Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-05-01T04:45:15.131Z Has data issue: false hasContentIssue false

Trypanosomiasis Control in African History: An Evaded Issue?

Published online by Cambridge University Press:  22 January 2009

James Giblin
Affiliation:
University of Iowa

Extract

Social control of trypanosomiasis in African history deserves further study. The pioneering work in this field is John Ford's respected but neglected The Role of the Trypanosomiases in African Ecology (1971). While Ford's arguments have received support from recent findings in immunological, epidemiological and epizootiological research, they have rarely met with evaluation or engagement, either in historical or scientific literature. Historians have tended to describe trypanosomiasis control as a matter of avoiding contact with tsetse fly. In so doing they have implicitly rejected the position of Ford, who regarded infrequent contacts between tsetse and mammalian hosts as necessary for the maintenance of host resistance. Ford believed that host resistance, rather than avoidance of tsetse, was the basis of trypanosomiasis control. The historical nature of Ford's work requires that a satisfactory evaluation of The Role of the Trypanosomiases make use of historical, as well as scientific, data. The evidence of trypanosomiasis and cattle-keeping from one region of north-eastern Tanzania supports Ford and suggests that other explanations of trypanosomiasis control are inadequate. The Tanzanian evidence shows that precolonial societies coexisted with, but could not avoid, tsetse. They could not eradicate tsetse because scarcity of water prevented permanent occupation of large areas. Tsetse and trypanosomiasis did not prevent cattle-keeping, but helped to keep the region's cattle population low and confined it to relatively densely settled neighbourhoods. Social control of trypanosomiasis collapsed during the pre-Second World War period of colonial rule. Economic and political developments were primarily responsible for a series of famines between 1894 and 1934. Famine-induced depopulation allowed steady spread of tsetse and wildlife reservoirs of trypanosomes into formerly cultivated areas which had been free of tsetse before the colonial period.

Type
Tsetse fly in East Africa
Copyright
Copyright © Cambridge University Press 1990

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

1 Duggan, A. J., ‘An historical perspective’, in Mulligan, H. W. (ed.), The African Trypanosomiases (London, 1970).Google Scholar

2 Ford, John, The Role of the Trypanosomiases in African Ecology: a Study of the Tsetse Fly Problem (Oxford, 1971).Google Scholar On gambian and rhodesian trypanosomiasis, see 66.

3 See Lamphear, J. E., review of Ford, The Role, J. Afr. Hist., XIV, 2 (1973), 335–7CrossRefGoogle Scholar, and the captious review by Soff, Harvey G., Int. J. Afr. Hist. Studies, VI, 1 (1973), 146–8.CrossRefGoogle Scholar

4 Iliffe, John, review of Kjekshus, H., Ecology Control and Economic Development in East African History (London, 1977)Google Scholar, J. Afr. Hist., XIX (1978), 139.Google Scholar In his discussions of trypanosomiasis in A Modern History of Tanganyika (Cambridge, 1979)Google Scholar, however, Iliffe did not use Ford's epidemiological reasoning: see below. Johnson, Douglas H. and Anderson, David M. (eds.), The Ecology of Survival: Case Studies from Northeast African History (London and Boulder, Col., 1988)Google Scholar, is an example of a recent work which credits Ford as having helped to ‘shape its perspectives’, 23.

5 Soff, review of Ford, 148; Lamphear, review of Ford, 335–6; T.A.M.N. [T.A.M. Nash], review of Ford, African Affairs, lxxi, 283 (04 1972), 221.Google Scholar

6 Jordan states that among trypanosomiasis specialists there is general agreement with Ford's arguments: Jordan, Anthony M., Trypanosomiasis Control and African Rural Development (New York, 1986), 303.Google Scholar A recent work which adheres closely to Ford's views is Matzke, Gordon, ‘A reassessment of the expected development consequences of tsetse control efforts in Africa’, Social Science and Medicine, XVII, 9 (1983), 531.CrossRefGoogle Scholar

7 An example of an ahistorical treatment is Lee, C. W. and Maurice, J. M. (eds.), The African Trypanosomiases: Methods and Concepts of Control and Eradication in Relation to Development (Washington: World Bank Technical Paper no. 4, 1983).Google Scholar Jordan comments that for his purposes the causes of the spread of tsetse in the colonial period are ‘of no concern’. Jordan, , Trypanosomiasis Control, 33.Google Scholar

8 Murray, Max et al. , ‘Host susceptibility to African trypanosomiasis: trypano-tolerance’, Advances in Parasitology, XXI (1982), 55Google Scholar, and Roelants, Georges E. and Williams, Richard O., ‘African trypanosomiasis’, Critical Reviews in Tropical Medicine, 1 (1982), 48.Google Scholar

