49
An overview of amyloidosis Author Peter D Gorevic, MD Section Editor Peter H Schur, MD Deputy Editor Paul L Romain, MD Disclosures Last literature review version 19.3: septiembre 2011 | This topic last updated: octubre 13, 2011 INTRODUCTION Although case descriptions dating back to the early 1600s probably recorded what we now recognize as amyloidosis, it was Rudolph Virchow in 1854 who adopted the term "amyloid", first introduced by Schleiden in 1838 to describe plant starch, to refer to tissue deposits of material that stained in a similar manner to cellulose when exposed to iodine. In these original descriptions, amyloid deposits were noted by Rokitansky to have a "waxy" or "lardaceous" appearance grossly, and by Virchow to be amorphous and hyaline on light microscopy. Congo red staining (later shown to confer typical apple-green birefringence with polarized microscopy) for the better demonstration of amyloid was introduced in the 1920s by Bennhold, and the use of thioflavine T (producing an intense yellow-green fluorescence) was popularized in the 1950s [1 ](picture 1A- D ). Electron microscopic examination of amyloid deposits, first performed in 1959, generally demonstrates straight and unbranching fibrils 8 to 10 nm in width, which may be composed of protofilaments at higher resolution [2,3 ]. Transmission electron and atomic force microscopy have had a role in elucidating the three dimensional structure of these macromolecular aggregates, and in defining folding intermediates [4 ]. In many instances, the type of amyloid fibril unit can be further defined by immunohistology

An Overview of Amyloidosis

  • Upload
    lisalas

  • View
    95

  • Download
    1

Embed Size (px)

Citation preview

Page 1: An Overview of Amyloidosis

An overview of amyloidosisAuthorPeter D Gorevic, MDSection EditorPeter H Schur, MDDeputy EditorPaul L Romain, MD

Disclosures

Last literature review version 19.3: septiembre 2011 | This topic last updated: octubre 13, 2011

INTRODUCTION — Although case descriptions dating back to the early 1600s probably recorded what we now recognize as amyloidosis, it was Rudolph Virchow in 1854 who adopted the term "amyloid", first introduced by Schleiden in 1838 to describe plant starch, to refer to tissue deposits of material that stained in a similar manner to cellulose when exposed to iodine. In these original descriptions, amyloid deposits were noted by Rokitansky to have a "waxy" or "lardaceous" appearance grossly, and by Virchow to be amorphous and hyaline on light microscopy. Congo red staining (later shown to confer typical apple-green birefringence with polarized microscopy) for the better demonstration of amyloid was introduced in the 1920s by Bennhold, and the use of thioflavine T (producing an intense yellow-green fluorescence) was popularized in the 1950s [1] (picture 1A-D).

Electron microscopic examination of amyloid deposits, first performed in 1959, generally demonstrates straight and unbranching fibrils 8 to 10 nm in width, which may be composed of protofilaments at higher resolution [2,3]. Transmission electron and atomic force microscopy have had a role in elucidating the three dimensional structure of these macromolecular aggregates, and in defining folding intermediates [4]. In many instances, the type of amyloid fibril unit can be further defined by immunohistology (immunofluorescence or immunoenzymatic techniques), or by immunoelectron microscopy [5].

A general overview of the pathogenesis, clinical manifestations, diagnosis, and treatment of the different amyloid disorders is presented here. More complete discussions of the individual disorders are presented elsewhere on the appropriate topic reviews.

PATHOGENESIS — Amyloidosis is a generic term that refers to the extracellular tissue deposition of fibrils composed of low molecular weight subunits (most of which are in the molecular weight range of 5 to 25 kD) of a variety of proteins, many of which circulate as constituents of plasma. At least 27 different human and nine different animal protein precursors of amyloid fibrils are now known. A listing of these molecules, current nomenclature for subunit proteins [6], and corresponding amyloid diseases is presented in the tables (table 1A-B). (See "Genetic factors in the amyloid diseases".)

Page 2: An Overview of Amyloidosis

As is apparent from these tables, several types of amyloidosis are clearly hereditary and, in most familial forms, clinical disease has been linked to missense mutations of the precursor proteins; in some instances, deletions or premature stop codon mutations have also been described [7,8]. Virtually all of the heredofamilial amyloidoses associated with nephropathic, neuropathic, or cardiopathic disease are dominantly inherited heterozygous disorders, and both the wild-type and mutant molecules can be identified in the amyloid deposits. In some instances (eg, transthyretin (TTR), apolipoprotein A-I (ApoAI); Alzheimer amyloid precursor protein (APP), prion protein (PRP)), both the wild-type and mutant molecules may separately form amyloid fibrils under different circumstances (eg, wild type TTR, ApoA1 and Ab deposits in association with organ-specific aging pathology in the heart, aorta, and brain, respectively) [8-10]. (See "Genetic factors in the amyloid diseases".)

Amyloid fibrils are insoluble polymers comprised of low molecular weight subunit proteins. These subunits are derived in turn from soluble precursors which undergo conformational changes that lead to the adoption of a predominantly antiparallel beta-pleated sheet configuration [7,11]. Oligomeric intermediates that are nonfibrillar may be responsible for tissue toxicity and disease pathogenesis in certain amyloid related disorders [12].

Routes to fibrillogenesis include partial folding or unfolding of the precursor protein that may be facilitated by acidification or proteolysis, and accelerated by nucleation [13]. Fibril formation is also associated with codeposition of other nonfibrillar substances, notably including glycosaminoglycans (GAGs, particularly heparan sulfate), serum amyloid P-component (SAP, a member of the pentraxin family that includes C-reactive protein), and specific apolipoproteins (E and J) [14]. Cofactors may significantly modulate fibrillogenesis at any of several steps involved in the conversion of soluble precursors to fibrils, and potentially may influence the deposition phase of amyloid in tissue, as well as resorption [15].

Genetic factors may be involved in one or more ways in predisposing to the development of fibrillogenesis and amyloidosis. (See "Genetic factors in the amyloid diseases".) Heritable effects include the following:

Mutant genes encoding abnormal proteins that are more prone to fibrillogenesis than their wild-type counterparts

Polymorphisms of cofactors (eg apolipoprotein E) or of subunit proteins (eg, serum amyloid A)

Heritable disorders that affect the level or accumulation of precursor proteins (eg presenilin mutations in familial Alzheimer disease)

Heritable disorders that may result in chronic inflammation and deposition of precursor protein in susceptible populations (eg, pyrin and cryopyrin mutations in Familial Mediterranean Fever (FMF) and Muckle Wells syndrome, tumor necrosis factor (TNF) receptor mutations in the TNF receptor associated periodic syndrome

Page 3: An Overview of Amyloidosis

(TRAPS), respectively, and serum amyloid A (SAA) polymorphisms in other forms of Secondary (AA) amyloidosis).

TYPES OF AMYLOIDOSIS — As mentioned above, the most frequent types of amyloidosis seen on internal medicine services and tertiary referral centers are the AL (primary) and AA (secondary) types. What follows is a brief review of the major types of amyloidosis with links to more complete discussions elsewhere in UpToDate.

AL amyloidosis — AL amyloid is due to deposition of protein derived from immunoglobulin light chain fragments. It is a plasma cell dyscrasia in which a monoclonal immunoglobulin is detectable in the serum or monoclonal light chains in the urine in approximately 80 percent of cases.

