Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Lancet Infect Dis. 2006 Apr; 6(4): 203–214.
Published online 2006 Mar 21. doi: 10.1016/S1473-3099(06)70435-2
PMCID: PMC7185836
PMID: 16554245

Crimean-Congo haemorrhagic fever

Abstract

Crimean-Congo haemorrhagic fever (CCHF) is an often fatal viral infection described in about 30 countries, and it has the most extensive geographic distribution of the medically important tickborne viral diseases, closely approximating the known global distribution of Hyalomma spp ticks. Human beings become infected through tick bites, by crushing infected ticks, after contact with a patient with CCHF during the acute phase of infection, or by contact with blood or tissues from viraemic livestock. Clinical features commonly show a dramatic progression characterised by haemorrhage, myalgia, and fever. The levels of liver enzymes, creatinine phosphokinase, and lactate dehydrogenase are raised, and bleeding markers are prolonged. Infection of the endothelium has a major pathogenic role. Besides direct infection of the endothelium, indirect damage by viral factors or virus-mediated host-derived soluble factors that cause endothelial activations and dysfunction are thought to occur. In diagnosis, enzyme-linked immunoassay and real-time reverse transcriptase PCR are used. Early diagnosis is critical for patient therapy and prevention of potential nosocomial infections. Supportive therapy is the most essential part of case management. Recent studies suggest that ribavirin is effective against CCHF, although definitive studies are not available. Health-care workers have a serious risk of infection, particularly during care of patients with haemorrhages from the nose, mouth, gums, vagina, and injection sites. Simple barrier precautions have been reported to be effective.

Introduction

Crimean-Congo haemorrhagic fever (CCHF) is a fatal viral infection described in parts of Africa, Asia, eastern Europe, and the middle east.1, 2 The virus belongs to the genus Nairovirus in the Bunyaviridae family and causes severe diseases in human beings, with a reported mortality rate of 3–30%.2, 3 The geographic range of CCHF virus is the most extensive one among the medically important tickborne viruses (figure 1 ). Human beings become infected through tick bites, by contact with a patient with CCHF during the acute phase of infection, or by contact with blood or tissues from viraemic livestock.4 The clinical features show common dramatic progression characterised by haemorrhage, myalgia, and fever, with some differences among different regions suggested but not well studied. Treatment with ribavirin has not yet been approved in many countries. However, there are reports that indicate the drug may be beneficial. The widespread geographic distribution of CCHF virus, its ability to produce severe human disease with high mortality rates, and fears about its intentional use as a bioterrorism agent5 make the virus an important human pathogen. Moreover, ecological complexity of vectorborne diseases, therapeutic controversy, and human-to-human transmission of a zoonotic infection make CCHF an interesting topic for research.

An external file that holds a picture, illustration, etc.
Object name is gr1_lrg.jpg

Worldwide distribution of CCHF virus

There has been a substantial increase in reports on CCHF virus over the past 5 years. Here I review published work on CCHF, with an emphasis on the recent outbreak in Turkey. CCHF virus-infected cases were first reported in Turkey in 2002,3, 6, 7, 8 although epidemics have been reported from neighbouring countries since the 1970s. Between 2002 and 2005, 500 cases were reported to Turkish Ministry of Health, and 26 (5·2%) of these cases died.9

Historical background

In the 12th century, a haemorrhagic syndrome was described in present day Tajikistan. The signs were presence of blood in the urine, rectum, gums, vomitus, sputum, and abdominal cavity.1 The arthropod that caused the disease was said to be tough, small, related to a louse or tick, and normally parasitising a black bird. In the modern era, Crimean haemorrhagic fever was first described as a clinical entity in 1944–45, when about 200 Soviet military personnel were infected while assisting peasants in Crimea in the wake of World War 2.1, 2 The virus was isolated from blood and tissues of patients using intracerebral inoculation of newborn white mice in 1967.1, 2 The virus responsible for Crimean haemorrhagic fever was later shown to be antigenically indistinguishable from Congo virus, isolated in 1956 from a febrile patient in Belgian Congo (present day Democratic Republic of the Congo).10 The common antigenic structure among Eurasian Crimean haemorrhagic fever strains11 and Asian12 and African strains of Congo virus9, 13, 14 led to the virus being called Crimean haemorrhagic fever-Congo virus,15 and then Crimean-Congo haemorrhagic fever virus.1

Epidemiology

The geographic range of CCHF virus is the most extensive among the tickborne viruses that affect human health, and the second most widespread of all medically important arboviruses, after dengue viruses (figure 1).2

The history of reported outbreaks is summarised in table 1 . Before 1970, most cases were reported from the former Soviet Union (Crimea, Astrakhan, Rostov, Uzbekistan, Kazakhstan, Tajikistan)1, 2 and Bulgaria,1, 2 as well as virus circulation in parts of Africa such as the Democratic Republic of the Congo and Uganda.10, 13 An outbreak in 1965 in China with a case fatality rate of 80% was noted, but not presented in detail.33 The initial recognition of haemorrhagic cases in Africa occurred in the 1960s, resulting in a series of in-depth studies in South Africa34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44 and reports of additional outbreaks from Congo,43 Mauritania,30 Burkina Faso,45 Tanzania,43 and Senegal.46 A substantial number of cases were also reported from middle eastern countries such as Iraq,25, 26 the United Arab Emirates (UAE),23, 24, 47 Saudi Arabia27 and Oman,28 and from Pakistan20 and China.19 By 2000, new outbreaks had been reported from Pakistan,21, 22, 48 Iran,29 Senegal,49 Albania,17 Yugoslavia,18, 50 Bulgaria,16 Turkey,3, 6, 7 Kenya,32 and Mauritania.51

Table 1

Reported outbreaks of CCHF since 1945

LocationYearsNumber of cases*Case fatality rate (%)Occupation
Southeast Europe
Crimea1944–45120010Military members
Astrakhan1953–63110417Agricultural workers
Rostov1963–69132315Agricultural workers
Bulgaria1953–742110517Agricultural workers, health-care workers
1975–961627911Agricultural workers
1997–031613821Agricultural workers
Albania20011770Agricultural workers, health-care workers
Kosovo2001181833Agricultural workers
Turkey2002–0595005Agricultural workers
Asia
China1965–941926021Agricultural workers
1997192624Agricultural workers
Kazakhstan1948–6817550Agricultural workers
Tajikistan1943–7019723Agricultural and laboratory workers
Pakistan1976201429Shepherd, health-care workers
1994213Not knownHealth-care workers
200022955Agricultural workers, health-care workers
Middle east
United Arab Emirates (UAE)197923, 24650Health-care workers
1994–95241173Agricultural workers
Sharjah, UAE1980210Storekeeper
Iraq1979–8025, 265564Agricultural workers
Saudi Arabia1990277Agricultural workers
Oman1995–96284Not knownAgricultural workers
Iran2003298118Agricultural workers
Africa
Zaire1956120Physician
Uganda1958–771128Laboratory workers
Mauritania19832, 3010Camel herd owner
2004313829Agricultural workers, health-care workers
Burkina Faso1983210..
South Africa1981–8623231Farmers, health-care workers
Tanzania1986210Student
Southwest Africa1986210..
Kenya2000321100Agricultural worker

..=not reported.

*The number of the confirmed cases is given if both suspected and confirmed cases were reported.
Present day Democratic Republic of the Congo.

Serological evidence for CCHF virus has been reported from Greece,31 India,52 Egypt,53 Portugal,54 Hungary,55 France,1 and Benin,1 although the virus was isolated only in Greece and the only reported human case was a Greek laboratory infection. CCHF virus is endemic in the Balkans, including Bulgaria,16 the former Yugoslavia,18, 50 and Albania.17 It is of interest that the strain that caused the laboratory-related infection in Greece was exceedingly mild, possibly reflecting chance variation; however, the virus has the greatest phylogenetic difference from other CCHF viruses and Greece is separated from Bulgaria by mountains approximately 1500–2500 m high.16

The microbiology and the life cycle of the virus

CCHF is a member of the Nairovirus genus of the family Bunyaviridae. Other genera within the family include Orthobunyavirus, Hantavirus, Phlebovirus, and Tospovirus. The Nairovirus genus includes 34 described viruses and is divided into seven different serogroups.56 The most important groups are the CCHF group, which includes CCHF virus and Hazara virus, and the Nairobi sheep disease group, which includes the Nairobi sheep disease and Dugbe viruses.4 CCHF virus, Dugbe virus, and Nairobi sheep disease virus are the only three members of the Nairovirus genus that are known to cause disease among human beings.

