SUPPORT THE WORK

GetWiki

0.999...

ARTICLE SUBJECTS
aesthetics  →
being  →
complexity  →
database  →
enterprise  →
ethics  →
fiction  →
history  →
internet  →
knowledge  →
language  →
licensing  →
linux  →
logic  →
method  →
news  →
perception  →
philosophy  →
policy  →
purpose  →
religion  →
science  →
sociology  →
software  →
truth  →
unix  →
wiki  →
ARTICLE TYPES
essay  →
feed  →
help  →
system  →
wiki  →
ARTICLE ORIGINS
critical  →
discussion  →
forked  →
imported  →
original  →
0.999...
[ temporary import ]
please note:
- the content below is remote from Wikipedia
- it has been imported raw for GetWiki
{{Short description|Alternative decimal expansion of 1}}{{Use dmy dates|date=May 2024}}(File:999 Perspective.svg|300px|class=skin-invert|thumbnail|Stylistic impression of the number, representing how its decimals go on infinitely)In mathematics, 0.999... (also written as 0.{{overline|9}}, 0.{{overset|.|9}} or 0.(9)) is a notation for the repeating decimal consisting of an unending sequence of 9s after the decimal point. This repeating decimal is a numeral that represents the smallest number no less than every number in the sequence (0.9, 0.99, 0.999, ldots) ; that is, the supremum of this sequence.{{efn|This definition is equivalent to the definition of decimal numbers as the limits of their summed components, which, in the case of 0.999..., is the limit of the sequence (0.9, 0.99, 0.999, ...). The equivalence is due to bounded increasing sequences having their limit always equal to their least upper bound.}} This number is equal to{{spaces}}1. In other words, “0.999...” is not “almost exactly 1” or “very, very nearly but not quite 1”; rather, “0.999...” and “1” represent {{em|exactly}} the same number.There are many ways of showing this equality, from intuitive arguments to mathematically rigorous proofs. The technique used depends on the target audience, background assumptions, historical context, and preferred development of the real numbers, the system within which 0.999... is commonly defined. In other systems, 0.999... can have the same meaning, a different definition, or be undefined.More generally, every nonzero terminating decimal has two equal representations (for example, 8.32000... and 8.31999...), which is a property of all positional numeral system representations regardless of base. The utilitarian preference for the terminating decimal representation contributes to the misconception that it is the only representation. For this and other reasons—such as rigorous proofs relying on non-elementary techniques, properties, or disciplines—some people can find the equality sufficiently counterintuitive that they question or reject it. This has been the subject of several studies in mathematics education.

Proof by adding and comparing decimal numbers

File:Archimedean_property_for_Achilles_and_Tortoise_example.svg|right|thumb|The (Archimedean property]]: any point {{math|1=x}} before the finish line lies between two of the points {{math|1=Pn}} (inclusive).|class=skin-invert)It is possible to prove the equation 0.999ldots = 1 using just the mathematical tools of comparison and addition of (finite) decimal numbers, without any reference to more advanced topics such as series, limits, or the formal construction of real numbers. The proof given below is a direct formalization of the intuitive fact that, if one draws 0.9 , 0.99 , 0.999 , etc. on the number line, there is no room left for placing a number between them and 1. The meaning of the notation 0.999... is the least point on the number line lying to the right of all of the numbers 0.9 , 0.99 , 0.999 , etc. Because there is ultimately no room between 1 and these numbers, the point 1 must be this least point, and so 0.999ldots = 1 .

Intuitive explanation

If one places 0.9 , 0.99 , 0.999 , etc. on the number line, one sees immediately that all these points are to the left of 1, and that they get closer and closer to 1. For any number x that is less than 1, the sequence 0.9 , 0.99 , 0.999 , and so on will eventually reach a number larger than x. So, it does not make sense to identify 0.999ldots with any number smaller than 1. Meanwhile, every number larger than 1 will be larger than any decimal of the form 0.999ldots9 for any finite number of nines. Therefore, 0.999ldots cannot be identified with any number larger than 1, either. Because 0.999ldots cannot be bigger than 1 or smaller than 1, it must equal 1 if it is to be any real number at all.{{sfnp|Cheng|2023|p=141}}{{sfnp|Diamond|1955}}

Rigorous proof

More precisely, the distance from 0.9 to 1 is 0.1 = frac{1}{10} , the distance from 0.99 to 1 is 0.01 = frac{1}{10^2} , and so on. The distance to 1 from the {{nowrap|1= n -th}} point (the one with n 9s after the decimal point) is frac{1}{10^n} .Therefore, if 1 were not the smallest number greater than 0.9 , 0.99 , 0.999 , etc., then there would be a point on the number line that lies between 1 and all these points. This point would be at a positive distance from 1 that is less than frac{1}{10^n} for every integer n . In the standard number systems (the rational numbers and the real numbers), there is no positive number that is less than frac{1}{10^n} for all n . This is (one version of) a property of these number systems known as the Archimedean property.{{sfnp|Baldwin|Norton|2012}}{{sfnp|Meier|Smith|2017|loc=§8.2}} Therefore, 1 is the smallest number that is greater than all 0.9 , 0.99 , 0.999 , etc., and so 1 = 0.999ldots .One way to show that the difference between 1 and 0.999ldots is less than the inverse of any positive integer is as follows. The number 0.999ldots9 , with n nines after the decimal point, is denoted 0.(9)_n . Thus 0.(9)_1 = 0.9 , 0.(9)_2 = 0.99 , 0.(9)_3 = 0.999 , and so on. As frac{1}{10^n} = 0.0ldots01 , with n digits after the decimal point, the addition rule for decimal numbers implies
begin{align}
0.(9)_n + 1/10^n &= 1,
0.(9)_n &< 1,
end{align} for every positive integer n . One has to show that 1 is the smallest number that is no less than all 0.(9)_n . For this, it suffices to prove that, if a number x is not larger than 1 and no less than all 0.(9)_n , then x = 1 . So let x such that
0.(9)_n le x le 1,
for every positive integer n . Therefore,
1 - 1le 1 - x le 1-0.(9)_n.
which, using basic arithmetic and the first equality established above, simplifies to
0le 1 - x le 1/10^n.
This implies that the difference between 1 and x is less than the inverse of any positive integer. Thus, this difference must be zero, and, thus x = 1 ; that is, 0.999ldots = 1 .This proof relies on the fact that zero is the only nonnegative number that is less than all inverses of integers, or equivalently that no number is larger than every integer. This is the Archimedean property, which holds for rational numbers and real numbers. Real numbers may be enlarged into number systems, such as hyperreal numbers, with infinitely small numbers (infinitesimals) and infinitely large numbers (infinite numbers).{{sfnp|Stewart|2009|p=175}}{{sfnp|Propp|2023}} When using such systems, the notation 0.999... is generally not used, as there is no smallest number among the numbers larger than all 0.(9)_n .{{efn|For example, one can show this as follows: if x is any number such that 0.(9)_n le x < 1 , then 0.(9)_{n - 1} le 10x - 9 < x < 1 . Thus if x has this property for all n, the smaller number 10x - 9 does, as well.}}

Least upper bounds and completeness

Part of what this argument shows is that there is a least upper bound of the sequence 0.9 , 0.99 , 0.999 , etc.: the smallest number that is greater than all of the terms of the sequence. One of the axioms of the real number system is the completeness axiom, which states that every bounded sequence has a least upper bound.{{sfnp|Stillwell|1994|p=42}}{{sfnp|Earl|Nicholson|2021|loc=“bound“}} This least upper bound is one way to define infinite decimal expansions: the real number represented by an infinite decimal is the least upper bound of its finite truncations.{{sfnp|Rosenlicht|1985|p=27}} The argument here does not need to assume completeness to be valid, because it shows that this particular sequence of rational numbers has a least upper bound and that this least upper bound is equal to one.