9 Vickerman, K. and Barry, J. D., ‘African trypanosomiasis’, in Cohen, Sydney and Warren, Kenneth S. (eds.), Immunology of Parasitic Infections, 2nd ed. (Oxford, 1982), 205Google Scholar; Diggs, Carter L., ‘Immunological research on African trypanosomiasis’, in Kallos, Paul (ed.), Immunity and Concomitant Immunity in Infectious Diseases, vol. 31 of Progress in Allergy (Basel and New York, 1982), 268300Google Scholar; Mansfield, John M., ‘Immunology and immunopathology of African trypanosomiasis’, in Mansfield, J. M. (ed.), The Immunology, vol., 1 of Parasitic Diseases (New York, 1981), 167226Google Scholar; and Roelants, Georges E. and Pinder, Margaret, ‘Immunobiology of African trypanosomiasis’, in Marchalonis, John J. (ed.), Immunobiology of Parasites and Parasitic Infections, vol. 12 of Contemporary Topics of Immunobiology (New York and London, 1984), 225–74.CrossRefGoogle Scholar

10 Terry, R. J., ‘Immunity to African trypanosomiasis’, in Cohen, Sydney and Sadun, Elvio H. (eds.), Immunology of Parasitic Infections (Oxford, 1976), 209 ff.Google Scholar

11 Vickerman, and Barry, , ‘African Trypanosomiasis’, 215Google Scholar; Roelants, and Pinder, ’Immunobiology’, 228.Google Scholar The significance of antigenic variation is not appreciated in historical writing about trypanosomiasis. One historian, overlooking the difficulties which it poses for the development of vaccines, attributes the lack of progress towards the development of trypanosomiasis vaccines entirely to uninterested pharmaceutical companies: Lyons, Maryinez, ‘The colonial disease: sleeping sickness in the social history of northern Zaire, 1903–1930’, (Ph.D. dissertation, University of California-Los Angeles, 1987), 321, fn. 100.Google Scholar

12 Genetic aspects of resistance are discussed by Mansfield, , ‘Immunology’, 190 ffGoogle Scholar, by Murray, , ’Host susceptibility’, 28 ff.Google Scholar, and by Murray, Max et al. , ‘Genetic resistance to African trypanosomiasis’, J Infectious Diseases, cxlix, 3 (05 1984), 311–19.CrossRefGoogle Scholar

13 Metacyclic development is described concisely by Kirchhoff, Louis V., ‘Agents of African trypanosomiasis (sleeping sickness)’, in Mandell, G. L. et al. (eds.), Principles and Practice of Infectious Diseases (New York, 1989), 2085.Google Scholar

14 Diggs, , ‘Immunological Research’, 291Google Scholar; Murray, ’Host susceptibility’, 37–8Google Scholar; Nantulya, V. M., ‘Immunological approaches to the control of animal trypanosomiasis’, Parasitology Today, 11, 6 (1986), 170–1.Google Scholar This theory is not accepted, however, by Vickerman and Barry, who argue that the tsetse vector inoculates a mixture of variant antigenic types into the host: ’African trypanosomiasis’, 215–17. Roelants, and Pinder, comment that ‘experimental proof of consistent reversion to basic [antigenic] types is lacking’, ’Immunobiology’, 229.Google Scholar

16 Henson, J. B. and Noel, Jan C., ‘Immunology and pathogenesis of African animal trypanosomiasis’, Advances in Veterinary Science and Comparative Medicine, XXIII (1979), 175.Google ScholarRoelants, and Pinder, , however, doubt this conclusion: ’Immunobiology’, 229.Google Scholar

16 Vickerman, and Barry, , ‘African trypanosomiasis’, 237Google Scholar; Murray, , ’Host susceptibility’, 37Google Scholar; Paling, R. W. et al. , ‘Epidemiology of animal trypanosomiasis on a cattle ranch in Kilifi, Kenya’, Acta Tropica, XLIV (1987), 6782.Google Scholar This assumption has been brought into question however. Murray and colleagues rejected their position of 1982 two years later: Murray, et al. , ’Genetic resistance’, 313.Google Scholar It is also doubted by Roelants, G. E. et al. , ‘Identification and selection of cattle naturally resistant to African trypanosomiasis’, Acta Tropica, XLIV (1987), 65.Google Scholar

17 Murray, , ‘Host susceptibility’, 56.Google Scholar

18 Roelants and Williams, ‘African trypanosomiasis’.

19 Ford, , The Role, 8690.Google Scholar Ford concluded that ‘a cattle population [may] live indefinitely in apparent good health, when subjected to continual light infection. So long as the infection persists the animals can resist further infection by antigenically related organisms’ (p. 90). Researchers who have questioned some elements of Ford's explanation for resistance, such as the role of antibodies and ‘basic’ antigens, nevertheless accept the existence of resistance to trypanosomiasis in cattle. See Roelants and Pinder, ‘Immunobiology’.

