Fact-checked by Grok 2 weeks ago

Surreal Number

Surreal numbers are a mathematical structure comprising a totally ordered proper class that extends the real numbers to include and quantities, forming the largest in which every real closed subfield can be embedded. Invented by British mathematician in the late 1960s while developing the theory of impartial games, surreal numbers were first detailed in his 1976 book On Numbers and Games. They gained public attention through computer scientist 's 1974 book Surreal Numbers: How Two Ex-Students Turned on to and Found Total Happiness, a fictional that presents their construction as a between characters decoding an ancient manuscript. The construction of surreal numbers proceeds recursively across "days," starting on day 0 with the number zero, denoted $0 = \{ \mid \}, where the empty left and right sets indicate no predecessors or successors. On subsequent days, new numbers are formed as x = \{ L \mid R \}, where L and R are sets of earlier surreals satisfying the condition that every element of L is less than every element of R, with equivalence under the "simplest" form ensuring uniqueness. This process yields all integers by day 2, dyadic rationals by finite days, and on day \omega (the first transfinite ordinal), it generates the entire real number line alongside infinitesimals like \varepsilon = \{ 0 \mid 1, \frac{1}{2}, \frac{1}{4}, \dots \} and infinities like \omega = \{ 0, 1, 2, \dots \mid \}. Arithmetic operations on surreal numbers—addition, , and more—are defined recursively to preserve the structure, with given by x + y = \{ L_x + y, x + L_y \mid R_x + y, x + R_y \} and similar forms for others, ensuring compatibility with operations. The surreals constitute a real closed field, meaning every positive surreal has a , every odd-degree has a , and the field is formally real (no equals zero except trivially). Beyond their algebraic properties, surreal numbers originated from Conway's analysis of combinatorial games, where each game position equates to a surreal number representing its value, enabling optimal play strategies in impartial games like . This connection bridges surreal numbers to , with applications in solving game-theoretic problems, though broader uses in analysis or physics remain exploratory.

History and Development

Origins in Combinatorial Game Theory

The conceptual foundations of surreal numbers emerged from early 20th-century analyses in , particularly the study of impartial games using set-theoretic tools and ordinal numbers. In 1913, applied to chess, demonstrating that in any finite, two-player game of without chance elements, either the first has a winning , the second has a winning , or both can force a draw; this proof relied on over the game's finite tree, implicitly ordering positions in a manner akin to ordinal structures. Building on this, Émile Borel explored strategic aspects of two-person games in the 1920s, including probabilistic elements and zero-sum formulations that highlighted the need for valuing positions quantitatively. advanced the field in 1928 with his , proving the existence of optimal mixed strategies in finite zero-sum games, while his concurrent work on von Neumann ordinals in provided a transfinite framework that later informed the hierarchical valuation of game positions. The 1930s saw further progress in impartial combinatorial games through the independent discoveries of Roland Sprague and Patrick Grundy. Sprague's 1936 paper analyzed sums of impartial games under the normal play convention—where the player making the last move wins—introducing a recursive assignment of values to positions based on the minimum excludant (mex) of options, effectively reducing complex games to equivalent heaps. Grundy formalized this in 1939, establishing the Sprague-Grundy theorem, which states that the (or ) of a of impartial games is the bitwise XOR of their individual nimbers, enabling the solution of diverse games via equivalents. These developments treated game positions as numerical values, motivating extensions beyond standard integers to capture finer distinctions in strategic advantage. Combinatorial game theory matured in the 1950s through the efforts of Richard Guy and Cedric Smith, who generalized the theory to partizan games (where players have distinct moves) and systematic sums thereof. Their 1956 paper introduced G-values—extensions of Grundy numbers to signed dyadic rationals—to evaluate positions in games like Dawson's Chess and Kayles, revealing that game outcomes could require fractional valuations not confined to non-negative integers, thus hinting at a broader numerical framework surpassing the reals in expressive power. For instance, in , a single heap of size n has Grundy number n, but sums yield XOR-based equivalents; in Dawson's Chess (a simplified on a chain of pawns), positions often evaluate to dyadic fractions like 1/2 or 3/2 under normal play, where small differences represent infinitesimal-like advantages in longer sums, marking the initial recognition of sub-unit strategic edges. The of these under the normal play convention—emphasizing symmetric rules and last-move —drove the quest for a unified number system capable of encoding both finite positional values and transfinite hierarchies, as recursive game lengths could exceed countable ordinals in theoretical extensions. This historical impetus from valuing simple impartial like , where infinitesimal distinctions first surfaced in fractional G-values, set the stage for later formalizations incorporating explicit and ordinals.

Conway's Formalization

developed the theory of surreal numbers in 1969 as part of his investigations into . This work formed a central component of his book On Numbers and Games, first published in 1976. Conway initially presented the surreal numbers through lectures delivered in 1970 at Cambridge University and the . The foundational insight behind surreal numbers was Conway's extension of the numerical values assigned to game positions—initially from impartial games—into a comprehensive class that incorporates all ordinal numbers, the real numbers, and infinitesimal quantities. This approach built briefly on prior advancements in from the mid-20th century. Conway's first formal definition represented each surreal number as a pair consisting of a left set and a right set of earlier-born surreal numbers, satisfying specific separation conditions. The construction proceeds recursively by "birthdays," with the number zero emerging as the inaugural element on day 0, possessing empty left and right sets. The name "surreal numbers" originated not from Conway but from Donald Knuth, who encountered the concept in 1974 and was struck by its dream-like, fantastical qualities, likening them to the surrealist art of Salvador Dalí. Knuth's enthusiasm led him to popularize the idea through his own book Surreal Numbers that same year. In his early exposition, Conway established that the surreal numbers constitute a totally ordered closed under , , and other operations, thereby forming a real closed field of characteristic zero that properly extends numbers. These proofs highlighted the surreal numbers' and their embedding of diverse mathematical structures within a unified framework.

Subsequent Extensions

Following Conway's foundational work, as popularized by in his 1974 book, subsequent developments in the 1970s focused on computational and representational aspects of surreal numbers. contributed significantly by exploring algorithmic methods for constructing and manipulating surreal numbers, emphasizing their tree-like hierarchical structure in a manner amenable to computer processing. This laid groundwork for later software realizations, including early prototypes that simulated the recursive birth of surreals to visualize their order and operations. In the and , efforts shifted toward axiomatic refinements and structural generalizations. Philip Ehrlich advanced the theory by developing an axiomatic framework that generalized 's construction to broader classes of ordered structures, establishing simplicity hierarchies as a core feature where each surreal's "birthday" corresponds to its complexity level in the inductive hierarchy. These hierarchies provide a canonical naming system via Conway names, enabling precise classification of surreals by their minimal representation in the construction process. Ehrlich's work demonstrated that such systems satisfy field axioms while incorporating transfinite and infinitesimal elements in a unified manner. Post-2000 research has connected surreal numbers to non-standard , showing that the surreal admits relational extensions isomorphic to hyperreal models, allowing infinitesimals to model non-Archimedean geometries consistently. Ehrlich further explored p-adic analogues within surreal frameworks, constructing non-Archimedean completions that embed p-adic numbers into surreal extensions while preserving valuation properties and structure. These developments enable surreal-based models for ultrametric spaces, bridging classical p-adic with transfinite orders. The concept has evolved into surreal analysis, incorporating advanced operations like and . Lou van den Dries and Philip Ehrlich established that the surreal exponential field shares first-order properties with the real field, supporting analytic functions over the entire surreal domain. This facilitates surreal interpretations of exponential towers and , where power towers of ordinals embed order-preservingly into surreals, allowing evaluation of expressions like \omega^{\omega^{\omega}} via hierarchical simplification. Transseries expansions, embeddable in surreals, further extend this to model superexponential growth in . More recently, in 2023, Matthias Aschenbrenner, Lou van den Dries, and Joris van der Hoeven proved that the surreal numbers constitute a universal H-field, unifying various ordered exponential structures.

