Mirifici Logarithmorum Canonis Descriptio

Last updated
John Napier (1550-1617), the inventor of logarithms John Napier.jpg
John Napier (1550–1617), the inventor of logarithms
Title page of Napier's 1614 table of logarithms of trigonometric functions Mirifici Logarithmorum Canonis Descriptio Mirifici Logarithmorum canonis Descriptio.jpg
Title page of Napier's 1614 table of logarithms of trigonometric functions Mirifici Logarithmorum Canonis Descriptio
The 19 degree pages from Napier's 1614 table. The left hand page covers angle increments of 0 to 30 minutes, the right hand page 30 to 60 minutes Napier's Mirici Logarithmorum table for 19 deg.agr.jpg
The 19 degree pages from Napier's 1614 table. The left hand page covers angle increments of 0 to 30 minutes, the right hand page 30 to 60 minutes

Mirifici Logarithmorum Canonis Descriptio (Description of the Wonderful Canon of Logarithms, 1614) and Mirifici Logarithmorum Canonis Constructio (Construction of the Wonderful Canon of Logarithms, 1619) are two books in Latin by John Napier expounding the method of logarithms. While others had approached the idea of logarithms, notably Jost Bürgi, it was Napier who first published the concept, along with easily used precomputed tables, in his Mirifici Logarithmorum Canonis Descriptio. [1] [2] [3]

Contents

Prior to the introduction of logarithms, high accuracy numerical calculations involving multiplication, division and root extraction were laborious and error prone. Logarithms greatly simplify such calculations. As Napier put it:

“…nothing is more tedious, fellow mathematicians, in the practice of the mathematical arts, than the great delays suffered in the tedium of lengthy multiplications and divisions, the finding of ratios, and in the extraction of square and cube roots… [with] the many slippery errors that can arise…I have found an amazing way of shortening the proceedings [in which]… all the numbers associated with the multiplications, and divisions of numbers, and with the long arduous tasks of extracting square and cube roots are themselves rejected from the work, and in their place other numbers are substituted, which perform the tasks of these rejected by means of addition, subtraction, and division by two or three only.” [2] :Preface

The book contains fifty-seven pages of explanatory matter and ninety pages of tables of trigonometric functions and their Napierian logarithms. [1] :p. 18 These tables greatly simplified calculations in spherical trigonometry, which are central to astronomy and celestial navigation and which typically include products of sines, cosines and other functions. Napier describes other uses, such a solving ratio problems, as well. [2] :Preface

John Napier spent 20 years calculating the tables. [4] :p. 16 He wrote a separate volume describing how he constructed his tables, but held off publication to see how his first book would be received. John died in 1617. His son, Robert, published his father's book, Mirifici Logarithmorum Canonis Constructio, with additions by Henry Briggs in 1619. [5] [6] The Constructio details how Napier created and used three tables of geometric progressions to facilitate the computation of logarithms of the sine function.

The tables

In Napier's time, decimal notation, as used in Europe, only represented integers. Including a fractional part with a decimal number had been proposed by Simon Stevin but his notation was awkward. The idea of using a period to separate the integer part of a decimal number from the fractional part was first proposed by Napier himself in his Constructio book, Section 5, “In numbers distinguished thus by a period in their midst, whatever is written after the period is a fraction, the denominator of which is unity with as many cyphers after it as there are figures after the period.“ He used the concept to facilitate more accurate computation of the tables, but not in the printed tables themselves.. [1] :p. 22 Thus an angle's sine value published in his table is a whole number representing the length of the side opposite that angle in a right triangle with hypotenuse of 10,000,000 units. The logarithm in the table, however, is of that sine value divided by 10,000,000. [1] :p. 19 The logarithm is again presented as an integer with an implied denominator of 10,000,000.