20 Roelants, G. E. et al. , ‘Trypanotolerance: an individual not a breed character’, Acta Tropica, XL (1983), 99104.Google Scholar

21 Roelants, , ‘Identification and selection’, 5566.Google Scholar Also, Roelants, G. E., ‘Natural resistance to African trypanosomiasis’, Parasite Immunology, VIII (1986), 110.CrossRefGoogle Scholar

22 Roelants, and Williams, , ‘African trypanosomiasis’, 3940.Google Scholar Murray, ‘Host susceptibility’, also writing in 1982, cited in defence of the acquired immunity thesis the work of six researchers who were not mentioned by Roelants and Williams. Roelants and Williams were not the first researchers who dismissed the possibility of acquired immunity. This was also the position of Weitz, B. G. F. in Mulligan, The African Trypanosomiases, 122.Google Scholar

23 Wellde, Bruce T. et al. , ‘Trypanosoma congolense: natural and acquired resistance in the bovine’, Experimental Parasitology, lii (1981), 219–32.CrossRefGoogle Scholar

24 Paling, ‘Epidemiology’.

25 See Morrison, W. Ivan, ‘Immune responses of cattle to African trypanosomes’, in Tizard, Ian (ed.), Immunology and Pathogenesis of Trypanosomiasis (Boca Raton, Fl., 1985), 122–3Google Scholar; Vickerman, and Barry, , ‘African trypanosomiasis’, 235 ff.Google Scholar; Stephen, Lorne E., Trypanosomiasis: a Veterinary Perspective (Oxford and New York, 1986), ch 12Google Scholar; Murray, Max and Trail, J. C. M., ‘Comparative epidemiology and control of trypanosomes’, Int. J. Parasitology, XVII, 2 (05 1987), 621–7.CrossRefGoogle Scholar The difficulties in accepting findings from this experimental work on cattle resistance, which they characterize as ‘rather crude’, are discussed by Roelants, and Pinder, , ‘Immunobiology’, 250–1 and 253.Google Scholar They note that the history of subject cattle may be unknown, that there may be inadvertent exposure to tsetse, and that experimental groups are small.

26 Roelants and colleagues spoke of ‘high Glossina challenge’ in their 1987 report: Roelants, , ‘Identification and selection’, 55 and 64.Google Scholar Roelants and Pinder, discussing experimental findings, distinguished between ‘light to moderate’ and ‘heavy’ fly challenge: ‘Immunobiology’, 252. Other reports assess the density of tsetse infestation by keeping a daily count of the tsetse caught in traps, but results are difficult to compare, partly because the distance between traps is not always noted: compare Roelants ‘Trypanotolerance’, and Paling ‘Epidemiology’. This problem is discussed in Murray, , ‘Host susceptibility’, 34.Google Scholar

27 Murray, M. et al. , ‘Trypanotolerance: a review’, World Animal Review, XXXI (1979), 5Google Scholar; Murray, , ‘Host susceptibility’, 34Google Scholar, and Murray, , ‘Genetic Resistance’, 313.Google Scholar

28 Roelants, , ‘Natural Resistance’, 8.Google Scholar

29 Paling, , ‘Epidemiology’, 7980.Google Scholar

30 Wellde, ‘Trypanosoma congolense’.

31 Vickerman, and Barry, , ‘African Trypanosomiasis’, 236Google Scholar, Roelants, and Pinder, , ‘Immunobiology’, 253Google Scholar, Jordan, , Trypanosomiasis Control, 79Google Scholar, and Molyneux, D. H. and Ashford, R. W., The Biology of Trypanosoma and Leishmania, Parasites of Man and Domestic Animals (London, 1983), 148–9.Google Scholar

32 Stephen, Trypanosomiasis.

33 See, for example, the discussion of Rhodesian sleeping sickness in Baker, J. R., ‘Epidemiology of African sleeping sickness’, in Trypanosomiasis and Leishmaniasis with Special Reference to Chaga's Disease, Ciba Foundation Symposium, N.S., xx (Amsterdam, London and New York, 1974), 3040.Google Scholar Other studies concerned with the distinction between Rhodesian and Gambian sleeping sickness which do not consider Ford's arguments are: Perich, P., ‘Les foyers de trypanosomiase humaine africaine à Trypanosoma rhodesiense (vecteur Glossina morsitans) au Burundi: aspects historiques et actuels’, Médecine Tropicale, XLII, I (01-02 1982), 3341Google Scholar, and Lambrecht, F. L., ‘Ecological and physiological factors in the cyclic transmission of African trypanosomiasis’, in Baldry, D. A. T. and Chaudhury, M. F. B. (eds.), Epidemiology of African Trypanosomiasis, vol. 1 of Insect Science and its Application (London, 1980), 4754.Google Scholar

