Skip to main content

Struggling with the Standard Model (1930–1941)

  • Chapter
  • First Online:
Explaining Photosynthesis

Part of the book series: History, Philosophy and Theory of the Life Sciences ((HPTL,volume 8))

  • 935 Accesses

Abstract

Otto Warburg’s manometric approach dominated photosynthesis research in the 1930s. By then, a number of scientists of very different disciplinary backgrounds had turned to the field, including William Arnold, Robert Emerson, James Franck, Charles Stacy French, Hans Gaffron, Robin Hill and Cornelis B. van Niel. The chapter follows their “investigative pathways” in order to clarify: why they chose their particular focus and method of photosynthesis research; how they (frequently research-opportunistically) stumbled into the field; and how they interacted with other researchers and groups to coordinate their interests and activities. In contrast to earlier decades, the 1930s became a period in which a wealth of exciting new findings and conceptual ideas were amassed. Some of these stood in clear contradiction to the standard model of the mechanism: such as, for instance, the suggestion of “photosynthetic units”, consisting of 2000 chlorophyll molecules that acted in coordination; and the finding that isolated chloroplasts were able to produce oxygen without reducing carbon dioxide at the same time. Today, these are elements of the body of commonly accepted knowledge. However, at the time it proved extremely difficult to integrate them into a coherent whole. A range of different strategies were employed in order to come to terms with this situation, ranging from stark conservatism to radically new forms of conceptualising the process. Still, no consensus emerged.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 89.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 119.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 119.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    In addition to these, the universities at Berkeley and at Urbana–Champaign in the USA would soon—in the 1940s—become equally important centres of photosynthesis research.

  2. 2.

    On Kautsky’s life and work, see, e.g., von Gerhard (2004).

  3. 3.

    See Kautsky and Hirsch (1931) and Kautsky et al. (1932) for the first reports of these phenomena, while Govindjee (1995) provides a historical review of the “Kautsky effect”. Govindjee (2004a) covers the phenomenon of chlorophyll a fluorescence from both a historical and a systematic viewpoint.

  4. 4.

    See Kautsky and Hirsch (1931) and Kautsky et al. (1932).

  5. 5.

    See Kautsky et al. (1932, 1933, 1935). This proposal was contested by Hans Gaffron, who argued that photosynthesis started without oxygen; see Gaffron (1935). It was presumably through his follow-up of this debate that James Franck first became acquainted with Gaffron’s work.

  6. 6.

    On Franck’s life and work see, e.g., the biographical memoir by Kuhn (1965) and the tribute by Rosenberg (2004). Beyerchen (1996) analyses Franck’s emigration from Germany and its consequences, notably his scientific migration to photosynthesis research. See also the extensive biography by Lemmerich (2007).

  7. 7.

    Quote taken from the Nobel Prize Announcement at http://www.nobelprize.org/nobel_prizes/physics/laureates/1925/. See Franck and Hertz (1914) for the pertinent publication. Hon and Goldstein (2013) provides a lucid account of the discovery.

  8. 8.

    This work included Franck’s well-known paper on the “elementary processes of photochemical reactions”, an analysis of the shape of molecular absorption and fluorescence spectra, which includes what later became known as the Franck–Condon principle; see Franck (1925).

  9. 9.

    An English translation of the pertinent documents (as well as perceptive commentaries and useful background information) can be found in Hentschel (1996, pp. 21–34).

  10. 10.

    Beyerchen (1996, pp. 77–79).

  11. 11.

    See Franck and Levi (1935a, b) for the resulting publications.

  12. 12.

    See Franck and Rabinowitch (1934), in which they formulated the hypothesis of the “cage effect”, based on their investigation of the photolysis of different compounds.

  13. 13.

    Beyerchen (1996, p. 80). Beyerchen refers to an interview that he conducted with Hilda Levi on 12 November 1980 in Copenhagen.

  14. 14.

    See Franck (1935a).

  15. 15.

    See Franck (1935b).

  16. 16.

    See Franck and Herzfeld (1937); Franck et al. (1941); Weller and Franck (1941); Franck (1945, 1949). Franck’s final attempt to solve the problem was completed shortly before his death: see Franck and Rosenberg (1964).

  17. 17.

    Rosenberg (2004, p. 73).

  18. 18.

    Quoted in Beyerchen (1996, p. 82).