Definition and Construction

Recursive Definition

The surreal numbers form a proper class constructed recursively via a transfinite induction process known as the "birth" of numbers on ordinal days, independent of any preexisting number system or arithmetic operations. This construction begins with the empty class S_{-1} = \emptyset, ensuring no presupposed structure beyond set theory. On each ordinal day \alpha, the surreals born on that day, denoted S_\alpha, consist of all possible forms \{ L \mid R \} where L and R are subsets of the earlier-born surreals \bigcup_{\beta < \alpha} S_\beta, subject to the condition that no element l \in L satisfies l \geq r for any r \in R (with the order relation defined recursively). The full class of surreal numbers is then the union \mathrm{No} = \bigcup_{\alpha} S_\alpha over all ordinals \alpha. The process commences on day 0 with the earliest surreal: $0 = \{ \mid \}, using empty left and right option sets. On day 1, utilizing the single previous surreal 0, the integers and are born as $1 = \{ 0 \mid \} and -1 = \{ \mid 0 \}. Subsequent finite days yield further integers and dyadic rationals; for instance, on day 2, $2 = \{ 1 \mid \}, -2 = \{ \mid -1 \}, and the dyadic \frac{1}{2} = \{ 0 \mid 1 \}. Positive integers arise recursively as n+1 = \{ n \mid \}, while negative integers follow symmetrically as -(n+1) = \{ \mid -n \}. Transfinite days introduce infinitesimals and infinities. The first positive infinitesimal, often denoted \varepsilon or \omega^{-1}, is born on day \omega as \varepsilon = \{ 0 \mid 1, \frac{1}{2}, \frac{1}{4}, \dots \}, where the right set comprises all positive dyadic rationals born on finite days; this \varepsilon satisfies $0 < \varepsilon < \delta for every positive dyadic rational \delta. Similarly, the first infinity \omega = \{ 0, 1, 2, \dots \mid \} emerges on day \omega, with the left set containing all finite integers. Further dyadics, such as \frac{3}{4} = \{ \frac{1}{2} \mid 1 \}, appear on later finite days. To establish completeness, a proof sketch proceeds by transfinite recursion: assume all surreals born before day \alpha have been constructed; then S_\alpha exhausts all valid forms from those predecessors, as the subsets L and R are well-defined by the axiom of choice and the well-ordering of ordinals prevents infinite descent in the recursive dependencies. Since every surreal form relies only on finitely many "ancestors" in its construction tree (via the finite support in normal forms, though not detailed here), and ordinals cofinally cover all possible recursion depths, every surreal number is born on some ordinal day \alpha. This ensures the class \mathrm{No} is closed under the recursive definition without gaps.

Normal Forms and Equivalence

Surreal numbers are typically represented in the form \{L \mid R\}, where L and R are sets of previously constructed surreal numbers satisfying the condition that every element of L is less than every element of R; however, multiple such forms may denote the same surreal number, necessitating a canonical representation for unique identification. The Conway normal form provides this standardization, expressing every surreal number x uniquely as x = \sum_{\beta < \alpha} \omega^{y_\beta} r_\beta, where \alpha is an ordinal, the exponents y_\beta form a strictly decreasing sequence of surreal numbers, and each r_\beta is a nonzero real number with |r_\beta| < 2. This form generalizes the for ordinals by incorporating real coefficients and extending to negative exponents for infinitesimal components, with the "finite part" captured by terms where the exponents are real numbers. Equivalence between two forms \{L \mid R\} and \{L' \mid R'\} holds if they represent the same surreal number, meaning no element of L is greater than or equal to any element of R' and no element of L' is greater than or equal to any element of R, ensuring the forms satisfy mutual order preservation across their options; more formally, \{L \mid R\} = \{L' \mid R'\} if \{L \mid R\} \geq \{L' \mid R'\} and \{L' \mid R'\} \geq \{L \mid R\}, where x \geq y if no right option of x is less than or equal to y and no left option of y is greater than or equal to x. To canonicalize a form, redundant options are simplified by removing any l \in L such that there exists l' \in L with l' > l and l' < r for all r \in R, or symmetrically for R, iteratively until the form is in simplest terms where no further reductions apply; this process leverages the simplicity rule, selecting the earliest-born (lowest birthday) number between the greatest left option and least right option as the representative. For example, the form \{0 \mid 1\} represents $1/2, and it is equivalent to \{0, 1/4 \mid 3/4, 1\} because the additional options $1/4 and $3/4 do not alter the bounding interval or the simplest mediant, which remains $1/2; upon simplification, both reduce to the unique normal form \{0 \mid 1\}. A fundamental theorem states that every surreal number possesses a unique normal form, ensuring that distinct representations converge to a single canonical expression under equivalence, as proven via transfinite induction on the construction process.

Induction and Birthday

The birthday of a surreal number x, denoted b(x), represents the earliest "day" in the recursive construction process on which x appears, where days are indexed by ordinals. This function measures the complexity of x relative to previously constructed numbers and induces a well-ordering on the class of all surreal numbers. Formally, for a surreal number x = \{ L \mid R \}, where L and R are sets of previously born surreal numbers with no member of L greater than or equal to any in R, the birthday is given by b(x) = \sup \{ b(z) + 1 \mid z \in L \cup R \}, with the base case b(0) = b(\{ \emptyset \mid \emptyset \}) = 0 (supremum over the empty set defined as 0). This ordinal-valued function establishes a hierarchy of simplicity among surreal numbers, where x is simpler than y if b(x) < b(y). All real numbers possess finite birthdays, reflecting their construction from finite iterations of the recursive process starting from 0; for instance, the integers and dyadic rationals emerge on successive finite days. Infinite surreal numbers like \omega = \{ 0,1,2,\dots \mid \emptyset \} have birthday \omega, marking the first transfinite day, while the positive infinitesimal \omega^{-1} = \{ 0 \mid 1, 2, 3, \dots \mid \} (simplifying appropriately with all positive integers, but equivalently with dyadics) also arises on day \omega. Specific examples include b(1) = 1 for $1 = \{ 0 \mid \emptyset \} and b(\omega) = \omega. The birthday function underpins the principle of transfinite induction for surreal numbers: for any property P satisfied by all surreal numbers, if P(y) holds for every y < x whenever it holds for all proper options of those y, then P(x) holds. More precisely, the induction rule states that if P holds for the empty set and, assuming P holds for all surreals born before day \alpha, it holds for all surreals born on day \alpha, then P holds for all surreal numbers. This follows directly from the birth order, as the ordinals indexing birthdays form a well-ordered class, ensuring that every nonempty subclass of surreals has a least element with minimal birthday, to which the inductive hypothesis applies without circularity.

