Fact-checked by Grok 2 weeks ago

Direct proof

A direct proof in is a logical used to establish the truth of a , particularly an implication of the form "if P, then Q," by assuming P is true and deriving the conclusion Q through a sequence of valid deductions based on axioms, definitions, previously proven theorems, and rules of inference, without additional assumptions or contradictions. This approach contrasts with indirect methods, such as proof by contradiction or contrapositive, by proceeding straightforwardly from the given hypotheses to the desired outcome, making it the most fundamental and commonly taught technique in mathematical reasoning. Direct proofs typically begin with the assumption of the and employ deductive steps, such as (if A implies B and A is true, then B is true) or algebraic manipulations, to bridge to the conclusion. For instance, to prove that the square of an even is even, one assumes the n = 2k for some k, substitutes to get n^2 = (2k)^2 = 4k^2 = 2(2k^2), and concludes n^2 is even since it is twice an . Similarly, proving the sum of two odd integers is even involves expressing the odds as $2j + 1 and $2m + 1, adding to $2(j + m + 1), which is even. These examples illustrate how direct proofs rely on precise definitions and basic properties of numbers, ensuring each step is justified. As a of , direct proofs are essential in fields like , , and , where they build foundational results incrementally and facilitate the of theorems through transparent . They are preferred when the logical path from to conclusion is clear, promoting rigor and clarity in arguments, though they may be supplemented by indirect methods for more statements.

Fundamentals

Definition

A direct proof is a fundamental method in mathematical logic and proof theory, wherein the conclusion is established by logically deriving it from the given premises through a sequence of valid implications, without invoking the negation of the conclusion or techniques such as reductio ad absurdum. This approach ensures that the truth of the statement follows directly from accepted axioms, definitions, and hypotheses, forming the backbone of deductive mathematics. The formal structure of a direct proof begins with the premises—typically axioms, previously proven theorems, or specific hypotheses—and proceeds step by step to the desired conclusion, known as the , using established rules of such as or . Each intermediate statement must be justified by a clear to ones, creating a transparent chain that leaves no gaps in reasoning. In contrast to informal arguments, which often depend on , diagrams, or explanations that may overlook subtle flaws, a direct proof demands rigorous justification at every , typically expressed through implications like P \rightarrow Q, where P encompasses the and Q the conclusion. This emphasis on formality distinguishes it as a of verifiable mathematical discourse. At its core, direct proof presupposes an understanding of deductive reasoning, which operates via the implication P \rightarrow Q: if the antecedent P holds, then the consequent Q necessarily follows. This mechanism underpins the method's reliability, providing certainty in conclusions drawn from true premises, unlike indirect alternatives such as proof by contradiction.

Key Characteristics

Direct proofs proceed in a unidirectional manner from the given hypotheses or assumptions to the desired conclusion, constructing a of logical deductions without reversing or assuming the of the conclusion. This forward-directed typically involves a sequence of statements, where each subsequent step follows directly from previous ones using established rules of inference, such as or tautologies. Every step in a direct proof must be grounded in mathematical definitions, axioms, previously proven theorems, or logical equivalences, ensuring that no unsubstantiated leaps occur and preventing . This reliance on foundational elements allows the proof to build rigorously upon accepted truths, such as defining an even as n = 2k for some k, and applying properties like under . The structure of direct proofs emphasizes , with each explicitly justified and traceable to prior statements, facilitating straightforward verification by readers familiar with the underlying . This explicitness makes direct proofs particularly amenable to , as one can check the validity of each independently without needing to reconstruct hidden assumptions. Common elements in direct proofs include the use of quantifiers, such as the universal quantifier \forall to address statements for all elements in a set (e.g., "for all integers n satisfying a property") and the existential quantifier \exists to assert the existence of specific instances, as well as equivalence relations denoted by \leftrightarrow to establish if-and-only-if conditions. In contexts involving , direct proofs may incorporate , such as working modulo 2 to demonstrate properties. While direct proofs are often concise for elementary theorems—spanning just a few lines or paragraphs—they can extend to considerable length for more complex results, potentially incorporating intermediate lemmas or sub-proofs to manage the overall argument. This variability in length reflects the theorem's complexity but maintains the proof's linear coherence throughout.