Bunyaviruses are enveloped particles with a single-stranded RNA genome of negative polarity (figure 2 ).57 The three genome segments encode four structural proteins—the RNA-dependent RNA polymerase (L protein) is encoded by the large (L) segment, the glycoproteins (GN and GC; previously referred to as G1 and G2) are encoded by the medium (M) segment, and the nucleocapsid protein (N) is encoded by the small (S) segment.58, 59

An external file that holds a picture, illustration, etc.
Object name is gr2_lrg.jpg

Schematic presentation of the virus structure

Copyright © 2006 Şebnem Eren

Emerging data on viral replication shows great potential for the development of new drugs. The viral glycoproteins are responsible for the recognition of receptor sites on susceptible cells. Following attachment, viruses are internalised by endocytosis.4 Replication occurs in the cytoplasm, and the virions mature by budding through the endoplasmic reticulum into cytoplasmic vesicles in the Golgi region.4 Bunyaviruses are known to bud from Golgi membranes and the budding site seems to be defined by retention of the glycoproteins GN and GC at that particular site.59 GN is localised to the Golgi compartment, whereas GC is found in the endoplasmic reticulum.59 Recently, the expression strategy and biosynthesis of the CCHF viral glycoproteins have been studied in more detail, including the identification of precursor cleavage sites and the determination of the exact amino termini of the two major cleavage products, GN and GC.60 The subtilase SKI-1 has been identified as the cellular protease responsible for the processing step that generates the amino terminus of mature GN.61, 62

Recent studies of the L RNA genome segment and predicted encoded L polymerase protein of CCHF virus demonstrate that they are approximately twice the size of those found in viruses of other bunyavirus genera. Regions containing ovarian tumour-like cysteine protease and helicase domains were identified in the L segments of CCHF and Dugbe viruses, suggesting an autoproteolytic cleavage process for nairovirus L proteins.63, 64

Phylogenetic studies and worldwide diversity

In 1970, when the virus was first named as CCHF virus,15 the antigenic structures of the viruses from various geographic regions were thought to be indistinguishable. However, the development of nucleic acid sequence analysis techniques revealed extensive genetic diversity. Most nucleic acid sequence analyses are based on the S segment of the genome, although some recent studies were done on the M RNA segment. According to these studies, there are eight genetically distinct clades.15

Because of their low genetic divergence, the European strains are grouped together, except the Greek strain AP92.16, 17 The strains from southeast Russia are closely related to the European strains.62, 65 Turkish CCHF virus isolates from the recent outbreak are clustered closely with CCHF viral strains from southwest Russia and Kosovo. Bootstrap analysis showed the clade containing the Russian, Balkan, and Turkish CCHF viruses to be well supported (99%), and these viruses are clearly distinct from those in other virus clades, including the clade containing the virus detected in the CCHF outbreak in neighbouring Iran in 2002.6

The AP92 strain, isolated from Rhipicephalus bursa ticks from Greece, differs from European strains66 and forms an independent clade. This genetic difference may be attributable to the different species of ticks vector and/or to genetic isolation by adjacent mountain ranges.16

The third clade is formed from the strains from central Asia—namely Kazahkstan,67 Tajikistan,68 Uzbekistan,68 and China68—which are closely related. Phylogenetic analysis of the M segment showed that the Chinese CCHF virus isolates were clustered into three groups, one of which was more closely related to a Nigerian isolate.69

Isolates from Iran, Madagascar, and Pakistan were found to be closely related, forming a fourth clade.16 Partial S-segment nucleotide sequences showed that the Iranian viral isolates clustered along with strains from Pakistan and Madagascar in one distinct lineage.70 Phylogenetic analysis also demonstrated that the Iranian isolates examined in this study and the CCHF virus strain ArTeh193-3 clustered into different genetic groups, indicating that at least two genetic lineages of CCHF virus could be co-circulating in Iran.70

The second group of strains from Iran is closely related to strains from Senegal and Mauritania, which together form a fifth clade.16

Finally, there are three distinct clades in Africa: first, strains from Senegal, Mauritania, and South Africa; second, strains from Nigeria and Central African Republic; and third, strains from Uganda.

Ecology

Vertebrate reservoir hosts

CCHF virus circulates in an enzootic tick–vertebrate–tick cycle, and there is no evidence that the virus causes disease in animals. CCHF viral infection has been commonly demonstrated among smaller wildlife species—eg, hares and hedgehogs—that act as hosts for the immature stages of the tick vectors.1, 2 Antibodies against CCHF virus have been detected in the sera of horses, donkeys, goats, cattle, sheep, and pigs in various regions of Europe, Asia, and Africa.2 It must be borne in mind that antibody studies, particularly if the prevalence is low, are not as meaningful as obtaining actual virus isolates. Although no ground-feeding birds have shown detectable viraemia,71 birds may have a role in the transportation of CCHF virus-infected ticks between different countries.

Tick vectors

CCHF viruses are transmitted by Hyalomma genus ticks, particularly by Hyalomma marginatum marginatum. CCHF virus was isolated from adult Hyalomma genus ticks in the 1960s.1, 2 Viral isolates were also obtained from field-collected eggs and unfed immature stages of H marginatum, providing evidence of transovarial (ie, from infected mother to egg stage), and transstadial (ie, from larvae to nymph to adult) transmissions.2

The known occurrence of CCHF in Europe, Asia, and Africa coincides with the global distribution of hyalomma ticks.1, 2 H marginatum marginatum is known as the Mediterranean hyalomma, and it may be the main vector of CCHF virus in Europe. CCHF virus has also been isolated from Hyalomma anatolicum anatolicum and other Hyalomma spp. Isolates from other tick genera—eg, Rhipicephalus, Ornithodoros, Boophilus, Dermacentor, and Ixodes spp1, 2—may be locally important because some have transmitted virus in the laboratory, but viral isolation alone does not incriminate them as vectors without additional laboratory and epidemiological studies.2

Climate change

Changes in climatic conditions have been suggested to be one of the factors that has facilitated reproduction of the tick population, and consequently the increased incidence of tick-borne infectious diseases.72, 73 In the northern hemisphere, H marginatum marginatum is usually activated by increasing temperature in the spring, particularly in April or May, and the immature stages are active in the summer between May and September.74 For example, in the Ukrainian steppes in 1963–64, the first adult hyalommas appeared when average daily temperatures reached 5–9°C on April 8 in 1963, and April 20 the following year.1 Tick densities were reduced by the severe winter of 1968–69 in Astrakhan Oblast (a federal republic of Russia), and consequently the number of the cases of CCHF was drastically reduced.1 The number of days with a temperature of over 5°C in April, and the daily mean temperature in April in the region of Turkey affected by the recent outbreak were reported to be increased in the years before the outbreak.75 However, climate change is not necessarily the cause of the marked increased incidence of a variety of tick-borne diseases in many parts of Europe over the past two decades.76

In general, CCHF outbreaks have developed against a background of favourable climatic factors and environmental changes beneficial for the survival of large numbers of Hyalomma spp ticks and of the hosts of both their immature and adult stages.1 In the former Soviet Union, environmental changes include wartime neglect of agricultural lands, introduction of susceptible military personnel or new settlers into an infected area, wide scale collectivisation of agriculture, changing pasture patterns, converting floodplains to farmland, and flood control.1 During World War 2, after the occupation of Crimea (1941–44), normal agricultural activities were disrupted and the common sport of hunting European hares was abandoned. When Soviet troops reoccupied the hilly Crimean steppes in 1944, hares had become excessively abundant and neglected pastures were overgrown with weeds, and the first outbreak of the modern era was documented.1 Interestingly, a similar explanation was suggested for the outbreak in Turkey:75 the fields in the affected region had been abandoned from hunting and pasturing between 1995 and 2001 because of terrorist activities in the region; in this period, the numbers of small mammals (eg, hares) and wild animals (eg, boars) increased. After 2001, the fields became available again for hunting and pasturing, and cattle and sheep were exposed to virus-carrying ticks.

The potential roles of migratory birds and the movement of livestock carrying ticks in the spread of the virus over distant geographic areas have been studied.1, 77, 78 Birds migrating from the Balkans were suggested to be the cause of the 2002 outbreak in Turkey.6 However, there is no precise data on CCHF virus in birds and on bird-parasitising ticks.

