Skip to main content

Posthumanist Performativity: Toward an Understanding of How Matter Comes to Matter

  • Chapter
  • First Online:
A Feminist Companion to the Posthumanities

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

Access this chapter

Institutional subscriptions

Notes

  1. 1.

    Dissatisfaction surfaces in the literature in the 1980s. See, e.g., Donna Haraway’s “Gender for a Marxist Dictionary: The Sexual Politics of a Word” (originally published 1987) and “Situated Knowledges: The Science Question in Feminism and the Privilege of Partial Perspective” (originally published 1988); both reprinted in Haraway (1991). See also Butler (1989).

  2. 2.

    Haraway proposes the notion of diffraction as a metaphor for rethinking the geometry and optics of relationality:

    [F]eminist theorist Trinh Minh-ha … was looking for a way to figure ‘difference’ as a ‘critical difference within,’ and not as special taxonomic marks grounding difference as apartheid. … Diffraction does not produce ‘the same’ displaced, as reflection and refraction do. Diffraction is a mapping of interference, not of replication, reflection, or reproduction. A diffraction pattern does not map where differences appear, but rather maps where the effects of differences appear. (1992, 300)

    Haraway (1997) promotes the notion of diffraction to a fourth semiotic category. Inspired by her suggestions for usefully deploying this rich and fascinating physical phenomenon to think about differences that matter, I further elaborate the notion of diffraction as a mutated critical tool of analysis (though not as a fourth semiotic category) in my forthcoming book (Barad 2007).

  3. 3.

    See Rouse (2002) on rethinking naturalism. The neologism intra-activity is defined below.

  4. 4.

    Pickering (1995) explicitly eschews the representationalist idiom in favor of a performative idiom. It is important to note, however, that Pickering’s notion of performativity would not be recognizable as such to poststructuralists, despite their shared embrace of performativity as a remedy to representationalism, and despite their shared rejection of humanism. Pickering’s appropriation of the term does not include any acknowledgement of its politically important – arguably inherently queer – genealogy (see Sedgwick 1993) or why it has been and continues to be important to contemporary critical theorists, especially feminist and queer studies scholars/activists. Indeed, he evacuates its important political historicity along with many of its crucial insights. In particular, Pickering ignores important discursive dimensions, including questions of meaning, intelligibility, significance, identity formation, and power, which are central to poststructuralist invocations of “performativity.” And he takes for granted the humanist notion of agency as a property of individual entities (such as humans, but also weather systems, scallops, and stereos), which poststructuralists problematize. On the other hand, poststructuralist approaches fail to take account of “nonhuman agency,” which is a central focus of Pickering’s account. See Barad (2007) for a more detailed discussion.

  5. 5.

    This notion of posthumanism differs from Pickering’s idiosyncratic assignment of a “posthumanist space [as] a space in which the human actors are still there but now inextricably entangled with the nonhuman, no longer at the center of the action calling the shots” (26). However, the decentering of the human is but one element of posthumanism. (Note that Pickering’s notion of “entanglement” is explicitly epistemological, not ontological. What is at issue for him in dubbing his account “posthumanist” is the fact that it is attentive to the mutual accommodation, or responsiveness, of human and nonhuman agents.)

  6. 6.

    It could be argued that “materialized refiguration” is an enterprised up (Haraway’s term) version of “materialization,” while the notion of “materialization” hints at a richer account of the former. Indeed, it is possible to read my posthumanist performative account along these lines, as a diffractive elaboration of Butler’s and Haraway’s crucial insights.

  7. 7.

    See also Butler (1989).

  8. 8.

    The conjunctive term material-discursive and other agential realist terms like intra-action are defined below.

  9. 9.

    This essay outlines issues I developed in earlier publications including Barad (1996, 1998a, 1998b, 2001b), and in my forthcoming book (Barad 2007).

  10. 10.

    Relata are would-be antecedent components of relations. According to metaphysical atomism, individual relata always preexist any relations that may hold between them.

  11. 11.

    Niels Bohr (1885–1962), a contemporary of Einstein, was one of the founders of quantum physics and also the most widely accepted interpretation of the quantum theory, which goes by the name of the Copenhagen interpretation (after the home of Bohr’s internationally acclaimed physics institute that bears his name). On my reading of Bohr’s philosophy-physics, Bohr can be understood as proposing a protoperformative account of scientific practices.

  12. 12.

    Bohr argues on the basis of this single crucial insight, together with the empirical finding of an inherent discontinuity in measurement “intra-actions,” that one must reject the presumed inherent separability of observer and observed, knower and known. See Barad (1996, 2007).

  13. 13.

    That is, relations are not secondarily derived from independently existing “relata,” but rather the mutual ontological dependence of “relata” – the relation – is the ontological primitive. As discussed below, relata only exist within phenomena as a result of specific intra-actions (i.e., there are no independent relata, only relata-within-relations).

