presentation of cerebral malaria

  • Subscribe to journal Subscribe
  • Get new issue alerts Get alerts

Secondary Logo

Journal logo.

Colleague's E-mail is Invalid

Your message has been successfully sent to your colleague.

Save my selection

Pathophysiology, clinical presentation, and treatment of coma and acute kidney injury complicating falciparum malaria

Plewes, Katherine a,b ; Turner, Gareth D.H. c,d ; Dondorp, Arjen M. a,d

a Faculty of Tropical Medicine, Mahidol Oxford Tropical Medicine Research Unit, Mahidol University, Bangkok, Thailand

b Division of Infectious Diseases, Department of Medicine, University of British Columbia, Vancouver, British Columbia, Canada

c Department of Cellular Pathology, John Radcliffe Hospital

d Nuffield Department of Clinical Medicine, Center for Tropical Medicine and Global Health, University of Oxford, Oxford, UK

Correspondence to Katherine Plewes, Clinical Investigator, Mahidol-Oxford Tropical Medicine Research Unit (MORU), Clinical Assistant Professor, UBC Division of Infectious Diseases, Heather Pavilion East 452D, Vancouver General Hospital, 2733 Heather Street, Vancouver, British Columbia, V5Z 3J5, Canada. Tel: +1 604 875 4588; fax: +1 604 875 4013; e-mail: [email protected]

This is an open access article distributed under the Creative Commons Attribution License 4.0 (CCBY), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. http://creativecommons.org/licenses/by/4.0

Purpose of review 

Cerebral impairment and acute kidney injury (AKI) are independent predictors of mortality in both adults and children with severe falciparum malaria. In this review, we present recent advances in understanding the pathophysiology, clinical features, and management of these complications of severe malaria, and discuss future areas of research.

Recent findings 

Cerebral malaria and AKI are serious and well recognized complications of severe malaria. Common pathophysiological pathways include impaired microcirculation, due to sequestration of parasitized erythrocytes, systemic inflammatory responses, and endothelial activation. Recent MRI studies show significant brain swelling in both adults and children with evidence of posterior reversible encephalopathy syndrome-like syndrome although targeted interventions including mannitol and dexamethasone are not beneficial. Recent work shows association of cell-free hemoglobin oxidation stress involved in the pathophysiology of AKI in both adults and children. Paracetamol protected renal function likely by inhibiting cell-free-mediated oxidative stress. It is unclear if heme-mediated endothelial activation or oxidative stress is involved in cerebral malaria.

Summary 

The direct causes of cerebral and kidney dysfunction remain incompletely understood. Optimal treatment involves prompt diagnosis and effective antimalarial treatment with artesunate. Renal replacement therapy reduces mortality in AKI but delayed diagnosis is an issue.

INTRODUCTION

Severe malaria incidence is approximately two million cases with nearly 430,000 deaths annually [1] . It is a medical emergency characterized by multisystem disease with different clinical manifestations between adults and children. However, recent studies show that cerebral involvement, kidney dysfunction, and acidosis are independent predictors of mortality in both adults and children ( Fig. 1 ) [2,3] . This is supported by a meta-analysis of children with severe malaria that found prognostic indicators with the strongest association with death to be acute kidney injury (AKI) (odds ratio 5.96, 95% confidence interval, CI: 2.9–12.11) and coma (4.83, 95% CI: 3.11–7.5) [4▪▪] .

F1

Cerebral malaria is a clinical syndrome of impaired consciousness associated with malaria in the absence of hypoglycemia, convulsions, drugs, and nonmalarial causes characterized by unrousable coma defined by a Glasgow Coma Score less than11 (adults) [5] or Blantyre Coma Score less than 3 (children) [6–8] . Two large intervention trials in Asian adults and African children with severe malaria found that 54% of adult and 34% of pediatric patients had cerebral malaria [9,10] . AKI in severe falciparum malaria is caused by acute tubular necrosis and defined as a creatinine more than 265 μmol/l or urea more than 20 mmol/l [6] . In adults with severe malaria, AKI develops in up to 40% of patients, whereas in children, the incidence is historically reported at approximately 10% [9,10] . As the WHO definition does not define AKI adequately for pediatric malaria, the reported incidence of AKI in children is likely underestimated. The recent Kidney Disease: Improving Global Outcomes (KDIGO) classification standardizes AKI for clinical practice and research [11] . In adult severe malaria, 58% had AKI as defined by KDIGO, of whom 40% died, accounting for 71% of overall mortality [12▪] . Among children with severe malaria 46% had AKI as defined by KDIGO, of whom 12–24% died with increasingly severe AKI [13▪] . In two large multicenter studies, approximately 25% of children with severe malaria had increased blood urea nitrogen, accounting for roughly 50% of total deaths [10,14] . These studies imply that AKI complicating pediatric severe malaria has been previously underdiagnosed [13▪] .

Notably, the majority of patients surviving these complications have complete recovery after appropriate treatment. The direct cause of coma and AKI complicating severe malaria are incompletely understood but likely share common pathophysiological mechanisms. This review will highlight recent developments in our understanding of the pathophysiologic and pathologic processes associated with cerebral malaria and malaria-associated AKI in addition to the clinical presentation, diagnosis and treatments of these complications.  

FB1

PATHOGENESIS

Severe malaria is predominantly caused by Plasmodium falciparum because of its ability to induce infected red blood cell (RBC) cytoadherence to the vascular endothelium and consequent end-organ dysfunction. Other plasmodium species can cause severe disease and AKI [15] , although their ability to cause coma is debated [6] .

Microvascular obstruction

Parasites developing within the infected RBC transport P. falciparum erythrocyte membrane protein 1 ( Pf EMP1) to the RBC membrane functioning as a key ligand for cytoadherence [16] . Pf EMP1 is expressed on RBC protrusions, or ‘knobs’, that confer points of attachment to the endothelium. Pf EMP1 is strain specific, encoded by a highly variable var gene family, which provides antigenic variation for immune evasion and differential endothelial receptor binding. CD36 is an endothelial receptor constitutively expressed on most vascular beds [17] . The key endothelial receptor in the brain is intercellular adhesion molecule-1 [18] . Recent studies have identified endothelial protein C receptor as an important receptor in the brain that binds to a specific Pf EMP1 domain (CIDRα1) [19▪,20–22] . Cytoadhesion results in sequestration of parasitized RBCs in the capillaries and postcapillary venules causing heterogeneous blockage of the microcirculation and tissue hypoxia [23] . In addition to flow obstruction by sequestered parasitized RBCs, microcirculatory flow is thought to be further compromised by increased rigidity of both infected and uninfected RBCs and clumping of infected RBCs (platelet-mediated autoagglutination) and uninfected RBCs adhering to infected RBCs (rosette formation) [24] .

Direct visualization of microvascular obstruction is observed in the retina of adults and children with cerebral malaria, termed ‘malaria retinopathy’ [25,26] . Cerebral blood flow is not decreased in adults [27,28] . Intracranial pressure is often increased in children, but less so in adult patients [29,30] . Indirect assessment of sequestration via estimated total parasite biomass, measured by P. falciparum histidine rich protein 2, was shown to contribute to AKI in adults with severe malaria [31] . Autopsy studies of adults and retinopathy-positive children dying from cerebral malaria show prominent sequestration in the brain microvasculature compared to adults with fatal noncerebral malaria and retinopathy-negative children [32,33] . Postmortem studies report sequestration of parasitized RBCs in renal glomerular and peritubular capillaries in adults and children [32,34] .

Endothelial activation

Microvascular obstruction-induced tissue hypoxia is compounded by microvascular dysfunction [35,36] and increased oxygen demand [36,37] . In adults with cerebral malaria, there is endothelial and astroglial activation in the brain [18] , with variable inflammatory responses [38] and mild functional change to the blood–brain barrier [39,40] . In children with strictly defined retinopathy-positive cerebral malaria, breakdown of the endothelial barrier is observed particularly in areas of sequestration [32] . Patterns of histopathological change within the brain in cerebral malaria vary between adults and children, with less inflammatory cellular infiltrates and edema in adult cases [41,42] . A recent pediatric autopsy study found that HIV coinfection influences histopathology, increasing the degree of platelet and monocyte infiltration around damaged microvasculature [43] .

In a recent MRI study of similarly defined children, 35% had evidence of brain swelling most commonly in fatal cases implicating brainstem herniation as the cause of death [44] . Another recent serial MRI study in India including adults and children found that 50% of patients had evidence of brain swelling with posterior vasogenic edema and vascular congestion in the basal nuclei [45▪] . All patients had rapid clinical improvement and radiological reversibility with hallmarks suggestive of posterior reversible encephalopathy syndrome. The exact cause of the brain swelling is yet unclear.

Studies of patients with severe malaria having AKI show reduced renal cortical blood flow [46] , increased kidney size [47] , and endothelial changes in both glomerular and peritubular capillaries on histopathology [34] . Cell-free hemoglobin and lipid peroxidation markers are strongly associated with AKI and renal replacement therapy (RRT) requirement in adults with severe malaria [12▪] . In children with severe malaria, an elevated heme-to-hemopexin ratio was associated with hemoglobinuria, stage 3 AKI, and 6-month mortality [48▪] .

There is an imbalance of proinflammatory and anti-inflammatory responses in severe malaria [49] . The role of cytokines and chemokines in cerebral malaria has been recently reviewed [50] . However, many of these studies are in the murine experimental cerebral malaria model, the relevance of which has been questioned [51] . Conflicting evidence has emerged from human studies as to the association between cerebral malaria and levels of numerous cytokines such as tumor necrosis factor α (TNFα) [49,52–56] . Although cytokines and/or chemokines are clearly involved in the pathogenesis of malarial fever and may be associated with disease severity and/or cerebral malaria, it is not established that they are a cause of coma.

The role of cytokines and chemokines in the pathophysiology of AKI in severe malaria was recently highlighted. Plasma-soluble urokinase-type plasminogen activator receptor, a marker of immune activation, was independently associated with AKI and RRT requirement [31] . Previously, it was shown that TNFα, but not inteleukin (IL)6 or IL6:IL10 ratio, was associated with AKI suggesting that TNFα may induce localized renal tubular cell injury [57] .

CLINICAL FEATURES

The clinical presentation of cerebral malaria is diffuse symmetrical encephalopathy with fever and absent or few focal neurological signs. In children, coma can rapidly develop after fever onset (mean, 2 days) [7] . In adults, coma is typically gradual with increasing drowsiness, confusion, obtundation, and high fevers (mean duration, 5 days). Convulsions are present in approximately 15% of adults and 80% of children with severe malaria [9,10] and frequently herald development of coma. Patients may recover full consciousness after a convulsion, thus transient postictal coma must be excluded [6] . Multiple convulsions are common and up to 50% of comatose children have subclinical seizures or status epilepticus. Ocular funduscopic findings include vessel color change, macular and extramacular whitening, and white-centered retinal hemorrhages [58] ( Fig. 2 ).

F2

Among survivors, the median time to coma recovery is roughly 24 h in children and 48 h in adults [9,10] . Retinal abnormalities resolve with no residual visual deficit. Neurologic sequelae occur in less than 1% of adults but up to 12% of children in the quinine-therapy era, including hemiplegia, cortical blindness, aphasia, and cerebellar ataxia [59] . Studies suggest that neurologic deficits may reflect slow neurological recovery [10,60] . Postmalaria neurological syndrome is self-limiting [61] ; however, longer term neurological sequelae, including cognitive deficits and epilepsy, are reported among children [62,63▪] .

  • (1) Few severity criteria with prerenal AKI that resolves with fluids.
  • (2) Several severity criteria including AKI that resolves without RRT.
  • (3) Progressive AKI that resolves with antimalarial treatment and RRT.
  • (4) Multiorgan dysfunction, often with anuric AKI and cerebral malaria, who die prior to or during RRT with hemodynamic shock and/or respiratory failure.

Any comatose patient with a history of fever and/or travel to malaria-endemic regions must be considered to have cerebral malaria until proven otherwise. In children, febrile convulsions should be distinguished from cerebral malaria, wherein coma will persist beyond 1 h after anticonvulsive treatment is administered. Absence of fever does not rule out malaria. Antimalarial treatment should not be delayed in severely ill patients if diagnostics are unavailable or delayed.

Parasitological diagnosis is by microscopy of stained thin and thick blood smears. A rapid diagnostic test for parasite antigens can be performed if microscopy is unavailable. Patients may have a low circulating parasitemia because of sequestration, thus a low parasitemia is not reassuring [67] . In high-transmission settings, children with partial immunity tolerate higher parasitemia without severe symptoms and may be asymptomatic at low parasitemia. A parasitemia of more than 1 000 000/μl in African children with cerebral malaria is associated with fatal outcomes [7] .

Funduscopy for malaria retinopathy improves specificity for diagnosis of cerebral malaria and is prognostic in patients with severe malaria [26,68,69] ( Fig. 2 ). Alternative causes of coma must be ruled out including hypoglycemia, and bacterial, fungal, or viral meningoencephalitis. Lumbar puncture does not increase mortality in stable, comatose children with suspected cerebral malaria even when MRI brain swelling or papilledema is present [70▪] .

There is no robust prognostic risk model or biomarker that can predict AKI or RRT requirement [71,72▪] . All patients with malaria should be considered at risk of developing AKI. To improve outcomes, early diagnosis and management is critical. AKI diagnosis requires quantification of creatinine (or urea) or observing low urine volume (<0.5 ml/kg/h) for 6 h ( Fig. 3 ) [6,11] . Urine output is difficult to accurately assess in malaria endemic countries and may delay diagnosis. As the WHO creatinine threshold is not applicable to children, the diagnosis of AKI must be considered using all available patient data. The KDIGO AKI definition of a creatinine rise at least 1.5 times baseline is the current standard, and baseline creatinine can be back-calculated using the Modified Diet in Renal Disease (>19 years) or Swartz equation (≤18 years) [11,73] .

F3

The two key pillars of severe malaria treatment are prompt antimalarial treatment and supportive management. Adjunctive therapies targeted at the underlying pathophysiology are unproven.

Antimalarial treatment

Two landmark trials in patients with severe malaria definitively showed that intravenous artesunate reduced mortality by 35 and 23% in adults and children, respectively, compared to quinine [9,10] . Intravenous artesunate is now the first-line treatment for severe malaria as recommended by the WHO. Artemether and quinine are the second-line therapies [74] . The mechanism of improved survival over quinine is the rapid cidal activity of artesunate on young ring forms, preventing parasite maturation and sequestration [75] . Once the patient is able to take oral medication, and after a minimum of 24 h of artesunate, an oral artemisinin-based combination therapy can be initiated to complete the treatment.

Supportive treatment

Despite the best available artemisinin therapy for malaria, mortality remains unacceptably high and supportive management is key to reducing this. Comatose patients require endotracheal intubation with mechanical ventilation for airway protection. Rapid sequence intubation should be performed to prevent transient hypercapnia and increased intracranial pressure. Routine care should be implemented including regular turning, lateral positioning (‘recovery position’), and catheterization. Nasogastric tube insertion and suctioning may protect against aspiration, however, enteral feeding in nonintubated patients should be delayed (>60 h) because of increased risk of aspiration pneumonia [76] .

The majority of children with cerebral malaria experience convulsions. Glucose replacement to ensure euglycemia and fever control with paracetamol are important. Prophylactic anticonvulsant therapy is not recommended. A randomized controlled trial (RCT) of phenobarbital in pediatric cerebral malaria showed increased mortality, likely caused by respiratory depression [77] .

Fluid management and nephrotoxic drug avoidance are cornerstones for management of malaria-associated AKI. Cautious fluid management is important, as patients with AKI are not necessarily hypovolemic and are at high risk of developing pulmonary edema [78–81] . Rapid infusions may exacerbate intracranial hypertension and precipitate cerebral herniation. The large multicenter Fluid Expansion as Supportive Therapy study of African children with severe febrile illness showed a relative risk for death of 1.59 (95% CI: 1.10–2.31) with fluid bolus therapy among those with malaria [14] . The WHO recommends individualized restrictive fluid management, keeping the patient slightly dry, using slow infusion of isotonic crystalloids [74] . Patients with BWF require creatinine and hemoglobin monitoring as resulting severe anemia requires whole blood transfusion.

Treatment of malaria-associated AKI with RRT reduces mortality from 75 to 26% [64] . In general, RRT is urgently indicated when biochemical disturbances and volume overload refractory to conventional therapy are present. The additional thresholds included in the WHO malaria guidelines are based on findings that anuria and elevated or rapidly rising creatinine are sensitive indicators for RRT [6,64] . As AKI in malaria rapidly progresses and is often compounded by multiorgan dysfunction, early RRT is recommended. Although intermittent hemodialysis and continuous venovenous hemofiltration have been shown to be superior to peritoneal dialysis in adults with severe malaria [82] , in the absence of hemodialysis, life-saving peritoneal dialysis should be initiated if this is the only modality available [83] . RRT has also been shown to be effective in the management of malaria-associated AKI in pediatric patients [84] .

Many adjunctive therapies have been suggested, mainly driven by studies in murine experimental cerebral malaria. However, none has proven benefit in humans. Evidence for exchange transfusion and the more recently employed RBC exchange transfusion remains limited [85▪,86,87] . Mannitol [88,89] , steroids [90,91] , and monoclonal antibodies to TNF [54,56] are not recommended as treatments in cerebral malaria as studies show no benefits and potential harm. Furosemide and mannitol are ineffective in preventing and treating AKI and BWF, respectively, and may be harmful [92,93] . On the basis of the ability of paracetamol to inhibit hemoprotein-mediated AKI [94] , a recent RCT of paracetamol in Bangladeshi adults with severe malaria found that acetaminophen improved kidney function and reduced the development of AKI, particularly in patients with high cell-free hemoglobin levels at enrollment [95▪,96] . Larger studies of paracetamol in adults and children with malaria are currently ongoing to further assess this renoprotective effect.

Cerebral malaria and AKI complicating severe malaria are prognostic for mortality in both adults and children. Microvascular obstruction and endothelial dysfunction are common mechanisms for both of these complications. Future study of adjunctive therapies should target reducing sequestration, improving endovascular function, and reducing hemoglobin-mediated oxidative stress.

Acknowledgements

Financial support and sponsorship.

This work was supported by the Wellcome Trust of Great Britain.

Conflicts of interest

There are no conflicts of interest.

REFERENCES AND RECOMMENDED READING

▪ of special interest

▪▪ of outstanding interest

  • Cited Here |
  • Google Scholar

cerebral malaria; malaria-associated acute kidney injury; pathophysiology; treatment

  • + Favorites
  • View in Gallery

Readers Of this Article Also Read

Severe malaria: update on pathophysiology and treatment, update on drug treatments for multidrug resistant tuberculosis, pseudomonas aeruginosa</em> infections', 'do rego hermann; timsit, jean-fran\u00e7ois', 'current opinion in infectious diseases', 'december 2023', '36', '6' , 'p 585-595');" onmouseout="javascript:tooltip_mouseout()" class="ejp-uc__article-title-link">management strategies for severe pseudomonas aeruginosa infections, acinetobacter baumannii</em> infections', 'bouza emilio; mu\u00f1oz, patricia; burillo, almudena', 'current opinion in infectious diseases', 'december 2023', '36', '6' , 'p 596-608');" onmouseout="javascript:tooltip_mouseout()" class="ejp-uc__article-title-link">how to treat severe acinetobacter baumannii infections, treatment of mucormycosis in transplant patients: role of surgery and of old....

Log in using your username and password

  • Search More Search for this keyword Advanced search
  • Latest content
  • Current issue
  • BMJ Journals

You are here

  • Volume 69, Issue 4
  • Cerebral malaria
  • Article Text
  • Article info
  • Citation Tools
  • Rapid Responses
  • Article metrics

Download PDF

  • Charles R J C Newton a ,
  • Tran Tinh Hien b ,
  • Nicholas White b , c
  • a Neurosciences Unit, Institute of Child Health, London, United Kingdom, b Centre for Tropical Diseases, Cho Quan Hospital, Ho Chi Minh City, Vietnam, c Faculty of Tropical Medicine, Mahidol University, Bangkok, Thialand
  • Dr C R J C Newton, Wellcome Trust/ KEMRI Centre, PO Box 230, Kilifi, Kenya cnewton{at}kilifi.mimcom.net

Cerebral malaria may be the most common non-traumatic encephalopathy in the world. The pathogenesis is heterogenous and the neurological complications are often part of a multisystem dysfunction. The clinical presentation and pathophysiology differs between adults and children. Recent studies have elucidated the molecular mechanisms of pathogenesis and raised possible interventions. Antimalarial drugs, however, remain the only intervention that unequivocally affects outcome, although increasing resistance to the established antimalarial drugs is of grave concern. Artemisinin derivatives have made an impact on treatment, but other drugs may be required. With appropriate antimalarial drugs, the prognosis of cerebral malaria often depends on the management of other complications—for example, renal failure and acidosis. Neurological sequelae are increasingly recognised, but further research on the pathogenesis of coma and neurological damage is required to develop other ancillary treatments.

  • antimalarial drugs
  • parasitic disease

https://doi.org/10.1136/jnnp.69.4.433

Statistics from Altmetric.com

Request permissions.

If you wish to reuse any or all of this article please use the link below which will take you to the Copyright Clearance Center’s RightsLink service. You will be able to get a quick price and instant permission to reuse the content in many different ways.

Read the full text or download the PDF:

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Review Article
  • Published: 19 October 2023

Pathogenetic mechanisms and treatment targets in cerebral malaria

  • Alexandros Hadjilaou   ORCID: orcid.org/0000-0002-2443-9611 1 , 2 ,
  • Johannes Brandi 1 ,
  • Mathias Riehn   ORCID: orcid.org/0000-0001-8737-5566 1 ,
  • Manuel A. Friese   ORCID: orcid.org/0000-0001-6380-2420 2   na1 &
  • Thomas Jacobs 1   na1  

Nature Reviews Neurology volume  19 ,  pages 688–709 ( 2023 ) Cite this article

1997 Accesses

3 Citations

18 Altmetric

Metrics details

  • Central nervous system infections
  • Cerebrovascular disorders
  • Epidemiology
  • Infectious diseases
  • Paediatrics

An Author Correction to this article was published on 06 November 2023

This article has been updated

Malaria, the most prevalent mosquito-borne infectious disease worldwide, has accompanied humanity for millennia and remains an important public health issue despite advances in its prevention and treatment. Most infections are asymptomatic, but a small percentage of individuals with a heavy parasite burden develop severe malaria, a group of clinical syndromes attributable to organ dysfunction. Cerebral malaria is an infrequent but life-threatening complication of severe malaria that presents as an acute cerebrovascular encephalopathy characterized by unarousable coma. Despite effective antiparasite drug treatment, 20% of patients with cerebral malaria die from this disease, and many survivors of cerebral malaria have neurocognitive impairment. Thus, an important unmet clinical need is to rapidly identify people with malaria who are at risk of developing cerebral malaria and to develop preventive, adjunctive and neuroprotective treatments for cerebral malaria. This Review describes important advances in the understanding of cerebral malaria over the past two decades and discusses how these mechanistic insights could be translated into new therapies.

Cerebral malaria is an acute, life-threatening cerebrovascular complication characterized by unarousable coma; children <5 years are most at risk, although cerebral malaria can develop at any age.

Effective treatments for blood-stage malaria include antiparasite drugs such as artesunates and therapies that increase pitting, opsonization and phagocytic clearance of infected erythrocytes.

Host–parasite interactions result in sequestration of infected erythrocytes and immune cells on brain endothelial cells, which can lead to vascular disintegration, disruption of the blood–brain barrier and potentially neuropathology.

CD8 + T cell-mediated immune responses during cerebral malaria contribute to endothelial injury and blood–brain barrier disintegration, highlighting potential roles for therapies that target T cell recruitment, activation and cytotoxicity.

Host–parasite interactions result in the activation of microglia and astrocytes; thus, manipulation of these brain-resident cells might beneficially influence the neuropathology of cerebral malaria.

Cerebral malaria can be alleviated by combinations of therapies that clear infected cells and parasites, ideally complemented in the future by therapies that increase the resilience of the brain to tissue damage.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 12 print issues and online access

195,33 € per year

only 16,28 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

presentation of cerebral malaria

Similar content being viewed by others

presentation of cerebral malaria

Targeting the CD146/Galectin-9 axis protects the integrity of the blood–brain barrier in experimental cerebral malaria

presentation of cerebral malaria

Positron emission tomography and magnetic resonance imaging of the brain in experimental human malaria, a prospective cohort study

presentation of cerebral malaria

Whole blood transfusion improves vascular integrity and increases survival in artemether-treated experimental cerebral malaria

Change history, 06 november 2023.

A Correction to this paper has been published: https://doi.org/10.1038/s41582-023-00899-8

WHO. World Malaria Report 2022 (WHO, 2022).

Lindblade, K. A., Steinhardt, L., Samuels, A., Kachur, S. P. & Slutsker, L. The silent threat: asymptomatic parasitemia and malaria transmission. Expert Rev. Anti Infect. Ther. 11 , 623–639 (2013).

Article   CAS   PubMed   Google Scholar  

Reyburn, H. et al. Association of transmission intensity and age with clinical manifestations and case fatality of severe Plasmodium falciparum malaria. J. Am. Med. Assoc. 293 , 1461–1470 (2005).

Article   CAS   Google Scholar  

Marsh, K. et al. Indicators of life-threatening malaria in African children. N. Engl. J. Med. 332 , 1399–1404 (1995).

Tack, B. et al. Health itinerary-related survival of children under-five with severe malaria or bloodstream infection, DR Congo. PLoS Negl. Trop. Dis. 17 , e0011156 (2023).

Article   PubMed   PubMed Central   Google Scholar  

Adeboye, M. A., Ojuawo, A., Ernest, S. K., Fadeyi, A. & Salisu, O. T. Mortality pattern within twenty-four hours of emergency paediatric admission in a resource-poor nation health facility. West Afr. J. Med. 29 , 249–252 (2010).

CAS   PubMed   Google Scholar  

Chiabi, A. et al. Severe malaria in Cameroon: pattern of disease in children at the Yaounde Gynaeco-Obstetric and Pediatric Hospital. J. Infect. Public Health 13 , 1469–1472 (2020).

Article   PubMed   Google Scholar  

Dondorp, A. M. et al. The relationship between age and the manifestations of and mortality associated with severe malaria. Clin. Infect. Dis. 47 , 151–157 (2008).

Dondorp, A. M. et al. Artesunate versus quinine in the treatment of severe falciparum malaria in African children (AQUAMAT): an open-label, randomised trial. Lancet 376 , 1647–1657 (2010).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Dondorp, A., Nosten, F., Stepniewska, K., Day, N. & White, N. Artesunate versus quinine for treatment of severe falciparum malaria: a randomised trial. Lancet 366 , 717–725 (2005).

Sahu, P. K. et al. Brain magnetic resonance imaging reveals different courses of disease in pediatric and adult cerebral malaria. Clin. Infect. Dis. 73 , e2387–e2396 (2021).

Johnson, H. et al. Multiple organ dysfunction syndrome and Pediatric Logistic Organ Dysfunction-2 score in pediatric cerebral malaria. Am. J. Trop. Med. Hyg. 107 , 820–826 (2022).

Kochar, D. K. et al. Clinical features of children hospitalized with malaria—a study from Bikaner, northwest India. Am. J. Trop. Med. Hyg. 83 , 981–989 (2010).

Tembo, D. et al. Risk factors for acute kidney injury at presentation among children with CNS malaria: a case control study. Malar. J. 21 , 310 (2022).

White, N. J. Severe malaria. Malar. J. 21 , 284 (2022).

Ashley, E. A., Pyae Phyo, A. & Woodrow, C. J. Malaria. Lancet 391 , 1608–1621 (2018).

World Health Organization. WHO Guidelines for Malaria, 14 March 2023. https://apps.who.int/iris/handle/10665/366432 (2023).

White, N. J. The parasite clearance curve. Malar. J. 10 , 278 (2011).

Conroy, A. L. et al. Parenteral artemisinins are associated with reduced mortality and neurologic deficits and improved long-term behavioral outcomes in children with severe malaria. BMC Med. 19 , 168 (2021).

Flannery, E. L., Chatterjee, A. K. & Winzeler, E. A. Antimalarial drug discovery — approaches and progress towards new medicines. Nat. Rev. Microbiol. 11 , 849–862 (2013).

Ashley, E. A. & Phyo, A. P. Drugs in development for malaria. Drugs 78 , 861–879 (2018).

Riley, E. M. & Stewart, V. A. Immune mechanisms in malaria: new insights in vaccine development. Nat. Med. 19 , 168–178 (2013).

Cockburn, I. A. & Seder, R. A. Malaria prevention: from immunological concepts to effective vaccines and protective antibodies. Nat. Immunol. 19 , 1199–1211 (2018).

De Koning-Ward, T. F., Dixon, M. W. A., Tilley, L. & Gilson, P. R. Plasmodium species: master renovators of their host cells. Nat. Rev. Microbiol. 14 , 494–507 (2016).

White, N. J., Turner, G. D. H., Day, N. P. J. & Dondorp, A. M. Lethal malaria: Marchiafava and Bignami were right. J. Infect. Dis. 208 , 192–198 (2013).

Bignami, A., Felkin, R. W., Mannaberg, J., Marchiafava, E. & Thompson, J. H. Two monographs on malaria and the parasites of malarial fevers I. Marchiafava and Bignami. II. Mannaberg (New Sydenham Society, 1894).

Varo, R., Erice, C., Johnson, S., Bassat, Q. & Kain, K. C. Clinical trials to assess adjuvant therapeutics for severe malaria. Malar. J. 19 , 268 (2020).

Wahlgren, M., Goel, S. & Akhouri, R. R. Variant surface antigens of Plasmodium falciparum and their roles in severe malaria. Nat. Rev. Microbiol. 15 , 479–491 (2017).

Mebius, R. E. & Kraal, G. Structure and function of the spleen. Nat. Rev. Immunol. 5 , 606–616 (2005).

Ndour, P. A. et al. Plasmodium falciparum clearance is rapid and pitting independent in immune Malian children treated with artesunate for malaria. J. Infect. Dis. 211 , 290–297 (2015).

Newton, P. N. et al. A comparison of the in vivo kinetics of Plasmodium falciparum ring-infected erythrocyte surface antigen-positive and -negative erythrocytes. Blood 98 , 450–457 (2001).

Angus, B. J., Chotivanich, K., Udomsangpetch, R. & White, N. J. In vivo removal of malaria parasites from red blood cells without their destruction in acute falciparum malaria. Blood 90 , 2037–2040 (1997).

Ayi, K. et al. CD47-SIRPα interactions regulate macrophage uptake of Plasmodium falciparum -infected erythrocytes and clearance of malaria in vivo. Infect. Immun. 84 , 2002–2011 (2016).

Torrez Dulgeroff, L. B. et al. CD47 blockade reduces the pathologic features of experimental cerebral malaria and promotes survival of hosts with Plasmodium infection. Proc. Natl Acad. Sci. USA 118 , e1907653118 (2021).

Banerjee, R., Khandelwal, S., Kozakai, Y., Sahu, B. & Kumar, S. CD47 regulates the phagocytic clearance and replication of the Plasmodium yoelii malaria parasite. Proc. Natl Acad. Sci. USA 112 , 3062–3067 (2015).

Feng, M. et al. Phagocytosis checkpoints as new targets for cancer immunotherapy. Nat. Rev. Cancer 19 , 568–586 (2019).

Ma, S. et al. A role of PIEZO1 in iron metabolism in mice and humans. Cell 184 , 969–982.e13 (2021).

Ma, S. et al. Common PIEZO1 allele in African populations causes RBC dehydration and attenuates Plasmodium infection. Cell 173 , 443–455.e12 (2018).

Rosa, T. F. A. et al. The Plasmodium falciparum blood stages acquire factor H family proteins to evade destruction by human complement. Cell. Microbiol. 18 , 573–590 (2016).

Kennedy, A. T. et al. Recruitment of factor H as a novel complement evasion strategy for blood-stage Plasmodium falciparum infection. J. Immunol. 196 , 1239–1248 (2016).

Kiyuka, P. K., Meri, S. & Khattab, A. Complement in malaria: immune evasion strategies and role in protective immunity. FEBS Lett. 594 , 2502–2517 (2020).

Patel, S. N. et al. C5 deficiency and C5a or C5aR blockade protects against cerebral malaria. J. Exp. Med. 205 , 1133–1143 (2008).

Kim, H. et al. Functional roles for C5a and C5aR but not C5L2 in the pathogenesis of human and experimental cerebral malaria. Infect. Immun. 82 , 371–379 (2014).

Biryukov, S. & Stoute, J. A. Complement activation in malaria: friend or foe? Trends Mol. Med. 20 , 293–301 (2014).

Schein, T. N. & Barnum, S. R. Role of complement in cerebral malaria. In Complement Activation in Malaria Immunity and Pathogenesis (ed. Stoute, J. A.) 65–90 (2018); https://doi.org/10.1007/978-3-319-77258-5_4 .

Lee, W. C. et al. Plasmodium -infected erythrocytes induce secretion of IGFBP7 to form type II rosettes and escape phagocytosis. eLife 9 , e51546 (2020).

Albrecht, L. et al. Rosettes integrity protects Plasmodium vivax of being phagocytized. Sci. Rep. 10 , 16706 (2020).

Lee, W. C. et al. Plasmodium falciparum rosetting protects schizonts against artemisinin. EBioMedicine 73 , 103680 (2021).

Moles, E. et al. Development of drug-loaded immunoliposomes for the selective targeting and elimination of rosetting Plasmodium falciparum -infected red blood cells. J. Control. Release 241 , 57–67 (2016).

Niang, M. et al. STEVOR is a Plasmodium falciparum erythrocyte binding protein that mediates merozoite invasion and rosetting. Cell Host Microbe 16 , 81–93 (2014).

Goel, S. et al. RIFINs are adhesins implicated in severe Plasmodium falciparum malaria. Nat. Med. 21 , 314–317 (2015).

Dorovini-Zis, K. et al. The neuropathology of fatal cerebral malaria in malawian children. Am. J. Pathol. 178 , 2146–2158 (2011).

Milner, D. A. Jr et al. The systemic pathology of cerebral malaria in African children. Front. Cell. Infect. Microbiol. 4 , 104 (2014).

Moxon, C. A. et al. Persistent endothelial activation and inflammation after Plasmodium falciparum infection in Malawian children. J. Infect. Dis. 209 , 610–615 (2014).

Nickerson, J. P., Tong, K. A. & Raghavan, R. Imaging cerebral malaria with a susceptibility-weighted MR sequence. AJNR Am. J. Neuroradiol. 30 , e85–e86 (2009).

Kraemer, S. M. & Smith, J. D. A family affair: var genes, PfEMP1 binding, and malaria disease. Curr. Opin. Microbiol. 9 , 374–380 (2006).

Turner, L. et al. Severe malaria is associated with parasite binding to endothelial protein C receptor. Nature 498 , 502–505 (2013).

Kessler, A. et al. Linking EPCR-binding PfEMP1 to brain swelling in pediatric cerebral malaria. Cell Host Microbe 22 , 601–614.e5 (2017).

Wichers, J. S. et al. Common virulence gene expression in adult first-time infected malaria patients and severe cases. eLife 10 , e69040 (2021).

Lau, C. K. Y. et al. Structural conservation despite huge sequence diversity allows EPCR binding by the PfEMP1 family implicated in severe childhood malaria. Cell Host Microbe 17 , 118–129 (2015).

Berendt, A. R., Simmons, D. L., Tansey, J., Newbold, C. I. & Marsh, K. Intercellular adhesion molecule-1 is an endothelial cell adhesion receptor for Plasmodium falciparum . Nature 341 , 57–59 (1989).

Fonager, J. et al. Reduced CD36-dependent tissue sequestration of Plasmodium -infected erythrocytes is detrimental to malaria parasite growth in vivo. J. Exp. Med. 209 , 93–107 (2012).

Vogt, A. M. et al. Heparan sulfate on endothelial cells mediates the binding of Plasmodium falciparum -infected erythrocytes via the DBL1α domain of PfEMP1. Blood 101 , 2405–2411 (2003).

Robert, C. et al. Chondroitin-4-sulphate (proteoglycan), a receptor for Plasmodium falciparum -infected erythrocyte adherence on brain microvascular endothelial cells. Res. Immunol. 146 , 383–393 (1995).

Rogerson, S. J., Chaiyaroj, S. C., Ng, K., Reeder, J. C. & Brown, G. V. Chondroitin sulfate A is a cell surface receptor for Plasmodium falciparum -infected erythrocytes. J. Exp. Med. 182 , 15–20 (1995).

Tuikue Ndam, N. et al. Parasites causing cerebral falciparum malaria bind multiple endothelial receptors and express EPCR and ICAM-1-binding PfEMP1. J. Infect. Dis. 215 , 1918–1925 (2017).

Lennartz, F. et al. Structure-guided identification of a family of dual receptor-binding PfEMP1 that is associated with cerebral malaria. Cell Host Microbe 21 , 403–414 (2017).

Duffy, F. et al. Meta-analysis of Plasmodium falciparum var signatures contributing to severe malaria in African children and Indian adults. MBio 10 , e00217 (2019).

Jensen, A. R., Adams, Y. & Hviid, L. Cerebral Plasmodium falciparum malaria: the role of PfEMP1 in its pathogenesis and immunity, and PfEMP1-based vaccines to prevent it. Immunol. Rev. 293 , 230–252 (2020).

Avril, M. et al. A restricted subset of var genes mediates adherence of Plasmodium falciparum -infected erythrocytes to brain endothelial cells. Proc. Natl Acad. Sci. USA 109 , E1782–E1790 (2012).

Claessens, A. et al. A subset of group A-like var genes encodes the malaria parasite ligands for binding to human brain endothelial cells. Proc. Natl Acad. Sci. USA 109 , E1772–E1781 (2012).

Bernabeu, M. et al. Severe adult malaria is associated with specific PfEMP1 adhesion types and high parasite biomass. Proc. Natl Acad. Sci. USA 113 , E3270–E3279 (2016).

Turner, G. D. H. 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. 145 , 1057–1069 (1994).

CAS   PubMed   PubMed Central   Google Scholar  

Ochola, L. B. et al. Specific receptor usage in Plasmodium falciparum cytoadherence is associated with disease outcome. PLoS One 6 , e14741 (2011).

Franke-Fayard, B. et al. Murine malaria parasite sequestration: CD36 is the major receptor, but cerebral pathology is unlinked to sequestration. Proc. Natl Acad. Sci. USA 102 , 11468–11473 (2005).

Bernabeu, M. & Smith, J. D. EPCR and malaria severity: the center of a perfect storm. Trends Parasitol. 33 , 295–308 (2017).

Utter, C., Serrano, A. E., Glod, J. W. & Leibowitz, M. J. Association of Plasmodium falciparum with human endothelial cells in vitro. Yale J. Biol. Med. 90 , 183–193 (2017).

Khaw, L. T. et al. Brain endothelial cells increase the proliferation of Plasmodium falciparum through production of soluble factors. Exp. Parasitol. 145 , 34–41 (2014).

Khoury, D. S. et al. Effect of mature blood-stage Plasmodium parasite sequestration on pathogen biomass in mathematical and in vivo models of malaria. Infect. Immun. 82 , 212–220 (2014).

Adams, Y. et al. Plasmodium falciparum erythrocyte membrane protein 1 variants induce cell swelling and disrupt the blood-brain barrier in cerebral malaria. J. Exp. Med. 218 , e20201266 (2021).

Fens, M. H. A. M. et al. A role for activated endothelial cells in red blood cell clearance: implications for vasopathology. Haematologica 97 , 500–508 (2012).

Sun, J. et al. Insights into the mechanisms of brain endothelial erythrophagocytosis. Front. Cell Dev. Biol. 9 , 672009 (2021).

Chang, R. et al. Brain endothelial erythrophagocytosis and hemoglobin transmigration across brain endothelium: implications for pathogenesis of cerebral microbleeds. Front. Cell. Neurosci. 12 , 279 (2018).

Cabrales, P., Zanini, G. M., Meays, D., Frangos, J. A. & Carvalho, L. J. M. Murine cerebral malaria is associated with a vasospasm-like microcirculatory dysfunction, and survival upon rescue treatment is markedly increased by nimodipine. Am. J. Pathol. 176 , 1306–1315 (2010).

van Meer, G. & Simons, K. The function of tight junctions in maintaining differences in lipid composition between the apical and the basolateral cell surface domains of MDCK cells. EMBO J. 5 , 1455–1464 (1986).

Dragsten, P. R., Blumenthal, R. & Handler, J. S. Membrane asymmetry in epithelia: is the tight junction a barrier to diffusion in the plasma membrane? Nature 294 , 718–722 (1981).

Turner, G. D. H. et al. Systemic endothelial activation occurs in both mild and severe malaria. Correlating dermal microvascular endothelial cell phenotype and soluble cell adhesion molecules with disease severity. Am. J. Pathol. 152 , 1477–1487 (1998).

Tripathi, A. K., Sullivan, D. J. & Stins, M. F. Plasmodium falciparum -infected erythrocytes increase intercellular adhesion molecule 1 expression on brain endothelium through NF-kappaB. Infect. Immun. 74 , 3262–3270 (2006).

Darling, T. K. et al. EphA2 contributes to disruption of the blood-brain barrier in cerebral malaria. PLoS Pathog. 16 , e1008261 (2020).

Conroy, A. L. et al. Endothelium-based biomarkers are associated with cerebral malaria in Malawian children: a retrospective case-control study. PLoS One 5 , e15291 (2010).

Yeo, T. W. et al. Glycocalyx breakdown is associated with severe disease and fatal outcome in Plasmodium falciparum malaria. Clin. Infect. Dis. 69 , 1712–1720 (2019).

Lyimo, E. et al. In vivo imaging of the buccal mucosa shows loss of the endothelial glycocalyx and perivascular hemorrhages in pediatric Plasmodium falciparum malaria. Infect. Immun. 88 , e00679–19 (2020).

Jakobsen, P. H. et al. Increased plasma concentrations of sICAM-1, sVCAM-1 and sELAM-1 in patients with Plasmodium falciparum or P. vivax malaria and association with disease severity. Immunology 83 , 665–669 (1994).

Tchinda, V. H. M. et al. Severe malaria in Cameroonian children: correlation between plasma levels of three soluble inducible adhesion molecules and TNF-α. Acta Trop. 102 , 20–28 (2007).

Saharinen, P., Eklund, L. & Alitalo, K. Therapeutic targeting of the angiopoietin–TIE pathway. Nat. Rev. Drug Discov. 16 , 635–661 (2017).

Gavard, J., Patel, V. & Gutkind, J. S. Angiopoietin-1 prevents VEGF-induced endothelial permeability by sequestering Src through mDia. Dev. Cell 14 , 25–36 (2008).

Fiedler, U. et al. Angiopoietin-2 sensitizes endothelial cells to TNF-α and has a crucial role in the induction of inflammation. Nat. Med. 12 , 235–239 (2006).

Wallez, Y. et al. Src kinase phosphorylates vascular endothelial-cadherin in response to vascular endothelial growth factor: identification of tyrosine 685 as the unique target site. Oncogene 26 , 1067–1077 (2006).

Gallego-Delgado, J. et al. Angiotensin receptors and β-catenin regulate brain endothelial integrity in malaria. J. Clin. Invest. 126 , 4016–4029 (2016).

Dhangadamajhi, G., Mohapatra, B. N., Kar, S. K. & Ranjit, M. Gene polymorphisms in angiotensin I converting enzyme (ACE I/D) and angiotensin II converting enzyme (ACE2 C–>T) protect against cerebral malaria in Indian adults. Infect. Genet. Evol. 10 , 337–341 (2010).

Saraiva, V. B. et al. Impairment of the Plasmodium falciparum erythrocytic cycle induced by angiotensin peptides. PLoS One 6 , e17174 (2011).

Gallego-Delgado, J. et al. Angiotensin II moderately decreases Plasmodium infection and experimental cerebral malaria in mice. PLoS One 10 , e0138191 (2015).

Maciel, C. et al. Anti- Plasmodium activity of angiotensin II and related synthetic peptides. PLoS One 3 , e3296 (2008).

Gillrie, M. R. et al. Src-family kinase dependent disruption of endothelial barrier function by Plasmodium falciparum merozoite proteins. Blood 110 , 3426–3435 (2007).

Kunkel, G. T., MacEyka, M., Milstien, S. & Spiegel, S. Targeting the sphingosine-1-phosphate axis in cancer, inflammation and beyond. Nat. Rev. Drug Discov. 12 , 688–702 (2013).

Finney, C. A. M. et al. S1P is associated with protection in human and experimental cerebral malaria. Mol. Med. 17 , 717–725 (2011).

Nacer, A. et al. Neuroimmunological blood brain barrier opening in experimental cerebral malaria. PLoS Pathog. 8 , e1002982 (2012).

Nacer, A. et al. Experimental cerebral malaria pathogenesis—hemodynamics at the blood brain barrier. PLoS Pathog. 10 , e1004528 (2014).

Lovegrove, F. E. et al. Serum angiopoietin-1 and -2 levels discriminate cerebral malaria from uncomplicated malaria and predict clinical outcome in African children. PLoS One 4 , e4912 (2009).

Conroy, A. L. et al. Angiopoietin-2 levels are associated with retinopathy and predict mortality in Malawian children with cerebral malaria: a retrospective case-control study. Crit. Care Med. 40 , 952–959 (2012).

Wassmer, S. C. et al. Investigating the pathogenesis of severe malaria: a multidisciplinary and cross-geographical approach. Am. J. Trop. Med. Hyg. 93 (Suppl), 42–56 (2015).

Papapetropoulos, A. et al. Angiopoietin-1 inhibits endothelial cell apoptosis via the Akt/survivin pathway. J. Biol. Chem. 275 , 9102–9105 (2000).

Higgins, S. J. et al. Dysregulation of angiopoietin-1 plays a mechanistic role in the pathogenesis of cerebral malaria. Sci. Transl. Med. 8 , 358ra128 (2016).

Ziegler, T. et al. Angiopoietin 2 mediates microvascular and hemodynamic alterations in sepsis. J. Clin. Invest. 123 , 3436–3445 (2013).

Han, S. et al. Amelioration of sepsis by TIE2 activation-induced vascular protection. Sci. Transl. Med. 8 , 335ra55 (2016).

Rasmussen, A. L. et al. Host genetic diversity enables Ebola hemorrhagic fever pathogenesis and resistance. Science 346 , 987–991 (2014).

Phanthanawiboon, S. et al. Acute systemic infection with dengue virus leads to vascular leakage and death through tumor necrosis factor-α and Tie2/angiopoietin signaling in mice lacking type I and II interferon receptors. PLoS One 11 , e0148564 (2016).

Stiehl, T. et al. Lung-targeted RNA interference against angiopoietin-2 ameliorates multiple organ dysfunction and death in sepsis. Crit. Care Med. 42 , e654–e662 (2014).

Ghosh, C. C. et al. Impaired function of the Tie-2 receptor contributes to vascular leakage and lethality in anthrax. Proc. Natl Acad. Sci. USA 109 , 10024–10029 (2012).

Serghides, L. et al. Rosiglitazone modulates the innate immune response to Plasmodium falciparum infection and improves outcome in experimental cerebral malaria. J. Infect. Dis. 199 , 1536–1545 (2009).

Serghides, L. et al. PPARγ agonists improve survival and neurocognitive outcomes in experimental cerebral malaria and induce neuroprotective pathways in human malaria. PLoS Pathog. 10 , e1003980 (2014).

Bopp, S. E. R. et al. Genome wide analysis of inbred mouse lines identifies a locus containing Ppar-gamma as contributing to enhanced malaria survival. PLoS One 5 , e10903 (2010).

Boggild, A. K. et al. Use of peroxisome proliferator-activated receptor γ agonists as adjunctive treatment for Plasmodium falciparum malaria: a randomized, double-blind, placebo-controlled trial. Clin. Infect. Dis. 49 , 841–849 (2009).

Opitz, B., Eitel, J., Meixenberger, K. & Suttorp, N. Role of Toll-like receptors, NOD-like receptors and RIG-I-like receptors in endothelial cells and systemic infections. Thromb. Haemost. 102 , 1103–1109 (2009).

Sisquella, X. et al. Malaria parasite DNA-harbouring vesicles activate cytosolic immune sensors. Nat. Commun. 8 , 1985 (2017).

Gowda, D. C. & Wu, X. Parasite recognition and signaling mechanisms in innate immune responses to malaria. Front. Immunol. https://doi.org/10.3389/fimmu.2018.03006 (2018).

Liehl, P. et al. Host-cell sensors for Plasmodium activate innate immunity against liver-stage infection. Nat. Med. 20 , 47–53 (2014).

Gazzinelli, R. T., Kalantari, P., Fitzgerald, K. A. & Golenbock, D. T. Innate sensing of malaria parasites. Nat. Rev. Immunol. 14 , 744–757 (2014).

Sharma, S. et al. Innate immune recognition of an AT-rich stem-loop DNA motif in the Plasmodium falciparum genome. Immunity 35 , 194–207 (2011).

Pais, T. F. et al. Brain endothelial STING1 activation by Plasmodium -sequestered heme promotes cerebral malaria via type I IFN response. Proc. Natl Acad. Sci. USA 119 , e2206327119 (2022).

Decout, A., Katz, J. D., Venkatraman, S. & Ablasser, A. The cGAS–STING pathway as a therapeutic target in inflammatory diseases. Nat. Rev. Immunol. 21 , 548–569 (2021).

Schrezenmeier, E. & Dörner, T. Mechanisms of action of hydroxychloroquine and chloroquine: implications for rheumatology. Nat. Rev. Rheumatol. 16 , 155–166 (2020).

An, J., Woodward, J. J., Sasaki, T., Minie, M. & Elkon, K. B. Cutting edge: antimalarial drugs inhibit IFN-β production through blockade of cyclic GMP-AMP synthase-DNA interaction. J. Immunol. 194 , 4089–4093 (2015).

Zhang, X. et al. The cytosolic DNA sensor cGAS forms an oligomeric complex with DNA and undergoes switch-like conformational changes in the activation loop. Cell Rep. 6 , 421–430 (2014).

He, X. et al. RTP4 inhibits IFN-I response and enhances experimental cerebral malaria and neuropathology. Proc. Natl Acad. Sci. USA 117 , 19465–19474 (2020).

Dunst, J., Kamena, F. & Matuschewski, K. Cytokines and chemokines in cerebral malaria pathogenesis. Front. Cell. Infect. Microbiol. 7 , 324 (2017).

Yañez, D. M., Manning, D. D., Cooley, A. J., Weidanz, W. P. & van der Heyde, H. C. Participation of lymphocyte subpopulations in the pathogenesis of experimental murine cerebral malaria. J. Immunol. 157 , 1620–1624 (1996).

Grau, G. E. et al. Monoclonal antibody against interferon γ can prevent experimental cerebral malaria and its associated overproduction of tumor necrosis factor. Proc. Natl Acad. Sci. USA 86 , 5572–5574 (1989).

Belnoue, E. et al. Control of pathogenic CD8 + T cell migration to the brain by IFN-γ during experimental cerebral malaria. Parasite Immunol. 30 , 544–553 (2008).

Amani, V. et al. Involvement of IFN-γ receptor-medicated signaling in pathology and anti-malarial immunity induced by Plasmodium berghei infection. Eur. J. Immunol. 30 , 1646–1655 (2000).

Villegas-Mendez, A. et al. Gamma interferon mediates experimental cerebral malaria by signaling within both the hematopoietic and nonhematopoietic compartments. Infect. Immun. 85 , e01035–16 (2017).

Torre, S. et al. USP15 regulates type I interferon response and is required for pathogenesis of neuroinflammation. Nat. Immunol. 18 , 54–63 (2016).

Vigário, A. M. et al. Recombinant human IFN-α inhibits cerebral malaria and reduces parasite burden in mice. J. Immunol. 178 , 6416–6425 (2007).

Xu, S. et al. The zinc finger transcription factor, KLF2, protects against COVID-19 associated endothelial dysfunction. Signal Transduct. Target. Ther. 6 , 266 (2021).

Waknine-Grinberg, J. H., McQuillan, J. A., Hunt, N., Ginsburg, H. & Golenser, J. Modulation of cerebral malaria by fasudil and other immune-modifying compounds. Exp. Parasitol. 125 , 141–146 (2010).

Wilson, N. O. et al. Pharmacologic inhibition of CXCL10 in combination with anti-malarial therapy eliminates mortality associated with murine model of cerebral malaria. PLoS One 8 , e60898 (2013).

Bienvenu, A. L. & Picot, S. Statins alone are ineffective in cerebral malaria but potentiate artesunate. Antimicrob. Agents Chemother. 52 , 4203–4204 (2008).

Souraud, J. B. et al. Atorvastatin treatment is effective when used in combination with mefloquine in an experimental cerebral malaria murine model. Malar. J. 11 , 13 (2012).

Dormoi, J. et al. Improvement of the efficacy of dihydroartemisinin with atorvastatin in an experimental cerebral malaria murine model. Malar. J. 12 , 302 (2013).

Reis, P. A. et al. Statins decrease neuroinflammation and prevent cognitive impairment after cerebral malaria. PLoS Pathog. 8 , e1003099 (2012).

Parihar, S. P., Guler, R. & Brombacher, F. Statins: a viable candidate for host-directed therapy against infectious diseases. Nat. Rev. Immunol. 19 , 104–117 (2018).

Article   Google Scholar  

Kopp, E. & Ghosh, S. Inhibition of NF-kappa B by sodium salicylate and aspirin. Science 265 , 956–959 (1994).

Kauppinen, A., Suuronen, T., Ojala, J., Kaarniranta, K. & Salminen, A. Antagonistic crosstalk between NF-κB and SIRT1 in the regulation of inflammation and metabolic disorders. Cell Signal . 25 , 1939–1948 (2013).

Rodriguez-Muñoz, D. et al. Hypothyroidism confers tolerance to cerebral malaria. Sci. Adv. 8 , eabj7110 (2022).

Abdrabou, W. et al. Upregulation of steroidogenesis is associated with coma in human cerebral malaria. Preprint at bioRxiv https://doi.org/10.1101/2023.05.01.538900 (2023).

Warrell, D. A. et al. Dexamethasone proves deleterious in cerebral malaria. A double-blind trial in 100 comatose patients. N. Engl. J. Med. 306 , 313–319 (1982).

Abdrabou, W. et al. Metabolome modulation of the host adaptive immunity in human malaria. Nat. Metab. 3 , 1001–1016 (2021).

Hoffman, S. L. et al. High-dose dexamethasone in quinine-treated patients with cerebral malaria: a double-blind, placebo-controlled trial. J. Infect. Dis. 158 , 325–331 (1988).

Prasad, K. & Garner, P. Steroids for treating cerebral malaria. Cochrane Database Syst. Rev. https://doi.org/10.1002/14651858.CD000972 (2000).

Cain, D. W. & Cidlowski, J. A. Immune regulation by glucocorticoids. Nat. Rev. Immunol. 17 , 233–247 (2017).

Rhen, T. & Cidlowski, J. A. Antiinflammatory action of glucocorticoids—new mechanisms for old drugs. N. Engl. J. Med. 353 , 1711–1723 (2005).

Moreira, D. R. et al. Dexamethasone increased the survival rate in Plasmodium berghei -infected mice. Sci. Rep. 11 , 2623 (2021).

Sanni, L. A. et al. Dramatic changes in oxidative tryptophan metabolism along the kynurenine pathway in experimental cerebral and noncerebral malaria. Am. J. Pathol. 152 , 611–619 (1998).

Farah, C., Michel, L. Y. M. & Balligand, J. L. Nitric oxide signalling in cardiovascular health and disease. Nat. Rev. Cardiol. 15 , 292–316 (2018).

Yeo, T. W. et al. Impaired nitric oxide bioavailability and L-arginine reversible endothelial dysfunction in adults with falciparum malaria. J. Exp. Med. 204 , 2693–2704 (2007).

Kwiatkowski, D. P. How malaria has affected the human genome and what human genetics can teach us about malaria. Am. J. Hum. Genet. 77 , 171–192 (2005).

Conroy, A. L. et al. Methemoglobin and nitric oxide therapy in Ugandan children hospitalized for febrile illness: results from a prospective cohort study and randomized double-blind placebo-controlled trial. BMC Pediatr. 16 , 177 (2016).

Mwanga-Amumpaire, J. et al. Inhaled nitric oxide as an adjunctive treatment for cerebral malaria in children: a phase II randomized open-label clinical trial. Open Forum Infect. Dis. 2 , ofv111 (2015).

Hawkes, M. T. et al. Inhaled nitric oxide as adjunctive therapy for severe malaria: a randomized controlled trial. Malar. J. 14 , 421 (2015).

Bangirana, P. et al. Inhaled nitric oxide and cognition in pediatric severe malaria: A randomized double-blind placebo controlled trial. PLoS One 13 , e0191550 (2018).

Gramaglia, I. et al. Low nitric oxide bioavailability contributes to the genesis of experimental cerebral malaria. Nat. Med. 12 , 1417–1422 (2006).

Cabrales, P., Zanini, G. M., Meays, D., Frangos, J. A. & Carvalho, L. J. M. Nitric oxide protection against murine cerebral malaria is associated with improved cerebral microcirculatory physiology. J. Infect. Dis. 203 , 1454–1463 (2011).

Zanini, G. M., Cabrales, P., Barkho, W., Frangos, J. A. & Carvalho, L. J. M. Exogenous nitric oxide decreases brain vascular inflammation, leakage and venular resistance during Plasmodium berghei ANKA infection in mice. J. Neuroinflammation 8 , 66 (2011).

Martins, Y. C., Zanini, G. M., Frangos, J. A. & Carvalho, L. J. M. Efficacy of different nitric oxide-based strategies in preventing experimental cerebral malaria by Plasmodium berghei ANKA. PLoS One 7 , e32048 (2012).

Serghides, L. et al. Inhaled nitric oxide reduces endothelial activation and parasite accumulation in the brain, and enhances survival in experimental cerebral malaria. PLoS One 6 , e27714 (2011).

Lopansri, B. K. et al. Low plasma arginine concentrations in children with cerebral malaria and decreased nitric oxide production. Lancet 361 , 676–678 (2003).

Olszewski, K. L. et al. Host-parasite interactions revealed by Plasmodium falciparum metabolomics. Cell Host Microbe 5 , 191–199 (2009).

Gramaglia, I. et al. Citrulline protects mice from experimental cerebral malaria by ameliorating hypoargininemia, urea cycle changes and vascular leak. PLoS One 14 , e0213428 (2019).

Elphinstone, R. E. et al. S-Nitrosoglutathione reductase deficiency confers improved survival and neurological outcome in experimental cerebral malaria. Infect. Immun. 85 , e00371–17 (2017).

Rubach, M. P. et al. Cerebrospinal fluid pterins, pterin-dependent neurotransmitters, and mortality in pediatric cerebral malaria. J. Infect. Dis. 224 , 1432–1441 (2021).

Grootaert, M. O. J. & Bennett, M. R. Sirtuins in atherosclerosis: guardians of healthspan and therapeutic targets. Nat. Rev. Cardiol. 19 , 668–683 (2022).

Pamplona, A. et al. Heme oxygenase-1 and carbon monoxide suppress the pathogenesis of experimental cerebral malaria. Nat. Med. 13 , 703–710 (2007).

Jeney, V. et al. Control of disease tolerance to malaria by nitric oxide and carbon monoxide. Cell Rep. 8 , 126–136 (2014).

Pena, A. C. & Pamplona, A. Heme oxygenase-1, carbon monoxide, and malaria – the interplay of chemistry and biology. Coord. Chem. Rev. 453 , 214285 (2022).

Campbell, N. K., Fitzgerald, H. K. & Dunne, A. Regulation of inflammation by the antioxidant haem oxygenase 1. Nat. Rev. Immunol. 21 , 411–425 (2021).

Motterlini, R. & Otterbein, L. E. The therapeutic potential of carbon monoxide. Nat. Rev. Drug Discov. 9 , 728–743 (2010).

Pena, A. C. et al. A novel carbon monoxide-releasing molecule fully protects mice from severe malaria. Antimicrob. Agents Chemother. 56 , 1281–1290 (2012).

Cuadrado, A. et al. Therapeutic targeting of the NRF2 and KEAP1 partnership in chronic diseases. Nat. Rev. Drug Discov. 18 , 295–317 (2019).

Mita-Mendoza, N. K. et al. Dimethyl fumarate reduces TNF and Plasmodium falciparum induced brain endothelium activation in vitro. Malar. J. 19 , 376 (2020).

Engelmann, B. & Massberg, S. Thrombosis as an intravascular effector of innate immunity. Nat. Rev. Immunol. 2012 13:1 13 , 34–45 (2012).

Google Scholar  

Stark, K. & Massberg, S. Interplay between inflammation and thrombosis in cardiovascular pathology. Nat. Rev. Cardiol. 18 , 666–682 (2021).

Grau, G. E. et al. Platelet accumulation in brain microvessels in fatal pediatric cerebral malaria. J. Infect. Dis. 187 , 461–466 (2003).

Grau, G. E. & Tacchini-Cottier, F. TNF-induced microvascular pathology: active role for platelets and importance of the LFA-1/ICAM-1 interaction. Eur. Cytokine Netw. 4 , 415–419 (1992).

von Zur Muhlen, C. et al. A contrast agent recognizing activated platelets reveals murine cerebral malaria pathology undetectable by conventional MRI. J. Clin. Invest. 118 , 1198–1207 (2008).

Darling, T. K. et al. Platelet α-granules contribute to organ-specific pathologies in a mouse model of severe malaria. Blood Adv. 4 , 1–8 (2020).

Srivastava, K. et al. Platelet factor 4 mediates inflammation in experimental cerebral malaria. Cell Host Microbe 4 , 179–187 (2008).

van der Heyde, H. C., Gramaglia, I., Sun, G. & Woods, C. Platelet depletion by anti-CD41 (alphaIIb) mAb injection early but not late in the course of disease protects against Plasmodium berghei pathogenesis by altering the levels of pathogenic cytokines. Blood 105 , 1956–1963 (2005).

Sun, G. et al. Inhibition of platelet adherence to brain microvasculature protects against severe Plasmodium berghei malaria. Infect. Immun. 71 , 6553–6561 (2003).

Gramaglia, I. et al. Platelets activate a pathogenic response to blood-stage Plasmodium infection but not a protective immune response. Blood 129 , 1669–1679 (2017).

Chapman, L. M. et al. Platelets present antigen in the context of MHC class I. J. Immunol. 189 , 916–923 (2012).

Gaertner, F. et al. Migrating platelets are mechano-scavengers that collect and bundle bacteria. Cell 171 , 1368–1382.e23 (2017).

Miu, J. et al. Chemokine gene expression during fatal murine cerebral malaria and protection due to CXCR3 deficiency. J. Immunol. 180 , 1217–1230 (2008).

Van Den Ham, K. M., Smith, L. K., Richer, M. J. & Olivier, M. Protein tyrosine phosphatase inhibition prevents experimental cerebral malaria by precluding CXCR3 expression on T cells. Sci. Rep. 7 , 5478 (2017).

Sachais, B. S. et al. Rational design and characterization of platelet factor 4 antagonists for the study of heparin-induced thrombocytopenia. Blood 119 , 5955–5962 (2012).

McMorran, B. J. et al. Platelets kill intraerythrocytic malarial parasites and mediate survival to infection. Science 323 , 797–800 (2009).

Papayannopoulos, V. Neutrophil extracellular traps in immunity and disease. Nat. Rev. Immunol. 18 , 134–147 (2018).

Brinkmann, V. et al. Neutrophil extracellular traps kill bacteria. Science 303 , 1532–1535 (2004).

Moxon, C. A. et al. Parasite histones are toxic to brain endothelium and link blood barrier breakdown and thrombosis in cerebral malaria. Blood Adv. 4 , 2851–2864 (2020).

Rodrigues, D. A. S. et al. CXCR4 and MIF are required for neutrophil extracellular trap release triggered by Plasmodium -infected erythrocytes. PLoS Pathog. 16 , e1008230 (2020).

Knackstedt, S. L. et al. Neutrophil extracellular traps drive inflammatory pathogenesis in malaria. Sci. Immunol. 4 , 336 (2019).

Kho, S. et al. Circulating neutrophil extracellular traps and neutrophil activation are increased in proportion to disease severity in human malaria. J. Infect. Dis. 219 , 1994–2004 (2019).

Vera, I. M. et al. Plasma cell-free DNA predicts pediatric cerebral malaria severity. JCI Insight 5 , e136279 (2020).

Armah, H. B. et al. Cerebrospinal fluid and serum biomarkers of cerebral malaria mortality in Ghanaian children. Malar. J. 6 , 147 (2007).

Feintuch, C. M. et al. Activated neutrophils are associated with pediatric cerebral malaria vasculopathy in Malawian children. MBio 7 , e01300–e01315 (2016).

Dhanesha, N. et al. PKM2 promotes neutrophil activation and cerebral thromboinflammation: therapeutic implications for ischemic stroke. Blood 139 , 1234–1245 (2022).

Wang, A. et al. Glucose metabolism mediates disease tolerance in cerebral malaria. Proc. Natl Acad. Sci. USA 115 , 11042–11047 (2018).

Amulic, B., Moxon, C. A. & Cunnington, A. J. A more granular view of neutrophils in malaria. Trends Parasitol. 36 , 501–503 (2020).

Branco, A. C. C. C., Yoshikawa, F. S. Y., Pietrobon, A. J. & Sato, M. N. Role of histamine in modulating the immune response and inflammation. Mediators Inflamm. 2018 , 9524075 (2018).

Enwonwu, C. O., Afolabi, B. M., Salako, L. O., Idigbe, E. O. & Bashirelah, N. Increased plasma levels of histidine and histamine in falciparum malaria: relevance to severity of infection. J. Neural Transm. (Vienna) 107 , 1273–1287 (2000).

Beghdadi, W. et al. Inhibition of histamine-mediated signaling confers significant protection against severe malaria in mouse models of disease. J. Exp. Med. 205 , 395–408 (2008).

Tiligada, E. & Ennis, M. Histamine pharmacology: from Sir Henry Dale to the 21st century. Br. J. Pharmacol. 177 , 469–489 (2020).

Beghdadi, W. et al. Histamine H(3) receptor-mediated signaling protects mice from cerebral malaria. PLoS One 4 , e6004 (2009).

Teuscher, C. et al. Central histamine H3 receptor signaling negatively regulates susceptibility to autoimmune inflammatory disease of the CNS. Proc. Natl Acad. Sci. USA 104 , 10146–10151 (2007).

Huang, B. et al. Activation of mast cells promote Plasmodium berghei ANKA infection in murine model. Front. Cell. Infect. Microbiol. 9 , 322 (2019).

Porcherie, A. et al. Critical role of the neutrophil-associated high-affinity receptor for IgE in the pathogenesis of experimental cerebral malaria. J. Exp. Med. 208 , 2225–2236 (2011).

Royo, J. et al. Kinetics of monocyte subpopulations during experimental cerebral malaria and its resolution in a model of late chloroquine treatment. Front. Cell. Infect. Microbiol. 12 , 952993 (2022).

Niewold, P. et al. Experimental severe malaria is resolved by targeting newly-identified monocyte subsets using immune-modifying particles combined with artesunate. Commun. Biol. 1 , 227 (2018).

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. 178 , 5779–5788 (2007).

Hermsen, C., van de Wiel, T., Mommers, E., Sauerwein, R. & Eling, W. Depletion of CD4 + or CD8 + T-cells prevents Plasmodium berghei induced cerebral malaria in end-stage disease. Parasitology 114 , 7–12 (1997).

Grau, G. E. et al. L3T4 + T lymphocytes play a major role in the pathogenesis of murine cerebral malaria. J. Immunol. 137 , 2348–2354 (1986).

Belnoue, E. et al. On the pathogenic role of brain-sequestered αβ CD8 + T cells in experimental cerebral malaria. J. Immunol. 169 , 6369–6375 (2002).

Finley, R. W., Mackey, L. J. & Lambert, P. H. Virulent P. berghei malaria: prolonged survival and decreased cerebral pathology in cell-dependent nude mice. J. Immunol. 129 , 2213–2218 (1982).

Wright, D. H. The effect of neonatal thymectomy on the survival of golden hamsters infected with Plasmodium berghei . Br. J. Exp. Pathol. 49 , 379–384 (1968).

Gramaglia, I. et al. Cell- rather than antibody-mediated immunity leads to the development of profound thrombocytopenia during experimental Plasmodium berghei malaria. J. Immunol. 175 , 7699–7707 (2005).

Nitcheu, J. et al. Perforin-dependent brain-infiltrating cytotoxic CD8 + T lymphocytes mediate experimental cerebral malaria pathogenesis. J. Immunol. 170 , 2221–2228 (2003).

Wright, D. H., Masembe, R. M. & Bazira, E. R. The effect of antithymocyte serum on golden hamsters and rats infected with Plasmodium berghei . Br. J. Exp. Pathol. 52 , 465–477 (1971).

Riggle, B. A. et al. CD8 + T cells target cerebrovasculature in children with cerebral malaria. J. Clin. Invest. 130 , 1128–1138 (2020).

Kurup, S. P., Butler, N. S. & Harty, J. T. T cell-mediated immunity to malaria. Nat. Rev. Immunol. 19 , 457–471 (2019).

Campanella, G. S. V. et al. Chemokine receptor CXCR3 and its ligands CXCL9 and CXCL10 are required for the development of murine cerebral malaria. Proc. Natl Acad. Sci. USA 105 , 4814–4819 (2008).

Swanson, P. A. II et al. CD8 + T cells induce fatal brainstem pathology during cerebral malaria via luminal antigen-specific engagement of brain vasculature. PLoS Pathog. 12 , e1006022 (2016).

Shaw, T. N. et al. Perivascular arrest of CD8 + T cells is a signature of experimental cerebral malaria. PLoS Pathog. 11 , e1005210 (2015).

Van Braeckel-Budimir, N. et al. A T cell receptor locus harbors a malaria-specific immune response gene. Immunity 47 , 835–847.e4 (2017).

Heide, J., Vaughan, K. C., Sette, A., Jacobs, T. & Schulze Zur Wiesch, J. Comprehensive review of human Plasmodium falciparum -specific CD8 + T cell epitopes. Front. Immunol. 10 , 397 (2019).

Junqueira, C. et al. Cytotoxic CD8 + T cells recognize and kill Plasmodium vivax –infected reticulocytes. Nat. Med. 24 , 1330–1336 (2018).

Pober, J. S., Merola, J., Liu, R. & Manes, T. D. Antigen presentation by vascular cells. Front. Immunol. 8 , 1907 (2017).

Howland, S. W., Poh, C. M. & Rénia, L. Activated brain endothelial cells cross-present malaria antigen. PLoS Pathog. 11 , e1004963 (2015).

Howland, S. W. et al. Brain microvessel cross-presentation is a hallmark of experimental cerebral malaria. EMBO Mol. Med. 5 , 984–999 (2013).

Fain, C. et al. Discrete class I molecules on brain endothelium differentially regulate neuropathology in experimental cerebral malaria. Brain https://doi.org/10.1093/brain/awad319 (2023).

