Next Article in Journal
More Time, Carrot-and-Stick, or Piling Coffins? Estimating the Role of Factors Overcoming COVID-19 Vaccine Hesitancy in Poland and Lithuania in the Years 2021–2022
Previous Article in Journal
Immunologic and Protective Properties of Subunit- vs. Whole Toxoid-Derived Anti-Botulinum Equine Antitoxin
Previous Article in Special Issue
Vaccination Accelerates Liver-Intrinsic Expression of Megakaryocyte-Related Genes in Response to Blood-Stage Malaria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mouse Models for Unravelling Immunology of Blood Stage Malaria

Department of Pathology, University of Utah, Emma Eccles Jones Medical Research Building, 15 N Medical Drive E, Room 1420A, Salt Lake City, UT 84112, USA
*
Author to whom correspondence should be addressed.
Vaccines 2022, 10(9), 1525; https://doi.org/10.3390/vaccines10091525
Submission received: 10 July 2022 / Revised: 5 September 2022 / Accepted: 6 September 2022 / Published: 14 September 2022
(This article belongs to the Special Issue Research on Malaria Immunology in Animal Models)

Abstract

:
Malaria comprises a spectrum of disease syndromes and the immune system is a major participant in malarial disease. This is particularly true in relation to the immune responses elicited against blood stages of Plasmodium-parasites that are responsible for the pathogenesis of infection. Mouse models of malaria are commonly used to dissect the immune mechanisms underlying disease. While no single mouse model of Plasmodium infection completely recapitulates all the features of malaria in humans, collectively the existing models are invaluable for defining the events that lead to the immunopathogenesis of malaria. Here we review the different mouse models of Plasmodium infection that are available, and highlight some of the main contributions these models have made with regards to identifying immune mechanisms of parasite control and the immunopathogenesis of malaria.

1. The Immune Response to Plasmodium Asexual Blood Stages Dictates Malarial Disease

Malaria is still a significant problem in the world with over 600,000 deaths resulting from 241 million cases in 2021, 95% of them concentrated in the African subcontinent [1]. The RTS, S Mosquirix™ vaccine in children has limited efficacy [2,3] but current efforts in improving this vaccine appear to be moving towards vaccine-mediated protection that is more durable [4]. Nonetheless, any further improvements in the development of efficacious therapeutics and vaccines require a better understanding of what constitutes an effective anti-malarial immune response.
Malaria is caused by infection with parasites of the species Plasmodium, deposited into the dermis of the skin by female Anopheles mosquitos while probing for a blood meal. The sporozoites travel through the blood circulation, invading the liver where they undergo several rounds of asexual division in hepatocytes before being released as merozoites into the blood stream. Whilst hepatocytes infected with P. falciparum always transition to liver schizogony, some species such as P. vivax and P. ovale can also differentiate to become a dormant stage known as a hypnozoite that can reactivate to cause malaria relapses [5]. Upon release into the blood stream parasites infect red blood cells (iRBCs) and replicate every 24–72 h depending on species. Clinical symptoms of malaria are exclusively caused by the erythrocytic lifecycle of Plasmodium.
The clinical manifestations of malaria are wide-ranging and include symptoms such as hypoglycemia, acidosis and anemia. Accumulation and sequestration of iRBCs on vascular endothelial cells is associated with vascular activation which is known to underlie organ-specific pathologies such as cerebral malaria, acute lung injury, hepatomegaly and liver fibrosis [6] (Table 1). While sterile immunity to malaria generally does not occur, years of repeated exposure to the parasite in endemic regions facilitates the development of clinical immunity that can be characterized by reduced parasite load (anti-parasite immunity) and controlled inflammatory responses to iRBCs (clinical immunity). Malarial disease encompasses a spectrum of virulence which is influenced by a number of factors including genetic variation of both the host and parasite [7,8,9], the make-up of the intestinal microbiome of the host [10] and environmental influences such transmission intensity or the presence of co-infections [11,12,13].
Table 1. The main clinical manifestations of Plasmodium infection.
Table 1. The main clinical manifestations of Plasmodium infection.
Disease ManifestationSpecies of PlasmodiumProbable MechanismSeverityReferences
FeverAll speciesSchizogony-induced inflammation from 24–72 h depending on parasite species. Mediated by endogenous pyrogens induced during schizogeny such as TNF-α, IL-6, IL-1β and prostaglandin E2).Not generally lethal[14]
AnemiaAll speciesErythrophagocytosis.
Dyserythropoiesis.
RBC destruction by parasite replication.
Can be lethal[15]
Cerebral malariaP. falciparum (children)
P. falciparum and P. vivax (Adults)
Vascular activation via parasite sequestration on the brain endothelium, followed by breakdown of the blood brain barrier, vasogenic odema and herniation.20% mortality[16,17,18,19]
Malaria-associated acute respiratory distress and Malaria-associated Acute Lung InjuryP. falciparum,
P. vivax
P. knowlsei
P. ovale
Vascular activation via parasite sequestration on the pulmonary endothelial followed by pulmonary leak.40% mortality[20,21,22]
Hepatomegaly/Liver fibrosisP. falciparum
P. vivax
P. ovale
Jaundice and hepatic dysfunction due to infiltration of iRBCs and sequestration of iRBCs in the liver. Results in activation of hepatic stellate cells to become myofibroblasts.Normally an indicator of severe malaria[23,24]
Acute Kidney Injury (AKI)P. falciparum
P. vivax
P. malariae
Glomerulonephritis, acute tubular necrosis and acute interstitial nephritis due to hemodynamic dysfunction and inflammation. Results in proteinuria, microalbuminuria and urinary casts along with hemolytic-uremic syndrome. Contributes to metabolic acidosis and can be exacerbated by liver damage.Normally an indicator of severe malaria and found in around 40% of those with severe disease[25]
Lactic acidosisP. falciparumTissue hypoperfusion and hypoxia resulting from capillary obstruction with sequestered iRBCs and anemia
Production of lactate by iRBCs.
Impaired lactate clearance by the liver and kidney.
Normally an indicator of severe malaria[26]
HypoglycemiaP. falciparum
P. vivax
Illness-induced fasting and inhibition of gluconeogenesis.An indicator of severe malaria and more common in children than adults. Predicts mortality in malaria[27]
Mouse models are commonly used to study the immunology of erythrocytic malaria. Given the well-characterized range of different rodent Plasmodium species and strains, as well as the plethora of mouse lines currently available to investigators, this article will outline some of the parasite-mouse combinations that are commonly used to study the different facets of blood stage malaria immunology. In addition, we will discuss novel models of rodent malaria that have not yet been fully harnessed to determine the environmental and genetic contributions to generating immune responses to Plasmodium iRBCs.

2. Utility of Rodent Plasmodium Species in the Investigation of Blood Stage Immunology

Human parasites cannot infect mice unless the mice are genetically humanized [28,29]. Whilst humanized mouse models have some utility in the investigation of immune responses to P. falciparum in a controlled environment, several species of Plasmodium exist that naturally infect rodents (Table 2). Isolated and cloned from Thamnomys thicket rats in the Central African Region in the 1960s [30], they have been instrumental in the study of the immunobiology of the erythrocytic stages of Plasmodium infection [31]. Although apparently asymptomatic in their original hosts, infection of mice gives rise to a number of different phenotypes of infection, many of which mimic various states of disease found in human Plasmodium infection. Rodent Plasmodium parasites cannot infect humans making them tractable and non-hazardous models of malaria. However rodent Plasmodium parasites have some differences to human Plasmodium parasites such as the variant antigen gene families expressed in the blood stage. Rodent Plasmodium parasites do not have var genes encoding P. falciparum erythrocyte membrane protein-1 (PfEMP1) [32,33], nor genes encoding the subtelomeric variant open reading frame (STEVOR) [34] or P. falciparum-encoded repetitive interspersed families of polypeptides (RIFINS) [35]. Instead, rodent Plasmodium parasites rely on genes encoded by Plasmodium interspersed repeat (pir) genes [36,37,38,39] which are the largest multigene family in many Plasmodium species [39]. It is important to note that no single rodent Plasmodium species replicates all features of human Plasmodium infection. Therefore, specificity of the focus of a particular study in combination with the correct choice of model is a key aspect of research into blood stage malaria immunology using rodent models of malaria.
Table 2. Disease phenotype and pathophysiology of the main rodent Plasmodium species used in biomedical research.
Table 2. Disease phenotype and pathophysiology of the main rodent Plasmodium species used in biomedical research.
SpeciesCloneRBC PreferencePhenotypesReferences
P. bergheiANKAReticulocyte preference but will invade normocytesAsynchronous life cycle, sequesters in the liver, lung and brain. Evidence of weight loss and anemia normally present at the time of death.
C57BL/6J: lethal infection with breakdown of the BBB and death between day 7–10 p.i. Pronounced pulmonary pathology.
BALB/c: Death from hyperparasitemia. No discernible cerebral complications. Less extensive lung pathology.
DBA/2J: No discernible cerebral complications. some pulmonary pathology but less pronounced than BALB/c mice. Death ~day 20 p.i from hyperparasitemia and anemia.
Pet shop mice: resistant to death by cerebral malaria. Death ~day 20 p.i. from hyperparasitemia.
[40,41,42,43]
NK65
New York (NY)
Reticulocyte preferenceAccumulates mostly in the lung, with very little accumulation in the brain. Causes anemia over the course of infection.
C57BL/6: death in ~20 days from respiratory distress.
BALB/c: no development of MA-ARDS.
[44,45]
NK65 Edinburgh (E)Normocytes and ReticulocytesAccumulates in the lung but not the brain with some evidence of anemia.
C57BL/6: death in 7–10 days from respiratory distress. Early increase in peripheral parasitemia.
BALB/c: Resistant to respiratory distress upon infection.
[44,45]
K173Reticulocyte preferenceVery little parasite accumulation/sequestration in the brain.
Does not produce gametocytes due to laboratory adaptation from passaging.
C57BL/6: Used as model for cerebral malaria. Early death after infection due to cerebral pathology accompanied with very high parasitemia. Causes lung pathology with increased pulmonary oedema.
[46,47]
P. yoeliiXL (also known as 17XL)Normocytes and reticulocytesC57BL/6: Lethal within ~10 days p.i. due to hyperparasitemia and severe anemia.
BALB/c: Lethal within ~10 days p.i.
DBA/2: Non-lethal infection.
[48]
XNL (also known as 17XNL)Strong reticulocyte preferenceC57BL/6: Resolving non-lethal infection accompanied by anemia.
BALB/c: Resolving non-lethal infection.
[48,49]
YMNormocytes and reticulocytesC57BL/6: Derivative of the XL line. Lethal within ~10 days p.i. due to hyperparasitemia and severe anemia.
DBA/2: Lethal infection in ~10 days p.i.
B10: Non-lethal infection.
[50,51,52,53]
nigeriensis N67Normocytes and reticulocytesC57BL/6: Lethal at ~15–20 days p.i. due to hyperparasitemia.[50,52,54]
nigeriensis N67CNormocytes and reticulocytesC57BL/6: Lethal within 7 days p.i.[50,54,55]
P. chabaudichabaudi ASNormocytes and reticulocytesSynchronous life cycle. Sequesters predominantly in the lung and liver.
C57BL/6: Resolving non-lethal infection accompanied by anemia, thrombocytopenia, hypoglycemia, weight loss and hypothermia. Recrudescent infections and sub-patent for up to 3 months.
BALB/c: More severe infection than in C57BL/6J mice but generally non-lethal in most BALB/c lines.
A/J mice: Lethal anemia due to poor control of iRBCs and hyperparasitemia.
[31,56]
chabaudi BCNormocytes and reticulocytesC57BL/6: similar symptoms to P. chabaudi AS but more severe.[57]
chabaudi CBNormocytes and reticulocytesC57BL/6: similar symptoms to P. chabaudi AS but more severe.[58]
chabaudi ERNormocytes and reticulocytesC57BL/6: Similar symptoms to those of P. chabaudi AS. Recrudescent infections occurs at 20 to 25 days p.i.[31,57,59]
chabaudi adamiPreference for younger normocytes
over reticulocytes
BALB/c mice: Non-lethal infection with single peak of infection around 10 days p.i.
A/J mice: Non-lethal resolving infection.
C57BL//6 mice: Non-lethal resolving infection.
[60,61,62,63,64]
P. vinckeivinckei CYNormocytes
Not thought to invade reticulocytes
CBA: Lethal infection by 6 days p.i. with hyperparasitemia.[65]
vinckei ATCC 30091Normocytes
Not thought to invade reticulocytes
ICR outbred mice: Lethal infection within 8–10 p.i.[66]
petteri ASNormocytes
Not thought to invade reticulocytes
AKR mice: Lethal in 5 days p.i. due to fast growing parasites.[67,68]
petteri BSNormocytes
Not thought to invade reticulocytes
ICR outbred mice: Non-lethal infection with a peak of parasitemia at 9 days p.i.
CBA: Non-lethal infection with a peak of parasitemia at 9 days p.i.
[65,66]
petteri ARNormocytes
Not thought to invade reticulocytes
AKR mice: Non-lethal with patent parasitemia not detectable by 22 days p.i.[67,68]
petteri CRNormocytes
Not thought to invade reticulocytes
BALB/c: single peak of non-lethal infection.
CBA: single peak at 6 days p.i.; non-lethal infection.
[65,69]
petteri HWNormocytes
Not thought to invade reticulocytes
C57BL/6: Lethal infection at 8–10 days p.i. from hyperparasitemia.[70]
Abbreviations: BBB: blood-brain barrier; MA-ARDS: malaria-associated acute respiratory distress syndrome; p.i.: post-infection.
Mice have been instrumental in elucidating the workings of the human immune system [71]. Nonetheless, there are fundamental differences between the physiology of mice and humans that should be noted such as a different balance of leukocyte subsets [72] between both species as well as in splenic architecture where human (sinusoidal) and mouse (non-sinusoidal) [73] differences would alter the direction of blood flow and possibly the timing or mechanisms by which iRBCs induce splenic immune responses [73,74]. Splenic sinusoids are blood vessels in the red pulp that drain into the white pulp of the spleen where most splenic immune cells reside. In humans the endothelial cells lining these sinuses form a barrier that RBCs have to squeeze through [74,75], a challenge for Plasmodium-iRBCs which are more rigid than uninfected RBCs. On the other hand, the mouse spleen endothelial cells have bigger gaps making it more likely for iRBCs to end up in the white pulp with a potentially expedited adaptive response that may be based on a higher antigenic load compared to humans with Plasmodium infection. Human spleens also lack a substantial marginal zone [73], a B-cell enriched layer that also contains unique macrophage populations that surrounds the follicles and periarteriolar sheath separating the white and red pulp.
Circulating Plasmodium parasites have been found to alter expression of variant antigens expressed on the RBC surface in both splenectomized squirrel monkeys infected with P. falciparum [76] and splenectomized mice infected with P. chabaudi [77] suggesting a splenic response occurs in primates and rodents infected with Plasmodium. Disorganization of the white pulp in both human autopsy samples of malaria fatalities [78] and in mice infected with P. chabaudi [79,80] also suggests involvement of the spleen in both species. However, the structural differences that exist may impact the timing of immunological events and possibly the importance of different cell types in orchestrating the immune response. As such, some caution should be exercised when extrapolating findings from rodent Plasmodium blood stage models with respect to the main players involved.
Despite these differences, the main features of the immune response to Plasmodium iRBCs (Figure 1, Table 3) are largely replicated with a strong inflammatory response characterized by interferon-γ (IFN-γ) producing CD4 T cells and the production of anti-parasite antibodies. Mouse models of Plasmodium infection provide a tractable and highly informative model to define how the immune system operates in human Plasmodium infection, in turn providing critical evidence of immune mechanisms in malaria that simply cannot be obtained in humans. Advancements in both rodent genetic engineering technology [81,82] and the ability to create transgenic rodent Plasmodium parasites [83,84] (Table 4) has facilitated dissection of immune responses to Plasmodium infection with unprecedented precision. As such, mouse models of blood stage Plasmodium infection are a key tool in understanding the immune responses driving Plasmodium parasite control and pathogenesis of malaria.
Selection of a mouse host and parasite species to study immune responses to blood stage Plasmodium infection is dependent on the question being asked. Some mouse-parasite combinations are lethal from around 7 days post-infection whereas others resolve to become a sub-patent infection that can only be detected by molecular methods and, in some cases, can be completely cleared. When selecting which combination to use it is important to determine whether the major goal of any study is to decipher anti-parasite immune responses, clinical immunity or a combination of both. Other considerations may involve the existence of comparative literature, the availability of transgenic tools (Table 4) or the existence of databases from “big data” sets available online (Table 5) that can be mined a priori to identify candidate molecules of importance.
Table 3. The main immunological features of the three most widely used Plasmodium species in the study of the immune response elicited by Plasmodium-infected RBCs.
Table 3. The main immunological features of the three most widely used Plasmodium species in the study of the immune response elicited by Plasmodium-infected RBCs.
Rodent SpeciesModel UsesMain FeaturesReferences
Plasmodium chabaudiInnate immune responsesControl of iRBCs by monocytes and γδ T cells, but not neutrophils or NK cells.
Direct activation of DC for activation of T cell responses.
[85,86,87,88,89,90,91,92]
T cell responsesRequires T cells for control of iRBCs.[59]
Immune regulationIL-10 is required for clinical immunity.
TGF-β provides some protection against pathogenesis.
[93,94]
Generation of humoral immunityParticipation of both IgG and IgM in control of iRBCs.
Does not require antibodies for control of acute infection with iRBCs.
Antibodies contribute to control of iRBCs during chronic infection.
[11,95,96,97]
T and B cell memory responsesGenerates memory T and B cell responses that expand upon secondary challenge infection.[98,99,100]
Immune basis of clonal virulenceClonal virulence is associated with differences in the immune response induction.[101,102]
Host genetic basis of immune resistance to infection by iRBCsGenetic control of host immune responses mediates immunological control of iRBCs and level of clinical immunity.[103,104]
Plasmodium yoeliiInnate immune responsesMacrophages are protective against iRBCs.
Minimal contribution of neutrophils,γδ-T cells or NK cells to control of iRBCs.
[105,106,107]
T cell responsesRequires T cell for control of iRBCs.[108]
Immune regulationIL-10 and TGF-β are required for clinical immunity.[93,109]
Generation of humoral immunityRequires antibodies for control of iRBCs during acute infection.[110]
T and B cell memory responsesGenerates memory T and B cell responses that expand upon secondary challenge infection.[111,112]
Immune correlates of lethal vs. non-lethal infectionLethality is correlated with faster parasite growth and an early burst of TGF-β.[113]
Plasmodium bergheiANKACD8 T cell induced vascular leak[114,115]
Pulmonary vascular leakage and leukocyte infiltration[116]
Clonal differences in the induction of experimental cerebral malaria[117]
NK65 EdinburghPulmonary vascular leakage and leukocyte infiltration[45,118]
NK65 New YorkPulmonary vascular leakage and leukocyte infiltration[45]
Table 4. Functionality of selected transgenic rodent Plasmodium parasites and mouse lines available to study of the immune response to Plasmodium-iRBCs. Repositories of rodent malaria strains available for use include the Rodent Malaria genetically modified Data Base (RMgmDB) (https://www.pberghei.eu/index.php), the European Malaria Reagent Repository (bank: http://www.malariaresearch.eu/) and the Malaria Research and Reference Reagent Resource Center (MR4) (https://www.beiresources.org/ProgramInformation.aspx). All links last accessed on 8 September 2022.
Table 4. Functionality of selected transgenic rodent Plasmodium parasites and mouse lines available to study of the immune response to Plasmodium-iRBCs. Repositories of rodent malaria strains available for use include the Rodent Malaria genetically modified Data Base (RMgmDB) (https://www.pberghei.eu/index.php), the European Malaria Reagent Repository (bank: http://www.malariaresearch.eu/) and the Malaria Research and Reference Reagent Resource Center (MR4) (https://www.beiresources.org/ProgramInformation.aspx). All links last accessed on 8 September 2022.
TransgeneFunctionalityTransgenic ParasitesProperties of the Transgenic ParasiteReferences
LuciferaseIn vivo and ex vivo visualization of organ-specific parasite sequestration.
Can be imaged after injection of D-luciferin using an IVIS imager
P. berghei ANKAExpression of luciferase under the eEF1α promoter (constitutive) or the AMA-1 promoter (schizont-specific).[119]
P. berghei NK65 EdinburghExpression of luciferase under the AMA-1 promoter (schizont-specific).[118]
P. chabaudi ASExpression of luciferase under the eEF1α promoter (constitutive).[56]
Allelic replacement of rodent Plasmodium proteins with human Plasmodium proteinsStudy of immune responses to human Plasmodium parasite vaccine targetsP. berghei ANKAExpress P. falciparum merozoite protein-119 (MSP-119).[120]
P. berghei ANKAExpress P. falciparum apical membrane antigen-1 (AMA-1).[121]
P. berghei ANKA
P. berghei NK65
Express P. vivax merozoite protein-119 (MSP-119).[122]
Fluorescent proteinsImaging of parasites ex-vivo by microscopy. Assessment of phagocytosis by flow cytometry or Imagestream.P. chabaudi ASExpression of GFP under the eEF1α promoter.[85]
P. chabaudi AJExpression of RFP under the eEF1α promoter.[83]
P. yoelii XNLExpression of GFP under the eEF1α promoter.[123]
P. berghei ANKAExpression of GFP under the eEF1α promoter.[124]
P. berghei NK65EExpression of GFP under the AMA-1 promoter (schizont-specific).[118]
P. berghei NK65NYExpression of GFP under the AMA-1 promoter (schizont-specific).[45]
Insertion of peptide epitopes recognized by available TCR transgenic miceQuantification of antigen-specific Plasmodium-induced T cells responsesP. yoelii XNLExpress a CD4 and CD8 immunodominant epitope from the glycoprotein LCMV. Can be tracked using SMARTA or P14 TCR transgenic mice.[112]
P. yoelii XNLExpress a CD4 and CD8 immunodominant epitope from the model antigen chicken ovalbumin. Can be tracked using OVA TCR transgenic mice (OT-I and OT-II).[125]
P. berghei ANKAExpress a CD4 and CD8 immunodominant epitope from the model antigen chicken ovalbumin. Can be tracked using OVA TCR transgenic mice (OT-I and OT-II).[126]
Endogenous parasite target epitopeParasite speciesTransgenic Mouse lineProperties of the transgenic mouse lineReferences
RPL6120–127P. berghei ANKAPb-I H2-Kb C57BL/6Transgenic Pb-I CD8 T cells with Vα8.3/Vβ10 TCRs that recognize PbRPL6120–127 in P. berghei ANKA liver stages but cross-react with blood stage antigens in P. berghei ANKA, PcRPLL6130–137 (P. chabaudi) and PyRPL6123–130 (P. yoelii).[127,128]
HSP90484–496P. berghei ANKAPb-II I-Ab C57BL/6Transgenic Pb-II CD4 T cells with Vα2/Vβ12 TCRs that react cross-react to P. berghei NK65, P. chabaudi AS and P. yoelii XNL.[129,130]
37/39 kDa fragment of MSP-11157–1171P. chabaudi ASB5 I-Ed BALB/cContains B5 MSP-1-reactive CD4 T cells with Vα2/Vβ8 TCRs reactive to MSP-11157–1171.[131]
21 kDa fragment of MSP-1P. chabaudi ASIghNIMP23/+ C57BL/6B cells express the NIMP23 Ig heavy chain and harbor B cells that react to the 21 kDa fragment of P. chabaudi AS MSP-1.[132]
Abbreviations: AMA-1: apical membrane antigen 1; eEF1: eukaryotic elongation factor; GFP: green fluorescent protein; HSP: heat shock protein; LCMV: lymphocytic choriomeningitis virus; MSP-1: merozoite surface protein 1; RFP: red fluorescent protein; RPL6: putative ribosomal protein L6.
Table 5. Published and publicly available databases of sequencing data from rodent Plasmodium infections.
Table 5. Published and publicly available databases of sequencing data from rodent Plasmodium infections.
Cell TypeMouse StrainParasite StrainSequencing TypeDatabase Access CodeReferences
Whole Blood and spleenC57BL/6P. chabaudi chabaudi AS and P. chabaudi chabaudi CBMicroarrayGSE93631[133]
Whole Blood and spleenC57BL6P. chabaudi chabaudi ASMicroarrayGSE123391[133]
GSE145781[134]
Bone marrowC57BL/6J and C57BL/6J Elf4−/− miceP. yoelii XNLRNAseqGSE121035[135]
Monocyte derived dendritic cellsC57BL/6P. berghei ANKARNAseqGSE126381[136]
Splenic MacrophagesC57BL/6P. berghei ANKAMicroarrayGSE111593[137]
MacrophagesBALB/cP. yoelii XNL-LucRNAseqGSE115906[138]
Red pulp macrophagesC57BL/6P. chabaudi chabaudi ASMicroarrayGSE23565[139]
NK cellsC57BL/6P. chabaudi chabaudi ASMicroarrayGSE12727[140]
γδ T cellsC57BL/6JP. chabaudi chabaudi AJSingle cell TCR sequencing and RNAseqGSE108478[89]
CD4 T cellsC57BL/6P. berghei ANKAmicroarrayGSE24903[141]
C57BL//6J and
Uba3fl/fl-Lck Cre+ (KO) Uba3fl/fl and Uba3ΔT mice on a C57BL/6J background
P. yoelii XNLRNAseqGSE111066[142]
C57BL/6P. chabaudi chabaudi ASMicroarray
Single-cell RNAseq
GSE81196
GSE81197
[143]
C57BL/6P. yoelii XNLMicroarrayGSE85896[144]
Regulatory T cellsBALB/cP. yoelii XNLMicroarrayGSE34621[49]
B cellsTbx21fl/flCd23Cre and CD23Cre+ mice on a C57BL/6J backgroundP. berghei ANKARNAseq
ATACseq
GSE120729
GSE120727
[145]
IFN-γR1−/− and C57BL/6JP. chabaudi chabaudi ASRNAseqGSE85205[146]
C57BL/6P. yoelii XNLRNAseqGSE134548[110]
C57BL/6P. chabaudi chabaudi ASRNAseqGSE115155[132]
MicrogliaC57BL/6 wildtype and IFNAR−/−P. berghei ANKAMicroarrayGSE119650
GSE86082
[147]
Abbreviations: Cre: cre recombinase; Elf4: E74 Like ETS Transcription Factor 4: IFNAR: interferon-α receptor 1; IFN-γR1: interferon-γ receptor 1; Tbx21: T box transcription factor 21; Uba3: Ubiquitin-Activating Enzyme 3.

3. Genetic Control of the Immune Responses to Plasmodium Infection

Population-wide genetic diversity and its effect on Plasmodium infection is evident in human populations. This can be clearly seen with hemoglobinopathy gene polymorphisms such as sickle cell, thalassemia or glucose-6-phosphate dehydrogenase (G6PD) [148,149,150,151,152] which offer resistance to infection of RBCs by Plasmodium and reduced clinical severity when infection does occur. Associations have been found with MHC haplotype [153,154] with varying results [155], and polymorphisms in immune genes and the promoters controlling their expression have been associated with malarial disease severity. For example, allele variants [156] and promoter polymorphisms [157,158] controlling the expression of the inflammatory cytokine tumor necrosis factor-α (TNF-α) have been associated with the propensity to develop cerebral malaria [157,158] and anemia [157,159].
There is now an increasing number of publications using Genome-Wide Association Studies (GWAS) for malaria [160,161,162,163]. These studies have found associations of polymorphisms encoding an erythrocyte calcium pump (ATP2B4) and an endothelial junction protein (MARVELD3) with severe malaria [162], and linkages to genes on chromosome 6q21.3 and possibly 19p13.12 to the development of uncomplicated (mild) malaria [164]. Linkages to asymptomatic malaria have been found on chromosome 5q31 [164]. Collectively these data suggest that the development of immune responses during Plasmodium infection is, in part, genetically controlled. In support of this notion, the Fulani tribe of western Africa who are generally more resistant to the clinical effects of Plasmodium infection have allelic variants of FcγRIIα [165], interleukin (IL)-10 [166] and IL-4 [167] which are not present in the more susceptible sympatric Dogon tribe [168,169]. It is thought that these variants allow the Fulani to mount a robust and protective immune response to Plasmodium that is characterized by early production of pro-inflammatory cytokines like IFN-γ [170].
In mice Plasmodium infections are also genetically controlled [171,172,173,174]. Between-strain genetic diversity can explain the variation in disease severity in Mus musculus infected with any of the rodent Plasmodium lines. For example, it is widely accepted that C57BL/6 mice are susceptible to P. berghei ANKA (PbA) infection succumbing to neurological manifestations of infection that resemble cerebral malaria, whereas BALB/c or DBA/2J are more resistant and survive significantly longer, dying much later from hyperparasitemia [173]. Linkage studies of P. chabaudi-AS-infected crossed lines of inbred, recombinant inbred and congenic inbred lines of mice has led to the identification of several gene regions termed Char (Chabaudi resistance) regions (reviewed in [103] and [104]) which include immune-associated genes such as tumor necrosis factor (TNF) and lymphotoxin-α (LT-α) amongst others demonstrating the importance of immunogenetics in the outcome of P. chabaudi infection in mice.
The majority of studies elucidating immune responses to the erythrocytic stage of Plasmodium infection are undertaken in genetically-uniform inbred mouse strains, many using mice with C57BL/6 and BALB/c backgrounds (Table 2). Whilst beneficial by virtue of uniform background genetics and MHC haplotype, such mice will not directly inform on the immunogenetic basis of disease severity. Although necessary with respect to studies incorporating genetically modified mouse lines, collectively such studies may bias our understanding of infection immunology due to their highly selected life histories in laboratory settings [175]. Many of these inbred strains have skewed immune responses, such as the Th1 skewed response in C57BL/6 mice or Th2-skewed response in BALB/c mice [176], which have been exploited to understand genetic influences on the differential immune responses mounted to Plasmodium parasites and the severity of infection [172]. As such, it is important to be cognizant of this limitation of mouse models when interpreting data that have been collected.
Several studies have endeavored to use outbred mouse lines, with Swiss Webster mice commonly used. Nonetheless, such mice have inbreeding coefficients of ~0.48 [177] which is rather removed from humans which are ~0.01–0.08 [178,179]. New endeavors to generate mouse resources that are more aligned with human diversity include wild-derived specific pathogen free (SPF) mice [180], or the Collaborative Cross (CC) mice [181,182,183,184] and Diversity Outbred (DO) mice [182,185,186]. These colonies allow assessments of the variation of immune responses to Plasmodium infections that may be more akin to humans. Given that all of these under-utilized mouse colonies are SPF, they have the advantage of assessing genetic diversity on anti-Plasmodium immune responses in the absence of contributing environmental factors.

