The Space Between: Literature and Culture 1914-1945

Cosmic Allegories: Post-Surrealism and Astronomy in Interwar Los Angeles

Marianne Kinkel
Washington State University


Abstract 
In 1934, The Los Angeles-based painters Helen Lundeberg and Lorser Feitelson founded Post-Surrealism, the first organized American response to European Surrealism. The Post-Surrealists distinguished their movement through their interest in science, particularly astronomy and biology.  This paper argues that as allegorical tableaux, many Post-Surrealist works evoke unexpected analogies and generate wonder, the desire to understand and in turn, heighten spectators’ awareness of their processes of interpretation. This essay examines Post-Surrealism in relation to modern cosmology and new ideas about space and time emanating from the observational astronomy conducted at the Mount Wilson Observatory in Pasadena. It focuses primarily on the work of Lundeberg, and argues that the artist’s inclusion of self-portraits troubles the notion that astronomy was largely a male enterprise. The paper contextualizes the Post-Surrealists’ engagement with astronomy through considering how the public was introduced to new theories of the universe and learned about astronomy through visual lessons presented in the Los Angeles Times and night sky performances at the newly opened Griffith Planetarium. 

Keywords: Post-Surrealism / Helen Lundeberg / Lorser Feitelson / astronomy / Los Angeles 



Anton Chekhov’s “Terror: My Friend’s Story” (1892) presents dialogue among four characters who contemplate the meaning of life in rural Russia. Beloved Marya Sergeyevna holds no pleasure or even satisfaction with daily life or the world around her: she is presented as devoid of curiosity and filled with boredom. Her husband, Dmitri Petrovitch Silin, described by the author as a worried and exhausted man, tethers his orientation to the everyday world through his love for Marya. The third character, Gavril Syeverov or Forty Martyrs, drunken and fallen from aristocratic beginnings, questions the causes of his fate, and the last, the story’s narrator, confesses that he constantly doubts his perceptions of the world and fears the blurring of fantasy and reality in his daily thoughts. He admits that he approaches life in terror, fearful about reality and the afterlife. Each character offers apprehensive attitudes toward being in the world: horror, fear of uncertainty, boredom, and resignation. 

Chekhov’s story fascinated the painter, Helen Lundeberg, before she undertook to become an artist. In a notebook, she transcribed several sections of the story that most intrigued her: “Tell me why is it that when we want to tell some terrible, mysterious, and fantastic story we draw our material not from life, but invariably from the world of ghosts and of the shadows beyond the grave?” “We are all frightened of what we don’t understand.” “And do you understand life?” “Our life and the life beyond the grave are equally incomprehensible and horrible.” Lundeberg wrote the following commentary on these passages: “Ideas of ‘the life beyond the grave,’ all these ideas of Heaven and Hell, and shadowy spirit worlds, have for a long time seemed to me childish, unsatisfactory, as compared with the incomprehensible wonders of the world in which I live” (Lundeberg, [On Chekhov]). Lundeberg initialed and noted the date, 1929, suggesting that her encounter with this story was significant to formulating her artistic sensibility. Shortly thereafter, in 1930, Lundeberg published her own prose work entitled Magic that counters feelings of resignation and despair when viewing life as struggling, suffering and ending in death. “Things happen so queerly,” Lundeberg claims, in mundane moments such as preparing toast for breakfast. Magic, she proposes, occurs when the banal blue and gray linoleum on the kitchen floor suddenly appears as patterns and colors that seem “to move, to change, they come alive and the world is new again” (Lundeberg, Magic). In a brief note entitled “Night Thoughts” (1981), written much later in her notebooks, Lundeberg declares: “I am not interested in photo realism or ‘magic realism’ either. But I wish to suggest the mysterious life of quite common place things, the magical quality of light and shadows falling on a wall, a floor, a table, a bowl of fruit seen at certain moments when we realize that an apple is as strange and beautiful as the solar system” (Lundeberg, Night). Instead of fear, boredom, or speculations of an afterlife, Lundeberg chose curiosity to navigate the known and the unknown in her everyday world. 

This paper argues that wonder prevails in Post-Surrealism, particularly in the work of Helen Lundeberg. She and Lorser Feitelson, her former teacher and later husband, developed Post-Surrealism by engaging allegory as a strategy for generating wonder, especially with respect to the science of astronomy. Curiously, wonder has not been taken up as an issue in prior analyses of Lundeberg’s works, including my own. In my previous scholarship, I discussed Lundeberg’s Post-Surrealist works in terms of the poetic possibilities of straddling classical and romantic tendencies within modern art. I argued for understanding Post-Surrealism as offering an allegorical approach to art, through which the viewer is invited to participate in the construction of meaning. I considered the Post-Surrealist aesthetic as conceptually related to cinematic techniques of montage. Perhaps most importantly, I drew attention to how Lundeberg’s inclusion of self-portraits within her work presented an alternative view of women in the history of early modernism, representing female figures as actively constructing worlds in which to examine philosophical questions of existence and the nature of reality.1 

In the following, I investigate how the Post-Surrealist works by Helen Lundeberg and Lorser Feitelson engaged interwar cosmology and new ideas about space and time emanating from observational astronomy conducted at the Mount Wilson Observatory in Pasadena (Fig. 2).



Many Post-Surrealist works evoke unexpected analogies among astronomy, biology, and the everyday world. As allegorical tableaux, they present an array of naturalistic depictions, seemingly photographic representations that converge with the logic of scientific diagrams. As such, they generate curiosity, the desire to understand, and, in turn, heighten spectators’ awareness of their processes of interpretation. The artworks, mixing the genres of landscape, portraiture and still life, offered an aesthetic experience which ran the risk of being dismissed as visual conundrums, cerebral puzzles disconnected from the social, economic, and political effects of the Depression. However, Post-Surrealism fostered an allegorical mode of thinking unconstrained by limitations of style, subject matter or political persuasion. It encouraged viewers to look attentively at everyday objects and make imaginative temporal and spatial connections among them. The first section of this essay introduces the Post-Surrealist movement and considers its relationship to European Surrealism. The second part examines how visual lessons from the Los Angeles Times and night sky performances at the newly-opened Griffith Planetarium introduced the public to new theories of the universe and astronomy. Part three considers several Post-Surrealist artworks in relation to these new theories of time and space. I focus my analysis on the works by Feitelson and Lundeberg, with more emphasis on the latter; the other Post-Surrealists’ engagement with astronomy will require a longer treatment elsewhere.
 

What Is Post-Surrealism? 

Since my initial research, the scholarship on Helen Lundeberg’s Post-Surrealist artworks has expanded considerably, especially after the artist’s notebooks, sketches and correspondence were donated to the Archives of American Art. Scholars have noted the importance of Lundeberg’s studies in science while pursuing a degree in literature at Pasadena City College.2 She took courses in astronomy, geometry, philosophy, and zoology, and considered a career in science illustration (Butterfield; Fort, “Imagined” 12). Ilene Susan Fort and Michael Duncan have made suggestive connections between her college drawings and her Post-Surrealist works.3 Her drawings of single-cell organisms such as paramecia and amoebae indicate that the artist developed drawing skills using a microscope in a zoological laboratory course and that she learned conventions associated with such scientific illustrations. 

Following graduation, Lundeberg received her initial artistic training under Lorser Feitelson at the Stickney Memorial School of Arts in Pasadena. As Diane Moran has shown, Feitelson’s teaching emphasized analyzing the pictorial structure of Renaissance, Mannerist and Baroque paintings. His pedagogical technique of diagramming artworks was intended for students to learn what he called the “classicism of suggestion,” whereby gestures and glances of figures carry meaning and direct the spectator’s eye through the pictorial composition (Moran, Painting 118, 121). Such diagramming of artworks no doubt complemented Lundeberg’s prior experience of drawing the form and structure of microscopic organisms. 