  19. 19.

    Quoted in Beyerchen (1996, pp. 82–83).

  20. 20.

    Franck to Meitner, quoted in Lemmerich (2007, p. 238); original German.

  21. 21.

    Cf. Franck (1935b, p. 433).

  22. 22.

    Cf. Willstätter and Stoll (1918).

  23. 23.

    Stoll (1932, p. 957).

  24. 24.

    This was in line with the general belief at the time that the Blackman reaction (according to its kinetics) consisted of a reaction between catalase and a peroxide. Warburg and Uyesugi (1924) was particularly influential in this respect. It was only in Emerson and Green (1937) that the supposed similarity between the Blackman reaction and the reaction between catalase and hydrogen peroxide was contested.

  25. 25.

    Willstätter (1933).

  26. 26.

    Franck (1935a, b).

  27. 27.

    Franck (1935b, p. 433).

  28. 28.

    Cf. Franck (1935b, p. 436).

  29. 29.

    Franck (1935b, p. 437).

  30. 30.

    The experiments and the many difficulties in realising the set-up have been described many times; see Myers (1994); Arnold (1991) and Govindjee (2001). Govindjee et al. (1996) is a special issue of the journal Photosynthesis Research dedicated to William Arnold; Govindjee (2014) provides a biography.

  31. 31.

    The biographical information on Emerson has been taken from the memoir by Rabinowitch (1961), complemented by the details given in Govindjee (2004b), and by Emerson’s own CV of 1936, which is held in his estate: Curriculum vitae and bibliography of Robert Emerson, Robert Emerson Papers, 1923–1961, Record Series 15/4/28, Box 1, University of Illinois Archives.

  32. 32.

    Cf. Govindjee (2004b, p. 184). On Osterhout, see Blinks (1974).

  33. 33.

    On the development of general physiology and biochemistry in the USA around 1900, see Hall (1975); Kohler (1982) and Pauly (1987b). On Jacques Loeb, see Osterhout (1928); Pauly (1987a) and Fangerau (2010).

  34. 34.

    Osterhout was the first to notice the induction period of photosynthesis and to attempt to study systematically the antagonism that exists between respiration and photosynthesis. See Osterhout (1918, 1919); Osterhout and Haas (1918, 1919).

  35. 35.

    Blinks (1974, p. 224).

  36. 36.

    Govindjee (2004b, p. 184). See also Willstätter’s autobiography for background information on his resignation. Wiesen (2000) discusses the ambiguous reception of Willstätter’s memoirs after 1945.

  37. 37.

    The title of the thesis was (translated into English) “On the effect of hydrocyanic acid, hydrogen sulphide and carbon monoxide on the respiration of different algae”. The thesis was officially handed in by the university’s botanist Hans Kniep, which at first glance implies that Kniep was Emerson’s supervisor. However, this (nominal) arrangement was due to the fact that only universities were authorised to award doctoral titles, while Kaiser Wilhelm Institutes and their members were not.

  38. 38.

    See French (1959, p. 437).

  39. 39.

    Borsook’s recollections of this period are preserved in the interview carried out with him in 1978 by Mary Terrall, as part of the Caltech Archives Oral History Project; see Borsook (1978). On Thimann see, e.g., Stowe (1999).

  40. 40.

    Kohler (1982, p. 318).

  41. 41.

    Quoted in Kohler (1982, p. 322). Emerson to William J. Crozier, 24 March 1931. The original is held by the Harvard University Archives, Pusey Library: Crozier Papers. On Crozier and how he became a central figure in general physiology, see Pauly (1987b, in particular pp. 201–204).

  42. 42.

    Arnold (1991, p. 74).

  43. 43.

    Emerson and Arnold (1932b, p. 417).

  44. 44.

    See Emerson and Arnold (1932b, p. 418).

  45. 45.

    Emerson and Arnold (1932b, p. 417).

  46. 46.

    Emerson and Arnold (1932a, p. 191).

  47. 47.

    Emerson and Arnold (1932a, p. 191).

  48. 48.

    Emerson and Arnold (1932a, pp. 193–194).

  49. 49.

    The findings were confirmed two years later, by Arnold and Kohn (1934), who discovered, as stated in the abstract, that “in six species of plants, representing four phyla, the minimum number of chlorophyll molecules present for each molecule of carbon dioxide reduced appears to lie between 2000 and 3000”.

