ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Biology of Anaplasmataceae

Biology of Anaplasmataceae
Author:
J Stephen Dumler, MD
Section Editor:
Daniel J Sexton, MD
Deputy Editor:
Keri K Hall, MD, MS
Literature review current through: Jan 2024.
This topic last updated: Jun 07, 2023.

INTRODUCTION — Bacteria in the family Anaplasmataceae, have been known to cause diseases of interest to veterinarians for over 100 years. However, their role as agents of human disease was not recognized until 1987. Since then, the number of genera and species recognized to cause human disease has markedly expanded, as has our knowledge about the epidemiology, clinical characteristics, and treatment of the diseases caused by these organisms.

The biology of pathogenic members of the family Anaplasmataceae will be reviewed in this topic. A discussion of the diseases caused by these bacteria is presented separately. (See "Human ehrlichiosis and anaplasmosis".)

TAXONOMY AND PHYLOGENY — The taxonomy of the Rickettsiales order and the two major families (Rickettsiaceae and Anaplasmataceae) underwent a radical change during the last two decades [1,2], based mostly on genomic data, but also on phenotypic features. A detailed discussion of Rickettsiaceae is presented separately. (See "Biology of Rickettsia rickettsii infection".)

The current configuration of the genus Anaplasmataceae provides for four to five clades that represent individual genera that contain known pathogens: Ehrlichia, Anaplasma, Neoehrlichia, Neorickettsia, and Wolbachia. These Anaplasmataceae can be divided into tick and non-tick transmitted bacteria:

Tick-transmitted – All members of the genera Ehrlichia, Anaplasma, and Neoehrlichia are transmitted by the bites of ticks. These include:

Ehrlichia: E. chaffeensis, E. canis, E. muris subsp muris, E. muris subsp eauclairensis, E. ruminantium, E. minasensis, and E. ewingii.

Anaplasma: A. phagocytophilum, A. platys, A. marginale, A. centrale, A. ovis, and A. bovis.

Neoehrlichia: Candidatus N. mikurensis.

Non-tick transmitted – Species in the Neorickettsia and Wolbachia genera have distinct and complex transmission mechanisms that do not involve ticks. Members of these genera include:

Neorickettsia: N. sennetsu, N. risticii, and N. helminthoeca.

Wolbachia: W. pipientis and other unnamed species.

HOST SPECIFICITY — Individual species of Anaplasmataceae infect a wide range of hosts and cause disease in humans, dogs, cats, sheep, goats, cattle, horses, and other animals (table 1). The most important causes of human disease are members of the genera Ehrlichia and Anaplasma. The majority of recognized human infections are caused by A. phagocytophilum and E. chaffeensis (see "Human ehrlichiosis and anaplasmosis"):

The white-tailed deer (Odocoileus virginianus) is the major reservoir host for E. chaffeensis. The natural cycle of E. chaffeensis involves lone star ticks (Amblyomma americanum), which may transmit to and acquire the organism from infected deer.

The major reservoir host for human-infective strains of A. phagocytophilum is generally a small mammal, such as the white-footed mouse (Peromyscus leucopus). P. leucopus is also the host for Ixodes scapularis, the main tick vector in North America. Similar small mammal/tick associations are seen with the related vectors I. pacificus (North America), I. ricinus and I. persulcatus (Europe and Asia), and Haemaphysalis species (Asia). Some variant strains of A. phagocytophilum that infect deer do not infect mice or humans but may be transmitted by Ixodes species ticks. While ticks often have a preference for a specific host, many are promiscuous in their feeding behavior and readily bite humans.

Other species that cause human infection include E. ewingii (which has a natural life cycle and tick vector similar to E. chaffeensis); E. muris subsp eauclairensis; the Panola Mountain Ehrlichia; "A. capra"; A. ovis; Candidatus N. mikurensis (a cause of an emerging potentially severe febrile disease found in Europe and Asia); and Neorickettsia sennetsu. Although evidence exists for human disease related to Wolbachia endosymbionts in filariasis [3], direct infection of humans has not been definitively identified.

CHARACTERISTICS OF ANAPLASMATACEAE

General characteristics — Ehrlichia and Anaplasma spp are obligate intracellular bacteria that grow within membrane-bound vacuoles in human and animal leukocytes. In humans, either monocytes (human monocytic ehrlichiosis) or granulocytes (human granulocytic anaplasmosis and Ehrlichia ewingii ehrlichiosis) can be infected. (See 'Taxonomy and phylogeny' above and 'Cell specificity' below.)

Unlike rickettsiae, which escape the endocytic vacuole to live freely in the cell cytoplasm, Anaplasmataceae spp replicate within modified vacuoles in the cytoplasm of the host cell. A microcolony of these bacteria within a vacuole is called a morula (picture 1A-B). Similar to forms described for Chlamydia species, Anaplasmataceae bacteria within morulae differentiate into both replicative "reticulate" forms and infectious but nonreplicative "dense-core" forms. (See "Biology of Rickettsia rickettsii infection".)

The genomes of Ehrlichia and Anaplasma species lack complete biosynthetic pathways for peptidoglycan and lipopolysaccharide biosynthesis, as suggested by their very simple cell wall ultrastructure, but can still synthesize some components of peptidoglycan. Their genomes also encode paralogous genes for major surface antigenic proteins that are both ligands by which the bacteria bind to host cells before entry, or that contribute to antigenic variation and immune evasion [4]. Likewise, both genera possess structurally unusual but functional type IV secretion systems that translocate bacterial proteins into host cells. These bacterial effector proteins: (1) trigger signaling in the host cell to induce modifications of the cell cytoskeleton that permit pathogen endocytic entry; (2) alter intracellular trafficking of vacuoles to avoid lysosome fusion; or (3) travel to the cell nucleus and alter host cell transcriptional programs and functions [5,6].

Anaplasmataceae spp are difficult to detect by light microscopy as single bacteria within a leukocyte; however, morulae are easily visualized when present within white blood cells. These can be observed in Giemsa and Wright or other Macchiavello method-stained samples. Individual organisms are generally coccoid but can be pleomorphic when visualized in tissue culture cells [1].

Growth characteristics — Individual organisms induce endocytic entry into host cell vacuoles where they grow, divide, and within a few days form small, tightly packed clusters observable as intracellular inclusions (morulae). These bacterial cells have an open chromatin configuration and are not highly infective. A. phagocytophilum and E. chaffeensis cells with condensed chromatin (dense-core cells) appear several days later. These dense-core cells express unique proteins APH_1235 (A. phagocytophilum) or TRP120 (E. chaffeensis) that are important for infectivity [7,8].

Like rickettsiae, Anaplasmataceae spp will not grow in cell-free culture media. Thus far, they have only been cultured in eukaryotic cells. E. chaffeensis and E. canis can be propagated in canine blood monocytes and in monocyte, endothelial, macrophage, and fibroblast cell lines from murine, human, primate, and canine origin. A. phagocytophilum can be cultivated in vitro using human myeloid leukemia cell lines and some primate and human endothelial cell lines. All can be cultivated in tick embryo cell lines. Axenic growth has not been achieved for any species in the Anaplasmataceae family.

Growth in vitro can take one week to one month or longer to detect. The duration of the organism's viability in human-donated blood is unknown; transfusion-transmitted infection is known to occur, even with leuko-reduced blood products, and organisms can be reisolated from spiked EDTA-anticoagulated blood stored at 4°C for 11 days for E. chaffeensis and stored for 18 days for A. phagocytophilum [9,10]. Fatal infection after transfusion-transmission of A. phagocytophilum has been reported [11].

The metabolism of Anaplasmataceae spp is primarily aerobic. Genes required for glycolysis are absent, but the presence of some intact metabolic pathways in their genomic repertoires suggests that A. phagocytophilum and E. chaffeensis can use simple substrates such as glyceraldehyde 3-phosphate and phosphoenolpyruvate for phospholipid and nucleotide synthesis. Carbon sources are proline and glutamine only, and the majority of amino acids must be imported from the host by membrane transporters [4].

E. chaffeensis and A. phagocytophilum evade intracellular defenses by detoxifying superoxide anions or by preventing assembly of phagocyte oxidase and by delaying normal cellular apoptosis. Many of the altered cellular functions in infected hosts are mediated by direct interactions of cellular proteins with microbial-secreted effector molecules [12,13]. However, there is evidence that microbial molecules that bind host nuclear DNA have a role in reprogramming of host cell gene transcription, ultimately altering cell function [5,6,14-16].

Antigenic characteristics — Serologic cross-reactions between E. chaffeensis and A. phagocytophilum occur. In addition, antibodies in the convalescent sera of humans infected with E. ewingii react with antigens of both E. chaffeensis and E. canis.

Cross-reactivity between Anaplasmataceae spp antigens and rickettsial antigens does not occur. Also, Anaplasmataceae spp do not elicit cross-reacting antibodies with Proteus antigens (Weil-Felix agglutinins).

All strains of E. chaffeensis and A. phagocytophilum have modest antigenic differences among strains that result from genetic variation, post-translational modifications, and perhaps variation in expression of the omp1/p28 and msp2/p44 multifamily genes encoded in their genomes. Genetic diversity varies by geographical region, and as a result, there are greater serologic differences among strains from distinct geographical regions.

The A. phagocytophilum genome contains at least 100 genes encoding Msp2/p44. As each organism expresses only a single Msp2 protein at a time, a population of organisms can produce substantial antigenic diversity and evade immune response by switching expression to an antigenically distinct paralog after negative selection by host immune response. Likewise, there are 22 tandemly arranged omp1/p28 genes on the circular genome of E. chaffeensis that are differentially expressed or undergo a variety of post-translational modifications [17]. This likely contributes to immune evasion and persistence among natural coadapted hosts, such as ticks, small mammals, and deer [18], but this is unlikely to contribute significantly in humans, in whom infection does not persist. There is evidence that T cell exhaustion could also contribute to persistent infection; whether this contributes to aberrant immune pathology or protection is unclear [19].

Cell specificity — Depending upon the species, the Ehrlichia or Anaplasma that cause human disease infect circulating granulocytes, monocytes, or tissue macrophages [20-25]. Individual species of Ehrlichia characteristically have different, highly specific cellular tropisms in vivo. As examples:

A. phagocytophilumA. phagocytophilum infects mostly neutrophils. Binding and entry into host cells in vitro involves several surface-exposed molecules, including the multigene family major surface protein (Msp2/p44), OmpA, Asp14, and AipA, mostly through interactions with neutrophil surface glycans and glycoproteins [26,27]. This interaction permits introduction of A. phagocytophilum protein AnkA via the type IV secretion apparatus that bridges the bacterial and cell membranes. AnkA interacts with and is phosphorylated by Abl-1 after cell entry. This interaction is critical for pathogen entry into the neutrophil vacuole [28].

Vacuoles that sustain replication of A. phagocytophilum accumulate a range of bacterial-secreted proteins that probably play a role in intracellular survival, vacuole trafficking, and acquisition of nutrients. At least one A. phagocytophilum protein (AnkA), enters the nucleus where it alters gene expression programs [6,14], perhaps including those that govern respiratory burst, apoptosis, and proinflammatory response.

Although monocytes resist infection by A. phagocytophilum, it is likely that induction of an innate immune response and activation of the inflammasome within monocytes and macrophages together mediate the inflammatory response that is the clinical hallmark of human A. phagocytophilum infection, while altering immune responsiveness of T lymphocytes for adaptive immunity [29-32].

E. chaffeensis Individual species of Ehrlichia characteristically have different, highly specific cellular tropisms in vivo. E. chaffeensis infects mostly monocytes and macrophages in vivo. As with A. phagocytophilum, it expresses specific surface adhesion proteins (eg, TRP120, EtpE) that mediate binding to monocyte and macrophage selectins or to DNase X in regions associated with lipid rafts [33-35]. Once within the vacuole, E. chaffeensis secretes protein effectors via its type IV secretion apparatus into the host cell that in turn influence intracellular signaling, vacuolar trafficking, apoptosis, and regulation of transcriptional programs that favor survival long enough for acquisition in a subsequent tick bite [12].