34 Turner, D. A., ‘Tsetse and trypanosomiasis in the Lambwe Valley, Kenya’, Transactions of the Royal Society of Tropical Medicine and Hygiene, lxxx (1986), 592–5.CrossRefGoogle Scholar One study very much in the Fordian tradition, with its historical perspective and postulation of acquired resistance in humans, is Buyst, Herman, ‘The epidemiology, clinical features, treatment and history of sleeping sickness on the northern edge of the Luangwa fly belt’, Medical Journal of Zambia, VIII (1974), 212.Google Scholar

35 Connor, R. J. and Halliwell, R. W., ‘Bovine trypanosomiasis in Southern Tanzania: parasitological and serological survey of prevalence’, Tropical Animal Health and Production, xix, 3 (1987), 165–72.CrossRefGoogle Scholar Snow and colleagues also present evidence on the different rates of T. congolense and T. vivax infections which directly contradicts Ford, but they do not note this: Snow, W. F. et al. , ‘The feeding habits of the tsetse, Glossina pallidipes Austen on the South Kenya coast, in the context of its host range and trypanosome infection rates in other parts of east Africa’, Acta Tropica, xlv (1988), 330–49.Google Scholar This may be compared with Ford, The Role, 85. Another study of this question which relies upon the argument advanced, and later rejected, by Ford is, Moloo, S. K., ‘Interacting factors in the epidemiology of trypanosomiasis in an endemic/enzootic region of Uganda and its contiguous area of Kenya’, in Baldry and Chaudhury, Epidemiology, 117–21.Google Scholar

36 Vale, G. A., ‘Prospects for tsetse control’, Int. J. Parasitology, XVII 2 (02 1987), 665–70.CrossRefGoogle Scholar

37 Matzke, Gordon, ‘Settlement and sleeping sickness control—a dual threshold model of colonial and traditional methods in East Africa’, Social Science and Medicine, XIII D (1979), 209–14.Google Scholar However, in Matzke, ‘Reassessment’, the author followed the approach of Ford, The Role. Another study which relies on epidemiological theory refuted by Ford, but which does not cite The Role, is Tarimo, C. S., ‘Contribution to the epidemiology of Trypanosoma rhodesiense sleeping sickness in Lower Kitete, Northern Tanzania, by the cultural practices of the Masai’, in Baldry and Chaudhury, Epidemiology, 73–6.Google Scholar

38 Kjekshus, Helge, Ecology Control and Economic Development in East African History: the Case of Tanganyika, 1850–1950 (London, 1977), 7.Google Scholar

39 Kjekshus, , Ecology Control, 53–5.Google Scholar

40 Kjekshus glosses over criticisms of this map (Ecology Control, 162–3), but the information on north-eastern Tanzania presented below shows that his confidence in it was misplaced; it entirely misrepresented the tsetse situation in the north-east. The 1913 map was criticized, with good reason, by Bax, S. Napier, ‘Notes on the presence of tsetse fly, between 1857 and 1915, in the Dar es Salaam Area’, Tanganyika Notes and Records, XVI (1943), 3348.Google Scholar The misleading map on which Kjekshus based his argument has been reproduced, along with its confusing implications, in Lewis, L. A. and Berry, L., African Environments and Resources (Boston and London, 1988), 236–7.Google Scholar Jordan, Trypanosomiasis Control, also uses Kjekshus's data uncritically.

41 McCracken, John, ‘Colonialism, capitalism and the ecological crisis in Malawi: a reassessment’, in Anderson, David and Grove, Richard (eds.), Conservation in Africa: People, Policies and Practice (Cambridge and New York, 1987), 6377Google Scholar, considers the approach of Kjekshus, but suggests that the expansion of tsetse belts evident in colonial Malawi began as the result of changes in pre-colonial settlement patterns. Although he cites Ford, The Role, he does not consider vector-human/bovine contacts as the source of trypanosomiasis control.

42 Iliffe, , A Modern History, 12, 164–6.Google Scholar

43 Vail, Leroy, ‘Ecology and history: the example of eastern Zambia’, J. Southern Afr. Studies, iii, 1 (1976), 138–41, 144, 147, 153.Google Scholar

44 Koponen, Juhani, People and Production in Late Precolonial Tanzania: History and Structures (Helsinki, 1988), esp. 157 and 249.Google Scholar Another work on Tanzania which invokes Ford's views is Jahnke, Hans E., Tsetse Flies and Livestock Development in East Africa (Munich, 1976).Google Scholar