Risk factors for infection among human beings

Epidemiologically, CCHF cases are distributed mainly among actively working age groups exposed to tick populations. The major at-risk group are farmers living in endemic areas; most of the affected cases deal with agriculture and/or animal husbandry. Almost 90% of the cases in the recent outbreak in Turkey were farmers.3, 6, 7 Although there is no evidence that the virus causes disease in animals, CCHF virus-infected individuals were reported after skin contact with livestock and other animals.1, 2, 37, 43 Veterinarians and abattoir workers who work with large domestic animals are also an at-risk group; acquisition of the virus usually takes place while slaughtering animals.27, 32, 37, 77, 79 Viraemic blood from subclinically infected animals was the most likely source of infection, but exposure to ticks during these processes is also likely, at least in some of the cases.38, 79 Meat itself is not a risk because the virus is inactivated by post-slaughter acidification of the tissues and would not survive cooking in any case.

Hiking, camping, and other rural activities are also a risk factor for tick exposure. Gender distribution varies between countries, depending on the participation of women in agricultural work.

Outbreaks have recently been reported in South Africa when heavily tick-infested ostriches were slaughtered.38, 80 Infection was reported to be acquired either by contact with ostrich blood or inadvertently crushing infected ticks while skinning ostriches. Although no antibody was detected among birds during the outbreak,81, 82 ostriches have been experimentally infected, and viraemia was observed for 1–4 days after infection.80 As a public-health measure, this study suggests that birds should be kept free of ticks for 14 days before slaughtering.

Health-care workers are the second most affected group. Hospital health-care workers are at serious risk of transmission of CCHF infection when caring for patients with haemorrhages from the nose, mouth, gums, vagina, and injection sites. The transmission of the CCHF infections and deaths among health-care workers has been reported in parallel with outbreaks in the general population (table 2 ).

Table 2

Nosocomial infections of health-care workers during outbreaks of CCHF

Location and referenceYear of outbreakNumber of casesFatal cases
Bulgaria21953–654217
Pakistan201976113
Iraq25197922
Dubai23197952
South Africa38198482
Pakistan2119943Not known
Pakistan48200221
Albania50200210
Mauritania49200355
Turkey (unpublished data)200520

In one hospital outbreak, it was reported that 8·7% of health-care workers who were exposed to infected blood and 33% of those who had a needlestick injury developed the disease.41 CCHF virus has repeatedly caused nosocomial outbreaks with high mortality, and percutaneous exposure presents the highest risk of transmission.11, 37, 38, 41 The most dangerous settings for acquiring CCHF virus are interventions to gastrointestinal bleedings, and emergency operations on patients that have yet to be diagnosed with CCHF.42 In general, these patients were diagnosed after the operation, and injuries to the operating team during the operation are usually under-reported. Airborne acquisition of the infection was suspected in several cases in Russia,1, 2 but were not documented. Horizontal transmission from a mother to her child has also been reported.83

Course of infection and clinical features

Human beings are the only known host of CCHF virus in which disease is manifested.2, 4 In one Russian study84 the probability of developing CCHF for people who had been infected was found to be 0·215—ie, one of every five infected people develops CCHF.

The typical course of CCHF infection has four distinct phases: incubation, prehaemorrhagic, haemorrhagic, and convalescence periods (figure 3 ).1 The incubation period that follows a tick bite is usually short—3–7 days—but it is difficult to obtain precise data.1, 44 The incubation period could differ depending on several factors including viral dose and route of exposure—eg, it could be shorter with bloodborne transmissions. In South Africa, the time to onset of disease after exposure to tick bite was 3·2 days, 5 days following exposure to blood or tissue of livestock, and 5·6 days after exposure to blood of infected human beings.43 The mean duration of the disease course before presenting at a hospital was reported to be 5·5 days in Turkey3 and 3·5 days in the UAE.24

An external file that holds a picture, illustration, etc.
Object name is gr3_lrg.jpg

Clinical and laboratory course of CCHF

DIC=disseminated intravascular coagulation.

The prehaemorrhagic period is characterised by the sudden onset of fever (39–41°C), headache, myalgia, and dizziness.1, 2, 3, 4, 6, 7, 24, 44 On average, fever persists for 4–5 days.1 Additional symptoms of diarrhoea, nausea, and vomiting are also seen in some cases.2, 24, 44 Hyperaemia of the face, neck, and chest, congested sclera, and conjunctivitis are commonly noted. The prehaemorrhagic period lasts an average of 3 days (range: 1–7 days).1

The haemorrhagic period is short (usually 2–3 days), develops rapidly, and usually begins between the third to fifth day of disease. There is no relation between the temperature of the feverish patient and onset of haemorrhage.1 Haemorrhagic manifestations range from petechiae to large haematomas appearing on the mucous membranes and skin (figure 4 ). Bleeding from other sites, including the vagina, gingival bleeding, and cerebral haemorrhage have been reported.43 The most common bleeding sites are the nose, gastrointestinal system (haematemesis, melena, and intra-abdominal), uterus (menometrorrhagia) and urinary tract (haematuria), and the respiratory tract (haemoptysis).2, 85 Atypical presentations of bleeding are also seen. For example, in one patient with stubborn abdominal pain, acute appendicitis was suspected, but haemorrhage and bleeding in the internal and external oblique muscles and caecum were detected, with no pathology of the appendix.86 Hepatomegaly and splenomegaly have been reported to be occur in one-third of patients.1 In Turkey, hepatomegaly was detected in 20–40% of cases,3, 6, 7, 8 and two studies reported splenomegaly, with frequencies of 14% and 23%.7, 8

An external file that holds a picture, illustration, etc.
Object name is gr4_lrg.jpg

CCHF patients from the Ankara Numune Education and Research Hospital

The convalescence period begins in survivors about 10–20 days after the onset of illness. Patients remain in hospital for around 9–10 days.3, 24 In the convalescent period, labile pulse, tachycardia, temporary complete loss of hair, polyneuritis, difficulty in breathing, xerostomia, poor vision, loss of hearing, and loss of memory have been reported,1 although none of these findings were noted in the recent outbreak in Turkey. Although cardiovascular changes—eg, bradycardia and low blood pressure—were reported in an earlier review,1 these have not been emphasised recently.3, 6, 7, 24 Hepatorenal insufficiency was reported in South Africa6 but not in Turkey. There is no known relapse of the infection, and a biphasic course of the disease—as noted in published work from the former Soviet Union1—was not observed in Turkey.

Biochemical tests

Thrombocytopenia appears to be a consistent feature of CCHF infection.2, 24, 44 Patients had leucopenia and raised levels of aspartate aminotransferase, alanine aminotransferase, lactate dehydrogenase, and creatinine phosphokinase. Coagulation tests such as prothrombin time and activated partial thromboplastin time are prolonged. The level of fibrinogen might be decreased, and fibrin degradation products could be increased. Laboratory tests, including complete blood count, and biochemical tests returned to normal levels within approximately 5–9 days among surviving patients (figure 5 ).85

An external file that holds a picture, illustration, etc.
Object name is gr5_lrg.jpg

The course of the disease in accordance with the thrombocyte level2

Predictors of mortality

Swanepoel and colleagues30 described clinical laboratory criteria that could be measured early in the course of disease (during the first 5 days) and that predicted a fatal outcome in 90% of patients with any of these findings: white blood cell count of 10 × 109 cells per L or above, platelet count of 20 × 109 per L or below, aspartate aminotransferase level of 200 U/L or over, alanine aminotransferase of 150 U/L or over, activated partial thromboplastin time of 60 seconds or more, or fibrinogen levels of 110 mg/dL or under. Other case series have confirmed that levels of aspartate aminotransferase and alanine aminotransferase are significantly higher among severe cases (p<0·05).3, 85 In a study from Turkey, higher aspartate aminotransferase and alanine aminotransferase levels (>700 and >900 IU/L, respectively) were found to have higher sensitivity for severe cases.85 Leucocytosis was observed in only one out of four fatal patients.85

In keeping with the prognostic importance of abnormalities of activated partial thromboplastin time and fibrinogen noted by Swanepoel and coworkers,44 patients with fatal outcomes in other series have had overt disseminated intravascular coagulation, according to criteria defined by the International Society of Thrombosis and Haemostasis.85, 87 Haematemesis, melena, and somnolence are significantly more common among patients with a fatal outcome (p=0·009, p=0·001, and p=0·022, respectively).7, 85 Of particular importance is the fact that in fatal cases there is little evidence of an antibody response.39, 85

Pathogenesis

The pathogenesis of CCHF is not well described. A common pathogenic feature of haemorrhagic fever viruses is their ability to disable the host immune response by attacking and manipulating the cells that initiate the antiviral response.88 This damage is characterised by marked replication of the virus together with dysregulation of the vascular system and lymphoid organs.89

Infection of the endothelium has an important role in CCHF pathogenesis.90, 91 The endothelium can be targeted in two ways—indirectly by viral factors or virus-mediated host-derived soluble factors that cause endothelial activations and dysfunction, and/or directly by virus infection and replication in endothelial cells.90 Endothelial damage contributes to haemostatic failure by stimulating platelet aggregation and degranulation, with consequent activation of the intrinsic coagulation cascade. Indeed, fatal CCHF cases had grossly abnormal indicators of coagulation system function from an early stage of illness, and disseminated intravascular coagulation is noted as an early and promiment feature of the disease process.