  50. 50.

    On van Niel see, e.g., Spath (1999); Barker and Hungate (1990) and Hungate (1986).

  51. 51.

    See, on Kluyver’s life and work, e.g., Woods (1957) and Kamp et al. (1959).

  52. 52.

    Kluyver and Donker (1926). The original German title reads “Die Einheit in der Biochemie”.

  53. 53.

    On this paper’s background and further implications, see Friedmann (2004).

  54. 54.

    See Werner (1997) for an analysis of the controversy between Warburg and Wieland.

  55. 55.

    See Spath (1999, Chapter 1, pp. 36–37).

  56. 56.

    See van Niel (1941, pp. 264–269), for a review of the field up to van Niel’s own work. Chemosynthetic bacteria are able to reduce carbon dioxide (or methane) in order to produce organic matter; while they use the oxidation of inorganic molecules (e.g. hydrogen gas, hydrogen sulfide) or methane as a source of energy, rather than sunlight.

  57. 57.

    See van Niel (1967, p. 11).

  58. 58.

    How he happened to develop this broad vision is explored in Chapter 1 of Spath (1999).

  59. 59.

    Opinions differ as to how to interpret this general shift to experimental biology. Kay (1993) suggests that there prevailed an agenda of “social control”, exerted through a concerted campaign by the Rockefeller Foundation, a conservative American elite and some influential scientists. Spath (1999) points to the fact that, although the effects of this move were convergent, the interests and aims of individual scientists at the numerous institutions were very different.

  60. 60.

    See van Niel (1967, p. 10).

  61. 61.

    See van Niel (1967, p. 19).

  62. 62.

    For a timeline of research in anoxygenic photosynthesis, see Gest and Blankenship (2004).

  63. 63.

    See Wurmser (1921, 1926) and, especially, Wurmser (1930). Ideas on photosynthesis being a redox process were also expressed in Thunberg (1923), although he was still looking for an acceptable pathway to formaldehyde.

  64. 64.

    van Niel (1935, pp. 138–139).

  65. 65.

    van Niel (1935, pp. 142–143).

  66. 66.

    van Niel (1935, p. 143).

  67. 67.

    van Niel (1967, p. 10).

  68. 68.

    van Niel (1930, p. 168). Quoted in Spath (1999, p. 117).

  69. 69.

    The biographical information on Gaffron was taken from Rürup (2008, pp. 199–201). For a tribute to Gaffron and his co-workers, with special emphasis on Gaffron’s work on the hydrogen metabolism in green algae, see Homann (2002).

  70. 70.

    Warburg was allegedly instrumental in securing Gaffron this position in the KWI for Biology through his good connections with Friedrich Glum, Administrative Director of the Kaiser Wilhelm Society at the time; cf. Werner (1988, p. 246).

  71. 71.

    Werner (1988, p. 246); Rürup (2008, p. 201).

  72. 72.

    Gaffron (1933b, p. 2). This general aim during these years was shared by Charles Stacy French, another giant in twentieth century photosynthesis research. In French (1937, p. 71), the latter wrote: “It is with the hope of finding a new approach to green plant photosynthesis that several workers are now studying the different kinds of photoassimilation in these bacteria. Probably by defining the differences between green plants and purple bacteria CO\(_2\) assimilation, the chemical mechanism of both will become clearer”.

  73. 73.

    van Niel (1931); Gaffron (1934); van Niel (1935).

  74. 74.

    See Homann (2002, p. 94). Gaffron (1963) provides an account of his dispute with van Niel.

  75. 75.

    See Roelofsen (1934).

  76. 76.

    Gaffron and Wohl (1936).

  77. 77.

    Gaffron (1933a).

  78. 78.

    See Gaffron and Wohl (1936, p. 81).

  79. 79.

    Timofeeff-Ressovsky et al. (1935).

  80. 80.

    It is not entirely clear how this name relates to the other famous “Three Men Paper”, likewise known in German as Dreimännerarbeit: the paper by Max Born, Werner Heisenberg and Pascual Jordan of 1926 in which they introduced the matrix mechanics formulation of quantum mechanics; see Born et al. (1926).

  81. 81.

    For more detailed information on the Delbrück seminar and the development of biophysics in Berlin at the time, see Sloan (2009).