45 Lyons, , ‘The colonial disease’; for example, 44, fn. 17.Google Scholar

46 Ford, The Role, 66; Lyons, , ‘The colonial disease’, 61.Google Scholar

47 Ford, , The Role, 472–3.Google Scholar

48 The view that undomesticated mammals are not involved in the transmission of ‘gambian’ forms of human trypanosomiasis is also expressed in Austen, Ralph A. and Headrick, Rita, ‘Equatorial Africa under Colonial Rule’, in Birmingham, David and Martin, Phyllis M. (eds.), History of Central Africa, vol. 2 (London and New York, 1983), 64.Google Scholar Historians who adopt this perspective are surely siding with majority opinion among scientific researchers, but some recent work indicates that it may be premature to dismiss Ford's dissident view. The minority view suggests that domesticated pigs, and possibly undomesticated mammals, may be reservoirs in areas where ‘gambian’ sleeping sickness prevails: D. H. Molyneux, ‘Animal reservoirs and residual “foci” of Trypanosoma Brucei Gambiense sleeping sickness in West Africa’, and D. A. T. Baldry, ‘Local distribution and ecology of Glossina Palpalis and Glossina Tachinoides in forest foci of West African human trypanosomiasis, with special reference to associations between peri-domestic tsetse and their hosts', both in Baldry, and Chaudhury, , Epidemiology, 5963 and 8593Google Scholar respectively. Baldry considers other similarities between ‘gambian’ and ‘rhodesian’ sleeping sickness. Other recent publications have also adopted this view: Lee and Maurice, African Trypanosomiasis, 11, and Jordan, , Trypanosomiasis Control, 37 and 74.Google Scholar

48 Lyons, , ‘The colonial disease’, 369 ff.Google Scholar Other studies quite similar in this respect to Lyons' are Vail, ‘Ecology and History’, 140–1, and Musambachime, Mwelwa C., ‘The social and economic effects of sleeping sickness in Mweru-Luapula, 1906–1922’, African Economic History, no. 10 (1981), 151–73.CrossRefGoogle ScholarPubMed

50 The British estimated the population of Handeni District to be 56,000 in 1928: Tanga Provincial Book, Tanzania National Archives [hereafter TNA]. German estimates had placed the population of the same area at about 53,500 from 1898 to 1905, but the fact that they take no account of the demographic consequences of the 1898–1900 famine is indication of their unreliability: Reichs Kolonialamt, Deutsches Zentralarchiv, Potsdam [hereafter RKA], 6467/296–298; 6478/135–136; 6479.

51 Conditions in the eastern morsitans fly-belt are discussed in Lichtenheld, Dr Georg, ‘Beobachtungen über Nagana und Glossinen in Deutsch-Ostafrika’, Archiv für Wissenschaftliche und Praktische Tierheilkunde, suppl. (1910), 272–82Google Scholar; Newham, Lt.-Col., ‘Trypanosomiasis in the East African campaign’, J. Royal Army Medical Corps, XXXIII (06-12 1919), 299311Google Scholar; Swynnerton, C. F. M., ‘An examination of the tsetse problem in North Mossurise, Portuguese East Africa’, Bulletin of Entomological Research, XI (1921), 315–85CrossRefGoogle Scholar; Dye, William H., ‘The relative importance of man and beast in human trypanosomiasis’, Transactions of the Royal Society of Tropical Medicine and Hygiene, XXI, 3 (11 1927), 187–98CrossRefGoogle Scholar; Swynnerton, C. F. M., ‘The tsetse flies of East Africa’, Transactions of the Royal Entomological Society of London, lxxxiv (1936), 348–51Google Scholar; Napier Bax, ‘Notes on the presence of tsetse’ Ford, J., ‘Distributions of Glossina and epidemiological patterns in the African trypanosomiases’, Journal of Tropical Medicine and Hygiene, lxviii (1965), 211–15Google Scholar; Matzke, ‘Settlement and sleeping sickness’; Murray and Trail, ‘Comparative epidemiology’; Connor and Halliwell, ‘Bovine trypanosomiasis’. Studies which consider the importance of bush pigs as natural hosts are Snow, ‘The feeding habits’, and Milligan, P. J. M. and Baker, R. D., ‘A model of tsetse-transmitted animal trypanosomiasis’, Parasitology, lxlvi (1988), 211–39.CrossRefGoogle Scholar For evidence that bushpigs are not effective reservoirs, see Jordan, , Trypanosomiasis Control, 26–8.Google Scholar

52 Ford, , The Role, 478–80.Google Scholar

53 ‘Report of Thomas Smee’ (25 September, 1811) and ‘Report of Lt. Hardy with accompaniements’, India Office Marine Records, Misc. vol. 586; Smee, Hardy and Wingham, , ‘Observations during a voyage of research’, Transactions of the Bombay Geographical Society, VI (1844), 2361.Google Scholar

54 The pattern of settlement, vegetation and tsetse infestation in more densely settled areas is described in Picarda, R. P. Cado, ‘Autour de Mandera’, Les Missions Catholiques, XVIII (1886), 197Google Scholar; Baumann, Oscar, Usambara und seine Nachbargebiete (Berlin, 1891), 272;Google ScholarStuhlmann, Franz, ‘Bericht über eine Reise durch Usegua und Unguu’, Mitteilungen der Geographischen Gesellschaft in Hamburg, x (18871888), 148 and 157Google Scholar; Wilson, Charles T. and Felkin, R. W., Uganda and the Egyptian Soudan, 1 (London, 1882), 32–3Google Scholar; Roy, A. Le, ‘Au Kilimandjaro’, Les Missions Catholiques, xxv (1893);Google ScholarFischer, G. A., Das Masai-Land (Hamburg, 1885), 6Google Scholar; Baumgarten, Dr Johannes, Deutsch-Afrika und seine Nachbarn in Schwarzen Erdteil (Berlin, 1887), 9.Google Scholar