In one study from Turkey, reactive haemophagocytosis was detected in seven (50%) of 14 patients, which suggested that haemophagocytosis could have a role in the cytopenia observed during CCHF infection.6 Because haemophagocytic lymphohistiocytosis92 has been attributed to excessive activation of monocytes by high levels of Th1 cytokines—eg, interferon gamma, tumour necrosis factor alpha, interleukin 1, or interleukin 6—this finding provides indirect evidence for the participation of cytokines in other aspects of CCHF pathogenesis. In one study of CCHF patients, the levels of interleukin 1, interleukin 6, and tumour necrosis factor alpha were higher among those patients that subsequently died compared with those that survived.93 The disseminated intravascular coagulation score was higher among fatal cases, correlating positively with interleukin 6 and tumour necrosis factor alpha levels, and negatively with interleukin 10 levels.93

Diagnosis

Early diagnosis is critical both for patient survival and for the prevention of potential nosocomial infections and transmission in the community. Suspected cases should be evaluated and their management carefully planned, including supportive care, particularly haematological support (panel ). The differential diagnosis list differs according to geographic location, and includes bacterial, viral, and non-infectious causes (table 3 ).

Panel

An algorithm for case management

Evaluation of a suspected case

  • Clinical symptoms (fever, myalgia, bleeding from various sites)
  • Patient history: referral from endemic area; outdoor activities (picnic, tracking, etc) in endemic area; history of tick exposure; exposure to potentially viraemic domestic animal blood
  • Laboratory tests (low platelet and high white blood cell count, raised levels of aspartate aminotransferase, alanine aminotransferase, lactate dehydrogenase, and creatinine phosphokinase)

Preventive measures

  • Isolate the patient
  • Inform and educate colleagues and staff
  • Use barrier precautions

Investigations for confirmation

  • Serum for PCR (early in disease) and ELISA (late in disease or convalescence): IgM positivity or PCR positive confirms diagnosis, IgG positivity cannot; sera for differential diagnosis

Decision making for therapy

  • Define the severity criteria and decide on ribavirin use or not.
  • Do not neglect other causes of clinical picture. Starting doxycycline or equivalent should be considered
  • Haematological support: fresh frozen plasma to improve haemostasis; thrombocyte solutions
  • Respiratory support

Follow up

  • No relapse occurs after the disease. Therefore there is no need for follow up of cases
  • Health-care workers exposed to the virus should be followed up with complete blood counts and biochemical tests for 14 days

Table 3

Differential diagnosis of CCHF

DiseaseGeographic locationTransmissionDifferentials with CCHF (clinical or laboratory findings)
Infections
BrucellosisWorldwide, particularly Mediterranean basin, the Arabian peninsula, the Indian subcontinent, and in parts of Mexico, Central America, and South AmericaPancytopenia, Wright agglutination
Q feverWorldwideTickSerology (ELISA or IFAT)
RickettsiaWorldwideTickWeil-Felix test
EhrlichiosisAmerica, Europe, middle east, southeast AsiaTickSerology (ELISA)
LymeWorldwide, mainly northern hemisphereTickSerology (ELISA), western blot
LeptospiraWorldwideRodentsAgglutination test
SalmonellaWorldwideNAWidal test
Tick-borne encephalitisNorthern hemisphereTickELISA
MalariaWorldwideMosquitoPeripheral smear
Other viral haemorrhagic infections
Arenaviridae
South America haemorrhagic feverArgentina, Bolivia, Brasil, and VenezuelaInterhumanNeurological symptoms
Lassa feverWest AfricaInterhumanPharyngitis, retrosternal pain, proteinuria, central nervous system involvement
Other Bunyaviridae
Rift Valley feverSub-Saharan AfricaMosquitoHepatitis, retinal vasculitis, encephalitis
Hanta fever with renal syndromeWorldwideRodentsRenal findings, serology, PCR
Hantavirus pulmonary syndromeAmericaInterhumanPulmonary findings, serology, PCR
Filoviridae
Marburg and EbolaAfrica, PhilippinesInterhumanMarked weight loss and prostration.
Hepatitis, uveitis, orchitis, arthralgia in convalescence
Filaviviridae
Yellow feverAfrica, South AmericaMosquitoJaundice
DengueTropics and subtropics, worldwideMosquitoGeneralised macular rash
Kyasanur forest diseaseIndiaTickHaemorrhagic pulmonary oedema, renal failure, neurological symptoms
Omsk haemorrhagic feverWestern SiberiaTickNeuropsychiatric sequelae
Al KhumrahMiddle east, AfricaTick (?), mosquito (?)Not known
Non-infectious
Vitamin B12 deficiencyWorldwidePancytopenia, and B12 level in serum
Febrile neutropeniaWorldwideUnderlying disease

NA=not applicable

Virus isolation

Virus isolation studies should be done in high-containment laboratories, generally recommended to be biosafety level four. Isolation in cell culture is simpler and more rapid, but less sensitive, than traditional methods such as intracranial inoculation of a sample into newborn mice.94 Virus can be isolated using cell lines including LLC-MK2, Vero, BHK-21, and SW-13.4 Virus isolation can be achieved in 2–5 days, but cell cultures lack sensitivity, and usually only allow detection of the relatively high viraemia encountered during the first 5 days of illness. The virus may produce little or no cytopathic effect, but can be identified by doing immunofluorescence assay tests with specific monoclonal antibodies.4 Although reverse transcriptase PCR is extremely useful for rapid diagnosis, only virus isolation yields a virus that can be subjected to further biological and sequencing studies.

Antigen capture ELISA has also been shown to be useful. In spite of its relative lack of sensitivity, this process can detect the most severe cases that would require antiviral therapy or could be candidates for a trial of an antiviral drug.94, 95

Molecular methods

Reverse transcriptase PCR is the method of choice for rapid laboratory diagnosis of CCHF virus infection.96 The method is highly specific, sensitive, and rapid.97 A further improvement has been the development of automated real-time assays, which have a lower contamination rate, higher sensitivity and specificity, and are more rapid than conventional reverse transcriptase PCR.4, 98

Serology

IgM and IgG antibodies are detectable by ELISA and immunofluorescence assays from about 7 days after the onset of disease.99 Specific IgM declines to undetectable levels by 4 months post-infection, but IgG remains detectable for at least 5 years. Recent or current infection is confirmed by demonstrating seroconversion, or a fourfold or greater increase in antibody titre in paired serum samples, or IgM antibodies with IgM antibody capture (MAC)-ELISA in a single sample.100 ELISA methods are quite specific and much more sensitive than immunofluorescence assays and neutralisation tests.95 Recently, a recombinant nucleoprotein-based IgG ELISA for serological diagnosis of CCHF virus infections was developed.101

Treatment

Supportive therapy is the most essential part of case management, and includes the administration of thrombocytes, fresh frozen plasma, and erythrocyte preparations. Replacement therapy with these blood products should be done after checking the patient's complete blood count, which should be done once or twice a day. Potential bleeding foci should be considered and conservative measures taken—eg, the use of histamine receptor blockers for peptic ulcer patients, avoidance of intramuscular injections, and not using aspirin or other drugs with actions on the coagulation system. Fluid and electrolyte balance should also be monitored meticulously.

