Albert Einstein’s Religiosity



Albert Einstein’s Religiosity

J. Patrick Woolley

_________________________________

[T]he Jewish tradition … contains … something which finds splendid expression in many of the Psalms, namely, a sort of intoxicated joy and amazement at the beauty and grandeur of this world, of which man can just form a faint notion. It is a feeling from which true scientific research draws its spiritual sustenance, but which also seems to find expression in the song of birds.[1]

______________________________________________________________________________

Einstein’s biographer, Abraham Pais, writes, “I am sure that Einstein’s strongest source of identity, after science, was to be a Jew, increasingly as the years went by.” [2] While Einstein had turned away from a deep childhood commitment to traditional Jewish teachings when he was twelve, by 1911, as an adult, during his time in Prague, he seems to have begun to rejuvenate his Jewish roots. There, he would frequently discuss philosophy and science with other Jewish intellectuals in weekly meetings held at the home of Berta and Otto Fanta, laying a foundation for his future activism for Jewish causes. In 1919, Kurt Blumenfeld, a friend and Zionist leader, [3] recruited Einstein to help fundraise for the Hebrew University of Jerusalem. Nearly four decades later, in a 1955 letter (the year of his death), Einstein would thank Blumenfeld for having “helped me become aware of my Jewish soul.” [4]

As Einstein’s Jewish identity developed, so did his engagement with the public on matters religious. In the 1930s Einstein’s high-profile publications on religion begin. On 9 November 1930, in an article written expressly for the New York Times Magazine entitled “Religion and Science,” Einstein describes what he calls “cosmic religion” and what he sees as its relation to organized religions. An interview with Einstein by the title “Science and God: A Dialogue” ran in Forum and Century, [5] followed by “What I Believe” in the following edition, where Einstein shares his religious credo.[6] High-profile conversations with Tagore[7] on religion, music, truth and the nature of reality were published the next year. [8] As Einstein shared his perspectives on religion he also wrote on pacifism, Zionism, and issues related to ethics and a morally responsible state. In 1939, Einstein delivered “Our Goal” at a conference held at the Princeton Theological Seminary. Here, he discusses the limits of science and rationality with respect to the establishment of values and the role of religion in culture. This address became the first of two parts, the second delivered at the Conference on Science, Philosophy, and Religion in Their Relation to the Democratic Way of Life, in New York, 1940. Together, these addresses are now entitled “Science and Religion: Parts I and II.” [9] These and many other writings, as well as various letters from his archive, paint a rich and complex picture of “Einstein’s God.”

However, despite this long-term and highly public effort to express his own religiosity and his views on religion in general, there is a widespread misconception that Einstein was antagonistic to religious thought as a whole. This is not the case. It is true that Einstein does criticize aspects of religious thought as it is commonly interpreted or portrayed, arguing that “superstitious” tendencies often obscure what is most valuable in religious teachings. However, he also praises many aspects of religious traditions,[10] holds that institutionalized religions are meant to provide the moral grounding of a culture, and points to “religious geniuses” within these traditions who epitomize the highest form of religious understanding, his own “cosmic religion.”

Einstein has wrongly been characterized as an atheist—both by atheists who attempt to conscript him for their cause, as well as by religious leaders who consider his liberal views to be too removed from their orthodox teachings. But such claims do not take into account what Einstein actually says, for he clearly did not subscribe to atheism. In an explicit stance against atheism Einstein says, “In view of such harmony in the cosmos which I, with my limited human mind, am able to recognize, there are yet people who say there is no God. But what makes me really angry is that they quote me for support of such views.”[11] At another time, Einstein says that he is not an atheist because “the insufficiency of the human mind to understand deeply the harmony of the Universe is not in the atheistic mentality.”[12]

Why is there such a broad discrepancy between the way in which Einstein portrays himself and the ways he is commonly portrayed by others? Einstein’s philosophy of religion proves to be problematic for interpreters, in part, because they too often fail to consider the philosophical contexts in which Einstein’s statements were made. And these contexts present imposing challenges. Einstein’s views—both his approach to science and the way in which he expresses his religiosity—do not follow the path usually taken in science and religion dialogue today. This is largely due to the fact that Einstein’s highly technical epistemology[13] does not easily lend itself to supporting either supernaturalistic religious ideas or the blunt scientific materialism which often stands opposed to them. Einstein’s epistemology simply does not fit into this dichotomous framework. If we want to understand him better, not only do we need to consider very closely what Einstein says in his many writings, but we need to do so while remaining cognizant of the epistemological and historical contexts in which his philosophy of religion developed.