Geppert, T. D. & Lipsky, P. E. Antigen presentation by interferon-gamma-treated endothelial cells and fibroblasts: differential ability to function as antigen-presenting cells despite comparable Ia expression. J. Immunol. 135 , 3750–3762 (1985).

Murata, S., Takahama, Y., Kasahara, M. & Tanaka, K. The immunoproteasome and thymoproteasome: functions, evolution and human disease. Nat. Immunol. 19 , 923–931 (2018).

Howland, S. W., Ng, G. X. P., Chia, S. K. & Rénia, L. Investigating proteasome inhibitors as potential adjunct therapies for experimental cerebral malaria. Parasite Immunol. 37 , 599–604 (2015).

Basler, M. et al. The immunoproteasome: a novel drug target for autoimmune diseases. Clin. Exp. Rheumatol. 33 , 74–79 (2015).

Baeza Garcia, A. et al. Suppression of Plasmodium MIF-CD74 signaling protects against severe malaria. FASEB J. 35 , e21997 (2021).

Andrews, S. P. & Cox, R. J. Small molecule CXCR3 antagonists. J. Med. Chem. 59 , 2894–2917 (2016).

Nie, C. Q. et al. IP-10-mediated T cell homing promotes cerebral inflammation over splenic immunity to malaria infection. PLoS Pathog. 5 , e1000369 (2009).

Wilson, N. O. et al. CXCL4 and CXCL10 predict risk of fatal cerebral malaria. Dis. Markers 30 , 39–49 (2011).

Lebwohl, M. et al. A novel targeted T-cell modulator, efalizumab, for plaque psoriasis. N. Engl. J. Med. 349 , 2004–2013 (2003).

Polman, C. H. et al. A randomized, placebo-controlled trial of natalizumab for relapsing multiple sclerosis. N. Engl. J. Med. 354 , 899–910 (2006).

Fauconnier, M. et al. Protein kinase C-theta is required for development of experimental cerebral malaria. Am. J. Pathol. 178 , 212–221 (2011).

Monks, C. R. F., Freiberg, B. A., Kupfer, H., Sciaky, N. & Kupfer, A. Three-dimensional segregation of supramolecular activation clusters in T cells. Nature 395 , 82–86 (1998).

Monks, C. R. F., Kupfer, H., Tamir, I., Barlow, A. & Kupfer, A. Selective modulation of protein kinase C-Θ during T-cell activation. Nature 385 , 83–86 (1997).

Potter, S. et al. Perforin mediated apoptosis of cerebral microvascular endothelial cells during experimental cerebral malaria. Int. J. Parasitol. 36 , 485–496 (2006).

Haque, A. et al. Granzyme B expression by CD8 + T cells is required for the development of experimental cerebral malaria. J. Immunol. 186 , 6148–6156 (2011).

Huggins, M. A. et al. Perforin expression by CD8 T cells is sufficient to cause fatal brain edema during experimental cerebral malaria. Infect. Immun. 85 , e00985–16 (2017).

Hermsen, C. C. et al. Circulating concentrations of soluble granzyme A and B increase during natural and experimental Plasmodium falciparum infections. Clin. Exp. Immunol. 132 , 467–472 (2003).

Kaminski, L.-C. et al. Cytotoxic T cell-derived granzyme B is increased in severe Plasmodium falciparum malaria. Front. Immunol. 10 , 2917 (2019).

Barrera, V. et al. Comparison of CD8 + T cell accumulation in the brain during human and murine cerebral malaria. Front. Immunol. 10 , 1747 (2019).

Skaro, A. I. et al. CD8 + T cells mediate aortic allograft vasculopathy by direct killing and an interferon-γ-dependent indirect pathway. Cardiovasc. Res. 65 , 283–291 (2005).

Gross, C. C. et al. CD8 + T cell-mediated endotheliopathy is a targetable mechanism of neuro-inflammation in Susac syndrome. Nat. Commun. 10 , 5779 (2019).

Bongfen, S. E. et al. An N -ethyl- N -nitrosourea (ENU)-induced dominant negative mutation in the JAK3 kinase protects against cerebral malaria. PLoS One 7 , e31012 (2012).

Thomis, D. C., Gurniak, C. B., Tivol, E., Sharpe, A. H. & Berg, L. J. Defects in B lymphocyte maturation and T lymphocyte activation in mice lacking Jak3. Science 270 , 794–797 (1995).

Torre, S. et al. THEMIS is required for pathogenesis of cerebral malaria and protection against pulmonary tuberculosis. Infect. Immun. 83 , 759–768 (2015).

Kennedy, J. M. et al. ZBTB7B (ThPOK) is required for pathogenesis of cerebral malaria and protection against pulmonary tuberculosis. Infect. Immun. 88 , e00845–19 (2020).

Kennedy, J. M. et al. CCDC88B is a novel regulator of maturation and effector functions of T cells during pathological inflammation. J. Exp. Med. 211 , 2519–2535 (2014).

Lesourne, R. et al. Themis, a T cell–specific protein important for late thymocyte development. Nat. Immunol. 10 , 840–847 (2009).

He, X. et al. The zinc finger transcription factor Th-POK regulates CD4 versus CD8 T-cell lineage commitment. Nature 433 , 826–833 (2005).

Chakravarty, S. et al. CD8 + T lymphocytes protective against malaria liver stages are primed in skin-draining lymph nodes. Nat. Med. 13 , 1035–1041 (2007).

Radtke, A. J. et al. Lymph-node resident CD8α + dendritic cells capture antigens from migratory malaria sporozoites and induce CD8 + T cell responses. PLoS Pathog. 11 , e1004637 (2015).

Lundie, R. J. et al. Blood-stage Plasmodium infection induces CD8 + T lymphocytes to parasite-expressed antigens, largely regulated by CD8α + dendritic cells. Proc. Natl Acad. Sci. USA 105 , 14509–14514 (2008).

Jung, S. et al. In vivo depletion of CD11c + dendritic cells abrogates priming of CD8 + T cells by exogenous cell-associated antigens. Immunity 17 , 211–220 (2002).

Salem, S., Salem, D. & Gros, P. Role of IRF8 in immune cells functions, protection against infections, and susceptibility to inflammatory diseases. Hum. Genet. 139 , 707–721 (2020).

Berghout, J. et al. Irf8-regulated genomic responses drive pathological inflammation during cerebral malaria. PLoS Pathog. 9 , e1003491 (2013).

Kuehlwein, J. M. et al. Protection of Batf3-deficient mice from experimental cerebral malaria correlates with impaired cytotoxic T-cell responses and immune regulation. Immunology 159 , 193–204 (2020).

Piva, L. et al. Cutting edge: Clec9A + dendritic cells mediate the development of experimental cerebral malaria. J. Immunol. 189 , 1128–1132 (2012).

Raulf, M. K. et al. The C-type lectin receptor CLEC12A recognizes plasmodial hemozoin and contributes to cerebral malaria development. Cell Rep. 28 , 30–38.e5 (2019).

Wiese, L., Hempel, C., Penkowa, M., Kirkby, N. & Kurtzhals, J. A. L. Recombinant human erythropoietin increases survival and reduces neuronal apoptosis in a murine model of cerebral malaria. Malar. J. 7 , 3 (2008).

Kaiser, K. et al. Recombinant human erythropoietin prevents the death of mice during cerebral malaria. J. Infect. Dis. 193 , 987–995 (2006).

Casals-Pascual, C. et al. High levels of erythropoietin are associated with protection against neurological sequelae in African children with cerebral malaria. Proc. Natl Acad. Sci. USA 105 , 2634–2639 (2008).

Wei, X. et al. Erythropoietin protects against murine cerebral malaria through actions on host cellular immunity. Infect. Immun. 82 , 165–173 (2014).

Guermonprez, P. et al. Inflammatory Flt3l is essential to mobilize dendritic cells and for T cell responses during Plasmodium infection. Nat. Med. 19 , 730–738 (2013).

Mejia, P. et al. A single rapamycin dose protects against late-stage experimental cerebral malaria via modulation of host immunity, endothelial activation and parasite sequestration. Malar. J. 16 , 455 (2017).

Burrack, K. S. et al. Interleukin-15 complex treatment protects mice from cerebral malaria by inducing interleukin-10-producing natural killer cells. Immunity 48 , 760–772.e4 (2018).

Wang, J. et al. PDL1 fusion protein protects against experimental cerebral malaria via repressing over-reactive CD8 + T cell responses. Front. Immunol. 9 , 3157 (2019).

Shen, Y. et al. The immunomodulatory effect of microglia on ECM neuroinflammation via the PD-1/PD-L1 pathway. CNS Neurosci. Ther. 28 , 46–63 (2022).

Gordon, E. B. et al. Targeting glutamine metabolism rescues mice from late-stage cerebral malaria. Proc. Natl Acad. Sci. USA 112 , 13075–13080 (2015).

Riggle, B. A. et al. MRI demonstrates glutamine antagonist-mediated reversal of cerebral malaria pathology in mice. Proc. Natl Acad. Sci. USA 115 , E12024–E12033 (2018).

Carr, E. L. et al. Glutamine uptake and metabolism are coordinately regulated by ERK/MAPK during T lymphocyte activation. J. Immunol. 185 , 1037–1044 (2010).

Chapman, N. M., Boothby, M. R. & Chi, H. Metabolic coordination of T cell quiescence and activation. Nat. Rev. Immunol. 20 , 55–70 (2019).

Leone, R. D. et al. Glutamine blockade induces divergent metabolic programs to overcome tumor immune evasion. Science 366 , 1013–1021 (2019).

Edington, G. M. Pathology of malaria in West Africa. BMJ 1 , 715–718 (1967).

Mejia, P. et al. Adipose tissue parasite sequestration drives leptin production in mice and correlates with human cerebral malaria. Sci. Adv. 7 , eabe2484 (2021).

Mejia, P. et al. Dietary restriction protects against experimental cerebral malaria via leptin modulation and T-cell mTORC1 suppression. Nat. Commun. 6 , 6050 (2015).

Robert, V. et al. Malaria and obesity: obese mice are resistant to cerebral malaria. Malar. J. 7 , 81 (2008).

Gordon, E. B. et al. Inhibiting the mammalian target of rapamycin blocks the development of experimental cerebral malaria. mBio 6 , e00725 (2015).

Bantug, G. R., Galluzzi, L., Kroemer, G. & Hess, C. The spectrum of T cell metabolism in health and disease. Nat. Rev. Immunol. 18 , 19–34 (2018).

O’Sullivan, D. & Pearce, E. L. Targeting T cell metabolism for therapy. Trends Immunol. 36 , 71–80 (2015).

Chapman, N. M. & Chi, H. Metabolic adaptation of lymphocytes in immunity and disease. Immunity 55 , 14–30 (2022).

Schluesener, H. J., Kremsner, P. G. & Meyermann, R. Widespread expression of MRP8 and MRP14 in human cerebral malaria by microglial cells. Acta Neuropathologica 96 , 575–580 (1998).

Guha, S. K. et al. Single episode of mild murine malaria induces neuroinflammation, alters microglial profile, impairs adult neurogenesis, and causes deficits in social and anxiety-like behavior. Brain Behav. Immun. 42 , 123–137 (2014).

Medana, I. M., Hunt, N. H. & Chan-Ling, T. Early activation of microglia in the pathogenesis of fatal murine cerebral malaria. Glia 19 , 91–103 (1997).

Capuccini, B. et al. Transcriptomic profiling of microglia reveals signatures of cell activation and immune response, during experimental cerebral malaria. Sci. Rep. 6 , 39258 (2016).

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 11 , 913 (2018).

Pais, T. F. & Chatterjee, S. Brain macrophage activation in murine cerebral malaria precedes accumulation of leukocytes and CD8 + T cell proliferation. J. Neuroimmunol. 163 , 73–83 (2005).

Goddery, E. N. et al. Microglia and perivascular macrophages act as antigen presenting cells to promote CD8 T cell infiltration of the brain. Front. Immunol. 12 , 726421 (2021).

Goddery, E. et al. Antigen presentation by CNS-resident microglia and macrophages regulates CD8 T cell infiltration of the brain during central nervous system (CNS) infection. J. Immunol. 206 (Suppl. 1), 11.18 https://doi.org/10.4049/jimmunol.206.Supp.11.18 (2021).

Velagapudi, R., Kosoko, A. M. & Olajide, O. A. Induction of neuroinflammation and neurotoxicity by synthetic hemozoin. Cell. Mol. Neurobiol. 39 , 1187–1200 (2019).

Strangward, P. et al. Targeting the IL33-NLRP3 axis improves therapy for experimental cerebral malaria. Proc. Natl Acad. Sci. USA 115 , 7404–7409 (2018).

Reimer, T. et al. Experimental cerebral malaria progresses independently of the Nlrp3 inflammasome. Eur. J. Immunol. 40 , 764–769 (2010).

Reverchon, F. et al. IL-33 receptor ST2 regulates the cognitive impairments associated with experimental cerebral malaria. PLoS Pathog. 13 , e1006322 (2017).

Besnard, A. G. et al. 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. 11 , e1004607 (2015).

Palomo, J. et al. Critical role of IL-33 receptor ST2 in experimental cerebral malaria development. Eur. J. Immunol. 45 , 1354–1365 (2015).

Shibui, A. et al. IL-25, IL-33 and TSLP receptor are not critical for development of experimental murine malaria. Biochem. Biophys. Rep. 5 , 191–195 (2015).

PubMed   PubMed Central   Google Scholar  

Ayimba, E. et al. Proinflammatory and regulatory cytokines and chemokines in infants with uncomplicated and severe Plasmodium falciparum malaria. Clin. Exp. Immunol. 166 , 218–226 (2011).

Fernander, E. M. et al. Elevated plasma soluble ST2 levels are associated with neuronal injury and neurocognitive impairment in children with cerebral malaria. Pathog. Immun. 7 , 60–80 (2022).

Belum, G. R., Belum, V. R., Chaitanya Arudra, S. K. & Reddy, B. S. N. The Jarisch-Herxheimer reaction: revisited. Travel Med. Infect. Dis. 11 , 231–237 (2013).

Kofoed, P. E. Jarisch-Herxheimer reaction in falciparum malaria? Br. Med. J. (Clin. Res. Ed.) 289 , 161 (1984).

Jerusalem, C., Polder, T., Kubat, K., Wijers-Rouw, M. & Trinh, P. Brain edema in cerebral malaria: a comparative clinical and experimental, ultrastructural and histochemical study. In Recent Progress in the Study and Therapy of Brain Edema 127–135 (Springer US, 1984).

Seydel, K. B. et al. Brain swelling and death in children with cerebral malaria. N. Engl. J. Med. 372 , 1126–1137 (2015).

Mohanty, S. et al. Evidence of brain alterations in noncerebral falciparum malaria. Clin. Infect. Dis. 75 , 11–18 (2022).

Kampondeni, S. et al. Amount of brain edema correlates with neurologic recovery in pediatric cerebral malaria. Pediatr. Infect. Dis. J. 39 , 277–282 (2020).

Moghaddam, S. M. et al. Diffusion-weighted MR imaging in a prospective cohort of children with cerebral malaria offers insights into pathophysiology and prognosis. AJNR Am. J. Neuroradiol. 40 , 1575–1580 (2019).

Maude, R. J. et al. Magnetic resonance imaging of the brain in adults with severe falciparum malaria. Malar. J. 13 , 177 (2014).

Mohanty, S. et al. Magnetic resonance imaging of cerebral malaria patients reveals distinct pathogenetic processes in different parts of the brain. MSphere 2 , e00193–17 (2017).

Jha, R. M. et al. Emerging therapeutic targets for cerebral edema. Expert Opin. Ther. Targets 25 , 917–938 (2021).

Mestre, H., Mori, Y. & Nedergaard, M. The brain’s glymphatic system: current controversies. Trends Neurosci. 43 , 458–466 (2020).

Papadopoulos, M. C. & Verkman, A. S. Aquaporin-4 and brain edema. Pediatr. Nephrol. 22 , 778–784 (2007).

Ampawong, S. et al. Quantitation of brain edema and localisation of aquaporin 4 expression in relation to susceptibility to experimental cerebral malaria. Int. J. Clin. Exp. Pathol. 4 , 566–574 (2011).

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 110 , 1035–1040 (2013).

de Souza, J. B., Hafalla, J. C. R., Riley, E. M. & Couper, K. N. Cerebral malaria: why experimental murine models are required to understand the pathogenesis of disease. Parasitology 137 , 755–772 (2010).

Shikani, H. J. et al. Cerebral malaria: we have come a long way. Am. J. Pathol. 181 , 1484–1492 (2012).

Craig, A. G. et al. The role of animal models for research on severe malaria. PLoS Pathog. 8 , e1002401 (2012).

White, N. J., Turner, G. D. H., Medana, I. M., Dondorp, A. M. & Day, N. P. J. The murine cerebral malaria phenomenon. Trends Parasitol. 26 , 11–15 (2010).

Georgiadou, A. et al. Comparative transcriptomic analysis reveals translationally relevant processes in mouse models of malaria. eLife 11 , e70763 (2022).

Hajal, C. et al. Engineered human blood-brain barrier microfluidic model for vascular permeability analyses. Nat. Protocols 17 , 95–128 (2022).

Cakir, B. et al. Engineering of human brain organoids with a functional vascular-like system. Nat. Methods 16 , 1169–1175 (2019).

Lippmann, E. S. et al. Derivation of blood-brain barrier endothelial cells from human pluripotent stem cells. Nat. Biotechnol. 30 , 783–791 (2012).

Krasemann, S. et al. The blood-brain barrier is dysregulated in COVID-19 and serves as a CNS entry route for SARS-CoV-2. Stem Cell Rep. 17 , 307–320 (2022).

Longo, S. K., Guo, M. G., Ji, A. L. & Khavari, P. A. Integrating single-cell and spatial transcriptomics to elucidate intercellular tissue dynamics. Nat. Rev. Genet. 22 , 627–644 (2021).

Elmentaite, R., Domínguez Conde, C., Yang, L. & Teichmann, S. A. Single-cell atlases: shared and tissue-specific cell types across human organs. Nat. Rev. Genet. 23 , 395–410 (2022).

Kim, J., Koo, B. K. & Knoblich, J. A. Human organoids: model systems for human biology and medicine. Nat. Rev. Mol. Cell Biol. 21 , 571–584 (2020).

Venugopal, K., Hentzschel, F., Valkiūnas, G. & Marti, M. Plasmodium asexual growth and sexual development in the haematopoietic niche of the host. Nat. Rev. Microbiol. 18 , 177–189 (2020).

Schuepbach, R. A., Madon, J., Ender, M., Galli, P. & Riewald, M. Protease-activated receptor-1 cleaved at R46 mediates cytoprotective effects. J. Thromb. Haemost. 10 , 1675–1684 (2012).

Sinha, R. K. et al. PAR1 biased signaling is required for activated protein C in vivo benefits in sepsis and stroke. Blood 131 , 1163–1171 (2018).

Moxon, C. A. et al. Loss of endothelial protein C receptors links coagulation and inflammation to parasite sequestration in cerebral malaria in African children. Blood 122 , 842–851 (2013).

Mohan Rao, L. V., Esmon, C. T. & Pendurthi, U. R. Endothelial cell protein C receptor: a multiliganded and multifunctional receptor. Blood 124 , 1553–1562 (2014).

Griffin, J. H., Zlokovic, B. V. & Mosnier, L. O. Activated protein C, protease activated receptor 1, and neuroprotection. Blood 132 , 159–169 (2018).

O’Regan, N. et al. Hemostatic and protein C pathway dysfunction in the pathogenesis of experimental cerebral malaria. Haematologica 107 , 1950–1954 (2022).

Gillrie, M. R. et al. Plasmodium falciparum histones induce endothelial proinflammatory response and barrier dysfunction. Am. J. Pathol. 180 , 1028–1039 (2012).

Petersen, J. E. V. et al. Protein C system defects inflicted by the malaria parasite protein PfEMP1 can be overcome by a soluble EPCR variant. Thromb. Haemost. 114 , 1038–1048 (2015).

Storm, J., Wu, Y., Davies, J., Moxon, C. A. & Craig, A. G. Testing the effect of PAR1 inhibitors on Plasmodium falciparum -induced loss of endothelial cell barrier function. Wellcome Open Res. 5 , 34 (2020).

Download references

Acknowledgements

A.H. was supported by an MD/PhD stipend (TI 07.001_Hadjilaou) and is currently supported by a Clinical Leave stipend of the Bundesministerium für Bildung und Forschung/Deutsches Zentrum für Infektionsforschung (TI 07.002_Hadjilaou).

Author information

These authors contributed equally: Manuel A. Friese, Thomas Jacobs.

Authors and Affiliations

Protozoen Immunologie, Bernhard-Nocht-Institut für Tropenmedizin (BNITM), Hamburg, Germany

Alexandros Hadjilaou, Johannes Brandi, Mathias Riehn & Thomas Jacobs

Institut für Neuroimmunologie und Multiple Sklerose, Universitätsklinikum Hamburg-Eppendorf, Hamburg, Germany

Alexandros Hadjilaou & Manuel A. Friese

You can also search for this author in PubMed   Google Scholar

Contributions

A.H. outlined the content of the manuscript, wrote the first draft of the manuscript and figures and researched data for the article. All authors contributed substantially to discussions of the article content, reviewed and/or edited the manuscript and figures before submission.

Corresponding author

Correspondence to Alexandros Hadjilaou .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Peer review

Peer review information.

Nature Reviews Neurology thanks Terrie Taylor and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Potent, low-molecular-weight molecules produced during immune responses that stimulate inflammation, attract white blood cells and modulate immune responses to infection or disease.

The clearance of damaged cells by phagocytosis.

A paravascular system thought to be responsible for waste clearance in the central nervous system, similar to the lymphatic system in other organs.

Small, round, basophilic, nuclear remnants of DNA within mature erythrocytes, which are typically removed in the spleen through a process called ‘pitting’.

Proteasomes in the cytoplasm of immune cells, especially antigen-presenting ones, formation of which is induced by pro-inflammatory cytokines or oxidative stress.

A collection of changes in the retina that can occur during Plasmodium infection.

Cellular metabolic changes that address increased bioenergetic and biosynthetic demands, such as those needed in the transition from a quiescent to an activated phenotype.

The process by which circulating, soluble proteins (opsonins) bind to a cell membrane, thereby marking those cells for phagocytosis.

A chemical modification that enables anchoring of a protein to the cell membrane.

Immature erythrocytes characterized by a reticular network of ribosomal RNA.

The increased cytoadhesion of erythrocytes infected with mature, asexual Plasmodium falciparum causes these cells to stick to uninfected erythrocytes, forming flower-like cell clusters.

The cytoadherence of Plasmodium -infected erythrocytes to the vascular endothelium, a feature predominantly seen in Plasmodium falciparum infections.

A post-translational peptide modification in which a nitrosyl group is added to the sulfur atom of a cysteine residue.

3-Hydroxy-3-methylglutaryl-CoA reductase inhibitors approved for the treatment of hypercholesterolaemia.

A measure of the frequency of malaria spread in a specific area, which is dictated by mosquito population density, local climate conditions and the level of human immunity (which shapes individual exposure to malaria antigens and the development of protective immunity).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Cite this article.

Hadjilaou, A., Brandi, J., Riehn, M. et al. Pathogenetic mechanisms and treatment targets in cerebral malaria. Nat Rev Neurol 19 , 688–709 (2023). https://doi.org/10.1038/s41582-023-00881-4

Download citation

Accepted : 11 September 2023

Published : 19 October 2023

Issue Date : November 2023

DOI : https://doi.org/10.1038/s41582-023-00881-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

presentation of cerebral malaria

Advertisement

Advertisement

Cerebral malaria – clinical manifestations and pathogenesis

  • Review Article
  • Published: 08 January 2016
  • Volume 31 , pages 225–237, ( 2016 )

Cite this article

presentation of cerebral malaria

  • Rachna Hora 1 ,
  • Payal Kapoor 1 ,
  • Kirandeep Kaur Thind 1 &
  • Prakash Chandra Mishra 2  

2610 Accesses

44 Citations

2 Altmetric

Explore all metrics

One of the most common central nervous system diseases in tropical countries is cerebral malaria (CM). Malaria is a common protozoan infection that is responsible for enormous worldwide mortality and economic burden on the society. Episodes of Plasmodium falciparum (Pf) caused CM may be lethal, while survivors are likely to suffer from persistent debilitating neurological deficits, especially common in children. In this review article, we have summarized the various symptoms and manifestations of CM in children and adults, and entailed the molecular basis of the disease. We have also emphasized how pathogenesis of the disease is effected by the parasite and host responses including blood brain barrier (BBB) disruption, endothelial cell activation and apoptosis, nitric oxide bioavailability, platelet activation and apoptosis, and neuroinflammation. Based on a few recent studies carried out in experimental mouse malaria models, we propose a basis for the neurological deficits and sequelae observed in human cerebral malaria, and summarize how existing drugs may improve prognosis in affected individuals.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

Similar content being viewed by others

presentation of cerebral malaria

Cerebral Malaria: Players in the Pathogenic Mechanism and Treatment Strategies

presentation of cerebral malaria

Pathogenetic mechanisms and treatment targets in cerebral malaria

presentation of cerebral malaria

Pathophysiology and neurologic sequelae of cerebral malaria

Anstey NM et al (1996) Nitric oxide in Tanzanian children with malaria: inverse relationship between malaria severity and nitric oxide production/nitric oxide synthase type 2 expression. J Exp Med 184:557–567

Article   CAS   PubMed   Google Scholar  

Armah H, Wiredu EK, Dodoo AK, Adjei AA, Tettey Y, Gyasi R (2005) Cytokines and adhesion molecules expression in the brain in human cerebral malaria. Int J Environ Res Public Health 2:123–131

Article   CAS   PubMed   PubMed Central   Google Scholar  

Bachmann J et al (2014) Affinity proteomics reveals elevated muscle proteins in plasma of children with cerebral malaria. PLoS Pathog 10:e1004038

Article   PubMed   PubMed Central   CAS   Google Scholar  

Bangirana P et al (2014) Severe malarial anemia is associated with long-term neurocognitive impairment. Clin Infect Dis :ciu293

Barbier M et al (2011) Platelets alter gene expression profile in human brain endothelial cells in an in vitro model of cerebral malaria. PLoS One 6:e19651

Barfod A, Persson T, Lindh J (2009) In vitro selection of RNA aptamers against a conserved region of the Plasmodium falciparum erythrocyte membrane protein 1. Parasitol Res 105:1557–1566

Article   PubMed   PubMed Central   Google Scholar  

Beare NA, Southern C, Chalira C, Taylor TE, Molyneux ME, Harding SP (2004) Prognostic significance and course of retinopathy in children with severe malaria. Arch Ophthalmol 122:1141–1147

Article   PubMed   Google Scholar  

Beare NA, Taylor TE, Harding SP, Lewallen S, Molyneux ME (2006) Malarial retinopathy: a newly established diagnostic sign in severe malaria. Am J Trop Med Hyg 75:790–797

PubMed   PubMed Central   Google Scholar  

Beare NA, Harding SP, Taylor TE, Lewallen S, Molyneux ME (2009) Perfusion abnormalities in children with cerebral malaria and malarial retinopathy. J Infect Dis 199:263–271

Beeson JG et al (2000) Adhesion of Plasmodium falciparum-infected erythrocytes to hyaluronic acid in placental malaria. Nat Med 6:86–90

Berendt A, Simmons D, Tansey J, Newbold C, Marsh K (1989) Intercellular adhesion molecule-1 is an endothelial cell adhesion receptor for Plasmodium falciparum. Nature 341:57–59

Boivin MJ, Bangirana P, Byarugaba J, Opoka RO, Idro R, Jurek AM, John CC (2007) Cognitive impairment after cerebral malaria in children: a prospective study. Pediatrics 119:e360–e366

Boivin MJ, Vokhiwa M, Sikorskii A, Magen JG, Beare N (2014) Cerebral malaria retinopathy predictors of persisting neurocognitive outcomes in Malawian children. Pediatr Infect Dis J

Bridges DJ et al (2010) Rapid activation of endothelial cells enables Plasmodium falciparum adhesion to platelet-decorated von Willebrand factor strings. Blood 115:1472–1474

Brown GC (2001) Regulation of mitochondrial respiration by nitric oxide inhibition of cytochrome c oxidase. Biochim Biophys Acta (BBA)-Bioenergetics 1504:46–57

Article   CAS   Google Scholar  

Brown H, Rogerson S, Taylor T, Tembo M, Mwenechanya J, Molyneux M, Turner G (2001) Blood–brain barrier function in cerebral malaria in Malawian children. Am J Trop Med Hyg 64:207–213

CAS   PubMed   Google Scholar  

Buchanan JE, Phillis JW (1993) The role of nitric oxide in the regulation of cerebral blood flow. Brain Res 610:248–255

Cabrales P, Zanini GM, Meays D, Frangos JA, Carvalho LJ (2011) Nitric oxide protection against murine cerebral malaria is associated with improved cerebral microcirculatory physiology. J Infect Dis :jir058

Carter J, Murira G, Ross A, Mung’ala-Odera V, Newton C (2003) Speech and language sequelae of severe malaria in Kenyan children. Brain Inj 17:217–224

Carter JA, Mung’ala-Odera V, Neville BG, Murira G, Mturi N, Musumba C, Newton CR (2005) Persistent neurocognitive impairments associated with severe falciparum malaria in Kenyan children. J Neurol Neurosurg Psychiatry 76:476–481. doi: 10.1136/jnnp.2004.043893

Carter JA, Lees JA, Gona JK, Murira G, Rimba K, Neville BG, Newton CR (2006) Severe falciparum malaria and acquired childhood language disorder. Dev Med Child Neurol 48:51–57

Charunwatthana P, Faiz MA, Ruangveerayut R, Maude R, Rahman MR, Roberts LJ (2009) N-acetylcysteine as adjunctive treatment in severe malaria: a randomized double blinded placebo controlled clinical trial. Crit Care Med 37:516

Combes V et al (2004) Circulating endothelial microparticles in malawian children with severe falciparum malaria complicated with coma. JAMA 291:2542–2544

Combes V et al (2005) ABCA1 gene deletion protects against cerebral malaria: potential pathogenic role of microparticles in neuropathology. Am J Pathol 166:295–302

Conroy AL et al (2010) Endothelium-based biomarkers are associated with cerebral malaria in Malawian children: a retrospective case–control study. PLoS One 5:e15291

Cooke BM, Berendt AR, Craig AG, MacGregor J, Newbold CI, Nash GB (1994) Rolling and stationary cytoadhesion of red blood cells parasitized by Plasmodium falciparum: separate roles for ICAM-1, CD36 and thrombospondin. Br J Haematol 87:162–170

Cramer JP et al (2005) Age dependent effect of plasma nitric oxide on parasite density in Ghanaian children with severe malaria. Tropical Med Int Health 10:672–680

Crawley J, Smith S, Muthinji P, Marsh K, Kirkham F (2001) Electroencephalographic and clinical features of cerebral malaria. Arch Dis Child 84:247–253

de Miranda AS et al (2015) Evidence for the contribution of adult neurogenesis and hippocampal cell death in experimental cerebral malaria cognitive outcome. Neuroscience 284:920–933

Article   PubMed   CAS   Google Scholar  

Dondorp AM, Kager PA, Vreeken J, White NJ (2000) Abnormal blood flow and red blood cell deformability in severe malaria. Parasitol Today 16:228–232

Duraisingh MT et al (2005) Heterochromatin silencing and locus repositioning linked to regulation of virulence genes in plasmodium falciparum. Cell 121:13–24

El-Assaad F, Wheway J, Hunt NH, Grau GER, Combes V (2014) Production, fate and pathogenicity of plasma microparticles in murine cerebral malaria. PLoS Pathog 10:e1003839

English M, Waruiru C, Lightowler C, Murphy S, Kirigha G, Marsh K (1996) Hyponatraemia and dehydration in severe malaria. Arch Dis Child 74:201–205

English M, Sauerwein R, Waruiru C, Mosobo M, Obiero J, Lowe B, Marsh K (1997) Acidosis in severe childhood malaria. QJM 90:263–270

English M, Wale S, Binns G, Mwangi I, Sauerwein H, Marsh K (1998) Hypoglycaemia on and after admission in Kenyan children with severe malaria. QJM 91:191–197

Epp C, Li F, Howitt CA, Chookajorn T, Deitsch KW (2009) Chromatin associated sense and antisense noncoding RNAs are transcribed from the var gene family of virulence genes of the malaria parasite Plasmodium falciparum. RNA 15:116–127

Favre N et al (1999) Role of ICAM-1 (CD54) in the development of murine cerebral malaria. Microbes Infect 1:961–968

Fiedler U et al (2006) Angiopoietin-2 sensitizes endothelial cells to TNF-alpha and has a crucial role in the induction of inflammation. Nat Med 12:235–239

Freitas-Junior LH et al (2000) Frequent ectopic recombination of virulence factor genes in telomeric chromosome clusters of P. falciparum. Nature 407:1018–1022

Freitas-Junior LH et al (2005) Telomeric heterochromatin propagation and histone acetylation control mutually exclusive expression of antigenic variation genes in malaria parasites. Cell 121:25–36

Fried M, Duffy PE (1996) Adherence of Plasmodium falciparum to chondroitin sulfate A in the human placenta. Science 272:1502–1504

Furlan M, Robles R, Lamie B (1996) Partial purification and characterization of a protease from human plasma cleaving von Willebrand factor to fragments produced by in vivo proteolysis. Blood 87:4223–4234

Gallo EF, Iadecola C (2011) Neuronal nitric oxide contributes to neuroplasticity-associated protein expression through cGMP, protein kinase G, and extracellular signal-regulated kinase. J Neurosci 31:6947–6955

Gramaglia I et al (2006) Low nitric oxide bioavailability contributes to the genesis of experimental cerebral malaria. Nat Med 12:1417–1422

Grau GE, Piguet PF, Gretener D, Vesin C, Lambert PH (1988) Immunopathology of thrombocytopenia in experimental malaria. Immunology 65:501

CAS   PubMed   PubMed Central   Google Scholar  

Grau GE, Tacchini-Cottier F, Vesin C, Milon G, Lou JN, Piguet PF, Juillard P (1992) TNF-induced microvascular pathology: active role for platelets and importance of the LFA-1/ICAM-1 interaction. Eur Cytokine Netw 4:415–419

Google Scholar  

Grau GE et al (2003) Platelet accumulation in brain microvessels in fatal pediatric cerebral malaria. J Infect Dis 187:461–466

Guha SK et al (2014) Single episode of mild murine malaria induces neuroinflammation, alters microglial profile, impairs adult neurogenesis, and causes deficits in social and anxiety-like behavior. Brain Behav Immun 42:123–137

Gyan B et al (2009) Cerebral malaria is associated with low levels of circulating endothelial progenitor cells in African children. AmJTrop Med Hyg 80:541–546

Havlik I et al (2005) Curdlan sulphate in human severe/cerebral Plasmodium falciparum malaria. Trans R Soc Trop Med Hyg 99:333–340

Hemmer CJ, Kern P, Holst FG, Nawroth PP, Dietrich M (1991) Neither heparin nor acetylsalicylic acid influence the clinical course in human Plasmodium falciparum malaria: a prospective randomized study. Am J Trop Med Hyg 45:608–612

Hempel C, Hyttel P, Staalso T, Nyengaard JR, Kurtzhals JAL (2012) Erythropoietin treatment alleviates ultrastructural myelin changes induced by murine cerebral malaria. Malar J 11:216

Hiller NL, Bhattacharjee S, van Ooij C, Liolios K, Harrison T, Lopez-Estrano C, Haldar K (2004) A host-targeting signal in virulence proteins reveals a secretome in malarial infection. Science 306:1934–1937