AL amyloid can occur alone or in association with multiple myeloma or, much less often, Waldenström’s macroglobulinemia or non-Hodgkin lymphoma. Light chain deposition disease has a similar pathogenesis and shares some clinical manifestations with AL amyloidosis; the primary difference is that deposited light chain fragments generally do not form fibrils and do not engender deposition of amyloid cofactors. (See "Pathogenesis of immunoglobulin light chain (AL) amyloidosis and light and heavy chain deposition diseases" and "Epidemiology, pathogenesis, clinical manifestations and diagnosis of Waldenstrom macroglobulinemia", section on 'Amyloidosis' and "Clinical features, laboratory manifestations, and diagnosis of multiple myeloma", section on 'AL amyloidosis'.)

AA amyloidosis — AA amyloidosis may complicate chronic diseases in which there is ongoing or recurring inflammation, such as rheumatoid arthritis (RA), spondyloarthropathy, or inflammatory bowel disease; chronic infections; or periodic fever syndromes [16]. The fibrils are composed of fragments of the acute phase reactant serum amyloid A. (See "Causes and diagnosis of secondary (AA) amyloidosis and relation to rheumatic diseases" and "Pathogenesis of secondary (AA) amyloidosis".)

Dialysis-related amyloidosis — Dialysis-related amyloidosis is due to deposition of fibrils derived from beta-2 microglobulin, which accumulates in patients with end-stage renal disease who are being maintained for prolonged periods of time by dialysis. This disorder has a predilection for osteoarticular structures. (See "Dialysis-related amyloidosis".)

Heritable amyloidoses — Many mutations lead to heritable types of amyloidosis and the hallmark clinical features are outlined in the tables (table 1A-B). An example of this heterogeneous group of disorders is heritable neuropathic and/or cardiomyopathic amyloidosis due to deposition of fibrils derived from transthyretin (also referred to as prealbumin). (See "Genetic factors in the amyloid diseases".)

Age-related (senile) systemic amyloidosis — Deposition of otherwise normal (wild-type) transthyretin in myocardium and other sites has been referred is referred to as Systemic Senile Amyloidosis (SSA) [17]. Compared with patients with AL amyloidosis, those with the senile systemic disease survive longer (75 versus 11 months) despite having ventricular free wall and septal thickening due to amyloid deposits [18]. Significant renal involvement

Page 4: An Overview of Amyloidosis

is rare in the senile systemic disorder; carpal tunnel syndrome may be seen. In one study, all 18 affected patients were elderly men. There may be considerable overlap clinically between senile cardiac amyloidosis due to deposition of wild-type TTR, and late-onset cardiomyopathy due to mutant TTR. A family history may not be apparent, and a screen for informative mutations (notably Ile122, particularly in African Americans) may be necessary to distinguish the two causes of restrictive cardiomyopathy in the elderly. (See 'Organ-specific amyloid' below and "Amyloid cardiomyopathy", section on 'ATTR cardiac amyloidosis'.)

Organ-specific amyloid — Amyloid deposition can be isolated to a single organ, such as the skin, eye, heart, pancreas, or genitourinary tract, resulting in specific syndromes. Forms of primary localized cutaneous amyloidosis include macular, nodular, and lichen amyloidosis, with the last occurring in some families with multiple endocrine neoplasia type 2 [19-21]. (See "Clinical manifestations and diagnosis of multiple endocrine neoplasia type 2".)

In some instances, organ-specific amyloidosis has been shown to be identical to systemic forms of amyloid biochemically. It is presumed that the precursor protein is synthesized and processed at sites contiguous to amyloid deposition [20,22].

The most common clinically important form of organ-specific amyloid occurs in patients with Alzheimer disease in which plaques and amyloid-laden cerebral vessels are composed of the beta protein (Ab), a 39 to 43 residue polypeptide that is cleaved out of the much larger amyloid precursor protein (APP) by secretases that are specific for its amino (beta-secretase) and carboxy- (gamma-secretase) terminal residues.

Isolated bladder amyloidosis can be the cause of life threatening hemorrhage and irritative voiding symptoms; in most instances, this appears to be a localized form immunoglobulin light chain (AL) amyloid disease [23].

CLINICAL MANIFESTATIONS — The type of precursor protein, the tissue distribution, and the amount of amyloid deposition largely determine the clinical manifestations. In the two most common forms of systemic amyloidosis, primary (AL) and secondary (AA amyloidosis), the major sites of clinically important amyloid deposition are in the kidneys, heart, and liver. In some disorders, clinically important amyloid deposition is limited to one organ. (See 'Types of amyloidosis' above and 'Organ-specific amyloid' above.).

Some clinical and laboratory features that suggest amyloidosis include: waxy skin and easy bruising, enlarged muscles (eg, tongue, deltoids), symptoms and signs of heart failure, cardiac conduction abnormalities, hepatomegaly, evidence of heavy proteinuria or the nephrotic syndrome, peripheral and/or autonomic neuropathy, and impaired coagulation.

The prevalence of the most common types of amyloidosis varies in different parts of the world. In developed countries AL is the most common type of systemic amyloidosis, while in developing countries AA amyloid is more frequent. This is a likely result of a higher burden of chronic infectious diseases such as tuberculosis, leprosy, and osteomyelitis in the latter [24].

Page 5: An Overview of Amyloidosis

Renal disease — Renal involvement most often presents as asymptomatic proteinuria or clinically apparent nephrotic syndrome. However, primary deposition can be limited to the blood vessels or tubules; such patients present with renal failure with little or no proteinuria [25]. End-stage renal disease is the cause of death in a minority of patients. (See "Renal amyloidosis".)

Cardiomyopathy — Cardiac involvement can lead to systolic or diastolic dysfunction and the symptoms of heart failure. Other manifestations that can occur include syncope due to arrhythmia or heart block, and angina or infarction due to accumulation of amyloid in the coronary arteries [26]. (See "Amyloid cardiomyopathy".)

Gastrointestinal disease — Hepatomegaly with or without splenomegaly is a common finding in some forms of amyloidosis. Other gastrointestinal manifestations include bleeding (due to vascular fragility and loss of vasomotor responses to injury), gastroparesis, constipation, bacterial overgrowth, malabsorption, and intestinal pseudo-obstruction resulting from dysmotility [27]. (See "Gastrointestinal amyloidosis".)

Neurologic abnormalities — Mixed sensory and motor peripheral neuropathy and/or autonomic neuropathy may occur and is a prominent feature in some of the heritable amyloidoses (called familial amyloidotic polyneuropathy) and in AL amyloidosis. Symptoms of numbness, paresthesia, and pain are frequently noted, as in peripheral neuropathy of many other etiologies. Compression of peripheral nerves, especially the median nerve within the carpal tunnel can cause more localized sensory changes. Symptoms of bowel or bladder dysfunction and findings of orthostatic hypotension may be due to autonomic nervous system damage [8].

Central nervous system involvement is unusual in patients with the more common AL and AA amyloidoses. Amyloid deposits can lead to extensive cortical pathology and dementia in patients with sporadic or familial Alzheimer disease, while cerebral amyloid angiopathy can cause spontaneous cortical and subcortical intracranial bleeding, primarily in the elderly [28]. (See "Genetic factors in the amyloid diseases", section on 'Presenilins and autosomal dominant early-onset AD' and "Cerebral amyloid angiopathy".)

Ischemic stroke may be the initial manifestation of amyloidosis. Risk factors for stroke are similar to those in the general population and include atrial fibrillation, hyperlipidemia, hypertension and diabetes mellitus; echocardiographic evidence of myocardial or valvular involvement is often present and supports a cardioembolic source in the majority those with ischemic stroke and biopsy proven amyloidosis [29]. (See "Amyloid cardiomyopathy".)

Musculoskeletal disease — Amyloid infiltration of muscles may cause visible enlargement (ie, pseudohypertrophy). A large tongue (ie, macroglossia), or lateral scalloping of the tongue from impingement on the teeth, is characteristic of AL amyloid. Arthropathy may be due to amyloid deposition in joints and surrounding structures. The "shoulder pad" sign is visible enlargement of the anterior shoulder due to fluid in the glenohumeral joint and/or amyloid infiltration of the synovial membrane and surrounding structures. (See "Musculoskeletal manifestations of amyloidosis".)

Page 6: An Overview of Amyloidosis

This type of shoulder involvement is characteristic of AL amyloid and dialysis related amyloidosis due to deposition of beta-2 microglobulin. Other musculoskeletal features of dialysis-related amyloidosis include scapulohumeral periarthritis, spondyloarthropathy, bone disease, and carpal tunnel syndrome. (See "Dialysis-related amyloidosis".)

Hematologic abnormalities — Amyloidosis may be directly associated with a bleeding diathesis [30-32]. In one report of 337 patients, abnormal bleeding and abnormal coagulation tests were seen in 28 and 51 percent, respectively [31]. Two major mechanisms have been described: factor X deficiency due to binding on amyloid fibrils primarily in the liver and spleen; and decreased synthesis of coagulation factors in patients with advanced liver disease. However, some patients with abnormal bleeding have no abnormality in any coagulation test [30]. In such patients, amyloid infiltration of blood vessels may contribute to the bleeding diathesis. Bleeding due to acquired von Willebrand disease has been described in AL amyloidosis [33].

Factor X deficiency results from binding of factor X to amyloid fibrils [34-37]. In a series of 368 consecutive patients with AL, 32 and 12 patients had factor X levels below 50 and 25 percent of normal, respectively [35]. Bleeding was noted in 18 patients, which was more severe in the 12 patients whose factor X levels were <25 percent of normal. Factor X levels improved in four of four patients obtaining complete remission and in one of two obtaining partial remission following high-dose melphalan chemotherapy and autologous hematopoietic cell transplantation.

Other hematologic manifestations are related to the degree of organ involvement. These include anemia in patients with renal failure or multiple myeloma, and thrombocytopenia due to splenomegaly.

Pulmonary disease — Pulmonary manifestations of amyloidosis include tracheobronchial infiltration, persistent pleural effusions, parenchymal nodules (amyloidomas), and (rarely) pulmonary hypertension [38-47]. Tracheobronchial infiltration can cause hoarseness, stridor, airway obstruction, and dysphagia; bronchoscopic or surgical resection of airway abnormalities may be required [48-53]. (See "Diagnosis and management of central airway obstruction".)

Persistent pleural effusions develop in 1 to 2 percent of patients with systemic amyloidosis, and appear to be caused by pleural infiltration with amyloid deposits [40,54]. However, it is difficult to distinguish primary effusions from those caused by amyloid-induced cardiomyopathy on the basis of echocardiographic findings alone, and the sensitivity of pleural biopsy in this setting has not been studied extensively [40]. Persistent pleural effusions are associated with a poor prognosis and limited response to treatment, although pleurodesis has been useful in some cases [40] and bevacizumab in others [55]. (See "Diagnostic evaluation of a pleural effusion in adults" and "Management of refractory nonmalignant pleural effusions".)

Skin manifestations — Signs of skin involvement in systemic amyloidosis include waxy thickening, easy bruising (ecchymoses), and subcutaneous nodules or plaques. Purpura, characteristically elicited in a periorbital distribution (raccoon eyes) by a valsalva maneuver

Page 7: An Overview of Amyloidosis

or minor trauma, is present in only a minority of patients, but is highly characteristic of AL amyloidosis (picture 2) [56].

Amyloidosis limited to the skin may also occur. (See 'Organ-specific amyloid' above.)

Infiltration of the subcutaneous fat is generally asymptomatic but provides a convenient site for biopsy.

Other — A variety of other manifestations can occur including visual and hearing loss in some heritable amyloidoses, in addition to gross hematuria and irritative urinary symptoms [22,23]. Ischemic symptoms and tissue infarction due to vascular infiltration have also been reported. Examples include: jaw claudication suggestive of giant cell (temporal) arteritis can occur in AL amyloidosis [57] and symptomatic ischemic coronary heart disease is associated with the presence of obstructive intramural deposits of AL amyloid [58].

DIAGNOSIS

Biopsy — The diagnosis of amyloidosis can be confirmed only by tissue biopsy, although the presence of amyloidosis may be suggested by the history and clinical manifestations (eg, nephrotic syndrome in a patient with multiple myeloma or long-standing, active rheumatoid arthritis). Biopsies may be directed to dysfunctional organs (eg. kidney, nerve) or to clinically uninvolved sites such as subcutaneous fat, minor salivary glands, or rectal mucosa.

We suggest a fat pad aspiration as the initial biopsy technique for patients with other than single organ involvement, because fat pad aspiration biopsy is less likely than liver, renal, or even rectal biopsy to be complicated by serious bleeding. Aspiration biopsy of subcutaneous fat with Congo red staining and examination using polarizing microscopy has a sensitivity of 57 to 85 percent and a specificity of 92 to 100 percent for AL or AA amyloidosis [59-62]. Fat pad aspiration biopsy has a low sensitivity for amyloidosis in patients with a single involved organ. The diagnostic sensitivity is higher in those with multiorgan involvement who are suspected to have systemic amyloidosis due to immunoglobulin light chain (primary, or AL), AA protein (secondary), or transthyretin (senile cardiac or familial amyloid polyneuropathy) deposition.

Biopsy of a specifically involved site is suggested for patients with a limited number of affected organs, because patients with single organ involvement are less likely to have biopsies of unaffected tissues reveal amyloid. This was illustrated by a study of 450 patients with peripheral neuropathy who had fat pad aspiration biopsies performed; among the 143 who had only peripheral neuropathy, none had amyloid deposits noted in the aspiration biopsy [63]. In contrast, among the 307 patients with at least one other clinical feature of amyloidosis, the rate of positive aspirations was 6 percent.

The sensitivity of rectal biopsy in one large series composed predominantly of patients with systemic amyloid ("primary amyloid" and myeloma associated in 193 patients, "secondary" or localized to a single organ in 41 patients) was 84 percent [64]. The sensitivity of kidney, liver, and carpal-tunnel biopsies were all 90 percent or more in this cohort.

Page 8: An Overview of Amyloidosis

Some patients have a predisposing underlying disorder before the presentation with amyloid, while others present with no such history but an amyloid manifestation such as otherwise unexplained heart failure or nephrotic syndrome. In such patients, the diagnosis of amyloidosis comes first, followed by an evaluation for a cause (eg, plasma cell dyscrasia for AL; chronic inflammatory disease for AA).

Tissue biopsy is important because assumptions regarding the type of amyloid may be incorrect. As an example, AA amyloidosis is only one cause of the nephrotic syndrome in patients with rheumatoid arthritis; other causes include drug side-effects, immune-complex disease, or an unrelated disorder. (See "Renal disease in patients with rheumatoid arthritis".)

Some noninvasive tests can provide supportive but not definitive findings. Examples include a speckled appearance of the myocardium on echocardiography, lytic bone lesions in multiple myeloma, and cystic bone lesions in dialysis-related amyloidosis.

Although liver biopsy is rarely complicated by life-threatening bleeding, less invasive approaches to diagnosis than percutaneous liver biopsy are generally preferred. (See "Percutaneous liver biopsy", section on 'Amyloid'.)

The amyloid deposits appear as amorphous hyaline material on light microscopy (picture 1A-D). The fibrils bind Congo red (leading to green birefringence under polarized light) and thioflavine T (producing an intense yellow-green fluorescence) and, on electron microscopy, are 8 to 10 nm in width and straight and unbranching [1,2].

In some cases, immunohistology can be used to identify the type of protein subunit [65]. This is most useful for AA and TTR amyloid, less so for AL amyloid; variable staining of AL deposits with standard antisera to kappa or lambda constant region-determinants is due to loss of antigenic epitopes in the course of proteolytic processing of the constant region that precedes fibrillogenesis. Variable region-specific antibodies to immunoglobulin light chain subclass determinants may provide an approach to circumventing this limitation [66]. (See "Clinical presentation, laboratory manifestations, and diagnosis of immunoglobulin light chain (AL) amyloidosis (primary amyloidosis)", section on 'Differential diagnosis'.)

Direct determination of the proteins present in the amyloid deposits, either by amino acid sequencing or mass spectroscopy, is the most straightforward approach to characterizing the type of amyloid present in the biopsy specimen. These direct methods of protein analysis are applicable to formalin-fixed paraffin-embedded biopsy specimens. However, these analyses are available only at specialized centers. The validation of this methodology for commercial purposes has been of particular interest. This method is performed using laser capture microdissection of amyloid from formalin-fixed amyloidotic tissue, followed by trypsin digestion, and mass spectroscopy and direct sequence analysis of peptides [67]. This methodology, which was originally used for the analysis of the proteomics of AL amyloid, has also been validated for AA, and specific tissues, including nerve in patients with amyloid neuropathy, and abdominal fat pad aspirates [68].

Page 9: An Overview of Amyloidosis

SAP scanning — Scintigraphy with radioisotope labeled serum amyloid P component (SAP) can identify the distribution of amyloid, and provide an estimate of the total body burden of fibrillar deposits [69]. Sensitivity of SAP scanning of 90 percent for AA and AL amyloid is contrasted with 48 percent for hereditary ATTR amyloidosis; the specificity is 93 percent in all three conditions [70]. However, the value of SAP scintigraphy is limited because it is currently obtained from blood donors (thereby carrying a potential infectious risk), and is less helpful in detecting cardiac amyloid. (See "Clinical presentation, laboratory manifestations, and diagnosis of immunoglobulin light chain (AL) amyloidosis (primary amyloidosis)", section on 'Serum amyloid P component scintigraphy'.)

Paraproteins — AL amyloidosis should be suspected in all patients with amyloid deposits who do not have a chronic infectious or inflammatory disease, end-stage renal disease, or a family history of amyloidosis. Patients who have biopsy-documented amyloidosis and meet criteria for a well-defined plasma cell dyscrasia, such as multiple myeloma or Waldenström’s macroglobulinemia, need not undergo further diagnostic testing. (See "Clinical features, laboratory manifestations, and diagnosis of multiple myeloma" and "Epidemiology, pathogenesis, clinical manifestations and diagnosis of Waldenstrom macroglobulinemia" and "Clinical presentation, laboratory manifestations, and diagnosis of immunoglobulin light chain (AL) amyloidosis (primary amyloidosis)", section on 'Evidence of monoclonal plasma cells'.)

In patients without a history of plasma cell dyscrasia, initial testing is aimed at determining whether a monoclonal population of plasma cells is present. This is usually accomplished by testing for a monoclonal protein by serum and urine protein electrophoresis (figure 1 and figure 2) and immunofixation (figure 3 and figure 4). Quantitation of serum free light chains has been utilized as an adjunctive diagnostic modality that may demonstrate clonality in patients with AL amyloid who do not have monoclonal proteins by immunofixation; this assay may also be used to follow response to treatment [71]. (See "Recognition of monoclonal proteins".)

Bone marrow biopsy can confirm the diagnosis of plasma cell dyscrasia by demonstrating a monoclonal population of plasma cells (picture 3). Biopsies should be specifically stained for amyloid, and immunohistology and flow cytometry performed with antibodies to kappa and lambda light chains to look for clonality of plasma cells. (See "Clinical features, laboratory manifestations, and diagnosis of multiple myeloma", section on 'Bone marrow examination'.)

However, the presence of a monoclonal protein alone is not sufficient to make a diagnosis of AL amyloid in a patient with documented amyloidosis unless immunofluorescence has demonstrated monoclonal light chains in the amyloid deposits [72,73]. The potential for misdiagnosis based upon a serum or urine monoclonal protein alone was illustrated in a study of 350 patients suspected of having AL amyloidosis by clinical and laboratory findings and the absence of a family history [73]. Ten percent of these patients had a mutant gene for an "amyloidogenic" protein, most often involving the alpha-chain of fibrinogen A or transthyretin. In 8 of these 34 patients, the presence of low concentrations of monoclonal immunoglobulins (all less than 0.2 g/dL) contributed to the misdiagnosis. (See "Genetic factors in the amyloid diseases".)

Page 10: An Overview of Amyloidosis

A thorough family history and exclusion of these heritable disorders is warranted if a plasma cell dyscrasia cannot be documented [73,74]. Treatment programs utilized for AL amyloidosis (eg, chemotherapy, hematopoietic cell transplantation) have no role in patients with hereditary forms of amyloidosis.

TREATMENT — Treatment of the different types of amyloidosis varies with the cause of fibril production. As examples, therapy is aimed at the underlying infectious or inflammatory disorder in AA amyloidosis, the underlying plasma cell dyscrasia in AL amyloidosis, and altering the mode of dialysis or considering renal transplantation in patients with dialysis-related amyloidosis. (See "Treatment of secondary (AA) amyloidosis" and "Prognosis and treatment of immunoglobulin light chain (AL) amyloidosis and light and heavy chain deposition diseases" and "Dialysis-related amyloidosis".)

For the hereditary amyloidoses in which the mutant amyloid precursor protein is produced by the liver (eg: transthyretin, apolipoprotein A-I, and fibrinogen Aa), liver transplantation may in some instances prevent further deposition of amyloid and lead to regression of established deposits [75-77]. Transplantation during the first year after appearance of symptoms is ideal. Patients with sporadic or undiagnosed hereditary amyloidosis who present with advanced end-organ damage may benefit from combined hepatorenal or hepatocardiac transplantation [76-78]. (See "Amyloid cardiomyopathy", section on 'Treatment'.)

Novel approaches to treatment are being investigated. Agents that interfere with fibril formation, prevent the formation of amyloidogenic precursors, hasten degradation of existing amyloid deposits, or disrupt the interaction between amyloidogenic proteins and accessory molecules such as SAP and heparitin sulfate proteoglycans have been shown to be effective in animal models, and some have progressed to human trials [79,80].

SUMMARY AND RECOMMENDATIONS

Amyloidosis is a generic term that refers to the extracellular tissue deposition of fibrils composed of low molecular weight subunits of a variety of proteins, many of which circulate as constituents of plasma. These subunit proteins are derived in turn from soluble precursors which undergo conformational changes that lead to the adoption of a predominantly antiparallel beta-pleated sheet configuration. At least 27 different human protein precursors of amyloid fibrils are now known (table 1A-B). (See 'Pathogenesis' above.)

The two major forms of amyloidosis are the AL (primary) and AA (secondary) types. AL amyloid, the most common form in developed countries, is due to deposition of protein derived from immunoglobulin light chain fragments. It is a plasma cell dyscrasia in which a monoclonal immunoglobulin is detectable in the serum or monoclonal light chains in the urine in approximately 80 percent of cases. (See 'AL amyloidosis' above and 'Clinical manifestations' above.)

AA amyloidosis, the most common form in developing countries, may complicate chronic diseases in which there is ongoing or recurring inflammation, such as

Page 11: An Overview of Amyloidosis

chronic infections; rheumatoid arthritis (RA), spondyloarthropathy, or inflammatory bowel disease; or periodic fever syndromes. (See 'AA amyloidosis' above and 'Clinical manifestations' above.)

Other major forms seen clinically include dialysis-related amyloidosis, heritable amyloidoses, age-related systemic amyloidosis and organ-specific amyloid. (See 'Dialysis-related amyloidosis' above and 'Heritable amyloidoses' above and 'Age-related (senile) systemic amyloidosis' above and 'Organ-specific amyloid' above and 'Clinical manifestations' above.)

Clinical manifestations vary depending upon the type of amyloid and the distribution of deposition. Some clinical and laboratory features that suggest amyloidosis include: waxy skin and easy bruising, enlarged muscles (eg, tongue, deltoids), symptoms and signs of heart failure, cardiac conduction abnormalities, hepatomegaly, evidence of heavy proteinuria or the nephrotic syndrome, peripheral and/or autonomic neuropathy, and impaired coagulation. (See 'Types of amyloidosis' above and 'Clinical manifestations' above.).

Tissue biopsy should be used to confirm the diagnosis in all cases of amyloidosis, although the diagnosis of amyloidosis may be suspected on the basis of history and clinical manifestations. Fat pad aspiration biopsy is less likely than liver, renal, or even rectal biopsy to be complicated by serious bleeding; we thus suggest it as the initial biopsy technique for patients with other than single organ involvement. In patients with single organ involvement biopsy of the clinically involved site is suggested because fat pad aspiration biopsy has a low sensitivity for amyloidosis in such patients. (See 'Biopsy' above.).

Patients who have biopsy-documented amyloidosis and meet criteria for a well-defined plasma cell dyscrasia, such as multiple myeloma or Waldenström’s macroglobulinemia, need not undergo further diagnostic testing. Patients who are not known to have a plasma cell disorder should be further tested to determine whether a monoclonal protein is present in serum, urine, or both. A combination of serum and urine protein electrophoresis, followed by immunofixation, is recommended. Quantitation of serum free light chains is suggested for AL patients who do not have monoclonal proteins by immunofixation. (See 'Paraproteins' above.)

The presence of a monoclonal protein alone is not sufficient to make a diagnosis of AL amyloid in a patient with documented amyloidosis unless immunofluorescence has demonstrated monoclonal light chains in the amyloid deposits. Heritable types of amyloidosis should be excluded if a plasma cell dyscrasia (multiple myeloma or Waldenström’s macroglobulinemia) cannot be documented. (See 'Paraproteins' above.)

Treatment of the different types of amyloidosis varies with the cause of fibril production. As examples, therapy is aimed at the underlying infectious or inflammatory disorder in AA amyloidosis, the underlying plasma cell dyscrasia in AL amyloidosis, and altering the mode of dialysis or considering renal

Page 12: An Overview of Amyloidosis

transplantation in patients with dialysis-related amyloidosis. Liver transplantation may be effective in certain of the hereditary amyloidoses. (See 'Treatment' above.)

Use of UpToDate is subject to the Subscription and License Agreement.

REFERENCES1. Kyle RA. Amyloidosis: a convoluted story. Br J Haematol 2001; 114:529. 2. COHEN AS, CALKINS E. Electron microscopic observations on a fibrous

component in amyloid of diverse origins. Nature 1959; 183:1202.

3. Stromer T, Serpell LC. Structure and morphology of the Alzheimer's amyloid fibril. Microsc Res Tech 2005; 67:210.

4. Gosal WS, Myers SL, Radford SE, Thomson NH. Amyloid under the atomic force microscope. Protein Pept Lett 2006; 13:261.

5. Picken MM. Amyloidosis-where are we now and where are we heading? Arch Pathol Lab Med 2010; 134:545.

6. Sipe JD, Benson MD, Buxbaum JN, et al. Amyloid fibril protein nomenclature: 2010 recommendations from the nomenclature committee of the International Society of Amyloidosis. Amyloid 2010; 17:101.

7. Bellotti V, Nuvolone M, Giorgetti S, et al. The workings of the amyloid diseases. Ann Med 2007; 39:200.

8. Benson MD, Kincaid JC. The molecular biology and clinical features of amyloid neuropathy. Muscle Nerve 2007; 36:411.

9. Mucchiano GI, Häggqvist B, Sletten K, Westermark P. Apolipoprotein A-1-derived amyloid in atherosclerotic plaques of the human aorta. J Pathol 2001; 193:270.

10. Yazaki M, Liepnieks JJ, Kincaid JC, Benson MD. Contribution of wild-type transthyretin to hereditary peripheral nerve amyloid. Muscle Nerve 2003; 28:438.

11. Jahn TR, Radford SE. Folding versus aggregation: polypeptide conformations on competing pathways. Arch Biochem Biophys 2008; 469:100.

12. Lindgren M, Hammarström P. Amyloid oligomers: spectroscopic characterization of amyloidogenic protein states. FEBS J 2010; 277:1380.

13. Rochet JC, Lansbury PT Jr. Amyloid fibrillogenesis: themes and variations. Curr Opin Struct Biol 2000; 10:60.

14. Kisilevsky R. The relation of proteoglycans, serum amyloid P and apo E to amyloidosis current status, 2000. Amyloid 2000; 7:23.

Page 13: An Overview of Amyloidosis

15. Dergunov AD. Role of ApoE in conformation-prone diseases and atherosclerosis. Biochemistry (Mosc) 2006; 71:707.

16. Lachmann HJ, Goodman HJ, Gilbertson JA, et al. Natural history and outcome in systemic AA amyloidosis. N Engl J Med 2007; 356:2361.

17. Westermark P, Bergström J, Solomon A, et al. Transthyretin-derived senile systemic amyloidosis: clinicopathologic and structural considerations. Amyloid 2003; 10 Suppl 1:48.

18. Ng B, Connors LH, Davidoff R, et al. Senile systemic amyloidosis presenting with heart failure: a comparison with light chain-associated amyloidosis. Arch Intern Med 2005; 165:1425.

19. Tanaka A, Arita K, Lai-Cheong JE, et al. New insight into mechanisms of pruritus from molecular studies on familial primary localized cutaneous amyloidosis. Br J Dermatol 2009; 161:1217.

20. Meijer JM, Schonland SO, Palladini G, et al. Sjögren's syndrome and localized nodular cutaneous amyloidosis: coincidence or a distinct clinical entity? Arthritis Rheum 2008; 58:1992.

21. Baykal C, Buyukbabani N, Boztepe H, et al. Multiple cutaneous neuromas and macular amyloidosis associated with medullary thyroid carcinoma. J Am Acad Dermatol 2007; 56:S33.

22. Hamidi Asl K, Liepnieks JJ, Nakamura M, Benson MD. Organ-specific (localized) synthesis of Ig light chain amyloid. J Immunol 1999; 162:5556.

23. Monge M, Chauveau D, Cordonnier C, et al. Localized amyloidosis of the genitourinary tract: report of 5 new cases and review of the literature. Medicine (Baltimore) 2011; 90:212.

24. Westermark P, Westermark GT. Review. Reflections on amyloidosis in Papua New Guinea. Philos Trans R Soc Lond B Biol Sci 2008; 363:3701.

25. Nishi S, Alchi B, Imai N, Gejyo F. New advances in renal amyloidosis. Clin Exp Nephrol 2008; 12:93.

26. Dubrey SW, Hawkins PN, Falk RH. Amyloid diseases of the heart: assessment, diagnosis, and referral. Heart 2011; 97:75.

27. Sattianayagam P, Hawkins P, Gillmore J. Amyloid and the GI tract. Expert Rev Gastroenterol Hepatol 2009; 3:615.

28. Revesz T, Holton JL, Lashley T, et al. Genetics and molecular pathogenesis of sporadic and hereditary cerebral amyloid angiopathies. Acta Neuropathol 2009; 118:115.

Page 14: An Overview of Amyloidosis

29. Zubkov AY, Rabinstein AA, Dispenzieri A, Wijdicks EF. Primary systemic amyloidosis with ischemic stroke as a presenting complication. Neurology 2007; 69:1136.

30. Yood RA, Skinner M, Rubinow A, et al. Bleeding manifestations in 100 patients with amyloidosis. JAMA 1983; 249:1322.

31. Mumford AD, O'Donnell J, Gillmore JD, et al. Bleeding symptoms and coagulation abnormalities in 337 patients with AL-amyloidosis. Br J Haematol 2000; 110:454.

32. Sucker C, Hetzel GR, Grabensee B, et al. Amyloidosis and bleeding: pathophysiology, diagnosis, and therapy. Am J Kidney Dis 2006; 47:947.

33. Kos CA, Ward JE, Malek K, et al. Association of acquired von Willebrand syndrome with AL amyloidosis. Am J Hematol 2007; 82:363.

34. Furie B, Greene E, Furie BC. Syndrome of acquired factor X deficiency and systemic amyloidosis in vivo studies of the metabolic fate of factor X. N Engl J Med 1977; 297:81.

35. Choufani EB, Sanchorawala V, Ernst T, et al. Acquired factor X deficiency in patients with amyloid light-chain amyloidosis: incidence, bleeding manifestations, and response to high-dose chemotherapy. Blood 2001; 97:1885.

36. Boggio L, Green D. Recombinant human factor VIIa in the management of amyloid-associated factor X deficiency. Br J Haematol 2001; 112:1074.

37. Thompson CA, Kyle R, Gertz M, et al. Systemic AL amyloidosis with acquired factor X deficiency: A study of perioperative bleeding risk and treatment outcomes in 60 patients. Am J Hematol 2010; 85:171.

38. Ross P Jr, Magro CM. Clonal light chain restricted primary intrapulmonary nodular amyloidosis. Ann Thorac Surg 2005; 80:344.

39. Howard ME, Ireton J, Daniels F, et al. Pulmonary presentations of amyloidosis. Respirology 2001; 6:61.

40. Berk JL, Keane J, Seldin DC, et al. Persistent pleural effusions in primary systemic amyloidosis: etiology and prognosis. Chest 2003; 124:969.

41. Maeno T, Sando Y, Tsukagoshi M, et al. Pleural amyloidosis in a patient with intractable pleural effusion and multiple myeloma. Respirology 2000; 5:79.

42. Ikeda S, Takabayashi Y, Maejima Y, et al. Nodular lung disease with five year survival and unilateral pleural effusion in AL amyloidosis. Amyloid 1999; 6:292.

43. Utz JP, Swensen SJ, Gertz MA. Pulmonary amyloidosis. The Mayo Clinic experience from 1980 to 1993. Ann Intern Med 1996; 124:407.

44. Dingli D, Utz JP, Gertz MA. Pulmonary hypertension in patients with amyloidosis. Chest 2001; 120:1735.

Page 15: An Overview of Amyloidosis

45. Onozato ML, Tojo A, Ogura S, et al. Primary amyloidosis with multiple pulmonary nodular lesions and IgA nephropathy-like renal involvement. Clin Nephrol 2003; 60:134.

46. Fujimoto N, Masuoka H, Kosaka H, et al. Primary amyloidosis with pulmonary involvement which presented exudative pleural effusion and high fever. Intern Med 2003; 42:756.

47. BoydKing A, Sharma O, Stevenson K. Localized interstitial pulmonary amyloid: a case report and review of the literature. Curr Opin Pulm Med 2009; 15:517.

48. Dahl KA, Kernstine KH, Vannatta TL, et al. Tracheobronchial amyloidosis: a surgical disease with long-term consequences. J Thorac Cardiovasc Surg 2004; 128:789.

49. O'Regan A, Fenlon HM, Beamis JF Jr, et al. Tracheobronchial amyloidosis. The Boston University experience from 1984 to 1999. Medicine (Baltimore) 2000; 79:69.

50. Capizzi SA, Betancourt E, Prakash UB. Tracheobronchial amyloidosis. Mayo Clin Proc 2000; 75:1148.

51. Pribitkin E, Friedman O, O'Hara B, et al. Amyloidosis of the upper aerodigestive tract. Laryngoscope 2003; 113:2095.

52. Monroe AT, Walia R, Zlotecki RA, Jantz MA. Tracheobronchial amyloidosis: a case report of successful treatment with external beam radiation therapy. Chest 2004; 125:784.

53. Dedo HH, Izdebski K. Laryngeal amyloidosis in 10 patients. Laryngoscope 2004; 114:1742.

54. Berk JL. Pleural effusions in systemic amyloidosis. Curr Opin Pulm Med 2005; 11:324.

55. Hoyer RJ, Leung N, Witzig TE, Lacy MQ. Treatment of diuretic refractory pleural effusions with bevacizumab in four patients with primary systemic amyloidosis. Am J Hematol 2007; 82:409.

56. Eder L, Bitterman H. Image in clinical medicine. Amyloid purpura. N Engl J Med 2007; 356:2406.

57. Gertz MA, Kyle RA, Griffing WL, Hunder GG. Jaw claudication in primary systemic amyloidosis. Medicine (Baltimore) 1986; 65:173.

58. Neben-Wittich MA, Wittich CM, Mueller PS, et al. Obstructive intramural coronary amyloidosis and myocardial ischemia are common in primary amyloidosis. Am J Med 2005; 118:1287.

Page 16: An Overview of Amyloidosis

59. Duston MA, Skinner M, Meenan RF, Cohen AS. Sensitivity, specificity, and predictive value of abdominal fat aspiration for the diagnosis of amyloidosis. Arthritis Rheum 1989; 32:82.

60. Westermark P, Davey E, Lindbom K, Enqvist S. Subcutaneous fat tissue for diagnosis and studies of systemic amyloidosis. Acta Histochem 2006; 108:209.

61. van Gameren II, Hazenberg BP, Bijzet J, van Rijswijk MH. Diagnostic accuracy of subcutaneous abdominal fat tissue aspiration for detecting systemic amyloidosis and its utility in clinical practice. Arthritis Rheum 2006; 54:2015.

62. Dhingra S, Krishnani N, Kumari N, Pandey R. Evaluation of abdominal fat pad aspiration cytology and grading for detection in systemic amyloidosis. Acta Cytol 2007; 51:860.

63. Andrews TR, Colon-Otero G, Calamia KT, et al. Utility of subcutaneous fat aspiration for diagnosing amyloidosis in patients with isolated peripheral neuropathy. Mayo Clin Proc 2002; 77:1287.

64. Kyle RA, Bayrd ED. Amyloidosis: review of 236 cases. Medicine (Baltimore) 1975; 54:271.

65. Arbustini, E, Morbini, P, Verga, L, Merlini, G. Light and electron microscopy immunohistochemical characterization of amyloid deposits. Amyloid 1997; 4:157.

66. Davern S, Tang LX, Williams TK, et al. Immunodiagnostic capabilities of anti-free immunoglobulin light chain monoclonal antibodies. Am J Clin Pathol 2008; 130:702.

67. Vrana JA, Gamez JD, Madden BJ, et al. Classification of amyloidosis by laser microdissection and mass spectrometry-based proteomic analysis in clinical biopsy specimens. Blood 2009; 114:4957.

68. Klein CJ, Vrana JA, Theis JD, et al. Mass spectrometric-based proteomic analysis of amyloid neuropathy type in nerve tissue. Arch Neurol 2011; 68:195.

69. Hawkins PN. Serum amyloid P component scintigraphy for diagnosis and monitoring amyloidosis. Curr Opin Nephrol Hypertens 2002; 11:649.

70. Hazenberg BP, van Rijswijk MH, Piers DA, et al. Diagnostic performance of 123I- labeled serum amyloid P component scintigraphy in patients with amyloidosis. Am J Med 2006; 119:355.e15.

71. Kumar S, Dispenzieri A, Katzmann JA, et al. Serum immunoglobulin free light- chain measurement in primary amyloidosis: prognostic value and correlations with clinical features. Blood 2010; 116:5126.

72. Comenzo RL, Zhou P, Fleisher M, et al. Seeking confidence in the diagnosis of systemic AL (Ig light-chain) amyloidosis: patients can have both monoclonal gammopathies and hereditary amyloid proteins. Blood 2006; 107:3489.

Page 17: An Overview of Amyloidosis

73. Lachmann HJ, Booth DR, Booth SE, et al. Misdiagnosis of hereditary amyloidosis as AL (primary) amyloidosis. N Engl J Med 2002; 346:1786.

74. Saraiva MJ. Sporadic cases of hereditary systemic amyloidosis. N Engl J Med 2002; 346:1818.

75. Herlenius G, Wilczek HE, Larsson M, et al. Ten years of international experience with liver transplantation for familial amyloidotic polyneuropathy: results from the Familial Amyloidotic Polyneuropathy World Transplant Registry. Transplantation 2004; 77:64.

76. Yamamoto S, Wilczek HE, Nowak G, et al. Liver transplantation for familial amyloidotic polyneuropathy (FAP): a single-center experience over 16 years. Am J Transplant 2007; 7:2597.

77. Tsuchiya A, Yazaki M, Kametani F, et al. Marked regression of abdominal fat amyloid in patients with familial amyloid polyneuropathy during long-term follow-up after liver transplantation. Liver Transpl 2008; 14:563.

78. Raichlin E, Daly RC, Rosen CB, et al. Combined heart and liver transplantation: a single-center experience. Transplantation 2009; 88:219.

79. Kolstoe SE, Wood SP. Drug targets for amyloidosis. Biochem Soc Trans 2010; 38:466.

80. Merlini G, Seldin DC, Gertz MA. Amyloidosis: pathogenesis and new therapeutic options. J Clin Oncol 2011; 29:1924.

GRAPHICSGlomerular amyloidosis

Light micrograph of glomerular amyloidosis shows nodular, amorphous material (arrows) extending from the mesangium into the capillary loops and narrowing or closing the capillary lumens. The nodules are more amorphous than those seen in diabetic nephropathy

Page 18: An Overview of Amyloidosis

but demonstration of amyloid fibrils on electron microscopy or demonstrating green birefringence on Congo red staining is required to confirm the diagnosis. Courtesy of Helmut Rennke, MD. Normal glomerulus

Light micrograph of a normal glomerulus. There are only 1 or 2 cells per capillary tuft, the capillary lumens are open, the thickness of the glomerular capillary wall (long arrow) is similar to that of the tubular basement membranes (short arrow), and the mesangial cells and mesangial matrix are located in the central or stalk regions of the tuft (arrows). Courtesy of Helmut G Rennke. Congo red stain in amyloidosis

Congo red stain viewed under polarized light of a renal biopsy from a patient with renal amyloidosis. Green birefringence (white arrows) of interstitial amyloid deposits can be seen. Courtesy of Helmut Rennke, MD. Renal amyloidosis

Page 19: An Overview of Amyloidosis

Electron micrograph showing expansion of the mesangium by amyoid fibrils measuring between 9 and 11 nanometers in diameter. The fibrillar appearance is best appreciated at the arrow. These fibrils are smaller than those seen in fibrillary glomerulonephritis. Courtesy of Helmut Rennke, MD. Normal glomerulus

Electron micrograph of a normal glomerular capillary loop showing the fenestrated endothelial cell (Endo), the glomerular basement membrane (GBM), and the epithelial cells with its interdigitating foot processes (arrow). The GBM is thin and no electron dense deposits are present. Two normal platelets are seen in the capillary lumen. Courtesy of Helmut Rennke, MD. Secondary (AA) amyloidosis

Page 20: An Overview of Amyloidosis

Immunofluorescence microscopy in secondary amyloidosis using an anti-AA antiserum. Markedly positive staining is present in the glomeruli. Although not shown, there is also substantial amyloid deposition in the tubules. Courtesy of Helmut Rennke, MD. Protein precursor of amyloid deposits-I

Protein classPrecursor protein (abbreviation)

Amyloid type

Clinical type

High density apolipoproteins

(Apo) serum AA AA

Associated with amyloid complicating chronic infections or inflammatory diseases, and some heredofamilial periodic fever syndromes, such as FMF

Apolipoprotein A-I (ApoAI)

AApoAI

Age-related amyloid occurring in the aortic intima, and some hereditary neuropathic or cardiopathic amyloidoses [1,2]

Apolipoprotein A-II (ApoAII)

AApoAIISome hereditary nephropathic amyloidoses [3]

Immunoglobulin (Ig) gene superfamily

Ig L chain/Ig H chains (IgL/IgH)

AL/AHPrimary and myeloma-associated amyloidosis

Beta-2 microglobulin

Aβ2m Dialysis amyloidosis

Neuroendocrine

(Pro)Calcitonin ACal Amyloid complicating C-cell thyroid tumors

Islet amyloid AIAPPIslet cell amyloid in Insulinomas, type II diabetes mellitus, and aging [4]

Atrial natriuretic peptide

AANF Isolated atrial amyloidosis of aging

Prolactin/Apro APro Prolactinomas/aging

Insulin AlnsLocal amyloid complicating use of the insulin pump

Cytoskeleton- related

Gelsolin AGel Hereditary neuropathic amyloid associated with corneal lattice dystrophy

Page 21: An Overview of Amyloidosis

and cutis laxa (Meretoja syndrome) [5]Keratin * Cutaneous amyloid

Keratoepithelin AKerHereditary granular, lattice, and avellino corneal dystrophies [6]

* Does not yet have nomenclature designated 1. Mucchiano, GI, Haggqvist, B, Sletten, K, Westermark, P, J Pathol 2001; 193:270.

Page 22: An Overview of Amyloidosis

2. de Sousa, MM, Vital, C, Ostler, D, et al, Am J Pathol 2000; 156:1911

Page 23: An Overview of Amyloidosis

3. Benson, MD, Liepnieks, JJ, Yazaki, M, et al, Genomics 2001; 72:272

Page 24: An Overview of Amyloidosis

4. Hoppener, JW, Ahren, B, Lips, CJ. N Engl J Med 2000; 343:411

Page 25: An Overview of Amyloidosis

5. Kiuru,S, Amyloid 1998; 5:55

Page 26: An Overview of Amyloidosis

6. Klintworth, GK. Front Biosci 2003; 8:d687 Protein precursor of amyloid deposits-II

Protein class Precursor protein (abbreviation)

Amyloid type

Clinical type

Transport protein Transthyretin (TTR; prealbumin)

ATTR

Hereditary neuropathic and/or cardiopathic amyloids; vitreous amyloidosis; leptomeningeal or renal amyloid in some kindreds; senile systemic amyloidosis [1]

Cerebrovascular/ neurdegeneration

Amyloid precursor protein (APP)

ABeta Hereditary and sporadic Alzheimer disease; congophilic cerebral angiopathy [2,3]

Prion protein (PRP)

APrPscHereditary and sporadic spongiform encephalopathies [4]

BRI gene product ABri/ADanHereditary dementias (British and Danish types) [5]

Cystatin C (Cys - C)

ACysHereditary cerebrovascular hemorrhage with amyloidosis (Icelandic type) [6]

Coagulation protein Fibrinogen alpha chain

AFib Hereditary nephropathic amyloidosis [7]

Enzyme Lysozome ALys Hereditary nephropathic amyloidosis; may have marked hepatic, splenic and gastrointestinal amyloid deposits [7]

Other

Keratoepithelin AKer Various familial corneal dystrophies [8]

Lactoferrin ALacCorneal amyloidosis associated with trichiasis

Odontogenic ameloblast-associated protein

AOaapCalcifying epithelial odontogenic tumors (CEOTs)

Semenogelin 1        

ASem1 Senile seminal vesicle amyloid

Lactadherin AMedSenile aortic amyloid; media deposition

Leukocyte chemotactic factor 2

ALect2 Amyloid nephropathy 

*Nomenclature not yet designated 1. Connors, LH, Lom, A,Prokaeva, T, et al. Amyloid 2003; 10:160 (TTR variants updated at http://medicine.bu.edu/amyloid/Amyloidl.htm)

Page 27: An Overview of Amyloidosis

2. Selkoe, DJ, Physiol Revs 2001; 81:741

Page 28: An Overview of Amyloidosis

3. Revesz, T, Ghiso, J, Lashley, T, et al. J Neuropathol Exp Neurol 2003; 62:885

Page 29: An Overview of Amyloidosis

4. Prusiner, SB. N Engl J Med 2001; 344:1516

Page 30: An Overview of Amyloidosis

5. Vidal, R, Frangione, B, Rostagno, A, et al, Nature 1999; 399:776 and Gibson, G, Gunasekera, N, Lee, M, et al. J Neurochem 2004; 88:281

Page 31: An Overview of Amyloidosis

6. Olafsson, I, Grubb, A, Amyloid 2000; 7:70

Page 32: An Overview of Amyloidosis

7. Hawkins, PN. J Nephrol 2003; 16:443

Page 33: An Overview of Amyloidosis

8. Klintworth, GK. Front Biosci 2003; 8:d687

Page 34: An Overview of Amyloidosis

9. Sipe, JD, Benson, MD, Buxbaum, JN, et al. Amyloid 2010; 17:101. Periorbital hemorrhage in a patient with light chain amyloidosis caused by tape used during general anesthesia

This image shows palpebral and periorbital purpura in a patient with amyloidosis. Reproduced with permission from: Weingarten TN, Hall BA, Richardson BF, et al. Periorbital ecchymoses during general anesthesia in a patient with primary amyloidosis: a harbinger for bleeding? Anesth Analg 2007; 105:1561. Copyright © 2007 Lippincott Williams & Wilkins. Monoclonal pattern SPEP

(A) Densitometer tracing of these findings reveals a tall, narrow-based peak (red asterisk) of gamma mobility and has been likened to a church spire. The monoclonal band has a densitometric appearance similar to that of albumin (alb) and a reduction in the normal polyclonal gamma band.

Page 35: An Overview of Amyloidosis

(B) A dense, localized band (red asterisk) representing a monoclonal protein of gamma mobility is seen on serum protein electrophoresis on agarose gel (anode on left). Reproduced with permission from: Kyle, RA, Rajkumar, SV. Plasma cell disorders. In: Cecil textbook of medicine, 22nd ed, Goldman, L, Ausiello, DA (Eds), WB Saunders, Philadelphia 2004. p.1184. Copyright © 2004 Elsevier. Urinary monoclonal protein

Panel B: This figure illustrates the cellulose acetate electrophoretic pattern of a urine sample. It reveals a dense band of protein with beta mobility. Panel A: Densitometer tracing shows a tall, narrow-based peak of beta mobility. These findings are consistent with a urine monoclonal protein (Bence Jones protein); confirmation of the diagnosis requires demonstration that the protein contains only a lambda or kappa light chain with no heavy chain reactivity. Monoclonal gammopathy on immunofixation

Page 36: An Overview of Amyloidosis

This figure shows the serum protein electrophoretic pattern (SPEP) and immunofixation pattern of a single serum sample with antisera to heavy chain determinants of IgG, IgA, and IgM, and to kappa and lambda light chains. It shows a discrete band on SPEP (red asterisk) and a band with similar mobility reacting only with the antisera to IgG (blue asterisk) and the kappa light chain (black asterisk), indicative of an IgG kappa monoclonal protein. Reproduced with permission from: Kyle, RA, Rajkumar, SV. Plasma cell disorders. In: Cecil textbook of medicine, 22nd ed, Goldman, L, Ausiello, DA (Eds), WB Saunders, Philadelphia 2004. p.1184. Copyright © 2004 Elsevier. Urinary monoclonal protein

Immunofixation of a concentrated urine specimen from the previous patient with antisera to kappa and lambda light chains shows a discrete lambda band, indicating a monoclonal lambda light chain (ie, Bence Jones protein of lambda specificity). Monoclonal plasma cells in primary amyloidosis

Page 37: An Overview of Amyloidosis

Immunoperoxidase stain of a bone marrow biopsy from a patient with primary (AL) amyloidosis. Evidence for a monoclonal plasma cell disease is provided by the intense staining for lambda light chains (left panel) with almost no staining for kappa light chains (right panel). Courtesy of Helmut Rennke, MD.