55 Ford argued that this is typically the nature of Grenzwildnis: The Role, 340. Recent work also adopts this view: Matzke, ‘Reassessment’, 536Google Scholar, and Jordan, , Trypanosomiasis Control, 198 and 275.Google Scholar

56 Picarda, , ‘Mandera’, 201Google Scholar; Wilson, and Felkin, , Uganda, 32–3Google Scholar; Stuhlmann, , ‘Usegua und Unguu’, 158Google Scholar; Horner, Anton, ‘De Bagamoyo À Mhonda (Oussigoua)’, Les Missions Catholiques, x (1878), 190Google Scholar; Marno, Ernest, ‘Bericht über eine Excursion von Zanzibar (Saadani) nach Koa-Kiora’, Mittheilungen der Kaiserlichen und Königlichen Geographischen Gesellschaft in Wien, XXI (1878), 373;Google Scholar P. Baur, ‘Dans l'Oudoe et l'Ouzigoua’, in Baur, P. and Leroy, P., A Travers le Zanguebar (Tours, 1893), 64;Google Scholar‘Einem Berichte des Bezirkamtmanns v. Rode über seine Betheiligung an dem Züge des Oberführers Freilherrn v. Manteuffel’, Deutsches Kolonialblatt [hereafter Dkb], IV, 15 (1 08 1893), 377 and 379.Google Scholar The District Officer from Pangani described this same Grenzwildnis in 1901: RKA, 217/7. In oral accounts this area is the hunting ground of famous chiefs: Elders of Magamba (interview at Magamba, 30 September 1983); Abedi Juma, Yusufu Selemani and others (Manga, 10 October 1983) and Ernest Mkomwa (Kwa Masaka, 29 September 1983).

57 Roy, A. Le, ‘Au Kilimandjaro’, 197200Google Scholar; Pfeil, Graf Joachim v., ‘Beobachtungen wahrend meiner letzten Reise in Ostafrika’, Petermanns Mitteilungen, XXXIV (1888), 19Google Scholar; Stuhlmann, RKA, 215/22–24. United Society for the Propagation of the Gospel, UMCA Archives, letter of John Hine (Korogwe, 10 February 1907), Box A. 1 (13), fo. 699. Bulletin de la Congregation [hereafter BC] XVI (1891–3), 727–8.Google Scholar

58 These developments are described in my ‘Famine, authority and the impact of foreign capital in Handeni District, Tanzania, 1840–1940’ (Ph.D. dissertation, University of Wisconsin-Madison, 1986), chs 3–5, and ‘Famine and social change during the transition to colonial rule in northeastern Tanzania, 1880–1896’, African Economic History, XX (1986), 84105.Google Scholar

59 The impact of rinderpest is discussed in Ford, The Role, 138 ff., 154 ff., 300 ff., 360, 479; Kjekshus, , Ecology Control, 126–32Google Scholar; Swynnerton, C. F. M., ‘An examination of the tsetse problem’, 343–4.Google Scholar

60 On depopulation during the 1890s, see BC, XVIII, 783–5; RKA, 237/1/79; Baumann, , Usambara, 276Google Scholar; Schoeller, Dr Max, Mitteilungen über meine Reise nach Äquatorial-Ost-Afrika und Uganda, 1896–1897, 1 (Berlin, 1901), 59, 61 and 75.CrossRefGoogle Scholar

61 BC, XXIII, 369; XXIV, 515; XXV (1909–10), 725 and 735; Vogt, François X., ‘Jahresbericht’, Echo aus den Missionen der Kongregation vom Heiligen Geist, X (19081909), 84.Google Scholar

62 RKA, 771, 5750/46–62, 6468/172, 6469/294–311, 6472/299–303 and 332, 6473/6 and 33; Jahresbericht über die Entwicklung der Deutschen Schutzgebiete [hereafter Jahresberichit] (18971998), 71, and (18981899), 242Google Scholar; BC, XX, 651 and 653; Kisbey, W. H., ‘A tour in the Zigua country’, Central Africa, XXIII (10 1904), 206.Google Scholar

63 Cultivators and pastoralists note high mortality among wildlife as the distinguishing characteristic of rinderpest: Mohamedi Lusingo (Mafisa, 1 July 1983); Salimu Kisailo (Kwa Maligwa, 19 May, 1983); Msulwa Mbega (Gombero, 21 September, 1983); Leiluli Namania, Tangano Orleerewi and Kari Orleerewi (Gitu, 24 June, 1983).

64 Mandera Mission Journal, February 1901; BC, XXII (1903–4), 110. Other sources note rinderpest outbreaks in this period throughout Uzigua: BC, XX, 654; Jahresbericht (1897–8), 72; Berichte über Land- und Forstwirtschaft in Deutsch Ostafrika (1902), 82.

65 On the cattle-trade and restocking: Dkb, XIII (1902), 257–9, XV (1904), 760Google Scholar; Usambara-Post, 17 July, 1909; Berichte über Land- (1902), 81–2Google Scholar; Deutsche Ostafrikanische Rundschau 11, 99 (15 December, 1909), and III, 93 (28 November, 1910); interviews with Saidi Hatibu (Mkonde, 29 May, 1983) and Leiluli Namania, Tangano Orleerewi and Kari Orleerewi (Gitu, 24 June, 1983).

66 BC, XXIV 515; XXV (1909–10), 735; XXVII (1913–14), 285; Mandera Journal, 12 August, 1908; Vogt, ‘Jahresbericht’, 84 and 153.

67 Jahresbericht (18971898), 72Google Scholar; Berichte über Land- (1904), 47Google Scholar; ‘Pangani’, Deutsche Ostafrikanische Rundschau 11, 97 (15 12, 1909)Google Scholar; Medizinal-Berichte (19081909), 86Google Scholar; Stuhlmann, Dr Franz, ‘Beiträge zur Kenntnis der Tsetse-fliege’, Arbeiten aus dem Kaiserlichen Gesundheitsamte, XXVI (1907), 303Google Scholar; Usambara-Post X, 32 (12 08, 1911).Google Scholar Examples of the opinion that tsetse were introduced into Uzigua during the German period, testimony in their own right to the spread of tsetse and epizootic, are the statements of Abdala Hamani Msede and Ali Omali Kipande (Kwa Dundwa, 12 September, 1983).

68 USPG, UMCA, letter of John Hine (10 February, 1907).

69 BC, XXVII, 305, and XXIX (1918–20), 100; Mandera Journal, 26 September, 1915. Young, Francis Brett, Marching on Tanga (New York, 1927), 237 and 261Google Scholar; Maskati Mission Journal, 8 August, 1915 and 3 January, 1917; Lugoba Mission Journal, 6 January, 1919; Vogt, François X., ‘Apostolisches Vikariat Bagamoyo’, Echo aus Afrika, XXXIII (1921), 75Google Scholar, and ‘Une conversation avec Monseigneur Vogt’, Annates Apostoliques de la Congrégation du Saìnt-Esprit, XXXVIII (1922), 53.Google Scholar Mrisho Kizenga (Kwediboma, 22 September, 1982) remembered the prevalence of cattle disease during the war, though he believed that it was brought with the cattle which British troops drove from Arusha.

70 The accounts of ‘Kwa Mkono’, Central Africa, xlii (February 1924) 23–4, and ‘Kwa Maizi, and its daughter Parishes in Zigualand’, Central Africa, xlii (11 1924), 240–3Google Scholar, may be compared with the following: letter of Hine (10 February, 1907); letter of Mr Herbert Lister (Korogwe, 1 March, 1892), Central Africa, X (1892), 68–9Google Scholar; Kisbey, W. H., ‘A journey in Zigualand’, Central Africa, XVIII (1900), 172–5;Google ScholarKisbey, W. H., ‘An advance in Zigualand’, Central Africa, XIX (04 1901), 64–6Google Scholar; Sehoza, Samwil, ‘A walk to Kwa Magome’, Central Africa, XXIX (06 1911), 160–3Google Scholar; and Baumann, , Usambara 273.Google Scholar Other sources are Young, , Marching on Tanga, 190Google Scholar; Mhonda Mission Journal, 18 August, 1920; BC, xxvm (19231924), 327Google Scholar, and Maskati Journal, July 1920.

71 Pangani District, Annual Report (1923), TNA.

72 Quote from SirDundas, Charles, African Crossroads (London, 1955), 101.Google Scholar Also, Pangani District, Annual Reports (1922, 1923, 1924), TNA; Handeni Sub-District, Annual Report (1925), TNA; Young, , Marching on Tanga, 237Google Scholar; USPG, UMCA Archives, letter of Petro Limo to Travers (Kwa Maizi, 12 April 1922), Box A. 5. Lion attacks on humans and livestock are noted in Lugoba Journal, 7–8 May and 24–25 July 1919; 16 September and 17 October, 1924. See also Te, Ilif, A Modern History, 270.Google Scholar

73 Mandera Journal, 1, 2 and 5 May, 30 August, 29 September and 31 December, 1922; 12 June and 27 November, 1925; BC, XXVIII (1923–4), 286Google Scholar; Mhonda Journal, 18 May, 1920; Lugoba Journal, 4, 24 and 27 October, 16–17 November, 1920; 26 August and 7 October, 1921; 24 April and 26 July, 1922; 1 and 5 August, 1923; 21 September, 5 and 17 October 1924.

74 ‘Life of Zumbe Majugwe Mkumbe of Mazingara’ (1949), TNA 4/6/5, v. 2; Young, , Marching on Tanga, 261Google Scholar; Maddocks, J. L., ‘Kwa Magome’, Central Africa, xlvi (04 1928), 6970Google Scholar; Zanzibar, Frank [Weston], ‘In Zigualand’, Central Africa, XXXIX (10 1921), 215;Google Scholar Limo to Travers (12 April, 1922), and Swynnerton to Chief Secretary (Dar es Salaam), 14 November, 1926, TNA Early Secretariat File 2702, vol. 4.

75 J. E. G. Ransome, tour reports of January and October 1929, Tanga Provincial Book. Inter-war famines in Uzigua and elsewhere in Tanzania are described by Iliffe, , A Modern History, 315.Google Scholar

76 On the decline of population during the famine: USPG, UMCA Archives, correspondence with Bishop of Zanzibar, Bishop of Zanzibar (16 December, 1933) and letter from Kideleko (n.d. but probably October 1934), S.F. 15, vol. 8; D.O. (Pangani) to P.C. (Tanga), 3 October 1934, TNA 4/401; Mandera Journal, 11 December 1933; Lee, Grace Bridges, ‘Famine in Zigualand’, Central Africa, II (12 1933), 247Google Scholar; letters of J. L. Maddocks (11 December, 1933) and P.C. (Tanga) to Chief Secretary (Dar es Salaam), 19 December, 1933 and 12 January, 1934 in TNA 13709, v. 1; D.O. (Handeni) to P.C. (Tanga), 12 March, 1935, TNA 13079, v. 2; Abedi Juma and others (Manga, 10 October, 1983); Ramadhani Semdili (Gombero, 20 September 1983); Mohamedi Chambo (Kiberashi, 17 May, 1983); Mohamedi Mjewe (Kwediboma, 27 September, 1982); Mohamedi Saidi Nyange (Kwediboma, 5 September, 1982); Pogwa Rashidi (Handeni, 14 August, 1982); Saidi Hatibu (Mkonde, 29 May, 1983); and M. A. Mavullah and others (Balanga, 23 May, 1983). Unfortunately there are no reliable census figures for Handeni District in the 1930s. A census taken in one neighbourhood shows a population decline of almost one-half before and after the 1932–5 famine: TNA 5/1161/1, v. 7. For impressions of the demographic situation in Handeni, see C. H. Grierson, ‘Handeni District’, notes of meeting of 25 June, 1934, C. F. M. Swynnerton, ‘Preliminary memorandum on the position in Handeni (23 July, 1934)’ and ‘Marginal notes’ in TNA 11299, v. 1, and Milne, G., ‘A soil reconnaissance journey through parts of Tanganyika Territory, December 1925 to February 1936’, Journal of Ecology, xxxv (1947), 199.Google Scholar

77 Swynnerton, ‘Preliminary memorandum’, NA 11299, v. 1.Google Scholar

78 In the published report of his findings in Handeni, Swynnerton cast the situation in much more optimistic terms than he had used in his original internal memorandum. Compare Swynnerton, , ‘Preliminary memorandum’ and ‘The tsetse flies’, 348.Google Scholar See also Giblin, , ‘Famine, authority’, 512–13.Google Scholar For another example of this tendency, compare J. E. G. Ransomew, tour report of 26 January, 1929, Tanga Provincial Book, and Tanga Province, Annual Reports (1928 and 1929), TNA. Government maps of tsetse distribution showed fly-belts only in those places where surveys had produced proof positive of infestation. In areas where no surveys had been done (and this included most of Uzigua) tsetse belts were not marked, even if their existence was well known to administrators. See map appended to Tanganyika Territory, Department of Veterinary Science and Animal Husbandry, Annual Report (1925), TNA.

79 Tanganyika Territory, Department of Veterinary Science, Annual Reports (1935, 1938); Korogwe District, Annual Report (1938), TNA. Mohamedi Lusingo (Mafisa, 1 July 1983) and Rajabu Sudi (Manga, 10 October, 1983) were among numerous people in various parts of Handeni District who remember the spread of tsetse after 1934.

80 Notes on baraza at Kiberashi (19 January, 1939), TNA 26650, v. 1; correspondence, esp. H. E. Hornby to Chief Secretary (Dar es Salaam), 11 October, 1940, in TNA 26650, v. 2.

81 P. C. (Tanga) to Chief Secretary (Dar es Salaam), 12 April, 1940 and Report of H. E. Hornby to Chief Secretary (Dar es Salaam), 24 September, 1940, in TNA 26650, v. 2.

82 Jordan, , Trypanosomiasis Control, 303.Google Scholar