The Importance of Epistemological Contexts

Before we delve into Einstein’s religiosity, it is important to consider what is at the foundation of his approach to science. If we do not do this first, we can easily fall into the error of substituting one’s own beliefs about the nature, scope, and relevance of science for Einstein’s. Then, when we read what he says about how his own religiosity relates to science, we inadvertently try to match up these views of religion with our own views of science. This is a common error in much of the popular literature. It fails to understand the epistemological dimension of Einstein’s thinking as a whole.

When we do examine Einstein’s epistemology—his theory of knowledge, how we come to know the world—we see that he practices remarkable epistemic restraint.[14] He seeks to strip out all unsubstantiated beliefs about “what the world is like.” He limits his epistemology of science to only those elements which are indispensable for constructing scientific models of the world. This results in great epistemic clarity and depth.

In his most refined account of his epistemology, Einstein walks us back through all distinctions made in the mind, all the way back to a consciousness which precedes the distinction between an individual self and an independent world. He considers all thought in terms of “categories” or “schemes of thought.” Even the distinction between “subjective” and “objective” marks only a distinction between schemes; they are “factors” that condition sense experience by which we build scientific models to describe the world. Einstein writes:

We represent the sense-impressions as conditioned by an “objective” and by a “subjective” factor. For this conceptual distinction there ... is no logical-philosophical justification. But if we reject it, we cannot escape solipsism. It is also the presupposition of every kind of physical thinking. Here too, the only justification lies in its usefulness. We are here concerned with “categories” or schemes of thought, the selection of which is, in principle, entirely open to us and whose qualification can only be judged by the degree to which its use contributes to making the totality of the contents of consciousness “intelligible.” The above mentioned “objective factor” is the totality of such concepts and conceptual relations as are thought of as independent of experience, viz., of perceptions. So long as we move within the thus programmatically fixed sphere of thought we are thinking physically. Insofar as physical thinking justifies itself, in the more than once indicated sense, by its ability to grasp experiences intellectually, we regard it as “knowledge of the real.”

The theoretical attitude here advocated is distinct from that of Kant only by the fact that we do not conceive of the "categories" as unalterable ... but as ... free conventions.[15]

This subtle scheme which underpins, as Einstein puts it above, the “programmatically fixed sphere of thought [in which] we are thinking physically,” underlies all which, in Einstein’s view, is to be considered “scientific.” It demarcates a distinction between scientific and other forms of knowledge.

Only when we have this epistemological framework firmly in mind can we consider Einstein’s religious writings without falling into common errors. Now we can consider in what ways his religiosity relates to the “fixed sphere of thought [in which] we are thinking physically,” and in what ways it does not. We find that, whereas scientific knowledge is achieved by scrutinizing relations within that “fixed sphere,” religious thought for Einstein begins when one recognizes a limit within the sphere. Consciousness of this limit forms the foundation of Einstein’s “cosmic religion.”[16]

We see the significance of this limit most directly in Einstein’s “religious credo.” Here, he claims that beyond all the conceptual structures by which we know the world, one can become aware of a profound mystery to existence. He writes:

The fairest thing we can experience is the mysterious … A knowledge of the existence of something we cannot penetrate, of the manifestations of the profoundest reason and the most radiant beauty … which are only accessible to our reason in their most elementary forms … Enough for me the mystery of the eternity of life, and the inkling of the marvellous structure of reality, together with the single-hearted endeavour to comprehend a portion, be it ever so tiny, of the reason that manifests itself in nature. [17]

In this credo, just as in Einstein’s epistemology of science, we see an emphasis on the conceptual means, or “reason,” by which we come to know the world around us. Yet here the focus is on a “mystery” amidst it all, amidst “the reason that manifests itself in nature.”