Comparison with Other Proof Techniques

Versus Proof by Contradiction

Proof by contradiction, also known as reductio ad absurdum, involves assuming the negation of the desired conclusion (¬Q in a statement P → Q) and deriving a logical contradiction from that assumption alongside the premises, thereby establishing that Q must hold. In contrast, a direct proof proceeds by assuming the premise P and logically deducing Q without invoking any negation or contradictory assumption, relying solely on forward inference from given conditions. Direct proofs offer advantages over proofs by by sidestepping the risk of overlooking subtle contradictions or introducing errors in negating statements, as the process remains grounded in constructive steps that explicitly build the conclusion. Moreover, direct proofs provide deeper constructive into the underlying the , revealing how and why the result follows from the premises, which can facilitate further applications or generalizations in . Proofs by are often preferred for existential statements, where demonstrating the of an object indirectly by refuting its non- avoids the need for an explicit , or when the route is obscured by intricate intermediate steps that are difficult to navigate sequentially. Under , proofs and proofs by are logically equivalent for establishing implications, as both leverage the ; however, proofs align more closely with the constructive principles of , which rejects non-constructive methods like assuming ¬Q to derive a . A classic example illustrating the preference for is the of √2, where assuming √2 = p/q for integers p and q in lowest terms leads to a via infinite descent on the denominator, whereas direct proofs are more naturally suited to statements like the rationality of sums of .

Versus Proof by Contraposition

A direct proof of an implication P \to Q proceeds by assuming P and deriving Q through a chain of logical deductions, establishing the forward direction explicitly. In contrast, proof by addresses the same implication by proving its contrapositive \neg Q \to \neg P, which reverses the conditional by negating both the antecedent and consequent. This approach is logically equivalent to the original statement, as the two forms share the same : if \neg Q implies \neg P, then it cannot be that P is true while Q is false. Both methods rely on the deductive rule of —given A \to B and A, conclude B—but specifically exploits the (P \to Q) \leftrightarrow (\neg Q \to \neg P), allowing the proof to start from the of the conclusion rather than . While a direct proof builds constructively from the toward the result, often yielding insight into the mechanism of the , can simplify reasoning when the forward is obscured or involves steps, such as in cases where assuming the failure of Q directly leads to the failure of P. For instance, in universal statements like "if n is even, then n^2 is even," the contrapositive "if n^2 is , then n is " may be easier to verify by assuming the undesired outcome. In practice, proofs are preferred for constructive arguments where the implication naturally unfolds from known properties, whereas suits scenarios involving or where hypothesizing the provides a clearer route to with the premise—though it remains a form of indirect reasoning akin to . However, is limited to implicational statements and cannot be applied to non-conditional propositions, such as existential claims or equalities, rendering methods more in those contexts.

Historical Context

Origins and Etymology

The conceptual roots of deductive proof practices, foundational to what is now termed direct proof, appear in various ancient traditions, including Mesopotamian procedural demonstrations around 2000 BCE and geometric justifications in and . Systematic axiomatic deductions, particularly in as seen in Euclid's Elements around 300 BCE, formalized theorems derived straightforwardly from axioms and prior propositions without explicit naming of the method. This approach emphasized logical progression from established premises, forming the foundation of and implicitly distinguishing it from indirect techniques like . Euclid's systematic use of such deductions established a paradigm for mathematical reasoning that persisted for over two , influencing later traditions. The term "direct proof" emerged in 19th- and early 20th-century mathematical to explicitly differentiate straightforward deductive methods from indirect ones, such as those formalized by and earlier by Aristotle's (reduction to absurdity), a technique involving assumption of the negation to derive a contradiction. Etymologically, "direct" derives from the Latin directus, meaning "straight" or "straightforward," highlighting the method's linear path from to conclusion, in contrast to the circuitous nature of reductio. Aristotle coined the conceptual basis for indirect proof in his Prior Analytics (ca. 350 BCE), but direct methods remained the default in axiomatic systems. Formalization of the term occurred in modern logic texts during the early , notably through David Hilbert's foundational studies, which categorized proof types within his program for securing via finitist, constructive methods akin to direct proofs. Hilbert's emphasis on direct consistency proofs for arithmetic systems, as in his 1928 Grundzüge der theoretischen Logik, helped delineate direct approaches from transfinite or indirect ones. The terminology standardized in English-speaking mathematical education after , influenced by foundational works like Alfred North Whitehead and Bertrand Russell's (), which employed formal deductive chains to rebuild from logical axioms.

Early Notable Applications

One of the earliest and most influential applications of direct proofs appears in 's Elements (c. 300 BCE), particularly in Book I, where he employs step-by-step deductions from axioms and postulates to establish congruence theorems for triangles. For instance, 4 demonstrates the side-angle-side () congruence criterion by constructing an auxiliary triangle and using previously proven properties of equality and superposition to show that the original triangles are congruent, laying foundational principles for plane geometry without reliance on contradiction or . This axiomatic approach influenced subsequent mathematical rigor, as builds each directly upon prior ones, culminating in results like the in 47. In non-Greek traditions, early deductive proofs include the Chinese mathematician Liu Hui's (ca. 263 CE) commentaries on The Nine Chapters on the Mathematical Art, where he used direct dissection methods, such as "layering the squares," to prove the volume of a by transforming it into prisms and applying area relations step-by-step. Similarly, Indian śulbasūtras (ca. 600–200 BCE) contain direct geometric constructions and verifications for altar designs, deriving lengths and areas from basic postulates without contradiction. In the 3rd century BCE, advanced geometric derivations in his treatise , where he computes and surface area formulas for spheres and cylinders by applying axioms and the , which involves indirect elements through double contradiction to establish bounds but builds on direct inequalities from prior propositions for exact ratios. These proofs highlight ' emphasis on logical deduction from elementary relations to three-dimensional results. The marked a revival of direct proofs through ' La Géométrie (1637), an appendix to his , where he introduces and uses algebraic s to directly prove of conic sections. By assigning coordinates to points and representing curves as s—such as the general second-degree for conics—Descartes demonstrates classifications and constructions (e.g., ellipses as sections of cones) through algebraic manipulations that translate geometric relations into arithmetic operations, verified step-by-step without geometric intuition alone. This fusion enabled direct proofs of tangency and points, transforming classical into a more algebraic, verifiable framework. In 19th-century , employed direct proofs in (1801) to establish the law of , a cornerstone relating solvability of quadratic congruences modulo primes. In Articles 125–146, Gauss directly constructs the and uses properties of least residues and cyclotomic polynomials to derive the reciprocity formula \left( \frac{p}{q} \right) \left( \frac{q}{p} \right) = (-1)^{\frac{p-1}{2} \cdot \frac{q-1}{2}} for odd primes p and q, building from axioms without contradiction. This systematic deduction, expanded in later sections, provided a rigorous foundation for higher reciprocity laws and influenced . The transition to the is evident in Giuseppe Peano's 1889 formulation of axioms for in Arithmetices Principia, where direct proofs from primitive notions (, successor, and ) establish core arithmetic properties like and . Peano's five axioms—starting with 0 as a and the successor function's injectivity—allow theorems such as commutativity of to be proven by direct on the successor, ensuring all properties derive transparently from the axiomatic base without external assumptions. This framework formalized arithmetic's logical structure, paving the way for set-theoretic foundations in mathematics.

Illustrative Examples

Sum of Even Integers

One classic example of a direct proof in elementary is the demonstration that the sum of two even integers is even. The theorem states: If m and n are even integers, then m + n is even. This result relies on the universal quantifier, asserting that the property holds for all even integers m and n. To prove this, begin with the definition of an even : an k is even there exists an l such that k = 2l. Assume m and n are even, so let m = 2a and n = 2b, where a and b are s. Then, substitute into the sum: m + n = 2a + 2b = 2(a + b). Here, a + b is an , so $2(a + b) is even by the definition of evenness. Each step follows directly: the first uses from the , and the second applies the of addition over multiplication (i.e., $2a + 2b = 2(a + b)), which holds for all s. This proof applies exclusively to integers, as the concept of evenness is defined within the \mathbb{Z}, and it does not extend straightforwardly to other number systems without adjustment. confirms its direct nature: no or is assumed; the argument proceeds forward solely from the definitions and basic arithmetic properties, establishing the conclusion without extraneous hypotheses.

Pythagorean Theorem

The Pythagorean theorem states that in a with legs of lengths a and b, and of length c, the equation a^2 + b^2 = c^2 holds. A classic direct proof of this theorem employs the construction of an altitude to the , which generates two smaller similar to the original, allowing the use of similarity ratios to establish the relationship. Consider ABC with the at C, where BC = a, AC = b, and AB = c. Draw the altitude from C to AB, intersecting at point H. This divides AB into segments AH and BH, and forms two smaller : \triangle ACH and \triangle CBH. Both smaller triangles are similar to the original \triangle [ABC](/page/ABC). Specifically, \triangle ACH \sim \triangle [ABC](/page/ABC) because they share the angle at A and each has a right (at H and C, respectively). Similarly, \triangle CBH \sim \triangle [ABC](/page/ABC) due to the shared angle at B and right s at H and C. The similarity \triangle ACH \sim \triangle [ABC](/page/ABC) implies that the ratios of corresponding sides are equal: \frac{AC}{AB} = \frac{AH}{AC}, or \frac{b}{c} = \frac{AH}{b}. Solving for AH yields AH = \frac{b^2}{c}. Likewise, the similarity \triangle CBH \sim \triangle [ABC](/page/ABC) gives \frac{BC}{AB} = \frac{BH}{BC}, or \frac{a}{c} = \frac{BH}{a}, so BH = \frac{a^2}{c}. Since AB = AH + BH = c, substitute the expressions for the segments: c = \frac{b^2}{c} + \frac{a^2}{c}. Multiplying through by c produces the desired equation: c^2 = a^2 + b^2. This proof assumes the axioms of Euclidean geometry, including those defining right angles, congruence of triangles, and the proportionality of sides in similar triangles derived from equal angles. An alternative direct variant, Euclid's original proof in the Elements, rearranges geometric figures built on the triangle's sides to demonstrate the area equality without invoking similarity.

Odd Number Squared

The theorem states that if n is an odd integer, then n^2 is odd. To prove this directly, assume n is an odd integer. By definition, there exists an integer k such that n = 2k + 1. Substituting this into the square gives: n^2 = (2k + 1)^2 = 4k^2 + 4k + 1 = 2(2k^2 + 2k) + 1. The expression $2(2k^2 + 2k) + 1 is of the form $2m + 1 where m = 2k^2 + 2k is an integer, confirming that n^2 is odd. This expansion relies on the applied to (a + b)^2 = a^2 + 2ab + b^2 with a = 2k and b = 1, which preserves the since even terms ($4k^2 and $4k) combine to form an even number, leaving the remainder of 1. In contrast, the square of an even integer is even, as (2k)^2 = 4k^2, which is divisible by 2; however, the direct proof here avoids reliance on or by working solely from the hypothesis of oddness. This result holds generally for all odd integers due to the existential representation n = 2k + 1, which captures the universal property of parity without exception.

Triangle Inequality

The theorem states that in any with side lengths a, b, and c, the inequalities a + b > c, a + c > b, and b + c > a hold. This result assumes a non-degenerate in , where the points are non-collinear and the space is equipped with the standard metric, ensuring positive side lengths and the formation of a valid triangular . A direct proof proceeds geometrically by constructing an auxiliary line and applying properties of isosceles triangles and angles, as in Euclid's Proposition I.20. Consider triangle ABC with sides BC = a, AC = b, and AB = c. To show b + c > a, extend side BA beyond A to a point D such that DA = b (equal to AC). Join D to C, forming triangle ACD and larger triangle BCD. Since DA = AC = b, triangle ACD is isosceles, so base angles are equal: \angle ACD = \angle ADC. Now, in triangle BCD, \angle BCD > \angle BDC because \angle BCD includes \angle ACD and \angle BCD is exterior to \angle ADC at point A. By the theorem that the side opposite the larger angle is longer ( I.19), BD > BC, or BD > a. But BD = BA + AD = c + b, so b + c > a. The other inequalities follow by . This proof relies solely on forward deductions from the axioms of , including equality of isosceles angles and the angle-side relationship, without assuming the result or using contradiction. An algebraic variant uses representations in , interpreting the sides as . Let \mathbf{u} and \mathbf{v} represent two sides from a common , with lengths \|\mathbf{u}\| = b and \|\mathbf{v}\| = c. The third side corresponds to \|\mathbf{u} + \mathbf{v}\|, which equals a. The follows from the of the Euclidean norm: \|\mathbf{u} + \mathbf{v}\| \leq \|\mathbf{u}\| + \|\mathbf{v}\|, or a \leq b + c. For the strict inequality in a non-degenerate , the vectors are not collinear in opposite directions, so the \mathbf{u} \cdot \mathbf{v} > -\|\mathbf{u}\|\|\mathbf{v}\|, ensuring \|\mathbf{u} + \mathbf{v}\|^2 = \|\mathbf{u}\|^2 + \|\mathbf{v}\|^2 + 2\mathbf{u} \cdot \mathbf{v} < (\|\mathbf{u}\| + \|\mathbf{v}\|)^2. This derivation uses the norm axioms and the Cauchy-Schwarz implicitly through the dot product bound.

Applications and Advantages

In Arithmetic and Algebra

In arithmetic and algebra, direct proofs are essential for establishing core theorems by systematically applying definitions, axioms, and logical deductions without relying on contradiction or other indirect methods. These proofs often leverage induction or constructive algorithms to demonstrate existence and uniqueness, providing clear insight into the structure of numbers and algebraic objects. For instance, they verify properties that form the basis for more advanced theories, such as factorization and , while highlighting the importance of rigorous verification from foundational principles. The Fundamental Theorem of Arithmetic states that every integer greater than 1 can be expressed as a product of primes, with the factorization being unique up to the order of factors. Its existence part is proved directly via strong mathematical induction: the base case holds for small integers, and assuming it for all numbers up to n, any composite n+1 factors into primes by dividing by its smallest prime divisor and applying the inductive hypothesis to the quotient. Uniqueness follows directly from the definition of primes and Euclid's lemma, which shows that if a prime divides a product, it divides one of the factors, ensuring no alternative factorization exists. In algebra, direct proofs verify key identities like the binomial theorem, which expands (x + y)^n = \sum_{k=0}^{n} \binom{n}{k} x^{n-k} y^k. This is established combinatorially by counting the number of ways to choose k factors of y from n terms in the product (x + y)^n, directly matching the coefficient \binom{n}{k} to the expansion terms without induction or other techniques. Similarly, in group theory, basic properties such as closure—ensuring the operation on any two elements yields another element in the set—and associativity—verifying (ab)c = a(bc) for all elements—are proved directly from the operation's definition, often by explicit computation in concrete examples like integers under addition. For polynomials over a field, the Euclidean algorithm serves as a direct constructive proof of the existence of the greatest common divisor (gcd). Starting with two polynomials f and g where \deg(f) \geq \deg(g), repeated division yields remainders until a zero remainder, with the last non-zero remainder being a gcd multiple; normalization ensures monicity, directly computing the gcd from the division algorithm's properties. A frequent pitfall in such direct proofs is assuming unverified properties, such as commutativity of the operation, which must be deduced from definitions to avoid invalidating the argument.

In Geometry and Beyond

In synthetic geometry, direct proofs play a central role in establishing fundamental theorems, such as the for triangles. This theorem states that the sum of the interior angles of a triangle equals 180 degrees, derived by constructing a line through one vertex parallel to the opposite side and applying properties of alternate interior angles with transversals. The proof relies on , which asserts that through a point not on a given line, exactly one parallel line can be drawn, ensuring the equality of corresponding angles without invoking contradictions. This direct approach highlights the constructive nature of Euclidean geometry, where axioms lead straightforwardly to spatial properties. Extending to calculus foundations, direct proofs via the epsilon-delta definition provide rigorous verification of limits and continuity. For instance, to prove that the limit of f(x) = x^2 as x approaches 2 is 4, one assumes \varepsilon > 0 and finds a \delta > 0 such that if $0 < |x - 2| < \delta, then |x^2 - 4| < \varepsilon, achieved by bounding |x^2 - 4| = |x-2||x+2| and choosing \delta = \min(1, \varepsilon/5). Similarly, continuity at a point c is directly shown if for every \varepsilon > 0, there exists \delta > 0 such that |x - c| < \delta implies |f(x) - f(c)| < \varepsilon, building from the limit definition without indirect assumptions. These proofs underscore the precision of real analysis, directly linking function behavior to quantitative closeness. In , direct proofs establish properties such as the connectedness of the union of connected sets that have a non-empty . If {C_i} is a family of connected subsets of a X all containing a fixed point p, then their is connected, proved by showing that any to {0,1} () must be on the union, as it is constant on each C_i and they intersect at p, directly using the definition of connectedness without contradiction. In physics, direct proofs derive conservation laws from symmetry principles, as in , which links continuous symmetries of the to conserved quantities. For time-translation invariance, the proof proceeds by varying the action integral under time shifts, yielding \frac{d}{dt} \left( \sum_i \frac{\partial L}{\partial \dot{q}_i} \dot{q}_i - L \right) = 0, directly implying . Spatial translation symmetry similarly leads to momentum conservation through analogous variational calculations, providing a foundational tool for classical and . Interdisciplinarily, direct proofs in verify algorithm correctness using invariants, which maintain a property true before, during, and after iterations. For example, in , the invariant states that the first i elements are the smallest i in sorted order; initialization holds for i=0, maintenance is shown by swapping the minimum into position i, and termination confirms the array is fully sorted. This method directly combines over iterations with precondition-postcondition analysis, ensuring partial correctness and termination without exhaustive testing.

Strengths Over Indirect Methods

Direct proofs are inherently constructive, providing explicit constructions or algorithms that demonstrate the existence of mathematical objects, whereas indirect proofs, such as those by or , often yield non-constructive results that merely establish existence without specifying how to find the object. This constructive quality aligns with the principles of constructive mathematics, where proofs must produce verifiable methods, offering practical utility in fields like and design. The intuitive appeal of direct proofs stems from their alignment with natural , progressing linearly from to conclusions without assuming negations, which facilitates easier comprehension and teaching in educational settings. In contrast to indirect methods like , which require working under assumed falsehoods, direct proofs mirror everyday problem-solving flows, enhancing student engagement and conceptual grasp. Direct proofs avoid potential paradoxes associated with indirect methods in non-classical logics, particularly , which rejects for positive assertions due to the lack of elimination, thereby ensuring proofs remain grounded in constructive evidence rather than mere inconsistency. This makes direct proofs more robust in frameworks like intuitionistic mathematics, where non-constructive arguments can lead to undecidable statements without explicit constructions. In terms of verification efficiency, direct proofs typically involve shorter logical chains for straightforward theorems, simplifying error detection and validation compared to the layered assumptions in indirect proofs. This streamlined structure aids in and , particularly for elementary results where brevity enhances reliability. Empirical studies in since 2000 indicate that direct proofs improve student retention and understanding over indirect ones, as learners report greater difficulty and lower convincingness with methods like due to their abstract negation-handling. For instance, pre-service teachers consistently find indirect proofs more challenging, leading to narrower proof repertoires and reduced long-term mastery when indirect methods dominate instruction. These findings underscore direct proofs' role in fostering deeper conceptual retention through accessible reasoning pathways.

References

  1. [1]
    2.1 Direct Proofs
    A direct proof is a sequence of statements which are either givens or deductions from previous statements, and whose last statement is the conclusion to be ...
  2. [2]
    [PDF] Ch 3.2: Direct proofs
    A direct proof is a way of showing that a given statement is true or false by using existing lemmas and theorems without making any further assumptions.
  3. [3]
    [PDF] 2. Methods of Proof 2.1. Types of Proofs. Suppose we wish to prove ...
    Direct Proof: Assume p, and then use the rules of inference, axioms, defi- nitions, and logical equivalences to prove q. • Indirect Proof or Proof by ...
  4. [4]
    [PDF] proof-templates.pdf - University of Washington Math Department
    The most straightforward type of proof is called a direct proof : This is one in which we assume the hypotheses, and then, using the rules of deduction that we ...
  5. [5]
    2.6 Indirect Proof
    Indirect proof is used when direct proof is difficult, and it starts by assuming the denial of the conclusion, using proof of the contrapositive or proof by ...
  6. [6]
    [PDF] Contents 1 Proofs, Logic, and Sets - Evan Dummit
    This type of reasoning is usually called deductive reasoning. ◦ Although ... conclusion directly from them: this is often referred to as a direct proof.
  7. [7]
    [PDF] Chapter 1: Logical systems and basic laws of reasoning - UNM Math
    1.7 Deductive Reasoning in an Elementary Proof. Most mathematical proofs consist of chains of simple deductions from definitions, axioms, and previous theorems.
  8. [8]
    [PDF] Direct Proof
    The simplest and most straightforward type of proof is a “direct” proof, which we'll call any proof that follows straight from the definitions or from a direct ...
  9. [9]
    [PDF] Course Notes MAT102H5 Introduction to Mathematical Proofs
    Jul 19, 2018 · An informal argument, that relies too heavily on diagrams, may be incomplete or flawed ... In a direct proof, we simply assume that our hypotheses ...
  10. [10]
    [PDF] Transition to Higher Mathematics: Structure and Proof Second Edition
    that the informal argument will be accepted by the mathematical com- munity ... theorem is a direct proof (Euclid's Elements I.47). (2) Contrapositive ...
  11. [11]
    [PDF] Direct Proofs - Stanford University
    Conclude that there is a general principle explaining the results. ● Even if all data are correct, the conclusion might be incorrect.
  12. [12]
    Mathematical Proofs - Department of Mathematics at UTSA
    Nov 6, 2021 · In direct proof, the conclusion is established by logically combining the axioms, definitions, and earlier theorems. For example, direct proof ...Missing: key characteristics
  13. [13]
    CS103 Guide to Proofs - Stanford University
    Oct 4, 2024 · Proof by Contradiction. (Covered in Lecture 02). One of the most common and most powerful forms of indirect proof is the proof by contradiction.
  14. [14]
    6.7. Mathematical Proof Techniques - OpenDSA
    In general, a direct proof is just a “logical explanation”. A direct proof is sometimes referred to as an argument by deduction. This is simply an argument in ...
  15. [15]
    [PDF] Writing Mathematical Proofs - Hamilton College
    In a direct proof, the first thing you do is explicitly assume that the hypothesis is true for your selected variable, then use this assumption with definitions ...Missing: advantages | Show results with:advantages
  16. [16]
    Proofs — Existential statements (CSCI 2824, Spring 2015)
    Existential statements can be proved in another way without producing an example. Typically this involves a proof by contradiction (we will study these types of ...
  17. [17]
    [PDF] Chapter 3 Proofs
    Proof by contradiction is typically used to prove claims that a certain type of object cannot exist. The negation of the claim then says that an object of this ...
  18. [18]
    Intuitionistic Logic - Stanford Encyclopedia of Philosophy
    Sep 1, 1999 · Intuitionistic logic encompasses the general principles of logical reasoning which have been abstracted by logicians from intuitionistic mathematics.
  19. [19]
    [PDF] Proof Complexity of Intuitionistic Propositional Logic - cs.Toronto
    Nov 29, 2006 · Intuitionistic logic is weaker than classical logic, since it disallows proof by contradiction; hence it is reasonable to conjecture that ...
  20. [20]
    A Geometric Proof That The Square Root Of Two Is Irrational
    Aug 28, 1997 · ... proof that is most commonly given for the irrationality of sqrt(2). ... irrational number (the geometric method and the contradiction ...
  21. [21]
    Proofs of irrationality - Computer Science at Indiana State University
    One of the classic proofs in mathematics is that the square root of 2 is irrational. The classic proof of this is the proof by "infinite descent" that is ...
  22. [22]
    Proofs - Discrete Mathematics - An Open Introduction
    It gives a direct proof of the contrapositive of the implication. This is enough because the contrapositive is logically equivalent to the original implication.
  23. [23]
    3.2: More Methods of Proof - Mathematics LibreTexts
    Sep 29, 2021 · A conditional statement is logically equivalent to its contrapositive. Use a direct proof to prove that ⌝ Q → ⌝ P is true. Caution: One ...
  24. [24]
    [PDF] The History and Concept of Mathematical Proof
    Feb 5, 2007 · The concept of “point” is an undefined term. 3The word “axiom” derives from the Greek axios, meaning “something worthy”. 4The word “postulate” ...
  25. [25]
    [PDF] The History of Mathematical Proof in Ancient Traditions
    ... direct proof with an indirect proof (a criterion notably explained by Heron – see Vitrac 2004 : 17–18 (regarding iii .9 aliter) – and Menelaus), the ...
  26. [26]
    Proof Theory - Stanford Encyclopedia of Philosophy
    Aug 13, 2018 · ... direct proof-theoretic treatment of set theories. Pioneering work on the proof theory of set theories is mainly due to Jäger (1980, 1982) ...
  27. [27]
    Euclid (325 BC - 265 BC) - Biography - University of St Andrews
    The Elements ends with book thirteen which discusses the properties of the five regular polyhedra and gives a proof that there are precisely five. This book ...
  28. [28]
    Archimedes - Biography - University of St Andrews
    In the first book of On the sphere and cylinder Archimedes shows that the surface of a sphere is four times that of a great circle, he finds the area of any ...
  29. [29]
    [PDF] The method of exhaustion - UBC Mathematics
    The method of exhaustion is a technique that the classical Greek mathematicians used to prove results that would now be dealt with by means of limits.
  30. [30]
    Descartes' Mathematics - Stanford Encyclopedia of Philosophy
    Nov 28, 2011 · In La Géométrie, Descartes details a groundbreaking program for geometrical problem-solving—what he refers to as a “geometrical calculus” ( ...Descartes' Early Mathematical... · La Géométrie (1637) · Book One: Descartes...
  31. [31]
    René Descartes (1596 - 1650) - Biography - University of St Andrews
    René Descartes was a French philosopher whose work, La géométrie, includes his application of algebra to geometry from which we now have Cartesian geometry. His ...
  32. [32]
    Carl Friedrich Gauss (1777 - 1855) - Biography
    He published the book Disquisitiones Arithmeticae Ⓣ. (Investigations in arithmetic). in the summer of 1801. There were seven sections, all but the last section, ...
  33. [33]
    Gauss: "Disquisitiones Arithmeticae" - University of St Andrews
    In 1801 Carl Friedrich Gauss published his classic work Disquisitiones Arithmeticae. He was 24 years old. A second edition of Gauss' masterpiece appeared in ...
  34. [34]
    Peano Axioms - MacTutor History of Mathematics
    Peano's axioms for the Natural numbers · 1 is a natural number. · Every natural number n n n has a natural number n ′ n^\prime n′ as a successor. · 1 is not the ...
  35. [35]
    Giuseppe Peano (1858 - 1932) - Biography - University of St Andrews
    In 1889 Peano published his famous axioms, called Peano axioms, which defined the natural numbers in terms of sets. ... The Peano axioms are listed at THIS LINK.
  36. [36]
    [PDF] Proofs
    In addition, we have some variations of these basic styles of proofs. Page 6. Direct Proof (1) Definition: An integer n is called even if and only if there ...<|control11|><|separator|>
  37. [37]
    Proof Theory - Kent
    Definitions. Even Integer. An integer n is even if, and only if, n = 2k for some integer k. Symbolically,. n is even ↔ ∃ an integer k such that n = 2k.
  38. [38]
    [PDF] Proofs of Pythagorean Theorem - OU Math
    3 Proof by similar triangles. Let CH be the perpendicular from C to the side ... Using (1) and (2), we rewrite this as c = a2 c. + b2 c. , which is.
  39. [39]
    Euclid's Elements, Book I, Proposition 47 - Clark University
    If the rectilinear figures on the sides of the triangle are similar, then the figure on the hypotenuse is the sum of the other two figures. A bit of history.
  40. [40]
    [PDF] Some Sample Proofs
    3. Prove that the square of an odd number is odd. Proof: Let x be an odd number. Then we can write x = 2k + 1 for some integer k.
  41. [41]
    [PDF] Proof Techniques - CS Stanford
    If we want to prove something is true for all odd numbers (for example, that the square of any odd number is odd), we can pick an arbitrary odd number x, and ...
  42. [42]
    [PDF] Discrete Mathematics - (Proof Techniques)
    Proof by contraposition. Computer-aided proofs. Page 6. Introduction to ... [Hint: Direct proof, contradiction.] Let n be a positive integer. Prove that ...Missing: textbook | Show results with:textbook
  43. [43]
    [PDF] A Unified Framework for Proof and Disproof - DePaul University
    May 19, 2009 · The square of any odd integer is odd. PROOF. Suppose that x is any odd integer. Then x = 2k + 1 for some integer k, and so x² = (2k + 1)² ...
  44. [44]
    Euclid's Elements, Book I, Proposition 20 - Clark University
    This proposition is known as the triangle inequality. It is part of the statement that the shortest path between two points is a straight line.
  45. [45]
    [PDF] Study Note—Euclid's Elements, Book I, Proposition 20
    Let ABC be a triangle. We aim to show that BC < BA + AC. To prove this inequality, we produce the side BA of the triangle ABC in a straight line beyond A to ...
  46. [46]
    [PDF] 1 Structures on Euclidean Space
    Therefore, kx + yk≤kxk + kyk. The triangle inequality for the usual Euclidean distance also follows this theorem. Inner products and norms can be defined on ...
  47. [47]
    The Fundamental Theorem of Arithmetic
    Proof: We can prove the first part of this theorem (that every integer n ≥ 2 can be factored into a product of primes) with strong induction.
  48. [48]
    The Fundamental Theorem of Arithmetic
    Proof. ( Fundamental Theorem of Arithmetic) First, I'll use induction to show that every integer greater than 1 can be expressed as a product of primes.
  49. [49]
    5.3 Combinatorial Proofs
    We will introduce a new method of proof called, “combinatorial proof” in which we're able to verify mathematical statements by counting!
  50. [50]
    [PDF] Math 120A — Introduction to Group Theory - UCI Mathematics
    Closure, Associativity, Identity, Inverse. The order of a group is the cardinality (size) |G| of the underlying set.2. In addition, a group G is said to be ...
  51. [51]
    [PDF] Math 310.003 Polynomial Euclidean Algorithm Fall 2018 Division ...
    The Euclidean Algorithm finds the greatest common divisor of polynomials using repeated division with remainder, similar to how it works for integers.
  52. [52]
    The most common errors in undergraduate mathematics
    Another common error is to assume that multiplication commutes with differentiation or integration. ... A direct proof is acceptable. A theorem has certain ...
  53. [53]
    [PDF] 1. Math 353 Angle Sum of Triangles Professor Richard Blecksmith ...
    In Euclidean geometry, given 스ABC. mZA + mZB + mZC = 180. 14. Angle Sum = 180 Proof The idea of the proof is to use the line m through C parallel to ←→ AB.
  54. [54]
    [PDF] Lecture notes for April 10
    Apr 12, 2017 · If we take a step back, we see that we have just proved the fact that the angle sum in a triangle is 180◦, starting from the parallel postulate.
  55. [55]
    Calculus I - Proof of Various Limit Properties
    Nov 16, 2022 · In this section we are going to prove some of the basic properties and facts about limits that we saw in the Limits chapter.
  56. [56]
    [PDF] Basic Proof Techniques 1 Basic Notation 2 Proofs
    Aug 13, 2013 · A lot of definitions and proofs in real analysis use the " and δ" concept. For example, recall the definition of the limit of a function: We ...
  57. [57]
    [PDF] Topology of the Real Numbers - UC Davis Mathematics
    The idea of the proof is that any open cover of [a, x] must cover a larger interval since the open set that contains x extends past x. Since C covers [a, b] ...Missing: direct connectedness
  58. [58]
    [PDF] CONNECTED SPACES AND HOW TO USE THEM 1. How to prove ...
    All open and half-open intervals are connected; all rays are connected; a line is connected. Proof. This follows from the theorem (argue by contradiction). For ...
  59. [59]
    [PDF] Emmy Noether and Symmetry
    Noether's (First) Theorem states that to each continuous symmetry group of the action functional there is a corresponding conservation law of the physical ...
  60. [60]
    [PDF] General Rules for Loop Invariant Proofs
    We use loop invariants to help us understand why an algorithm is correct. We must show three things about a loop invariant: Initialization: It is true prior ...
  61. [61]
    [PDF] How to use induction and loop invariants to prove correctness 1 ...
    The loop invariant needs to have two properties: it needs to be self-justifying (it's not enough that it's true at all iterations; it's truth at one iteration ...
  62. [62]
  63. [63]
  64. [64]
    Why are direct proofs often considered better than indirect proofs?
    Sep 1, 2019 · The indirect case might be considered more elegant, but the direct proof provides an extra quantitative information. Theorem. Any continuous map ...
  65. [65]
    [PDF] Sources of Students' Difficulties with Proof By Contradiction
    Jul 20, 2021 · this proof technique is “more difficult” for students than direct proof, and offers ... Learning and teaching indirect proof. The Mathematics ...
  66. [66]
    Are indirect proofs less convincing? A study of students' comparative ...
    Aug 6, 2025 · Mathematics education researchers are nearly unanimous in maintaining that PBC is "more difficult" for students than direct proof (DP) (Tall, ...Missing: retention | Show results with:retention
  67. [67]
    Pre-Service Teachers' Knowledge of and Beliefs About Direct and ...
    Jun 20, 2024 · First, for situation-specific beliefs and knowledge, we found that indirect proofs seem to be more challenging than direct proofs. Second, for ...Missing: advantages | Show results with:advantages
  68. [68]
    [PDF] Proof construction and evaluation practices of prospective ... - ERIC
    indirect proof report that “students' difficulties with indirect proof seem to greater than those related with direct proof”, and assuming that what needs ...