viral infections. Lancet 1986:1351–3.10.1016/S0140-6736(86)92003-9Search in Google Scholar

17. Ruíz-Giardín JM, Cremades C, Romero J, Marazuela M. Insulin autoimmune syndrome in a patient with viral hepatitis C. Clin Endocrinol 2002;57:411–2.10.1046/j.1365-2265.2002.01604.xSearch in Google Scholar

18. Fineberg SE, Kawabata TT, Finco-Kent D, Fountain RJ, Finch GL, Krasner AS. Immunological responsonses to exogenous insulin. Endocr Rev 2007;28:625–52.10.1210/er.2007-0002Search in Google Scholar

19. Uchigata Y, Hirata Y, Iwamoto Y. Drug-induced insulin autoimmune syndrome. Diabetes Res Clin Practice 2009;83:e19–20.10.1016/j.diabres.2008.10.015Search in Google Scholar

20. Gomez Cruz MJ, Jabbar M, Saini E, Crawford B, Vazquez DM, Menon R, et al. Severe hypoglycaemia secondary to methimazole-induced insulin autoimmune syndrome in a 16 years old African-American male. Paediatr Diabetes 2012;13:652–5.10.1111/j.1399-5448.2012.00884.xSearch in Google Scholar

21. Ma W-Y, Won JG, Tang K-T, Lin H-D. Severe hypoglycaemic coma due to insulin autoimmune syndrome. J Chin Med Assoc 2005;68:82–6.10.1016/S1726-4901(09)70140-6Search in Google Scholar

22. Uchigata Y. The novel agent, alpha lipoic acid, can cause the development of insulin autoimmune syndrome. Internal Med 2007;46:1321–22.10.2169/internalmedicine.46.0221Search in Google Scholar PubMed

23. Takeuchi Y, Miyamoto T, Kakizawa T, Shigematsua S, Hashizume K. Insulin autoimmune syndrome possibly caused by alpha lipoic acid. Internal Med 2007;46:237–39.10.2169/internalmedicine.46.1893Search in Google Scholar PubMed

24. Ishida Y, Ohara T, Okuno Y, Ito T, Hirota Y, Fukukawa K, et al. α-Lipoic Acid and insulin autoimmune syndrome. Diabetes Care 2007;30:2240–1.10.2337/dc07-0689Search in Google Scholar PubMed

25. Furukawa N, Miyamura N, Nishida K, Motoshima H, Taketa K, Araki E. Possible relevance of alpha lipoic acid contained in a health supplement in a case of insulin autoimmune syndrome. Diabetes Res Clin Pract 2007;75:366–7.10.1016/j.diabres.2006.07.005Search in Google Scholar

26. Bae SM, Bae MN, Kim EY, Kim II, Seo MW, Shin JK, et al. Recurrent insulin autoimmune syndrome caused by α-lipoic acid in type 2 diabetes. Endocrinol Metab (Seoul) 2013;28:326–30.10.3803/EnM.2013.28.4.326Search in Google Scholar

27. Bresciani E, Bussi A, Bazzigaluppi E, Balestrieri G. Insulin autoimmune syndrome induced by alpha-lipoic acid in a Caucasian woman: case report. Diabetes Care 2011;34:e146.10.2337/dc11-0600Search in Google Scholar

28. Gullo D, Evans JL, Sortino G, Goldfine ID, Vigneri R. Insulin autoimmune syndrome (Hirata disease) in European Caucasians taking alpha-lipoic acid. Clin Endocrinol 2014;81:204–9.10.1111/cen.12334Search in Google Scholar

29. Hirata Y, Tominaga M, Ito J-I, Noguchi A. Spontaneous hypoglycaemia with insulin autoimmunity in Graves’ disease. Ann Intern Med 1974;81:214–8.10.7326/0003-4819-81-2-214Search in Google Scholar

30. Yaturu S, DePrisco C, Lurie A. Severe autoimmune hypoglycemia with insulin antibodies necessitating plasmapheresis. Endocr Pract 2004;10:49–54.10.4158/EP.10.1.49Search in Google Scholar