Algebraic arguments{{anchor|Proofs}}{{anchor|Algebraic}}

Many algebraic arguments have been provided, which suggest that 1=0.999ldots They are not rigorous mathematical proofs since they are typically based on the assumption that the rules for adding and multiplying finite decimals extend to infinite decimals. The extension of these rules to infinite decimals is both intuitive and correct, but it requires justification.Simple algebraic illustrations of equality are a subject of pedagogical discussion and critique. {{harvtxt|Byers|2007}} discusses the argument that, in elementary school, one is taught that frac{1}{3} = 0.333ldots , so, ignoring all essential subtleties, “multiplying” this identity by 3 gives 1 = 0.999ldots . He further says that this argument is unconvincing, because of an unresolved ambiguity over the meaning of the equals sign; a student might think, “It surely does not mean that the number 1 is identical to that which is meant by the notation 0.999ldots .” Most undergraduate mathematics majors encountered by Byers feel that while 0.999ldots is “very close” to 1 on the strength of this argument, with some even saying that it is “infinitely close”, they are not ready to say that it is equal to 1.{{sfnp|Byers|2007|p=39}} {{harvtxt|Richman|1999}} discusses how “this argument gets its force from the fact that most people have been indoctrinated to accept the first equation without thinking”, but also suggests that the argument may lead skeptics to question this assumption.{{sfnp|Richman|1999}}Byers also presents the following argument.{{block indent|1=begin{align}
x &= 0.999ldots
10x &= 9.999ldots && text{by multiplying by }10
10x &= 9+0.999ldots && text{by splitting off integer part}
10x &= 9 + x && text{by definition of }x
9x &= 9 && text{by subtracting }x
x &= 1 && text{by dividing by }9
end{align}}}Students who did not accept the first argument sometimes accept the second argument, but, in Byers’s opinion, still have not resolved the ambiguity, and therefore do not understand the representation of infinite decimals. {{harvtxt|Peressini|Peressini|2007}}, presenting the same argument, also state that it does not explain the equality, indicating that such an explanation would likely involve concepts of infinity and completeness.{{sfnp|Peressini|Peressini|2007|p=186}} {{harvtxt|Baldwin|Norton|2012}}, citing {{harvtxt|Katz|Katz|2010a}}, also conclude that the treatment of the identity based on such arguments as these, without the formal concept of a limit, is premature.{hide}sfnmp
| 1a1 = Baldwin | 1a2 = Norton | 1y = 2012
| 2a1 = Katz | 2a2 = Katz | 2y = 2010a
{edih} {{harvtxt|Cheng|2023}} concurs, arguing that knowing one can multiply 0.999ldots by 10 by shifting the decimal point presumes an answer to the deeper question of how one gives a meaning to the expression 0.999ldots at all.{{sfnp|Cheng|2023|p=136}}The same argument is also given by {{harvtxt|Richman|1999}}, who notes that skeptics may question whether x is cancellable{{snd}} that is, whether it makes sense to subtract x from both sides.{{sfnp|Richman|1999}}

Analytic proofs{{anchor|Analytic}}

Since the question of 0.999... does not affect the formal development of mathematics, it can be postponed until one proves the standard theorems of real analysis. One requirement is to characterize real numbers that can be written in decimal notation, consisting of an optional sign, a finite sequence of one or more digits forming an integer part, a decimal separator, and a sequence of digits forming a fractional part. To discuss 0.999..., the integer part can be summarized as b_0 and one can neglect negatives, so a decimal expansion has the form
b_0.b_1b_2b_3b_4b_5 dots.
The fraction part, unlike the integer part, is not limited to finitely many digits. This is a positional notation, so for example the digit 5 in 500 contributes ten times as much as the 5 in 50, and the 5 in 0.05 contributes one tenth as much as the 5 in 0.5.

Infinite series and sequences

{{further|Decimal representation}}A common development of decimal expansions is to define them as sums of infinite series. In general:
b_0 . b_1 b_2 b_3 b_4 ldots = b_0 + b_1left({tfrac{1}{10}}right) + b_2left({tfrac{1}{10}}right)^2 + b_3left({tfrac{1}{10}}right)^3 + b_4left({tfrac{1}{10}}right)^4 + cdots .
For 0.999... one can apply the convergence theorem concerning geometric series, stating that if |r| < 1 , then:{{sfnmp
| 1a1 = Rudin | 1y = 1976 | 1p = 61 | 1loc = Theorem 3.26
| 2a1 = Stewart | 2y = 1999 | 2p = 706
}}
ar + ar^2 + ar^3 + cdots = frac{ar}{1-r}.
Since 0.999... is such a sum with a = 9 and common ratio r = frac{1}{10} , the theorem makes short work of the question:
0.999ldots = 9left(tfrac{1}{10}right) + 9left({tfrac{1}{10}}right)^2 + 9left({tfrac{1}{10}}right)^3 + cdots = frac{9left({tfrac{1}{10}}right)}{1-{tfrac{1}{10}}} = 1.
This proof appears as early as 1770 in Leonhard Euler’s Elements of Algebra.{{sfnp|Euler|1822|p=170}}File:base4 333.svg|right|thumb|200px|Limits: The unit interval, including the base-4base-4The sum of a geometric series is itself a result even older than Euler. A typical 18th-century derivation used a term-by-term manipulation similar to the algebraic proof given above, and as late as 1811, Bonnycastle’s textbook An Introduction to Algebra uses such an argument for geometric series to justify the same maneuver on 0.999...{hide}sfnmp
| 1a1 = Grattan-Guinness | 1y = 1970 | 1p = 69
| 2a1 = Bonnycastle |2y=1806 | 2p = 177
{edih} A 19th-century reaction against such liberal summation methods resulted in the definition that still dominates today: the sum of a series is defined to be the limit of the sequence of its partial sums. A corresponding proof of the theorem explicitly computes that sequence; it can be found in any proof-based introduction to calculus or analysis.{hide}sfnmp
| 1a1 = Stewart | 1y = 1999 | 1p = 706
| 2a1 = Rudin | 2y = 1976 | 2p = 61
| 3a1 = Protter | 3a2 = Morrey | 3y = 1991 | 3p = 213
| 4a1 = Pugh | 4y = 2002 | 4p = 180
| 5a1 = Conway | 5y = 1978 | 5p = 31
{edih}A sequence (x_0, x_1, x_2, ldots) has the value x as its limit if the distance left | x - x_n right | becomes arbitrarily small as n increases. The statement that 0.999ldots = 1 can itself be interpreted and proven as a limit:{{efn|The limit follows, for example, from {{harvtxt|Rudin|1976}}, p. 57, Theorem 3.20e. For a more direct approach, see also {{harvtxt|Finney|Weir|Giordano|2001}}, section 8.1, example 2(a), example 6(b).}}
0.999ldots overset{underset{mathrm{def}}{}}{=} lim_{ntoinfty}0.underbrace{ 99ldots9 }_{n} overset{underset{mathrm{def}}{}}{=} lim_{ntoinfty}sum_{k = 1}^nfrac{9}{10^k} = lim_{ntoinfty}left(1-frac{1}{10^n}right) = 1-lim_{ntoinfty}frac{1}{10^n} = 1 - 0 = 1.
The first two equalities can be interpreted as symbol shorthand definitions. The remaining equalities can be proven. The last step, that 10^n approaches to 0 as n approaches to infinity ( infty ), is often justified by the Archimedean property of the real numbers. This limit-based attitude towards 0.999ldots is often put in more evocative but less precise terms. For example, the 1846 textbook The University Arithmetic explains, ”.999 +, continued to infinity = 1, because every annexation of a 9 brings the value closer to 1”; the 1895 Arithmetic for Schools says, “when a large number of 9s is taken, the difference between 1 and .99999... becomes inconceivably small”.{hide}sfnmp
| 1a1 = Davies | 1y = 1846 | 1p = 175
| 2a1 = Smith | 2a2 = Harrington | 2y = 1895 | 2p = 115
{edih} Such heuristics are often incorrectly interpreted by students as implying that {{math|0.999...}} itself is less than 1.