  14. 14.

    A concrete example may be helpful. When light passes through a two-slit diffraction grating and forms a diffraction pattern it is said to exhibit wavelike behavior. But there is also evidence that light exhibits particlelike characteristics, called photons. If one wanted to test this hypothesis, the diffraction apparatus could be modified in such a way as to allow a determination of which slit a given photon passes through (since particles only go through a single slit at a time). The result of running this experiment is that there is no longer a diffraction pattern! Classically, these two results together seem contradictory – frustrating efforts to specify the true ontological nature of light. Bohr resolves this wave-particle duality paradox as follows: the objective referent is not some abstract, independently existing entity but rather the phenomenon of light intra-acting with the apparatus. The first apparatus gives determinate meaning to the notion of “wave,” while the second provides determinate meaning to the notion of “particle.” The notions of “wave” and “particle” do not refer to inherent characteristics of an object that precedes its intra-action. There are no such independently existing objects with inherent characteristics. The two different apparatuses effect different cuts, that is, draw different distinctions delineating the “measured object” from the “measuring instrument.” In other words, they differ in their local material resolutions of the inherent ontological indeterminacy. There is no conflict because the two different results mark different intra-actions. See Barad (1996), forthcoming for more details.

  15. 15.

    This elaboration is not based on an analogical extrapolation. Rather, I argue that such anthropocentric restrictions to laboratory investigations are not justified and indeed defy the logic of Bohr’s own insights. See Barad (2007).

  16. 16.

    Because phenomena constitute the ontological primitives, it makes no sense to talk about independently existing things as somehow behind or as the causes of phenomena. In essence, there are no noumena, only phenomena. Agential realist phenomena are neither Kant’s phenomena nor the phenomenologist’s phenomena.

  17. 17.

    The nature of causal intra-actions is discussed further in the next section.

  18. 18.

    See Barad (1998b, 2001a, 2001b, 2007) for examples.

  19. 19.

    I am grateful to Joe Rouse for putting this point so elegantly (private conversation). Rouse (2002) suggests that measurement need not be a term about laboratory operations, that before answering whether or not something is a measurement a prior question must be considered, namely, What constitutes a measurement of what?

  20. 20.

    Geometry is concerned with shapes and sizes (this is true even of the non-Euclidean varieties, such as geometries built on curved surfaces like spheres rather than on flat planes), whereas topology investigates questions of connectivity and boundaries. Although spatiality is often thought of geometrically, particularly in terms of the characteristics of enclosures (like size and shape), this is only one way of thinking about space. Topological features of manifolds can be extremely important. For example, two points that seem far apart geometrically may, given a particular connectivity of the spatial manifold, actually be proximate to one another (as, e.g., in the case of cosmological objects called “wormholes”).

  21. 21.

    In contrast to Butler’s “constitutive outside,” for example.

  22. 22.

    For example, the space of agency is much larger than that postulated by Butler’s or Louis Althusser’s theories. There is more to agency than the possibilities of linguistic resignification, and the circumvention of deterministic outcome does not require a clash of apparatuses/ discursive demands (i.e., overdetermination).

  23. 23.

    This is true at the atomic level as well. Indeed, as Bohr emphasizes, the mutual exclusivity of “position” and “momentum” is what makes the notion of causality in quantum physics profoundly different from the determinist sense of causality of classical Newtonian physics.

  24. 24.

    Others have made this point as well, e.g., Haraway (1991); Kirby (1997); Rouse (2002); and Bohr.

  25. 25.

    The notion of agential separability, which is predicated on the agential realist notion of intra-actions, has far-reaching consequences. Indeed, it can be shown to play a critical role in the resolution of the “measurement problem” and other long-standing problems in quantum theory. See Barad (2007).

  26. 26.

    Vicki Kirby (private communication, 2002). Kirby’s sustained interrogation of the tenacious nature/culture binary is unparalleled. See Kirby 1997 for a remarkable “materialist” (my description) reading of Derridean theory.

References

  • Barad, Karen. 1995. A feminist approach to teaching quantum physics. In Teaching the majority: breaking the gender barrier in science, mathematics, and engineering, ed. Sue V. Rosser, 43–75. Athene Series. New York: Teacher’s College Press.

    Google Scholar 

  • Barad, Karen. 1996. Meeting the universe halfway: realism and social constructivism without contradiction. In Feminism, Science, and the Philosophy of Science, eds. Lynn Hankinson Nelson and Jack Nelson, 161–94. Dordrecht, Holland: Kluwer Press.

    Chapter  Google Scholar 

  • Barad, Karen. 1998a. Agential realism: feminist interventions in understanding scientific practices. In The science studies reader, ed. Mario Biagioli, 1–11. New York: Routledge.

    Google Scholar 

  • Barad, Karen. 1998b. Getting real: technoscientific practices and the materialization of reality. Differences: A Journal of Feminist Cultural Studies 10(2):87–126.

    Google Scholar 

  • Barad, Karen. 2001a. Performing culture/performing nature: using the piezoelectric crystal of ultrasound technologies as a transducer between science studies and queer theories. In Digital anatomy, ed. Christina Lammar, 98–114. Vienna: Turia & Kant.

    Google Scholar 

  • Barad, Karen. 2001b. Re(con)figuring space, time, and matter. In Feminist locations: global and local, theory and practice, ed. Marianne DeKoven, 75–109. New Brunswick, NJ: Rutgers University Press.

    Google Scholar 

  • Barad, Karen. 2007. Meeting the Universe Halfway: Quantum Physics and the Entanglement of Matter and Meaning. Durham, NC: Duke University Press.

    Chapter  Google Scholar 

  • Butler, Judith. 1989. Foucault and the paradox of bodily inscriptions. Journal of Philosophy 86(11):601–7.

    Article  Google Scholar 

  • Butler, Judith. 1990. Gender trouble: feminism and the subversion of identity. New York: Routledge.

    Google Scholar 

  • Butler, Judith. 1993. Bodies that matter: on the discursive limits of “sex.”. New York: Routledge.

    Google Scholar 

  • Deleuze, Gilles. 1988. Foucault. Trans. Sea´n Hand. Minneapolis, MN: University of Minnesota Press.

    Google Scholar 

  • Foucault, Michel. 1970. The order of things: an archaeology of the human sciences. New York: Vintage Books.

    Google Scholar 

  • Foucault, Michel. 1972. The archaeology of knowledge and the discourse on language. Trans. A. M. Sheridan Smith. New York: Pantheon Books.

    Google Scholar 

  • Foucault, Michel. 1980a. The history of sexuality. vol. 1. an introduction. Trans. Robert Hurley. New York: Vintage Books.

    Google Scholar 

  • Foucault, Michel. 1980b. Power/knowledge: selected interviews and other writings, 1972–1977, ed. Gordon Colin, New York: Pantheon Books.

    Google Scholar 

  • Hacking, Ian. 1983. Representing and intervening: introductory topics in the philosophy of natural science. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Haraway, Donna. 1991. Simians, cyborgs, and women: the reinvention of nature. New York: Routledge.

    Google Scholar 

  • Haraway, Donna. 1992. The promises of monsters: a regenerative politics for inappropriate/d others. In Cultural studies, ed. Lawrence Grossberg, Cory Nelson, Paula Treichler, 295–337. New York: Routledge.

    Google Scholar 

  • Haraway, Donna. 1997. Modest_witness@second_millennium.femaleman_meets_oncomouse_: feminism and technoscience. New York: Routledge.

    Google Scholar 

  • Hennessey, Rosemary. 1993. Materialist feminism and the politics of discourse. New York: Routledge.

    Google Scholar 

  • Kirby, Vicki. 1997. Telling flesh: the substance of the corporeal. New York: Routledge.

    Google Scholar 

  • Pickering, Andrew. 1995. The mangle of practice: time, agency, and science. Chicago, IL: University of Chicago Press.

    Book  Google Scholar 

  • Rouse, Joseph. 1987. Knowledge and power: toward a political philosophy of science. Ithaca, NY: Cornell University Press.

    Google Scholar 

  • Rouse, Joseph. 1996. Engaging science: how to understand its practices philosophically. Ithaca, NY: Cornell University Press.

    Google Scholar 

  • Rouse, Joseph. 2002. How scientific practices matter: reclaiming philosophical naturalism. Chicago, IL: University of Chicago Press.

    Google Scholar 

  • Sedgwick, Eve Kosofsky. 1993. Queer performativity: Henry James’s the art of the novel. A Journal of Lesbian and Gay Studies 1(1):1–16.

    Google Scholar 

  • Shaviro, Steve. 1997. Doom patrols: a theoretical fiction about postmodernism. New York: Serpent’s Tail. Available on-line at http://www.dhalgren.com/Doom/.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Karen Barad .

Editor information

Editors and Affiliations

Rights and permissions

Reprints and permissions

Copyright information

© 2018 Springer International Publishing AG, part of Springer Nature

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Barad, K. (2018). Posthumanist Performativity: Toward an Understanding of How Matter Comes to Matter. In: Åsberg, C., Braidotti, R. (eds) A Feminist Companion to the Posthumanities. Springer, Cham. https://doi.org/10.1007/978-3-319-62140-1_19

Download citation

Publish with us

Policies and ethics