 In 1934, Lundeberg and Feitelson founded Post-Surrealism, which they also called Subjective Classicism or New Classicism, the first organized American response to European Surrealism.4 Other artists associated with Post-Surrealism are Grace Clements, Ethel Evans, Philip Guston, Reuben Kadish, Helen Klokke, Lucien Labaudt, Harold Lehman, Knud Merrild, and Elizabeth Mills, many of whom were former students of Lorser Feitelson. The first nominally Post-Surrealist exhibition, held at Stanley Rose’s Centaur Gallery in Hollywood in November 1934, was followed by a show at the Stanley Rose Gallery in March 1935. As Diane Moran notes, Feitelson was involved in developing these exhibition spaces at the bookstore in alliance with Murray Youlin and Stanley Rose (Moran, Lorser 67). Harold Lehman recalled that Stanley Rose’s Bookshop stocked many recent books and journals on art and literature and that he first encountered the work of Salvador Dali at the bookstore, where he purchased Dali’s Conquest of the Irrational (Polcari). Lundeberg stated that the bookshop was a “hangout for a lot of Hollywood people, the industry” next to Al Levy’s Tavern (Butterfield). At the time, Stanley Rose was cultivating interest in Russian films, which were viewed as avant-garde in contrast to Hollywood movies. In 1932, he became the publisher of Experimental Cinema, a journal devoted to avant-garde film, which featured essays on montage and cinematic close-up techniques written by directors and critics such as Sergei Eisenstein and Seymour Stern (Robe).

Following the exhibitions at Stanley Rose’s galleries, the Post-Surrealists exhibited in the fall of 1935 at the Hollywood Gallery of Modern Art, which Feitelson opened and operated with Lundberg (“Helen Lundeberg – American Painter”). Marcel Rivet, a screenwriter and Grace Clements’s husband, gave a lecture on “Cine-Montage and Modern Art” at the gallery.5 Northern California was introduced to the movement when several Post-Surrealist artworks were included in a group show at the Paul Elder Modern Gallery in San Francisco in June 1935, and later that year, a Post-Surrealism exhibition was held at the San Francisco Museum of Art. These shows were followed by an exhibition at the Brooklyn Museum during the summer of 1936 where the Post-Surrealists received substantial attention from East Coast art critics. Several Post-Surrealists were then included in Alfred Barr’s Fantastic Art, Dada, Surrealism at the Museum of Modern Art in December 1936. Thereafter, Post-Surrealists continued to hold exhibitions at Los Angeles galleries over the course of the next three years or so and several of them had works included in the American Art exhibition at the 1939 World’s Fair in New York.

Five out of the eleven artists involved in Post-Surrealism were women, a high number considered in relation to other art movements in the history of early modernism. Women were significantly represented in Post-Surrealist exhibitions and in articulating the Post-Surrealist aesthetic to the public; such visibility and participation, as noted by Susan Ehrlich, stands in contrast to women associated with European Surrealism.6 Although Lundeberg later credited Feitelson for developing Post-Surrealism, she wrote the movement’s manifesto, “New Classicism,” in a privately published pamphlet featuring a reproduction of Plant and Animal Analogies (Fig. 3).7 At the beginning of her essay, Lundeberg encouraged readers to look at the painting and to toggle back and forth between the visual and the verbal, an interpretative stance conducive to the Post-Surrealist aesthetic. In the manifesto, Lundeberg claimed that Post-Surrealism rejected mimesis as the sole function of art and that it offered a new aesthetic based on viewers’ processes of interpretation: a rhythm is experienced, she argued, “through contemplation of the subjective relationships and sequences of forms and groups of forms” (M. Duncan, “Helen” 104). Lundeberg suggested considering a Post-Surrealist painting like Plant and Animal Analogies as a site of signification, based on the normal functions of the mind: analogy, the recognition of idea-content forms, and the logical sequence of forms unrelated in size, time, or space (M. Duncan, “Helen” 104). These forms and sequences could create what Lundeberg called an idea entity or a mood entity

Plant and Animal Analogies, which she stated as presenting “simultaneously extremely dissimilar experiences of form: those directly visual-perceptive and those purely conceptive,” exemplified an idea entity (Lundeberg, “New” 19). The work suggests the genre of still life with a window framing a landscape; Lundeberg referred to this kind of format as an “H” form of composition.8 Superposed on the naturalistically-rendered objects are enlarged, grey and black painted, conceptive forms. She intended viewers to notice similarities between the naturalistically depicted halved bell pepper and the conceptual form of the cross-sectional embryonic brain on the left, or between the cross-sectional diagrams of the human embryo within a woman’s uterus and a cherry in the center. 

The conceptive forms reflect conventions associated with scientific illustrations intended to reveal unseen structures and functions of living things. Some of these cross-sectional diagrams derive from illustrations featured in Frederick Randolph Bailey and Adam Marion Miller’s Text-book of Embryology of 1921 (M. Duncan, “Helen” 104). Such scientific illustrations can engage spectators, as Matthew Allen has recently argued, by “drawing viewers in, confounding them, and prodding them to ask questions.” Following the diagrammatic lines and arrows, viewers compare a series of dissimilar images, while the dotted lines with circles initiate a shift to considering the cross-sectional view of the embryonic brain and the branching placenta form as magnifications of the cross-sectional view of the uterus on the right. Contiguity also encourages linkages, such as superposing the cross-sectional uterine diagram as an internal view of the sculptural female torso depicted behind it. Such prodding directives do not lead to a rational conclusion, but instead encourage speculation about gendered processes of growth and sustenance. Perhaps the most compelling proposition for viewers is to think about the similitude among all living things, and to become aware that we ourselves are implicated subjects in this series of comparisons. Considering what it means to be human, we become conscious of our own processes of deciphering and analogic leaps in reasoning. A significant element that fosters this perception is the surface of the painting: the paint has been thinly applied on a Celotex board, and the texture of the board calls attention to the painting’s surface. As such, viewers are never in doubt that they are looking at a constructed proposition. 

The Red Planet (1934) (Fig. 4) presents an unsettling, wondrous disjunctive experience that relates to Lundeberg’s concept of a mood entity. Painted on Celotex, it represents a visual paradox between worlds, in which viewers shift from perceiving an illusionistically rendered interior scene with a still life on a table to an allegorical reading of these everyday objects as a planetary system.



Directional allusions are less overt, functional, and diagrammatic than in Plant and Animal Analogies, as the golden doorknob seems to transform into a sun, a red marble on a table morphs into a planet, and the ellipse of the table becomes an imaginary orbit. This process is what Lundeberg explained as the Post-Surrealist aesthetic, “in The Red Planet as in The Mirror the introspective activity is not recorded but takes place entirely in the mind of the spectator” (Lundeberg, “Explanatory” 21). For this viewer, the painting evokes textbook explanations of the dimensions and relative distances of the solar system. For example, John C. Duncan, in his Astronomy: A Textbook (1927), quoted John Herschel’s well-known model of the planetary system and encouraged readers to: 

Choose any well-leveled field. On it place a globe, two feet in diameter; this will represent the Sun; Mercury will be represented by a grain of mustard-seed, on the circumference of a circle 164 feet in diameter for its orbit; Venus a pea, on a circle 284 feet in diameter; the Earth also a pea, on a circle of 430 feet; Mars a rather large pin's head, on a circle of 654 feet; Juno, Ceres, Vesta, and Pallas, grains of sand, in orbits of from 1,000 to 1,200 feet; Jupiter a moderate-sized orange, in a circle nearly half a mile across; Saturn a small orange, on a circle of four fifths of a mile; Uranus a full-sized cherry, or small plum, upon the circumference of a circle more than a mile and a half, and Neptune a good-sized plum on a circle about two miles and a half in diameter. (J. Duncan, Astronomy 257)

 

This analogical model dramatically reduces the gigantic scale of the planets through the use of familiar objects and fosters a way of envisioning the solar system from outside and from multiple viewpoints. In so doing, vastness of space and distance become comprehensible. In The Red Planet, Lundeberg reverses the intent of Herschel’s model; she suggests the possibilities of analogies to transform spectators’ perceptions of everyday life, where a room filled with familiar objects imaginatively becomes the universe. Joseph Solman, an artist and art critic for Art Front, praised the painting on these grounds: Lundeberg “dreams about planets, attracts them to her room and drinks in their topaz light.”