The table consists of 45 pairs of facing pages. Each pair is labeled at the top with an angle, from 0 to 44 degrees, and at the bottom from 90 to 45 degrees. (The 44–45 degree page being a single side.) The first column on each page of the table is an angle increment in minutes, to be added to the degree value at the top of the page. The far right column is minutes to be added to the degree value at the bottom of each page. This arrangement is such that for each line on a page, the full angle represented by column 7 is the co-angle of column 1 (90° – column_1). Moving inward, adjacent to each angle column is the sine of that angle, followed by the absolute value of Napierian log of that sine. One can get log-cosines for column 1 easily by reading across the page to column 5, and vice versa. The middle column shows the difference between the two logs, which is the Napierian log of the tangent function (cotangent if you reverse signs). [2] :Ch. III

The tables can also be used as a table of Napierian logarithms for positive numbers less than one, using the sine values (Columns 2 and 6) as the argument and the log sine values (Columns 3 and 5) as the resulting logarithm. Reversing the procedure gives anti-logarithms.

The first three rows in the table below reproduce the headings and first two data rows of the left-side 19 degree page of Napier's table, see photograph. They are followed by values computed using modern algorithms for the same angles. [6] These are truncated to 8-digit accuracy, one more than the nominal 7-digit accuracy of Napier's table. Column numbers are shown for clarity.

Napier's page (left side) for Gr. 19 (19 degrees)
mi.SinusLogarithmiDifferentieLogarithmi Sinus
032556821122183010661613560217945518660
132584321121338610652167561219945423959
θsin(θ)−ln(sin(θ))−ln(tan(θ))−ln(sin(90°-θ))sin(90°  θ)90°  θ
19°0.325568151.122183451.06616170.056021740.9455185771°
19° 1'0.325843181.121339051.06521710.056121950.9454238370° 59'
Column 1234567

The Descriptio text

Napier's Descriptio is divided into two books. The first describes his invention and some applications and includes the tables. The second discusses applications to trigonometry.

Book I

Chapter 1 contains a series of definitions and propositions that explains Napier's conception of logarithms. He conceived logarithms in terms of two models of motion. In the first, a particle starts at a point and moves along a straight line at a constant speed. In the second, a particle moves along a straight line, starting with the same initial speed, but its speed decreases in proportion to its distance from the starting point.:Ch. I,Def. 2 The logarithm of a number, a, is then the distance traveled by a particle in the constant speed model, during the time it takes the particle in the second, inverse proportional model to reach a. [2] :Ch. I,Def. 6 He defines the logarithm of 10,000,000 to be zero and that the logarithm of values less than that to be positive, while greater numbers have a negative logarithm, a reversal of sign from modern logarithms. (This sign change is sometimes expressed by saying Napier used logarithms with base 1/e.) [1] :p. 19 He notes that he has freedom to choose any value to have a zero log, in modern terms the base, but chooses 10,000,000 for ease of calculation as it matches the "total sine" (the hypotenuse), in his sine tables. See Naperian logarithm.

The second chapter describes the properties of logarithms and give some formulas (in text form) for working with ratios. It ends with a note stating he is delaying publication of his work on constructing logarithms until he sees how his invention is received. Chapter 3 describes the tables and their seven columns, see above. The fourth chapter explains how to use the tables and gives worked examples for sines tangents and secants. He also explains how to get logarithms of numbers directly by using the sine values as the argument and the log-sine values as result and vice versa. He discusses how to deal with different multiples of ten and introduces a notation, similar to modern scientific notation where he appends a number of zeros after the logarithm of a quantity to indicate a need for correction by decades. He give specific logarithm quantities to be added or subtracted in different cases:

These correspond to 10,000,000*ln(10), 10,000,000*ln(100), etc.

Chapter 5 presents four problems in proportionality and their solution using Napier's logarithms. He concludes by asking "how great a benefit is bestowed by these logarithms : since by the addition of these for multiplication, by subtraction for divisional, by division by two for the extraction of square roots, and by three for cube roots ... all the heavier work of calculation is avoided."

Book II

Book II deals with "that noble kind of Geometry, that is called Trigonometry." The first chapter deals with using logarithms to solve problems in plane trigonometry with right triangles and, in particular, with small angles, where his trigonometric logarithms become large. The next chapter cover plane oblique triangles. The remaining chapters cover spherical trigonometry, starting with quadrants. He also describes his Pentagramma mirificum, a five-pointed star on the sphere whose angles are all 90 degrees. [2] :p. 32

Mirifici Logarithmorum Canonis Constructio

Title page of English translation of Constructio Construction of the Wonderful Canon Logarithms-0025.jpg
Title page of English translation of Constructio

Napier was reluctant to publish the theory and details of how he created his table of logarithms pending feedback from the mathematics community on his ideas, and he died shortly after publication of the Discriptio. His son Robert published the Constructio in 1619. The volume has a preface by Robert and several appendices, including a section on John Napier's methods for more easily solving spherical triangles, and a section by Henry Biggs on “another and better kind of logarithms,” namely base 10 or common logarithms. An English translation by William Rae Macdonald was published, with annotations, in 1889. [6]

Napier's describes logarithms via a correspondence between two points moving under different speed profiles. The first point, P, moves along a finite line segment P0 to Q, with an initial speed that decreases proportional to P's distance to Q. The second point, L, moves along an unbounded line segment starting at L0 at the same time as P and with the same initial speed, but maintaining that speed without change. For each possible position of P, measured by its distance from Q, there is a corresponding simultaneous position of L. Napier defined the logarithm of the distance from P to Q to be distance from L0 to that L. [7]

Computation approach

Napier's first auxiliary table, a geometric progression computed by successive subtraction Napier's first table.agr.png
Napier's first auxiliary table, a geometric progression computed by successive subtraction
Napier's second auxiliary table. The final value should be 9995001.224804 Napier's second table.agr.png
Napier's second auxiliary table. The final value should be 9995001.224804
Napier's third "Radical" table Napier's third table.agr.png
Napier's third "Radical" table

Napier relies on several insights to compute his table of logarithms. To achieve high accuracy he starts with a large base of 10,000,000. But he then gets additional precision by using decimal fractions in a notation that he invented, but now universally familiar, namely using a decimal point. He goes on to explain how his notation works with some examples. He also introduces a form of interval arithmetic to bound any errors that occur in his calculations. [4] :p.8 [6] :Sec.3ff [4] :p.8

Another now familiar fact he observes is that fractions with denominators that are powers of 10 can be computed easily in decimal notation by shifting the number right relative the decimal point. As he puts it: "We call easy parts of a number, any parts the denominators of which are made up of unity and a number of cyphers, such parts being obtained by rejecting as many of the figures at the end of the principal number as there are cyphers in the denominator." [6] :Sec.14

He notes that arithmetic progressions are easy to calculate since they only involve addition and subtraction but that geometric progressions are, in general, harder to compute because they involve multiplication, division and possibly roots. However he observes that geometric progressions involving multipliers of the form 1 − 1/10m (i.e. of the form 0.99...9, with m nines) can be computed to arbitrary precision using just one shift and one subtraction per stage. [6] :Sec.16 Similarly, multipliers of the form 1 − 1/(2*10m) (i.e. 0.99...95) only require one shift, one division by two and one subtraction per stage for full precision, which he calls "tolerably easy." [6] :Sec.15

Napier also observes that logarithms of a geometric progression differ by a constant value at each stage, namely the logarithm of the multiplier. So if one knows the logarithm of the initial value of a geometric progression and of the multiplier, one can compute the logarithm of each member of the progression by repeated addition of the multiplier's logarithm.

Calculating first logarithm

Using his two line model, Napier finds lower and upper bounds for the logarithm of 0.9999999. His lower bound assumes the point P does not slow down, in which case L will move a distance of 1-0.9999999. His upper bound assumes P started out at its final speed of 0.9999999 in which case L will have moved the distance of (1-0.9999999)/0.9999999. Scaled up by his radius of 10,000,000, the lower bound is 1 and his upper bound is 1.0000001. He suggests that since the difference between these values is tiny, any value between them will present an "insensible error" of less than one part in 10 million, but he chooses, without much explanation, the midpoint, 1.00000005. [6] :Sec. 30,31 This choice gives him far greater precision, as his translator, William Rae Macdonald. points out in an appendix, noting that Napier's scaled up value for the logarithm of .9999999 is very close to the correct value, 1.000000050000003333333583..., [4] :9 and that all his subsequent computations of logarithms derive from the 1.00000005 value. Macdonald suggests that Napier must have had a better reason for picking the midpoint. [6] :p. 90ff

Auxiliary tables

Napier uses these insights to construct three tables. The first table, in modern notation, consists of the numbers 10000000*(0.9999999)n for n ranging from 0 to 100. The second consists of the numbers 100000*(0.99999)n for n ranging from 0 to 50. [6] :p.14 He then applies his value for the log of .9999999 to fill in logarithms for all the entries in his first table. He can use the last entry to compute the log of .99999, since .9999999100 is very close to .99999. He then the uses his second table, which is essentially 50 powers of .99999 to compute the logarithm of .9995. Macdonald also points out that an error crept into Napier's calculation of the second table; Napier's fiftieth value is 9995001.222927, but should be 9995001.22480. Macdonald discusses the consequences of this error in his appendix. [6] :p. 90ff [1] :p. 32

The radical table

Napier then constructs a third table of proportions with 69 columns and 21 rows, which he calls his "radical table." [1] :p. 35 The proportion along the top rows, starting with 1 is 0.99. The entries in each column are in proportion 0.9995. (Note that 0.9995 = 1-1/2000, allowing "tolerably easy" multiplication by halving, shifting and subtracting.) Napier uses the first column to computing the logarithm of .99, using log of .9995, which he already has. He can now fill in the logarithm of each entry in the third table because, by proportionality, the difference in logarithms between entries is constant. [6] :p.33 The third table now provides logarithms for a set of 1,449 values that cover the range from roughly 5,000,000 to 10,000,000, which corresponds to values of the sine function from 30 to 90 degrees, assuming a radius of 10,000,000. [4] :p. 7 Napier then explains how to use the tables to calculate a bounding interval for logarithms in that range.

Constructing the published tables

Latin title page of Constructio, 1619 Napier's Constructio Latin title page.jpg
Latin title page of Constructio, 1619

Napier then gives instructions for reproducing his published tables, with their seven columns and coverage of each minute of arc. He does not compute the sines themselves, the values for which are to be filled in from an already available table. "Reinhold's common table of sines, or any other more exact, will supply you with these values." [6] :Sec. 59,also p. 156 [4] :16 Logarithms of sines for angles from 30 degrees to 90 degrees are then computed by finding the closest number in the radical table and its logarithm and calculating the logarithm of the desired sine by linear interpolation. He suggests several ways for computing logarithms for sines of angles less than 30 degrees. For example, one can multiply a sine that is less than 0.5 by some power of two or ten to bring it into the range [0.5,1]. After finding that logarithm in the radical table, one adds the logarithm of the power of two or ten that was used (he gives a short table), to get the required logarithm. [1] :p. 36

Napier ends by pointing out that two of his methods for extending his table produce results with small differences. He proposes that others “who perchance may have plenty of pupils and computers” construct a new table with a larger scale factor of 10,000,000,000, by the same methods but using a radical table with only 35 columns, enough to cover angles from 45 to 90 degrees. [6] :Sec. 60,p.46

After matter

In an appendix, Napier discusses construction of “another and better kind of logarithm” where the logarithm of one is zero and the logarithm of ten is 10,000,000,000, the index. This is essentially base 10 logarithms with the large scale factor. [6] :p.97 He discusses various ways to compute such a table and ends by describing the logarithm of 2 as the number of digits in 210,000,000, which he computes as 301029995. [6] :p.48 The appendix is followed by remarks Henry Briggs on Napier's concepts and base-10 logarithms. The next section [6] :64 is a 12 page essay by Napier titled “Some very remarkable propositions for the solution of spherical triangles with wonderful ease,” where he describes how to solve them without dividing them into two right triangles. This section is also followed by commentary from Briggs.

The translator, Macdonald, includes some notes at this point, [6] :p.84 discussing the spelling of Napier's name, references to delays in publishing the second volume, the development of decimal arithmetic, the error in Napiers second table and the accuracy of Napier's method, and methods for computing base-10 logarithms.

The last section [6] :p.101 is a catalog by Macdonald of Napier's works in public libraries, including religious works, editions in different languages. and other books related to the work of John Napier and logarithms.

Reception

Napier's novel method of calculation spread quickly in Britain and abroad. Kepler dedicated his 1620 Ephereris to Napier, with a letter congratulating him on his invention and its benefits to astronomy. Kepler found no essential errors except for some inaccuracies at small angles. [1] :p. 16 Edward Wright, an authority on celestial navigation, translated Napier's Latin Descriptio into English in 1615, the next year, though publication was delayed by Wright's death. [2] Briggs extended the concept to the more convenient base 10, or common logarithm. [1] :pp.16-18 Ursinus called Napier "a mathematician without equal." [6] :p.158 Three hundred years later, in 1914, E. W. Hobson called logarithms "one of the very greatest scientific discoveries that the world has seen." [1] :p.5

In 1620 Edmund Gunter developed a ruler with a logarithmic scale; with a pair of dividers it could be used to multiply and divide. [8] In c. 1622, William Oughtred combined two handheld Gunter rules to make a calculating device that was essentially the first slide rule. [9]

The logarithm function became a staple of mathematical analysis, but printed tables of logarithms gradually diminished in importance in the twentieth century as multiplying mechanical calculators and, later, electronic computers took over high accuracy computation needs. [10] The introduction of hand-held scientific calculators in the 1970s ended the era of slide rules. [11] The current, 2002, edition of The American Practical Navigator (Bowditch) still contains tables of logarithms and logarithms of trigonometric functions. [12] :p.565ff

See also

Related Research Articles

<span class="mw-page-title-main">John Napier</span> Scottish mathematician (1550–1617)

John Napier of Merchiston, nicknamed Marvellous Merchiston, was a Scottish landowner known as a mathematician, physicist, and astronomer. He was the 8th Laird of Merchiston. His Latinized name was Ioannes Neper.

<span class="mw-page-title-main">Logarithm</span> Mathematical function, inverse of an exponential function

In mathematics, the logarithm is the inverse function to exponentiation. That means that the logarithm of a number x to the base b is the exponent to which b must be raised to produce x. For example, since 1000 = 103, the logarithm base 10 of 1000 is 3, or log10 (1000) = 3. The logarithm of x to base b is denoted as logb (x), or without parentheses, logbx, or even without the explicit base, log x, when no confusion is possible, or when the base does not matter such as in big O notation.

<span class="mw-page-title-main">Slide rule</span> Mechanical analog computer

A slide rule is a hand-operated mechanical calculator consisting of slidable rulers for evaluating mathematical operations such as multiplication, division, exponents, roots, logarithms, and trigonometry. It is one of the simplest analog computers.

<span class="mw-page-title-main">Neper</span> Logarithmic unit for ratios of measurements of physical field and power quantities

The neper is a logarithmic unit for ratios of measurements of physical field and power quantities, such as gain and loss of electronic signals. The unit's name is derived from the name of John Napier, the inventor of logarithms. As is the case for the decibel and bel, the neper is a unit defined in the international standard ISO 80000. It is not part of the International System of Units (SI), but is accepted for use alongside the SI.

<span class="mw-page-title-main">Common logarithm</span> Mathematical function

In mathematics, the common logarithm is the logarithm with base 10. It is also known as the decadic logarithm and as the decimal logarithm, named after its base, or Briggsian logarithm, after Henry Briggs, an English mathematician who pioneered its use, as well as standard logarithm. Historically, it was known as logarithmus decimalis or logarithmus decadis. It is indicated by log(x), log10(x), or sometimes Log(x) with a capital L; on calculators, it is printed as "log", but mathematicians usually mean natural logarithm (logarithm with base e ≈ 2.71828) rather than common logarithm when writing "log". To mitigate this ambiguity, the ISO 80000 specification recommends that log10(x) should be written lg(x), and loge(x) should be ln(x).

<span class="mw-page-title-main">Henry Briggs (mathematician)</span> British mathematician (1561–1630), inventor of common logarithms

Henry Briggs was an English mathematician notable for changing the original logarithms invented by John Napier into common logarithms, which are sometimes known as Briggsian logarithms in his honour. The specific algorithm for long division in modern use was introduced by Briggs c. 1600 AD.

<span class="mw-page-title-main">Mathematical table</span> List of values of a mathematical function

Mathematical tables are lists of numbers showing the results of a calculation with varying arguments. Trigonometric tables were used in ancient Greece and India for applications to astronomy and celestial navigation, and continued to be widely used until electronic calculators became cheap and plentiful, in order to simplify and drastically speed up computation. Tables of logarithms and trigonometric functions were common in math and science textbooks, and specialized tables were published for numerous applications.

The multiplication sign, also known as the times sign or the dimension sign, is the symbol ×, used in mathematics to denote the multiplication operation and its resulting product. While similar to a lowercase X, the form is properly a four-fold rotationally symmetric saltire.

<span class="mw-page-title-main">Napier's bones</span> 1617 device for calculating products and quotients

Napier's bones is a manually-operated calculating device created by John Napier of Merchiston, Scotland for the calculation of products and quotients of numbers. The method was based on lattice multiplication, and also called rabdology, a word invented by Napier. Napier published his version in 1617. It was printed in Edinburgh and dedicated to his patron Alexander Seton.

<span class="mw-page-title-main">Identity (mathematics)</span> Equation that is satisfied for all values of the variables

In mathematics, an identity is an equality relating one mathematical expression A to another mathematical expression B, such that A and B produce the same value for all values of the variables within a certain range of validity. In other words, A = B is an identity if A and B define the same functions, and an identity is an equality between functions that are differently defined. For example, and are identities. Identities are sometimes indicated by the triple bar symbol instead of =, the equals sign. Formally, an identity is a universally quantified equality.

<span class="mw-page-title-main">CORDIC</span> Algorithm for computing trigonometric, hyperbolic, logarithmic and exponential functions

CORDIC, also known as Volder's algorithm, or: Digit-by-digit methodCircular CORDIC, Linear CORDIC, Hyperbolic CORDIC, and Generalized Hyperbolic CORDIC, is a simple and efficient algorithm to calculate trigonometric functions, hyperbolic functions, square roots, multiplications, divisions, and exponentials and logarithms with arbitrary base, typically converging with one digit per iteration. CORDIC is therefore also an example of digit-by-digit algorithms. CORDIC and closely related methods known as pseudo-multiplication and pseudo-division or factor combining are commonly used when no hardware multiplier is available, as the only operations it requires are additions, subtractions, bitshift and lookup tables. As such, they all belong to the class of shift-and-add algorithms. In computer science, CORDIC is often used to implement floating-point arithmetic when the target platform lacks hardware multiply for cost or space reasons.

<span class="mw-page-title-main">Rabdology</span> The first of three devices to aid arithmetic calculation, described by John Napier

In 1617 a treatise in Latin titled Rabdologiæ and written by John Napier was published in Edinburgh. Printed three years after his treatise on the discovery of logarithms and in the same year as his death, it describes three devices to aid arithmetic calculations.

<span class="mw-page-title-main">Napierian logarithm</span>

The term Napierian logarithm or Naperian logarithm, named after John Napier, is often used to mean the natural logarithm. Napier did not introduce this natural logarithmic function, although it is named after him. However, if it is taken to mean the "logarithms" as originally produced by Napier, it is a function given by :

Prosthaphaeresis was an algorithm used in the late 16th century and early 17th century for approximate multiplication and division using formulas from trigonometry. For the 25 years preceding the invention of the logarithm in 1614, it was the only known generally applicable way of approximating products quickly. Its name comes from the Greek prosthesis (πρόσθεσις) and aphaeresis (ἀφαίρεσις), meaning addition and subtraction, two steps in the process.

<span class="mw-page-title-main">Bartholomaeus Pitiscus</span> German astronomer and mathematician (1561–1613)

Bartholomaeus Pitiscus was a 16th-century German trigonometrist, astronomer and theologian who first coined the word trigonometry.

<span class="mw-page-title-main">History of trigonometry</span>

Early study of triangles can be traced to the 2nd millennium BC, in Egyptian mathematics and Babylonian mathematics. Trigonometry was also prevalent in Kushite mathematics. Systematic study of trigonometric functions began in Hellenistic mathematics, reaching India as part of Hellenistic astronomy. In Indian astronomy, the study of trigonometric functions flourished in the Gupta period, especially due to Aryabhata, who discovered the sine function. During the Middle Ages, the study of trigonometry continued in Islamic mathematics, by mathematicians such as Al-Khwarizmi and Abu al-Wafa. It became an independent discipline in the Islamic world, where all six trigonometric functions were known. Translations of Arabic and Greek texts led to trigonometry being adopted as a subject in the Latin West beginning in the Renaissance with Regiomontanus. The development of modern trigonometry shifted during the western Age of Enlightenment, beginning with 17th-century mathematics and reaching its modern form with Leonhard Euler (1748).

<span class="mw-page-title-main">History of logarithms</span> Development of the mathematical function

The history of logarithms is the story of a correspondence between multiplication on the positive real numbers and addition on the real number line that was formalized in seventeenth century Europe and was widely used to simplify calculation until the advent of the digital computer. The Napierian logarithms were published first in 1614. E. W. Hobson called it "one of the very greatest scientific discoveries that the world has seen." Henry Briggs introduced common logarithms, which were easier to use. Tables of logarithms were published in many forms over four centuries. The idea of logarithms was also used to construct the slide rule, which became ubiquitous in science and engineering until the 1970s. A breakthrough generating the natural logarithm was the result of a search for an expression of area against a rectangular hyperbola, and required the assimilation of a new function into standard mathematics.

A timeline of numerals and arithmetic.

This is a timeline of pure and applied mathematics history. It is divided here into three stages, corresponding to stages in the development of mathematical notation: a "rhetorical" stage in which calculations are described purely by words, a "syncopated" stage in which quantities and common algebraic operations are beginning to be represented by symbolic abbreviations, and finally a "symbolic" stage, in which comprehensive notational systems for formulas are the norm.

Madhava's sine table is the table of trigonometric sines constructed by the 14th century Kerala mathematician-astronomer Madhava of Sangamagrama. The table lists the jya-s or Rsines of the twenty-four angles from 3.75° to 90° in steps of 3.75°. Rsine is just the sine multiplied by a selected radius and given as an integer. In this table, as in Aryabhata's earlier table, R is taken as 21600 ÷ 2π ≈ 3437.75.

References

  1. 1 2 3 4 5 6 7 8 9 10 11 12 Ernest William Hobson (1914), John Napier and the invention of logarithms, 1614, Cambridge: The University Press
  2. 1 2 3 4 5 6 7 Napier, John (1614). The Description of the Wonderful Canon of Logarithms. Translated by Wright, Edward; Bruce, Ian. 17centurymaths.com. Retrieved March 14, 2022.
  3. Napier, John (1614), Mirifici Logarithmorum Canonis Descriptio [The Description of the Wonderful Rule of Logarithms] (in Latin), Edinburgh, Scotland: Andrew Hart
  4. 1 2 3 4 5 6 Roegel, Denis (2010). "Napier's ideal construction of the logarithms, Research Report inria-00543934". HAL. INRIA. Retrieved 13 March 2022. from the Loria Collection of Mathematical Tables
  5. Napier, John; Briggs, Henry (1619). Mirifici Logarithmorum Canonis Constructio (in Latin). Edinburgh: Andres Hart.
  6. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 Napier, John (1889) [1620]. The Construction of the Wonderful Canon of Logarithms. Translated by Macdonald, William Rae. Edinburgh: Blackwood & Sons. Also available on Wikisource
  7. "Napier's approach to logarithms".
  8. Smith, David E. (1958). History of Mathematics. Courier Corporation. p. 205. ISBN   9780486204307.
  9. Applebaum, Wilbur (2003-12-16). "Slide Rule". Encyclopedia of the Scientific Revolution: From Copernicus to Newton. Routledge. Bibcode:2000esrc.book.....A. ISBN   9781135582555.
  10. Logarithms, Encyclopedia Britanica online edition, accessed June 17, 2022
  11. Slide rule, Encyclopedia Britanica online edition, accessed June 17, 2022
  12. American Practical Navigator, 2002, Publications/APN Current and previous editions at National Geospatial-Intelligence Agency