4. Modelling the Influence of Environmental Factors on Immune Responses to Plasmodium Infection

Genetic diversity in immune responsiveness has been studied in wild-captured mice. However responses in such mice are confounded by environmental factors [187] which include differences in microbiota [188], a community of microorganisms including bacteria, fungi, viruses and protozoans. These microorganisms colonize a number of external facing environments of humans such as the respiratory tract, gastrointestinal tract and skin. Studies in humans and mice have found associations between gut microbiota and the susceptibility to Plasmodium infection [189,190]. Normally dominated by members of the Firmicutes and Bacteroides taxa, the diversity of the bacterial microbiota has been shown to increase upon P. falciparum infection [191,192]. Furthermore, upon infection with Plasmodium the level of malarial disease may be affected by the increased composition of bacterial microbiota. Specifically, it has been shown that severe malarial anemia in P. falciparum-infected children in Uganda is associated with a greater alpha diversity of bacteria in the gut [191].
Mouse models of malaria have been instrumental with respect to parsing out possible mechanisms underlying the effects of bacterial microbiota composition on immune responses to Plasmodium blood stages. In general, bacterial microbiota play a key role in glycan metabolism in the gut giving rise to monosaccharides that are fermented to short-chain fatty acids and used as an energy source for the host, but also have immunomodulatory functions. Differences in the metabolic capacity of the bacterial microbiome correlate with disease susceptibility to P. yoelii XNL infections [193]. Evidence that this may result from the immunomodulatory effects of the bacterial microbiome comes from studies showing that mice that are more resistant to P. yoelii XNL have microbiomes enriched for Lactobacillus and Bifidobactierum [10] and a better humoral response to P. yoelii XNL with a higher magnitude of anti-parasite antibodies produced. Germinal Center (GC) reactions to Plasmodium infection, as discussed below, are a key event resulting in an efficacious anti-Plasmodium humoral response. Highly dynamic in nature, splenic GC reactions have been shown to be continuously modulated by gut microbiota in P. yoelii XNL infections [191,194] demonstrating the utility of mouse models with regards to dissecting mechanisms by which microbiota may influence malarial disease in those infected with Plasmodium. The implications of this work for those undertaking research using mouse models of malaria to dissect the immunology of blood stage infections is the choice of mouse vendor; the severity of erythrocytic Plasmodium infection in genetically similar mice obtained from different vendors is significantly altered in response owing to the differences in gut microbiome [10].
Inbred mice that are removed from an SPF environment and have been exposed to natural environments (“re-wilded mice”) are found to have a different immune landscape modulated by the microbiota [188,195]. Recent work has studied the role of genetic diversity in Plasmodium immune responses in the context of environmental exposures through co-housing specific pathogen free (SPF) mice with mice obtained from pet shops that were not SPF (so-called “dirty mice”) [196]. Influencing the environmental exposure of mice in this way induced a less susceptible phenotype to P. berghei ANKA infections but did not alter immune responses sufficiently to fully protect all mice [196]. However, it should be noted that mice obtained from pet shops are highly inbred and do not recapitulate the genetic diversity conferred by wild-derived, CC or DO mice described above.
In addition to the microbiota, there are other environmental factors that collectively can influence the immune responses to Plasmodium iRBCs that are hard to consistently replicate in laboratory mice. These include the alteration of the immune landscape of humans by prior and current co-infections including the influence of immune responses to existing liver stage Plasmodium parasites [197]. However there has been some success in modelling co-infection scenarios in mice and measuring how immune responses to Plasmodium are influenced when co-infections are present (for examples see [11,198]). In addition, there are likely effects of the human biting rate (HBR) which would alter amounts of mosquito saliva exposure [199,200,201] and has been shown to influence Plasmodium infection in mice [202]. The entomological inoculation rate (EIR) may also differ and be associated with a varying number and/or multiplicity of Plasmodium infection in an individual [203]. Without use of mouse models of blood stage infection where each aspect can be dissected individually, it would be virtually impossible to determine the relative influence each of these environmental factors has on immune responses mounted to blood stage Plasmodium infection.

5. Mechanisms of Anti-Parasite Immunity: What Have We Learnt about Control of iRBCs from Mouse Models of Plasmodium Infection?

Successful control of intraerythrocytic Plasmodium parasites requires a robust cellular and humoral immune response that generates broadly-reactive antibodies. Rodent malaria models of Plasmodium erythrocytic infection have been instrumental in revealing some of the mechanisms governing the cellular immune responses to Plasmodium blood stage parasites, as well as spatial information related to immune responses generated in different organs where Plasmodium iRBCs sequester. It is challenging to obtain this level of information from human infections where the main available source for immune analysis is the peripheral blood. Here we discuss some of the main findings from use of rodent models of Plasmodium blood stage infection.

5.1. Invasion Blocking Is a Key Mechanism of Anti-Parasite Antibodies for the Control of iRBCs

The importance of humoral immunity in host defense against Plasmodium infection was first demonstrated in rhesus monkeys [204] and later in P. falciparum-infected children [205,206] when passive transfer of immune sera limited parasite growth and symptoms associated with the disease. These observations formed the basis of the hypothesis that there is a requirement for sustained antibody production in the control of Plasmodium blood stage of infection where the clinical manifestations of the disease occur.
The possible effector functions of these antibodies have been elucidated with careful in vitro culture studies. These range from recognition and uptake of iRBCs by phagocytes [207,208,209], blocking of parasite adhesion and invasion [210], to inhibition of parasite growth [211]. The targets of these antibodies are numerous and involve proteins expressed on the surface of merozoite required for RBC invasion such as merozoite surface protein 1 (MSP-1) or apical membrane antigen 1 (AMA) [212,213] as well as variant surface antigens such as Plasmodium falciparum erythrocyte membrane protein 1 (PfEMP-1) [214]. Positive correlations between the breadth, as well as magnitude, of the antibody response and successful control of iRBCs [212,213,215] provide further evidence of the importance of the humoral response in controlling Plasmodium iRBCs.
The relative contributions of these mechanisms to parasite control are hard to assign in humans; mouse models of infection have been instrumental in identifying the importance of invasion blocking as a key mechanism of iRBC control in vivo. Studies using FcγR−/− mice which are deficient in the FcγR used by phagocytes to detect IgG-opsonized iRBCs demonstrate that IgG-dependent phagocytosis is not a key mechanism of control of iRBCs, at least in the avirulent P. yoelii XNL model [216]. This conclusion is supported by a recent study whereby in vivo tracking of a single generation of labeled iRBCs of either P. chabaudi or P. yoelii adoptively transferred into mice demonstrated that parasite-specific IgG does not affect the rate at which iRBC are cleared, but rather it limits the progression of the iRBC to a new RBC by blocking invasion [210]. This is not a surprising finding when most IgG was reactive to merozoites found within schizonts, the terminal stage of iRBCs prior to release of merozoites that will infect new RBCs. The observation that infection of mice with P. yoelii XNL line becomes lethal in a B cell-deficient host [11,217] does not differentiate the role of IgG from other isotypes. There is a growing appreciation for the role that IgM plays in the control of iRBCs. IgM may mediate antibody dependent phagocytosis through the Fcμ receptor, although this receptor is expressed only on B cells in mice [218]. It is also possible that complement-mediated lysis of opsonized iRBCs could contribute to parasite control as shown in P. falciparum infections [219,220] although the effects of complement depletion has been shown to be minimal in the P. chabaudi AS mouse model [221].
It is important to note that antibodies do not appear to be an absolute requirement to control all species of Plasmodium infection in mice. Unlike in P. yoelii XNL infections, B cell-deficient mice infected with P. chabaudi are able to control acute infection via antibody-independent mechanisms [217,222]. Depletion of γδ T cells in B cell-deficient mice following P. chabaudi AS infection led to exacerbated parasitemia, indicating a more critical role for γδ T cells in cell-mediated immune response against P. chabaudi [223]. Thus, mouse models of Plasmodium infection indicate that some immune mechanisms of iRBC control may be differentially important for different species or clones of Plasmodium.

5.2. Plasmodium Blood Stage Infection Leads to the Development of Memory B Cells That Respond during Secondary Challenge Infection

Humoral immune responses against malaria develop slowly, inefficiently and wane over time in the absence of reinfection [224,225,226]. Antibodies are derived from antibody-secreting cells (ASCs) that are comprised of plasmablasts and plasma cells. These cell types are generated from a specialized compartment called the germinal center (GC) in secondary lymphoid tissue [227]. ASCs can also be found out-with the B cell follicles of the spleen. Mouse models of infection have been instrumental in demonstrating that this early response may have some protective qualities [228]. Memory B cells (MBCs) and long-lived plasma cells (LLPC) offering protection against re-challenge infections are also thought to develop in the GC after infection but some evidence suggests that these are heterogeneous populations and some sub-populations of MBCs may form prior to GC formation [229,230,231]. Given that secondary lymphoid tissue is not readily accessible for study in Plasmodium-infected humans, mouse Plasmodium infections have allowed dissection of cellular responses in humoral immunity to malaria. The investigation of B cell responses to blood stage Plasmodium has typically utilized both P. chabaudi AS and P. yoelii XNL. Despite differences in the importance of antibodies for iRBC control between these species, the cellular mechanisms underlying B cell responses to Plasmodium blood stage parasites appear similar regardless of which species was used to initiate infection [232,233,234].
The cellular basis underpinning the lack of efficacious long-lived humoral responses to Plasmodium in those living in endemic areas is still incompletely understood. Indeed, it has been shown that a large proportion of memory B cells in P. falciparum infection are IgM+ with IgG+ memory B cells developing with age [235]. IgM+ memory B cells in malaria harbor somatically hypermutated B cell receptors [97] suggestive of affinity maturation and can develop into plasma cells that secrete IgM neutralizing antibodies that have high invasion-blocking capability against P. falciparum in vitro [236], suggesting that they may play a key role in controlling iRBCs.
Following infection with P. chabaudi, Plasmodium-reactive memory B cells and plasma cells can be detected over eight months post-primary infection [97,99]. Upon secondary infection with homologous parasites, a more rapid production of IgG isotypes can be observed [99] indicating recall responses are active and functional. B cell tetramers that bind to MSP-1 have been used to show both class-switched IgG+ as well as somatically hypermutated IgM+ memory B cells participate in recall responses [97]. The utility of IgM+ memory B cells in the memory response to malaria has been demonstrated in the P. chabaudi mouse model of malaria where IgM+ B cells were the dominant MBCs expanding on challenge infections leading to the early protection against re-infections [97]. Work in the P. yoelii XNL [111] and P. berghei ANKA [237] models have taken these results further demonstrating participation of IgM+ memory B cells in splenic secondary germinal centers upon challenge [111] with upregulation of the necessary CD80 and CD73 co-stimulatory molecules resulting in differentiation into antibody secreting cells and the secretion of iRBC-reactive antibodies. Thus, mouse models of malaria have demonstrated that IgM+ memory B cells are a critical player in the secondary responses to malaria.
One of the key mechanisms underpinning an impaired memory B cell response may be related to apoptosis induced by blood stage infections. MSP-1 vaccinated BALB/c mice infected with the lethal P. yoelii YM strain led to ablation of memory B cells and LLPCs, including those that developed prior to vaccination with MSP-1 or unrelated antigens [238]. These data suggest that, although memory B cells and LLPCs can develop to blood stage infection, more lethal Plasmodium infections may have a deleterious effect on these cell subsets via induction of apoptosis, albeit by an unknown mechanism.

5.3. Development of Functional Anti-Plasmodium Blood Stage GC Responses

While there is evidence of the formation of GCs in individuals with malaria, there are some indications that GC reactions might not be optimal during human Plasmodium infection [239]. Mouse models have been instrumental in demonstrating that fully functional GCs can develop in a primary blood stage Plasmodium infection leading to protective B cell responses. In GCs, follicular helper T (Tfh) cells interact with B cells and help push differentiation of B cells into plasma cells (short-lived and long-lived) and memory B cells [227]. The P. yoelii mouse model of infection has been used to show that B cells are the primary cell type expanding Tfh cells [240]. Upon expansion, IL-21, one of the major Tfh cell-derived cytokines, has been shown to be important in the development of robust and durable class-switched B cell responses following blood stage infection with P. chabaudi AS and P. yoelii XNL [232]. Disruption of IL-21-derived signals on B cells led to a diminished level of Plasmodium-specific antibodies and resulted in increased parasitemia which was correlated with a deficiency in the development of plasma cells and memory B cells [232]. Furthermore, Tfh-deficient CD4CrexBcl6fl animals or SAP−/− deficient animals were unable to clear chronic infection with P. chabaudi AS [95] demonstrating that, although the establishment of chronic infection appears to be antibody-independent, antibodies are critical for control of chronic infection.
During Plasmodium infection GCs form in the context of innate-derived inflammatory responses as well as during responses to existing Plasmodium infections, particularly in higher transmission areas where simultaneous multiclonal infection is common [12,13]. Mouse models of Plasmodium infection have demonstrated that B cell priming of Tfh cells in the spleen after blood stage Plasmodium infection is dampened by type 1 interferon via downregulation of T-cell expressed Inducible CO-Stimulator (ICOS) and interruption of ICOS-ICOSR signaling between GC Tfh cells and GC B cells, respectively [234]. The interaction between ICOS-ICOSR is critical for Tfh cell development against blood stage Plasmodium infection in mice [241,242]. Furthermore, upregulation of the inhibitor molecules Programmed Death-Ligand 1 (PD-L1) on antigen-presenting cells and Lymphocyte Activation Gene-3 (LAG-3) [243] on T cells negatively regulate the development and function of Tfh cells. IL-6 also plays a role in Tfh cell differentiation in blood stage Plasmodium infection, albeit IL-6R signaling appears to be more important for plasma cell development [233]. Thus, Tfh cell development and the ensuing GC reactions are highly dynamic processes controlled by positive and negative molecular regulators. These can be soluble mediators such as cytokines, but also cellular mediators such as Cytotoxic Lymphocyte Associated Antigen-4 (CTLA4)-expressing T follicular regulatory cells [244] that downregulate B cell responsiveness or reactions outside of the follicle such as the rapid of expansion of extrafollicular plasmablasts that deplete the nutrients required for cells participation in germinal center reactions with the follicles [110].
A key feature of the immune response in Plasmodium-infected individuals is the induction of a strong production of pro-inflammatory cytokines, with IFN-γ a defining cytokine. P. falciparum-induced IFN-γ in human infection has been shown to drive the expansion of T-box Expressed in T cells (T-bet)+ CD21-CD27- atypical B cells [245] which upregulate inhibitory receptors. In some [245], but not all [246], studies atypical B cells appear to have reduced functionality with respect to antibody production. Atypical B cells develop in both the P. chabaudi AS and the P. yoelii XNL models of Plasmodium infections [132,146] making these mouse models a crucial tool in dissecting these contrasting observations in human studies. The expansion of atypical B cells in acute febrile P. falciparum infection are transcriptionally distinct from activated B cells or classical memory B cells and able to interact with Tfh cells to differentiate into antibody secreting cells [246]. Mouse models of non-lethal Plasmodium infection have been used to define the function of atypical B cells in vivo, particularly with respect to their contribution to the memory B cell compartment that remains upon resolution of acute Plasmodium infection. Using a transgenic mouse that harbored antigen-specific B cells with BCRs that react to MSP-1 atypical B cells were considered to be short-lived disappearing upon resolution of infection [132]. However using Fc Receptor-Like 5 (FCRL5) as a marker of long-lived memory B cells [132] it has also been shown in a secondary challenge infection in the P. chabaudi model that FCRL5-expressing memory B cells have a robust recall response and that some of these cells adopt an atypical phenotype in a T-bet-independent fashion [146].
Although there is evidence that IFN-γ both supports [112,145] or impairs [247,248,249] GC B cell responses in mouse models of Plasmodium infection, the effects are likely contextual. T-bet intrinsic expression in B cells, induced by signaling from the IFN-γ receptor is needed for IgG2c isotype class switching during Plasmodium blood stage infection and also enhances affinity maturation [145]. This IFN-γ likely comes from IL-21/IFN-γ expressing Tfh (Tfh1) cells [98]. Although first described in the periphery of P. falciparum-infected individuals [239], rodent models of malaria have been instrumental in demonstrating the lineage and function of Tfh1 cells [98] with interferon-mediated signaling via Interferon Regulatory Factor 3 (IRF3) supporting a developmental shift away from Tfh cells to Th1 cells [250]. More recent studies have comprehensively dissected the intracellular signals governing plasticity of the Tfh/Th1 cell phenotype in CD4+T cells responding to blood stage Plasmodium infection [251]. Accordingly, molecules that down-regulate T-bet-mediated IFN-γ secretion in B cells, such as IL-10 [252], promote humoral responses to blood stage Plasmodium infection [247]. T cell-derived IL-10 can also act directly on B cells early on in infection, influencing B cell survival, their interactions with Tfh and ultimately the formation of germinal centers [253].
There is still much to be learned regarding the factors that regulate the development of B cell responses to blood stage Plasmodium. The discovery of a novel population of NK1.1 T cells supporting antibody production from short-lived extrafollicular plasma blasts [254] demonstrates the complexity in the development and control of humoral responses to blood stage Plasmodium infections. The main rodent models utilized in investigating humoral responses to malaria involve the species P. chabaudi and P. yoelii due to their non-lethal phenotype in many backgrounds of mice, including C57BL/6. However, modelling the role of antibodies in severe malaria has been accomplished using P. berghei ANKA infections normally employed for immunopathogenesis studies. One study observed that the pro-inflammatory mediators that enhance the onset of pathology associated with severe malaria also affect the development of efficacious humoral immune responses through inhibition of Tfh cell differentiation and consequently compromised GC reactions [248]. With the development of B cell tetramers [97] and BCR-transgenic mice [132] to identify malaria-reactive B cells, mouse models of malaria will be needed for moving forward the analysis and definition of factors that influence a robust and efficacious humoral response with the precision required to identify more relevant metrics to signify efficacious candidate vaccines against malaria that elicit a long-lived protective response.

5.4. The Importance of Innate Immune Cells in Control of iRBCs

Antibody-mediated control of parasites via blocking of invasion is not the only immune mechanism of iRBC control. Ample data have been gathered on human Plasmodium infections clearly demonstrate functional activity of innate cells against iRBCs. The contribution of innate immune responses to P. falciparum in the Fulani tribe in sub-Saharan Africa has been attributed to their greater resistance to infection compared with more susceptible sympatric ethnic groups [255] and innate responses to P. falciparum in CHMI studies have been associated with subsequent control of both iRBCs and clinical symptoms [256]. In P. falciparum blood stage infection, innate cells of the myeloid lineage [105,257,258], neutrophils [259,260], natural killer (NK) [28,261,262,263,264] and γδ-T cells [265] have all been shown to neutralize iRBCs. Correlations of innate cell function with parasitemia or clinical symptoms have suggested the importance of these cells in control of iRBCs. However, mouse models have played a key role in deciphering how innate cells modulate adaptive responses and exert protection against the blood stages of Plasmodium in the context of the global response.
Studies of isolated antigen-presenting cells from human peripheral blood mononuclear cells (PBMCs) and cell lines derived from human myeloid lineages were initially used in combination with cultured Plasmodium-iRBCs from P. falciparum lines to investigate how iRBCs are recognized by the immune system. The biological significance of initial studies demonstrating a role for Pattern Recognition Receptors (PRRs) such as Toll-Like Receptors (TLRs) in the recognition of iRBCs, merozoites and products of schizogony such as GlycosylPhosphatidylInositol (GPI) [266] has only been possible using in vivo mouse models of malaria. There are several PRRs that recognize Plasmodium-iRBCs and these have been reviewed by Gowda and Wu [267]. Here we will focus on the contribution of mouse models to understanding the significance of GPI-TLR2 recognition and DNA/hemozoin-TLR9 recognition in the immunology of blood stage infections.
The inflammatory nature of GPI was shown using purified from P. falciparum but it was the injection of this molecule into mice that demonstrated that GPI-induce inflammatory responses may influence malarial symptoms [268]. Complementary studies blocking responses to GPI in mice immunized with the glycan portion of this molecule were protected from P. berghei ANKA-induced experimental cerebral malaria (ECM) [269]. In vitro macrophage cultures were used to demonstrate a receptor-mediated mechanism for the effects of P. falciparum-derived GPI [270] and to identify TLR2 in a heterodimer with TLR1 as the principal PRR in GPI recognition [271]. However it was infection of TLR2−/− mice with the non-lethal P. chabaudi AS line [272] or the non-lethal P. yoelii XNL line [273] that indicated there was very little impact on control of circulating iRBCs and the pathogenesis of infection. This is in contrast to a stark pathological role in experimental cerebral malaria (ECM) where, in agreement with antibody-mediated GPI blocking, TLR2−/− mice were protected from death from neurological symptoms during P. berghei ANKA infection [274]. GPI-mediated TLR2 signaling may be more pathogenic in this system given the effects of GPI signaling on endothelial cells [266], and the pathogenic role this may play in areas of iRBC sequestration and accumulation such as the brain in CM/ECM.
Similarly, mouse models have been instrumental in determining the role of innate sensing of parasite DNA and hemozoin, an insoluble crystalline by-product of hemoglobin digestion [275], on the immunology and pathogenesis of acute Plasmodium blood stage infections. It was initially demonstrated in 2004 that P. falciparum schizont extract contained a TLR9 ligand [276], classically thought to be unmethylated CpG motifs in pathogen DNA [277]. However, it has subsequently been shown that Plasmodium-derived DNA complexed with protein [278] or hemozoin [279] is able to signal through TLR9. The role of hemozoin in Plasmodium infection is complex: the P. chabaudi model suggests detrimental effects on splenic DCs that have internalized hemozoin crystals with respect to limited ability to activate T cell effector function in the spleen [280], possibly due to the induction of anti-inflammatory cytokines as shown in P. yoelii XNL model [273]. The consequences of this can be seen in P. yoelii XNL infection of TLR9−/− mice which have higher parasitemia [273], although control of P. chabaudi AS appears to be less affected [272]. This could be explained by the effects of hemozoin and TLR9 on B cells and antibody production. With respect to parasite control P. yoelii XNL is more dependent on antibody-mediated control of iRBCs than P. chabaudi AS [11]. More recent work in the P. yoelii model suggests that DNA sensing through TLR9 promotes the development of autoreactive Tbet+ B cells that may produce antibodies that recognize components of uninfected RBCs contributing to malarial anemia [281]. As for lethal infection and similar to TLR2, TLR9 sensing is a critical pathogenic factor in the development of ECM in the P. berghei ANKA model [274].
In addition to identifying the significance of PRR recognition of different Plasmodium-derived molecules during Plasmodium infections, mouse models of malaria have been instrumental in dissecting the relative importance of different cells from the myeloid lineage in control of iRBCs. Circulating monocytes are able to phagocytose P. falciparum [207,282] and P. vivax [283] iRBCs in both an opsonic and non-opsonic [284] manner. The P. chabaudi AS model has been used to demonstrate a significant contribution of monocytes [85] compared to neutrophils [86,87] in control of iRBCs. In agreement with these data, neutrophils also do not seem to be a dominant cell type involved in the control of iRBCs in non-lethal P. yoelii XNL infections [106]. However, neutrophils have been shown to play an important role in the P. berghei ANKA model where iRBCs are less controlled when neutrophils cannot make Neutrophil Extracellular Traps (NETs) [285]. Thus, it seems that differences in the function of myeloid cells may exist amongst rodent parasite species that may be related to parasite life cycle preferences such as infection of reticulocytes (P. yoelii XNL) over normocytes (P. chabaudi AS) or differences in inflammatory potential from iRBCs of different species.
Along a similar line, mouse models of malaria have been instrumental in understanding the contributions of NK cells and γδ-T cells with regard to control of iRBCs and the pathogenesis of infection. Both NK cells and γδ-T cells have been shown to be able to target Plasmodium-iRBCs. NK cells are able to directly recognize iRBCs to produce IFN-γ [286,287] and have been shown to confer protection in P. falciparum infection via cytokine production and ADCC that subsequently inhibits P. falciparum growth in RBCs [261,263]. However, NK activation has been shown to depend on accessory cells of the myeloid lineage [288] as well as T cells [289]. Infection of humanized mice with P. falciparum confirmed in vitro observations of contact-dependent NK cells in elimination of iRBCs [28]. Furthermore, using the non-lethal P. yoelii XNL mouse model of Plasmodium infection it has been shown in an in vivo setting that iRBCs induce activation of NK cells via synergistic effects of IL-18 and IL-12 to induce the expression of CD25 and IFN-γ production [290].
In acute malaria infection, an increase in the γδ-T cell numbers correlate with protection from high parasitemia [89,265]. However repeated exposure to malaria has been shown to lead to a decrease in circulating γδ-T cell numbers, particularly the subset that expresses the Vδ2+ chain of the γδ TCR [265,291,292]. Although it is not possible to define the role of the Vδ2+γδ-T cell subset in mice due to the absence of the Vδ2 chain in mice, in general γδ-T cells produce IL-21 and IFN-γ that may enhance humoral immune response against blood stage infection [293]. γδ-T cells produce IFN-γ, granzymes and granulysin that collectively inhibit parasite growth in a contact-dependent manner [88]. γδ-T cells displaying CD16+ Vδ2+ TCRs are able to respond to opsonized P. falciparum iRBCs through engagement of CD16 receptors [294] facilitating Vδ2+ T cell effector function with respect to ADCC cytotoxicity and in removal of iRBCs by phagocytosis [294,295]. CD16+ Vδ2+ T cells are shown to exhibit some of the features of NK cells and data also suggests they are more cytolytic than their CD16 Vδ2+ T cell counterparts [294]. In addition, Vδ2+γδ-T cells appear to induce contact-dependent lysis of iRBCs via granulysin [295].
Mouse models of malaria have been used to assess the relative contributions of NK cells and γδ-T cells with respect to the control of iRBCs. Despite an evident early expansion of NK cells in the spleen upon infection with P. chabaudi [140] and the contribution activated NK cells play in ramping up the inflammatory response during P. chabaudi infection [296], NK cells only have a small impact on control of parasitemia during primary infection [223,297,298]. On the other hand, γδ T cells have been shown to exert significant control of iRBCs during acute infection with P. chabaudi AS, exceeding that mediated by antibodies [223]. γδ-T cells can also express M-CSF that protect against recurrence of P. chabaudi parasitemia in mice at the later stage of the infection [89]. In line with these findings, depletion studies in P. chabaudi adami 556KA-infected mice found a more prominent role for γδ T cells compared to NK cells in controlling iRBCs [223], a finding supported by a second study in P. yoelii-infected mice which found no significant role for NK cells in parasite control [105].
These data illustrate the power of undertaking in vivo experiments in uniform settings such as in laboratory mice where the contribution of different immune cells can be parsed out in a way that is not easy to undertake in culture studies or human infection studies, an exception being controlled human malaria infections [299,300]. As heterogenous innate cell populations become better defined with advances in technologies such as single-cell sequencing, mouse models of malaria will be instrumental in defining how cell sub-populations of innate cells enhance iRBC lysis and phagocytosis in a temporal setting and in relation to unfolding and established adaptive responses.