Order and Topology

Total Order

The total order on the class of surreal numbers is defined recursively, building on their construction as forms x = \{ X^L \mid X^R \}, where X^L and X^R are sets of earlier-born surreal numbers with no element of X^L greater than or equal to any element of X^R. The basic relation x < y holds if x \in X_y^L (i.e., x is a left option of y) or y \in X_x^R (i.e., y is a right option of x); this "simplest form" is then extended recursively to all pairs by induction on their birthdays, ensuring x < y if every right option of x is less than y and every left option of y is greater than x. To establish totality, consider any two distinct surreal numbers x and y with birthdays \alpha and \beta, assuming without loss of generality that \alpha \leq \beta. By the recursive construction and the no-overlap condition (no left option exceeds a right option), the options of x and y cannot interleave in a way that prevents ordering; specifically, if neither x < y nor y < x holds via direct option membership, induction on the maximum birthday shows that all options of the earlier-born number lie entirely to one side of the later-born one, forcing exactly one of x < y or y < x. This yields the trichotomy property: for any x, y, precisely one of x < y, x = y, or y < x obtains, where equality means x \not< y and y \not< x. Transitivity follows similarly by induction: if x \leq y and y \leq z, then the options of x are bounded above by those of y, which are bounded above by those of z, ensuring x \leq z without violations of the no-overlap axiom. The rational numbers embed densely within the surreals via finite-birthday constructions, such as integers built day-by-day (e.g., $0 = \{ \mid \}, $1 = \{ 0 \mid \}) and rationals like \frac{1}{2} = \{ 0 \mid 1 \}; the reals embed at birthday \omega through Dedekind-like cuts of rationals. Ordinals up to the class \Omega (the largest surreal) embed as "purely infinite" surreals, such as \omega = \{ 0, 1, 2, \dots \mid \}. For illustration, consider \frac{1}{\omega} = \{ 0 \mid 1, \frac{1}{2}, \frac{1}{3}, \dots \}, $0 = \{ \mid \}, and \omega = \{ 0,1,2,\dots \mid \}: here, $0 < \frac{1}{\omega} since 0 is a left option of \frac{1}{\omega}; moreover, \frac{1}{\omega} is less than every positive rational, filling an infinitesimal gap above 0. Similarly, $0 < \omega as 0 is among \omega's left options, creating a structure with infinitesimal separations beyond the reals.

Gaps and Dedekind Cuts

The field of surreal numbers is real-closed, meaning every non-negative surreal has a square root and every odd-degree polynomial equation with surreal coefficients has a root in the surreals. Unlike the real numbers, however, the surreals lack Dedekind completeness: not every nonempty class of surreals that is bounded above has a least upper bound in the surreals. This incompleteness manifests as gaps in the order, particularly for Dedekind cuts involving proper classes, contrasting with the reals' completeness where every bounded subset has a supremum. A Dedekind cut in the surreals is defined as a partition of the class of all surreal numbers into two proper classes L and R such that every element of L is less than every element of R, L has no maximum element, and R has no minimum element. The recursive construction of the surreals ensures that every such cut where L and R are sets (born earlier in the transfinite hierarchy) is realized by a surreal born later, filling more cuts than the do over the . For instance, there is no largest finite surreal number, creating a gap between the class of all finite surreals and the infinites; this gap is filled by the surreal \omega = \{ n \mid \}, where n ranges over all finite surreals and the right set is empty. Topologically, the surreals form a linear order under their natural total order, inducing an order topology that is dense in local segments (such as the reals embedded within) but globally discontinuous due to these gaps at transfinite levels. Subsets born before a given ordinal are dense in their induced order, but the full class lacks the least upper bound property for arbitrary bounded classes. A fundamental theorem states that the surreals realize all over any of their proper subclasses ordered by birthday, ensuring the hierarchy progressively fills gaps in earlier structures.

Transfinite Induction

Transfinite induction on relies on the birthday function b(x), which assigns to each surreal number x the smallest ordinal \alpha at which x appears in the recursive construction of the class \mathrm{No}. The formal principle states: a property P holds for all x \in \mathrm{No} if whenever P(y) is true for all y with b(y) < b(x), then P(x) follows. This induction leverages the well-ordered structure of the ordinals governing birthdays, ensuring the recursive definition propagates properties across the entire class. The standard proof technique for such implications proceeds by contradiction. Suppose there exists some z \in \mathrm{No} such that \neg P(z); among all counterexamples, select z with minimal b(z). The left and right options of z all have strictly smaller birthdays, so by the induction hypothesis, P holds for them. Since the surreal construction and the property P are defined recursively from these options, this forces P(z) to hold, yielding a contradiction. Thus, no counterexamples exist. This method exploits the minimal-birthday property inherent to the ordinal-indexed construction. Key applications include establishing the transitivity of the surreal order: for all a, b, c \in \mathrm{No}, if a < b and b < c, then a < c. The proof inducts on birthdays, verifying the recursive comparison rule \neg (a^R \leq b^L) and \neg (b^R \leq c^L) imply \neg (a^R \leq c^L). Similarly, it proves arithmetic compatibility with the order, such as: if a < b, then a + d < b + d for any d \in \mathrm{No}, by inducting on the birthdays of a, b, d and using the recursive addition formula to preserve inequalities in options. A representative example is the proof that x < x + 1 for every x \in \mathrm{No}, where $1 = \{0 \mid \}. Proceed by induction on b(x): the base case x = 0 gives $0 < 1, as $0^L = \emptyset \not\leq 1^L = \emptyset fails vacuously while the right-side condition holds. Assuming the inequality for all y with b(y) < b(x), the form of x + 1 inherits options from x and $1, and the induction hypothesis on those options ensures no right option of x is \leq any left option of $1, confirming x < x + 1. In contrast to transfinite induction on ordinals, which builds only nonnegative structures in a one-directional manner, the surreal version accommodates bidirectional extension: the symmetric recursive construction generates negative numbers alongside positives from the outset, allowing induction to verify properties uniformly across the signed order without separate treatments for directions.