Hiratsuka M, Katayama T, Uematsu K, Kiyomura M, Ito M (2009) In vivo visualization of nitric oxide and interactions among platelets, leukocytes, and endothelium following hemorrhagic shock and reperfusion. Inflamm Res 58:463–471

Hossain M, Qadri SM, Liu L (2012) Inhibition of nitric oxide synthesis enhances leukocyte rolling and adhesion in human microvasculature. J Inflamm 9:28–35

Idro R, Karamagi C, Tumwine J (2004) Immediate outcome and prognostic factors for cerebral malaria among children admitted to Mulago Hospital Uganda. Ann Trop Paediatr: Int Child Health 24:17–24

Article   Google Scholar  

Idro R, Jenkins NE, Newton CR (2005) Pathogenesis, clinical features, and neurological outcome of cerebral malaria. Lancet Neurol 4:827–840

Idro R, Carter JA, Fegan G, Neville BGR, Newton CRJC (2006) Risk factors for persisting neurological and cognitive impairments following cerebral malaria. Arch Dis Child 91:142–148

Idro R, Kakooza-Mwesige A, Balyejjussa S, Mirembe G, Mugasha C, Tugumisirize J, Byarugaba J (2010) Severe neurological sequelae and behaviour problems after cerebral malaria in Ugandan children. BMC Res Notes 3:104

Jain K, Sood S, Gowthamarajan K (2013) Modulation of cerebral malaria by curcumin as an adjunctive therapy. Braz J Infect Dis 17:579–591

John CC et al (2008) Cerebral malaria in children is associated with long-term cognitive impairment. Pediatrics 122:e92–e99

John CC, Kutamba E, Mugarura K, Opoka RO (2010) Adjunctive therapy for cerebral malaria and other severe forms of Plasmodium falciparum malaria 8(9):997–1008

Kawachi S, Leffer DJ, Van Der Heyde HC, Laroux S, Gray L, Bharwani SS, Grisham MB (2000) Role of different nitric oxide synthese (NOS) isoforms in the regulation of endothelial cell adhesion molecule (ECAM) expression in vivo. Gastroenterology 118:A827

Kim H, Higgins S, Liles WC, Kain KC (2011) Endothelial activation and dysregulation in malaria: a potential target for novel therapeutics. Curr Opin Hematol 18:177–185

Kochar D, Kumawat B, Kochar S (1997) Seizures in cerebral malaria. QJM 90:605–607

Kochar DK, Kumawat BL, Vyas SP (2000) Prognostic significance of eye changes in cerebral malaria. J Assoc Physicians India 48:473–477

Kochar D, Kumawat B, Kochar S, Halwai M, Makkar R, Joshi A, Thanvi I (2002) Cerebral malaria in Indian adults: a prospective study of 441 patients from Bikaner, north-west India. J Assoc Physicians India 50:234–241

Kraisin S et al (2011) Association of ADAMTS13 polymorphism with cerebral malaria. Malar J 10:366

Kubes P, Suzuki M, Granger DN (1991) Nitric oxide: an endogenous modulator of leukocyte adhesion. Proc Natl Acad Sci 88:4651–4655

Lacerda MVG, Mourão MPG, Coelho HCC, Santos JB (2011) Thrombocytopenia in malaria: who cares? Mem Inst Oswaldo Cruz 106:52–63

Larkin D et al (2009) Severe Plasmodium falciparum malaria is associated with circulating ultra-large von Willebrand multimers and ADAMTS13 inhibition. PLoS Pathog 5:e1000349

Looareesuwan S et al (1983) Retinal hemorrhage, a common sign of significance in cerebral malaria Am J Trop Med Hyg September 1983 32: 911–915. Am J Trop Med Hyg 32:1002–1012

Lopansri BK et al (2003) Low plasma arginine concentrations in children with cerebral malaria and decreased nitric oxide production. Lancet 361:676–678

Lopez-Rubio JJ, Gontijo AM, Nunes MC, Issar N, Hernandez Rivas R, Scherf A (2007) 5′ flanking region of var genes nucleate histone modification patterns linked to phenotypic inheritance of virulence traits in malaria parasites. Mol Microbiol 66:1296–1305

Marchi N, Tierney W, Alexopoulos AV, Puvenna V, Granata T, Janigro D (2011) The etiological role of blood–brain barrier dysfunction in seizure disorders Cardiovascular psychiatry and neurology 2011

Matsushita K, Yamakuchi M, Morrell CN, Ozaki M, O’Rourke B, Irani K, Lowenstein CJ (2004) Vascular endothelial growth factor regulation of Weibel-Palade body exocytosis. Blood

Maude RJ et al (2009) The spectrum of retinopathy in adults with Plasmodium falciparum malaria. Trans R Soc Trop Med Hyg 103:665–671

Maude RJ et al. (2014) Magnetic resonance imaging of the brain in adults with severe falciparum malaria Renal failure (creatinine > 3 g/dL or anuria) 5:12

Mayer C, Slater L, Erat MC, Konrat R, Vakonakis I (2012) Structural analysis of the Plasmodium falciparum erythrocyte membrane protein 1 (PfEMP1) intracellular domain reveals a conserved interaction epitope. J Biol Chem 287:7182–7189

Medana IM, Turner GDH (2006) Human cerebral malaria and the blood brain barrier. Int J Parasitol 36:555–568

Medana IM et al (2002) Axonal injury in cerebral malaria. Am J Pathol 160:655–666

Miranda AS et al (2013) Further evidence for an anti-inflammatory role of artesunate in experimental cerebral malaria. Malar J 12:1–13

Mohanty S, Mishra SK, Pati SS, Pattnaik J, Das BS (2003) Complications and mortality patterns due to Plasmodium falciparum malaria in hospitalized adults and children, Rourkela, Orissa, India. Trans R Soc Trop Med Hyg 97:69–70

Molyneux M, Taylor T, Wirima J, Borgsteinj A (1989) Clinical features and prognostic indicators in paediatric cerebral malaria: a study of 131 comatose Malawian children. QJM 71:441–459

Morel O et al. (2005) [The significance of circulating microparticles in physiology, inflammatory and thrombotic diseases] La Revue de medecine interne/fondee par la Societe nationale francaise de medecine interne 26:791–801

Mwanga-Amumpaire J et al (2015) Inhaled nitric oxide as an adjunctive treatment for cerebral malaria in children: a phase II randomized open-label clinical trial. In: Open forum infectious diseases. Oxford University Press, p ofv111

Nacer A, Movila A, Baer K, Mikolajczak SA, Kappe SHI, Frevert U (2012) Neuroimmunological blood brain barrier opening in experimental cerebral malaria. PLoS Pathog 8:e1002982

N’Dilimabaka N et al (2014) P. Falciparum isolate-specific distinct patterns of induced apoptosis in pulmonary and brain endothelial cells. PLoS One 9:e90692

Newton C et al (1997) Intracranial hypertension in Africans with cerebral malaria. Arch Dis Child 76:219–226

Newton CR, Hien TT, White N (2000) Cerebral malaria. J Neurol Neurosurg Psychiatry 69:433–441

Ockenhouse CF et al (1992) Human vascular endothelial cell adhesion receptors for Plasmodium falciparum-infected erythrocytes: roles for endothelial leukocyte adhesion molecule 1 and vascular cell adhesion molecule 1. J Exp Med 176:1183–1189

Okoromah C, Afolabi BB, Wall E (2011) Mannitol and other osmotic diuretics as adjuncts for treating cerebral malaria. Cochrane Database Syst Rev 4

Olliaro P (2008) Mortality associated with severe Plasmodium falciparum malaria increases with age. Clin Infect Dis 47:158–160

Olumese P, Adeyemo A, Gbadegesin R, Walker O (1997) Retinal haemorrhage in cerebral malaria. East Afr Med J 74:285–287

Oluwayemi IO, Brown BJ, Oyedeji OA, Oluwayemi MA (2014) Neurological sequelae in survivors of cerebral malaria. Pan Afr Med J 15(1)

Organisation W (2013) World malaria report 2013. WHO, Geneva

Pacher P, Beckman JS, Liaudet L (2007) Nitric oxide and peroxynitrite in health and disease. Physiol Rev 87:315–424

Penet M-F, Abou-Hamdan M, Coltel N, Cornille E, Grau GE, De Reggi M, Gharib B (2008) Protection against cerebral malaria by the low-molecular-weight thiol pantethine. Proc Natl Acad Sci 105:1321–1326

Percãrio S et al (2012) Oxidative stress in malaria. Int J Mol Sci 13:16346–16372

Persidsky Y, Ramirez SH, Haorah J, Kanmogne GD (2006) Blood brain barrier: structural components and function under physiologic and pathologic conditions. J Neuroimmune Pharmacol 1:223–236

Petter M et al (2011) Expression of P. falciparum var genes involves exchange of the histone variant H2A. Z at the promoter. PLoS Pathog 7:e1001292

Phiri HT et al (2011) Elevated plasma von Willebrand factor and propeptide levels in Malawian children with malaria. PLoS One 6:e25626

Piguet PF, Kan CD, Vesin C (2002) Thrombocytopenia in an animal model of malaria is associated with an increased caspase-mediated death of thrombocytes. Apoptosis 7:91–98

Piguet PF, Da Laperrousaz C, Vesin C, Tacchini-Cottier F, Senaldi G, Grau GE (2000) Delayed mortality and attenuated thrombocytopenia associated with severe malaria in urokinase-and urokinase receptor-deficient mice. Infect Immun 68:3822–3829

Pino P et al (2003) Plasmodium falciparum-infected erythrocyte adhesion induces caspase activation and apoptosis in human endothelial cells. J Infect Dis 187:1283–1290

Pino P, Taoufiq Z, Nitcheu J, Vouldoukis I, Mazier D (2005) Blood–brain barrier breakdown during cerebral malaria: suicide or murder? THROMBOSIS AND HAEMOSTASIS-STUTTGART- 94:336

Pino P et al (2004) Induction of the CD23/nitric oxide pathway in endothelial cells downregulates ICAM-1 expression and decreases cytoadherence of Plasmodium falciparum infected erythrocytes. Cell Microbiol 6:839–848

Pongponratn E, Riganti M, Harinasuta T, Bunnag D (1985) Electron microscopy of the human brain in cerebral malaria. Southeast Asian J Trop Med Public Health 16:219–227

Ralph SA, Scheidig-Benatar C, Scherf A (2005) Antigenic variation in Plasmodium falciparum is associated with movement of var loci between subnuclear locations. Proc Natl Acad Sci U S A 102:5414–5419

Reis PA et al (2010) Cognitive dysfunction is sustained after rescue therapy in experimental cerebral malaria, and is reduced by additive antioxidant therapy. PLoS Pathog 6:e1000963

Reis PA et al (2012) Statins decrease neuroinflammation and prevent cognitive impairment after cerebral malaria. PLoS Pathog 8:e1003099

Rockett KA, Awburn MM, Cowden WB, Clark IA (1991) Killing of Plasmodium falciparum in vitro by nitric oxide derivatives. Infect Immun 59:3280–3283

Rowe JA, Claessens A, Corrigan RA, Arman M (2009) Adhesion of Plasmodium falciparum-infected erythrocytes to human cells: molecular mechanisms and therapeutic implications. Expert Rev Mol Med 11:e16

Sattar MA, Hoque HW, Amin MR, Faiz MA, Rahman MR (2009) Neurological findings and outcome in adult cerebral malaria. Bangladesh Med Res Counc Bull 35:15–17

Schindler SM, Little JP, Klegeris A (2014) Microparticles: a New perspective in central nervous system disorders. BioMed Res Int 2014:756327

Senczuk AM, Reeder JC, Kosmala MM, Ho M (2001) Plasmodium falciparum erythrocyte membrane protein 1 functions as a ligand for P-selectin. Blood 98:3132–3135

Serirom S, Raharjo WH, Chotivanich K, Loareesuwan S, Kubes P, Ho M (2003) Anti-adhesive effect of nitric oxide on plasmodium falciparum cytoadherence under flow. Am J Pathol 162:1651–1660

Smith JD et al (1995) Switches in expression of plasmodium falciparum var genes correlate with changes in antigenic and cytoadherent phenotypes of infected erythrocytes. Cell 82:101–110

Su X-z et al (1995) The large diverse gene family var encodes proteins involved in cytoadherence and antigenic variation of plasmodium falciparum-infected erythrocytes. Cell 82:89–100

Susomboon P et al (2006) Down-regulation of tight junction mRNAs in human endothelial cells co-cultured with Plasmodium falciparum infected erythrocytes. Parasitol Int 55:107–112

Toure FS et al (2008) Apoptosis: a potential triggering mechanism of neurological manifestation in Plasmodium falciparum malaria. Parasite Immunol 30:47–51

Turner GD et al (1994) 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 145:1057

Turner GD et al (1998) Systemic endothelial activation occurs in both mild and severe malaria. Correlating dermal microvascular endothelial cell phenotype and soluble cell adhesion molecules with disease severity. Am J Pathol 152:1477

van der Heyde HC, Nolan J, Combes V, Gramaglia I, Grau GE (2006) A unified hypothesis for the genesis of cerebral malaria: sequestration, inflammation and hemostasis leading to microcirculatory dysfunction. Trends Parasitol 22:503–508

van Hensbroek MB et al (1996) The effect of a monoclonal antibody to tumor necrosis factor on survival from childhood cerebral malaria. J Infect Dis 174:1091–1097

Waller KL, Cooke BM, Nunomura W, Mohandas N, Coppel RL (1999) Mapping the binding domains involved in the interaction between the plasmodium falciparum knob-associated histidine-rich protein (KAHRP) and the cytoadherence ligand P. FalciparumErythrocyte membrane protein 1 (PfEMP1). J Biol Chem 274:23808–23813

Warrell DA, Looareesuwan S, Warrell MJ, Kasemsarn P, Intaraprasert R, Bunnag D, Harinasuta T (1982) Dexamethasone proves deleterious in cerebral malaria: a double-blind trial in 100 comatose patients. N Engl J Med 306:313–319

Wassmer SC, Combes V, Candal FJ, Juhan-Vague I, Grau GE (2006a) Platelets potentiate brain endothelial alterations induced by Plasmodium falciparum. Infect Immun 74:645–653

Wassmer SC, Lépolard C, Traoré B, Pouvelle B, Gysin J, Grau GE (2004) Platelets reorient Plasmodium falciparum-infected erythrocyte cytoadhesion to activated endothelial cells. J Infect Dis 189:180–189

Wassmer SC, de Souza JB, Frère C, Candal FJ, Juhan-Vague I, Grau GE (2006b) TGF-beta-1 released from activated platelets can induce TNF-stimulated human brain endothelium apoptosis: a new mechanism for microvascular lesion during cerebral malaria. J Immunol 176:1180–1184

Weatherall DJ, Miller LH, Baruch DI, Marsh K, Doumbo OK, Casals-Pascual C, Roberts DJ (2002) Malaria and the red cell ASH Education Program Book 2002:35–57

Weinberg JB, et al. (2014) Dimethylarginines: endogenous inhibitors of nitric oxide synthesis in children with falciparum malaria. J Infect Dis :jiu156

White NJ, Phillips RE, Looareesuwan S, Chanthavanich P, Warrell DA (1988) Single dose phenobarbitone prevents convulsions in cerebral malaria. Lancet 332:64–66

White VA, Lewallen S, Beare N, Kayira K, Carr RA, Taylor TE (2001) Correlation of retinal haemorrhages with brain haemorrhages in children dying of cerebral malaria in Malawi. Trans R Soc Trop Med Hyg 95:618–621

World Health O (2000) Severe falciparum malaria. Trans R Soc Trop Med Hyg 94:1–90

Yeo TW et al (2008a) Recovery of endothelial function in severe falciparum malaria: relationship with improvement in plasma L-arginine and blood lactate concentrations. J Infect Dis 198:602–608

Yeo TW et al (2008b) Angiopoietin-2 is associated with decreased endothelial nitric oxide and poor clinical outcome in severe falciparum malaria. Proc Natl Acad Sci 105:17097–17102

Zanini GM, Cabrales P, Barkho W, Frangos JA, Carvalho L (2011) Exogenous nitric oxide decreases brain vascular inflammation, leakage and venular resistance during Plasmodium berghei ANKA infection in mice. J Neuroinflammation 8:66

Download references

Acknowledgments

Laboratories of Rachna Hora and Prakash Chandra Mishra are funded by Department of Biotechnology (DBT), Govt. of India, University Grants Commission (UGC, Govt. of India) and Department of Science and technology (DST, Govt. of India). Payal Kapoor is a DST INSPIRE (Department of Science and Technology - Innovation in Science Pursuit for Inspired Research) senior research fellow. Kirandeep Kaur Thind was initially funded by DBT and later supported by junior research fellowship from CSIR (Council of Scientific and Industrial Research), Govt. of India.

Author information

Authors and affiliations.

Department of Molecular Biology and Biochemistry, Guru Nanak Dev University, Amritsar, 143005, India

Rachna Hora, Payal Kapoor & Kirandeep Kaur Thind

Department of Biotechnology, Guru Nanak Dev University, Amritsar, 143005, India

Prakash Chandra Mishra

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Rachna Hora .

Rights and permissions

Reprints and permissions

About this article

Hora, R., Kapoor, P., Thind, K.K. et al. Cerebral malaria – clinical manifestations and pathogenesis. Metab Brain Dis 31 , 225–237 (2016). https://doi.org/10.1007/s11011-015-9787-5

Download citation

Received : 09 July 2015

Accepted : 22 December 2015

Published : 08 January 2016

Issue Date : April 2016

DOI : https://doi.org/10.1007/s11011-015-9787-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Cerebral malaria
  • Cytoadherence
  • Neurological deficits
  • Neuroinflammation
  • Find a journal
  • Publish with us
  • Track your research
  • Submit a Manuscript
  • Advanced search

American Journal of Neuroradiology

American Journal of Neuroradiology

Advanced Search

Cerebral Malaria

  • Cerebral malaria is a life-threatening complication of P. falciparum infestation that occurs in approximately 2% of patients.
  • Pathogenesis may be explained by 2 mechanisms: vascular sequestration of parasitized erythrocytes and the potential cerebral toxicity by cytokines.
  • Progressive clinical changes occur, along with high fever and chills.
  • Neurologic manifestations are nonspecific because of diffuse involvement of the brain.
  • Patients can become drowsy and disorientated, with associated seizures and progressive deterioration in consciousness before they become comatose.
  • Appearances may be normal on CT.
  • Restricted diffusion representing infarcts
  • No restricted diffusion, as with this patient, thought to be secondary to an inflammatory/demyelinating etiology 
  • Acute infarcts, which are frequently hemorrhagic, tend to involve the thalami/basal ganglia due to involvement of perforator arteries or cortical/subcortical involvement, represented by T2-FLAIR hyperintensity, diffusion restriction, and susceptibility effects.
  • Presence of enhancement may reflect the evolution of acute infarcts or be related to active inflammation/demyelination.
  • Demyelination including ADEM
  • Neuropsychiatric SLE
  • Antimalarial therapy such as artemisinin derivatives
  • Supportive management based on clinical and biochemical abnormalities

presentation of cerebral malaria

A 41-year-old woman on HAART with fever and seizures before entering a coma after recent travel to Uganda. MRI performed 2 weeks into admission. 

Subscribe to RSS - Cerebral Malaria

U.S. flag

A .gov website belongs to an official government organization in the United States.

A lock ( ) or https:// means you've safely connected to the .gov website. Share sensitive information only on official, secure websites.

  • How It Spreads
  • Where Malaria Occurs
  • World Malaria Day 2024
  • Clinical Guidance: Malaria Diagnosis & Treatment in the U.S.
  • Clinical Features
  • Clinical Testing and Diagnosis
  • Malaria Risk Assessment for Travelers
  • Choosing a Drug to Prevent Malaria
  • Malaria Surveillance, United States 2019 – 2020
  • How to Report a Case of Malaria
  • Public Health Strategy
  • Malaria's Impact Worldwide
  • Communication Resources
  • Malaria Surveillance & Case Investigation Best Practices
  • View All Home

Clinical Features of Malaria

At a glance.

  • Malaria infection is caused by Plasmodium parasite species.
  • Malaria disease can be categorized as uncomplicated or severe (complicated).
  • Severity of symptoms and duration of disease can depend on the species of the malaria parasite and level of immunity.
  • Generally, malaria is curable if diagnosed and treated promptly and correctly.

Senegal PARMA Hub

Clinical presentation

Infection with malaria parasites may result in a wide variety of symptoms, ranging from absent or very mild symptoms to severe disease and even death. Malaria disease can be categorized as uncomplicated or severe (complicated). In general, malaria is a curable disease if diagnosed and treated promptly and correctly.

All the clinical symptoms associated with malaria are caused by the asexual erythrocytic or blood stage parasites . When the parasite develops in the erythrocyte, numerous known and unknown waste substances such as hemozoin pigment and other toxic factors accumulate in the infected red blood cell. These are dumped into the bloodstream when the infected cells lyse and release invasive merozoites. The hemozoin and other toxic factors such as glucose phosphate isomerase (GPI) stimulate macrophages and other cells to produce cytokines and other soluble factors which act to produce fever and rigors and probably influence other severe pathophysiology associated with malaria.

Plasmodium falciparum- infected erythrocytes, particularly those with mature trophozoites, adhere to the vascular endothelium of venular blood vessel walls and do not freely circulate in the blood. When this sequestration of infected erythrocytes occurs in the vessels of the brain it is believed to be a factor in causing the severe disease syndrome known as cerebral malaria, which is associated with high mortality.

Incubation Period

Following the infective bite by the female Anopheles mosquito , a period of time (the "incubation period") goes by before the first symptoms appear in a patient. The incubation period in most cases varies from 7 to 30 days. The shorter periods most frequently observed with P. falciparum and the longer ones with P. malariae .

Antimalarial drugs taken for prophylaxis by travelers are highly effective if taken as prescribed. However, antimalarial prophylaxis occasionally can delay the appearance of malaria symptoms by weeks or months, long after the traveler has left the malaria-endemic area. (This can happen particularly with P. vivax and P. ovale , both of which can produce dormant liver stage parasites; the liver stages may reactivate and cause disease months after the infective mosquito bite.)

Such long delays between exposure and development of symptoms can result in misdiagnosis or delayed diagnosis because of reduced clinical suspicion by the health-care provider. Returned travelers should always remind their healthcare providers of any travel during the past 12 months in areas where malaria occurs.

Uncomplicated Malaria

Commonly, the patient initially presents with a combination of the following symptoms (which may be mild):

  • Nausea and vomiting
  • General malaise

In countries where cases of malaria are infrequent, these symptoms may be attributed to influenza, a cold, or other common infections, especially if malaria is not suspected. Conversely, in countries where malaria is frequent, residents often recognize the symptoms as malaria and treat themselves without seeking diagnostic confirmation ("presumptive treatment"). Physical findings may include the following:

  • Elevated temperatures
  • Perspiration
  • Enlarged spleen
  • Mild jaundice
  • Enlargement of the liver
  • Increased respiratory rate

Diagnosis of malaria depends on the demonstration of parasites in the blood, usually by microscopy. Additional laboratory findings may include mild anemia, mild decrease in blood platelets (thrombocytopenia), elevation of bilirubin, and elevation of aminotransferases.

Severe Malaria

Progression to severe malaria occurs when infections are complicated by serious organ failures or abnormalities in the patient's blood or metabolism, usually following delays in diagnosis and treatment. Criteria for severe malaria may differ slightly depending on the country where you are practicing, see WHO Guidelines for Malaria .

For healthcare providers practicing in the U.S., the criteria for severe malaria include any one or more of the following:

  • High percent parasitemia (≥5%)
  • Impaired consciousness
  • Circulatory collapse/shock
  • Pulmonary edema or acute respiratory distress syndrome (ARDS)
  • Acute kidney injury
  • Abnormal bleeding or disseminated intravascular coagulation (DIC)
  • Jaundice (must be accompanied by at least one other sign)
  • Severe anemia (Hb <7 g/dL)

Malaria Relapse

In P. vivax and P. ovale infections, patients having recovered from the first episode of illness may suffer several additional attacks ("relapses") after months or even years without symptoms. Relapses occur because P. vivax and P. ovale have dormant liver stage parasites ( "hypnozoites" ) that may reactivate, infect peripheral erythrocytes, and begin a new symptomatic episode of malaria. Treatment to reduce the chance of such relapses is available and should follow treatment of the first attack.

Other manifestations of malaria

  • Neurologic defects may occasionally persist (sometimes life-long) following cerebral malaria, especially in children. Such defects include trouble with movements (ataxia), palsies, speech difficulties, deafness, neurocognitive deficits, and blindness.
  • Recurrent infections with P. falciparum may result in severe anemia. This occurs especially in young children in tropical Africa with frequent infections that are inadequately treated.
  • Malaria during pregnancy (especially P. falciparum ) may cause severe disease in the mother and may lead to premature delivery or delivery of a low-birth-weight baby.
  • On rare occasions, P. vivax malaria can cause rupture of the spleen.
  • Nephrotic syndrome (a chronic, severe kidney disease) can result from chronic or repeated infections with P. malariae .
  • Hyperreactive malarial splenomegaly (also called "tropical splenomegaly syndrome") occurs infrequently and is attributed to an abnormal immune response to repeated malarial infections. The disease is marked by a very enlarged spleen and liver, abnormal immunologic findings, anemia, and a susceptibility to other infections (such as skin or respiratory infections).

Malaria is a serious disease caused by a parasite that infects the Anopheles mosquito. You get malaria when bitten by an infective mosquito.

For Everyone

Health care providers, public health.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Pattern of Clinical and Laboratory Presentation of Cerebral Malaria among Children in Nigeria

Affiliations.

  • 1 Department of Paediatrics, Federal Medical Centre, Owo, Ondo, Nigeria.
  • 2 Department of Paediatrics and Child Health, Obafemi Awolowo University, Ile-Ife, Osun, Nigeria.
  • 3 Department of Paediatrics, University of Medical Sciences Teaching Hospital, Akure, Ondo, Nigeria.
  • 4 Department of Psychiatry, Federal Neuropsychiatry Hospital, Benin, Edo, Nigeria.
  • PMID: 38680759
  • PMCID: PMC11045150
  • DOI: 10.4103/jgid.jgid_100_23

Introduction: Cerebral malaria (CM) is the most lethal form of severe malaria with high case fatality rates. Overtime, there is an inherent risk in changing pattern of presentation of CM which, if the diagnosis is missed due to these changing factors, may portend a poor outcome. Variations in the pattern of clinic-laboratory presentations also make generalization difficult. This study was, therefore, set out to report the pattern of clinical and laboratory presentation of CM.

Methods: This was a cross-sectional study among children aged 6 months to 14 years admitted with a diagnosis of CM as defined by the World Health Organization criteria. A pretested pro forma was filled, and detailed neurological examination and laboratory (biochemical, microbiology, and hematology) investigations were done. P <5% was considered statistically significant.

Results: Sixty-four children were recruited with a mean age of 34.9 ± 24.9 months and a male-to-female ratio of 1.9:1. There were 87.5% of under-five children. Fever (96.9%) was the major presenting feature closely followed by convulsions (92.2%). Convulsions were mainly generalized (94.9%) and multiple (76.5%). Profound coma (Blantyre coma score of 0) was present in 12.5% of cases, and the leading features on examination were fever (84.4%) and pallor (75.0%). Retinal vessel whitening (48.4%) was the most common funduscopic abnormality. Metabolic acidosis (47.9%), severe anemia (14.1%), hyperglycemia (17.2%), and hypoglycemia (7.8%) were seen among the children. Few (1.6%) had hyperparasitemia and bacteremia (3.2%).

Conclusion: Early recognition of the clinical presentation and prompt management may improve the outcome of cerebral malaria.

Keywords: Cerebral malaria; Nigeria; clinical features; coma; hyperparasitemia.

Copyright: © 2024 Journal of Global Infectious Diseases.

PubMed Disclaimer

Conflict of interest statement

There are no conflicts of interest.

Types and patterns of convulsions…

Types and patterns of convulsions recorded among the children

Similar articles

  • Folic acid supplementation and malaria susceptibility and severity among people taking antifolate antimalarial drugs in endemic areas. Crider K, Williams J, Qi YP, Gutman J, Yeung L, Mai C, Finkelstain J, Mehta S, Pons-Duran C, Menéndez C, Moraleda C, Rogers L, Daniels K, Green P. Crider K, et al. Cochrane Database Syst Rev. 2022 Feb 1;2(2022):CD014217. doi: 10.1002/14651858.CD014217. Cochrane Database Syst Rev. 2022. PMID: 36321557 Free PMC article.
  • Severe anaemia in childhood cerebral malaria is associated with profound coma. Idro R. Idro R. Afr Health Sci. 2003 Apr;3(1):15-8. Afr Health Sci. 2003. PMID: 12789083 Free PMC article.
  • Retinopathy in severe malaria in Ghanaian children--overlap between fundus changes in cerebral and non-cerebral malaria. Essuman VA, Ntim-Amponsah CT, Astrup BS, Adjei GO, Kurtzhals JA, Ndanu TA, Goka B. Essuman VA, et al. Malar J. 2010 Aug 12;9:232. doi: 10.1186/1475-2875-9-232. Malar J. 2010. PMID: 20704742 Free PMC article.
  • Diagnosis and Treatment of the Febrile Child. Herlihy JM, D’Acremont V, Hay Burgess DC, Hamer DH. Herlihy JM, et al. In: Black RE, Laxminarayan R, Temmerman M, Walker N, editors. Reproductive, Maternal, Newborn, and Child Health: Disease Control Priorities, Third Edition (Volume 2). Washington (DC): The International Bank for Reconstruction and Development / The World Bank; 2016 Apr 5. Chapter 8. In: Black RE, Laxminarayan R, Temmerman M, Walker N, editors. Reproductive, Maternal, Newborn, and Child Health: Disease Control Priorities, Third Edition (Volume 2). Washington (DC): The International Bank for Reconstruction and Development / The World Bank; 2016 Apr 5. Chapter 8. PMID: 27227231 Free Books & Documents. Review.
  • Neurological sequelae of cerebral malaria in children. Brewster DR, Kwiatkowski D, White NJ. Brewster DR, et al. Lancet. 1990 Oct 27;336(8722):1039-43. doi: 10.1016/0140-6736(90)92498-7. Lancet. 1990. PMID: 1977027 Review.
  • Garcia LS. Malaria. Clin Lab Med. 2010;30:93–129. - PubMed
  • Trivedi S, Chakravarty A. Neurological complications of malaria. Curr Neurol Neurosci Rep. 2022;22:499–513. - PMC - PubMed
  • World Health Organization. The “World Malaria Report 2022”at a Glance. [[Last accessed on 2022 Dec 23]]. Available from: https://www.who.int/teams/global-malaria-programme/reports/world-malaria... .
  • World Health Organization. Fact Sheet about Malaria. Geneva: World Health Organization; 2020. [[Last accessed on 2022 Dec 20]]. Available from: https://www.who.int/news-room/fact-sheets/detail/malaria .
  • Olumese PE, Gbadegesin RA, Adeyemo AA, Brown B, Walker A. Neurological features of cerebral malaria in Nigerian children. Ann Trop Paediatr. 1999;19:321–5. - PubMed

Related information

Linkout - more resources, full text sources.

  • Europe PubMed Central
  • Ovid Technologies, Inc.
  • PubMed Central
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

What is Cerebral Malaria?

  • Download PDF Copy

Benedette Cuffari, M.Sc.

Cerebral malaria is a serious neurological complication of severe malaria that affects about 1% of children under the age of 5 who have been infected with Plasmodium falciparum.

Malaria

Malaria. Image Credit: Christoph Burgstedt/Shutterstock.com

An overview of malaria

The Plasmodium genus of unicellular protozoan parasites are responsible for causing the disease malaria in humans, the most deadly of which includes Plasmodium falciparum. Each year, malaria affects about 219 million people, with about 500,000 people dying each year from this disease. It is estimated that about 93% of the people who die from malaria reside in sub-Saharan Africa, many of whom are under the age of 5.

Humans typically acquire malaria after being bitten by an infected female mosquito of the Anopheles genus. In addition, blood-borne transmission through the transfusion of blood products, transplantation, or needle-sharing, as well as congenital transmission of malaria can also occur. Once infected, malaria can either evolve into simple or severe forms, the latter of which is a less frequent presentation of this disease.

Severity of malaria

The severity of malaria is often dependent upon the immune status of the host, as well as the area in which malaria has been acquired. For example, areas that have a stable endemic of P. falciparum will often find that severe malaria commonly occurs in children up to 5 years of age, whereas older children and adults will experience less severe forms of the infection due to partial immunity. Comparatively, areas with lower endemic rates will have a less defined age distribution of severe malaria.

Initially, malaria will cause nonspecific flu-like symptoms including malaise, anorexia, lassitude, dizziness, headache, body aches, nausea, vomiting, and chills. The progression of malaria into more severe symptoms will often depend upon the infecting parasite. For example, infection by P. vivax and P. ovale will typically cause a classical malaria paroxysm that includes three symptom stages, beginning with a cold stage, followed by a hot stage, and ending with a sweating stage.  

Most of the severe complications of malaria will occur in individuals who have been infected with P. falciparum. Severe malaria is often defined as the presence of Plasmodium in peripheral blood. Some of the complications of severe malaria can involve the central nervous system, which is otherwise referred to as cerebral malaria, the pulmonary system, renal system, and/or the hematopoietic system. The progression to these complications is often rapid and can lead to death in many cases.

What is cerebral malaria?

According to the World Health Organization (WHO), cerebral malaria is defined as a severe form of P. falciparum malaria that causes cerebral manifestations. Typically, patients with cerebral malaria will experience a coma that persists for more than 30 minutes after a seizure occurs. While this may be true, patients with any degree of altered consciousness and other signs of cerebral dysfunction should be treated for severe malaria.

It is estimated that about 1% of children who have been infected with P. falciparum will develop cerebral malaria, with the incidence in adults being highly rare. However, low transmission areas will often find that cerebral malaria occurs more commonly in older children and adults. Studies have shown that younger children may be partly protected from cerebral malaria as a result of a process known as premunition, during which maternal immunity is transferred to the child in utero.

Symptoms of cerebral malaria      

As previously mentioned, coma is a hallmark symptom of cerebral malaria; however, the onset of this neurological complication can vary from occurring suddenly or gradually. When coma develops gradually, patients will often initially present with drowsiness, which should always be considered a worrying symptom, confusion, disorientation, delirium, or agitation. Addition observations that might be seen with cerebral malaria include:

Related Stories

  • Using AI to revolutionize mosquito surveillance to help combat malaria in Africa
  • Spatial transcriptomics unlocks malaria's liver stage secrets
  • SRI’s novel injectable formulation aims to provide a better option to fight malaria
  • Open-eyed but non-seeing
  • Disconjugate gaze
  • Sustained ocular deviation, usually upward or lateral
  • Decerebrate rigidity
  • Decorticate rigidity
  • Opisthotonos
  • Neck rigidity
  • Electroencephalographic abnormalities

Some other symptoms that are often common manifestations of cerebral malaria include fixed jaw closure and tooth grinding. Cerebral malaria is also often accompanied by non-neurological symptoms including enlargement of the liver and spleen, jaundice, pulmonary edema, renal dysfunction, pallor, hypoglycemia, bleeding, hypotension, and severe anemia.

Another notable complication of cerebral malaria that can be used to differentiate this condition from other encephalopathies that often occur in malaria-endemic areas includes malignant retinopathy. Malignant retinopathy in malaria arises due to the cerebral sequestration of parasites in the brain and can be diagnosed based on the presence of four major symptoms, which include:

  • Retinal whitening
  • Vessel changes
  • Retinal hemorrhages
  • papilledema

Treatment for cerebral malaria

Antimalarial treatment can be in the form of two classes of drugs, which include derivatives of artemisinin and those generated from cinconcha alkaloids. Whereas derivatives of artemisinin include artesunate and artemether, those that originate from cinconcha alkaloids include quinine and quinidine.