6. Immunopathogenesis of Malaria and Clinical Immunity

While sterile immunity preventing Plasmodium infection does not commonly occur, people living in malaria endemic regions ultimately develop clinical immunity that protects against symptoms associated with Plasmodium blood-stage infection. Clinical immunity to malaria is characterized by reduced parasitemia and attenuated inflammatory responses [301,302]. As such, people who develop clinical immunity to malaria often carry Plasmodium iRBCs asymptomatically with a low-grade pro-inflammatory immune response that limits blood stage parasite replication.
Based on human studies, clinical immunity has long been thought to center on the acquisition of strong immunomodulatory mechanisms that fine tune the inflammatory response necessary for control of the parasite burden while controlling the inflammation-induced pathology. Clinical symptoms of malaria are driven by Th1 inflammation characterized by IFN-γ, a cytokine known to be important in the development of immune effector mechanisms including high affinity class-switched anti-parasite antibody [245,303] and activation of phagocytes [304]. The main, but not only, sources of IFN-γ found in P. falciparum infection include Th1 cells, cytolytic CD8 T cells, NK cells and γδ T cells [304], in particular Vδ2+ γδ T cells [265], as high production of pro-inflammatory cytokines by Vδ2+ γδ T cells has been shown to protect from infection by P. falciparum in children living in a high transmission setting. Analysis of T cell responses after controlled human malaria infection (CHMI) with P. falciparum demonstrated that higher blood stage parasitemia was associated with an expansion of T regulatory (T reg) cells that express CD25 and FoxP3 after schizogony from the liver [300], suggesting downregulation of the inflammatory response supports parasite replication in the blood stage.
At the same time, systemic inflammation appears to correlate with the pathogenesis of malaria. Higher IFN-γ production from γδ T cells diminishes clinical immunity in response to subsequent infections with P. falciparum [265], presumably due to inflammation-induced pathogenesis. As such, decreased Vδ2+ T cell numbers, and an upregulation of immunoregulatory markers such as Tim-3 and CD57 on γδ T cells, is associated with clinical immunity to malaria [265,291]. Along the same lines, the identification of CD4 T cells producing both IFN-γ and IL-10, termed Tr1 (Foxp3-ve regulatory) cells, have been identified in P. falciparum-infected individuals [305] and associated with uncomplicated disease in children. Indeed, a longitudinal analysis of children from an endemic region of Mali indicates that a recent exposure to Plasmodium changes the cytokine profile of a subsequent Plasmodium-infection, specifically upregulation of IL-10 in children with persistent asymptomatic infection [301]. These data suggest that IL-10 offers protection from clinical symptoms of malaria. Furthermore, there are multiple correlative studies suggesting a protective role of transforming growth factor- β (TGF-β) against clinical symptoms of malaria [306,307,308]. Collectively, these data support a role for inflammation-suppressing cytokines in protecting against the pathogenesis of malaria.
Mouse models of non-lethal malaria have been used to confirm the importance of inflammatory responses in contributing to the control of iRBCs, including the contribution of CD8 T cells which can recognize P. yoelii XNL-infected erythroblasts that express peptide-loaded MHC-I on their surface [309] and kill them via Fas-FasL dependent cytotoxicity [310] and ganulysin-mediated mechanisms of cytotoxicity [48]. A transgenic mouse with Pb-I CD8+ T cells reactive to the 60s ribosomal protein L6 (RPL6) has provided evidence that CD8 T cells can be primed in the liver stage and boosted when there is a shared antigen with blood stage malaria [127,128]. The ability of CD8 T cells to contribute to parasite control has been replicated in the P. chabaudi AS model [311,312,313] but ultimately these data suggest that this contribution is not essential for the control of parasitemia in several mouse models of malaria blood stage infection [314] and CD8 T cells in cannot confer protection without the help of CD4 T cells [315].
Given the well-defined pathologies that can be measured in the mouse models of malaria there have been several studies confirming the importance of immunomodulatory cytokines such as TGF-β [93] and IL-10 [94,109] against malaria pathogenesis. Nonetheless, IL-10 and TGF-β are both pleiotropic cytokines with several possible sources. Thus, the main contribution of studies in mouse models of Plasmodium infection has been the ability to dissect the roles of these pleiotropic cytokines throughout the course of the infection, as well as identify the most potent sources of these cytokines mediating clinical immunity to the blood stages of Plasmodium infection. For example, comparison of the lethal (P. yoelii XL) and non-lethal (P. yoelii XNL) strains of P. yoelii revealed that early production of TGF-β (within 24 h) in lethal infection is associated with delayed IFN-γ and TNF-α production, leading to uncontrolled parasite growth and 100% fatality [113]. This was in contrast with a later (day 5 post-infection) production of TGF-β in non-lethal infection which was associated with reduced parasitemia and resolution of the infection [113]. In a similar vein, the timing of IL-10 during the progression of malarial disease seems to be crucial for control of severe immunopathology [316].
With regards to IL-10, the absence of which turns a non-lethal P. chabaudi AS infection into a lethal one [94], mouse models have been used to determine that this cytokine is essential in the control of pathogenic TNF-α production [94,109]. Mouse models have challenged the notion that these immunoregulatory cytokines were produced by classical CD4+ CD25+ FoxP3+ T reg cells. Early studies comparing lethal P. yoelii XL and nonlethal P. yoelii XNL infections demonstrated a similar expansion and activation of Treg cells following infection with these two strains, indicating that the early activation of Treg cells does not contribute to the virulence [109]. Indeed, studies of TGF-β induction by P. yoelii indicated that the main producers of TGF-β were in fact CD8+ CD25+ T reg cells.
On the other hand, the main source of IL-10 has been found in both the P. chabaudi [317] and P. yoelii [109] models to come not from classical T reg cells, but rather from FoxP3-ve T cells that have been shown to simultaneously-produce IFN-γ [317]. The presence of IL-10/IFN-γ Tr1 cells has been shown in human infection [318] but it is in mouse models that the production of IL-10 and IFN-γ in Tr1 cells has been shown to be dependent on IL-27 signaling [317,319]. IL-10 production by Tr1 cells (CD4+ CD25 Foxp3 CD127) was able to down-regulate iRBC-induced pro-inflammatory responses [109] resulting in the increased growth of parasitemia in non-lethal P. yoelii XNL infection. It has been subsequently shown that the immune regulatory role of IL-10-producing Tr1 may differ between primary and secondary infection in blood stage Plasmodium infection: IL-10 may exert a more suppressive effect on the innate immune system, specifically MHC-II expression on APCs, during primary infection while it suppresses the adaptive system, specifically expansion of antigen-experienced memory CD4 T cells, in secondary infections [320]. The use of double IFN-γ-YFP+ and IL-10-GFP+ reporter mice have indicated that following resolution of primary infection, the stability and potential of CD4+ IFNγ+ IL-10+ T cells to become memory is limited [320], in part because they exhibit an exhaustion phenotype and are generally unresponsive at the early stage of secondary infection.

6.1. Organ-Specific Pathologies

Severe cases of malaria are associated with organ dysfunction which can be caused by sequestration of iRBCs via vascular adhesion [321] and trapping of iRBCs in the capillaries due to reduced deformability [322,323,324]. With the advent of luciferase expressing constructs, the rodent Plasmodium parasites have been shown to sequester in different organs in mice [56,118,325,326]. Given the relative inaccessibility of human organs from patients experiencing severe malaria syndromes, the rodent models of Plasmodium infections have been instrumental in dissecting immunopathological mechanisms associated with localized inflammation from sequestered and accumulated iRBCs [45,56,119]. Models of particular malaria-associated syndromes can be achieved using different combinations of rodent parasite species and mouse backgrounds (Table 2). Here, we focus on how mouse models of blood stage malaria have contributed to our understanding of the immunological underpinnings of three of the most well-studied sequelae of malaria: Severe Malarial Anemia (SMA), Cerebral Malaria (CM) and Acute Respiratory Distress Syndrome (ARDS).

6.2. Mechanisms of Inflammation-Induced SMA

SMA in children is defined as a hemoglobin value < 5 g/dL and detectable parasitemia in the blood stream [327]. Although Plasmodium replication in RBCs results in physical destruction of the RBC, SMA is more likely caused by mechanisms that result in hemolysis [328] and clearance of both uninfected and iRBCs via erythrophagocytosis [329] in combination with disrupted erythropoiesis in the bone marrow [330]. The relative contributions and mechanisms underlying these different contributors to a reduction in circulating RBCs are difficult to assess in humans without splenic or bone marrow biopsies. Sequestration/accumulation of iRBCs in the inflamed bone marrow has been shown [331]. However, mouse models of SMA, principally the non-lethal P. chabaudi model, have been instrumental in demonstrating the underlying mechanisms of anemia and in dissecting the relative contributions of each.
Early work in using P. chabaudi as a model for SMA established that dyserythropoiesis in malaria may result from stalling of late erythroid progenitor cells [332] and be related to bone marrow inflammation, in particular production of the pro-inflammatory cytokines IL-12 [333] and macrophage migration inhibitory factor (MIF) [334]. A role for type 2 cytokines, specifically IL-4, has also been shown to suppress late erythroid progenitor cells [335]. Inflammation is likely derived from iRBCs that accumulate in the bone marrow, but early studies suggested that “malaria toxins”, free GPI anchors that are released during iRBC schizogeny [336], can directly lead to dyserythropoiesis [337,338]. Hemozoin has also been shown to induce anemia [339], demonstrating a contribution from parasite products in the suppression of erythrocyte production. Nonetheless, the density of circulating iRBCs is not necessarily related to level of anemia in the P. chabaudi model [58] suggesting that direct parasite destruction of RBCs during replication and release of inflammatory products during schizogony plays a more minor role in the severity of malarial anemia. Given the insoluble and persistent nature of hemozoin, the contribution of hemozoin may be cumulative over time during chronic infection.
Whilst existing data using rodent models of Plasmodium infection point to a direct suppression in the development of late erythroid progenitor cells via inflammatory cytokine induction, there may also be an indirect effect via cytokine modulation of erythropoietin produced by the kidney [340]. Other studies have investigated whether a defect in iron handling also contributes to suppression of erythropoiesis [341] and how this may be reversed [342]. Other than production of new RBCs during the process of erythropoiesis, removal of both infected and uninfected circulating RBCs has been shown to occur in the liver via erythrophagocytosis in rodent infections [343]. The removal of uninfected RBCs via autoimmune antibody-dependent mechanisms has been suggested in human malaria [344], and these include antibodies that recognize band 3, an anion transporter that mediates RBC flexibility which is the most abundant erythrocyte surface receptor [345], phosphatidyl serine (PS) that becomes exposed on the surface RBCs [346] and spectrin, a cytoskeletal molecule important for maintaining the stability and structure of the cell membrane [347]. Autoantibodies to band-3 can occur naturally but appear to be elevated in Plasmodium-infected individuals [348,349].
The contribution of anti-band-3 antibodies to malarial anemia is uncertain and may depend on the infecting species of Plasmodium. However, anti-PS antibodies have been demonstrated to contribute to RBC removal in the P. yoelii XNL model [350] providing a mouse model in which to test the significance of this mechanism with respect to malarial anemia. Plasma cells producing anti-PS antibodies in the P. yoelii model have been identified as CD11c+ Tbet+ atypical B cells that are induced in response to DNA sensed by TLR9 and IFN-γ [281]. Work in the P. chabaudi model has been used to show that malaria appears to prime B cell clones that already exist, producing autoantibodies with a strong IgG response to band-3 and spectrin [351]. Thus, in addition to defining relative contributions of autoantibodies in the severity of malarial anemia relative to other mechanisms, mouse models of malarial anemia allow interrogation of the cellular mechanisms underlying how autoantibodies arise.
Moving forward, mouse models of blood stage Plasmodium infection will be instrumental in determining how iRBCs interact in the bone marrow niche [352] and the mechanisms by which extramedullary erythropoiesis are established in an attempt to remedy diminished circulating RBCs, particularly in the red pulp of the spleen [353,354]. It will also be pertinent to determine the degree to which removal of uninfected RBCs by autoantibodies impacts on the levels of anemia.

6.3. T Cell-Mediated Breakdown of the Blood-Brain Barrier in Cerebral Malaria

Pediatric cerebral malaria (CM) is almost always fatal when not treated with antimalarials, and still has mortality rates between 15% and 20% with treatment [355]. The initiation of CM is thought to occur as a result of sequestration and adherence of iRBCs to the brain vasculature leading to disruption of the blood-brain barrier (BBB), a complex of cells and extracellular structures that regulates the exchange of molecules between the blood and the central nervous system. BBB disruption occurs upon activation of brain microvascular endothelial cells. Although markers of vascular activation can be measured in the bloodstream of individuals with CM [356], the mechanism by which the BBB breaks down is poorly understood, in part due to the paucity of brain tissue availability from victims of CM and other control groups for comparison. As such, rodent models of CM are essential to enable cellular mechanisms of BBB breakdown in CM to be elucidated and rationally targeted therapeutically.
Infection of C57Bl/6 mice with P. berghei ANKA recapitulates many of the features that characterize human CM (Table 2) and is a commonly used model described as ECM [357]. Compared to some of the other ECM models, P. berghei ANKA infection of C57BL/6 mice does not rely on extremely high parasitemia to cause disease. Infected mice usually die between 6 and 10 days after infection [358] with accumulation of iRBCs to the brain microvasculature [359] and the activation of brain endothelial cells [360,361]. Human studies suggest that both host and parasite factors mediate the development of CM in P. falciparum-infected children. As such there is some debate on the utility of the rodent ECM model, particularly since there are some differences in the expression of parasite adhesion molecules such as CD36 on human brain microvascular endothelial cells [362,363] compared with mouse [364]. Furthermore P. berghei ANKA lacks PfEMP-1, a ligand of both Intercellular Adhesion Molecule-1 (ICAM-1) and Endothelial Protein C Receptor (EPCR) on brain endothelial cells, and this interaction is thought to facilitate sequestration [365]. P. berghei ANKA iRBC accumulation in organs relies on the expression of the Schizont Membrane-Associated Cytoadherence protein (SMAC) on the surface of P. berghei ANKA iRBCs [364]. Nonetheless, there is ample microscopic evidence, particularly using 2-photon techniques, of iRBC accumulation in brain microvessels with some sequestration of iRBCs on the endothelial lining [115,366,367]. Experiments with luciferase-expressing P. berghei ANKA strain clearly show focused accumulation in the brain, particularly in the brain stem and olfactory bulb [40,368,369,370,371,372].
Inflammation related to sequestered iRBCs is thought to be a central facet of the pathogenesis of CM, and is necessary for the pathogenesis of ECM. Neuroinflammation often involves the production of TNF-α but ablation of TNF-α using Etanercept did not reduce the mortality rate of pediatric CM [373]. Data using the ECM model of CM concurs with this finding whereby infection of TNF-α deficient mice still die from BBB breakdown in the same time frame as intact animals [374]. Indeed, ECM has been critical in demonstrating the importance of lymphotoxin-α (LT-α), rather than TNF-α, in mediating breakdown of the BBB [374,375].
In addition to TNF-α and LT-α, interferons are a key facet of the neuroinflammatory response to iRBCs. Plasmodium parasites are known to induce type 1 IFNs (IFN-I) which, depending on context, have the capacity to both suppress and activate innate and adaptive immune cells, promote pro-inflammatory cytokine production and enhance parasite clearance. Both IFN-I subtypes a signal through the heterodimeric IFNAR functioning in both an autocrine and paracrine manner. The binding of IFN-I to IFNAR induces a signal cascade that initiates the transcription of interferon stimulated genes (ISGs). Host genetic variation can lead to differences in gene regulatory regions of the IFNAR1 subunit of IFNAR. The development of cerebral malaria in children has previously been associated with a variant of IFNAR1 associated with a higher expression of IFNAR1 [376,377,378,379]. This suggests that Type 1 interferon signaling is a pathogenic event [376] and is a finding supported by studies in ECM [141,380].
Similarly, there have been associations with polymorphisms in the IFN-γ receptor [381] and lower levels of plasma IFN-γ [382] with development of CM suggesting that IFN-γ is protective. This is in agreement with polymorphisms in the IFN-γ gene promoter which are associated with increased transcription of IFN-γ and protection from CM [382]. The ECM model relays a different story: IFN-γ has been shown to be necessary for death to occur [358,383]. IFN-γ derived from CD4 T cells [384] leads to trafficking of pathogenic CD8 T cells to the brain [358,384] and cross-presentation of merozoite-derived epitopes on major histocompatibility complex (MHC)-I [360] for recognition by infiltrating parasite-reactive CD8 T cells and BBB disruption. In ECM, IFN-γ leads to upregulation of adhesion molecules on brain microvascular endothelial cells enhancing the adhesive properties of P. berghei ANKA iRBCs [385]. The reason for the apparent difference in the role of IFN-γ in BBB disruption is unknown, but may be related to the differences exerted on parasite control mechanisms initiated by IFN-γ in the context of a more chronic setting than that studied in the ECM model.
One of the significant breakthroughs in our understanding of the immunological underpinning of CM from the ECM model was the demonstration that CXCR3 [386] and CCR5 [387]-dependent CD8 T cell infiltration into the brain is necessary for disruption of the BBB [114]. In mice IFN-γ, including that secreted by NK cells [386], induces production of CXCR3- and CCR5-responsive chemokines in the neurovascular unit thus facilitating recruitment of pathogenic CD8 T cells and other immune cells to the Central Nervous System (CNS) [388]. Whilst initial studies on human autopsy samples indicated a cellular infiltrate that was largely devoid of CD8 T cells [389,390], suggesting a potential fundamental difference in the etiology of BBB breakdown between human CM and mouse ECM, more in depth studies from pediatric CM victims in Malawi have provided evidence that CD8 T cells do infiltrate the brain [390,391], and this increase in CD8 T cells is correlated to density of iRBC sequestration [391]. As such, CD8 T cell involvement in human CM can no longer be ruled out.
The mechanisms by which CD8 T cells mediate breakdown of the BBB via effects in the endothelium are still poorly understood but the ECM model has been critical in elucidating some of the parameters by which this occurs. It has been shown in ECM that lytic molecules perforin and granzyme B [392,393] are essential components in this process. Evidence of apoptosis in brain endothelial cells can be observed in autopsy samples of pediatric CM cases [388,394] as may be expected via the lytic action of incoming primed CD8 T cell recognition of cross-presenting brain endothelial cells. Although apoptosis can also be seen in brain sections of P. berghei ANKA in addition to 2-photon microscopy [367], it is minimal. Furthermore, infected mice do not have a significantly increased cleaved caspase compared with naïve mice [393]. Whilst other mechanisms of brain microvascular endothelial cells such as necrosis, ferroptosis and pyroptosis have not been extensively investigated, these data do suggest that perforin and granzyme are acting through non-cell death-inducing pathway that disrupts the BBB. In this regard findings in the ECM model are similar to Theiler’s murine encephalomyelitis in mice, another model of CD8 T cell dependent disruption of the BBB, where perforin but not FasL is required to mediate vascular leakage and death [395].
The findings that CD8-derived lytic enzymes might not act via induction of apoptosis leads to the idea that they may act to disrupt the BBB via another mechanism. They may have a role in downregulating tight-junction and adherens-junction proteins which normally enable endothelial cells to dynamically control the passage of solutes and other molecules across the BBB [396]. Disassembly and downregulation of junction proteins on brain microvascular endothelial cells has been observed in both pediatric CM autopsy samples [397] and in ECM [40,366]. In ECM CD8 T cell-degranulation may induce downregulation of junction proteins via release of perforin [393,398] which could augment expression of vascular activation-induced molecules such as the tyrosine kinase receptor EphA2 which has been shown to mediate the loss of tight junctions on both human and mouse brain microvascular endothelial cells [40]. In the Theiler’s murine encephalomyelitis model, leakage and downregulation of tight junction proteins occurs before an increase in apoptosis markers [398]. Thus, the timing of BBB breakdown in ECM and CM relative to initial Plasmodium infection may be important in the interpretation of ECM studies as applied to CM.
Given that BBB disfunction is a feature of both CM in humans and ECM in mice, the rodent model of ECM is a crucial tool in unravelling the most important mechanisms that lead to fatal pathogenesis. Endothelial cells are only one player in the neurovascular unit that also includes mural cells (pericytes), astrocytes and microglia [399]. It is hard to discount the potential role of these accessory cells in disassembly of inter-endothelial junction proteins in CM given that astrocytes and microglia are both activated in ECM [400,401,402] and the known role they play in regulation of BBB integrity. Indeed, molecules secreted from these cells upon activation can be measured in the cerebral spinal fluid of children with CM [403] and pediatric autopsy samples demonstrate activation of microglia and astrocytes in fatal CM [404]. The mechanisms by which these accessory cells become activated, and the mechanisms which control endothelial cell junction protein modulation in CM, remain to be discovered. Given the difficulty in studying these cells, it is likely that the ECM model will be instrumental in disentangling the cellular and molecular basis of endothelial cell junction disassembly.
Identification of immunodominant epitopes from P. berghei ANKA allows for more in-depth studies on the characteristics of the CD8 T cells that are pathogenic in ECM. In addition to the glideosome associated protein 50 (GAP50)40–48 epitope in the context of H2-Db [360], two further immunodominant epitopes in the P. berghei ANKA model have been identified: bergheilysin protein (Pb2)592–599 and replication protein A1 (F4199–206) [405] both in the context of MHC Class I H2-Kb. This work has allowed the development of tetramers to track Plasmodium-reactive CD8 T cells in vivo and opens up the possibility of developing TCR transgenic animals, such as the PbI transgenic mice that recognize parasite-derived 60S Ribosomal protein L6 (RPL6)[127] in several Plasmodium species, including P. falciparum, that can be used to investigate the mechanisms by which CD8 T cells might mediate BBB breakdown. In addition, the ECM model is likely to be important in the identification of possible avenues for therapeutic targeting, such as possible IL-33 administration to induce anti-inflammatory cytokine expression and the expansion of anti-inflammatory macrophage and regulatory T cell populations [406] or IL-15 complex treatment to protect BBB leak by expanding a population of IL-10 producing NK cells [407].

6.4. Mechanisms of Malaria-Associated Acute Lung Injury (MA-ALI) in Malaria-Associated Acute Respiratory Distress Syndrome (MA-ARDS)

Pulmonary complications arising from Plasmodium infection can occur with all species but in particular upon infection with falciparum, vivax and knowlsei species. This is a syndrome of severe malaria resulting in up to 40% mortality even with treatment [20]. Though more common in adults infected with vivax malaria, in children, MA-ARDS can often present along with cerebral complications [22]. MA-ARDS is characterized by increased permeability of pulmonary capillary endothelial cells and alveolar epithelial cells, with Pulmonary Edema (PE), hypoxia [20] and in some cases fibrosis [408,409,410]. Most data related to the pathogenesis of MA-ARDS and MA-ALI comes from post-mortem studies of lung tissue from adult fatalities of Plasmodium infection showing apoptosis of alveolar cells [21]. However, the immunological mechanisms underlying MA-ARDS and MA-ALI in Plasmodium patients are relatively understudied and poorly understood.
Pulmonary vascular activation is thought to arise in response to the sequestration of iRBCs resulting in inflammation in the lung microvasculature [17,21] characterized by expression of TNF-α [22], Von Willebrand Factor (VWF) and ANGiopoietin-2 (ANG2) [411]. However, sequestration of iRBCs is likely to occur via a different suite of adhesion molecules upregulated on the pulmonary vasculature compared with the BBB. For example, EPCR expression which is a key molecule mediating adhesion of iRBCs on brain microvascular endothelial cells [412] has been found to be significantly downregulated on pulmonary vasculature endothelial cells in those who have succumbed to MA-ARDS [22] compared with those dying of other malaria-related syndromes.
Some studies employ the P. berghei ANKA strain used to study ECM by virtue of the fact this strain sequesters in the lung [413,414] and the ultrastructure of the infected lung looks similar to postmortem samples from victims of MA-ARDS [116]. However, MA-ARDS and MA-ALI are more commonly studied using infection with the NK65 strains of P. berghei [45]. The advantages of the P. berghei NK65 models are that they do not appear to result in neurological manifestations of infection and have higher ARDS clinical scores, than P. berghei ANKA infection [45]. P. berghei NK65 iRBCs accumulate in the lung vasculature, with an increase in VWF expression [415] as also found in human Plasmodium infections. There are two primary strains of P. berghei NK65 used for studies in the pathogenesis of MA-ARDS and MA-ALI: the Edinburgh strain (P. berghei NK65E) and the New York strain (P. berghei NK65NY). Possibly due to slower growth of iRBCs due to the predilection of P. berghei NK65NY to infect reticulocytes, the P. berghei NK65NY does not recapitulate MA-ARDS despite sequestering in the lung tissue. However, the Edinburgh strain results in rapid death of C57BL/6 mice from days 6–10 post-infection and recapitulates features of MA-ARDS seen in humans such as extensive neutrophil infiltration, an increase in pulmonary VWF expression [415] and an increase in protein concentration in lungs [45]. P. berghei-NK65E has been used to demonstrate the critical role of VWF in alveolar leakage [415].
In mice, MA-ARDS and MA-ALI appear to have similarities regarding the underlying pathogenesis of ECM. Studies on MA-ARDS/ALI using P. berghei ANKA infection have demonstrated that IFN-γ, upregulation of chemokines [416] and functioning CD8 T cells are all necessary for lung sequestration of iRBCs and pulmonary edema [417]. In addition, pulmonary vascular leak and BBB breakdown are dependent on the presence of platelets [41]. Unlike the BBB where molecules such as ICAM-1 and EPCR have been shown to play a key role, sequestration in the lung appears to be more dependent on the scavenger receptor CD36 [364]. There also appears to be a difference in the importance of myeloid cells with marked infiltration of neutrophils [418] and monocytes [419] to the lung which, at least for monocytes, appear to play a key role in controlling iRBC numbers.
The suite of P. berghei strains available to study this syndrome of malaria will be of some help in the interpretation of pulmonary autopsy samples from patients who have died of MA-ARDS and MA-ALI, a necessary endeavor given the lack of other tractable options to study this in Plasmodium-infected humans.

7. Conclusions

In summary this review has highlighted the utility of the rodent models of Plasmodium infection with regards to understanding the immunology of blood stage malaria. Several models exist although none completely recapitulate all aspects of malaria. However, this reflects the heterogeneity of this disease. Choosing the correct model to investigate specific aspects of this disease is essential in order to be able to extrapolate to human Plasmodium infections. There is still a plethora of key outstanding questions that remain in the field of blood stage immunology of malaria (see Outstanding Questions box). With the advent of genetically-modified rodent Plasmodium strains and an ever-increasing catalog of genetically-modified and transgenic mice in addition to SPF genetically diverse mice available through the CC and DO mouse resources, these questions can only be answered with the employment of mouse models of blood stage malaria. Analysis of spatial aspects of anti-malarial immune responses can only be studied in the context of infection rather than employing in vitro studies. The mechanisms at play in immunologically-driven organ-specific pathogenesis of malaria can only be holistically studied using a whole-organism approach and in organs that are not readily accessible in humans. For all of these reasons, the tractable rodent models of malaria described here will be a critical tool with respect to answering these outstanding questions in the field of blood stage malaria. In turn, the information gained will be instrumental in the rational design of novel immunologically-based therapeutic strategies that are badly needed in the fight against this disease.

Author Contributions

A.C.O.: Drafted the work, revised the work and created the tables; D.H.C.: Drafted the work, revised the work and created Figure 1; M.R.: Drafted the work and revised the work; T.J.L.: Drafted the work, revised the work and created the tables and Figure 1. All authors have read and agreed to the published version of the manuscript.

