This file is indexed.

/usr/share/doc/yacas-doc/html/Algochapter4.html is in yacas-doc 1.3.1-1.

This file is owned by root:root, with mode 0o644.

The actual contents of the file can be viewed below.

   1
   2
   3
   4
   5
   6
   7
   8
   9
  10
  11
  12
  13
  14
  15
  16
  17
  18
  19
  20
  21
  22
  23
  24
  25
  26
  27
  28
  29
  30
  31
  32
  33
  34
  35
  36
  37
  38
  39
  40
  41
  42
  43
  44
  45
  46
  47
  48
  49
  50
  51
  52
  53
  54
  55
  56
  57
  58
  59
  60
  61
  62
  63
  64
  65
  66
  67
  68
  69
  70
  71
  72
  73
  74
  75
  76
  77
  78
  79
  80
  81
  82
  83
  84
  85
  86
  87
  88
  89
  90
  91
  92
  93
  94
  95
  96
  97
  98
  99
 100
 101
 102
 103
 104
 105
 106
 107
 108
 109
 110
 111
 112
 113
 114
 115
 116
 117
 118
 119
 120
 121
 122
 123
 124
 125
 126
 127
 128
 129
 130
 131
 132
 133
 134
 135
 136
 137
 138
 139
 140
 141
 142
 143
 144
 145
 146
 147
 148
 149
 150
 151
 152
 153
 154
 155
 156
 157
 158
 159
 160
 161
 162
 163
 164
 165
 166
 167
 168
 169
 170
 171
 172
 173
 174
 175
 176
 177
 178
 179
 180
 181
 182
 183
 184
 185
 186
 187
 188
 189
 190
 191
 192
 193
 194
 195
 196
 197
 198
 199
 200
 201
 202
 203
 204
 205
 206
 207
 208
 209
 210
 211
 212
 213
 214
 215
 216
 217
 218
 219
 220
 221
 222
 223
 224
 225
 226
 227
 228
 229
 230
 231
 232
 233
 234
 235
 236
 237
 238
 239
 240
 241
 242
 243
 244
 245
 246
 247
 248
 249
 250
 251
 252
 253
 254
 255
 256
 257
 258
 259
 260
 261
 262
 263
 264
 265
 266
 267
 268
 269
 270
 271
 272
 273
 274
 275
 276
 277
 278
 279
 280
 281
 282
 283
 284
 285
 286
 287
 288
 289
 290
 291
 292
 293
 294
 295
 296
 297
 298
 299
 300
 301
 302
 303
 304
 305
 306
 307
 308
 309
 310
 311
 312
 313
 314
 315
 316
 317
 318
 319
 320
 321
 322
 323
 324
 325
 326
 327
 328
 329
 330
 331
 332
 333
 334
 335
 336
 337
 338
 339
 340
 341
 342
 343
 344
 345
 346
 347
 348
 349
 350
 351
 352
 353
 354
 355
 356
 357
 358
 359
 360
 361
 362
 363
 364
 365
 366
 367
 368
 369
 370
 371
 372
 373
 374
 375
 376
 377
 378
 379
 380
 381
 382
 383
 384
 385
 386
 387
 388
 389
 390
 391
 392
 393
 394
 395
 396
 397
 398
 399
 400
 401
 402
 403
 404
 405
 406
 407
 408
 409
 410
 411
 412
 413
 414
 415
 416
 417
 418
 419
 420
 421
 422
 423
 424
 425
 426
 427
 428
 429
 430
 431
 432
 433
 434
 435
 436
 437
 438
 439
 440
 441
 442
 443
 444
 445
 446
 447
 448
 449
 450
 451
 452
 453
 454
 455
 456
 457
 458
 459
 460
 461
 462
 463
 464
 465
 466
 467
 468
 469
 470
 471
 472
 473
 474
 475
 476
 477
 478
 479
 480
 481
 482
 483
 484
 485
 486
 487
 488
 489
 490
 491
 492
 493
 494
 495
 496
 497
 498
 499
 500
 501
 502
 503
 504
 505
 506
 507
 508
 509
 510
 511
 512
 513
 514
 515
 516
 517
 518
 519
 520
 521
 522
 523
 524
 525
 526
 527
 528
 529
 530
 531
 532
 533
 534
 535
 536
 537
 538
 539
 540
 541
 542
 543
 544
 545
 546
 547
 548
 549
 550
 551
 552
 553
 554
 555
 556
 557
 558
 559
 560
 561
 562
 563
 564
 565
 566
 567
 568
 569
 570
 571
 572
 573
 574
 575
 576
 577
 578
 579
 580
 581
 582
 583
 584
 585
 586
 587
 588
 589
 590
 591
 592
 593
 594
 595
 596
 597
 598
 599
 600
 601
 602
 603
 604
 605
 606
 607
 608
 609
 610
 611
 612
 613
 614
 615
 616
 617
 618
 619
 620
 621
 622
 623
 624
 625
 626
 627
 628
 629
 630
 631
 632
 633
 634
 635
 636
 637
 638
 639
 640
 641
 642
 643
 644
 645
 646
 647
 648
 649
 650
 651
 652
 653
 654
 655
 656
 657
 658
 659
 660
 661
 662
 663
 664
 665
 666
 667
 668
 669
 670
 671
 672
 673
 674
 675
 676
 677
 678
 679
 680
 681
 682
 683
 684
 685
 686
 687
 688
 689
 690
 691
 692
 693
 694
 695
 696
 697
 698
 699
 700
 701
 702
 703
 704
 705
 706
 707
 708
 709
 710
 711
 712
 713
 714
 715
 716
 717
 718
 719
 720
 721
 722
 723
 724
 725
 726
 727
 728
 729
 730
 731
 732
 733
 734
 735
 736
 737
 738
 739
 740
 741
 742
 743
 744
 745
 746
 747
 748
 749
 750
 751
 752
 753
 754
 755
 756
 757
 758
 759
 760
 761
 762
 763
 764
 765
 766
 767
 768
 769
 770
 771
 772
 773
 774
 775
 776
 777
 778
 779
 780
 781
 782
 783
 784
 785
 786
 787
 788
 789
 790
 791
 792
 793
 794
 795
 796
 797
 798
 799
 800
 801
 802
 803
 804
 805
 806
 807
 808
 809
 810
 811
 812
 813
 814
 815
 816
 817
 818
 819
 820
 821
 822
 823
 824
 825
 826
 827
 828
 829
 830
 831
 832
 833
 834
 835
 836
 837
 838
 839
 840
 841
 842
 843
 844
 845
 846
 847
 848
 849
 850
 851
 852
 853
 854
 855
 856
 857
 858
 859
 860
 861
 862
 863
 864
 865
 866
 867
 868
 869
 870
 871
 872
 873
 874
 875
 876
 877
 878
 879
 880
 881
 882
 883
 884
 885
 886
 887
 888
 889
 890
 891
 892
 893
 894
 895
 896
 897
 898
 899
 900
 901
 902
 903
 904
 905
 906
 907
 908
 909
 910
 911
 912
 913
 914
 915
 916
 917
 918
 919
 920
 921
 922
 923
 924
 925
 926
 927
 928
 929
 930
 931
 932
 933
 934
 935
 936
 937
 938
 939
 940
 941
 942
 943
 944
 945
 946
 947
 948
 949
 950
 951
 952
 953
 954
 955
 956
 957
 958
 959
 960
 961
 962
 963
 964
 965
 966
 967
 968
 969
 970
 971
 972
 973
 974
 975
 976
 977
 978
 979
 980
 981
 982
 983
 984
 985
 986
 987
 988
 989
 990
 991
 992
 993
 994
 995
 996
 997
 998
 999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
1071
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
1083
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
1099
1100
1101
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
1113
1114
1115
1116
1117
1118
1119
1120
1121
1122
1123
1124
1125
1126
1127
1128
1129
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
1158
1159
1160
1161
1162
1163
1164
1165
1166
1167
1168
1169
1170
1171
1172
1173
1174
1175
1176
1177
1178
1179
1180
1181
1182
1183
1184
1185
1186
1187
1188
1189
1190
1191
1192
1193
1194
1195
1196
1197
1198
1199
1200
1201
1202
1203
1204
1205
1206
1207
1208
1209
1210
1211
1212
1213
1214
1215
1216
1217
1218
1219
1220
1221
1222
1223
1224
1225
1226
1227
1228
1229
1230
1231
1232
1233
1234
1235
1236
1237
1238
1239
1240
1241
1242
1243
1244
1245
1246
1247
1248
1249
1250
1251
1252
1253
1254
1255
1256
1257
1258
1259
1260
1261
1262
1263
1264
1265
1266
1267
1268
1269
1270
1271
1272
1273
1274
1275
1276
1277
1278
1279
1280
1281
1282
1283
1284
1285
1286
1287
1288
1289
1290
1291
1292
1293
1294
1295
1296
1297
1298
1299
1300
1301
1302
1303
1304
1305
1306
1307
1308
1309
1310
1311
1312
1313
1314
1315
1316
1317
1318
1319
1320
1321
1322
1323
1324
1325
1326
1327
1328
1329
1330
1331
1332
1333
1334
1335
1336
1337
1338
1339
1340
1341
1342
1343
1344
1345
1346
1347
1348
1349
1350
1351
1352
1353
1354
1355
1356
1357
1358
1359
1360
1361
1362
1363
1364
1365
1366
1367
1368
1369
1370
1371
1372
1373
1374
1375
1376
1377
1378
1379
1380
1381
1382
1383
1384
1385
1386
1387
1388
1389
1390
1391
1392
1393
1394
1395
1396
1397
1398
1399
1400
1401
1402
1403
1404
1405
1406
1407
1408
1409
1410
1411
1412
1413
1414
1415
1416
1417
1418
1419
1420
1421
1422
1423
1424
1425
1426
1427
1428
1429
1430
1431
1432
1433
1434
1435
1436
1437
1438
1439
1440
1441
1442
1443
1444
1445
1446
1447
1448
1449
1450
1451
1452
1453
1454
1455
1456
1457
1458
1459
1460
1461
1462
1463
1464
1465
1466
1467
1468
1469
1470
1471
1472
1473
1474
1475
1476
1477
1478
1479
1480
1481
1482
1483
1484
1485
1486
1487
1488
1489
1490
1491
1492
1493
1494
1495
1496
1497
1498
1499
1500
1501
1502
1503
1504
1505
1506
1507
1508
1509
1510
1511
1512
1513
1514
1515
1516
1517
1518
1519
1520
1521
1522
1523
1524
1525
1526
1527
1528
1529
1530
1531
1532
1533
1534
1535
1536
1537
1538
1539
1540
1541
1542
1543
1544
1545
1546
1547
1548
1549
1550
1551
1552
1553
1554
1555
1556
1557
1558
1559
1560
1561
1562
1563
1564
1565
1566
1567
1568
1569
1570
1571
1572
1573
1574
1575
1576
1577
1578
1579
1580
1581
1582
1583
1584
1585
1586
1587
1588
1589
1590
1591
1592
1593
1594
1595
1596
1597
1598
1599
1600
1601
1602
1603
1604
1605
1606
1607
1608
1609
1610
1611
1612
1613
1614
1615
1616
1617
1618
1619
1620
1621
1622
1623
1624
1625
1626
1627
1628
1629
1630
1631
1632
1633
1634
1635
1636
1637
1638
1639
1640
1641
1642
1643
1644
1645
1646
1647
1648
1649
1650
1651
1652
1653
1654
1655
1656
1657
1658
1659
1660
1661
1662
1663
1664
1665
1666
1667
1668
1669
1670
1671
1672
1673
1674
1675
1676
1677
1678
1679
1680
1681
1682
1683
1684
1685
1686
1687
1688
1689
1690
1691
1692
1693
1694
1695
1696
1697
1698
1699
1700
1701
1702
1703
1704
1705
1706
1707
1708
1709
1710
1711
1712
1713
1714
1715
1716
1717
1718
1719
1720
1721
1722
1723
1724
1725
1726
1727
1728
1729
1730
1731
1732
1733
1734
1735
1736
1737
1738
1739
1740
1741
1742
1743
1744
1745
1746
1747
1748
1749
1750
1751
1752
1753
1754
1755
1756
1757
1758
1759
1760
1761
1762
1763
1764
1765
1766
1767
1768
1769
1770
1771
1772
1773
1774
1775
1776
1777
1778
1779
1780
1781
1782
1783
1784
1785
1786
1787
1788
1789
1790
1791
1792
1793
1794
1795
1796
1797
1798
1799
1800
1801
1802
1803
1804
1805
1806
1807
1808
1809
1810
1811
1812
1813
1814
1815
1816
1817
1818
1819
1820
1821
1822
1823
1824
1825
1826
1827
1828
1829
1830
1831
1832
1833
1834
1835
1836
1837
1838
1839
1840
1841
1842
1843
1844
1845
1846
1847
1848
1849
1850
1851
1852
1853
1854
1855
1856
1857
1858
1859
1860
1861
1862
1863
1864
1865
1866
1867
1868
1869
1870
1871
1872
1873
1874
1875
1876
1877
1878
1879
1880
1881
1882
1883
1884
1885
1886
1887
1888
1889
1890
1891
1892
1893
1894
1895
1896
1897
1898
1899
1900
1901
1902
1903
1904
1905
1906
1907
1908
1909
1910
1911
1912
1913
1914
1915
1916
1917
1918
1919
1920
1921
1922
1923
1924
1925
1926
1927
1928
1929
1930
1931
1932
1933
1934
1935
1936
1937
1938
1939
1940
1941
1942
1943
1944
1945
1946
1947
1948
1949
1950
1951
1952
1953
1954
1955
1956
1957
1958
1959
1960
1961
1962
1963
1964
1965
1966
1967
1968
1969
1970
1971
1972
1973
1974
1975
1976
1977
1978
1979
1980
1981
1982
1983
1984
1985
1986
1987
1988
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010
2011
2012
2013
2014
2015
2016
2017
2018
2019
2020
2021
2022
2023
2024
2025
2026
2027
2028
2029
2030
2031
2032
2033
2034
2035
2036
2037
2038
2039
2040
2041
2042
2043
2044
2045
2046
2047
2048
2049
2050
2051
2052
2053
2054
2055
2056
2057
2058
2059
2060
2061
2062
2063
2064
2065
2066
2067
2068
2069
2070
2071
2072
2073
2074
2075
2076
2077
2078
2079
2080
2081
2082
2083
2084
2085
2086
2087
2088
2089
2090
2091
2092
2093
2094
2095
2096
2097
2098
2099
2100
2101
2102
2103
2104
2105
2106
2107
2108
2109
2110
2111
2112
2113
2114
2115
2116
2117
2118
2119
2120
2121
2122
2123
2124
2125
2126
2127
2128
2129
2130
2131
2132
2133
2134
2135
2136
2137
2138
2139
2140
2141
2142
2143
2144
2145
2146
2147
2148
2149
2150
2151
2152
2153
2154
2155
2156
2157
2158
2159
2160
2161
2162
2163
2164
2165
2166
2167
2168
2169
2170
2171
2172
2173
2174
2175
2176
2177
2178
2179
2180
2181
2182
2183
2184
2185
2186
2187
2188
2189
2190
2191
2192
2193
2194
2195
2196
2197
2198
2199
2200
2201
2202
2203
2204
2205
2206
2207
2208
2209
2210
2211
2212
2213
2214
2215
2216
2217
2218
2219
2220
2221
2222
2223
2224
2225
2226
2227
2228
2229
2230
2231
2232
2233
2234
2235
2236
2237
2238
2239
2240
2241
2242
2243
2244
2245
2246
2247
2248
2249
2250
2251
2252
2253
2254
2255
2256
2257
2258
2259
2260
2261
2262
2263
2264
2265
2266
2267
2268
2269
2270
2271
2272
2273
2274
2275
2276
2277
2278
2279
2280
2281
2282
2283
2284
2285
2286
2287
2288
2289
2290
2291
2292
2293
2294
2295
2296
2297
2298
2299
2300
2301
2302
2303
2304
2305
2306
2307
2308
2309
2310
2311
2312
2313
2314
2315
2316
2317
2318
2319
2320
2321
2322
2323
2324
2325
2326
2327
2328
2329
2330
2331
2332
2333
2334
2335
2336
2337
2338
2339
2340
2341
2342
2343
2344
2345
2346
2347
2348
2349
2350
2351
2352
2353
2354
2355
2356
2357
2358
2359
2360
2361
2362
2363
2364
2365
2366
2367
2368
2369
2370
2371
2372
2373
2374
2375
2376
2377
2378
2379
2380
2381
2382
2383
2384
2385
2386
2387
2388
2389
2390
2391
2392
2393
2394
2395
2396
2397
2398
2399
2400
2401
2402
2403
2404
2405
2406
2407
2408
2409
2410
2411
2412
2413
2414
2415
2416
2417
2418
2419
2420
2421
2422
2423
2424
2425
2426
2427
2428
2429
2430
2431
2432
2433
2434
2435
2436
2437
2438
2439
2440
2441
2442
2443
2444
2445
2446
2447
2448
2449
2450
2451
2452
2453
2454
2455
2456
2457
2458
2459
2460
2461
2462
2463
2464
2465
2466
2467
2468
2469
2470
2471
2472
2473
2474
2475
2476
2477
2478
2479
2480
2481
2482
2483
2484
2485
2486
2487
2488
2489
2490
2491
2492
2493
2494
2495
2496
2497
2498
2499
2500
2501
2502
2503
2504
2505
2506
2507
2508
2509
2510
2511
2512
2513
2514
2515
2516
2517
2518
2519
2520
2521
2522
2523
2524
2525
2526
2527
2528
2529
2530
2531
2532
2533
2534
2535
2536
2537
2538
2539
2540
2541
2542
2543
2544
2545
2546
2547
2548
2549
2550
2551
2552
2553
2554
2555
2556
2557
2558
2559
2560
2561
2562
2563
2564
2565
2566
2567
2568
2569
2570
2571
2572
2573
2574
2575
2576
2577
2578
2579
2580
2581
2582
2583
2584
2585
2586
2587
2588
2589
2590
2591
2592
2593
2594
2595
2596
2597
2598
2599
2600
2601
2602
2603
2604
2605
2606
2607
2608
2609
2610
2611
2612
2613
2614
2615
2616
2617
2618
2619
2620
2621
2622
2623
2624
2625
2626
2627
2628
2629
2630
2631
2632
2633
2634
2635
2636
2637
2638
2639
2640
2641
2642
2643
2644
2645
2646
2647
2648
2649
2650
2651
2652
2653
2654
2655
2656
2657
2658
2659
2660
2661
2662
2663
2664
2665
2666
2667
2668
2669
2670
2671
2672
2673
2674
2675
2676
2677
2678
2679
2680
2681
2682
2683
2684
2685
2686
2687
2688
2689
2690
2691
2692
2693
2694
2695
2696
2697
2698
2699
2700
2701
2702
2703
2704
2705
2706
2707
2708
2709
2710
2711
2712
2713
2714
2715
2716
2717
2718
2719
2720
2721
2722
2723
2724
2725
2726
2727
2728
2729
2730
2731
2732
2733
2734
2735
2736
2737
2738
2739
2740
2741
2742
2743
2744
2745
2746
2747
2748
2749
2750
2751
2752
2753
2754
2755
2756
2757
2758
2759
2760
2761
2762
2763
2764
2765
2766
2767
2768
2769
2770
2771
2772
2773
2774
2775
2776
2777
2778
2779
2780
2781
2782
2783
2784
2785
2786
2787
2788
2789
2790
2791
2792
2793
2794
2795
2796
2797
2798
2799
2800
2801
2802
2803
2804
2805
2806
2807
2808
2809
2810
2811
2812
2813
2814
2815
2816
2817
2818
2819
2820
2821
2822
2823
2824
2825
2826
2827
2828
2829
2830
2831
2832
2833
2834
2835
2836
2837
2838
2839
2840
2841
2842
2843
2844
2845
2846
2847
2848
2849
2850
2851
2852
2853
2854
2855
2856
2857
2858
2859
2860
2861
2862
2863
2864
2865
<html>
<head>
  <title>Numerical algorithms I: basic methods</title>
  <link rel="stylesheet" href="yacas.css" TYPE="text/css" MEDIA="screen">
</head>
<body>
<a name="c4">

</a>
<h1>
4. Numerical algorithms I: basic methods
</h1>
This and subsequent chapters document the numerical algorithms used in
Yacas for exact integer calculations as well as for multiple precision
floating-point calculations. We give self-contained descriptions of
the non-trivial algorithms and estimate their computational cost.
Most of the algorithms were taken from referenced literature; the remaining algorithms were developed by us.

<p> </p>

<p>

<a name="c4s1">

</a>
<h2>
<hr>4.1 Adaptive function plotting
</h2>
Here we consider plotting of functions <b>y=f(x)</b>.


<p>
There are two tasks related to preparation of plots of functions: first, to produce the numbers required for a plot, and second, to draw a plot with axes, symbols, a legend, perhaps additional illustrations and so on.
Here we only concern ourselves with the first task, that of preparation of the numerical data for a plot.
There are many plotting programs that can read a file with numbers and plot it in any desired manner.


<p>
Generating data for plots of functions generally does not require
high-precision calculations.
However, we need an algorithm that can be adjusted to produce data to different levels of precision.
In some particularly ill-behaved cases, a precise plot will not be possible and we would not want to waste time producing data that is too accurate for what it is worth.


<p>

<a name="plotting!adaptive algorithms">

</a>
<a name="plotting!two-dimensional">

</a>
<a name="Plot2D'adaptive@\relax <b><tt>Plot2D'adaptive</tt></b>">

</a>
A simple approach to plotting would be to divide the interval into many equal subintervals and to evaluate the function on the resulting grid.
Precision of the plot can be adjusted by choosing a larger or a smaller number of points.


<p>
However, this approach is not optimal. Sometimes a function changes rapidly near one point but slowly everywhere else.
For example, <b>f(x)=1/x</b> changes very quickly at small <b> x</b>.
Suppose we need to plot this function between <b> 0</b> and <b> 100</b>.
It would be wasteful to use the same subdivision interval everywhere: a finer grid is only required over a small portion of the plotting range near <b> x=0</b>.


<p>
The adaptive plotting routine <b><tt>Plot2D'adaptive</tt></b> uses a simple algorithm
to select the optimal grid to approximate a function of one argument <b> f(x)</b>.
The algorithm repeatedly subdivides the grid intervals near points where the existing grid does not represent the function well enough.
A similar algorithm
for adaptive grid refinement could be used for numerical integration. The
idea is that plotting and numerical integration require the same kind of
detailed knowledge about the behavior of the function.


<p>
The algorithm first splits the interval into a specified initial number of
equal subintervals, and then repeatedly splits each subinterval in half
until the function is well enough approximated by the resulting grid. The
integer parameter <b><tt>depth</tt></b> gives the maximum number of binary splittings for
a given initial interval; thus, at most <b>2^depth</b> additional grid points
will be generated. The function <b><tt>Plot2D'adaptive</tt></b> should return a list of
pairs of points <b><tt>{{x1,y1}, {x2,y2}, ...}</tt></b> to be used directly for plotting.


<p>
The adaptive plotting algorithm works like this:


<p>
<ul><li> 1.  Given an interval (</li><b> a</b>, <b> c</b>), we split
it in half, <b> b:=(a+c)/2</b> and first compute <b> f(x)</b> at five grid
points <b>a</b>, <b> a[1]:=(a+b)/2</b>, <b> b</b>, <b> b[1]:=(b+c)/2</b>, <b> c</b>. 
<li> 2. If currently </li><b> depth&lt;=0</b>, return this list of five points and
values because we cannot refine the grid any more.
<li> 3. Otherwise, check that the function does not oscillate too
rapidly on the interval [</li><b> a</b>, <b> c</b>]. 
The formal criterion is that the five values are all finite and do not make a "zigzag" pattern
such as (1,3,2,3,1).
More formally, we use the following procedure:
For each three consecutive values, write "1" if the middle value is larger
than the other two, or if it is smaller than the other two,
or if one of them is not a number (e.g. <b><tt>Infinity</tt></b> or <b><tt>Undefined</tt></b>).
If we have at most two ones now, then we consider the change of values to be
"slow enough". Otherwise it is not "slow enough".
In this case we need to refine the grid; go to step 5. 
Otherwise, go to step 4.
<li> 4. Check that the function values are smooth enough through the
interval. Smoothness is controlled by a parameter </li><b> epsilon</b>. The
meaning of the parameter <b> epsilon</b> is the (relative) error of the
numerical approximation of the integral of <b> f(x)</b> by the grid. A good heuristic
value of <b>epsilon</b> is 1/(the number of pixels on the screen)
because it means that no pixels will be missing in the area under
the graph. For this to work we need to make sure that we are actually computing the area <i>under</i> the graph; so we define <b> g(x):=f(x)-f[0]</b> where <b>f[0]</b> is the minimum of the values of <b>f(x)</b> on the five grid points <b>a</b>, <b> a[1]</b>, <b>b</b>, <b> b[1]</b>, and <b>c</b>; the function <b> g(x)</b> is nonnegative and has the minimum value 0.
Then we compute two different Newton-Cotes quadratures
for <b>Integrate(x,b,b[1])g(x)</b> using these five points. (Asymmetric
quadratures are chosen to avoid running into an accidental symmetry of the
function; the first quadrature uses points <b>a</b>, <b> a[1]</b>, <b>b</b>, <b> b[1]</b> and the second
quadrature uses <b>b</b>, <b> b[1]</b>, <b>c</b>.) If the
absolute value of the difference between these quadratures is less
than <b> epsilon</b> * (value of the second quadrature), then we
are done and we return the list of these five points and values.
<li> 5. Otherwise, we need to refine the grid. We compute
</li><b><tt>Plot2D'adaptive</tt></b> recursively for the two halves of the interval,
that is, for (<b> a</b>, <b> b</b>) and (<b> b</b>, <b> c</b>).
We also decrease <b><tt>depth</tt></b> by 1 and multiply <b> epsilon</b> by 2 because we need to maintain a constant <i>absolute</i> precision and this means that the relative error for the two subintervals can be twice as large.
The resulting two lists for the two subintervals are
concatenated (excluding the double value at point <b> b</b>) and
returned.
</ul>

<p>
This algorithm works well if the initial number of points and the <b><tt>depth</tt></b>
parameter are large enough.
These parameters can be adjusted to balance the available computing time and the desired level of detail in the resulting plot.


<p>
Singularities in the function are handled by the step 3.
Namely, the change in the sequence <b> a</b>, <b> a[1]</b>, <b>b</b>, <b> b[1]</b>, <b>c</b> is always considered to be "too rapid" if one of these values is a non-number (e.g. <b><tt>Infinity</tt></b> or <b><tt>Undefined</tt></b>).
Thus, the interval immediately adjacent to a singularity will be plotted at the highest allowed refinement level. When preparing the plotting data, the singular points are simply not printed to the data file, so that
a plotting programs does not encounter any problems.


<p>

<h3>
<hr>Newton-Cotes quadratures
</h3>
<a name="Newton-Cotes quadratures">

</a>
The meaning of Newton-Cotes quadrature coefficients is that an integral of a function <b> f(x)</b> is approximated by a sum,

<p><center><b>(Integrate(x,a[0],a[n])f(x))&lt;=&gt;h*Sum(k,0,n,c[k]*f(a[k])),</b></center></p>

where <b>a[k]</b> are the grid points, <b>h:=a[1]-a[0]</b> is the grid step, and
<b>c[k]</b> are the quadrature coefficients.
It may seem surprising, but these coefficients <b>c[k]</b> are independent
of the function <b>f(x)</b> and can be precomputed
in advance for a given grid <b>a[k]</b>.
[The quadrature coefficients do depend on the relative separations of the grid.
Here we assume a uniform grid with a constant step <b>h=a[k]-a[k-1]</b>.
Quadrature coefficients can also be found for non-uniform grids.]


<p>
The coefficients <b>c[k]</b> for
grids with a constant step <b>h</b> can be found, for example, by solving the following system of equations,

<p><center><b> Sum(k,0,n,c[k]*k^p)=n^(p+1)/(p+1) </b></center></p>

for <b>p=0</b>, 1, ..., <b> n</b>. This system of equations means that the quadrature correctly gives the integrals of <b> p+1</b> functions <b> f(x)=x^p</b>, <b> p=0</b>, 1, ..., <b> n</b>, over the interval (0, <b> n</b>).
The solution of this system always exists and gives quadrature coefficients as rational numbers. For example, the well-known Simpson quadrature <b> c[0]=1/3</b>, <b> c[1]=4/3</b>, <b> c[2]=1/3</b> is obtained with <b> n=2</b>.
An example of using this quadrature is the approximation

<p><center><b>(Integrate(x,0,2)f(x))&lt;=&gt;(f(0)+f(2))/3+4/3*f(1).</b></center></p>



<p>

<a name="Newton-Cotes quadratures!for partial intervals">

</a>
In the same way it is possible to find quadratures for the integral over a subinterval rather than over the whole interval of <b>x</b>. In the current implementation of the adaptive plotting algorithm, two quadratures are used: the 3-point quadrature (<b> n=2</b>) and the 4-point quadrature (<b> n=3</b>) for the integral over the first subinterval, <b> Integrate(x,a[0],a[1])f(x)</b>. Their coefficients are (<b>5/12</b>, <b> 2/3</b>, <b>-1/12</b>) and (<b> 3/8</b>, <b> 19/24</b>, <b>-5/24</b>, <b> 1/24</b>).
An example of using the first of these subinterval quadratures would be the approximation

<p><center><b>(Integrate(x,0,1)f(x))&lt;=&gt;5/12*f(0)+2/3*f(1)-1/12*f(2).</b></center></p>

These quadratures are intentionally chosen to be asymmetric to avoid an accidental cancellation when the function <b>f(x)</b> itself is symmetric.
(Otherwise the error estimate could accidentally become exactly zero.)


<p>

<a name="c4s2">

</a>
<h2>
<hr>4.2 Surface plotting
</h2>
Here we consider plotting of functions <b>z=f(x,y)</b>.


<p>

<a name="plotting!of surfaces">

</a>
<a name="plotting!three-dimensional">

</a>
The task of surface plotting is to obtain a picture of a two-dimensional surface as if it were a solid object in three dimensions.
A graphical representation of a surface is a complicated task.
Sometimes it is required to use particular coordinates or projections, to colorize the surface, to remove hidden lines and so on.
We shall only be concerned with the task of obtaining the data for a plot from a given function of two variables <b>f(x,y)</b>.
Specialized programs can take a text file with the data and let the user interactively produce a variety of surface plots.


<p>
The currently implemented algorithm in the function <b><tt>Plot3DS</tt></b> is very similar to the adaptive plotting algorithm for two-dimensional plots.
A given rectangular plotting region <b>a[1]&lt;=x&lt;=a[2]</b>, <b>b[1]&lt;=y&lt;=b[2]</b> is subdivided to produce an equally spaced rectangular grid of points.
This is the initial grid which will be adaptively refined where necessary.
The refinement algorithm will divide a given rectangle in four quarters if the available function values indicate that the function does not change smoothly enough on that rectangle.


<p>
The criterion of a "smooth enough" change is very similar to the procedure outlined in the previous section.
The change is "smooth enough" if all points are finite, nonsingular values, and if the integral of the function over the rectangle is sufficiently well approximated by a certain low-order "cubature" formula.


<p>
The two-dimensional integral of the function is estimated using the following 5-point Newton-Cotes cubature:


<p>
<table cellpadding="0" width="100%">
<tr><td width=100% bgcolor="#DDDDEE"><pre>
1/12   0   1/12

 0    2/3   0

1/12   0   1/12
</pre></tr>
</table>


<p>
An example of using this cubature would be the approximation

<p><center><b>(Integrate(y,0,1)Integrate(x,0,1)f(x,y))&lt;=&gt;(f(0,0)+f(0,1)+f(1,0)+f(1,1))/12 </b></center></p>


<p><center><b>+2/3*f(1/2,1/2).</b></center></p>



<p>
Similarly, an 8-point cubature with zero sum is used to estimate the error:


<p>
<table cellpadding="0" width="100%">
<tr><td width=100% bgcolor="#DDDDEE"><pre>
-1/3   2/3   1/6

-1/6  -2/3  -1/2

 1/2    0    1/3
</pre></tr>
</table>
This set of coefficients was intentionally chosen to be asymmetric to avoid possible exact cancellations when the function itself is symmetric.


<p>
One minor problem with adaptive surface plotting is that the resulting set of points may not correspond to a rectangular grid in the parameter space (<b>x</b>,<b> y</b>).
This is because some rectangles from the initial grid will need to be bisected more times than others.
So, unless adaptive refinement is disabled, the function <b><tt>Plot3DS</tt></b> produces a somewhat disordered set of points.
However, most surface plotting programs require that the set of data points be a rectangular grid in the parameter space.
So a smoothing and interpolation procedure is necessary to convert a non-gridded set of data points ("scattered" data) to a gridded set.


<p>

<a name="c4s3">

</a>
<h2>
<hr>4.3 Parametric plots
</h2>
<a name="plotting!parametric">

</a>
Currently, parametric plots are not directly implemented in Yacas.
However, it is possible to use Yacas to obtain numerical data for such plots.
One can then use external programs to produce actual graphics.


<p>
A two-dimensional parametric plot is a line in a two-dimensional space, defined by two equations such as <b>x=f(t)</b>, <b>y=g(t)</b>.
Two functions <b>f</b>, <b> g</b> and a range of the independent variable <b> t</b>, for example, <b> t[1]&lt;=t&lt;=t[2]</b>, need to be specified.


<p>

<a name="plotting!non-Euclidean coordinates">

</a>
Parametric plots can be used to represent plots of functions in non-Euclidean coordinates.
For example, to plot the function <b>rho=Cos(4*phi)^2</b> in polar coordinates (<b> rho</b>,<b> phi</b>), one can rewrite the Euclidean coordinates through the polar   coordinates,
<b> x=rho*Cos(phi)</b>,
<b>y=rho*Sin(phi)</b>,
and use the equivalent parametric plot with <b>phi</b> as the parameter:
<b> x=Cos(4*phi)^2*Cos(phi)</b>,
<b>y=Cos(4*phi)^2*Sin(phi)</b>.


<p>

<a name="plotting!three-dimensional">

</a>
Sometimes higher-dimensional parametric plots are required.
A line plot in three dimensions is defined by three functions of one variable, for example, <b>x=f(t)</b>, <b>y=g(t)</b>, <b>z=h(t)</b>, and a range of the parameter <b>t</b>.
A surface plot in three dimensions is defined by three functions of two variables each, for example, <b> x=f(u,v)</b>, <b>y=g(u,v)</b>, <b>z=h(u,v)</b>, and a rectangular domain in the (<b>u</b>,<b> v</b>) space.


<p>
The data for parametric plots can be generated separately using the same adaptive plotting algorithms as for ordinary function plots,
as if all functions such as <b> f(t)</b> or <b>g(u,v)</b> were unrelated functions.
The result would be several separate data sets for the <b>x</b>, <b> y</b>, ... coordinates.
These data sets could then be combined using an interactive plotting program.


<p>

<a name="c4s4">

</a>
<h2>
<hr>4.4 The cost of arbitrary-precision computations
</h2>
A computer algebra system absolutely needs to be able to perform
computations with very large <i>integer</i>  numbers. Without this
capability, many symbolic computations (such as exact GCD of
polynomials or exact solution of polynomial equations) would be
impossible.


<p>

<a name="arbitrary-precision computation">

</a>
A different question is whether a CAS really needs to be able to
evaluate, say, 10,000 digits of the value of a Bessel function of some
10,000-digit complex argument.
It seems likely that no applied problem of
natural sciences would need floating-point computations of special
functions with such a high precision. However, arbitrary-precision
computations are certainly useful in some mathematical applications;
e.g. some mathematical identities can be first guessed by a floating-point
computation with many digits and then proved.


<p>
Very high precision computations of special functions <i>might</i> be useful in the future.
But it is already quite clear that computations with moderately high
precision (say, 50 or 100 decimal digits) are useful for applied problems.
For example, to obtain the leading asymptotic of an analytic function, we could expand it in series and take the first term.
But we need to check that the coefficient at what we think is the leading term of the series does not vanish.
This coefficient could be a certain "exact" number such as <b>(Cos(355)+1)^2</b>.
This number is "exact" in the sense that it is made of integers and elementary functions.
But we cannot say <i>a priori</i> that this number is nonzero.
The problem of "zero determination" (finding out whether a certain "exact" number is zero)
is known to be algorithmically unsolvable if we allow transcendental functions.
The only practical general approach seems to be to compute the number in question with many digits.
Usually a few digits are enough, but occasionally several hundred digits are needed.


<p>
Implementing an efficient algorithm that computes 100 digits of
<b> Sin(3/7)</b> already involves many of the issues that would also be
relevant for a 10,000 digit computation. 
Modern algorithms allow evaluations of all elementary functions in time
that is asymptotically logarithmic in the number of digits <b>P</b> and
linear in the cost of long multiplication (usually denoted <b> M(P)</b>).
Almost all special functions can be evaluated in time that is asymptotically linear in <b>P</b> and in <b> M(P)</b>.
(However, this asymptotic cost sometimes applies only to very high precision, e.g., <b>P&gt;1000</b>, and different algorithms need to be implemented for calculations in lower precision.)


<p>
In <b><tt>Yacas</tt></b> we strive to implement all numerical functions to arbitrary precision.
All integer or rational functions return exact
results, and all floating-point functions return their value with <b> P</b>
correct decimal digits (assuming sufficient precision of the arguments).
The current value of <b> P</b> is accessed as
<b><tt>Builtin'Precision'Get()</tt></b> and may be changed by <b><tt>Builtin'Precision'Set(...)</tt></b>.


<p>

<a name="arbitrary-precision computation!requirements">

</a>
Implementing an arbitrary-precision floating-point computation of a
function <b> f(x)</b>, such as <b>f(x)=Exp(x)</b>, typically needs the following:


<p>
<ul><li>An algorithm that will compute </li><b>f(x)</b> for a given value <b>x</b> to a
user-specified precision of <b> P</b> (decimal) digits. Often, several
algorithms must be implemented for different subdomains of the
(<b> x</b>,<b> P</b>) space.
<li>An estimate of the computational cost of the algorithm(s), as a function of </li><b> x</b> and <b> P</b>. This is needed to select the best algorithm for given <b> x</b>, <b> P</b>.
<li>An estimate of the round-off error.
This is needed to select the "working precision" which will typically be somewhat higher than the precision of the final result.
</li></ul>

<p>
In calculations with machine precision where the number of digits is fixed, the problem of round-off errors is quite prominent.
Every arithmetic operation causes a small loss of precision;
as a result, a few last digits of the final value are usually incorrect.
But if we have an arbitrary precision capability, we can always increase precision by a few more digits during intermediate computations and thus eliminate all round-off error in the final result.
We should, of course, take care not to increase the working precision unnecessarily, because any increase of precision means slower calculations.
Taking twice as many digits as needed and hoping that the result is precise is not a good solution.


<p>
Selecting algorithms for computations is the most non-trivial part of
the implementation.
We want to achieve arbitrarily high precision, so
we need to find either a series, or a continued fraction, or a
sequence given by explicit formula, that converges to the function in a
controlled way.
It is not enough to use a table of precomputed values
or a fixed approximation formula that has a limited precision.


<p>
In the last 30 years, the interest in arbitrary-precision
computations grew and many efficient algorithms for elementary and
special functions were published.
Most algorithms are iterative.
Almost always it is very important to know in advance how many iterations
are needed for given <b> x</b>, <b> P</b>.
This knowledge allows to estimate the computational cost, in terms of
the required precision <b> P</b> and of the cost of long multiplication
<b> M(P)</b>, and choose the best algorithm.


<p>
Typically all operations will fall into one of the following categories
(sorted by the increasing cost):


<p>

<a name="arbitrary-precision computation!speed estimates">

</a>
<ul><li>addition, subtraction: linear in </li><b>P</b>;
<li>multiplication, division, integer power, integer root: linear in </li><b> M(P)</b>;
<li>elementary functions: </li><b>Exp(x)</b>, <b>Ln(x)</b>, <b>Sin(x)</b>, <b>ArcTan(x)</b> etc.: <b>M(P)*Ln(P)</b> or slower by some powers of <b>Ln(P)</b>;
<li>transcendental functions: </li><b>Erf(x)</b>, <b>Gamma(x)</b> etc.: typically <b>P*M(P)</b> or slower.
</ul>

<p>
The cost of long multiplication <b>M(P)</b> is between <b>O(P^2)</b> for low precision and <b>O(P*Ln(P))</b> for very high precision.
In some cases, a different algorithm should be chosen if the precision is high enough to allow <b>M(P)</b> faster than <b>O(P^2)</b>.


<p>
Some algorithms also need storage space
(e.g. an efficient algorithm for summation of the Taylor series uses <b>O(Ln(P))</b> temporary <b>P</b>-digit numbers).


<p>
Below we shall normally denote by <b> P</b> the required number of decimal digits.
The formulae frequently contain conspicuous factors of <b> Ln(10)</b>, so it will be clear how to obtain analogous expressions for another base.
(Most implementations use a binary base rather than a decimal base since it is more convenient for many calculations.)


<p>

<a name="c4s5">

</a>
<h2>
<hr>4.5 Estimating convergence of a series
</h2>
<a name="Taylor series!required number of terms">

</a>
Analyzing convergence of a power series is usually not difficult.
Here is a worked-out example of how we could estimate the required number of terms in the power series

<p><center><b>Exp(x)=1+x+x^2/2! +...+x^n/n! +O(x^(n+1)) </b></center></p>

if we need <b>P</b> decimal digits of precision in the result.
To be specific, assume that
<b> Abs(x)&lt;1</b>. (A similar calculation can be done for any other bound on <b> x</b>.)


<p>
Suppose we truncate the series after <b> n</b>-th term and
the series converges "well enough" after that term. Then the error will be
approximately equal to the first term we dropped. (This is what we
really mean by "converges well enough" and this will generally be the case in all applications, because we would not want to use a series that does not converge well enough.)


<p>
The term we dropped is <b> x^(n+1)/(n+1)!</b>.
To estimate <b>n!</b> for large <b>n</b>, one can use the inequality

<p><center><b> e^(e-1)*(n/e)^n&lt;n! &lt;(n/e)^(n+1) </b></center></p>

(valid for all <b>n&gt;=47</b>) which provides tight bounds for the growth of
the factorial, or a weaker inequality which is somewhat easier to use,

<p><center><b>(n/e)^n&lt;n! &lt;((n+1)/e)^(n+1) </b></center></p>

(valid for all <b>n&gt;=6</b>). The latter inequality is sufficient for most purposes.


<p>
If we use the upper bound on <b> n!</b> from this estimate, we find that the term we dropped is bounded by

<p><center><b>x^(n+1)/(n+1)! &lt;(e/(n+2))^(n+2).</b></center></p>

We need this number to be smaller than <b>10^(-P)</b>. This leads to an inequality

<p><center><b>(e/(n+2))^(n+2)&lt;10^(-P),</b></center></p>

which we now need to solve for <b>n</b>. The left hand side decreases with growing <b> n</b>. So it is clear that the inequality will hold for large enough <b> n</b>, say for <b> n&gt;=n0</b> where <b> n0</b> is an unknown (integer) value. We can take a logarithm of both sides, replace <b> n</b> with <b> n0</b> and obtain the following equation for <b> n0</b>:

<p><center><b>(n0+2)*Ln((n0+2)/e)=P*Ln(10).</b></center></p>

This equation cannot be solved exactly in terms of elementary functions; this is a typical situation in such estimates. However, we do not really need a very precise solution for <b>n0</b>; all we need is an estimate of its integer part.
This is also a typical situation.
It is acceptable if our approximate value of <b> n0</b> comes out a couple of units higher than necessary, because a couple of extra terms of the Taylor series will not significantly slow down the algorithm (but it is important that we do not underestimate <b> n0</b>).
Finally, we are mostly interested in having a good enough answer for
large values of <b> P</b>.


<p>
We can try to guess the result.
The largest term on the LHS
grows as <b> n0*Ln(n0)</b> and it should be approximately equal to
<b>P*Ln(10)</b>; but <b>Ln(n0)</b> grows very slowly, so this gives us a hint that
<b>n0</b> is proportional to <b> P*Ln(10)</b>.
As a first try, we set <b>n0=P*Ln(10)-2</b>
and compare the RHS with the LHS; we find that we have overshot by a
factor <b> Ln(P)-1+Ln(Ln(10))</b>, which is not a large factor. We
can now compensate and divide <b>n0</b> by this factor, so our second try is

<p><center><b> n0=(P*Ln(10))/(Ln(P)-1+Ln(Ln(10)))-2.</b></center></p>

(This approximation procedure is equivalent to solving the equation

<p><center><b> x=(P*Ln(10))/(Ln(x)-1) </b></center></p>

by direct iteration, starting from <b>x=P*Ln(10)</b>.)
If we substitute our second try for <b>n0</b> into the equation, we shall find that we undershot a little bit (i.e. the LHS is a little smaller than the RHS), but our <b> n0</b> is now smaller than it should be by a quantity that is smaller than 1 for large enough <b> P</b>.
So we should stop at this point and simply add 1 to this approximate answer. We should also replace <b> Ln(Ln(10))-1</b> by 0 for simplicity (this is safe because it will slightly increase <b> n0</b>.)


<p>
Our final result is that it is enough to take

<p><center><b> n=(P*Ln(10))/Ln(P)-1 </b></center></p>

terms in the Taylor series to compute <b> Exp(x)</b> for <b>Abs(x)&lt;1</b> to <b> P</b>
decimal digits. (Of course, if <b> x</b> is much smaller than 1, many fewer
terms will suffice.)


<p>

<a name="c4s6">

</a>
<h2>
<hr>4.6 Estimating the round-off error
</h2>
<a name="arbitrary-precision computation!round-off error estimates">

</a>
<h3>
<hr>Unavoidable round-off errors
</h3>
As the required precision <b>P</b> grows, an arbitrary-precision algorithm
will need more iterations or more terms of the series. So the round-off
error introduced by every floating-point operation will increase. When
doing arbitrary-precision computations, we can always perform all
calculations with a few more digits and compensate for round-off error.
It is however imperative to know in advance how many more digits we
need to take for our "working precision". We should also take that
increase into account when estimating the total cost of the method.
(In most cases this increase is small.)


<p>
Here is a simple estimate of the normal round-off error in a computation of
<b> n</b> terms of a power series.
Suppose that the sum of the series is of order <b> 1</b>, that the terms monotonically decrease in magnitude, and that adding one term requires two
multiplications and one addition. If all calculations are performed
with absolute precision <b> epsilon=10^(-P)</b>, then the total accumulated
round-off error is <b>3*n*epsilon</b>. If the relative error is <b> 3*n*epsilon</b>, it means that our
answer is something like <b> a*(1+3*n*epsilon)</b> where <b>a</b> is the correct
answer. We can see that out of the total <b> P</b> digits of this answer,
only the first <b> k</b> decimal digits are correct, where
<b> k= -Ln(3*n*epsilon)/Ln(10)</b>. In other words, we have lost

<p><center><b>P-k=Ln(3*n)/Ln(10) </b></center></p>

digits because of accumulated round-off error. So we found that we need
<b>Ln(3*n)/Ln(10)</b> extra decimal digits to compensate for this
round-off error.


<p>
This estimate assumes several things about the series (basically, that the series is "well-behaved").
These assumptions must be verified in each particular case.
For example, if the series begins with some large
terms but converges to a very small value, this estimate is wrong (see the next
subsection).


<p>
In the previous exercise we found the number of terms <b>n</b> for <b> Exp(x)</b>. So now we know how many extra digits of working precision we need for this particular case.


<p>
Below we shall have to perform similar estimates of the required number
of terms and of the accumulated round-off error in our analysis of the
algorithms.


<p>

<h3>
<hr>Catastrophic round-off error
</h3>
<a name="arbitrary-precision computation!catastrophic round-off error">

</a>
Sometimes the round-off error of a particular method of computation becomes so large that the method becomes highly inefficient.


<p>
Consider the computation of <b>Sin(x)</b> by the truncated Taylor series

<p><center><b>Sin(x)&lt;=&gt;Sum(k,0,N-1,(-1)^k*x^(2*k+1)/(2*k+1)!),</b></center></p>

when <b>x</b> is large.
We know that this series converges for all <b> x</b>, no matter how large.
Assume that <b> x=10^M</b> with <b> M&gt;=1</b>, and that we need <b> P</b> decimal digits of precision in the result.


<p>
First, we determine the necessary number of terms <b> N</b>.
The magnitude of the sum is never larger than <b> 1</b>.
Therefore we need the <b> N</b>-th term of the series to be smaller than <b> 10^(-P)</b>.
The inequality is
<b>(2*N+1)! &gt;10^(P+M*(2*N+1))</b>.
We obtain that <b>2*N+2&gt;e*10^M</b> is a necessary condition, and if <b> P</b> is large, we find approximately

<p><center><b> 2*N+2&lt;=&gt;((P-M)*Ln(10))/(Ln(P-M)-1-M*Ln(10)).</b></center></p>



<p>
However, taking enough terms does not yet guarantee a good result.
The terms of the series grow at first and then start to decrease.
The sum of these terms is, however, small.
Therefore there is some cancellation and we need to increase the working precision to avoid the round-off.
Let us estimate the required working precision.


<p>
We need to find the magnitude of the largest term of the series.
The ratio of the next term to the previous term is <b>x/(2*k*(2*k+1))</b> and therefore the maximum will be when this ratio becomes equal to <b>1</b>, i.e. for <b> 2*k&lt;=&gt;Sqrt(x)</b>.
Therefore the largest term is of order <b>x^Sqrt(x)/Sqrt(x)!</b> and so we need about <b>M/2*Sqrt(x)</b> decimal digits before the decimal point to represent this term.
But we also need to keep at least <b>P</b> digits after the decimal point, or else the round-off error will erase the significant digits of the result.
In addition, we will have unavoidable round-off error due to <b> O(P)</b> arithmetic operations.
So we should increase precision again by <b>P+Ln(P)/Ln(10)</b> digits plus a few guard digits.


<p>
As an example, to compute <b>Sin(10)</b> to <b>P=50</b> decimal digits with this method, we need a working precision of about <b> 60</b> digits, while to compute <b> Sin(10000)</b> we need to work with about <b>260</b> digits.
This shows how inefficient the Taylor series for <b> Sin(x)</b> becomes for large arguments <b>x</b>.
A simple transformation <b> x=2*Pi*n+x'</b> would have reduced <b> x</b> to at most 7, and the unnecessary computations with <b> 260</b> digits would be avoided.
The main cause of this inefficiency is that we have to add and subtract extremely large numbers to get a relatively small result of order <b> 1</b>.


<p>
We find that the method of Taylor series for <b> Sin(x)</b> at large <b>x</b> is highly inefficient because of round-off error and should be complemented by other methods.
This situation seems to be typical for Taylor series.


<p>

<a name="c4s7">

</a>
<h2>
<hr>4.7 Basic arbitrary-precision arithmetic 
</h2>
<a name="arbitrary-precision computation!Yacas internal math">

</a>
Yacas uses an internal math library (the <b><tt>yacasnumbers</tt></b> library) which comes with the source
code. This reduces the dependencies of the Yacas system and improves portability.
The internal math library is simple and does not necessarily use the most optimal algorithms.


<p>
If <b>P</b> is
the number of digits of precision, then multiplication and division take
<b> M(P)=O(P^2)</b> operations in the internal math. (Of course, multiplication and division by a short integer takes time linear in <b>P</b>.)
Much faster algorithms (Karatsuba, Toom-Cook, FFT multiplication, Newton-Raphson division etc.) are
implemented in <b><tt>gmp</tt></b>, <b><tt>CLN</tt></b> and some other libraries.
The asymptotic cost of multiplication for very large precision is <b> M(P)&lt;=&gt;O(P^1.6)</b> for the Karatsuba method and <b>M(P)=O(P*Ln(P)*Ln(Ln(P)))</b> for the FFT method.
In the estimates of computation cost in this book we shall assume that <b>M(P)</b> is at
least linear in <b>P</b> and maybe a bit slower.


<p>
The costs of multiplication may be different in various arbitrary-precision arithmetic libraries and on different computer platforms.
As a rough guide, one can assume that the straightforward <b> O(P^2)</b> multiplication is good until 100-200 decimal digits,
the asymptotically fastest method of FFT multiplication is good at the precision of about 5,000 or more decimal digits,
and the Karatsuba multiplication is best in the middle range.


<p>
Warning: calculations with internal Yacas math using precision exceeding 10,000 digits are currently impractically slow.


<p>

<a name="continued fraction approximation!estimating <b>Ln(2)</b>">

</a>
In some algorithms it is necessary to compute the integer parts of expressions such as <b>a*Ln(b)/Ln(10)</b> or <b>a*Ln(10)/Ln(2)</b>, where <b>a</b>, <b> b</b> are short integers of order <b> O(P)</b>. Such expressions are frequently needed to estimate the number of terms in the Taylor series or similar parameters of the algorithms. In these cases, it is important that the result is not underestimated. However, it would be wasteful to compute <b>1000*Ln(10)/Ln(2)</b> in great precision only to discard most of that information by taking the integer part of that number.
It is more efficient to approximate such constants from above by short rational numbers, for example, <b>Ln(10)/Ln(2)&lt;28738/8651</b> and <b> Ln(2)&lt;7050/10171</b>. The error of such an approximation will be small enough for practical purposes. The function <b><tt>BracketRational</tt></b> can be used to find optimal rational approximations.


<p>
The function <b><tt>IntLog</tt></b> (see below) efficiently computes the integer part of a logarithm (for an integer base, not a natural logarithm). If more precision is desired in calculating <b> Ln(a)/Ln(b)</b> for integer <b>a</b>, <b> b</b>, one can compute <b> IntLog(a^k,b)</b> for some integer <b>k</b> and then divide by <b> k</b>.


<p>

<a name="c4s8">

</a>
<h2>
<hr>4.8 How many digits of <b> Sin(Exp(Exp(1000)))</b> do we need?
</h2>
<a name="arbitrary-precision computation!very large numbers">

</a>
Arbitrary-precision math is not omnipotent against overflow.
Consider the problem of representing very large numbers such as <b>x=Exp(Exp(1000))</b>.
Suppose we need a floating-point representation of the number <b>x</b> with <b> P</b> decimal digits of precision.
In other words, we need to express
<b> x&lt;=&gt;M*10^E</b>,
where the mantissa <b> 1&lt;M&lt;10</b> is a floating-point number and the exponent
<b> E</b> is an integer, chosen so that the relative precision is <b> 10^(-P)</b>.
How much effort is needed to find <b>M</b> and <b> E</b>?


<p>
The exponent <b> E</b> is easy to obtain:

<p><center><b> E=Floor(Ln(x)/Ln(10))=Floor(Exp(1000)/Ln(10))&lt;=&gt;8.55*10^433.</b></center></p>

To compute the integer part <b> Floor(y)</b> of a number <b>y</b> exactly, we need to approximate <b> y</b> with at least <b> Ln(y)/Ln(10)</b> floating-point digits.
In our example, we find that we need 434 decimal digits to represent <b>E</b>.


<p>
Once we found <b> E</b>, we can write <b> x=10^(E+m)</b> where <b>m=Exp(1000)/Ln(10)-E</b> is a floating-point number, <b> 0&lt;m&lt;1</b>.
Then <b> M=10^m</b>.
To find <b> M</b> with <b> P</b> (decimal) digits, we need <b> m</b> with also at least <b> P</b> digits.
Therefore, we actually need to evaluate <b> Exp(1000)/Ln(10)</b> with <b>434+P</b> decimal digits before we can find <b> P</b> digits of the mantissa of <b> x</b>.
We ran into a perhaps surprising situation: one needs a high-precision calculation even to find the first digit of <b> x</b>, because it is necessary to find the exponent <b> E</b> exactly as an integer, and <b> E</b> is a rather large integer.
A normal double-precision numerical calculation would give an overflow error at this point.


<p>
Suppose we have found the number <b> x=Exp(Exp(1000))</b> with some precision.
What about finding <b>Sin(x)</b>?
Now, this is extremely difficult, because to find even the first digit of <b>Sin(x)</b> we have to evaluate <b>x</b> with <i>absolute</i> error of at most <b> 0.5</b>.
We know, however, that the number <b> x</b> has approximately <b> 10^434</b> digits <i>before</i> the decimal point.
Therefore, we would need to calculate <b> x</b> with at least that many digits.
Computations with <b> 10^434</b> digits is clearly far beyond the capability of modern computers.
It seems unlikely that even the sign of <b> Sin(Exp(Exp(1000)))</b> will be obtained in the near future.
<h6>It seems even less likely that the sign of <b>Sin(Exp(Exp(1000)))</b> would be of any use to anybody even if it could be computed.</h6>

<p>
Suppose that <b>N</b> is the largest integer that our arbitrary-precision facility can reasonably handle.
(For Yacas internal math library, <b> N</b> is about <b> 10^10000</b>.)
Then it follows that numbers <b> x</b> of order <b> 10^N</b> can be calculated with at most one (1) digit of floating-point precision,
while larger numbers cannot be calculated with any precision at all.


<p>
It seems that very large numbers can be obtained in practice only through exponentiation or powers.
It is unlikely that such numbers will arise from sums or products of reasonably-sized numbers in some formula.
<h6>A factorial function can produce rapidly growing results, but exact factorials <b> n!</b> for large <b>n</b> are well represented by the Stirling formula which involves powers and exponentials.</h6>For example, suppose a program operates with numbers <b> x</b> of size <b> N</b> or smaller;
a number such as <b> 10^N</b> can be obtained only by multiplying <b> O(N)</b> numbers <b>x</b> together.
But since <b> N</b> is the largest representable number, it is certainly not feasible to perform <b> O(N)</b> sequential operations on a computer.
However, it is feasible to obtain <b>N</b>-th power of a small number, since it requires only <b> O(Ln(N))</b> operations.


<p>

<a name="exponentially large numbers">

</a>
If numbers larger than <b>10^N</b>  are created only by exponentiation operations, then special exponential notation could be used to represent them.
For example, a very large number <b> z</b> could be stored and manipulated as an unevaluated exponential <b> z=Exp(M*10^E)</b> where <b>M&gt;=1</b> is a floating-point number with <b> P</b> digits of mantissa and <b> E</b> is an integer, <b> Ln(N)&lt;E&lt;N</b>.
Let us call such objects "exponentially large numbers" or "exp-numbers" for short.


<p>
In practice, we should decide on a threshold value <b> N</b> and promote a number to an exp-number when its logarithm exceeds <b> N</b>.


<p>
Note that an exp-number <b> z</b> might be positive or negative, e.g.
<b> z= -Exp(M*10^E)</b>.


<p>
Arithmetic operations can be applied to the exp-numbers.
However, exp-large arithmetic is of limited use because of an almost certainly critical loss of precision.
The power and logarithm operations can be meaningfully performed on exp-numbers <b>z</b>.
For example, if <b> z=Exp(M*10^E)</b> and <b>p</b> is a normal floating-point number, then <b> z^p=Exp(p*M*10^E)</b> and <b>Ln(z)=M*10^E</b>.
We can also multiply or divide two exp-numbers.
But it makes no sense to multiply an exp-number <b> z</b> by a normal number because we cannot represent the difference between <b> z</b> and say <b> 2.52*z</b>.
Similarly, adding <b> z</b> to anything else would result in a total underflow, since we do not actually know a single digit of the decimal representation of <b> z</b>.
So if <b> z1</b> and <b> z2</b> are exp-numbers, then <b> z1+z2</b> is simply equal to either <b> z1</b> or <b> z2</b> depending on which of them is larger.


<p>
We find that an exp-number <b> z</b> acts as an effective "infinity" compared with normal numbers.
But exp-numbers cannot be used as a device for computing limits: the unavoidable underflow will almost certainly produce wrong results.
For example, trying to verify 

<p><center><b>(Limit(x,0)(Exp(x)-1)/x)=1 </b></center></p>

by substituting <b> x=1/z</b> with some exp-number <b> z</b> gives 0 instead of 1.


<p>
Taking a logarithm of an exp-number brings it back to the realm of normal, representable numbers.
However, taking an exponential of an exp-number results in a number which is not representable even as an exp-number.
This is because an exp-number <b> z</b> needs to have its exponent <b> E</b> represented exactly as an integer,
but <b> Exp(z)</b> has an exponent of order <b>O(z)</b> which is not a representable number.
The monstrous number <b>Exp(z)</b> could be only written as <b>Exp(Exp(M*10^E))</b>, i.e. as a "doubly exponentially large" number, or "2-exp-number" for short.
Thus we obtain a hierarchy of iterated exp-numbers.
Each layer is "unrepresentably larger" than the previous one.


<p>

<a name="arbitrary-precision computation!very small numbers">

</a>
<a name="exponentially small numbers">

</a>
The same considerations apply to very small numbers of the order <b>10^(-N)</b> or smaller.
Such numbers can be manipulated as "exponentially small numbers", i.e. expressions of the form <b>Exp(-M*10^E)</b> with floating-point mantissa <b>M&gt;=1</b> and integer <b> E</b> satisfying <b> Ln(N)&lt;E&lt;N</b>.
Exponentially small numbers act as an effective zero compared with normal numbers.


<p>
Taking a logarithm of an exp-small number makes it again a normal representable number.
However, taking an exponential of an exp-small number produces 1 because of underflow.
To obtain a "doubly exponentially small" number, we need to take a reciprocal of a doubly exponentially large number, or take the exponent of an exponentially large negative power.
In other words, <b> Exp(-M*10^E)</b> is exp-small, while <b>Exp(-Exp(M*10^E))</b> is 2-exp-small.


<p>
The practical significance of exp-numbers is rather limited. 
We cannot obtain even a single significant digit of an exp-number.
A "computation" with exp-numbers is essentially a floating-point computation with logarithms of these exp-numbers.
A practical problem that needs numbers of this magnitude can probably be restated in terms of more manageable logarithms of such numbers.
In practice, exp-numbers could be useful not as a means to get a numerical answer,
but as a warning sign of critical overflow or underflow.
<h6>Yacas currently does not implement exp-numbers or any other guards against overflow and underflow. If a decimal exponential becomes too large, an incorrect answer may result.</h6>
<a name="c4s9">

</a>
<h2>
<hr>4.9 Continued fractions
</h2>
A "continued fraction" is an expression of the form

<p><center><b>a[0]+b[0]/(a[1]+b[1]/(a[2]+b[2]/(a[3]+...))).</b></center></p>

The coefficients <b>a[i]</b>, <b>b[i]</b> are called the "terms" of the fraction.
(Usually one has <b>a[i]!=0</b>, <b> b[i]!=0</b>.)
The above continued fraction is sometimes written as

<p><center><b> a[0]+b[0]/(a[1]+...)*b[1]/(a[2]+...)*b[2]/(a[3]+...)*... </b></center></p>



<p>
Usually one considers infinite continued fractions, i.e. the sequences <b>a[i]</b>, <b>b[i]</b> are infinite.
The value of an infinite continued fraction is defined as the limit of the fraction truncated after a very large number of terms.
(A continued traction can be truncated after <b>n</b>-th term if one replaces <b> b[n]</b> by <b>0</b>.)


<p>
An infinite continued fraction does not always converge.
Convergence depends on the values of the terms.


<p>
The representation of a number via a continued fraction is not unique because we could, for example, multiply the numerator and the denominator of any simple fraction inside it by any number.
Therefore one may consider some normalized representations.
A continued fraction is called "regular" if <b> b[k]=1</b> for all <b> k</b>, all <b> a[k]</b> are integer and <b>a[k]&gt;0</b> for <b> k&gt;=1</b>.
Regular continued fractions always converge.


<p>

<h3>
<hr>Approximation of numbers by continued fractions
</h3>
<a name="continued fraction approximation!of rational numbers">

</a>
The function <b><tt>ContFrac</tt></b> converts a (real) number <b> r</b> into a regular
continued fraction with integer terms,

<p><center><b> r=n[0]+1/(n[1]+1/(n[2]+...)).</b></center></p>

Here all numbers <b>n[i]</b> are integers and all except <b>n[0]</b> are positive.
This representation of a real number <b>r</b> is unique.
We may write this representation as <b> r=[n[0];n[1];n[2];...;]</b>.
If <b>r</b> is a rational number, the corresponding regular continued fraction is finite, terminating at some <b> n[N]</b>.
Otherwise the continued fraction will be infinite.
It is known that the truncated fractions will be in some sense "optimal" rational representations of the real number <b>r</b>.


<p>
The algorithm for converting a rational number <b> r=n/m</b> into a regular continued
fraction is simple. First, we determine the integer part of <b> r</b>, which is
<b> Div(n,m)</b>. If it is negative, we need to subtract one, so that <b>r=n[0]+x</b> and
the remainder <b> x</b> is nonnegative and less than <b> 1</b>. The remainder
<b> x=Mod(n,m)/m</b> is then inverted, <b> r[1]:=1/x=m/Mod(n,m)</b> and so we have
completed the first step in the decomposition, <b>r=n[0]+1/r[1]</b>; now <b>n[0]</b>
is integer but <b>r[1]</b> is perhaps not integer. We repeat the same procedure on
<b>r[1]</b>, obtain the next integer term <b>n[1]</b> and the remainder <b>r[2]</b> and so on,
until such <b>n</b> that <b> r[n]</b> is an integer and there is no more work to do. This
process will always terminate.


<p>
If <b>r</b> is a real number which is known by its floating-point representation at some precision, then we can use the same algorithm because all floating-point values are actually
rational numbers.


<p>
Real numbers known by their exact representations can sometimes be expressed as infinite continued fractions, for example

<p><center><b> Sqrt(11)=[3;3;6;3;6;3;6;...;];</b></center></p>


<p><center><b>Exp(1/p)=[1;p-1;1;1;3*p-1;1;1;5*p-1;...;].</b></center></p>



<p>

<a name="GuessRational@\relax <b><tt>GuessRational</tt></b>">

</a>
The functions <b><tt>GuessRational</tt></b> and <b><tt>NearRational</tt></b> take a real number <b>x</b> and use
continued fractions to find rational approximations <b> r=p/q&lt;=&gt;x</b> with "optimal" (small) numerators and denominators <b> p</b>, <b> q</b>.


<p>
Suppose we know that a
certain number <b> x</b> is rational but we have only a floating-point representation of
<b> x</b> with a limited precision, for example, <b> x&lt;=&gt;1.5662650602409638</b>.
We would like to guess a
rational form for <b> x</b> (in this example <b> x=130/83</b>).
The function <b><tt>GuessRational</tt></b> solves this problem.


<p>
Consider the following example. The number <b> 17/3</b> has a continued fraction
expansion <b><tt>{5,1,2}</tt></b>. Evaluated as a floating point number with limited
precision, it may become something like <b> 17/3+0.00001</b>, where the small number
represents a round-off error. The continued fraction expansion of this number is
<b><tt>{5, 1, 2, 11110, 1, 5, 1, 3, 2777, 2}</tt></b>. The presence of an unnaturally large
term <b> 11110</b> clearly signifies the place where the floating-point error was
introduced; all terms following it should be discarded to recover the continued fraction <b><tt>{5,1,2}</tt></b> and from it the initial number <b> 17/3</b>.


<p>
If a continued fraction for a number <b> x</b> is
cut right before an unusually large term and
evaluated, the resulting rational number has a small denominator and is very close to <b> x</b>.
This works because partial continued fractions provide "optimal"
rational approximations for the final (irrational) number, and because the
magnitude of the terms of the partial fraction is related to the magnitude of
the denominator of the resulting rational approximation.


<p>
<b><tt>GuessRational(x, prec)</tt></b> needs to choose the place where it should cut the
continued fraction.
The algorithm for this is somewhat heuristic but it works well enough.
The idea is to cut the continued fraction
when adding one more term would change the result by less than the specified
precision. To realize this in practice, we need an estimate of how much a
continued fraction changes when we add one term.


<p>

<a name="continued fraction approximation!error bound">

</a>
The routine <b><tt>GuessRational</tt></b> uses a (somewhat weak) upper bound for the difference of continued fractions that differ only by an additional last term:

<p><center><b> Abs(delta):=Abs(1/(a[1]+1/(...+1/a[n]))-1/(a[1]+1/(...+1/a[n+1])))&lt;1/((a[1]*...*a[n])^2*a[n+1]).</b></center></p>

(The derivation of this inequality is given below.)
Thus we should compute the product of successive terms <b>a[i]</b> of the continued fraction and stop at <b>a[n]</b> at which this product exceeds the maximum number of digits. The routine <b><tt>GuessRational</tt></b> has a second parameter <b><tt>prec</tt></b> which is by default 1/2 times the number of decimal digits of current precision; it stops at <b>a[n]</b> at which the product <b>a[1]*...*a[n]</b> exceeds <b>10^prec</b>.


<p>
The above estimate for <b> delta</b> hinges on the inequality

<p><center><b> 1/(a+1/(b+...))&lt;1/a </b></center></p>

and is suboptimal if some terms <b> a[i]=1</b>, because the product of <b> a[i]</b> does not increase when one of the terms is equal to 1, whereas in fact these terms do make <b>delta</b> smaller. A somewhat better estimate would be obtained if we use the inequality

<p><center><b> 1/(a+1/(b+1/(c+...)))&lt;1/(a+1/(b+1/c)).</b></center></p>

(See the next section for more explanations of precision of continued fraction approximations.)
This does not lead to a significant improvement if <b>a&gt;1</b> but makes a difference when <b> a=1</b>. In the product <b> a[1]*...*a[n]</b>, the terms <b>a[i]</b> which are equal to 1 should be replaced by

<p><center><b>a[i]+1/(a[i+1]+1/a[i+2]).</b></center></p>

Since the comparison of <b>a[1]*...*a[n]</b> with <b>10^prec</b> is qualitative, it it enough to perform calculations of <b> a[1]*...*a[n]</b> with limited precision.


<p>
This algorithm works well if <b>x</b> is computed with enough precision; namely, it
must be computed to at least as many digits as there are in the numerator and
the denominator of the fraction combined. Also, the parameter <b><tt>prec</tt></b> should not
be too large (or else the algorithm will find another rational number with a larger
denominator that approximates <b> x</b> "better" than the precision to which you know <b> x</b>).


<p>

<a name="NearRational@\relax <b><tt>NearRational</tt></b>">

</a>
The related function <b><tt>NearRational(x, prec)</tt></b> works somewhat differently. The
goal is to find an "optimal" rational number, i.e. with smallest numerator and
denominator, that is within the distance <b> 10^(-prec)</b> of a given value <b>x</b>.
The function <b><tt>NearRational</tt></b> does not always give the same answer as <b><tt>GuessRational</tt></b>.


<p>
The
algorithm for <b><tt>NearRational</tt></b> comes from the HAKMEM [Beeler <i>et al.</i> 1972], Item 101C. Their
description is terse but clear:


<p>
<table cellpadding="0" width="100%">
<tr><td width=100% bgcolor="#DDDDEE"><pre>
Problem: Given an interval, find in it the
rational number with the smallest numerator and
denominator.
Solution: Express the endpoints as continued
fractions.  Find the first term where they differ
and add 1 to the lesser term, unless it's last. 
Discard the terms to the right.  What's left is
the continued fraction for the "smallest"
rational in the interval.  (If one fraction
terminates but matches the other as far as it
goes, append an infinity and proceed as above.)
</pre></tr>
</table>


<p>
The HAKMEM text [Beeler <i>et al.</i> 1972] contains several interesting insights relevant to continued fractions and other numerical algorithms.


<p>

<h3>
<hr>Accurate computation of continued fractions
</h3>
Sometimes an analytic function <b> f(x)</b> can be approximated using a continued
fraction that contains <b>x</b> in its terms. Examples include the inverse tangent
<b> ArcTan(x)</b>, the error function <b>Erf(x)</b>, and the incomplete gamma function
<b>Gamma(a,x)</b> (see below for details).
For these functions, continued fractions
provide a method of numerical calculation that works when the Taylor series
converges slowly or not at all, or is not easily available.
However, continued fractions may converge
quickly for one value of <b>x</b> but slowly for another.
Also, it is not as easy to
obtain an analytic error bound for a continued fraction approximation as it is for power series.


<p>
In this section we describe some methods for computing general continued fractions and for estimating the number of terms needed to achieve a given precision.


<p>
Let us introduce some notation. A continued fraction

<p><center><b> a[0]+b[0]/(a[1]+b[1]/(a[2]+...)) </b></center></p>

is specified by a set of terms (<b>a[i]</b>, <b>b[i]</b>).
[If continued fractions are used to approximate analytic functions such as <b>ArcTan(x)</b>, then (<b>a[i]</b>, <b>b[i]</b>) will depend on <b>x</b>.]
Let us denote by <b> F[m][n]</b> the truncated fraction containing only the terms from <b>m</b> to <b> n</b>,

<p><center><b> F[m][n]:=a[m]+b[m]/(a[m+1]+b[m+1]/(...+b[n]/a[n])).</b></center></p>

In this notation, the continued fraction that we need to compute is <b>F[0][n]</b>.
Our task is first, to select a large enough <b>n</b> so that <b> F[0][n]</b> gives enough precision, and second, to compute that value.


<p>

<h3>
<hr>Method 1: bottom-up with straightforward division
</h3>
<a name="continued fraction approximation!bottom-up computation">

</a>
All "bottom-up" methods need to know the number of terms <b>n</b> in advance.
The simplest method is to start evaluating the fraction from the bottom upwards.
As the initial approximation we take <b> F[n][n]=a[n]</b>.
Then we use the obvious relation of backward recurrence,

<p><center><b>F[m][n]=a[m]+b[m]/F[m+1][n],</b></center></p>

to obtain successively <b>F[n-1][n]</b>, ..., <b>F[0][n]</b>.


<p>
This method requires one long division at each step.
There may be significant round-off error if <b>a[m]</b> and <b>b[m]</b> have opposite signs, but otherwise the round-off error is very small because a convergent continued fraction is not sensitive to small changes in its terms.


<p>

<h3>
<hr>Method 2: bottom-up with two recurrences
</h3>
An alternative implementation may be faster in some cases.
The idea is to obtain the numerator and the denominator of <b>F[0][n]</b>
separately as two simultaneous backward recurrences. If 
<b>F[m+1][n]=p[m+1]/q[m+1]</b>, then 
<b>p[m]=a[m]*p[m+1]+b[m]*q[m+1]</b> and <b>q[m]=p[m+1]</b>.
The recurrences start with
<b>p[n]=a[n]</b>, <b>q[n]=1</b>.
The method requires two long multiplications at each step; the only
division will be performed at the very end.
Sometimes this method reduces the round-off error as well.


<p>

<h3>
<hr>Method 3: bottom-up with estimated remainders
</h3>
<a name="continued fraction approximation!bottom-up computation!estimated remainder">

</a>
There is an improvement over the bottom-up method that can sometimes increase the achieved
precision without computing more terms.
This trick is suggested in [Tsimring 1988], sec. 2.4, where it is also claimed that the remainder estimates improve convergence.


<p>
The idea is that the starting value of the backward recurrence should be chosen not as
<b> a[n]</b> but as another number that more closely approximates the infinite remainder of the fraction.
The infinite remainder, which we can symbolically write as <b>F[n][Infinity]</b>, can be sometimes estimated analytically (obviously, we are unable to compute the remainder exactly).
In simple cases, <b>F[n][Infinity]</b> changes very slowly at large <b>n</b>
(warning: this is not always true and needs to be verified in each particular case!).
Suppose that <b> F[n][Infinity]</b> is approximately constant; then it must be approximately equal to <b>F[n+1][Infinity]</b>.
Therefore, if we solve the (quadratic) equation

<p><center><b>x=a[n]+b[n]/x,</b></center></p>

we shall obtain the (positive) value <b> x</b> which may be a much better
approximation for <b> F[n][Infinity]</b> than <b>a[n]</b>. But this depends on the
assumption of the way the continued fraction converges. It may happen,
for example, that for large <b>n</b> the value <b> F[n][Infinity]</b> is almost
the same as <b>F[n+2][Infinity]</b> but is significantly different from
<b>F[n+1][Infinity]</b>. Then we should instead solve the (quadratic) equation

<p><center><b>x=a[n]+b[n]/(a[n+1]+b[n+1]/x) </b></center></p>

and take the positive solution <b>x</b> as <b> F[n][Infinity]</b>.


<p>
We may use more terms of the original continued fraction starting from <b>a[n]</b> and obtain a more precise estimate of the remainder.
In each case we will only have to solve one quadratic equation.


<p>

<h3>
<hr>Method 4: top-down computation
</h3>
The "bottom-up" method obviously requires to know the number of terms <b>n</b> in
advance; calculations have to be started all over again if more terms are
needed. Also, the bottom-up method provides no error estimate.


<p>

<a name="continued fraction approximation!top-down computation">

</a>
The "top-down" method is slower but provides an automatic error estimate and can be used to
evaluate a continued fraction with more and more terms until the desired
precision is achieved.
The idea
<h6>This is a well-known result in the theory of continued fractions. We give an elementary derivation below.</h6>is to
replace the continued fraction <b> F[0][n]</b> with a sum of a certain series,

<p><center><b>a[0]+b[0]/(a[1]+b[1]/(...+b[n-1]/a[n]))=Sum(k,0,n,f[k]).</b></center></p>

Here

<p><center><b>f[k]:=F[0][k]-F[0][k-1] </b></center></p>

(<b>k&gt;=1</b>) is a sequence that will be calculated in the forward direction, starting from <b> k=1</b>.
If we manage to find a formula for this sequence, then adding one
more term <b> f[k]</b> will be equivalent to recalculating the
continued fraction with <b>k</b> terms instead of <b> k-1</b> terms. This will
automatically give an error estimate and allow to compute with more
precision if necessary without having to repeat the calculation from
the beginning. (The transformation of the continued fraction into a
series is exact, not merely an approximation.)


<p>
The formula for <b> f[k]</b> is the following.
First the auxiliary sequence <b>P[k]</b>, <b>Q[k]</b> for <b>k&gt;=1</b> needs to be
defined by <b> P[1]=0</b>, <b> Q[1]</b>=1, and <b>P[k+1]:=b[k]*Q[k]</b>,
<b>Q[k+1]:=P[k]+a[k]*Q[k]</b>.
Then define <b>f[0]:=a[0]</b> and 

<p><center><b>f[k]:=((-1)^k*b[0]*...*b[k-1])/(Q[k]*Q[k+1]) </b></center></p>

for <b>k&gt;=1</b>.
The "top-down" method consists of computing <b> f[n]</b>
sequentially and adding them together, until <b>n</b> is large enough so
that <b> f[n]/f[0]</b> is less than the required precision.


<p>
Evaluating the next element <b>f[k]</b> requires four long multiplications
and one division.
This is significantly slower, compared with just one long division or two long multiplications in the bottom-up method.
Therefore it is desirable to have an a priori estimate of the convergence rate and to be able to choose the number of terms before the computation.
Below we shall consider some examples when the formula for <b>f[k]</b> allows to estimate the required number of terms analytically.


<p>

<h3>
<hr>Method 5: top-down with two steps at once
</h3>
<a name="continued fraction approximation!top-down computation!non-alternating signs">

</a>
If all coefficients <b>a[i]</b>, <b>b[i]</b> are positive, then the
series we obtained in the top-down method will have terms <b>f[k]</b> with alternating signs.
This leads to a somewhat larger round-off error.
We can convert the alternating series into a monotonic series
by adding together two adjacent elements, say
<b>f[2*k]+f[2*k+1]</b>.
<h6>This method is used by [Thacher 1963], who refers to a suggestion by Hans Maehly.</h6>The relevant formulae can be derived from the
definition of <b>f[k]</b> using the recurrence relations for <b>P[k]</b>, <b>Q[k]</b>:

<p><center><b>f[2*k-1]+f[2*k]= -(b[0]*...*b[2*k-2]*a[2*k])/(Q[2*k-1]*Q[2*k+1]),</b></center></p>


<p><center><b>f[2*k]+f[2*k+1]=(b[0]*...*b[2*k-1]*a[2*k+1])/(Q[2*k]*Q[2*k+2]).</b></center></p>

Now in the series <b>f[0]</b>+(<b>f[1]+f[2]</b>)+(<b>f[3]+f[4]</b>)+... the first term is positive and all subsequent terms will be negative.


<p>

<h3>
<hr>Which method to use
</h3>
We have just described the following methods of computing a continued fraction:
<ul><li>Bottom-up, straight division.
</li><li>Bottom-up, separate recurrences for numerators and denominators.
</li><li>Bottom-up, with an estimate of the remainder.
</li><li>Top-down, with ordinary step.
</li><li>Top-down, with two steps at once.
</li></ul>

<p>
The bottom-up methods are simpler and faster than the top-down methods but require to know the number of terms in advance.
In many cases the required number of terms can be estimated analytically, and then the bottom-up methods are always preferable.
But in some cases the convergence analysis is very complicated.


<p>
The plain bottom-up method requires one long division at each step, while the bottom-up method with two recurrences requires two long multiplications at each step.
Since the time needed for a long division is usually about four times that for a long multiplication (e.g. when the division is computed by Newton's method), the second variation of the bottom-up method is normally faster.


<p>
The estimate of the remainder improves the convergence of the bottom-up method and should always be used if available.


<p>
If an estimate of the number of terms is not possible, the top-down methods should be used, looping until the running error estimate shows enough precision.
This incurs a performance penalty of at least 100<b><tt>%</tt></b> and at most 300<b><tt>%</tt></b>.
The top-down method with two steps at once should be used only when the formula for <b>f[k]</b> results in alternating signs.


<p>

<h3>
<hr>Usefulness of continued fractions
</h3>
Many mathematical functions have a representation as a continued fraction.
Some systems of "exact arithmetic" use continued fractions as a primary internal representation of real numbers.
This has its advantages (no round-off errors, lazy "exact" computations) and disadvantages (it is slow, especially with some operations).
Here we consider the use of continued fractions with a traditional implementation of arithmetic (floating-point numbers with variable precision).


<p>
Usually, a continued fraction representation of a function will converge geometrically or slower, i.e. at least <b>O(P)</b> terms are needed for a precision of <b>P</b> digits.
If a geometrically convergent Taylor series representation is also available, the continued fraction method will be slower because it requires at least as many or more long multiplications per term.
Also, in most cases the Taylor series can be computed much more efficiently using the rectangular scheme.
(See, e.g., the section on <b> ArcTan(x)</b> for a more detailed consideration.)


<p>
However, there are some functions for which a Taylor series is not easily computable or does not converge but a continued fraction is available.
For example, the incomplete Gamma function and related functions can be computed using continued fractions in some domains of their arguments.


<p>

<h3>
<hr>Derivation of the formula for <b>f[k]</b>
</h3>
<a name="continued fraction approximation!top-down computation!derivation">

</a>
Here is a straightforward derivation of the formula for <b>f[k]</b> in the top-down method.
We need
to compute the difference between successive approximations <b>F[0][n]</b>
and <b>F[0][n+1]</b>.
The recurrence relation we shall use is

<p><center><b>F[m][n+1]-F[m][n]= -(b[m]*(F[m+1][n+1]-F[m+1][n]))/(F[m+1][n+1]*F[m+1][n]).</b></center></p>

This can be shown by manipulating the two fractions, or by using
the recurrence relation for <b>F[m][n]</b>.


<p>
So far we have reduced the difference between <b>F[m][n+1]</b> and <b>F[m][n]</b>
to a similar difference on the next level <b>m+1</b> instead of <b> m</b>; i.e. we
can increment <b> m</b> but keep <b> n</b> fixed. We can apply this formula to
<b> F[0][n+1]-F[0][n]</b>, i.e. for <b>m=0</b>, and continue applying the same
recurrence relation until <b> m</b> reaches <b> n</b>. The result
is

<p><center><b> F[0][n+1]-F[0][n]=((-1)^n*b[0]*...*b[n])/(F[1][n+1]*...*F[n+1][n+1]*F[1][n]*...*F[n][n]).</b></center></p>



<p>
Now the problem is to simplify the two long products in the
denominator. We notice that <b>F[1][n]</b> has <b>F[2][n]</b> in the denominator,
and therefore <b>F[1][n]*F[2][n]=F[2][n]*a[1]+b[1]</b>. The next product is
<b>F[1][n]*F[2][n]*F[3][n]</b> and it simplifies to a linear function of
<b>F[3][n]</b>, namely <b>F[1][n]*F[2][n]*F[3][n]</b> =
<b>(b[1]+a[1]*a[2])*F[3][n]+b[1]*a[2]</b>. So we can see that there is a
general formula

<p><center><b>F[1][n]*...*F[k][n]=P[k]+Q[k]*F[k][n] </b></center></p>

with some coefficients <b>P[k]</b>, <b>Q[k]</b> which do not actually depend on
<b>n</b> but only on the terms of the partial fraction up to <b> k</b>. In other
words, these coefficients can be computed starting with <b> P[1]=0</b>,
<b> Q[1]=1</b> in the forward direction. The recurrence relations for <b> P[k]</b>,
<b>Q[k]</b> that we have seen above in the definition of <b>f[k]</b> follow from the identity <b>(P[k]+Q[k]*F[k][n])*F[k+1][n]</b> =
<b>P[k+1]+Q[k+1]*F[k+1][n]</b>.


<p>
Having found the coefficients <b>P[k]</b>, <b>Q[k]</b>, we can now rewrite the long products in the denominator, e.g.

<p><center><b>F[1][n]*...*F[n][n]=P[n]+Q[n]*F[n][n]=Q[n+1].</b></center></p>

(We have used the recurrence relation for <b>Q[n+1]</b>.) Now it follows that

<p><center><b>f[n+1]:=F[0][n+1]-F[0][n]=((-1)^n*b[0]*...*b[n])/(Q[n+1]*Q[n+2]).</b></center></p>

Thus we have converted the continued fraction into a series, i.e. <b>F[0][n]=Sum(k,0,n,f[k])</b> with <b>f[k]</b> defined above.


<p>

<h3>
<hr>Examples of continued fraction representations
</h3>
We have already mentioned that continued fractions give a computational advantage only when other methods are not available.
There exist continued fraction representations of almost all functions, but in most cases the usual methods (Taylor series, identity transformations, Newton's method and so on) are superior.


<p>

<a name="continued fraction approximation!of <b>ArcTan(x)</b>">

</a>
For example, the continued fraction

<p><center><b>ArcTan(x)=x/(1+x^2/(3+(2*x)^2/(5+(3*x)^2/(7+...)))) </b></center></p>

converges geometrically at all <b>x</b>.
However, the Taylor series also converges geometrically and can be computed much faster than the continued fraction.


<p>

<a name="continued fraction approximation!of <b> Erfc(x)</b>">

</a>
There are some cases when a continued fraction representation is efficient.
The complementary error function <b>Erfc(x)</b> can be computed using the
continued fraction due to Laplace (e.g. [Thacher 1963]),

<p><center><b>Sqrt(Pi)/2*x*Exp(x^2)*Erfc(x)=1/(1+v/(1+(2*v)/(1+(3*v)/(1+...)))),</b></center></p>

where <b>v:=(2*x^2)^(-1)</b>.
This continued fraction converges for all (complex) <b>x</b> except pure imaginary <b> x</b>.


<p>

<a name="continued fraction approximation!of <b> Gamma(a,z)</b>">

</a>
The error function is a particular case of the incomplete Gamma function

<p><center><b>Gamma(a,z):=(Integrate(x,z,+Infinity)x^(a-1)*Exp(-x)).</b></center></p>

There exists an analogous continued fraction representation due to Legendre (e.g. [Abramowitz <i>et al.</i>], 6.5.31),

<p><center><b>Gamma(a,z)=(z^(a-1)*Exp(-z))/(1+((1-a)*v)/(1+v/(1+((2-a)*v)/(1+(2*v)/(1+...))))),</b></center></p>

where <b>v:=z^(-1)</b>.


<p>

<a name="c4s10">

</a>
<h2>
<hr>4.10 Estimating convergence of continued fractions
</h2>
Elsewhere in this book we have used elementary considerations to find the required number of terms in a power series.
It is much more difficult
to estimate the convergence rate of a continued fraction.
In many cases this can be done using the theory of complex variable.
Below we shall consider some cases when this computation is analytically tractable.


<p>

<a name="continued fraction approximation!convergence rate">

</a>
Suppose we are given the terms <b>a[k]</b>, <b>b[k]</b> that define an infinite continued fraction, and we need to estimate its convergence rate.
We have to find the number of terms <b>n</b> for which the error of approximation is less than a given <b> epsilon</b>.
In our notation, we need to solve
<b> Abs(f[n+1])&lt;epsilon</b>
for <b> n</b>.


<p>
The formula that we derived for <b> f[n+1]</b> gives an error estimate for
the continued fraction truncated at the <b>n</b>-th term.
But this formula contains the numbers <b> Q[n]</b> in the denominator.
The main problem is to find how quickly the sequence <b>Q[n]</b> grows.
The recurrence relation for this sequence can be rewritten as

<p><center><b>Q[n+2]=a[n+1]*Q[n+1]+b[n]*Q[n],</b></center></p>

for <b>k&gt;=0</b>, with initial values <b> Q[0]=0</b> and <b> Q[1]=1</b>.
It is not always easy to get a handle on this sequence, because in most cases there is no closed-form expression for <b> Q[n]</b>.


<p>

<a name="continued fraction approximation!error bound">

</a>
<h3>
<hr>Simple bound on <b>Q[n]</b>
</h3>
A simple lower bound on the growth of <b>Q[n]</b> can be obtained from the recurrence relation for <b>Q[n]</b>.
Assume that <b>a[k]&gt;0</b>, <b> b[k]&gt;0</b>.
It is clear that all <b> Q[n]</b> are positive, so <b>Q[n+1]&gt;=a[n]*Q[n]</b>.
Therefore <b>Q[n]</b> grows at least as the product of all <b>a[n]</b>:

<p><center><b>Q[n+1]&gt;=Factorize(i,1,n,a[n]).</b></center></p>

This result gives the following upper bound on the precision,

<p><center><b>Abs(f[n+1])&lt;=(b[0]*...*b[n])/((a[1]*...*a[n])^2*a[n+1]).</b></center></p>



<p>
We have used this bound to estimate the relative error of the
continued fraction expansion for <b>ArcTan(x)</b> at small <b>x</b> (elsewhere in this book).
However, we found that at large <b> x</b> this bound becomes greater than 1.
This does not mean that the continued fraction does not converge and cannot be used to compute <b> ArcTan(x)</b> when <b>x&gt;1</b>, but merely
indicates that the "simple bound" is too weak.
The sequence <b> Q[n]</b> actually grows faster than the product of all <b>a[k]</b> and we
need a tighter bound on this growth.
In many cases such a bound can be obtained by the method of generating functions.


<p>

<h3>
<hr>The method of generating functions
</h3>
<a name="continued fraction approximation!convergence rate!from generating functions">

</a>
The idea is to find a generating function <b>G(s)</b> of the sequence <b>Q[n]</b> and then use an explicit form of <b>G(s)</b> to obtain an asymptotic estimate of <b>Q[n]</b> at large <b>k</b>.


<p>

<a name="method of steepest descent">

</a>
The asymptotic growth of the sequence <b> Q[n]</b> can be estimated by the method of steepest descent, also known as Laplace's method.
(See, e.g., [Olver 1974],
ch. 3, sec. 7.5.)
This method is somewhat complicated but quite powerful.
The method requires that we find an integral representation for <b>Q[n]</b> (usually a contour integral in the complex plane).
Then we can convert the integral into an asymptotic series in <b>k^(-1)</b>.


<p>

<a name="continued fraction approximation!of <b>Erfc(x)</b>">

</a>
Along with the general presentation of the method, we shall consider an example when the convergence rate can be
obtained analytically.
The example is the representation of the complementary error function <b>Erfc(x)</b>,

<p><center><b>Sqrt(Pi)/2*x*Exp(x^2)*Erfc(x)=1/(1+v/(1+(2*v)/(1+(3*v)/(1+...)))),</b></center></p>

where <b>v:=(2*x^2)^(-1)</b>.
We shall assume that <b>Abs(v)&lt;1/2</b> since the continued fraction representation will not be used for small <b> x</b> (where the Taylor series is efficient). 
The terms of this continued fraction are: <b> a[k]=1</b>, <b> b[k]=k*v</b>, for <b> k&gt;=1</b>, and <b> a[0]=0</b>, <b> b[0]=1</b>.


<p>
The "simple bound" would give <b> Abs(f[n+1])&lt;=v^n*n!</b> and this expression grows with <b>n</b>.
But we know that the above continued fraction actually converges for any <b> v</b>, so <b> f[n+1]</b> must tend to zero for large <b>n</b>.
It seems that the "simple bound" is not strong enough for any <b> v</b> and we need a better bound.


<p>

<a name="generating function of a sequence">

</a>
An integral representation for <b> Q[n]</b> can be obtained using the method of generating functions.
Consider a function <b>G(s)</b> defined by the infinite series

<p><center><b>G(s)=Sum(n,0,Infinity,Q[n+1]*s^n/n!).</b></center></p>

<b>G(s)</b> is usually called the "generating function" of a sequence.
We shifted the index to <b>n+1</b> for convenience, since <b> Q[0]=0</b>, so now <b> G(0)=1</b>.


<p>
Note that the above series for the function <b> G(s)</b> may or may not converge for any given <b>s</b>;
we shall manipulate <b> G(s)</b> as a formal power series until we obtain an explicit representation.
What we really need is an analytic continuation of <b>G(s)</b> to the complex <b>s</b>.


<p>

<a name="generating function of a sequence!obtaining">

</a>
It is generally the case that if we know a simple linear recurrence relation for a sequence, then we can also easily find its generating function.
The generating function will satisfy a linear differential equation.
To guess this equation, we write down the series for <b> G(s)</b> and its
derivative <b>G'(s)</b> and try to find their linear combination which is
identically zero because of the recurrence relation.
(There is, of course, a computer algebra algorithm for doing this automatically.)


<p>
Taking the derivative <b>G'(s)</b> produces the forward-shifted series

<p><center><b>G'(s)=Sum(n,0,Infinity,Q[n+2]*s^n/n!).</b></center></p>

Multiplying <b>G(s)</b> by <b>s</b> produces a back-shifted series with each term multiplied by <b> n</b>:

<p><center><b> s*G(s)=Sum(n,0,Infinity,n*Q[n]*s^n/n!).</b></center></p>

If the recurrence relation for <b>Q[n]</b> contains only constants and polynomials in <b>n</b>, then we can easily convert that relation into a differential equation for <b> G(s)</b>.
We only need to find the right combination of back- and forward-shifts and multiplications by <b>n</b>.


<p>
In the case of our sequence <b> Q[n]</b> above, the recurrence relation is

<p><center><b>Q[n+2]=Q[n+1]+v*n*Q[n].</b></center></p>

This is equivalent to the differential equation

<p><center><b>G'(s)=(1+v*s)*G(s).</b></center></p>

The solution with the obvious initial condition <b>G(0)=1</b> is

<p><center><b> G(s)=Exp(s+(v*s^2)/2).</b></center></p>



<p>

<a name="generating function of a sequence!integral representation">

</a>
The second step is to obtain an integral representation for <b>Q[n]</b>, so that we could use the method of steepest descents and find its asymptotic at large <b>n</b>.


<p>
In our notation <b> Q[n+1]</b> is equal to the <b>n</b>-th derivative of the generating function at <b> s=0</b>:

<p><center><b> Q[n+1]=(D(s,n)G(s=0)),</b></center></p>

but it is generally not easy to estimate the growth of this derivative at large <b>n</b>.


<p>
There are two ways to proceed.
One is to obtain an integral representation for <b> G(s)</b>, for instance

<p><center><b>G(s)=(Integrate(t,-Infinity,Infinity)F(t,s)),</b></center></p>

where <b>F(t,s)</b> is some known function.
Then an integral representation for <b>Q[n]</b> will be found by differentiation.
But it may be difficult to find such <b>F(t,s)</b>.


<p>
The second possibility is to express <b>Q[n]</b> as a contour integral in the complex plane around <b>s=0</b> in the counter-clockwise direction:

<p><center><b> Q[n]=(n-1)! /(2*Pi*I)*(Integrate(s)G(s)*s^(-n)).</b></center></p>

If we know the singularities and of <b>G(s)</b>, we may transform the contour of the integration into a path that is more convenient for the method of the steepest descent.
This procedure is more general but requires a detailed understanding of the behavior of <b>G(s)</b> in the complex plane.


<p>

<a name="Stirling's formula">

</a>
<a name="continued fraction approximation!of <b>Erfc(x)</b>!error estimate">

</a>
In the particular case of the continued fraction for <b>Erfc(x)</b>, the calculations are somewhat easier if <b>Re(v)&gt;0</b> (where <b> v:=1/(2*x^2)</b>).
Full details are given in a separate section.
The result for <b>Re(v)&gt;0</b> is

<p><center><b> Q[n]&lt;=&gt;(v*n)^(n/2)/Sqrt(2*n*v)*Exp(Sqrt(n/v)-1/(4*v)-n/2).</b></center></p>

This, together with Stirling's formula

<p><center><b>n! &lt;=&gt;Sqrt(2*Pi*n)*(n/e)^n,</b></center></p>

allows to estimate the error of the continued fraction approximation:

<p><center><b> f[n+1]&lt;=&gt;2*(-1)^(n+1)*Sqrt((2*Pi)/v)*Exp(-2*Sqrt(n/v)+1/(2*v)).</b></center></p>



<p>
Note that this is not merely a bound but an actual asymptotic estimate of <b>f[n+1]</b>.
(Stirling's formula can also be derived using the method of
steepest descent from an integral representation of the Gamma function,
in a similar way.)


<p>
Defined as above, the value of <b>f[n+1]</b> is in general a complex number. The absolute value of <b>f[n+1]</b> can be found using the formula

<p><center><b>Re(Sqrt(n/v))=Sqrt(n/2)*Sqrt(1+Re(v)/Abs(v)).</b></center></p>

We obtain

<p><center><b>Abs(f[n+1])&lt;=&gt;2*Sqrt((2*Pi)/Abs(v))*Exp(-Sqrt(2*n)*Sqrt(1+Re(v)/Abs(v))+Re(v)/(2*Abs(v)^2)).</b></center></p>
 


<p>
When <b>Re(v)&lt;=0</b>, the same formula can be used (this can be shown by a more careful consideration of the branches of the square roots).
The continued fraction does not converge when <b> Re(v)&lt;0</b> and <b> Im(v)=0</b> (i.e. for pure imaginary <b> x</b>).
This can be seen from the above formula: in that case <b> Re(v)= -Abs(v)</b> and <b>Abs(f[n+1])</b> does not decrease when <b>n</b> grows.


<p>
These estimates show that the error of the continued fraction approximation to <b> Erfc(x)</b> (when it converges)
decreases with <b>n</b> slower than in a geometric progression.
This means that we need to take <b> O(P^2)</b> terms to get <b>P</b> digits of precision.


<p>

<h3>
<hr>Derivations for the example with <b> Erfc(x)</b>
</h3>
Here we give a detailed calculation of the convergence rate of the continued fraction for <b>Erfc(x)</b> using the method of generating functions.


<p>

<h5>
A simpler approach
</h5>
<a name="method of steepest descent!example for real <b>x</b>">

</a>
In our case, <b> G(s)</b> is a function with a known Fourier transform and we can obtain a straightforward representation valid when <b>Re(v)&gt;0</b>,

<p><center><b> Q[n+1]=1/Sqrt(2*Pi*v)*(Integrate(t,-Infinity,Infinity)(1+t)^n*Exp(-t^2/(2*v))).</b></center></p>

We shall first apply the method of steepest descent to this integral (assuming real <b>v&gt;0</b> for simplicity) and then consider the more general procedure with the contour integration.


<p>
To use the method of steepest descent, we represent the integrand as an exponential of some function <b> g(t,n)</b> and find "stationary points" where this function has local maxima:

<p><center><b>Q[n+1]=1/Sqrt(2*Pi*v)*(Integrate(t,-Infinity,Infinity)Exp(g(t,n))),</b></center></p>


<p><center><b>g(t,n):= -t^2/(2*v)+n*Ln(1+t).</b></center></p>

(Note that the logarithm here must acquire an imaginary part <b>I*Pi</b> for <b> t&lt;-1</b>,
and we should take the real part which is equal to <b> Ln(Abs(1+t))</b>.
We shall see that the integral over negative <b>t</b> is negligible.)
We expect that when <b> n</b> is large, <b> Re(g(t,n))</b> will have a peak or several peaks at certain values of <b>t</b>.
When <b> t</b> is not close to the peaks, the value of <b> Re(g(t,n))</b> is smaller and,
since <b>g</b> is in the exponential, the integral is dominated by the
contribution of the peaks.
This is the essence of the method of steepest descent on the real line.


<p>
We only need to consider very large values of <b> n</b>, so we can neglect terms of order <b> O(1/Sqrt(n))</b> or smaller.
We find that, in our case, two peaks of <b>Re(g)</b> occur at approximately <b>t1&lt;=&gt; -1/2+Sqrt(n*v)</b> and <b>t2&lt;=&gt; -1/2-Sqrt(n*v)</b>.
We assume that <b>n</b> is large enough so that <b> n*v&gt;1/2</b>. Then the first peak is at a positive <b> t</b> and the second peak is at a negative <b> t</b>.
The contribution of the peaks can be computed from the Taylor approximation of <b> g(t,n)</b> near the peaks.
We can expand, for example,

<p><center><b>g(t,n)&lt;=&gt;g(t1,n)+(Deriv(t,2)g(t1,n))*(t-t1)^2/2 </b></center></p>

near <b> t=t1</b>.
The values <b> g(t1,n)</b> and <b>Deriv(t,2)g(t1,n)</b>, and likewise for <b>t2</b>, are constants that we already know since we know <b> t1</b> and <b> t2</b>.
Then the integral of <b> Exp(g)</b> will be approximated by the integral

<p><center><b>Integrate(t,-Infinity,Infinity)Exp(g(t1,n)+(Deriv(t,2)g(t1,n))*(t-t1)^2/2).</b></center></p>

(Note that <b>Deriv(t,2)g(t1,n)</b> is negative.)
This is a Gaussian integral that can be easily evaluated, with the result

<p><center><b>exp(g(t1,n))*Sqrt(-(2*Pi)/(Deriv(t,2)g(t1,n))).</b></center></p>

This is the leading term of the contribution of the peak at <b>t1</b>;
there will be a similar expression for the contribution of <b> t2</b>.
We find that the peak at <b> t1</b> gives a larger contribution, by the factor <b> Exp(2*Sqrt(n/v))</b>.
This factor is never small since <b>n&gt;1</b> and <b> v&lt;1/2</b>.
So it is safe to ignore the peak at <b> t2</b> for the purposes of our analysis.


<p>
Then we obtain the estimate

<p><center><b> Q[n+1]&lt;=&gt;1/Sqrt(2)*Exp(Sqrt(n/v)-1/(4*v)-n/2)*(v*n)^(n/2).</b></center></p>



<p>

<h5>
The contour integral approach
</h5>
In many cases it is impossible to compute the Fourier transform of the generating function <b>G(s)</b>.
Then one can use the contour integral approach.
One should represent the integrand as

<p><center><b>G(s)*s^(-n)=Exp(g(s)) </b></center></p>

where

<p><center><b>g(s):=Ln(G(s))-n*Ln(s),</b></center></p>

and look for stationary points of <b>g(s)</b> in the complex plane (<b>(D(s)g)=0</b>).
The original contour is around the pole <b> s=0</b> in the counter-clockwise direction.
We need to deform that contour so that the new contour passes through the stationary points.
The contour should cross each stationary point in a certain direction in the complex plane.
The direction is chosen to make the stationary point the sharpest local maximum of <b> Re(g(s))</b> on the contour.


<p>
Usually one of the stationary points has the largest value of <b>Re(g(s))</b>; this is the dominant stationary point.
If <b>s0</b> is the dominant stationary point and <b> g2=(Deriv(s,2)g(s0))</b> is the second derivative of <b>g</b> at that point, then the asymptotic of the integral is

<p><center><b> 1/(2*Pi)*(Integrate(s)Exp(g(s)))=1/Sqrt(2*Pi*Abs(g2))*Exp(g(s0)).</b></center></p>

(The integral will have a negative sign if the contour crosses the point <b>s0</b> in the negative imaginary direction.)


<p>
We have to choose a new contour and check the convergence of the resulting integral separately.
In each case we may need to isolate the singularities of <b> G(s)</b> or to find the regions of infinity where <b>G(s)</b> quickly decays (so that the infinite parts of the contour might be moved there).
There is no prescription that works for all functions <b>G(s)</b>.


<p>
Let us return to our example.
For
<b>G(s)=Exp(s+(v*s^2)/2)</b>,
the function <b>g(s)</b> has no singularities except the pole at <b>s=0</b>.
There are two stationary points located at the (complex) roots <b> s1</b>, <b> s2</b> of the quadratic equation
<b> v*s^2+s-n=0</b>.
Note that <b> v</b> is an arbitrary (nonzero) complex number.
We now need to find which of the two stationary points gives the dominant contribution.
By comparing <b> Re(g(s1))</b> and <b>Re(g(s2))</b> we find
that the point <b>s</b> with the largest real part gives the dominant contribution.
However, if <b> Re(s1)=Re(s2)</b> (this happens only if <b>v</b> is real and <b> v&lt;0</b>, i.e. if <b> x</b> is pure imaginary), then both stationary points contribute equally.
Barring that possibility, we find
(with the usual definition of the complex square root) that
the dominant contribution for large <b> n</b> is from the stationary point at

<p><center><b> s1=(Sqrt(4*n*v+1)-1)/(2*v).</b></center></p>

The second derivative of <b>g(s)</b> at the stationary point is approximately <b>2*v</b>.
The contour of integration can be deformed into a line passing through the dominant stationary point in the positive imaginary direction.
Then the leading asymptotic is found using the Gaussian approximation (assuming <b> Re(v)&gt;0</b>):

<p><center><b> Q[n]=((n-1)! *v^(n/2))/Sqrt(4*Pi*v)*Exp((n*(1-Ln(n)))/2+Sqrt(n/v)-1/(4*v)).</b></center></p>



<p>

<a name="Stirling's formula">

</a>
This formula agrees with the asymptotic for <b>Q[n+1]</b> found above for real <b>v&gt;0</b>, when we use Stirling's formula for <b>(n-1)!</b>:

<p><center><b>(n-1)! =Sqrt(2*Pi)*e^(-n)*n^(n-1/2).</b></center></p>



<p>
The treatment for <b>Re(v)&lt;0</b> is similar.


<p>

<a name="c4s11">

</a>
<h2>
<hr>4.11 Newton's method and its improvements
</h2>
Newton's method (also called the Newton-Raphson method) of numerical solution of algebraic equations and its generalizations
can be used to obtain multiple-precision values of several elementary functions.


<p>

<h3>
<hr>Newton's method
</h3>
<a name="Newton's method">

</a>
The basic formula is widely known: If <b>f(x)=0</b> must be solved, one starts with a value of <b> x</b> that is close to some root and iterates 
<p><center><b> x'=x-f(x)*(D(x)f(x))^(-1).</b></center></p>

This formula is based on the approximation of the function <b>f(x)</b> by a tangent line at some point <b>x</b>. A Taylor expansion in the neighborhood of the root shows that (for an initial value <b> x[0]</b> sufficiently close to the root) each iteration gives at least twice as many correct digits of the root as the previous one ("quadratic convergence"). Therefore the complexity of this algorithm is proportional to a logarithm of the required precision and to the time it takes to evaluate the function and its derivative. Generalizations of this method require computation of higher derivatives of the function <b>f(x)</b> but successive approximations to the root converge several times faster (the complexity is still logarithmic).


<p>

<a name="Newton's method!initial value">

</a>
Newton's method sometimes suffers from a sensitivity to the initial guess.
If the initial value <b>x[0]</b> is not chosen sufficiently close to the root, the iterations may converge very slowly or not converge at all. To remedy this, one can combine Newton's iteration with simple bisection. Once the root is bracketed inside an interval (<b>a</b>, <b> b</b>), one checks whether <b>(a+b)/2</b> is a better approximation for the root than that obtained from Newton's iteration. This guarantees at least linear convergence in the worst case.


<p>

<a name="Newton's method!cubic convergence">

</a>
For some equations <b> f(x)=0</b>, Newton's method converges faster than quadratically.
For example, solving <b> Sin(x)=0</b> in the neighborhood of <b> x=3.14159</b> gives "cubic" convergence, i.e. the number of correct digits is tripled at each step.
This happens because <b> Sin(x)</b> near its root <b>x=Pi</b> has a vanishing second derivative and thus the function is particularly well approximated by a straight line.


<p>

<h3>
<hr>Halley's method
</h3>
<a name="Halley's method">

</a>
<i>Halley's method</i> is an improvement over Newton's method that makes each equation well approximated by a straight line near the root.
Edmund Halley computed fractional powers, <b> x=a^(1/n)</b>, by the iteration

<p><center><b>x'=x*(n*(a+x^n)+a-x^n)/(n*(a+x^n)-(a-x^n)).</b></center></p>

This formula is equivalent to Newton's method applied to the equation
<b>x^(n-q)=a*x^(-q)</b> with <b>q=(n-1)/2</b>. This iteration has a cubic convergence rate. This is the fastest method to compute <b> n</b>-th roots (<b> n&gt;=3</b>) with multiple precision. Iterations with higher order of convergence, for example, the method with quintic convergence rate

<p><center><b> x'=x*((n-1)/(n+1)*(2*n-1)/(2*n+1)*x^(2*n)+2*(2*n-1)/(n+1)*x^n*a+a^2)/(x^(2*n)+2*(2*n-1)/(n+1)*x^n*a+(n-1)/(n+1)*(2*n-1)/(2*n+1)*a^2),</b></center></p>

require more arithmetic operations per step and are in fact less efficient at high precision.


<p>
Halley's method can be generalized to any function <b>f(x)</b>. A cubically convergent iteration is always obtained if we replace the equation <b>f(x)=0</b> by an equivalent equation

<p><center><b> g(x):=f(x)/Sqrt(Abs(D(x)f(x)))=0 </b></center></p>

and use the standard Newton's method on it.
Here the function <b> g(x)</b> is chosen so that its second derivative vanishes (<b>(Deriv(x,2)g(x))=0</b>) at the root of the equation <b> f(x)=0</b>, independently of where this root is.
For example, the equation <b> Exp(x)=a</b> is transformed into <b> g(x):=Exp(x/2)-a*Exp(-x/2)=0</b>.
(There is no unique choice of the function <b> g(x)</b> and sometimes another choice will make the iteration more quickly computable.)


<p>

<a name="Halley's method!explicit formula">

</a>
The Halley iteration for the equation <b>f(x)=0</b> can be written as

<p><center><b> x'=x-(2*f(x)*(D(x)f(x)))/(2*(D(x)f(x))^2-f(x)*(Deriv(x,2)f(x))).</b></center></p>



<p>
Halley's iteration, despite its faster convergence rate, may be more cumbersome to evaluate than Newton's iteration and so it may not provide a more efficient numerical method for a given function.
Only in some special cases is Halley's iteration just as simple to compute as Newton's iteration.


<p>
Halley's method is sometimes less sensitive to the choice of the initial point <b>x[0]</b>.
An extreme example of sensitivity to the initial point is the equation <b>x^(-2)=12</b> for which Newton's iteration <b> x'=3/2*x-6*x^3</b> converges to the root only from initial points <b> 0&lt;x[0]&lt;0.5</b> and wildly diverges otherwise, while Halley's iteration converges to the root from any <b> x[0]&gt;0</b>.


<p>
It is at any rate not true that Halley's method always converges better than Newton's method. For instance, it diverges on the equation <b> 2*Cos(x)=x</b> unless started at <b> x[0]</b> within the interval (<b>-1/6*Pi</b>,<b> 7/6*Pi</b>). Another example is the equation <b> Ln(x)=a</b>. This equation allows to compute <b> x=Exp(a)</b> if a fast method for computing <b>Ln(x)</b> is available (e.g. the AGM-based method). For this equation, Newton's iteration

<p><center><b>x'=x*(1+a-Ln(x)) </b></center></p>

converges for any <b>0&lt;x&lt;Exp(a+1)</b>, while Halley's iteration
converges only if <b>Exp(a-2)&lt;x&lt;Exp(a+2)</b>.


<p>

<a name="Halley's method!when to use">

</a>
When it converges, Halley's iteration can still converge very slowly for certain functions <b>f(x)</b>, for example, for <b>f(x)=x^n-a</b> if <b> n^n&gt;a</b>. For such functions that have very large and rapidly changing derivatives, no general method can converge faster than linearly. In other words, a simple bisection will generally do just as well as any sophisticated iteration, until the root is approximated very precisely.
Halley's iteration combined with bisection seems to be a good choice for such problems.


<p>

<h3>
<hr>When to use Halley's method
</h3>
Despite its faster convergence, Halley's iteration frequently gives no advantage over Newton's method.
To obtain <b> P</b> digits of the result, we need about <b> Ln(P)/Ln(2)</b> iterations of a quadratically convergent method and about <b>Ln(P)/Ln(3)</b> iterations of a cubically convergent method.
So the cubically convergent iteration is faster only if the time taken by cubic one iteration is less than about <b>Ln(3)/Ln(2)&lt;=&gt;1.6</b> of the time of one quadratic iteration.


<p>

<h3>
<hr>Higher-order schemes
</h3>
<a name="Newton's method!higher-order schemes">

</a>
Sometimes it is easy to generalize Newton's iteration to higher-order schemes.
There are general formulae such as Shroeder's and Householder's iterations.
We shall give some examples where the construction is very straightforward.
In all examples <b> x</b> is the initial approximation and the next approximation is obtained by truncating the given series.


<p>
<ul><li>Inverse </li><b> 1/a</b>.
Set <b> y=1-a*x</b>, then

<p><center><b> 1/a=x/(1-y)=x*(1+y+y^2+...).</b></center></p>

<li>Square root </li><b>Sqrt(a)</b>.
Set <b>y=1-a*x^2</b>, then

<p><center><b> Sqrt(a)=Sqrt(1-y)/x=1/x*(1-1/2*y-1/8*y^2-...).</b></center></p>

<li>Inverse square root </li><b>1/Sqrt(a)</b>.
Set <b>y=1-a*x^2</b>, then

<p><center><b> 1/Sqrt(a)=x/Sqrt(1-y)=x*(1+1/2*y+3/8*y^2+...).</b></center></p>

<li></li><b>n</b>-th root <b> a^(1/n)</b>.
Set <b>y=1-a*x^n</b>, then

<p><center><b> a^(1/n)=(1-y)^(1/n)/x=1/x*(1-1/n*y-(n-1)/(2*n^2)*y^2-...).</b></center></p>

<li>Exponential </li><b>Exp(a)</b>.
Set <b>y=a-Ln(x)</b>, then

<p><center><b>Exp(a)=x*Exp(y)=x*(1+y+y^2/2! +y^3/3! +...).</b></center></p>

<li>Logarithm </li><b>Ln(a)</b>.
Set <b>y=1-a*Exp(-x)</b>, then

<p><center><b>Ln(a)=x+Ln(1-y)=x-y-y^2/2-y^3/3-....</b></center></p>

</ul>

<p>
In the above examples, <b>y</b> is a small quantity and the series represents corrections to the initial value <b> x</b>, therefore the order of convergence is equal to the first discarded order of <b> y</b> in the series.


<p>
These simple constructions are possible because the functions satisfy simple identities, such as <b> Exp(a+b)=Exp(a)*Exp(b)</b> or <b>Sqrt(a*b)=Sqrt(a)*Sqrt(b)</b>.
For other functions the formulae quickly become very complicated and unsuitable for practical computations.


<p>

<h3>
<hr>Precision control
</h3>
<a name="Newton's method!precision control">

</a>
Newton's method and its generalizations are particularly convenient for multiple precision calculations because of their insensitivity to accumulated errors:
if at some iteration <b>x[k]</b> is found with a small error, the error will be corrected at the next iteration.
Therefore it is not necessary to compute all iterations with the full required precision; each iteration needs to be performed at the precision of the root expected from that iteration (plus a few more digits to allow for round-off error).
For example, if we know that the initial approximation is accurate to 3 digits, then (assuming quadratic convergence)
<h6>This disregards the possibility that the convergence might be slightly slower. For example, when the precision at one iteration is <b>n</b> digits, it might be <b> 2*n-10</b> digits at the next iteration. In these (fringe) cases, the initial approximation must be already precise enough (e.g. to at least 10 digits in this example).</h6>it is enough to perform the first iteration to 6 digits, the second iteration to 12 digits and so on. In this way, multiple-precision calculations are enormously speeded up.


<p>
For practical evaluation, iterations must be supplemented with "quality control".
For example, if <b> x0</b> and <b> x1</b> are two consecutive approximations that are
already very close, we can quickly compute the achieved (relative) precision by
finding the number of leading zeros in the number 
<p><center><b> Abs(x0-x1)/Max(x0,x1).</b></center></p>

This
is easily done using the bit count function. After performing a small number of
initial iterations at low precision, we can make sure that <b>x1</b> has at least a
certain number of correct digits of the root. Then we know which precision to
use for the next iteration (e.g. triple precision if we are using a cubically
convergent scheme). It is important to perform each iteration at the precision
of the root which it will give and not at a higher precision; this saves a
great deal of time since all calculations are very slow at high precision.


<p>

<h3>
<hr>Fine-tuning the working precision
</h3>
To reduce the computation time, it is important to write the iteration formula with explicit separation of higher-order quantities.
For example, Newton's iteration for the inverse square root <b> 1/Sqrt(a)</b> can be written either as

<p><center><b>x'=x*(3-a*x^2)/2 </b></center></p>

or equivalently as

<p><center><b> x'=x+x*(1-a*x^2)/2.</b></center></p>

At first sight the first formula seems simpler because it saves one long addition.
However, the second formula can be computed significantly faster than the first one, if we are willing to exercise a somewhat more fine-grained control of the working precision.


<p>
Suppose <b> x</b> is an approximation that is correct to <b> P</b> digits; then we expect the quantity <b> x'</b> to be correct to <b> 2*P</b> digits.
Therefore we should perform calculations in the first formula with <b> 2*P</b> digits;
this means three long multiplications, <b> 3*M(2*P)</b>.
Now consider the calculation in the second formula.
First, the quantity <b>y:=1-a*x^2</b> is computed using two <b> 2*P</b>-digit multiplications.
<h6>In fact, both multiplications are a little shorter, since <b> x</b> is a number with only <b> P</b> correct digits; we can compute <b> a*x</b> and then <b> a*x^2</b> as products of a <b> 2*P</b>-digit number and a <b> P</b>-digit number, with a <b> 2*P</b>-digit result. We ignore this small difference.</h6>Now, the number <b> y</b> is small and has only <b> P</b> nonzero digits.
Therefore the third multiplication <b> x*y</b> costs only <b> M(P)</b>, not <b>M(2*P)</b>.
This is a significant time savings, especially with slower multiplication.
The total cost is now <b>2*M(2*P)+M(P)</b>.


<p>
The advantage is even greater with higher-order methods.
For example, a fourth-order iteration for the inverse square root can be written as

<p><center><b>x'=x+1/2*x*y+3/8*x*y^2+5/16*x*y^3,</b></center></p>

where <b> y:=1-a*x^2</b>.
Suppose <b> x</b> is an approximation that is correct to <b> P</b> digits; we expect <b> 4*P</b> correct digits in <b> x'</b>.
We need two long multiplications in precision <b> 4*P</b> to compute <b> y</b>, then <b> M(3*P)</b> to compute <b>x*y</b>, <b> M(2*P)</b> to compute <b>x*y^2</b>, and <b> M(P)</b> to compute <b>x*y^3</b>.
The total cost is <b> 2*M(4*P)+M(3*P)+M(2*P)+M(P)</b>.


<p>

<a name="Newton's method!asymptotic cost">

</a>
The asymptotic cost of finding the root <b>x</b> of the equation <b> f(x)=0</b> with <b> P</b> digits of precision is usually the same as the cost of computing <b> f(x)</b> [Brent 1975].
The main argument can be summarized by the following simple example.
To get the result to <b>P</b> digits, we need <b> O(Ln(P))</b> Newton's iterations.
At each iteration we shall have to compute the function <b>f(x)</b> to a certain number of digits.
Suppose that we start with one correct digit and that each iteration costs us <b>c*M(2*P)</b> operations where <b>c</b> is a given constant, while the number of correct digits grows from <b> P</b> to <b> 2*P</b>.
Then the total cost of <b> k</b> iterations is

<p><center><b> c*M(2)+c*M(4)+c*M(8)+...+c*M(2^k).</b></center></p>

If the function <b>M(P)</b> grows linearly with <b>P=2^k</b>, then we can estimate this sum roughly as <b> 2*c*M(P)</b>; if <b>M(P)=O(P^2)</b> then the result is about <b>4/3*c*M(P)</b>.
It is easy to see that when <b>M(P)</b> is some power of <b>P</b> that grows faster than linear, the sum is not larger than a small multiple of <b> M(P)</b>.


<p>
Thus, if we have a fast method of computing, say, <b>ArcTan(x)</b>, then we immediately obtain a method of computing <b>Tan(x)</b> which is asymptotically as fast (up to a constant).


<p>

<h3>
<hr>Choosing the optimal order
</h3>
<a name="Newton's method!optimal order">

</a>
Suppose we need to compute a function by Newton's method to precision <b>P</b>.
We can sometimes find iterations of any order of convergence.
For example, a <b> k</b>-th order iteration for the reciprocal <b> 1/a</b> is

<p><center><b> x'=x+x*y+x*y^2+...+x*y^(k-1),</b></center></p>

where <b>y:=1-a*x</b>.
The cost of one iteration with final precision <b> P</b> is

<p><center><b> C(k,P):=M(P/k)+M((2*P)/k)+M((3*P)/k)+...+c*M(P).</b></center></p>

(Here the constant <b>c:=1</b> is introduced for later convenience.
It denotes the number of multiplications needed to compute <b> y</b>.)


<p>
Increasing the order by <b> 1</b> costs us comparatively little, and
we may change the order <b> k</b> at any time.
Is there a particular order <b> k</b> that gives the smallest computational cost and should be used for all iterations, or the order needs to be adjusted during the computation?
A natural question is to find the optimal computational strategy.


<p>
It is difficult to fully analyze this question, but it seems that choosing a particular order <b> k</b> for all iterations is close to the optimal strategy.


<p>
A general "strategy" is a set of orders <b> S(P,P[0])</b>=(<b>k[1]</b>, <b>k[2]</b>, ..., <b>k[n]</b>) to be chosen at the first, second, ..., <b>n</b>-th iteration, given the initial precision <b> P[0]</b> and the required final precision <b>P</b>.
At each iteration, the precision will be multiplied by the factor <b> k[i]</b>.
The optimal strategy <b>S(P,P[0])</b> is a certain function of <b>P[0]</b> and <b>P</b> such that the required precision is reached, i.e.

<p><center><b> P[0]*k[1]*...*k[n]=P,</b></center></p>

and the cost

<p><center><b> C(k[1],P[0]*k[1])+C(k[2],P[0]*k[1]*k[2])+...+C(k[n],P) </b></center></p>

is minimized.


<p>
If we assume that the cost of multiplication <b>M(P)</b> is proportional to some power of <b>P</b>, for instance <b> M(P)=P^mu</b>, then
the cost of each iteration and the total cost are homogeneous functions of <b> P</b> and <b> P[0]</b>.
Therefore the optimal strategy is a function only of the ratio <b>P/P[0]</b>.
We can multiply both <b>P[0]</b> and <b>P</b> by a constant factor and the optimal strategy will remain the same.
We can denote the optimal strategy <b> S(P/P[0])</b>.


<p>
We can check whether it is better to use several iterations at smaller orders instead of one iteration at a large order.
Suppose that  <b>M(P)=P^mu</b>, the initial precision is 1 digit, and the final precision <b> P=k^n</b>.
We can use either <b> n</b> iterations of the order <b> k</b> or 1 iteration of the order <b> P</b>.
The cost of one iteration of order <b> P</b> at target precision <b> P</b> is
<b> C(P,P)</b>, whereas the total cost of <b>n</b> iterations of order <b> k</b> is

<p><center><b> C(k,k)+C(k,k^2)+...+C(k,k^n).</b></center></p>

With <b>C(k,P)</b> defined above, we can take approximately

<p><center><b>C(k,p)&lt;=&gt;p^mu*(c-1+k/(mu+1)).</b></center></p>

Then the cost of one <b>P</b>-th order iteration is

<p><center><b> P^mu*(c-1+P/(mu+1)),</b></center></p>

while the cost of <b>n</b> iterations of the order <b> k</b> is clearly smaller since <b> k&lt;P</b>,

<p><center><b> P^mu*(c-1+k/(mu+1))*k^mu/(k^mu-1).</b></center></p>

At fixed <b>P</b>, the best value of <b> k</b> is found by minimizing this function.
For <b> c=1</b> (reciprocal) we find <b> k=(1+mu)^(1/mu)</b> which is never above <b>2</b>.
This suggests that <b> k=2</b> is the best value for finding the reciprocal <b> 1/a</b>.
However, larger values of <b> c</b> can give larger values of <b> k</b>.
The  equation for the optimal value of <b> k</b> is

<p><center><b> k^(mu+1)/(mu+1)-k=mu*(c-1).</b></center></p>



<p>
So far we have only considered strategies that use the same order <b>k</b> for all iterations, and we have not yet shown that such strategies are the best ones.
We now give a plausible argument (not quite a rigorous proof) to justify this claim.


<p>
Consider the optimal strategy <b> S(P^2)</b> for the initial precision <b>1</b> and the final precision <b> P^2</b>, when <b> P</b> is very large.
Since it is better to use several iterations at lower orders, we may assume that the strategy <b> S(P^2)</b> contains many iterations and that one of these iterations reaches precision <b>P</b>.
Then the strategy <b> S(P^2)</b> is equivalent to a sequence of the two optimal strategies to go from <b>1</b> to <b> P</b> and from <b> P</b> to <b> P^2</b>.
However, both strategies must be the same because the optimal strategy only depends on the ratio of precisions.
Therefore, the optimal strategy <b> S(P^2)</b> is a sequence of two identical strategies (<b>S(P)</b>, <b>S(P)</b>).


<p>
Suppose that <b>k[1]</b> is the first element of <b>S(P)</b>.
The optimal strategy to go from precision <b>k[1]</b> to precision <b>P*k[1]</b> is also <b>S(P)</b>.
Therefore the second element of <b>S(P)</b> is also equal to <b>k[1]</b>, and by extension
all elements of <b>S(P)</b> are the same.


<p>
A similar consideration gives the optimal strategy for other iterations that compute inverses of analytic functions, such as Newton's iteration for the inverse square root or for higher roots.
The difference is that the value of <b>c</b> should be chosen as the equivalent number of multiplications needed to compute the function.
For instance, <b> c=1</b> for division and <b> c=2</b> for the inverse square root iteration.


<p>
The conclusion is that in each case we should compute the optimal order <b> k</b> in advance and use this order for all iterations.


<p>

<a name="c4s12">

</a>
<h2>
<hr>4.12 Fast evaluation of Taylor series
</h2>
<a name="Taylor series">

</a>
Taylor series for analytic functions  is a common method of evaluation.


<p>

<h3>
<hr>Method 1: simple summation
</h3>
If we do not know the required number of terms in advance, we cannot do any better than just evaluate each term and check if it is already small enough.
Take, for example, the series for <b>Exp(x)</b>.
To straightforwardly evaluate

<p><center><b>Exp(x)&lt;=&gt;Sum(k,0,N-1,x^k/k!) </b></center></p>

with <b>P</b> decimal digits of precision and <b> x&lt;2</b>, one would need about <b> N&lt;=&gt;P*Ln(10)/Ln(P)</b> terms of the series.


<p>
Divisions by large integers <b>k!</b> and separate evaluations of powers <b>x^k</b> are avoided if we store the previous term.
The next term can be obtained  by a short division of the previous term by <b> k</b> and a long multiplication by <b> x</b>.
Then we only need <b> O(N)</b> long multiplications to evaluate the series.
Usually the required number of terms <b>N=O(P)</b>, so the total cost is <b>O(P*M(P))</b>.


<p>
There is no accumulation of round-off error in this method if <b>x</b> is small enough (in the case of <b> Exp(x)</b>, a sufficient condition is <b>Abs(x)&lt;1/2</b>).
To see this, suppose that <b> x</b> is known to <b> P</b> digits (with relative error <b> 10^(-P)</b>).
Since <b>Abs(x)&lt;1/2</b>, the <b> n</b>-th term <b> x^n/n! &lt;4^(-n)</b> (this is a rough estimate but it is enough).
Since each multiplication by <b>x</b> results in adding 1 significant bit of relative round-off error, the relative error of <b> x^n/n!</b> is about <b>2^n</b> times the relative error of <b> x</b>, i.e. <b> 2^n*10^(-P)</b>.
So the absolute round-off error of <b>x^n/n!</b> is not larger than

<p><center><b>Delta&lt;4^(-n)*2^n*10^(-P)=2^(-n)*10^(-P).</b></center></p>

Therefore all terms with <b>n&gt;1</b> contribute less than <b> 10^(-P)</b> of absolute round-off error, i.e. less than was originally present in <b>x</b>.


<p>
In practice, one could truncate the precision of <b> x^n/n!</b> as <b>n</b> grows, leaving a few guard bits each time to keep the round-off error negligibly small and yet to gain some computation speed.
This however does not change the asymptotic complexity of the method---it remains <b> O(P*M(P))</b>.


<p>
However, if <b>x</b> is a small rational number, then the multiplication by <b> x</b> is short and takes <b> O(P)</b> operations.
In that case, the total complexity of the method is <b>O(P^2)</b> which is always faster than <b>O(P*M(P))</b>.


<p>

<h3>
<hr>Method 2: Horner's scheme
</h3>
<a name="Taylor series!by Horner's scheme">

</a>
<a name="Horner's scheme">

</a>
Horner's scheme is widely known and consists of the following rearrangement,

<p><center><b>Sum(k,0,N-1,a[k]*x^k)=a[0]+x*(a[1]+x*(a[2]+x*(...+x*a[N-1]))) </b></center></p>

The calculation is started from the last coefficient <b>a[N-1]</b> toward the first.
If <b>x</b> is small, then the round-off error generated during the summation is constantly being multiplied by a small number <b> x</b> and thus is always insignificant.
Even if <b> x</b> is not small or if the coefficients of the series are not small, Horner's scheme usually results in a smaller round-off error than the simple summation.


<p>
If the coefficients <b> a[k]</b> are related by a simple ratio, then Horner's scheme may be modified to simplify the calculations.
For example, the Horner scheme for the Taylor series for <b>Exp(x)</b> may be written as

<p><center><b>Sum(k,0,N-1,x^k/k!)=1+x*(1+x/2*(1+x/3*(...+x/(N-1)))).</b></center></p>

This avoids computing the factorial function.


<p>
Similarly to the simple summation method, the working precision for Horner's scheme may be adjusted to reduce the computation time: for example, <b>x*a[N-1]</b> needs to be computed with just a few significant digits if <b>x</b> is small.
This does not change the asymptotic complexity of the method: it requires <b> O(N)=O(P)</b> long multiplications by <b>x</b>, so for general real <b> x</b> the complexity is again <b> O(P*M(P))</b>.
However, if <b>x</b> is a small rational number, then the multiplication by <b> x</b> is short and takes <b> O(P)</b> operations.
In that case, the total complexity of the method is <b>O(P^2)</b>.


<p>

<h3>
<hr>Method 3: "rectangular" or "baby step/giant step"
</h3>
<a name="Taylor series!"rectangular" method">

</a>
<a name="Taylor series!"baby step/giant step" method">

</a>
We can organize the calculation much more efficiently if we are able to estimate the necessary number of terms and to afford some storage (see [Smith 1989]).


<p>
The "rectangular" algorithm uses <b>2*Sqrt(N)</b> long multiplications (assuming that the coefficients of the series are short rational numbers) and <b>Sqrt(N)</b> units of storage.
For high-precision floating-point <b>x</b>, this method provides a significant advantage over Horner's scheme.


<p>
Suppose we need to evaluate <b> Sum(k,0,N,a[k]*x^k)</b> and we know the number of terms <b>N</b> in advance.
Suppose also that the coefficients <b> a[k]</b> are rational numbers with small numerators and denominators, so a  multiplication <b>a[k]*x</b> is not a long multiplication (usually, either <b> a[k]</b> or the ratio <b>a[k]</b>/<b>a[k-1]</b> is a short rational number). Then we can organize the calculation in a rectangular array with <b>c</b> columns and <b> r</b> rows like this,

<p><center><b> a[0]+a[r]*x^r+...+a[(c-1)*r]*x^((c-1)*r)+</b></center></p>


<p><center><b>x*(a[1]+a[r+1]*x^r+...+a[(c-1)*r+1]*x^((c-1)*r))+</b></center></p>


<p><center><b>...+</b></center></p>


<p><center><b>x^(r-1)*(a[r-1]+a[2*r+1]*x^r+...).</b></center></p>

To evaluate this rectangle, we first compute <b>x^r</b> (which, if done by the fast
binary algorithm, requires <b> O(Ln(r))</b> long multiplications). Then we compute
the <b>c-1</b> successive powers of <b> x^r</b>, namely
<b> x^(2*r)</b>, <b>x^(3*r)</b>, ..., <b>x^((c-1)*r)</b> in <b>c-1</b> long
multiplications. The partial sums in the <b> r</b> rows are evaluated column by column as more powers of <b> x^r</b> become available. This requires storage of <b> r</b> intermediate results but no more long multiplications by <b> x</b>. If a simple formula relating the coefficients <b> a[k]</b> and <b>a[k-1]</b> is available, then a whole column can be computed and added to the accumulated row values using only short operations, e.g. <b>a[r+1]*x^r</b> can be computed from <b> a[r]*x^r</b> (note that each column contains some consecutive terms of the series). Otherwise, we would need to multiply each coefficient <b> a[k]</b> separately by the power of <b>x</b>; if the coefficients <b> a[k]</b>  are short numbers, this is also a short operation. After this, we need <b>r-1</b> more
multiplications for the vertical summation of rows (using the Horner scheme). We have potentially saved time because we do not
need to evaluate powers such as <b> x^(r+1)</b> separately, so we do not have to multiply <b>x</b> by itself quite so many
times.


<p>
The total required number of long multiplications is <b> r+c+Ln(r)-2</b>. The minimum number of multiplications, given that <b> r*c&gt;=N</b>, is around  <b> 2*Sqrt(N)</b> at <b>r&lt;=&gt;Sqrt(N)-1/2</b>.
Therefore, by arranging the Taylor series in a rectangle with sides <b> r</b> and <b> c</b>,
we obtain an algorithm which costs <b> O(Sqrt(N))</b> instead of <b>O(N)</b> long multiplications and requires <b>Sqrt(N)</b> units of storage.


<p>
One might wonder if we should not try to arrange the Taylor series in a cube or another multidimensional matrix instead of a rectangle. However, calculations show that this does not save time: the optimal arrangement is the two-dimensional rectangle.


<p>
The rectangular method saves the number of long multiplications by <b>x</b> but increases the number of short multiplications and additions.
If <b> x</b> is a small integer or a small rational number, multiplications by <b> x</b> are fast and it does not make sense to use the rectangular method.
Direct evaluation schemes are more efficient in that case.


<p>

<h3>
<hr>Truncating the working precision
</h3>
At the <b> k</b>-th step of the rectangular method, we are evaluating the <b> k</b>-th column with terms containing <b> x^(r*k)</b>.
Since a power series in <b>x</b> is normally used at small <b> x</b>, the number <b> x^(r*k)</b> is typically much smaller than <b>1</b>.
This number is to be multiplied by some <b> a[i]</b> and added to the previously computed part of each row, which is not small.
Therefore we do not need all <b>P</b> floating-point digits of the number <b> x^(r*k)</b>,
and the precision with which we obtain it can be gradually decreased from column to column.
For example, if <b>x^r&lt;10^(-M)</b>, then we only need <b>P-k*M</b> decimal digits of <b> x^(r*k)</b> when evaluating the <b>k</b>-th column.
(This assumes that the coefficients <b> a[i]</b> do not grow, which is the case for most of the practically useful series.)


<p>
Reducing the working precision saves some computation time.
(We also need to estimate <b>M</b> but this can usually be done quickly by bit counting.)
Instead of <b> O(Sqrt(P))</b> long multiplications at precision <b>P</b>, we now need one long multiplication at precision <b> P</b>, another long multiplication at precision <b> P-M</b>, and so on.
This technique will not change the asymptotic complexity which remains <b> O(Sqrt(P)*M(P))</b>, but it will reduce the constant factor in front of the <b>O</b>.


<p>
Like the previous two methods, there is no accumulated round-off error if <b> x</b> is small.


<p>

<h3>
<hr>Which method to use
</h3>
There are two cases: first, the argument <b> x</b> is a small integer or rational number with very few digits and the result needs to be found as a floating-point number with <b> P</b> digits;
second, the argument <b> x</b> itself is a floating-point number with <b> P</b> digits.


<p>
In the first case, it is better to use either Horner's scheme (for small <b> P</b>, slow multiplication) or the binary splitting technique (for large <b> P</b>, fast multiplication).
The rectangular method is actually <i>slower</i> than Horner's scheme if <b> x</b> and the coefficients <b> a[k]</b> are small rational numbers.
In the second case (when <b>x</b> is a floating-point number), it is better to use the "rectangular" algorithm.


<p>
In both cases we need to know the number of terms in advance, as we will have to repeat the whole calculation if a few more terms are needed.
The simple summation method rarely gives an advantage over Horner's scheme, because it is almost always the case that one can easily compute the number of terms required for any target precision.


<p>
Note that if the argument <b> x</b> is not small, round-off error will become significant and needs to be considered separately for a given series.


<p>

<h3>
<hr>Speed-up for some functions
</h3>
<a name="Taylor series!<b> O(N^(1/3))</b> method">

</a>
An additional speed-up is possible if the function allows a transformation that reduces <b>x</b> and makes the Taylor series converge faster.
For example, <b> Ln(x)=2*Ln(Sqrt(x))</b>, <b>Cos(2*x)=2*Cos(x)^2-1</b> (bisection), and <b> Sin(3*x)=3*Sin(x)-4*Sin(x)^3</b> (trisection) are such transformations.
It may be worthwhile to perform a number of such transformations before evaluating the Taylor series, if the time saved by its quicker convergence is more than the time needed to perform the transformations.
The optimal number of transformations can be estimated.
Using this technique in principle reduces the cost of Taylor series from <b> O(Sqrt(N))</b> to <b>O(N^(1/3))</b> long multiplications. However, additional round-off error may be introduced by this procedure for some <b>x</b>.


<p>
For example, consider the Taylor series for <b> Sin(x)</b>,

<p><center><b>Sin(x)&lt;=&gt;Sum(k,0,N-1,(-1)^k*x^(2*k+1)/(2*k+1)!).</b></center></p>

It is sufficient to be able to evaluate <b>Sin(x)</b> for <b>0&lt;x&lt;Pi/2</b>.
Suppose we perform <b> l</b> steps of the trisection and then use the Taylor series with the rectangular method.
Each step of the trisection needs two long multiplications.
The value of <b> x</b> after <b> l</b> trisection steps becomes much smaller, <b> x'=x*3^(-l)</b>.
For this <b>x'</b>, the required number of terms in the Taylor series for <b> P</b> decimal digits of precision is

<p><center><b> N&lt;=&gt;(P*Ln(10))/(2*(Ln(P)-Ln(x')))-1.</b></center></p>

The number of long multiplications in the rectangular method is <b> 2*Sqrt(N)</b>.
The total number of long multiplications, as a function of <b>l</b>, has its minimum at

<p><center><b> l&lt;=&gt;(32*Ln(10)/Ln(3)*P)^(1/3)-(Ln(P)-Ln(x))/Ln(3),</b></center></p>

where it has a value roughly proportional to <b>P^(1/3)</b>.
Therefore we shall minimize the total number of long multiplications if we first perform <b>l</b> steps of trisection and then use the rectangular method to compute  <b> N</b> terms of the Taylor series.


<p>

<a name="c4s13">

</a>
<h2>
<hr>4.13 Using asymptotic series for calculations
</h2>
<a name="asymptotic series">

</a>
Several important analytic functions have asymptotic series expansions.
For example, the complementary error function <b>Erfc(x)</b> and Euler's Gamma function <b>Gamma(x)</b> have the following asymptotic expansions at large (positive) <b>x</b>:


<p>

<p><center><b> Erfc(x)=e^(-x^2)/(x*Sqrt(Pi))*(1-1/(2*x^2)+...+(2*n-1)!! /(-2*x^2)^n+...),</b></center></p>



<p>

<p><center><b>Ln(Gamma(x))=(x-1/2)*Ln(x)-x+Ln(2*Pi)/2 </b></center></p>


<p><center><b>+Sum(n,1,Infinity,B[2*n]/(2*n*(2*n-1)*x^(2*n-1))) </b></center></p>

(here <b>B[k]</b> are Bernoulli numbers).


<p>
The above series expansions are asymptotic in the following sense:
if we truncate the series and then take the limit of very large <b>x</b>,
then the difference between the two sides of the equation goes to zero.


<p>
It is important that the series be first truncated and then the limit of large <b> x</b> be taken.
Usually, an asymptotic series, if taken as an infinite series,
does not actually converge for any finite <b> x</b>.
This can be seen in the examples above.
For instance, in the asymptotic series for <b> Erfc(x)</b> the <b>n</b>-th term has <b>(2*n-1)!!</b> in the numerator which grows faster than the <b>n</b>-th power of any number.
The terms of the series decrease at first but then eventually start to grow,
even if we select a large value of <b> x</b>.


<p>
The way to use an
asymptotic series for a numerical calculation is to truncate the series
<i>well before</i> the terms start to grow.


<p>
Error estimates of the asymptotic series are sometimes difficult, but the rule of thumb seems to be that the error of the
approximation is usually not greater than the first discarded term of the series.
This can be understood intuitively as follows.
Suppose we truncate the asymptotic series at a point where the terms still decrease, safely before they start to grow.
For example, let the terms around the 100-th term be <b> A[100]</b>, <b>A[101]</b>, <b>A[102]</b>, ...,
each of these numbers being significantly smaller than the previous one,
and suppose we retain <b>A[100]</b> but drop the terms after it.
Then our approximation would have been a lot better if we retained <b>A[101]</b> as well.
(This step of the argument is really an assumption about the behavior of the series; it seems that this assumption is correct in many practically important cases.)
Therefore the error of the approximation is approximately equal to <b>A[101]</b>.


<p>
The inherent limitation of the method of asymptotic series is that for any given
<b>x</b>, there will be a certain place in the series where the term has the minimum
absolute value (after that, the series is unusable), and the error of the
approximation cannot be smaller than that term.


<p>

<a name="error function <b> Erf(x)</b>!by asymptotic series ">

</a>
<a name="asymptotic series!estimate of precision">

</a>
For example, take the above asymptotic series for <b>Erfc(x)</b>.
The logarithm of the absolute value of the <b>n</b>-th term can be estimated using Stirling's formula for the factorial as

<p><center><b> Ln((2*n-1)!! /(2*x^2)^n)&lt;=&gt;n*(Ln(n)-1-2*Ln(x)).</b></center></p>

This function of <b>n</b> has its minimum at <b> n=x^2</b> where it is equal to <b>-x^2</b>.
Therefore the best we can do with this series is to truncate it before this term.
The resulting approximation to <b> Erfc(x)</b> will have relative precision of order <b>Exp(-x^2)</b>.
Suppose that <b>x</b> is large and we need to compute <b> Erfc(x)</b> with <b>P</b> decimal digits of floating point.
Then it follows that we can use the asymptotic series only if <b> x&gt;Sqrt(P*Ln(10))</b>.


<p>
We find that for a given finite <b>x</b>, no matter how large, there is a maximum
precision that can be achieved with the asymptotic series; if we need more
precision, we have to use a different method.


<p>
However, sometimes the function we are evaluating allows identity transformations that relate <b> f(x)</b> to <b>f(y)</b> with <b>y&gt;x</b>.
For example, the Gamma function satisfies <b> x*Gamma(x)=Gamma(x+1)</b>.
In this case we can transform the function so that we would need to
evaluate it at  large enough <b>x</b> for the asymptotic series to give us
enough precision.


<p>

<a name="c4s14">

</a>
<h2>
<hr>4.14 The AGM sequence algorithms
</h2>
<a name="AGM sequence">

</a>
Several algorithms are based on the arithmetic-geometric mean (AGM)
sequence. If one takes two numbers <b>a</b>, <b> b</b> and computes their arithmetic
mean <b>(a+b)/2</b> and their geometric mean <b> Sqrt(a*b)</b>, then one finds that the two means
are generally much closer to each other than the original numbers. Repeating
this process creates a rapidly converging sequence of pairs.


<p>
More formally, one can define the function of two
arguments <b>AGM(x,y)</b> as the limit of the sequence <b>a[k]</b> where
<b>a[k+1]=1/2*(a[k]+b[k])</b>, <b>b[k+1]=Sqrt(a[k]*b[k])</b>, and the initial values
are <b>a[0]=x</b>, <b> b[0]=y</b>. (The limit of the sequence <b> b[k]</b> is the same.)
This function is obviously linear, <b>AGM(c*x,c*y)=c*AGM(x,y)</b>, so in
principle it is enough to compute <b>AGM(1,x)</b> or arbitrarily select <b>c</b> for convenience.


<p>

<a name="AGM sequence!integral representation">

</a>
Gauss and Legendre knew that
the limit of the AGM sequence is related to the complete elliptic integral,

<p><center><b> Pi/2*1/AGM(a,Sqrt(a^2-b^2))=(Integrate(x,0,Pi/2)1/Sqrt(a^2-b^2*Sin(x)^2)).</b></center></p>

(Here <b>0&lt;b&lt;a</b>.) This integral can be rearranged to provide some other useful functions. For
example, with suitable parameters <b> a</b> and <b> b</b>, this integral is equal to
<b> Pi</b>.
Thus, one obtains a fast method of computing <b> Pi</b> (the Brent-Salamin method).


<p>
The AGM sequence is also defined for complex values <b> a</b>, <b> b</b>.
One needs to take a square root <b> Sqrt(a*b)</b>, which requires a branch cut to be well-defined.
Selecting the natural cut along the negative real semiaxis (<b>Re(x)&lt;0</b>, <b> Im(x)=0</b>), we obtain an AGM sequence that converges for any initial values <b> x</b>, <b> y</b> with positive real part.


<p>

<a name="AGM sequence!convergence rate">

</a>
Let us estimate the convergence rate of the AGM sequence starting from <b> x</b>, <b> y</b>, following the paper [Brent 1975]. Clearly the worst case is when
the numbers <b> x</b> and <b> y</b> are very different (one is much larger than another). In this case
the numbers <b> a[k]</b>, <b>b[k]</b> become approximately equal after about
<b>k=1/Ln(2)*Ln(Abs(Ln(x/y)))</b> iterations (note: Brent's paper online mistypes
this as <b>1/Ln(2)*Abs(Ln(x/y))</b>).
This is easy to see: if <b>x</b> is much larger than <b> y</b>, then at each step the ratio <b> r:=x/y</b> is transformed into <b> r'=1/2*Sqrt(r)</b>.
When the two numbers become roughly equal to each other, one needs about <b>Ln(n)/Ln(2)</b> more
iterations to make the first <b>n</b> (decimal) digits of <b> a[k]</b> and <b>b[k]</b>
coincide, because the relative error <b>epsilon=1-b/a</b> decays approximately as
<b> epsilon[k]&lt;=&gt;1/8*Exp(-2^k)</b>.


<p>
Unlike Newton's iteration, the AGM sequence does not correct errors, so all numbers need to be computed with full precision.
Actually, slightly more precision is needed to compensate for accumulated round-off error.
Brent (in [Brent 1975]) says that <b>O(Ln(Ln(n)))</b> bits of accuracy are lost to round-off error if there are total of <b>n</b> iterations.


<p>
The AGM sequence can be used for fast computations of <b> Pi</b>, <b> Ln(x)</b> and <b>ArcTan(x)</b>.
However, currently the limitations of Yacas internal math make these methods less efficient than simpler methods based on Taylor series and Newton iterations.


<p>

<a name="c4s15">

</a>
<h2>
<hr>4.15 The binary splitting method
</h2>
<a name="binary splitting">

</a>
The method of binary splitting is well explained in [Haible <i>et al.</i> 1998].
Some examples are also given in [Gourdon <i>et al.</i> 2001].
This method applies to power series of rational numbers and to hypergeometric series.
Most series for transcendental functions belong to this category.


<p>
If we need to take <b>O(P)</b> terms of the series to obtain <b>P</b> digits of precision, then ordinary methods would require <b> O(P^2)</b> arithmetic operations.
(Each term needs <b>O(P)</b> operations because all coefficients are rational numbers with <b>O(P)</b> digits and we need to perform a few short multiplications or divisions.)
The binary splitting method requires <b>O(M(P*Ln(P))*Ln(P))</b> operations instead of the <b>O(P^2)</b> operations.
In other words, we need to perform long multiplications of integers of size <b>O(P*Ln(P))</b> digits, but we need only <b>O(Ln(P))</b> such multiplications.
The binary splitting method performs better than the straightforward summation method if the cost of multiplication is lower than <b>O(P^2)/Ln(P)</b>.
This is usually true only for large enough precision (at least a thousand digits).


<p>
Thus there are two main limitations of the binary splitting method:
<ul><li>As a rule, we can only compute functions of small integer or rational arguments.
For instance, the method works for the calculation of a Bessel function </li><b>J0(1/3)</b> but not for <b>J0(Pi)</b>.
(As an exception, certain elementary functions <i>can</i> be computed by the binary splitting method for general floating-point arguments, with some clever tricks.)
<li>The method is fast only at high enough precision, when advanced multiplication methods become more efficient than simple </li><b>O(P^2)</b> methods.
The binary splitting method is actually <i>slower</i> than the simple summation when the long integer multiplication is <b>M(P)=O(P^2)</b>.
</ul>

<p>
The main advantages of the method are:
<ul><li>The method is asymptotically fast and, when applicable, outperforms most other methods at very high precision.
The best applications of this method are for computing various constants.
</li><li>There is no accumulated round-off error since the method uses only exact integer arithmetic.
</li><li>The sum of a long series can be split into many independent partial sums which can be computed in parallel.
One can store exact intermediate results of a partial summation (a few long integers), which provides straightforward checkpointing: a failed partial summation can be repeated without repeating all other parts.
One can also resume the summation later to get more precision and reuse the old results, instead of starting all over again.
</li></ul>

<p>

<h3>
<hr>Description of the method
</h3>
We follow [Haible <i>et al.</i> 1998].
The method applies to any series of rational numbers of the form

<p><center><b>S=Sum(n,0,N-1,A(n)/B(n)),</b></center></p>

where <b>A</b>, <b> B</b> are integer coefficients with <b> O(n*Ln(n))</b> bits.
Usually the series is of the particular form

<p><center><b>S(0,N):=Sum(n,0,N-1,a(n)/b(n)*(p(0)*...*p(n))/(q(0)*...*q(n))),</b></center></p>

where <b>a</b>, <b> b</b>, <b> p</b>, <b> q</b> are polynomials in <b> n</b> with small integer coefficients and values that fit into <b> O(Ln(n))</b> bits.


<p>
For example, the Taylor series for <b>ArcSin(x)</b> (when <b>x</b> is a short rational number) is of this form:

<p><center><b> ArcSin(x)=x+1/2*x^3/3+(1*3)/(2*4)*x^5/5+(1*3*5)/(2*4*6)*x^7/7+... </b></center></p>

This example is of the above form with the definitions <b>a=1</b>, <b> b(n)=2*n+1</b>, <b> p(n)=x^2*(2*n-1)</b>, <b>q(n)=2*n</b> for <b> n&gt;=1</b> and <b> p(0)=x</b>, <b> q(0)=1</b>.
(The method will apply only if <b> x</b> is a rational number with <b> O(Ln(N))</b> bits in the numerator and the denominator.)
The Taylor series for the hypergeometric function is also of this form.


<p>
The goal is to compute the sum <b>S(0,N)</b> with a chosen number of terms <b>N</b>.
Instead of computing the rational number <b> S</b> directly, the binary splitting method propose to compute the following four integers <b> P</b>, <b> Q</b>, <b> B</b>, and <b> T</b>:

<p><center><b> P(0,N):=p(0)*...*p(N-1),</b></center></p>


<p><center><b>Q(0,N):=q(0)*...*q(N-1),</b></center></p>


<p><center><b>B(0,N):=b(0)*...*b(N-1),</b></center></p>
 and

<p><center><b>T(0,N):=B(0,N)*Q(0,N)*S(0,N).</b></center></p>

At first sight it seems difficult to compute <b>T</b>, but the computation is organized recursively.
These four integers are computed for the left (<b> l</b>) half and for the right (<b> r</b>) half of the range [<b> 0</b>, <b> N</b>) and then combined using the obvious recurrence relations <b> P=P[l]*P[r]</b>, <b>Q=Q[l]*Q[r]</b>, <b>B=B[l]*B[r]</b>, and the slightly less obvious relation

<p><center><b>T=B[r]*Q[r]*T[l]+B[l]*P[l]*T[r].</b></center></p>

Here we used the shorthand <b>P[l]:=P(0,N/2-1)</b>, <b>P[r]:=P(N/2,N-1)</b> and so on.


<p>
Thus the range [<b>0</b>, <b> N</b>) is split in half on each step.
At the base of recursion the four integers <b> P</b>, <b> Q</b>, <b> B</b>, and <b> T</b> are computed directly.
At the end of the calculation (top level of recursion), one floating-point division is performed to recover <b> S=T/(B*Q)</b>.
It is clear that the four integers carry the full information needed to continue the calculation with more terms.
So this algorithm is easy to checkpoint and parallelize.


<p>
The integers <b>P</b>, <b> Q</b>, <b> B</b>, and <b> T</b> grow during the calculation to <b> O(N*Ln(N))</b> bits, and we need to multiply these large integers.
However, there are only <b>O(Ln(N))</b> steps of recursion and therefore <b>O(Ln(N))</b>
long multiplications are needed.
If the series converges linearly, we need <b>N=O(P)</b> terms to obtain <b>P</b> digits of precision.
Therefore, the total asymptotic cost of the method is <b> O(M(P*Ln(P))*Ln(P))</b> operations.


<p>
A more general form of the binary splitting technique is also given in [Haible <i>et al.</i> 1998].
The generalization applies to series for the form

<p><center><b>Sum(n,0,N-1,a(n)/b(n)*(p(0)*...*p(n))/(q(0)*...*q(n))*(c(0)/d(0)+...+c(n)/d(n))),</b></center></p>

Here <b>a(n)</b>, <b>b(n)</b>, <b>c(n)</b>, <b>d(n)</b>, <b>p(n)</b>, <b>q(n)</b> are integer-valued functions with "short" values of size <b>O(Ln(n))</b> bits.
For example, the Ramanujan series for Catalan's constant is of this form.


<p>
The binary splitting technique can also be used for series with complex integer coefficients, or more generally for coefficients in any finite algebraic extension of integers, e.q. <b>Z</b>[<b> Sqrt(2)</b>] (the ring of numbers of the form <b>p+q*Sqrt(2)</b> where <b>p</b>, <b> q</b> are integers).
Thus we may compute the Bessel function <b> J0(Sqrt(3))</b> using the binary splitting method and obtain exact intermediate results of the form <b>p+q*Sqrt(3)</b>.
But this will still not help compute <b>J0(Pi)</b>.
This is a genuine limitation of the binary splitting method.


<p>


<script src="http://www.google-analytics.com/urchin.js" type="text/javascript">
</script>
<script type="text/javascript">
_uacct = "UA-2425144-1";
urchinTracker();
</script>
</body>

</html>