Anda di halaman 1dari 86

c 

      
      
This page attempts to show some of the ambiguities in defining some of the mathematical terms
that might be encountered at the high school level. A number of these issues may seem quite
trivial, and some are the result of consulting out-of-date texts. But many of these issues have led
to disputed answers in mathematics competitions. Contributions and suggestions are welcome.

c
  c !"#    
by Glenn James and Robert C. James
(1949) defines 
   as "the greatest numerical value of the ordinates of a
periodic curve." However, the third edition of this dictionary (1968) has the definition "half the
difference between the greatest and the least values of the ordinates of a periodic curve." The
more recent definition seems generally to be in use today.

c!  c$ c$
c!  c$% !c$& # Some textbooks give the range of the
Arcsecant function to be [0, ʌ] except for ʌ/2. However, by defining Arcsec  to be Arccos (1/),
ʌ/2 can be included in the range.

Other textbooks give the range of the Arcsecant function to be [0, ʌ/2) U [ʌ, 3ʌ/2).

Some textbooks give the range of the Arccosecant function to be [-ʌ/2, 0) U (0, ʌ/2). However,
by defining Arccsc  to be Arcsin (1/), 0 can be included in the range.

Other textbooks give the range of the Arccosecant function to be (0, ʌ/2] U (pi, 3ʌ/2].

r$% !$% # There are two systems for naming large numbers. For example, in
the American system a billion is 109, whereas in a system formerly used in Britain, a billion is
1012 and a milliard is 109.

c!c ! c$
c$c# These terms have less significance now that
logarithmic tables are no longer used. Older textbooks would write the logarithm of 0.56 as
.7482 - 1 and identify .7482 as the mantissa and -1 as the characteristic. Since nowadays a
calculator gives the logarithm of 0.56 as -0.2518 and modern dictionaries define the
characteristic and mantissa as the integer and decimal parts of a logarithm respectively, it would
seem that a requirement to identify the characteristic or mantissa of the logarithm of a number
between 0 and 1 is ambiguous.

$"!&$  
"!&$ !# According to the third edition of c  

c (1892), an oscillating series is called "divergent by Cauchy, Bertrand, Laurent, and
others." In 1893 in c  


 
 James Harkness and Frank Morley
write: "Strictly speaking, an oscillating series is distinct from a divergent series, but it is usual to
speak of non-convergent series as divergent." In 1898 in  




  
 
 by Harkness and Morley a footnote says, "Most English text-books regard oscillating
series as not divergent." The 1968 third edition of    
by James and James
defines a divergent series as: "A sequence which does not converge. It might either be properly
divergent, or oscillate ..." This seems to be a generally accepted definition today.

$cr has a mathematical definition that is very different from its meaning in other
contexts. For example,      2nd ed., Course I defines an infinite set as one
whose elements "cannot be counted." But in mathematics a countable set can be countably
infinite.

! c $#   c  


 by Sherman K. Stein and Anthony
Barcellos (1992) defines a critical number as a number for which the derivative is zero, and the
critical point as the corresponding point, and has the following remark: "Some texts define a
critical number as a number where the derivative is 0 or else is not defined. Since we emphasize
differentiable functions, a critical number is a number where the derivative is 0." Some textbooks
include endpoints as critical points. For example,   (1991) by Gilbert Strang has: "The
maximum always occurs at a stationary point (where df/dx=0) or a rough point (no derivative) or
an endpoint of the domain. These are the three types of critical points. All maxima and minima
occur at critical points!"

!"c!  c $ $# According to Roman Arce, "Sometimes it's defined as


equivalent of its second derivative and other times as the inverse of the radius of a circle that fits
the curve (the osculating circle), the relation is 1/þ =  '' ()/(1+ ' ())3/2 so the two definitions are
proportional but not equal and some times it's used without specifying if they mean 1/þ or  ''
()."

&$!c $  $# The degenerate conic sections are normally considered
to be a point, a line, and a pair of lines. True or false: "2/ 2 + 2/2 = 0 is an ellipse." It is a point,
but it is also a degenerate ellipse.    
by James and James, both in the 1949
and 1968 editions, calls this a 
 

Various web pages say that a degenerate ellipse is a circle or a parabola. Steve Wilson, Professor
of Mathematics at Johnson County Community College, writes, "Personally, I could consider
circles and parabolas as special or limiting cases, but not as degenerates."

c$  c $ $# @ c  by Larson, Kanold, and Stiff (1998) says the
domain of () = sqrt ( - 7) is all real numbers greater than or equal to 7. The book defines the

  
 
  
as "the -values of a function = ()." Is 0 in the domain of
?

Steve Wilson, Professor of Mathematics at Johnson County Community College, writes,


"Concerning domain, I would say that zero was not in the domain of . However, zero
would be in the domain of sqrt( 2 ( - 1) ), an almost equivalent function. I take this position
because the domain of the product function  should only be the intersection of the domain of 
with the domain of  (e.g., see Larson & Hostetler's 
c  2nd edition, 1989).
Unfortunately, some newer textbooks will ignore the domain issue, but I didn't find any that took
a different view. I would be interested to know if any other textbook does differ."
c&$c!' $r!# There are two modern meanings of the term      In
   
  
 10th ed., an imaginary number is a number of the
form +  where  is not equal to 0. In   c  
 (1992) by Stein and
Barcellos, "a complex number that lies on the axis is called   #"

Steve Wilson, referring to the 1968 Third edition of    
by James and
James, writes:

JJ3's discussion of imaginary numbers is interesting. Under complex numbers, they say "Any
number, real or imaginary, of the form + , where and  are real numbers and 2 = -1. Called
imaginary numbers when  is not equal to 0, and pure imaginary when = 0 and  is not equal to
0 (although complex numbers are not imaginary in the usual sense)." If that parenthetical
comment had not been made, their position would have been clear. Most current college algebra
texts appear to agree with JJ3, although some will avoid the term "imaginary number" altogether,
and Gustafson & Frisk's "College Algebra", 8th ed., 2004, agrees with Stein & Barcellos.

c&$c!' c!# According to    


  
 11th ed., the
imaginary part of 2 + 3 is 3 However, according to the definition found on several web sites,
the imaginary part is 3. Roman Arce points out that the function Im[] of Mathematica 3.0 gives
3.

$ !c$&(
 !c$& $ $# In Stein and Barcellos (1992), "a function  is
said to be      if whenever 1 < 2, one has (1) <  (2)." The definition in Ellis and Gulick
(1986) requires only that (1) is less than or equal to  (2). The latter text used       
to describe the other situation.

$ $ $# Some texts require that a tangent line exist at an inflection point, but
other texts do not.

$"! !&$! $ $% !c$& # =  !"c


  c
 3rd. edition, by Larson, Hostetler, and Edwards has: "In Example 4, if you
had set the calculator to   mode, the display would have been in degrees rather than in
radians. This convention is peculiar to calculators. By definition, the values of inverse
trigonometric functions are always in  " Other textbooks, however, allow degrees.

  !c$&# Older definitions specified that an isosceles triangle has exactly two
congruent sides, as, for example, a 1570 translation of Euclid: "Isosceles, is a triangle, which
hath onely two sides equall." Many texts define an isosceles triangle as a triangle having two
sides equal; such a definition could perhaps be interpreted either way. @ 
 "c 
  c
 (1998) has: "An 
   has at least two congruent sides."

% )$  # Some texts say carefully that, although we write lim () = oo, the
limit 

 Ellis and Gulick (1986) says that " has an infinite limit at " although it also
has "Caution: If  has an infinite limit at  then 

 have a limit at in the sense of
Definition 2.1."
$c!c $r!# Modern dictionaries, except for the þ 
@
 dictionaries,
exclude 0 in the definition of this term. The 1771 #  
  $   excludes 0. In 1889
# 
c by G. A. Wentworth includes 0. In 1919 in  


   
=

 Bertrand Russell includes 0.

$ # This is a term which should probably be avoided, since it could be interpreted to
mean "slope of zero" or "undefined slope."

c c* !c$&# In @ c "c   c


 (1998), the row of
Pascal's Triangle consisting only of 1 is Row 0. However, on the following page, it has: "The
binomial coefficients in Row 4 of Pascal's Triangle are 1, 3, 3, 1." Presumably, this is an
oversight, but some texts call the row consisting only of 1 Row 1.

Paul Sisson's 
c  2003, says, "Pascal's triangle is a useful way of generating
binomial coefficients, with the th row containing the coefficients of a binomial raised to the ( -
1)st power."

In December 2003, Steve Wilson wrote, "Having just examined 15 college algebra textbooks,
they seem to come in the ratio 2 agreeing with Sisson to 1 recognizing a zeroth row to 1
finessing the issue (e.g., "the row for = 6")."

! $r!# Is 1 a prime number? Most textbooks today call it neither prime nor
composite, but older texts generally considered it to be prime. In 1859, Lebesgue stated
explicitly that 1 is prime in #    % & It is prime in =  # 

c 


'

 c  (1866) by Joseph Ray and '  c 
(1892) by William J. Milne. A list of primes to 10,006,721 published in 1914 by D. N. Lehmer
includes 1. [David Cantrell suggests this is more a historical issue and is inappropriate for this
page. He writes, "No high-school students (unless they're doing historical research) are going to
encounter a legitimate definition of prime which would include 1." This entry may be removed.]

!) c$&# Older geometry texts sometimes defined a   as an angle with
measure greater than 180 degrees, whereas modern geometry textbooks seem to consider an
angle in such a way that no angle measures greater than 180 degrees.

!c" )!# Textbook authors differ on whether an endpoint can be a local


minimum or maximum. According to a post in the AP Calculus mailing list, this issue is avoided
in AP exam questions because of the inconsistency.

!r#    


in 1949 and in the 1968 third edition has: "Some authors
require that a rhombus not be a square, but the preference seems to be to call the square a special
case of the rhombus." [Can any readers of this page cite any modern writer who makes this
requirement?]

!c$c '!'# Two definitions are: 1. A figure has rotational symmetry if a


rotation less than 360 degrees about a fixed point rotates the figure onto itself. 2. A figure has
rotational symmetry if a rotation of 180 degrees or less rotates the figure onto itself. [Vince
DeBlase]

!$
$&# There is not a universally accepted procedure for rounding, for example, 2.315 to
the nearest hundredth. However, one commonly accepted rule, called the computer's rule or the
odd-even rule, is to make the digit before the 5 even, thus to add 1 if it is odd and leave it alone if
it is already even.

 c$# =  @ 


 defines   as "Line, ray, or segment that contains a chord
of a circle." Is a chord also a secant? If a chord contains itself, it would be a secant. This text
seems not to define 
  


!#    
(1949) calls this term "probably the most
indefinite term used seriously in mathematics." One particular case is that some newer textbooks
seem not to require rationalizing the denominators of some fractions.

+c! !# This term probably should be avoided. Some Internet web pages say it is a
prism with a square base; others say it is a cube.

The 1968 third edition of    


by James and James has: "A prism with a
triangle as base is a triangular prism; one with a quadrilateral as base is a quadrangular prism;
etc." This might imply that a square prism should be a prism with a square base.

c$
c!
!# Some writers require that for the equation  +  =  to be in   

 should not be negative. Other writers do not mention this requirement.

!c,
# The usual definition of a trapezoid requires that it have   one pair of
parallel sides. However, the UCSMP textbook 
 (1997) has the definition "a
quadrilateral with at least one pair of parallel sides." (In this textbook, an 
 (
 is
defined as "a trapezoid with a pair of base angles equal in measure.") According to Chris White,
the    

   by William Karush (1962) defines  (
 as "A
quadrilateral which has exactly one pair of parallel sides (sometimes, the parallelogram, with
both pairs of sides parallel, is included)."

!c,
(!c,# In the United States, a  (
 is generally defined as a
quadrilateral with exactly one pair of parallel sides and a  ( is a quadrilateral with no
sides parallel. However, until the end of the eighteenth century, the definitions for these two
terms were reversed, and outside the United States even today the definitions are often reversed.

!&$! $ $%


c$ # When using radians, are the arguments of
the trigonometric functions numbers of radians or just real numbers? Textbook writers seem to
differ.

The definition of radian in the third edition of    


by James and James is: "A
central angle subtended in a circle by an arc whose length is equal to the radius of the circle.
Thus the radian measure of an angle is the ratio of the arc it subtends to the radius of the circle in
which it is the central angle." Thus technically this dictionary distinguishes between "radian" and
"radian measure," and since radian measure is a ratio of lengths, it would be a real number.

,!   ,! -!# Most textbook writers either leave 00 undefined, or state that
it is undefined. Others say it should be defined as 1. Euler argued for 00 = 1; Cauchy considered
it undefined. Most calculators give "error" for this expression, although some give "1."

 c  -    +     


   
.     /
  
1. What is the sum of the
The item should specify whether there is only one angle at
measures of the exterior angles
each vertex, as this is usually the intent.
of a triangle?
2. If you roll a pair of dice, what
The item should say "a sum of six," since if you roll a 6 and a
is the probability you will get a
3 you did get a 6.
6?
Are the digits to the left of the decimal point counted? All of
3. What is the fourth decimal
the digits are decimal digits, since the number is written in
digit of 4?
decimal notation (base 10).

        


     à  was first used by Leonhard Euler (1707-1783) in 1734 in

  c  '  =

  (Cajori, vol. 2, page 268).

c    # Karl Weierstrass (1815-1897) used | | in an 1841 essay "Zur Theorie
der Potenzreihen," in which the symbol appears on page 67. He also used the symbol in 1859 in
"Neuer Beweis des Fundamentalsatzes der Algebra," in which the symbol appears on page 252.
This latter essay was submitted to the Berlin Academy of Sciences on December 12, 1859. These
are the two reference shown by Cajori (vol. 2, page 123).

Cajori says that the first essay was not printed at the time, and Julio González Cabillón believes
neither paper was published until 1894, "when the welcome #  $  [vol. I] of Karl
Weierstrass "Mathematische Werke" [Berlin: Mayer & Mueller], saw the light. I do not know to
what extent the editors could have interfered with Weierstrass manuscripts. In both papers the
notation under discussion does not appear with a definition or with a further comment; thus I am
speculating that their subsequent published typesetting might differ from that of Weierstrass
original."

The memoir "Zur Theorie der eindeutigen analytischen Functionen," which appeared in
c    )
 c     [pp. 11-60, Berlin 1876, and was
reprinted in *+ $  (volume II) of Weierstrass "Mathematische Werke" (1895)] has a
footnote on page 78 in which Weierstrass remarks:

Ich bezeichne den absoluten Betrag einer complexen Groesse  mit ||. [I denote the absolute
value of complex number  by ||]
In this memoir, Weierstrass applied the absolute value symbolism to complex numbers.

r    # The use of ȕ (for the function created by Euler) was introduced by Jacques P. M.
Binet (1786-1856) in 1839 (Cajori, vol. 2, page 272).

Julio González Cabillón says the capital letter $ is a common one in the Greek and Latin
alphabets. If, after Legendre, the second Eulerian integral was known as the Gamma function,
why Binet could not choose the initial of his name to denote the first Eulerian integral (Beta
function), conventionally written as $,&- The citation: "Memoire sur les intégrales définies
euleriennes, et sur leur application a la theorie des suites, ansi qu'a l'evaluation des fonctions des
grands nombres," .
 ! #
þ
=
 % &
/0 pp. 123-343, Paris, 1839.

On page 131 of his "Memoire...", Binet states:

Je designerai la premiere de ces fonctions par $,&- et pour la seconde j'adoptarai la notacion
Gamma,- proposee par M. Legendre.
See also BETA and GAMMA FUNCTIONS on the math words page.

&    # The use of ī (for the function created by Euler) was introduced by Adrien-
Marie Legendre (1752-1833) (Cajori vol. 2, page 271). On page 277 of his "Exercices de Calcul
integral sur divers ordres de transcendantes et sur les quadrantes," Tome Premier, Paris, 1811,
Legendre states:

... Cette quantité étant simplement fonction de  nous la designerons par ī( ), et nous ferons
ī( ) = [dx(log 1/)( -1)
].
It is unknown why Legendre chose that letter, but Julio González Cabillón says compare capital
letter ! (Le Gendre) and the upside-down ! Or the relation between  (in Gendre) and  in
Gamma. And there is also a nice relation between the gamma function and the contant  (=
0.577...). Letter  (the one that Euler actually used in his 
 
 


  
) is third in our alphabet; Ȗ is also third in the Greek alphabet. Please mind that
Legendre also used capital  to represent the famous Euler-Mascheroni constant (= 0.577...): On
page 295 (ibidem) Legendre says:
 étant une constant dont la valeur calculée avec précision par une autre voie est  =
0,5772156649015325 donc enfin on aura,  étant très-petit log = -log  - "
See also BETA and GAMMA FUNCTIONS on the math words page.

!  * 0    # The use of ȗ for this function was introduced by Bernhard Riemann
(1826-1866) as early as 1857 (Cajori vol. 2, page 278). It appears in "Über die Anzahl der
Primzahlen unter einer gegebenen Grösse" (1859) Werke (p. 138) English translation (p. 4). See
RIEMANN HYPOTHESIS and RIEMANN ZETA FUNCTION on the Words page.

r     # P. A. Hansen used the letter . for this function in 1843 in #   

 '1    although the designation of the index and argument has varied since then.
Bessel himself used the letter  (Cajori vol. 2, page 279).

   # Log. appears as an abbreviation for logarithm in c  




c  
!
  (1616), an English translation by Edward Wright of Napier's
work.

Log. (with a period, capital "L") was used by Johannes Kepler (1571-1630) in 1624 in  

 
 (Cajori vol. 2, page 105)

log. (with a period, lower case "l") was used by Bonaventura Cavalieri (1598-1647) in
 
  0
  in 1632 (Cajori vol. 2, page 106).

log (without a period, lower case "l") appears in the 1647 edition of      by
William Oughtred (1574-1660) (Cajori vol. 1, page 193).

Kline (page 378) says Leibniz introduced the notation log  (showing no period), but he does not
give a source.

was introduced by Edmund Gunter (1581-1626) according to an Internet source. [I do not


see a reference for this in Cajori.]

ln (for natural logarithm) was used in 1893 by Irving Stringham (1847-1909) in  
c (Cajori vol. 2, page 107).

William Oughtred (1574-1660) used a minus sign over the characteristic of a logarithm in the
     (Key to Mathematics), "except in the 1631 edition which does not consider
logarithms" (Cajori vol. 2, page 110). The      was composed around 1628 and
published in 1631 (Smith 1958, page 393). Cajori shows a use from the 1652 edition.

&         .    /# Until recently [] has been the standard symbol for
the greatest integer function. According to Grinstein (1970), "The use of the bracket notation,
which has led some authors to term this the   
 stems back to the work of Gauss
(1808) in number theory. The function is also referred to by Legendre who used the now
obsolete notation #()." The Gauss reference is to 
  
 

 .
 0
"$ p. 5.
Recently the symbol has come into use. It was introduced in 1962 by Kenneth E. Iverson
who also coined the name 

 
. See CEILING FUNCTION and FLOOR FUNCTION on the
Words page.

  1# Saunders Mc Lane, in  



+
    (Springer-
Verlag, 1971, p. 29), says: "The fundamental idea of representing a function by an arrow first
appeared in topology about 1940, probably in papers or lectures by W. Hurewicz on relative
homotopy groups. (Hurewicz, W.: "On duality theorems," $c '
 47, 562-563) His
initiative immediately attracted the attention of R. H. Fox and N. E. Steenrod, whose ... paper
used arrows and (implicitly) functors... The arrow f: : X ²> Y rapidly displaced the occasional
notation f(X) (subset of ) Y for a function. It expressed well a central interest of topology. Thus a
notation (the arrow) led to a concept (category)". [Arturo Mena]

 %    %   # The symbol [ ], to represent 0, 1, or -1, according to whether is 0,


positive, or negative, was introduced by Leopold Kronecker (1823-1891). He wrote:

Bezeichnet man naemlich mit [ ] den Werth Null oder +1 oder -1, je nachdem die reelle Groesse
a selbst gleich Null oder positiv oder negativ ist ... [February 14, 1878]
This citation was provided by Julio González Cabillón

     &2   


"   1# The first use of the vinculum was in 1484 by Nicolas Chuquet (1445?-1500?)
in his !     ' 2
  The bar was placed under the parts affected (Cajori
vol. 1, pages 101 and 385). Chuquet wrote:

The above expression in modern notation is . This use of a vinculum appears to


be the earliest use of a grouping symbol of any kind mentioned by Cajori.

"   # According to Cajori, the first use of the vinculum above the parts affected was
by Frans van Schooten (c. 1615-1660), who "in editing Vieta's collected works, discarded the
parentheses and placed a horizontal bar above the parts affected." In Van Schooten's 1646 edition
of Vieta, is used to represent $( 2 + $ ).

Ball (page 242) says the vinculum was introduced by Francois Vieta (1540-1603) in 1591. This
information may be incorrect.

&2  32      # In the late fifteenth century and in the sixteenth century various
writers used letters or words to indicate grouping. The earliest use of such a device mentioned by
Cajori (vol. 1, page 385) is the use of the letter  for     by Luca Paciolo (or Pacioli) (c.
1445 - prob. after 1509) in his ' of 1494 and 1523.

   # Parentheses ( ) are "found in rare instances as early as the sixteenth century"
(Cajori vol. 1, page 390). Apparently the earliest work Cajori names in which round parentheses
are found is     
    by Nicolo Tartaglia (c. 1506-1557) in 1556.
Round parentheses occur once in c   by Cardan, as printed in 3 (1663) (Cajori vol.
1, page 392; Cajori does not indicate whether the parentheses occur in the original 1545 edition).

Cajori (vol. 1, page 391) says that Michael Stifel (1487 or 1486 - 1567) does not use parentheses
as signs of aggregation in his printed works, but that they are found in one of his handwritten
marginal notes. Cajori expresses the opinion that these parentheses are actually punctuation
marks rather than mathematical symbols.

Kline says parentheses appear in 1544. He presumably refers to c    by Michael
Stifel.

r4 # Brackets [ ] are found in the manuscript edition of c by Rafael Bombelli (1526-
1573) from about 1550 (Cajori vol. 1, page 391).

Ball (page 242) and Lucas say brackets were introduced by Albert Girard (1595-1632) in 1629.
This information appears to be inaccurate.

Kline says square brackets were introduced by Vieta (1540-1603). He presumably refers to the
1593 edition of *  which according to Cajori uses both braces and brackets.

r # Braces { } are found in the 1593 edition of Francois Vieta's * (Cajori vol. 1, page
391).

&2       # In the writing of large numbers, various methods have been
used to separate numerals into groups, including dots, vertical bars, commas, arcs, colons, and
accent marks.

In 1202, Leonardo of Pisa in ! c  directs that the hundreds, hundred thousands, hundred
millions, etc., be marked with an accent mark above, and that thousands, millions, thousands of
millions, etc., be marked with an accent below (Cajori vol. 1, page 58).

The earliest example of the modern system of simply separating the numeral into groups of three
with commas shown by Cajori is in 1795 in the article "Numeration" in     
=

  
by Charles Hutton.

          


ʌ  5#67689### Early writers indicated this constant as a ratio of two values. William Oughtred
(1574-1660) designated the ratio by the fraction ʌ over į in      The
symbolism appears in the editions of this book of 1647, 1648, 1652, 1667, 1693, and 1694
(Cajori vol. 2, page 9).

Cajori writes that "perhaps the earliest use of a single letter to represent the ratio of the length of
a circle to its diameter" occurs in 1689 in     by J. Christoph Sturm, who used 
for 3.14159....

si diameter alicuius circuli ponatur  circumferentiam appellari posse  (quaecumque enim


inter eas fuerit ratio, illius nomen potest designari littera ).
Cajori cites a note by A. Krazer in # 
 
  as a reference for the above.

The first person to use ʌ to represent the ratio of the circumference to the diameter (3.14159...)
was William Jones (1675-1749) in 1706 in '
  
 
 It is believed he
used the Greek letter pi because it is the first letter in  
(= perimeter). From Cajori (vol.
2, page 9):

The modern notation for 3.14159 .... was introduced in 1706. It was in that year that William
Jones made himself noted, without being aware that he was doing anything noteworthy, through
his designation of the ratio of the length of the circle to its diameter by the letter ʌ. He took this
step without ostentation. No lengthy introduction prepares the reader for the bringing upon the
stage of mathematical history this distinguished visitor from the field of Greek letters. It simply
came, unheralded, in the following prosaic statement (p. 263):

"There are various other ways of finding the !  or c   of particular  !  or
=  which may very much facilitate the Practice; as for instance, in the   the Diameter

is to the Circumference as 1 to , &c. = 3.14159, &c. = ʌ. This  


(among others for the same purpose, and drawn from the same Principle) I received from the
Excellent Analyst, and my much esteem'd Friend Mr. .
   4 and by means thereof, 0 
  Number, or that in Art. 64.38 may be Examin'd with all desirable Ease and Dispatch."

In 1734 Leonhard Euler (1707-1783) employed  instead of ʌ in "De summis serierum


reciprocarum."

In a letter of April 16, 1738, from Stirling to Euler, as well as in Euler's reply, the letter  is used.

In 1736 in   


    
  Euler used 1 : ʌ and "thus either
consciously adopted the notation of Jones or independently fell upon it" (Cajori vol. 2, page 10).
Euler wrote, "Si enim est  = 1/2 terminus respondens inuenitur ʌ/2 denotante 1 : ʌ rationem
diametri ad peripheriam." But the letter is not restricted to this use in his    and the
definition of ʌ is repeated when it is taken for 3.14159...

In 1737 Euler used ʌ for 3.14159... in a letter, and again in various letters in 1737, 1738, and
1739.
Johann Bernoulli used  in 1739, in his correspondence with Euler, but in a letter of 1740 he
began to use ʌ.

In 1741 ʌ was used in H. Sherwin's     

Nikolaus Bernoulli employed ʌ in his letters to Euler of 1742.

Euler popularized the use of ʌ by employing it in 1748 in  



 c     
:

Satis liquet Peripheriam hujus Circuli in numeris rationalibus exacte exprimi non posse, per
approximationes autem inventa est .. esse = 3,14159 [etc., to 128 places], pro quo numero,
brevitatis ergo, scribam ʌ, ita ut sit ʌ = Semicircumferentiae Circuli, cujus Radius = 1, seu ʌ erit
longitudo Arcus 180 graduum.
     # This constant, 2.71828..., was referred to in Edward Wright's
English translation of Napier's work on logarithms, published in 1618.

The first symbol used for the constant mentioned by Cajori is the letter  used by Leibniz in
letters to Huygens in 1690 and 1691.

Leonhard Euler (1707-1783) introduced  for this constant in a manuscript,  


 
#   


 
    (Meditation on experiments made recently
on the firing of cannon), written at the end of 1727 or the beginning of 1728 (when Euler was
just 21 years old). The manuscript was first printed in 1862 in Euler's 3 
 
      Petropoli, edited by P. H. Fuss and N. Fuss (vol ii, pp. 800-804). The
manuscript describes seven experiments performed between August 21 and September 2, 1727:

For the number whose logarithm is unity, let  be written, which is 2,7182817... [sic] whose
logarithm according to Vlacq is 0,4342944... [translated from Latin by Florian Cajori].
Euler next used  in a letter addressed to Goldbach on November 25, 1731, writing that 
"denotes that number whose hyperbolic logarithm is = 1."

The earliest appearance of  in a  work was in Euler's   (1736), in which he
laid the foundations of analytical mechanics (Maor, p. 156).

Maor writes (page 156):

Why did he choose the letter ? There is no general consensus. According to one view, Euler
chose it because it is the first letter of the word 
   More likely, the choice came to him
naturally as the first "unused" letter of the alphabet, since the letters  and  frequently
appear elsewhere in mathematics. It seems unlikely that Euler chose the letter because it is the
initial of his own name, as occasionally been suggested: he was an extremely modest man and
often delayed publication of his own work so that a colleague or student of his would get due
credit. In any event, his choice of the symbol  like so many other symbols of his, became
universally accepted.
Ball says: "It is probable that the choice of  for a particular base was determined by its being the
vowel consecutive to "

According to Boyer (page 494), this notation was "suggested perhaps by the first letter of the
word 'exponential.'"

In a post in sci.math in 1995, Wei-hwa Huang wrote: "I believe that  was not named because it
was the first letter in Euler's name, but rather because he was using vowels for constants in a
proof of his and  happened to be the second one."

In a post to a history of mathematics list in 1999, Olivier Gerard wrote: "The hypothesis made by
my friend Etienne Delacroix de La Valette was that  was for 'ein' (one in German) or 'Einheit'
(unity), which would be matching the sentence Euler uses to define it (whose logarithm is unity).
As always, many explanations may be true at the same time."

Several textbooks claim that the letter  was chosen to honor Euler. Cajori has no information to
support this claim, and in fact the earliest uses of  were by Euler himself.

The early uses of symbols for 2.718... mentioned by Cajori are as follows:

1690  Leibniz Letter to Huygens


1691  Leibniz Letter to Huygens
1703 A reviewer c  

1727/8  Euler  
 #   


 
   
1736  Euler   
    

1747  D'Alembert @
  c %
1747  Euler various articles
1751  Euler various articles
1760  Daniel Bernoulli @
  c %  
1763  J. A. Segner    
1764  D'Alembert @
  c %
1764  J. H. Lambert @
  c %   
1771  Condorcet @
  c %
1774  Abbé Sauri 
  % &
1775  J. A. Fas   
)     3  ) 
1782  P. Frisi 3 
 
1787  Daniel Melandri 2
 c @  
  
The term 2  
  has been suggested for 2.718... The name # 
  may be
inappropriate for this number, as the number was known before Euler's birth and # 
 
more frequently is used to refer to 0.577....

Benjamin Peirce suggested the innovative notation for ʌ and  shown below:

From J. D. Runkin's    


  vol. I, No. 5, Feb. 1859

A somewhat modified notation, shown below, appears in the    


(1889-1897) in
the entry
 
:

 :     # Leonhard Euler (1707-1783) used  in 


 

 

  

       

  7 (1734-35),
published in 1740, pp.150-161. Reprinted in 3 
  (1) 14, pp. 87-100.

According to Cajori (vol. 2, page 32), Mascheroni used c in c


 
  
  #  (1790-1792).

According to William Dunham in #   


 c (1999), Mascheroni introduced the
symbol Ȗ for the Euler-Mascheroni constant. Dunham's source is "On the History of Euler's
Constant," by J. W. L. Glaisher, which appeared in 1872 in   
   In
the paper, Glaisher does not specify where Mascheroni used the symbol, but seems to imply it is
in c
 
 #     which Glaisher indicates in a footnote is a work
he has not seen but which is referred to in volume 3 of Lacroix's       
 
According to the þ
#  
 
   (2003), Euler used Ȗ in 1781.

Gauss used ȥ.

Julio González Cabillón has found Ȗ in "Theoriae logarithmi integralis lineamenta nova," an
essay submitted by Carl Anton Bretschneider (1808-1878) on October 13, 1835, to   
.
  The article was published in volume 17, pp. 257-285, 1837. The symbol itself can be
found on page 260.

DeMorgan used Ȗ, according to J. W. L. Glaisher in "On the History of Euler's Constant" (1872)
in   
  

W. Shanks used "E. or Eul. constant" in =


þ
'

!

 Vol XV (1867).

The letter # was adopted by J. W. L. Glaisher in 1871 and J. C. Adams in 1878.

Ernst Pascal retained Mascheroni's notation c in 1900 (Cajori vol. 2, page 32).

In vol. VII of "L'Intermediaire des mathematiciens" (1900), G. Vacca (from Turin) asks who
introduced Ȗ. His question #1998 reads as follows:

In the German "Encyclopaedie" (1900, vol. II, p. 171) it says that Mascheroni has denoted the
Euler's constant 0.577... by Ȗ. According to my research, this author designated it by letter "A".

In "Synopsis" of Mr. Hagen (1891, vol. I, p. 86), it is said that Euler has introduced this symbol
in "Acta Petr." (1769, vol. XIV).

Mr. E. Pascal, in his "Repertorium" (1900, vol. I, p. 478 of the German edition) reproduces that
suggestion. But, in the quoted volume, and in many memoirs of Euler, I have found that this
author has just used the symbol "C", 'et parfois' "O".

Who is the first mathematician that has introduced the symbol Ȗ for the Euler's constant?

In his c
 
    #  (1790), Lorenzo Mascheroni (1750-1800)
calculated the constant to 32 decimal places:
.57721 56649 01532 86061 81120 90082 39...

In 1809 Johann von Soldner (1766-1833) published his "Théorie d'un nouvelle fonction
transcendante," in which his value of the constant given on page 13 is:

.57721 56649 01532 86060 6065...

which differs from Mascheroni's value at the twentieth decimal place. In 1812 Gauss asked F. G.
B. Nicolai (1793-1846), "juvenem in calculo indefessum," to check their results, and he agreed
with von Soldner. A note in a memoir by Gauss which contained the results of Nicolai's
calculation apparently attracted little attention and Mascheroni's value was repeatedly quoted
thereafter (Glaisher).

ij    # According to  


!"$  c
 
' 
 

 c 


+ 2  
'  
c "' þ  

  
!

 0  (1914) by Sir Theodore Andrea Cook (1867-1928), page 420:

Mr. Mark Barr . . . suggested . . . that this ratio should be called the phi proportion for reasons
given below . . . The symbol phi was given to this proportion partly because it has a familiar
sound to those who wrestle constantly with pi and partly because it is the 1st letter of the name of
Pheidias, in whose sculpture this sculpture is seen to prevail when the distance between salient
points are measured.
The above quotation and citation were provided by Samuel S. Kutler and Julio González
Cabillón. Barr was an American mathematician.

According to Gardner (1961) and Huntley, the letter phi was chosen because it is the first letter
in the name of Phidias who is believed to have used the golden proportion frequently in his
sculpture. However, Schwartzman (page 164) implies the letter stands for Fibonacci.

The Greek letter tau is also used for this constant. Tau is found in 1948 in þ =
 
 by
Harold Scott MacDonald Coxeter, according to John Conway, who believes Coxeter may have
used the symbol in his papers of the 1920s and 1930s. Ball and Coxeter (1987, page 57) write,
"The symbol [tau] is appropriate because it is the initial of tomh\ ("section") [Antreas P.
Hatzipolakis].

H. v. Baravalle used  for 0.618... in "The Geometry of the Pentagon and the Golden Section,"
which appeared in     in January 1948. He may have used the same
symbol in his "Die Geometrie des Pentagrammes und der Goldene Schnitt" in 1932.

In ' 
  =  (1999), Roger Herz-Fischler uses  for 1.618... and  for
.618....

Ê        was first used by Leonhard Euler (1707-1783) in a memoir presented in
1777 but not published until 1794 in his "Institutionum calculi integralis."

On May 5, 1777, Euler addressed to the 'Academiae' the paper "De Formulis Differentialibus
Angularibus maxime irrationalibus quas tamen per logarithmos et arcus circulares integrare
licet," which was published posthumously in his "Institutionum calculi integralis," second ed.,
vol. 4, pp. 183-194, Impensis Academiae Imperialis Scientiarum, Petropoli, 1794.

Quoniam mihi quidem alia adhuc via non patet istud praestandi nisi per imaginaria procedendo,
formulam littera  in posterum designabo, ita ut sit  = -1 ideoque 1/ = -.
According to Cajori, the next appearance of  in print is by Gauss in 1801 in the &

c   Carl Boyer believes that Gauss' adoption of  made it the standard. By 1821, when
Cauchy published 
  c   the use of  was rather standard, and Cauchy defines  as "as
if was a real quantity whose square is equal to -1."

Throughout his  

 Euler consistently writes , denoting by  the "numerus infinite
magnus" [namely, an infinitely large number]. Nonetheless, there are very few occasions where
Euler chose  with a different meaning. Thus, chapter XXI (volume 2) of Euler's  


contains the first appearance of  as &     :

Cum enim numerorum negativorum Logarithmi sint imaginarii (...) erit log(- ) quantitas
imaginaria, quae sit = 
The citation above is from "Introductio in analysin infinitorum," Lausannae, Apud Marcum-
Michaelem Bousquet & socios, M.DCC.XLVIII (1748).

Please note that, in this fascinating passage about logarithms, Euler does
 introduce the
symbol  such that 2 = -1.

[This entry was contributed by Julio González Cabillón.]

     0 # The following is taken from a paper "Africa, Cradle of Mathematics"
by Beatrice Lumpkin:

It is well known that a zero placeholder was not used or needed in Egyptian numerals, a system
of numerals without place value. Still historians such as Boyer and Gillings have found examples
of the use of the zero concept in ancient Egypt. But Gillings added, "Of course zero, which had
not yet been invented, was not written down by the scribe or clerk; in the papyri, a blank space
indicates zero." However, some Egyptologists did know that the ancient Egyptians used a zero
symbol, but it may have been missed by historians of mathematics because the symbol did not
appear in the surviving mathematical papyri.

The Egyptian zero symbol was a triliteral hieroglyph, with consonant sounds [symbol]. This was
the same hieroglyph used to represent beauty, goodness, or completion. There are two major
sources of evidence for an Egyptian zero symbol:

6# ,             # Massive stone structures such as the


ancient Egyptian pyramids required deep foundations and careful leveling of the courses of
stone. Horizontal leveling lines were used to guide the construction. One of these lines, often at
pavement level, was used as a reference and was labeled [symbol], or zero. Other horizontal
leveling lines were spaced 1 cubit apart and labeled as 1 cubit above [symbol], 2 cubits above
[symbol], or 1 cubit, 2 cubits, 3 cubits, and so forth, below [symbol]. Here zero was used as a
reference for directed or signed numbers.

In 1931, George Reisner described the terms used to label the leveling lines at the Mycerinus
(Menkure) pyramid at Giza, built c. 2600 BCE. He gave the following list collected earlier by
Borchardt and Petrie from their study of Old Kingdom pyramids. [...]
 # r44 2 % 0     # A bookkeeper's record from the 13th dynasty c 1700 BCE
shows a monthly balance sheet for items received and disbursed by the royal court during its
travels. On subtracting total disbursements from total income, a zero remainder was left in many
columns. This zero remainder was represented with the same symbol, [symbol], as used for the
zero reference line in construction.

These practical applications of a zero symbol in ancient Egypt, a society which conventional
wisdom believed did not have a zero, may encourage historians to reexamine the everyday
records of ancient cultures for mathematical ideas that have been overlooked.

According to Milo Gardner, Mesoamericans used a fully positional base 4, 5 system, with zero as
a place holder, counting 0-19, as early as 1,000 BC.

As early as the fourth century BC, the Chinese represented zero as a blank space on a counting
board (Johnson, page 160).

Babylonians in the Seleucid period (300 BC onward) used a symbol for zero, mainly in
astronomical texts, and never in final positions. According to Neugebauer in # ' 
 c &  Dover, 2nd edition, 1969, p. 20:

The Babylonian place value notation shows in its earlier development two disadvantages which
are due to the lack of a symbol for zero. The first difficulty consists in the possibility of
misreading a number 1 20 as 1,20 = 80 when actually 1,0,20 = 3620 was meant. Occasionally
this ambiguity is overcome by separating the two numbers very clearly if a whole sexagesimal
place is missing. But this method is by no means strictly applied and we have many cases where
numbers are spaced widely without any significance. In the latest period, however, when
astronomical texts were computed, a special symbol for "zero" was used. This symbol also
occurs earlier as a separation mark between sentences, and I therefore transcribe it by a "period".
Thus we find in Seleucid astronomical texts many instances of numbers like 1, . , 20 or even 1, . ,
. ,20 which apply exactly the same principle as, e.g.,our 201 or 2001.

But even in the final phase of Babylonian writing we do not find any examples of zero signs at
the end of numbers. Though there are many instances of cases like . ,20 there is no safe example
of a writing like 20, . known to me. In others words, in all periods the context alone decides the
absolute value of a sexagesimal written number.

However, Georges Iffrah (@


     Seghers, Paris, 1981, p. 400-401)
writes that Babylonian astronomers used the zero not only in the intermediate positions but also
in initial or final positions, and he gives Neugebauer as the source of this information.

According to Boyer (p. 29-30), "The Babylonians seem at first to have had no clear way in which
to indicate an "empty" position--that is, they did not have a zero symbol, although they
sometimes left a space where a zero was intended. ... By about the time of the conquest by
Alexander the Great, however, a special sign, consisting of two small wedges placed obliquely,
was invented to serve as a placeholder where a numeral was missing. ... The Babylonian zero
symbol apparently did not end all ambiguity, for the sign seems to have been used for
intermediate empty positions only. There are no extant tablets in which the zero sign appears in a
terminal position. This means that the Babylonians in antiquity never achieved an absolute
positional system."

Burton says that about A. D. 150, the Alexandrian astronomer Ptolemy began using the omicron
[which looks something like a zero] in the manner of our zero, not only in a medial but also in a
terminal position. He says there is no evidence that Ptolemy regarded the symbol as a number by
itself that could enter into computations with other numbers. Omicron is the first letter of the
Greek word for "nothing." However, Len Berggren says, "Ptolemy probably did not use omicron
to denote 0. Papyri from the period when Ptolemy lived show a small 'o' +  
  as the
symbol for 0, and the small 'o' alone doesn't come in until the Byzantine period. Even in that
period Neugebauer considers it unlikely that the small 'o' stood for the Greek word
 (=
nothing). See the discussion in the second edition of Neugebauer's Exact Sciences in Antiquity,
esp. pp. 13 - 14."

The oldest Maya artifact employing both positional notation and a zero is Pestac, Stela 1, with a
contemporaneous date of Feb. 8, AD 665. The oldest Maya artifacts employing a zero but not
positional notation are Uaxactun, Stelae 18 and 19, with a contemporaneous date of AD 357. The
oldest Maya artifact employing the same chronological system as in the previous cases but
without a zero and without positional notation is Tikal, Stela 29, with a contemporaneous date of
July 8, AD 292 (Michael Closs).

[This website previously showed what seemed to be an occurrence of a symbol for zero on a
tablet from the Old Babylonian period, about 1800 B. C. However, it is now believed that this
sole appearance was not real and may have been due to an error of the scribe who made the
tablet.]

   0   # The #  
  $   says, "Hindu literature gives evidence
that the zero may have been known before the birth of Christ, but no inscription has been found
with such a symbol before the 9th century."

According to Johnson (page 160), by the third century A.D., Hindu mathematicians were using a
heavy dot to mark its place in calculations and the dot was eventually replaced by an empty
circle.

The earliest date Bell has found to be fairly proven for the use of zero was A. D. 505 in a
=    by Varahamihira.

According to Iffrah (op. cit., p. 468), a heavy point for zero appears in a Khmer inscription of
683 found at Trapeang Bay, Sambor Province, Cambodia. A small circle for zero appears in 683
in an inscription found in Kedukan Bukit, Palebang, Sumatra. A small circle is also found in 684
at Talang Tuwo, Palebang, Sumatra, and in 686 at Kota Kapur, Banka Island.

According to Menninger (p. 400):


This long journey begins with the Indian inscription which contains the earliest true zero known
thus far (Fig. 226). This famous text, inscribed on the wall of a small temple in the vicinity of
Gvalior (near Lashkar in Central India) first gives the date 933 (A.D. 870 in our reckoning) in
words and in Brahmi numerals. Then it goes on to list four gifts to a temple, including a tract of
land "270 royal hastas long and 187 wide, for a flower-garden." Here, in the number 270 the zero
first appears as a small circle (fourth line in the Figure); in the twentieth line of the inscription it
appears once more in the expression "50 wreaths of flowers" which the gardeners promise to
give in perpetuity to honor the divinity.

Manoel Almeida reports that, following Iffrah, the date of the Gvalior inscription is A. D. 876.
According to Johnson (page 160), the date of the inscription is A. D. 840. Johnson says this
inscription is the oldest surviving use of an empty circle for zero.

Michael Closs writes that the Gvalior inscription is not the "earliest known written zero" from
India:

There are several more ancient inscriptions which also exhibit written zeros within the context of
positional notation. Rabindra Nath Mukherjee, an Indian historian of mathematics, has noted
several such artifacts. Unfortunately, he has provided neither photographs of the artifacts nor
descriptions of the texts. I am hoping to collect this information. The oldest hard evidence which
Mukherjee gives is an inscription from AD 672 in which the zero is written as a small dot. Next
is an inscription from AD 683 in which the zero is written as a large dot. This is followed by an
inscription from AD 684 in which the zero is written as a small circle. This is the earliest known
(by me!) antecedent having the same form as our modern zero symbol.

Iffrah (op. cit., p. 464-465) mentions two cooper's letters, dated from VIII AD, with small circles
to represent zeros. Both are earlier than Gwalior inscriptions, but consensus about its authenticity
is lacking.

Manoel Almeida, who contributed much of the information for this article, says he suspects, due
to the puzzling coincidence of dates and of the use of the same symbols for zero, that the AD 683
and AD 684 inscriptions mentioned by Mukherjee can be the Trapeang Bay and the Talang
Tuwo inscriptions, quoted by Iffrah. He is interested in additional information from Mukherjee's
articles.

Another source says the first zero in the Hindu system was represented by a dot and was found in
a text written by Bakhshali, the date of which is unknown. The Bakhshali manuscript may have
been written in the 8th or 9th century, but may have been written later and may not even be of
Hindu origin (Smith, vol. 1, page 164).

According to Georges Iffrah in @


     Seghers, Paris, 1981, p.489-
490, the first European to advocate the use of the Hindu zero was Abraham ben Meir ibn Ezra
(1092-1167), who wrote ' @ (The Book of the Number), in which he used the circle to
represent zero. He preferred to use the first nine letters of the Hebrew alphabet rather than the
nine Hindu-Arabic numbers. He called the zero    (Hebrew for wheel) or Sifra (apud the
arab Sifr, certainly). He also changed the old Hebrew alphabetic numeration system into a new
decimal positional system like ours.

Leonardo of Pisa (1180-1250) (or Fibonacci) also advocated the use of zero, using the term
(  in ! c 

The zero symbol first appears in print in the 1200s, according to Burton. Smith says that Ch'in
Kiu-shao (or Tsin Kiu tschaou or Ts'in K'ieou-Chao) of China used 0 in 1247 or 1257 in 
2 '

  .

Rida A. K. Irani, in a paper in Centaurus, vol. 4 (1955), pp. 1-12 gives forms of the Hindu-
Arabic characters as they occur in dated Arabic manuscripts, and shows that the medieval Arabs
used a small 'o' for zero. This tends to degenerate to a dot in late (e.g. 17th century) manuscripts.

         


      # The Babylonians wrote numbers in a system which was almost
a place-value (positional) system, using base 60 rather than base 10. Their place value system of
notation made it easy to write fractions. The numeral

has been found on an old Babylonian tablet from the Yale collection. It is an approximation for
the square root of two. The symbols are 1, 24, 51, and 10. Because the Babylonians used a base
60, or sexagesimal, system, this number is 1 x 600 + 24 x 60-1 + 51 x 60-2 + 10 x 60-3, or about
1.414222.

The Babylonian system of numeration was not a pure positional system because of the absence
of a symbol for zero. In the older tablets, a space was placed in the appropriate place in the
numeral; in some later tablets, a symbol for zero does appear but in the tablets which have been
discovered, this symbol only used between other symbols and never in a terminal position.

The earliest Egyptian and Greek fractions were usually unit fractions (having a numerator of 1),
so that the fraction was shown simply by writing a numeral with a mark above or to the right
indicating that the numeral was the denominator of a fraction.

c   ! # The Romans did not use numerals to indicate fractions, but instead used words
to indicate parts of a whole. A unit of weight was the  and the   (from which we have the
word "ounce") was a twelfth part of the  The following words were used to indicate parts of
the  or, more generally, parts of any quantity:

11/12   for    1/12 taken away


10/12   for   1/6 taken away
9/12 
  for &   1/4 taken away
8/12    for    2/3
7/12   for   
6/12 
5/12 &   for & &  
4/12   
3/12 &  
2/12  
1/12  
1/24  
1/48 
1/72  
1/144  
1/288  

Multiples of the  were indicated using the following scheme, in which a denarius represents 16
asses.     represented 1/24 + 1/48 of a denarius or 1/16 denarius, or 1 as.
     represented 1/12 + 1/24 of a denarius or 1/8 denarius, or 2 asses.  
  represented 1/6 + 1/48 of a denarius, or 3/16 denarius, or 3 asses.   
 represented 11/12 + 1/48 of a denarius, or 15/16 denarius, or 15 asses [Smith vol. 2,
pages 208-209].

    1  0  # According to Smith (vol. 2, page 215), it is
probable that our method of writing common fractions is due essentially to the Hindus, although
they did not use the bar. Brahmagupta (c. 628) and Bhaskara (c. 1150) wrote fractions as we do
today but without the bar.

 0    was introduced by the Arabs. "The Arabs at first copied the Hindu
notation, but later improved on it by inserting a horizontal bar between the two numbers"
(Burton).

Several sources attribute the horizontal fraction bar to al-Hassar around 1200.

When Rabbi ben Ezra (c. 1140) adopted the Moorish forms he generally omitted the bar.
Fibonacci (c.1175-1250) was the first European mathematician to use the fraction bar as it is
used today. He followed the Arab practice of placing the fraction to the left of the integer (Cajori
vol. 1, page 311).

According to the '$ Abu Abdallah Yaish ibn Ibrahim ibn Yusuf ibn Simak al-Umawi (14th or
15th century) insisted that the horizontal fraction bar be used, whereas easterners continued to
write it without the bar.

The bar is generally found in Latin manuscripts of the late Middle Ages, but when printing was
introduced it was frequently omitted, doubtless owing to typographical difficulties. This
inference is confirmed by such books as Rudolff's )     (1526), where the bar is
omitted in all ordinary fractions but is inserted in fractions printed in larger type and those
having large numbers (Smith vol. 2, page 216).

Michael Closs points out that if we define a horizontal fraction bar to be a horizontal line that
separates the numerator from the denominator and demarcates them as such, then this type of
notation was used with exactly that purpose more than a millennium before al-Hassar. In

   =   (Brown University Press, London, 1972, pages 8-9) Richard A.
Parker writes that in three papyri dating from the third century B. C. to the Roman period, "the
numerator is written first, and the denominator follows on the same line. In problems 2, 3, 10,
and 13 (the Cairo papyrus) the numerator is underlined. In problems 51 and 72 the denominator
is underlined."

Some writers use the term   for the horizontal fraction bar. This term originally applied
to the mark when used as a grouping symbol. Fibonacci used the Latin word   for the
horizontal fraction bar.

     (also called a solidus or virgule) was introduced because the
horizontal fraction bar was difficult typographically, requiring three terraces of type.

An early handwritten document with forward slashes in lieu of fraction bars is Thomas Twining's
Ledger of 1718, where quantities of tea and coffee transactions are listed, e.g. 1/4 pound green
tea. This usage of the horizontal fraction bar was found by Hans Lausch, who believes there are
likely even earlier occurrences.

Lausch has also found the horizontal fraction bar in c  $
 a Berlin
review journal which was started in 1765. A precise reference may be forthcoming.

The earliest instance of a diagonal fraction bar shown by Cajori (vol. 1, page 313) is in 1784,
when a curved line resembling the sign of integration was used in the  ( 
by
Manuel Antonio Valdes.

In 1843, a curved line was used by Henri Cambuston in  


    

 
  (Cajori vol. 1, page 313)
In 1845, the use of the solidus was recommended by De Morgan in an article "The Calculus of
Functions" published in the #  
  

 of 1845 (Cajori vol. 1, page 313).

In 1852, the solidus was used by Antonio Serra Y Oliveres in   
  # 5

(Cajori vol. 1, page 313).

   # Abu'l Hasan Ahmad ibn Ibrahim Al-Uqlidisi (c. 920-c. 980) wrote the
earliest known text offering a direct treatment of decimal fractions. "Al-Uqlidisi uses decimal
fractions as such, appreciates the importance of a decimal sign, and suggests a good one,"
according to A. S. Saidan, "The earliest extant Arabic arithmetic,"  57 (1966), 475-490.

The idea of decimal fractions had been present in the work of several mathematicians of al-
Karaji's school, in particular Ibn Yahya al-Maghribi Al-Samawal (c. 1130-c. 1180), according to
the University of St. Andrews website.

In ) 
c  Ghiyath al-Din Jamshid Mas'ud al-Kashi (c. 1380-1429) gave a clear
description of decimal fractions, according to P. Luckey, þ    
  )  (1951).

Al-Kashi in his þ  


6 (  
   ) wrote the value of pi using
Arabic characters as follows:

mmmmmmmmmmm m
mmmmmmmmmmmmmmmmmmmm
   m
The word    meant complete, correct, integral. (The modern Turkish form is  ) Thus
the part at the right is the decimal, although there is no decimal point. According to Smith (vol.
2, page 240), "Manifestly it is, therefore, a clear case of a decimal fraction, and it seems to be
earlier than any similar one to be found in Europe."

Yang Hui (c. 1238-c. 1298) was a minor Chinese official who wrote two books, dated 1261 and
1275, which use decimal fractions (in the modern form). The 1275 work is called  

$ $ 
[University of St. Andrews website].

Smith (vol. 1, page 255) writes, "Francesco Pellos or Pellizzati, a native of Nice, published a
commercial arithmetic at Turin in 1492 in which, as will be shown in Volume II, use is made of
a decimal point to denote the division of a number by a power of ten." In vol. 2 (page 138) Smith
says Pellos "unwittingly made use of the decimal point for the first time in a printed work" and
that he "did not recognize the significance of the decimal point." Cajori (vol. 1, page 315) says
Pellos "used a point and came near the invention of decimal fractions."

In 1530, Christoff Rudolff (1499?-1545?) used a vertical bar exactly as we use a decimal point
today in setting up a compound interest table in the #$7 (Cajori vol. 1, page 316).

Smith (vol. 2, page 240) writes:

The first man who gave evidence of having fully comprehended the significance of all this
preliminary work seems to have been Christoff Rudolff, whose #$7 appeared at
Augsburg in 1530. In this work he solved an example in compound interest, and used the bar
precisely as we should use a decimal point today. If any particular individual were to be named
as having the best reason to be called the inventor of decimal fractions, Rudolff would seem to
be the man, because he apparently knew how to operate with these forms as well as merely to
write them, as various predecessors had done. His work, however, was not appreciated, and
apparently was not understood, and it was not until 1585 that a book upon the subject appeared.

In 1579 Francois Vieta (1540-1603) published a work which included a systematic use of
decimal fractions, using a vertical stroke as a separator; "from the vertical stroke to the actual
comma there is no great change" (Cajori vol. 1, page 316).

In 1585 Simon Stevin (or Stevinus) (1548-1620) published !   ("The Tenth") and ! 
 ("The Decimal"), both of which explained the use of decimal fractions. He is credited
with introducing decimal fractions into common use, although he did not use the notation we use
today. He wrote 5.912 as or .

Boyer writes (on page 340):

The use of a decimal point separatrix generally is attributed either to G. A. Magini (1555-1617),
a map-making friend of Kepler and rival of Galileo for a chair at Bologna, in his  
   of 1592, or to Christoph Clavius (1537-1612), a Jesuit friend of Kepler, in a table of
sines of 1593. But the decimal point did not become popular until Napier used it more than
twenty years later.
Jobst Bürgi (1552-1632) "was not clear as to the best method of representing these fractions,
however, and in his manuscript of 1592 he used both a period and a comma for the decimal
point" (Smith vol. 2, page 243-244). He also used instead a small circle placed above or below
the units digit (Smith vol. 2, page 244 and Cajori (vol. 1, page 317).

In 1593 Christopher Clavius (1537-1612) used a period to separate the units and tenths digits in a
table of sines in c
  However, he used the period for other reasons in his works, and his
purpose in using the period in this case is not clear (Cajori vol. 1, page 322). Carl Boyer says
Clavius was the first person to use the decimal point with a clear idea of its significance.

William Oughtred (1574-1660) did not use a decimal point, but instead wrote 0.56 as 0/56, with
the 56 underlined.

The dot as a separator occurs in 1616 in E. Wright's translation of John Napier's  

Boyer refers to this as the first appearance of a decimal point separating the whole number part
from the decimal part, in the notation we use today. However Cajori (vol. 1, page 323) says "no
evidence has been advanced, thus far, to show that the sign was intended as a separator of units
and tenths, and not as a more general separator as in Pitiscus." According to Scott (p. 128),
"Wright's translation of his treatise on logarithms, which was published in 1616 shows the
decimal point on the first page."
In 1617 in his Latin þ 

 , Napier used both the comma and the period as separators of
units and tenths. Before 1617, he used the period in his 
 
 which was not published
until 1619 (Cajori vol. 1, page 324).

 2     is believed to have evolved from a symbol introduced in an anonymous


Italian manuscript of about 1425, according to D. E. Smith in þ   in 1898.

           


" 
Most of the basic notation for matrices and vectors in use today was available by the early 20th
century. Its development is traced in volume 2 of Florian Cajori¶s @

   
2
 
 published in 1929. Cajori made much use of Thomas Muir¶s 4-volume 


   @
 3  
 
  (1906-24) which covered the years 1990-
1900; a supplement (1930) brought the story up to 1920.

The modern reader of Muir will be struck that he invested so much in a history of determinants
but determinants seemed so much more central at the end of the 19th century when Muir began
work than they do now. The modern reader of Cajori will be struck by how     the
material was organised and the emphasis distributed in his time. Thus matrices appear in a sub-
section of ³Determinant notations´ and vectors in ³Symbolism for imaginaries and vector
analysis.´ Muir¶s investment and Cajori¶s organisation reflect how the subject(s) developed. The
emerging field of abstract linear spaces is hardly noticed by Cajori.

Notation for Systems of Linear Equations


The modern notation for systems of linear equations did not become established until the early
twentieth century, though by then such systems had been studied continuously in the West for
perhaps 200 years. MacTutor: Matrices and determinants describes how GAUSSIAN ELIMINATION
(see below) was known to Chinese mathematicians of the Han dynasty.

A few of the many notations presented by Muir and Cajori will be illustrated here.

Muir and Cajori quote a remarkable passage from a letter of 1693 from Leibniz to L¶Hospital in
which something like the modern scheme of literal coefficients with numerical subscripts ( 88,
8, etc.) is suggested. The letter appears ! (  '  $ 9 (ed. C.
I. Gerhardt) p. 239
However Leibniz¶s economical notation remained an isolated curiosity and there was no
standardisation on this or any other scheme. The schemes described by Muir and Cajori did not
just use numbers but exploited the ordering provided by the alphabet and the possibilities of
adding primes.

Two important papers from 1772 by Vandermonde and Laplace illustrate some possibilities.
Vandermonde treats the constants in a similar way to Leibniz but replaces the alphabetic
ordering of the variables by a numerical one. Laplace combines numerical ordering, alphabetical
ordering and repeated primes applied to a single symbol representing an unknown.

Vandermonde¶s paper in @


  c %
  Ann. 1772 p. 523 presents a
pair of equations in the unknowns ȟ 1 and ȟ 2. The constants are on two levels with the upper
level representing the equation (1 or 2) and the lower level representing the number of the
variable²with 3 representing the constant. (I apologise for the poor reproduction):

Laplace¶s memoir in Histoire de l'Académie royale des sciences Ann. 1772 p. 294. shows how
the alphabet, repeated primes and generous use of ³&c.´ (after the third instance) could be
combined to good effect²the use of the numerical pre-fix to indicate the number of the
equation, as in 1 , 2 , 3 did
 catch on:
(This was the paper in which Laplace described what is now called the LAPLACE EXPANSION.)
Elements of this notation remained in use for a long time: see Gauss¶s notation from 1811
(below) and Cayley¶s matrix and determinant notation from the 1850s (below).

Forty years after these contributions from Vandermonde and Laplace, what is essentially the
modern notation appears in Cauchy¶s 
  of 1815 (the paper that introduced the term
DETERMINANT). See, for instance, this set of equations (Oeuvres (2) i: p. 130). Cauchy used
numerical subscripts for the variables and doubly indexed numerical subscripts, e.g. Ô6%6, for the
coefficients

Although Cauchy¶s 


  was recognised as a very important paper, its notation did not
become established as  notation for almost another century: at the beginning of the 20th
century it appears in the textbooks by M. Bôcher  


@ c (1909) and G.
Kowalewski # 7       
 (1909).

Gaussian Elimination & Least Squares. One of the most long-lived notations
for One of the most long-lived notations for expressing linear equations is
 discussed by Muir
or Cajori. It belongs to the numerical analysis of linear equations and was devised by Gauss for
the method now called GAUSSIAN ELIMINATION. Gauss used the method for solving the NORMAL
EQUATIONS associated with the METHOD OF LEAST SQUARES. Both in Gauss¶s time and
subsequently the least squares problem has had a major influence on the theory and formalism of
linear algebra. See the review of different least squares formalisms in J. Aldrich Doing Least
Squares: Perspectives from Gauss and Yule,   
'  þ+, ÑÑ, (1998), 61-81.
Gauss describes the method and the associated notation in section 13 of the &

# #=   (1811) in  
: pp. 20-22. (Section 2 in P. M. Lee¶s
translation Application of the Method of Least Squares to the Elements of the Planet Pallas.)

In the use of primes and alphabetical ordering Gauss¶s practice resembles Laplace but the
abbreviations were an innovation.

(In the modern regression notation (below) V, VV, VVV, etc. are the elements of the V vector
and the Vs, Vs, etc. are elements of the ' matrix.)

Using his abbreviations Gauss writes the normal equations (the equations to be solved) as

(where , &, ,  etc. are the elements of the i vector. )

When Gauss reduces this system to triangular form he uses constructions of the form

for the new constant when  is eliminated and then for the new constant when & is eliminated
The notation survived well into the twentieth century when it was replaced by matrix
formulations of least squares calculations. For a typical textbook presentation see Chauvenet
  
'   =  c



Symbols for determinants. See DETERMINANT on 



All the authors who produced notation for systems of equations produced notation for
determinants. A great variety of notations are displayed by Muir and Cajori.

The use of a single vertical line on both sides of the entries seems to have introduced by Cayley
writing in 1841: see Muir vol. 1 p. and Cajori vol. 2, p. 92.

The notation appears in the      .


  Vol. II (1841), p. 267-271,
reprinted in =   vol. 1, p.1. On p. 1 Cayley writes

Apart from the use of commas to separate entries within rows, this is how determinants are
written today.

By the early twentieth century the modern combination of bar lines and the Ô; symbols for
elements are found in textbooks: see Bôcher¶s  


@ c (1909) and
Kowalewski¶s # 7       
 (1909).

Symbols for Matrices. See MATRIX on 



Determinants had been studied for well over a century and a half before the idea of a matrix
appeared. Cayley wrote about matrices on several occasions without seeming to settle on a fixed
notation.

In his ³Mémoire sur les Hyperdéterminants´ of 1846,  , 8< p. 2 Cayley uses the double
vertical line notation found in works from the 1930s.

In his most substantial work on matrices, ³A Memoir on the Theory of Matrices´ of 1855
(=  , II, 475-96), Cayley does not use such nice notation. He writes (p. 475).

A set of quantities arranged in the form of a square, e.g.


is said to be a matrix.

Cayley explained the motivation for introducing matrices:

The notion of such a matrix arises naturally from an abbreviated notation for a set of linear
equations, viz. the equations

However Cayley did


 manipulate such matrix equations nor did he introduce a symbol for
non-square matrices like (/, ;, *): his matrix manipulations were confined to forming powers of
square matrices.

Cayley¶s main interest was in powers of square matrices (below) and matrix polynomials. The
big result in the 1855 paper is the CAYLEY-HAMILTON THEOREM. Cayley uses upper-case letters
for matrices but there is nothing systematic about his notation.

Matrix algebra was rediscovered by Frobenius; see ³Ueber lineare Substitutionen und bilineare
Formen,´ .   +  Vol. 84 (1878) pp.1-63. The notation in this and in his later
works was much more sophisticated than that used by Cayley.

The printing of Cayley¶s matrices always looks improvised but more permanent forms followed.
Cajori (vol. 2, p. 103) writes that round parentheses were used for matrices by many, including
Maxine Bocher in 1909 in  


@ c and G. Kowalewski in 1909 in
   
 (although Kowalewski also used double vertical lines and a single brace).
In the 1930s books on matrices written in English started to appear. The two leading ones were
C. C. MacDuffee 

   Springer (1933) and J. H. M. Wedderburn ! 

   (1934). They both wrote matrices with double vertical lines as in

Associated symbols
Matrix inverse. See INVERSE on 


In the ³Memoir´ Cayley describes the formation of powers of matrices including the inverse of
the matrix: see Papers, I, 480.

Identity matrix and zero matrix. See IDENTITY MATRIX and ZERO MATRIX on 
.

Cayley used 1 for the identity matrix and 0 for the zero matrix. Both Wedderburn ! 

   (1934, p. 8) and MacDuffee use  and 3.

Transpose. See TRANSPOSE on 


.

MacDuffee 

   (1933, p. 5) reports, ³Many different notations for the transpose
have been used, as cV, M, ,c, c1, 1c´ His preference is for cT: ³The present notation is in
keeping with a systematic notation which, it is hoped, may find favour.´ Wedderburn ! 

   (1934, p. 8) writes cV for what he calls the   .

Kronecker product. See KRONECKER PRODUCT on 


.

F.D. Murnaghan 




þ   
 (1938, p. 68) writes c ' $ for the Kronecker
product. The same symbols is used byWedderburn Lectures on Matrices (1934, p. 74) although
he uses the term  
 as does MacDuffee 

   (1933, p. 81).
O
MacDuffee writes c ' $although the dot is not printed quite so high.

Generalized inverse. See GENERALIZED INVERSE on 




The generalized inverse of c is written as c‚ by R. Penrose in ³A Generalized Inverse for


Matrices,´ =
  =
'
., 86, (1955), 406-413. However the symbol c quickly
became popular; see e.g. C. R. Rao (1962) Note on a Generalized Inverse of a Matrix with
Applications to Problems in Mathematical Statistics, .
 
þ
'  '
 $,
 7, 152-158.

Symbols for Vectors.


Vector notation and indeed the idea of a vector developed independently of matrix ideas; the
vector development had more to do with generalising complex numbers. M. J. Crowe c@


0
c   (1967/87) describes the nineteenth century work of Hamilton, Grassmann and
Gibbs.

Of these Gibbs seems to have had the greatest influence on the notation, but not through his own
writing. His vector analysis became widely known through E. B. Wilson¶s 0
c  "c
$


 
' 
   = 
 
! 
.
  (1901).

Many symbols have come and gone: e.g. it is now standard to use the ordinary equals sign to
represent equality between vectors but Cajori (pp. 133-4) shows earlier writers reluctant to do
this and using a variety of equals-like symbols.

Vectors and scalars

In his # 


0
c   (1881) Gibbs writes, ³we shall use small Greek letters to
denote vectors and the small English letters to denote scalars.´ (' =  ,  , p. 17).

One of the innovations in Wilson¶s book was very influential, the use of   ./ 2
for vectors and ordinary type for scalars. E.g. ³Thus if c be the vector and  the scalar, the
product is denoted by  c or c .´

Wilson used the bold/ordinary pair to representing a vector and its length: if c is a vector, then c
is the scalar denoting its  . Later writers used the bold/ordinary type possibility to mark
other distinctions. In Hermann Weyl¶s þ *   ('    4th edition
1921) the bold type  represents the vector but the ordinary type i represents a component of the
vector. German writers also used old German (Gothic) characters to represent vectors and
modern script for components: see Courant and Hilbert¶s 
     = 
(1924).
Following Hamilton¶s example, Wilson writes the ³3 fundamental unit vectors´ as , ; and 4.
Hamilton was extending the notation for complex numbers. The corresponding notation in
Weyl¶s þ *   (for -dimensional spaces) is 1 ,.., n .

Associated symbols
Scalar product and inner product. See INNER PRODUCT and DOT PRODUCT on 
.

In the #  (1881) Gibbs wrote <= for what he called the  
 of the vectors < and
=

In 0
c   Wilson wrote cðr. The notation inspired the alternative term for the direct
product, the 



Other authors used symbols resembling Gauss¶s symbol: see above. Cajori (p. 135) reports that
(,) was used in Henrici & Turner¶s 0
 þ

 (1903). In his paper axiomatising
HILBERT SPACE, ³Allgemeine Eigenwerttheorie Hermitescher Funktionaloperatoren,´ Math.
Ann. 102, (1929), 49-131, von Neumann used the symbol (,)

Vector product. See VECTOR PRODUCT on 


.

In the #  (1881) Gibbs wrote < ' = for what he called the +
 of the vectors<
and =

In 0
c   Wilson wrote the skew product as c ' r. The notation inspired the term 


. See CROSS PRODUCT.

Length or norm. See NORM

Wilson¶s use of ordinary type has been mentioned. The absolute value symbol of Weierstrass
(see Function Symbols), or some modification of it, has often been used. The symbols |/| and
ŒijŒ both appear in Banach¶s ³Sur les opérations dans les ensembles abstraits et leur application
aux équations integrales´,        , 5, (1922) 133-181.

Matrix notation in least squares/regression.


For the past fifty years or so it has been common to write the REGRESSION model in matrix
notation as

where  is the vector of values of the DEPENDENT VARIABLE, ) is the DESIGN MATRIX and = is
the vector of ERRORS.
Such notation can be found in J. Durbin and G. S. Watson ³Testing for Serial Correlation in
Least Squares Regression: I,´ $
  , 5>, (1950) and O. Kempthorne's   c  

#   (1952). Matrix notation was first used in the 1920s but the most noticed of the
early contributions was a paper by Aitken, ³On least squares and linear combinations of
observations,´ =
þ
'
#  , 55, (1935), 42-48.

See also !   and 3  in Symbols in Probability & Statistics. The use of
matrix theory in statistics is described by R. W. Farebrother ³A. C. Aitken and the Consolidation
of Matrix Theory,´ !  c  c 
,  Ñ7, (1997), 3-12. For the earlier Gauss
notation that was used in works on the combination of observations, or theory of errors, see
above.

          


  2    
 # In 1583, Thomas Fincke (or Finck) used sin. (with a period) in Book 14 of his 
  

  Cajori writes that "perhaps the first use of abbreviations for the trigonometric lines goes
back to ... Finck" (Cajori vol. 2, page 150).

In 1624, Edmund Gunter (1581-1626) used sin (without a period) in a drawing representing
Gunter's scale (Cajori vol. 2, page 156). However, the symbol does not appear in Gunter's work
published the same year.

In 1626, Girard designated the sine of c by c and the cosine of c by (Smith vol. 2, page 618).

In a trigonometry published by Richard Norwood in London in 1631, the author states that "in
these examples  stands for  :  for   :  for  
 :  for   

 :  for  
 :  for   
  :  for  " (Smith vol. 2,
page 618).

In 1632, William Oughtred (1574-1660) used sin (without a period) in c


 
0

     
=


(Cajori vol. 1, page 193, and vol. 2, page 158).

According to Smith (vol. 2, page 618), "the first writer to use the symbol  for   in a book
seems to have been the French mathematician Hérigone (1634)." However, the use by Oughtred
would seem to predate that of Hérigone.

 # In his 
  
  (1588) Thomas Fincke used sin. com. for the cosine.

In 1674, Sir Jonas Moore (1617-1679) used Cos. in    
  (Cajori vol. 2,
page 163).
Samuel Jeake (1623-1690) used cos. in c  published in 1696 (Cajori vol. 2, page 163).

The earliest use of cos Cajori shows apparently is by Leonhard Euler in 1729 in 
  
c  ' =

   8>? (Cajori vol. 2, page 166).

Ball and Asimov say cos was first used by William Oughtred (1574-1660). Ball gives the date
1657; Asimov gives 1631. However, Cajori indicates Oughtred used only sco for cosine (Cajori
vol. 1, page 193) and Cajori reports Glaisher says Oughtred did not even use the word cosine.

  # In 1583, Thomas Fincke (1561-1656) used tan. in Book 14 of his 


  
 
(Cajori vol. 2, page 150). Fincke also used tang.

In 1632, William Oughtred (1574-1660) used tan in  


=


(Cajori vol. 1,
page 193).

  # In 1583, Thomas Fincke (1561-1656) used sec. in Book 14 of his 


  
 
(Cajori vol. 2, page 150).

In 1632, William Oughtred (1574-1660) used sec in  


=


(Cajori vol. 1,
page 193).

  # In his 
  
  (1588), Thomas Fincke used sec. com. for the cosecant
(Cajori vol. 2, page 150).

Samuel Jeake in his c  (1696) used cosec. for cosecant (Cajori vol. 2, page 163).

Simon Klügel in c   



  (1770) used cosec for cosecant.

It appears that the earliest use Cajori shows for csc is in 1881 in   
 

  by
Oliver, Wait, and Jones (Cajori vol. 2, page 171).

  # In his 
  
  (1588), Thomas Fincke used tan. com. for the cotangent
(Cajori vol. 2, page 150).

In 1674, Sir Jonas Moore (1617-1679) used Cot. in    
  (Cajori vol. 2,
page 163).

Samuel Jeake in his c  (1696) used cot. for cotangent (Cajori vol. 2, page 163).

A. G. Kästner in c   7  c 


  

  (1758) used cot for
cotangent (Cajori vol. 2, page 166).

  # Irving Stringham used cis ȕ for cos ȕ +  sin ȕ in 1893 in  c
(Cajori vol. 2, page 133).
         # Except where noted otherwise, the following citations
are from Cajori vol. 2, page 175-76.

Daniel Bernoulli was the first to use symbols for the inverse trigonometric functions. In 1729 he
used A S. for arcsine in 
   =
 Vol. II.

In 1736 Leonhard Euler (1707-1783) used A t for arctangent in   


  
Later in the same publication he used simply A

In 1737 Euler used A sin for arcsine in 


     =

   8>@>
Vol. IX.

In 1769 Antoine-Nicolas Caritat Marquis de Condorcet (1743-1794) used arc (sin. = ) (Katz,
page 42).

In 1772 Carl Scherffer used arc. tang. in  


      

In 1772 Joseph Louis Lagrange (1736-1813) used arc. sin in 2


 
   
  

According to Cajori (vol. 2, page 176) the inverse trigonometric function notation utilizing the
exponent -1 was introduced by John Frederick William Herschel in 1813 in the =

 
  

!

 A full-page footnote explained his choice of notation for the inverse
trigonometric functions, such as cos.-1  which he used in the body of the article (Cajori vol. 2,
page 176).

However, according to         (1908) by Daniel A. Murray, "this
notation was explained in England first by J. F. W. Herschell in 1813, and at an earlier date in
Germany by an analyst named Burmann. See Herschell, c


# 

c 

 
     (Cambridge, 1820), page 5, note."

According to Cajori, in France, Jules Houël (1823-1886) used arcsin.

In 1914, =  

 by George Wentworth and David Eugene Smith has:

In American and English books the symbol sin-1 is generally used; on the continent of Europe
the symbol arc sin is the one that is met. The symbol sin-1 is read "the inverse sine of " "the
antisine of " or "the angle whose sine is " The symbol arc sin is read "the arc whose sine is
" or "the angle whose sine is "
In 1922 in  


  William F. Osgood wrote, "The usual notation on the
Continent for sin-1,  tan-1  etc., is arc sin  arc tan  etc. It is clumsy, and is followed for a
purely academic reason; namely, that sin-1  might be misunderstood as meaning the minus first
power of sin  It is seldom that one has occasion to write the reciprocal of sin  in terms of a
negative exponent. When one wishes to do so, all ambiguity can be avoided by writing (sin )-1."
1        # The practice of placing the exponent beside the symbol for
the trigonometric function to indicate, for example, the square of the sine of  was used by
William Jones in 1710. He wrote cs2 (for cosine) and s2 (for sine) (Cajori vol. 2, page 179).

 %   %   # See the 


 page.

! # G. B. Halstead in   
(1881) suggested using the Greek letter rho to indicate
radians.

G. N. Bauer and W. E. Brooke in =  '   



 (1907) suggested the use of

(the lower case r in a raised position) to indicate radians.

A. G. Hall and F. G. Frink in =  



 (1909) suggested the use of þ (the capital R in
a raised position).

P. R. Rider and A. Davis in =  



 (1923) suggested the use of (r) (the lower case r
in parentheses) to indicate radians.

2    # Vincenzo Riccati (1707-1775), who introduced the hyperbolic functions,
used Sh. and Ch. for hyperbolic sine and cosine (Cajori vol. 2, page 172). He used Sc. and Cc.
for the circular functions.

Johann Heinrich Lambert (1728-1777) further developed the theory of hyperbolic functions in
@
   %þ
  $   vol. XXIV, p. 327
(1768). According to Cajori (vol. 2, page 172), Lambert used sin h and cos h.

According to Scott (page 190), Lambert began using   and 


 in 1771.

According to  2+  


 
 2nd. ed., sinh is an abbreviation for  
  


In 1902, George M. Minchin proposed using    


   etc.: "If the prefix  were put
to each of the trigonometric functions, all the names would be pronounceable and not too long."
The proposal appeared in 2   vol. 65 (April 10, 1902).

        " 


& 4   # The use of letters to represent general numbers goes back to Greek antiquity.
Aristotle frequently used single capital letters or two letters for the designation of magnitude or
number (Cajori vol. 2, page 1).

Diophantus (fl. about 250-275) used a Greek letter with an accent to represent an unknown. G.
H. F. Nesselmann takes this symbol to be the final sigma and remarks that probably its selection
was prompted by the fact that it was the only letter in the Greek alphabet which was not used in
writing numbers. However, differing opinions exist (Cajori vol. 1, page 71).

In 1463, Benedetto of Florence used the Greek letter rho for an unknown in   
  
    (Franci and Rigatelli, p. 314)

!    # In Leonardo of Pisa's !    (1202) the representation of given numbers


by small letters is found (Cajori vol. 2, page 2).

Jordanus Nemorarius (1225-1260) used letters to replace numbers.

Christoff Rudolff used the letters  and  to represent numbers, although not in algebraic
equations, in $  @þ   (1525) (Cajori vol. 1, page 136).

Michael Stifel used & (abbreviation for &  (which Cardan had already done) but he also
used c$  and  for unknowns in 1544 in c    (Cajori vol. 1, page 140).

Girolamo Cardan (1501-1576) used the letters and  to designate known numbers in   
( (1570) (Cajori vol. 1, page 120).

In 1575 Guilielmus Xylander translated the c  of Diophantus from Greek into Latin
and used 2 (  ) for unknowns in equations (Cajori vol. 1, page 380).

In 1591 Francois Vieta (1540-1603) was the first person to use letters for unknowns and
constants in algebraic equations. He used vowels for unknowns and consonants for given
numbers (all capital letters) in     

 Vieta wrote:

Quod oopus, ut arte aliqua juventur, symbolo constanti et perpetuo ac bene conspicuo date
magnitudines ab incertis quaesititiis distinguantur ut [illegible in Cajori] magnitudines
quaesititias elemento A aliave litera volcali, #30; [illegible in Cajori] elementis $ 
aliisve consonis designando. [As one needs, in order that one may be aided by a particular
device, some unvarying, fixed and clear symbol, the given magnitudes shall be distinguished
from the unknown magnitudes with the letter c or with another vowel #3 ;, the given
ones with the letters $ or other consonants.]
(Cajori vol. 1, page 183, and vol. 2, page 5).

Thomas Harriot (1560-1621) in c c   =  c& 


c   used lower
case vowels for unknowns and lower case consonants for known quantities.

 *   * V O The following is from Cajori (vol 1, page 381):

The use of (  . . . to represent unknowns is due to René Descartes, in his ! %
 
(1637). Without comment, he introduces the use of the first letters of the alphabet to signify

+ quantities and the use of the last letters to signify  
+ quantities. His own langauge
is: "...l'autre, !2 est (1/2) la moitié de l'autre quantité connue, qui estoit multipliée par ( que ie
suppose estre la ligne inconnue." Again: "...ie considere ... Que le segment de la ligne c$ qui est
entre les poins c et $ soit nommé  et quie $ soit nommé ; ... la proportion qui est entre les
costés c$ et $þ est aussy donnée, et ie la pose comme de ( a ; de façon qu' c$ estant  þ$ sera
/( et la toute þ sera = /(. ..." Later he says: "et pour ce que $ et $c sont deux quantités
indeterminées et inconnuës, ie les nomme, l'une ; et l'autre  Mais, affin de trouver le rapport de
l'une a l'autre, ie considere aussy les quantités connuës qui determinent la description de cete
ligne courbe: comme c que je nomme  )! que je nomme  et 2! parallele a c que ie
nomme " As co-ordinates he uses later only  and  In equations, in the third book of the
%
%   predominates. In manuscripts written in the interval 1629-40, the unknown ( occurs
only once. In the other places  and occur. In a paper on Cartesian ovals, prepared before 1629,
 alone occurs as unknown, being used as a parameter. This is the earliest place in which
Descartes used one of the last letters of the alphabet to represent an unknown. A little later he
used  ( again as known quantities.

Some historical writers have focused their attention upon the  disregarding the and ( and the
other changes in notation made by Descartes; these wrtiers have endeavored to connect this 
with older symbols or with Arabic words. Thus, J. Tropfke, P. Treutlein, and M. Curtze
advanced the view that the symbol for the unknown used by early German writers, looked so
much like an  that it could easily have been taken as such, and that Descartes actually did
interpret and use it as an  But Descartes' mode of introducing the knowns  etc., and the
unknowns (  makes this hypothesis improbable. Moreover, G. Eneström has shown that in a
letter of March 26, 1619, addressed to Isaac Beeckman, Descartes used the symbol as a
symbol in form distinct from  hence later could not have mistaken it for an . At one time,
before 1637, Descartes used  along the side of ; at that time  ( are still used by him as
symbols for known quantities. German symbols including the for  as they are found in the
algebra of Clavius, occur regularly in a manuscript due to Descartes, the 38:8?
8:8

All these facts caused Tropfke in 1921 to abandon his old view on the origin of  but he now
argues with force that the resemblance of  and , and Descartes' familiarity with , may
account for the fact that in the latter part of Descartes' %
%  the  occurs more frequently
than ( and  Eneström, on the other hand, inclines to the view that the predominance of  over
and ( is due to typographical reasons, type for  being more plentiful because of the more
frequent occurrence of the letter  to and ( in the French and Latin languages.

Descartes introduced the equation  +  =  which is still used to describe the equation of a
line (Johnson, page 145).

Johnson says (on page 145):

The predominant use of the letter  to represent an unknown value came about in an interesting
way. During the printing of ! 
  and its appendix, 
 ! 
 which
introduced coordinate geometry, the printer reached a dilemma. While the text was being typeset,
the printer began to run short of the last letters of the alphabet. He asked Descartes if it mattered
whether   or ( was used in each of the book's many equations. Descartes replied that it made
no difference which of the three letters was used to designate an unknown quantity. The printer
selected  for most of the unknowns, since the letters and ( are used in the French language
more frequently than is 
There are, however, other explanations for Descartes' use of   and ( for unknowns. For
example, the in the definition of  in  2+  
 
(1909-1916) and
the subsequent second edition of the same dictionary, it is claimed that "/ was used as an
abbreviation for Arabic  a thing, something, which, in the Middle Ages, was used to
designate the unknown, and was then prevailingly transcribed as " Cajori says there is no
evidence for this.

According to the 3


#  
(2nd ed.):

The introduction of  ( as symbols of unknown quantities is due to Descartes (%
% 
1637), who, in order to provide symbols of unknowns corresponding to the symbols  of
knowns, took the last letter of the alphabet, ( for the first unknown and proceeded backwards to
and  for the second and third respectively. There is no evidence in support of the hypothesis
that  is derived ultimately from the mediaeval transliteration  of "thing", used by the
Arabs to denote the unknown quantity, or from the compendium for L.  "thing" or 
"root" (resembling a loosely-written ), used by mediaeval mathematicians.
Descartes used letters to represent only positive numbers; a negative number could be
represented as - (Cajori vol. 2, page 5).

John Hudde (1633-1704) was first to allow a letter to represent a positive or negative number, in
1657 in  
 & 
 published at the end of the first volume of F. Van
Schooten's second Latin edition of René Descartes' %
%  (Cajori vol. 2, page 5).

Jonas Moore wrote in c  (1660): "Note alwayes the given quantities or numbers with
Consonants, and those which are sought with Vowels, or else the given quantities with the
former letters in the Alphabet, and the sought with the last sort of letters, as (  &c. lest you
make a confusion in your work."

 2 3   # The +  notation was introduced by Leonhard Euler (1707-1783).

        &  


         # The designation of points, lines, and planes by a letter or letters
was in vogue among the ancient Greeks and has been traced back to Hippocrates of Chios (about
440 B. C.) (Cajori vol. 1, page 420, attributed to Moritz Cantor).

       # Richard Rawlinson in a pamphlet prepared at Oxford sometime between


1655 and 1668 used c$ for the sides of a triangle and  for the opposite angles. In his
notation, c was the largest side and  the smallest (Cajori vol. 2, page 162).
Leonhard Euler and Thomas Simpson reintroduced this scheme many years later, Euler using it
in 1753 in @
   %$  (Cajori vol 2., page 162). Euler used capital letters
for the angles.

In 1866, Karl Theodor Reye (1838-1919) proposed the plan of using capital letters for points,
lower case letters for lines, and lower case Greek letters for planes in a remarkable two-volume
work on geometry, 
  !  (Cajori vol. 1, page 423).

As early as 1618, an anonymous writer of the "Appendix" in the 1618 edition of Edward
Wright's translation of John Napier's "Mirifici logarithmorum canonis descriptio" labeled the
right angle of a triangle with the letter c:

It will bee conuenient in euery calculation, to haue in your view a triangle, described according
to the present occasion: and if it bee a right angled triangle, to note it with Letters A.B.C: so that
A may bee alwayes the right angle; B the angle at the Base B.A and C the angle at the Cathetus
CA [sic].
Cf. page 3 of "An Appendix to the Logarithmes of the Calculation of Triangles, and also a new
and ready way for the exact finding out of such lines and logarithmes as are not precisely to be
found in the Canons", in John Napier's "A description of the Admirable Table of Logarithmes
...", London, Printed for Simon Waterson, 1618

James W. L. Glaisher (1848-1928) has remarked that the letter c is taken to be the right angle in
the right-angled triangle c$ in order that $c may represent the BAse, and c the CAthetus, the
first two initials indicating the words. The fact that this lettering was also employed by William
Oughtred (1574-1660) in his books is one of the many arguments in support that Oughtred might
be the author of the "Appendix" (Cajori vol. 2, p. 154).

c  # Pierre Hérigone (1580-1643) used both and < in     This work was
published in 1634 and in a second edition in 1644. Cajori lists the symbols from the 1644
edition, which shows both angle symbols (Cajori vol. 1, page 202).

c# The arc symbol appears in the middle of the twelfth century in Plato of Tivoli's translation
of the !  
 by Savasorda (Cajori vol. 1, page 402).

 # Heron used a modified circle with a dot in the center to represent a circle around A. D.
150 (Cajori vol. 1, page 401).

Pappus used a circle with and without a dot in the center to represent a circle in the fourth
century A. D. (Cajori vol. 1, page 401).

  # Heron about A. D. 150 used a triangle as a symbol for triangle (Cajori).

   # Gottfried Wilhelm Leibniz (1646-1716) introduced for congruence in an


unpublished manuscript of 1679 (Cajori vol. 1, page 414).
The first appearance in print of Leibniz' sign for congruence was in 1710 in the   
$
   in the anonymous article "Monitum," which is attributed to Leibniz (Cajori vol. 2,
page 195).

In 1777, Johann Friedrich Häseler (1732-1797) used (with the tilde reversed) in c   7 
 c c
   (Lemgo), #  
  (Cajori vol. 1, page 415).

In 1824 Carl Brandan Mollweide (1774-1825) used the modern congruent symbol in # 
#  (Cajori vol. 1, page 415).

!# Leonhard Euler introduced the use of þ for the radius of the circumscribed circle and
for the radius of the inscribed circle (Boyer, page 495).

 # The symbols for degrees, minutes, and seconds were used by Claudius Ptolemy (c. 85-
c. 165) in the c . However, the notation differed somewhat from the modern notation, and
according to Cajori (vol. 2, page 143), "it is difficult to uphold" the view that our signs for
degrees, minutes, and seconds are of Greek origin.

The first modern appearance of the degree symbol ° Cajori found is in the revised 1569 edition
of   c    

  by Gemma Frisius (1508-1555),
although the symbol appears in the Appendix on astronomical fractions due to Jacques Peletier
(1517-1582) and dated 1558. Cajori writes:

This is the first modern appearance that I have found of ° for   or "degrees." It is explained
that the denomination of the product of two such denominate numbers is obtained by combining
the denominations of the factors; minutes times seconds give thirds, because 1+2=3. The
denomination ° for integers or degrees is necessary to impart generality to this mode or
procedure. "Integers when multiplied by seconds make seconds, when multiplied by thirds make
thirds" (fol. 62, 76). It is possible that Peletier is the originator of the ° for degrees. But nowhere
in this book have I been able to find the modern angular notation ° ' " used in writing angles. The
° is used only in multiplication.
Erasmus Reinhold (1511-1553) used ° ' " in =     

 published in
1571 (Cajori).

   # A bar above c$ to indicate line segment c$ was used in 1647 by Bonaventura


Cavalieri (1598-1647) in 
     and #  

   according
to Cajori.

2 # The earliest known use of  for slope appears in Vincenzo Riccati¶s memoir 


@   


 
 
 
 which is chapter XII of the first part of his book
0  þ 3
  =  A       (1757):

Propositio prima. Aequationes primi gradus construere. Ut Hermanni methodo utamor, danda est
aequationi hujusmodi forma =  +  quod semper fieri posse certum est. (p. 151)
The reference is to the Swiss mathematician Jacob Hermann (1678-1733). This use of  was
found by Dr. Sandro Caparrini of the Department of Mathematics at the University of Torino.
C. B. Boyer in his @

c  
 (1956, p. 205) points to a passage in Monge
where the ³modern point-slope form of plane equation of the straight line is given explicitly,
perhaps for the first time.´ See ³Mémoire sur la Théorie des Déblais & des Remblais,´ @
 
 c %
  1784 (Année 1781) p. 669.

In 1830,  #    c &c   ! #


  B  2 
"
= '! 
 @#
has =  +  [Karen Dee Michalowicz].

Another use of  occurs in 1842 in c #    


      by Rev.
Matthew O'Brien, from the bottom of page 1: "Thus in the general equation to a right line,
namely =  +  if we suppose the line..." [Dave Cohen].

O'Brien used  for slope again in 1844 in c  


= 
3  
 [V.
Frederick Rickey].

George Salmon (1819-1904), an Irish mathematician, used =  +  in his c  




'
 which was published in several editions beginning in 1848. Salmon referred in several
places to O'Brien's 
'
 and it may be that he adopted O'Brien's notation. Salmon used
to denote the -intercept, and gave the equation (x/a) + (y/b) = 1 [David Wilkins].

Karen Dee Michalowicz has found an 1848 British analytic geometry text which has =  + 

The 1855 edition of Isaac Todhunter's   


= 
3  
 has =  + 
[Dave Cohen].

In 1891,         by George A. Osborne has - = ( -  ).

In  2+  
 
(1909), the "slope form" is =  + 

In 1921, in c  


   c   by Frank Loxley Griffin, the equation is
written =  + 

In c  
 (1924) by Arthur M. Harding and George W. Mullins, the "slope-intercept
form" is =  + 

In c$ 
  c c by Buchanan and others (1937), the "slope form" is =
 + 

According to Erland Gadde, in Swedish textbooks the equation is usually written as =  + 


He writes that the technical Swedish word for "slope" is "riktningskoefficient", which literally
means "direction coefficient," and he supposes  comes from "koefficient."

According to Dick Klingens, in the Netherlands the equation is usually written as =  +  or


 + & or  +  He writes that the Dutch word for slope is "richtingscoëfficiënt", which literally
means "direction coefficient."
In Austria  is used for the slope, and  for the y-intercept.

According to Julio González Cabillón, in Uruguay the equation is usually written as =  +  or


=  +  and slope is called "pendiente," "coeficiente angular," or "parametro de direccion."

According to George Zeliger, "in Russian textbooks the equation was frequently written as = 
+  especially when plotting was involved. Since in Russian the slope is called 'the angle
coefficient' and the word 
  is spelled with  in the Cyrillic alphabet, usually nobody
questioned the use of  The use of  is less clear."

It is not known why the letter  was chosen for slope; the choice may have been arbitrary. John
Conway has suggested  could stand for "modulus of slope." One high school algebra textbook
says the reason for  is unknown, but remarks that it is interesting that the French word for "to
climb" is 
  However, there is no evidence to make any such connection. Descartes, who
was French, did not use  In     þ (1971) mathematics historian
Howard W. Eves suggests "it just happened."

  # Two vertical bars, written horizontally and resembling the modern equal sign, were
used by Heron about A. D. 150 and by Pappus (Cajori).

Thee parallel symbol written vertically was first used by William Oughtred (1574-1660) in
3    @     which was published posthumously in 1677 (Cajori
vol. 1, page 193).

John Kersey (1616-1677) also used the vertical parallel symbol. He used it after Oughtred, but in
a work which was published before Oughtred. He used the symbol in c  which was
published in 1673. Kersey switched the lines from horizontal to vertical because of the adoption
of the equal symbol (Cajori vol. 1, page 411).

 2 # was first used by Pierre Hérigone (1580-1643) in 1634 in  


   which was published in five volumes from 1634 to 1637 (Cajori vol. 1, page 408).
Johnson (page 149) says, "Herigone introduced so many new symbols in this six-volume work
that some suggest that the introduction of these symbols, rather than an effective mathematics
text, was his goal."

!   # was used by Pappus (Cajori vol. 1, page 401).

 :2   # A capital ' was first used by Leonhard Euler (1707-1783) in 1750 (Cajori
1919, page 235).

 # ~ was introduced by Gottfried Wilhelm Leibniz (1646-1716) in a manuscripts of


1679 which were not published by him. The symbol was an S for  written sideways. The
original manuscripts do not survive and it is uncertain whether the symbol Leibniz first used
resembled the tilde or the tilde inverted (Cajori vol. 1, page 414).
In the manuscript of his    
  he wrote: "similitudinem ita notabimus: ~
" (Cajor vol. 1, page 414).

The first appearance in print of Leibniz' sign for similarity was in 1710 in the   
$
   in the anonymous article "Monitum," which is attributed to Leibniz (Cajori vol. 2,
page 195).

###% #-##% and -##-# for the triangle congruence theorems and axioms were invented by
Julius Worpitzky (1835-1895), professor at the Friedrich Werder Gymnasium in Berlin (Cajori
vol. 1, page 424). (W for  =angle)

An article in     in March 1938 uses a.s.a. = a.s.a. and s.s.s. = s.s.s. and
s.a.a. = s.a.a. as reasons in a proof. An article in the same journal in 1940 uses C.p.c.t.e., which is
written out as "Corresponding parts of congruent triangles are equal."

An article in     in April 1948 has: "The three common theorems on
congruence of triangles (SAS = SAS; ASA = ASA; SSS = SSS) are 'proved' by superposition."

          



  
The entries below are organised as follows. Because of the interconnectedness of the histories the
categories overlap.

Vm Combinatorial analysisü Many of the symbols of elementary combinatorial analysis found in


modern probability and statistics books were created in the 19th century. In 19th century Britain
probability was most visible in algebra textbooks as an application of combinatorial analysis.
Vm 2  normal distribution˜ also known as the Gaussian distribution˜ the second law of Laplace˜ the
law of error ... ˜ has been studied since the 18th century and many people have left their tracks
on the notation.
Vm =robabilityü At the turn of the 20th century there was a revival of interest in probability in
continental Europe. The central limit theorem was one of their main concerns. The main
contributors were Russian and French and they created much of the modern terminology and
notation around 1930.
Vm átatisticsü None of the notation used by Laplace and Gauss and their followers has survived into
modern statistics. The oldest notation still in use comes from the period 1890-1940 when the
British biometrician/statisticians Karl Pearson and R. A. Fisher introduced many of the basic
symbols and many of the principles for constructing new ones.

The languages of English statistics and continental European probability came together in the
1940s.
á á
     áá
¢actorial. An early factorial symbol˜ ˜ was suggested by Rev. Thomas Jarrett (1805-1882) in 1827. It
occurs in a paper "On Algebraic Notation" that was printed in 1830 in the Transactions of the Cambridge
Philosophical Society and it appears in 1831 in An Essay on Algebraic Development containing the
Principal Expansions in Common Algebra, in the Differential and Integral Calculus and in the Calculus of
Finite Differences (Cajori vol. 2˜ pages 69˜ 75).

The notation ! was introduced by Christian Kramp (1760-1826) in 1808 as a convenience to the
printer. In his C%  %&   (1808), Kramp wrote:

Je me sers de la notation trés simple n! pour désigner le produit de nombres décroissans depuis n
jusqu'à l'unité˜ savoir n(n - 1)(n - 2) ... 3.2.1. L'emploi continuel de l'analyse combinatoire que je fais dans
la plupart de mes démonstrations˜ a rendu cette notation indispensable.

In "Mémoire sur les facultés numériques˜" published in J. D. Gergonne's Annales de Mathématiques [vol.
III˜ 1812 and 1813]˜ Kramp writesü

1. [...] Je donne le nom de Facultés aux produits dont les facteurs constituent une progression
arithmétique˜ tels que

a(a + r)(a + 2r)...[a + (m-1)r];

et, pour désigner un pareil produit, j'ai proposé la notation

am$r.

Les facultés forment une classe de fontions très-élementaires, tant que leur exposant est un
nombre entier, soit positif soit négatif; mais, dans tous les autres cas, ces mêmes fonctions
deviennent absolument transcendantes. [page 1]

2. J'observe que toute faculté numérique quelconque est constamment réductible ô la forme trés-
simple

1m$1 = 1 . 2 . 3 ... m

ou à cette autre forme plus simple [page 2]

m!˜

si l'on veut adopter la notation dont j'ai fait usage dans mes C%  %&  
no. 289. [page 3]

[Julio González Cabillón; Cajori vol. 2, p. 72]


In # 
c '
$ (1958), Albert Eagle advocated writing ! rather
than !, so that the operator would precede the argument, as it does in most cases [Daren Scot
Wilson].

In his article "Symbols" in the =  


  (1842) De Morgan complained: "Among the
worst of barabarisms is that of introducing symbols which are quite new in mathematical, but
perfectly understood in common, language. Writers have borrowed from the Germans the
abbreviation ! to signify 1.2.3.( - 1).  which gives their pages the appearance of expressing
surprise and admiration that 2, 3, 4, &c. should be found in mathematical results" [Cajori vol. 2,
p. 328].

      2   # Leonhard Euler (1707-1783) designated the binomial


coefficients by over within parentheses and using a horizontal fraction bar in a paper written
in 1778 but not published until 1806. He used used the same device except with brackets in a
paper written in 1781 and published in 1784 (Cajori vol. 2, page 62).

The modern notation, using parentheses and no fraction bar, appears in 1826 in 

 
c   by Andreas von Ettingshausen [Henry W. Gould]. According to Cajori
(vol. 2, page 63) this notation was introduced in 1827 by Andreas von Ettingshausen in
0
  7 1    Vol. I.

Harvey Goodwin used = for the number of permutations of things taken at a time in 1869
and earlier. The notation appears in his #  
 
   3rd ed. (Cajori vol.
2, page 79).

G. Chrystal used  for the number of combinations of things taken at a time in c 
Part II (1899) (Cajori vol. 2, page 80).

á á áá


    á


The normal distribution has been studied for nearly 300 yearsü see the entries CENTRAL LIMIT THEOREM˜
GAUSSIAN and NORMAL. It has been written in many forms and the forms used today are relatively
recent.

The normal distribution was first obtained as the limiting form of the binomial distribution in the
early 18th century by Abraham De Moivre (the 20th century editor provides the equation in
footnote 2). From the early 19th century the normal distribution was the foundation of the theory
of errors, developed for use in astronomy and geodesy. The normal distribution went by various
names, including the law of error and the probability curve. Although the most important early
contributor was Laplace, the most common way of writing the normal distribution--at least in the
English literature--came from Gauss.

2  
C. F. Gauss's (1777-1855) Theoria Motus Corporum Coelestium in Sectionibus Conicis Solem Ambientum
(The Theory of the Motion of Heavenly Bodies moving around the Sun in Conic Sections) Werke 7 of 1809
was extremely influential. It presented the normal distribution in conjunction with the method of least
squares.

The following equation appears on p. 244

Using modern conventions for brackets and squares this would be written

The errors (ǻ) are centred on D. In the English literature the quantity  or its reciprocal, was
often called the modulus. See MODULUS

A typical presentation of Gauss's ideas can be found in Chauvenet's A Manual of Spherical and
Practical Astronomy ... with an Appendix on the Method of Least Squares (4th edition, 1871).
The section on "the probability curve" (pp. 478-485) discusses Gauss's function which appears
on p. 484.

2  

From the middle of the 19th century the error distribution was used in the theory of gases (and later
statistical mechanics). J. Clerk Maxwell (1831-1879) wrote˜ instead of a density˜ "The number of particles
whose velocity˜ resolved in a certain direction˜ lies between x and x + dx is given by

where 2 is the total number of particles. "Illustrations of the Dynamical Theory of Gases,"
=

   ( , 69, (1860), 19-32.

r  
Biometry (q.v.) appeared at the end of the 19th century. Karl Pearson (1857-1936) was responsible for
most of the mathematical machinery. Pearson was a follower of Maxwell rather than Gauss. His
principal innovation was a new measure of dispersion˜ the standard deviation (q.v.) but he also
popularised a new name for the distribution (see the entry normal). Pearson wrote
where "c is the total number of units measured˜ or the area of the probability curve." See p. 80 of
"Contributions to the Mathematical Theory of Evolution˜" Philosophical Transactions of the Royal Society
of London A˜ ^ ˜ 71-110.

In most of the biometric literature the number of units was represented by 2. See e.g. the
equation for the sample mean in Section III of Student¶s "The Probable Error of a Mean",
$
  , Ñ, (1908), 1-25.


   
  
R. A. Fisher (1890-1962)˜ the most influential statistician of the first half of the 20th century˜ changed the
form of the normal distribution principally by presenting the case with non-zero mean as the typical
case. Fisher had learnt the theory of errors as a student and in his first paper "On an Absolute Criterion
for Fitting Frequency Curves" (1912˜ p. 157) uses the Gauss notation but with a change

where  is the mean.

Fisher soon went over to the biometric notation (but without the  or 2). He wrote the bivariate
density in his 1915 paper on correlation (p. 508). When he next needs the univariate form he
writes "the chance of any observation falling in the range  is

from A Mathematical Examination of the Methods of Determining the Accuracy of an


Observation by the Mean Error, and by the Mean Square Error (1920, p. 758.) Fisher generally
used  to denoted this chance--see the expression on p. 508 of his 1915 paper.

Fisher wrote the normal density like this (see section 12 of his Statistical Methods for Research
Workers) until the mid-1930s when he replaced  with ȝ. The new symbol appears in The
Fiducial Argument in Statistical Inference (1935) and it went into the 1936 (sixth) edition of the
'  

þ 
 

2 
   

hile the biometricians and statisticians were using the normal distribution˜ the probability theorists
were developing ever more refined versions of the central limit theorem. (Although the term central
limit theorem (q.v.) dates from the early 20th century˜ the subject begins with Laplace.) here modern
statistical notation for the normal differs from Fisher's˜ the changes mainly reflect the influence and
prestige of modern probability theory. (See symbols of probability.)

#    

Fisher usually wrote the density in the form D but more recently  as been reserved for
the density function and  for the distribution function and so a more "correct" way of writing
would be D(See symbols in probability.) However the differential notation has gone
out of fashion and it is more usual to write some variation of

The subscript / is used if there a danger of confusion with other random variables. (See symbols
in probability.)

#     

Modern texts often write the density function for the standard normal as

in accordance with Halperin, Hartley & Hoel's "Recommended Standards for Statistical Symbols
and Notation. ..., (c  '  , 69, (1965), p. 12). They recommend ĭ for the
distribution function and the corresponding lower case letter for the density, however "the use of
the variable, z, as argument, is optional." ĭ had been used in influential probability works by
Cramér (1937) and Feller (1950). (See symbols in probability.) In recent decades ( has come to
be very widely used, particularly in the expression "z score." In earlier decades ( was not
available as it was established with a different meaning in the analysis of variance and in
correlation.

Information on the history of the term   


  is here.

á á  

Apart from the combinatorial symbols very little of the notation of modern probability dates from
before the 20th century.

# Symbols for the probability of an event c on the pattern of =(c) or = (c) are a
relatively recent development given that probability has been studied for centuries. A. N.
Kolmogorov's            (1933) used the symbol (c).
The use of upper-case letters for events was taken from set theory where they referred to sets H.
Cramér's þ 
0   =
    
 (1937), "the first modern book on
probability in English," used =(c). In the same year J. V. Uspensky ( 



   =
  ) wrote simply (c), following A. A. Markov
       (1912, p. 179) W. Feller's influential c  



=
  
 c 

8 (1950) uses = {A} and {c}in later editions.

See also the entry PROBABILITY and the "Earliest Uses of Symbols of Set Theory and Logic"
page of this website.

   2# Kolmogorov's (1933) symbol for conditional probability ("die


bedingte Wahrscheinlichkeit") was B(c). Cramér (1937) wrote P$ (c) and referred to the
"relative probability." Uspensky (1937) used the term "conditional probability" and took the
symbol (c$) from A. A. Markov¶s        (1912, p. 179). The vertical
stroke notation = {c | $} was made popular by Feller (1950), though it was used earlier by H.
Jeffreys. In Jeffreys¶s '     (1931) =( | &) stands for "the probability of the
proposition p on the data &" Jeffreys mentions that Keynes and Johnson, earlier Cambridge
writers, had used E&4 Jeffreys himself had used =( : &). The symbols  and & came from
Whitehead and Russell's =      

See also the entries CONDITIONAL PROBABILITY and POSTERIOR PROBABILITY and the "Earliest
Uses of Symbols of Set Theory and Logic" page of this website.

32  # A large script E was used for the expectation in W. A. Whitworth's well-known
textbook 
   (fifth edition) of 1901 but neither the symbol nor the calculus of
expectations became established in the #  literature until much later. For example, Rietz
   '  (1927) used the symbol # and commented that "the expected value of the
variable is a concept that has been much used by various continental European writers..." For the
continental European writers # signified "Erwartung" or "'éspérance."

!   # The use of upper and lower case letters to distinguish a random variable from
the value it takes, as in = {/ = 6 }, became popular around 1950. The convention is used in
Feller's  


=
  


         # The use of  for the generic distribution function
has been established in the probabillity literature since the 1920s. Paul Lévy  
=
 % (1925) (p. 136), conforming to the usual notation for the Stieltjes integral.

Lévy uses  for the density function but its use in that role was not automatic--thus Cramér
(1937) uses  for the characteristic function corresponding to . Since the 1940s the  for
distribution function and  for density convention (within the broader convention of using the
upper-case and corresponding lower-case letters in these roles) has been widely adopted,
particularly by statisticians, following the treatises by M. G. Kendall c 


' (1943) and S. S. Wilks    ' (1944).
 and  are often adorned with affixes to register the random variable concerned. Kolmogorov
(1933) wrote  but now / is more common--in accordance with the convention that upper case
letters are for the names of random variables.

The      2 symbol plim was introduced by H. B. Mann and A Wald "On
Stochastic Limit and Order Relationships," c 
   ' , 67, (1943), 217-
226. The      3 and
, modelled on the 3 and
, or Landau, symbols
(see Symbols used in number theory), were introduced in the same paper.

á áá


á
ÿotation for =aramtrs and Estimats. In the 1890s Karl Pearson used some Greek letters as symbols
but the use of Greek letters to represent parameters only became systematic in the 1920s and -30s. By
then R. A. Fisher was insisting on clearly distinguishing parameters and estimates (see PARAMETER on the
Math ords page). Two conventions for representing parameters and the corresponding estimates
came into use. One forms the estimate by adding a circumflex/caret/hat (or other accent) to the Greek
character representing the parameter. The other uses corresponding Greek and Latin characters for
parameter and estimate. Fisher tended to favour the Graeco-Latin convention but he also used the hat
device˜ mainly in conjunction with in general discussions of estimation as in his Phil. Trans. R. Soc. 1922
paper. The most familiar example of the Graeco-Latin convention˜ s for an estimate of ʍ˜ dates from
1908 but the general convention came much later. The entries on specific symbols show the
development of the conventions. (Based on Aldrich (2003 and 1998))

   2  is a relic of a convention that has otherwise vanished from probability


and statistics. It derives from the practice of applied mathematicians of representing kind of
average by a bar. J. Clerk Maxwell's "On the Dynamical Theory of Gases (=

 
  

þ
'
 , 68>, (1867) p. 64) uses for the "mean velocity" of
molecules while W. Thomson & P. G. Tait's   
2  =

 (1879) uses for
the centre of inertia, ( = + / ) Karl Pearson, the leading statistician of the early 20th
century, had such a physics background. Pearson and his contemporaries used the bar for sample
averages  for expected values but eventually # replaced it in the latter role. The survival of
for the sample mean is probably due to the influential example of R. A. Fisher who used it in
all his works; the first of these was "On an Absolute Criterion for Fitting Frequency Curves"
(1912). See Expectation in ' 
 =
  above and also AVERAGE, MEAN and
EXPECTATION on the Math Words page.

        # (See STANDARD DEVIATION and VARIANCE on the Math
Words page.) The use of ı for standard deviation first occurs in Karl Pearson's 1894 paper,
"Contributions to the Mathematical Theory of Evolution," =

   


þ
'
 
!

' c 6?8, 71-110. On page 80, he wrote, " Then ı will be termed its
standard-deviation (error of mean square)" (David, 1995). When Fisher introduced variance in
1918 he did not introduce a new symbol but instead used ı2.
Pearson's notation did not distinguish between parameter and estimate. Student (W. S. Gosset) in
"The Probable Error of a Mean", $
   Ñ, (1908), 1-25, used  for an estimate of ı, though
contrary to modern practice his divisor was , not ( - 1). Fisher eventually adopted Student's 2
(with adjusted ) as an estimate of ı2 beginning with his 1922 paper, "The goodness of fit of
regression formulae, and the distribution of regression coefficients" (.þ
' '
, ?8,
597-612).

 # Pearson introduced the basic symbol ¼ to which numerical subscripts would be added
to indicate the order and a prime could be added to indicate about which value the moment is
taken. Originally the moment was given by an expression of the form <¼ where < is the "area of
the entire system;" see e.g. Contributions to the Mathematical Theory of Evolution. II. Skew
Variation in Homogeneous Material, =

   

þ
'
 c, 6?Ñ, p.
347. Eventually the area was normalised to unity and the moment coefficient became the
moment. Fisher applied the Graeco-Latin convention and twinned the ¼'s with 's in his paper on
cumulants (1929). See MOMENT on the Math Words page.

  # (See CORRELATION on the Math Words page.) When Galton introduced correlation
in "Co-Relations and Their Measurement", =
þ'
 45, 135-145, 1888 (also on Galton
website) he chose the symbol for the index of co-relation, perhaps for its affinity with
regression. The use of F for the population linear correlation coefficient is found in 1892 in F. Y
Edgeworth, "Correlated Averages," =

   ( 9'  57, 190-204. The
symbol appears on page 190 (David, 1995).

Karl Pearson, who dominated correlation research from the mid-1890s, favoured (for both
parameter and estimate), using F only if a second correlation symbol was required; thus both
symbols appear on p. 302 of Contributions to the Mathematical Theory of Evolution. Note on
Reproductive Selection," =
þ'
, 89, (1895-6), 301-305. Student (W. S. Gosset) in "The
Probable Error of the Correlation Coefficient" ($
   6, 302-310 1908) had different
symbols for the parameter value (þ) and for the estimate ( ). H. E. Soper ($
   9, 91-115,
1913) used F and in these roles. R. A. Fisher used the Soper symbols from his first work in
correlation (1915).

G. Udny Yule introduced the notation 12#3 for the 2   between 1 and 2 holding
3 fixed in his 1907 "On the Theory of Correlation for any Number of Variables, Treated by a
New System of Notation," =
þ'
' c 79, pp. 182-193. The Greek forms, including F
12#3, followed in M. S. Bartlett's 1933 "On the theory of statistical regression," =
þ
'

#  , 85, 260-283.

þ has been used for the double, triple, ..., -fold or 2   coefficient, at least
since Yule used it in 1896. þ is now generally used for the   coefficient. This is awkward
because the upper-case F the natural choice for the population coefficient, is the unappealing
letter, =

!   # (See REGRESSION and METHOD OF LEAST SQUARES on the Math Words page.)
Modern regression analysis has its roots in Gauss's work (1809/-23) on the use of least squares
for combining observations and in the work of Galton and Pearson on heredity. Gauss's notation
can be seen in Chauvenet's Manual pp. 509ff with the special notation for Gaussian elimination
on pp. 530ff. Pearson¶s correlation-based notation can be seen in the equation for @1 on p. 241 of
his "Note on Regression and Inheritance in the Case of Two Parents," =
þ'
, 8?, (1895),
240-2. The notational highpoint of the correlation/regression development was Yule¶s "On the
Theory of Correlation for any Number of Variables, Treated by a New System of Notation,"
=
þ'
, c, >9, (1907), 182-193 where b12#.3 stands for the partial regression of 1 on 2
holding 3 fixed. (Cf. correlation notation above).

Yule¶s regression notation is used sometimes in multivariate analysis but the most familiar
modern regression notation dates from the 1920s when R. A. Fisher drew the Gauss and Pearson
lines together. In his '  

þ 
  (1925) Fisher presents regression
using and  and the terms "dependent variable" and "independent variable." For the population
values of the intercept and slope Fisher uses Į and ȕ, for the estimates he uses and  This
textbook exposition was based on a 1922 paper, "The goodness of fit of regression formulae, and
the distribution of regression coefficients" (.þ
' '
, ?8, 597-612.

3  in regression was first used in the 1920s but only came into wide use in the
1950s. The most noticed of the early contributions was a paper by A. C. Aitken, "On least
squares and linear combinations of observations," =
þ
'
#  , 88, (1935), 42-
48. This paper is also notable for its account of what has been called "Aitken¶s generalised least
squares." Aitken appears not to have regarded this work highly; it belonged with the "mere
applications ... to standard problems." The practice of writing an error term in the equation also
became common around 1950. See the entry ERROR on the Math Words page.

@     A 4 1 A 2  # R. A. Fisher established the role and ș in it in "On the


Mathematical Foundations of Theoretical Statistics" (= þ'
 1922) and the papers
that followed. However Fisher had already used the notation in his first publication, a paper he
wrote as a third year undergraduate, "On an Absolute Criterion for Fitting Frequency Curves "
(  
   1912, 76: 155-160).

ê     .      /     2   :# Fisher


introduced this notation in his 1929 paper "Moments and Product Moments of Sampling
Distributions", =
 
!

   '
 '  30, 199-238. He
introduced the cumulant notation into the 1932 (fourth) edition of the '  


þ 
 

¼         # (See ' 


 
 +
 
 
) ¼, as the symbol for the mean of the normal distribution, was surprisingly late in
becoming established. Fisher adopted it in the 1936 (sixth) edition of the '  


þ 
  He had been using  since 1912. He had always used for the sample mean.

    1    2  #

=: . Please see the entry on the mathematical words page here.
@ was used to represent "the hypothesis in which we are particularly interested" in J. Neyman
and E. S. Pearson¶s "On the Problem of the Most Efficient Tests of Statistical Hypotheses, "
=

   

þ
'
 
!

' c  56. (1933), pp. 289-337.
They had referred to "Hypothesis A" in their 1928 paper, "On the use and Interpretation of
Certain Test Criteria for Purposes of Statistical Inference. Part I," $
    < c, 175-240.
See the entry HYPOTHESIS & HYPOTHESIS TESTING on the mathematical words page here.

B   . 3  / 4  # This symbol was introduced by J. Neyman and E. S.
Pearson in their "On the Use of Certain Test Criteria for Purposes of Statistical Inference, Part II"
$
   (1928),  <c% 263-294. They called the quantity it denoted the 

 but later
authors called it the likelihood ratio. See the entry LIKELIHOOD RATIO on the mathematical
words page here.

  0      appears in J. Neyman and E. S. Pearson¶s "Contributions to


the Theory of Testing Statistical Hypotheses," '  þ 
  6, (1936), 1-37. In
their 1933 paper, which introduced "size," they had used the symbol İ. See the entry SIZE on the
mathematical words page here.

i   21    was introduced by J. Neyman "Tests of Statistical Hypotheses Which


are Unbiased in the Limit," c 
   ' , 9, (1938), p. 79: "Let + be any
critical region ... [W]e may introduce a new symbol ... = (+ |ș} ... [where] + is kept constant
and ș varied." See the entry POWER on the mathematical words page here.

¢  # Please see the entry on the mathematical words page here.

  .:C /# Please see the entry on the mathematical words page here.

. */ # Please see the entry Student's -Distribution on the math words page here.

$          . In 1921, when R. A. Fisher introduced the concept of degrees


of freedom, he used for the number of degrees of freedom, following Pearson's (1900) chi-
square goodness of fit paper. Pearson derived the distribution of a quadratic form in jointly
normal variables where this normal is the limiting form of a multinomial with ' = +1 cells.
When Fisher used for the number of degrees of freedom in the -distribution, he could not use
it for the number of observations and so he used for that number; see e.g. chapter V of his
'  

þ 
  (1925, with 13 further editions to 1970.) C. P.
Eisenhart (1979, p. 8n) relates in his "On the Transition from Student's ( to Student's ,"
c  '  , 55, 6-10 how there was an aversion among many "not Fisherian"
statisticians to Fisher¶s use of and how E. S. Pearson and some colleagues decided on the
Greek letter Ȟ. This letter appears in M. G. Kendall¶s c 

'  (1943)
and in Halperin, Hartley & Hoel's "Recommended Standards for Statistical Symbols and
Notation. ..., (c  '  , 69, (1965), p. 12). For references and further details see the
entries on DEGREES OF FREEDOM, CHI SQUARE and STUDENT'S t DISTRIBUTION on the math
words page.
2  was introduced by Harold Hotelling in "The Generalization of Student's Ratio," c 

   ' ,  , (1931), 360-378.

0 has played several roles. Today it most often stands for the standard normal; see ' 


 +2
  
above. R. A. Fisher used ( in the analysis of variance
(see the entry z AND z DISTRIBUTION here) and in transforming the correlation coefficient (see the
entry FISHER¶S z TRANSFORMATION OF THE CORRELATION COEFFICIENT here.) Student had used
originally z for the test statistic that was turned into "Student¶s ". (See the entry STUDENT¶S -
DISTRIBUTION here.)

        

„rivativ. The symbols dx, dy, and dx/dy were introduced by Gottfried ilhelm Leibniz (1646-1716) in
a manuscript of November 11˜ 1675 (Cajori vol. 2˜ page 204).

 ,- for the first derivative,  ,- for the second derivative, etc., were introduced by Joseph Louis
Lagrange (1736-1813). In 1797 in %


  & the symbols   and   are
found; in the 3 0
/, "which purports to be a reprint of the 1806 edition, on p. 15, 17,
one finds the corresponding parts given as ,- ,- ,- ,-" (Cajori vol. 2, page 207).

In 1770 Joseph Louis Lagrange (1736-1813) wrote for in his memoir 2


%


  %
 %& 
%  
 %  (Oeuvres, Vol. III, pp. 5-76). The
notation also occurs in a memoir by François Daviet de Foncenex in 1759 believed actually to
have been written by Lagrange (Cajori 1919, page 256).

In 1772 Lagrange wrote  = / and  =   in "Sur une nouvelle espèce de calcul relatif à
la différentiation et à l'integration des quantités variables," 2
 
  c 

' $! $  (Oeuvres, Vol. III, pp. 451-478).

 was introduced by Louis François Antoine Arbogast (1759-1803) in "De Calcul des
dérivations et ses usages dans la théorie des suites et dans le calcul différentiel," Strasbourg, xxii,
pp. 404, Impr. de Levrault, fréres, an VIII (1800). (This information comes from Julio González
Cabillón; Cajori indicates in his @

   that Arbogast introduced this symbol,
but it seems he does not show this symbol in c@

   2
 
)

was used by Arbogast in the same work, although this symbol had previously been used by
Johann Bernoulli (Cajori vol. 2, page 209). Bernoulli used the symbol in a non-operational sense
(Maor, page 97).

   # The "curly d" was used in 1770 by Antoine-Nicolas Caritat, Marquis de
Condorcet (1743-1794) in "Memoire sur les Equations aux différence partielles," which was
published in @
 ! c þ
'  pp. 151-178, Annee M. DCCLXXIII
(1773). On page 152, Condorcet says:

Dans toute la suite de ce Memoire˜ dz & z désigneront ou deux differences partielles de z,˜ dont une
par rapport a x, l'autre par rapport a y, ou bien dz sera une différentielle totale˜ & z une difference
partielle. [Throughout this paper˜ both dz & z will either denote two partial differences of z, where one
of them is with respect to x, and the other˜ with respect to y, or dz and z will be employed as symbols
of total differential˜ and of partial difference˜ respectively.]

However˜ the "curly d" was first used in the form by Adrien Marie Legendre in 1786 in his "Memoire
sur la manière de distinguer les maxima des minima dans le Calcul des Variations˜" Histoire de
l'Academie Royale des Sciences, Annee M. DCCLXXXVI (1786)˜ pp. 7-37˜ Paris˜ M. DCCXXXVIII (1788). On
page 8˜ it readsü

Pour éviter toute ambiguité˜ je répresentarie par le coefficient de x dans la différence de u, & par

la différence complète de u divisée par dx.

Legendre abandoned the symbol and it was re-introduced by Carl Gustav Jacob Jacobi in 1841. Jacobi
used it extensively in his remarkable paper "De determinantibus Functionalibus" Crelle's Journal˜ Band
22˜ pp. 319-352˜ 1841 ( pp. 393-438 of vol. 1 of the Collected Works).

Sed quia uncorum accumulatio et legenti et scribenti molestior fieri solet˜ praetuli characteristica

differentialia vulgaria, differentialia autem partialia characteristica

denotare.

The "curly d" symbol is sometimes called the "rounded d" or "curved d" or 0acobi's delta. It corresponds
to the cursive "dey" (equivalent to our d) in the Cyrillic alphabet.

  # Before introducing the integral symbol, Leibniz wrote omn. for "omnia" in front of the
term to be integrated.

The integral symbol was first used by Gottfried Wilhelm Leibniz (1646-1716) on October 29,
1675, in an unpublished manuscript. Several weeks later, on Nov. 21, he first placed  after the
integral symbol (Burton, page 359). Later in 1675, he proposed the use of the symbol in a letter
to Henry Oldenburg, secretary of the Royal Society: "Utile erit scribi pro omnia, ut  = omn. ,
id est summa ipsorum " [It will be useful to write for omn. so that  = omn. , or the sum of all
the 's.] The first appearance of the integral symbol in print was in a paper by Leibniz in the c 
# 
 The integral symbol was actually a long letter S for "summa."

In his B      of 1704, Newton wrote a small vertical bar above x to indicate the
integral of x. He wrote two side-by-side vertical bars over x to indicate the integral of (x with a
single bar over it). Another notation he used was to enclose the term in a rectangle to indicate its
integral. Cajori writes that Newton's symbolism for integration was defective because the x with
a bar could be misinterpreted as x-prime and the placement of a rectangle about the term was
difficult for the printer, and that therefore Newton's symbolism was never popular, even in
England.

      # Limits of integration were first indicated only in words. Euler was the
first to use a symbol in  
    where he wrote the limits in brackets and
used the Latin words  and  (Cajori vol. 2, page 249).

The modern definite integral symbol was originated by Jean Baptiste Joseph Fourier (1768-
1830). In 1822 in his famous c   

@  he wrote:

Nous désignons en général par le signe l'intégrale qui commence lorsque la variable équivaut à a, et
qui est complète lorsque la variable équivaut à b. . .

The citation above is from "Théorie analytique de la chaleur" [The Analytical Theory of Heat]˜ Firmin
Didot˜ Paris˜ 1822˜ page 252 (paragraph 231).

Fourier had used this notation somewhat earlier in the %


  of the French Academy for
1819-20, in an article of which the early part of his book of 1822 is a reprint (Cajori vol. 2 page
250).

The   to indicate evaluation of an antiderivative at the two limits of integration was
first used by Pierre Frederic Sarrus (1798-1861) in 1823 in Gergonne's c  Vol. XIV. The
notation was used later by Moigno and Cauchy (Cajori vol. 2, page 250).

        2# Dan Ruttle, a reader of this page, has found a use of the
integral symbol with a circle in the middle by Arnold Sommerfeld (1868-1951) in 1917 in
c   =  "Die Drudesche Dispersionstheorie vom Standpunkte des Bohrschen
Modelles und die Konstitution von H2, O2 und N2." This use is earlier than the 1923 use shown
by Cajori. Ruttle reports that J. W. Gibbs used only the standard integral sign in his # 

0
c   (1881-1884), and that and E. B. Wilson used a small circle below the standard
integral symbol to denote integration around a closed curve in his 0
c   (1901, 1909)
and in c   (1911, 1912).

 # lim. (with a period) was used first by Simon-Antoine-Jean L'Huilier (1750-1840). In
1786, L'Huilier gained much popularity by winning the prize offered by *l'Academie royale des
Sciences et Belles-Lettres de Berlin*. His essay, "Exposition élémentaire des principes des
calculs superieurs," accepted the challenge thrown by the Academy -- a clear and precise theory
on the nature of infinity. On page 31 of this remarkable paper, L'Huilier states:

Pour abreger & pour faciliter le calcul par une notation plus commode˜ on est convenu de désigner
autrement que par

la limite du rapport des changements simultanes de = & de  favoir par

en sorte que

ou

designent la même chose

lim (without a period) was written in 1841 by Karl eierstrass (1815-1897) in one of his papers
published in 1894 in Mathematische Werke, Band I˜ page 60 (Cajori vol. 2˜ page 255).

The arrow notation for limits was introduced by Godfrey Harold Hardy (1877-1947) in his
remarkable "A Course of Pure Mathematics," Cambridge: At the University Press, xv, pp. 428,
1908. Check the preface of this first edition (Julio González Cabillón and Cajori vol. 2, page
257).

  # The infinity symbol was introduced by John Wallis (1616-1703) in 1655 in his 


 (On Conic Sections) as follows:

Suppono in limine (juxta^ Bonaventurae Cavallerii eometriam Indivisibilium) Planum quodlibet quasi ex
infinitis lineis parallelis conflariü Vel potiu\s (quod ego mallem) ex infinitis Prallelogrammis [sic] aeque\
altis; quorum quidem singulorum altitudo sit totius altitudinis 1/ ˜ sive alicuota pars infinite parva;
(esto enim nota numeri infiniti;) adeo/q; omnium simul altitude aequalis altitudini figurae.

allis also used the infinity symbol in various passages of his Arithmetica infinitorum (Arithmetic of
Infinites) (1655 or 1656). For instance˜ he wrote (p. 70)ü
Cum enim primus terminus in serie Primanorum sit 0˜ primus terminus in serie reciproca erit vel
infinitusü (sicut˜ in divisione˜ si diviso sit 0˜ quotiens erit infinitus)

In ƒero to Lazy Eight, Alexander Humez˜ Nicholas Humez˜ and Joseph Maguire writeü "allis was a
classical scholar and it is possible that he derived from the old Roman sign for 1˜000˜ CD˜ also written
M--though it is also possible that he got the idea from the lowercase omega˜ omega being the last letter
of the Greek alphabet and thus a metaphor of long standing for the upper limit˜ the end."

Cajori (vol. 2, p 44) says the conjecture has been made that Wallis adopted this symbol from the
late Roman symbol for 1,000. He attributes the conjecture to Wilhelm Wattenbach (1819-
1897), c  (    = G
  2. Aufl., Leipzig: S. Hirzel, 1872. Appendix: p.
41.

This conjecture is lent credence by the labels inscribed on a Roman hand abacus stored at the
Bibliothèque Nationale in Paris. A plaster cast of this abacus is shown in a photo on page 305 of
the English translation of Karl Menninger's 2 
 2 ' 
; at the time, the
cast was held in the Cabinet des Médailles in Paris. The photo reveals that the column devoted to
1000 on this abacus is inscribed with a symbol quite close in shape to the lemniscate symbol, and
which Menninger shows would easily have evolved into the symbol M, the eventual Roman
symbol for 1000 [Randy K. Schwartz].

[Julio González Cabillón contributed to this entry.]

       C # In 1706, Johann Bernoulli used į to denote the difference of
functions. Julio González Cabillón believes this is probably one of the first if not the first use of
delta in this sense.

   2 # Augustin-Louis Cauchy (1789-1857) used İ in 1821 in 


   
(Oeuvres II.3), and sometimes used į instead (Cajori vol. 2, page 256). According to Finney and
Thomas (page 113), "į meant 'différence' (French for    and İ meant 'erreur' (French for

)."

The first theorem on limits that Cauchy sets out to prove in the 
  c   (Oeuvres II.3, p.
54) has as hypothesis that

for increasing values of x, the difference f(x + 1) - f(x) converges to a certain limit k.

The proof then begins by saying

denote by ɸ a number as small as one may wish. Since the increasing values of x make the difference f(x
+ 1) - f(x) converge to the limit k, one can assign a sufficiently substantial value to a number h so that˜ for
x bigger than or equal to h˜ the difference in question is always between the bounds k - ɸ˜ k + ɸ.

[illiam C. aterhouse]
The first delta-epsilon proof is Cauchy's proof of what is essentially the mean-value theorem for
derivatives. It comes from his lectures on the Calcul infinitesimal, 1823, Leçon 7, in Oeuvres,
Ser. 2, vol. 4, pp. 44-45. The proof translates Cauchy's verbal definition of the derivative as the
limit (when it exists) of the quotient of the differences into the language of algebraic inequalities
using both delta and epsilon. In the 1820s Cauchy did not specify on what, given an epsilon, his
delta or n depended, so one can read his proofs as holding for all values of the variable. Thus he
does not make the distinction between converging to a limit pointwise and convering to it
uniformly.

[Judith V. Grabiner, author of 3  


  þ

  (MIT, 1981)]

h 
  áá á

           2 %      #

The vector differential operator, now written and called  or , was introduced by
William Rowan Hamilton (1805-1865). Hamilton wrote the operator as and it was P. G. Tait
who established as the conventional symbol--see his c #    
B  

(1867). Tait was also responsible for establishing the term  . See NABLA on the Earliest
Uses of Words page.

David Wilkins suggests that Hamilton may have used the nabla as a general purpose symbol or
abbreviation for whatever operator he wanted to introduce at any time. In 1837 Hamilton used
the nabla, in its modern orientation, as a symbol for any arbitrary function in  þ 
c  XVII. 236. (OED.) He used the nabla to signify a permutation operator in "On the
Argument of Abel, respecting the Impossibility of expressing a Root of any General Equation
above the Fourth Degree, by any finite Combination of Radicals and Rational Functions,"
  

þ
 c   18 (1839), pp. 171-259.

Hamilton used the rotated nabla, i.e. , for the vector differential operator in the "Proceedings
of the Royal Irish Academy" for the meeting of July 20, 1846. This paper appeared in volume 3
(1847), pp. 273-292. Hamilton also used the rotated nabla as the vector differential operator in
"On Quaternions; or on a new System of Imaginaries in Algebra"; which he published in
instalments in the =

   (  between 1844 and 1850. The relevant portion of the
paper consists of articles 49-50, in the instalment which appeared in October 1847 in volume 31
(3rd series, 1847) of the Philosophical Magazine, pp. 278-283. A footnote in vol. 31, page 291,
reads:

In that paper designed for Southampton the characteristic was written ; but this more common sign
has been so often used with other meanings˜ that it seems desirable to abstain from appropriating it to
the new signification here proposed.

ilkins writes that "that paper" refers to an unpublished paper that Hamilton had prepared for a
meeting of the British Association for the Advancement of Science˜ but which had been forwarded by
mistake to Sir John Herschel's home address˜ not to the meeting itself in Southampton˜ and which
therefore was not communicated at that meeting. The footnote indicates that Hamilton had originally
intended to use the nabla symbol that is used today but then decided to rotate it to avoid confusion
with other uses of the symbol.

The rotated form appears in Hamilton's magnum opus, the ! 
B  
 (1853, p.
610).

Cajori (vol. 2, page 135) and the OED give this reference.

According to Stein and Barcellos (page 836), Hamilton denoted the gradient with an ordinary
capital delta in 1846. However, this information may be incorrect, as David Wilkins writes that
he has never seen the gradient denoted by an ordinary capital delta in any paper of Hamilton
published in his lifetime.

David Wilkins of the School of Mathematics at Trinity College in Dublin has made available
texts of the mathematical papers published by Hamilton in his lifetime at his History of
Mathematics website.

grad as a symbol for gradient appears in H. Weber's         
     þ 0
  of 1900 (Cajori vol. 2, page 135). See
GRADIENT on the Earliest Uses of Words page.

William Kingdon Clifford (1845-1879) used  or  as symbols for divergence (Cajori vol.
2, page 135).

The symbol ǻ for the Laplacian operator (also represented by 2) was introduced by Robert
Murphy in 1833 in #  =  


#   (Kline, page 786). See
LAPLACE'S OPERATOR on the Earliest Uses of Words page.

Front Page $ Operation $ Grouping $ Relation $ Fractions $ Constants $ Variables $ Functions $ Geometry $
Trigonometry $ Calculus $ Set Theory and Logic $ Number theory $ Statistics $ Sources

        $    


!  " :DD>

      # The congruent symbol used in number theory was introduced in
print in 1801 by Carl Friedrich Gauss (1777-1855) in &
  :
Numerorum congruentiam hoc signo, Ł, in posterum denotabimus, modulum ubi opus erit in
clausulis adiungentes, -16 Ł 9 (mod. 5), -7 Ł 15 (mod. 11).
The citation above is from &
   (Leipzig, 1801), art. 2;   Vol. I
(Gottingen, 1863), p. 10 (Cajori vol. 2, page 35).

However, Gauss had used the symbol much earlier in his personal writings (Francis, page 82).

See MODULUS on Words page.

     2     O Edmund Landau used () for the number of primes less
than or equal to  in 1909 in @  ! 
 0   = (  (Cajori vol.
2, page 36).

                # The authors of classical textbooks such as


Weber and Fricke did not denote particular domains of computation with letters.

Richard Dedekind (1831-1916) denoted the rationals by R and the reals by gothic R in '
  
*  (1872) (
    
   
 5, 315-334. Dedekind
also used K for the integers and J for complex numbers.

Giuseppe Peano (1858-1932) 2þ and B in in Arithmetices prinicipia nova methodo exposita,
1889; ed. in: Peano, opere scelte, 2, Rom 1958; p. 23. This page (p. 23) shows the table of all
symbols in his famous work, which includes:

2 numerus integer positivus


þ num. rationalis positivus
B quantitas, sive numerus realis positivus

[This information was provided by Wilfried Neumaier.]

In 1895 in his 
   % & Peano used 2 for the positive integers, for
integers, 2D for the positive integers and zero, þ for positive rational numbers, for rational
numbers, B for positive real numbers, & for real numbers, and BD for positive real numbers and
zero [Cajori vol. 2, page 299].

Helmut Hasse (1898-1979) used ī for the integers and ȇ (capital rho) for the rationals in @1 
c I and II, Berlin 1926. He kept to this notation in his later books on number theory.
Hasse's choice of gamma and rho may have been determined by the initial letters of the German
terms "ganze Zahl" (integer) and "rationale Zahl" (rational).

Otto Haupt used 0 for the integers and ȇ0 (capital rho) for the rationals in # 7   
c   Leipzig 1929.

Bartel Leendert van der Waerden (1903-1996) used C for the integers and ī for the rationals in

 c  Berlin 1930, but in editions during the sixties, he changed to Z and Q.
Edmund Landau (1877-1938) denoted the set of integers by a fraktur Z with a bar over it in
     c   (1930, p. 64). He does not seem to introduce symbols for the sets of
rationals, reals, or complex numbers.

B for the set of rational numbers and * for the set of integers are apparently due to N. Bourbaki.
(N. Bourbaki was a group of mostly French mathematicians which began meeting in the 1930s,
aiming to write a thorough unified account of all mathematics.) The letters stand for the German
B
  and *   These notations occur in Bourbaki's c%  Chapter 1.

Julio González Cabillón writes that he believes Bourbaki was responsible for both of the above
symbols, quoting Weil, who wrote, "...it was high time to fix these notations once and for all, and
indeed the ones we proposed, which introduced a number of modifications to the notations
previously in use, met with general approval."

[Walter Felscher, Stacy Langton, Peter Flor, and A. J. Franco de Oliveira contributed to this
entry.]

       2 3   # William C. Waterhouse wrote to a history of mathematics


mailing list in 2001:

Checking things I have available, I found C used for the complex numbers in an early paper by
Nathan Jacobson:
Structure and Automorphisms of Semi-Simple Lie Groups in the Large, c 
  40
(1939), 755-763.
The second edition of Birkhoff and MacLane, '  

 c (1953), also uses C
(but is not using the Bourbaki system: it has J for integers, R for rationals, R^# for reals). I have't
seen the first edition (1941), but I would expect to find C used there too. I'm sure I remember C
used in this sense in a number of other American books published around 1950.

I think the first Bourbaki volume published was the results summary on set theory, in 1939, and
it does not contain any symbol for the complex numbers. Of course Bourbaki had probably
chosen the symbols by that time, but I think in fact the first appearance of (bold-face) C in
Bourbaki was in the formal introduction of complex numbers in Chapter 8 of the topology book
(first published in 1947).

 * 2   .    /# The symbol ij() for the number of integers less than 
that are relatively prime to  was introduced by Carl Friedrich Gauss (1777-1855) in 1801 in his
&
   articles 38, 39 (p. 30) (Cajori vol. 2, page 35, and Dickson, page
113-115).

The function was first studied by Leonhard Euler (1707-1783), although Dickson (page 113) and
Cajori (vol. 2, p. 35) say that Euler did not use a functional notation in 2

c

8, 1760-1, 74, and 
c  1, 274, and that Euler used ʌ2 in c c=
 4 II (or 8),
1780 (1755), 18, and 
c  2, 127-133. Shapiro agrees, writing: "He did not employ any
symbol for the function until 1780, when he used the notation ʌ "
Sylvester, who introduced the name 
  for the function, seems to have believed that Euler
had used ij. He writes in 1888 ( vol. IV p. 589 of his 
   =  ) "I am in
the habit of representing the totient of by the symbol IJ , IJ (taken from the initial of the word it
denotes) being a less hackneyed letter than Euler's ij, which has no claim to preference over any
other letter of the Greek alphabet, but rather the reverse." This information was taken from a post
in sci.math by Robert Israel.

See TOTIENT on Words page.

     .C  2/# Adrien-Marie Legendre introduced the notation

= 1 if is a quadratic residue of  and = -1 if is a quadratic non-residue of 

According to Hardy & Wright's c  






2  "Legendre introduced
'Legendre's symbol' in his #   

  first published in 1798. See, for
example, §135 of the second edition (1808)." In the third edition on Gallica this is on p.197.

However, according to William J. Leveque in     


2 
 "Legendre
introduced his symbol in an article in 1785, and at the same time stated the reciprocity law
without using the symbol."

[Both of these citations were provided by Paul Pollack.]

    # Mersenne numbers are marked  by Allan Cunningham in 1911 in


   B
 '


# 
 (Cajori vol. 2, page 41).

     # Fermat numbers are marked  in 1919 in L. E. Dickson's @






2  (Cajori vol. 2, page 42).

   Ô D ÊO Dirichlet used 2( +) for the norm 2+2 of the complex number + in
  .
  Vol. XXIV (1842) (Cajori vol. 2, page 33). See NORM on Words page

&  # Eliakim Hastings Moore used the symbol [& ] to represent the Galois field of
order & in 1893. The modern notation is "Galois-field of order & " (Julio González Cabillón and
Cajori vol. 2, page 41).

     ÷O Euler introduced the symbol in a paper published in 1750 (DSB,


article: "Euler").

In 1888, James Joseph Sylvester continued the use of Euler's notation (Shapiro).

Allan Cunningham used ı(2) to represent the sum of the proper divisors of 2 in =
 

!

   '
 35 (1902-03):
The Repetition of the Sum-Factor Operation. Abstract of an informal communication made by
Lieut.-Col. A. Cuningham, June 12th, 1902.

Let ı(2) denote the sum of the sub-factors of 2 (including 1, but excluding 2). It was found that,
with most numbers, ı 2 = 1, when the operation (ı) is repeated often enough. There is a small
class for which ı 2 = = (a   number), and then repeats; another small class for which ı 2 =
c, ı + 12 = $, where c$ are   numbers, and then repeats (c$ alternately); another
small class for which (even when 2 is  , < 1000) ı 2 increases beyond the practical power
of calculation.

[Cajori, vol. 2, page 29, and Paul Pollack]

In 1927 Landau chose the notation '( ) (Shapiro).

L. E. Dickson used ( ) for the sum of the divisors of (Cajori vol. 2, page 29).

 E   # Möbius' work appeared in 1832 but the µ symbol was not used.

The notation µ( ) was introduced by Franz Mertens (1840-1927) in 1874 in "Über einige
asymptotische Gesetze der Zahlentheorie,"   .
  (Shapiro).

r: and  :  # According to Wladyslaw Narkiewicz in  


 
= 
2 
:

The symbols O(ð) and o(ð) are usually called the Landau symbols. This name is only partially
correct, since it seems that the first of them appeared first in the second volume of P. Bachmann's
treatise on number theory (Bachmann, 1894). In any case Landau (1909a, p. 883) states that he
had seen it for the first time in Bachmann's book. The symbol o(ð) appears first in Landau
(1909a). Earlier this relation has been usually denoted by {ð}.
The references are to Paul Bachmann (1837-1920) and his c  *  
 and to
Edmund Landau (1877-1938) and his @  ! 
 0   = (  

[Paul Pollack contributed to this entry.]

        2 


  á

á á
=lus (+) and minus (-). Nicole d' Oresme (1323-1382) may have used a figure which looks like a plus
symbol as an abbreviation for the Latin et (meaning "and") in Algorismus proportionum, believed to
have been written between 1356 and 1361. The symbol appears in a manuscript of this work believed to
have been written in the fourteenth century˜ but perhaps by a copyist and not Oresme himself. The
symbol appears˜ for example˜ in the sentenceü "Primi numeri sesquiterti sunt .4. et .3.˜ et primi numeri
sev termini sesquialtere sunt .3. et .2." [Dic Sonneveld].

The plus symbol as an abbreviation for the Latin , though appearing with the downward stroke
not quite vertical, was found in a manuscript dated 1417 (Cajori).

The + and - symbols first appeared in print in   c  or $  7
þ     )   , by Johannes Widmann (born c. 1460), published in
Leipzig in 1489. However, they referred not to addition or subtraction or to positive or negative
numbers, but to surpluses and deficits in business problems (Cajori vol. 1, page 128).

           2    D   :  % from Widman's $  


7þ   This image is taken from the Augsburg edition of 1526.

Widman wrote, "Was - ist / das ist minus ... vnd das + das ist mer." He also wrote, "4 centner + 5
pfund" and "5 centner ² 17 pfund," thus showing the excess or deficiency in the weight of
boxes or bales (Smith vol. 2, page 399).

Smith (vol. 2, page 398) explains the origin of the + sign by connecting it to the Latin word for
"and":

In a manuscript of 1456˜ written in Germany˜ the word et is used for addition and is generally written so
that it closely resembles the symbol +. The et is also found in many other manuscripts˜ as in "5 et 7" for
5 + 7˜ written in the same contracted form˜ as when we write the ligature & rapidly. There seems˜
therefore˜ little doubt that this sign is merely a ligature for et.

Cajori says, "There is clear evidence that, as a lecturer at the University of Leipzig, Widmann
had studied manuscripts in the Dresden library in which + and - signify operations, some of these
having been written as early as 1486." Johnson (page 144) says a series of notes from 1481,
annotated by Widmann, contain the + and - symbols, and he asks whether Widman could have
copied these symbols from some unknown professor at the University of Leipzig. Johnson also
says that a student's notes from one of Widmann's 1486 lectures show the + and - signs.

Giel Vander Hoecke used + and - as symbols of operation in # 


  
  

c   published at Antwerp in 1514 (Smith 1958, page 341). Burton (page 335)
says Vander Hoecke was the first person to use + and - in writing algebraic expressions, but
Smith (page 341) says he followed Grammateus.

Henricus Grammateus (also known as Henricus Scriptor and Heinrich Schreyber or Schreiber)
published an arithmetic and algebra, entitled c  +) $ printed in 1518, in which
he used + and - in a technical sense for addition and subtraction (Cajori vol. 1, page 131).

The plus and minus symbols only came into general use in England after they were used by
Robert Recorde in in 1557 in 

 Recorde wrote, "There be other 2 signes in
often use of which the first is made thus + and betokeneth more: the other is thus made - and
betokeneth lesse."
The plus and minus symbols were in use before they appeared in print. For example, they were
painted on barrels to indicate whether or not the barrels were full. Some have attempted to trace
the minus symbol as far back as Heron and Diophantus.



á á
' was used by illiam Oughtred (1574-1660) in the Clavis Mathematicae (Key to Mathematics)˜
composed about 1628 and published in London in 1631 (Smith). Cajori calls ' St. Andrew's Cross.

' actually appears earlier, in 1618 in an anonymous appendix to Edward Wright's translation of
John Napier's  
(Cajori vol. 1, page 197). However, this appendix is believed to have
been written by Oughtred.

  .F/ was advocated by Gottfried Wilhelm Leibniz (1646-1716). According to Cajori (vol.
1, page 267):

The dot was introduced as a symbol for multiplication by G. . Leibniz. On July 29˜ 1698˜ he wrote in a
letter to John Bernoulliü "I do not like ' as a symbol for multiplication˜ as it is easily confounded with x;
... often I simply relate two quantities by an interposed dot and indicate multiplication by ƒC · LM.
Hence˜ in designating ratio I use not one point but two points˜ which I use at the same time for division."

Cajori shows the symbol as a raised dot. However˜ according to Margherita Barile˜ consulting Gerhardt's
edition of Leibniz's Mathematische Schriften (G. Olms˜ 1971)˜ the dot is never raised˜ but is located at
the bottom of the line. She writes that the non-raised dot as a symbol for multiplication appears in all
the letters of 1698˜ and earlier˜ and˜ according to the same edition˜ it already appears in a letter by
Johann Bernoulli to Leibniz dated September˜ 2nd 1694 (see vol. III˜ part 1˜ page 148).

The dot was used earlier by Thomas Harriot (1560-1621) in c   =  c& 

c  þ
   which was published posthumously in 1631, and by Thomas Gibson in
1655 in '      However Cajori says, "it is doubtful whether Harriot or Gibson
meant these dots for multiplication. They are introduced without explanation. It is much more
probable that these dots, which were placed after numerical coefficients, are survivals of the dots
habitually used in old manuscripts and in early printed books to separate or mark off numbers
appearing in the running text" (Cajori vol. 1, page 268).

However, Scott (page 128) writes that Harriot was "in the habit of using the dot to denote
multiplication." And Eves (page 231) writes, "Although Harriot on occasion used the dot for
multiplication, this symbol was not prominently used until Leibniz adopted it."

  4 .G/ was used by Johann Rahn (1622-1676) in 1659 in c (Cajori vol.
1, page 211).
r ;32 # In a manuscript found buried in the earth near the village of Bakhshali, India,
and dating to the eighth, ninth, or tenth century, multiplication is normally indicated by placing
numbers side-by-side (Cajori vol. 1, page 78).

Multiplication by juxtaposition is also indicated in "some fifteenth-century manuscripts" (Cajori


vol. 1, page 250). Juxtaposition was used by al-Qalasadi in the fifteenth century (Cajori vol. 1,
page 230).

According to Lucas, Michael Stifel (1487 or 1486 - 1567) first showed multiplication by
juxtaposition in 1544 in c    

In 1553, Michael Stifel brought out a revised edition of Rudolff's 


 in which he showed
multiplication by juxtaposition and repeating a letter to designate powers (Cajori vol. 1, pages
145-147).

háá á
Clos parnt sis. The arrangement 8)24 was used by Michael Stifel (1487-1567 or 1486-1567) in
Arithmetica integra, which was completed in 1540 and published in 1544 in Nuernberg (Cajori vol. 1˜
page 269; DSB).

  .H/ was used in 1633 in a text entitled .


 
c 4 +
$

 (2nd ed.:


London, 1633). However Johnson only used the symbol to indicate fractions (for example three-
fourths was written 3:4); he did not use the symbol for division "dissociated from the idea of a
fraction" (Cajori vol. 1, page 276).

Gottfried Wilhelm Leibniz (1646-1716) used : for both ratio and division in 1684 in the c 
 
 (Cajori vol. 1, page 295).

   .I/ was first used as a division symbol by Johann Rahn (or Rhonius) (1622-1676) in
1659 in c (Cajori vol. 2, page 211).

Here is the page in which the division symbol first appears in print, as reproduced in Cajori.

Rahn's book was translated into English and published, with additions by John Pell, in London in
1668, with the division symbol retained. According to some recent sources, John Pell was a
major influence on Rahn and he may in fact be responsible for the invention of the symbol.
However, according to Cajori there is no evidence to support this claim. The division symbol
was used by many writers before Rahn as a minus sign.

!     # In nineteenth century U. S. textbooks, long division is typically shown with


the divisor, dividend, and quotient on the same line, separated by parentheses, as 36)116(3. See
the figure below.
Division shown in Arithmetick by John Hill (1772)

This same 1772 book shows a vinculum above the dividend when dividing decimals:

Decimal division in Arithmetick by John Hill (1772)

In 1882 in 
 c  by James B. Thomson, the 36)116(3 notation is used for
long division. However, in examples for short division, a vinculum is placed under the dividend
and the vinculum is almost attached to the bottom of the close parenthesis. The quotient is
written under the vinculum, as shown below.
From Complete raded Arithmetic, 1882

In 1888 in the teacher's edition of # 


c by G. A. Wentworth the vinculum is
almost attached to the top of the close parenthesis and the quotient is written above the vinculum,
as shown below.

From Elements of Algebra, 1888

The symbolism shown above may also appear in the earlier 1882 edition of Wentworth, which
has not been seen. I welcome earlier uses of this symbol that may be found by readers of this
page.

In 1901, the second edition of þ


 

c  by Daniel W. Fish uses the same
notations for short and long division as Thomson (1882) above, except that the vinculum under
the dividend is actually attached to the close parenthesis. This notation may appar in the earlier
1873 edition, which has not been seen.

David E. Smith writes, "It is impossible to fix an exact date for the origin of our present
arrangement of figures in long division, partly because it developed gradually" (Smith vol. 2).

The symbol is not mentioned by Cajori. The symbol does not have a name.

 
á
=ositiv intrs as ponnts. Nicole Oresme (c. 1323-1382) used numbers to indicate powering in the
fourteenth century˜ although he did not use raised numbers.

Nicolas Chuquet (1445?-1500?) used raised numbers in !     ' 2


 
in 1484. However, in Chuquet's notation, 123 actually meant 123 (Cajori vol. 1, page 102).
In 1634, Pierre Hérigone (or Herigonus) (1580-1643) wrote a, a2, a3, etc., in  
   which was published in several volumes from 1634 to 1637; the numerals were
not raised, however (Cajori vol. 1, page 202, and Ball).

In 1636 James Hume used Roman numerals as exponents in ! cH 0H  


    Cajori writes (vol. 1, pages 345-346):

In 1636 James Hume brought out an edition of the algebra of Vieta˜ in which he introduced a superior
notation˜ writing down the base and elevating the exponent to a position above the regular line and a
little to the right. The exponent was expressed in Roman numerals. Thus˜ he wrote Aiii for A3. Except for
the use of Roman numerals˜ one has here our modern notation. Thus˜ this Scotsman˜ residing in Paris˜
had almost hit upon the exponential symbolism which has become universal through the writings of
Descartes.

In 1637 exponents in the modern notation (although with positive integers only) were used by Rene
Descartes (1596-1650) in eometrie. Descartes tended not to use 2 as an exponent˜ however˜ usually
writing aa rather than a2˜ perhaps because aa occupies no less space than a2.

Descartes wrote: " ou 2 pour multiplier à par soiméme; et 3


pour le multiplier encore une fois
par  et ainsi à l'infini" (Cajori 1919, page 178).

$       32  were used by Nicolas Chuquet (1445?-1500?) in 1484 in !


     ' 2
  Chuquet wrote to indicate 12-1 (Cajori vol. 1, page
102).

Negative integers as exponents were first used with the modern notation by Isaac Newton in June
1676 in a letter to Henry Oldenburg, secretary of the Royal Society, in which he described his
discovery of the general binomial theorem twelve years earlier (Cajori 1919, page 178).

Before Newton, John Wallis suggested the use of negative exponents but did not actually use
them (Cajori vol. 1, page 216).

   32 # The first use of fractional exponents (although not with the modern

notation) is by Nicole Oresme (c. 1323-1382) in c





 Oresme used
to represent 91/3. According to Cajori (1919), this notation remained unnoticed.

Simon Stevin (1548-1620) "had no occasion to use the fractional index notation," but "he clearly
stated that 1/2 in a circle would mean square root and 3/2 in a circle would indicate the square
root of the cube" (Boyer, page 356).

John Wallis (1616-1703), in his c    


 which was published in 1656, speaks of
fractional "indices" but does not actually write them (Cajori vol. 1, page 354).
Fractional exponents in the modern notation were first used by Isaac Newton in the 1676 letter
referred to above (Cajori 1919, page 178).

    # The earliest use of scientific notation is not known. However, some
physicists working with electricity in in the decade or so up to 1873, when our modern volt, ohm,
etc., were standardized, used scientific notation. James A. Landau has found only two usages of
scientific notation in Maxwell's collected papers, and could find no other physicists of mid-
century using scientific notation.

In 1863, in c  #  


  þ 

 # 
; 8I: the
article on "Electricity" has on page 404:

The aim should be to make this standard [of electrical resistance] correspond to a current force equal to
10˜000˜000˜000 times the value given by the quotient of 1 metre by 1 second of time˜ that is˜ 1010
mètre/seconds.

In 1868 Rep. Brit. Assoc. 1867 hasü 105 EMF˜ acting on a circuit of 1013˜ will pass in one second 10-8
absolute units of quantity; and similarly˜ 105 EMF will charge a condenser of absolute capacity equal to
10-13 absolute units with 10-8 absolute units of quantity... Mr. Clark calls the unit of quantity thus defined
(10-8) one Farad˜ and similarly says that the unit of capacity has a capacity of one Farad˜ it being
understood that this is the capacity when charged with unit electromotive force (105).

The above quotation was taken from the OED2.

In 1885 Johann Jakob Balmer in "Notiz uber die Spectrallinien des Wasserstoffs" (c   
=   Vol. 25, p. 80, 1885) wrote in English translation:

From the formula we obtained for a fifth hydrogen line

49/45 · 3645.6 = 3969.65 · 10-7 mm

In 1887 Albert Abraham Michelson and Edward Williams Morley wrote in =

 
  (  Series 5, December, 1887:

Considering the motion of the earth in its orbit only˜ this displacement should be

2 D v2/V2 = 2D x 10-8

The distance D was about eleven metres˜ or 2 x 107 wave-lengths of yellow light.

Both of these citations were taken from A Source Book in Physics by illiam Francis Magie.

An even earlier possible use of scientific notation is by Robert Whillhelm Bunsen in 1857 in
=

   
 where these formulae appear on page 357:

I = I0 10-hɲ
I = I1 10-hĮ + I2 10-hĮ + ...

However the objection is that Bunsen was measuring the intensity of light before and after going
through a tube of chlorine˜ and the alpha above is defined as "The value of 1/alpha˜ which signifies...the
depth of chlorine to which the chemical rays must penetrate in order to be reduced to one-tenth of
their original amount..." Therefore the 10 is not necessarily part of scientific notation but comes from
the fact that Bunsen elected to measure a reduction of light to one-tenth of the original.

This entry was largely contributed by James A. Landau.

 
á
The convention that multiplication precedes addition and subtraction was in use in the earliest books
employing symbolic algebra in the 16th century. The convention that exponentiation precedes
multiplication was used in the earliest books in which exponents appeared.

In 1892 in   c  M. A. Bailey advises avoiding expressions containing both ÷ and
×.

In 1898 in $


c by G. E. Fisher and I. J. Schwatt, ÷× is interpreted as
( ÷)×

In 1907 in @'

c #  
  by Slaught and Lennes, it is recommended
that multiplications in any order be performed first, then divisions as they occur from left to
right.

In 1910 in  
 
c by Hawkes, Luby, and Touton, the authors write that ÷ and ×
should be taken in the order in which they occur.

In 1912,  ; c by Webster Wells and Walter W. Hart has: "Indicated operations are
to be performed in the following order: first, all multiplications and divisions in their order from
left to right; then all additions and subtractions from left to right."

In 1913, '

  c by Webster Wells and Walter W. Hart has: "3  


 
 In a sequence of the fundamental operations on numbers, it is agreed that operations
under radical signs or within symbols of grouping shall be performed before all others; that,
otherwise, all multiplications and divisions shall be performed first, proceeding from left to right,
and afterwards all additions and subtractions, proceeding again from left to right."

In 1917, "The Report of the Committee on the Teaching of Arithmetic in Public Schools,"
    ( 8, p. 238, recommended the use of brackets to avoid ambiguity in such
cases.
In c@

   2
 
 (1928-1929) Florian Cajori writes (vol. 1, page 274), "If
an arithmetical or algebraical term contains ÷ and ×, there is at present no agreement as to which
sign shall be used first."

Modern textbooks seem to agree that all multiplications and divisions should be performed in
order from left to right.


 á á

„ot for scalar product was used in 1902 in J. . Gibbs's hector Analysis by E. B. ilson. However the dot
was written at the baseline and was not a "raised dot."

'    2 was used in 1902 in J. W. Gibbs's 0


c   by E. B. Wilson.

::     .J/ was used by William Oughtred (1574-1660) in     
published in 1631 (Cajori vol. 1, page 245).

 3  or !$ began as prefix notation, a mathematical notation which did away with
grouping symbols. It was proposed by Jan Lukasiewicz (1878-1956). "Prefix" meant that the
operators ( * + - etc.) preceded the operands or variables they were meant to operate on. Then it
was discovered that it is much more convenient to place the operands first and operators last, so
"Postfix" or "reverse Lukasiewicz" or "reverse Polish" notation was created. With postfix
notation the operators themselves become delimiters between operations. Thus the sequence
U*(V^(W + 3))/(X - Y) becomes

UV3+^*XY-/

RPN is used in the computer language Forth. [Axel Harvey]

 2   ( ) was introduced by Rene Descartes, according to Gullberg.

Cajori says this symbol was introduced by Gauss in 1812 (vol. 2, page 78).

C # The first use of was in 1220 by Leonardo of Pisa in =  
   where
the symbol meant "square root" (Cajori vol. 1, page 90).

The radical symbol first appeared in 1525 in 


 by Christoff Rudolff (1499-1545). He
used (without the vinculum) for square roots. He did not use indices to indicate higher roots,
but instead modified the appearance of the radical symbol for higher roots.

It is often suggested that the origin of the modern radical symbol is that it is an altered letter 
the first letter in the word  This is the opinion of Leonhard Euler in his  
 
    (1775). However, Florian Cajori, author of c@

   2
 

argues against this theory.
In 1637 Rene Descartes used , adding the vinculum to the radical symbol ! 
 
(Cajori vol. 1, page 375).

Placement of the index within the opening of the radical sign was suggested in 1629 by Albert
Girard (1595-1632) in   

 He suggested this notation for the cube root (DSB;
Cajori vol. 1, page 371).

According to Cajori (vol. 1, page 372) the first person to adopt Girard's suggestion and place the
index within the opening of the radical sign was Michel Rolle (1652-1719) in 1690 in  % 
c% 

However, a history note in a high school textbook states that the symbol was first used by Girard
"around 1633" ( '=c c  2nd ed., 1996, page 496).

According to Tropfke,  #     4th edition, vol. 1: c  
c  Berlin/New York: De Gruyter, 1980, p. 296, Girard introduced this use with examples
for the cubic and the quintic root on fol. 1r° of the   

 and suggested on the same
page also (3/2)49 for the fractional exponent of 49 [3/2 written as fraction]. [This information
contributed by Siegmund Probst.]

Examples of Leibniz's use of the radical sign with index placed in the opening:

VII,3 N. 57_2 p. 736 (End of April 1676)


VII,3 N. 58 p. 752 (May 3, 1676)
VII,3 N. 60 p. 760 (June 26, 1676)

In the     ( of Feb. 1895, G. Heppel wrote, "Following Chrystal, Todhunter,
Hall and Knight, and the majority of writers [sqrt] should be considered a quantity having one
and not two values, although the algebra of C. Smith and the article by Professor Kelland in the
#  
 $   make [sqrt] have two values."

  # The summation symbol ( ) was first used by Leonhard Euler (1707-1783) in 1755:

luemadmodum ad differentiam denotandam vsi sumus signo [capital delta]˜ ita summam indicabimus
signo ( ).

The citation above is from Institutiones calculi differentialis (St. Petersburg˜ 1755)˜ Cap. I˜ para. 26˜ p. 27.

This symbol was used by Lagrange, but otherwise received little attention during the eighteenth
century (Cajori vol. 2, pages 61 and 265.)

c       # The tilde was introduced for this purpose by William Oughtred
(1574-1660) in the      (Key to Mathematics), composed about 1628 and
published in London in 1631, according to Smith, who shows a reversed tilde (Smith 1958, page
394).
 # In 1841, Arthur Cayley (1821-1895) used the modern notation for the determinant of a
matrix, a single vertical line on both sides of the entries. The notation appeared in the   
   .
  Vol. II (1841), p. 267-271. However, Cayley used commas to separate
entries within rows (Cajori vol. 2, page 92).

The double vertical line notation was introduced by Cayley in 1843 (Cajori vol. 2, page 95).

In 1846, the first occurrence of both the single vertical line notation for determinants and double
vertical lines for matrices is found in "Mémoire sur les hyperdéterminants" by Arthur Cayley in
  .
  (Cajori vol. 2, page 93).

Cajori (vol. 2, p. 103) writes that round parentheses were used for matrices by many, including
Maxine Bocher in 1919 in  


@ c and G. Kowalewski in 1909 in
   
 (although Kowalewski also used double vertical lines and a single brace).

Cajori also shows a use of brackets for matrices (and no commas within) by C. E. Cullis in
     
 in 1913.

c1  # In 1936 in ! c c  Oystein Ore wrote:

Nous dirons que deux systemes algebriques S et S' sont homomorphes (par rapport a l'addition et a la
multiplication) s'il existe une correspondance a Ͷ> a' entre les elements de S et S' donnant a chaque
element a de S une image unique a' dans S' telle que chaque element de S' soit l'image d'au moins un
element de S et en outre telle que de a Ͷ> a'˜ b --> b' on puisse conclure

a + b Ͷ> a' + b' ˜ ab Ͷ> a'b' .

As early as 1939 Bourbaki used the arrow in element-to-element notation ["la application x ²>
f(x)"].

Saunders Mac Lane wrote:

At first the vivid arrow notation f ü X Ͷ> Y for a map was not available˜ and homomorphisms of
homology groups (or rings) were always expressed in terms of the corresponding quotient group or
rings. Thus the familiar long exact sequence of the homotopy groups of a fibration was originally
described in terms of subgroups and quotient groups; this is the style used by all three discoveries of the
sequence and of the covering homotopy theorem [...] The occurrence of exact sequences of homology
groups (though not the name 'exact') was first noted by . Hurewicz in 1941 [...] The practice of using
an arrow to represent a map f ü X Ͷ> Y arose at the same time. I have not been able to determine who
first introduced this convenient notation; it may well have appeared first on the blackboard˜ perhaps in
lectures by Hurewicz and it is used in the Hurewicz-Steenrod paper˜ submitted November 1940 ...

The above quotation is from Saunders Mac Lane˜ "Concepts and Categories in Perspective˜" A Century of
Mathematics in America, Part I, AMS˜ vol 1˜ 1988 [Julio González Cabillón].
        ! 
C# In printed books before the modern equal sign, equality was usually expressed with a
word, such as &  &     6 or  and sometimes by the
abbreviated form & (Cajori vol. 1, page 297).

The equal symbol (=) was first used by Robert Recorde (c. 1510-1558) in 1557 in 


 He wrote, "I will sette as I doe often in woorke use, a paire of parralles, or Gemowe
lines of one lengthe, thus : ==, bicause noe 2, thynges, can be moare equalle." Recorde used an
elongated form of the present symbol. He proposed no other algebraic symbol (Cajori vol. 1,
page 164).

        2  2 ÷ àÊ  1  C  


  #

The equal symbol did not appear in print again until 1618, when it appeared in an anonymous
Appendix, very probably due to Oughtred, printed in Edward Wright's English translation of
Napier's  
 It reappeared 1631, when it was used by Thomas Harriot and William
Oughtred (Cajori vol. 1, page 298).

Cajori states (vol. 1, page 126):

A manuscript, kept in the Library of the University of Bologna, contains data regarding the sign
of equality (=). These data have been communicated to me by Professor E. Bortolotti and tend to
show that (=) as a sign of equality was developed at Bologna independently of Robert Recorde
and perhaps earlier.
Cajori elsewhere writes that the manuscript was probably written between 1550 and 1568.

         # The symbols < and > first appear in c c   =  
c& 
c  þ
   (The Analytical Arts Applied to Solving Algebraic
Equations) by Thomas Harriot (1560-1621), which was published posthumously in 1631:
"Signum majoritatis ut a > b significet a majorem quam b" and "Signum minoritatis ut a < b
significet a minorem quam b."

According to Johnson (page 144), while Harriot was surveying North America, he saw a native
American with this symbol on his arm: . Johnson says it is likely he developed the two
symbols from this symbol.

However, Seltman and Mizzy say that Harriot himself did not use the symbols which appear in
the work, which was published after his death:

We now know that Harriot was not directly responsible for the =  which was put together
after his death from papers which are no longer extant by Walter Warner (and, perhaps, one or
two others), when Nathaniel Torporley had failed to complete the task which had been assigned
to him in Harriot's will. Torporley was a respected mathematician of the day, reputed to have
been associated with Viéte himself. The manuscripts that we do have (in the British Library and
Petworth) cannot have been the origin of the =  not only on account of their disorder and
incoherence, but also because there are significant differences between them and the published
work. Notably, the inequality signs associated with his name are never found in his handwriting
in the manuscripts but appear throughout as and . Similarity, equality is denoted in the
manuscripts by II and not by = (the sign introduced by Robert Recorde), as in the =  The
significance of the inequality signs lies in the fact that this is the first time that such signs were
used and accorded the same status as the equality sign.
    C %      C # Pierre Bouguer (1698-1758) used and in
1734 (Ball). In 1670, John Wallis used similar symbols each with a single horizontal bar, but the
bar was above the < and > rather than below it (Cajori vol. 2, page 118). Cajori apparently does
not show a use of the modern symbols with the single horizontal bar.

$ C %      %     # These symbols were "employed, if not invented
by, Euler" (Ball, page 242). Ball shows the symbol rather than .

  C # The symbol was used in 1875 by Anton Steinhauser in !  
   "Algebra" (Cajori vol. 2, page 256). The same symbol was used in 1832 by
Wolfgang Bolyai to signify absolute equality (Cajori vol. 1, page 307).

2 # The symbol :: was introduced by William Oughtred (1574-1660) in  


   , composed about 1628 and published in London in 1631. He wrote a proportion as
:: (Gullberg).

The astronomer Vincent Wing (1619-1668) used colons to write a proportion in the modern
notation, as c:$:::  in 1651 in @ 

 (Cajori vol. 1, page 286).

The symbol for variation (an eight lying on its side with a piece removed) was introduced in
1768 by W. Emerson in
  

 (3d ed., London) (Cajori vol. 1, page 297).

             



The study of logic goes back more than two thousand years and in that time many symbols and
diagrams have been devised. Around 300 BC Aristotle introduced letters as term-variables, a
"new and epoch-making device in logical technique." (W. & M. Kneale  
 

!
 (1962, p. 61). The modern era of mathematical notation in logic began with George Boole
(1815-1864), although none of his notation survives. Set theory came into being in the late 19th
and early 20th centuries, largely a creation of Georg Cantor (1845-1918). See MacTutor's A
history of set theory or, for more detail, Set theory from the ' 
#  
 

=

 .
Most of the basic symbols of logic and set theory in use today were introduced between 1880
and 1920. The main contributors were Ernst Schröder (1841-1902), Giuseppe Peano (1858-
1932), Alfred North Whitehead (1861-1947) and Bertrand Russell (1872-1970). Peano had a
strong influence on Whitehead and Russell and their joint work, =      (1910-
1913), was itself very influential. Today Gottlob Frege (1848-1925) is the most admired logician
of that age but his notation was not taken up. Most of the symbols described here are treated in
the chapter on mathematical logic in Cajori's c@

   2
 

 (1929).
The ideas of the period are covered in I. Grattan-Guinness (2000) ' 
   
þ

8I>D8?JD"!
'
 
  

  
 


þ
1. Many of the classic works are available in English in Jan van
Heijenoort (1967) 
 
1"c'
 $

    !


8I>?8?@8.

              2 %      #

        # The symbols ŀ and ¢ were first used by Hermann Grassmann in 
c   
8IJJ §2+§5. There Grassmann used the two symbols still as general
operation symbols, not specialized for intersection and union. [Wilfried Neumaier]. From there
Giuseppe Peano (1858-1932) took both symbols and used them for intersection and union in
1888 in  


 


 c   @  (Cajori vol. 2, page
298); the logical part of this work with this symbols is ed. in: Peano, opere scelte, 2, Rom 1958,
p. 3-19.

Peano also created the large symbols for general intersection and union of more than two classes
in 1908. These symbols can be seen in the Wikipedia article on Union (set theory). The source is:
G. Peano: Formulario mathematico, tomo V, Torino 1908, Facsimile-Reprint, Rom 1960, p. 82.
[This information was provided by Wilfried Neumaier, who reports that Cajori vol. 2 does not
have the correct earliest use for these symbols.]

3  . 3  C   /# Peano used in volume II, number 1, of his 
  
  &% which was published in 1897 (Cajori vol. 2, page 300).

Kevin C. Klement writes, "While Peano had the backwards E for a predicate of classes, Russell
was the first to use the backwards E as a variable binding operator, and there are the wonderful
manuscripts printed in CPBR vol 4 in which Russell's makes large dots out of Peano's backwards
epsilons to change over from the Peano-notation for existence to a more Fregean one."

  2# Peano used for membership in ³Arithmetices prinicipia nova methodo exposita,´
1889; ed. in: =

   % Rom 1958, p. 27. [This citation was contributed by Wilfried
Neumaier.] The website at the University of St. Andrews states that Peano introduced the symbol
in 1889 and that it comes from the first letter of the Greek word meaning "is."

Peano also used this symbol in the introduction to volume I of his 


    &%
which was published in Turin in 1895, although the introduction itself is dated 1894 (Cajori vol.
2, page 300).
Peano's symbol for membership was an ordinary epsilon ; the stylized epsilon now used was
adopted by Bertrand Russell in =  
   in 1903 (Julio González Cabillón).

Kevin C. Klement writes, "I think Russell always intended to use epsilon, but the way both POM
and PM are typeset, it does come off looking more like the modern membership sign. Certainly
Russell's manuscripts after 1903 continue to use epsilon. Is typographical accident the root of the
modern sign?"

 # According to Julio González Cabillón, Peano introduced the backwards lower-case
epsilon for "such that" in "Formulaire de Mathematiques vol. II, #2" (p. iv, 1898).

Peano introduced the backwards lower-case epsilon for "such that" in his 1889 "Principles of
arithmetic, presented by a new method," according van Heijenoort's 
 
1"c
'
 $

    !


8I>?8?@8 [Judy Green].

 # According to M. J. Cresswell and Irving H. Anellis, originated in Gerhard Gentzen,
"Untersuchungen ueber das logische Schliessen,"  *., 59, (1935), p, 178. In footnote 4 on
that page, Gentzen explains how he came to use the sign. It is the "All-Zeichen," an analogy with
for the existential quantifier which Gentzen says that he borrowed from Russell.

Cajori, however, shows that Peano used before Russell and Whitehead (whose backwards E
had serifs, unlike Peano's).

Russell used the notation ./ for "for all ". See his "Mathematical Logic as Based on the Theory
of Types," c  .
 
  , 5<, (1908), 222-262. [Denis Roegel].

r          # Georg Cantor used the set brackets {a}, {a,b},... in 1878
in ³Ein Beitrag zur Mannifaltigkeitslehre´ in  .
 7    ?7 (1878), p. 242-
258. [This citation was contributed by Wilfried Neumaier.]

Cantor later used this symbolism as follows:

Unter einer 'Menge' verstehen wir jede Zusammenfassung  von bestimmten


wohlunterschiedenen Objecten  unsrer Anschauung oder unseres Denkens (welche die
'Elemente' von  genannt werden) zu einem Ganzen.

In Zeichen druecken wir dies so aus:

 = {}.
The citation above is from "Beiträge zur Begründung der transfiniten Mengenlehre"
[Contributions to the founding of the theory of transfinite numbers],   c  
Band XLVI [vol. 46], pp. 481-512, B. G. Teubner, Leipzig, 1895. Please recall that Cantor's
"Contributions to the founding of the theory of transfinite numbers" [first published by The Open
Court publishing Company, Chicago-London, 1915] is a translation of the two memoirs which
had appeared in   c  for 1895 and 1897 under the title: "Beiträge zur
Begründung der transfiniten Mengenlehre" -- translation from the German, introduction, and
notes by Philip Edward Bertrand Jourdain (1879-1919). An unabridged and unaltered
republication of the English translation mentioned was edited also by Dover Publications, Inc.,
New York, 1955 [ISBN: 0486600459].

 stands for the German term "Menge." Cantor may have used this notation earlier in his
correspondence with the mathematicians of his day. (This entry was contributed by Julio
González Cabillón.)

is used for propositions in 1897 in Peano's "Studii di logica matematica," according to Kevin
C. Klement.

    were used as propositional letters by Bertrand Russell in 1903 in =  



   (OED).

$  # The tilde ~ for negation was used by Peano in 1897. See Peano, ³Studii di logica
matematica,´ ed. in: Peano, opere scelte, 2, Rom 1958, p. 212. [Kevin C. Klement]

K
for "the negation of " appears in 1908 in the article "Mathematical logic as based on the
theory of types" by Bertrand Russell [Denis Roegel].

The symbolism was also used in 1910 by Alfred North Whitehead and Bertrand Russell in the
first volume of =      (Cajori vol. 2, page 307).

The main symbol for negation wich is used today is ¬. It was first used in 1930 by Arend
Heyting in ³Die formalen Regeln der intuitionistischen Logik,´ '(    K 
c      phys.-math. Klasse, 1930, p. 42-65. The ¬ appears on p. 43.
[Wilfried Neumaier]

;  # ç for disjunction is found in Russell's manuscripts from 1902-1903 and in 1906 in
Russell's paper "The Theory of Implication," in c  .
 
   vol. 28, pp.
159-202, according to Kevin C. Klement.

L(  for " or &" appears in 1908 in the article (1908) "Mathematical Logic as Based on the
Theory of Types," c  .
 
  , 30, 222-262. by Bertrand Russell [Denis
Roegel].

The symbolism was also used in 1910 by Alfred North Whitehead and Bertrand Russell in the
first volume of =       (These authors used & for " and &") (Cajori vol. 2,
page 307)

 ;  # The symbol ˆ for logical conjunction "and" was first used in 1930 by Arend
Heyting in the same source as shown for the negation symbol above. [Wilfried Neumaier]

 2 # The arrow symbol ĺ for the logical implication was first used in 1922 by David
Hilbert in: Hilbert: Neubegründung der Mathematik, 1922, in: Abhandlungen aus dem
Mathematischen Seminar der Hamburger Universität, Band I (1922), 157-177. The symbol is
found on p. 166. [Wilfried Neumaier]

The arrow with double lines ¶ was first used 1954 by Nicholas Bourbaki, in: Bourbaki: Theorie
des ensembles, 3. edition, Paris, 1954. The symbol appears on p. 14. [Wilfried Neumaier]

C  # The double arrow symbol ļ for the logical equivalence was apparently first used
in 1936 by Wilhelm Ackermann in: Ackermann, W.: Die Widerspruchsfreiheit der allgemeinen
Mengenlehre, in: Mathematische Annalen 114 (1937), 305-315. The symbol appears on p. 306.
[Wilfried Neumaier]

The double arrow with double line ä was first used 1954 by Nicholas Bourbaki, in: Bourbaki:
Theorie des ensembles, 3. edition, Paris, 1954. The symbol appears on p. 32. [Wilfried
Neumaier]

      .M/ first appeared in N. Bourbaki C%  % & 8"!
  
     4!8"
  (   )
(1939): "certaines propriétés... ne sont vraies pour  élément de E... la partie qu¶elles
définissent est appelée la   de E, et designée par la notation Ø." (p. 4.)

André Weil (1906-1998) says in his autobiography that he was responsible for the symbol:

Wisely, we had decided to publish an installment establishing the system of notation for set
theory, rather than wait for the detailed treatment that was to follow: it was high time to fix these
notations once and for all, and indeed the ones we proposed, which introduced a number of
modifications to the notations previously in use, met with general approval. Much later, my own
part in these discussions earned me the respect of my daughter Nicolette, when she learned the
symbol Ø for the empty set at school and I told her that I had been personally responsible for its
adoption. The symbol came from the Norwegian alphabet, with which I alone among the
Bourbaki group was familiar.
The citation above is from page 114 of André Weil's c  
    
Birkhaeuser Verlag, Basel-Boston-Berlin, 1992. Translated from the French by Jennifer Gage.
The citation was provided by Julio González Cabillón.

This letter is used in the Norwegian, Danish and Faroese alphabets.

 A   A   ( )was first published in 1659 in the original German edition of
c by Johann Rahn (1622-1676) (Cajori vol. 1, page 212, and vol 2., page 282).

 A  A   ( ). Cajori writes, "We have not been able to find the use of for
'because' in the eighteenth century. This usage seems to have been introduced in Great Britain
and the United States in the nineteenth century. It is found in the      

 
(1805). It did not meet with as wide acceptance in Great Britain and America as did
for 'therefore'" (pages 282-283).
   . 3        2 /# On the last page of his autobiography, Paul
R. Halmos writes:

My most nearly immortal contributions are an abbreviation and a typographical symbol. I


invented "iff", for "if and only if" -- but I could never believe that I was really its first inventor. I
am quite prepared to beieve that it existed before me, but I don't 
+ that it did, and my
invention (re-invention?) of it is what spread it thorugh the mathematical world. The symbol is
definitely not my invention -- it appeared in popular magazines (not mathematical ones) before I
adopted it, but, once again, I seem to have introduced it into mathematics. It is the symbol that
sometimes looks like [an empty square], and is used to indicate an end, usually the end of a
proof. It is most frequently called the "tombstone", but at least one generous author referred to it
as the "halmos".
This quote is from  
$    "c c
 
   by Paul R. Halmos,
Springer-Verlag, New York, Berlin, Heidelberg, Tokyo, 1985, page 403.

  2    was conceived by Georg Cantor (1845-1918) around 1893, and became
widely known after "Beiträge zur Begründung der transfiniten Mengenlehre" [Contributions to
the Foundation of Transfinite Set Theory] saw the light in   c   Band XLVI
[vol. 46], B. G. Teubner, Leipzig, 1895.

On page 492 of this prestigious journal we find the paragraph    
  ( c  [The minimum transfinite cardinal number Aleph null], and the
following:

...wir nennen die ihr zukommende Cardinalzahl, in Zeichen, ... [We call the cardinal number
related to that (set); in symbol, ]

In a letter dated April 30, 1895, Cantor wrote, "it seemed to me that for this purpose, other
alphabets were [already] over-used" (translation by Martin Davis).

In 
 
 Dauben (page 179) says that Cantor did not want to use Roman or Greek
alphabets, because they were already widely used, and "His new numbers deserved something
unique. ... Not wishing to invent a new symbol himself, he chose the aleph, the first letter of the
Hebrew alphabet...the aleph could be taken to represent new beginnings...." Avinoam Mann
points out that aleph is also the first letter of the Hebrew word "Einsof," which means infinity
and that the Kabbalists use "einsof" for the Godhead.

Although his father was a Lutheran and his mother was a Roman Catholic, Cantor had at least
some Jewish ancestry.

(Julio González Cabillón. contributed to this entry.)

    # According to Cajori (vol. 2, page 294), the symbols for "is included in"
(  
 ) and for "includes" (7 
 ) were introduced by Schröder 0
  
7 c  !

8 (1890). Previously the symbols < and > had been used.
According to a web page by Paul Taylor, "Joseph Gergonne introduced C for 
  and ) for
its converse in 1817, and these symbols were used by Peano and by Russell and Whitehead. (In
fact Russell and Whitehead 
used @ for containment in our sense.)"

According W. V. Quine, 



!
 4th ed., Harvard University Press, 1982, page 132:
³The inclusion signs and , now current in set theory, are derived from Gergonne's use in
1816 of 'C' for containment.´ In the bibliography he lists the following: Gergonne, J. D. ³Essai
de dialectique rationelle.´ c  % &  &% vol. 7 (1816-17), pp.
189-228. [Susanna Epp]

C     # =Def is found in C. Burali-Forti !


    (1894) (Grattan-
Guinness (2000) p. 216).

     "    


  2  1        1          %   
      22  # The most important written source is the definitive c@

   
2
 
 by Florian Cajori.

Vm Symbols of operation, including +, -, ', division, exponents, radical symbol, dot and vector product
Vm Grouping symbols, including (), [], {}, vinculum
Vm Symbols of relation, including =, >, <
Vm Fractions, including decimals
Vm Symbols for various constants, such as ʌ,  0
Vm Symbols for variables
Vm Symbols to represent various functions, such as log, ln, Ȗ, absolute value; also the ,- notation
Vm Symbols used in geometry
Vm Symbols used in trigonometry; also symbols for hyperbolic functions
Vm Symbols used in calculus
Vm Symbols for matrices and vectors
Vm Set notation and logic
Vm Symbols used in number theory
Vm Symbols used in probability and statistics
Vm Written sources for these pages

Please see also Earliest Known Uses of Some of the Words of Mathematics, Images of Mathematicians on Postage
Stamps, and Ambiguously Defined Mathematical Terms at the High School Level.

Assistance for this page has been provided by Julio González Cabillón, John Aldrich, Avinoam Mann, Eddie Mizzi,
Fred E. Wadley, Giovanni Ferraro, Judy Ann Brown, Len Berggren, Manoel Almeida, Michael Closs, Milo Gardner,
Paul Pollack, and Samuel S. Kutler. The page is maintained by Jeff Miller, a teacher at Gulf High School in New
Port Richey, Florida.

Anda mungkin juga menyukai