Funding

This work is generously supported by funding from the National Institute for Allergy and Infectious Diseases (Grant numbers: R01AI167422; R01AI147641 and 1R21AI152578) and the National Institute for Neurological Disorders and Strokes (Grant number: 1R01NS097819).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We thank Jenna Reed and Jared Andersen for critical reading of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. WHO. World Malaria Report; WHO: Geneva, Switzerland, 2021. [Google Scholar]
  2. Olotu, A.; Fegan, G.; Wambua, J.; Nyangweso, G.; Leach, A.; Lievens, M.; Kaslow, D.C.; Njuguna, P.; Marsh, K.; Bejon, P. Seven-year efficacy of RTS, S/AS01 malaria vaccine among young African children. N. Engl. J. Med. 2016, 374, 2519–2529. [Google Scholar] [CrossRef] [PubMed]
  3. Tinto, H.; Otieno, W.; Gesase, S.; Sorgho, H.; Otieno, L.; Liheluka, E.; Valéa, I.; Sing’oei, V.; Malabeja, A.; Valia, D. Long-term incidence of severe malaria following RTS, S/AS01 vaccination in children and infants in Africa: An open-label 3-year extension study of a phase 3 randomised controlled trial. Lancet Infect. Dis. 2019, 19, 821–832. [Google Scholar] [CrossRef]
  4. Datoo, M.S.; Natama, M.H.; Some, A.; Traore, O.; Rouamba, T.; Bellamy, D.; Yameogo, P.; Valia, D.; Tegneri, M.; Ouedraogo, F.; et al. Efficacy of a low-dose candidate malaria vaccine, R21 in adjuvant Matrix-M, with seasonal administration to children in Burkina Faso: A randomised controlled trial. Lancet 2021, 397, 1809–1818. [Google Scholar] [CrossRef]
  5. Vaughan, A.M.; Kappe, S.H.I. Malaria Parasite Liver Infection and Exoerythrocytic Biology. Cold Spring Harb. Perspect. Med. 2017, 7, a025486. [Google Scholar] [CrossRef]
  6. WHO Severe Malaria. 2014. Available online: https://onlinelibrary.wiley.com/doi/full/10.1111/tmi.12313_2 (accessed on 8 September 2022).
  7. Ndila, C.M.; Uyoga, S.; Macharia, A.W.; Nyutu, G.; Peshu, N.; Ojal, J.; Shebe, M.; Awuondo, K.O.; Mturi, N.; Tsofa, B. Human candidate gene polymorphisms and risk of severe malaria in children in Kilifi, Kenya: A case-control association study. Lancet Haematol. 2018, 5, e333–e345. [Google Scholar] [CrossRef]
  8. Marquet, S.; Conte, I.; Poudiougou, B.; Argiro, L.; Dessein, H.; Couturier, C.; Burté, F.; Oumar, A.A.; Brown, B.J.; Traore, A. A functional IL22 polymorphism (rs2227473) is associated with predisposition to childhood cerebral malaria. Sci. Rep. 2017, 7, 41636. [Google Scholar] [CrossRef]
  9. Jenkins, N.E.; Mwangi, T.W.; Kortok, M.; Marsh, K.; Craig, A.G.; Williams, T.N. A polymorphism of intercellular adhesion molecule-1 is associated with a reduced incidence of nonmalarial febrile illness in Kenyan children. Clin. Infect. Dis. 2005, 41, 1817–1819. [Google Scholar] [CrossRef]
  10. Villarino, N.F.; LeCleir, G.R.; Denny, J.E.; Dearth, S.P.; Harding, C.L.; Sloan, S.S.; Gribble, J.L.; Campagna, S.R.; Wilhelm, S.W.; Schmidt, N.W. Composition of the gut microbiota modulates the severity of malaria. Proc. Natl. Acad. Sci. USA 2016, 113, 2235–2240. [Google Scholar] [CrossRef]
  11. Matar, C.G.; Anthony, N.R.; O’Flaherty, B.M.; Jacobs, N.T.; Priyamvada, L.; Engwerda, C.R.; Speck, S.H.; Lamb, T.J. Gammaherpesvirus Co-Infection with Malaria Suppresses Anti-Parasitic Humoral Immunity. PLoS Pathog. 2015, 11, e1004858. [Google Scholar] [CrossRef]
  12. Weiss, G.E.; Traore, B.; Kayentao, K.; Ongoiba, A.; Doumbo, S.; Doumtabe, D.; Kone, Y.; Dia, S.; Guindo, A.; Traore, A.; et al. The Plasmodium falciparum-specific human memory B cell compartment expands gradually with repeated malaria infections. PLoS Pathog. 2010, 6, e1000912. [Google Scholar] [CrossRef] [Green Version]
  13. Wipasa, J.; Suphavilai, C.; Okell, L.C.; Cook, J.; Corran, P.H.; Thaikla, K.; Liewsaree, W.; Riley, E.M.; Hafalla, J.C. Long-lived antibody and B Cell memory responses to the human malaria parasites, Plasmodium falciparum and Plasmodium vivax. PLoS Pathog. 2010, 6, e1000770. [Google Scholar] [CrossRef]
  14. Oakley, M.S.; Gerald, N.; McCutchan, T.F.; Aravind, L.; Kumar, S. Clinical and molecular aspects of malaria fever. Trends Parasitol. 2011, 27, 442–449. [Google Scholar] [CrossRef]
  15. Perkins, D.J.; Were, T.; Davenport, G.C.; Kempaiah, P.; Hittner, J.B.; Ong’echa, J.M. Severe malarial anemia: Innate immunity and pathogenesis. Int. J. Biol. Sci. 2011, 7, 1427–1442. [Google Scholar] [CrossRef]
  16. Dvorin, J.D. Getting Your Head around Cerebral Malaria. Cell Host Microbe 2017, 22, 586–588. [Google Scholar] [CrossRef]
  17. Storm, J.; Jespersen, J.S.; Seydel, K.B.; Szestak, T.; Mbewe, M.; Chisala, N.V.; Phula, P.; Wang, C.W.; Taylor, T.E.; Moxon, C.A. Cerebral malaria is associated with differential cytoadherence to brain endothelial cells. EMBO Mol. Med. 2019, 11, e9164. [Google Scholar] [CrossRef]
  18. Kessler, A.; Dankwa, S.; Bernabeu, M.; Harawa, V.; Danziger, S.A.; Duffy, F.; Kampondeni, S.D.; Potchen, M.J.; Dambrauskas, N.; Vigdorovich, V. Linking EPCR-binding PfEMP1 to brain swelling in pediatric cerebral malaria. Cell Host Microbe 2017, 22, 601–614.e5. [Google Scholar] [CrossRef]
  19. Ponsford, M.J.; Medana, I.M.; Prapansilp, P.; Hien, T.T.; Lee, S.J.; Dondorp, A.M.; Esiri, M.M.; Day, N.P.; White, N.J.; Turner, G.D. Sequestration and microvascular congestion are associated with coma in human cerebral malaria. J. Infect. Dis. 2012, 205, 663–671. [Google Scholar] [CrossRef]
  20. Koh, Y. Update in acute respiratory distress syndrome. J. Intensive Care 2014, 2, 2. [Google Scholar] [CrossRef]
  21. Punsawad, C.; Viriyavejakul, P.; Setthapramote, C.; Palipoch, S. Enhanced expression of Fas and FasL modulates apoptosis in the lungs of severe P. falciparum malaria patients with pulmonary edema. Int. J. Clin. Exp. Pathol. 2015, 8, 10002. [Google Scholar]
  22. Maknitikul, S.; Luplertlop, N.; Grau, G.E.; Ampawong, S. Dysregulation of pulmonary endothelial protein C receptor and thrombomodulin in severe falciparum malaria-associated ARDS relevant to hemozoin. PLoS ONE 2017, 12, e0181674. [Google Scholar] [CrossRef]
  23. Blumberg, L.; Lee, R.P.; Lipman, J.; Beards, S. Predictors of mortality in severe malaria: A two year experience in a non-endemic area. Anaesth. Intensive Care 1996, 24, 217–223. [Google Scholar] [CrossRef] [Green Version]
  24. Viriyavejakul, P.; Khachonsaksumet, V.; Punsawad, C. Liver changes in severe Plasmodium falciparum malaria: Histopathology, apoptosis and nuclear factor kappa B expression. Malar. J. 2014, 13, 106. [Google Scholar] [CrossRef]
  25. Silva, G.B.D.J.; Pinto, J.R.; Barros, E.J.G.; Farias, G.M.N.; Daher, E.F. Kidney involvement in malaria: An update. Rev. Inst. Med. Trop. Sao Paulo 2017, 59, e53. [Google Scholar]
  26. Possemiers, H.; Vandermosten, L.; Van den Steen, P.E. Etiology of lactic acidosis in malaria. PLoS Pathog. 2021, 17, e1009122. [Google Scholar] [CrossRef]
  27. Thien, H.V.; Kager, P.A.; Sauerwein, H.P. Hypoglycemia in falciparum malaria: Is fasting an unrecognized and insufficiently emphasized risk factor? Trends Parasitol. 2006, 22, 410–415. [Google Scholar] [CrossRef]
  28. Chen, Q.; Amaladoss, A.; Ye, W.; Liu, M.; Dummler, S.; Kong, F.; Wong, L.H.; Loo, H.L.; Loh, E.; Tan, S.Q.; et al. Human natural killer cells control Plasmodium falciparum infection by eliminating infected red blood cells. Proc. Natl. Acad. Sci. USA 2014, 111, 1479–1484. [Google Scholar] [CrossRef]
  29. Zhang, L.L.; Li, J.L.; Ji, M.X.; Tian, D.; Wang, L.Y.; Chen, C.; Tian, M. Attenuated P. falciparum Parasite Shows Cytokine Variations in Humanized Mice. Front. Immunol. 2020, 11, 1801. [Google Scholar] [CrossRef]
  30. Killick-Kendrick, R.; Peters, W. Rodent Malaria; Academic Press: London, UK, 1978. [Google Scholar]
  31. Stephens, R.; Culleton, R.L.; Lamb, T.J. The contribution of Plasmodium chabaudi to our understanding of malaria. Trends Parasitol. 2012, 28, 73–82. [Google Scholar] [CrossRef]
  32. Smith, J.D.; Chitnis, C.E.; Craig, A.G.; Roberts, D.J.; Hudson-Taylor, D.E.; Peterson, D.S.; Pinches, R.; Newbold, C.I.; Miller, L.H. Switches in expression of Plasmodium falciparum var genes correlate with changes in antigenic and cytoadherent phenotypes of infected erythrocytes. Cell 1995, 82, 101–110. [Google Scholar] [CrossRef]
  33. Su, X.Z.; Heatwole, V.M.; Wertheimer, S.P.; Guinet, F.; Herrfeldt, J.A.; Peterson, D.S.; Ravetch, J.A.; Wellems, T.E. The large diverse gene family var encodes proteins involved in cytoadherence and antigenic variation of Plasmodium falciparum-infected erythrocytes. Cell 1995, 82, 89–100. [Google Scholar] [CrossRef]
  34. Niang, M.; Bei, A.K.; Madnani, K.G.; Pelly, S.; Dankwa, S.; Kanjee, U.; Gunalan, K.; Amaladoss, A.; Yeo, K.P.; Bob, N.S.; et al. STEVOR is a Plasmodium falciparum erythrocyte binding protein that mediates merozoite invasion and rosetting. Cell Host Microbe 2014, 16, 81–93. [Google Scholar] [CrossRef] [PubMed]
  35. Goel, S.; Palmkvist, M.; Moll, K.; Joannin, N.; Lara, P.; Akhouri, R.R.; Moradi, N.; Ojemalm, K.; Westman, M.; Angeletti, D.; et al. RIFINs are adhesins implicated in severe Plasmodium falciparum malaria. Nat. Med. 2015, 21, 314–317. [Google Scholar] [CrossRef] [PubMed]
  36. Cunningham, D.A.; Jarra, W.; Koernig, S.; Fonager, J.; Fernandez-Reyes, D.; Blythe, J.E.; Waller, C.; Preiser, P.R.; Langhorne, J. Host immunity modulates transcriptional changes in a multigene family (yir) of rodent malaria. Mol. Microbiol. 2005, 58, 636–647. [Google Scholar] [CrossRef] [PubMed]
  37. Fonager, J.; Cunningham, D.; Jarra, W.; Koernig, S.; Henneman, A.A.; Langhorne, J.; Preiser, P. Transcription and alternative splicing in the yir multigene family of the malaria parasite Plasmodium y. yoelii: Identification of motifs suggesting epigenetic and post-transcriptional control of RNA expression. Mol. Biochem. Parasitol. 2007, 156, 1–11. [Google Scholar] [CrossRef]
  38. Lawton, J.; Brugat, T.; Yan, Y.X.; Reid, A.J.; Bohme, U.; Otto, T.D.; Pain, A.; Jackson, A.; Berriman, M.; Cunningham, D.; et al. Characterization and gene expression analysis of the cir multi-gene family of Plasmodium chabaudi chabaudi (AS). BMC Genom. 2012, 13, 125. [Google Scholar] [CrossRef]
  39. Yam, X.Y.; Brugat, T.; Siau, A.; Lawton, J.; Wong, D.S.; Farah, A.; Twang, J.S.; Gao, X.; Langhorne, J.; Preiser, P.R. Characterization of the Plasmodium Interspersed Repeats (PIR) proteins of Plasmodium chabaudi indicates functional diversity. Sci. Rep. 2016, 6, 23449. [Google Scholar] [CrossRef]
  40. Darling, T.K.; Mimche, P.N.; Bray, C.; Umaru, B.; Brady, L.M.; Stone, C.; Eboumbou Moukoko, C.E.; Lane, T.E.; Ayong, L.S.; Lamb, T.J. EphA2 contributes to disruption of the blood-brain barrier in cerebral malaria. PLoS Pathog. 2020, 16, e1008261. [Google Scholar] [CrossRef]
  41. Darling, T.K.; Schenk, M.P.; Zhou, C.C.; Maloba, F.M.; Mimche, P.N.; Gibbins, J.M.; Jobe, S.M.; Lamb, T.J. Platelet alpha-granules contribute to organ-specific pathologies in a mouse model of severe malaria. Blood Adv. 2020, 4, 1–8. [Google Scholar] [CrossRef]
  42. Epiphanio, S.; Campos, M.G.; Pamplona, A.; Carapau, D.; Pena, A.C.; Ataide, R.; Monteiro, C.A.; Felix, N.; Costa-Silva, A.; Marinho, C.R.; et al. VEGF promotes malaria-associated acute lung injury in mice. PLoS Pathog. 2010, 6, e1000916. [Google Scholar] [CrossRef]
  43. Hanum, P.S.; Hayano, M.; Kojima, S. Cytokine and chemokine responses in a cerebral malaria-susceptible or -resistant strain of mice to Plasmodium berghei ANKA infection: Early chemokine expression in the brain. Int. Immunol. 2003, 15, 633–640. [Google Scholar] [CrossRef]
  44. Niikura, M.; Kamiya, S.; Kita, K.; Kobayashi, F. Coinfection with nonlethal murine malaria parasites suppresses pathogenesis caused by Plasmodium berghei NK65. J. Immunol. 2008, 180, 6877–6884. [Google Scholar] [CrossRef]
  45. Vandermosten, L.; Pham, T.T.; Possemiers, H.; Knoops, S.; Van Herck, E.; Deckers, J.; Franke-Fayard, B.; Lamb, T.J.; Janse, C.J.; Opdenakker, G.; et al. Experimental malaria-associated acute respiratory distress syndrome is dependent on the parasite-host combination and coincides with normocyte invasion. Malar. J. 2018, 17, 102. [Google Scholar] [CrossRef] [Green Version]
  46. Hee, L.; Dinudom, A.; Mitchell, A.J.; Grau, G.E.; Cook, D.I.; Hunt, N.H.; Ball, H.J. Reduced activity of the epithelial sodium channel in malaria-induced pulmonary oedema in mice. Int. J. Parasitol. 2011, 41, 81–88. [Google Scholar] [CrossRef]
  47. Janse, C.J.; Boorsma, E.G.; Ramesar, J.; Grobbee, M.J.; Mons, B. Host cell specificity and schizogony of Plasmodium berghei under different in vitro conditions. Int. J. Parasitol. 1989, 19, 509–514. [Google Scholar] [CrossRef]
  48. Hojo-Souza, N.S.; de Azevedo, P.O.; de Castro, J.T.; Teixeira-Carvalho, A.; Lieberman, J.; Junqueira, C.; Gazzinelli, R.T. Contributions of IFN-gamma and granulysin to the clearance of Plasmodium yoelii blood stage. PLoS Pathog. 2020, 16, e1008840. [Google Scholar] [CrossRef]
  49. Abel, S.; Luckheide, N.; Westendorf, A.M.; Geffers, R.; Roers, A.; Muller, W.; Sparwasser, T.; Matuschewski, K.; Buer, J.; Hansen, W. Strong impact of CD4+ Foxp3+ regulatory T cells and limited effect of T cell-derived IL-10 on pathogen clearance during Plasmodium yoelii infection. J. Immunol. 2012, 188, 5467–5477. [Google Scholar] [CrossRef]
  50. Xia, L.; Wu, J.; Pattaradilokrat, S.; Tumas, K.; He, X.; Peng, Y.C.; Huang, R.; Myers, T.G.; Long, C.A.; Wang, R.; et al. Detection of host pathways universally inhibited after Plasmodium yoelii infection for immune intervention. Sci. Rep. 2018, 8, 15280. [Google Scholar] [CrossRef]
  51. Yoeli, M.; Hargreaves, B.; Carter, R.; Walliker, D. Sudden increase in virulence in a strain of Plasmodium berghei yoelii. Ann. Trop. Med. Parasitol. 1975, 69, 173–178. [Google Scholar] [CrossRef]
  52. Nair, S.C.; Xu, R.; Pattaradilokrat, S.; Wu, J.; Qi, Y.; Zilversmit, M.; Ganesan, S.; Nagarajan, V.; Eastman, R.T.; Orandle, M.S.; et al. A Plasmodium yoelii HECT-like E3 ubiquitin ligase regulates parasite growth and virulence. Nat. Commun. 2017, 8, 223. [Google Scholar] [CrossRef]
  53. Sayles, P.C.; Wassom, D.L. Immunoregulation in murine malaria. Susceptibility of inbred mice to infection with Plasmodium yoelii depends on the dynamic interplay of host and parasite genes. J. Immunol. 1988, 141, 241–248. [Google Scholar]
  54. Wu, J.; Tian, L.; Yu, X.; Pattaradilokrat, S.; Li, J.; Wang, M.; Yu, W.; Qi, Y.; Zeituni, A.E.; Nair, S.C.; et al. Strain-specific innate immune signaling pathways determine malaria parasitemia dynamics and host mortality. Proc. Natl. Acad. Sci. USA 2014, 111, E511–E520. [Google Scholar] [CrossRef]
  55. Lacerda-Queiroz, N.; Riteau, N.; Eastman, R.T.; Bock, K.W.; Orandle, M.S.; Moore, I.N.; Sher, A.; Long, C.A.; Jankovic, D.; Su, X.Z. Mechanism of splenic cell death and host mortality in a Plasmodium yoelii malaria model. Sci. Rep. 2017, 7, 10438. [Google Scholar] [CrossRef]
  56. Brugat, T.; Cunningham, D.; Sodenkamp, J.; Coomes, S.; Wilson, M.; Spence, P.J.; Jarra, W.; Thompson, J.; Scudamore, C.; Langhorne, J. Sequestration and histopathology in Plasmodium chabaudi malaria are influenced by the immune response in an organ-specific manner. Cell. Microbiol. 2014, 16, 687–700. [Google Scholar] [CrossRef] [Green Version]
  57. Mackinnon, M.J.; Read, A.F. Genetic Relationships between Parasite Virulence and Transmission in the Rodent Malaria Plasmodium Chabaudi. Evolution 1999, 53, 689–703. [Google Scholar] [CrossRef]
  58. Lamb, T.J.; Langhorne, J. The severity of malarial anaemia in Plasmodium chabaudi infections of BALB/c mice is determined independently of the number of circulating parasites. Malar. J. 2008, 7, 68. [Google Scholar] [CrossRef]
  59. Fonseca, L.; Seixas, E.; Butcher, G.; Langhorne, J. Cytokine responses of CD4+ T cells during a Plasmodium chabaudi chabaudi (ER) blood-stage infection in mice initiated by the natural route of infection. Malar. J. 2007, 6, 77. [Google Scholar] [CrossRef]
  60. Yadava, A.; Kumar, S.; Dvorak, J.A.; Milon, G.; Miller, L.H. Trafficking of Plasmodium chabaudi adami-infected erythrocytes within the mouse spleen. Proc. Natl. Acad. Sci. USA 1996, 93, 4595–4599. [Google Scholar] [CrossRef]
  61. Scorza, T.; Grubb, K.; Cambos, M.; Santamaria, C.; Tshikudi Malu, D.; Spithill, T.W. Vaccination with a Plasmodium chabaudi adami multivalent DNA vaccine cross-protects A/J mice against challenge with P. c. adami DK and virulent Plasmodium chabaudi chabaudi AS parasites. Int. J. Parasitol. 2008, 38, 819–827. [Google Scholar] [CrossRef]
  62. Gaudreault, V.; Wirbel, J.; Jardim, A.; Rohrbach, P.; Scorza, T. Red Blood Cells Preconditioned with Hemin Are Less Permissive to Plasmodium Invasion In Vivo and In Vitro. PLoS ONE 2015, 10, e0140805. [Google Scholar]
  63. Cavacini, L.A.; Guidotti, M.; Parke, L.A.; Melancon-Kaplan, J.; Weidanz, W.P. Reassessment of the role of splenic leukocyte oxidative activity and macrophage activation in expression of immunity to malaria. Infect. Immun. 1989, 57, 3677–3682. [Google Scholar] [CrossRef]
  64. Van der Heyde, H.C.; Elloso, M.M.; Roopenian, D.C.; Manning, D.D.; Weidanz, W.P. Expansion of the CD4-, CD8- gamma delta T cell subset in the spleens of mice during non-lethal blood-stage malaria. Eur. J. Immunol. 1993, 23, 1846–1850. [Google Scholar] [CrossRef] [PubMed]
  65. Ramaprasad, A.; Klaus, S.; Douvropoulou, O.; Culleton, R.; Pain, A. Plasmodium vinckei genomes provide insights into the pan-genome and evolution of rodent malaria parasites. BMC Biol. 2021, 19, 69. [Google Scholar] [CrossRef] [PubMed]
  66. LaCrue, A.N.; Scheel, M.; Kennedy, K.; Kumar, N.; Kyle, D.E. Effects of artesunate on parasite recrudescence and dormancy in the rodent malaria model Plasmodium vinckei. PLoS ONE 2011, 6, e26689. [Google Scholar] [CrossRef] [PubMed]
  67. Bhardwaj, J.; Siddiqui, A.J.; Goyal, M.; Prakash, K.; Soni, A.; Puri, S.K.; Srivastava, M. Host immune response is severely compromised during lethal Plasmodium vinckei infection. Parasitol. Res. 2015, 114, 3445–3457. [Google Scholar] [CrossRef]
  68. Viens, P.; Chevalier, J.L.; Sonea, S.; Yoeli, M. The effect of reticulocytosis on Plasmodium vinckei infection in white mice. Action of phenylhydrazine and of repeated bleedings. Can. J. Microbiol. 1971, 17, 257–261. [Google Scholar] [CrossRef]
  69. Cavacini, L.A.; Parke, L.A.; Weidanz, W.P. Resolution of acute malarial infections by T cell-dependent non-antibody-mediated mechanisms of immunity. Infect. Immun. 1990, 58, 2946–2950. [Google Scholar] [CrossRef]
  70. Vigario, A.M.; Belnoue, E.; Cumano, A.; Marussig, M.; Miltgen, F.; Landau, I.; Mazier, D.; Gresser, I.; Renia, L. Inhibition of Plasmodium yoelii blood-stage malaria by interferon alpha through the inhibition of the production of its target cell, the reticulocyte. Blood 2001, 97, 3966–3971. [Google Scholar] [CrossRef]
  71. Masopust, D.; Sivula, C.P.; Jameson, S.C. Of mice, dirty mice, and men: Using mice to understand human immunology. J. Immunol. 2017, 199, 383–388. [Google Scholar] [CrossRef]
  72. Mestas, J.; Hughes, C.C. Of mice and not men: Differences between mouse and human immunology. J. Immunol. 2004, 172, 2731–2738. [Google Scholar] [CrossRef]
  73. Steiniger, B.S. Human spleen microanatomy: Why mice do not suffice. Immunology 2015, 145, 334–346. [Google Scholar] [CrossRef]
  74. Weiss, L. The spleen in malaria: The role of barrier cells. Immunol. Lett. 1990, 25, 165–172. [Google Scholar] [CrossRef]
  75. Drenckhahn, D.; Wagner, J. Stress fibers in the splenic sinus endothelium in situ: Molecular structure, relationship to the extracellular matrix, and contractility. J. Cell Biol. 1986, 102, 1738–1747. [Google Scholar] [CrossRef]
  76. Hommel, M.; David, P.H.; Oligino, L.D. Surface alterations of erythrocytes in Plasmodium falciparum malaria. Antigenic variation, antigenic diversity, and the role of the spleen. J. Exp. Med. 1983, 157, 1137–1148. [Google Scholar] [CrossRef]
  77. Gilks, C.F.; Walliker, D.; Newbold, C.I. Relationships between sequestration, antigenic variation and chronic parasitism in Plasmodium chabaudi chabaudi—A rodent malaria model. Parasite Immunol. 1990, 12, 45–64. [Google Scholar] [CrossRef]
  78. Urban, B.C.; Hien, T.T.; Day, N.P.; Phu, N.H.; Roberts, R.; Pongponratn, E.; Jones, M.; Mai, N.T.; Bethell, D.; Turner, G.D.; et al. Fatal Plasmodium falciparum malaria causes specific patterns of splenic architectural disorganization. Infect. Immun. 2005, 73, 1986–1994. [Google Scholar] [CrossRef]
  79. Achtman, A.H.; Khan, M.; MacLennan, I.C.; Langhorne, J. Plasmodium chabaudi chabaudi infection in mice induces strong B cell responses and striking but temporary changes in splenic cell distribution. J. Immunol. 2003, 171, 317–324. [Google Scholar] [CrossRef]
  80. Cadman, E.T.; Abdallah, A.Y.; Voisine, C.; Sponaas, A.M.; Corran, P.; Lamb, T.; Brown, D.; Ndungu, F.; Langhorne, J. Alterations of splenic architecture in malaria are induced independently of Toll-like receptors 2, 4, and 9 or MyD88 and may affect antibody affinity. Infect. Immun. 2008, 76, 3924–3931. [Google Scholar] [CrossRef]
  81. Lanigan, T.M.; Kopera, H.C.; Saunders, T.L. Principles of genetic engineering. Genes 2020, 11, 291. [Google Scholar] [CrossRef]
  82. Clark, J.F.; Dinsmore, C.J.; Soriano, P. A most formidable arsenal: Genetic technologies for building a better mouse. Genes Dev. 2020, 34, 1256–1286. [Google Scholar] [CrossRef]
  83. Spence, P.J.; Cunningham, D.; Jarra, W.; Lawton, J.; Langhorne, J.; Thompson, J. Transformation of the rodent malaria parasite Plasmodium chabaudi. Nat. Protoc. 2011, 6, 553–561. [Google Scholar] [CrossRef]
  84. Janse, C.J.; Ramesar, J.; Waters, A.P. High-efficiency transfection and drug selection of genetically transformed blood stages of the rodent malaria parasite Plasmodium berghei. Nat. Protoc. 2006, 1, 346–356. [Google Scholar] [CrossRef]
  85. Sponaas, A.M.; Freitas do Rosario, A.P.; Voisine, C.; Mastelic, B.; Thompson, J.; Koernig, S.; Jarra, W.; Renia, L.; Mauduit, M.; Potocnik, A.J.; et al. Migrating monocytes recruited to the spleen play an important role in control of blood stage malaria. Blood 2009, 114, 5522–5531. [Google Scholar] [CrossRef]
  86. Knackstedt, S.L.; Georgiadou, A.; Apel, F.; Abu-Abed, U.; Moxon, C.A.; Cunnington, A.J.; Raupach, B.; Cunningham, D.; Langhorne, J.; Kruger, R.; et al. Neutrophil extracellular traps drive inflammatory pathogenesis in malaria. Sci. Immunol. 2019, 4, eaaw0336. [Google Scholar] [CrossRef]
  87. Rocha, B.C.; Marques, P.E.; Leoratti, F.M.S.; Junqueira, C.; Pereira, D.B.; Antonelli, L.; Menezes, G.B.; Golenbock, D.T.; Gazzinelli, R.T. Type I Interferon Transcriptional Signature in Neutrophils and Low-Density Granulocytes Are Associated with Tissue Damage in Malaria. Cell Rep. 2015, 13, 2829–2841. [Google Scholar] [CrossRef] [Green Version]
  88. Hernández-Castañeda, M.A.; Happ, K.; Cattalani, F.; Wallimann, A.; Blanchard, M.; Fellay, I.; Scolari, B.; Lannes, N.; Mbagwu, S.; Fellay, B. γδ T Cells Kill Plasmodium falciparum in a Granzyme-and Granulysin-Dependent Mechanism during the Late Blood Stage. J. Immunol. 2020, 204, 1798–1809. [Google Scholar] [CrossRef]
  89. Mamedov, M.R.; Scholzen, A.; Nair, R.V.; Cumnock, K.; Kenkel, J.A.; Oliveira, J.H.M.; Trujillo, D.L.; Saligrama, N.; Zhang, Y.; Rubelt, F. A macrophage colony-stimulating-factor-producing γδ T cell subset prevents malarial parasitemic recurrence. Immunity 2018, 48, 350–363.e7. [Google Scholar] [CrossRef]
  90. Deroost, K.; Langhorne, J. Gamma/Delta T Cells and Their Role in Protection against Malaria. Front. Immunol. 2018, 9, 2973. [Google Scholar] [CrossRef]
  91. Seixas, E.; Moura Nunes, J.F.; Matos, I.; Coutinho, A. The interaction between DC and Plasmodium berghei/chabaudi-infected erythrocytes in mice involves direct cell-to-cell contact, internalization and TLR. Eur. J. Immunol. 2009, 39, 1850–1863. [Google Scholar] [CrossRef]
  92. Voisine, C.; Mastelic, B.; Sponaas, A.M.; Langhorne, J. Classical CD11c+ dendritic cells, not plasmacytoid dendritic cells, induce T cell responses to Plasmodium chabaudi malaria. Int. J. Parasitol. 2010, 40, 711–719. [Google Scholar] [CrossRef] [PubMed]
  93. Omer, F.M.; Riley, E.M. Transforming growth factor β production is inversely correlated with severity of murine malaria infection. J. Exp. Med. 1998, 188, 39–48. [Google Scholar] [CrossRef] [PubMed]
  94. Li, C.; Sanni, L.A.; Omer, F.; Riley, E.; Langhorne, J. Pathology of Plasmodium chabaudi chabaudi infection and mortality in interleukin-10-deficient mice are ameliorated by anti-tumor necrosis factor alpha and exacerbated by anti-transforming growth factor beta antibodies. Infect. Immun. 2003, 71, 4850–4856. [Google Scholar] [CrossRef] [PubMed]
  95. Perez-Mazliah, D.; Nguyen, M.P.; Hosking, C.; McLaughlin, S.; Lewis, M.D.; Tumwine, I.; Levy, P.; Langhorne, J. Follicular Helper T Cells are Essential for the Elimination of Plasmodium Infection. EBioMedicine 2017, 24, 216–230. [Google Scholar] [CrossRef] [PubMed]
  96. Couper, K.N.; Phillips, R.S.; Brombacher, F.; Alexander, J. Parasite-specific IgM plays a significant role in the protective immune response to asexual erythrocytic stage Plasmodium chabaudi AS infection. Parasite Immunol. 2005, 27, 171–180. [Google Scholar] [CrossRef] [PubMed]
  97. Krishnamurty, A.T.; Thouvenel, C.D.; Portugal, S.; Keitany, G.J.; Kim, K.S.; Holder, A.; Crompton, P.D.; Rawlings, D.J.; Pepper, M. Somatically Hypermutated Plasmodium-Specific IgM(+) Memory B Cells Are Rapid, Plastic, Early Responders upon Malaria Rechallenge. Immunity 2016, 45, 402–414. [Google Scholar] [CrossRef] [Green Version]
  98. Carpio, V.H.; Opata, M.M.; Montanez, M.E.; Banerjee, P.P.; Dent, A.L.; Stephens, R. IFN-gamma and IL-21 Double Producing T Cells Are Bcl6-Independent and Survive into the Memory Phase in Plasmodium chabaudi Infection. PLoS ONE 2015, 10, e0144654. [Google Scholar] [CrossRef]
  99. Ndungu, F.M.; Cadman, E.T.; Coulcher, J.; Nduati, E.; Couper, E.; Macdonald, D.W.; Ng, D.; Langhorne, J. Functional memory B cells and long-lived plasma cells are generated after a single Plasmodium chabaudi infection in mice. PLoS Pathog. 2009, 5, e1000690. [Google Scholar] [CrossRef]
  100. Opata, M.M.; Carpio, V.H.; Ibitokou, S.A.; Dillon, B.E.; Obiero, J.M.; Stephens, R. Early effector cells survive the contraction phase in malaria infection and generate both central and effector memory T cells. J. Immunol. 2015, 194, 5346–5354. [Google Scholar] [CrossRef]
  101. Fairlie-Clarke, K.J.; Allen, J.E.; Read, A.F.; Graham, A.L. Quantifying variation in the potential for antibody-mediated apparent competition among nine genotypes of the rodent malaria parasite Plasmodium chabaudi. Infect. Genet. Evol. 2013, 20, 270–275. [Google Scholar] [CrossRef]
  102. Long, G.H.; Chan, B.H.; Allen, J.E.; Read, A.F.; Graham, A.L. Blockade of TNF receptor 1 reduces disease severity but increases parasite transmission during Plasmodium chabaudi chabaudi infection. Int. J. Parasitol. 2008, 38, 1073–1081. [Google Scholar] [CrossRef]
  103. Fortin, A.; Stevenson, M.M.; Gros, P. Complex genetic control of susceptibility to malaria in mice. Genes Immun. 2002, 3, 177–186. [Google Scholar] [CrossRef]
  104. Huang, H.M.; McMorran, B.J.; Foote, S.J.; Burgio, G. Host genetics in malaria: Lessons from mouse studies. Mamm. Genome 2018, 29, 507–522. [Google Scholar] [CrossRef]
  105. Couper, K.N.; Blount, D.G.; Hafalla, J.C.; van Rooijen, N.; de Souza, J.B.; Riley, E.M. Macrophage-mediated but gamma interferon-independent innate immune responses control the primary wave of Plasmodium yoelii parasitemia. Infect. Immun. 2007, 75, 5806–5818. [Google Scholar] [CrossRef]
  106. Theess, W.; Sellau, J.; Steeg, C.; Klinke, A.; Baldus, S.; Cramer, J.P.; Jacobs, T. Myeloperoxidase Attenuates Pathogen Clearance during Plasmodium yoelii Nonlethal Infection. Infect. Immun. 2017, 85, e00475-16. [Google Scholar] [CrossRef]
  107. Miyakoda, M.; Bayarsaikhan, G.; Kimura, D.; Akbari, M.; Udono, H.; Yui, K. Metformin Promotes the Protection of Mice Infected with Plasmodium yoelii Independently of gammadelta T Cell Expansion. Front. Immunol. 2018, 9, 2942. [Google Scholar] [CrossRef] [Green Version]
  108. Mogil, R.J.; Patton, C.L.; Green, D.R. Cellular subsets involved in cell-mediated immunity to murine Plasmodium yoelii 17X malaria. J. Immunol. 1987, 138, 1933–1939. [Google Scholar]
  109. Couper, K.N.; Blount, D.G.; Wilson, M.S.; Hafalla, J.C.; Belkaid, Y.; Kamanaka, M.; Flavell, R.A.; de Souza, J.B.; Riley, E.M. IL-10 from CD4CD25Foxp3CD127 adaptive regulatory T cells modulates parasite clearance and pathology during malaria infection. PLoS Pathog. 2008, 4, e1000004. [Google Scholar] [CrossRef]
  110. Vijay, R.; Guthmiller, J.J.; Sturtz, A.J.; Surette, F.A.; Rogers, K.J.; Sompallae, R.R.; Li, F.; Pope, R.L.; Chan, J.A.; de Labastida Rivera, F.; et al. Infection-induced plasmablasts are a nutrient sink that impairs humoral immunity to malaria. Nat. Immunol. 2020, 21, 790–801. [Google Scholar] [CrossRef]
  111. Brown, S.L.; Bauer, J.J.; Lee, J.; Ntirandekura, E.; Stumhofer, J.S. IgM(+) and IgM(−) memory B cells represent heterogeneous populations capable of producing class-switched antibodies and germinal center B cells upon rechallenge with P. yoelii. J. Leukoc. Biol. 2022, in press. [Google Scholar] [CrossRef]
  112. Zander, R.A.; Vijay, R.; Pack, A.D.; Guthmiller, J.J.; Graham, A.C.; Lindner, S.E.; Vaughan, A.M.; Kappe, S.H.I.; Butler, N.S. Th1-like Plasmodium-Specific Memory CD4(+) T Cells Support Humoral Immunity. Cell Rep. 2017, 21, 1839–1852. [Google Scholar] [CrossRef]
  113. Omer, F.M.; de Souza, J.B.; Riley, E.M. Differential induction of TGF-beta regulates proinflammatory cytokine production and determines the outcome of lethal and nonlethal Plasmodium yoelii infections. J. Immunol. 2003, 171, 5430–5436. [Google Scholar] [CrossRef]
  114. McQuillan, J.A.; Mitchell, A.J.; Ho, Y.F.; Combes, V.; Ball, H.J.; Golenser, J.; Grau, G.E.; Hunt, N.H. Coincident parasite and CD8 T cell sequestration is required for development of experimental cerebral malaria. Int. J. Parasitol. 2011, 41, 155–163. [Google Scholar] [CrossRef] [PubMed]
  115. Swanson, P.A., 2nd; Hart, G.T.; Russo, M.V.; Nayak, D.; Yazew, T.; Pena, M.; Khan, S.M.; Janse, C.J.; Pierce, S.K.; McGavern, D.B. CD8+ T Cells Induce Fatal Brainstem Pathology during Cerebral Malaria via Luminal Antigen-Specific Engagement of Brain Vasculature. PLoS Pathog. 2016, 12, e1006022. [Google Scholar] [CrossRef] [PubMed]
  116. Aitken, E.H.; Negri, E.M.; Barboza, R.; Lima, M.R.; Alvarez, J.M.; Marinho, C.R.; Caldini, E.G.; Epiphanio, S. Ultrastructure of the lung in a murine model of malaria-associated acute lung injury/acute respiratory distress syndrome. Malar. J. 2014, 13, 230. [Google Scholar] [CrossRef] [PubMed]
  117. Amani, V.; Boubou, M.I.; Pied, S.; Marussig, M.; Walliker, D.; Mazier, D.; Renia, L. Cloned lines of Plasmodium berghei ANKA differ in their abilities to induce experimental cerebral malaria. Infect. Immun. 1998, 66, 4093–4099. [Google Scholar] [CrossRef] [PubMed]
  118. Pham, T.T.; Verheijen, M.; Vandermosten, L.; Deroost, K.; Knoops, S.; Van den Eynde, K.; Boon, L.; Janse, C.J.; Opdenakker, G.; Van den Steen, P.E. Pathogenic CD8(+) T Cells Cause Increased Levels of VEGF-A in Experimental Malaria-Associated Acute Respiratory Distress Syndrome, but Therapeutic VEGFR Inhibition Is Not Effective. Front. Cell. Infect. Microbiol. 2017, 7, 416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Franke-Fayard, B.; Janse, C.J.; Cunha-Rodrigues, M.; Ramesar, J.; Buscher, P.; Que, I.; Lowik, C.; Voshol, P.J.; den Boer, M.A.; van Duinen, S.G.; et al. Murine malaria parasite sequestration: CD36 is the major receptor, but cerebral pathology is unlinked to sequestration. Proc. Natl. Acad. Sci. USA 2005, 102, 11468–11473. [Google Scholar] [CrossRef]
  120. De Koning-Ward, T.F.; O’Donnell, R.A.; Drew, D.R.; Thomson, R.; Speed, T.P.; Crabb, B.S. A new rodent model to assess blood stage immunity to the Plasmodium falciparum antigen merozoite surface protein 119 reveals a protective role for invasion inhibitory antibodies. J. Exp. Med. 2003, 198, 869–875. [Google Scholar] [CrossRef]
  121. Kocken, C.H.; van der Wel, A.M.; Dubbeld, M.A.; Narum, D.L.; van de Rijke, F.M.; van Gemert, G.J.; van der Linde, X.; Bannister, L.H.; Janse, C.; Waters, A.P.; et al. Precise timing of expression of a Plasmodium falciparum-derived transgene in Plasmodium berghei is a critical determinant of subsequent subcellular localization. J. Biol. Chem. 1998, 273, 15119–15124. [Google Scholar] [CrossRef]
  122. Dobrescu, I.; de Camargo, T.M.; Gimenez, A.M.; Murillo, O.; Amorim, K.; Marinho, C.R.F.; Soares, I.S.; Boscardin, S.B.; Bargieri, D.Y. Protective Immunity in Mice Immunized with P. vivax MSP119-Based Formulations and Challenged with P. berghei Expressing PvMSP119. Front. Immunol. 2020, 11, 28. [Google Scholar] [CrossRef]
  123. Ono, T.; Tadakuma, T.; Rodriguez, A. Plasmodium yoelii yoelii 17XNL constitutively expressing GFP throughout the life cycle. Exp. Parasitol. 2007, 115, 310–313. [Google Scholar] [CrossRef]
  124. Franke-Fayard, B.; Trueman, H.; Ramesar, J.; Mendoza, J.; van der Keur, M.; van der Linden, R.; Sinden, R.E.; Waters, A.P.; Janse, C.J. A Plasmodium berghei reference line that constitutively expresses GFP at a high level throughout the complete life cycle. Mol. Biochem. Parasitol. 2004, 137, 23–33. [Google Scholar] [CrossRef]
  125. Dookie, R.S.; Villegas-Mendez, A.; Kroeze, H.; Barrett, J.R.; Draper, S.J.; Franke-Fayard, B.M.; Janse, C.J.; MacDonald, A.S.; Couper, K.N. Combinatorial Tim-3 and PD-1 activity sustains antigen-specific Th1 cell numbers during blood-stage malaria. Parasite Immunol. 2020, 42, e12723. [Google Scholar] [CrossRef]
  126. Miyakoda, M.; Kimura, D.; Yuda, M.; Chinzei, Y.; Shibata, Y.; Honma, K.; Yui, K. Malaria-specific and nonspecific activation of CD8+ T cells during blood stage of Plasmodium berghei infection. J. Immunol. 2008, 181, 1420–1428. [Google Scholar] [CrossRef]
  127. Valencia-Hernandez, A.M.; Ng, W.Y.; Ghazanfari, N.; Ghilas, S.; de Menezes, M.N.; Holz, L.E.; Huang, C.; English, K.; Naung, M.; Tan, P.S.; et al. A Natural Peptide Antigen within the Plasmodium Ribosomal Protein RPL6 Confers Liver TRM Cell-Mediated Immunity against Malaria in Mice. Cell Host Microbe 2020, 27, 950–962.e7. [Google Scholar] [CrossRef]
  128. Lau, L.S.; Fernandez-Ruiz, D.; Mollard, V.; Sturm, A.; Neller, M.A.; Cozijnsen, A.; Gregory, J.L.; Davey, G.M.; Jones, C.M.; Lin, Y.H.; et al. CD8+ T cells from a novel T cell receptor transgenic mouse induce liver-stage immunity that can be boosted by blood-stage infection in rodent malaria. PLoS Pathog. 2014, 10, e1004135. [Google Scholar] [CrossRef] [Green Version]
  129. Enders, M.H.; Bayarsaikhan, G.; Ghilas, S.; Chua, Y.C.; May, R.; de Menezes, M.N.; Ge, Z.; Tan, P.S.; Cozijnsen, A.; Mollard, V.; et al. Plasmodium berghei Hsp90 contains a natural immunogenic I-A(b)-restricted antigen common to rodent and human Plasmodium species. Curr. Res. Immunol. 2021, 2, 79–92. [Google Scholar] [CrossRef]
  130. Fernandez-Ruiz, D.; Lau, L.S.; Ghazanfari, N.; Jones, C.M.; Ng, W.Y.; Davey, G.M.; Berthold, D.; Holz, L.; Kato, Y.; Enders, M.H.; et al. Development of a Novel CD4(+) TCR Transgenic Line That Reveals a Dominant Role for CD8(+) Dendritic Cells and CD40 Signaling in the Generation of Helper and CTL Responses to Blood-Stage Malaria. J. Immunol. 2017, 199, 4165–4179. [Google Scholar] [CrossRef]
  131. Stephens, R.; Albano, F.R.; Quin, S.; Pascal, B.J.; Harrison, V.; Stockinger, B.; Kioussis, D.; Weltzien, H.U.; Langhorne, J. Malaria-specific transgenic CD4(+) T cells protect immunodeficient mice from lethal infection and demonstrate requirement for a protective threshold of antibody production for parasite clearance. Blood 2005, 106, 1676–1684. [Google Scholar] [CrossRef]
  132. Perez-Mazliah, D.; Gardner, P.J.; Schweighoffer, E.; McLaughlin, S.; Hosking, C.; Tumwine, I.; Davis, R.S.; Potocnik, A.J.; Tybulewicz, V.L.; Langhorne, J. Plasmodium-specific atypical memory B cells are short-lived activated B cells. eLife 2018, 7, e39800. [Google Scholar] [CrossRef]
  133. Talavera-López, C.; Bediako, Y.; Lin, J.-W.; Valletta, J.J.; Recker, M.; Langhorne, J. Comparison of whole blood and spleen transcriptional signatures over the course of an experimental malaria infection. Sci. Rep. 2019, 9, 15853. [Google Scholar] [CrossRef]
  134. Zhao, Y.; Hosking, C.; Cunningham, D.; Langhorne, J.; Lin, J.-W. Transcriptome analysis of blood and spleen in virulent and avirulent mouse malaria infection. Sci. Data 2020, 7, 253. [Google Scholar] [CrossRef]
  135. Wang, D.; Zhang, Z.; Cui, S.; Zhao, Y.; Craft, S.; Fikrig, E.; You, F. ELF4 facilitates innate host defenses against Plasmodium by activating transcription of Pf4 and Ppbp. J. Biol. Chem. 2019, 294, 7787–7796. [Google Scholar] [CrossRef]
  136. Pereira, L.M.; Assis, P.A.; de Araújo, N.M.; Durso, D.F.; Junqueira, C.; Ataíde, M.A.; Pereira, D.B.; Lien, E.; Fitzgerald, K.A.; Zamboni, D.S. Caspase-8 mediates inflammation and disease in rodent malaria. Nat. Commun. 2020, 11, 4596. [Google Scholar] [CrossRef]
  137. Oakley, M.S.; Chorazeczewski, J.K.; Aleshnick, M.; Anantharaman, V.; Majam, V.; Chawla, B.; Myers, T.G.; Su, Q.; Okoth, W.A.; Takeda, K. TCRβ-expressing macrophages induced by a pathogenic murine malaria correlate with parasite burden and enhanced phagocytic activity. PLoS ONE 2018, 13, e0201043. [Google Scholar] [CrossRef]
  138. Lai, S.M.; Sheng, J.; Gupta, P.; Renia, L.; Duan, K.; Zolezzi, F.; Karjalainen, K.; Newell, E.W.; Ruedl, C. Organ-specific fate, recruitment, and refilling dynamics of tissue-resident macrophages during blood-stage malaria. Cell Rep. 2018, 25, 3099–3109.e3. [Google Scholar] [CrossRef] [Green Version]
  139. Kim, C.C.; Nelson, C.S.; Wilson, E.B.; Hou, B.; DeFranco, A.L.; DeRisi, J.L. Splenic red pulp macrophages produce type I interferons as early sentinels of malaria infection but are dispensable for control. PLoS ONE 2012, 7, e48126. [Google Scholar]
  140. Kim, C.C.; Parikh, S.; Sun, J.C.; Myrick, A.; Lanier, L.L.; Rosenthal, P.J.; DeRisi, J.L. Experimental malaria infection triggers early expansion of natural killer cells. Infect. Immun. 2008, 76, 5873–5882. [Google Scholar] [CrossRef]
  141. Haque, A.; Best, S.E.; Ammerdorffer, A.; Desbarrieres, L.; de Oca, M.M.; Amante, F.H.; de Labastida Rivera, F.; Hertzog, P.; Boyle, G.M.; Hill, G.R.; et al. Type I interferons suppress CD4(+) T-cell-dependent parasite control during blood-stage Plasmodium infection. Eur. J. Immunol. 2011, 41, 2688–2698. [Google Scholar] [CrossRef]
  142. Cheng, Q.; Liu, J.; Pei, Y.; Zhang, Y.; Zhou, D.; Pan, W.; Zhang, J. Neddylation contributes to CD4+ T cell-mediated protective immunity against blood-stage Plasmodium infection. PLoS Pathog. 2018, 14, e1007440. [Google Scholar] [CrossRef]
  143. Fontana, M.F.; de Melo, G.L.; Anidi, C.; Hamburger, R.; Kim, C.Y.; Lee, S.Y.; Pham, J.; Kim, C.C. Macrophage colony stimulating factor derived from CD4+ T cells contributes to control of a blood-borne infection. PLoS Pathog. 2016, 12, e1006046. [Google Scholar] [CrossRef] [PubMed]
  144. Ames, R.Y.; Ting, L.-M.; Gendlina, I.; Kim, K.; Macian, F. The transcription factor NFAT1 participates in the induction of CD4+ T cell functional exhaustion during Plasmodium yoelii infection. Infect. Immun. 2017, 85, e00364-17. [Google Scholar] [CrossRef] [PubMed]
  145. Ly, A.; Liao, Y.; Pietrzak, H.; Ioannidis, L.J.; Sidwell, T.; Gloury, R.; Doerflinger, M.; Triglia, T.; Qin, R.Z.; Groom, J.R. Transcription factor T-bet in B cells modulates germinal center polarization and antibody affinity maturation in response to malaria. Cell Rep. 2019, 29, 2257–2269.e6. [Google Scholar] [CrossRef] [PubMed]
  146. Kim, C.C.; Baccarella, A.M.; Bayat, A.; Pepper, M.; Fontana, M.F. FCRL5+ memory B cells exhibit robust recall responses. Cell Rep. 2019, 27, 1446–1460.e4. [Google Scholar] [CrossRef] [PubMed]
  147. Talavera-López, C.; Capuccini, B.; Mitter, R.; Lin, J.-W.; Langhorne, J. Transcriptomes of microglia in experimental cerebral malaria in mice in the presence and absence of Type I Interferon signaling. BMC Res. Notes 2018, 11, 913. [Google Scholar] [CrossRef]
  148. Friedman, M.J. Erythrocytic mechanism of sickle cell resistance to malaria. Proc. Natl. Acad. Sci. USA 1978, 75, 1994–1997. [Google Scholar] [CrossRef]
  149. Aidoo, M.; Terlouw, D.J.; Kolczak, M.S.; McElroy, P.D.; ter Kuile, F.O.; Kariuki, S.; Nahlen, B.L.; Lal, A.A.; Udhayakumar, V. Protective effects of the sickle cell gene against malaria morbidity and mortality. Lancet 2002, 359, 1311–1312. [Google Scholar] [CrossRef]
  150. Peters, A.L.; Noorden, C.J.V. Glucose-6-phosphate dehydrogenase deficiency and malaria: Cytochemical detection of heterozygous G6PD deficiency in women. J. Histochem. Cytochem. 2009, 57, 1003–1011. [Google Scholar] [CrossRef]
  151. Avalos, S.; Mejia, R.E.; Banegas, E.; Salinas, C.; Gutierrez, L.; Fajardo, M.; Galo, S.; Pinto, A.; Mejia, A.; Fontecha, G. G6PD deficiency, primaquine treatment, and risk of haemolysis in malaria-infected patients. Malar. J. 2018, 17, 415. [Google Scholar] [CrossRef]
  152. Mockenhaupt, F.P.; Ehrhardt, S.; Gellert, S.; Otchwemah, R.N.; Dietz, E.; Anemana, S.D.; Bienzle, U. α+-thalassemia protects African children from severe malaria. Blood 2004, 104, 2003–2006. [Google Scholar] [CrossRef]
  153. Hill, A.V.; Allsopp, C.E.; Kwiatkowski, D.; Anstey, N.M.; Twumasi, P.; Rowe, P.A.; Bennett, S.; Brewster, D.; McMichael, A.J.; Greenwood, B.M. Common west African HLA antigens are associated with protection from severe malaria. Nature 1991, 352, 595–600. [Google Scholar] [CrossRef]
  154. May, J.; Lell, B.; Luty, A.J.; Meyer, C.G.; Kremsner, P.G. HLA-DQB1*0501-restricted Th1 type immune responses to Plasmodium falciparum liver stage antigen 1 protect against malaria anemia and reinfections. J. Infect. Dis. 2001, 183, 168–172. [Google Scholar] [CrossRef]
  155. Troye-Blomberg, M.; Olerup, O.; Larsson, A.; Sjoberg, K.; Perlmann, H.; Riley, E.; Lepers, J.P.; Perlmann, P. Failure to detect MHC class II associations of the human immune response induced by repeated malaria infections to the Plasmodium falciparum antigen Pf155/RESA. Int. Immunol. 1991, 3, 1043–1051. [Google Scholar] [CrossRef]
  156. Gichohi-Wainaina, W.N.; Melse-Boonstra, A.; Feskens, E.J.; Demir, A.Y.; Veenemans, J.; Verhoef, H. Tumour necrosis factor allele variants and their association with the occurrence and severity of malaria in African children: A longitudinal study. Malar. J. 2015, 14, 249. [Google Scholar] [CrossRef]
  157. McGuire, W.; Knight, J.C.; Hill, A.V.; Allsopp, C.E.; Greenwood, B.M.; Kwiatkowski, D. Severe malarial anemia and cerebral malaria are associated with different tumor necrosis factor promoter alleles. J. Infect. Dis. 1999, 179, 287–290. [Google Scholar] [CrossRef]
  158. McGuire, W.; Hill, A.V.; Allsopp, C.E.; Greenwood, B.M.; Kwiatkowski, D. Variation in the TNF-α promoter region associated with susceptibility to cerebral malaria. Nature 1994, 371, 508–511. [Google Scholar] [CrossRef]
  159. Clark, T.G.; Diakite, M.; Auburn, S.; Campino, S.; Fry, A.E.; Green, A.; Richardson, A.; Small, K.; Teo, Y.Y.; Wilson, J. Tumor necrosis factor and lymphotoxin-α polymorphisms and severe malaria in African populations. J. Infect. Dis. 2009, 199, 569–575. [Google Scholar] [CrossRef] [Green Version]
  160. Ravenhall, M.; Campino, S.; Sepulveda, N.; Manjurano, A.; Nadjm, B.; Mtove, G.; Wangai, H.; Maxwell, C.; Olomi, R.; Reyburn, H.; et al. MalariaGen, Novel genetic polymorphisms associated with severe malaria and under selective pressure in North-eastern Tanzania. PLoS Genet. 2018, 14, e1007172. [Google Scholar] [CrossRef]
  161. Malaria Genomic Epidemiology Network Writing Group; Band, G.; Rockett, K.A.; Spencer, C.C.; Kwiatkowski, D.P. A novel locus of resistance to severe malaria in a region of ancient balancing selection. Nature 2015, 526, 253–257. [Google Scholar]
  162. Timmann, C.; Thye, T.; Vens, M.; Evans, J.; May, J.; Ehmen, C.; Sievertsen, J.; Muntau, B.; Ruge, G.; Loag, W.; et al. Genome-wide association study indicates two novel resistance loci for severe malaria. Nature 2012, 489, 443–446. [Google Scholar] [CrossRef]
  163. Jallow, M.; Teo, Y.Y.; Small, K.S.; Rockett, K.A.; Deloukas, P.; Clark, T.G.; Kivinen, K.; Bojang, K.A.; Conway, D.J.; Pinder, M.; et al. Genome-wide and fine-resolution association analysis of malaria in West Africa. Nat. Genet. 2009, 41, 657–665. [Google Scholar] [CrossRef]
  164. Brisebarre, A.; Kumulungui, B.; Sawadogo, S.; Atkinson, A.; Garnier, S.; Fumoux, F.; Rihet, P. A genome scan for Plasmodium falciparum malaria identifies quantitative trait loci on chromosomes 5q31, 6p21.3, 17p12, and 19p13. Malar. J. 2014, 13, 198. [Google Scholar] [CrossRef]
  165. Maiga, B.; Dolo, A.; Touré, O.; Dara, V.; Tapily, A.; Campino, S.; Sepulveda, N.; Corran, P.; Rockett, K.; Clark, T. Fc gamma Receptor II a-H 131 R Polymorphism and Malaria Susceptibility in Sympatric Ethnic Groups, Fulani and Dogon of Mali. Scand. J. Immunol. 2014, 79, 43–50. [Google Scholar] [CrossRef] [PubMed]
  166. Vafa, M.; Maiga, B.; Berzins, K.; Hayano, M.; Bereczky, S.; Dolo, A.; Daou, M.; Arama, C.; Kouriba, B.; Färnert, A. Associations between the IL-4-590 T allele and Plasmodium falciparum infection prevalence in asymptomatic Fulani of Mali. Microbes Infect. 2007, 9, 1043–1048. [Google Scholar] [CrossRef]
  167. Vafa, M.; Maiga, B.; Israelsson, E.; Dolo, A.; Doumbo, O.K.; Troye-Blomberg, M. Impact of the IL-4-590 C/T transition on the levels of Plasmodium falciparum specific IgE, IgG, IgG subclasses and total IgE in two sympatric ethnic groups living in Mali. Microbes Infect. 2009, 11, 779–784. [Google Scholar] [CrossRef]
  168. Dolo, A.; Modiano, D.; Maiga, B.; Daou, M.; Dolo, G.; Guindo, H.; Ba, M.; Maiga, H.; Coulibaly, D.; Perlman, H. Difference in susceptibility to malaria between two sympatric ethnic groups in Mali. Am. J. Trop. Med. Hyg. 2005, 72, 243–248. [Google Scholar] [CrossRef]
  169. Hedrick, P.W. Population genetics of malaria resistance in humans. Heredity 2011, 107, 283–304. [Google Scholar] [CrossRef]
  170. McCall, M.B.; Hopman, J.; Daou, M.; Maiga, B.; Dara, V.; Ploemen, I.; Nganou-Makamdop, K.; Niangaly, A.; Tolo, Y.; Arama, C. Early interferon-γ response against Plasmodium falciparum correlates with interethnic differences in susceptibility to parasitemia between sympatric Fulani and Dogon in Mali. J. Infect. Dis. 2010, 201, 142–152. [Google Scholar] [CrossRef] [Green Version]
  171. Laroque, A.; Min-Oo, G.; Tam, M.; Radovanovic, I.; Stevenson, M.M.; Gros, P. Genetic control of susceptibility to infection with Plasmodium chabaudi chabaudi AS in inbred mouse strains. Genes Immun. 2012, 13, 155–163. [Google Scholar] [CrossRef]
  172. Stevenson, M.; Lyanga, J.; Skamene, E. Murine malaria: Genetic control of resistance to Plasmodium chabaudi. Infect. Immun. 1982, 38, 80–88. [Google Scholar] [CrossRef]
  173. Bopp, S.E.; Ramachandran, V.; Henson, K.; Luzader, A.; Lindstrom, M.; Spooner, M.; Steffy, B.M.; Suzuki, O.; Janse, C.; Waters, A.P. Genome wide analysis of inbred mouse lines identifies a locus containing ppar-γ as contributing to enhanced malaria survival. PLoS ONE 2010, 5, e10903. [Google Scholar] [CrossRef]
  174. Burt, R.A.; Marshall, V.M.; Wagglen, J.; Rodda, F.R.; Senyschen, D.; Baldwin, T.M.; Buckingham, L.A.; Foote, S.J. Mice that are congenic for the char2 locus are susceptible to malaria. Infect. Immun. 2002, 70, 4750–4753. [Google Scholar] [CrossRef] [PubMed]
  175. Guénet, J.-L.; Benavides, F.; Panthier, J.-J.; Montagutelli, X. The different categories of genetically standardized populations of laboratory mice. In Genetics of the Mouse; Springer: Berlin/Heidelberg, Germany, 2015; pp. 319–359. [Google Scholar]
  176. Yang, X.; HayGlass, K.T.; Brunham, R.C. Genetically determined differences in IL-10 and IFN-gamma responses correlate with clearance of Chlamydia trachomatis mouse pneumonitis infection. J. Immunol. 1996, 156, 4338–4344. [Google Scholar] [PubMed]
  177. Cui, S.; Chesson, C.; Hope, R. Genetic variation within and between strains of outbred Swiss mice. Lab. Anim. 1993, 27, 116–123. [Google Scholar] [CrossRef] [PubMed]
  178. Carothers, A.D.; Rudan, I.; Kolcic, I.; Polasek, O.; Hayward, C.; Wright, A.F.; Campbell, H.; Teague, P.; Hastie, N.D.; Weber, J.L. Estimating human inbreeding coefficients: Comparison of genealogical and marker heterozygosity approaches. Ann. Hum. Genet. 2006, 70 Pt 5, 666–676. [Google Scholar] [CrossRef]
  179. Lemes, R.B.; Nunes, K.; Carnavalli, J.E.P.; Kimura, L.; Mingroni-Netto, R.C.; Meyer, D.; Otto, P.A. Inbreeding estimates in human populations: Applying new approaches to an admixed Brazilian isolate. PLoS ONE 2018, 13, e0196360. [Google Scholar] [CrossRef]
  180. Abolins, S.; King, E.C.; Lazarou, L.; Weldon, L.; Hughes, L.; Drescher, P.; Raynes, J.G.; Hafalla, J.C.R.; Viney, M.E.; Riley, E.M. The comparative immunology of wild and laboratory mice, Mus musculus domesticus. Nat. Commun. 2017, 8, 14811. [Google Scholar] [CrossRef]
  181. Aylor, D.L.; Valdar, W.; Foulds-Mathes, W.; Buus, R.J.; Verdugo, R.A.; Baric, R.S.; Ferris, M.T.; Frelinger, J.A.; Heise, M.; Frieman, M.B.; et al. Genetic analysis of complex traits in the emerging Collaborative Cross. Genome Res. 2011, 21, 1213–1222. [Google Scholar] [CrossRef] [Green Version]
  182. Chesler, E.J. Out of the bottleneck: The Diversity Outcross and Collaborative Cross mouse populations in behavioral genetics research. Mamm. Genome 2014, 25, 3–11. [Google Scholar] [CrossRef]
  183. Collaborative Cross Consortium. The genome architecture of the Collaborative Cross mouse genetic reference population. Genetics 2012, 190, 389–401. [Google Scholar] [CrossRef]
  184. Threadgill, D.W.; Miller, D.R.; Churchill, G.A.; de Villena, F.P. The collaborative cross: A recombinant inbred mouse population for the systems genetic era. ILAR J. 2011, 52, 24–31. [Google Scholar] [CrossRef]
  185. Chesler, E.J.; Gatti, D.M.; Morgan, A.P.; Strobel, M.; Trepanier, L.; Oberbeck, D.; McWeeney, S.; Hitzemann, R.; Ferris, M.; McMullan, R.; et al. Diversity Outbred Mice at 21: Maintaining Allelic Variation in the Face of Selection. G3 Genes Genomes Genet. 2016, 6, 3893–3902. [Google Scholar] [CrossRef]
  186. Churchill, G.A.; Gatti, D.M.; Munger, S.C.; Svenson, K.L. The Diversity Outbred mouse population. Mamm. Genome 2012, 23, 713–718. [Google Scholar] [CrossRef]
  187. Bar, J.; Leung, J.M.; Hansen, C.; Loke, P.; Hall, A.R.; Conour, L.; Graham, A.L. Strong effects of lab-to-field environmental transitions on the bacterial intestinal microbiota of Mus musculus are modulated by Trichuris muris infection. FEMS Microbiol. Ecol. 2020, 96, fiaa167. [Google Scholar] [CrossRef]
  188. Lin, J.D.; Devlin, J.C.; Yeung, F.; McCauley, C.; Leung, J.M.; Chen, Y.H.; Cronkite, A.; Hansen, C.; Drake-Dunn, C.; Ruggles, K.V.; et al. Rewilding Nod2 and Atg16l1 Mutant Mice Uncovers Genetic and Environmental Contributions to Microbial Responses and Immune Cell Composition. Cell Host Microbe 2020, 27, 830–840.e4. [Google Scholar] [CrossRef]
  189. Yilmaz, B.; Portugal, S.; Tran, T.M.; Gozzelino, R.; Ramos, S.; Gomes, J.; Regalado, A.; Cowan, P.J.; d’Apice, A.J.; Chong, A.S.; et al. Gut microbiota elicits a protective immune response against malaria transmission. Cell 2014, 159, 1277–1289. [Google Scholar] [CrossRef]
  190. Yooseph, S.; Kirkness, E.F.; Tran, T.M.; Harkins, D.M.; Jones, M.B.; Torralba, M.G.; O’Connell, E.; Nutman, T.B.; Doumbo, S.; Doumbo, O.K.; et al. Stool microbiota composition is associated with the prospective risk of Plasmodium falciparum infection. BMC Genom. 2015, 16, 631. [Google Scholar] [CrossRef]
  191. Mandal, R.K.; Denny, J.E.; Namazzi, R.; Opoka, R.O.; Datta, D.; John, C.C.; Schmidt, N.W. Dynamic modulation of spleen germinal center reactions by gut bacteria during Plasmodium infection. Cell Rep. 2021, 35, 109094. [Google Scholar] [CrossRef]
  192. Mandal, R.K.; Crane, R.J.; Berkley, J.A.; Gumbi, W.; Wambua, J.; Ngoi, J.M.; Ndungu, F.M.; Schmidt, N.W. Longitudinal Analysis of Infant Stool Bacteria Communities before and after Acute Febrile Malaria and Artemether-Lumefantrine Treatment. J. Infect. Dis. 2019, 220, 687–698. [Google Scholar] [CrossRef]
  193. Stough, J.M.; Dearth, S.P.; Denny, J.E.; LeCleir, G.R.; Schmidt, N.W.; Campagna, S.R.; Wilhelm, S.W. Functional Characteristics of the Gut Microbiome in C57BL/6 Mice Differentially Susceptible to Plasmodium yoelii. Front. Microbiol. 2016, 7, 1520. [Google Scholar] [CrossRef]
  194. Waide, M.L.; Polidoro, R.; Powell, W.L.; Denny, J.E.; Kos, J.; Tieri, D.A.; Watson, C.T.; Schmidt, N.W. Gut Microbiota Composition Modulates the Magnitude and Quality of Germinal Centers during Plasmodium Infections. Cell Rep. 2020, 33, 108503. [Google Scholar] [CrossRef]
  195. Yeung, F.; Chen, Y.H.; Lin, J.D.; Leung, J.M.; McCauley, C.; Devlin, J.C.; Hansen, C.; Cronkite, A.; Stephens, Z.; Drake-Dunn, C.; et al. Altered Immunity of Laboratory Mice in the Natural Environment Is Associated with Fungal Colonization. Cell Host Microbe 2020, 27, 809–822.e6. [Google Scholar] [CrossRef]
  196. Beura, L.K.; Hamilton, S.E.; Bi, K.; Schenkel, J.M.; Odumade, O.A.; Casey, K.A.; Thompson, E.A.; Fraser, K.A.; Rosato, P.C.; Filali-Mouhim, A. Normalizing the environment recapitulates adult human immune traits in laboratory mice. Nature 2016, 532, 512–516. [Google Scholar] [CrossRef]
  197. Nahrendorf, W.; Scholzen, A.; Bijker, E.M.; Teirlinck, A.C.; Bastiaens, G.J.; Schats, R.; Hermsen, C.C.; Visser, L.G.; Langhorne, J.; Sauerwein, R.W. Memory B-cell and antibody responses induced by Plasmodium falciparum sporozoite immunization. J. Infect. Dis. 2014, 210, 1981–1990. [Google Scholar] [CrossRef]
  198. Coomes, S.M.; Pelly, V.S.; Kannan, Y.; Okoye, I.S.; Czieso, S.; Entwistle, L.J.; Perez-Lloret, J.; Nikolov, N.; Potocnik, A.J.; Biro, J.; et al. IFNgamma and IL-12 Restrict Th2 Responses during Helminth/Plasmodium Co-Infection and Promote IFNgamma from Th2 Cells. PLoS Pathog. 2015, 11, e1004994. [Google Scholar] [CrossRef]
  199. Schleicher, T.R.; Yang, J.; Freudzon, M.; Rembisz, A.; Craft, S.; Hamilton, M.; Graham, M.; Mlambo, G.; Tripathi, A.K.; Li, Y. A mosquito salivary gland protein partially inhibits Plasmodium sporozoite cell traversal and transmission. Nat. Commun. 2018, 9, 2908. [Google Scholar] [CrossRef]
  200. Isawa, H.; Yuda, M.; Orito, Y.; Chinzei, Y. A mosquito salivary protein inhibits activation of the plasma contact system by binding to factor XII and high molecular weight kininogen. J. Biol. Chem. 2002, 277, 27651–27658. [Google Scholar] [CrossRef]
  201. Depinay, N.; Hacini, F.; Beghdadi, W.; Peronet, R.; Mécheri, S. Mast cell-dependent down-regulation of antigen-specific immune responses by mosquito bites. J. Immunol. 2006, 176, 4141–4146. [Google Scholar] [CrossRef] [Green Version]
  202. Donovan, M.J.; Messmore, A.S.; Scrafford, D.A.; Sacks, D.L.; Kamhawi, S.; McDowell, M.A. Uninfected mosquito bites confer protection against infection with malaria parasites. Infect. Immun. 2007, 75, 2523–2530. [Google Scholar] [CrossRef]
  203. Bediako, Y.; Adams, R.; Reid, A.J.; Valletta, J.J.; Ndungu, F.M.; Sodenkamp, J.; Mwacharo, J.; Ngoi, J.M.; Kimani, D.; Kai, O.; et al. Repeated clinical malaria episodes are associated with modification of the immune system in children. BMC Med. 2019, 17, 60. [Google Scholar] [CrossRef]
  204. Coggeshall, L.; Kumm, H. Demonstration of passive immunity in experimental monkey malaria. J. Exp. Med. 1937, 66, 177–190. [Google Scholar] [CrossRef]
  205. Cohen, S.; McGregor, I.; Carrington, S.J.N. Gamma-globulin and acquired immunity to human malaria. Nature 1961, 192, 733–737. [Google Scholar] [CrossRef] [PubMed]
  206. Sabchareon, A.; Burnouf, T.; Ouattara, D.; Attanath, P.; Bouharoun-Tayoun, H.; Chantavanich, P.; Foucault, C.; Chongsuphajaisiddhi, T.; Druilhe, P. Parasitologic and clinical human response to immunoglobulin administration in falciparum malaria. Am. J. Trop. Med. Hyg. 1991, 45, 297–308. [Google Scholar] [CrossRef] [PubMed]
  207. Osier, F.H.; Feng, G.; Boyle, M.J.; Langer, C.; Zhou, J.; Richards, J.S.; McCallum, F.J.; Reiling, L.; Jaworowski, A.; Anders, R.F.; et al. Opsonic phagocytosis of Plasmodium falciparum merozoites: Mechanism in human immunity and a correlate of protection against malaria. BMC Med. 2014, 12, 108. [Google Scholar] [CrossRef] [PubMed]
  208. Turrini, F.; Ginsburg, H.; Bussolino, F.; Pescarmona, G.P.; Serra, M.V.; Arese, P. Phagocytosis of Plasmodium falciparum-infected human red blood cells by human monocytes: Involvement of immune and nonimmune determinants and dependence on parasite developmental stage. Blood 1992, 80, 801–808. [Google Scholar] [CrossRef]
  209. Yoneto, T.; Waki, S.; Takai, T.; Tagawa, Y.; Iwakura, Y.; Mizuguchi, J.; Nariuchi, H.; Yoshimoto, T. A critical role of Fc receptor-mediated antibody-dependent phagocytosis in the host resistance to blood-stage Plasmodium berghei XAT infection. J. Immunol. 2001, 166, 6236–6241. [Google Scholar] [CrossRef]
  210. Akter, J.; Khoury, D.S.; Aogo, R.; Lansink, L.I.; SheelaNair, A.; Thomas, B.S.; Laohamonthonkul, P.; Pernold, C.P.; Dixon, M.W.; Soon, M.S. Plasmodium-specific antibodies block in vivo parasite growth without clearing infected red blood cells. PLoS Pathog. 2019, 15, e1007599. [Google Scholar] [CrossRef]
  211. McCallum, F.J.; Persson, K.E.; Mugyenyi, C.K.; Fowkes, F.J.; Simpson, J.A.; Richards, J.S.; Williams, T.N.; Marsh, K.; Beeson, J.G. Acquisition of growth-inhibitory antibodies against blood-stage Plasmodium falciparum. PLoS ONE 2008, 3, e3571. [Google Scholar] [CrossRef] [Green Version]
  212. Osier, F.H.; Fegan, G.; Polley, S.D.; Murungi, L.; Verra, F.; Tetteh, K.K.; Lowe, B.; Mwangi, T.; Bull, P.C.; Thomas, A.W.; et al. Breadth and magnitude of antibody responses to multiple Plasmodium falciparum merozoite antigens are associated with protection from clinical malaria. Infect. Immun. 2008, 76, 2240–2248. [Google Scholar] [CrossRef]
  213. Rono, J.; Osier, F.H.; Olsson, D.; Montgomery, S.; Mhoja, L.; Rooth, I.; Marsh, K.; Färnert, A. Breadth of anti-merozoite antibody responses is associated with the genetic diversity of asymptomatic Plasmodium falciparum infections and protection against clinical malaria. Clin. Infect. Dis. 2013, 57, 1409–1416. [Google Scholar] [CrossRef]
  214. Chan, J.-A.; Howell, K.B.; Reiling, L.; Ataide, R.; Mackintosh, C.L.; Fowkes, F.J.; Petter, M.; Chesson, J.M.; Langer, C.; Warimwe, G.M. Targets of antibodies against Plasmodium falciparum-infected erythrocytes in malaria immunity. J. Clin. Investig. 2012, 122, 3227–3238. [Google Scholar] [CrossRef]
  215. Nogaro, S.I.; Hafalla, J.C.; Walther, B.; Remarque, E.J.; Tetteh, K.K.; Conway, D.J.; Riley, E.M.; Walther, M. The breadth, but not the magnitude, of circulating memory B cell responses to P. falciparum increases with age/exposure in an area of low transmission. PLoS ONE 2011, 6, e25582. [Google Scholar] [CrossRef]
  216. Rotman, H.L.; Daly, T.M.; Clynes, R.; Long, C.A. Fc receptors are not required for antibody-mediated protection against lethal malaria challenge in a mouse model. J. Immunol. 1998, 161, 1908–1912. [Google Scholar]
  217. Grun, J.; Weidanz, W. Antibody-independent immunity to reinfection malaria in B-cell-deficient mice. Infect. Immun. 1983, 41, 1197–1204. [Google Scholar] [CrossRef]
  218. Kubagawa, H.; Honjo, K.; Ohkura, N.; Sakaguchi, S.; Radbruch, A.; Melchers, F.; Jani, P.K. Functional Roles of the IgM Fc Receptor in the Immune System. Front. Immunol. 2019, 10, 945. [Google Scholar] [CrossRef]
  219. Kurtovic, L.; Behet, M.C.; Feng, G.; Reiling, L.; Chelimo, K.; Dent, A.E.; Mueller, I.; Kazura, J.W.; Sauerwein, R.W.; Fowkes, F.J.I.; et al. Human antibodies activate complement against Plasmodium falciparum sporozoites, and are associated with protection against malaria in children. BMC Med. 2018, 16, 61. [Google Scholar] [CrossRef]
  220. Boyle, M.J.; Reiling, L.; Feng, G.; Langer, C.; Osier, F.H.; Aspeling-Jones, H.; Cheng, Y.S.; Stubbs, J.; Tetteh, K.K.; Conway, D.J.; et al. Human antibodies fix complement to inhibit Plasmodium falciparum invasion of erythrocytes and are associated with protection against malaria. Immunity 2015, 42, 580–590. [Google Scholar] [CrossRef]
  221. Taylor, P.R.; Seixas, E.; Walport, M.J.; Langhorne, J.; Botto, M. Complement contributes to protective immunity against reinfection by Plasmodium chabaudi chabaudi parasites. Infect. Immun. 2001, 69, 3853–3859. [Google Scholar] [CrossRef]
  222. Grun, J.L.; Weidanz, W.P. Immunity to Plasmodium chabaudi adami in the B-cell-deficient mouse. Nature 1981, 290, 143–145. [Google Scholar] [CrossRef]
  223. Weidanz, W.P.; LaFleur, G.; Brown, A.; Burns, J.M.; Gramaglia, I.; van der Heyde, H.C. γδ T cells but not NK cells are essential for cell-mediated immunity against Plasmodium chabaudi malaria. Infect. Immun. 2010, 78, 4331–4340. [Google Scholar] [CrossRef]
  224. Kinyanjui, S.M.; Bejon, P.; Osier, F.H.; Bull, P.C.; Marsh, K. What you see is not what you get: Implications of the brevity of antibody responses to malaria antigens and transmission heterogeneity in longitudinal studies of malaria immunity. Malar. J. 2009, 8, 242. [Google Scholar] [CrossRef]
  225. Kinyanjui, S.M.; Conway, D.J.; Lanar, D.E.; Marsh, K. IgG antibody responses to Plasmodium falciparum merozoite antigens in Kenyan children have a short half-life. Malar. J. 2007, 6, 82. [Google Scholar] [CrossRef] [PubMed]
  226. Portugal, S.; Pierce, S.K.; Crompton, P.D. Young lives lost as B cells falter: What we are learning about antibody responses in malaria. J. Immunol. 2013, 190, 3039–3046. [Google Scholar] [CrossRef] [PubMed]
  227. Olatunde, A.C.; Hale, J.S.; Lamb, T.J. Cytokine-skewed Tfh cells: Functional consequences for B cell help. Trends Immunol. 2021, 42, 536–550. [Google Scholar] [CrossRef] [PubMed]
  228. Elsner, R.A.; Shlomchik, M.J. Germinal Center and Extrafollicular B Cell Responses in Vaccination, Immunity, and Autoimmunity. Immunity 2020, 53, 1136–1150. [Google Scholar] [CrossRef]
  229. Inamine, A.; Takahashi, Y.; Baba, N.; Miyake, K.; Tokuhisa, T.; Takemori, T.; Abe, R. Two waves of memory B-cell generation in the primary immune response. Int. Immunol. 2005, 17, 581–589. [Google Scholar] [CrossRef]
  230. Takemori, T.; Kaji, T.; Takahashi, Y.; Shimoda, M.; Rajewsky, K. Generation of memory B cells inside and outside germinal centers. Eur. J. Immunol. 2014, 44, 1258–1264. [Google Scholar] [CrossRef]
  231. Toyama, H.; Okada, S.; Hatano, M.; Takahashi, Y.; Takeda, N.; Ichii, H.; Takemori, T.; Kuroda, Y.; Tokuhisa, T. Memory B cells without somatic hypermutation are generated from Bcl6-deficient B cells. Immunity 2002, 17, 329–339. [Google Scholar] [CrossRef]
  232. Perez-Mazliah, D.; Ng, D.H.; Freitas do Rosario, A.P.; McLaughlin, S.; Mastelic-Gavillet, B.; Sodenkamp, J.; Kushinga, G.; Langhorne, J. Disruption of IL-21 signaling affects T cell-B cell interactions and abrogates protective humoral immunity to malaria. PLoS Pathog. 2015, 11, e1004715. [Google Scholar] [CrossRef] [Green Version]
  233. Sebina, I.; Fogg, L.G.; James, K.R.; Soon, M.S.F.; Akter, J.; Thomas, B.S.; Hill, G.R.; Engwerda, C.R.; Haque, A. IL-6 promotes CD4(+) T-cell and B-cell activation during Plasmodium infection. Parasite Immunol. 2017, 39, e12455. [Google Scholar] [CrossRef]
  234. Sebina, I.; James, K.R.; Soon, M.S.; Fogg, L.G.; Best, S.E.; Labastida Rivera, F.; Montes de Oca, M.; Amante, F.H.; Thomas, B.S.; Beattie, L.; et al. IFNAR1-Signalling Obstructs ICOS-Mediated Humoral Immunity during Non-Lethal Blood-Stage Plasmodium Infection. PLoS Pathog. 2016, 12, e1005999. [Google Scholar] [CrossRef]
  235. Hopp, C.S.; Sekar, P.; Diouf, A.; Miura, K.; Boswell, K.; Skinner, J.; Tipton, C.M.; Peterson, M.E.; Chambers, M.J.; Andrews, S.; et al. Plasmodium falciparum-specific IgM B cells dominate in children, expand with malaria, and produce functional IgM. J. Exp. Med. 2021, 218, e20200901. [Google Scholar] [CrossRef]
  236. Thouvenel, C.D.; Fontana, M.F.; Netland, J.; Krishnamurty, A.T.; Takehara, K.K.; Chen, Y.; Singh, S.; Miura, K.; Keitany, G.J.; Lynch, E.M.; et al. Multimeric antibodies from antigen-specific human IgM+ memory B cells restrict Plasmodium parasites. J. Exp. Med. 2021, 218, e20200942. [Google Scholar] [CrossRef]
  237. Pietrzak, H.M.; Ioannidis, L.J.; Hansen, D.S. IgM(+) memory B cells induced in response to Plasmodium berghei adopt a germinal centre B cell phenotype during secondary infection. Parasitology 2020, 147, 994–998. [Google Scholar] [CrossRef]
  238. Wykes, M.N.; Zhou, Y.-H.; Liu, X.Q.; Good, M.F. Plasmodium yoelii can ablate vaccine-induced long-term protection in mice. J. Immunol. 2005, 175, 2510–2516. [Google Scholar] [CrossRef]
  239. Obeng-Adjei, N.; Portugal, S.; Tran, T.M.; Yazew, T.B.; Skinner, J.; Li, S.; Jain, A.; Felgner, P.L.; Doumbo, O.K.; Kayentao, K.; et al. Circulating Th1-cell-type Tfh cells that exhibit impaired B cell help are preferentially activated during acute malaria in children. Cell Rep. 2015, 13, 425–439. [Google Scholar] [CrossRef]
  240. Arroyo, E.N.; Pepper, M. B cells are sufficient to prime the dominant CD4+ Tfh response to Plasmodium infection. J. Exp. Med. 2020, 217, e20190849. [Google Scholar] [CrossRef]
  241. Wikenheiser, D.J.; Ghosh, D.; Kennedy, B.; Stumhofer, J.S. The Costimulatory Molecule ICOS Regulates Host Th1 and Follicular Th Cell Differentiation in Response to Plasmodium chabaudi chabaudi AS Infection. J. Immunol. 2016, 196, 778–791. [Google Scholar] [CrossRef]
  242. O’Neal, K.A.; Latham, L.E.; Ntirandekura, E.; Foscue, C.L.; Stumhofer, J.S. ICOS Expression Is Required for Maintenance but Not the Formation of Germinal Centers in the Spleen in Response to Plasmodium yoelii Infection. Infect. Immun. 2022, 90, e0046821. [Google Scholar] [CrossRef]
  243. Butler, N.S.; Moebius, J.; Pewe, L.L.; Traore, B.; Doumbo, O.K.; Tygrett, L.T.; Waldschmidt, T.J.; Crompton, P.D.; Harty, J.T. Therapeutic blockade of PD-L1 and LAG-3 rapidly clears established blood-stage Plasmodium infection. Nat. Immunol. 2012, 13, 188. [Google Scholar] [CrossRef]
  244. Kurup, S.P.; Obeng-Adjei, N.; Anthony, S.M.; Traore, B.; Doumbo, O.K.; Butler, N.S.; Crompton, P.D.; Harty, J.T. Regulatory T cells impede acute and long-term immunity to blood-stage malaria through CTLA-4. Nat. Med. 2017, 23, 1220–1225. [Google Scholar] [CrossRef]
  245. Obeng-Adjei, N.; Portugal, S.; Holla, P.; Li, S.; Sohn, H.; Ambegaonkar, A.; Skinner, J.; Bowyer, G.; Doumbo, O.K.; Traore, B.; et al. Malaria-induced interferon-gamma drives the expansion of Tbethi atypical memory B cells. PLoS Pathog. 2017, 13, e1006576. [Google Scholar] [CrossRef]
  246. Hopp, C.S.; Skinner, J.; Anzick, S.L.; Tipton, C.M.; Peterson, M.E.; Li, S.; Doumbo, S.; Kayentao, K.; Ongoiba, A.; Martens, C.; et al. Atypical B cells up-regulate costimulatory molecules during malaria and secrete antibodies with T follicular helper cell support. Sci. Immunol. 2022, 7, eabn1250. [Google Scholar] [CrossRef]
  247. Guthmiller, J.J.; Graham, A.C.; Zander, R.A.; Pope, R.L.; Butler, N.S. Cutting Edge: IL-10 Is Essential for the Generation of Germinal Center B Cell Responses and Anti-Plasmodium Humoral Immunity. J. Immunol. 2017, 198, 617–622. [Google Scholar] [CrossRef]
  248. Ryg-Cornejo, V.; Ioannidis, L.J.; Ly, A.; Chiu, C.Y.; Tellier, J.; Hill, D.L.; Preston, S.P.; Pellegrini, M.; Yu, D.; Nutt, S.L.; et al. Severe Malaria Infections Impair Germinal Center Responses by Inhibiting T Follicular Helper Cell Differentiation. Cell Rep. 2016, 14, 68–81. [Google Scholar] [CrossRef]
  249. Zander, R.A.; Obeng-Adjei, N.; Guthmiller, J.J.; Kulu, D.I.; Li, J.; Ongoiba, A.; Traore, B.; Crompton, P.D.; Butler, N.S. PD-1 Co-inhibitory and OX40 Co-stimulatory Crosstalk Regulates Helper T Cell Differentiation and Anti-Plasmodium Humoral Immunity. Cell Host Microbe 2015, 17, 628–641. [Google Scholar] [CrossRef]
  250. James, K.R.; Soon, M.S.F.; Sebina, I.; Fernandez-Ruiz, D.; Davey, G.; Liligeto, U.N.; Nair, A.S.; Fogg, L.G.; Edwards, C.L.; Best, S.E.; et al. IFN Regulatory Factor 3 Balances Th1 and T Follicular Helper Immunity during Nonlethal Blood-Stage Plasmodium Infection. J. Immunol. 2018, 200, 1443–1456. [Google Scholar] [CrossRef]
  251. Carpio, V.H.; Aussenac, F.; Puebla-Clark, L.; Wilson, K.D.; Villarino, A.V.; Dent, A.L.; Stephens, R. T Helper Plasticity Is Orchestrated by STAT3, Bcl6, and Blimp-1 Balancing Pathology and Protection in Malaria. iScience 2020, 23, 101310. [Google Scholar] [CrossRef]
  252. Cope, A.; Le Friec, G.; Cardone, J.; Kemper, C. The Th1 life cycle: Molecular control of IFN-γ to IL-10 switching. Trends Immunol. 2011, 32, 278–286. [Google Scholar] [CrossRef]
  253. Surette, F.A.; Guthmiller, J.J.; Li, L.; Sturtz, A.J.; Vijay, R.; Pope, R.L.; McClellan, B.L.; Pack, A.D.; Zander, R.A.; Shao, P.; et al. Extrafollicular CD4 T cell-derived IL-10 functions rapidly and transiently to support anti-Plasmodium humoral immunity. PLoS Pathog. 2021, 17, e1009288. [Google Scholar] [CrossRef]
  254. Wikenheiser, D.J.; Brown, S.L.; Lee, J.; Stumhofer, J.S. NK1.1 Expression Defines a Population of CD4(+) Effector T Cells Displaying Th1 and Tfh Cell Properties That Support Early Antibody Production During Plasmodium yoelii Infection. Front. Immunol. 2018, 9, 2277. [Google Scholar] [CrossRef]
  255. Troye-Blomberg, M.; Arama, C.; Quin, J.; Bujila, I.; Ostlund Farrants, A.K. What will studies of Fulani individuals naturally exposed to malaria teach us about protective immunity to malaria? Scand. J. Immunol. 2020, 92, e12932. [Google Scholar] [CrossRef] [PubMed]
  256. Walther, M.; Woodruff, J.; Edele, F.; Jeffries, D.; Tongren, J.E.; King, E.; Andrews, L.; Bejon, P.; Gilbert, S.C.; De Souza, J.B.; et al. Innate immune responses to human malaria: Heterogeneous cytokine responses to blood-stage Plasmodium falciparum correlate with parasitological and clinical outcomes. J. Immunol. 2006, 177, 5736–5745. [Google Scholar] [CrossRef] [PubMed]
  257. Franklin, B.S.; Parroche, P.; Ataide, M.A.; Lauw, F.; Ropert, C.; de Oliveira, R.B.; Pereira, D.; Tada, M.S.; Nogueira, P.; da Silva, L.H.; et al. Malaria primes the innate immune response due to interferon-gamma induced enhancement of toll-like receptor expression and function. Proc. Natl. Acad. Sci. USA 2009, 106, 5789–5794. [Google Scholar] [CrossRef] [PubMed]
  258. Fu, Y.; Ding, Y.; Zhou, T.; Fu, X.; Xu, W. Plasmodium yoelii blood-stage primes macrophage-mediated innate immune response through modulation of toll-like receptor signalling. Malar. J. 2012, 11, 104. [Google Scholar] [CrossRef]
  259. Sun, T.; Chakrabarti, C. Schizonts, merozoites, and phagocytosis in falciparum malaria. Ann. Clin. Lab. Sci. 1985, 15, 465–469. [Google Scholar]