The focus on the limits to scientific knowledge in Einstein’s religious credo should not be interpreted as a rejection or a turn away from scientific pursuits.[18] Quite the contrary. Religious consciousness results from turning directly into the depths and limits of what is scientifically known. One does not become aware of this “mystery” by rejecting scientific knowledge, but by embracing it—embracing its foundations, its scope, and ultimately its self-revealed logical limits.[19]

The awareness of this mystery, this limit to knowledge, is what engenders Einstein’s profoundly felt religiosity.[20] For instance, when asked if he believes in God, Einstein responds:

I am not an atheist. I do not know if I can define myself as a pantheist. The problem involved is too vast for our limited minds . . . The human mind, no matter how highly trained, cannot grasp the universe.[21]

And in a letter, he writes, “That deeply emotional conviction of the presence of a superior reasoning power, which is revealed in the incomprehensible universe, forms my idea of God.” [22]

The Importance of Historical Contexts

The excerpts above are a mere representative sampling of the stance that Einstein lays out. When we read the full body of his writings on religion, we see that his view presents philosophical challenges for traditional theists and atheists alike. To truly appreciate what is entailed in interpreting Einstein’s religious writings—the interwoven themes of “mystery,” “reason in nature,” “religious feeling,” the ethical dimension, as well as its many influences, including Judaism, Spinoza, Schopenhauer, and so on—one does need to go into considerable philosophical depth. And one needs to do so by becoming fully cognizant of the nineteenth- and early twentieth-century philosophical movements in which Einstein’s views developed, movements with which we are largely unfamiliar today.

For instance, the prominent role of ethics and the way it relates to other themes in Einstein’s religious thought cannot be well understood outside of the contexts of these movements in which traditional metaphysics frequently takes a backseat to meta-ethics. In this final part of our discussion, I will attempt to introduce some of that context by drawing from the work of three Einstein scholars—Gerald Holton, Max Jammer, and Don Howard—all philosophers and historians of physics who have broached the topic of Einstein’s religiosity in order to better interpret his philosophy of science. In this brief whirlwind tour, our objective is simply to gain a sense for the complexities entailed in any thorough reading of Einstein. Each scholar adds to our understanding by emphasizing Einstein’s own philosophical and historical contexts in different ways.

Gerald Holton begins by considering the significance of “transcendence” in Einstein’s thought.[23] He argues that over the course of Einstein’s adult life, as religious issues came increasingly to the forefront in his thinking, Einstein’s writings reveal an intensifying struggle to relate what Holton refers to as scientific and transcendental impulses. Holton stresses the importance of understanding this struggle from within the context of Einstein’s own historical period. It is a period in which German Idealists [24] had made their mark and the influence of Spinoza, the Enlightenment philosopher whom Einstein loved, was central. Spinoza had had a profound affect on the development of the liberal theologies that emerged in the wake of Kant’s Critiques. Debates concerning how God’s transcendence relates to knowledge of the immanent world resulted in transformative methodological shifts which continued to spur philosophical movements well into Einstein’s era. Holton underscores the importance of Einstein’s involvement in these movements when interpreting his views on religion as well as science.[25]

Max Jammer, similarly, stresses the importance of Spinoza.[26] And he too recognizes the importance of post-Kantian philosophies of religion.[27] But instead of pursuing an interpretive approach which focuses on nineteenth-century influences, he emphasizes the continuity of Einstein’s religiosity with his Jewish heritage. Jammer offers a reading of Einstein in which “mystery” is comparable to the medieval rationalism of Moses Maimonides. Here, the rational limits which prevent direct knowledge of God are constructive in that they reveal and clarify what it means to say that God transcends nature. In this reading of Einstein, speaking of God as a “mystery” amidst the “reason that manifests in nature” is perfectly in keeping with centuries of traditional apophatic[28] Jewish philosophy. It also suggests continuity with the much earlier Hellenistic figure, Philo, and the ancient doctrine of logos.[29]

Though Einstein did hold Maimonides in high regard,[30] Einstein was certainly no scholar of medieval or ancient thought, so we should be careful not to draw the parallels here too directly. Nonetheless, when we learn more about the intellectual climate in which Einstein was raised and the intellectual circles in which he participated, we see that there are highly influential post-Kantian philosophers through which Einstein would have been exposed to these ideas. Through the work of Hermann Cohen, a founder of neo-Kantianism and a leader of liberal Judaism,[31] and Edward Zeller, a leading nineteenth-century Greek scholar,[32] the ideas of Maimonides, Philo, and others, re-emerged in cultural consciousness, but were now presented in neo-Kantian and Hegelian terms rather than medieval and ancient ones. In these figures, we see many longstanding themes of religious philosophy—problems in defining God, the meaning of transcendence, the significance of the transcendent-immanent relationship, the relationship between scientific, religious, and ethical knowledge, and so on—newly recast into post-Kantian frameworks and systems.