The ability of severe malaria to affect multiple organ systems, therefore, warrants additional medical interventions to treat these complications. For example, patients who experience a coma should, when necessary, be intubated and supported through mechanical ventilation. Since seizures are also a common complication of cerebral malaria, their medical management through the use of anticonvulsants is also crucial.

If left untreated, cerebral malaria is almost always fatal; therefore, aggressive treatment immediately should be initiated immediately after a diagnosis of cerebral malaria is made. Even after treatment is initiated, cerebral malaria still has a mortality rate of 20% and 15% in adults and children, respectively.

Fortunately, many of the patients who do survive cerebral malaria will typically experience a rapid recovery and a complete reversal of their neurological symptoms. However, some neurological symptoms may persist after recovery from cerebral malaria for several days or several weeks after their onset. These manifestations might include psychosis, cranial nerve lesions, extrapyramidal tremor, ataxia, polyneuropathy, and seizures.

References:

  • Luzolo, A. L., & Ngoyi, D. M. (2019). Cerebral malaria. Brain Research Bulletin 145 ; 53-58. doi:10.1016/j.brainresbull.2019.01.010.
  • Bruneel, F. (2019). Human cerebral malaria: 2019 mini-review. Revue Neurologique 175 (7-8); 445-450. doi:10.1016/j.neurol.2019.07.008.
  • BArtoloni, A., & Zammarchi, L. (2012). Clinical Aspects of Uncomplicated and Severe Malaria. Mediterranean Journal of Hematology and Infectious Diseases 4 (1). doi:10.4084/MJHID.2012.026.

Further Reading

  • All Malaria Content
  • The Effect of Climate Change on Malaria?
  • What is Malaria?
  • Malaria Causes
  • World Malaria Day 2021: An interview with Sir Brian Greenwood

Last Updated: Aug 17, 2021

Benedette Cuffari

Benedette Cuffari

After completing her Bachelor of Science in Toxicology with two minors in Spanish and Chemistry in 2016, Benedette continued her studies to complete her Master of Science in Toxicology in May of 2018. During graduate school, Benedette investigated the dermatotoxicity of mechlorethamine and bendamustine; two nitrogen mustard alkylating agents that are used in anticancer therapy.

Please use one of the following formats to cite this article in your essay, paper or report:

Cuffari, Benedette. (2021, August 17). What is Cerebral Malaria?. News-Medical. Retrieved on August 26, 2024 from https://www.news-medical.net/health/What-is-Cerebral-Malaria.aspx.

Cuffari, Benedette. "What is Cerebral Malaria?". News-Medical . 26 August 2024. <https://www.news-medical.net/health/What-is-Cerebral-Malaria.aspx>.

Cuffari, Benedette. "What is Cerebral Malaria?". News-Medical. https://www.news-medical.net/health/What-is-Cerebral-Malaria.aspx. (accessed August 26, 2024).

Cuffari, Benedette. 2021. What is Cerebral Malaria? . News-Medical, viewed 26 August 2024, https://www.news-medical.net/health/What-is-Cerebral-Malaria.aspx.

Suggested Reading

CRISPR-based genetic technique eradicates malaria mosquitoes with over 99% efficiency

Cancel reply to comment

  • Trending Stories
  • Latest Interviews
  • Top Health Articles

Study reveals genetic link between Alzheimer's disease, lipid metabolism, and coronary artery disease

Global and Local Efforts to Take Action Against Hepatitis

Lindsey Hiebert and James Amugsi

In this interview, we explore global and local efforts to combat viral hepatitis with Lindsey Hiebert, Deputy Director of the Coalition for Global Hepatitis Elimination (CGHE), and James Amugsi, a Mandela Washington Fellow and Physician Assistant at Sandema Hospital in Ghana. Together, they provide valuable insights into the challenges, successes, and the importance of partnerships in the fight against hepatitis.

Global and Local Efforts to Take Action Against Hepatitis

Addressing Important Cardiac Biology Questions with Shotgun Top-Down Proteomics

In this interview conducted at Pittcon 2024, we spoke to Professor John Yates about capturing cardiomyocyte cell-to-cell heterogeneity via shotgun top-down proteomics.

Addressing Important Cardiac Biology Questions with Shotgun Top-Down Proteomics

A Discussion with Hologic’s Tim Simpson on the Future of Cervical Cancer Screening

Tim Simpson

Hologic’s Tim Simpson Discusses the Future of Cervical Cancer Screening.

A Discussion with Hologic’s Tim Simpson on the Future of Cervical Cancer Screening

Latest News

GluFormer outperforms existing AI models in predicting blood sugar levels

Newsletters you may be interested in

Malaria

Your AI Powered Scientific Assistant

Hi, I'm Azthena, you can trust me to find commercial scientific answers from News-Medical.net.

A few things you need to know before we start. Please read and accept to continue.

  • Use of “Azthena” is subject to the terms and conditions of use as set out by OpenAI .
  • Content provided on any AZoNetwork sites are subject to the site Terms & Conditions and Privacy Policy .
  • Large Language Models can make mistakes. Consider checking important information.

Great. Ask your question.

Azthena may occasionally provide inaccurate responses. Read the full terms .

While we only use edited and approved content for Azthena answers, it may on occasions provide incorrect responses. Please confirm any data provided with the related suppliers or authors. We do not provide medical advice, if you search for medical information you must always consult a medical professional before acting on any information provided.

Your questions, but not your email details will be shared with OpenAI and retained for 30 days in accordance with their privacy principles.

Please do not ask questions that use sensitive or confidential information.

Read the full Terms & Conditions .

Provide Feedback

presentation of cerebral malaria

On this page

When to see a doctor, risk factors, complications.

Malaria is a disease caused by a parasite. The parasite is spread to humans through the bites of infected mosquitoes. People who have malaria usually feel very sick with a high fever and shaking chills.

While the disease is uncommon in temperate climates, malaria is still common in tropical and subtropical countries. Each year nearly 290 million people are infected with malaria, and more than 400,000 people die of the disease.

To reduce malaria infections, world health programs distribute preventive drugs and insecticide-treated bed nets to protect people from mosquito bites. The World Health Organization has recommended a malaria vaccine for use in children who live in countries with high numbers of malaria cases.

Protective clothing, bed nets and insecticides can protect you while traveling. You also can take preventive medicine before, during and after a trip to a high-risk area. Many malaria parasites have developed resistance to common drugs used to treat the disease.

Products & Services

  • A Book: Mayo Clinic Family Health Book
  • Newsletter: Mayo Clinic Health Letter — Digital Edition

Signs and symptoms of malaria may include:

  • General feeling of discomfort
  • Nausea and vomiting
  • Abdominal pain
  • Muscle or joint pain
  • Rapid breathing
  • Rapid heart rate

Some people who have malaria experience cycles of malaria "attacks." An attack usually starts with shivering and chills, followed by a high fever, followed by sweating and a return to normal temperature.

Malaria signs and symptoms typically begin within a few weeks after being bitten by an infected mosquito. However, some types of malaria parasites can lie dormant in your body for up to a year.

Talk to your doctor if you experience a fever while living in or after traveling to a high-risk malaria region. If you have severe symptoms, seek emergency medical attention.

From Mayo Clinic to your inbox

Malaria is caused by a single-celled parasite of the genus plasmodium. The parasite is transmitted to humans most commonly through mosquito bites.

Mosquito transmission cycle

Malaria transmission cycle

  • Malaria transmission cycle

Malaria spreads when a mosquito becomes infected with the disease after biting an infected person, and the infected mosquito then bites a noninfected person. The malaria parasites enter that person's bloodstream and travel to the liver. When the parasites mature, they leave the liver and infect red blood cells.

  • Uninfected mosquito. A mosquito becomes infected by feeding on a person who has malaria.
  • Transmission of parasite. If this mosquito bites you in the future, it can transmit malaria parasites to you.
  • In the liver. Once the parasites enter your body, they travel to your liver — where some types can lie dormant for as long as a year.
  • Into the bloodstream. When the parasites mature, they leave the liver and infect your red blood cells. This is when people typically develop malaria symptoms.
  • On to the next person. If an uninfected mosquito bites you at this point in the cycle, it will become infected with your malaria parasites and can spread them to the other people it bites.

Other modes of transmission

Because the parasites that cause malaria affect red blood cells, people can also catch malaria from exposure to infected blood, including:

  • From mother to unborn child
  • Through blood transfusions
  • By sharing needles used to inject drugs

The greatest risk factor for developing malaria is to live in or to visit areas where the disease is common. These include the tropical and subtropical regions of:

  • Sub-Saharan Africa
  • South and Southeast Asia
  • Pacific Islands
  • Central America and northern South America

The degree of risk depends on local malaria control, seasonal changes in malaria rates and the precautions you take to prevent mosquito bites.

Risks of more-severe disease

People at increased risk of serious disease include:

  • Young children and infants
  • Older adults
  • Travelers coming from areas with no malaria
  • Pregnant women and their unborn children

In many countries with high malaria rates, the problem is worsened by lack of access to preventive measures, medical care and information.

Immunity can wane

Residents of a malaria region may be exposed to the disease enough to acquire a partial immunity, which can lessen the severity of malaria symptoms. However, this partial immunity can disappear if you move to a place where you're no longer frequently exposed to the parasite.

Malaria can be fatal, particularly when caused by the plasmodium species common in Africa. The World Health Organization estimates that about 94% of all malaria deaths occur in Africa — most commonly in children under the age of 5.

Malaria deaths are usually related to one or more serious complications, including:

  • Cerebral malaria. If parasite-filled blood cells block small blood vessels to your brain (cerebral malaria), swelling of your brain or brain damage may occur. Cerebral malaria may cause seizures and coma.
  • Breathing problems. Accumulated fluid in your lungs (pulmonary edema) can make it difficult to breathe.
  • Organ failure. Malaria can damage the kidneys or liver or cause the spleen to rupture. Any of these conditions can be life-threatening.
  • Anemia. Malaria may result in not having enough red blood cells for an adequate supply of oxygen to your body's tissues (anemia).
  • Low blood sugar. Severe forms of malaria can cause low blood sugar (hypoglycemia), as can quinine — a common medication used to combat malaria. Very low blood sugar can result in coma or death.

Malaria may recur

Some varieties of the malaria parasite, which typically cause milder forms of the disease, can persist for years and cause relapses.

If you live in or are traveling to an area where malaria is common, take steps to avoid mosquito bites. Mosquitoes are most active between dusk and dawn. To protect yourself from mosquito bites, you should:

  • Cover your skin. Wear pants and long-sleeved shirts. Tuck in your shirt, and tuck pant legs into socks.
  • Apply insect repellent to skin. Use an insect repellent registered with the Environmental Protection Agency on any exposed skin. These include repellents that contain DEET, picaridin, IR3535, oil of lemon eucalyptus (OLE), para-menthane-3,8-diol (PMD) or 2-undecanone. Do not use a spray directly on your face. Do not use products with oil of lemon eucalyptus (OLE) or p-Menthane-3,8-diol (PMD) on children under age 3.
  • Apply repellent to clothing. Sprays containing permethrin are safe to apply to clothing.
  • Sleep under a net. Bed nets, particularly those treated with insecticides, such as permethrin, help prevent mosquito bites while you are sleeping.

Preventive medicine

If you'll be traveling to a location where malaria is common, talk to your doctor a few months ahead of time about whether you should take drugs before, during and after your trip to help protect you from malaria parasites.

In general, the drugs taken to prevent malaria are the same drugs used to treat the disease. What drug you take depends on where and how long you are traveling and your own health.

The World Health Organization has recommended a malaria vaccine for use in children who live in countries with high numbers of malaria cases.

Researchers are continuing to develop and study malaria vaccines to prevent infection.

Feb 09, 2023

  • AskMayoExpert. Malaria. Rochester, Minn.: Mayo Foundation for Medical Education and Research; 2018.
  • Jameson JL, et al., eds. Malaria. In: Harrison's Principles of Internal Medicine. 20th ed. New York, N.Y.: The McGraw-Hill Companies; 2018. https://accessmedicine.mhmedical.com. Accessed Oct. 9, 2018.
  • Tintinalli JE, et al., eds. Malaria. In: Tintinalli's Emergency Medicine: A Comprehensive Study Guide. 8th ed. New York, N.Y.: McGraw-Hill Education; 2016. http://www.accessmedicine.mhmedical.com. Accessed Oct. 9, 2018.
  • Malaria. Merck Manual Professional Version. http://www.merckmanuals.com/professional/infectious-diseases/extraintestinal-protozoa/malaria. Accessed Oct. 9, 2018.
  • Malaria. Centers for Disease Control and Prevention. http://wwwnc.cdc.gov/travel/diseases/malaria. Accessed Nov. 6, 2015.
  • Breman JG. Clinical manifestations of malaria in nonpregnant adults and children. https://www.uptodate.com/contents/search. Accessed Oct. 9, 2018.
  • Daily J. Treatment of uncomplicated falciparum malaria in nonpregnant adults and children. https://www.uptodate.com/contents/search. Accessed Oct. 9, 2018.
  • Key points: World malaria report 2017. World Health Organization. https://www.who.int/malaria/media/world-malaria-report-2017/en/. Accessed Oct. 9, 2018.
  • Malaria. World Health Organization. https://www.who.int/malaria/en/. Accessed Oct. 9, 2018.
  • Mutebi JP, et al. Protection against mosquitoes, ticks, & other arthropods. In: CDC Yellow Book 2018: Health Information for International Travelers. https://wwwnc.cdc.gov/travel/yellowbook/2018/the-pre-travel-consultation/protection-against-mosquitoes-ticks-other-arthropods. Accessed Oct. 27, 2018.
  • Diseases & Conditions
  • Malaria symptoms & causes

News from Mayo Clinic

presentation of cerebral malaria

More Information

CON-XXXXXXXX

Help transform healthcare

Your donation can make a difference in the future of healthcare. Give now to support Mayo Clinic's research.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Front Cell Infect Microbiol

Cerebral malaria induced by plasmodium falciparum : clinical features, pathogenesis, diagnosis, and treatment

Xiaonan song.

1 School of Basic Medical Sciences, Hubei University of Medicine, Shiyan, China

2 Beijing School of Chemistry and Bioengineering, University of Science and Technology Beijing, Beijing, China

Weijia Cheng

3 Key Laboratory of National Health Commission on Technology for Parasitic Diseases Prevention and Control, Jiangsu Provincial Key Laboratory on Parasites and Vector Control Technology, Jiangsu Institute of Parasitic Diseases, Wuxi, China

Haifeng Dong

4 Guangdong Key Laboratory for Genome Stability and Human Disease Prevention, Department of Biochemistry and Molecular Biology, School of Medicine, Shenzhen University, Shenzhen, China

Cerebral malaria (CM) caused by Plasmodium falciparum is a fatal neurological complication of malaria, resulting in coma and death, and even survivors may suffer long-term neurological sequelae. In sub-Saharan Africa, CM occurs mainly in children under five years of age. Although intravenous artesunate is considered the preferred treatment for CM, the clinical efficacy is still far from satisfactory. The neurological damage induced by CM is irreversible and lethal, and it is therefore of great significance to unravel the exact etiology of CM, which may be beneficial for the effective management of this severe disease. Here, we review the clinical characteristics, pathogenesis, diagnosis, and clinical therapy of CM, with the aim of providing insights into the development of novel tools for improved CM treatments.

Introduction

Malaria, a mosquito-transmitted infectious disease caused by Plasmodium species, remains a significant public health concern globally ( XP D, L S, 2021 ). In 2020, 241 million malaria cases were estimated, and 627,000 deaths occurred, with 77% found among children under five years of age ( Jiang et al., 2021 ). Currently, five Plasmodium spp. are reported to infect humans, including P. falciparum , P. ovale , P. vivax , P. malariae , and P. knowlesi . As we know, P. falciparum is considered the most severe species and the primary cause of mortality, notably in young children ( Su and Wu, 2021 ). Cerebral malaria (CM) is a fatal neurological complication of P. falciparum malaria ( Luzolo and Ngoyi, 2019 ), and children aged under 3 years and pregnant women are most susceptible ( World Health Organization, 2021 ). The mortality of CM is estimated to be 20% in children and 30% in adults ( Solomon et al., 2014 ). Furthermore, 15-20% of survivors suffer long-term neurological sequelae, such as hemiplegia, ataxia, speech disorders, and epilepsy, resulting in lifelong neurological deficits and even death ( Birbeck et al., 2010 ). Hereby, we review the clinical manifestations, pathogenesis, diagnosis, and treatment of CM so as to provide insights into the management of CM.

Clinical manifestations of CM

CM is clinically characterized as a diffuse encephalopathy with a history of fever for 2 to 3 days, subsequent seizures, and loss of consciousness (coma). Previous studies have demonstrated substantial differences in the clinical manifestations of CM between children and adults ( Table 1 ). Although this may be attributed to immune status and age, there are still many questions that remain to be answered ( Olliaro, 2008 ; Sahu et al., 2021 ). Pediatric CM usually manifests with coma, seizures, and severe anemia, while renal failure and respiratory distress rarely occur in African children ( Waller et al., 1995 ; Newton et al., 2000 ). Nevertheless, adult CM is frequently associated with multiple organ complications, including central nervous system (CNS) and liver dysfunction, respiratory failure, and acute kidney failure ( Mishra et al., 2007 ; Wassmer et al., 2015 ).

Table 1

Clinical manifestations of pediatric and adult cerebral malaria.

Clinical featuresChildrenAdults
Preceding symptomsFever, failure to eat or drink, vomiting and cough, and convulsions ( ).General malaise, head, back, and limb pain, dizziness, anorexia, nausea, vomiting, and mild diarrhea ( ).
Neurological systemComaIt develops rapidly, often after a seizure, and lasts for 1 to 2 days, reversible ( ; ).Develops gradually following delirium, Disorientation, and agitation over 2 to 3 days or follows a generalized seizure, lasts longer (2 days) ( ).
Nerve reflexMore common ( ).Rare.
Neurological impairmentsAtaxia (43%), hemiplegia (39%), speech disorders (39%) and blindness (30%). Other sequelae include behavioral disturbances, hypotonia, generalized spasticity, and a variety of tremors ( ).Psychosis, psychosis, ataxia, transitory cranial nerve palsies, mononeuritis multiplex, polyneuropathy, extrapyramidal and extrapyramidal tremors, and other cerebellar signs ( ).
Motor systemSeizuresHigh incidence, frequently mostly partial motor ( ).Low incidence, generalized seizures frequently, less focal ( ).
Status epilepticusUsual ( ).Rare ( ).
Abnormal behaviorHyperactivity, impulsiveness, and inattentiveness or conduct disorders ( ; ).Ataxia of gait, intention tremor, dysmetria, dysdiadochokinesis, nystagmus, and cerebellar dysarthria ( ).
Systemic complicationsHyponatremia, anemia, hypoglycemia, jaundice, metabolic acidosis, respiratory distress, hepatosplenomegaly, and intracranial pressure ( ; ; ; ).Anemia, hypoglycemia,
hemoglobinuria, jaundice, shock, renal failure, severe lactic acidosis, abnormal bleeding, pulmonary edema, and adult respiratory distress syndrome, Kussmaul’s breathing ( ; ).
RetinopathyRetinal whitening, orange or white discoloration of the retinal vessels, retinal hemorrhages, and infrequent papilledema ( ).Less prominent. Characterized by retinal hemorrhages and retinal whitening, no change in retinal vessel discoloration ( ).

Neurological system

Compared with adults, children have a higher incidence rate of seizures ( Postels and Birbeck, 2013 ). In children, focal motor and generalized tonic–clonic convulsions are the most common clinically detected seizures; however, subtle or subclinical seizures detected with electroencephalography (EEG) are also common ( Newton et al., 2000 ; Postels and Birbeck, 2013 ). Subtle seizures manifest as nystagmoid eye movements, irregular breathing, excessive salivation, and conjugate eye deviation ( Crawley et al., 1996 ). Most seizures in adult CM patients are generalized seizures; however, focal motor seizures may also occur. Occasionally, the sign of seizure activity is subtle, such as repetitive eye or hand movements, and may be easily overlooked. Subtle seizure activity seems to be more common in children than in adults ( Newton et al., 2000 ). The level of consciousness after a seizure is usually lower than that preceding it. Status epilepticus is unusual in adults, although more than one seizure is frequent ( Vespa et al., 1999 ). Previous studies reported an association between status epilepticus and neurological sequelae among CM patients, which occur in 5-15% of survivors ( Brewster et al., 1990 ), and it has been shown that prolonged seizure activity may damage the brain, causing deficits in both motor and cognitive functions ( Stafstrom et al., 1993 ).

Coma usually develops rapidly after seizures among children living in malaria-endemic areas, and consciousness recovers to normal rapidly (within 24-48 h) ( Genton et al., 1997 ). Different disease processes may affect awareness in children with malaria, including convulsions, hypoglycemia, hyperpyrexia, acidosis, severe anemia, and sedative drugs. Although the cause of impaired consciousness or coma remains unclear, it is likely to result from several interacting mechanisms ( Newton et al., 2000 ). Adhesion of malaria parasite-infected red blood cells (iRBCs) reduces microvascular blood flow ( Kaul et al., 1998 ), which may be the cause of organ tissue dysfunction, such as coma. High concentrations of tumor necrosis factor-α (TNF-α) are associated with coma ( Kwiatkowski et al., 1990 ; Kaul et al., 1998 ). Compared to children, coma gradually develops in adults following drowsiness, disorientation, delirium, and agitation within 2 to 3 days ( Kochar et al., 2002 ). Convulsion leads to the development of a coma and occurs in approximately 15% of adults and 80% of children ( Plewes et al., 2018 ).

Neurologic features

Abnormal corneal and oculocephalic reflexes (doll’s eye) are likely to occur in children with deep coma. Abnormal plantar reflexes are also detected, and abdominal reflexes are almost invariably absent. In adults with profound coma, corneal and eyelash reflexes are usually intact unless in a state of deep coma, and the pupils are normal. Forcible jaw closure and teeth grinding (bruxism) are commonly seen in CM. Pout reflex usually indicates a “frontal release”; however, the grasp reflex is frequently absent. In addition, increased muscle tone and tendon reflexes are found. CM may elicit ankle and patellar clonus, and extensor plantar responses. Nevertheless, abdominal and cremasteric reflexes are invariably absent ().

Neurological impairments

CM affects the CNS, and although most survivors have a full recovery, 3-31% of patients still develop neurological deficits and cognitive sequelae ( Oluwayemi et al., 2013 ). The prevalence of neurological deficits is higher in children than in adults, ranging from 6% to 29% at the time of discharge ( Idro et al., 2004 ; Hawkes et al., 2013 ). Children with CM frequently present long-term neurologic deficits, and episodes of CM imply the development of long-term sequelae in children. In children, the most common sequelae include ataxia, paralysis, paresis, cortical blindness, epilepsy, deafness, behavioral disorders, language disorders, and cognitive impairment ( Brewster et al., 1990 ). Sequelae are less common in adults. During the acute phase of CM, neurologic abnormalities include psychosis, ataxia, transitory cranial nerve palsies or tremor ( Peixoto and Kalei, 2013 ).

Retinopathy

The characteristic features of retinopathy due to CM include retinal whitening (macula whitening sparing central fovea and peripheral whitening of the fundus), retinal vessel discoloration to pink–orange or white, retinal hemorrhages, and papilledema ( Hora et al., 2016 ). The first two abnormalities are considered specific symptoms of CM. Commonalities between pediatric and adult patients include retinal hemorrhage, a common manifestation but a less distinctive feature. Retinal hemorrhage correlates with disease severity and cerebral hemorrhage in the microvascular dissection of the brain ( White et al., 2001 ). Papilledema is rare in children and adults. Although it is a nonspecific symptom of CM, it reflects increased intracranial pressure and portends a poor prognosis in children ( Beare et al., 2004 ). A prominent difference between children and adults is vessel discoloration. Orange or white discoloration of the retinal vessels has been attributed to the hemoglobinization of stationary erythrocytes infected with mature parasites ( Beare et al., 2011 ). The degree of retinal microvascular damage is comparable to cerebral damage ( Beare et al., 2004 ; Lewallen et al., 2008 ).

Non-CNS abnormalities in CM

Systemic complications include anemia (20% to 50% incidence), hypoglycemia (30% incidence), hyponatremia (>50% incidence), jaundice (8% incidence), metabolic acidosis characterized by respiratory distress, and hepatosplenomegaly in children living with CM ( White et al., 1987 ; English et al., 1996 ; Idro et al., 2005 ; Maitland and Newton, 2005 ). Renal failure and pulmonary edema are unusual in children ( Newton et al., 1991 ). CM predominantly manifests as CNS dysfunction in children; however, it is mainly present in multisystem and organ (circulatory, hepatic, coagulation, renal, and pulmonary) dysfunctions in adults ( Day et al., 2000 ; Krishnan and Karnad, 2003 ).

In adults, anemia is an inevitable consequence of CM and develops exceptionally rapidly. CM has been reported in patients together with pulmonary edema, adult respiratory distress syndrome and hemoglobinuria, and Kussmaul’s breathing occurs with acute renal failure and severe lactic acidosis ( Newton and Warrell, 1998 ). Hypoglycemia occurs in 8% of patients aged 26 to 28 years ( White et al., 1983 ). Other complications included jaundice, shock, abnormal bleeding, and coagulopathy.

Pathogenesis of CM

Although the pathophysiology of CM has been extensively investigated, the exact pathogenesis remains unclear. Currently, CM is widely accepted as a multifactorial process related to the adhesion and sequestration of iRBCs, immunological responses, endothelial cell (EC) activation, and loss of BBB integrity ( Idro et al., 2005 ). Nevertheless, any of these mechanisms alone fail to explain the pathogenesis of human CM, and they jointly participate in this potentially fatal infection. A mouse model of experimental cerebral malaria (ECM) has been used to simulate and explain the pathogenesis of human CM ( Figure 1 ).

An external file that holds a picture, illustration, etc.
Object name is fcimb-12-939532-g001.jpg

Schematic of experimental cerebral malaria (ECM) pathogenesis. The ECM is initiated by dendritic cells (DCs) processing and presenting infected red blood cell (iRBC) antigens to CD4 + and CD8 + T cells in the spleen (1). NK cells and macrophages are activated by iRBCs to secrete inflammatory cytokines (2). The iRBCs adhere to endothelial cells (ECs) of the brain microvasculature through the interaction between P. falciparum erythrocyte membrane protein-1 (PfEMP-1) of iRBCs and cell adhesion molecules of ECs (3). The adhesion of iRBCs to the cerebral microvascular endothelium is also further accompanied by agglutination to other iRBCs, platelets, white blood cells (WBCs), and the rosetting effect formed by the adhesion of iRBCs and RBCs. ECs are activated by interactions with iRBCs and responses to inflammatory cytokines. Activated ECs promote the upregulation of cell adhesion molecules (CAMs) on brain ECs and release cytokines and chemokines simultaneously (4). Activated CD8 + T cells express CXCR3 and CCR5 chemokine receptors, which bind to chemokines such as CXCL9, CXCL10, and CXCL4, inducing T-cell migration to the brain (5). Meanwhile, LFA-1 on CD8 + T cells promotes their adhesion to endothelial ICAM-1 (6). Parasitic antigens can be transferred from the vascular lumen to brain ECs. Brain ECs can cross-present parasitic antigens on MHC-1 molecular antigens and bind with antigen receptors (TCRs) on CD8 + T cells (7). The interaction induces apoptosis of ECs, leading to the destruction of the BBB (8). Meanwhile, the iRBCs directly activate platelets and stimulate the release of CXCL4. CXCL4 induces the production of TNF by T cells and macrophages, which causes more platelets to adhere to ECs (9). As leukocytes and platelets are recruited and activated, a local proinflammatory cycle ensues, with a positive feedback loop of EC activation, leukocyte/platelet sequestration, and parasite accumulation (10).

Adhesion and sequestration

Cerebral iRBCs adherence is an indicative marker of CM in adults and children, and it is considered a starting point during the development of CM. Sequestration is thought to be a specific interaction between iRBCs and vascular ECs, which is not limited to brain tissues but also occurs on ECs in different organs, including the lung, kidney, liver, and intestine.

The adhesion of iRBCs to the vascular endothelium is mediated by P. falciparum erythrocyte membrane protein 1 (PfEMP1) ( Jensen et al., 2020 ), a specific cell-surface ligand expressed by iRBCs. PfEMP1 belongs to the antigen-variant protein family, and the var genes encoding the protein are a large multigene family ( Kim, 2012 ). To date, 60 different var genes have been characterized, and var gene-encoded proteins have shown dual functions in regulating antigen variation and cell adhesion ( Tembo et al., 2014 ). PfEMP1 contains a host molecule binding domain and binds to several cell adhesion molecules (CAMs) on ECs, such as CD36 ( Berendt et al., 1989 ; Ockenhouse et al., 1989 ), intercellular adhesion molecule 1 (ICAM-1) ( Berendt et al., 1989 ), vascular adhesion molecule 1 (VCAM-1) ( Ockenhouse et al., 1989 ; Ockenhouse et al., 1992 ), endothelial protein C receptor (EPCR) ( Mohan Rao et al., 2014 ), thrombospondin, E-selectin ( Turner et al., 1994 ) and chondroitin sulphate A ( Rogerson et al., 1995 ; Fried and Duffy, 1996 ). Adhesion of iRBCs to the cerebral microvascular endothelium is further accompanied by agglutination to other iRBCs, platelets, white blood cells (WBCs), and rosetting produced by adhesion of iRBCs and uninfected erythrocytes ( Fried and Duffy, 1996 ). Sequestration of iRBCs in microvessels may protect iRBCs from clearance by the spleen. In addition, it weakens the capability of iRBCs and RBCs to denature, leading to blood vessel blockage. Previous studies reported a significant correlation between sequestration of iRBCs in cerebral vessels and coma in CM patients ( Silamut et al., 1999 ; Ponsford et al., 2012 ; Storm et al., 2019 ). Taken together, sequestration of iRBCs leads to increased vasoconstriction and vascular obstruction, as well as decreased cerebral blood flow and hypoxia.

Inflammatory responses

Excessive immune responses and the release of a large number of inflammatory factors play important roles in the pathogenesis of CM ( Shikani et al., 2012 ). The humoral response to malaria parasites includes immune activation of macrophages and lymphocytes (CD8 + , CD4 + , natural killer (NK) cells) and activation of monocytes, resulting in accumulation of immune cells in the microvasculature and a systemic inflammatory response secreted by proinflammatory cytokines, including tumor necrosis factor (TNF)-α, interferon (IFN)-γ, and interleukin-1β (IL-1β), which are elevated in an episode of acute CM.

At the early stage of malaria infection, CD4 + and CD8 + T cells are activated by antigen-presenting cells (APCs) to initiate antimalarial protective cellular immune responses. The chemotaxis of T cells to peripheral cerebral vessels is one of the prominent features of CM. Recruitment of CD8 + T cells is the most predominant characteristic ( Riggle et al., 2020 ), and priming of CD4 + and CD8 + T cells initiates CM in the spleen by dendritic cells (DCs) presenting iRBCs antigens. NK cells and macrophages are activated by iRBCs to release inflammatory cytokines, such as TNF-α, IFN-γ, IL-1β, IL-12 and chemokines ( Dunst et al., 2017 ). Adhesion of iRBCs and the release of inflammatory cytokines can activate brain ECs, triggering ECs to produce chemokines and inflammatory cytokines and upregulate CAM expression. Activation of CD8 + T cells results in the expression of chemokine receptors, including CXCR3 and CCR5. Subsequently, chemokine receptors bind to chemokine ligands expressed by ECs to induce CD8 + T-cell migration and infiltration into brain ECs. CD11a (LFA-1) on CD8 + T cells promotes adhesion to endothelial ICAM-1 ( Howland et al., 2015 ; Dunst et al., 2017 ), and upregulated expression of CAMs induces increased recruitment of iRBCs, WBCs, and platelets in brain capillaries, which enhances cerebral microvascular sequestration ( McEver, 2001 ; Shikani et al., 2012 ). The rupture of iRBCs releases merozoites, which are endocytosed by ECs and then cross-presented on major histocompatibility complex class 1 (MHC-1) molecules. MHC-1 binds to antigen receptors (TCRs) on effector CD8 + T cells to activate CD8 + T cells ( Howland et al., 2013 ). Activated CD8 + T cells release perforin, granzyme-B, and chemokines, triggering NK cells and macrophages to migrate toward the brain. Immune cell accumulation and perforin release induce apoptotic signaling in ECs and alter the tight junctions of ECs, resulting in EC dysfunction and increased cerebral vascular permeability ( Yañez et al., 1996 ; Belnoue et al., 2002 ; Haque et al., 2011 ). Disruption of BBB integrity frequently results in perivascular space enlargement, edema formation, and increased intracranial pressure, eventually resulting in death.

Activation of vascular ECs

Activation of microvascular ECs is a central component of brain microvascular pathology, resulting from the sequestration of iRBCs on the surface of vascular ECs and systematic release of inflammatory cytokines ( Siddiqui et al., 2020 ). Activated ECs are well characterized by aggravation of brain microvascular sequestration, breakdown of tight junctions, and initiation of coagulation cascading reactions.

EPCR, a host receptor involved in anticoagulation and endothelial protection, has been identified as a receptor of PfEMP1 ( Turner et al., 2013 ). It is speculated that EPCR mediates iRBCs sequestration and participates in thrombin-induced disruption of the BBB. EPCR plays a crucial role in stabilizing ECs by activating activated protein C, an inhibitor of thrombin production that prevents EC activation ( Mohan Rao et al., 2014 ). In CM, some variants of the Plasmodium adhesins PfEMP-1 (called DC8 and DC13) preferentially bind to EPCR. Upon binding to EPCR, iRBCs reduce the level of available EPCR binding sites and block the activation of activated protein C by EPCR ( Shabani et al., 2017 ). Induction of the coagulation pathway by reducing the synthesis of EPCR and activated protein C leads to increased thrombin production and EC activation, as well as decreased protective effects of ECs.

Platelets are considered effector cells of the hemostasis system and contribute to CM. It is actively involved in sequestration, inflammation, and coagulation dysfunction and is identified as their joint point ( Cox and McConkey, 2010 ). Platelets bind to iRBCs (agglutination) and ECs via adhesion receptors (CD36, ICAM-1, P-selectin). In addition, platelets promote immune activation by binding Toll-like receptors to parasite-derived molecules, expressing chemokine receptors, and releasing cytokines, chemokines, and other immunomodulatory molecules. All these activated cells (ECs, platelets, monocytes) release microparticles (TNF-α, IFN-γ) ( Combes et al., 2004 ). Taken together, microparticles alter EC functions and are regarded as proinflammatory factors and cellular activation markers.

BBB disruption

The BBB is a semipermeable membrane that separates the peripheral blood from the cerebral parenchyma and maintains balance by protecting the brain from potentially harmful blood pathogens and chemicals. The BBB consists of the microvascular endothelium, pericytes, microglia, astrocyte end-feet, neurons, and basement membrane. Microvascular ECs have tight junctions that impede the passive paracellular diffusion of small and large molecules ( Abbott et al., 2010 ; Moura et al., 2017 ).

Binding of PfEMP1 to receptors on ECs, including ICAM-1, VCAM-1, and EPCR, may trigger multiple signaling pathways in ECs, leading to reorganization of the tight junction complex and ultimately resulting in BBB leakage. ICAM-1 induces endothelial cytoskeletal remodeling via Rho-dependent phosphorylation of cytoskeleton-associated proteins, including FAK, paxillin, p130Cas, and cortactin, thereby promoting BBB opening ( Wittchen, 2009 ). In addition, VCAM-1 cross-linking results in the activation of Rac1 signaling, which induces the attenuation of tight junctions through Rho-dependent induction of stress fibers. Binding of PfEMP1 to EPCR fosters activation of tissue factors Va and VIIIa, thereby disrupting the anticoagulant pathway. Activation of these tissue factors results in thrombin generation, leading to fibrin deposition. In addition, PfEMP1 binding to EPCR activates the Rho A and NF-κB pathways through thrombin-mediated cleavage of PAR1, which induces a proinflammatory response, leading to BBB disruption ( Bernabeu and Smith, 2017 ; Kessler et al., 2017 ). Microglia also disrupt the BBB by producing TNF and IL-1β. Adhesion of iRBCs, leukocytes, and platelets to ECs also causes EC damage and irreversible changes ( Nishanth and Schlüter, 2019 ). iRBCs stimulate leukocytes (monocytes, NK cells) to release inflammatory cytokines (TNF-α, IL-1α, IL-1β) by releasing parasitic toxins ( Medana and Turner, 2006 ; Nishanth and Schlüter, 2019 ). TNF-α upregulates miR-155 expression in ECs, leading to dysfunction of BBB integrity by altering tight junctions ( Barker et al., 2017 ). IL-1α and IL-1β activate ECs to release the chemokines CCL2, CCL4, CXCL8, and CXCL10, which promote leukocyte accumulation ( Dunst et al., 2017 ), and infiltrated leukocytes induce EC apoptosis through granzyme-B and perforin-mediated cytotoxicity ( Rénia et al., 2012 ). CD8 + T cells directly induce neuronal cell death through their cytotoxic function and activation of neurons. Due to increased BBB permeability, cytokines, chemokines, immune cells, and plasma factors enter the brain parenchyma and activate neurons and astrocytes, resulting in nerve injury and neurological sequelae ( Schiess et al., 2020 ). Kynurenic acid produced by macrophages and ECs during tryptophan metabolism is further converted into cytotoxic quinoline ( Bosco et al., 2003 ; Medana et al., 2003 ; Guillemin, 2012 ). All othese molecules induce disruption of the BBB ( Figure 2 ).

An external file that holds a picture, illustration, etc.
Object name is fcimb-12-939532-g002.jpg

Molecular mechanisms of blood–brain barrier dysfunction. The binding of P. falciparum erythrocyte membrane protein-1 (PfEMP-1) to the receptors on the ECs, including ICAM-1, VCAM-1, and EPCR, may trigger multiple signaling pathways in ECs, leading to the change to cytoskeleton-associated proteins, ultimately resulting in the disruption of the BBB. Meanwhile, signaling pathways triggered by PfEMP1 lead to activation and injury of astrocytes, microglia, neurons, and perivascular macrophages and initiate the process of neuropathological injury. The binding of PfEMP1 to EPCR fosters the activation of tissue factors Va and VIIIa, thereby disrupting the anticoagulant pathway. Activation of these tissue factors results in thrombin generation, leading to fibrin deposition. Microglia also disrupt the BBB by producing TNF and IL-1β. Astrocytes retract their end feet from ECs, resulting in reduced vascular wrapping. Angiopoietin-2 produced by ECs also leads to reduced vascular wrapping by inducing pericyte dysfunction. The iRBCs stimulate leukocytes to release inflammatory cytokines (TNF-α, IL-1α, IL-1β) by releasing parasitic toxins. These cytokines disrupt BBB integrity by altering tight junctions and activating ECs to release chemokines (CCL2, CCL4, CXCL4, CXCL8, and CXCL10), which promote leukocyte accumulation, including CD4 + T cells and CD8 + T cells. Infiltrated leukocytes induce EC apoptosis through granzyme B and perforin-mediated cytotoxicity. Granzyme B and perforin directly induce neuronal cell death. Adhesion of iRBCs, leukocytes, and platelets to ECs also causes EC damage and irreversible changes. Due to the increased permeability of the BBB, cytokines, chemokines, immune cells, and plasma factors flood into the brain parenchyma and activate neurons and astrocytes, resulting in nerve injury and neurological sequelae. Kynurenic acid produced by macrophages and ECs during tryptophan metabolism is further converted into cytotoxic quinoline, which plays a vital role in stromal cells and microglia. These molecules induce the disruption of the BBB.

Recently, multiomics platforms, including genomics, transcriptomics, proteomics and metabolomics, have been widely used to unravel the underlying pathogenesis of cancer and design therapeutic strategies ( Nam et al., 2021 ). To date, there has been no combined use of multiomics approaches for CM studies, which has inspired the joint analysis of individual omics data. Analysis of DNA markers, RNA transcripts, proteins, and metabolites generated during the progression of CM contributes to understanding CM pathogenesis, which facilitates the precise diagnosis of CM and the discovery of novel therapeutic targets.

Diagnosis of CM

Diagnosis is central to malaria control, and early diagnosis is one of the crucial factors affecting the prognosis of CM. Unfortunately, there is no gold standard for the diagnosis of CM because of its complex and nonspecific clinical manifestations. Currently, the primary clinical symptoms that are available for CM diagnosis include (1) nonarousal coma (no local responses to pain) that persists for more than six hours after experiencing a generalized convulsion; (2) presence of asexual forms of P. falciparum on both thick and thin blood smears; and (3) exclusion of other causes of encephalopathy. To improve the accuracy of CM diagnosis, state-of-the-art cerebral imaging tools are available to assist the diagnosis of CM ( Table 2 ).

Table 2

Advantages and disadvantages of different approaches for the diagnosis of cerebral malaria.

Diagnostic approachesAdvantages and disadvantages
Imaging approachesMalaria retinopathyFundoscopyAdvantage: relatively low cost and simple, accurate distinction between malarial and nonmalarial comas ( ; ).
Disadvantage: requiring trained ophthalmologists and expensive equipment, subject to environmental conditions ( ).
Optical coherence tomography (OCT)Advantage: requiring qualitative and quantitative evaluation, noninvasive nature, and high-resolution output ( ).
Disadvantage: High cost as well as practical issues ( ).
TeleophthalmologyInexpensive, portable, require little additional training, and suitable for bedside patients in a variety of settings ( ).
Fluorescein fundus angiography (FFA)Advantage: Reflect the integrity of retinal blood perfusion and blood–retinal barrier by intraretinal fluorescein, and high-resolution digital imaging ( ).
Disadvantage: Large size, bulky and inconvenient to use ( ).
Electroencephalography (EEG) and Micro-EEGEEGAdvantage: Useful, noninvasive, and relatively inexpensive diagnostic tests make it possible to detect delayed cerebral malaria sequelae ( ). EEG abnormalities in cerebral malaria patients are manifested by diffuse slowing, atypical sleep elements (fusiform and parietal waves), and epileptiform activity ( ).
Disadvantage: Require continuous postdischarge follow-up assessment.
Micro-EEGMiniature, portable, easier continuous recording after patient discharge ( ),
OtherMagnetic resonance imaging (MRI) ( ), computed tomography (CT) ( ), intravital microscopy (IVM) ( ), and bioluminescent imaging devices ( ).
BiomarkersHigh levels of soluble ICAM-1 ( ), decreased Ang-1 and increased Ang-2 and Ang-2/Ang-1 ( ; ), the elevation of specific smooth muscle proteins in plasma, including carbonic anhydrase III (CA3), creatine kinase (CK), creatine kinase muscle (CKM), and myoglobin (MB) ( ), enhanced plasma levels of CXCL10 and CXCL4 ( ).
Hsa-miR-3158-3p represents a promising biomarker candidate for CM prognosis ( ) and the relative expression levels of miR-19a-3p, miR-19b-3p, miR-146a, miR-193b, miR-467a, miR-27a, and miR-146a may be associated with CM ( ; ; ).

Malarial retinopathy

The presence of malarial retinopathy facilitates the improvement in the specificity for the clinical diagnosis of CM and offers strong evidence for CM diagnosis in both adults and children ( MacCormick et al., 2014 ). In pediatric patients, the degree of retinal microcirculation is comparable to that of the brain, making it an easily observable surrogate marker to assess the severity of cerebral pathology during CM ( Bearden, 2012 ). It has been shown that malarial retinopathy presents 100% specificity and 95% sensitivity for the detection of CM, with autopsy as the diagnostic gold standard ( Beare et al., 2006 ).

Fundoscopy is a relatively low-cost and simple technique for the detection of retinopathy, which allows accurate differentiation between malarial and nonmalarial comas. The diagnosis of malarial retinopathy depends on the presence of peripheral retinal whitening, orange and white discoloration of retinal vessels, white-centered hemorrhages, and mild papilledema. The unique retinopathy of patchy retinal whitening and focal changes in vascular color are highly specific for CM diagnosis ( Beare et al., 2006 ; MacCormick et al., 2014 ). In addition, retinal hemorrhage is a common but less distinctive feature, while papilledema is not specific to CM and is unavailable for CM diagnosis alone.

Optical coherence tomography

OCT is an in vivo imaging tool that detects retinal changes and is feasible for qualitatively and quantitatively evaluating high-resolution cross-sectional retinal images, papilla of the optic nerve, and even retinal nerve fiber layer thickness ( Spaide et al., 2018 ). OCT is a noninvasive, high-resolution measure; however, this technique fails to diagnose malarial retinopathy.

Teleophthalmology

The introduction of fundoscopy improves the accuracy of CM diagnosis; however, it requires well-trained ophthalmologists and expensive equipment, which restrains its applications in resource-limited settings ( Abu Sayeed et al., 2011 ). To overcome these problems, an innovative approach, teleophthalmology, has emerged for retinal assessment ( Salongcay and Silva, 2018 ). This technique uses a simple and inexpensive portable fundus camera to capture images by well-trained professionals, and then, the images are transferred to ophthalmologists for rapid diagnosis. Teleophthalmology requires little additional training, minimizes healthcare-seeking inconvenience and is feasible in various settings ( Maude et al., 2011 ).

Fluorescein fundus angiography

With improvements in optical technology and high-resolution digital imaging, FFA has been extensively used by ophthalmologists across the world. FFA measures the integrity of retinal blood perfusion and the blood–retinal barrier by observing a map of the intraretinal fluorescein. CM patients have nonperfusion in the central retina and extensive nonperfusion in the peripheral retina ( Glover et al., 2010 ). Nevertheless, FFA requires a bulky tabletop retinal camera, whose weight and stillness make it difficult to capture clear images from conscious CM patients.

EEG and micro-EEG

EEG pulses are recorded by measuring voltage fluctuations caused by ionic currents within the neural tissues. This noninvasive technique has made it possible to detect delayed CM sequelae ( Sahu et al., 2015 ), including neurological disorders such as status epilepticus. CM patients’ EEG abnormalities manifest as diffuse slowing, atypical sleep elements (fusiform and parietal waves), and epileptiform activity ( Postels et al., 2018 ).

Although EEG is a noninvasive and relatively inexpensive diagnostic method, a significant limitation is continuous follow-up assessment of brain activity after discharge from the hospital. To address this concern, micro-EEG, a miniature, wireless, and battery-powered portable headset, was developed, and this device achieved a comparable accuracy for the diagnosis of status epilepticus with standard EEG systems ( Grant et al., 2014 ). This new tool facilitates the recording of brain activity after discharge from the hospital and may provide an option for CM diagnosis.

In addition, other imaging tools, including magnetic resonance imaging (MRI) ( Grant et al., 2014 ; Sahu et al., 2021 ), computed tomography (CT) ( Mohanty et al., 2011 ; Sahu et al., 2021 ), intravital microscopy (IVM) ( Volz, 2013 ), and in vivo bioluminescent imaging ( Franke-Fayard et al., 2006 ), may serve as additional diagnostic approaches for CM.

In addition to imaging tools, biomarkers have been extensively used for the rapid diagnosis of CM. Soluble ICAM-1, which is strongly associated with CM, was reported to be upregulated in the brain microvasculature ( Ramos et al., 2013 ). The soluble EPCR (sEPCR) level at admission is positively correlated with CM and malaria-related mortality, and admission sEPCR was identified as an early biomarker of prognosis among CM patients ( Ramos et al., 2013 ). Angiopoietin-1 (Ang-1) and Ang-2 have been characterized as mediators of endothelial activation and integrity, and Ang-1 maintains vascular quiescence, while Ang-2 displaces Ang-1 upon endothelial activation and sensitizes cells to become responsive to subthreshold concentrations of TNF. Reduced Ang-1 and Ang-2 and increased Ang-2/Ang-1 are detected in patients with CM ( Conroy et al., 2012 ; Eisenhut, 2012 ), which is consistent with the pathophysiological changes of activation and dysfunction of ECs among CM patients. In addition, elevation of specific plasma smooth muscle proteins, including carbonic anhydrase III (CA3), creatine kinase (CK), creatine kinase muscle (CKM), and myoglobin (MB), indicates muscular damage and microvasculature lesions during CM ( Bachmann et al., 2014 ). These proteins may serve as novel biomarkers for predicting CM severity and therapeutic targets for CM.

Previous reports have demonstrated that the expression of circulating microRNAs (miRNAs) is highly sensitive to physiological and pathological stimuli ( Paul et al., 2018 ). As a consequence, their changes in response to P. falciparum infection raise the possibility of new diagnostic and potentially prognostic tools for CM. Hsa-miR-3158-3p was found to be effective for the diagnosis of severe/cerebral malaria across all age groups, and hsa-miR-3158-3p represents a promising biomarker candidate for predicting CM prognosis in all age groups ( Gupta et al., 2021 ). In addition, previous studies have shown associations of the relative expression of miR-19a-3p, miR-19b-3p, miR-146a, miR-193b, miR-467a, miR-27a, and miR-146a with CM ( Martin-Alonso et al., 2018 ; Wah et al., 2019 ; Assis et al., 2020 ).

Spatial metabolomics is an emerging omics tool that provides precise determination of species, contents, and differential spatial distributions of metabolites in animal/plant tissues ( Martin-Alonso et al., 2018 ; Geier et al., 2020 ). In the ECM, both kidney and spleen metabolism are differentially perturbed in CM compared with noncerebral malaria, and lipid metabolism and the TCA cycle are altered in the kidney and spleen ( Ghosh et al., 2012 ). Spatial metabolomics is beneficial for the diagnosis, biomarker discovery, and prognosis prediction of CM.

Antimalarial therapy

Early standard antimalarial treatment is crucial for CM. In 2011, parenteral artesunate was recommended as the first-line treatment for CM by the World Health Organization (WHO). Although artesunate is effective in clearing malaria parasites, administration with potent artemisinin derivatives alone is insufficient to protect against cell death, nerve damage, and cognitive impairment ( Brejt and Golightly, 2019 ), and the long-term and widespread use of artemisinins alone may lead to the emergence of drug-resistant strains. Artemisinin-based combination therapies (ACTs) are therefore introduced to improve clinical outcomes, reduce mortality, prevent long-term neurocognitive deficits and delay the emergence of artemisinin resistance.

Potential adjuvant therapy

Targeting a single signaling pathway may be insufficient to reduce mortality or improve neurological conditions among CM patients, since CM is a multiprocess disorder. Therefore, adjuvant therapy targeting multiple physiological processes of CM is needed to improve clinical outcomes, prolong survival, and reduce neurological damage in survivors ( John et al., 2010 ). Adjuvant therapy aims to decrease cytoadherence and sequestration, modulate immune responses and improve endothelial functions, with neuroprotection given as a priority, and previous studies have shown the effectiveness of adjuvant therapy in reducing mortality due to CM in ECM models ( Wei et al., 2022 ). However, the results from clinical trials are disappointing.

Targeting parasite adhesion to vascular endothelium

Clinical episodes of CM are associated with the expression of var genes encoding the specific PfEMP1 protein, while Var genes are independently observed to bind to the brain endothelium in vitro ( Avril et al., 2012 ; Claessens et al., 2012 ). Once the crucial var ligand and its endothelial receptors are identified, high-throughput screening may be used to identify small molecules that block the binding or activation of microvascular endothelium by iRBCs. Levamisole was found to interrupt CD36-dependent binding by inhibiting CD36 dephosphorylation, which is required for high-affinity binding ( Miller et al., 2013 ). It is therefore suggested that blockade of malaria parasite adhesion to the vascular endothelium may be a promising strategy for CM treatment.

Regulating immune responses

Preventive measures prior to malaria may alter the immune system status and delay CM development; therefore, adjuvant therapy targeting immune regulation is difficult. Previous animal studies have identified modulators of host targets as potential adjuvant therapies, opening up new avenues for developing highly selective adjuvant therapies for CM. Targeting mammalian targets of rapamycin (mTOR) with rapamycin has been proven to be effective in suppressing immune responses ( Mejia et al., 2015 ), thus supporting the potential of rapamycin as an adjuvant treatment for CM. 6-Diazo-5-oxo-L-norleucine (DON), a glutamine analog, was found to block the glutaminase-mediated conversion of glutamine to glutamate, thereby inhibiting T-cell activation ( Crunkhorn, 2015 ), and administration of DON resulted in survival from CM and brain recovery in ECM ( Gordon et al., 2015 ). These data demonstrate that regulation of immune balance may be effective for CM treatment.

Improving endothelial functions and maintaining endothelial barrier integrity

Several therapeutics have been found to target endothelial dysfunction, including a platelet-activating factor receptor antagonist ( Lacerda-Queiroz et al., 2012 ), statins such as atorvastatin ( Souraud et al., 2012 ) and lovastatin ( Reis et al., 2012 ), activated protein C ( Mohan Rao et al., 2014 ), and erythropoietin ( Kaiser et al., 2006 ). In addition, Ang protein was reported to regulate endothelial barrier integrity and is associated with CM-induced retinopathy and death ( Conroy et al., 2012 ). In response to TNF stimulation, Ang-2 causes destruction of endothelial barrier integrity and triggers endothelial adhesion molecule expression. Secretion of Ang-2 in endothelial Weibel-Palade bodies may lead to vascular leakage, inflammation, and encephaledema associated with CM. Endothelium-targeted therapy that inhibits Weibel-Palade extracellular secretion may block the pathogenic autocrine activity of Ang-2 ( Yeo et al., 2008 ).

Neuroprotection

CM is a severe neurological syndrome that may cause epilepsy, coma and death, and survivors may present with neurological and cognitive deficits. Protection of nerve cells is therefore highly essential. Among the potential neuroprotective agents, erythropoietin (EPO) is one of the most promising. In addition to stimulating erythropoiesis, EPO has neuroprotective functions and increases the stability of endothelial barriers ( Ghezzi and Brines, 2004 ; Maiese et al., 2005 ). Artesunate plus recombinant human erythropoietin (rhEPO) has been found to reduce endothelial activation and improve BBB integrity in murine ECM models, resulting in faster recovery, increased survival rates, and high neuroprotective effects ( Du et al., 2017 ). Administration of peroxisome proliferator-activated receptor-gamma (PPARγ) has been proven to improve long-term cognitive ability and prolong survival ( Serghides et al., 2014 ). In addition, PPARγ has shown neuroprotective effects via various pathways and promotes neuronal repair, making it an attractive adjuvant therapy. Dysregulation of the limk-1/cofilin-1 pathway might lead to alterations in neuronal morphology and is considered the cause of cognitive defects in patients surviving CM ( Simhadri et al., 2017 ); therefore, the LIMK-1/cofilin-1 pathway is considered a potential therapeutic target for CM. In addition, granzyme-B produced by CD8 + T cells directly kills neurons through cytotoxic function and activation of caspase-3 and calpain1 ( Kaminski et al., 2019 ). Therefore, targeting granzyme-B may be an option to prevent neuronal cell death.

Unfortunately, the clinical efficacy and safety of these adjuvant treatments have not been tested until now. Inclusion of specific PfEMP-1 receptors on the surface of iRBCs may allow its connection with T cells to yield the ability to kill iRBCs, thus inhibiting the downstream pathological reactions initiated by iRBC adhesion. Chimeric antigen receptor T (CAR-T) immune cell therapy is a breakthrough for cancer therapy ( Herzig et al., 2019 ; Bertoletti and Tan, 2020 ). Since iRBC adhesion is the initial step during the development of CM, the efficacy and safety of CAR-T immune cell therapy for CM deserve further investigation.

Conclusions and perspectives

CM is a multifactorial and multiprocess disorder. Administration of antimalarials alone is effective in clearing malaria parasites; however, such a treatment fails to protect against nerve cell death, neurological damage and cognitive impairment. This urges the development of novel treatment for improved outcomes of CM. In addition, the rapid developments of -omics offer an opportunity for understanding the etiology of CM and provide insights into the clinical diagnosis and therapy of this potentially fatal disorder.

Author contributions

JL, HD, and WWe conceived and designed the study. XS, WW, WC, HZ, WWa, HD, and JL wrote the paper. All the authors read and approved the final manuscript.

This study was supported by the National Natural Science Foundation of China (Grant Number 81802046), the Principle Investigator Program of Hubei University of Medicine (Grant Number HBMUPI202101) and the Advantages Discipline Group (Public health) Project in Higher Education of Hubei Province (2022PHXKQ1).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

  • Abbott N. J., Patabendige A. A., Dolman D. E., Yusof S. R., Begley D. J. (2010). Structure and function of the blood-brain barrier . Neurobiol. Dis. 37 ( 1 ), 13–25. doi:  10.1016/j.nbd.2009.07.030 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Abu Sayeed A., Maude R. J., Hasan M. U., Mohammed N., Hoque M. G., Dondorp A. M., et al.. (2011). Malarial retinopathy in Bangladeshi adults . Am. J. Trop. Med. Hyg. 84 ( 1 ), 141–147. doi:  10.4269/ajtmh.2011.10-0205 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Assis P. A., Fernandes Durso D., Chacon Cavalcante F., Zaniratto R., Carvalho-Silva A. C., Cunha-Neto E., et al.. (2020). Integrative analysis of microRNA and mRNA expression profiles of monocyte-derived dendritic cells differentiation during experimental cerebral malaria . J. Leukoc. Biol. 108 ( 4 ), 1183–1197. doi:  10.1002/jlb.1ma0320-731r [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Avril M., Tripathi A. K., Brazier A. J., Andisi C., Janes J. H., Soma V. L., et al.. (2012). A restricted subset of var genes mediates adherence of plasmodium falciparum-infected erythrocytes to brain endothelial cells . Proc. Natl. Acad. Sci. U. S. A. 109 ( 26 ), E1782–E1790. doi:  10.1073/pnas.1120534109 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bachmann J., Burté F., Pramana S., Conte I., Brown B. J., Orimadegun A. E., et al.. (2014). Affinity proteomics reveals elevated muscle proteins in plasma of children with cerebral malaria . PLoS Pathog. 10 ( 4 ), e1004038. doi:  10.1371/journal.ppat.1004038 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Barker K. R., Lu Z., Kim H., Zheng Y., Chen J., Conroy A. L., et al.. (2017). miR-155 modifies inflammation, endothelial activation and blood-brain barrier dysfunction in cerebral malaria . Mol. Med. 23 , 24–33. doi:  10.2119/molmed.2016.00139 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bearden D. (2012). Retinopathy in cerebral malaria: new answers, new puzzles . Neurology. 79 ( 12 ), 1196–1197. doi:  10.1212/WNL.0b013e31826aad9d [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Beare N. A., Lewallen S., Taylor T. E., Molyneux M. E. (2011). Redefining cerebral malaria by including malaria retinopathy . Future Microbiol. 6 ( 3 ), 349–355. doi:  10.2217/fmb.11.3 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Beare N. A., Southern C., Chalira C., Taylor T. E., Molyneux M. E., Harding S. P. (2004). Prognostic significance and course of retinopathy in children with severe malaria . Arch. Ophthalmol. 122 ( 8 ), 1141–1147. doi:  10.1001/archopht.122.8.1141 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Beare N. A., Taylor T. E., Harding S. P., Lewallen S., Molyneux M. E. (2006). Malarial retinopathy: a newly established diagnostic sign in severe malaria . Am. J. Trop. Med. Hyg. 75 ( 5 ), 790–797. doi: 10.4269/ajtmh.2006.75.790 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Belnoue E., Kayibanda M., Vigario A. M., Deschemin J. C., van Rooijen N., Viguier M., et al.. (2002). On the pathogenic role of brain-sequestered alphabeta CD8+ T cells in experimental cerebral malaria . J. Immunol. 169 ( 11 ), 6369–6375. doi:  10.4049/jimmunol.169.11.6369 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Berendt A. R., Simmons D. L., Tansey J., Newbold C. I., Marsh K. (1989). Intercellular adhesion molecule-1 is an endothelial cell adhesion receptor for plasmodium falciparum . Nature. 341 ( 6237 ), 57–59. doi:  10.1038/341057a0 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bernabeu M., Smith J. D. (2017). EPCR and malaria severity: The center of a perfect storm . Trends Parasitol. 33 ( 4 ), 295–308. doi:  10.1016/j.pt.2016.11.004 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bertoletti A., Tan A. T. (2020). Challenges of CAR- and TCR-T cell-based therapy for chronic infections . J. Exp. Med. 217 ( 5 ), e20191663. doi:  10.1084/jem.20191663. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Birbeck G. L., Molyneux M. E., Kaplan P. W., Seydel K. B., Chimalizeni Y. F., Kawaza K., et al.. (2010). Blantyre Malaria project epilepsy study (BMPES) of neurological outcomes in retinopathy-positive paediatric cerebral malaria survivors: a prospective cohort study . Lancet Neurol. 9 ( 12 ), 1173–1181. doi:  10.1016/s1474-4422(10)70270-2 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bosco M. C., Rapisarda A., Reffo G., Massazza S., Pastorino S., Varesio L. (2003). Macrophage activating properties of the tryptophan catabolite picolinic acid . Adv. Exp. Med. Biol. 527 , 55–65. doi:  10.1007/978-1-4615-0135-0_6 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Brejt J. A., Golightly L. M. (2019). Severe malaria: update on pathophysiology and treatment . Curr. Opin. Infect. Dis. 32 ( 5 ), 413–418. doi:  10.1097/qco.0000000000000584 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Brewster D. R., Kwiatkowski D., White N. J. (1990). Neurological sequelae of cerebral malaria in children . Lancet. 336 ( 8722 ), 1039–1043. doi:  10.1016/0140-6736(90)92498-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Claessens A., Adams Y., Ghumra A., Lindergard G., Buchan C. C., Andisi C., et al.. (2012). A subset of group a-like var genes encodes the malaria parasite ligands for binding to human brain endothelial cells . Proc. Natl. Acad. Sci. U. S. A. 109 ( 26 ), E1772–E1781. doi:  10.1073/pnas.1120461109 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Combes V., Taylor T. E., Juhan-Vague I., Mège J. L., Mwenechanya J., Tembo M., et al.. (2004). Circulating endothelial microparticles in malawian children with severe falciparum malaria complicated with coma . JAMA 291 ( 21 ), 2542–2544. doi:  10.1001/jama.291.21.2542-b [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Conroy A. L., Glover S. J., Hawkes M., Erdman L. K., Seydel K. B., Taylor T. E., et al.. (2012). Angiopoietin-2 levels are associated with retinopathy and predict mortality in Malawian children with cerebral malaria: a retrospective case-control study* . Crit. Care Med. 40 ( 3 ), 952–959. doi:  10.1097/CCM.0b013e3182373157 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cox D., McConkey S. (2010). The role of platelets in the pathogenesis of cerebral malaria . Cell Mol. Life Sci. 67 ( 4 ), 557–568. doi:  10.1007/s00018-009-0211-3 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Crawley J., Smith S., Kirkham F., Muthinji P., Waruiru C., Marsh K. (1996). Seizures and status epilepticus in childhood cerebral malaria . Qjm. 89 ( 8 ), 591–597. doi:  10.1093/qjmed/89.8.591 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Crunkhorn S. (2015). Infectious disease: Glutamine analogue reverses cerebral malaria . Nat. Rev. Drug Discovery 14 ( 12 ), 820. doi:  10.1038/nrd4786 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Day N. P., Phu N. H., Mai N. T., Chau T. T., Loc P. P., Chuong L. V., et al.. (2000). The pathophysiologic and prognostic significance of acidosis in severe adult malaria . Crit. Care Med. 28 ( 6 ), 1833–1840. doi:  10.1097/00003246-200006000-00025 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Du Y., Chen G., Zhang X., Yu C., Cao Y., Cui L. (2017). Artesunate and erythropoietin synergistically improve the outcome of experimental cerebral malaria . Int. Immunopharmacol. 48 , 219–230. doi:  10.1016/j.intimp.2017.05.008 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dunst J., Kamena F., Matuschewski K. (2017). Cytokines and chemokines in cerebral malaria pathogenesis . Front. Cell Infect. Microbiol. 7 . doi:  10.3389/fcimb.2017.00324 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Eisenhut M. (2012). Low angiopoietin-1 as a predisposing factor for cerebral vasospasm in cerebral malaria . Crit. Care Med. 40 ( 12 ), 3333–3334. doi:  10.1097/CCM.0b013e318267a8d4 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • English M., Sauerwein R., Waruiru C., Mosobo M., Obiero J., Lowe B., et al.. (1997). Acidosis in severe childhood malaria . Qjm. 90 ( 4 ), 263–270. doi:  10.1093/qjmed/90.4.263 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • English M., Wale S., Binns G., Mwangi I., Sauerwein H., Marsh K. (1998). Hypoglycaemia on and after admission in Kenyan children with severe malaria . Qjm. 91 ( 3 ), 191–197. doi:  10.1093/qjmed/91.3.191 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • English M. C., Waruiru C., Lightowler C., Murphy S. A., Kirigha G., Marsh K. (1996). Hyponatraemia and dehydration in severe malaria . Arch. Dis. Child 74 ( 3 ), 201–205. doi:  10.1136/adc.74.3.201 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Essuman V. A., Ntim-Amponsah C. T., Astrup B. S., Adjei G. O., Kurtzhals J. A., Ndanu T. A., et al.. (2010). Retinopathy in severe malaria in ghanaian children–overlap between fundus changes in cerebral and non-cerebral malaria . Malar J. 9 , 232. doi:  10.1186/1475-2875-9-232 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Franke-Fayard B., Waters A. P., Janse C. J. (2006). Real-time in vivo imaging of transgenic bioluminescent blood stages of rodent malaria parasites in mice . Nat. Protoc. 1 ( 1 ), 476–485. doi:  10.1038/nprot.2006.69 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fried M., Duffy P. E. (1996). Adherence of plasmodium falciparum to chondroitin sulfate a in the human placenta . Science. 272 ( 5267 ), 1502–1504. doi:  10.1126/science.272.5267.1502 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Garg R. K., Karak B., Misra S. (1999). Neurological manifestations of malaria : an update . Neurol. India 47 ( 2 ), 85–91. [ PubMed ] [ Google Scholar ]
  • Geier B., Sogin E. M., Michellod D., Janda M., Kompauer M., Spengler B., et al.. (2020). Spatial metabolomics of in situ host-microbe interactions at the micrometre scale . Nat. Microbiol. 5 ( 3 ), 498–510. doi:  10.1038/s41564-019-0664-6 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Genton B., MP A., Mokela D. (1997). Indicators of fatal outcome in paediatric cerebral malaria: a study of 134 comatose Papua new guinean children . Int. J. Epidemiol 26 ( 3 ), 670–676. doi:  10.1093/ije/26.3.670 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ghezzi P., Brines M. (2004). Erythropoietin as an antiapoptotic, tissue-protective cytokine . Cell Death Differ. 11 Suppl 1 , S37–S44. doi:  10.1038/sj.cdd.4401450 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ghosh S., Sengupta A., Sharma S., Sonawat H. M. (2012). Metabolic fingerprints of serum, brain, and liver are distinct for mice with cerebral and noncerebral malaria: a ¹H NMR spectroscopy-based metabonomic study . J. Proteome Res. 11 ( 10 ), 4992–5004. doi:  10.1021/pr300562m [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Glover S. J., Maude R. J., Taylor T. E., Molyneux M. E., Beare N. A. (2010). Malarial retinopathy and fluorescein angiography findings in a Malawian child with cerebral malaria . Lancet Infect. Dis. 10 ( 6 ), 440. doi:  10.1016/s1473-3099(10)70073-6 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gordon E. B., Hart G. T., Tran T. M., Waisberg M., Akkaya M., Kim A. S., et al.. (2015). Targeting glutamine metabolism rescues mice from late-stage cerebral malaria . Proc. Natl. Acad. Sci. U. S. A. 112 ( 42 ), 13075–13080. doi:  10.1073/pnas.1516544112 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Grant A. C., Abdel-Baki S. G., Omurtag A., Sinert R., Chari G., Malhotra S., et al.. (2014). Diagnostic accuracy of microEEG: a miniature, wireless EEG device . Epilepsy Behav. 34 , 81–85. doi:  10.1016/j.yebeh.2014.03.015 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Guillemin G. J. (2012). Quinolinic acid, the inescapable neurotoxin . FEBS J. 279 ( 8 ), 1356–1365. doi:  10.1111/j.1742-4658.2012.08485.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gupta H., Sahu P. K., Pattnaik R., Mohanty A., Majhi M., Mohanty A. K., et al.. (2021). Plasma levels of hsa-miR-3158-3p microRNA on admission correlate with MRI findings and predict outcome in cerebral malaria . Clin. Transl. Med. 11 ( 6 ), e396. doi:  10.1002/ctm2.396 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Haque A., Best S. E., Unosson K., Amante F. H., de Labastida F., Anstey N. M., et al.. (2011). Granzyme b expression by CD8+ T cells is required for the development of experimental cerebral malaria . J. Immunol. 186 ( 11 ), 6148–6156. doi:  10.4049/jimmunol.1003955 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hawkes M., Elphinstone R. E., Conroy A. L., Kain K. C. (2013). Contrasting pediatric and adult cerebral malaria: the role of the endothelial barrier . Virulence. 4 ( 6 ), 543–555. doi:  10.4161/viru.25949 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Herzig E., Kim K. C., Packard T. A., Vardi N., Schwarzer R., Gramatica A., et al.. (2019). Attacking latent HIV with convertibleCAR-T cells, a highly adaptable killing platform . Cell. 179 ( 4 ), 880–894. doi:  10.1016/j.cell.2019.10.002 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hora R., Kapoor P., Thind K. K., Mishra P. C. (2016). Cerebral malaria–clinical manifestations and pathogenesis . Metab. Brain Dis. 31 ( 2 ), 225–237. doi:  10.1007/s11011-015-9787-5 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Howland S. W., Claser C., Poh C. M., Gun S. Y., Rénia L. (2015). Pathogenic CD8+ T cells in experimental cerebral malaria . Semin. Immunopathol. 37 ( 3 ), 221–231. doi:  10.1007/s00281-015-0476-6 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Howland S. W., Poh C. M., Gun S. Y., Claser C., Malleret B., Shastri N., et al.. (2013). Brain microvessel cross-presentation is a hallmark of experimental cerebral malaria . EMBO Mol. Med. 5 ( 7 ), 984–999. doi:  10.1002/emmm.201202273 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Idro R., Jenkins N. E., Newton C. R. (2005). Pathogenesis, clinical features, and neurological outcome of cerebral malaria . Lancet Neurol. 4 ( 12 ), 827–840. doi:  10.1016/s1474-4422(05)70247-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Idro R., Kakooza-Mwesige A., Balyejjussa S., Mirembe G., Mugasha C., Tugumisirize J., et al.. (2010). Severe neurological sequelae and behaviour problems after cerebral malaria in Ugandan children . BMC Res. Notes 3 , 104. doi:  10.1186/1756-0500-3-104 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Idro R., Karamagi C., Tumwine J. (2004). Immediate outcome and prognostic factors for cerebral malaria among children admitted to mulago hospital, Uganda . Ann. Trop. Paediatr. 24 ( 1 ), 17–24. doi:  10.1179/027249304225013240 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jensen A. R., Adams Y., Hviid L. (2020). Cerebral plasmodium falciparum malaria: The role of PfEMP1 in its pathogenesis and immunity, and PfEMP1-based vaccines to prevent it . Immunol. Rev. 293 ( 1 ), 230–252. doi:  10.1111/imr.12807 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jiang T., Huang Y., Cheng W., Sun Y., Wei W., Wu K., et al.. (2021). Multiple single-nucleotide polymorphism detection for antimalarial pyrimethamine resistance via allele-specific pcr coupled with gold nanoparticle-based lateral flow biosensor . Antimicrob. Agents Chemother. 65 ( 3 ), e01063-20. doi:  10.1128/aac.01063-20. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • John C. C., Kutamba E., Mugarura K., Opoka R. O. (2010). Adjunctive therapy for cerebral malaria and other severe forms of plasmodium falciparum malaria . Expert Rev. Anti Infect. Ther. 8 ( 9 ), 997–1008. doi:  10.1586/eri.10.90 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kaiser K., Texier A., Ferrandiz J., Buguet A., Meiller A., Latour C., et al.. (2006). Recombinant human erythropoietin prevents the death of mice during cerebral malaria . J. Infect. Dis. 193 ( 7 ), 987–995. doi:  10.1086/500844 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kaminski L. C., Riehn M., Abel A., Steeg C., Yar D. D., Addai-Mensah O., et al.. (2019). Cytotoxic T cell-derived granzyme b is increased in severe plasmodium falciparum malaria . Front. Immunol. 10 . doi:  10.3389/fimmu.2019.02917 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kaul D. K., Liu X. D., Nagel R. L., Shear H. L. (1998). Microvascular hemodynamics and in vivo evidence for the role of intercellular adhesion molecule-1 in the sequestration of infected red blood cells in a mouse model of lethal malaria . Am. J. Trop. Med. Hyg. 58 ( 2 ), 240–247. doi:  10.4269/ajtmh.1998.58.240 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kessler A., Dankwa S., Bernabeu M., Harawa V., Danziger S. A., Duffy F., et al.. (2017). Linking EPCR-binding PfEMP1 to brain swelling in pediatric cerebral malaria . Cell Host Microbe 22 ( 5 ), 601–614.e5. doi:  10.1016/j.chom.2017.09.009 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kim K. (2012). Malaria var gene expression: keeping up with the neighbors . Cell Host Microbe 11 ( 1 ), 1–2. doi:  10.1016/j.chom.2012.01.002 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kochar D. K., Shubhakaran, Kumawat B. L., Kochar S. K., Halwai M., Makkar R. K., et al.. (2002). Cerebral malaria in Indian adults: a prospective study of 441 patients from bikaner, north-west India . J. Assoc. Physicians India 50 , 234–241. [ PubMed ] [ Google Scholar ]
  • Krishnan A., Karnad D. R. (2003). Severe falciparum malaria: an important cause of multiple organ failure in Indian intensive care unit patients . Crit. Care Med. 31 ( 9 ), 2278–2284. doi:  10.1097/01.Ccm.0000079603.82822.69 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kwiatkowski D., Hill A. V., Sambou I., Twumasi P., Castracane J., Manogue K. R., et al.. (1990). TNF concentration in fatal cerebral, non-fatal cerebral, and uncomplicated plasmodium falciparum malaria . Lancet. 336 ( 8725 ), 1201–1204. doi:  10.1016/0140-6736(90)92827-5 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lacerda-Queiroz N., Rodrigues D. H., Vilela M. C., Rachid M. A., Soriani F. M., Sousa L. P., et al.. (2012). Platelet-activating factor receptor is essential for the development of experimental cerebral malaria . Am. J. Pathol. 180 ( 1 ), 246–255. doi:  10.1016/j.ajpath.2011.09.038 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lewallen S., Bronzan R. N., Beare N. A., Harding S. P., Molyneux M. E., Taylor T. E. (2008). Using malarial retinopathy to improve the classification of children with cerebral malaria . Trans. R Soc. Trop. Med. Hyg. 102 ( 11 ), 1089–1094. doi:  10.1016/j.trstmh.2008.06.014 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Luzolo A. L., Ngoyi D. M. (2019). Cerebral malaria . Brain Res. Bull. 145 , 53–58. doi:  10.1016/j.brainresbull.2019.01.010 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • MacCormick I. J., Beare N. A., Taylor T. E., Barrera V., White V. A., Hiscott P., et al.. (2014). Cerebral malaria in children: using the retina to study the brain . Brain 137 ( Pt 8 ), 2119–2142. doi:  10.1093/brain/awu001. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Maiese K., Li F., Chong Z. Z. (2005). New avenues of exploration for erythropoietin . JAMA 293 ( 1 ), 90–95. doi:  10.1001/jama.293.1.90 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Maitland K., Newton C. R. (2005). Acidosis of severe falciparum malaria: heading for a shock ? Trends Parasitol. 21 ( 1 ), 11–16. doi:  10.1016/j.pt.2004.10.010. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Martin-Alonso A., Cohen A., Quispe-Ricalde M. A., Foronda P., Benito A., Berzosa P., et al.. (2018). Differentially expressed microRNAs in experimental cerebral malaria and their involvement in endocytosis, adherens junctions, FoxO and TGF-β signalling pathways . Sci. Rep. 8 ( 1 ), 11277. doi:  10.1038/s41598-018-29721-y [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Maude R. J., Plewes K., Dimock J., Dondorp A. M. (2011). Low-cost portable fluorescein angiography . Br. J. Ophthalmol. 95 ( 9 ), 1213–1215. doi:  10.1136/bjo.2010.200576 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • McEver R. P. (2001). Adhesive interactions of leukocytes, platelets, and the vessel wall during hemostasis and inflammation . Thromb. Haemost 86 ( 3 ), 746–756. doi: 10.1055/s-0037-1616128 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Medana I. M., Day N. P., Salahifar-Sabet H., Stocker R., Smythe G., Bwanaisa L., et al.. (2003). Metabolites of the kynurenine pathway of tryptophan metabolism in the cerebrospinal fluid of Malawian children with malaria . J. Infect. Dis. 188 ( 6 ), 844–849. doi:  10.1086/377583 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Medana I. M., Turner G. D. (2006). Human cerebral malaria and the blood-brain barrier . Int. J. Parasitol. 36 ( 5 ), 555–568. doi:  10.1016/j.ijpara.2006.02.004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mejia P., Treviño-Villarreal J. H., Hine C., Harputlugil E., Lang S., Calay E., et al.. (2015). Dietary restriction protects against experimental cerebral malaria via leptin modulation and T-cell mTORC1 suppression . Nat. Commun. 6 , 6050. doi:  10.1038/ncomms7050 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Miller L. H., Ackerman H. C., Su X. Z., Wellems T. E. (2013). Malaria biology and disease pathogenesis: insights for new treatments . Nat. Med. 19 ( 2 ), 156–167. doi:  10.1038/nm.3073 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mishra S. K., Mohanty S., Satpathy S. K., Mohapatra D. N. (2007). Cerebral malaria in adults – a description of 526 cases admitted to ispat general hospital in rourkela, India . Ann. Trop. Med. Parasitol. 101 ( 3 ), 187–193. doi:  10.1179/136485907x157004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mohan Rao L. V., Esmon C. T., Pendurthi U. R. (2014). Endothelial cell protein c receptor: a multiliganded and multifunctional receptor . Blood. 124 ( 10 ), 1553–1562. doi:  10.1182/blood-2014-05-578328 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mohanty S., Benjamin L. A., Majhi M., Panda P., Kampondeni S., Sahu P. K., et al.. (2017). Magnetic resonance imaging of cerebral malaria patients reveals distinct pathogenetic processes in different parts of the brain . mSphere 2 ( 3 ), e00193-17. doi:  10.1128/mSphere.00193-17. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mohanty S., Mishra S. K., Patnaik R., Dutt A. K., Pradhan S., Das B., et al.. (2011). Brain swelling and mannitol therapy in adult cerebral malaria: a randomized trial . Clin. Infect. Dis. 53 ( 4 ), 349–355. doi:  10.1093/cid/cir405 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Molyneux M. E., Taylor T. E., Wirima J. J., Borgstein A. (1989). Clinical features and prognostic indicators in paediatric cerebral malaria: a study of 131 comatose Malawian children . Q J. Med. 71 ( 265 ), 441–459. [ PubMed ] [ Google Scholar ]
  • Moura R. P., Almeida A., Sarmento B. (2017). The role of non-endothelial cells on the penetration of nanoparticles through the blood brain barrier . Prog. Neurobiol. 159 , 39–49. doi:  10.1016/j.pneurobio.2017.09.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nam A. S., Chaligne R., Landau D. A. (2021). Integrating genetic and non-genetic determinants of cancer evolution by single-cell multi-omics . Nat. Rev. Genet. 22 ( 1 ), 3–18. doi:  10.1038/s41576-020-0265-5 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Newton C. R., Hien T. T., White N. (2000). Cerebral malaria . J. Neurol. Neurosurg. Psychiatry 69 ( 4 ), 433–441. doi:  10.1136/jnnp.69.4.433 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Newton C. R., Kirkham F. J., Winstanley P. A., Pasvol G., Peshu N., Warrell D. A., et al.. (1991). Intracranial pressure in African children with cerebral malaria . Lancet. 337 ( 8741 ), 573–576. doi:  10.1016/0140-6736(91)91638-b [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • NIH . (2014). Severe malaria . Trop. Med. Int. Health 19 Suppl 1 , 7–131. doi:  10.1111/tmi.12313_2 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Newton C. R., Warrell D. A. (1998). Neurological manifestations of falciparum malaria . Ann. Neurol. 43 ( 6 ), 695–702. doi:  10.1002/ana.410430603 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nishanth G., Schlüter D. (2019). Blood-brain barrier in cerebral malaria: Pathogenesis and therapeutic intervention . Trends Parasitol. 35 ( 7 ), 516–528. doi:  10.1016/j.pt.2019.04.010 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ockenhouse C. F., Tandon N. N., Magowan C., Jamieson G. A., Chulay J. D. (1989). Identification of a platelet membrane glycoprotein as a falciparum malaria sequestration receptor . Science. 243 ( 4897 ), 1469–1471. doi:  10.1126/science.2467377 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ockenhouse C. F., Tegoshi T., Maeno Y., Benjamin C., Ho M., Kan K. E., et al.. (1992). Human vascular endothelial cell adhesion receptors for plasmodium falciparum-infected erythrocytes: roles for endothelial leukocyte adhesion molecule 1 and vascular cell adhesion molecule 1 . J. Exp. Med. 176 ( 4 ), 1183–1189. doi:  10.1084/jem.176.4.1183 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Olliaro P. (2008). Editorial commentary: mortality associated with severe plasmodium falciparum malaria increases with age . Clin. Infect. Dis. 47 ( 2 ), 158–160. doi:  10.1086/589288 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Oluwayemi I. O., Brown B. J., Oyedeji O. A., Oluwayemi M. A. (2013). Neurological sequelae in survivors of cerebral malaria . Pan Afr Med. J. 15 , 88. doi:  10.11604/pamj.2013.15.88.1897 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Paul P., Chakraborty A., Sarkar D., Langthasa M., Rahman M., Bari M., et al.. (2018). Interplay between miRNAs and human diseases . J. Cell Physiol. 233 ( 3 ), 2007–2018. doi:  10.1002/jcp.25854 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Peixoto B., Kalei I. (2013). Neurocognitive sequelae of cerebral malaria in adults: a pilot study in benguela central hospital, Angola . Asian Pac J. Trop. Biomed. 3 ( 7 ), 532–535. doi:  10.1016/s2221-1691(13)60108-2 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Plewes K., Turner G. D. H., Dondorp A. M. (2018). Pathophysiology, clinical presentation, and treatment of coma and acute kidney injury complicating falciparum malaria . Curr. Opin. Infect. Dis. 31 ( 1 ), 69–77. doi:  10.1097/qco.0000000000000419 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ponsford M. J., Medana I. M., Prapansilp P., Hien T. T., Lee S. J., Dondorp A. M., et al.. (2012). Sequestration and microvascular congestion are associated with coma in human cerebral malaria . J. Infect. Dis. 205 ( 4 ), 663–671. doi:  10.1093/infdis/jir812 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Postels D. G., Birbeck G. L. (2013). Cerebral malaria . Handb. Clin. Neurol. 114 , 91–102. doi:  10.1016/b978-0-444-53490-3.00006-6 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Postels D. G., Wu X., Li C., Kaplan P. W., Seydel K. B., Taylor T. E., et al.. (2018). Admission EEG findings in diverse paediatric cerebral malaria populations predict outcomes . Malar J. 17 ( 1 ), 208. doi:  10.1186/s12936-018-2355-9 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ramos T. N., Bullard D. C., Darley M. M., McDonald K., Crawford D. F., Barnum S. R. (2013). Experimental cerebral malaria develops independently of endothelial expression of intercellular adhesion molecule-1 (icam-1) . J. Biol. Chem. 288 ( 16 ), 10962–10966. doi:  10.1074/jbc.C113.457028 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Reis P. A., Estato V., da Silva T. I., d'Avila J. C., Siqueira L. D., Assis E. F., et al.. (2012). Statins decrease neuroinflammation and prevent cognitive impairment after cerebral malaria . PLoS Pathog. 8 ( 12 ), e1003099. doi:  10.1371/journal.ppat.1003099 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rénia L., Howland S. W., Claser C., Charlotte Gruner A., Suwanarusk R., Hui Teo T., et al.. (2012). Cerebral malaria: mysteries at the blood-brain barrier . Virulence. 3 ( 2 ), 193–201. doi:  10.4161/viru.19013 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Riggle B. A., Manglani M., Maric D., Johnson K. R., Lee M. H., Neto O. L. A., et al.. (2020). CD8+ T cells target cerebrovasculature in children with cerebral malaria . J. Clin. Invest 130 ( 3 ), 1128–1138. doi:  10.1172/jci133474 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rogerson S. J., Chaiyaroj S. C., Ng K., Reeder J. C., Brown G. V. (1995). Chondroitin sulfate a is a cell surface receptor for plasmodium falciparum-infected erythrocytes . J. Exp. Med. 182 ( 1 ), 15–20. doi:  10.1084/jem.182.1.15 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sahu P. K., Hoffmann A., Majhi M., Pattnaik R., Patterson C., Mahanta K. C., et al.. (2021). Brain magnetic resonance imaging reveals different courses of disease in pediatric and adult cerebral malaria . Clin. Infect. Dis. 73 ( 7 ), e2387–e2396. doi:  10.1093/cid/ciaa1647 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sahu P. K., Satpathi S., Behera P. K., Mishra S. K., Mohanty S., Wassmer S. C. (2015). Pathogenesis of cerebral malaria: new diagnostic tools, biomarkers, and therapeutic approaches . Front. Cell Infect. Microbiol. 5 . doi:  10.3389/fcimb.2015.00075 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Salongcay R. P., Silva P. S. (2018). The role of teleophthalmology in the management of diabetic retinopathy . Asia Pac J. Ophthalmol. (Phila) 7 ( 1 ), 17–21. doi:  10.22608/apo.2017479 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schiess N., Villabona-Rueda A., Cottier K. E., Huether K., Chipeta J., Stins M. F. (2020). Pathophysiology and neurologic sequelae of cerebral malaria . Malar J. 19 ( 1 ), 266. doi:  10.1186/s12936-020-03336-z [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Senanayake N. (1987). Delayed cerebellar ataxia: a new complication of falciparum malaria ? Br. Med. J. (Clin Res. Ed) 294 ( 6582 ), 1253–1254. doi:  10.1136/bmj.294.6582.1253 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Serghides L., McDonald C. R., Lu Z., Friedel M., Cui C., Ho K. T., et al.. (2014). PPARγ agonists improve survival and neurocognitive outcomes in experimental cerebral malaria and induce neuroprotective pathways in human malaria . PLoS Pathog. 10 ( 3 ), e1003980. doi:  10.1371/journal.ppat.1003980 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shabani E., Hanisch B., Opoka R. O., Lavstsen T., John C. C. (2017). Plasmodium falciparum EPCR-binding PfEMP1 expression increases with malaria disease severity and is elevated in retinopathy negative cerebral malaria . BMC Med. 15 ( 1 ), 183. doi:  10.1186/s12916-017-0945-y [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shikani H. J., Freeman B. D., Lisanti M. P., Weiss L. M., Tanowitz H. B., Desruisseaux M. S. (2012). Cerebral malaria: we have come a long way . Am. J. pathology 181 ( 5 ), 1484–1492. doi:  10.1016/j.ajpath.2012.08.010 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Siddiqui A. J., Adnan M., Jahan S., Redman W., Saeed M., Patel M. (2020). Neurological disorder and psychosocial aspects of cerebral malaria: what is new on its pathogenesis and complications? a minireview . Folia Parasitol. (Praha) 67 , 2020.015. doi:  10.14411/fp.2020.015. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Silamut K., Phu N. H., Whitty C., Turner G. D., Louwrier K., Mai N. T., et al.. (1999). A quantitative analysis of the microvascular sequestration of malaria parasites in the human brain . Am. J. Pathol. 155 ( 2 ), 395–410. doi:  10.1016/s0002-9440(10)65136-x [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Simhadri P. K., Malwade R., Vanka R., Nakka V. P., Kuppusamy G., Babu P. P. (2017). Dysregulation of LIMK-1/cofilin-1 pathway: A possible basis for alteration of neuronal morphology in experimental cerebral malaria . Ann. Neurol. 82 ( 3 ), 429–443. doi:  10.1002/ana.25028 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Solomon W., Wilson N. O., Anderson L., Pitts S., Patrickson J., Liu M., et al.. (2014). Neuregulin-1 attenuates mortality associated with experimental cerebral malaria . J. Neuroinflammation 11 , 9. doi:  10.1186/1742-2094-11-9 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Souraud J. B., Briolant S., Dormoi J., Mosnier J., Savini H., Baret E., et al.. (2012). Atorvastatin treatment is effective when used in combination with mefloquine in an experimental cerebral malaria murine model . Malar J. 11 , 13. doi:  10.1186/1475-2875-11-13 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Spaide R. F., Fujimoto J. G., Waheed N. K., Sadda S. R., Staurenghi G. (2018). Optical coherence tomography angiography . Prog. Retin Eye Res. 64 , 1–55. doi:  10.1016/j.preteyeres.2017.11.003 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Stafstrom C. E., Chronopoulos A., Thurber S., Thompson J. L., Holmes G. L. (1993). Age-dependent cognitive and behavioral deficits after kainic acid seizures . Epilepsia. 34 ( 3 ), 420–432. doi:  10.1111/j.1528-1157.1993.tb02582.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Storm J., Jespersen J. S., Seydel K. B., Szestak T., Mbewe M., Chisala N. V., et al.. (2019). Cerebral malaria is associated with differential cytoadherence to brain endothelial cells . EMBO Mol. Med. 11 ( 2 ), e9164. doi:  10.15252/emmm.201809164. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Su XZ., Wu J. (2021). Zoonotic transmission and host switches of malaria parasites . Zoonoses 1 ( 1 ), 11. doi:  10.1021/acs.nanolett.1c03514 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tembo D. L., Nyoni B., Murikoli R. V., Mukaka M., Milner D. A., Berriman M., et al.. (2014). Differential PfEMP1 expression is associated with cerebral malaria pathology . PLoS Pathog. 10 ( 12 ), e1004537. doi:  10.1371/journal.ppat.1004537 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Turner L., Lavstsen T., Berger S. S., Wang C. W., Petersen J. E., Avril M., et al.. (2013). Severe malaria is associated with parasite binding to endothelial protein c receptor . Nature. 498 ( 7455 ), 502–505. doi:  10.1038/nature12216 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Turner G. D., Morrison H., Jones M., Davis T. M., Looareesuwan S., Buley I. D., et al.. (1994). 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. 145 ( 5 ), 1057–1069. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • van Hensbroek M. B., Palmer A., Jaffar S., Schneider G., Kwiatkowski D. (1997). Residual neurologic sequelae after childhood cerebral malaria . J. Pediatr. 131 ( 1 Pt 1 ), 125–129. doi:  10.1016/s0022-3476(97)70135-5. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Vespa P. M., Nuwer M. R., Nenov V., Ronne-Engstrom E., Hovda D. A., Bergsneider M., et al.. (1999). Increased incidence and impact of nonconvulsive and convulsive seizures after traumatic brain injury as detected by continuous electroencephalographic monitoring . J. Neurosurg. 91 ( 5 ), 750–760. doi:  10.3171/jns.1999.91.5.0750 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Volz J. C. (2013). Looking through a cranial window: intravital microscopy for in vivo study of cerebral malaria . Virulence. 4 ( 8 ), 661–663. doi:  10.4161/viru.26802 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wah S. T., Hananantachai H., Patarapotikul J., Ohashi J., Naka I., Nuchnoi P. (2019). microRNA-27a and microRNA-146a SNP in cerebral malaria . Mol. Genet. Genomic Med. 7 ( 2 ), e00529. doi:  10.1002/mgg3.529 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Waller D., Krishna S., Crawley J., Miller K., Nosten F., Chapman D., et al.. (1995). Clinical features and outcome of severe malaria in Gambian children . Clin. Infect. Dis. 21 ( 3 ), 577–587. doi:  10.1093/clinids/21.3.577. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wassmer S. C., Taylor T. E., Rathod P. K., Mishra S. K., Mohanty S., Arevalo-Herrera M., et al.. (2015). Investigating the pathogenesis of severe malaria: A multidisciplinary and cross-geographical approach . Am. J. Trop. Med. Hyg. 93 ( 3 Suppl ), 42–56. doi:  10.4269/ajtmh.14-0841 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wei W., Cheng W., Dai W., Lu F., Cheng Y., Jiang T., et al.. (2022). A nanodrug coated with membrane from brain microvascular endothelial cells protects against experimental cerebral malaria . Nano Lett. 22 ( 1 ), 211–219. doi:  10.1021/acs.nanolett.1c03514 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • White V. A., Lewallen S., Beare N., Kayira K., Carr R. A., Taylor T. E. (2001). Correlation of retinal haemorrhages with brain haemorrhages in children dying of cerebral malaria in Malawi . Trans. R Soc. Trop. Med. Hyg. 95 ( 6 ), 618–621. doi:  10.1016/s0035-9203(01)90097-5 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • White N. J., Miller K. D., Marsh K., Berry C. D., Turner R. C., Williamson D. H., et al.. (1987). Hypoglycaemia in African children with severe malaria . Lancet. 1 ( 8535 ), 708–711. doi:  10.1016/s0140-6736(87)90354-0 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • White N. J., Warrell D. A., Chanthavanich P., Looareesuwan S., Warrell M. J., Krishna S., et al.. (1983). Severe hypoglycemia and hyperinsulinemia in falciparum malaria . N Engl. J. Med. 309 ( 2 ), 61–66. doi:  10.1056/nejm198307143090201 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wilson N. O., Jain V., Roberts C. E., Lucchi N., Joel P. K., Singh M. P., et al.. (2011). CXCL4 and CXCL10 predict risk of fatal cerebral malaria . Dis. Markers 30 ( 1 ), 39–49. doi:  10.3233/dma-2011-0763 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wittchen E. S. (2009). Endothelial signaling in paracellular and transcellular leukocyte transmigration . Front. Biosci. (Landmark Ed) 14 , 2522–2545. doi:  10.2741/3395 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • World Health Organization (2021). Switzerland: Geneva, World malaria report 2021 . [ Google Scholar ]
  • XP D, L S (2021). Emerging and re-emerging zoonoses are major and global challenges for public health . Zoonoses 1 , 1. doi:  10.1016/j.brainresbull.2019.01.010 [ CrossRef ] [ Google Scholar ]
  • Yañez D. M., Manning D. D., Cooley A. J., Weidanz W. P., van der Heyde H. C. (1996). Participation of lymphocyte subpopulations in the pathogenesis of experimental murine cerebral malaria . J. Immunol. 157 ( 4 ), 1620–1624. [ PubMed ] [ Google Scholar ]
  • Yeo T. W., Lampah D. A., Gitawati R., Tjitra E., Kenangalem E., Piera K., et al.. (2008). Angiopoietin-2 is associated with decreased endothelial nitric oxide and poor clinical outcome in severe falciparum malaria . Proc. Natl. Acad. Sci. U. S. A. 105 ( 44 ), 17097–17102. doi:  10.1073/pnas.0805782105 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Open access
  • Published: 23 August 2024

Longitudinal associations of plasma amino acid levels with recovery from malarial coma

  • Donald L. Granger 1 , 2 ,
  • Daniel Ansong 3 ,
  • Tsiri Agbenyega 3 ,
  • Melinda S. Liddle 4 ,
  • Benjamin A. Brinton 5 ,
  • Devon C. Hale 1 ,
  • Bert K. Lopansri 4 ,
  • Richard Reithinger 6 &
  • Donal Bisanzio 6  

Malaria Journal volume  23 , Article number:  253 ( 2024 ) Cite this article

70 Accesses

1 Altmetric

Metrics details

Disordered amino acid metabolism is observed in cerebral malaria (CM). This study sought to determine whether abnormal amino acid concentrations were associated with level of consciousness in children recovering from coma. Twenty-one amino acids and coma scores were quantified longitudinally and the data were analysed for associations.

In a prospective observational study, 42 children with CM were enrolled. Amino acid levels were measured at entry and at frequent intervals thereafter and consciousness was assessed by Blantyre Coma Scores (BCS). Thirty-six healthy children served as controls for in-country normal amino acid ranges. Logistic regression was employed using a generalized linear mixed-effects model to assess associations between out-of-range amino acid levels and BCS.

At entry 16/21 amino acid levels were out-of-range. Longitudinal analysis revealed 10/21 out-of-range amino acids were significantly associated with BCS. Elevated phenylalanine levels showed the highest association with low BCS. This finding held when out-of-normal-range data were analysed at each sampling time.

Longitudinal data is provided for associations between abnormal amino acid levels and recovery from CM. Of 10 amino acids significantly associated with BCS, elevated phenylalanine may be a surrogate for impaired clearance of ether lipid mediators of inflammation and may contribute to CM pathogenesis.

Of the complications associated with Plasmodium falciparum infection, unarousable coma, the hallmark of cerebral malaria (CM), may lead to long-term disability and death [ 1 , 2 ]. Recent research aided by autopsy and imaging indicates that coma may be the result of generalized brain swelling due to accumulation of sequestered infected erythrocytes in the microvasculature [ 3 , 4 , 5 , 6 ]. Sequestration is the physical adherence of infected cells to activated endothelial cells within brain capillaries and venules. Microscopic, sequestered foci with patchy distribution restrict blood flow with downstream consequences including anoxia and acidosis. Leakage of intravascular fluid and, in some cases, erythrocytes into adjacent brain parenchyma through endothelial junctions adds to pathogenic events [ 7 , 8 ].

In addition to the anatomical pathology in the brain, there are systemic abnormalities involving carbohydrate [ 9 , 10 ] lipid [ 11 ] and amino acid [ 12 , 13 ] metabolism. None are specific for CM—indeed, they are also seen in septic syndrome and other severe inflammatory diseases [ 14 , 15 , 16 ]. Metabolic causes of coma in CM have been sought, but no convincing pathogenic mechanism has been found. Nevertheless, these changes may contribute to coma in ways not yet recognized.

Research on the role of nitric oxide in malaria pathogenesis noted abnormalities in the plasma levels of several amino acids in children presenting with CM [ 17 , 18 ]. All abnormal amino acid levels were below the normal ranges, except for one, phenylalanine, which was consistently above 80 micromolar . Healthy controls (HC) rarely exceeded this cutoff concentration. With resolution of coma in treated survivors, phenylalanine levels normalized. This study questioned whether longitudinal measurements of abnormal amino acid levels might be associated with the level of consciousness as children recovered. Such analysis may uncover an amino acid or group of amino acids involved in P. falciparum pathogenesis.

To address this question, a prospective observational study was conducted in Ghana, in which children entering hospital in coma due to malaria were observed frequently (every 12–24 h) for 60 h and their level of consciousness was quantified by Blantyre coma score(s) (BCS) [ 19 ]. At the same time intervals, blood samples were obtained and processed for 21 plasma amino acid levels. The a priori hypothesis was that the levels of one or more amino acid(s) would be most closely associated with the BCS as participants regained consciousness. A healthy control group was enrolled to determine whether amino acid levels in Ghanaian children matched reference laboratory normal ranges for healthy children in the US. The control group also provided disease versus normal comparisons at the time of presentation to hospital.

Study design and site

The study was a prospective, observational, longitudinal investigation at Komfo Anokye Teaching Hospital in Kumasi, Ashanti Region, Ghana (6.697479° N 1.631690° W) from 2004 to 2006. Enrolled were children who presented to the hospital with CM, as well as healthy control children who were asymptomatic hospital visitors or attendees at well-clinic check-ups. Children with uncomplicated malaria were not included, as the study was designed to address amino acid levels associated with recovery from coma. Study participants largely resided in Kumasi and its surrounding districts. The study was conducted by senior Ghanaian physicians in the Department of Pediatrics assisted by infectious diseases faculty, fellows, and medical students from the University of Utah. Plasma samples for amino acid analysis were obtained at enrollment (i.e., at admission) and at least every 24 h during hospitalization.

Enrollment criteria

Children were 6 months to 6 years of age (median age, 2.8). WHO case definition for CM was used as inclusion criteria [ 20 ]: (1) any level of P. falciparum parasitaemia on peripheral blood film; (2) unarousable coma as assessed by BCS ≤ 2 not attributable to hypoglycaemia (i.e., blood glucose level < 40 mg/dl); (3) coma persisting more than 60 min after any convulsion; and (4) no other identifiable cause of coma. Exclusion criteria were any of the following: (1) microscopic or culture evidence of bacterial or viral co-infection; (2) oral or intravenous quinine or oral artemisinin-based combination therapy initiated > 18 h prior to enrollment; (3) haemoglobin < 5 mg/dl when blood transfusion was unavailable at the study site.

Similar aged healthy children were prospectively enrolled as HC. Weight was not included as criterion for enrollment and therefore there is significant missing data for this variable (Table  1 ). Eligible HC had no symptoms or signs of active illness, no febrile illness within the past 2 weeks, no history or evidence of an active inflammatory condition, and negative blood film for malaria parasites. Availability of HCs for enrollment at the study site was limited and resulted in fewer than anticipated children in this group.

Clinical evaluation and management

Demographic information, clinical history, and physical examination were documented using standardized case report forms. History of last food or liquid intake for CM participants was recorded to assess potential confounding influence of protein intake on plasma amino acid concentrations. The severity of CM resulting in anorexia ruled out this possible confounder for the majority of participants. Capillary blood samples were obtained for malaria thick and thin blood films and were prepared by Giemsa staining. Venous samples for routine laboratory analysis included complete blood count, electrolytes, creatinine, and lactate. Because laboratory service was not always available, there are significant missing data (Table  2 ) for these analytes. Urine was obtained for dipstick analysis and culture. Blood and urine laboratory results were immediately available to clinicians. Blood cultures were obtained on all participants with CM. Lumbar puncture was done to investigate possible bacterial meningitis unless it was clear to the presiding clinicians that this diagnosis was unlikely. Thirty-four CM cases received lumbar puncture. Cerebrospinal fluid analysis included: (a) determination of glucose and protein concentrations; (b) cell count with differential performed by personnel trained in microscopy; (c) Gram stain and bacterial and fungal cultures.

Children with CM received anti-malarial therapy and supportive care as per standard Ghanaian Ministry of Health protocols for the years during which the study was conducted (intravenous quinine or intravenous artesunate as recommended by WHO protocols). Treatment was initiated as soon as the diagnosis of malaria was suspected.

BCS was assessed at presentation and at least every 24 h until hospital discharge or death. For some participants (year 1 of the study) BCS was measured at admission and at 12-h intervals for up to 60 h. For years 2 and 3, midnight blood draws became impractical due to transportation and safety concerns, and samples were taken every 24 h. For each child at each interval, three clinicians independently assessed BCS. The assessments were done by a Ghanaian faculty physician, a US infectious diseases faculty or fellow, and a Ghanaian or US medical student. After assessment, the three clinicians met to obtain a consensus for BCS. For some patients this required returning to the bedside to review the findings. In all cases, consensus was reached by the three examiners and one BCS was recorded. At the time of BCS assessments the clinician examiners were unaware of any data pertaining to plasma amino acid concentrations.

Plasma amino acid analysis

Blood samples were collected into heparin tubes, mixed, and immediately centrifuged to sediment blood cells. Supernatant plasma was placed into polypropylene freezer tubes and stored at − 80 °C until shipment. Samples were transported in a liquid nitrogen dry shipper to the US for amino acid analysis. All amino acid analyses for the 3-year study were performed within 1 month after collection for each year. Amino acid analyses were performed at the Biochemical Genetics Section, ARUP Laboratories, University of Utah School of Medicine in collaboration with Dr. Marzia Pasquali. The amino acid analyzer employed ion exchange chromatography for separation and quantification. With two exceptions, all plasma amino acids were quantified. Exceptions were tryptophan, which emerged from the column lastly at a long retention time unsuitable for analysis, and hydroxyproline, which gave frequent values of zero or below the level of sensitivity of the assay. Computer output results from the amino acid analyzer were electronically transferred to Excel spread sheets for subsequent data analysis.

Statistical methods

Data was compared for CM cases and healthy controls at entry (Tables 1 , 2 , 3 ). Data was analysed using GraphPad Prism version 10.1.1. Data sets for each variable were tested for normality. Parametric data (Tables 1 , 2 ) for both CM and healthy control groups were compared for significant difference using two-tailed Student’s t-test. For non-parametric data (Table  3 ) the two-tailed Mann–Whitney test was used. A significant difference was defined as P ≤ 0.05.

Association between BCS values and amino acid levels was investigated using statistical modeling. The aim was to assess if an out-of-range level of each amino acid was linked to a low BCS. Since BCS is based on an ordinal discrete scale from 0 to 5, an ordinal logistic regression, which is the most appropriate modeling technique for ordinal variables, was employed [ 21 , 22 ]. The model was formulated using a generalized linear mixed-effects model (GLM-EM) [ 23 ]. It incorporated children as a random effect to account for variations in the sampling period. The ordinal logistic regression model describes the association between the dependent variable and independent variables by estimating odds ratios (OR). Odds ratio was used to identify the association between out-of-range amino acid levels and BCS. An OR significantly below 1.0 indicated that having an amino acid level outside the normal range was associated with low BCS. An OR above 1.0 indicated that having an amino acid level within the normal range was associated with a high BCS. The model equation was the following:

where BCS was children’s BCS expressed as an ordinal variable from 0 to 5, Year was the sampling year, and Child ID random was children’s ID included as a random effect.

A logistic regression GLM-EM was also built to investigate the effect of time on the levels of amino acids [ 24 ]. This model aimed to identify the time threshold after which the levels of amino acids returned to the normal range. An OR significantly above 1.0 indicated a high probability that the amino acid level was normal at a given time, while an OR significantly below 1.0 indicated a high probability that the amino acid level was abnormal at a given time. The equation for the GLM-EM was the following:

where Amino acid level (0,1) was a dichotomous variable reporting if the amino acid level was normal (value = 1) or out of normal range (value = 0), Sampling time was the time intervals of sampling, Year was the sampling year, and Child ID random was children’s ID included as a random effect.

All statistical modelling was performed using BayesX software through R language interface [ 25 , 26 ]. All P values were adjusted using the Bonferroni correction [ 27 ]. A significant association was defined as P ≤ 0.05.

The study was approved by the ethics committee at Komfo Anokye Teaching Hospital and the Institutional Review Board at the University of Utah, USA. Written informed consent was obtained from either parent or guardian of all participants. Consent forms were presented in Twi or in English, depending on the consenting parent or guardian preference. US Department of Health and Human Services guidelines for human subjects research, the University of Utah guidelines, and the guidelines for the Komfo Anokye Teaching Hospital were followed.

Clinical and laboratory findings

Forty-two children were enrolled with CM due to P. falciparum; of CM cases, 6 were enrolled in 2004, 19 in 2005, and 17 in 2006; note: fewer children were enrolled in year 1 due to lack of dedicated study personnel and difficulties for setting up the Study Protocol requirements. Thirty-six healthy controls were enrolled during these same years. The enrollment periods for each year were the same, i.e., during the long rainy season. Two CM participants were excluded from the analyses due to missing data; six children with CM died (14.3%) during hospitalization, all within the first 48 h of admission.

Clinical characteristics for the two groups are listed in Table  1 . The groups were closely matched for age, gender, and weight. Physical findings revealed the degree of illness in the CM group, including significant differences in pulse, respiratory rate, and temperature. Of those CM cases for whom clinical data was available, 68% experienced a witnessed seizure and 59% received diagnostic lumbar puncture. Historical data on most recent food intake indicated that plasma amino acid levels in CM participants were unlikely to be confounded by recent protein ingestion. The length of pre-admission symptoms (mean, 4 days) was consistent with the acute course of CM leading to coma prompting presentation at hospital.

Laboratory findings in CM cases were consistent with CM, including anemia, thrombocytopenia, and metabolic acidosis (Table  2 ). Elevated creatinine was likely due to acute kidney injury [ 28 ]. Hypoglycaemia in CM participants was obscured by immediate intravenous glucose-containing fluid begun on all children presenting in coma.

Plasma amino acid levels at entry

Plasma samples collected from 40 CM enrollees at various time-points during hospitalization were available for amino acid analysis with one exception: a single sample at time zero. Of the 36 healthy control enrollees there were 6 samples unavailable for analysis because of insufficient venipuncture blood for plasma separation or because samples went missing.

For healthy controls, normal distribution was usually the case. However, all amino acid data for the CM group showed non-parametric distributions. For 5 of the 21 amino acids (Asp, Cys, Leu, Tyr, Val), there were no significant differences between CM versus healthy controls at entry (Table  3 ). In all cases but one, the levels of the remaining amino acids (Ala, Arg, Cit, Glu, Gln, Gly, His, Ilu, Lys, Met, Orn, Pro, Ser, Tau, Thr) were significantly lower in the CM group compared to healthy controls. One amino acid (Phe) was significantly elevated in CM participants compared to healthy controls. Also shown in Table  3 are the normal ranges (mean ± 2SD) for each amino acid, which were established at the ARUP diagnostic laboratory based on a large age-dependent database of healthy US children. For Ghanaian healthy control children, amino acid levels were largely within the US normal ranges.

Longitudinal association between BCS and amino acid levels outside the normal range

Results obtained from the ordinal GLM-EM showed that BCS was significantly associated with amino acid levels outside the normal range for 10 of the 21 amino acids (Table  4 ). Nine of the 10 amino acids were associated with significant probability for having a low BCS; the lower the odds ratio the greater the probability of having a low BCS. A phenylalanine level outside the normal range was associated with the highest probability of having a low BCS (lowest odds ratio). Conversely, valine out-of-range levels were significantly associated with a high probability of having a high BCS (highest odds ratio). Odds ratios calculated using statistical modeling (GLM-EM) do not specify whether an amino acid association significance is above or below its normal range. In this regard it is notable that only one of the ten amino acids (phenylalanine) with significant association for a low BCS is above its normal range. All other significant amino acid associations are below their normal ranges.

For completeness box plots showing raw data for out-of-range versus normal range individual samples for each amino acid at a given BCS are shown in Supplementary data (Supplementary Fig. 1, panels A–D).

Logistic regression assessment of time to reach normal range levels for the 10 amino acids significantly associated with BCS

For clarity, the time course was plotted for plasma levels (mean ± SEM) of each of the ten amino acids (Table  4 ) significantly associated with a low BCS (Fig.  1 ). In each panel the right axis reproduces the mean ± SEM BCS. A dashed horizontal line marks the normal range lower limit for nine of the ten amino acids. For these nine all abnormal levels were below the normal range. The one exception is phenylalanine where the dashed line shows the upper limit of normal, commensurate with hyperphenylalaninaemia for this amino acid. Some amino acids (Arg, Cit, Gln) exhibit a time lag before returning to within their normal ranges while BCS rose to a near conscious level (BCS ~ 3–4) by 24 h. Others (Gly, Ilu, Lys, Ser, Thr) showed borderline low levels, but rose to within their normal ranges as time progressed. Exceptionally, high phenylalanine levels promptly fell in concert with rising BCS.

figure 1

Out-of-range amino acid levels significantly associated with Blantyre Coma Score at sampling time intervals. Plasma levels of each amino acid (circles, mean ± SEM) significantly associated with BCS (Table  4 ) are shown at sampling times. Blantyre Coma Score (triangles, mean ± SEM) at the same sampling times are reproduced for each amino acid panel

A logistic regression model was used to determine the time after which there was a high probability that amino acid levels reached the normal range (Fig.  2 ). Bar graphs for each of the ten amino acids show odds ratios (± 95% CI) at each sampling time. Odds ratios below 1.0 with Asterix denote significant association with levels outside normal ranges. Odds ratios above 1.0 with Asterix denote significant association with levels within normal ranges. Bars without Asterix are without significance. Almost all amino acids normalized by 48–60 h post admission and initiation of treatment. However, the analyses showed marked variability across amino acids: for example, phenylalanine and isoleucine reached their normal range at 36 h, while for other amino acids it took 48 and 60 h to reach the normal range. An outlier was valine, with significant normal range values at zero and 12 h and abnormal range values at 36, 48 and 60 h. The changes were associated with increasing BCS as time elapsed. For reference, Fig.  2 (lower right-most panel) shows BCS data at each sampling time.

figure 2

Effect of sampling time on having an amino acid level outside of, or within, the normal range. Bar graphs for each amino acid show odds ratios (± 95% CI) at each sampling time. Odds ratios below 1.0 with Asterix denote significant association with levels outside normal ranges. Odds ratios above 1.0 with Asterix denote significant association with levels within normal ranges. Bars without Asterix are without significance. Box plot in the lower right hand panel shows median BCS with variance measures at sampling intervals. Boxes represent interquartile ranges (IQR). Dark lines within boxes are medians. Whiskers indicate minimum and maximum values within 1.5 times the IQR

Plasma amino acid abnormalities in CM at hospital entry

Abnormal amino acid levels based on age-dependent normal ranges were defined for 21 amino acids as established for healthy US children in the reference laboratory. To determine whether these normal ranges applied to healthy Ghanaian children, an age-matched healthy control group was enrolled and their amino acid plasma levels were measured. The healthy control levels were within normal ranges save for slight median decreases of cystine and glutamine (Table  3 ).

Amino acid levels were compared in CM versus healthy control cohorts at hospital entry. Sixteen of the 21 amino acids were significantly different. The results mirror those reported by others for children [ 13 ] and adults [ 12 ] with severe malaria, including CM. Fifteen of the 16 were below the normal range and one (phenylalanine) was elevated. The hypercatabolic state induced by the inflammatory response in severe malaria, in which amino acids are oxidatively degraded to yield chemical energy likely contributes to the low levels observed [ 29 , 30 ] Additionally, gluconeogenesis in liver consumes glucogenic amino acids. Acute kidney injury associated with amino acid reabsorption abnormalities poses yet another source of loss [ 13 ]. Significantly low levels of glutamine, glutamate, proline, ornithine, citrulline and arginine comprise the pathway of de novo arginine synthesis [ 18 ] a finding consistent with low nitric oxide production [ 31 ]. Longitudinal data found low arginine levels associated with nitric oxide-dependent endothelial dysfunction [ 32 ].

Longitudinal associations between out-of-range amino acid levels and BCS

These results extended amino acid abnormalities by longitudinal measurements with analysis for association with BCS. By employing the GLM-EM model the data showed that ten of the sixteen out-of-range amino acids at entry were significantly associated with BCS (P ≤ 0.05, Table  4 ). Six of the ten (arginine, glycine, isoleucine, phenylalanine, threonine, and valine) showed high significance for association with BCS (P ≤ 0.01). Of this set, phenylalanine and valine stood out. Phenylalanine was the only amino acid with levels above the normal range in CM participants compared to healthy controls. Of all significant BCS-associated amino acids, phenylalanine showed the highest association (lowest odds ratio, 0.25). Valine out-of-range values were associated with high BCS.

For the ten amino acids significantly associated with BCS, the out-of-range values at each sampling time were analysed by calculating odds ratios with the GLM-EM model. With time, 9 of 10 amino acids normalized by 60 h post admission and initiation of treatment (odds ratios > 1.0). At 24 to 36 h post admission and initiation of treatment, levels transitioned to normal ranges for isoleucine, lysine, phenylalanine, and threonine. At these same time-points median BCS rose from about 1 to 4–5. Thus, regaining consciousness occurred at the time when these four amino acids normalized. The outlier was valine with significant normal range values at entry and at 12 h with transition to abnormally low levels at 36, 48, and 60 h post admission and initiation of treatment.

Significance of association of abnormal valine levels with high BCS

Of the ten amino acids associated with BCS, a unique association between BCS and valine was found. Out-of-range valine levels bore significant probability to have high, rather than low, BCS. This finding could relate to the complex catabolism of the branched-chain amino acids (Leu, Ilu, Val). In experimental animal models and possibly in humans, high leucine levels antagonize valine degradation at the decarboxylation (keto acid dehydrogenase) step by competitive inhibition [ 33 , 34 ]. This might delay valine catabolic kinetics resulting in out-of-range levels at later times during recovery, when BCS had risen to 4 or 5.

Possible significance of hyperphenylalaninaemia as a surrogate for malarial coma

Like other amino acids, the phenylalanine degradation supplies carbon to the TCA cycle for oxidation [ 35 ]. Why then did investigators find consistently elevated phenylalanine levels in CM? In a previous study, a possible mechanism to explain hyperphenylalaninaemia was found [ 36 ]. The liver enzyme, phenylalanine hydroxylase (PHA), requires the cofactor tetrahydrobiopterin (BH 4 ) for mono-oxygenation of phenylalanine to yield its product, tyrosine [ 37 ]. In health biopterin is poised in the reduced state (BH 4 ) such that with each phenylalanine to tyrosine reaction, the oxidized cofactors, biopterin (B) and dihydrobiopterin (BH 2 ), are reduced back to BH 4 via two pathways (recycling and salvage) [ 37 ]. The reducing equivalents for restoring biopterin to its reduced state are supplied by the nicotinamide adenine dinucleotides NADH and NADPH [ 37 ]. In the systemic compartment (i.e., plasma and liver) about 2/3 of biopterin is poised in the reduced state (BH 4 ) [ 38 ]. This provides sufficient reducing power for tight regulation of plasma phenylalanine below 80 micromolar [ 39 ]. Phenylalanine regulation prevents hyperphenylalaninaemia, which is toxic to the brain (e.g., in the congenital disease, phenylketonuria [ 39 ]). While total biopterins (B + BH2 + BH 4 ) were slightly increased in CM, the percentage of reduced biopterin fell by 75% [ 36 ]. This suggests that hyperphenylalaninaemia in CM results from a diminution of PAH catalysis in the liver due to insufficient BH 4 availability, possibly a result of oxidative stress.

Potential role of BH 4 in CM pathogenesis

BH 4 is a unique cofactor in that there are only 6 enzymes for which it is required [ 37 , 40 ]. These are the enzymes for (a) synthesis of catecholamines and serotonin in the brain, (b) systemic and CNS synthesis of nitric oxide, (c) PAH in liver, and (d) a little studied liver enzyme for catabolism of membrane ether lipids, alkylglycerol mono-oxygenase (AGMO) [ 40 , 41 , 42 ]. Accumulation of membrane-derived ether lipids from impaired AGMO activity resulting from limited BH 4 availability might adversely affect the CNS [ 43 ]. Blood–brain barrier permeability is markedly increased in experimental animals treated with short chain alkyl (ether) glycerols [ 44 , 45 ]. Limiting AGMO activity could enhance circulating ether lipid mediators such as platelet activation factor (PAF). This possibility is yet to be investigated [ 46 , 47 ]. There is evidence from malaria models that PAF could enhance sequestration of infected red cells if regulation control mechanisms fail [ 48 , 49 ]. The possible role for an ether lipid(s) in malaria pathogenesis should be explored .

Limitations

Because a group of children with severe malaria without coma was not included, one cannot infer that amino acid abnormalities described here are specific for the depth of coma itself, but rather could reflect broader associations with severe disease states. The cohort of 40 CM cases is relatively small. BCS are subjective, therefore open to observer bias. The linear mixed effects model can only assign an association between out-of-range amino acid levels and BCS, and thus cannot establish cause and effect. Diagnosis of CM carries a significant false positive fraction [ 50 ]. However, 59% of the CM cohort received lumbar puncture to rule out meningitis, thus diminishing the likelihood of non-malarial causes of coma.

Longitudinal data on normalization of amino acid levels with recovery from malarial coma found a highly significant association for phenylalanine. Unlike other coma-associated amino acids, abnormal phenylalanine levels were above, not below, normal range. Hyperphenylalaninaemia likely results from insufficient reducing power normally provided by tetrahydrobiopterin. A previous study found a marked diminution of tetrahydrobiopterin. Therefore, the single other liver tetrahydrobiopterin-dependent enzyme, responsible for degrading biologically active ether lipids (AGMO), may be impaired. Elevated bioactive ether lipids are known to permeabilize the CNS blood–brain barrier and lead to platelet activation, events shown to be critical in CM pathogenesis. If impaired, restoration of AGMO activity could provide efficacious adjunctive therapy.

Data availability

No datasets were generated or analysed during the current study.

Marsh K, Forster D, Waruiru C, Mwangi I, Winstanley M, Marsh V, et al. Indicators of life-threatening malaria in African children. N Engl J Med. 1995;332:1399–404.

Article   CAS   PubMed   Google Scholar  

WHO. World Malaria Report 2021. Geneva: World Health Organization; 2021.

Google Scholar  

Seydel KB, Kampondeni SD, Valim C, Potchen MJ, Milner DA, Muwalo FW, et al. Brain swelling and death in children with cerebral malaria. N Engl J Med. 2015;372:1126–37.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Potchen MJ, Kampondeni SD, Seydel KB, Haacke EM, Sinyangwe SS, Mwenechanya M, et al. 1.5 Tesla magnetic resonance imaging to investigate potential etiologies of brain swelling in pediatric cerebral malaria. Am J Trop Med Hyg. 2018;98:497–504.

Kampondeni S, Seydel KB, Zhang B, Small DS, Birbeck GL, Hammond CA, et al. Amount of brain edema correlates with neurologic recovery in pediatric cerebral malaria. Pediatr Infect Dis J. 2020;39:277–82.

Article   PubMed   Google Scholar  

Poespoprodjo JR, Douglas NM, Ansong D, Kho S, Anstey NM. Malaria. Lancet. 2023;402:2328–45.

Darling TK, Mimche PN, Bray C, Umaru B, Brady LM, Stone C, et al. EphA2 contributes to disruption of the blood-brain barrier in cerebral malaria. PLoS Pathog. 2020;16: e1008261.

Moxon C, Alhamdi Y, Storm J, Toh J, Ko JY, Murphy G, et al. Parasite histones mediate blood-brain barrier disruption in cerebral malaria. Clin Med. 2020;20(Suppl 2):s96–7.

Article   Google Scholar  

White NJ, Warrell DA, Chanthavanich P, Looareesuwan S, Warrell MJ, Krishna S, et al. Severe hypoglycemia and hyperinsulinemia in falciparum malaria. N Engl J Med. 1983;309:61–6.

Planche T, Krishna S. Severe malaria: metabolic complications. Curr Mol Med. 2006;6:141–53.

Herdman MT, Sriboonvorakul N, Leopold SJ, Douthwaite S, Mohanty S, Hassan MM, et al. The role of previously unmeasured organic acids in the pathogenesis of severe malaria. Crit Care. 2015;19:317.

Article   PubMed   PubMed Central   Google Scholar  

Leopold SJ, Apinan S, Ghose A, Kingston HW, Plewes KA, Hossain A, et al. Amino acid derangements in adults with severe falciparum malaria. Sci Rep. 2019;9:6602.

Conroy AL, Tran TM, Bond C, Opoka RO, Datta D, Liechty EA, et al. Plasma amino acid concentrations in children with severe malaria are associated with mortality and worse long-term kidney and cognitive outcomes. J Infect Dis. 2022;226:2215–25.

Dinarello CA. Interleukin-1 and the pathogenesis of the acute-phase response. N Engl J Med. 1984;311:1413–8.

Lundblad RL. Biotechnology of plasma proteins. Ist. Boca Raton: CRC Press; 2012. p. 460.

Book   Google Scholar  

van Gassel RJJ, Baggerman MR, van de Poll MCG. Metabolic aspects of muscle wasting during critical illness. Curr Opin Clin Nutr Metab Care. 2020;23:96–101.

Lopansri BK, Anstey NM, Stoddard GJ, Mwaikambo ED, Boutlis CS, Tjitra E, et al. Elevated plasma phenylalanine in severe malaria and implications for pathophysiology of neurological complications. Infect Immun. 2006;74:3355–9.

Rubach MP, Zhang H, Florence SM, Mukemba JP, Kalingonji AR, Anstey NM, et al. Kinetic and cross-sectional studies on the genesis of hypoargininemia in severe pediatric Plasmodium falciparum malaria. Infect Immun. 2019;87:e00655-e718.

Molyneux ME, Taylor TE, Wirima JJ, Borgstein A. Clinical features and prognostic indicators in paediatric cerebral malaria: a study of 131 comatose Malawian children. Q J Med. 1989;71:441–59.

CAS   PubMed   Google Scholar  

WHO. Severe falciparum malaria. Trans R Soc Trop Med Hyg. 2000;94:S1–90.

Armstrong BG, Sloan M. Ordinal regression models for epidemiologic data. Am J Epidemiol. 1989;129:191–204.

Norris CM, Ghali WA, Saunders LD, Brant R, Galbraith D, Faris P, et al. Ordinal regression model and the linear regression model were superior to the logistic regression models. J Clin Epidemiol. 2006;59:448–56.

Hedeker D. A mixed-effects multinomial logistic regression model. Stat Med. 2003;22:1433–46.

Ugwu CLJ, Zewotir TT. Using mixed effects logistic regression models for complex survey data on malaria rapid diagnostic test results. Malar J. 2018;17:453.

Umlauf N, Adler D, Kneib T, Lang S, Zeileis A. Structured additive regression models: an R interface to BayesX. Stat Softw. 2015;63:1–46.

Team RC. R: a language and environment for statistical computing. Vienna: R Foundation for Statistical Computing; 2021.

Cabin RJ, Mitchell RJ. To Bonferroni or not to Bonferroni: when and how are the questions. Ecol Soc Am. 2000;81:246–8.

Batte A, Berrens Z, Murphy K, Mufumba I, Sarangam ML, Hawkes MT, et al. Malaria-associated acute kidney injury in African children: prevalence, pathophysiology, impact, and management challenges. Int J Nephrol Renovasc Dis. 2021;14:235–53.

Basler T, Meier-Hellmann A, Bredle D, Reinhart K. Amino acid imbalance early in septic encephalopathy. Intensive Care Med. 2002;28:293–8.

Bröer S, Bröer A. Amino acid homeostasis and signalling in mammalian cells and organisms. Biochem J. 2017;474:1935–63.

Anstey NM, Weinberg JB, Hassanali MY, Mwaikambo ED, Manyenga D, Misukonis MA, et al. Nitric oxide in Tanzanian children with malaria: inverse relationship between malaria severity and nitric oxide production/nitric oxide synthase type 2 expression. J Exp Med. 1996;184:557–67.

Yeo TW, Lampah DA, Gitawati R, Tjitra E, Kenangalem E, McNeil YR, et al. Recovery of endothelial function in severe falciparum malaria: relationship with improvement in plasma L-arginine and blood lactate concentrations. J Infect Dis. 2008;198:602–8.

Harper AE, Miller RH, Block KP. Branched-chain amino acid metabolism. Annu Rev Nutr. 1984;4:409–54.

Staten MA, Bier DM, Matthews DE. Regulation of valine metabolism in man: a stable isotope study. Am J Clin Nutr. 1984;40:1224–34.

Castillo L, Yu YM, Marchini JS, Chapman TE, Sanchez M, Young VR, et al. Phenylalanine and tyrosine kinetics in critically ill children with sepsis. Pediatr Res. 1994;35:580–8.

Rubach MP, Mukemba J, Florence S, Lopansri BK, Hyland K, Volkheimer AD, et al. Impaired systemic tetrahydrobiopterin bioavailability and increased oxidized biopterins in pediatric falciparum malaria: association with disease severity. PLoS Pathog. 2015;11: e1004655.

Blau N, Thony B, Cotton RGH, Hyland K. Disorders of tetrahydrobiopterin and related biogenic amines. In: CR Scriver AB, WS Sly, D Valle, (eds.) The Metabolic and molecular bases of inherited disease. 8 edn. Vol II. New York: McGraw-Hill; 2001.

Hyland K. Estimation of tetrahydro, dihydro and fully oxidised pterins by high-performance liquid chromatography using sequential electrochemical and fluorometric detection. J Chromatogr. 1985;343:35–41.

Kaufman CS. Hyperphenylalaninemia: phenylalanine hydroxylase deficiency. In: CR Scriver AB, WS Sly, D Valle, (eds). The metabolic and molecular bases of inherited disease. 8 th edn VII. New York: McGraw-Hill; 2001, 1667–724.

Werner ER. Three classes of tetrahydrobiopterin-dependent enzymes. Pteridines. 2013;24:7–11.

Article   CAS   Google Scholar  

Watschinger K, Werner ER. Alkylglycerol monooxygenase. IUBMB Life. 2013;65:366–72.

Sailer S, Keller MA, Werner ER, Watschinger K. The emerging physiological role of AGMO 10 years after its gene identification. Life. 2021;11:88.

Dorninger F, Forss-Petter S, Wimmer I, Berger J. Plasmalogens, platelet-activating factor and beyond - ether lipids in signaling and neurodegeneration. Neurobiol Dis. 2020;145: 105061.

Erdlenbruch B, Alipour M, Fricker G, Miller DS, Kugler W, Eibl H, et al. Alkylglycerol opening of the blood-brain barrier to small and large fluorescence markers in normal and C6 glioma-bearing rats and isolated rat brain capillaries. Br J Pharmacol. 2003;140:1201–10.

Erdlenbruch B, Jendrossek V, Eibl H, Lakomek M. Transient and controllable opening of the blood-brain barrier to cytostatic and antibiotic agents by alkylglycerols in rats. Exp Brain Res. 2000;135:417–22.

Stafforini DM, McIntyre TM, Zimmerman GA, Prescott SM. Platelet-activating factor, a pleiotrophic mediator of physiological and pathological processes. Crit Rev Clin Lab Sci. 2003;40:643–72.

Gupta S, Seydel K, Miranda-Roman MA, Feintuch CM, Saidi A, Kim RS, et al. Extensive alterations of blood metabolites in pediatric cerebral malaria. PLoS ONE. 2017;12: e0175686.

Lacerda-Queiroz N, Rodrigues DH, Vilela MC, Rachid MA, Soriani FM, Sousa LP, et al. Platelet-activating factor receptor is essential for the development of experimental cerebral malaria. Am J Pathol. 2012;180:246–55.

Lacerda-Queiroz N, Rachid MA, Teixeira MM, Teixeira AL. The role of platelet-activating factor receptor (PAFR) in lung pathology during experimental malaria. Int J Parasitol. 2013;43:11–5.

Taylor TE, Fu WJ, Carr RA, Whitten RO, Mueller JS, Fosiko NG, et al. Differentiating the pathologies of cerebral malaria by postmortem parasite counts. Nat Med. 2004;10:143–5.

Download references

Acknowledgements

Research reported in this publication was supported by the National Center for Advancing Translational Sciences of the National Institutes of Health under Award Number UL1TR002538. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health or their employers. John Hibbs Jr., J. Brice Weinberg, and Guy Zimmerman (deceased) provided useful comments. Tsin Yeo and Nick Anstey reviewed the manuscript and raised important suggestions for statistical analysis. Mary MacFarland aided in the preparation of the manuscript. Thanks to Will J. Fennelly for data entry and to Arthur L. Granger for arranging statistical assistance. Special appreciations go to Justice Sylverken, Clinical Assistant, and the pediatric nursing staff at Komfo Anokye Teaching Hospital for their dedicated patient care and service to the study.

Presentation

This work has not been presented in any form at a scientific meeting.

NIH/NIAID: K23 AI 116869 BKL (awardee), DLG (sponsor).

Author information

Authors and affiliations.

Division of Infectious Diseases, Department of Internal Medicine, University of Utah Spencer Fox Eccles School of Medicine, 2761 E. Swasont Way, Holladay, Salt Lake City, UT, 84117, USA

Donald L. Granger & Devon C. Hale

George H. Wahlen Veterans Affairs Medical Center, Salt Lake City, UT, USA

Donald L. Granger

Department of Pediatrics, Komfo Anokye Teaching Hospital, Kumasi, Ghana

Daniel Ansong & Tsiri Agbenyega

Intermountain Health Care, Salt Lake City, UT, USA

Melinda S. Liddle & Bert K. Lopansri

Department of Psychiatry, North Shore University Hospital, Glen Oaks, NY, USA

Benjamin A. Brinton

Research Triangle Institute International, Washington, DC, USA

Richard Reithinger & Donal Bisanzio

You can also search for this author in PubMed   Google Scholar

Contributions

D.L.G wrote the main manuscript text. B.K.L., R.R., and D.B provided edits to the manuscript. D.L.G. and D.B. assembled the Tables and Figures. D.A., T.A., M.S.L., B.A.B., D.C.H., and B.K.L. carried out the study protocol in Kumasi, Ghana. D.B., R.R., and D.L.G. analyzed the data. D.B. and R.R. provided statistical expertise. All authors reviewed and approved the manuscript.

Corresponding author

Correspondence to Donald L. Granger .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

12936_2024_5077_moesm1_esm.zip.

Additional file1Supplementary Fig. 1. Box plots show out-of-range versus normal-range amino acid plasma levels at given BCS. Fig. 1 A–1D show results for all 21 amino acids analysed. Coloured circles represent individual amino acid levels for participants with cerebral malaria. An Asterix denotes significant association between amino acid levels within or outside the normal ranges at given Blantyre Coma Scores as determined by the generalized linear mixed-effects model

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/ .

Reprints and permissions

About this article

Cite this article.

Granger, D.L., Ansong, D., Agbenyega, T. et al. Longitudinal associations of plasma amino acid levels with recovery from malarial coma. Malar J 23 , 253 (2024). https://doi.org/10.1186/s12936-024-05077-9

Download citation

Received : 14 May 2024

Accepted : 13 August 2024

Published : 23 August 2024

DOI : https://doi.org/10.1186/s12936-024-05077-9

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Cerebral malaria
  • Blantyre coma score
  • Amino acids
  • Generalized linear mixed-effects model
  • Tetrahydrobiopterin
  • Glyceryl lipid ethers

Malaria Journal

ISSN: 1475-2875

presentation of cerebral malaria

IMAGES

  1. PPT

    presentation of cerebral malaria

  2. PPT

    presentation of cerebral malaria

  3. PPT

    presentation of cerebral malaria

  4. PPT

    presentation of cerebral malaria

  5. Neurological and behavioral manifestations of cerebral malaria: An update

    presentation of cerebral malaria

  6. PPT

    presentation of cerebral malaria

COMMENTS

  1. Pathophysiology, clinical presentation, and treatment of com

    The clinical presentation of cerebral malaria is diffuse symmetrical encephalopathy with fever and absent or few focal neurological signs. In children, coma can rapidly develop after fever onset (mean, 2 days) . In adults, coma is typically gradual with increasing drowsiness, confusion, obtundation, and high fevers (mean duration, 5 days).

  2. Cerebral Malaria; Mechanisms Of Brain Injury And Strategies For Improved Neuro-Cognitive Outcome

    Cerebral malaria is the most severe neurological manifestation of severe malaria. With an incidence of 1,120/100,000/year in the endemic areas of Africa, children in this region bear the brunt. Peak incidence is in pre-school children and at a minimum, 575,000 children in Africa develop cerebral malaria annually( 2 ).

  3. Cerebral malaria: insight into pathogenesis, complications and

    Cerebral malaria is a medical emergency. All patients with Plasmodium falciparum malaria with neurologic manifestations of any degree should be urgently treated as cases of cerebral malaria. Pathogenesis of cerebral malaria is due to damaged vascular endothelium by parasite sequestration, inflammatory cytokine production and vascular leakage, which result in brain hypoxia, as indicated by ...

  4. Cerebral Malaria: Current Clinical and Immunological Aspects

    Treatment of Severe and Cerebral Malaria. Artemisinins, which have the fastest parasite clearing time of all anti-malarial drugs, have become the drugs of the moment, and artesunate is the first-line therapy for treating severe and cerebral malaria in both children and adults ().. The main difference between the treatment of uncomplicated and severe malaria is the route of administration of ...

  5. Pathophysiology and neurologic sequelae of cerebral malaria

    Cerebral malaria (CM), results from Plasmodium falciparum infection, and has a high mortality rate. CM survivors can retain life-long post CM sequelae, including seizures and neurocognitive deficits profoundly affecting their quality of life. As the Plasmodium parasite does not enter the brain, but resides inside erythrocytes and are confined to the lumen of the brain's vasculature, the ...

  6. Cerebral malaria

    Cerebral malaria may be the most common non-traumatic encephalopathy in the world. The pathogenesis is heterogenous and the neurological complications are often part of a multisystem dysfunction. The clinical presentation and pathophysiology differs between adults and children. Recent studies have elucidated the molecular mechanisms of pathogenesis and raised possible interventions ...

  7. Cerebral Malaria: Mechanisms of Brain Injury and Strategies for

    Cerebral malaria is the most severe neurological complication of infection with Plasmodium falciparum. With >575,000 cases annually, children in sub-Saharan Africa are the most affected.

  8. Pathogenetic mechanisms and treatment targets in cerebral malaria

    In mice with cerebral malaria, cross-presentation of Plasmodium antigens by dendritic cells results in the activation of cytotoxic CD8 + T cells 277,278,279,280.

  9. PDF Pathogenetic mechanisms and treatment targets in cerebral malaria

    treatment targets in cerebral malaria ... presentation Decreased intercellular contacts Increased Degranulation cytoadhesion C3a C4a C5a Cytokines Cytoadherence Haemoglobin MHC-I Blood vessel lumen

  10. Cerebral malaria

    P. falciparum caused malaria is the most dreaded form of this infectious disease that is responsible for significant global mortality. One of its most common clinical manifestations is cerebral malaria, which is common in children under 5 years of age. Though curable, drug action is limited by timely diagnosis and hospitalization.

  11. Cerebral Malaria

    Cerebral malaria is a life-threatening complication of P. falciparum infestation that occurs in approximately 2% of patients. Pathogenesis may be explained by 2 mechanisms: vascular sequestration of parasitized erythrocytes and the potential cerebral toxicity by cytokines. Progressive clinical changes occur, along with high fever and chills.

  12. PDF Pathophysiology, clinical presentation and treatment of cerebral malaria

    cerebral malaria is functionally grossly intact. 35 ,36 Studies in African children with cerebral malaria do show a subtle increase in BBB permeability with a disruption of endothelial intercellular tight junctions on autopsy.37 ,38 Imaging studies reveal that most adults with cerebral malaria have no evidence of cerebral oedema.39 ,40 In African

  13. PDF Pathophysiology and neurologic sequelae of cerebral malaria

    Cerebral malaria (CM), results from Plasmodium falciparum infection, and has a high mortality rate. CM survivors can retain life‐long post CM sequelae, including seizures and neurocognitive deficits profoundly afecting their quality of life. As the Plasmodium parasite does not enter the brain, but resides inside erythrocytes and are confined ...

  14. Pathophysiology and neurologic sequelae of cerebral malaria

    These variances in CM disease presentation may arise due to differences in the immature brain, including differences in host responses of the cerebral vasculature in different brain regions to sequestration and the magnitude of inflammation. ... Cerebral malaria pathology manifests itself differently in white matter and gray matter of the brain ...

  15. Clinical Features of Malaria

    Clinical presentation. Infection with malaria parasites may result in a wide variety of symptoms, ranging from absent or very mild symptoms to severe disease and even death. ... occurs in the vessels of the brain it is believed to be a factor in causing the severe disease syndrome known as cerebral malaria, which is associated with high mortality.

  16. Pattern of Clinical and Laboratory Presentation of Cerebral Malaria

    Introduction: Cerebral malaria (CM) is the most lethal form of severe malaria with high case fatality rates. Overtime, there is an inherent risk in changing pattern of presentation of CM which, if the diagnosis is missed due to these changing factors, may portend a poor outcome.

  17. What is Cerebral Malaria?

    Cerebral malaria is a serious neurological complication of severe malaria that affects about 1% of children under the age of 5 who have been infected with Plasmodium falciparum.

  18. Malaria

    Malaria is a disease caused by a parasite. The parasite is spread to humans through the bites of infected mosquitoes. People who have malaria usually feel very sick with a high fever and shaking chills. ... Cerebral malaria. If parasite-filled blood cells block small blood vessels to your brain (cerebral malaria), swelling of your brain or ...

  19. Malaria: Clinical manifestations and diagnosis in ...

    The clinical manifestations of malaria vary with parasite species, epidemiology, immunity, and age. Issues related to clinical manifestations and diagnosis of malaria will be reviewed here. Technical aspects of laboratory tools for diagnosis of malaria are discussed further separately. The epidemiology, pathogenesis, diagnosis, and treatment of ...

  20. Cerebral malaria

    Cerebral malaria is the most severe pathology caused by the malaria parasite, Plasmodium falciparum. The pathogenic mechanisms leading to cerebral malaria are still poorly defined as studies have been hampered by limited accessibility to human tissues. Nevertheless, histopathology of post-mortem human tissues and mouse models of cerebral ...

  21. Cerebral malaria induced by plasmodium falciparum: clinical features

    Cerebral malaria (CM) is a fatal neurological complication of P. falciparum malaria (Luzolo and Ngoyi, 2019), and children aged under 3 years and pregnant women are most susceptible ... Brain microvessel cross-presentation is a hallmark of experimental cerebral malaria. EMBO Mol. Med. 5 (7), ...

  22. PDF GUIDELINES

    Guidelines for the treatment of malaria - 3rd edition. 1.Malaria - drug therapy. 2.Malaria - diagnosis. 3.Antimalarials - administration and dosage. ... The designations employed and the presentation of the material in this publication do not ... Cerebral malaria. Severe P. falciparum malaria with coma (Glasgow coma scale < 11,

  23. Longitudinal associations of plasma amino acid levels with recovery

    Disordered amino acid metabolism is observed in cerebral malaria (CM). This study sought to determine whether abnormal amino acid concentrations were associated with level of consciousness in children recovering from coma. Twenty-one amino acids and coma scores were quantified longitudinally and the data were analysed for associations. In a prospective observational study, 42 children with CM ...

  24. Pathogenesis, clinical features, and neurological outcome of cerebral

    Even though this type of malaria is most common in children living in sub-Saharan Africa, it should be considered in anybody with impaired consciousness that has recently travelled in a malaria-endemic area. Cerebral malaria has few specific features, but there are differences in clinical presentation between African children and non-immune adults.