31. Goldman J, Baldwin D, Rubenstein AH, Klink DD, Blackard WG, Fisher LK, et al. Characterization of circulating insulin and proinsulin-binding antibodies in autoimmune hypoglycaemia. J Clin Invest 1978;63:1050–9.10.1172/JCI109374Search in Google Scholar

32. Schlemper RJ, Uchigata Y, Frolich M, Vingerhoeds AC, Meinders AE. Recurrent hypoglycaemia caused by the insulin autoimmune syndrome: the first Dutch case. Neth J Med 1996;48:188–92.10.1016/0300-2977(95)00085-2Search in Google Scholar

33. Vogeser M, Parhofer KG, Furst H, Jacob K, Brodl C, Seidel D. Autoimmune hypoglycaemia presenting as seizure one week after surgery. Clin Chem 2001;47:795–6.10.1093/clinchem/47.4.795Search in Google Scholar

34. Ismail AA. The double whammy of endogenous insulin antibodies in non-diabetic subjects. Clin Chem Lab Med 2008;46:153–6.10.1515/CCLM.2008.031Search in Google Scholar PubMed

35. Tompkins CV, Shivastava MC, Sonksen PH, Nabarro DN. A comparison study of the distribution of monocomponent human insulin and porcine proinsulin in man. Biochem J 1971;125:65P.10.1042/bj1250064PaSearch in Google Scholar

36. Glauber HS, Revers RR, Henry R, Schmeiser L, Wallace P, Kolterman OG, et al. In vivo deactivation of proinsulin action on glucose disposal and hepatic glucose production in normal man. Diabetes 1986;35:311–7.10.2337/diab.35.3.311Search in Google Scholar

37. Arky RH. Antibodies and disorders of glucose metabolism: a courting couple. N Engl J Med 1982;307:1445–7.10.1056/NEJM198212023072310Search in Google Scholar

38. Reverberi R, Reverberi L. Factors affecting the antigen-antibody reaction. Blood Transfus 2007;5:227–40.Search in Google Scholar

39. Egawa T, Tsuneshige A, Suematsu M. Method for determination of association and dissociation rate constants of reversible bimolecular reactions by isothermal titration calorimeters. Anal Chem 2007;79:2972–8.10.1021/ac062183zSearch in Google Scholar

40. Sklenar I, Wilkin TJ, Diaz J-L, Erb P, Keller U. Spontaneous hypoglycaemia associated with autoimmunity specific to human insulin. Diabetic Care 1987;10:2152–9.10.2337/diacare.10.2.152Search in Google Scholar

41. Anderson JH, Blackard WG, Goldman J, Rubenstein AH. Diabetes and hypoglycaemia due to insulin antibodies. Am J Med 1978;64:868–73.10.1016/0002-9343(78)90530-2Search in Google Scholar

42. Reisch N, Zwermann O, Reincke M. Hypoglycaemia and transient diabetes mellitus. Dtsch Med Wochenschr 2004;129:1135–8.10.1055/s-2004-824862Search in Google Scholar

43. Elias D, Cohen IR, Schechter Y, Spirer Z, Golander A. Antibodies to insulin receptor followed by anti-idiotype antibodies to insulin in child with hypoglycaemia. Diabetes 1987;36:348–54.10.2337/diab.36.3.348Search in Google Scholar

44. Kato T, Itoh M, Hanashita J, Matsumoto T, Ono Y, Imamura N, et al. Severe hypoglycaemia in a person with insulin autoimmune syndrome accompanied by insulin receptor anomaly type B. Diab Med 2007;24:1279–81.10.1111/j.1464-5491.2007.02232.xSearch in Google Scholar

45. Bortolotti D, Mothe-Satney I, Ferrari P, Gautier N, Sonke J, Pallé S, et al. Spontaneous hypoglycaemia in the presence of both anti-insulin antibody and anti-insulin receptor antibody. Diabetes Metab 2006;32:598–603.10.1016/S1262-3636(07)70314-2Search in Google Scholar

46. Shim MS, Kim MY, Kim MJ, Lee Y, Lee BJ, Chung CH, et al. A case of autoimmune hypoglycaemia complicated with diabetic ketoacidosis. Yonsei Med J 2004;45:140–4.10.3349/ymj.2004.45.1.140Search in Google Scholar

47. Kim CH, Park H, Park TS, Baek HS. Autoimmune hypoglycemia in a type 2 diabetic patient with anti-insulin and insulin receptor antibodies. Diabetes Care 2004;27:288–9.10.2337/diacare.27.1.288Search in Google Scholar

48. Kim CH, Park TS. Autoimmune hypoglycaemia in a type 2 diabetic patient with anti-insulin and insulin receptor antibodies. Diabetes Care 2004;27:1247.10.2337/diacare.27.5.1247Search in Google Scholar

49. Ismail AA, Walker PL, Cawood ML, Barth JH. Interference in immunoassay is an underestimated problem. Ann Clin Biochem 2002;39:366–73.10.1258/000456302760042128Search in Google Scholar

50. Marks V, Teale JD. Hypoglycaemic disorders. Clin Lab Med 2001;21:79–97.Search in Google Scholar

51. Glaser B, Leibowitz G. Hypoglycemia. In: Joslin EP, Kahn CR, editors. Diagnostic and therapeutic approach to the hypoglycaemic patient in Joslin’s Diabetes Mellitus, 14th ed. Lippincott, Philadelphia, Pennsylvania, USA: Williams and Wilkins (LWW), 2005:1148–75. ISBN-10: 0781727960.Search in Google Scholar

52. Peavy DE, Brunner MR, Duckworth WC, Hooker CS, Frank BH. Receptor binding and biological potency of several split forms (conversion intermediates) of human proinsulin. Studies in cultured IM-9 lymphocytes and in vivi and in vitro rats. J Biol Chem 1985;260:13989–94.10.1016/S0021-9258(17)38673-8Search in Google Scholar

53. Heath WE, Belagaje RM, Brooke GS, Chance RE, Hoffmann JA, Long HB, et al. (A-C-B) human proinsulin, a novel insulin agonist and intermediates in the synthesis of biosynthetic human insulin. J Biol Chem 1992;267:419–25.10.1016/S0021-9258(18)48511-0Search in Google Scholar

54. Clark PM. Assays for insulin, proinsulin(s) and C-peptide. Ann Clin Biochem 1999;36:541–64.10.1177/000456329903600501Search in Google Scholar PubMed

55. Clark P, McDonald T. Diabetes mellitus. In: David Wild, editor. The immunoassay handbook; theory and applications of ligand binding, ELISA and related techniques, 4th ed. Elsevier science, 2013;783–94. ISBN-10:0080970370; ISBN-13: 978-0080970370.10.1016/B978-0-08-097037-0.00064-6Search in Google Scholar

56. Heding LG. Radioimmunoassays for insulin, C-peptide and proinsulin. Lancaster, England: MTP press Ltd, 1988. DOI: 10.1007/978-94-011-7094-9.Search in Google Scholar

57. Sapin R. Anti-insulin antibodies in insulin immunometric assays: a still possible pitfall. Eur J Clin Chem Clin Biochem 1997;35:365–7.10.1515/cclm.1997.35.5.365Search in Google Scholar PubMed

58. Keilacker H, Rjasanowski I, Woltanski KP, Ziegler B, Kohnert KD, Michaelis D, et al. Antibodies against insulin (IAA), C-peptide (CAA), and glucagon (GAA) in new-onset type 1 diabetic patients. Exp Clin Endocrinol 1990;95:123–8.10.1055/s-0029-1210944Search in Google Scholar PubMed

59. Casenoves A, Mauri M, Dominguez JR, Alfayate R, Pico AM. Influence of anti-insulin antibodies on insulin immunoassays in the autoimmune insulin syndrome. Ann Clin Biochem 1998;35:768–74.10.1177/000456329803500610Search in Google Scholar PubMed

60. Houssa P, Dinesen B, Deberg M, Frank BH, Schravendijk CV, Sodoyez-Gaffaux F, et al. First direct assay for intact human proinsulin. Clin Chem 1998;44:1514–9.10.1093/clinchem/44.7.1514Search in Google Scholar

61. Tillil H, Frank BH, Pekar AH, Broelsch C, Rubenstein AH, Polonsky KS. Hypoglycaemic potency and metabolic clearance rate of intravenously administered human proinsulin and metabolites. Endocrinology 1990;127:2418–22.10.1210/endo-127-5-2418Search in Google Scholar PubMed

62. Moreira RO, Lima GA, Peixoto PC, Farias ML, Vaisman M. Insulin autoimmune syndrome: case report. Sao Paulo Med J 2004;122:178–80.10.1590/S1516-31802004000400010Search in Google Scholar

63. Vezzosi D, Bennet A, Fauvel J, Carton P. Insulin, C-peptide and proinsulin for the biochemical diagnosis of hypoglycaemia related to endogenous hyperinsulinaemia. Eur J Endocrinol 2007;157:76–83.10.1530/EJE-07-0109Search in Google Scholar PubMed

64. Guettier J-M, Lungu A, Goodling A, Cohran C, Gorden P. The role of proinsulin and insulin in the diagnosis of insulinoma: a critical evaluation of the endocrine society clinical practice guidline. J Clin Endocrinol Metab 2013;98:4752–8.10.1210/jc.2013-2182Search in Google Scholar PubMed PubMed Central

65. Lohmann T, Kratzsch J, Kellner K, Witzigmann H, Hauss J, Paschke R. Severe hypoglycemia due to insulin autoimmune syndrome with insulin antibodies cross reactive to proinsulin. Exp Clin Endocrinol Diabetes 2001;109:245–8.10.1055/s-2001-15113Search in Google Scholar PubMed

66. Lupsa BC, Cheng AY, Cochran EK, Soos MA, Semple PK, Gorden P. Autoimmune forms of hypoglycaemia. Medicine (Baltimore) 2009;88:141–53.10.1097/MD.0b013e3181a5b42eSearch in Google Scholar PubMed

67. Duckworth WC, Bennett RG, Hamel FG. Insulin degradation: progress and potential. Endocr Rev 1998;19:608–24.Search in Google Scholar

68. Kahn CR, Flier JS, Bar RS, Archer JA, Gordon P, Martin MM, et al. The syndrome of insulin resistance and acanthosis nigricans. Insulin-receptor disorder in man. N Eng J Med 1976;294:439–45.10.1056/NEJM197604012941401Search in Google Scholar PubMed

69. Bourron O, Caron-Debarle M, Hie M, Halbron M, Andreelli F, Fonfrede M, et al. Type B insulin-resistance syndrome: a cause of reversible autoimmune hypoglycaemia. Lancet 2014;384:1548–9.10.1016/S0140-6736(14)61833-XSearch in Google Scholar

70. Kang SM, Jin HY, Lee KA, Park JH, Baek HS, Park TS. Type B insulin-resistance syndrome presenting as autoimmune hypoglycaemia associated with systemic lupus erythematosus and interstitial lung disease. Korean J Intern Med 2013;28:98–102.10.3904/kjim.2013.28.1.98Search in Google Scholar

71. Venugopal Y, Vethankkon S, Sockalingam S, Jasmin R, Choong K. A case of persistent hypoglycaemia: when to think outside the box. Clin Diabetes 2013;31:130–33.10.2337/diaclin.31.3.130Search in Google Scholar

72. Yamada H, Asano T, Kusaka I, Kakei M, Ishikawa S. Type-B insulin resistance syndrome with fasting hypoglycaemia and postprandial hyperglycaemia. Diabetology Int 2015;6:144–8.10.1007/s13340-014-0190-ySearch in Google Scholar

73. De Pirro PR, Roth RA, Rossetti L. Characterization of the serum from a patient with insulin resistance and hypoglycaemia. Evidence of multiple populations of insulin receptor antibodies with different receptor binding and insulin-mimicking activities. Diabetes 1984;33:301–4.10.2337/diab.33.3.301Search in Google Scholar

74. Taylor SI, Barbetti F, Accili D, Roth J, Gorden P. Syndromes of autoimmunity and hypoglycemia. Antibodies directed against insulin and its receptor. Endocrinol Metab Clin North Am 1989;18:123–43.10.1016/S0889-8529(18)30392-XSearch in Google Scholar

75. Chon S, Choi MC, Lee YJ, Hwang YC, Jeong IK, Oh S, et al. Autoimmune hypoglycaemia in a patient with characterization of insulin receptor antibodies. Diabetes Metab J 2011;35:80–5.10.4093/dmj.2011.35.1.80Search in Google Scholar

76. Yamasaki H, Yamaguchi Y, Fujita N, Kato C, Kuwahara H, Yamauchi MD, et al. Anti-insulin receptor antibodies in a patient with type B insulin resistance and fasting hypoglycaemia. Acta Diabetol 2000;37:189–96.10.1007/s005920070004Search in Google Scholar

77. Rodriguez O, Collier E, Arakaki R, Gorden P. Characterization of purified antibodies to the insulin receptor from six patients with type B insulin resistance. Metabolism 1992;41:325–31.10.1016/0026-0495(92)90279-JSearch in Google Scholar

78. Batarseh H, Thompson RA, Odugbesan O, Barnett AH. Insulin receptor antibodies in diabetes mellitus. Clin Exp Immunol 1988;71:85–90.Search in Google Scholar

79. Walters EG, Denton RM, Tavare JM, Walters G. Hypoglycaemia due to an insulin-receptor antibodies in Hodgikin’s disease. Lancet 1987;329:241–3.10.1016/S0140-6736(87)90064-XSearch in Google Scholar

80. Wabl M, Steinberg C. Affinity maturation and class switching. Curr Opinion Immunol 1996;8:89–92.10.1016/S0952-7915(96)80110-5Search in Google Scholar

81. Ismail AA. Identifying and reducing potentially wrong immunoassay results when plausible and “not-unreasonable”. Advances Clin Chem 2014;66:241–94.10.1016/B978-0-12-801401-1.00007-4Search in Google Scholar

82. Ismail AA. Interference from endogenous antibodies in automated immunoassays: what laboratorians need to know. J Clin Pathol 2009;62:673–8.10.1136/jcp.2008.055848Search in Google Scholar

83. Uchigata Y, Eguchi Y, Takayama-Hasumi S, Omori Y. Insulin autoimmune syndrome (Hirata disease): clinical features and epidemiology in Japan. Diabetes Res Clin Pract 1994;22:89–94.10.1016/0168-8227(94)90040-XSearch in Google Scholar

84. Cryer PE, Axelrod L, Grossman AB, Heller SR, Montori VM, Seaquist ER, et al. Evaluation and management of adult hypoglycemic disorders: an endocrine society clinical practice guideline. J Clin Endocrinol Metab 2009;94:709–28.10.1210/jc.2008-1410Search in Google Scholar PubMed

85. Ismail AA. Testing for insulin antibodies is mandatory in the differential diagnosis of hypoglycaemia in non-diabetic subjects. Clin Endocrinol 2012;76:603–4.10.1111/j.1365-2265.2011.04259.xSearch in Google Scholar PubMed

86. Ismail AA, Ismail AA, Ismail Y. Probabilistic Bayesian reasoning can help identifying potentially wrong immunoassays results in clinical practice: even when they appear not-unreasonable. Ann Clin Biochem 2011;48:65–71.10.1258/acb.2010.010197Search in Google Scholar PubMed

87. Mills F, Jeffery J, Mackenzie P, Cranfield A, Ayling RM. An immunoglobulin G complexed form of thyroid stimulating hormone (macro thyroid-stimulating hormone) is a cause of elevated serum thyroid-stimulating hormone concentration. Ann Clin Biochem 2013;50:416–20.10.1177/0004563213476271Search in Google Scholar PubMed

88. Sapin R. The interference of insulin antibodies in insulin immunometric assays. Clin Chem Lab Med 2002;40:705–8.10.1515/CCLM.2002.121Search in Google Scholar PubMed

89. Sapin R. Insulin assays: previously known and new analytical features. Clin Lab 2003;49:113–21.Search in Google Scholar

90. Manley SE, Stratton IM, Clark PM, Luzio SD. Comparison of 11 human insulin assays: implications for clinical investigations and research. Clin Chem 2007;53:922–32.10.1373/clinchem.2006.077784Search in Google Scholar PubMed

91. Sudano M, Turchi F, Sossai P. Insulin autoimmune syndrome (Hirata disease): case report in a Caucasian patient with new-onset diabetes. Clin Med Diagnosis 2012;2:51–3.10.5923/j.cmd.20120205.02Search in Google Scholar

92. Arzamendi AE, Rajamani U, Jialal I. Pseudoinsulinoma in a white man with autoimmune hypoglycemia due to anti-insulin antibodies: value of the free C-peptide assay. Am J Clin Pathol 2014;142:689–93.10.1309/AJCPX56JQBJHUBGJSearch in Google Scholar PubMed

93. Philippon M, Séjil S, Mugnier M, Rocher L, Guibergia C, Vialettes B, et al. Use of the continuous glucose monitoring system to treat insulin autoimmune syndrome: quantification of glucose excursions and evaluation of treatment efficacy. Diabet Med 2014;31:e20–4.10.1111/dme.12418Search in Google Scholar PubMed

94. Gopal K, Priya G, Gupta N, Praveen EP, Khadgawat R. A case of autoimmune hypoglycaemia outside Japan: rare but in the era of expanding drug-list, important to suspect. Indian J Endocrinol Metab 2013;17:1117–9.10.4103/2230-8210.122644Search in Google Scholar PubMed PubMed Central

95. Grotemeyer KC, Zimmer V, Friesenhaln-Ochs B, Lammert F. 75-Jähriger patient mit rezidivierenden hypoglykämien bei insulinexzess. Der Internist 2013;54:762–4.10.1007/s00108-013-3278-8Search in Google Scholar PubMed

96. Wong SL, Priestman A, Holmes DT. Recurrunt hypoglycaemia from insulin autoimmune syndrome. J Gen Intern Med 2014;29:250–4.10.1007/s11606-013-2588-9Search in Google Scholar PubMed PubMed Central

97. Savas-Erdeve S, Agladioglu SY, Onder A, Kendirici HN, Bas VN, Sagsak E, et al. An uncommon cause of hypoglycaemia: insulin autoimmune syndrome. Horm Res Paediatr 2014;82:278–82.10.1159/000362758Search in Google Scholar PubMed

98. White BP, Southwood R. Persistent hypoglycaemia of unknown etiology in a patient without diabetes: a case report and review. J Pharm Pract 2013;26:138–43.10.1177/0897190012465984Search in Google Scholar PubMed

99. Nettanande C, de Silva HJ, Fernando R. Hypoglycaemia and fits in a thyrotoxic man. BMJ Case Rep 2009;2009. (DOI: 10.1136/bcr.07.2008.0448).Search in Google Scholar PubMed PubMed Central

100. Jassam N, Amin A, Holland P, Semple RK, Halsall DJ, Barth JH. Analytical and clinical challenges in a patient with concurrent type 1 diabetes, subcutaneous insulin resistance and insulin autoimmune syndrome. Endocrinol Diabetes Metab Case Rep 2014;2014. DOI: 10:1530/EDM-13-0086.Search in Google Scholar

101. Matsuzoe K, Shimoda S, Tsuruzoe K, Taketa K, Chirioka T, Sakamoto F, et al. A case of slowly progressive type 1 diabetes with unstable glycaemic control caused by unusual insulin antibody and successfully treated with steroid therapy. Diabetes Res Clin Practice 2006;72:238–43.10.1016/j.diabres.2005.10.018Search in Google Scholar PubMed

102. Trabucchi A, Faccinetti NI, Krochik AG, Arriazu MC, Poskus E, Valdez SN. Characterisation of insulin antibodies by surface plasmon resonance in two clinical cases: brittle diabetes and insulin autoimmune syndrome. PLoS One 2013;8:e84099.10.1371/journal.pone.0084099Search in Google Scholar PubMed PubMed Central

103. Ismail AA. Insulin analogues as a new example of interference in insulin assays. Ann Clin Biochem 2016;53:181–2.10.1177/0004563215590165Search in Google Scholar PubMed

Received: 2015-12-21
Accepted: 2016-2-23
Published Online: 2016-4-12
Published in Print: 2016-11-1

©2016 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 26.4.2024 from https://www.degruyter.com/document/doi/10.1515/cclm-2015-1255/html
Scroll to top button