Lundeberg claimed that the Post-Surrealist aesthetic was a “new, unprecedented” form of classical order. Yet, as suggested above, wonder is not excluded from contemplating such visual analogies that hinge on perceiving formal and functional similarities and following sequences of forms. As Diane Moran notes, “the viewer realizes that the subjective meaning of the object changes according to its disposition in the composition. This quality, the potential to go beyond a fixed meaning, illustrates the open-ended character of the literary content of Post-Surrealism” (qtd. in M. Duncan, Post 9). An image or diagram in a Post-Surrealist artwork does not serve to convey one iconographic meaning, but encourages spectators to generate additional meanings or associations and complete the aesthetic experience. Indeed, Lundeberg indicated such an interpretative process in her description of another Post-Surrealist work, The Mirror: 

Every element functions to create the emotional entity of the picture. The realistic rendering of a possible arrangement of commonplace objects establishes a norm, which is disturbed by the subjective incongruity of the scallop shell and reflected electric light. Seeking, and failing to find, a rational significance and relationship of these forms to each other and to their surroundings, the mind takes refuge in a sort of mysticism. The commonplace objects take on a strangeness, a meaning which cannot be analyzed. The pleasure of this introspective activity determines the aesthetic value of the picture’s subjective form. (Lundeberg, “Explanatory” 21)

 

In her presentation of Post-Surrealist aesthetic in Art Front (March 1936), Grace Clements claimed that it provided a new means to create a socially conscious art. The unity of form and content, she stated, could be attained through “multiple perception—visual, psychological and philosophical.”9 Acknowledging that Cubism had challenged conventions of a fixed point of view, Clements claimed that Surrealism, with its interest in psychology, had engaged the “complexity of objects, both related and unrelated in form, function or idea.” (9) Arguing that Surrealists had provided the key to developing a new aesthetic by their use of associative ideas, Clements posed that Post-Surrealists directed such associations and created art that was calculated for comprehension. Clements referred to montage to explain how Post-Surrealists directed spectators’ processes of interpretation: “Montage, a term borrowed from the cinema, which, by conflict of two or more factors unlike in kind but united in idea, create a climax, is a method not unsympathetic to adaptability in the field of painting” (9). 

Lundeberg and Clements stress exploiting the functioning of the conscious mind to encourage the viewer to participate in the construction of meaning. They acknowledged the importance of Surrealism in developing the Post-Surrealist aesthetic in terms of associative ideas and use of analogy, but they also made clear distinctions to mark how Post-Surrealism departed from European Surrealism. In her manifesto, Lundeberg asserted that Post-Surrealists were not interested in exploring the subconscious (Lundeberg, “Explanatory” 21). For Clements, the weakness of Surrealism was its “attempt to make an automatic art, uncontrolled by the conscious mind” (9). Lorser Feitelson amplified this distinction and insisted that “the very term ‘artist’ still implies one who imposes order upon the images of either the outside world or the inner world. We believe that the wonders of the cerebral world are not only as marvelous as the fantasy of the subconscious world of basic sensual memories, but that their intellectual manipulations have incomparably profound possibilities” (qtd. in O’Connor 342–43). 

Framing Post-Surrealism in antithetical terms catered to art critics, such as Arthur Millier at the Los Angeles Times, who viewed Surrealism quite negatively and stressed the role of analogy in Post-Surrealism in his reviews. It also helped to remove Post-Surrealism from sensationalism surrounding the International Surrealist exhibition held in London (1936), and from the contemptuous responses to the work of Salvador Dalí at the time (Miller 63). However, the Post-Surrealists did absorb how Surrealists engaged analogy, juxtaposition, and dramatic shifts in scale. Helen Lundeberg, for example, sought out reviews and articles on Surrealism and drew lessons from them.10 The Post-Surrealists’ interest in perceptions of space, time, and scale paralleled concerns in the cultural and scientific communities of Los Angeles, specifically observational astronomy conducted at Mount Wilson. While there are no notebooks or materials relating to astronomy in the artists’ papers, Lundeberg indicated that she consulted “astronomy and picture books” at the time, and one such source for astronomical images and diagrams in the work of Lundeberg and Feitelson was John C. Duncan’s Astronomy: A Textbook (1927).11 A professor of astronomy at Wellesley College, Duncan frequently visited Mount Wilson Observatory to use the telescopes to photograph nebulae and stellar objects such as variable stars or Cepheids (see Fig. 2).
 

Astronomy, Cosmology and Visual Culture of Los Angeles 

Radical changes in the understanding of the universe occurred following World War One, when cosmologists engaged Albert Einstein’s theory of general relativity to generate new mathematical models of the universe. The new mathematical models, contributions of studies of thermodynamics, chemistry and observational astronomy, created what is now known as modern cosmology during the interwar period, with Los Angeles at the center of its formation.12 According to the geographers Denis Cosgrove and Veronica della Dora, the science of astronomy shaped habits of seeing within Los Angeles communities during the 1930s.13 Founded by George Ellery Hale with the financial support of the Carnegie Institution in 1904, Mount Wilson Observatory, in the San Gabriel Mountains overlooking Pasadena, boasted the world’s largest reflector telescopes (the 60-inch telescope erected in 1908 and the 100-inch telescope constructed in 1917). Astronomy at Mount Wilson supported claims that Los Angeles had become the Mecca of world science. The telescopes, in particular, were frequently referenced in the Los Angeles Times as technological marvels that allowed for seeing and pioneering new worlds, and yoked wonder to the frontier spirit of Manifest Destiny (Hampton; Sutton, “To Pioneer”). 

Equipped with cameras and spectroscopes, these gigantic telescopes facilitated the audacious project of Edwin Hubble, Milton Humason, Frederick Seares and Harlow Shapley to measure the size, shape, and motion of the universe. Their astronomical studies offered dramatic changes in perceiving humanity’s place in the galaxy and the universe. Our solar system was no longer thought to be at the center of the Milky Way; this profound shift in thinking had made front-page news in the New York Times (May 31, 1921), which quoted Shapely as saying, “Personally I am glad to see man sink into such physical nothingness, and it is wholesome for human beings to realize of what small importance they are in comparison with the universe” (qtd. in Bartusiak 131–32). Even more disturbing was the revelation described by James Stokley in his newspaper article “Newest Ideas about Space and the Size of Everything” (Fig. 5) that the Milky Way was not the only galaxy but merely one of millions of galaxies, part of a much larger system of “island universes.”14



The science writer took recourse to various analogies, photographs and diagrams to explain the astronomic research of Hubble, Humason, and Shapley, the latter now director of the Harvard College Observatory. The illustration at the upper right side of the page, for example, suggested that readers imagine themselves as if they were an infinitely large giant looking through a magnifying glass at the entire universe; they would then see different systems of stars starting with the largest island universes to a series of smaller clusters, to a magnification of a “million million times” where they would discover the smallest local systems made up of suns surrounded by planets like the earth. Stokley’s essay was filled with textual and visual analogies, such as conceiving light as a yardstick, visualizing a series of electric light bulbs in a corridor to indicate perceptual differences in the brightness and distance of stars or imagining an ant walking on a globe to illustrate the idea of curved space in a finite but unbounded universe.15 This latter analogy, intended to explain Einstein’s theory of relativity, was presented alongside a discussion of Hubble and Humason’s spectroscopic investigations of “spiral nebulae” (as galaxies were then known). Spectrographs, which revealed a “red shift,” helped Hubble to demonstrate that individual galaxies are moving away from each other, and that the universe is expanding, contradicting earlier theories of Albert Einstein and others who posited a static universe. 

The Los Angeles Times and other California newspapers regularly followed discussions of new and competing theories of evolution of the universe put forward by local researchers like Robert Millikan and Richard Tolman at Cal Tech, but also visiting cosmologists at Mount Wilson such as Willem de Sitter, Arthur Eddington, Albert Einstein, Georges Lemaitre, and James Jeans. For example, in the full-page article, “Latest Theories about the Beginning of the Universe and How It Will End” (1932), science writer Watson Davis provided an overview of theoretical responses to the new view of the expanding universe stemming from the astronomical research of Hubble and Humason. The Catholic cleric, Abbé Georges Lemaître, theorized that the universe had a definite moment of “birth” from a single “primordial atom” (his hypothesis later became foundational to the big bang theory). In his account of cosmic evolution, Lemaître claimed that from this primordial atom a rapid expansion like “fireworks” occurred, which was then followed by two stages: one of slowing down, then an accelerated expansion of space in which stars separated into extra-galactic nebulae.16 This theory would account for observations by the astronomers. Davis indicated that De Sitter and Eddington, both of whom had devised other theories of the universe, supported Lemaître's theory.17 Informed by studies of relativistic thermodynamics, Richard Tolman proposed an unending cyclical model of the universe fueled with perpetual energy, where expansion would continue until it “begins to contract, reducing in size until it undergoes a sort of rebirth after which it would continue to expand and contract again and again.” To visualize this cyclical process of contraction and expansion, Davis explained that “this Tolman universe is like a toy rubber balloon that is forever being blown up and deflated.” Davis briefly contrasted the Tolman universe to James Jeans’s entropic theory, in which the universe will eventually run down due to a loss of energy: it will retain its size but all material bodies are shrinking uniformly and would end “when all matter had shrunk to nothing.” Although not mentioned in Davis’s article, Robert Millikan viewed the universe in terms of his studies of the atom and theorized an eternal cycle of cosmic forces of destruction and creation. Millikan linked his idea of cosmic evolution to his Christian religious views, offering his studies of cosmic rays as evidence.18 The Times presented his theory as offering an optimistic view of the universe, while entropic theories suggesting annihilation or an ultimate heat death such as that posed by Jeans were framed in pessimistic terms, and according to historian Helge Kragh, were sometimes linked to nihilism and communism.19 Debates concerning these competing evolutionary theories of the universe offered radically different views of time and space to the general public. Far from providing a secure, reassuring position from which to contemplate the universe, these theories of the evolution of the cosmos would have provoked questions about the fate of the universe and humanity’s place within it. 

Surrealists were not alone in deploying dramatic shifts in scale, incongruent juxtapositions, and analogies; such visual tactics appeared repeatedly in the Los Angeles Times’ extensive coverage of astronomy and accompanying diagrams, photographs and illustrations, often drawn by Charles H. Owens. The Times staff artist provided on an almost weekly basis lessons in deciphering scientific illustrations featuring arrows, cutaways, oblique or bird’s-eye views, sequences of enlargements and a mixture of text and images. His two-page spread for the proposed 200-inch reflector telescope (Fig. 6) was intended to convey visually various kinds of highly complex information through a structure akin to montage in film.20 Through dramatic shifts in scale and varying points of view, Owens shaped how the public understood the planetary system, such as his imagined “trip to the moon” with the 100-inch Hooker telescope at Mount Wilson (Fig. 7), which brought the moon’s surface within 60 miles of an earthly observer.





These magnified views of the lunar landscape were likened to the relative proportions of familiar architectural landmarks such as the Hollywood Bowl and Yankee Stadium as if “seen from an airplane a mile overhead.” Although their focus is on how Owens’s later pictorial “air-age” cartography ideologically formed Americans’ understanding of World War Two global politics, Cosgrove and della Dora’s analysis of his repertoire of graphic techniques led them to argue that Owens’s representations of astronomy complemented Edwin Hubble’s view of a dynamic, expanding universe.21 

The Los Angeles Times’ s coverage of the planning and construction of the Griffith Planetarium and Observatory, which opened in May 1935, similarly offered the public new ways to consider relations among the Earth, the planetary system, and the universe in terms of spectacle and wonder. Atop Mount Hollywood, the domes of the planetarium and observatory at Griffith Park were likened to a Roman temple, Moorish mosque or a mausoleum.22 Others humorously claimed that at night the planetarium above Hollywood gleamed “like a birthday cake someone has tipped over.”23 Visitors strolled by Foucault's pendulum, saw a working model of the universe as well as a miniature model of the Earth in cross-section, and viewed new photographs of the night sky taken at Mount Wilson. One exhibit offered a Jules Verne “Trip to the Moon” in which a giant plaster replica of the moon could be viewed from an “airship” observation platform. The newspaper described the new Zeiss projector, one of only two in the United States, by way of metaphor: it was a colossal “dumbbell” (sometimes compared to a giant ant) which astonished visitors by shining stellar and planetary images on the interior dome with its painted silhouettes of the city skyline. 

A newspaper heading claiming “They have brought the universe to Los Angeles” suggests a dramatic miniaturization of the night sky to be mastered or possessed, yet many reports framed the immersive experience of watching the program as an almost religious experience. Programs consisted of identifying familiar constellations of the night sky, but audiences were enthralled by the awesome and ever-stupefying experience of watching stars and planets move at great speeds in dramatic visual compressions of space and time (Turner; “Park”). Philip Fox, director of the Adler Planetarium in Chicago and an official at the opening ceremonies, remarked that the new planetarium offered a wondrous spectacle, which he stated, was “the drama of creation. This mystery of life is akin to the sweep of the stars in the infinite age of the skies. Here in this theater one can lose himself for awhile with the suns, moons, and stars as the actors” (“Park”). Following these comments, Fox activated the projector and made the instrument perform “astounding tricks like hurling the stars around the ceiling and completing their circuits for an entire year in the space of four minutes” and the crowd gasped (“Sky”). Philip K. Scheuer, the film critic of the Los Angeles Times, also claimed the spectacle was breathtaking. In one of his daily columns, Scheuer interviewed the director of Griffith Planetarium, Dinsmore Alter, who compared the planetary program to a moving picture, which he asserted was “far more magnificent than anything Hollywood conceived.” Scheuer lent credence to Alter’s claim, as the film critic faulted the Hollywood movie industry for relying on clichéd rapid dissolve techniques to signify dramatic temporal and spatial shifts. Hollywood, Scheuer seemed to imply, should learn from the planetarium’s presentation and develop new devices to compress time and space to advance a storyline, a concern he had earlier expressed in columns addressing Sergei Eisenstein’s film techniques and montage. 
 

Cosmic Allegories: Post-Surrealism 

In his review of the Post-Surrealist exhibition at the Brooklyn Museum, Joseph Solman faulted Lorser Feitelson for negating the “great class struggle today” through his preoccupation with “religious-philosophical connotations.” However, Feitelson believed that in considering a theme such as Genesis, he was engaging a subject that is “the most important that man is capable of contemplating, for metaphorically it deals with the scheme of all things” (Feitelson, “Notes”). Like Lundeberg’s Plant and Animal Analogies, his allegorical tableaux, Genesis #2 (Fig. 8), posits links between the microcosmic and macrocosmic realms in terms of the female body, but framed by religion and science.



The overriding theme engages the origin of the word “galaxy,” as in the ancient Greek word galaktos for milk and its related term in Latin “via lactea” or milky way. Based on resemblance of the cloudy appearance of the night sky to mammary fluids, both analogies attempt to give tangible form to the overwhelming immensity of cosmic clouds and stars. Feitelson no doubt would have been aware of Tintoretto’s Origin of the Milky Way (1575) and Rubens’s Birth of the Milky Way (1637) which represent the generative power of the mythical goddess Hera by representing milk spurting out of her breasts to form the cloudy stream of stars (Kuberski 61–62). The artist conflates these mythical allusions with Christian iconography of the Virgin Mary through the depiction of an Annunciation scene on the right side of the “H” form of composition.24 This image is overlaid with a schematic drawing of a female head and torso, which is then linked to a pair of increasing enlarged white lines representing curved breasts. These lines are connected to a larger cross-sectional diagram of mammary ducts and glands to suggest their biological function. This visual sequence of enlargements shifts conceptually from religion to science to a reference to the cultural practice of infant feeding through the final object, the outmoded baby bottle. Near the nursing bottle is the naturalistic depiction of a broken eggshell, perhaps an alchemical reference to the primordial cosmic egg. The central series of facial masks, suggesting a temporal theme of ontogeny or the human life cycle, is connected to the celestial sky through a collapsible hand telescope, an amateur optical device common in the artist’s youth (Feitelson, “To Diane”). Symbolic associations suggesting generative powers of the female body, on the left side of the painting are representations of natural objects, a conch shell and cantaloupe, linked to a mass-produced cultural object, the electric light bulb with pull cord. Echoing the representation of divine energy emanating from the dove on the right, the bulb radiates light to illuminate the sequence of forms in the central foreground area. Feitelson hoped that such pictorial sequences of “optically apprehended thoughts” would prompt spectators to consider “all things in terms of cosmic relationships” (“News Notes”; Feitelson, “To Diane”). 

Feitelson continued to explore the theme of creation in his Life Begins (Fig. 9), but without mythological or religious connotations. Instead, he experimented in collage, a departure from his painting practice, by gluing photographic reproductions that were then circulating within American mass culture.



This irregularly-shaped work on a sheet of blue painted Masonite heightens its montage structure; as Moran notes, Feitelson began using shaped canvases to direct the viewer’s gaze across the depicted objects. She asserts that Post-Surrealists were “convinced that internal structure and the meaning of the pictures should determine its peripheral shape” and thus contribute to the process of signification.25 The three sections of the collage suggest finding commonality among different representations of creation. On the left side of the work is a page from the inaugural November 23, 1936, issue of Life magazine.26 This photographic image represents the moment of a birth, in which the viewing position is distant and from a bird’s-eye view of a baby held upside-down by a male doctor surrounded by nurses (the mother is not visible in the scene) in a hospital delivery room. Across from the glued-in magazine page is a painted section of a peach pit in front of a halved peach resting on a plate, metaphorically suggesting a prior moment in time of being in a mother’s womb. Above this painted area is another photographic image, an enlargement and slightly blurry section of Duncan’s well-known astronomical photograph of the western portion of the Veil Nebula (N.G.C. 6960 Cygni) to suggest creation as a sudden explosion or ejaculation of life. Featured as the frontispiece image in his Astronomy: A Textbook (1927), the half-tone photograph, part of the author’s prior study of the structure of nebulae at Mount Wilson, was taken with the 100-inch Hooker reflector telescope. Hubble considered the Veil Nebula to be a “wreath” form that occurred after the formation of a nova.27 Feitelson’s sequence of images would thus relate to contemporary theories of the birth of a new solar system through the sudden explosion of a nova and engage Lemaître’s theory of the evolution of the universe involving a rapid expansion or “fireworks” from a primeval atom (egg). 

Like Feitelson’s Life Begins, Lundeberg’s Relative Magnitude (Fig. 10) is a montage of images that engages astronomic knowledge of space and time, but poses them in a humorous way (Helen Lundeberg: American Painter).



Instead of adhering photographic reproductions to the work, Lundeberg transforms images illustrating Duncan’s textbook on astronomy. When I interviewed the artist, Lundeberg stated that she used the smooth side of a Masonite board to heighten the spectators’ awareness of the four interrelated rectangular pictures, representing two parallel viewpoints of a male astronomer and an ant. Lundeberg does not engage Owens’s illustrations of the existing telescopes at Mount Wilson or the future 200” telescope (see Figs. 6 and 7); instead, she turns to nineteenth-century technology.28 Dressed in antiquated clothing, the comical male figure is shown seated between two piers supporting a telescope, gazing intently through the instrument, but he is a passive spectator to the celestial sky above. The circular view of stars and guide wires in the reticle calls attention to the astronomer’s act of viewing through the telescope. The transit instrument, in combination with a sidereal clock, was used by astronomers in the nineteenth century and afterward to locate the direction of a star on the celestial globe but not its distance. Mapping stars in this manner thus constituted a projection of a three-dimensional volume onto a two-dimensional (spherical) surface. Transit instruments were also used to determine time precisely using “clock stars” with known locations (J. Duncan, Astronomy 27–28). However, in Lundeberg’s painting no time is indicated on the face of the clock, nor are forms depicted on the celestial globe, thus indicating impotency or an incapacity to understand. 

The series of images reveals an ant’s perception of the world. The lower left panel depicts an interior scene: resting on a table is a piece of cloth or drapery, a marble, and an opened book, all proportionally scaled to one another and to a human-scaled environment. A red line is painted and leads to a red circle that presents a magnified view of an ant. Through recognizing a shift in scale, the viewer is led to the assumption that the dot on the book is an ant. In the upper panel, the ant’s perception of the world is presented, where the curved page of the book is transformed into a flat plane, the cloth becomes a mountainous landscape and the solid red marble becomes a striated planet Jupiter looming on the horizon.29

The analogical structure, the improbable comparison between the view of the ant and that of the astronomer, veers toward irony. The ant seems to be attempting to understand its world, but its vantage point prevents it from understanding that the brightly colored marble is not the planet Jupiter. The astronomer’s relation to the cosmos is similar to the ant’s position to the marble. Like the ant, the astronomer is confined to an earthbound perspective. Knowledge of the cosmic through visual observation and the telescope is suggested to lead equally to misapprehensions. This ironic reading, however, is available to an omniscient vantage point of the viewer, who is invited to shift between the various points of views and perceive the two orders from a distant view. 

Several of Lundeberg’s Post-Surrealist paintings rework the image of the male astronomer; her Microcosm and Macrocosm (Fig. 11), The Veil (Fig. 12) and Self-Portrait (see Fig. 1), explicitly reference the artist and gender in posing questions about the universe.  This is distinct from how the Los Angeles Times and other publications represented astronomy and science as largely a male enterprise (for an image of the male astronomer, see Margaret Bourke-White’s 1937 photograph of Hubble featured in Life magazine—“Carnegie”). Although women astronomers and researchers known as computers worked at Mount Wilson, they were largely invisible to the public and seldom recognized as peers.30 As a 1944 publication described them, women were not telescope observers of the night sky, but daytime interpreters of photographic images produced by these instruments (Richardson). Key achievements in astronomy developed out of such research, for instance Henrietta Leavitt’s standard photographic magnitude scale. Her period-luminosity studies of variable stars provided a foundational method which spurred the interwar astronomical studies at Mount Wilson. Harlow Shapley used her method to measure the size of the Milky Way, while Edwin Hubble applied it to observations of galaxies lying beyond the Milky Way, confirming the theory of “island universes” (Bartusiak 57). 