  82. 82.

    See the interview with Delbrück, carried out as part of the Caltech Archives Oral History Project, for the latter’s recollections of these years; Delbrück (1978, in particular p. 41).

  83. 83.

    Delbrück (1978, p. 55).

  84. 84.

    French (1979, p. 7). On the life and work of French, see Govindjee and Fork (2006).

  85. 85.

    Gaffron and Wohl (1936); Wohl (1940). Being of Jewish origin, also Wohl had to leave Germany in 1933 and emigrated to Oxford, England, where he was able to find a position at Balliol college; see Jost (1963).

  86. 86.

    Gaffron and Wohl (1936, p. 82).

  87. 87.

    Gaffron and Wohl (1936, p. 86).

  88. 88.

    Gaffron and Wohl (1936, p. 87).

  89. 89.

    Gaffron and Wohl (1936, p. 88).

  90. 90.

    Franck and Herzfeld (1937, p. 238).

  91. 91.

    Franck and Herzfeld (1937, p. 237).

  92. 92.

    Franck and Herzfeld (1937, p. 240).

  93. 93.

    Franck and Herzfeld (1937, p. 239).

  94. 94.

    Franck and Herzfeld (1937, p. 240).

  95. 95.

    Emerson (1936). In his obituary of Emerson in 1961, Rabinowitch considered this to have been the most generally accepted interpretation. See Rabinowitch (1961, pp. 118–119). Emerson also used the terms “catalyst” or “photoenzyme”. The concept aligns neatly with today’s conception of “reaction centres”.

  96. 96.

    Wohl (1940, 1941).

  97. 97.

    Manning (1938, p. 156).

  98. 98.

    On Teller see, e.g., Rhodes (1995); Teller (2001).

  99. 99.

    Franck and Teller (1938, p. 861).

  100. 100.

    Again, it is striking how closely this resembles the idea, which was elaborated in the Three Man Paper, of the gene being similar to a crystal; cf. Timofeeff-Ressovsky et al. (1935).

  101. 101.

    On the paper’s argument, see also the review by Franck and Gaffron (1941, p. 210). However, Franck and Teller assumed that there existed a one-dimensional structure (a linear chain of chlorophyll molecules); the application of two- or three-dimensional models, in, e.g., Bay and Pearlstein (1963), led to very different results. See Pearlstein (2002) for a short review.

  102. 102.

    See Franck and Herzfeld (1941).

  103. 103.

    Franck and Herzfeld (1941, p. 979).

  104. 104.

    Franck and Herzfeld (1941, p. 982).

  105. 105.

    Franck and Herzfeld (1941, p. 985).

  106. 106.

    A special 1992 issue of Photosynthesis Research was dedicated to the memory of Hill; see Rich (1992). See Bendall (1994) for a biographical account and Walker (2002) for a tribute to Hill’s work on chloroplasts. Hill (1965) provides an autobiographical perspective.

  107. 107.

    On Hopkins and his institute, see, e.g., Needham et al. (1949) as well as Kohler (1982); Chapter 4.

  108. 108.

    On the history of biochemistry in the early twentieth century, see Holmes (1986).

  109. 109.

    See Hopkins (1949). He argued along similar lines in Hopkins (1926).

  110. 110.

    Hill was admitted as a scholar to Emmanuel College in 1917; however, he only started seriously reading the Natural Sciences Tripos after the end of the First World War. Bendall (1994, pp. 145–146).

  111. 111.

    See Kohler (1982, p. 81).

  112. 112.

    Kohler (1982, p. 83).

  113. 113.

    Hill never lost this interest in plant pigments, and he became a well-known expert in the chemistry of natural dyes. Hill invariably grew the material for these and other studies in his own garden. He was also very skilled in extracting pigments and used them, among other things, for his own watercolour paintings. See Bendall (1994, p. 143).

  114. 114.

    Keilin was made director of the institute in 1931. On Keilin’s life and work, see Mann (1964). Keilin started his career with a strong interest in beetles and became a proficient entomologist. Even during the years of his research into cytochromes, Keilin never gave up his pursuit of questions on the morphology and physiology of insects.

  115. 115.

    See the papers by Keilin et al. (1931) and Keilin and Hill (1933).

  116. 116.

    Keilin (1925b, p. 315).

  117. 117.

    Keilin (1925a). See Keilin (1966) on the history of research into cytochromes.

  118. 118.

    Hill (1965, p. 124).

  119. 119.

    Bendall (1994, p. 153).

  120. 120.

    See Hill (1937, 1939) for the first publications and Bendall (1994, pp. 153–154), for an illuminating retrospective description. Hill himself never used the term “myoglobin” but always spoke of “muscle haemoglobin”, which is frequently (and misleadingly) abbreviated to “haemoglobin”.

  121. 121.

    Hill (1939, p. 207).

  122. 122.

    Hill (1939, p. 209).

  123. 123.

    Cf. Hill and Scarisbrick (1940a, b).

  124. 124.

    Hill and Scarisbrick (1940a, p. 61).

  125. 125.

    Hill and Scarisbrick (1940b, p. 254).

  126. 126.

    French (1979, p. 10). The review Franck and Gaffron (1941, p. 219), states, however, that Hill’s findings only came to their notice upon publication of Hill and Scarisbrick (1940a).

  127. 127.

    French (1979, p. 10).

  128. 128.

    See French and Anson (1941) for the abstract of the paper. In the accounts of this episode in French (1979) as well as in Myers (1974), the name of the society was inaccurately reported.

  129. 129.

    French and Anson (1941). Incidentally, in these experiments French and Anson were the first scientists to use spinach as a source of chloroplast; it remains a popular source to this day.

  130. 130.

    Myers (1974, p. 422).

  131. 131.

    Gaffron (1969, p. 11).

  132. 132.

    Cf. Myers (1974, p. 422).

  133. 133.

    Of the wealth of literature on this topic, see, in particular, Kohler (1991) and Kay (1993). Having suggested that photosynthesis, which was still considered a marginal subject, greatly profited from the advancement of “new biology”; one could even turn it the other way round and claim that photosynthesis research paved the way for the development of the “new” or molecular biology. See, e.g., the argument brought forward in Zallen (1993b).

  134. 134.

    See Haber and Willstätter (1931), which argues that biological oxidation should be seen as a dehydration process. The elimination of hydrogen, Haber and Willstätter suggested, usually resulted in the formation of radicals (since only one of the two corresponding electrons would be removed at the same time). Cf. Willstätter (1973, p. 378); Werner and Irmscher (1995, pp. 30–31).

  135. 135.

    Werner and Irmscher (1995, pp. 122–123); letter from Haber to Willstätter, 24 February 1933.

  136. 136.

    See van Niel (1967, p. 9).

  137. 137.

    This discrepancy was later taken by Franck and Gaffron (1941) as an argument for the assumption that “the anaerobic type of photosynthesis is the same in all cells but that it is supplemented in green plants by the capacity of liberating gaseous oxygen. […] Photosynthesis in plants, therefore, is the exception to the general rule” (p. 252). By contrast, van Niel (1941) assumed that in all types of photosynthesis water is reduced, while in bacteria the liberated oxygen immediately underwent secondary reactions and the dehydrogenation of the specific hydrogen donor took place in later stages of the process.

  138. 138.

    Today it is known that this “neat balance” is no longer tenable and the quantities as well as proportions can vary substantially.

  139. 139.

    It was Sam Ruben and Martin Kamen’s experiments in 1941 with “heavy” water, incorporating the oxygen isotope \(^{18}\)O, that provided the strongest evidence for the hypothesis that photosynthetic oxygen came exclusively from water. See Ruben et al. (1941) for the publication, which will also be discussed in Chapter 6.

  140. 140.

    Gaffron (1940).

  141. 141.

    This shift in the conception of photosynthesis neatly coincides with a general change in biochemistry. Up to the 1930s, a conception of metabolism as a set of linear processes prevailed. However, when ATP and various coenzymes were discovered in the 1930s, scientists slowly realised the close entanglement of the cell’s metabolism, which came to be conceived of as a highly integrated system; cf. Bechtel (1986a).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kärin Nickelsen .

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer Science+Business Media Dordrecht

About this chapter

Cite this chapter

Nickelsen, K. (2015). Struggling with the Standard Model (1930–1941). In: Explaining Photosynthesis. History, Philosophy and Theory of the Life Sciences, vol 8. Springer, Dordrecht. https://doi.org/10.1007/978-94-017-9582-1_4

Download citation

Publish with us

Policies and ethics