Don Howard was the first to recognize the importance of post- and neo-Kantian contexts in interpreting Einstein. He stresses the formative influence of debates between neo-Kantians and logical positivists on Einstein’s epistemology of science.[33] In addition, for clues as to the source and reasons for Einstein’s “separation principle,”[34] Howard draws our attention to Einstein’s love of Arthur Schopenhauer, a crucial figure in his religious thought.[35] In Schopenhauer, we see many of the same themes of Einstein’s religiosity expressed. Schopenhauer’s religious thought culminates in an apophatic understanding of transcendence in much the way in which Einstein’s culminates in “mystery.” After “mystery,” a concern for ethics is the most prominent aspect of Einstein’s religious writings, as it is in Schopenhauer’s. And, like Einstein, Schopenhauer considered Spinoza to be a kindred soul.[36]

With Schopenhauer, we also unearth an important historical link to early twentieth-century liberal Judaism. Schopenhauer’s seminal critique of Kant helped to give rise to neo-Kantian philosophies like Cohen’s. Schopenhauer and Cohen each, like Kant, turn their focus away from metaphysics and toward meta-ethics for the purposes of establishing the formative principles of their religious philosophies. Within these post- and neo-Kantian contexts, we can trace out quite clearly how it is that Einstein’s religious views, while still being influenced by traditional religious systems, do not lend themselves easily to questions stemming from a supernaturalist metaphysics. And once we do so, we will not make the error of hastily assuming that, because the focus is not supernaturalist metaphysics, Einstein’s “cosmic religion” is merely romanticized scientism.

In summary, questions related to supernaturalist metaphysics, however that may be defined, are not those best suited to bring to interpretations of Einstein’s religiosity. Such questions undermine the way in which Einstein frames his statements on religion. They tend to overlook the subtleties involved in his epistemology of science and ignore the post-Kantian era in which his religious thinking developed. But, if not via questions of metaphysics, how should one proceed? Just as was for the case for Schopenhauer and Cohen, how we transition from an epistemology of science to an epistemology of ethics becomes the essential question to put to Einstein’s philosophy of religion.

Einstein emphasizes that religion, not science, is where a culture finds the source and expression for what should be, rather than what is. With ethics, we enter into a dimension of thought that does not fall within the scope of science, or, as Einstein puts it, within the “programmatically fixed sphere of thought [in which] we are thinking physically.” We are dealing with something of a completely different order than “sense-impressions” and scientific ways of thinking. Therefore, as we no longer live in a world driven by post- and neo-Kantian debates, perhaps the best way to enquire as to the relevance of Einstein’s religious views today, at least in non-academic circles, lies in questions related to, not metaphysics, but meta-ethics.

References

Calaprice, Alice. The Einstein Almanac. Baltimore, Md.: Johns Hopkins University Press, 2005.

Cohen, Hermann. Religion of Reason: Out of the Sources of Judaism. New York: F. Ungar Pub. Co., 1972.

Einstein, Albert. Ideas and Opinions. Edited by Carl Seelig. New York: Crown Publishers, 1982.

———. "Is There a Jewish Point of View?." In The World as I See It, x, 214. London: J. Lane, 1935.

———. "Moses Maimonides (1935)." In Out of My Later Years, vii, 282 p. New York: Philosophical Library, 1950.

———. "Out of My Later Years." vii, 282 p. New York: Philosophical Library, 1950.

———. "Religion and Science (1930)." In The World as I See It, x, 214. London: J. Lane, 1935.

———. "Remarks to the Essays Appearing in This Collective Volume." In Albert Einstein: Philosopher-Scientist, edited by Paul Arthur Schilpp, xviii, 781 p. La Salle, Ill.: Open Court, 1970.

———. "Science and God: A Dialogue." The Forum and Century 1930 373-79.

———. "What I Believe." The Forum and Century 1931, 193-94.

———. The World as I See It. London: J. Lane, 1935.

———. "The World as I See It (1931)." In The World as I See It, x, 214. London: J. Lane, 1935.

Einstein, Albert, Alice Calaprice, and Albert Einstein. The New Quotable Einstein. Enl. commemorative ed. Princeton, N.J.: Princeton University Press, 2005.

Holton, Gerald James. Victory and Vexation in Science: Einstein, Bohr, Heisenberg, and Others. Cambridge, Massachusetts: Harvard University Press, 2005.

Howard, Don. "Einstein's Philosophy of Science." In The Stanford Encyclopedia of Philosophy edited by Edward N. Zalta (ed.), URL = , 2008.

———. "A Peek Behind the Veil of Maya." In The Cosmos of Science: Essays of Exploration, xvi, 581. Pittsburgh, Pa., Konstanz: University of Pittsburgh Press; Universitèatsverlag Konstanz, 1997.