  260. Wickramasinghe, S.N.; Phillips, R.E.; Looareesuwan, S.; Warrell, D.A.; Hughes, M. The bone marrow in human cerebral malaria: Parasite sequestration within sinusoids. Br. J. Haematol. 1987, 66, 295–306. [Google Scholar] [CrossRef]
  261. Arora, G.; Hart, G.T.; Manzella-Lapeira, J.; Doritchamou, J.Y.; Narum, D.L.; Thomas, L.M.; Brzostowski, J.; Rajagopalan, S.; Doumbo, O.K.; Traore, B.; et al. NK cells inhibit Plasmodium falciparum growth in red blood cells via antibody-dependent cellular cytotoxicity. eLife 2018, 7, e36806. [Google Scholar] [CrossRef]
  262. Burrack, K.S.; Hart, G.T.; Hamilton, S.E. Contributions of natural killer cells to the immune response against Plasmodium. Malar. J. 2019, 18, 321. [Google Scholar] [CrossRef]
  263. Hart, G.T.; Tran, T.M.; Theorell, J.; Schlums, H.; Arora, G.; Rajagopalan, S.; Sangala, A.D.J.; Welsh, K.J.; Traore, B.; Pierce, S.K.; et al. Adaptive NK cells in people exposed to Plasmodium falciparum correlate with protection from malaria. J. Exp. Med. 2019, 216, 1280–1290. [Google Scholar] [CrossRef] [Green Version]
  264. Sherratt, S.; Patel, A.; Baker, D.A.; Riley, E.M.; Goodier, M.R. Differential IL-18 Dependence of Canonical and Adaptive NK Cells for Antibody Dependent Responses to P. falciparum. Front. Immunol. 2020, 11, 533. [Google Scholar] [CrossRef]
  265. Jagannathan, P.; Lutwama, F.; Boyle, M.J.; Nankya, F.; Farrington, L.A.; McIntyre, T.I.; Bowen, K.; Naluwu, K.; Nalubega, M.; Musinguzi, K. Vδ2+ T cell response to malaria correlates with protection from infection but is attenuated with repeated exposure. Sci. Rep. 2017, 7, 11487. [Google Scholar] [CrossRef]
  266. Schofield, L.; Novakovic, S.; Gerold, P.; Schwarz, R.T.; McConville, M.J.; Tachado, S.D. Glycosylphosphatidylinositol toxin of Plasmodium up-regulates intercellular adhesion molecule-1, vascular cell adhesion molecule-1, and E-selectin expression in vascular endothelial cells and increases leukocyte and parasite cytoadherence via tyrosine kinase-dependent signal transduction. J. Immunol. 1996, 156, 1886–1896. [Google Scholar]
  267. Gowda, D.C.; Wu, X. Parasite Recognition and Signaling Mechanisms in Innate Immune Responses to Malaria. Front. Immunol. 2018, 9, 3006. [Google Scholar] [CrossRef]
  268. Schofield, L.; Hackett, F. Signal transduction in host cells by a glycosylphosphatidylinositol toxin of malaria parasites. J. Exp. Med. 1993, 177, 145–153. [Google Scholar] [CrossRef]
  269. Schofield, L.; Hewitt, M.C.; Evans, K.; Siomos, M.A.; Seeberger, P.H. Synthetic GPI as a candidate anti-toxic vaccine in a model of malaria. Nature 2002, 418, 785–789. [Google Scholar] [CrossRef]
  270. Vijaykumar, M.; Naik, R.S.; Gowda, D.C. Plasmodium falciparum glycosylphosphatidylinositol-induced TNF-alpha secretion by macrophages is mediated without membrane insertion or endocytosis. J. Biol. Chem. 2001, 276, 6909–6912. [Google Scholar] [CrossRef]
  271. Krishnegowda, G.; Hajjar, A.M.; Zhu, J.; Douglass, E.J.; Uematsu, S.; Akira, S.; Woods, A.S.; Gowda, D.C. Induction of proinflammatory responses in macrophages by the glycosylphosphatidylinositols of Plasmodium falciparum: Cell signaling receptors, glycosylphosphatidylinositol (GPI) structural requirement, and regulation of GPI activity. J. Biol. Chem. 2005, 280, 8606–8616. [Google Scholar] [CrossRef]
  272. Franklin, B.S.; Rodrigues, S.O.; Antonelli, L.R.; Oliveira, R.V.; Goncalves, A.M.; Sales-Junior, P.A.; Valente, E.P.; Alvarez-Leite, J.I.; Ropert, C.; Golenbock, D.T.; et al. MyD88-dependent activation of dendritic cells and CD4(+) T lymphocytes mediates symptoms, but is not required for the immunological control of parasites during rodent malaria. Microbes Infect. 2007, 9, 881–890. [Google Scholar] [CrossRef]
  273. Gowda, N.M.; Wu, X.; Gowda, D.C. TLR9 and MyD88 are crucial for the development of protective immunity to malaria. J. Immunol. 2012, 188, 5073–5085. [Google Scholar] [CrossRef]
  274. Coban, C.; Ishii, K.J.; Uematsu, S.; Arisue, N.; Sato, S.; Yamamoto, M.; Kawai, T.; Takeuchi, O.; Hisaeda, H.; Horii, T.; et al. Pathological role of Toll-like receptor signaling in cerebral malaria. Int. Immunol. 2007, 19, 67–79. [Google Scholar] [CrossRef] [Green Version]
  275. Pham, T.T.; Lamb, T.J.; Deroost, K.; Opdenakker, G.; Van den Steen, P.E. Hemozoin in Malarial Complications: More Questions Than Answers. Trends Parasitol. 2021, 37, 226–239. [Google Scholar] [CrossRef]
  276. Pichyangkul, S.; Yongvanitchit, K.; Kum-arb, U.; Hemmi, H.; Akira, S.; Krieg, A.M.; Heppner, D.G.; Stewart, V.A.; Hasegawa, H.; Looareesuwan, S.; et al. Malaria blood stage parasites activate human plasmacytoid dendritic cells and murine dendritic cells through a Toll-like receptor 9-dependent pathway. J. Immunol. 2004, 172, 4926–4933. [Google Scholar] [CrossRef]
  277. Lamphier, M.S.; Sirois, C.M.; Verma, A.; Golenbock, D.T.; Latz, E. TLR9 and the recognition of self and non-self nucleic acids. Ann. N. Y. Acad. Sci. 2006, 1082, 31–43. [Google Scholar] [CrossRef]
  278. Wu, X.; Gowda, N.M.; Kumar, S.; Gowda, D.C. Protein-DNA complex is the exclusive malaria parasite component that activates dendritic cells and triggers innate immune responses. J. Immunol. 2010, 184, 4338–4348. [Google Scholar] [CrossRef]
  279. Parroche, P.; Lauw, F.N.; Goutagny, N.; Latz, E.; Monks, B.G.; Visintin, A.; Halmen, K.A.; Lamphier, M.; Olivier, M.; Bartholomeu, D.C.; et al. Malaria hemozoin is immunologically inert but radically enhances innate responses by presenting malaria DNA to Toll-like receptor 9. Proc. Natl. Acad. Sci. USA 2007, 104, 1919–1924. [Google Scholar] [CrossRef]
  280. Millington, O.R.; Di Lorenzo, C.; Phillips, R.S.; Garside, P.; Brewer, J.M. Suppression of adaptive immunity to heterologous antigens during Plasmodium infection through hemozoin-induced failure of dendritic cell function. J. Biol. 2006, 5, 5. [Google Scholar] [CrossRef]
  281. Rivera-Correa, J.; Guthmiller, J.J.; Vijay, R.; Fernandez-Arias, C.; Pardo-Ruge, M.A.; Gonzalez, S.; Butler, N.S.; Rodriguez, A. Plasmodium DNA-mediated TLR9 activation of T-bet(+) B cells contributes to autoimmune anaemia during malaria. Nat. Commun. 2017, 8, 1282. [Google Scholar] [CrossRef]
  282. Jaschke, A.; Coulibaly, B.; Remarque, E.J.; Bujard, H.; Epp, C. Merozoite Surface Protein 1 from Plasmodium falciparum Is a Major Target of Opsonizing Antibodies in Individuals with Acquired Immunity against Malaria. Clin. Vaccine Immunol. 2017, 24, e00155-17. [Google Scholar] [CrossRef]
  283. Antonelli, L.R.; Leoratti, F.M.; Costa, P.A.; Rocha, B.C.; Diniz, S.Q.; Tada, M.S.; Pereira, D.B.; Teixeira-Carvalho, A.; Golenbock, D.T.; Goncalves, R.; et al. The CD14+CD16+ inflammatory monocyte subset displays increased mitochondrial activity and effector function during acute Plasmodium vivax malaria. PLoS Pathog. 2014, 10, e1004393. [Google Scholar] [CrossRef]
  284. McGilvray, I.D.; Serghides, L.; Kapus, A.; Rotstein, O.D.; Kain, K.C. Nonopsonic monocyte/macrophage phagocytosis of Plasmodium falciparum-parasitized erythrocytes: A role for CD36 in malarial clearance. Blood 2000, 96, 3231–3240. [Google Scholar] [CrossRef]
  285. Rodrigues, D.A.S.; Prestes, E.B.; Gama, A.M.S.; Silva, L.S.; Pinheiro, A.A.S.; Ribeiro, J.M.C.; Campos, R.M.P.; Pimentel-Coelho, P.M.; De Souza, H.S.; Dicko, A.; et al. CXCR4 and MIF are required for neutrophil extracellular trap release triggered by Plasmodium-infected erythrocytes. PLoS Pathog. 2020, 16, e1008230. [Google Scholar] [CrossRef] [PubMed]
  286. Korbel, D.S.; Newman, K.C.; Almeida, C.R.; Davis, D.M.; Riley, E.M. Heterogeneous human NK cell responses to Plasmodium falciparum-infected erythrocytes. J. Immunol. 2005, 175, 7466–7473. [Google Scholar] [CrossRef] [PubMed]
  287. Artavanis-Tsakonas, K.; Riley, E.M. Innate immune response to malaria: Rapid induction of IFN-gamma from human NK cells by live Plasmodium falciparum-infected erythrocytes. J. Immunol. 2002, 169, 2956–2963. [Google Scholar] [CrossRef] [PubMed]
  288. Newman, K.C.; Korbel, D.S.; Hafalla, J.C.; Riley, E.M. Cross-talk with myeloid accessory cells regulates human natural killer cell interferon-gamma responses to malaria. PLoS Pathog. 2006, 2, e118. [Google Scholar] [CrossRef] [PubMed]
  289. McCall, M.B.; Roestenberg, M.; Ploemen, I.; Teirlinck, A.; Hopman, J.; de Mast, Q.; Dolo, A.; Doumbo, O.K.; Luty, A.; van der Ven, A.J.; et al. Memory-like IFN-gamma response by NK cells following malaria infection reveals the crucial role of T cells in NK cell activation by P. falciparum. Eur. J. Immunol. 2010, 40, 3472–3477. [Google Scholar] [CrossRef]
  290. Stegmann, K.A.; De Souza, J.B.; Riley, E.M. IL-18-induced expression of high-affinity IL-2R on murine NK cells is essential for NK-cell IFN-γ production during murine Plasmodium yoelii infection. Eur. J. Immunol. 2015, 45, 3431–3440. [Google Scholar] [CrossRef]
  291. Jagannathan, P.; Kim, C.C.; Greenhouse, B.; Nankya, F.; Bowen, K.; Eccles-James, I.; Muhindo, M.K.; Arinaitwe, E.; Tappero, J.W.; Kamya, M.R. Loss and dysfunction of Vδ2+ γδ T cells are associated with clinical tolerance to malaria. Sci. Transl. Med. 2014, 6, 251ra117. [Google Scholar] [CrossRef]
  292. Farrington, L.A.; Jagannathan, P.; McIntyre, T.I.; Vance, H.M.; Bowen, K.; Boyle, M.J.; Nankya, F.; Wamala, S.; Auma, A.; Nalubega, M. Frequent malaria drives progressive Vδ2 T-cell loss, dysfunction, and CD16 up-regulation during early childhood. J. Infect. Dis. 2016, 213, 1483–1490. [Google Scholar] [CrossRef]
  293. Inoue, S.I.; Niikura, M.; Asahi, H.; Kawakami, Y.; Kobayashi, F. γδ T cells modulate humoral immunity against Plasmodium berghei infection. Immunology 2018, 155, 519–532. [Google Scholar] [CrossRef]
  294. Farrington, L.A.; Callaway, P.C.; Vance, H.M.; Baskevitch, K.; Lutz, E.; Warrier, L.; McIntyre, T.I.; Budker, R.; Jagannathan, P.; Nankya, F. Opsonized antigen activates Vδ2+ T cells via CD16/FCγRIIIa in individuals with chronic malaria exposure. PLoS Pathog. 2020, 16, e1008997. [Google Scholar] [CrossRef]
  295. Junqueira, C.; Polidoro, R.B.; Castro, G.; Absalon, S.; Liang, Z.; Sen Santara, S.; Crespo, A.; Pereira, D.B.; Gazzinelli, R.T.; Dvorin, J.D.; et al. Gammadelta T cells suppress Plasmodium falciparum blood-stage infection by direct killing and phagocytosis. Nat. Immunol. 2021, 22, 347–357. [Google Scholar] [CrossRef]
  296. Mohan, K.; Moulin, P.; Stevenson, M.M. Natural killer cell cytokine production, not cytotoxicity, contributes to resistance against blood-stage Plasmodium chabaudi AS infection. J. Immunol. 1997, 159, 4990–4998. [Google Scholar]
  297. Ing, R.; Stevenson, M.M. Dendritic cell and NK cell reciprocal cross talk promotes gamma interferon-dependent immunity to blood-stage Plasmodium chabaudi AS infection in mice. Infect. Immun. 2009, 77, 770–782. [Google Scholar] [CrossRef]
  298. Ng, S.S.; Souza-Fonseca-Guimaraes, F.; Rivera, F.L.; Amante, F.H.; Kumar, R.; Gao, Y.; Sheel, M.; Beattie, L.; Montes de Oca, M.; Guillerey, C.; et al. Rapid loss of group 1 innate lymphoid cells during blood stage Plasmodium infection. Clin. Transl. Immunol. 2018, 7, e1003. [Google Scholar] [CrossRef]
  299. Teirlinck, A.C.; McCall, M.B.; Roestenberg, M.; Scholzen, A.; Woestenenk, R.; de Mast, Q.; van der Ven, A.J.; Hermsen, C.C.; Luty, A.J.; Sauerwein, R.W. Longevity and composition of cellular immune responses following experimental Plasmodium falciparum malaria infection in humans. PLoS Pathog. 2011, 7, e1002389. [Google Scholar] [CrossRef]
  300. Walther, M.; Tongren, J.E.; Andrews, L.; Korbel, D.; King, E.; Fletcher, H.; Andersen, R.F.; Bejon, P.; Thompson, F.; Dunachie, S.J.; et al. Upregulation of TGF-beta, FOXP3, and CD4+CD25+ regulatory T cells correlates with more rapid parasite growth in human malaria infection. Immunity 2005, 23, 287–296. [Google Scholar] [CrossRef]
  301. Portugal, S.; Moebius, J.; Skinner, J.; Doumbo, S.; Doumtabe, D.; Kone, Y.; Dia, S.; Kanakabandi, K.; Sturdevant, D.E.; Virtaneva, K. Exposure-dependent control of malaria-induced inflammation in children. PLoS Pathog. 2014, 10, e1004079. [Google Scholar] [CrossRef]
  302. Portugal, S.; Tran, T.M.; Ongoiba, A.; Bathily, A.; Li, S.; Doumbo, S.; Skinner, J.; Doumtabe, D.; Kone, Y.; Sangala, J. Treatment of chronic asymptomatic Plasmodium falciparum infection does not increase the risk of clinical malaria upon reinfection. Clin. Infect. Dis. 2017, 64, 645–653. [Google Scholar] [CrossRef]
  303. Su, Z.; Stevenson, M.M. Central role of endogenous gamma interferon in protective immunity against blood-stage Plasmodium chabaudi AS infection. Infect. Immun. 2000, 68, 4399–4406. [Google Scholar] [CrossRef]
  304. King, T.; Lamb, T. Interferon-γ: The Jekyll and Hyde of malaria. PLoS Pathog. 2015, 11, e1005118. [Google Scholar] [CrossRef]
  305. Walther, M.; Jeffries, D.; Finney, O.C.; Njie, M.; Ebonyi, A.; Deininger, S.; Lawrence, E.; Ngwa-Amambua, A.; Jayasooriya, S.; Cheeseman, I.H.; et al. Distinct roles for FOXP3 and FOXP3 CD4 T cells in regulating cellular immunity to uncomplicated and severe Plasmodium falciparum malaria. PLoS Pathog. 2009, 5, e1000364. [Google Scholar] [CrossRef]
  306. Wenisch, C.; Parschalk, B.; Burgmann, H.; Looareesuwan, S.; Graninger, W. Decreased serum levels of TGF-beta in patients with acute Plasmodium falciparum malaria. J. Clin. Immunol. 1995, 15, 69–73. [Google Scholar] [CrossRef]
  307. Perkins, D.J.; Weinberg, J.B.; Kremsner, P.G. Reduced interleukin-12 and transforming growth factor-beta1 in severe childhood malaria: Relationship of cytokine balance with disease severity. J. Infect. Dis. 2000, 182, 988–992. [Google Scholar] [CrossRef]
  308. Chaiyaroj, S.C.; Rutta, A.S.; Muenthaisong, K.; Watkins, P.; Na Ubol, M.; Looareesuwan, S. Reduced levels of transforming growth factor-beta1, interleukin-12 and increased migration inhibitory factor are associated with severe malaria. Acta Trop. 2004, 89, 319–327. [Google Scholar] [CrossRef]
  309. Imai, T.; Shen, J.; Chou, B.; Duan, X.; Tu, L.; Tetsutani, K.; Moriya, C.; Ishida, H.; Hamano, S.; Shimokawa, C.; et al. Involvement of CD8+ T cells in protective immunity against murine blood-stage infection with Plasmodium yoelii 17XL strain. Eur. J. Immunol. 2010, 40, 1053–1061. [Google Scholar] [CrossRef]
  310. Imai, T.; Ishida, H.; Suzue, K.; Taniguchi, T.; Okada, H.; Shimokawa, C.; Hisaeda, H. Cytotoxic activities of CD8(+) T cells collaborate with macrophages to protect against blood-stage murine malaria. eLife 2015, 4, e04232. [Google Scholar] [CrossRef]
  311. Horne-Debets, J.M.; Faleiro, R.; Karunarathne, D.S.; Liu, X.Q.; Lineburg, K.E.; Poh, C.M.; Grotenbreg, G.M.; Hill, G.R.; MacDonald, K.P.; Good, M.F.; et al. PD-1 dependent exhaustion of CD8+ T cells drives chronic malaria. Cell Rep. 2013, 5, 1204–1213. [Google Scholar] [CrossRef]
  312. Horne-Debets, J.M.; Karunarathne, D.S.; Faleiro, R.J.; Poh, C.M.; Renia, L.; Wykes, M.N. Mice lacking Programmed cell death-1 show a role for CD8(+) T cells in long-term immunity against blood-stage malaria. Sci. Rep. 2016, 6, 26210. [Google Scholar] [CrossRef]
  313. Podoba, J.E.; Stevenson, M.M. CD4+ and CD8+ T lymphocytes both contribute to acquired immunity to blood-stage Plasmodium chabaudi AS. Infect. Immun. 1991, 59, 51–58. [Google Scholar] [CrossRef]
  314. Van der Heyde, H.C.; Manning, D.D.; Roopenian, D.C.; Weidanz, W.P. Resolution of blood-stage malarial infections in CD8+ cell-deficient beta 2-m0/0 mice. J. Immunol. 1993, 151, 3187–3191. [Google Scholar]
  315. Vinetz, J.M.; Kumar, S.; Good, M.F.; Fowlkes, B.J.; Berzofsky, J.A.; Miller, L.H. Adoptive transfer of CD8+ T cells from immune animals does not transfer immunity to blood stage Plasmodium yoelii malaria. J. Immunol. 1990, 144, 1069–1074. [Google Scholar] [PubMed]
  316. Kobayashi, F.; Morii, T.; Matsui, T.; Fujino, T.; Watanabe, Y.; Weidanz, W.P.; Tsuji, M. Production of interleukin 10 during malaria caused by lethal and nonlethal variants of Plasmodium yoelii yoelii. Parasitol. Res. 1996, 82, 385–391. [Google Scholar] [CrossRef] [PubMed]
  317. Do Rosário, A.P.F.; Lamb, T.; Spence, P.; Stephens, R.; Lang, A.; Roers, A.; Muller, W.; O’Garra, A.; Langhorne, J. IL-27 promotes IL-10 production by effector Th1 CD4+ T cells: A critical mechanism for protection from severe immunopathology during malaria infection. J. Immunol. 2012, 188, 1178–1190. [Google Scholar] [CrossRef] [PubMed]
  318. Jagannathan, P.; Eccles-James, I.; Bowen, K.; Nankya, F.; Auma, A.; Wamala, S.; Ebusu, C.; Muhindo, M.K.; Arinaitwe, E.; Briggs, J.; et al. IFNgamma/IL-10 co-producing cells dominate the CD4 response to malaria in highly exposed children. PLoS Pathog. 2014, 10, e1003864. [Google Scholar] [CrossRef] [PubMed]
  319. Sukhbaatar, O.; Kimura, D.; Miyakoda, M.; Nakamae, S.; Kimura, K.; Hara, H.; Yoshida, H.; Inoue, S.-I.; Yui, K. Activation and IL-10 production of specific CD4+ T cells are regulated by IL-27 during chronic infection with Plasmodium chabaudi. Parasitol. Int. 2020, 74, 101994. [Google Scholar] [CrossRef] [PubMed]
  320. Villegas-Mendez, A.; Inkson, C.A.; Shaw, T.N.; Strangward, P.; Couper, K.N. Long-lived CD4+ IFN-γ+ T cells rather than short-lived CD4+ IFN-γ+ IL-10+ T cells initiate rapid IL-10 production to suppress anamnestic T cell responses during secondary malaria infection. J. Immunol. 2016, 197, 3152–3164. [Google Scholar] [CrossRef]
  321. Chen, Q.; Schlichtherle, M.; Wahlgren, M. Molecular aspects of severe malaria. Clin. Microbiol. Rev. 2000, 13, 439–450. [Google Scholar] [CrossRef]
  322. Dondorp, A.M.; Kager, P.A.; Vreeken, J.; White, N.J. Abnormal blood flow and red blood cell deformability in severe malaria. Parasitol. Today 2000, 16, 228–232. [Google Scholar] [CrossRef]
  323. Shelby, J.P.; White, J.; Ganesan, K.; Rathod, P.K.; Chiu, D.T. A microfluidic model for single-cell capillary obstruction by Plasmodium falciparum-infected erythrocytes. Proc. Natl. Acad. Sci. 2003, 100, 14618–14622. [Google Scholar] [CrossRef]
  324. Barber, B.E.; Russell, B.; Grigg, M.J.; Zhang, R.; William, T.; Amir, A.; Lau, Y.L.; Chatfield, M.D.; Dondorp, A.M.; Anstey, N.M. Reduced red blood cell deformability in Plasmodium knowlesi malaria. Blood Adv. 2018, 2, 433–443. [Google Scholar] [CrossRef]
  325. Mwakingwe, A.; Ting, L.M.; Hochman, S.; Chen, J.; Sinnis, P.; Kim, K. Noninvasive real-time monitoring of liver-stage development of bioluminescent Plasmodium parasites. J. Infect. Dis. 2009, 200, 1470–1478. [Google Scholar] [CrossRef]
  326. Franke-Fayard, B.; Waters, A.P.; Janse, C.J. Real-time in vivo imaging of transgenic bioluminescent blood stages of rodent malaria parasites in mice. Nat. Protoc. 2006, 1, 476–485. [Google Scholar] [CrossRef]
  327. Nallandhighal, S.; Park, G.S.; Ho, Y.-Y.; Opoka, R.O.; John, C.C.; Tran, T.M. Whole-Blood Transcriptional Signatures Composed of Erythropoietic and NRF2-Regulated Genes Differ Between Cerebral Malaria and Severe Malarial Anemia. J. Infect. Dis. 2019, 219, 154–164. [Google Scholar] [CrossRef]
  328. White, N.J. Anaemia and malaria. Malar. J. 2018, 17, 371. [Google Scholar] [CrossRef]
  329. Jakeman, G.; Saul, A.; Hogarth, W.; Collins, W. Anaemia of acute malaria infections in non-immune patients primarily results from destruction of uninfected erythrocytes. Parasitology 1999, 119, 127–133. [Google Scholar] [CrossRef]
  330. Abdalla, S.; Weatherall, D.; Wickramasinghe, S.; Hughes, M. The anaemia of P. falciparum malaria. Br. J. Haematol. 1980, 46, 171–183. [Google Scholar] [CrossRef]
  331. Aguilar, R.; Moraleda, C.; Achtman, A.H.; Mayor, A.; Quinto, L.; Cistero, P.; Nhabomba, A.; Macete, E.; Schofield, L.; Alonso, P.L.; et al. Severity of anaemia is associated with bone marrow haemozoin in children exposed to Plasmodium falciparum. Br. J. Haematol. 2014, 164, 877–887. [Google Scholar] [CrossRef]
  332. Chang, K.H.; Tam, M.; Stevenson, M.M. Inappropriately low reticulocytosis in severe malarial anemia correlates with suppression in the development of late erythroid precursors. Blood 2004, 103, 3727–3735. [Google Scholar] [CrossRef]
  333. Mohan, K.; Stevenson, M.M. Dyserythropoiesis and severe anaemia associated with malaria correlate with deficient interleukin-12 production. Br. J. Haematol. 1998, 103, 942–949. [Google Scholar] [CrossRef]
  334. McDevitt, M.A.; Xie, J.; Ganapathy-Kanniappan, S.; Griffith, J.; Liu, A.; McDonald, C.; Thuma, P.; Gordeuk, V.R.; Metz, C.N.; Mitchell, R.; et al. A critical role for the host mediator macrophage migration inhibitory factor in the pathogenesis of malarial anemia. J. Exp. Med. 2006, 203, 1185–1196. [Google Scholar] [CrossRef]
  335. Thawani, N.; Tam, M.; Stevenson, M.M. STAT6-mediated suppression of erythropoiesis in an experimental model of malarial anemia. Haematologica 2009, 94, 195–204. [Google Scholar] [CrossRef]
  336. O’Dea, K.P.; Pasvol, G. Optimal tumor necrosis factor induction by Plasmodium falciparum requires the highly localized release of parasite products. Infect. Immun. 2003, 71, 3155–3164. [Google Scholar] [CrossRef] [Green Version]
  337. Rudin, W.; Quesniaux, V.; Favre, N.; Bordmann, G. Malaria toxins from P. chabaudi chabaudi AS and P. berghei ANKA cause dyserythropoiesis in C57BL/6 mice. Parasitology 1997, 115 Pt 5, 467–474. [Google Scholar] [CrossRef]
  338. Bordmann, G.; Favre, N.; Rudin, W. Malaria toxins: Effects on murine spleen and bone marrow cell proliferation and cytokine production in vitro. Parasitology 1997, 115 Pt 5, 475–483. [Google Scholar] [CrossRef]
  339. Thawani, N.; Tam, M.; Bellemare, M.J.; Bohle, D.S.; Olivier, M.; de Souza, J.B.; Stevenson, M.M. Plasmodium products contribute to severe malarial anemia by inhibiting erythropoietin-induced proliferation of erythroid precursors. J. Infect. Dis. 2014, 209, 140–149. [Google Scholar] [CrossRef]
  340. Chang, K.H.; Stevenson, M.M. Effect of anemia and renal cytokine production on erythropoietin production during blood-stage malaria. Kidney Int. 2004, 65, 1640–1646. [Google Scholar] [CrossRef]
  341. Spottiswoode, N.; Armitage, A.E.; Williams, A.R.; Fyfe, A.J.; Biswas, S.; Hodgson, S.H.; Llewellyn, D.; Choudhary, P.; Draper, S.J.; Duffy, P.E.; et al. Role of Activins in Hepcidin Regulation during Malaria. Infect. Immun. 2017, 85, e00191-17. [Google Scholar] [CrossRef]
  342. Maretty, L.; Sharp, R.E.; Andersson, M.; Kurtzhals, J.A. Intravenous ferric carboxymaltose accelerates erythropoietic recovery from experimental malarial anemia. J. Infect. Dis. 2012, 205, 1173–1177. [Google Scholar] [CrossRef]
  343. Harris, J.V.; Bohr, T.M.; Stracener, C.; Landmesser, M.E.; Torres, V.; Mbugua, A.; Moratz, C.; Stoute, J.A. Sequential Plasmodium chabaudi and Plasmodium berghei infections provide a novel model of severe malarial anemia. Infect. Immun. 2012, 80, 2997–3007. [Google Scholar] [CrossRef]
  344. Mourao, L.C.; Cardoso-Oliveira, G.P.; Braga, E.M. Autoantibodies and Malaria: Where We Stand? Insights into Pathogenesis and Protection. Front. Cell. Infect. Microbiol. 2020, 10, 262. [Google Scholar] [CrossRef]
  345. Van den Akker, E.; Satchwell, T.J.; Williamson, R.C.; Toye, A.M. Band 3 multiprotein complexes in the red cell membrane; of mice and men. Blood Cells Mol. Dis. 2010, 45, 1–8. [Google Scholar] [CrossRef] [PubMed]
  346. Wesseling, M.C.; Wagner-Britz, L.; Huppert, H.; Hanf, B.; Hertz, L.; Nguyen, D.B.; Bernhardt, I. Phosphatidylserine Exposure in Human Red Blood Cells Depending on Cell Age. Cell. Physiol. Biochem. 2016, 38, 1376–1390. [Google Scholar] [CrossRef] [PubMed]
  347. Zhang, R.; Zhang, C.; Zhao, Q.; Li, D. Spectrin: Structure, function and disease. Sci. China Life Sci. 2013, 56, 1076–1085. [Google Scholar] [CrossRef] [PubMed]
  348. Mourao, L.C.; Baptista, R.P.; de Almeida, Z.B.; Grynberg, P.; Pucci, M.M.; Castro-Gomes, T.; Fontes, C.J.F.; Rathore, S.; Sharma, Y.D.; da Silva-Pereira, R.A.; et al. Anti-band 3 and anti-spectrin antibodies are increased in Plasmodium vivax infection and are associated with anemia. Sci. Rep. 2018, 8, 8762. [Google Scholar] [CrossRef]
  349. Hogh, B.; Petersen, E.; Crandall, I.; Gottschau, A.; Sherman, I.W. Immune responses to band 3 neoantigens on Plasmodium falciparum-infected erythrocytes in subjects living in an area of intense malaria transmission are associated with low parasite density and high hematocrit value. Infect. Immun. 1994, 62, 4362–4366. [Google Scholar] [CrossRef]
  350. Fernandez-Arias, C.; Rivera-Correa, J.; Gallego-Delgado, J.; Rudlaff, R.; Fernandez, C.; Roussel, C.; Gotz, A.; Gonzalez, S.; Mohanty, A.; Mohanty, S.; et al. Anti-Self Phosphatidylserine Antibodies Recognize Uninfected Erythrocytes Promoting Malarial Anemia. Cell Host Microbe 2016, 19, 194–203. [Google Scholar] [CrossRef]
  351. Ternynck, T.; Falanga, P.B.; Unterkirscher, C.; Gregoire, J.; da Silva, L.P.; Avrameas, S. Induction of high levels of IgG autoantibodies in mice infected with Plasmodium chabaudi. Int. Immunol. 1991, 3, 29–37. [Google Scholar] [CrossRef]
  352. Romero, H.; Warburton, C.; Sanchez-Dardon, J.; Scorza, T. Osteoclasts Are Required for Hematopoietic Stem and Progenitor Cell Mobilization but Not for Stress Erythropoiesis in Plasmodium chabaudi adami Murine Malaria. Mediat. Inflamm. 2016, 2016, 3909614. [Google Scholar] [CrossRef]
  353. Yap, G.S.; Stevenson, M.M. Plasmodium chabaudi AS: Erythropoietic responses during infection in resistant and susceptible mice. Exp. Parasitol. 1992, 75, 340–352. [Google Scholar] [CrossRef]
  354. Asami, M.; Owhashi, M.; Abe, T.; Nawa, Y. A comparative study of the kinetic changes of hemopoietic stem cells in mice infected with lethal and non-lethal malaria. Int. J. Parasitol. 1992, 22, 43–47. [Google Scholar] [CrossRef]
  355. Bruneel, F. Human cerebral malaria: 2019 mini review. Rev. Neurol. 2019, 175, 445–450. [Google Scholar] [CrossRef]
  356. Sahu, P.K.; Satpathi, S.; Behera, P.K.; Mishra, S.K.; Mohanty, S.; Wassmer, S.C. Pathogenesis of cerebral malaria: New diagnostic tools, biomarkers, and therapeutic approaches. Front. Cell. Infect. Microbiol. 2015, 5, 75. [Google Scholar] [CrossRef] [Green Version]
  357. Ghazanfari, N.; Mueller, S.N.; Heath, W.R. Cerebral Malaria in Mouse and Man. Front. Immunol. 2018, 9, 2016. [Google Scholar] [CrossRef]
  358. Belnoue, E.; Potter, S.M.; Rosa, D.S.; Mauduit, M.; Gruner, A.C.; Kayibanda, M.; Mitchell, A.J.; Hunt, N.H.; Renia, L. Control of pathogenic CD8+ T cell migration to the brain by IFN-gamma during experimental cerebral malaria. Parasite Immunol. 2008, 30, 544–553. [Google Scholar] [CrossRef]
  359. Strangward, P.; Haley, M.J.; Shaw, T.N.; Schwartz, J.-M.; Greig, R.; Mironov, A.; de Souza, J.-B.; Cruickshank, S.M.; Craig, A.G.; Milner, D.A., Jr. A quantitative brain map of experimental cerebral malaria pathology. PLoS Pathog. 2017, 13, e1006267. [Google Scholar] [CrossRef]
  360. Howland, S.W.; Poh, C.M.; Gun, S.Y.; Claser, C.; Malleret, B.; Shastri, N.; Ginhoux, F.; Grotenbreg, G.M.; Renia, L. Brain microvessel cross-presentation is a hallmark of experimental cerebral malaria. EMBO Mol. Med. 2013, 5, 984–999. [Google Scholar] [CrossRef]
  361. Howland, S.W.; Poh, C.M.; Renia, L. Activated Brain Endothelial Cells Cross-Present Malaria Antigen. PLoS Pathog. 2015, 11, e1004963. [Google Scholar] [CrossRef]
  362. Newbold, C.; Craig, A.; Kyes, S.; Rowe, A.; Fernandez-Reyes, D.; Fagan, T. Cytoadherence, pathogenesis and the infected red cell surface in Plasmodium falciparum. Int. J. Parasitol. 1999, 29, 927–937. [Google Scholar] [CrossRef]
  363. Turner, G.D.; Morrison, H.; Jones, M.; Davis, T.M.; Looareesuwan, S.; Buley, I.D.; Gatter, K.C.; Newbold, C.I.; Pukritayakamee, S.; Nagachinta, B.; et al. An immunohistochemical study of the pathology of fatal malaria. Evidence for widespread endothelial activation and a potential role for intercellular adhesion molecule-1 in cerebral sequestration. Am. J. Pathol. 1994, 145, 1057–1069. [Google Scholar]
  364. Fonager, J.; Pasini, E.M.; Braks, J.A.; Klop, O.; Ramesar, J.; Remarque, E.J.; Vroegrijk, I.O.; van Duinen, S.G.; Thomas, A.W.; Khan, S.M.; et al. Reduced CD36-dependent tissue sequestration of Plasmodium-infected erythrocytes is detrimental to malaria parasite growth in vivo. J. Exp. Med. 2012, 209, 93–107. [Google Scholar] [CrossRef]
  365. Tuikue Ndam, N.; Moussiliou, A.; Lavstsen, T.; Kamaliddin, C.; Jensen, A.T.; Mama, A.; Tahar, R.; Wang, C.W.; Jespersen, J.S.; Alao, J.M. Parasites causing cerebral falciparum malaria bind multiple endothelial receptors and express EPCR and ICAM-1-binding PfEMP1. J. Infect. Dis. 2017, 215, 1918–1925. [Google Scholar] [CrossRef]
  366. Nacer, A.; Movila, A.; Baer, K.; Mikolajczak, S.A.; Kappe, S.H.; Frevert, U. Neuroimmunological blood brain barrier opening in experimental cerebral malaria. PLoS Pathog. 2012, 8, e1002982. [Google Scholar] [CrossRef]
  367. Frevert, U.; Nacer, A.; Cabrera, M.; Movila, A.; Leberl, M. Imaging Plasmodium immunobiology in the liver, brain, and lung. Parasitol. Int. 2014, 63, 171–186. [Google Scholar] [CrossRef] [Green Version]
  368. Rénia, L.; Howland, S.W. Targeting the olfactory bulb during experimental cerebral malaria. Trends Parasitol. 2014, 30, 375–376. [Google Scholar] [CrossRef]
  369. Franke-Fayard, B.; Djokovic, D.; Dooren, M.; Ramesar, J.; Waters, A.; Falade, M.; Kranendonk, M.; Martinelli, A.; Cravo, P.; Janse, C. Simple and sensitive antimalarial drug screening in vitro and in vivo using transgenic luciferase expressing Plasmodium berghei parasites. Int. J. Parasitol. 2008, 38, 1651–1662. [Google Scholar] [CrossRef]
  370. Ploemen, I.H.; Prudêncio, M.; Douradinha, B.G.; Ramesar, J.; Fonager, J.; van Gemert, G.-J.; Luty, A.J.; Hermsen, C.C.; Sauerwein, R.W.; Baptista, F.G. Visualisation and quantitative analysis of the rodent malaria liver stage by real time imaging. PLoS ONE 2009, 4, e7881. [Google Scholar] [CrossRef]
  371. Franke-Fayard, B.; Fonager, J.; Braks, A.; Khan, S.M.; Janse, C.J. Sequestration and tissue accumulation of human malaria parasites: Can we learn anything from rodent models of malaria? PLoS Pathog. 2010, 6, e1001032. [Google Scholar] [CrossRef]
  372. Claser, C.; Malleret, B.; Gun, S.Y.; Wong, A.Y.W.; Chang, Z.W.; Teo, P.; See, P.C.E.; Howland, S.W.; Ginhoux, F.; Rénia, L. CD8+ T cells and IFN-γ mediate the time-dependent accumulation of infected red blood cells in deep organs during experimental cerebral malaria. PLoS ONE 2011, 6, e18720. [Google Scholar] [CrossRef]
  373. Kwiatkowski, D.; Molyneux, M.E.; Stephens, S.; Curtis, N.; Klein, N.; Pointaire, P.; Smit, M.; Allan, R.; Brewster, D.R.; Grau, G.E.; et al. Anti-TNF therapy inhibits fever in cerebral malaria. Q. J. Med. 1993, 86, 91–98. [Google Scholar]
  374. Togbe, D.; de Sousa, P.L.; Fauconnier, M.; Boissay, V.; Fick, L.; Scheu, S.; Pfeffer, K.; Menard, R.; Grau, G.E.; Doan, B.T.; et al. Both functional LTbeta receptor and TNF receptor 2 are required for the development of experimental cerebral malaria. PLoS ONE 2008, 3, e2608. [Google Scholar] [CrossRef]
  375. Engwerda, C.R.; Mynott, T.L.; Sawhney, S.; De Souza, J.B.; Bickle, Q.D.; Kaye, P.M. Locally up-regulated lymphotoxin alpha, not systemic tumor necrosis factor alpha, is the principle mediator of murine cerebral malaria. J. Exp. Med. 2002, 195, 1371–1377. [Google Scholar] [CrossRef] [PubMed]
  376. Feintuch, C.M.; Tare, A.; Cusumano, L.R.; Benayoun, J.; Ryu, S.; Sixpence, A.; Seydel, K.; Laufer, M.; Taylor, T.; Suh, Y. Type I interferon receptor variants in gene regulatory regions are associated with susceptibility to cerebral malaria in Malawi. Am. J. Trop. Med. Hyg. 2018, 98, 1692–1698. [Google Scholar] [CrossRef] [PubMed]
  377. Aucan, C.; Walley, A.; Hennig, B.; Fitness, J.; Frodsham, A.; Zhang, L.; Kwiatkowski, D.; Hill, A. Interferon-alpha receptor-1 (IFNAR1) variants are associated with protection against cerebral malaria in the Gambia. Genes Immun. 2003, 4, 275–282. [Google Scholar] [CrossRef] [PubMed]
  378. Ball, E.A.; Sambo, M.R.; Martins, M.; Trovoada, M.J.; Benchimol, C.; Costa, J.; Gonçalves, L.A.; Coutinho, A.; Penha-Gonçalves, C. IFNAR1 controls progression to cerebral malaria in children and CD8+ T cell brain pathology in Plasmodium berghei-infected mice. J. Immunol. 2013, 190, 5118–5127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  379. Krupka, M.; Seydel, K.; Feintuch, C.M.; Yee, K.; Kim, R.; Lin, C.-Y.; Calder, R.B.; Petersen, C.; Taylor, T.; Daily, J. Mild Plasmodium falciparum malaria following an episode of severe malaria is associated with induction of the interferon pathway in Malawian children. Infect. Immun. 2012, 80, 1150–1155. [Google Scholar] [CrossRef]
  380. Sharma, S.; DeOliveira, R.B.; Kalantari, P.; Parroche, P.; Goutagny, N.; Jiang, Z.; Chan, J.; Bartholomeu, D.C.; Lauw, F.; Hall, J.P.; et al. Innate immune recognition of an AT-rich stem-loop DNA motif in the Plasmodium falciparum genome. Immunity 2011, 35, 194–207. [Google Scholar] [CrossRef]
  381. Koch, O.; Awomoyi, A.; Usen, S.; Jallow, M.; Richardson, A.; Hull, J.; Pinder, M.; Newport, M.; Kwiatkowski, D. IFNGR1 gene promoter polymorphisms and susceptibility to cerebral malaria. J. Infect. Dis. 2002, 185, 1684–1687. [Google Scholar] [CrossRef]
  382. Cabantous, S.; Poudiougou, B.; Traore, A.; Keita, M.; Cisse, M.B.; Doumbo, O.; Dessein, A.J.; Marquet, S. Evidence that interferon-gamma plays a protective role during cerebral malaria. J. Infect. Dis. 2005, 192, 854–860. [Google Scholar] [CrossRef]
  383. Villegas-Mendez, A.; Strangward, P.; Shaw, T.N.; Rajkovic, I.; Tosevski, V.; Forman, R.; Muller, W.; Couper, K.N. Gamma Interferon Mediates Experimental Cerebral Malaria by Signaling within Both the Hematopoietic and Nonhematopoietic Compartments. Infect. Immun. 2017, 85, e01035-16. [Google Scholar] [CrossRef]
  384. Villegas-Mendez, A.; Greig, R.; Shaw, T.N.; de Souza, J.B.; Gwyer Findlay, E.; Stumhofer, J.S.; Hafalla, J.C.; Blount, D.G.; Hunter, C.A.; Riley, E.M.; et al. IFN-gamma-producing CD4+ T cells promote experimental cerebral malaria by modulating CD8+ T cell accumulation within the brain. J. Immunol. 2012, 189, 968–979. [Google Scholar] [CrossRef]
  385. Van den Steen, P.E.; Deroost, K.; Aelst, I.V.; Geurts, N.; Martens, E.; Struyf, S.; Nie, C.Q.; Hansen, D.S.; Matthys, P.; Damme, J.V. CXCR3 determines strain susceptibility to murine cerebral malaria by mediating T lymphocyte migration toward IFN-γ-induced chemokines. Eur. J. Immunol. 2008, 38, 1082–1095. [Google Scholar] [CrossRef]
  386. Hansen, D.S.; Bernard, N.J.; Nie, C.Q.; Schofield, L. NK cells stimulate recruitment of CXCR3+ T cells to the brain during Plasmodium berghei-mediated cerebral malaria. J. Immunol. 2007, 178, 5779–5788. [Google Scholar] [CrossRef]
  387. Belnoue, E.; Kayibanda, M.; Deschemin, J.-C.; Viguier, M.; Mack, M.; Kuziel, W.A.; Rénia, L. CCR5 deficiency decreases susceptibility to experimental cerebral malaria. Blood 2003, 101, 4253–4259. [Google Scholar] [CrossRef]
  388. Dorovini-Zis, K.; Schmidt, K.; Huynh, H.; Fu, W.; Whitten, R.O.; Milner, D.; Kamiza, S.; Molyneux, M.; Taylor, T.E. The neuropathology of fatal cerebral malaria in malawian children. Am. J. Pathol. 2011, 178, 2146–2158. [Google Scholar] [CrossRef] [Green Version]
  389. Barrera, V.; Haley, M.J.; Strangward, P.; Attree, E.; Kamiza, S.; Seydel, K.; Taylor, T.; Milner, D.; Craig, A.; Couper, K.N. Comparison of CD8+ T cell accumulation in the brain during human and murine cerebral malaria. Front. Immunol. 2019, 10, 1747. [Google Scholar] [CrossRef]
  390. Riggle, B.A.; Manglani, M.; Maric, D.; Johnson, K.R.; Lee, M.-H.; Neto, O.L.A.; Taylor, T.E.; Seydel, K.B.; Nath, A.; Miller, L.H. CD8+ T cells target cerebrovasculature in children with cerebral malaria. J. Clin. Investing. 2020, 130, 1128–1138. [Google Scholar] [CrossRef]
  391. Haque, A.; Best, S.E.; Unosson, K.; Amante, F.H.; de Labastida, F.; Anstey, N.M.; Karupiah, G.; Smyth, M.J.; Heath, W.R.; Engwerda, C.R. Granzyme B expression by CD8+ T cells is required for the development of experimental cerebral malaria. J. Immunol. 2011, 186, 6148–6156. [Google Scholar] [CrossRef]
  392. Huggins, M.A.; Johnson, H.L.; Jin, F.; N’ Songo, A.; Hanson, L.M.; LaFrance, S.J.; Butler, N.S.; Harty, J.T.; Johnson, A.J. Perforin Expression by CD8 T Cells Is Sufficient To Cause Fatal Brain Edema during Experimental Cerebral Malaria. Infect. Immun. 2017, 85, e00985-16. [Google Scholar] [CrossRef]
  393. Punsawad, C.; Maneerat, Y.; Chaisri, U.; Nantavisai, K.; Viriyavejakul, P. Nuclear factor kappa B modulates apoptosis in the brain endothelial cells and intravascular leukocytes of fatal cerebral malaria. Malar. J. 2013, 12, 260. [Google Scholar] [CrossRef]
  394. Pino, P.; Vouldoukis, I.; Kolb, J.P.; Mahmoudi, N.; Desportes-Livage, I.; Bricaire, F.; Danis, M.; Dugas, B.; Mazier, D. Plasmodium falciparum-infected erythrocyte adhesion induces caspase activation and apoptosis in human endothelial cells. J. Infect. Dis. 2003, 187, 1283–1290. [Google Scholar] [CrossRef]
  395. Tsunoda, I.; Kuang, L.-Q.; Fujinami, R.S. Induction of autoreactive CD8+ cytotoxic T cells during Theiler’s murine encephalomyelitis virus infection: Implications for autoimmunity. J. Virol. 2002, 76, 12834–12844. [Google Scholar] [CrossRef]
  396. Cong, X.; Kong, W. Endothelial tight junctions and their regulatory signaling pathways in vascular homeostasis and disease. Cell. Signal. 2020, 66, 109485. [Google Scholar] [CrossRef]
  397. Brown, H.; Hien, T.; Day, N.; Mai, N.; Chuong, L.; Chau, T.; Loc, P.; Phu, N.; Bethell, D.; Farrar, J. Evidence of blood-brain barrier dysfunction in human cerebral malaria. Neuropathol. Appl. Neurobiol. 1999, 25, 331–340. [Google Scholar] [CrossRef]
  398. Suidan, G.L.; McDole, J.R.; Chen, Y.; Pirko, I.; Johnson, A.J. Induction of blood brain barrier tight junction protein alterations by CD8 T cells. PLoS ONE 2008, 3, e3037. [Google Scholar] [CrossRef] [Green Version]
  399. Tunon-Ortiz, A.; Lamb, T.J. Blood brain barrier disruption in cerebral malaria: Beyond endothelial cell activation. PLoS Pathog. 2019, 15, e1007786. [Google Scholar] [CrossRef]
  400. Shrivastava, S.K.; Dalko, E.; Delcroix-Genete, D.; Herbert, F.; Cazenave, P.A.; Pied, S. Uptake of parasite-derived vesicles by astrocytes and microglial phagocytosis of infected erythrocytes may drive neuroinflammation in cerebral malaria. Glia 2017, 65, 75–92. [Google Scholar] [CrossRef]
  401. Promeneur, D.; Lunde, L.K.; Amiry-Moghaddam, M.; Agre, P. Protective role of brain water channel AQP4 in murine cerebral malaria. Proc. Natl. Acad. Sci. USA 2013, 110, 1035–1040. [Google Scholar] [CrossRef]
  402. Capuccini, B.; Lin, J.; Talavera-Lopez, C.; Khan, S.M.; Sodenkamp, J.; Spaccapelo, R.; Langhorne, J. Transcriptomic profiling of microglia reveals signatures of cell activation and immune response, during experimental cerebral malaria. Sci. Rep. 2016, 6, 39258. [Google Scholar] [CrossRef]
  403. Medana, I.M.; Idro, R.; Newton, C.R. Axonal and astrocyte injury markers in the cerebrospinal fluid of Kenyan children with severe malaria. J. Neurol. Sci. 2007, 258, 93–98. [Google Scholar] [CrossRef]
  404. Schluesener, H.J.; Kremsner, P.G.; Meyermann, R. Widespread expression of MRP8 and MRP14 in human cerebral malaria by microglial cells. Acta Neuropathol. 1998, 96, 575–580. [Google Scholar] [CrossRef]
  405. Poh, C.M.; Howland, S.W.; Grotenbreg, G.M.; Renia, L. Damage to the blood-brain barrier during experimental cerebral malaria results from synergistic effects of CD8+ T cells with different specificities. Infect. Immun. 2014, 82, 4854–4864. [Google Scholar] [CrossRef] [PubMed]
  406. Besnard, A.-G.; Guabiraba, R.; Niedbala, W.; Palomo, J.; Reverchon, F.; Shaw, T.N.; Couper, K.N.; Ryffel, B.; Liew, F.Y. IL-33-mediated protection against experimental cerebral malaria is linked to induction of type 2 innate lymphoid cells, M2 macrophages and regulatory T cells. PLoS Pathog. 2015, 11, e1004607. [Google Scholar] [CrossRef] [PubMed]
  407. Burrack, K.S.; Huggins, M.A.; Taras, E.; Dougherty, P.; Henzler, C.M.; Yang, R.; Alter, S.; Jeng, E.K.; Wong, H.C.; Felices, M.; et al. Interleukin-15 Complex Treatment Protects Mice from Cerebral Malaria by Inducing Interleukin-10-Producing Natural Killer Cells. Immunity 2018, 48, 760–772.e4. [Google Scholar] [CrossRef] [PubMed]
  408. Charoenpan, P.; Indraprasit, S.; Kiatboonsri, S.; Suvachittanont, O.; Tanomsup, S. Pulmonary edema in severe falciparum malaria: Hemodynamic study and clinicophysiologic correlation. Chest 1990, 97, 1190–1197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  409. Feldman, R.M.; Singer, C. Noncardiogenic pulmonary edema and pulmonary fibrosis in falciparum malaria. Clin. Infect. Dis. 1987, 9, 134–139. [Google Scholar] [CrossRef] [PubMed]
  410. Ando, M.; Miyazaki, E.; Abe, T.; Ehara, C.; Goto, A.; Masuda, T.; Nishio, S.; Fujisaki, H.; Yamasue, M.; Ishii, T. Angiopoietin-2 expression in patients with an acute exacerbation of idiopathic interstitial pneumonias. Respir. Med. 2016, 117, 27–32. [Google Scholar] [CrossRef] [PubMed]
  411. Pham, T.-T.; Punsawad, C.; Glaharn, S.; De Meyer, S.F.; Viriyavejakul, P.; Van den Steen, P.E. Release of endothelial activation markers in lungs of patients with malaria-associated acute respiratory distress syndrome. Malar. J. 2019, 18, 395. [Google Scholar] [CrossRef] [PubMed]
  412. Avril, M.; Bernabeu, M.; Benjamin, M.; Brazier, A.J.; Smith, J.D. Interaction between endothelial protein C receptor and intercellular adhesion molecule 1 to mediate binding of Plasmodium falciparum-infected erythrocytes to endothelial cells. MBio 2016, 7, e00615-16. [Google Scholar] [CrossRef]
  413. Lovegrove, F.E.; Gharib, S.A.; Peña-Castillo, L.; Patel, S.N.; Ruzinski, J.T.; Hughes, T.R.; Liles, W.C.; Kain, K.C. Parasite burden and CD36-mediated sequestration are determinants of acute lung injury in an experimental malaria model. PLoS Pathog. 2008, 4, e1000068. [Google Scholar] [CrossRef]
  414. Togbe, D.; Schofield, L.; Grau, G.E.; Schnyder, B.; Boissay, V.; Charron, S.; Rose, S.; Beutler, B.; Quesniaux, V.F.; Ryffel, B. Murine cerebral malaria development is independent of toll-like receptor signaling. Am. J. Pathol. 2007, 170, 1640–1648. [Google Scholar] [CrossRef]
  415. Kraisin, S.; Verhenne, S.; Pham, T.T.; Martinod, K.; Tersteeg, C.; Vandeputte, N.; Deckmyn, H.; Vanhoorelbeke, K.; Van den Steen, P.E.; De Meyer, S.F. Von Willebrand factor in experimental malaria-associated acute respiratory distress syndrome. J. Thromb. Haemost. 2019, 17, 1372–1383. [Google Scholar] [CrossRef]
  416. Zielinska, K.A.; de Cauwer, L.; Knoops, S.; Van der Molen, K.; Sneyers, A.; Thommis, J.; De Souza, J.B.; Opdenakker, G.; De Bosscher, K.; Van den Steen, P.E. Plasmodium berghei NK65 in Combination with IFN-gamma Induces Endothelial Glucocorticoid Resistance via Sustained Activation of p38 and JNK. Front. Immunol. 2017, 8, 1199. [Google Scholar] [CrossRef]
  417. Claser, C.; Nguee, S.Y.T.; Balachander, A.; Wu Howland, S.; Becht, E.; Gunasegaran, B.; Hartimath, S.V.; Lee, A.W.; Theng Theng Ho, J.; Bing Ong, C.; et al. Lung endothelial cell antigen cross-presentation to CD8(+)T cells drives malaria-associated lung injury. Nat. Commun. 2019, 10, 4241. [Google Scholar] [CrossRef]
  418. Sercundes, M.K.; Ortolan, L.S.; Debone, D.; Soeiro-Pereira, P.V.; Gomes, E.; Aitken, E.H.; Condino-Neto, A.; Russo, M.; Neto, A.C.; Alvarez, J.M.; et al. Targeting Neutrophils to Prevent Malaria-Associated Acute Lung Injury/Acute Respiratory Distress Syndrome in Mice. PLoS Pathog. 2016, 12, e1006054. [Google Scholar] [CrossRef] [Green Version]
  419. Lagasse, H.A.; Anidi, I.U.; Craig, J.M.; Limjunyawong, N.; Poupore, A.K.; Mitzner, W.; Scott, A.L. Recruited monocytes modulate malaria-induced lung injury through CD36-mediated clearance of sequestered infected erythrocytes. J. Leukoc. Biol. 2016, 99, 659–671. [Google Scholar] [CrossRef]
Figure 1. Contribution of mechanisms influencing immune control to Plasmodium-infected red blood cells. Mouse models of malaria have been used to demonstrate how innate and adaptive immune mechanisms synergize to effectuate control of iRBCs. In the white pulp of the spleen activation of T helper cells including Th1 cells, Tfh cells and regulatory cells (both Treg cells and Tr1 cells) are activated by antigen presenting cells which have phagocytosed iRBCs by-products of parasite development that are released during schizogony. Tfh cells are critical in providing help to B cells to produce anti-parasite antibodies which play important roles via interaction with innate cells antibody-dependent cellular cytotoxicity and in invasion blocking upon schizogony. The generation of anti-parasite antibodies both within and out-with germinal centers has been interrogated in mouse models of malaria gleaning important contributions towards what constitutes an efficacious cellular response generating efficacious antibodies reactive against iRBCs. Figure created with Biorender. Abbreviations: ICOS: inducible T cell costimulatory; IFN: interferon; Ig: immunoglobulin; IL: interleukin; iRBC: infected red blood cell; LLPC: long-lived plasma cell; M-CSF: macrophage colony-stimulating factor; NO: nitric oxide; ROS: reactive oxygen species; TGF: transforming growth factor; TNF: tumor necrosis factor.
Figure 1. Contribution of mechanisms influencing immune control to Plasmodium-infected red blood cells. Mouse models of malaria have been used to demonstrate how innate and adaptive immune mechanisms synergize to effectuate control of iRBCs. In the white pulp of the spleen activation of T helper cells including Th1 cells, Tfh cells and regulatory cells (both Treg cells and Tr1 cells) are activated by antigen presenting cells which have phagocytosed iRBCs by-products of parasite development that are released during schizogony. Tfh cells are critical in providing help to B cells to produce anti-parasite antibodies which play important roles via interaction with innate cells antibody-dependent cellular cytotoxicity and in invasion blocking upon schizogony. The generation of anti-parasite antibodies both within and out-with germinal centers has been interrogated in mouse models of malaria gleaning important contributions towards what constitutes an efficacious cellular response generating efficacious antibodies reactive against iRBCs. Figure created with Biorender. Abbreviations: ICOS: inducible T cell costimulatory; IFN: interferon; Ig: immunoglobulin; IL: interleukin; iRBC: infected red blood cell; LLPC: long-lived plasma cell; M-CSF: macrophage colony-stimulating factor; NO: nitric oxide; ROS: reactive oxygen species; TGF: transforming growth factor; TNF: tumor necrosis factor.
Vaccines 10 01525 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Olatunde, A.C.; Cornwall, D.H.; Roedel, M.; Lamb, T.J. Mouse Models for Unravelling Immunology of Blood Stage Malaria. Vaccines 2022, 10, 1525. https://doi.org/10.3390/vaccines10091525

AMA Style

Olatunde AC, Cornwall DH, Roedel M, Lamb TJ. Mouse Models for Unravelling Immunology of Blood Stage Malaria. Vaccines. 2022; 10(9):1525. https://doi.org/10.3390/vaccines10091525

Chicago/Turabian Style

Olatunde, Adesola C., Douglas H. Cornwall, Marshall Roedel, and Tracey J. Lamb. 2022. "Mouse Models for Unravelling Immunology of Blood Stage Malaria" Vaccines 10, no. 9: 1525. https://doi.org/10.3390/vaccines10091525

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop