Complement Deficiencies

Back

Background

The complement system is part of the innate immune system. The complement system plays an important part in defense against pyogenic organisms. It promotes the inflammatory response, eliminates pathogens, and enhances the immune response. Deficiencies in the complement cascade can lead to overwhelming infection and sepsis.

In addition to playing an important role in host defense against infection, the complement system is a mediator in both the pathogenesis and prevention of immune complex diseases, such as systemic lupus erythematosus (SLE). These findings underscore the duality of the complement system. It has a protective effect when functioning in moderation against pathogens; at the same time, the inflammation promoted by complement activation can result in cellular damage when not kept in check.

Complement deficiencies are said to comprise between 1 and 10% of all primary immunodeficiencies.[1] The genetic deficiency of early components of the classical pathway (C1q, C1r/s, C2, C4) tend to be linked with autoimmune diseases[2] , whereas C5 to C9 may have enhanced susceptibility to meningococcal disease. Some new clinical entities are linked with partial complement defects.

Cases of complement deficiency have helped defined the role of complement in host defense.[3] A registry of complement deficiencies has been established as a means to promote joint projects on treatment and prevention of diseases associated with defective complement function. Knowledge about the complement system is expanding. New studies point to the complex interplay between the complement cascade and adaptive immune response, and complement is also being studied in association with ischemic injury as a target of therapy. Although the complement system is part of the body's innate, relatively nonspecific defense against pathogens, its role is hardly primitive or easily understood. This article outlines some of the disease states associated with complement deficiencies and their clinical implications.[4]

Genes that encode the proteins of complement components or their isotypes are distributed throughout different chromosomes, with 19 genes comprising 3 significant complement gene clusters in the human genome.[5] Genetic deficiency of C1q, C1r/s, C2, C4, and C3 is associated with autoimmune diseases, whereas deficiency of C5, C6, C7, C8, C9 increase susceptibility to infections.

Pathophysiology

The complement cascade consists of 3 separate pathways that converge in a final common pathway. The pathways include the classical pathway (C1qrs, C2, C4), the alternative pathway (C3, factor B, properdin), and the lectin pathway (mannan-binding lectin [MBL]). The classical pathway is triggered by interaction of the Fc portion of an antibody (immunoglobulin [Ig] M, IgG1, IgG2, IgG3) or C-reactive protein with C1q. The alternative pathway is activated in an antibody-independent manner. Lectins activate the lectin pathway in a manner similar to the antibody interaction with complement in the classical pathway. These 3 pathways converge at the component C3. Although each branch is triggered differently, the common goal is to deposit clusters of C3b on a target. This deposition provides for the assembly of the membrane attack complex (MAC), components C5b-9. The MAC exerts powerful killing activity by creating perforations in cellular membranes. See the image below.



View Image

Complement pathways and deficiencies.

Deficiencies in complement predispose patients to infection via 2 mechanisms: (1) ineffective opsonization and (2) defects in lytic activity (defects in MAC). Specific complement deficiencies are also associated with an increased risk of developing autoimmune disease, such as SLE.

An intricate system regulates complement activity. The important components of this system are various cell membrane–associated proteins such as complement receptor 1 (CR1), complement receptor 2 (CR2), and decay accelerating factor (DAF). A North African study of molecular basis of complement factor I deficiency in atypical hemolytic and uremic syndrome patients suggested that the Ile357Met mutation may be a founding effect.[6]

In addition to these cell surface–associated proteins, other plasma proteins regulate specific steps of the classic or alternative pathway; for example, the proteins factor H and factor I inhibit the formation of the enzyme C3 convertase of the alternative pathway. Similarly, the enzyme C1q esterase acts as an inhibitor of the classic pathway serine proteases C1r and C1s. Deficiency of any of these regulatory proteins results in a state of overactivation of the complement system, with damaging inflammatory effects.[7] Two clinical manifestations of such deficiencies are paroxysmal nocturnal hemoglobinuria and hereditary angioedema, both of which are discussed in other Medscape Reference articles (see Paroxysmal Nocturnal Hemoglobinuria and Angioedema).

Epidemiology

Frequency

International

Complement deficiencies are relatively rare worldwide, and estimates of prevalence are based on results from screening high-risk populations. Retrospective studies of persons with frequent meningococcal infections report varying prevalence based on geographic location. In populations with recurrent meningococcal infection, the prevalence rate is as high as 30%. Individuals with C1q deficiency have a 93% chance of developing SLE. Similarly, C1rs deficiency has a 57% association with SLE and C4 deficiency has a 75% association with SLE.

Mortality/Morbidity

Individuals with complement deficiencies that hinder opsonization present with frequent recurrent infections and a high rate of morbidity and mortality. Deficiency of C3, the major opsonin, results in recurrent pyogenic infections, particularly with encapsulated bacteria.

Deficiencies of early classical pathway components (C1, C4, C2) do not usually predispose individuals to severe infections but are associated with autoimmune disorders, especially SLE.

Patients with a defect in formation of the MAC have a lesser degree of morbidity and mortality than, for example, patients with a defect in C3; the deficiency in the lytic component of the complement cascade is thought to have some protective effect against the generation of full-blown sepsis. These patients are at high risk for recurrent infection with Neisseria gonorrhoeae or Neisseria meningitidis. Severe pyogenic infections and sepsis occur in children and neonates who have a deficiency of a MAC component.

Race

While no definitive racial patterns of association have been established for the majority of complement deficiencies, ethnic predispositions have been described for certain complement deficiencies. For example, deficiencies in properdin and C2 have been associated with the white race, C6 deficiencies have been shown to have a possible predisposition in African populations, and deficiencies in C8 and C9 have been associated with an Asian racial background. More specifically, 2 functionally distinct C8 deficiency states have been described: C8 alpha-gamma deficiency seen mostly in persons of Afro-Caribbean, Hispanic, and Japanese descent; and C8beta, mainly evident in persons of Caucasian descent.[8] However, for most of these deficiencies, the absolute number of patients studied has been quite small.

Sex

Most complement deficiencies affect both sexes equally.

The majority of complement deficiencies are inherited in an autosomal recessive pattern (although MBL deficiency has been described as having both an autosomal dominant and recessive pattern). An exception to the autosomal pattern of inheritance is properdin deficiency, which is an X-linked trait.

Age

Individuals with complement deficiencies that hinder opsonization often present at an early age (months to a few years) because of increased susceptibility to overwhelming infection.

Patients with deficiencies in formation of the MAC tend to present when slightly older (late-teenage years).

Complement deficiencies associated with immune complex diseases, such as SLE, do not show a clear pattern of age at first presentation.

History

Infants may have Leiner disease, which manifests as recurrent diarrhea, wasting, and generalized seborrheic dermatitis. The defect in persons with Leiner disease is usually attributed to a defect of the fifth component of complement (C5). However, a child was described by Sonea and associates who had Leiner disease associated with diminished C3, and another was described by Goodyear and Harper with a low level of the fourth component of complement and reduced neutrophil mobility.[9, 10] Thus, the C5 defect may not be the sole cause of Leiner disease, as has been suggested; diminished C3 or C4, or C5 dysfunction or deficiency with hypogammaglobulinemia or other lymphoid deficiency, is also required for its expression.

One family from the Arabian Gulf region with multiple members affected by meningococcemia and abscent serum complement 5 (C5) was found to have a homozygous nonsense mutation in exon 1, with the change of cytosine to thymine at position 55 (55C > T) leading to change of the glutamine amino acid at position 19 to a stop codon (Q19X), and serologically absence of C5 in the serum.[11]

The 3 major sequelae of complement deficiencies, based on the pathophysiology of each defect, are (1) defects that result in inadequate opsonization, (2) defects in cell lysis, and (3) the association of complement deficiencies with immune complex diseases.

Defects that result in inadequate opsonization

Opsonization is the process of coating a pathogenic organism so that it is more easily ingested by the macrophage system. The complement protein C3b, along with its cleavage product C3bi, is a potent agent of opsonization in the complement cascade. Any defect that causes decreased production of C3b results in inadequate opsonization ability. Such opsonization defects can be caused by deficiencies in components of the classic, alternative, or MBL pathways, or defects may be caused by deficiencies of the C3b component itself.

The clinical history of patients with classic pathway deficiencies varies slightly from other complement-deficient patients. In the small number of patients studied, patients with classic pathway deficiencies (ie, deficiency of C1qrs, C2, or C4) are similar in presentation to patients with primary immunoglobulin deficiencies. For example, patients tend to have frequent sinopulmonary infections with organisms such as Streptococcus pneumoniae. More commonly, these patients develop autoimmune syndromes.

In order to generate an antibody response, an antigen must bind to the complement receptor (CR2) on B cells and the complement protein C3d. A deficiency of C1-C4 proteins leads to an inadequate humoral response in these patients. Patients also have a decrease in classic pathway production of the opsonin C3b, but the alternative and MBL pathways seem to compensate for this defect because opsonin is not completely absent.

Opsonization defects can also be caused by alternative pathway deficiencies. In the alternative pathway, a deficiency of factor B, factor D, or properdin can result in a decreased amount of C3b. Deficiencies in properdin have been described in some detail. Properdin is a protein encoded on the X chromosome. Properdin stabilizes the C3 convertase (C3bBb) of the alternative pathway. Stabilization of C3 convertase increases the half-life of the complex from 5 minutes to 30 minutes, exponentially increasing the amount of C3b that can be deposited on a microbial surface. The role of C3b as an opsonin is essential in defense against neisserial infection, and the risk of overwhelming neisserial infection increases in the absence of properdin.

The third pathway whose deficiencies can result in opsonization defects is the MBL pathway. MBL is one of the collectin proteins. These proteins share specific structural characteristics, namely the presence of a collagenlike region and a Ca2+ -dependent lectin domain. Of all the lectin proteins, only MBL has been shown to have the ability to activate the complement system. The MBL protein can activate the C4 and C2 components of complement by forming a complex with serine proteases known as MASP1 and MASP2. MASP1 and MASP2 activation results in the protein products C3 and C3b. The MBL protein is versatile because it can bind to a variety of substrates, prompting some to describe the MBL as a kind of universal antibody. Clinically, MBL deficiencies increase risk of infection with the yeast Saccharomyces cerevisiae and encapsulated bacteria such as Neisseria meningitides and S pneumoniae.[12]

Finally, absolute deficiencies of C3 itself also result in defective opsonization. The C3 component occupies an important place at the junction of both the classic and alternative pathways. As such, C3 deficiency results in severe opsonization dysfunction. C3 deficiency also causes deficient leukocyte chemotaxis because of decreased C3a concentrations and decreased bactericidal killing secondary to decreased formation of MAC. Clinically, patients present at an early age with overwhelming infections from encapsulated bacteria. In addition to opsonization problems, C3 deficiency also impairs adequate clearance of circulating immune complexes, and 79% of patients with C3 deficiency develop some form of collagen vascular disease.

Deficiencies of the inhibitory proteins of the classic and alternative pathways can also result in a functional C3 deficiency through uncontrolled consumption of C3. Factors H and I are proteins that inhibit C3 formation in the alternative and classic pathways, respectively. Deficiencies in either of these C3 inhibitors can result in an overactivation of C3 and subsequent C3 depletion. Clinically, these patients are similar to patients with absolute C3 deficiency.

While deficiencies in complement proteins can predispose patients to infections such as the clinical conditions described above, a deficiency in regulation of complement can also lead to disease. Deficiencies or defective regulation of the alternative complement pathway can occur because of genetic mutations or deficiencies in the regulatory protein Factor H. This defective regulation of the alternative pathway can be associated with diseases such as an atypical form of hemolytic uremic syndrome, membranoproliferative glomerulonephritis (type I and II), and age-related macular degeneration.

Defects in cell lysis

Complement deficiencies of the terminal cascade proteins also predispose patients to infection, but the clinical history of these patients is different. The terminal complement proteins are the proteins in the cascade that form the MAC, ie, complement proteins C5-C9. These proteins are responsible for bactericidal killing of organisms such as N meningitidis. The frequency rate of meningococcal infection in patients with terminal complement deficiency is as high as 66%. In addition to this high rate of first-time infection, the frequency rate of recurrence with the same organism is also as high as 50%. The serogroups of N meningitidis responsible for infections in this group tend to be the more rare serogroups Y and W135, rather than the more common serogroups B, A, and C.

Clinically, patients with terminal deficiency tend to present with infection at an older age compared with patients with other complement deficiencies. These individuals also have less morbidity and mortality associated with infection. Unlike patients with a classic pathway deficiency, humoral immunity is intact but lysis of pathogenic organisms is impaired.

Both the opsonization and lytic function of complement protect against a variety of other nonbacterial pathogens, such fungi, viruses, and mycobacteria. The role of complement in defense against viral infection is sufficiently important that pathogenic viruses have had to develop strategies to evade complement activation. For example, human immunodeficiency virus type 1 has recently been described as escaping complement-mediated lysis through the incorporation of regulatory proteins, such as DAF, into the viral envelope. Similarly, other viruses have also evolved complement-specific means of escape.

Complement deficiencies and associated immune complex diseases

Patients with complement deficiencies of the classic pathway are predisposed to develop immune complex diseases.

Patients with deficiencies of the classic pathway components C1qrs, C2, or C4 have been shown to have an increased likelihood of developing SLE. C1q deficiency is less commonly linked with neuropsychiatric SLE, which may be first evident with seizures.[13]  Homozygous deficiency of C1q has the highest association with SLE, with a recently quoted prevalence rate of 93%. Subsequent components of the classic pathway have respective prevalence rates of 57% for C1rs deficiency, 75% association with homozygous C4 deficiencies, and 10% prevalence in patients with C2 deficiencies.

The reason complement deficiency increases the risk of developing SLE is that complement helps in the prevention of immune complex disease by decreasing the number of circulating immune complexes; the greater the concentration of these precipitating immune complexes, the higher the likelihood that they will deposit in nearby tissues and cause an inflammatory response.

Complement aids in neutralization and clearance of antigen-antibody complexes in several ways. The classic pathway acts to inhibit immune complex precipitation by physically interfering with immune complex aggregation. Secondly, complement enhances the clearance of circulating immune complexes by binding to complement receptors (CR1) on cells such as erythrocytes, B lymphocytes, T lymphocytes, and macrophages. When complement (specifically C3b) binds to CR1 on erythrocytes, the immune complex can be transported through the circulation to be presented to the macrophage systems in the spleen and liver.

Components of the classical pathway also play an important role in the recognition and clearance of apoptotic cells. Normally, intracellular proteins are displayed on the surface of cells undergoing apoptosis. If these apoptotic cells are not cleared efficiently by the complement system, these cell surface proteins have the potential to act as autoantigens, acting as potential triggers for autoimmune diseases such as systemic lupus.

In addition to its role in the development of diseases such as SLE, complement activation also likely plays a role in the pathogenesis of the antiphospholipid antibody syndrome (APS), a thrombophilic inflammatory disorder that can be associated with SLE or can occur independently. In a mouse model of antiphospholipid-antibody associated fetal death, mice who were deficient in C3 and mice who were treated with a regulatory protein that inhibits C3 cleavage were protected from fetal loss. Recent studies in human subjects have also found a positive association between the presence of C4d deposition on activated platelets and the presence of arterial thrombosis. Further studies in human subjects are ongoing.[14]

Complement component C8, when entirely absent, results in increased susceptibility to gram-negative bacteria such as Neisseria species.[8] Two functionally distinct C8 deficiency states have been described: C8 alpha-gamma deficiency and C8beta deficiency. A duplication mutation in C8-beta deficiency was recently documented, extending the molecular heterogeneity of this disorder. Complement screening would detect this rare primary immunodeficiency and allow prophylaxis to prevent recurrent Neisseria infections with this potentially severe outcome.[15]

A relatively small sampling of Finnish non-tuberculous mycobacteria patients had significantly more often C4 deficiencies than the healthy control subjects, suggesting that both a deficiency of complement C4 and bronchiectasis in healthy females as risk factors for pulmonary NTM infections.[16]

Physical

No specific physical findings are pathognomonic for complement deficiencies. Rather, clinical manifestations are representative of the infections and immune complex diseases to which patients are predisposed.

Because N meningitides is the overwhelmingly prevalent bacterial pathogen in these patients, knowledge of the physical characteristics of disseminated meningococcal disease is important.[17] The characteristic maculopapular rash that occurs in up to 75% of individuals with meningococcemia occurs soon after disease onset. The rash consists of pink lesions on the trunk and extremities; lesions are approximately 2-10 mm in diameter. The rash can quickly progress to hemorrhagic lesions. Petechiae are also a prominent finding and can occur on the skin of the trunk and extremities or on mucous membranes, such as the palate and conjunctivae.

Noninfectious diseases, such as SLE, that are associated with complement deficiencies can also have a characteristic physical presentation. Complement deficiencies associated with the deposition of immune complexes in various tissues can result in many of the sequelae of SLE, such as glomerulonephritis, arthralgia, uveitis, and vasculitic rash.

Causes

Most complement deficiencies are caused by a genetic defect in one of the genes that code for the various complement proteins.

No clear environmental or drug-related causes have been identified.

Laboratory Studies

One can screen for deficiencies in complement by performing the total serum classic hemolytic complement (CH50) test or the alternative hemolytic complement (AP50) test. The CH50 test specifically tests for deficiencies in the classic pathway by measuring the ability of the patient's serum to lyse antibody-coated sheep erythrocytes. A deficiency in any of the classic proteins results in a CH50 of zero. Similarly, the AP50 tests for alternative pathway activity. Direct measurement of individual serum complement proteins, such as C3 and C4, can also be performed and is helpful in determining the diagnosis.

Dried blood spot samples from newborns, which are already widely used in neonatal screening for selected metabolic diseases, may be employed in the future using reverse phase protein microarrays for determination of complement component C3 levels collected at birth.[18] In one recent study, normal levels of C3 were detected from healthy newborns, while no C3 was documented in sera and dried blood samples from patients who were C3 deficient in C3.[18]

Imaging Studies

No specific imaging studies are indicated. Consider performing a head CT scan prior to a lumbar puncture in a patient thought to have meningitis.

Other Tests

Patients with classic complement pathway deficiencies should be screened for sequelae of immune complex diseases. Urinalysis and a complete blood cell count should be performed on these patients.

Procedures

A lumbar puncture should be performed on patients with possible meningitis to assist in the definitive diagnosis of bacterial meningitis.

Medical Care

Definitive treatment of complement deficiencies requires replacing the missing component of the cascade, either through direct infusion of the protein or through gene therapy. Because neither of these options is currently available, treatment of these patients focuses on managing the sequelae of the particular complement deficiencies.

For many patients, treatment must be focused on eradicating a particular infection, especially with encapsulated organisms such as N meningitidis. In most cases of meningococcal disease, treatment with meningeal doses of a third-generation cephalosporin covers most strains of N meningitidis.

For other patients, the complement deficiency may manifest as episodic flares of autoimmune diseases; treatment of these patients focuses on immunosuppressive therapy of these diseases.

Importantly, note that some overlap often exists between an increased susceptibility to infection and the greater tendency to develop autoimmune disease; both of these clinical situations may need to be addressed simultaneously in any one patient.

Consultations

In a patient with a possible complement deficiency, consider consultation with an allergist and immunologist to determine appropriate diagnostic tests.

Also, consider consultation with a rheumatologist or infectious disease specialist to help manage acute complications of the complement deficiency.

Diet

No specific diet restrictions are required.

Activity

Activity can continue as tolerated by the patient.

Prevention

Complement screening in relatives of patients with complement deficiencies might detect this primary immunodeficiency state and allow prophylaxis to prevent potentially disabling or life-threatening infections.[15]

Complement deficiencies may become apparent in adults by meningococcal and other infections. Adults with N meningitidis infection should be evaluated for a complement deficiency.[19]

Medication Summary

Cephalosporins are often used for treatment of N meningitidis infection in patients with complement deficiency. Third- or fourth-generation cephalosporins are used for coverage of infection with any of the encapsulated bacteria.

Ceftriaxone (Rocephin)

Clinical Context:  Third-generation cephalosporin with broad-spectrum, gram-negative activity; lower efficacy against gram-positive organisms; higher efficacy against resistant organisms. Arrests bacterial growth by binding to one or more penicillin-binding proteins.

Cefepime (Maxipime)

Clinical Context:  Fourth-generation cephalosporin with good gram-negative coverage. Similar to third-generation cephalosporins but has better gram-positive coverage.

Class Summary

Therapy must cover all likely pathogens in the context of this clinical setting. Antibiotic selection should be guided by blood culture sensitivity results whenever feasible.

Further Outpatient Care

Patients with a known complement deficiency should be screened for glomerular or immune complex disease. Obtain urinalysis results to check for proteinuria and rheumatologic serology results to screen for SLE.

Further Inpatient Care

Serious infectious states warrant hospitalization for treatment.

Inpatient treatment is not necessarily needed to screen for complement deficiencies if the patient is asymptomatic.

Inpatient & Outpatient Medications

Cephalosporins (third- or fourth-generation) are needed for treatment of meningeal infection.

Deterrence/Prevention

Administration of the multivalent meningococcal vaccine is recommended in patients with known complement deficiency, especially those patients deficient in the MAC proteins. Similarly, administration of the pneumococcal vaccine and the Haemophilus influenzae vaccine also may provide protection against these encapsulated organisms.

Complications

Complications of complement deficiencies can be serious; severe CNS damage and death from meningitis are among the worst possible adverse outcomes.

Prognosis

In general, the prognosis for patients with C3 deficiencies is poorer than that of individuals with other complement deficiencies. Patients may have severe, recurrent episodes of pyogenic infection beginning when as young as a few months. Many can die from sepsis early in life.

Patients with a deficiency of one of the early components of the classical pathway (C1, C4, C2) are at high risk for autoimmune disease but at lower risk for overwhelming sepsis with pyogenic infections.

Deficiency of a MAC component (C5, C6, C7, C8) or of properdin increases the risk for recurrent infections caused by Neisseria organisms.

Mannan-binding lectin (MBL) deficiency has been linked to an increased frequency of pyogenic infections and sepsis, especially in neonates and children.

Patient Education

Patients with an identified complement deficiency should be counseled regarding possible complications and risks associated with this deficiency.

Family members should be screened for complement deficiencies and counseled regarding possible risks.

What are complement deficiencies?What is the pathophysiology of complement deficiencies?What is the prevalence of complement deficiencies?What is the mortality and morbidity associated with complement deficiencies?What are the racial predilections of complement deficiencies?What are the sexual predilections of complement deficiencies?At what age do complement deficiencies typically present?Which clinical history findings are characteristic of complement deficiencies?Which clinical history findings are characteristic of opsonization-caused complement deficiencies?Which clinical history findings are characteristic of defects in cell lysis causing complement deficiencies?Which clinical history findings are characteristic of immune complex diseases in complement deficiencies?Which physical findings are characteristic of complement deficiencies?What causes complement deficiencies?What are the differential diagnoses for Complement Deficiencies?What is the role of lab tests in the workup of complement deficiencies?What is the role of imaging studies in the workup of complement deficiencies?Which tests are performed in the workup of classic complement pathway deficiencies?What is the role of lumbar puncture in the workup of complement deficiencies?How are complement deficiencies treated?Which specialist consultations are beneficial to patients with complement deficiencies?Which dietary modifications are used in the treatment of complement deficiencies?Which activity modifications are used in the treatment of complement deficiencies?Who should undergo complement screening?What is the role of medications in the treatment of complement deficiencies?Which medications in the drug class Antibiotics are used in the treatment of Complement Deficiencies?What screenings should be performed for patients with known complement deficiency?When is inpatient care indicated for complement deficiencies?What is the role of cephalosporins in the treatment of complement deficiencies?Which vaccines should be given to patients with complement deficiencies?What are the possible complications of complement deficiencies?What is the prognosis of complement deficiencies?What is included in patient education about complement deficiencies?

Author

Robert A Schwartz, MD, MPH, Professor and Head of Dermatology, Professor of Pathology, Professor of Pediatrics, Professor of Medicine, Rutgers New Jersey Medical School

Disclosure: Nothing to disclose.

Coauthor(s)

R Krishna Chaganti, MD, Assistant Clinical Professor, Division of Rheumatology, UCSF

Disclosure: Nothing to disclose.

Specialty Editors

Francisco Talavera, PharmD, PhD, Adjunct Assistant Professor, University of Nebraska Medical Center College of Pharmacy; Editor-in-Chief, Medscape Drug Reference

Disclosure: Received salary from Medscape for employment. for: Medscape.

Chief Editor

Michael A Kaliner, MD, Clinical Professor of Medicine, George Washington University School of Medicine; Medical Director, Institute for Asthma and Allergy

Disclosure: Nothing to disclose.

Acknowledgements

Margaret R Donohoe, MD Consulting Staff, Department of Allergy and Immunology, Albemarle Hospital

Disclosure: Nothing to disclose.

Samuel R Marney, Jr, MD Director, Associate Professor, Department of Internal Medicine, Division of Allergy and Immunology, Vanderbilt University School of Medicine

Samuel R Marney, Jr, MD is a member of the following medical societies: American Academy of Allergy Asthma and Immunology, American College of Allergy, Asthma and Immunology, American College of Physicians, and Tennessee Medical Association

Disclosure: Nothing to disclose.

Darrilyn Moyer, MD Associate Program Director, Associate Professor, Department of Internal Medicine, Temple University School of Medicine

Disclosure: Nothing to disclose.

References

  1. Grumach AS, Kirschfink M. Are complement deficiencies really rare? Overview on prevalence, clinical importance and modern diagnostic approach. Mol Immunol. 2014 Oct. 61(2):110-117. [View Abstract]
  2. Bryan AR, Wu EY. Complement deficiencies in systemic lupus erythematosus. Curr Allergy Asthma Rep. 2014 Jul. 14(7):448. [View Abstract]
  3. Botto M, Kirschfink M, Macor P, Pickering MC, Wurzner R, Tedesco F. Complement in human diseases: Lessons from complement deficiencies. Mol Immunol. 2009 Sep. 46(14):2774-83. [View Abstract]
  4. Skattum L, van Deuren M, van der Poll T, Truedsson L. Complement deficiency states and associated infections. Mol Immunol. 2011 Aug. 48(14):1643-55. [View Abstract]
  5. Mayilyan KR. Complement genetics, deficiencies, and disease associations. Protein Cell. 2012 Jul. 3(7):487-96. [View Abstract]
  6. Jlajla H, Dehman F, Jallouli M, Khedher R, Ayadi I, Zerzeri Y, et al. Molecular basis of complement factor I deficiency in Tunisian atypical hemolytic and uremic syndrome Patients. Nephrology (Carlton). 2018 Jan 2. [View Abstract]
  7. Roumenina LT, Sène D, Radanova M, Blouin J, Halbwachs-Mecarelli L, Dragon-Durey MA, et al. Functional complement C1q abnormality leads to impaired immune complexes and apoptotic cell clearance. J Immunol. 2011 Oct 15. 187(8):4369-73. [View Abstract]
  8. Arnold DF, Roberts AG, Thomas A, Ferry B, Morgan BP, Chapel H. A novel mutation in a patient with a deficiency of the eighth component of complement associated with recurrent meningococcal meningitis. J Clin Immunol. 2009 Sep. 29(5):691-5. [View Abstract]
  9. Sonea MJ, Moroz BE, Reece ER. Leiner's disease associated with diminished third component of complement. Pediatr Dermatol. 1987 Aug. 4(2):105-7. [View Abstract]
  10. Goodyear HM, Harper JI. Leiner's disease associated with metabolic acidosis. Clin Exp Dermatol. 1989 Sep. 14(5):364-6. [View Abstract]
  11. Arnaout R, Al Shorbaghi S, Al Dhekri H, et al. C5 Complement Deficiency in a Saudi Family, Molecular Characterization of Mutation and Literature Review. J Clin Immunol. 2013 Feb 1. [View Abstract]
  12. Degn SE, Jensenius JC, Thiel S. Disease-causing mutations in genes of the complement system. Am J Hum Genet. 2011 Jun 10. 88(6):689-705. [View Abstract]
  13. van Schaarenburg RA, Magro-Checa C, Bakker JA, Teng YK, Bajema IM, Huizinga TW, et al. C1q Deficiency and Neuropsychiatric Systemic Lupus Erythematosus. Front Immunol. 2016 Dec 27. 7:647. [View Abstract]
  14. Peerschke EI, Yin W, Alpert DR, Roubey RA, Salmon JE, Ghebrehiwet B. Serum complement activation on heterologous platelets is associated with arterial thrombosis in patients with systemic lupus erythematosus and antiphospholipid antibodies. Lupus. 2009 May. 18(6):530-8. [View Abstract]
  15. Dellepiane RM, Dell'Era L, Pavesi P, Macor P, Giordano M, De Maso L, et al. Invasive meningococcal disease in three siblings with hereditary deficiency of the 8(th) component of complement: evidence for the importance of an early diagnosis. Orphanet J Rare Dis. 2016 May 17. 11 (1):64. [View Abstract]
  16. Kotilainen H, Lokki ML, Paakkanen R, Seppänen M, Tukiainen P, Meri S, et al. Complement C4 deficiency--a plausible risk factor for non-tuberculous mycobacteria (NTM) infection in apparently immunocompetent patients. PLoS One. 2014. 9(3):e91450. [View Abstract]
  17. Brostowski LE, Graf EH. The Brief Case: Meningococcemia Leading to a Diagnosis of Complement Deficiency in a 23-Month-Old. J Clin Microbiol. 2019 Feb. 57 (2):[View Abstract]
  18. Janzi M, Sjoberg R, Wan J, et al. Screening for C3 deficiency in newborns using microarrays. PLoS One. 2009. 4(4):e5321. [View Abstract]
  19. Audemard-Verger A, Descloux E, Ponard D, Deroux A, Fantin B, et al. Infections Revealing Complement Deficiency in Adults: A French Nationwide Study Enrolling 41 Patients. Medicine (Baltimore). 2016 May. 95 (19):e3548. [View Abstract]
  20. Beynon HL, Davies KA, Haskard DO, Walport MJ. Erythrocyte complement receptor type 1 and interactions between immune complexes, neutrophils, and endothelium. J Immunol. 1994 Oct 1. 153(7):3160-7. [View Abstract]
  21. Brandtzaeg P, Mollnes TE, Kierulf P. Complement activation and endotoxin levels in systemic meningococcal disease. J Infect Dis. 1989 Jul. 160(1):58-65. [View Abstract]
  22. Carroll MC. The role of complement in B cell activation and tolerance. Adv Immunol. 2000. 74:61-88. [View Abstract]
  23. Davies KA, Erlendsson K, Beynon HL, et al. Splenic uptake of immune complexes in man is complement-dependent. J Immunol. 1993 Oct 1. 151(7):3866-73. [View Abstract]
  24. Davies KA, Peters AM, Beynon HL, Walport MJ. Immune complex processing in patients with systemic lupus erythematosus. In vivo imaging and clearance studies. J Clin Invest. 1992 Nov. 90(5):2075-83. [View Abstract]
  25. Dempsey PW, Allison ME, Akkaraju S, et al. C3d of complement as a molecular adjuvant: bridging innate and acquired immunity. Science. 1996 Jan 19. 271(5247):348-50. [View Abstract]
  26. Densen P, Weiler JM, Griffiss JM, Hoffmann LG. Familial properdin deficiency and fatal meningococcemia. Correction of the bactericidal defect by vaccination. N Engl J Med. 1987 Apr 9. 316(15):922-6. [View Abstract]
  27. Doody GM, Dempsey PW, Fearon DT. Activation of B lymphocytes: integrating signals from CD19, CD22 and Fc gamma RIIb1. Curr Opin Immunol. 1996 Jun. 8(3):378-82. [View Abstract]
  28. Fauci A, Braunwald E, Isselbacher KJ. Meningococcal Infections. Fauci A, Braunwald E, Isselbacher KJ, Wilson JD, Martin JB, Kasper DL, Hauser SL, Longo DL, Harrison TR. Harrison's Principles of Internal Medicine. 14th ed. New York, NY: McGraw-Hill; 1998. 910-5.
  29. Figueroa JE, Densen P. Infectious diseases associated with complement deficiencies. Clin Microbiol Rev. 1991 Jul. 4(3):359-95. [View Abstract]
  30. Frank MM. Complement deficiencies. Pediatr Clin North Am. 2000 Dec. 47(6):1339-54. [View Abstract]
  31. Girardin E, Grau GE, Dayer JM, et al. Tumor necrosis factor and interleukin-1 in the serum of children with severe infectious purpura. N Engl J Med. 1988 Aug 18. 319(7):397-400. [View Abstract]
  32. Goodnow CC. Pathways for self-tolerance and the treatment of autoimmune diseases. Lancet. 2001 Jun 30. 357(9274):2115-21. [View Abstract]
  33. Kang HJ, Kim HS, Lee YK, Cho HC. High incidence of complement C9 deficiency in Koreans. Ann Clin Lab Sci. 2005. 35(2):144-8. [View Abstract]
  34. Kuby J. The Complement System. Immunology. 3rd ed. New York, NY: WH Freeman; 1997. 335-55.
  35. Lehner PJ, Davies KA, Walport MJ, et al. Meningococcal septicaemia in a C6-deficient patient and effects of plasma transfusion on lipopolysaccharide release. Lancet. 1992 Dec 5. 340(8832):1379-81. [View Abstract]
  36. Leitao MF, Vilela MM, Rutz R, et al. Complement factor I deficiency in a family with recurrent infections. Immunopharmacology. 1997 Dec. 38(1-2):207-13. [View Abstract]
  37. Lutz HU. How pre-existing, germline-derived antibodies and complement may help induce a primary immune response to nonself. Scand J Immunol. 1999 Mar. 49(3):224-8. [View Abstract]
  38. Manzi S, Ahearn JM, Salmon J. New insights into complement: a mediator of injury and marker of disease activity in systemic lupus erythematosus. Lupus. 2004. 13(5):298-303. [View Abstract]
  39. Medof ME, Iida K, Mold C, Nussenzweig V. Unique role of the complement receptor CR1 in the degradation of C3b associated with immune complexes. J Exp Med. 1982 Dec 1. 156(6):1739-54. [View Abstract]
  40. Pickering MC, Botto M, Taylor PR, et al. Systemic lupus erythematosus, complement deficiency, and apoptosis. Adv Immunol. 2000. 76:227-324. [View Abstract]
  41. Prodinger WM. Complement receptor type two (CR2,CR21): a target for influencing the humoral immune response and antigen-trapping. Immunol Res. 1999. 20(3):187-94. [View Abstract]
  42. Ross GD, Yount WJ, Walport MJ, et al. Disease-associated loss of erythrocyte complement receptors (CR1, C3b receptors) in patients with systemic lupus erythematosus and other diseases involving autoantibodies and/or complement activation. J Immunol. 1985 Sep. 135(3):2005-14. [View Abstract]
  43. Rother K, Till GO, Hansch GM, eds. The Complement System. New York, NY: Springer-Verlag; 1997.
  44. Schlesinger M, Nave Z, Levy Y, et al. Prevalence of hereditary properdin, C7 and C8 deficiencies in patients with meningococcal infections. Clin Exp Immunol. 1990 Sep. 81(3):423-7. [View Abstract]
  45. Soderstrom C, Braconier JH, Danielsson D, Sjoholm AG. Bactericidal activity for Neisseria meningitidis in properdin-deficient sera. J Infect Dis. 1987 Jul. 156(1):107-12. [View Abstract]
  46. Sullivan KE, Winkelstein JA. Genetically Determined Deficiencies of the Complement System. Ochs HD, Edvard Smith CI, Puck J. Primary Immunodeficiency Diseases: A Molecular and Genetic Approach. New York, NY: Oxford University Press; 1999. 397-416.
  47. Sumiya M, Super M, Tabona P, et al. Molecular basis of opsonic defect in immunodeficient children. Lancet. 1991 Jun 29. 337(8757):1569-70. [View Abstract]
  48. Summerfield JA, Sumiya M, Levin M, Turner MW. Association of mutations in mannose binding protein gene with childhood infection in consecutive hospital series. BMJ. 1997 Apr 26. 314(7089):1229-32. [View Abstract]
  49. Super M, Thiel S, Lu J, et al. Association of low levels of mannan-binding protein with a common defect of opsonisation. Lancet. 1989 Nov 25. 2(8674):1236-9. [View Abstract]
  50. Tolnay M, Tsokos GC. Complement receptor 2 in the regulation of the immune response. Clin Immunol Immunopathol. 1998 Aug. 88(2):123-32. [View Abstract]
  51. Turner MW. Mannose-binding lectin: the pluripotent molecule of the innate immune system. Immunol Today. 1996 Nov. 17(11):532-40. [View Abstract]
  52. Volanakis JE, Frank MM, eds. The Human Complement System in Health and Disease. New York, NY: Marcel Dekker; 1998.
  53. Walport MJ. Complement. First of two parts. N Engl J Med. 2001 Apr 5. 344(14):1058-66. [View Abstract]
  54. Walport MJ. Complement. Second of two parts. N Engl J Med. 2001 Apr 12. 344(15):1140-4. [View Abstract]
  55. Walport MJ, Davies KA, Botto M. C1q and systemic lupus erythematosus. Immunobiology. 1998 Aug. 199(2):265-85. [View Abstract]
  56. Walport MJ, Lachmann PJ. Erythrocyte complement receptor type 1, immune complexes, and the rheumatic diseases. Arthritis Rheum. 1988 Feb. 31(2):153-8. [View Abstract]
  57. Wisnieski JJ, Baer AN, Christensen J, et al. Hypocomplementemic urticarial vasculitis syndrome. Clinical and serologic findings in 18 patients. Medicine (Baltimore). 1995 Jan. 74(1):24-41. [View Abstract]
  58. Zipfel PF, Heinen S, Jozsi M, Skerka C. Complement and diseases: defective alternative pathway control results in kidney and eye diseases. Mol Immunol. 2006 Jan. 43(1-2):97-106.

Complement pathways and deficiencies.

Complement pathways and deficiencies.