Jammer, Max. Einstein and Religion: Physics and Theology. Princeton, N.J.; Chichester: Princeton University Press, 1999.

Pais, Abraham. 'Subtle Is the Lord - ': The Science and the Life of Albert Einstein. Oxford: Oxford University Press, 1982.

Schilpp, Paul Arthur. Albert Einstein: Philosopher-Scientist. 3rd ed. Vol. 7, The Library of Living Philosophers. La Salle, Ill.: Open Court, 1970.

Schopenhauer, Arthur. The World as Will and Representation, Vol. I. Translated by E. F. J. Payne. New York: Dover Publications, 1969.

Tillich, Paul, and Robert C. Kimball. "Science and Theology: A Discussion with Einstein." In Theology of Culture. New York: Oxford University Press, 1959.

Viereck, George Sylvester. Glimpses of the Great. London: Duckworth, 1930.

Zeller, Eduard, Wilhelm Nestle, and Leonard Robert Palmer. Outlines of the History of Greek Philosophy. 13th ed. New York, London: Harcourt Brace and co.; K. Paul Trench Trubner & co., 1931.

-----------------------

[1] Albert Einstein, "Is There a Jewish Point of View?," in The World as I See It (London: J. Lane, 1935), vii-viii

[2] Abraham Pais, 'Subtle Is the Lord - ': The Science and the Life of Albert Einstein (Oxford: Oxford University Press, 1982), 314-15.

[3] The secretary general of the World Zionist Organization from 1911 to 1914 as well as secretary of the German Federation of Zionist from 1909-1911 and its president in the 1920’s

[4] Alice Calaprice, The Einstein Almanac (Baltimore, Md.: Johns Hopkins University Press, 2005), 56.

[5] Albert Einstein, "Science and God: A Dialogue," The Forum and Century 1930

[6] ———, "What I Believe," The Forum and Century 1931.

[7] Tagore, a Nobel Laureate, was one of the luminaries of the Brahmo Samaj. Founded in 1828 in Calcutta, the Brahmo Samaj hold that there is one God, as opposed to many.

[8] “The Nature of Reality.” Modern Review (Calcutta) 49 (1931): 42-43.

[9] Most of these pieces, and many more, can be found in Albert Einstein, The World as I See It (London: J. Lane, 1935)., Albert Einstein, Ideas and Opinions, ed. Carl Seelig (New York: Crown Publishers, 1982)., and Albert Einstein, "Out of My Later Years," (New York: Philosophical Library, 1950).

[10] Einstein refers to Judaism, Christianity, and Buddhism with high regard in several of his addresses.

[11] Max Jammer, Einstein and Religion: Physics and Theology (Princeton, N.J.; Chichester: Princeton University Press, 1999), 97. Quoting Prinz Hubertus zu Löwenstein, Toward the Further Shore, (Victor Gollancz, London, 1968), 156.

[12] Ibid., 121-22. Quoting Einstein to BF, 17 December 1952. Einstein Archive, reel 59-797.

[13] The theory of knowledge by which we come to justify claims about the world

[14] Einstein credits reading David Hume for helping him to recognize unsubstantiated assumptions about time while developing his special theory of relativity. Yet Einstein was not a Humean sceptic. He refers to himself as a “rationalist,” but he means this in a very specific sense which is dependent upon empiricism. See Albert Einstein, "Remarks to the Essays Appearing in This Collective Volume," in Albert Einstein: Philosopher-Scientist, ed. Paul Arthur Schilpp, The Library of Living Philosophers (La Salle, Ill.: Open Court, 1970), 680.

[15] Ibid., 673-74.

[16] It is “cosmic” in that it stems from knowledge of the cosmos. Einstein separates religious consciousness into three levels, or cultural stages, which build upon one another. He ascribes “cosmic religion” to the highest and rarest level. Just below it is a level defined by the primacy of ethical awareness and behaviour. The lowest level is where fear, “superstition,” and a poor sense of causal connection determine one’s religious beliefs. See ———, "Religion and Science (1930)," in The World as I See It (London: J. Lane, 1935).

[17] ———, "The World as I See It (1931)," in The World as I See It (London: J. Lane, 1935), 4-5.

[18] On the other hand, note that Einstein does not deify science or nature. Nor does he turn science toward scientistic ideologies or belief systems. The reliance on such beliefs would amount to no more than an intellectual crutch which would undercut the epistemic depth and clarity by which “mystery” becomes known.

[19] It is important that one not interpret Einstein’s “mystery” as a “God of the Gaps” scenario. It is mystery itself that is significant; we are not speaking of a mere “intellectual space” within which one might shelter belief systems from challenges emerging from the natural sciences.

[20] In other writings Einstein refers to this as a “cosmic religious feeling,” a profound sense of awe whose meaning, he says, is difficult to communicate to anyone who is without it. ———, "Religion and Science (1930)," in The World as I See It (London: J. Lane, 1935), 26.

[21] George Sylvester Viereck, Glimpses of the Great (London: Duckworth, 1930), 372-73.

[22] Albert Einstein, Alice Calaprice, and Albert Einstein, The New Quotable Einstein, Enl. commemorative ed. (Princeton, N.J.: Princeton University Press, 2005), 195-96. Quoting letter from Einstein to a banker in Colorado.

[23] See Gerald James Holton, Victory and Vexation in Science: Einstein, Bohr, Heisenberg, and Others (Cambridge, Massachusetts: Harvard University Press, 2005).

[24] e.g., Schleiermacher, Hegel, Schelling, Schopenhauer, and many others

[25] Holton highlights Paul Tillich’s work as the type of theology by which Einstein’s religious views can be made more explicit. He argues that an opportunity was lost when Einstein and Tillich, who attended some of the same conferences as Einstein, did not dialogue on religious philosophy. Tillich’s critique of Einstein’s religious views offers a promising place to begin science and religion dialogue. See Paul Tillich and Robert C. Kimball, "Science and Theology: A Discussion with Einstein," in Theology of Culture (New York: Oxford University Press, 1959), 130.

[26] See Jammer, Einstein and Religion: Physics and Theology.

[27] and of Tillich

[28] A method that spans ancient to modern religious philosophy which seeks to clarify the meaning of “transcendence” through the systematic application of negative statements, for example, by articulating why God cannot be described, why God is not temporal, why God is not finite, why God is not “a thing,” and so on.

[29] Philo held logos to be that which mediates the immanent world to a transcendent God. He is credited with being one of the first to attempt a synthesis of Jewish and Greek thought, two traditions to which Einstein refers with respect and admiration.

[30] See Albert Einstein, "Moses Maimonides (1935)," in Out of My Later Years (New York: Philosophical Library, 1950).

[31] See Hermann Cohen, Religion of Reason: Out of the Sources of Judaism (New York: F. Ungar Pub. Co., 1972).

[32] See Eduard Zeller, Wilhelm Nestle, and Leonard Robert Palmer, Outlines of the History of Greek Philosophy, 13th ed. (New York, London: Harcourt Brace and co.; K. Paul Trench Trubner & co., 1931).

[33] See Don Howard, "Einstein's Philosophy of Science," in The Stanford Encyclopedia of Philosophy ed. Edward N. Zalta (ed.) (2008).

[34] Einstein presents his separation principle as a condition necessary for the very possibility of framing empirical laws. He developed this principle as part of the E.P.R. Paradox, a thought experiment designed to counter arguments made in favour of certain interpretations of quantum mechanics. To account for the philosophical justification of this principle, Howard looks to Schopenhauer’s concept of the principium individuationis. The principium individuationis is the a priori principle of individuation in Schopenhauer’s philosophy that provides the foundation upon which to anchor intellectual knowledge to the sensory world so that scientific thinking becomes possible. See Don Howard, "A Peek Behind the Veil of Maya," in The Cosmos of Science: Essays of Exploration, Pittsburgh-Konstanz Series in the Philosophy and History of Science (Pittsburgh, Pa., Konstanz: University of Pittsburgh Press; Universitèatsverlag Konstanz, 1997). For a discussion of Einstein’s E.P.R. Paradox see, for example, Paul Arthur Schilpp, Albert Einstein: Philosopher-Scientist, 3rd ed., vol. 7, The Library of Living Philosophers (La Salle, Ill.: Open Court, 1970), 681-82..)

[35] Einstein cites Schopenhauer as a prime example of “cosmic religion.” Albert Einstein, "Religion and Science (1930)," in The World as I See It (London: J. Lane, 1935), 25.

[36] See Arthur Schopenhauer, The World as Will and Representation, Vol. I, trans. E. F. J. Payne (New York: Dover Publications, 1969), 179 and 422, footnote 2.

................
................

In order to avoid copyright disputes, this page is only a partial summary.

Google Online Preview   Download