Arithmetic Operations

Addition and Subtraction

Addition in the surreal numbers is defined recursively using the options of the operands. For surreal numbers x = \{ X^L \mid X^R \} and y = \{ Y^L \mid Y^R \}, where X^L and X^R are the left and right options of x, and similarly for y, the sum is given by x + y = \{ x^L + y, x + y^L \mid x^R + y, x + y^R \}. This construction ensures that addition is well-defined for all earlier-born surreal numbers and extends the usual addition of real numbers. Subtraction is derived from addition and negation. The additive inverse of a surreal number x = \{ X^L \mid X^R \} is -x = \{ -X^R \mid -X^L \}, which swaps the left and right options with negation applied recursively. Then, x - y = x + (-y). This definition preserves the field's structure and aligns with real subtraction when restricted to the reals. The operations satisfy key algebraic properties, proven via transfinite induction on the birthdays of the numbers involved. Addition is commutative, so x + y = y + x for all surreal numbers x and y; it is also associative, with (x + y) + z = x + (y + z). Moreover, addition is strictly order-preserving: if x < y, then x + z < y + z for any z, and similarly for \leq. These properties ensure that the surreals form an ordered group under addition, extending the ordered additive group of the reals. Illustrative examples highlight how surreal addition behaves differently from real addition in the presence of infinities and infinitesimals. For instance, adding the real number 1 to the infinite surreal \omega = \{ 0, 1, 2, \dots \mid \} yields $1 + \omega = \omega, as the finite increment is absorbed by the infinite magnitude. In contrast, adding the positive infinitesimal \varepsilon = \{ 0 \mid 1, \frac{1}{2}, \frac{1}{4}, \dots \} to 1 gives $1 + \varepsilon > 1, showing that infinitesimals are not absorbed by finites in this system—unlike in some non-standard analyses. These examples demonstrate the compatibility of surreal addition with the while introducing novel behaviors beyond the reals.

Multiplication

The multiplication of two surreal numbers x = \{x^L \mid x^R\} and y = \{y^L \mid y^R\} is defined recursively as x \cdot y = \left\{ x^L \cdot y + x \cdot y^L - x^L \cdot y^L, \, x^R \cdot y + x \cdot y^R - x^R \cdot y^R \;\middle|\; x^L \cdot y + x \cdot y^R - x^L \cdot y^R, \, x^R \cdot y + x \cdot y^L - x^R \cdot y^L \right\}, where the left options consist of the first two sets and the right options consist of the last two sets. This construction ensures the product is a valid surreal number by on the birthdays of x and y, with all terms involving multiplications and additions of "simpler" (earlier-born) surreals. Multiplication distributes over addition, satisfying x \cdot (y + z) = x \cdot y + x \cdot z for all surreals x, y, z. This property follows from verifying the definition of the product against the recursive , using to show that the left and right sets of the triple product match those of the sum of pairwise products. Similarly, multiplication preserves in the positive direction: if x < y and $0 < z, then x \cdot z < y \cdot z. The proof proceeds by , confirming that the options of x \cdot z and y \cdot z align with the order via the monotonicity of addition and the base cases for earlier surreals. The sign of the product adheres to standard rules: x \cdot y > 0 if both x > 0 and y > 0 or both x < 0 and y < 0; x \cdot y < 0 otherwise; and x \cdot y = 0 if at least one factor is zero. This positivity preservation for positive factors is established inductively from the order properties and the structure of the options, ensuring no overlap between positive and non-positive products. Representative examples illustrate these operations. The infinite surreal \omega = \{0, 1, 2, \dots \mid \} satisfies \omega \cdot 2 = \omega + \omega, where $2 = \{1 \mid 3\}, reflecting the ordinal-like scaling in the positive direction. Additionally, \omega \cdot (1/\omega) = 1, confirming the multiplicative inverse property for \omega, and similarly \omega \cdot \varepsilon = 1 since \varepsilon = 1/\omega, showing that the product of an infinity and its reciprocal yields a finite number. In John H. Conway's original development, surreal numbers emerged from the analysis of combinatorial games, where multiplication captures the more intricate disjunctive combination of game positions compared to the simpler additive sum, leading to the involved recursive definition.

Division and Consistency

Division in the surreal numbers is defined for nonzero y as x / y = x \cdot (1/y), where the reciprocal $1/y is the unique surreal number z such that y \cdot z = 1. The construction of the reciprocal relies on the recursive definition using the left and right options of y: specifically, $1/y = \{ L' \mid R' \}, where the sets L' and R' are formed from the reciprocals of the options of y, ensuring z is the simplest number satisfying the multiplication equation. This definition extends the arithmetic operations to include division, completing the field structure. The consistency theorem asserts that all arithmetic operations on surreal numbers, including division, produce results independent of the specific representation of the operands as forms \{ L \mid R \}. That is, if two surreal numbers are equal but expressed differently, their sum, product, quotient, or other operation yields the same result. The proof proceeds by transfinite induction on the birthdays of the numbers involved: assuming consistency holds for all numbers born earlier, one verifies it for numbers born on a given day by showing that alternative representations lead to equivalent forms via the simplicity criterion. This induction leverages the recursive construction of surreals, ensuring operations are well-defined across the entire class. For example, dividing the infinite ordinal \omega by 2 yields \omega / 2, a dyadic surreal number born on day \omega + 1, which can be represented as \{ n \mid \omega - n \} for finite n, but simplifies under the equivalence relation to the unique form satisfying $2 \cdot (\omega / 2) = \omega. Similarly, $1 / \omega is an infinitesimal surreal number, given by \{ 0 \mid 1/n \mid n < \omega \}, which multiplies with \omega to produce 1, illustrating how division captures magnitudes smaller than any positive real. The surreal numbers satisfy all field axioms, forming a totally ordered field under addition, multiplication, and division, with additive and multiplicative identities and inverses for nonzero elements. Moreover, they constitute a real-closed field, meaning every positive surreal has a square root, every odd-degree polynomial with surreal coefficients has a surreal root, and the ordering is compatible with the field operations. This structure partitions the surreals into archimedean classes, where two numbers are equivalent if neither is infinitely larger than the other, providing a hierarchy of infinitesimal, finite (isomorphic to ), and infinite scales.

Advanced Algebraic Structures

Negation and Closure Properties

The negation of a surreal number x = \{ L \mid R \}, where L and R are the left and right option sets, is defined recursively by -x = \{ -r \mid r \in R \} \mid \{ -l \mid l \in L \}. This construction swaps and negates the option sets, ensuring that negation is well-defined within the class of surreal numbers. In particular, the zero surreal $0 = \{ \mid \} satisfies -0 = 0, as its empty option sets remain empty after negation. The order on surreal numbers is reversed under negation: for distinct surreals x and y, x < y if and only if -x > -y. This follows from the definition, as the left options of -x become the negated right options of x, which are greater than those of y, and similarly for the right options. The (earliest construction stage) of -x equals that of x, preserving simplicity under this . The class S of all surreal numbers is closed under negation, addition, multiplication, and division by nonzero elements, forming a real closed field. Starting from the empty surreal $0, repeated application of these operations generates increasingly complex surreals: for instance, integers arise from additions and negations of $1 = \{ 0 \mid \}, and further operations yield dyadic rationals, dense subsets approximating reals, and eventually the full class S. This closure ensures that any finite arithmetic expression in surreals produces another surreal, with no elements escaping the class. A key result is the arithmetic closure theorem: for any ordinal \alpha, every surreal born by stage \alpha (i.e., with birthday < \alpha) that appears in an arithmetic expression is mapped to a surreal born by some later stage \beta > \alpha. This is proved by on the birthday and the expression's structure. preserves the birthday exactly, of x and y (with birthdays \gamma_x and \gamma_y) yields a result with birthday \gamma_x + \gamma_y, and follows a recursive bound f(\gamma_x, \gamma_y) satisfying f(n, m) = f(n, m-1) + f(n-1, m) + f(n-1, m-1) + 1 for positive finite arguments, extended ordinally. , being with , follows similarly. Since the inverse of a nonzero surreal exists uniquely within S, division is closed, though specific birthday bounds for division are more involved. These birthday bounds for , , , and confirm that operations only increase complexity by a controlled ordinal amount, generating all surreals iteratively from simpler precursors.

Powers of Omega

The surreal number \omega is defined as \{0, 1, 2, \dots \mid \}, making it the simplest infinite surreal number greater than every finite surreal. This representation positions \omega as the first transfinite ordinal within the surreal class, with no right options, ensuring its infinitude. Powers of \omega are constructed recursively for any surreal exponent y. Specifically, \omega^y is the simplest surreal satisfying $0 < \omega^y and, for positive reals r, s, r \cdot \omega^{y_L} < \omega^y < s \cdot \omega^{y_R}, where y_L and y_R are typical left and right options of y. For sums of exponents, \omega^{x_1 + x_2 + \dots + x_n} = \omega^{x_1} \cdot \omega^{x_2} \cdot \dots \cdot \omega^{x_n}, with the ordering of exponents preserved in the product via surreal multiplication. This recursive structure extends the ordinal exponentiation familiar from set theory into the surreal domain. A key example is \omega^\omega = \{\omega^n \mid n < \omega\}, the simplest surreal larger than all finite powers \omega^n for natural numbers n. For negative exponents, \omega^{-1} = 1/\omega, an infinitesimal positive surreal smaller than every positive finite number. Fractional powers also arise naturally, such as \omega^{1/2}, the surreal square root of \omega, which satisfies $0 < \omega^{1/2} < \omega and (\omega^{1/2})^2 = \omega. Every surreal number admits a unique representation in Conway normal form: x = \sum_{i=1}^k c_i \omega^{y_i}, where the exponents satisfy y_1 > y_2 > \dots > y_k, the coefficients c_i are nonzero integers, and k is finite. This theorem ensures a expansion analogous to normal form for ordinals but extended to the full surreal field, with uniqueness guaranteed by the strict decreasing order of exponents and the properties of surreal addition and multiplication.

Exponential Function

The exponential function on surreal numbers, denoted \exp: \mathrm{No} \to \mathrm{No}, is defined by transfinite recursion along the birthday ordering of the surreals, extending the classical real exponential while preserving its core analytic properties within the surreal framework. This construction, originally conceived by Martin Kruskal and formalized by Harry Gonshor, ensures the function is well-defined for every surreal number by inducting on the day of birth \alpha of each x \in \mathrm{No}. For x born on day \alpha, \exp(x) is the earliest surreal simplifying the power series \sum_{n=0}^\infty x^n / n! in the sense of surreal approximation, avoiding later-born terms. The precise inductive definition for a surreal x = \{ x_L \mid x_R \} with left options x_L and right options x_R is given by \exp(x) = \left\{ 0,\ \exp(x_L) \cdot [x - x_L]_n,\ \exp(x_R) \cdot [x - x_R]_{2n+1} \ \middle|\ \exp(x_L) \cdot [x_L - x]_{2n+1},\ \exp(x_R) \cdot [x_R - x]_n \right\}, where the index n ranges over the natural numbers, and _k = \sum_{i=0}^k y^i / i! denotes the k-th partial sum of the exponential series, omitting any terms that would yield non-positive values on the appropriate side to maintain the order. This recursion leverages the surreal construction process, ensuring convergence in the surreal order by including only earlier-born approximations. This definition yields a surjective ordered group isomorphism from (\mathrm{No}, +) to (\mathrm{No}_{>0}, \cdot), with \exp(x) = e^x for all real x. In particular, the functional equation \exp(x + y) = \exp(x) \cdot \exp(y) holds for all x, y \in \mathrm{No}, established by transfinite induction on the pair of birthdays of x and y: the base cases align with the real exponential, and the inductive step verifies the equation using the recursive options of x + y relative to those of x and y. For infinitesimals \epsilon \prec 1, the definition reduces exactly to the power series \exp(\epsilon) = \sum_{n=0}^\infty \epsilon^n / n!. Examples include \exp(0) = 1 and \exp(\omega) > \omega^n for every finite natural number n, reflecting the rapid growth beyond polynomial scales in \omega. The exponential function is strictly order-preserving: x < y implies \exp(x) < \exp(y), as verified inductively from the real case and the positive increments in the recursive options. It is also continuous with respect to the order topology on \mathrm{No}, where basic open intervals are preserved under the isomorphism, ensuring limits of surreal sequences map appropriately. These properties confirm \exp as a natural extension of real analysis to the full surreal class.

Extensions and Generalizations

Surcomplex Numbers

Surcomplex numbers extend the field of surreal numbers \mathrm{No} by adjoining an imaginary unit i satisfying i^2 = -1, forming the field \mathrm{No}. This construction is analogous to the complex numbers as an extension of the reals, and surcomplex numbers can be represented concretely as ordered pairs (x, y) where x, y \in \mathrm{No}, identified with the formal expression x + y i. Addition in the surcomplex numbers is defined componentwise: for z_1 = x_1 + y_1 i and z_2 = x_2 + y_2 i, z_1 + z_2 = (x_1 + x_2) + (y_1 + y_2) i, leveraging the existing addition in \mathrm{No}. Multiplication follows the standard complex rule, incorporating the surreal multiplication: z_1 z_2 = (x_1 x_2 - y_1 y_2) + (x_1 y_2 + x_2 y_1) i. These operations make \mathrm{No} a field, with the surreals embedding naturally as the subfield of elements with y = 0, and it contains a canonical copy of the complex numbers \mathbb{C}. Unlike \mathrm{No}, which is real-closed, the surcomplex field is algebraically closed. The surcomplex numbers lack a total order compatible with their field structure, as adjoining i disrupts the linear ordering of \mathrm{No}. However, partial orders can be defined, such as one induced by the real part (ordering via x in x + y i) or by the modulus squared |z|^2 = x^2 + y^2, which is well-defined since \mathrm{No} admits squares and is real-closed. Examples include the Gaussian integers within the surcomplex numbers, formed by pairs (m, n) where m, n are surreal integers (the subring \mathbb{Z} \subset \mathrm{No}), generalizing the classical Gaussian integers \mathbb{Z}. A surreal analog of Euler's formula arises via the canonical exponential function on \mathrm{No}, such as e^{i \omega}, where \omega is the first infinite surreal ordinal, extending the periodicity of the complex exponential with kernel $2\pi i \mathbb{Z}, now involving surreal integers.

Infinitesimals and Infinities

In surreal numbers, infinitesimals are positive elements \epsilon such that $0 < \epsilon < 1/n for every positive integer n, making them smaller than any positive real number yet greater than zero. A canonical example is \epsilon = 1/\omega = \{0 \mid 1, 1/2, 1/4, \dots \}, the reciprocal of the first infinite surreal number, born on or before day \omega in the constructive hierarchy. These infinitesimals arise naturally from the recursive definition, filling gaps between existing surreals, and they satisfy field properties such as \epsilon \cdot \omega = 1. Infinite surreal numbers, conversely, are those x where |x| > n for every positive integer n, extending beyond the reals in both directions. The simplest positive infinity is \omega = \{0, 1, 2, 3, \dots \mid \}, greater than all finite numbers and born on day \omega, with its negative counterpart -\omega = \{\mid 0, -1, -2, \dots \}. Larger infinities include \omega^\omega, constructed as \{\omega^n \mid n < \omega\}, which exceeds \omega raised to any finite power and represents a higher order of magnitude. The class S_\omega, comprising all surreal numbers born by day \omega, encompasses the entire real line—constructed via infinite descending chains of dyadic rationals, such as $1/3 = \{0, 1/4, 5/16, \dots \mid 1, 1/2, 3/8, \dots \}—along with the initial infinitesimals like \epsilon and infinities like \pm \omega. This stage marks the first inclusion of non-real extremes, unifying finite precision with transfinite and infinitesimal scales in a single ordered field. Monads in the surreal numbers refer to infinitesimal neighborhoods around points, particularly the monad at 0, which consists of all surreals x satisfying |x| < \delta for some \delta > 0, populated exclusively by and excluding nonzero reals. For instance, dividing an by an , such as \epsilon / \omega = 1/\omega^2 = \{0 \mid \epsilon\}, yields another smaller than \epsilon but still within the monad at 0, preserving the structure's without collapsing to zero. Such operations highlight how and interact additively and multiplicatively while maintaining the . Normal forms provide a canonical representation for these numbers, expressing them as sums of terms involving powers of \omega with real coefficients.

Alternative Realizations

One alternative realization of surreal numbers employs sign expansions, introduced by Gonshor, where each surreal number x is represented as a function from an ordinal \alpha to the set \{+, -\}, or equivalently as a formal series x = \sum_{\beta < \alpha} s_\beta \omega^\beta with coefficients s_\beta \in \{ -1, 0, +1 \}, where the sum converges in the surreal topology and the support is well-ordered. This representation leverages the base-\omega expansion but restricts coefficients to signs, enabling a direct encoding of the simplicity hierarchy via the length of the sequence. There exists a bijection between these sign sequences and the Conway normal forms, preserving the order and arithmetic operations, as established by the canonical isomorphism theorem in Gonshor's framework. For example, the surreal number $1/2, born on day 2 in the Conway construction, corresponds to the sign sequence + - , which alternates signs to position it midway between 0 and 1 in the lexicographic order on sequences. This finite expansion highlights how dyadic rationals arise from short, alternating patterns. Another realization embeds surreal numbers into Hahn series fields, specifically the field \mathbb{R}((t^{\mathrm{No}})) of generalized power series \sum r_\gamma t^\gamma with r_\gamma \in \mathbb{R} \setminus \{0\}, exponents \gamma \in \mathrm{No}, and well-ordered support under the reverse lexicographic order on supports. The surreal field \mathrm{No} is isomorphic to this Hahn field via the Conway normal form, where each surreal admits a unique such expansion with decreasing exponents, providing a valuation-theoretic perspective on infinitesimals and infinities. An approach characterizes the surreals as the unique (up to ) real-closed equipped with a birth b: \mathrm{No} \to \mathrm{On} (the ordinals) that is surjective and satisfies the Conway inductive construction , ensuring every surreal is born as the simplest number between two earlier-born sets. This formulation, developed by Ehrlich, abstracts away recursive definitions while preserving the class-sized nature and simplicity hierarchy of \mathrm{No}.

Applications

Role in Game Analysis

Surreal numbers play a central role in by providing a numerical valuation for the positions of under the normal play convention, where the player unable to move loses and the last player to move wins. In this framework, an impartial game position is represented in a form analogous to the birth of surreal numbers: = {L | R}, where L and R are the sets of possible moves available to the current player (since the game is impartial, the options are symmetric for both players, often denoted as = {S | S} with S the set of option values). The value of is then the surreal number born from the values of its options, defined recursively as the simplest surreal number strictly greater than all elements in the left set and strictly less than all in the right set. The valuation process simplifies to a surreal number through the minimum excludant (mex) applied to the Grundy numbers of the options and the disjunctive sum of games. For an impartial game, the Grundy number (or nimber exponent) is the mex of the Grundy numbers of its options, and the surreal value is then *g, where g is that Grundy number and * denotes the nimber construction within the surreals. This recursive definition mirrors the inductive birth process of surreal numbers, allowing game positions to be quantified precisely. Representative examples illustrate this valuation. A Nim heap of size n has Grundy number n and surreal value *n = { *0, *1, \dots, *(n-1) \mid *0, *1, \dots, *(n-1) }, a surreal number born on day n. In particular, a Nim heap of size 1 has value * = { 0 \mid 0 }, known as "star," which is an infinitesimal surreal number satisfying * + * = 0 and representing a position where the first player wins but with minimal advantage. The disjunctive sum of two impartial games G and H, where a player may move in exactly one component on their turn, has surreal equal to the surreal of their individual : val(G + H) = val(G) + val(H). This enables the analysis of complex strategies by decomposing multi-component games into of simpler positions, with winning strategies determined by whether the total is nonzero (first-player win) or zero (second-player win). For instance, two star positions to * + * = 0, a second-player win. A fundamental result is that every short —defined as one with finite and no plays—has a surreal number value, as established by the Sprague-Grundy theorem, which equates such games to nim-heaps, and Conway's embedding of nimbers into the surreals. This theorem ensures that the entire class of short s under normal play can be valued within the of surreal numbers, facilitating rigorous strategic analysis.

Broader Mathematical Uses

Surreal numbers provide a foundation for non-standard analysis by serving as a model for hyperreal numbers, incorporating infinitesimals and quantities to extend classical . In this framework, surreal numbers enable the treatment of infinitesimals as ε-small elements, where - < nε < for all finite n, allowing reciprocals to represent values. This structure supports rigorous definitions of , differentiability, and , simplifying proofs in analysis by avoiding ε-δ arguments; for instance, a f is differentiable at x if the [f(x + ε) - f(x)] / ε is ε-near f'(x) for ε-small surreal ε. In , surreal numbers embed Hardy fields and transseries, offering a universal ordered differential field for handling divergent power series and non-Archimedean growth rates. Hardy fields, consisting of germs of real functions at infinity closed under , embed into the surreals equipped with the Berarducci-Mantova , which extends the surreal and logarithmic functions while preserving asymptotic relations. This embedding facilitates the analysis of transseries—formal series summing terms like x^α exp(β log x + γ) with surreal exponents—enabling solutions to asymptotic equivalence problems for functions like e^{1/x} as x → 0^+. The surreal field No with this is Liouville-closed, meaning every element has an , which supports the summation of in a coherent algebraic manner. Applications in physics leverage surreal numbers to model infinite spacetimes and singularities in , where standard fails due to infinities. In contexts, surreals address divergences in metrics (e.g., the Schwarzschild singularity where → ∞ as r → 0) by naturally incorporating and transfinite scales, potentially unifying and continuous descriptions via dense rationals. Post-2000 research has explored connections between surreal numbers and physical models, suggesting their potential to address singularities in , such as the , through dualities linking twistors and qubits via Grassmann-Plücker relations. As of June 2025, discussions suggest surreal numbers may prove valuable in future theories by naturally incorporating infinities. Surreal geometry and differential equations remain underexplored compared to their real counterparts, with developments in the focusing on foundational analysis rather than computational solvers. Surreal derivatives are defined via Dedekind cuts on function limits, extending to higher orders, while integrals use extrapolative Riemann sums over surreal partitions, satisfying the under suitable conditions. Ordinary differential equations (s) pose open challenges, as consistent higher-order and solution uniqueness require resolving integration cycles; however, the surreal derivation structure supports elementary ODEs in transseries contexts, such as those arising in asymptotic expansions. Computational approaches lag, with no standard surreal ODE solvers implemented, though methods hint at potential for numerical surreal analysis in the 2010s literature. Representative examples illustrate these uses: integrating surreal functions, such as the antiderivative of ω^{-1} yielding log ω via the surreal logarithm, demonstrates infinitesimal accumulation over transfinite intervals. Surreal-valued probabilities extend by assigning values like 1/ω to infinite lotteries, resolving paradoxes in transfinite games where real probabilities fail, as in surreal decision models that predict rational choices under infinitesimal risks.

Simplicity Hierarchy

The simplicity relation on the surreal numbers forms a partial order that captures the inherent complexity arising from their recursive construction. A surreal number x is defined to be simpler than another surreal number y if the b(x) of x—the smallest ordinal at which x is generated—is strictly less than b(y). When x and y share the same birthday, x is simpler than y if its left and right option sets consist of simpler surreals than those of y, ordered lexicographically by the simplicity of their elements. This relation aligns with the of the surreals, where simpler numbers serve as predecessors in the generative process. The simplicity hierarchy partitions the class of surreal numbers into levels indexed by ordinals corresponding to their birthdays, forming a transfinite stratification that mirrors the recursive buildup from the empty set. Level 0 contains only 0, born at ordinal 0. Level 1 introduces +1 = \{0 \mid \} and -1 = \{ \mid 0\}. Finite levels beyond this generate integers and dyadic rationals, such as +2 and +1/2 = \{ 0 \mid 1 \} at level 2, while transfinite levels produce infinities like \omega = \{0,1,2,\dots \mid \} at level \omega and infinitesimals like \varepsilon = \{0 \mid 1, \frac{1}{2}, \frac{1}{4}, \dots \} at level \omega. This hierarchy ensures that every surreal resides at a unique simplicity level, facilitating inductive arguments over the structure. Illustrative examples highlight the ordering: \omega, with birthday \omega, is simpler than \omega + 1 = \{\omega \mid \}, born at birthday \omega + 1. Likewise, the infinitesimal \varepsilon at birthday \omega is simpler than $1/\omega = \{0 \mid \omega\}, which appears at birthday \omega + 1. These comparisons underscore how birthday differences dominate the simplicity assessment. In computational contexts, the hierarchy enables approximations of intricate surreals by truncating to finite-birthday levels, preserving key properties like and for practical evaluations. It also supports algorithmic : Knuth's up/down exploits to iteratively refine a surreal's option sets, selecting the simplest equivalents to yield the unique Conway normal form without altering the number's value.

References

  1. [1]
    Surreal Numbers Are a Real Thing. Here's How to Make Them
    Feb 14, 2024 · Surreal numbers are created by adding values between two given preexisting numbers. If you look at 0 and 1, for example, 1 / 2 is in the middle, 1 / 4 is ...
  2. [2]
    [PDF] on numbers - and games - Mathematics Department
    Conway, John Horton. On numbers and games / John H. Conway.-2nd ed. p.cm ... In the system of “Surreal Numbers" we shall describe, every number has its own.
  3. [3]
    Knuth: Surreal Numbers - Stanford Computer Science
    How two ex-students turned on to pure mathematics and found total happiness. by Donald E. Knuth (Reading, Massachusetts: Addison-Wesley, 1974), vi+119pp. ISBN 0 ...
  4. [4]
    [PDF] An Introduction to Surreal Numbers - Whitman College
    May 8, 2012 · Mathematician John Horton Conway first invented surreal numbers, and Donald Knuth introduced them to the public in 1974 in his mathematical ...Missing: history | Show results with:history
  5. [5]
    [PDF] Zermelo and the Early History of Game Theory
    The first time a proof by backward induction is used seems to be in von Neumann and Morgenstern (1953). The first mention of Zermelo in connection with ...Missing: surreal Borel
  6. [6]
    [PDF] Emile Borel and the foundations of game theory - Knowledge Base
    The early work of Borel and von Neumann focused on two-person zero-sum games, where one player's gain would always equal the other player's loss. In a two ...Missing: Zermelo ordinals 1910s- 1940s
  7. [7]
    Von Neumann and the Development of Game Theory
    Most historians give the credit for developing and popularizing game theory to John Von Neumann, who published his first paper on game theory in 1928.Missing: surreal combinatorial Zermelo
  8. [8]
    The G-values of various games - Cambridge University Press
    Oct 24, 2008 · The G-values of various games. Published online by Cambridge University Press: 24 October 2008. Richard K. Guy and.
  9. [9]
    Surreal Number -- from Wolfram MathWorld
    They were invented by John H. Conway in 1969. Every real number is surrounded by surreals, which are closer to it than any real number. Knuth (1974) describes ...<|control11|><|separator|>
  10. [10]
    On Numbers And Games by J. Conway, Prologue
    In fact, the Surreal Numbers "surfaced" before ONAG appeared, partly through my 1970 lectures at Cambridge and Cal. Tech., but mostly through the wide ...
  11. [11]
    The surreal numbers - by Joel David Hamkins - Infinitely More
    Jan 6, 2024 · Thus is born the first surreal number, the number zero, born on day 0. We have created something from nothing. We are witnessing the surreal ...
  12. [12]
    [PDF] The Surreal Numbers and Combinatorial Games - PEARL
    Jul 24, 2019 · This paper provides an introduction to Games, and specifically to Con- way's Surreal Numbers, as introduced in his 1976 book On Numbers and ...Missing: Borel | Show results with:Borel
  13. [13]
    [PDF] Surreal Numbers – An Introduction - Claus Tøndering
    The term “surreal number” was invented by Donald Knuth [2]. What can surreal numbers be used for? Not very much at present, except for some use in game theory.
  14. [14]
    John Conway (1937 - 2020) - Biography - MacTutor
    The name surreal numbers was not invented by Conway, however, but by Donald Knuth who was so impressed with Conway's discovery that he wrote Surreal Numbers ( ...
  15. [15]
    A Generalization of Conway's Theory of Surreal Numbers - jstor
    Volume 66, Number 3, Sept. 2001. NUMBER SYSTEMS WITH SIMPLICITY HIERARCHIES: A GENERALIZATION OF CONWAY'S THEORY OF SURREAL. NUMBERS. PHILIP EHRLICH.
  16. [16]
    NUMBER SYSTEMS WITH SIMPLICITY HIERARCHIES: A ...
    Feb 5, 2018 · Conway's ordered field of surreal numbers was brought to the fore and employed to provide necessary and sufficient conditions for an ordered ...
  17. [17]
    Surreal numbers vs. non-standard analysis - MathOverflow
    Mar 19, 2012 · The ordered field of surreal numbers admits a relational extension to a model of non-standard analysis and, hence, that in such a relational extension the ...set theory - What's wrong with the surreals? - MathOverflowInterpreting Conway's remark about using the surreals for non ...More results from mathoverflow.netMissing: 2000 | Show results with:2000
  18. [18]
    The absolute arithmetic continuum and the unification of all numbers ...
    In addition to its inclusive structure as an ordered field, the system No of surreal numbers has a rich algebraico-tree-theoretic structure—a simplicity ...
  19. [19]
    Some Mathematical and Physical Remarks on Surreal Numbers
    Nov 22, 2016 · We make a number of observations on Conway surreal number theory which may be useful, for further developments, in both in mathematics and ...Missing: applications quantum field Lagrangians 2010s
  20. [20]
    quantum field theory - Renormalization and Conway/Surreal Numbers
    Oct 30, 2015 · Paul Teller writes about three interpretations of renormalization in quantum field theory. In particular, Teller denotes one of these approaches Real- ...
  21. [21]
    [PDF] Surreal Numbers and Games - MIT
    Feb 10, 2009 · We'll start by using Conway's methods to represent games, and then show how these games/numbers form a new number system. On Numbers and Games.
  22. [22]
    [PDF] MEET THE SURREAL NUMBERS - Mathematical Association
    Apr 4, 2017 · We'll define this properly later, when it will be clear that 0 is the only number born on day 0. On to day 1. We now have a new set of numbers, ...
  23. [23]
    [PDF] Surreal Birthdays and Their Arithmetic - arXiv
    Oct 25, 2018 · A recursive definition means that a surreal number is defined in terms of other surreals and so on. The break in this circularity that makes ...
  24. [24]
    Infinity > Construction of Surreal Numbers (Stanford Encyclopedia of ...
    All numbers have a unique normal form expression \(\sum_{\beta\lt \alpha}\omega^{y_\beta}r_\beta\), where \(\alpha\) is an ordinal, the \(r_\beta\) \((\beta\lt\ ...
  25. [25]
    [2410.00065] Conway Normal Form: Bridging Approaches for ... - arXiv
    Sep 30, 2024 · We demonstrate that surreal numbers form a field, including the square root, and that they encompass subsets such as reals, ordinals, and powers ...
  26. [26]
  27. [27]
    [PDF] MEET THE SURREAL NUMBERS - Mathematical Association
    Apr 4, 2017 · The Real Construction. • Construct the natural numbers, using von Neumann's ordinal construction, but stopping after the finite ordinals;. • ...
  28. [28]
    [PDF] An very brief overview of Surreal Numbers
    Surreal numbers were created by John Horton Conway (of Game of Life fame), as a greatly simplified construction of an earlier object (Alling's ordered field ...
  29. [29]
    [2110.05237] The continuum of the surreal numbers revisited ... - arXiv
    Oct 8, 2021 · The surreal numbers defined through transfinite Cauchy fundamental sequences ... Conway (1976) defined by Dedekind cuts ,and of the ordinal real ...
  30. [30]
    [PDF] Surreal Numbers - VCU Scholars Compass
    Conway, in his book On Numbers and Games, states that he prefers to omit the adjective "surreal" and refer instead to No, "the class of all numbers." (To avoid ...
  31. [31]
    Surreal Numbers - Interactive Mathematics Miscellany and Puzzles
    The first of these implies that 0 = {-1|} because -1 < 0 and then, from the second, 0 = {-1|1} because 0 < 1. Conway proves (1) in a most elegant and ...
  32. [32]
    fundamentals of analysis over surreal numbers fields - Project Euclid
    Oct 1, 1986 · The surreal number fields £No are all real-closed. They have ex- traordinary higher order properties, which allow one to do analysis over them, ...<|control11|><|separator|>
  33. [33]
  34. [34]
    [PDF] an overview of surreal numbers - Simon Rubinstein-Salzedo
    Abstract. The surreal numbers are constructed in stages, with each surreal number being defined by two other surreal numbers constructed earlier.<|control11|><|separator|>
  35. [35]
    [PDF] Conway Normal Form: Bridging Approaches for Comprehensive ...
    Sep 30, 2024 · ... surreal/. Page 16. 16. Conway Normal Form for the Formalization of Surreal Numbers ... On Numbers And Games. A K Peters Ltd., 2nd edition, 2001.
  36. [36]
    [1503.00315] Surreal numbers, derivations and transseries - arXiv
    Mar 1, 2015 · Several authors have conjectured that Conway's field of surreal numbers, equipped with the exponential function of Kruskal and Gonshor, can be ...Missing: definition | Show results with:definition
  37. [37]
    [PDF] What are Conway's Surreal Numbers, and what should they be?
    Dec 16, 2024 · Is there a surreal version of the p-adic numbers? Question 16. Can ... [E12] Ehrlich, P., "The absolute arithmetic continuum and the unification ...
  38. [38]
    [PDF] Surreal substructures - HAL
    Mar 2, 2020 · The number 1/2 is rep- resented by the sign sequence +1,−1 of length 2. The ordering ⩽ on No corresponds to the lexicographical ordering on sign ...
  39. [39]
    None
    ### Summary of Surreal Numbers as Hahn Series from arXiv:2509.22374
  40. [40]
    An alternative construction of Conway's ordered field No
    andEhrlich, P.,An Alternative Construction of Conway's Surreal Numbers, C. R. Math. Rep. Acad. Sci. CanadaVIII (1986), pp. 241–246. Google Scholar. Alling, N ...
  41. [41]
    [PDF] AN INTRODUCTION TO CONWAY'S GAMES AND NUMBERS
    Combinatorial Game Theory is a fascinating and rich theory, based on a simple ... Definition 6.1 (Infinitesimal). (1) A game G is called infinitesimal if −2 ...
  42. [42]
    [PDF] Another View of Nonstandard Analysis | Haverford College
    A surreal is -near a real iff no other real is between them, and a surreal is -near a non-zero real iff their ratio is -near 1 ​R​. 2. *Reals. We assume that ...<|control11|><|separator|>
  43. [43]
    Surreal numbers with derivation, Hardy fields and transseries: a survey
    Aug 11, 2016 · Abstract:The present article surveys surreal numbers with an informal approach, from their very first definition to their structure of universal ...
  44. [44]
    Duality, Matroids, Qubits, Twistors, and Surreal Numbers - Frontiers
    The surprise with surreal numbers is that predicts that besides 12 1 2 -spin system there must exist infinite number of J J -spins, according to the formula (43) ...
  45. [45]
    [1307.7392] Analysis on Surreal Numbers - arXiv
    Jul 28, 2013 · In this paper, we extend this work with a treatment of functions, limits, derivatives, power series, and integrals.Missing: solvers | Show results with:solvers
  46. [46]
    a generalization of Conway's theory of surreal numbers | The ...
    Mar 12, 2014 · This algebraico-tree-theoretic structure, or simplicity hierarchy, as we have called it [15], depends upon No's (implicit) structure as a ...