Ribavirin is the recommended antiviral agent for infected patients, although its mechanism of action is not clear. In one in-vitro study,102 ribavirin was shown to inhibit viral activity, and some CCHF viral strains appeared more sensitive than others. In an experimental study done in mice,103 ribavirin treatment substantially reduced infant mouse mortality and extended the mean time to death. It should be noted that there is no evidence from randomised clinical trials for the use of ribavirin to treat human CCHF—its effectiveness has only been described in observational studies.3, 22, 104

Mild cases do not need to be treated with ribavirin.3 In case management, severe cases should be defined and treated. Severe cases in Turkey are defined according to a revised form of the Swanepoel criteria.85 Oral and intravenous forms of ribavirin are available in many countries. Patients should be treated for 10 days (30 mg/kg as an initial loading dose, then 15 mg/kg every 6 hours for 4 days, and then 7·5 mg/kg every 8 hours for 6 days).105 Haemolytic anaemia, hypocalcaemia, and hypomagnesaemia were reported in patients that received ribavirin to treat severe acute respiratory syndrome.106, 107 However, no adverse events related to ribavirin therapy were noted among CCHF patients in Turkey. The use of ribavirin is contraindicated in pregnant women.

One study suggested treatment using passive immunotherapy, transferring the plasma of convalescing survivors to infected patients.108 However, the study had no control groups and was limited to seven patients.

Paragas and colleagues109 screened drugs for potential activity against CCHF virus and found that ribavirin inhibited the replication of CCHF virus, ribamidine had antiviral activity that was 4·5-fold to eightfold less than that of ribavirin, and three other drugs (6-azauridine, selenazofurin, and tiazofurin) had no significant antiviral activity. A newly identified molecule known as MxA, which is a member of the interferon-induced GTPases that belong to the dynamin superfamily, prevented the replication of CCHF viral RNA when present intracellularly,110 and inhibited the production of new infectious virus particles by interacting with a component of the nucleocapsid.

Prevention

People living in endemic areas should use personal protective measures that include the avoidance of areas where tick vectors are abundant, particularly when they are active; regular examination of clothing and skin for ticks, and their removal; and the use of repellents.105 People who are exposed to potentially viraemic animal blood should take practical measures to protect themselves, including the use of repellents on the skin and clothing and wearing gloves or other protective clothing to prevent skin contact with infected tissue or blood.105

The recommended safety measures for health-care workers include barrier nursing and isolation, and the use of gloves, gowns, face-shields, and goggles with side shields when in contact with patients or soiled environmental surfaces. Simple barrier precautions have been reported to be effective.111 Strict adherence to universal barrier precautions in hospitals in Turkey during the recent outbreak meant that no antibodies against CCHF virus were found in at-risk health-care workers when serologically screened after the outbreak.112 In a recent study, prophylactic ribavirin was administered to a health-care worker who had a needlestick injury. The health-care worker did not subsequently develop CCHF.22 The administration of oral ribavirin prophylaxis was offered to anyone at high risk of infection, such as those who were directly exposed to the blood of CCHF patients through contact or needlestick injury. However, rigorous daily follow-up of any such individual by checking white blood cell counts and biochemical tests for at least 14 days after the exposure should be sufficient; ribavirin should only be administered if fever develops.

In 1974, an immunisation programme was introduced for medical workers and military personnel in CCHF-endemic areas.18 The vaccine consisted of a mouse brain preparation inactivated by chloroform, heated at 58°C, and adsorbed on aluminium hydroxide. It was proposed that the vaccine was helpful in reducing the number of cases and the case fatality rate in this area.18 Vaccine was also given to 583 human volunteers in Bulgaria, and was reported to elicit antibody production in 96·6% of these individuals.2 However, experiences with vaccines against CCHF virus are limited of limited duration and confined to a few locations. The vaccine would not be suitable for use in many countries because of its method of preparation.

Future research areas

The dynamics of the enzootic environment and transmission cycle of the CCHF virus need to be further detailed. The role of climatic factors, reservoir hosts, and vectors should be described. These studies need multidisciplinary team work, including entomologists, microbiologists, epidemiologists, veterinarians, and clinicians. New data on viral replication offers substantial potential for the development of new drugs. Further studies on the pathogenesis of viral haemorrhagic fevers will shed light on the mechanisms of disseminated intravascular coagulation and probably bacterial sepsis. Understanding new mechanisms of CCHF viral infection or other viral haemorrhagic fevers will assist in the development of new therapeutic molecules. Agents used to treat disseminated intravascular coagulation—eg, heparin or other coagulation blockers—could be tried in the control of clinical course of CCHF.

Search strategy and selection criteria

Data for this review were identified by searches of PubMed, Google, and references from relevant articles, using the search terms “Crimean Congo haemorrhagic fever”, “viral haemorrhagic fever”, and “ribavirin”. No date limits were set; only the abstracts of non-English language papers were included.

Acknowledgments

I thank my colleagues Başak Dokuzoğuz, Aysel Çelikbaş, Nurcan Baykam, Şebnem Eren, Harika Esener, Mustafa Eroğlu, and Salim Yaprakçi, for their invaluable contributions. I also thank the residents, the nurses, and the staff of the Infectious Diseases and Clinical Microbiology Clinic of Ankara Numune Education and Research Hospital.

Conflicts of interest

I declare that I have no conflicts of interest.

References

1. Hoogstraal H. The epidemiology of tick borne Crimean-Congo hemorrhagic fever in Asia, Europe, and Africa. J Med Entomol. 1979;15:307–417. [PubMed] [Google Scholar]
2. Watts DM, Ksiazek TG, Linthicum KJ, Hoogstraal H. Crimean-Congo hemorrhagic fever. In: Monath TP, editor. volume 2. CRC Press; Boca Raton, FL, USA: 1988. pp. 177–260. (The arboviruses: epidemiology and ecology). [Google Scholar]
3. Ergonul O, Celikbas A, Dokuzoguz B, Eren S, Baykam N, Esener H. The characteristics of Crimean-Congo hemorrhagic fever in a recent outbreak in Turkey and the impact of oral ribavirin therapy. Clin Infect Dis. 2004;39:285–289. [PubMed] [Google Scholar]
4. Whitehouse CA. Crimean-Congo hemorrhagic fever. Antivir Res. 2004;64:145–160. [PubMed] [Google Scholar]
5. Centers for Disease Control and Prevention Bioterrorism agents/diseases. http://www.bt.cdc.gov/Agent/Agentlist.asp (accessed Feb 16, 2006)
6. Karti SS, Odabasi Z, Korten V. Crimean-Congo hemorrhagic fever in Turkey. Emerg Infect Dis. 2004;19:1379–1384. [PMC free article] [PubMed] [Google Scholar]
7. Bakir M, Ugurlu M, Dokuzoguz B, Bodur H, Tasyaran MA, Vahaboglu H. Crimean-Congo haemorrhagic fever outbreak in Middle Anatolia: a multicentre study of clinical features and outcome measures. J Med Microbiol. 2005;54:385–389. [PubMed] [Google Scholar]
8. Ozkurt Z, Kiki I, Erol S. Crimean-Congo hemorrhagic fever in Eastern Turkey: clinical features, risk factors and efficacy of ribavirin therapy. J Infect. 2006;52:207–215. [PubMed] [Google Scholar]
9. Ministry of Health Turkey. Reports of the Communicable Diseases Department, Ankara, 2005 (in Turkish).
10. Simpson DIH, Knight EM, Courtois G, Williams MC, Weinbern MP, Kibukamusoke JW. Congo virus: a hitherto undescribed virus occuring in Africa. Human isolations-clinical notes. East Afr Med J. 1967;44:87. [PubMed] [Google Scholar]
11. Casals J. Antigenic similarity between the virus causing Crimean hemorrhagic fever and Congo virus. Proc Soc Exp Biol Med. 1969;131:233–236. [PubMed] [Google Scholar]
12. Begum F, Wisseman CL, Jr, Casals J. Tick-borne viruses of west Pakistan. IV. Viruses similar to or identical with, Crimean hemorrhagic fever (Congo-Semunya), Wad Medani and Pak Argas 461 isolated from ticks of the Changa Manga Forest, Lahore District, and of Hunza, Gilgit Agency, W. Pakistan. Am J Epidemiol. 1970;92:197–202. [PubMed] [Google Scholar]
13. Woodall JP, Williams MC, Simpson DI. Congo virus: a hitherto undescribed virus occurring in Africa. II. Identification studies. East Afr Med J. 1967;44:93–98. [PubMed] [Google Scholar]
14. Causey OR, Kemp GE, Madbouly MH, David-West TS. Congo virus from domestic livestock, African hedgehog, and arthropods in Nigeria. Am J Trop Med Hyg. 1970;19:846–850. [PubMed] [Google Scholar]
15. Casals J, Henderson BE, Hoogstraal H, Johnson KM, Shelokov A. A review of Soviet viral hemorrhagic fevers, 1969. J Infect Dis. 1970;122:437–453. [PubMed] [Google Scholar]
16. Papa A, Christova I, Papadimitriou E, Antoniadis A. Crimean-Congo hemorrhagic fever in Bulgaria. Emerg Infect Dis. 2004;10:1465–1467. [PMC free article] [PubMed] [Google Scholar]
17. Papa A, Bino S, Llagami A. Crimean-Congo hemorrhagic fever in Albania, 2001. Eur J Clin Microbiol Infect Dis. 2002;21:603–606. [PubMed] [Google Scholar]
18. Papa A, Bozovic B, Pavlidou V, Papadimitriou E, Pelemis M, Antoniadis A. Genetic detection and isolation of Crimean-Congo hemorrhagic fever virus, Kosovo, Yugoslavia. Emerg Infect Dis. 2002;8:852–854. [PMC free article] [PubMed] [Google Scholar]
19. Papa A, Ma B, Kouidou S, Tang Q, Hang C, Antoniadis A. Genetic characterization of the M RNA segment of Crimean Congo hemorrhagic fever virus strains, China. Emerg Infect Dis. 2002;8:50–53. [PubMed] [Google Scholar]
20. Burney MI, Ghafoor A, Saleen M, Webb PA, Casals J. Nosocomial outbreak of viral hemorrhagic fever caused by Crimean hemorrhagic fever-Congo virus in Pakistan, January 1976. Am J Trop Med Hyg. 1980;29:941–947. [PubMed] [Google Scholar]
21. Sheikh AS, Sheikh AA, Sheikh NS. Bi-annual surge of Crimean-Congo haemorrhagic fever (CCHF): a five-year experience. Int. J Infect Dis. 2005;9:37–42. [PubMed] [Google Scholar]
22. Smego RA, Sarwari AR, Siddiqui AR. Crimean-Congo hemorrhagic fever: Prevention and control limitations in a resource poor country. Clin Infect Dis. 2004;38:1731–1735. [PubMed] [Google Scholar]
23. Suleiman MN, Muscat-Baron JM, Harries JR. Congo/Crimean haemorrhagic fever in Dubai. An outbreak at the Rashid hospital. Lancet. 1980;2:939–941. [PubMed] [Google Scholar]
24. Schwarz TF, Nsanze H, Ameen AM. Clinical features of Crimean-Congo haemorrhagic fever in the United Arab Emirates. Infection. 1997;25:364–367. [PubMed] [Google Scholar]
25. Al-Tikriti SK, Al-Ani F, Jurji FJ. Congo/Crimean haemorrhagic fever in Iraq. Bull World Health Organ. 1981;59:85–90. [PMC free article] [PubMed] [Google Scholar]
26. Tantawi HH, Al-Moslih MI, Al-Janabi NY. Crimean-Congo haemorrhagic fever virus in Iraq: isolation, identification and electron microscopy. Acta Virol. 1980;24:464–467. [PubMed] [Google Scholar]
27. El-Azazy OM, Scrimgeour EM. Crimean-Congo haemorrhagic fever virus infection in the western province of Saudi Arabia. Trans R Soc Trop Med Hyg. 1997;91:275–278. [PubMed] [Google Scholar]
28. Williams RJ, Al-Busaidy S, Mehta FR. Crimean-Congo haemorrhagic fever: a seroepidemiological and tick survey in the Sultanate of Oman. Trop Med Int Health. 2000;5:99–106. [PubMed] [Google Scholar]
29. Mardani M, Jahromi MK, Naieni KH, Zeinali M. The efficacy of oral ribavirin in the treatment of crimean-congo hemorrhagic fever in Iran. Clin Infect Dis. 2003;36:1613–1618. [PubMed] [Google Scholar]
30. Saluzzo JF, Aubry P, McCormick J, Digoutte JP. Haemorrhagic fever caused by Crimean Congo haemorrhagic fever virus in Mauritania. Trans R Soc Trop Med Hyg. 1985;79:268. [PubMed] [Google Scholar]
31. Antoniadis A, Casals J. Serological evidence of human infection with Congo-Crimean hemorrhagic fever virus in Greece. Am J Trop Med Hyg. 1982;31:1066–1067. [PubMed] [Google Scholar]
32. Dunster L, Dunster M, Ofula V. First documentation of human Crimean-Congo hemorrhagic fever, Kenya. Emerg Infect Dis. 2002;8:1005–1006. [PMC free article] [PubMed] [Google Scholar]
33. Yen YC, Kong LX, Lee L, Zhang YQ, Li F, Cai BJ, Gao SY. Characteristics of Crimean-Congo hemorrhagic fever virus (Xinjiang strain) in China. Am J Trop Med Hyg. 1985;34:1179–1182. [PubMed] [Google Scholar]
34. Gear JH, Thomson PD, Hopp M. Congo-Crimean haemorrhagic fever in South Africa. Report of a fatal case in the Transvaal. S Afr Med J. 1982;62:576–580. [PubMed] [Google Scholar]
35. Swanepoel R, Struthers JK, Shepherd AJ, McGillivray GM, Nel MJ, Jupp PG. Crimean-congo hemorrhagic fever in South Africa. Am J Trop Med Hyg. 1983;32:1407–1417. [PubMed] [Google Scholar]
36. Swanepoel R, Shepherd AJ, Leman PA, Shepherd SP. Investigations following initial recognition of Crimean-Congo haemorrhagic fever in South Africa and the diagnosis of 2 further cases. S Afr Med J. 1985;68:638–641. [PubMed] [Google Scholar]
37. Swanepoel R, Shepherd AJ, Leman PA, Shepherd SP, Miller GB. A common-source outbreak of Crimean-Congo haemorrhagic fever on a dairy farm. S Afr Med J. 1985;68:635–637. [PubMed] [Google Scholar]
38. van Eeden PJ, Joubert JR, van de Wal BW, King JB, de Kock A, Groenewald JH. A nosocomial outbreak of Crimean-Congo haemorrhagic fever at Tygerberg Hospital. Part I. Clinical features. S Afr Med J. 1985;68:711–717. [PubMed] [Google Scholar]
39. van Eeden PJ, van Eeden SF, Joubert JR, King JB, van de Wal BW, Michell WL. A nosocomial outbreak of Crimean-Congo haemorrhagic fever at Tygerberg Hospital. Part II. Management of patients. S Afr Med J. 1985;68:718–721. [PubMed] [Google Scholar]
40. Joubert JR, King JB, Rossouw DJ, Cooper R. A nosocomial outbreak of Crimean-Congo haemorrhagic fever at Tygerberg Hospital. Part III. Clinical pathology and pathogenesis. S Afr Med J. 1985;68:722–728. [PubMed] [Google Scholar]
41. van de Wal BW, Joubert JR, van Eeden PJ, King JB. A nosocomial outbreak of Crimean-Congo haemorrhagic fever at Tygerberg Hospital. Part IV. Preventive and prophylactic measures. S Afr Med J. 1985;68:729–732. [PubMed] [Google Scholar]
42. Shepherd AJ, Swanepoel R, Shepherd SP, Leman PA, Blackburn NK, Hallet AF. A nosocomial outbreak of Crimean-Congo haemorrhagic fever at Tygerberg Hospital. Part V. Virological and serological observations. S Afr Med J. 1985;68:733–736. [PubMed] [Google Scholar]
43. Swanepoel R, Shepherd AJ, Leman PA. Epidemiologic and clinical features of Crimean-Congo hemorrhagic fever in southern Africa. Am J Trop Med Hyg. 1987;36:120–132. [PubMed] [Google Scholar]
44. Swanepoel R, Gill DE, Shepherd AJ. The clinical pathology of Crimean-Congo hemorrhagic fever. Rev Infect Dis. 1989;11:794–800. [PubMed] [Google Scholar]
45. Saluzzo JF, Digoutte JP, Cornet M, Baudon D, Roux J, Robert V. Isolation of Crimean-Congo haemorrhagic fever and Rift Valley fever viruses in Upper Volta. Lancet. 1984;1:1179. [PubMed] [Google Scholar]
46. Chapman LE, Wilson ML, Hall DB. Risk factors for Crimean-Congo hemorrhagic fever in rural northern Senegal. J Infect Dis. 1991;164:686–692. [PubMed] [Google Scholar]
47. Baskerville A, Satti A, Murphy FA, Simpson DI. Congo-Crimean haemorrhagic fever in Dubai: histopathological studies. J Clin Pathol. 1981;34:871–874. [PMC free article] [PubMed] [Google Scholar]
48. Athar MN, Baqai HZ, Ahmad M. Short report: Crimean-Congo hemorrhagic fever outbreak in Rawalpindi, Pakistan, February 2002. Am J Trop Med Hyg. 2003;69:284–287. [PubMed] [Google Scholar]
49. Nabeth P, Thior M, Faye O, Simon F. Human Crimean-Congo hemorrhagic fever, Senegal. Emerg Infect Dis. 2004;10:1881–1882. [PMC free article] [PubMed] [Google Scholar]
50. Drosten C, Minnak D, Emmerich P, Schmitz H, Reinicke T. Crimean-Congo hemorrhagic fever in Kosovo. J Clin Microbiol. 2002;40:1122–1123. [PMC free article] [PubMed] [Google Scholar]
51. Nabeth P, Cheikh DO, Lo B. Crimean-Congo hemorrhagic fever, Mauritania. Emerg Infect Dis. 2004;10:2143–2149. [PMC free article] [PubMed] [Google Scholar]
52. Shanmugam J, Smirnova SE, Chumakov MP. Presence of antibody to arboviruses of the Crimean haemorrhagic fever-Congo (CHF-Congo) group in human beings and domestic animals in India. Indian J Med Res. 1976;64:1403–1413. [PubMed] [Google Scholar]
53. Darwish MA, Imam IZ, Omar FM, Hoogstraal H. Results of a preliminary seroepidemiological survey for Crimean-Congo hemorrhagic fever virus in Egypt. Acta Virol. 1978;22:77. [PubMed] [Google Scholar]
54. Filipe AR, Calisher CH, Lazuick J. Antibodies to Congo-Crimean haemorrhagic fever, Dhori, Thogoto and Bhanja viruses in southern Portugal. Acta Virol. 1985;29:324–328. [PubMed] [Google Scholar]
55. Horvath LB. Precipitating antibodies to Crimean haemorrhagic fever virus in human sera collected in Hungary. Acta Microbiol Acad Sci Hung. 1976;23:331–335. [PubMed] [Google Scholar]
56. van Regenmortel MHV, Fauquet CM, Bishop DML. 7th report of the International Committee of Taxonomy of Viruses. http://www.virustaxonomyonline.com/virtax/lpext.dll?f=templates&fn=main-h.htm (accessed Feb 17, 2006)
57. Elliott RM, Schmaljohn CS, Collett MS. Bunyaviridae genome structure and gene expression. Curr Top Microbiol Immunol. 1991;169:91–141. [PubMed] [Google Scholar]
58. Schmaljohn CS, Hooper JW. Bunyaviridae: the viruses and their replication. In: Knipe DM, Howley PM, Griffin DE, editors. Fields virology. Lippincott Williams & Wilkins; Philadelphia: 2001. pp. 1581–1602. [Google Scholar]
59. Haferkamp S, Fernando L, Schwarz TF, Feldmann H, Flick R. Intracellular localization of Crimean-Congo hemorrhagic fever (CCHF) virus glycoproteins. Virol J. 2005;2:42. [PMC free article] [PubMed] [Google Scholar]
60. Sanchez AJ, Vincent MJ, Nichol ST. Characterization of the glycoproteins of Crimean-Congo hemorrhagic fever virus. J Virol. 2002;76:7263–7275. [PMC free article] [PubMed] [Google Scholar]
61. Vincent MJ, Sanchez AJ, Erickson BR. Crimean-Congo hemorrhagic fever virus glycoprotein proteolytic processing by subtilase SKI-1. J Virol. 2003;77:8640–8649. [PMC free article] [PubMed] [Google Scholar]
62. Papa A, Papadimitriou E, Bozovic B, Antoniadis A. Genetic characterization of the M RNA segment of a Balkan Crimean-Congo hemorrhagic fever virus strain. J Med Virol. 2005;75:466–469. [PubMed] [Google Scholar]
63. Honig JE, Osborne JC, Nichol ST. Crimean-Congo hemorrhagic fever virus genome L RNA segment and encoded protein. Virology. 2004;321:29–35. [PubMed] [Google Scholar]
64. Kinsella E, Martin SG, Grolla A, Czub M, Feldmann H, Flick R. Sequence determination of the Crimean-Congo hemorrhagic fever virus L segment. Virology. 2004;321:23–28. [PubMed] [Google Scholar]
65. Yashina L, Vyshemirskii O, Seregin S. Genetic analysis of Crimean-Congo hemorrhagic fever virus in Russia. J Clin Microbiol. 2003;41:860–862. [PMC free article] [PubMed] [Google Scholar]
66. Papadopoulos O, Koptopoulos G. Crimean-Congo hemorrhagic fever (CCHF) in Greece: isolation of the virus from Rhipicephalus bursa ticks and a preliminary serological survey. Zentbl Bakteriol Hyg Abt. 1980;(suppl 9):189–193. [Google Scholar]
67. Yashina L, Petrova I, Seregin S. Genetic variability of Crimean-Congo haemorrhagic fever virus in Russia and central Asia. J Gen Virol. 2003;84:1199–1206. [PubMed] [Google Scholar]
68. Seregin SV, Samokhvalov EI, Petrova ID. Genetic characterization of the M RNA segment of Crimean-Congo hemorrhagic fever virus strains isolated in Russia and Tajikistan. Virus Genes. 2004;28:187–193. [PubMed] [Google Scholar]
69. Morikawa S, Qing T, Xinqin Z, Saijo M, Kurane I. Genetic diversity of the M RNA segment among Crimean-Congo hemorrhagic fever virus isolates in China. Virology. 2002;296:159–164. [PubMed] [Google Scholar]
70. Chinikar S, Persson SM, Johansson M. Genetic analysis of Crimean-Congo hemorrhagic fever virus in Iran. J Med Virol. 2004;73:404–411. [PubMed] [Google Scholar]
71. Zeller HG, Cornet JP, Camicas JL. Experimental transmission of Crimean-Congo hemorrhagic fever virus by west African wild ground-feeding birds to Hyalomma marginatum rufipes ticks. Am J Trop Med Hyg. 1994;50:676–681. [PubMed] [Google Scholar]
72. Gubler DJ, Reiter P, Ebi KL, Yap W, Nasci R, Patz JA. Climate variability and change in the United States: potential impacts on vector- and rodent-borne diseases. Environ Health Perspect. 2001;109:223–233. [PMC free article] [PubMed] [Google Scholar]
73. Estrada-Pena A. Forecasting habitat suitability for ticks and prevention of tick-born diseases. Vet Parasitol. 2001;98:111–132. [PubMed] [Google Scholar]
74. Walker RA, Bouttaour A, Camicas JL. Ticks of domestic animals in Africa. A guide to identification of species. Bioscience Reports; Edinburgh, UK: 2003. p. 114. [Google Scholar]
75. Ergonul O, Akgunduz S, Kocaman I, Vatansever Z, Korten V. Changes in temperature and the Crimean Congo hemorrhagic fever outbreak in Turkey. Clin Microbiol Infect. 2005;11(suppl 2):360. [Google Scholar]
76. Randolph SE. Evidence that climate change has caused ‘emergence’ of tick-borne diseases in Europe? Int J Med Microbiol. 2004;293(suppl 37):5–15. [PubMed] [Google Scholar]
77. Khan AS, Maupin GO, Rollin PE. An outbreak of Crimean-Congo hemorrhagic fever in the United Arab Emirates, 1994–1995. Am J Trop Med Hyg. 1997;57:519–525. [PubMed] [Google Scholar]
78. Rodriguez LL, Maupin GO, Ksiazek TG. Molecular investigation of a multisource outbreak of Crimean-Congo hemorrhagic fever in the United Arab Emirates. Am J Trop Med Hyg. 1997;57:512–518. [PubMed] [Google Scholar]
79. Fisher-Hoch SP, McCormick JB, Swanepoel R, Van Middlekoop A, Harvey S, Kustner HG. Risk of human infections with Crimean-Congo hemorrhagic fever virus in a South African rural community. Am J Trop Med Hyg. 1992;47:337–345. [PubMed] [Google Scholar]
80. Swanepoel R, Leman PA, Burt FJ. Experimental infection of ostriches with Crimean-Congo haemorrhagic fever virus. Epidemiol Infect. 1998;121:427–432. [PMC free article] [PubMed] [Google Scholar]
81. Shepherd AJ, Swanepoel R, Leman PA, Shepherd SP. Field and laboratory investigation of Crimean-Congo haemorrhagic fever virus (Nairovirus, family Bunyaviridae) infection in birds. Trans R Soc Trop Med Hyg. 1987;81:1004–1007. [PubMed] [Google Scholar]
82. Shepherd AJ, Swanepoel R, Shepherd SP, McGillivray GM, Searle LA. Antibody to Crimean-Congo hemorrhagic fever virus in wild mammals from southern Africa. Am J Trop Med Hyg. 1987;36:133–142. [PubMed] [Google Scholar]
83. Saijo M, Tang Q, Shimayi B. Possible horizontal transmission of crimean-congo hemorrhagic fever virus from a mother to her child. Jpn J Infect Dis. 2004;57:55–57. [PubMed] [Google Scholar]
84. Goldf arb LG, Chumakov MP, Myskin AA, Kondratenko VF, Reznikov OY. An epidemiological model of Crimean hemorrhagic fever. Am J Trop Med Hyg. 1980;29:260. [PubMed] [Google Scholar]
85. Ergonul O, Celikbas A, Baykam N, Eren S, Esener H, Dokuzoguz B. Analysis of the mortality among the patients with Crimean Congo hemorrhagic fever virus infection. Clin Microbiol Infect (in press). [PubMed]
86. Celikbas A, Ergonul O, Dokuzoguz B, Eren S, Baykam N, Polat-Duzgun A. Crimean Congo hemorrhagic fever infection simulating acute appendicitis. J Infect. 2005;50:363–365. [PubMed] [Google Scholar]
87. Bakhtiari K, Meijers JC, de Jonge E, Levi M. Prospective validation of the International Society of Thrombosis and Haemostasis scoring system for disseminated intravascular coagulation. Crit Care Med. 2004;32:2416–2421. [PubMed] [Google Scholar]
88. Geisbert TW, Jahrling PB. Exotic emerging viral diseases: progress and challenges. Nat Med. 2004;10:S110–S121. [PubMed] [Google Scholar]
89. Feldman H, Jones S, Klenk HD, Schnittler HJ. Ebola virus: from discovery to vaccine. Nat Immunol. 2003;3:677–685. [PubMed] [Google Scholar]
90. Schnittler HJ, Feldman H. Viral hemorrhagic fever—a vascular disease? Thromb Haemost. 2003;89:967–972. [PubMed] [Google Scholar]
91. Burt FJ, Swanepoel R, Shieh WJ. Immunohistochemical and in situ localization of Crimean-Congo hemorrhagic fever virus in human tissues and implications for CCHF pathogenesis. Arch Pathol Lab Med. 1997;121:839–846. [PubMed] [Google Scholar]
92. Fisman DN. Hemophagocytic syndromes and infection. Emerg Infect Dis. 2000;6:601–608. [PMC free article] [PubMed] [Google Scholar]
93. Ergonul O, Tuncbilek S, Baykam N, Celikbas A, Dokuzoguz B. Evaluation of serum levels of IL-6, IL-10, and TNF-alpha in patients with Crimean-Congo hemorrhagic fever. J Infect Dis (in press). [PubMed]
94. Shepherd AJ, Swanepoel R, Leman PA, Shepherd SP. Comparison of methods for isolation and titration of Crimean-Congo hemorrhagic fever virus. J Clin Microbiol. 1986;24:654–656. [PMC free article] [PubMed] [Google Scholar]
95. Burt FJ, Leman PA, Abbott JC, Swanepoel R. Serodiagnosis of Crimean-Congo haemorrhagic fever. Epidemiol Infect. 1994;113:551–562. [PMC free article] [PubMed] [Google Scholar]
96. Drosten C, Kummerer BM, Schmitz H, Gunther S. Molecular diagnostics of viral hemorrhagic fevers. Antiviral Res. 2003;57:61–87. [PubMed] [Google Scholar]
97. Schwarz TF, Nsanze H, Longson M. Polymerase chain reaction for diagnosis and identification of distinct variants of Crimean-Congo hemorrhagic fever virus in the United Arab Emirates. Am J Trop Med Hyg. 1996;55:190–196. [PubMed] [Google Scholar]
98. Drosten C, Gottig S, Schilling S. Rapid detection and quantification of RNA of Ebola and Marburg viruses, Lassa virus, Crimean-Congo hemorrhagic fever virus, Rift Valley fever virus, dengue virus, and yellow fever virus by real-time reverse transcription-PCR. J Clin Microbiol. 2002;40:2323–2330. [PMC free article] [PubMed] [Google Scholar]
99. Shepherd AJ, Swanepoel R, Leman PA. Antibody response in Crimean-Congo hemorrhagic fever. Rev Infect Dis. 1989;11:S801–S806. [PubMed] [Google Scholar]
100. Charrel RN, Attoui H, Butenko AM. Tick-borne virus diseases of human interest in Europe. Clin Microbiol Infect. 2004;10:1040–1055. [PubMed] [Google Scholar]
101. Saijo M, Tang Q, Shimayi B. Recombinant nucleoprotein-based serological diagnosis of Crimean-Congo hemorrhagic fever virus infections. J Med Virol. 2005;75:295–299. [PubMed] [Google Scholar]
102. Watts DM, Ussery MA, Nash D, Peters CJ. Inhibition of Crimean-Congo hemorrhagic fever viral infectivity yields in vitro by ribavirin. Am J Trop Med Hyg. 1989;41:581–585. [PubMed] [Google Scholar]
103. Tignor GH, Hanham CA. Ribavirin efficacy in an in vivo model of Crimean-Congo hemorrhagic fever virus (CCHF) infection. Antiviral Res. 1993;22:309–325. [PMC free article] [PubMed] [Google Scholar]
104. Fisher-Hoch SP, Khan JA, Rehman S, Mirza S, Khurshid M, McCormick JB. Crimean Congo-haemorrhagic fever treated with oral ribavirin. Lancet. 1995;346:472–475. [PubMed] [Google Scholar]
105. WHO Crimean-Congo haemorrhagic fever. http://www.who.int/mediacentre/factsheets/fs208/en/ (accessed Feb 17, 2006)
106. Knowles SR, Phillips EJ, Dresser L, Matukas L. Common adverse events associated with the use of ribavirin for severe acute respiratory syndrome in Canada. Clin Infect Dis. 2003;37:1139–1142. [PMC free article] [PubMed] [Google Scholar]
107. Chiou HE, Liu CL, Buttrey MJ. Adverse effects of ribavirin and outcome in severe acute respiratory syndrome: experience in two medical centers. Chest. 2005;128:263–272. [PMC free article] [PubMed] [Google Scholar]
108. Vassilenko SM, Vassilev TL, Bozadjiev LG, Bineva IL, Kazarov GZ. Specific intravenous immunoglobulin for Crimean-Congo haemorrhagic fever. Lancet. 1990;335:791–792. [PubMed] [Google Scholar]
109. Paragas J, Whitehouse CA, Bray M, Endy TP. A simple assay for determining antiviral activity against Crimean-Congo hemorrhagic fever virus. Antiviral Res. 2004;62:21–25. [PubMed] [Google Scholar]
110. Andersson I, Bladh L, Mousavi-Jazi M. Human MxA protein inhibits the replication of Crimean-Congo hemorrhagic fever virus. J Virol. 2004;78:4323–4329. [PMC free article] [PubMed] [Google Scholar]
111. Athar MN, Khalid MA, Ahmad AM. Crimean-Congo hemorrhagic fever outbreak in Rawalpindi, Pakistan, February 2002: contact tracing and risk assessment. Am J Trop Med Hyg. 2005;72:471–473. [PubMed] [Google Scholar]
112. Ergonul O, Zeller H, Celikbas A, Dokuzoguz B. The lack of Crimean-Congo haemorrhagic fever virus antibodies among health care workers in an endemic region. Int J Infect Dis (in press). [PubMed]

Articles from The Lancet. Infectious Diseases are provided here courtesy of Elsevier

-