Other species, such as Neorickettsia risticii, infect intestinal epithelial cells in horses; Candidatus N. mikurensis likely infects endothelial cells, and "Anaplasma capra" likely infects erythrocytes. In contrast, the host target cells for infection by the E. muris subsp eauclairensis and the Panola Mountain Ehrlichia have not been specifically identified.

MECHANISMS OF DISEASE — Laboratory-based investigation of pathogenetic mechanisms has primarily been focused on E. chaffeensis and A. phagocytophilum infections. Despite genetic, antigenic, and morphologic similarities, these studies suggest that disease mechanisms caused by these two pathogens are distinct. Thus, they are discussed separately below.

E. chaffeensis and human monocytic ehrlichiosis — After introduction into the skin through a tick bite, E. chaffeensis likely spreads via the lymphatics or blood to mononuclear phagocyte-enriched tissues, such as blood, bone marrow, liver, lymph nodes, and spleen. Human monocytic ehrlichiosis does not produce endothelial cell infection manifesting as vasculitis and thrombosis, which are the hallmarks of spotted fever rickettsial infections such as Rocky Mountain spotted fever. (See "Human ehrlichiosis and anaplasmosis" and "Clinical manifestations and diagnosis of Rocky Mountain spotted fever" and "Other spotted fever group rickettsial infections".)

Bone marrow aspirates in patients with E. chaffeensis infection demonstrate hypercellularity of all hematopoietic elements, despite clinical findings such as leukopenia, thrombocytopenia, and anemia [25]. Other histopathologic findings include: noncaseating granulomas or mononuclear phagocyte-rich infiltrates, sometimes with hemophagocytosis in bone marrow; focal hepatocellular apoptosis ranging from mild lobular hepatitis and cholestasis to severe hepatic lobular necrosis; hypercellularity and dissolution of follicular architecture in lymph nodes and spleen, including occasional regions of marked apoptosis or necrosis; and lymphohistiocytic inflammatory infiltrates in tissues including the meninges, brain, and pulmonary interstitium [25,36-39].

The discordance between the marked degree of histopathologic injury that can occur in human monocytic ehrlichiosis and the relatively small burden of bacteria in tissues in infected hosts suggests that human disease is secondary to immunopathologic mechanisms [40]. This is supported by studies in murine models of infection, where infection leads to marked increases in proinflammatory and anti-inflammatory cytokine production and diminished Th1 cytokine production [41]. In addition, TNFa produced by natural killer T (NKT) and CD8 T lymphocytes results in killing of CD4 T cells, and therefore reduced production of IFN-gamma, a key cytokine for control of intracellular E. chaffeensis infection [42,43]. E. chaffeensis may also directly modulate host defense genes of the infected macrophage, or trigger inflammasome activation and marked cellular injury via pathogen- or damage-associated molecular patterns [44,45]. An increasing number of severe infections meet the criteria for diagnosis of hemophagocytic lymphohistiocytosis [46-48]. (See "Clinical features and diagnosis of hemophagocytic lymphohistiocytosis".)

A. phagocytophilum and human granulocytic anaplasmosis — A. phagocytophilum gains access to neutrophils that are recruited to the site of the tick bite before entering the vasculature and disseminating hematogenously [49]. Thereafter, infected cells predominantly remain in the intravascular compartment.

A. phagocytophilum alters normal neutrophil cell functions and their transcriptional programs. As a result, infected neutrophils no longer bind to or emigrate effectively through endothelial barriers and have defects in phagocytosis and intracellular killing. However, infected neutrophils still become activated, degranulate, and produce proinflammatory responses and chemokines that recruit new inflammatory host cells to sustain infection [50].

The histopathologic findings seen in human granulocytic anaplasmosis (HGA) include: normocellular to hypercellular bone marrow despite leukopenia, thrombocytopenia, and anemia; periportal lymphohistiocytic inflammatory infiltrates with focal lobular hepatic lesions and hepatocyte apoptosis (explaining increased serum hepatic transaminase activities); increased splenic red pulp cellularity and focal splenic necrosis; mild to moderate interstitial pneumonitis; and pulmonary hemorrhage. Hemophagocytosis is occasionally observed in spleen, bone marrow, and liver sinusoids, but vasculitis, granulomas, and meningeal inflammation are rare or absent [51].

Some of the clinical findings (eg, fever, leukopenia, and thrombocytopenia) and histologic features (lymphohistiocytic infiltrates) seen in HGA are suggestive of the macrophage activation or hemophagocytic syndrome. As an example, a study of 29 patients with HGA showed that the severity of disease in infected patients was correlated with higher serum levels of triglycerides, ferritin, interleukin-12 p70, and interleukin-10:IFN-gamma ratios compared with matched, uninfected controls [52]. In addition, several patients with severe disease fulfilled the diagnostic criteria for macrophage activation syndrome. Defects in lymphocyte cytotoxicity (CD8 T cell, NK, and NKT cells) in mouse models suggest infection-related dysfunction of MCH class I antigen-presenting cells in vivo [29,31,32,50]. Detailed discussions of HGA and hemophagocytic syndrome are presented separately. (See "Human ehrlichiosis and anaplasmosis", section on 'Clinical manifestations' and "Clinical features and diagnosis of hemophagocytic lymphohistiocytosis".)

IMMUNE RESPONSES — Specific antibody responses are elicited following infection for all members of the Anaplasmataceae family. Pretreatment with convalescent antibodies protects mice with severe combined immunodeficiency from fatal experimental E. chaffeensis infection and partially protects wild-type mice from experimental infection with A. phagocytophilum [53,54].

However, for full protection and recovery from infection, a contribution of cellular immunity by CD4 T lymphocytes is also required [41,55]. For some infected hosts, antibodies may not coincide with the disappearance of organisms or protective immunity. As an example, dogs infected with E. canis, deer infected by E. chaffeensis, small mammals infected with A. phagocytophilum, and ruminants infected by A. marginale can continue to have bacteremia for months or years after acute infection and the appearance of pathogen-specific antibodies [56-59].

SUMMARY

Ehrlichia and Anaplasma spp are obligate intracellular bacteria that grow within vacuoles in human and animal leukocytes. (See 'General characteristics' above.)

Ehrlichia and Anaplasma spp are classified in the Rickettsiales order and Anaplasmataceae family and are among five genera (Ehrlichia, Anaplasma, Neoehrlichia, Neorickettsia, Wolbachia), for which individual species can cause or contribute to human disease (table 1). (See 'Taxonomy and phylogeny' above.)

Individual species of the Anaplasmataceae family infect a wide range of hosts, and disease can be recognized in humans, dogs, cats, sheep, goats, cattle, and horses (table 1). The most important causes of human disease are members of the genus Ehrlichia and Anaplasma. The majority of recognized human infections are caused by A. phagocytophilum and E. chaffeensis. (See 'Host specificity' above.)

Individual organisms enter host cells by inducing endocytosis after introducing bacterial proteins into the host cells. The organisms redirect the vacuole from lysosome fusion by modifying intracellular trafficking to allow microbial growth and division. Within a few days, small, tightly packed clusters of organisms are observable as intracellular inclusions mostly in leukocytes. (See 'Characteristics of Anaplasmataceae' above.)

The major target cells of E. chaffeensis, which causes human monocytic ehrlichiosis, are monocytes and macrophages. The major target cell of A. phagocytophilum, which causes human granulocytic anaplasmosis, and for E. ewingii is the neutrophil. (See 'Cell specificity' above and 'Mechanisms of disease' above.)

Disease severity in Ehrlichia and Anaplasma infections appears to be related to immune pathology and innate immune responses. (See 'Mechanisms of disease' above and 'Immune responses' above.)

  1. Dumler JS, Rikihisa Y, Dasch GA. Genus I. Anaplasma. In: Bergey's Manual of Systematic Bacteriology, 2, Garrity GM, Brenner DJ, Krieg NR, Staley JT (Eds), Springer, East Lansing 2005. Vol 2, p.117.
  2. Dumler JS, Walker DH. Order II: Rickettsiales. In: Bergey's Manual of Systematic Bacteriology, 2, Garrity GM, Brenner DJ, Krieg NR, Staley JT (Eds), Springer, New York 2005. Vol 2, p.96.
  3. Slatko BE, Luck AN, Dobson SL, Foster JM. Wolbachia endosymbionts and human disease control. Mol Biochem Parasitol 2014; 195:88.
  4. Dunning Hotopp JC, Lin M, Madupu R, et al. Comparative genomics of emerging human ehrlichiosis agents. PLoS Genet 2006; 2:e21.
  5. Zhu B, Kuriakose JA, Luo T, et al. Ehrlichia chaffeensis TRP120 binds a G+C-rich motif in host cell DNA and exhibits eukaryotic transcriptional activator function. Infect Immun 2011; 79:4370.
  6. Rennoll-Bankert KE, Garcia-Garcia JC, Sinclair SH, Dumler JS. Chromatin-bound bacterial effector ankyrin A recruits histone deacetylase 1 and modifies host gene expression. Cell Microbiol 2015; 17:1640.
  7. Troese MJ, Kahlon A, Ragland SA, et al. Proteomic analysis of Anaplasma phagocytophilum during infection of human myeloid cells identifies a protein that is pronouncedly upregulated on the infectious dense-cored cell. Infect Immun 2011; 79:4696.
  8. Zhang JZ, Popov VL, Gao S, et al. The developmental cycle of Ehrlichia chaffeensis in vertebrate cells. Cell Microbiol 2007; 9:610.
  9. McKechnie DB, Slater KS, Childs JE, et al. Survival of Ehrlichia chaffeensis in refrigerated, ADSOL-treated RBCs. Transfusion 2000; 40:1041.
  10. Kalantarpour F, Chowdhury I, Wormser GP, Aguero-Rosenfeld ME. Survival of the human granulocytic ehrlichiosis agent under refrigeration conditions. J Clin Microbiol 2000; 38:2398.
  11. Goel R, Westblade LF, Kessler DA, et al. Death from Transfusion-Transmitted Anaplasmosis, New York, USA, 2017. Emerg Infect Dis 2018; 24:1548.
  12. Rikihisa Y. Molecular events involved in cellular invasion by Ehrlichia chaffeensis and Anaplasma phagocytophilum. Vet Parasitol 2010; 167:155.
  13. Truchan HK, Seidman D, Carlyon JA. Breaking in and grabbing a meal: Anaplasma phagocytophilum cellular invasion, nutrient acquisition, and promising tools for their study. Microbes Infect 2013; 15:1017.
  14. Garcia-Garcia JC, Rennoll-Bankert KE, Pelly S, et al. Silencing of host cell CYBB gene expression by the nuclear effector AnkA of the intracellular pathogen Anaplasma phagocytophilum. Infect Immun 2009; 77:2385.
  15. Sinclair SH, Rennoll-Bankert KE, Dumler JS. Effector bottleneck: microbial reprogramming of parasitized host cell transcription by epigenetic remodeling of chromatin structure. Front Genet 2014; 5:274.
  16. Rennoll-Bankert KE, Dumler JS. Lessons from Anaplasma phagocytophilum: chromatin remodeling by bacterial effectors. Infect Disord Drug Targets 2012; 12:380.
  17. Singu V, Liu H, Cheng C, Ganta RR. Ehrlichia chaffeensis expresses macrophage- and tick cell-specific 28-kilodalton outer membrane proteins. Infect Immun 2005; 73:79.
  18. Scorpio DG, Leutenegger C, Berger J, et al. Sequential analysis of Anaplasma phagocytophilum msp2 transcription in murine and equine models of human granulocytic anaplasmosis. Clin Vaccine Immunol 2008; 15:418.
  19. Brown WC, Barbet AF. Persistent Infections and Immunity in Ruminants to Arthropod-Borne Bacteria in the Family Anaplasmataceae. Annu Rev Anim Biosci 2016; 4:177.
  20. Chen SM, Dumler JS, Bakken JS, Walker DH. Identification of a granulocytotropic Ehrlichia species as the etiologic agent of human disease. J Clin Microbiol 1994; 32:589.
  21. Bakken JS, Krueth J, Wilson-Nordskog C, et al. Clinical and laboratory characteristics of human granulocytic ehrlichiosis. JAMA 1996; 275:199.
  22. Goodman JL, Nelson C, Vitale B, et al. Direct cultivation of the causative agent of human granulocytic ehrlichiosis. N Engl J Med 1996; 334:209.
  23. Buller RS, Arens M, Hmiel SP, et al. Ehrlichia ewingii, a newly recognized agent of human ehrlichiosis. N Engl J Med 1999; 341:148.
  24. Fishbein DB, Sawyer LA, Holland CJ, et al. Unexplained febrile illnesses after exposure to ticks. Infection with an Ehrlichia? JAMA 1987; 257:3100.
  25. Dumler JS, Dawson JE, Walker DH. Human ehrlichiosis: hematopathology and immunohistologic detection of Ehrlichia chaffeensis. Hum Pathol 1993; 24:391.
  26. Herron MJ, Nelson CM, Larson J, et al. Intracellular parasitism by the human granulocytic ehrlichiosis bacterium through the P-selectin ligand, PSGL-1. Science 2000; 288:1653.
  27. Seidman D, Hebert KS, Truchan HK, et al. Essential domains of Anaplasma phagocytophilum invasins utilized to infect mammalian host cells. PLoS Pathog 2015; 11:e1004669.
  28. Lin M, den Dulk-Ras A, Hooykaas PJ, Rikihisa Y. Anaplasma phagocytophilum AnkA secreted by type IV secretion system is tyrosine phosphorylated by Abl-1 to facilitate infection. Cell Microbiol 2007; 9:2644.
  29. Walker DH, Dumler JS. The role of CD8 T lymphocytes in rickettsial infections. Semin Immunopathol 2015; 37:289.
  30. Chen G, Severo MS, Sakhon OS, et al. Anaplasma phagocytophilum dihydrolipoamide dehydrogenase 1 affects host-derived immunopathology during microbial colonization. Infect Immun 2012; 80:3194.
  31. Choi KS, Webb T, Oelke M, et al. Differential innate immune cell activation and proinflammatory response in Anaplasma phagocytophilum infection. Infect Immun 2007; 75:3124.
  32. Scorpio DG, Choi KS, Dumler JS. Anaplasma phagocytophilum-Related Defects in CD8, NKT, and NK Lymphocyte Cytotoxicity. Front Immunol 2018; 9:710.
  33. Zhang JZ, McBride JW, Yu XJ. L-selectin and E-selectin expressed on monocytes mediating Ehrlichia chaffeensis attachment onto host cells. FEMS Microbiol Lett 2003; 227:303.
  34. Mohan Kumar D, Yamaguchi M, Miura K, et al. Ehrlichia chaffeensis uses its surface protein EtpE to bind GPI-anchored protein DNase X and trigger entry into mammalian cells. PLoS Pathog 2013; 9:e1003666.
  35. Lin M, Rikihisa Y. Obligatory intracellular parasitism by Ehrlichia chaffeensis and Anaplasma phagocytophilum involves caveolae and glycosylphosphatidylinositol-anchored proteins. Cell Microbiol 2003; 5:809.
  36. Walker DH, Dumler JS. Human monocytic and granulocytic ehrlichioses. Discovery and diagnosis of emerging tick-borne infections and the critical role of the pathologist. Arch Pathol Lab Med 1997; 121:785.
  37. Sehdev AE, Dumler JS. Hepatic pathology in human monocytic ehrlichiosis. Ehrlichia chaffeensis infection. Am J Clin Pathol 2003; 119:859.
  38. Dunn BE, Monson TP, Dumler JS, et al. Identification of Ehrlichia chaffeensis morulae in cerebrospinal fluid mononuclear cells. J Clin Microbiol 1992; 30:2207.
  39. Dierberg KL, Dumler JS. Lymph node hemophagocytosis in rickettsial diseases: a pathogenetic role for CD8 T lymphocytes in human monocytic ehrlichiosis (HME)? BMC Infect Dis 2006; 6:121.
  40. Dumler JS. Anaplasma and Ehrlichia infection. Ann N Y Acad Sci 2005; 1063:361.
  41. Ismail N, Walker DH, Ghose P, Tang YW. Immune mediators of protective and pathogenic immune responses in patients with mild and fatal human monocytotropic ehrlichiosis. BMC Immunol 2012; 13:26.
  42. Stevenson HL, Crossley EC, Thirumalapura N, et al. Regulatory roles of CD1d-restricted NKT cells in the induction of toxic shock-like syndrome in an animal model of fatal ehrlichiosis. Infect Immun 2008; 76:1434.
  43. Ismail N, Crossley EC, Stevenson HL, Walker DH. Relative importance of T-cell subsets in monocytotropic ehrlichiosis: a novel effector mechanism involved in Ehrlichia-induced immunopathology in murine ehrlichiosis. Infect Immun 2007; 75:4608.
  44. Luo T, Kuriakose JA, Zhu B, et al. Ehrlichia chaffeensis TRP120 interacts with a diverse array of eukaryotic proteins involved in transcription, signaling, and cytoskeleton organization. Infect Immun 2011; 79:4382.
  45. Tominello TR, Oliveira ERA, Hussain SS, et al. Emerging Roles of Autophagy and Inflammasome in Ehrlichiosis. Front Immunol 2019; 10:1011.
  46. Agudelo Higuita NI, Yuen C. Hemophagocytic Lymphohistiocytosis Secondary to Ehrlichia Chaffeensis in Adults: A Case Series From Oklahoma. Am J Med Sci 2021; 361:269.
  47. Singh NS, Pagano AL, Hays AJ, et al. Ehrlichia-induced hemophagocytic lymphohistiocytosis in a pediatric kidney transplant recipient. Pediatr Transplant 2022; 26:e14134.
  48. Hammoud K, Fulmer R, Hamner M, El Atrouni W. Ehrlichiosis-Associated Hemophagocytic Lymphohistiocytosis: A Case Series and Review of the Literature. Case Rep Hematol 2023; 2023:5521274.
  49. Granquist EG, Aleksandersen M, Bergström K, et al. A morphological and molecular study of Anaplasma phagocytophilum transmission events at the time of Ixodes ricinus tick bite. Acta Vet Scand 2010; 52:43.
  50. Dumler JS. The biological basis of severe outcomes in Anaplasma phagocytophilum infection. FEMS Immunol Med Microbiol 2012; 64:13.
  51. Lepidi H, Bunnell JE, Martin ME, et al. Comparative pathology, and immunohistology associated with clinical illness after Ehrlichia phagocytophila-group infections. Am J Trop Med Hyg 2000; 62:29.
  52. Dumler JS, Barat NC, Barat CE, Bakken JS. Human granulocytic anaplasmosis and macrophage activation. Clin Infect Dis 2007; 45:199.
  53. Yager E, Bitsaktsis C, Nandi B, et al. Essential role for humoral immunity during Ehrlichia infection in immunocompetent mice. Infect Immun 2005; 73:8009.
  54. Sun W, IJdo JW, Telford SR 3rd, et al. Immunization against the agent of human granulocytic ehrlichiosis in a murine model. J Clin Invest 1997; 100:3014.
  55. Birkner K, Steiner B, Rinkler C, et al. The elimination of Anaplasma phagocytophilum requires CD4+ T cells, but is independent of Th1 cytokines and a wide spectrum of effector mechanisms. Eur J Immunol 2008; 38:3395.
  56. Hess PR, English RV, Hegarty BC, et al. Experimental Ehrlichia canis infection in the dog does not cause immunosuppression. Vet Immunol Immunopathol 2006; 109:117.
  57. Magnarelli LA, Stafford KC 3rd, Ijdo JW, Fikrig E. Antibodies to whole-cell or recombinant antigens of Borrelia burgdorferi, Anaplasma phagocytophilum, and Babesia microti in white-footed mice. J Wildl Dis 2006; 42:732.
  58. Paddock CD, Yabsley MJ. Ecological havoc, the rise of white-tailed deer, and the emergence of Amblyomma americanum-associated zoonoses in the United States. Curr Top Microbiol Immunol 2007; 315:289.
  59. Kocan KM, de la Fuente J, Blouin EF, et al. The natural history of Anaplasma marginale. Vet Parasitol 2010; 167:95.
Topic 7903 Version 17.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