Lundeberg addresses gender in her painting Microcosm and Macrocosm (Fig. 11; see right) through a curious self-portrait. Only a fragmented view of the artist is represented in the upper left corner of the painting, yet the figure magically appears to conjure up an analogic world where microcosmic organisms are compared to celestial bodies of the universe. The self-portrait’s proportional relations with the depicted sea and sky are out of scale, as if the figure embodies the size and point of view of the “infinite giant” as described by Stokley. As in her earlier Post-Surrealist paintings, Lundeberg incorporates self-portraiture to suggest intense thought and here, the image of the artist serves as a diegetic agent that actively orchestrates the comparisons.31 In contrast to the distant view fostered in Relative Magnitude, the portrait offers a surrogate for the beholder to inhabit, for the viewer is invited to mirror the figure’s contemplative stance and to follow the figure’s downcast eyes that gaze through a magnifying glass held in the figure’s right hand. Lundeberg intended the magnifying glass to symbolize forms of seeing associated with both microscopes and telescopes (Helen Lundeberg: American Painter, 17:20). Her arrangement of microscopic views of the sea moves from a small grouping of indistinguishable forms to a magnified grouping, culminating in the large “blow up” image of an amoeba. These shifts in scale recall the use of close-ups in montage filmic sequences that encourage spectators switching from close-up views to the overall view of the painting and back again. 

Spectators are directed to the section representing the vast space of the cosmos in the upper part of the painting to the red diagrammatic circle enclosing stars and clouds. An arrow links this section of stars to an enlarged view of Saturn, which probably derives from a nineteenth-century drawing reproduced in Duncan’s textbook on astronomy. In his book, the astronomer marvels at Saturn: “To the naked eye, Saturn appears as a star of the first magnitude, but not eminent in any way above several other stars. By a good telescope, it is revealed as a delicately banded globe poised within a shining ring—a spectacle which, once seen, can never be forgotten” (J. Duncan, Astronomy 252). He explains that although the rings of Saturn appear as continuous bands, they are neither solid nor liquid, but are similar to a “swarm of independent bodies like meteors” (J. Duncan, Astronomy 254). His recourse to analogy was due to the uncertain knowledge concerning the planet; astronomers at the time offered competing theories concerning the origin of the rings. Henry Norris Russell proposed that the rings were too close to Saturn for a moon to form and that gravity would have broken any large solid mass into smaller and smaller pieces (Russell 31–32).32 James Jeans suggested that the rings represented a moon broken up by tidal forces near Saturn (Sutton, “What’s New”), and warned that Earth’s moon would eventually share the same fate (Millard). Lundeberg suggested such destructive possibilities in a related Post-Surrealist work, Cosmicide (1935), where death, rather than life, serves as the linkage between microcosmic and macrocosmic realms. But in Microcosm and Macrocosm, Lundeberg offers speculative and wondrous possibilities for Saturn through emphasizing microcosmic life in the lower section of the painting. What would we see if a magnifying glass could magically reveal the planet and its rings? Would we see complexity comparable to that in the sea below? 

The aquatic scene reveals wonders through magnification. The fish at the bottom pursue a cloudlike form of plankton, gradually revealed through magnified, encircled forms. The largest magnified protozoan is Amoeba proteus, a single-cell organism which the artist had sketched during her studies at Pasadena City College. Instead of representing the amoeba in accordance with conventions associated with scientific diagrams, as she had done previously, Lundeberg represents the amoeba, an organism that can assume any shape, as a semi-transparent, almost gelatinous three-dimensional form.33 Below this image is a red-circled enlarged view with various illustrations of protozoa with flagella, Euglena viridis, around the Amoeba proteus. The latter is depicted surrounding, perhaps devouring, another protozoan form. These various images engage what most interested zoologists at the time: processes of reproduction and movement.34 We are shown Euglena self-replicating through division, while another image depicts how this single-cell organism moves through contraction. The Euglena was a subject of much curiosity in zoology at the time, as it defied established categories of plant and animal and was understood as possessing one of nature’s most elementary eyes, a primitive “eyespot.”35 As Hegner noted, both the Amoeba and Euglena exhibit phototropism (phototaxis), in that they move in response to light (Hegner 37–38, 43–44). Representing the wonders of unseen life in water encourages viewers to follow the artist in making imaginative associations with the cosmic realm above.

The suggestion that the artist is an orchestrator of analogies operates in several other Post-Surrealist paintings, sometimes through representations of fragmented hands (Fig. 12).36 In the small painting, The Veil, Lundeberg represented a life-size woman’s hand, in a one-to-one accordance with the viewer, but it appears gigantic when considered in relation to the depicted landscapes. The hand, the artist explained, “is a useful introduction of something human...it can hold a flower, stone, a shell, what have you. It can point at something. I used my own hands because I was the most convenient model” (Stein 15). The image of a disembodied hand was a potent sign at the time. Hands served as a substitute for a portrait of an individual in early modernist photography. Surrealists’ use of hands was linked to this practice in visionary terms of palm reading in the journal Minotaure. Male surrealists also invested representations of women’s disembodied hands with an uncanny erotic charge.37 Beauty columnists and advertisers included severed graceful hands as markers of white femininity in the pages of the Los Angeles Times and other mainstream periodicals, while fragmented male hands served to represent work and to symbolize labor unions. 

The gigantic well-manicured feminine hand in The Veil functions ambiguously within the realm of such meanings, but its gesture of pulling aside a curtain is one of revealing that which is concealed. Astronomers frequently used the metaphor of a veil to refer to the Earth’s atmosphere, which hindered observational astronomy. “The most maddening thing to astronomers,” according to one Mount Wilson researcher, was the existence of a heavy layer of ozone 50,000 feet above the earth’s surface (“Gold of Sun”). On the right side of the painting, Lundeberg presents a celestial view of planets stripped of atmosphere. Joseph Young interpreted this painting as a depiction of the artist’s own hand revealing an “identical” landscape in day and night, in which the sun-filled landscape serves as a veil hiding another world. The left side of the painting does show a sunlit, flowered desert landscape. However, the night scene on the right is not identical: the darkened horizon curves down toward the lower right corner of the painting, suggesting a non-terrestrial perspective. This convention of the curved horizon is deployed in the work of Howard Russell Butler, Mars as Seen from Deimos (n.d.), and Charles Owens likewise used to represent the terrain of the moon to suggest a “moon man’s” view of the earth (see Fig. 7). The Veil’s night sky reveals celestial objects which would not be visible without magnification, in particular a fanciful rendering of the famous Horsehead Nebula in Orion, which Duncan had photographed at Mount Wilson in 1920. Dark nebulae such as the Horsehead, like the stratospheric ozone, served to remind observers that the seemingly transparent media of the air around us or of interstellar space constantly intervene in our processes of vision. 

In her Self-Portrait (see Fig. 1), Lundeberg experimented with an accordion-like spatial composition format to shift scale among illusionistic worlds.38 This format rhythmically directs the viewer to consider perceptual boundaries between the landscape depicted on an easel and the interior setting occupied by the artist. Formal similarities move the eye across the painting. The horizontal orange paintbrush simultaneously points to the body of the artist and the depicted cosmic landscape. The turquoise paint at the end of the paintbrush refers to the color of the painted ball and represented planet. The ball held in the artist’s hand corresponds to the planet represented in the painting, and the shadow across the face of the artist echoes the edge of the depicted cloudlike form in the painting. 

The portrait of the artist is grounded in American contemporary life as the painted fingernails, red lipstick and hairdo references fashion. In the painting, Lundeberg foregrounds her role as an artist and generator of such cosmic analogies. The figure looks directly outward to the viewer as if to posit her skill in investigating and reimagining the natural world and the cosmos. This self-assurance grew from her scholarly and artistic studies of the interwar period; as Lundeberg later stated, “My college-day interest in the biological sciences and the technical aspects of literature, undoubtedly motivated my eagerness to investigate the intricacies of aesthetic structure and mechanisms in all the significant phases of art” (Lundeberg, unedited statement). One might add that her curiosity towards astronomy and zoology fueled her Post-Surrealist speculations about the wondrous aspects of everyday life. 


Notes 

1 See Kinkel (104, 127). For a recap of my scholarship, see Stein.

2 For more on Lundeberg’s biography, see Muchnic; Fort, “Helen Lundeberg: Imagined”; Fort, “Helen Lundeberg, an Oxymoron”; M. Duncan, Post.

3 See Fort, Helen; M. Duncan, “Helen” (104).

4 See M. Duncan, Post (5).

5 See Muchnic (40); Moran, Painting (132–33).

6 See Ehrlich.

7 Muchnic states that it was published in October 1934 (40, 194) and notes that Lundeberg’s “Explanatory Text for Six Paintings by Helen Lundeberg” was also privately published in 1934 (194). The Los Angeles Times featured a reproduction of Plant and Animal Analogies in their announcement of the Post-Surrealist exhibition at the Stanley Rose Gallery (March 24, 1935).

8 Lundeberg, in her interview with Jan Butterfield, stated it was difficult not to consider the traditional visual compositional principles in painting.

9 See Clements (8). 

10 The art historian Ilene Susan Fort discusses how Lundeberg took notes on an essay concerning the work of Max Ernst.

11 Interview with author. Fort recognizes that astronomy was an important subject for Lundeberg and notes the prominence of the science during the 1930s (“Adventuresome” 96). See also G. Allen on Duncan and Feitelson’s Life Begins

12 See Fernlund (30); and Kragh, Conceptions.

13 See Cosgrove and della Dora (387).

14 See Stokely.

15 See Stokely.

16 See Kragh, Conceptions (153–54).

17 See H.D. (700_.

18 See “Millikan.”

19 See Kragh, Entropic (202).

20 Cosgrove and della Dora take recourse to terms such as collage, comic strips and montage to describe Owen’s various graphic techniques. 

21 See Cosgrove and della Dora (387). 

22 See “Sky” (1). 

23 See Scheuer.

24 Feitelson’s allegorical tableaux is reminiscent of 17th century vanitas paintings. Artists such as Pieter Claesz, Pieter van Roestraten, Jacques de Gheyn II, and Simon Luttichuys created still life paintings which harnessed everyday objects to evoke meanings concerning the passage of time. On still life, see Bryson, Brusati, and the J. Paul Getty Museum.

25 See Moran, Lorser (73–4); Moran quoted in Kinkel, Visual Poetics (127, n. 241). 

26 The Life photograph was credited to Andre Da Miano. 

27 See J. Duncan, Astronomy (402). Duncan called the Veil Nebula “one of the beautiful filamentous nebulae in Cygnus” (“Photographic” 146). Burnham includes images and descriptions of N.G.C. 6960. 

28 The images of the astronomer/telescope and the reticle are based on plate 1.4 and figure 16 in M. Duncan, Astronomy. The astronomer/telescope drawing depicts Hartwell Transit House in Great Britain in the mid-19th century; see Smyth.

29 The left side of the painting with an ant’s point of view on Jupiter seems to be based on studies of Jupiter (3 photos taken with the Hooker Telescope, Lowell Refractor, from J. Duncan, Astronomy 240). 

30 Beginning in the 1930s, women astronomers receive belated recognition; in 1934, the American Astronomical Society inaugurated its Annie Jump Cannon Award, named after one of Harvard’s many female astronomers, to honor outstanding women astronomers.

31 See, for example, Lundeberg’s Artist, Flowers and Hemispheres.

32 The rings of Saturn were the subject of much debate at the time, especially when a white spot suddenly appeared on the planet in August of 1933. This unexplainable event, visible only through telescopes, was subject of much speculation in the press. Was it a cloud or volcanic eruption?

33 Her image resembles zoologist Oris Dellinger’s photographs of amoeba in locomotion, commonly featured in textbooks See Hegner (35, 42) and Menge (123 and 127) for the Amoeba and Euglena drawings.

34 These forms, based on anatomist Gilbert Bourne’s drawings, were reproduced in many zoology textbooks)

35 See Hegner (43). The eyespot was thought to offer clues to the evolution of vision (Bailey 404).

36 Other Lundeberg works, as in Poetic Justice, also include large, well-manicured hands, and Grace Clements similarly incorporated fragmented hands in her work.

37 See Hoving (104–106).

38 Lundeberg, Feitelson and other Post-Surrealists used this accordion-like or inverted “W” composition format to suggest dramatic shifts in time and space. Here, Lundeberg uses it in a more subtle way.

 

Works Cited

Adamowsky, Natascha. The Mysterious Science of the Sea, 1775–1943. Routledge, 2016.

Allen, Greg. “On Lorser Feitelson’s Life Begins.” http://greg.org/archive/2014/12/22/on_lorser_feitelsons_life_begins.html. Last accessed 19 December 2018.

Allen, Matthew. “Compelled by the Diagram: Thinking through C. H. Waddington’s Epigenetic Landscape.” Contemporaneity: Historical Presence in Visual Culture, vol. 4, no. 1, 2015. DOI 10.5195/contemp.2015.143.

Bailey, W. E. “Evolution of the Human Eye.” Scientific American, vol. 137, no. 5, November 1927, pp. 404–406.

Bann, Stephen. Ways around Modernism. Routledge, 2007.

Bartusiak, Marcia. The Day We Found the Universe. Vintage, 2009.

Bassett, James. “The Scientists Speak.” Los Angeles Times, 3 December 1947, p. 20.

Brusati, Celeste. “Stilled Lives: Self-Portraiture and Self-Reflection in Seventeenth-Century Netherlandish Still-Life Painting.” Simiolus: Netherlands Quarterly for the History of Art, vol. 20, nos. 2-3, 1999, pp. 168–82.

Bryson, Norman. Looking at the Overlooked: Four Essays on Still Life Painting. Harvard UP, 1990.

Burnham, Robert. Burnham’s Celestial Handbook: An Observer’s Guide to the Universe beyond the Solar System. Dover Publications, 1978.

Butterfield, Jan. “Oral History Interview with Helen Lundeberg, 1980 July 19–Aug. 29.” Archives of American Art, Smithsonian Institution, Washington, D.C. https://www.aaa.si.edu/collections/interviews/oral-history-interview-helen-lundeberg-12910.  Last accessed 19 December 2018.

"Carnegie Institution's Mt. Wilson Observatory Makes Astronomical History." Life, 8 November 1937, pp. 44–53.

Chekhov, Anton. “Terror: My Friend’s Story.” The Party and Other Stories, translated by Constance Garnett, The Macmillan Company, 1917, pp. 63–86.

Clements, Grace. “New Content: New Form.” Art Front, vol. 2, no. 4, March 1936, pp. 8–9.

Cosgrove, Denis E., and Veronica della Dora. “Mapping Global War: Los Angeles, the Pacific, and Charles H. Owens’ Pictorial Cartography.” Annals of the Association of American Geographers, vol. 95, no. 2, 2005, pp. 373–90.

Davis, Watson. “Latest Theories about the Beginning of the Universe – and How It Will End.” Modesto News-Herald, 3 January 1932, n.p.

Doane, Mary Ann. “The Close-Up: Scale and Detail in the Cinema.” differences: A Journal of Feminist Cultural Studies, vol. 14, no. 3, Fall 2003, pp. 89–111.

Duncan, John Charles. Astronomy: A Textbook. Harper and Brothers, 1927.

---. “Photographic Studies of Nebulae: Third Paper.” Astrophysical Journal, vol. 57, 1923, pp. 137–48. http://adsabs.harvard.edu/full/1923ApJ....57..137D

Duncan, Michael. “Helen Lundeberg’s New Classicism.” Fort, Helen Lundeberg: A Retrospective, pp. 98–110.

Duncan, Michael, curator. Post Surrealism. Pasadena Museum of California Art / Nora Eccles Harrison Museum of Art, 2002.

Ehrlich, Susan. Pacific Dreams: Currents of Surrealism and Fantasy in California Art, 1934–1957. UCLA at the Armand Hammer Museum of Art and Cultural Center, 1995.