Nested intervals and least upper bounds

{{further|Nested intervals}}(File:999 Intervals C.svg|right|thumb|250px|Nested intervals: in base 3, {{math|1=1 = 1.000... = 0.222...}}.|class=skin-invert)The series definition above defines the real number named by a decimal expansion. A complementary approach is tailored to the opposite process: for a given real number, define the decimal expansion(s) to name it.If a real number {{mvar|x}} is known to lie in the closed interval [0, 10] (that is, it is greater than or equal to 0 and less than or equal to 10), one can imagine dividing that interval into ten pieces that overlap only at their endpoints: [0, 1] , [1, 2] , [2, 3] , and so on up to [9, 10] . The number x must belong to one of these; if it belongs to [2, 3], then one records the digit “2” and subdivides that interval into [2, 2.1] , [2.1, 2.2] , ..., [2.8, 2.9] , [2.9, 3] . Continuing this process yields an infinite sequence of nested intervals, labeled by an infinite sequence of digits b_1, b_2, b_3, ldots , and one writes
x = b_0.b_1b_2b_3 ldots.
In this formalism, the identities 1 = 0.999... and 1 = 1.000... reflect, respectively, the fact that 1 lies in both [0, 1] . and [1, 2] , so one can choose either subinterval when finding its digits. To ensure that this notation does not abuse the “=” sign, one needs a way to reconstruct a unique real number for each decimal. This can be done with limits, but other constructions continue with the ordering theme.{hide}sfnmp
| 1a1 = Beals | 1y = 2004 | 1p = 22
| 2a1 = Stewart | 2y = 2009 | 2p = 34
{edih}One straightforward choice is the nested intervals theorem, which guarantees that given a sequence of nested, closed intervals whose lengths become arbitrarily small, the intervals contain exactly one real number in their intersection. So b_1, b_2, b_3, ldots is defined to be the unique number contained within all the intervals [b_0, b_0 + 1] , [b_0. b_1, b_0. b_1 + 0.1] , and so on. 0.999... is then the unique real number that lies in all of the intervals [0, 1] , [0.9, 1] , [0.99, 1] , and [0.99...9, 1] for every finite string of 9s. Since 1 is an element of each of these intervals, 0.999... = 1 .{hide}sfnmp
| 1a1 = Bartle | 1a2 = Sherbert | 1y = 1982 | 1pp = 60–62
| 2a1 = Pedrick | 2y = 1994 | 2p = 29
| 3a1 = Sohrab | 3y = 2003 | 3p = 46
{edih}The nested intervals theorem is usually founded upon a more fundamental characteristic of the real numbers: the existence of least upper bounds or suprema. To directly exploit these objects, one may define b_0.b_1 b_2 b_3 ldots to be the least upper bound of the set of approximants b_0, b_0. b_1, b_0. b_1 b_2, ldots .{hide}sfnmp
| 1a1 = Apostol | 1y = 1974 | 1pp = 9, 11–12
| 2a1 = Beals | 2y= 2004 | 2p = 22
| 3a1 = Rosenlicht | 3y = 1985 | 3p = 27
{edih} One can then show that this definition (or the nested intervals definition) is consistent with the subdivision procedure, implying 0.999... = 1 again. Tom Apostol concludes, “the fact that a real number might have two different decimal representations is merely a reflection of the fact that two different sets of real numbers can have the same supremum.“{{sfnp|Apostol|1974|p=12}}

Proofs from the construction of the real numbers{{anchor|Based on the construction of the real numbers}}

{{further|Construction of the real numbers}}Some approaches explicitly define real numbers to be certain structures built upon the rational numbers, using axiomatic set theory. The natural numbers {0, 1, 2, 3, ldots} begin with 0 and continue upwards so that every number has a successor. One can extend the natural numbers with their negatives to give all the integers, and to further extend to ratios, giving the rational numbers. These number systems are accompanied by the arithmetic of addition, subtraction, multiplication, and division.{{sfnp|Cheng|2023|pp=153–156}}{{sfnp|Conway|2001|pp=25–27}} More subtly, they include ordering, so that one number can be compared to another and found to be less than, greater than, or equal to another number.{{sfnp|Rudin|1976|pp=3,8}}The step from rationals to reals is a major extension. There are at least two popular ways to achieve this step, both published in 1872: Dedekind cuts and Cauchy sequences. Proofs that 0.999ldots = 1 that directly uses these constructions are not found in textbooks on real analysis, where the modern trend for the last few decades has been to use an axiomatic analysis. Even when a construction is offered, it is usually applied toward proving the axioms of the real numbers, which then support the above proofs. However, several authors express the idea that starting with a construction is more logically appropriate, and the resulting proofs are more self-contained.{{efn|The historical synthesis is claimed by {{harvtxt|Griffiths|Hilton|1970}}, p. xiv and again by {{harvtxt|Pugh|2002}}, p. 10; both actually prefer Dedekind cuts to axioms. For the use of cuts in textbooks, see {{harvtxt|Pugh|2002}}, p. 17 or {{harvtxt|Rudin|1976}}, p. 17. For viewpoints on logic, see {{harvtxt|Pugh|2002}}, p. 10, {{harvtxt|Rudin|1976}}, p.ix, or {{harvtxt|Munkres|2000}}, p. 30.}}

Dedekind cuts

{{further|Dedekind cut}}In the Dedekind cut approach, each real number x is defined as the infinite set of all rational numbers less than x .{{efn|{{harvtxt|Enderton|1977}}, p. 113 qualifies this description: “The idea behind Dedekind cuts is that a real number x can be named by giving an infinite set of rationals, namely all the rationals less than x . We will in effect define x to be the set of rationals smaller than x . To avoid circularity in the definition, we must be able to characterize the sets of rationals obtainable in this way...“}} In particular, the real number 1 is the set of all rational numbers that are less than 1.{{efn|{{harvtxt|Rudin|1976}}, pp. 17–20, {{harvtxt|Richman|1999}}, p. 399, or {{harvtxt|Enderton|1977}}, p. 119. To be precise, Rudin, Richman, and Enderton call this cut 1* , 1_- , and 1_R , respectively; all three identify it with the traditional real number 1. Note that what Rudin and Enderton call a Dedekind cut, Richman calls a “non-principal Dedekind cut”.}} Every positive decimal expansion easily determines a Dedekind cut: the set of rational numbers that are less than some stage of the expansion. So the real number 0.999... is the set of rational numbers r such that r < 0 , or r < 0.9 , or r < 0.99 , or r is less than some other number of the form{{sfnp|Richman|1999|p=399}}
1-frac{1}{10^n} = 0.(9)_n = 0.underbrace{99ldots 9}_{ntext{ nines}}.
Every element of 0.999... is less than 1, so it is an element of the real number 1. Conversely, all elements of 1 are rational numbers that can be written as
frac{a}{b} 0 and b > a . This implies
1-frac{a}{b} = frac{b-a}{b} ge frac{1}{b} > frac{1}{10^b},
and thus
frac{a}{b} < 1 - frac{1}{10^b}.
Since
1-frac{1}{10^b} = 0.(9)_b < 0.999ldots,
by the definition above, every element of 1 is also an element of 0.999..., and, combined with the proof above that every element of 0.999... is also an element of 1, the sets 0.999... and 1 contain the same rational numbers, and are therefore the same set, that is, 0.999... = 1 .The definition of real numbers as Dedekind cuts was first published by Richard Dedekind in 1872.{{sfnp|O’Connor|Robertson|2005}}The above approach to assigning a real number to each decimal expansion is due to an expository paper titled “Is 0.999 ... = 1?” by Fred Richman in Mathematics Magazine.{{sfnp|Richman|1999}} Richman notes that taking Dedekind cuts in any dense subset of the rational numbers yields the same results; in particular, he uses decimal fractions, for which the proof is more immediate. He also notes that typically the definitions allow {x colon x < 1 } to be a cut but not {x colon x le 1 } (or vice versa).{{harvtxt|Richman|1999}}, p. 398–399. “Why do that? Precisely to rule out the existence of distinct numbers 0.overline{9} and 1 . [...] So we see that in the traditional definition of the real numbers, the equation 0.overline{9} = 1 is built in at the beginning.” A further modification of the procedure leads to a different structure where the two are not equal. Although it is consistent, many of the common rules of decimal arithmetic no longer hold, for example, the fraction frac{1}{3} has no representation; see {{slink||Alternative number systems}} below.

Cauchy sequences

{{further|Cauchy sequence}}Another approach is to define a real number as the limit of a Cauchy sequence of rational numbers. This construction of the real numbers uses the ordering of rationals less directly. First, the distance between x and y is defined as the absolute value left | x - y right | , where the absolute value left | z right | is defined as the maximum of z and -z , thus never negative. Then the reals are defined to be the sequences of rationals that have the Cauchy sequence property using this distance. That is, in the sequence x_0, x_1, x_2, ldots , a mapping from natural numbers to rationals, for any positive rational delta there is an N such that left | x_m - x_n right | le delta for all m, n > N ; the distance between terms becomes smaller than any positive rational.{{sfnp|Griffiths|Hilton|1970|p=386|loc=§24.2 “Sequences“}}If (x_n) and (y_n) are two Cauchy sequences, then they are defined to be equal as real numbers if the sequence (x_n - y_n) has the limit 0. Truncations of the decimal number b_0. b_1 b_2 b_3 ldots generate a sequence of rationals, which is Cauchy; this is taken to define the real value of the number.{{sfnp|Griffiths|Hilton|1970|pp=388, 393}} Thus in this formalism the task is to show that the sequence of rational numbers
left(1 - 0, 1 - {9 over 10}, 1 - {99 over 100}, dotsright) = left(1, {1 over 10}, {1 over 100}, dots right)has a limit 0. Considering the {{nowrap|1= n -}}th term of the sequence, for n in mathbb{N}, it must therefore be shown that
lim_{nrightarrowinfty}frac{1}{10^n} = 0.
This can be proved by the definition of a limit. So again, 0.999ldots = 1 .{{sfnp|Griffiths|Hilton|1970|p=395}}The definition of real numbers as Cauchy sequences was first published separately by Eduard Heine and Georg Cantor, also in 1872.{{sfnp|O’Connor|Robertson|2005}} The above approach to decimal expansions, including the proof that 0.999ldots = 1 , closely follows Griffiths & Hilton’s 1970 work A comprehensive textbook of classical mathematics: A contemporary interpretation.{{sfnp|Griffiths|Hilton|1970|pp=viii, 395}}

Infinite decimal representation

{{further|Construction of the real numbers#Stevin’s construction}}Commonly in secondary schools’ mathematics education, the real numbers are constructed by defining a number using an integer followed by a radix point and an infinite sequence written out as a string to represent the fractional part of any given real number. In this construction, the set of any combination of an integer and digits after the decimal point (or radix point in non-base 10 systems) is the set of real numbers. This construction can be rigorously shown to satisfy all of the real axioms after defining an equivalence relation over the set that defines 1 =eq 0.999... as well as for any other nonzero decimals with only finitely many nonzero terms in the decimal string with its trailing 9s version.{{sfnp|Li|2011}} With this construction of the reals, all proofs of the statement “1 = 0.999...” can be viewed as implicitly assuming equality when any operations are performed on the real numbers.

Dense order

{{further|Dense order}}One of the notions that can resolve the issue is the requirement that real numbers be densely ordered. Dense ordering implies that if there is no new element strictly between two elements of the set, the two elements must be considered equal. Therefore, if 0.99999... were to be different from 1, there would have to be another real number in between them but there is none: a single digit cannot be changed in either of the two to obtain such a number.{{sfnp|Artigue|2002|p=212|loc=“... the ordering of the real numbers is recognized as a dense order. However, depending on the context, students can reconcile this property with the existence of numbers just before or after a given number (0.999... is thus often seen as the predecessor of 1).“}}

Generalizations

The result that 0.999ldots = 1 generalizes readily in two ways. First, every nonzero number with a finite decimal notation (equivalently, endless trailing 0s) has a counterpart with trailing 9s. For example, 0.24999... equals 0.25, exactly as in the special case considered. These numbers are exactly the decimal fractions, and they are dense.{{sfnp|Petkovšek|1990|p=408}}{{sfnp|Rosenlicht|1985|p=27}}Second, a comparable theorem applies in each radix or base. For example, in base 2 (the binary numeral system) 0.111... equals 1, and in base 3 (the ternary numeral system) 0.222... equals 1. In general, any terminating base b expression has a counterpart with repeated trailing digits equal to b - 1 . Textbooks of real analysis are likely to skip the example of 0.999... and present one or both of these generalizations from the start.{hide}sfnmp
| 1a1 = Protter | 1a2 = Morrey | 1y = 1991 | 1p = 503
| 2a1 = Bartle | 2a2 = Sherbert | 2y = 1982 | 2p = 61
{edih}Alternative representations of 1 also occur in non-integer bases. For example, in the golden ratio base, the two standard representations are 1.000ldots and 0.101010ldots , and there are infinitely many more representations that include adjacent 1s. Generally, for almost all q between 1 and 2, there are uncountably many {{nowrap|1=base- q }} expansions of 1. In contrast, there are still uncountably many q , including all natural numbers greater than 1, for which there is only one {{nowrap|1=base- q }} expansion of 1, other than the trivial 1.000ldots . This result was first obtained by Paul Erdős, Miklos Horváth, and István Joó around 1990. In 1998 Vilmos Komornik and Paola Loreti determined the smallest such base, the Komornik–Loreti constant q = 1.787231650ldots . In this base, 1 = 0.11010011001011010010110011010011ldots ; the digits are given by the Thue–Morse sequence, which does not repeat.{{sfnp|Komornik|Loreti|1998|p=636}}A more far-reaching generalization addresses the most general positional numeral systems. They too have multiple representations, and in some sense, the difficulties are even worse. For example:{hide}sfnmp
| 1a1 = Kempner | 1y = 1936 | 1p = 611
| 2a1 = Petkovšek | 2y = 1990 | 2p = 409
{edih}
  • In the balanced ternary system, frac{1}{2} = 0.111ldots = 1.underline{111}ldots .
  • In the reverse factorial number system (using bases 2!, 3!, 4!, ldots for positions after the decimal point), 1 = 1.000ldots = 0.1234ldots .

Impossibility of unique representation

That all these different number systems suffer from multiple representations for some real numbers can be attributed to a fundamental difference between the real numbers as an ordered set and collections of infinite strings of symbols, ordered lexicographically. Indeed, the following two properties account for the difficulty:
  • If an interval of the real numbers is partitioned into two non-empty parts L , R , such that every element of L is (strictly) less than every element of R , then either L contains a largest element or R contains a smallest element, but not both.
  • The collection of infinite strings of symbols taken from any finite “alphabet”, lexicographically ordered, can be partitioned into two non-empty parts L , R , such that every element of L is less than every element of R , while L contains the largest element and R contains the smallest element. Indeed, it suffices to take two finite prefixes (initial substrings) p_1, p_2 of elements from the collection such that they differ only in their final symbol, for which symbol they have successive values, and take for L the set of all strings in the collection whose corresponding prefix is at most p_1 , and for R the remainder, the strings in the collection whose corresponding prefix is at least p_2 . Then L has the largest element, starting with p_1 and choosing the largest available symbol in all following positions. In contrast, R has the smallest element obtained by following p_2 by the smallest symbol in all positions.
The first point follows from the basic properties of real numbers: L has a supremum and R has an infimum, which are easily seen to be equal; being a real number it either lies in R or in L , but not both since L and R are supposed to be disjoint. The second point generalizes the 0.999.../1.000... pair obtained for p_1 = 0 , p_2 = 1 . One need not use the same alphabet for all positions (so that for instance mixed radix systems can be included) or consider the full collection of possible strings; the only important points are that at each position a finite set of symbols (which may even depend on the previous symbols) can be chosen from (this is needed to ensure maximal and minimal choices), and that making a valid choice for any position should result in a valid infinite string (so one should not allow “9” in each position while forbidding an infinite succession of “9“s). Under these assumptions, the above argument shows that an order preserving map from the collection of strings to an interval of the real numbers cannot be a bijection: either some numbers do not correspond to any string, or some of them correspond to more than one string.{{harvtxt|PetkovÅ¡ek|1990}} has proven that for any positional system that names all the real numbers, the set of reals with multiple representations is always dense. He calls the proof “an instructive exercise in elementary point-set topology”; it involves viewing sets of positional values as Stone spaces and noticing that their real representations are given by continuous functions.{{sfnp|PetkovÅ¡ek|1990|pp=410–411}}

Applications

One application of 0.999... as a representation of 1 occurs in elementary number theory. In 1802, H. Goodwyn published an observation on the appearance of 9s in the repeating-decimal representations of fractions whose denominators are certain prime numbers.{{sfnmp|1a1=Goodwyn|1y=1802|2a1=Dickson|2y=1919|2pp=161}} Examples include:
  • frac{1}{7} = 0.overline{142857} and 142 + 857 = 999 .
  • frac{1}{73} = 0.overline{01369863} and 0136 + 9863 = 9999 .
E. Midy proved a general result about such fractions, now called Midy’s theorem, in 1836. The publication was obscure, and it is unclear if his proof directly involved 0.999..., but at least one modern proof by W. G. Leavitt does. If it can be proved that if a decimal of the form 0.b_1 b_2 b_3 ldots is a positive integer, then it must be 0.999..., which is then the source of the 9s in the theorem.{{sfnp|Leavitt|1984|p=301}} Investigations in this direction can motivate such concepts as greatest common divisors, modular arithmetic, Fermat primes, order of group elements, and quadratic reciprocity.{hide}sfnmp
| 1a1 = Ginsberg | 1y = 2004 | 1pp = 26–30
| 2a1 = Lewittes | 2y = 2006 | 2pp = 1–3
| 3a1 = Leavitt | 3y = 1967 | 3pp = 669, 673
| 4a1 = Shrader-Frechette | 4y = 1978 | 4pp = 96–98
{edih}File:Cantor base 3.svg|right|class=skin-invert|thumb|Positions of {{frac|1|4}}, {{frac|2|3}}, and 1 in the Cantor setCantor setReturning to real analysis, the base-3 analogue 0.222ldots = 1 plays a key role in the characterization of one of the simplest fractals, the middle-thirds Cantor set: a point in the unit interval lies in the Cantor set if and only if it can be represented in ternary using only the digits 0 and 2.The {{nowrap|1= n -}}th digit of the representation reflects the position of the point in the {{nowrap|1= n -}}th stage of the construction. For example, the point frac{2}{3} is given the usual representation of 0.2 or 0.2000..., since it lies to the right of the first deletion and the left of every deletion thereafter. The point frac{1}{3} is represented not as 0.1 but as 0.0222..., since it lies to the left of the first deletion and the right of every deletion thereafter.{hide}sfnmp
| 1a1 = Pugh | 1y = 2002 | 1p = 97
| 2a1 = Alligood | 2a2 = Sauer | 2a3 = Yorke | 2y = 1996 | 2pp = 150–152
| 3a1 = Protter | 3a2 = Morrey | 3y = 1991 | 3p = 507
| 4a1 = Pedrick | 4y = 1994 | 4p = 29
{edih}Repeating nines also turns up in yet another of Georg Cantor’s works. They must be taken into account to construct a valid proof, applying his 1891 diagonal argument to decimal expansions, of the uncountability of the unit interval. Such a proof needs to be able to declare certain pairs of real numbers to be different based on their decimal expansions, so one needs to avoid pairs like 0.2 and 0.1999... A simple method represents all numbers with nonterminating expansions; the opposite method rules out repeating nines.{{efn|{{harvtxt|Maor|1987}}, p. 60 and {{harvtxt|Mankiewicz|2000}}, p. 151 review the former method; Mankiewicz attributes it to Cantor, but the primary source is unclear. {{harvtxt|Munkres|2000}}, p. 50 mentions the latter method.}} A variant that may be closer to Cantor’s original argument uses base 2, and by turning base-3 expansions into base-2 expansions, one can prove the uncountability of the Cantor set as well.{hide}sfnmp
| 1a1 = Rudin | 1y = 1976 | 1p = 50
| 2a1 = Pugh | 2y = 2002 | 2p = 98
{edih}

Skepticism in education

Students of mathematics often reject the equality of 0.999... and 1, for reasons ranging from their disparate appearance to deep misgivings over the limit concept and disagreements over the nature of infinitesimals. There are many common contributing factors to the confusion:
  • Students are often “mentally committed to the notion that a number can be represented in one and only one way by a decimal.” Seeing two manifestly different decimals representing the same number appears to be a paradox, which is amplified by the appearance of the seemingly well-understood number 1.{{efn|{{harvtxt|Bunch|1982}}, p. 119; {{harvtxt|Tall|Schwarzenberger|1978}}, p. 6. The last suggestion is due to {{harvtxt|Burrell|1998}}, p. 28: “Perhaps the most reassuring of all numbers is 1 ... So it is particularly unsettling when someone tries to pass off 0.9~ as 1.“}}
  • Some students interpret “0.999...” (or similar notation) as a large but finite string of 9s, possibly with a variable, unspecified length. If they accept an infinite string of nines, they may still expect a last 9 “at infinity”.{hide}sfnmp


| 1a1 = Tall | 1a2 = Schwarzenberger | 1y = 1978 | 1pp = 6–7
| 2a1 = Tall | 2y = 2000 | 2p = 221
{edih}
  • Intuition and ambiguous teaching lead students to think of the limit of a sequence as a kind of infinite process rather than a fixed value since a sequence need not reach its limit. Where students accept the difference between a sequence of numbers and its limit, they might read “0.999...” as meaning the sequence rather than its limit.{hide}sfnmp


| 1a1 = Tall | 1a2 = Schwarzenberger | 1y = 1978 | 1p = 6
| 2a1 = Tall | 2y = 2000 | 2p = 221
{edih}These ideas are mistaken in the context of the standard real numbers, although some may be valid in other number systems, either invented for their general mathematical utility or as instructive counterexamples to better understand 0.999...Many of these explanations were found by David Tall, who has studied characteristics of teaching and cognition that lead to some of the misunderstandings he has encountered with his college students. Interviewing his students to determine why the vast majority initially rejected the equality, he found that “students continued to conceive of 0.999... as a sequence of numbers getting closer and closer to 1 and not a fixed value, because ‘you haven’t specified how many places there are’ or ‘it is the nearest possible decimal below 1{{’“}}.{{sfnp|Tall|2000|p=221}}The elementary argument of multiplying 0.333ldots = frac{1}{3} by 3 can convince reluctant students that 0.999ldots = 1 . Still, when confronted with the conflict between their belief in the first equation and their disbelief in the second, some students either begin to disbelieve the first equation or simply become frustrated.{{sfnp|Tall|1976|pp=10–14}} Nor are more sophisticated methods foolproof: students who are fully capable of applying rigorous definitions may still fall back on intuitive images when they are surprised by a result in advanced mathematics, including 0.999ldots . For example, one real analysis student was able to prove that 0.333ldots = frac{1}{3} using a supremum definition but then insisted that 0.999ldots < 1 based on her earlier understanding of long division.{hide}sfnmp
| 1a1 = Pinto | 1a2 = Tall | 1y = 2001 | 1p = 5
| 2a1 = Edwards | 2a2 = Ward | 2y = 2004 | 2pp = 416–417
{edih} Others still can prove that frac{1}{3} = 0.333ldots , but, upon being confronted by the fractional proof, insist that “logic” supersedes the mathematical calculations.{{harvtxt|Mazur|2005}} tells the tale of an otherwise brilliant calculus student of his who “challenged almost everything I said in class but never questioned his calculator,” and who had come to believe that nine digits are all one needs to do mathematics, including calculating the square root of 23. The student remained uncomfortable with a limiting argument that 9.99ldots = 10 , calling it a “wildly imagined infinite growing process.“{{sfnp|Mazur|2005|pp=137–141}}As part of Ed Dubinsky’s APOS theory of mathematical learning, he and his collaborators (2005) propose that students who conceive of 0.999... as a finite, indeterminate string with an infinitely small distance from 1 have “not yet constructed a complete process conception of the infinite decimal”. Other students who have a complete process conception of 0.999... may not yet be able to “encapsulate” that process into an “object conception”, like the object conception they have of 1, and so they view the process 0.999... and the object 1 as incompatible. Dubinsky et al. also link this mental ability of encapsulation to viewing frac{1}{3} as a number in its own right and to dealing with the set of natural numbers as a whole.{{sfnp|Dubinsky|Weller|McDonald|Brown|2005|pp=261–262}}

Cultural phenomenon

With the rise of the Internet, debates about 0.999... have become commonplace on newsgroups and message boards, including many that nominally have little to do with mathematics. In the newsgroup {{mono|sci.math}}, arguing over 0.999... is described as a “popular sport”, and it is one of the questions answered in its FAQ.{{sfnp|Richman|1999|p=396}}{{sfnp|de Vreught|1994}} The FAQ briefly covers frac{1}{3} , multiplication by 10, and limits, and it alludes to Cauchy sequences as well.A 2003 edition of the general-interest newspaper column The Straight Dope discusses 0.999... via frac{1}{3} and limits, saying of misconceptions,{{blockquote|text=The lower primate in us still resists, saying: .999~ doesn’t really represent a number, then, but a process. To find a number we have to halt the process, at which point the .999~ = 1 thing falls apart.Nonsense.{{sfnp|Adams|2003}}}}A Slate article reports that the concept of 0.999... is “hotly disputed on websites ranging from World of Warcraft message boards to Ayn Rand forums”.WEB,www.slate.com/blogs/how_not_to_be_wrong/2014/06/06/does_0_999_1_and_are_divergent_series_the_invention_of_the_devil.html, Slate, Ellenberg, Jordan, Does 0.999... = 1? And Are Divergent Series the Invention of the Devil?, 6 June 2014, In the same vein, the question of 0.999... proved such a popular topic in the first seven years of Blizzard Entertainment’s Battle.net forums that the company issued a “press release” on April Fools’ Day 2004 that it is 1:{{blockquote|We are very excited to close the book on this subject once and for all. We’ve witnessed the heartache and concern over whether .999~ does or does not equal 1, and we’re proud that the following proof finally and conclusively addresses the issue for our customers.PRESS RELEASE,us.blizzard.com/en-us/company/press/pressreleases.html?040401, Blizzard Entertainment Announces .999~ (Repeating) = 1, Blizzard Entertainment, 1 April 2004, 2009-11-16,us.blizzard.com/en-us/company/press/pressreleases.html?040401," title="web.archive.org/web/20091104222830us.blizzard.com/en-us/company/press/pressreleases.html?040401,">web.archive.org/web/20091104222830us.blizzard.com/en-us/company/press/pressreleases.html?040401, 4 November 2009, }}Two proofs are then offered, based on limits and multiplication by 10.{{blockquote|Q: How many mathematicians does it take to screw in a lightbulb?A: 0.999999....}}

In alternative number systems{{anchor|Alternative number systems}}

Although the real numbers form an extremely useful number system, the decision to interpret the notation “0.999...” as naming a real number is ultimately a convention, and Timothy Gowers argues in Mathematics: A Very Short Introduction that the resulting identity 0.999ldots = 1 is a convention as well:{{blockquote|However, it is by no means an arbitrary convention, because not adopting it forces one either to invent strange new objects or to abandon some of the familiar rules of arithmetic.{{sfnp|Gowers|2002|p=60}}}}

Infinitesimals

Some proofs that 0.999ldots = 1 rely on the Archimedean property of the real numbers: that there are no nonzero infinitesimals. Specifically, the difference 1 - 0.999ldots must be smaller than any positive rational number, so it must be an infinitesimal; but since the reals do not contain nonzero infinitesimals, the difference is zero, and therefore the two values are the same.However, there are mathematically coherent ordered algebraic structures, including various alternatives to the real numbers, which are non-Archimedean. Non-standard analysis provides a number system with a full array of infinitesimals (and their inverses).{{efn|For a full treatment of non-standard numbers, see {{harvtxt|Robinson|1996}}.}} A. H. Lightstone developed a decimal expansion for hyperreal numbers in (0, 1)∗. Lightstone shows how to associate each number with a sequence of digits,
0.d_1d_2d_3 dots;dots d_{infty - 1}d_infty d_{infty + 1}dots,
indexed by the hypernatural numbers. While he does not directly discuss 0.999..., he shows the real number frac{1}{3} is represented by 0.333ldots;ldots 333 ldots , which is a consequence of the transfer principle. As a consequence the number 0.999ldots ; ldots 999 ldots = 1 . With this type of decimal representation, not every expansion represents a number. In particular ” 0.333ldots ; ldots 000 ldots ” and ” 0.999 ldots ; ldots 000 ldots ” do not correspond to any number.{{sfnp|Lightstone|1972|pp=245–247}}The standard definition of the number 0.999... is the limit of the sequence 0.9, 0.99, 0.999, ldots . A different definition involves an ultralimit, i.e., the equivalence class [(0.9, 0.99, 0.999, ldots)] of this sequence in the ultrapower construction, which is a number that falls short of 1 by an infinitesimal amount.{{sfnp|Tao|2012|pp=156–180}} More generally, the hyperreal number u_H = 0.999 ldots ; ldots 999000 ldots , with last digit 9 at infinite hypernatural rank H , satisfies a strict inequality u_H < 1 . Accordingly, an alternative interpretation for “zero followed by infinitely many 9s” could be{{sfnp|Katz|Katz|2010a}}
underset{H}{0.underbrace{999ldots}}; = 1;-;frac{1}{10^{H}}.
All such interpretations of “0.999...” are infinitely close to 1. Ian Stewart characterizes this interpretation as an “entirely reasonable” way to rigorously justify the intuition that “there’s a little bit missing” from 1 in 0.999....{{efn|{{harvtxt|Stewart|2009}}, p. 175; the full discussion of 0.999... is spread through pp. 172–175.}} Along with {{harvtxt|Katz|Katz|2010b}}, Robert Ely also questions the assumption that students’ ideas about 0.999ldots < 1 are erroneous intuitions about the real numbers, interpreting them rather as nonstandard intuitions that could be valuable in the learning of calculus.{hide}sfnmp
| 1a1 = Katz | 1a2 = Katz | 1y = 2010b
| 2a1 = Ely | 2y = 2010
{edih}

Hackenbush

Combinatorial game theory provides a generalized concept of number that encompasses the real numbers and much more besides.{{sfnp|Conway|2001|pp=3–5,12–13,24–27}} For example, in 1974, Elwyn Berlekamp described a correspondence between strings of red and blue segments in Hackenbush and binary expansions of real numbers, motivated by the idea of data compression. For example, the value of the Hackenbush string LRRLRLRL ldots is 0.010101_2 ldots = frac{1}{3} . However, the value of LRLLL ldots (corresponding to 0.111 ldots_2 is infinitesimally less than 1. The difference between the two is the surreal number frac{1}{omega} , where omega is the first infinite ordinal; the relevant game is LRRRRldots or 0.000ldots _2 .{{efn|{{harvtxt|Berlekamp|Conway|Guy|1982}}, pp. 79–80, 307–311 discuss 1 and frac{1}{3} and touch on frac{1}{omega} . The game for 0.111dots_2 follows directly from Berlekamp’s Rule.}}This is true of the binary expansions of many rational numbers, where the values of the numbers are equal but the corresponding binary tree paths are different. For example, 0.10111ldots_2 = 0.11000ldots _2 , which are both equal to frac{3}{4} , but the first representation corresponds to the binary tree path LRLRLLLldots , while the second corresponds to the different path LRLLRRRldots .

Revisiting subtraction

Another manner in which the proofs might be undermined is if 1 - 0.999ldots simply does not exist because subtraction is not always possible. Mathematical structures with an addition operation but not a subtraction operation include commutative semigroups, commutative monoids, and semirings. Richman considers two such systems, designed so that 0.999ldots < 1 .First, {{harvtxt|Richman|1999}} defines a nonnegative decimal number to be a literal decimal expansion. He defines the lexicographical order and an addition operation, noting that 0.999ldots < 1 simply because 0 < 1 in the ones place, but for any nonterminating x , one has 0.999ldots + x = 1 + x . So one peculiarity of the decimal numbers is that addition cannot always be canceled; another is that no decimal number corresponds to frac{1}{3} . After defining multiplication, the decimal numbers form a positive, totally ordered, commutative semiring.{{sfnp|Richman|1999|pp=397–399}}In the process of defining multiplication, Richman also defines another system he calls “cut D ”, which is the set of Dedekind cuts of decimal fractions. Ordinarily, this definition leads to the real numbers, but for a decimal fraction d he allows both the cut (-infty, d) and the “principal cut” (-infty, d] . The result is that the real numbers are “living uneasily together with” the decimal fractions. Again 0.999ldots < 1 . There are no positive infinitesimals in cut D , but there is “a sort of negative infinitesimal,” 0^- , which has no decimal expansion. He concludes that 0.999ldots = 1 + 0^- , while the equation ” 0.999ldots + x = 1 ” has no solution.{{efn|{{harvtxt|Richman|1999}}, pp. 398–400. {{harvtxt|Rudin|1976}}, p. 23 assigns this alternative construction (but over the rationals) as the last exercise of Chapter 1.}}p-}}}}adic numbers“>

{{nowrap|1{{math|p-}}}}adic numbers

When asked about 0.999..., novices often believe there should be a “final 9”, believing 1 - 0.999ldots to be a positive number which they write as “0.000...1”. Whether or not that makes sense, the intuitive goal is clear: adding a 1 to the final 9 in 0.999... would carry all the 9s into 0s and leave a 1 in the ones place. Among other reasons, this idea fails because there is no “final 9” in 0.999....{hide}sfnmp
| 1a1 = Gardiner | 1y = 2003 | 1p = 98
| 2a1 = Gowers | 2y = 2002 | 2p = 60
{edih} However, there is a system that contains an infinite string of 9s including a last 9.(File:4adic 333.svg|right|thumb|class=skin-invert|200px|The 4-adic integers (black points), including the sequence {{math|1=(3, 33, 333, ...)}} converging to {{math|1=−1}}. The 10-adic analogue is {{math|1=...999 = −1}}.)The {{nowrap|1= p -}}adic numbers are an alternative number system of interest in number theory. Like the real numbers, the {{nowrap|1= p -}}adic numbers can be built from the rational numbers via Cauchy sequences; the construction uses a different metric in which 0 is closer to p , and much closer to p^n , than it is to 1.{{sfnp|Mascari|Miola|1988|p=83–84}} The {{nowrap|1= p -}}adic numbers form a field for prime p and a ring for other p , including 10. So arithmetic can be performed in the {{nowrap|1= p -}}adics, and there are no infinitesimals.In the 10-adic numbers, the analogues of decimal expansions run to the left. The 10-adic expansion ...999 does have a last 9, and it does not have a first 9. One can add 1 to the ones place, and it leaves behind only 0s after carrying through: 1 + ldots999 = ldots 000 = 0 , and so ldots 999 = -1 .{{sfnp|Fjelstad|1995|p=11}} Another derivation uses a geometric series. The infinite series implied by “...999” does not converge in the real numbers, but it converges in the 10-adics, and so one can re-use the familiar formula:{{sfnp|Fjelstad|1995|pp=14–15}}ldots999 = 9 + 9(10) + 9(10)^2 + 9(10)^3 + cdots = frac{9}{1-10} = -1.Compare with the series in the section above. A third derivation was invented by a seventh-grader who was doubtful over her teacher’s limiting argument that 0.999... = 1 but was inspired to take the multiply-by-10 proof above in the opposite direction: if x = ldots 999, then 10x = ldots 990 , so 10x = x - 9 , hence x = -1 again.{{sfnp|Fjelstad|1995|p=11}}As a final extension, since 0.999ldots = 1 (in the reals) and ldots 999 = -1 (in the 10-adics), then by “blind faith and unabashed juggling of symbols“{{sfnp|DeSua|1960|p=901}} one may add the two equations and arrive at ldots 999.999ldots = 0 . This equation does not make sense either as a 10-adic expansion or an ordinary decimal expansion, but it turns out to be meaningful and true in the doubly infinite decimal expansion of the 10-adic solenoid, with eventually repeating left ends to represent the real numbers and eventually repeating right ends to represent the 10-adic numbers.{{sfnp|DeSua|1960|p=902–903}}

Related questions

Zeno’s paradoxes, particularly the paradox of the runner, are reminiscent of the apparent paradox that 0.999... and 1 are equal. The runner paradox can be mathematically modelled and then, like 0.999..., resolved using a geometric series. However, it is not clear if this mathematical treatment addresses the underlying metaphysical issues Zeno was exploring.{hide}sfnmp
| 1a1 = Wallace | 1y = 2003 | 1p = 51
| 2a1 = Maor | 2y = 1987 | 2p = 17
{edih}Negative zero is another redundant feature of many ways of writing numbers. In number systems, such as real numbers, where “0” denotes the additive identity and is neither positive nor negative, the usual interpretation of “−0” is that it should denote the additive inverse of 0, which forces -0 = 0 .{{sfnp|Munkres|2000|p=34|loc=Exercise 1(c)}} Nonetheless, some scientific applications use separate positive and negative zeroes, as do some computing binary number systems (for example integers stored in the sign and magnitude or ones’ complement formats, or floating point numbers as specified by the IEEE floating-point standard).{{harvtxt|Kroemer|Kittel|1980}}, p. 462. WEB,msdn.microsoft.com/library/en-us/csspec/html/vclrfcsharpspec_4_1_6.asp, Floating point types, Microsoft Developer Network, MSDN C# Language Specification, 29 August 2006,msdn.microsoft.com/library/en-us/csspec/html/vclrfcsharpspec_4_1_6.asp," title="web.archive.org/web/20060824085452msdn.microsoft.com/library/en-us/csspec/html/vclrfcsharpspec_4_1_6.asp,">web.archive.org/web/20060824085452msdn.microsoft.com/library/en-us/csspec/html/vclrfcsharpspec_4_1_6.asp, 24 August 2006,

See also

Notes

{{notelist|colwidth=30em}}{{reflist|colwidth=30em}}

References

  • WEB,www.straightdope.com/columns/030711.html, An infinite question: Why doesn’t .999~ = 1?, 11 July 2003, Cecil Adams, Cecil, Adams, The Straight Dope, Chicago Reader, 2006-09-06,www.straightdope.com/columns/030711.html," title="web.archive.org/web/20060815010844www.straightdope.com/columns/030711.html,">web.archive.org/web/20060815010844www.straightdope.com/columns/030711.html, 15 August 2006, live,
  • BOOK, Alligood, K. T., Sauer, T. D., Yorke, J. A., James A. Yorke, 1996, Chaos: An introduction to dynamical systems, 4.1 Cantor Sets, Springer, 978-0-387-94677-1,
  • : This introductory textbook on dynamical systems is aimed at undergraduate and beginning graduate students. (p. ix)
  • BOOK, Apostol, Tom M., Tom M. Apostol, 1974,archive.org/details/mathematicalanal02edtomm, Mathematical Analysis, 2e, Addison-Wesley, 978-0-201-00288-1,
  • : A transition from calculus to advanced analysis, Mathematical analysis is intended to be “honest, rigorous, up to date, and, at the same time, not too pedantic.” (pref.) Apostol’s development of the real numbers uses the least upper bound axiom and introduces infinite decimals two pages later. (pp. 9–11)
  • BOOK, Artigue, Michèle, Derek, Holton, Michèle, Artigue, Urs, Kirchgräber, Joel, Hillel, Mogens, Niss, Alan, Schoenfeld, The Teaching and Learning of Mathematics at University Level, New ICMI Study Series, 2002, 7, Springer, Dordrecht, 10.1007/0-306-47231-7, 978-0-306-47231-2,
  • JOURNAL, Baldwin, Michael, Norton, Anderson, Does 0.999... Really Equal 1?,eric.ed.gov/?id=EJ961516, 2012, The Mathematics Educator, 21, 2, 58–67,
  • BOOK, Bartle, R. G., Robert G. Bartle, Sherbert, D. R., 1982,archive.org/details/introductiontore0000bart, Introduction to Real Analysis, Wiley, 978-0-471-05944-8,
  • : This text aims to be “an accessible, reasonably paced textbook that deals with the fundamental concepts and techniques of real analysis.” Its development of the real numbers relies on the supremum axiom. (pp. vii–viii)
  • BOOK, Beals, Richard, Richard Beals (mathematician),books.google.com/books?id=cXAqJUYqXx0C, Analysis: An Introduction, 2004, Cambridge UP, 978-0-521-60047-7,
  • BOOK, 1982, Winning Ways for your Mathematical Plays, Academic Press, 978-0-12-091101-1, Berlekamp, E. R., Conway, J. H., Guy, R. K., Elwyn Berlekamp, John Horton Conway, Winning Ways for your Mathematical Plays,
  • BOOK, Bonnycastle, John, 1806, An introduction to algebra; with notes and observations: designed for the use of schools and places of public education,hdl.handle.net/2027/mdp.39015063620382, First American, Philadelphia, 2027/mdp.39015063620382,
  • BOOK, Bunch, Bryan H.,archive.org/details/mathematicalfall0000bunc, Mathematical Fallacies and Paradoxes, 1982, Van Nostrand Reinhold, 978-0-442-24905-2,
  • : This book presents an analysis of paradoxes and fallacies as a tool for exploring its central topic, “the rather tenuous relationship between mathematical reality and physical reality”. It assumes first-year high-school algebra; further mathematics is developed in the book, including geometric series in Chapter 2. Although 0.999... is not one of the paradoxes to be fully treated, it is briefly mentioned during a development of Cantor’s diagonal method. (pp. ix-xi, 119)
  • BOOK, Burrell, Brian, Merriam-Webster’s Guide to Everyday Math: A Home and Business Reference, 1998, Merriam-Webster, 978-0-87779-621-3,archive.org/details/merriamwebstersg00burr,
  • BOOK, Byers, William, How Mathematicians Think: Using Ambiguity, Contradiction, and Paradox to Create Mathematics, 2007, Princeton UP, 978-0-691-12738-5,archive.org/details/howmathematician00byer,
  • BOOK, Cheng, Eugenia, Eugenia Cheng, Is Math Real? How Simple Questions Lead Us To Mathematics’ Deepest Truths, 2023, Basic Books, 978-1-541-6-01826,
  • BOOK, Conway, John B., John B. Conway,archive.org/details/isbn_9781461263142, Functions of One Complex Variable I, 2e, Springer-Verlag, 1973, 1978, 978-0-387-90328-6,
  • BOOK, Conway, John H., John Horton Conway, On Numbers and Games, On Numbers and Games, A K Peters, 2001, 2nd, 1-56881-127-6,
  • BOOK, Davies, Charles, Charles Davies (professor), 1846, The University Arithmetic: Embracing the Science of Numbers, and Their Numerous Applications, A.S. Barnes,archive.org/details/universityarith00davigoog, 175, 4 July 2011,
  • WEB,www.faqs.org/faqs/sci-math-faq/specialnumbers/0.999eq1/, Hans, de Vreught, 1994, sci.math FAQ: Why is 0.9999... = 1?, 29 June 2006,www.faqs.org/faqs/sci-math-faq/specialnumbers/0.999eq1/," title="web.archive.org/web/20070929122649www.faqs.org/faqs/sci-math-faq/specialnumbers/0.999eq1/,">web.archive.org/web/20070929122649www.faqs.org/faqs/sci-math-faq/specialnumbers/0.999eq1/, 29 September 2007, live,
  • JOURNAL, DeSua, Frank C., A System Isomorphic to the Reals,archive.org/details/sim_american-mathematical-monthly_1960-11_67_9/page/900, 2309468, The American Mathematical Monthly, 67, 900–903, 10.2307/2309468, 9, November 1960,
  • JOURNAL, Diamond, Louis E., Irrational Numbers, Mathematics Magazine, Mathematical Association of America, 1955, 29, 2, 89–99, 10.2307/3029588, 3029588,
  • BOOK, Dickson, Leonard Eugene, History of the Theory of Numbers, 1, Carnegie Institution of Washington, 1919,
  • JOURNAL, Dubinsky, Ed, Weller, Kirk, McDonald, Michael, Brown, Anne, Some historical issues and paradoxes regarding the concept of infinity: an APOS analysis: part 2,archive.org/details/sim_educational-studies-in-mathematics_2005_60_2/page/253, Educational Studies in Mathematics, 2005, 60, 253–266, 10.1007/s10649-005-0473-0, 2, 45937062,
  • BOOK, Earl, Richard, Nicholson, James, The Concise Oxford Dictionary of Mathematics, 6th, 2021, Oxford University Press, 978-0-192-58405-2,
  • JOURNAL, Edwards, Barbara, Ward, Michael, Surprises from mathematics education research: Student (mis)use of mathematical definitions, The American Mathematical Monthly, 111, 411–425,www.wou.edu/~wardm/FromMonthlyMay2004.pdf, 10.2307/4145268, 4145268, 4 July 2011,www.wou.edu/~wardm/FromMonthlyMay2004.pdf," title="web.archive.org/web/20110722153906www.wou.edu/~wardm/FromMonthlyMay2004.pdf,">web.archive.org/web/20110722153906www.wou.edu/~wardm/FromMonthlyMay2004.pdf, 22 July 2011, 5, May 2004, 10.1.1.453.7466,
  • JOURNAL, Ely, Robert, 2010, Nonstandard student conceptions about infinitesimals,archive.org/details/sim_journal-for-research-in-mathematics-education_2010-03_41_2/page/117, Journal for Research in Mathematics Education, 41, 2, 117–146, 10.5951/jresematheduc.41.2.0117,
  • : This article is a field study involving a student who developed a Leibnizian-style theory of infinitesimals to help her understand calculus, and in particular to account for {{nowrap|0.999...}} falling short of 1 by an infinitesimal {{nowrap|0.000...1.}}
  • BOOK, Enderton, Herbert B., Herbert Enderton, 1977,archive.org/details/elementsofsetthe0000ende, Elements of Set Theory, Elsevier, 978-0-12-238440-0,
  • : An introductory undergraduate textbook in set theory that “presupposes no specific background”. It is written to accommodate a course focusing on axiomatic set theory or on the construction of number systems; the axiomatic material is marked such that it may be de-emphasized. (pp. xi–xii)
  • BOOK, Euler, Leonhard, Leonhard Euler, 1770, 1822, 3rd English, Elements of Algebra, John Hewlett and Francis Horner, English translators, Orme Longman,archive.org/details/elementsalgebra00lagrgoog, 170, 978-0-387-96014-2, 4 July 2011,
  • BOOK, Finney, Ross L., Weir, Muarice D., Giordano, Frank tR., 2001


, Thomas’ Calculus: Early Transcendentals, 10th, Addison-Wesley, New York,

Further reading

  • JOURNAL, Why Does 0.999... = 1?: A Perennial Question and Number Sense,eric.ed.gov/?id=EJ717818, Beswick, Kim, Australian Mathematics Teacher, 60, 7–9, 2004, 4, none,
  • JOURNAL, Journal of Statistical Physics, 47, 3/4, 1987, One-dimensional model of the quasicrystalline alloy, S. E., Burkov, 10.1007/BF01007518, 409–438, 1987JSP....47..409B, 120281766, none,
  • JOURNAL, 81.15 A Case of Conflict,archive.org/details/sim_mathematical-gazette_1997-03_81_490/page/109, Bob, Burn, The Mathematical Gazette, 81, 490, 109–112, 3618786, 10.2307/3618786, March 1997, 187823601, none,
  • JOURNAL, The Age of Newton: An Intensive Interdisciplinary Course,archive.org/details/sim_history-teacher_1981-02_14_2/page/167, J. B., Calvert, E. R., Tuttle, Michael S., Martin, Peter, Warren, The History Teacher, 14, 2, 167–190, 493261, 10.2307/493261, February 1981, none,
  • JOURNAL, Younggi, Choi, Jonghoon, Do, Equality Involved in 0.999... and (-8)1/3, For the Learning of Mathematics, 25, 3, 13–15, 36, 40248503, November 2005, none,
  • JOURNAL, Rational Approximations to Ï€,archive.org/details/sim_mathematics-of-computation_1971-04_25_114/page/387, K. Y., Choong, D. E., Daykin, C. R., Rathbone, Mathematics of Computation, 25, 114, 387–392, 2004936, 10.2307/2004936, April 1971, none,
  • BOOK, Edwards, B., 1997, An undergraduate student’s understanding and use of mathematical definitions in real analysis, Dossey, J., Swafford, J.O., Parmentier, M., Dossey, A.E., Proceedings of the 19th Annual Meeting of the North American Chapter of the International Group for the Psychology of Mathematics Education, 1, ERIC Clearinghouse for Science, Mathematics and Environmental Education, Columbus, OH, 17–22, none,
  • JOURNAL, Eisenmann, Petr, 2008, Why is it not true that 0.999... < 1?, The Teaching of Mathematics, 11, 1, 35–40,elib.mi.sanu.ac.rs/files/journals/tm/20/tm1114.pdf, 4 July 2011, none,
  • JOURNAL, Ferrini-Mundy, J., Graham, K., 1994, Research in calculus learning: Understanding of limits, derivatives and integrals, MAA Notes: Research Issues in Undergraduate Mathematics Learning, 33, 31–45, J., Kaput, E., Dubinsky, none,
  • JOURNAL, Infinite processes in elementary mathematics: How much should we tell the children?,archive.org/details/sim_mathematical-gazette_1985-06_69_448/page/77, Tony, Gardiner, The Mathematical Gazette, 69, 448, 77–87, 3616921, 10.2307/3616921, June 1985, 125222118, none,
  • JOURNAL, Real Mathematics: One Aspect of the Future of A-Level,archive.org/details/sim_mathematical-gazette_1988-12_72_462/page/276, John, Monaghan, The Mathematical Gazette, 72, 462, 276–281, 3619940, 10.2307/3619940, December 1988, 125825964, none,
  • BOOK, Núñez, Rafael, Rafael E. Núñez, Do Real Numbers Really Move? Language, Thought, and Gesture: The Embodied Cognitive Foundations of Mathematics, 2006, 18 Unconventional Essays on the Nature of Mathematics, Springer, 160–181,www.cogsci.ucsd.edu/~nunez/web/publications.html, 978-0-387-25717-4, 4 July 2011,www.cogsci.ucsd.edu/~nunez/web/publications.html," title="web.archive.org/web/20110718014351www.cogsci.ucsd.edu/~nunez/web/publications.html,">web.archive.org/web/20110718014351www.cogsci.ucsd.edu/~nunez/web/publications.html, 18 July 2011, none,
  • JOURNAL, Malgorzata, Przenioslo, Images of the limit of function formed in the course of mathematical studies at the university, Educational Studies in Mathematics, 55, 1–3, 10.1023/B:EDUC.0000017667.70982.05, 103–132, March 2004, 120453706, none,
  • JOURNAL, Using Self-Similarity to Find Length, Area, and Dimension,archive.org/details/sim_american-mathematical-monthly_1996-02_103_2/page/107, James T., Sandefur, The American Mathematical Monthly, 103, 2, 107–120, 2975103, 10.2307/2975103, February 1996, none,
  • JOURNAL, SierpiÅ„ska, Anna, Anna SierpiÅ„ska, Humanities students and epistemological obstacles related to limits, Educational Studies in Mathematics, 18, 4, 371–396, 10.1007/BF00240986, 3482354, November 1987, 144880659, none,
  • JOURNAL, Starbird, M., Michael Starbird, Starbird, T., Required Redundancy in the Representation of Reals, 114, 769–774, Proceedings of the American Mathematical Society, 2159403, 3, 10.1090/S0002-9939-1992-1086343-5, March 1992, free, none,
  • JOURNAL, Mathematical Beliefs and Conceptual Understanding of the Limit of a Function, Jennifer Earles, Szydlik, Journal for Research in Mathematics Education, 31, 3, 258–276, 749807, 10.2307/749807, May 2000, none,
  • JOURNAL, David O., Tall, Dynamic mathematics and the blending of knowledge structures in the calculus, ZDM Mathematics Education, 2009, 41, 4, 481–492, 10.1007/s11858-009-0192-6, 14289039, none,
  • JOURNAL, David O., Tall, Intuitions of infinity,archive.org/details/sim_mathematics-in-school_1981-05_10_3/page/30, Mathematics in School, May 1981, 10, 3, 30–33, 30214290, none,

External links

{{Spoken Pseudopedia|En-0.999...-article.oga|date=2006-10-19}}{{commons category}} {{Real numbers}}{{featured article}}

- content above as imported from Wikipedia
- "0.999..." does not exist on GetWiki (yet)
- time: 10:16am EDT - Wed, May 22 2024
[ this remote article is provided by Wikipedia ]
LATEST EDITS [ see all ]
GETWIKI 21 MAY 2024
GETWIKI 09 JUL 2019
Eastern Philosophy
History of Philosophy
GETWIKI 09 MAY 2016
GETWIKI 18 OCT 2015
M.R.M. Parrott
Biographies
GETWIKI 20 AUG 2014
CONNECT