Feitelson, Lorser. “Notes on Genesis II.” In M. Duncan, Post Surrealism, p. 16.

---. “To Diane Moran.” 17 March 1976. Box 3, Folder 57, Lorser Feitelson and Helen Lundeberg Papers, Archives of American Art, Smithsonian Institution, Washington, D.C. https://www.aaa.si.edu/collections/lorser-feitelson-and-helen-lundeberg-papers-7341 (hereafter Feitelson-Lundeberg Papers).  Last accessed 19 December 2018.

Fernlund, Kevin J. “To Think like a Star: The American West, Modern Cosmology, and Big History.” Montana: The Magazine of Western History, vol. 59, no. 2, Summer 2009, pp. 23–44, 93–95.

Fort, Ilene Susan. “The Adventuresome, the Eccentrics and the Dreamers: Women Modernists of Southern California.” Independent Spirits: Women Painters of the American West, 1890–1945, edited by Patricia Trenton, U of California P / The Autry Museum of Western Heritage, 1995.

---. “Helen Lundeberg, an Oxymoron.” Intersections: Women Artists/Surrealism/Modernism, edited by Patricia Allmer, Manchester UP, 2016.

---. “Helen Lundeberg: Imagined Rather than Seen.” Fort, Helen Lundeberg: A Retrospective, pp. 10–36.

Fort, Ilene Susan, et al. Helen Lundeberg: A Retrospective. Laguna Art Museum, 2016, pp. 193–210.

The J. Paul Getty Museum. “In Focus: Still Life.” http://www.getty.edu/art/exhibitions/focus_still_life/.  Last accessed 19 December 2018.

“Gold of Sun Poetic Myth.” Los Angeles Times, 22 February 1929, pt. II, p. 9

H.D. “The Evolution of the Universe.” Nature, vol. 128, 24 October 1931, pp. 699–701.

Hampton, Edgar Lloyd. “At the Foot of the Rainbow.” Los Angeles Times, 2 January 1929, Annual Midwinter number, section 1, p. 3.

Hegner, Robert W. College Zoology. Macmillan, 1912.

Helen Lundeberg: American Painter. Narrated by Helen Lundeberg. Edited by John Amodeo. Atmosphere Productions, 1987.

“Helen Lundeberg—American Painter.” Tobey C. Moss Gallery, http://www.tobeycmossgallery.com/hlvideo.html.  Last accessed 19 December 2018.

Hoving, Kirsten A. “’Blond Hands over the Magic Fountain’: Photography in Surrealism’s Uneasy Grip.” Jennifer Blessing, editor. Speaking with Hands: Photographs from the Buhl Collection, Guggenheim Museum, 2004, pp. 93–113.

Kinkel, Marianne. Visual Poetics: The Post-Surrealist Paintings by Helen Lundeberg. 1991. California State University, Chico, MA thesis.

Kragh, Helge S. Conceptions of Cosmos: From Myths to the Accelerating Universe: A History of Cosmology. Oxford UP, 2007.

---. Entropic Creation: Religious Contexts of Thermodynamics and Cosmology, Ashgate, 2008.

Kuberski, Philip. The Persistence of Memory: Organism, Myth, Text. U of California P, 1992.

“Life Begins.” Life, vol. 1, 23 November 1936, p. 2.

Lundeberg, Helen. “Explanatory Text for Six Paintings by Helen Lundeberg.” In M. Duncan, Post Surrealism, pp. 21–22.

---. Magic Litera, Box 11, Folder 29, Feitelson-Lundeberg Papers.

---. “New Classicism.” In M. Duncan, Post Surrealism, p. 19.

---. Night Thoughts – 10/28/1981, Box 8, Folder 69, Feitelson-Lundeberg Papers.

---. [On Chekhov], Box 8, Folder 71, Feitelson-Lundeberg Papers.

---. Personal interview. 17 November 1988.

---. unedited statement for the Americans 1942 exhibition. Roll 1103, frame 366, Feitelson-Lundeberg Papers.

Menge, Edward J. General and Professional Biology. Bruce Publishing, 1922.

Millard, Bailey. "What, No Moon?" Los Angeles Times, 17 November 1935, pt. II, p. 4.

Miller, Angela. “’With Eyes Wide Open’: The American Reception of Surrealism.” Caught by Politics: Hitler Exiles and American Visual Culture, edited by Sabine Eckmann and Lutz Koepnick, Palgrave Macmillan, 2007, pp. 61–94.

“Millikan Finds Creation Still Goes on While Creator Directs the Universe.” New York Times, 30 December 1930, p. 1.

Moran, Diane. Lorser Feitelson: Eternal Recurrence. The Feitelson / Lundeberg Art Foundation / Louis Stern Fine Arts, 2014.

---. “Post-Surrealism: The Art of Lorser Feitelson and Helen Lundeberg.” Arts, vol. 57, no. 4, 1982, pp. 124–28.

Moran, Diane De Gasis. The Painting of Lorser Feitelson. 1979. University of Virginia, Ph.D. dissertation.

Muchnic, Suzanne. Helen Lundeberg: Poetry Space Silence. The Feitelson / Lundeberg Art Foundation / Louis Stern Fine Arts, 2014.

"News Notes." Brooklyn Museum Quarterly, vol. 23, no. 3, July 1936, pp. 144–49.

O’Connor, Francis V. “Philip Guston and Political Humanism.” Art and Architecture in the Service of Politics, edited by Henry A. Millon and Linda Nochlin, MIT Press, 1978, pp. 343–55.

“On a Mexican Wall.” Time, 1 April 1935, pp. 46–48.

“Park Gift Dedicated.” Los Angeles Times, 15 May 1935, pt. II, p. 2.

Polcari, Stephen. “Oral History Interview with Harold Lehman, 1997 Mar. 28.” Archives of American Art, Smithsonian Institution, Washington, D.C. https://www.aaa.si.edu/collections/interviews/oral-history-interview-harold-lehman-12894.  Last accessed 19 December 2018.

Richardson, Charles S. “Astronomy – The Distaff Side.” Astronomical Society of the Pacific Leaflets, vol. 4, no. 181, 1944, pp. 238–45.

Robe, Chris. “Eisenstein in America: The Que Viva Mexico! Debates and the Emergent Popular Front in U.S. Film Theory and Criticism.” The Velvet Light Trap, no. 54, Fall 2004, pp. 18–31.

Russell, Henry Norris. The Solar System and Its Origin. Macmillan, 1935.

Scheuer, Philip K. “A Town Called Hollywood.” Los Angeles Times, 3 November 1935, part II, p. 3.

“Sky Theater Awes Crowd.” Los Angeles Times, 9 February 1936, pt. II, pp. 1–2.

Smyth, William Henry. Aedes Hartwellianae, or Notices of the Manor and Mansion of Hartwell. John Bowyer Nichols and Son, 1851.

Solman, Joe. “The Post Surrealists of California.” Art Front, vol. 2, no. 7, p. 12.

Stein, Donna. “The Art of Helen Lundeberg: Illuminating Portraits.” Woman’s Art Journal, vol. 27, no. 1, Spring–Summer 2006, pp. 10–16.

Stokley, James. “Newest Ideas about Space and the Size of Everything.” Oakland Tribune, 28 April 1929, magazine section, p. 4.

Sutton, Ransome. “To Pioneer New Worlds.” Los Angeles Times, 2 January 1929, Annual Midwinter number, section 1, p. 29.

---. "What's New in Science." Los Angeles Times, 17 September 1933, Sunday magazine, p. 15.

Turner, Timothy G. “Griffith Park Includes 3761 Acres of Scenic Beauty.” Los Angeles Times, 2 May 1937, p. 6.

Young, Joseph E. “Helen Lundeberg: An American Independent.” Art International, September 1971, pp. 46–53.

This page has paths:

This page references: