Distributed Hydrologic Modeling Using GIS (Water Science and Technology Library) [2nd ed.] 1402024592, 9781402024597, 9781402024603

From the foreword of the second edition: "All of the above make this a unique, invaluable book for the student, pro

253 111 5MB

English Pages 294 [312] Year 2004

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
00000___42d9a7205358f8e4d59c860bd8f25833......Page 1
00001___aeef12492e7e7c68183edc6ac4864d8d......Page 2
00002___c06e3d1e95a0f1a53ad6979c85c9684b......Page 3
00003___009ef804fd4543e2607aafee2379efa6......Page 4
00004___eb6167e66baa37724e6a5a6edd7fc4b0......Page 5
00005___300e8d01287ac264252ecfcee3d6d1c7......Page 6
00006___edf3d3c1c7c1c55b7da9fb1a6f13d652......Page 7
00007___48b80b3078e6824e1e3056359ea27114......Page 8
00008___a6b96d8ae2957fe753d356a3ae4a877b......Page 9
00009___a356e18c5a1e96cb2eddc7fbe5c9c605......Page 10
00010___c174112cea6b56da2f31ea68df641a1b......Page 11
00011___ec555e37495a301467fe304c9c6567ef......Page 12
00012___ca9be9da4f48ce96933771e478228574......Page 13
00013___f93d006d02d6e2b8bf485de84281b48c......Page 14
00014___4ece26d66c1ab6c6a90ea8ef42048d1b......Page 15
00015___7eada6b42f08abab54e12cfb0f5ed110......Page 16
00016___3c37825a301dfb02465b4ae94b3889b5......Page 17
00017___e63d292207f35de53c8eb36a4e5974b4......Page 18
00018___0ad1e9695cc22c06767087bc8f06625c......Page 19
00019___22bb2d808f25111e0180bacbd779d848......Page 20
00020___846db5b3f1c01788500d2d695fb2d39b......Page 21
00021___0da531a8f6fa7c0049e937210bc08139......Page 22
00022___41f977817b6fde74f566597b73153aa5......Page 23
00023___ff304c9c987dd870de28f16430f13141......Page 24
00024___40ed3b1bf797035c0851f8d244c99ee2......Page 25
00025___c74d5bccd6abe4665347dc177831568e......Page 26
00026___d1affd66808f9637b91c456a1082511e......Page 27
00027___e9d10405cac408d9ec5bade856f432c7......Page 28
00028___520a6b10876061ba58b729042712e8c6......Page 29
00029___c1deea68ba8d69dd5c847aaaf2506c06......Page 30
00030___ca2711dd67b3d6208547ef72021e08a8......Page 31
00031___c3f9062481464b9f9b5e4091742d99ed......Page 32
00032___85bc2d31aa26682fa6ecf683debc9e6f......Page 33
00033___5a662b32c83fb127c9441a409bc12baa......Page 34
00034___d4c4d85660e082299525b06682be90a8......Page 35
00035___acdc137b9a8d3d41e9bd9d273d4d9dcd......Page 36
00036___b77c7ea865d281ca1dcf3ee5acd40edc......Page 37
00037___26482d9241ee5ea5c9a0bca78e48d223......Page 38
00038___6184a79c0b0ab82d892268afcb034d1b......Page 39
00039___6a4538ffef0feeda6b027772d3672b4e......Page 40
00040___62bb56afdb6fadd6edacdda8ac746ca6......Page 41
00041___84dacb15c4f2f2ce530d7513e421a82f......Page 42
00042___b9418d33a008e07b5a39bb932f089058......Page 43
00043___a7ef295ddc832b08a40a913929c7753c......Page 44
00044___da08887a69db63e6ad528fe3ab1fe7ee......Page 45
00045___3bfd5d67770e2d8fe8f1715e5e424b50......Page 46
00046___10ae1b5e794940cc9cd72f8d70973ec8......Page 47
00047___3aca58a0754532f0db485c4b223313bd......Page 48
00048___75d75addf8325da248a4c316ac17dca5......Page 49
00049___9a40ffddffbb6b146908ee424a74306c......Page 50
00050___5d7981ec8d0185c15d4624e8a154a2a0......Page 51
00051___57acf2e287c2925c01977740196d977d......Page 52
00052___90641bfed041b9f3fbe6c7c215e987de......Page 53
00053___4aac20d09f505ce30c20f1609d6f956e......Page 54
00054___f165f8809e48d9b2df43394b6c0ebef7......Page 55
00055___c9bc7db7d84645bfe157b740ee6543b6......Page 56
00056___d17b494ba33c84f2b78e6faef88b3c9a......Page 57
00057___1713f3abd1c61411eee49ff74bfa039a......Page 58
00058___887502fbc0c671c7bf4586deb174c829......Page 59
00059___17533f638da4be83e4a4b1450db5d2da......Page 60
00060___66d1177192abd5eb93337d1ee96b8007......Page 61
00061___b791df6ab9a5272629afc1638c435c6a......Page 62
00062___f98c4b8b0dc0385ec0597d8fb2bde7d5......Page 63
00063___a46bd2cfde9335c67498bfbd10b6d54f......Page 64
00064___cf16cca668f8c392cabf6a6926fdd5f8......Page 65
00065___330888fde9d7ed038ebf281a4270865c......Page 66
00066___151a2392d2d5ccaf39f3743e74a31a3d......Page 67
00067___8a2ceb5d933bce45ad940a829713e279......Page 68
00068___175a672166369e4c5af583269ba2a64d......Page 69
00069___1a48c20ef1e881867323999792bba509......Page 70
00070___20287594da9f19a660d604c2d6b14ac6......Page 71
00071___7f3a9a1a687595292b6b0d24154d4226......Page 72
00072___d597f0781444be9501fad1a62dd1add6......Page 73
00073___7a647daf85d0b02bbb51b57ec39c8e7e......Page 74
00074___8ce46f38abe51eaa9dd8c3797efdcdbd......Page 75
00075___302fbe8da047ab945455e272d7335e0b......Page 76
00076___b84883d78914aea3a07e83d610ae51b9......Page 77
00077___e05377601260fc1b855d91719d530d0e......Page 78
00078___00af8d206568bac7c3f555be835ea74d......Page 79
00079___951b61a5ebe07e54b15d5567cdb3c303......Page 80
00080___fa1d08e4d82e3bfe7689a927ba1aea41......Page 81
00081___1a418a322937d1bf97cc84a613f09fe4......Page 82
00082___c02b161ce4e93fb444f2682e775d9c00......Page 83
00083___6bc081a32046e9e05f358497ad771d3a......Page 84
00084___998b0fe193c61eca43c961d12a9e6663......Page 85
00085___1842e126b3f393cfce1f36051765ba48......Page 86
00086___0a107219e06bd2e0abe43d0f8b122f8e......Page 87
00087___b7d47e77fdd682ac04d5045c5c02e870......Page 88
00088___9e5d6fbf3c6ad951c667e8ef585a9bc2......Page 89
00089___2786ce289d3e26f38ba9c267d0716abf......Page 90
00090___4e31c2bc89b0cb71b817e9fecfb41313......Page 91
00091___8e92645d74bb27bb2e12acbaeffec462......Page 92
00092___ccc4930318dc313a10576be7ab2af7b9......Page 93
00093___83808080a4173fa6dbacbc6a12e9e045......Page 94
00094___a1628c9342ca3334e0f2a125c0bb82a9......Page 95
00095___af5c0b5707d23c3c797573a9c71513ba......Page 96
00096___d95c9778dca41da4c1f5a11b41027780......Page 97
00097___1ff0c2a98fa21a3751988226ba8d47e5......Page 98
00098___33895c1d96345878e0e853a036efc580......Page 99
00099___1e3dbe266f93d1cc1a5134a4653ba9b3......Page 100
00100___e6eb3cf3e089823ded19007acbd97222......Page 101
00101___21fba988a9296a0a4443f7970abfeab9......Page 102
00102___cfdfd2bab3aeff30f84545efaaa5540c......Page 103
00103___457dc938922994b4a9226b2aa999f0dd......Page 104
00104___d1cccfb9f2e4a82caa38218a48569a6c......Page 105
00105___76264395a90c9824832126e8c7333b0c......Page 106
00106___74c2c2b82ec92221cbb016564aa25a7f......Page 107
00107___675dc183f60df65da00aeccdd6e350f7......Page 108
00108___3c6a4c82e7c107904e6bcc49fe5500ba......Page 109
00109___8a81fea318bc6a8f5b1dcc9cbaab6777......Page 110
00110___f1fd5e842edec94b50c64c13faf89599......Page 111
00111___8af8a99d746a4f3eb260ddfc8ed64995......Page 112
00112___006ab4641e4a4618580aff8f230a01e0......Page 113
00113___27115f4e01d0dc44215ea5312b4747f5......Page 114
00114___3c8183e85d462e4b007c9d188e329738......Page 115
00115___9a385edac0e8ac5057ae632128f4bee6......Page 116
00116___83fa8a1bc7a12b771869519af3d3f148......Page 117
00117___16a74e65a7441e19885d6fb91389907a......Page 118
00118___b6986584f43c625a0ccc812559b49343......Page 119
00119___3df08886a2a2bf546fa96b4c276613c8......Page 120
00120___a3f4c6f31ce102834ddcff51cdfb5ab0......Page 121
00121___498e2fc7946dcad3aeed118bca5a1e00......Page 122
00122___9a596029d9e5663bd69e7808dd231b1a......Page 123
00123___2724834eb536904a5e1845664fbf7749......Page 124
00124___c619bcc27e4bbb6b25461759583b094b......Page 125
00125___b512f6585a697e1334a0dcae9acd1cfb......Page 126
00126___43dce214d8ec8f99c9f9b8df5ebdd021......Page 127
00127___05e57371aca3793d90d75380296ba74f......Page 128
00128___07accd3a9bb3c0d100b0a4f46ddd1382......Page 129
00129___b156ee4bf0ea6d15d0ea017a570566e5......Page 130
00130___d6216d2cbdd3e36ea0a4a03070c664cd......Page 131
00131___d900c4f9f86dee44c5b7fd76eceaa5c7......Page 132
00132___9a2ba5a38b80756c172964ceb1c745d0......Page 133
00133___d3c4faae3fb71af165156bbb77858a66......Page 134
00134___2e69ed8efdd8b23c378a7a56defa4f31......Page 135
00135___85ba1bc07f2f4f9875afd7547f5b0124......Page 136
00136___eb963537fe9ecc143d7c3391ed24ffc9......Page 137
00137___994fa5deb903609c1984343f5e2a2a10......Page 138
00138___ca905a4a4ce858781862ad9a16ed858c......Page 139
00139___d70cbf80298601a739fa55b79c57fbf5......Page 140
00140___017df9058177b6c86eccab1f210ad095......Page 141
00141___ad8de5905c925c6e309f1d468143d813......Page 142
00142___ae90553d554065b40fcfb9b23ae43f13......Page 143
00143___343c1a1ac4f334828d774a3731a5f5cd......Page 144
00144___0a38cea913ac2b75b36ffef930d35b4a......Page 145
00145___00646265f6cde2f8d1abe6b8f216a725......Page 146
00146___f5f819ad8304ef6312ef2085a1f65577......Page 147
00147___8b7378167ab51f8b226b88c5d829582d......Page 148
00148___4dd9f739f548b3c545b4d378b22a109c......Page 149
00149___44feb416751abc50d9c0ed655ca5ec21......Page 150
00150___62670d2a609ed83e3a0a3f9cfcef92d2......Page 151
00151___432d8fe35029d91d54bc7943a7f4d192......Page 152
00152___4fdd418ae60b0f3d15e7e8e2ff315bf1......Page 153
00153___eede86fb9114f210fa27b30ffbe68bb0......Page 154
00154___dbd37166d14326713841b784c2b49656......Page 155
00155___a36f0815b3c1f71936f01bdaea7bb5fb......Page 156
00156___d3758a38c4a54892e5fde7510016484f......Page 157
00157___a5cd0a16d6f2f7b8d6e3b5de8f20b0be......Page 158
00158___60c57572b4d92b83209c39081f498e22......Page 159
00159___a0e4f02f140a80f52d0c1acbbd296c19......Page 160
00160___b816c3f71f75e5d344f59a5e7b33b387......Page 161
00161___1ee0cb47c2a0a8c3a5599619c21c6e01......Page 162
00162___888b2c8c6d785bddb642782741dd1e4c......Page 163
00163___ddc0dffa11cee332b0593a8a52545694......Page 164
00164___51050e434570398c284ede2151b82342......Page 165
00165___7bfc52ecba93364c50c74081a3766697......Page 166
00166___d691c9ad720f8e8f14b3426ff0819f1d......Page 167
00167___b3cb865818bb19ab220aa4b96802bbfa......Page 168
00168___d6573945c73a3ce1aca95bfe6784030b......Page 169
00169___939812e7f084a0144e4a767a87ab5c8f......Page 170
00170___c6d0935469226b28a7518aac070224ff......Page 171
00171___8c59bdeade498fe892d38e451ca34352......Page 172
00172___6b70f66f548d1cec64f3ddbed3a5f5b5......Page 173
00173___64845c696ca3084438d9c3c4ece35c07......Page 174
00174___aa41f4a8f6a8d3d76f4502d9725c2675......Page 175
00175___7fbd87812bd0f4be9cb106a02ed12d55......Page 176
00176___d191e00af449e01fe0fe7b06e6b39eea......Page 177
00177___5a77215cf74d758e0e4361cfdbc68062......Page 178
00178___fac0479a120d4c02be00ea785e8a4828......Page 179
00179___a8c7646330c0dca94c0a4b7106cfa46f......Page 180
00180___5de192c5d2e70aeb733fc65c35d4a20b......Page 181
00181___adfcc4bdbeac6f52d59dea0f7ed3c57b......Page 182
00182___c0a740bb7ab12b4da8ecca1ba136c868......Page 183
00183___0a4234d9034e8d4b64c03f1d55dcd89b......Page 184
00184___b2d39b0315c7fa57d8eb317a1101acd2......Page 185
00185___0c297c29b708556167c711b48a2b298b......Page 186
00186___58d39c0b2eb0d62264a5e8a872711ed8......Page 187
00187___8101e0399ca4aa7c662660d647bb0f07......Page 188
00188___781913099f215cc6e7abc83d9d086235......Page 189
00189___153f52135693318eee76ed7a055de5f2......Page 190
00190___ef8465bead589dd6429a3796f42ee506......Page 191
00191___e569b60328861a1d5d8a767dc3a5f82a......Page 192
00192___6e46687ae98cccde140da06e87480def......Page 193
00193___1b35d80cb2d7508b5952d438844ae586......Page 194
00194___1896bfeb34141c4e37dabe16bfae6d53......Page 195
00195___1c0934588bfa1874b9acf37d0e2e35fb......Page 196
00196___17522d5f0d3060644ecdf81e03ac048f......Page 197
00197___802210f7681740a6e1fa01ffc02b5654......Page 198
00198___9e4a778ee8bf34deb43c3105828e5465......Page 199
00199___f27067c23210cbc0c03091958e4f4b03......Page 200
00200___92b81798b3c76d95784a2a76397b7533......Page 201
00201___d1a26c7ba5c44304657608be1c37684c......Page 202
00202___c7b6fa3981bc3526cf79a3ee98521807......Page 203
00203___d413fc2b1dfcccf4b6c00a1f2fe949ca......Page 204
00204___5ef0fe41e08d234cc7a05aa8cab71abf......Page 205
00205___4e60cafa43c1a575771b6d98bfb2cfba......Page 206
00206___34d53bf28c549f75f35fe909c1ba2247......Page 207
00207___7050182b27a750efafb28f5d4dd1a205......Page 208
00208___030c475f4e9ea0a49cc1eab7a3f487ce......Page 209
00209___87d43e130373ac4925f888b2f1bf3ac2......Page 210
00210___fd6b922e5918ab6c8f3876bfb8192524......Page 211
00211___3f0537d0ebf68a19ed2c50e34d46be3c......Page 212
00212___03621514c7cbf4395da87ad163ec3929......Page 213
00213___d8234196f9cae21559bf89a1332ed6ee......Page 214
00214___76b370ed39cdb0d1abce7cf9637f4a83......Page 215
00215___9a8113e297b8bb061a609e248cca00e6......Page 216
00216___536c5300850def80b5c0bd6bbbe7da1b......Page 217
00217___c8234d3b4e8eea39240b4c2e9c3b932c......Page 218
00218___306b035452aeb775bce8bc468d8b9b23......Page 219
00219___f123016da1c630445ca2cdc462ab68a4......Page 220
00220___5c64197b62dba9a39b2de8b3d41d5d62......Page 221
00221___a054ae453fc853d20fa8fde4d091fcb2......Page 222
00222___a6373be5bc10bf8d47fea8626fe5ae00......Page 223
00223___43f5fd1168a6af7dfc055d13ba8dae10......Page 224
00224___66f847d0bdf80699aa8203a834fd7aca......Page 225
00225___417e5ebde8ffd65f617f17ebee3ac34d......Page 226
00226___cf3bc6dd446260c81eaaab6dc07921dd......Page 227
00227___391e00f27f9a6178e1798ac7c3dd0031......Page 228
00228___5316d7f06e1d95ad3d089437c79ad602......Page 229
00229___c8b6fc65a2362511e9bb6fda52673352......Page 230
00230___05218c76154cfc438df04ebea9e42475......Page 231
00231___8134a87ad020b078016c81416b85b75c......Page 232
00232___75a278b035b19d7d7f021181016c7edd......Page 233
00233___ca4e942670204b50ae21c51aa2ac14ea......Page 234
00234___b4a67b8066caaf39f282666d2fc18f32......Page 235
00235___b2472553d8fa5d31be91cc86b3bfdbb9......Page 236
00236___5430ee3fd55d1188a021fd30a852f613......Page 237
00237___ec3b9845c249ab3390f6c013f3fdd583......Page 238
00238___5f206488a2b153610da05375aeb8cc7e......Page 239
00239___f40cbeb8a407d44e91349680d259ef74......Page 240
00240___815a33d9a668be86fa643782b9c1094b......Page 241
00241___7a03fdca9a0049871c0d6d90434a6f4e......Page 242
00242___ddbffa1ada175adfec5d280f2fabab81......Page 243
00243___fd54ec2e7d2fc1e3e01b114e034bfa5a......Page 244
00244___47e9037f935f3f7329b09cb26647bace......Page 245
00245___fd8023f29e18805eb1fead6cb2422e8b......Page 246
00246___d4d78750389d6c1115dfc94511b610e6......Page 247
00247___ccf67e7734ce56d4d553a60c0de4f7f3......Page 248
00248___f6d41242114e5ba087304ad327fb61c6......Page 249
00249___87223e01857a2d39042a48f3f47d59b0......Page 250
00250___d3e2f600aae17ef59c4aab9583e2987d......Page 251
00251___3cacfc8af3e0108055109f04e4b3c30f......Page 252
00252___c9f828bad3aca3f972083c33f2f62df1......Page 253
00253___6a8002bbffdccb30c65e8995089e33e6......Page 254
00254___c6b5502345d7ef32be5183c00208f07f......Page 255
00255___6bbd0235c8b8a7d493064baac99a2b4d......Page 256
00256___de8fe2b4cc99c2fe53c9fd70fa00b58a......Page 257
00257___ce5797893f849eebc5e05d7ca58ab9b6......Page 258
00258___730883e6f83b4a9322e73177b8295c1c......Page 259
00259___5378044027b9696e84fb6523f4002af8......Page 260
00260___ae0f840391fcd394f5d20ff73e8cf012......Page 261
00261___f4a7af1f97d23e5e7dd0f65260b2db78......Page 262
00262___09c3ee499221e958557c287572569c67......Page 263
00263___fa18de2eac575361f550fcc432292ace......Page 264
00264___ffab616e6a82490ba1a69f5542028d4b......Page 265
00265___c5306e640e84274dc199de591e6f97b7......Page 266
00266___436b48215ff5430577950677ad990df7......Page 267
00267___47841b43c989c988b6ec48457ae84e76......Page 268
00268___5255bdf5c16acc00adf37dcb516c5b9c......Page 269
00269___e282b3323a8334b475d8b5a119a5858e......Page 270
00270___9e5665d4eeb57bf19bf93b18f977aaf5......Page 271
00271___f7f6e281fe5d1dc439e9a410f34f295e......Page 272
00272___c459da56548b985fa969d869bdca9ae4......Page 273
00273___bdd72b974bedf2ad3cd1763e6ba38619......Page 274
00274___7f14749655c95a8ab8f33ec7106fa7e1......Page 275
00275___7c0f6feede3e7c7b4b253b7baf3248cb......Page 276
00276___491473d7da7bb078d6e7e5c5636353ba......Page 277
00277___5ce5aae479b392fcce03503be4b4d025......Page 278
00278___9908371311e20eadddec11575cf82af9......Page 279
00279___f7017df5d45eb844869f3843877c58b2......Page 280
00280___9eae1c568212165cc7df2409d505a7f6......Page 281
00281___5fafa748669bd59fb6432f83c0a41514......Page 282
00282___1c50aaba4f66f7936378e335477e8676......Page 283
00283___1393007776845a13af8f03d0b800e92d......Page 284
00284___8f6a79ac09f34de36cfa7cf3787c2b34......Page 285
00285___f0b8c666ce5e350d29bf879cb7a0cc90......Page 286
00286___5c57c3159893aee8ce7ad324774084f0......Page 287
00287___5ff297bc91e505c3d9388e18f3b900bd......Page 288
00288___dee0b316dae7292bceffca1d6a752eea......Page 289
00289___fad2091ce9b8027eb673fd9fb6a5285e......Page 290
00290___3f1789626b169bbbc9dbfd9b93d56183......Page 291
00291___8451c6e0bfdbf4ebfadec69ed7713697......Page 292
00292___9380f16b7dcea72b302e156c6413322f......Page 293
00293___bf0972162cda6eda3a818054333f90c3......Page 294
00294___bec7cba707e2afd139c419e015696da0......Page 295
00295___53b63c71ba98afc615882950ab888956......Page 296
00296___8cd5e2051c7bf5181d816e1c12a9c930......Page 297
00297___d219b1c8eb0b2e826543676176b28ae3......Page 298
00298___bc816f4000656172ccbfd0675d3f265a......Page 299
00299___559104160f6cc5fa63c4af2d773d90a0......Page 300
00300___fcb24ee62159f472b79312f7b4b56518......Page 301
00301___d5fb7a6f3c17de3a529f34646cec703f......Page 302
00302___e9623f99262765b8ff406763ff7f67ec......Page 303
00303___a833c0b213a58d7e0e229b32c43f4329......Page 304
00304___d19c4e16baa47ececd1e8412fe74dbdd......Page 305
00305___6d8f62459f2e99de9561d8e792b29901......Page 306
00306___2066a6dfeb12fb9182d6d7703bde7d5b......Page 307
00307___9976c7ae0b05edcad8b08f8285c6f456......Page 308
00308___cb2d697cbc653de6b3c452b79b2b0a52......Page 309
00309___119d592f5cabe3a8e41b598af658f05d......Page 310
00310___95dbb93cbc66ac48e058ea2eb33004eb......Page 311
00311___a1d536166dfb68858728cfb144e91777......Page 312
page_xviii......Page 0
Recommend Papers

Distributed Hydrologic Modeling Using GIS (Water Science and Technology Library) [2nd ed.]
 1402024592, 9781402024597, 9781402024603

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

DISTRIBUTED HYDROLOGIC MODELING USING GIS Second Edition

Water Science and Technology Library VOLUME 48

Editor-in-Chief V. P. Singh, Louisiana State University, Baton Rouge, U.S.A. Editorial Advisory Board M. Anderson, Bristol, U.K. L. Bengtsson, Lund, Sweden J. F. Cruise, Huntsville, U.S.A. U. C. Kothyari, Roorkee, India S.E. Serrano, Philadelphia, U.S.A. D. Stephenson, Johannesburg, South Africa W.G. Strupczewski, Warsaw, Poland

The titles published in this series are listed at the end of this volume.

DISTRIBUTED HYDROLOGIC MODELING USING GIS Second Edition

by

BAXTER E. VIEUX School of Civil Engineering and Environmental Science, University of Oklahoma, Norman, U.S.A.

KLUWER ACADEMIC PUBLISHERS NEW YORK, BOSTON, DORDRECHT, LONDON, MOSCOW

CD-ROM available only in print edition eBook ISBN: 1-4020-2460-6 Print ISBN: 1-4020-2459-2

©2005 Springer Science + Business Media, Inc.

Print ©2004 Kluwer Academic Publishers Dordrecht All rights reserved

No part of this eBook may be reproduced or transmitted in any form or by any means, electronic, mechanical, recording, or otherwise, without written consent from the Publisher

Created in the United States of America

Visit Springer's eBookstore at: and the Springer Global Website Online at:

http://ebooks.springerlink.com http://www.springeronline.com

Dedication

This book is dedicated to my wife, Jean and to our children, William, Ellen, Laura, Anne, and Kimberly, and to my parents.

This page intentionally left blank

Contents

Dedication

v

Preface

xi

Foreword

xv

Acknowledgments 1 DISTRIBUTED HYDROLOGIC MODELING 1.1 INTRODUCTION 1.2 WHY DISTRIBUTED HYDROLOGIC MODELING? 1.3 DISTRIBUTED MODEL REPRESENTATION 1.4 MATHEMATICAL ANALOGY 1.5 GIS DATA STRUCTURES AND SOURCES 1.6 SURFACE GENERATION 1.7 SPATIAL RESOLUTION AND INFORMATION CONTENT 1.8 RUNOFF PROCESSES 1.9 HYDRAULIC ROUGHNESS 1.10 DRAINAGE NETWORKS AND RESOLUTION 1.11 SPATIALLY VARIABLE PRECIPITATION 1.12 DISTRIBUTED HYDROLOGIC MODEL FORMULATION 1.13 DISTRIBUTED MODEL CALIBRATION 1.14 CASE STUDIES 1.15 HYDROLOGIC ANALYSIS AND PREDICTION 1.16 SUMMARY 1.17 REFERENCES

xvii 1 1 2 5 8 9 10 10 11 14 15 15 16 16 17 18 18 19

viii

Distributed Hydrologic Modeling Using GIS

2 DATA SOURCES AND STRUCTURE 1.1 INTRODUCTION 1.2 DIMENSIONALITY 1.3 MAP SCALE AND SPATIAL DETAIL 1.4 DATUM AND SCALE 1.5 GEOREFERENCED COORDINATE SYSTEMS 1.6 MAP PROJECTIONS 1.7 DATA REPRESENTATION 1.8 WATERSHED DELINEATION 1.9 SOIL CLASSIFICATION 1.10 LAND USE/COVER CLASSIFICATION 1.11 SUMMARY 1.12 REFERENCES

21 21 23 23 24 26 26 31 37 42 43 45 46

3 SURFACE GENERATION 1.1 INTRODUCTION 1.2 SURFACE GENERATORS 1.3 SURFACE GENERATION APPLICATION 1.4 SUMMARY 1.5 REFERENCES

47 48 49 66 70 71

4 SPATIAL VARIABILITY 1.1 INTRODUCTION 1.2 INFORMATION CONTENT 1.3 FRACTAL INTERPRETATION 1.4 RESOLUTION EFFECTS ON DEMS 1.5 SUMMARY 1.6 REFERENCES

73 74 78 80 82 88 89

5 INFILTRATION MODELING 1.1 INTRODUCTION 1.2 INFILTRATION PROCESS 1.3 APPROACHES TO INFILTRATION MODELING 1.4 GREEN-AMPT THEORY 1.5 ESTIMATION OF GREEN-AMPT PARAMETERS 1.6 ATTRIBUTE ERROR 1.7 SUMMARY 1.8 REFERENCES

91 92 93 93 101 103 108 111 111

6 HYDRAULIC ROUGHNESS 1.1 INTRODUCTION 1.2 HYDRAULICS OF SURFACE RUNOFF 1.3 APPLICATION TO THE ILLINOIS RIVER BASIN

115 116 117 123

Distributed Hydrologic Modeling Using GIS 1.4 1.5

SUMMARY REFERENCES

ix 127 127

7 DIGITAL TERRAIN 1.1 INTRODUCTION 1.2 DRAINAGE NETWORK 1.3 DEFINITION OF CHANNEL NETWORKS 1.4 RESOLUTION DEPENDENT EFFECTS 1.5 CONSTRAINING DRAINAGE DIRECTION 1.6 SUMMARY 1.7 REFERENCES

129 129 130 135 138 141 145 146

8 PRECIPITATION MEASUREMENT 1.1 INTRODUCTION 1.2 RAIN GAUGE ESTIMATION OF RAINFALL 1.3 RADAR ESTIMATION OF PRECIPITATION 1.4 WSR-88D RADAR CHARACTERISTICS 1.5 INPUT FOR HYDROLOGIC MODELING 1.6 SUMMARY 1.7 REFERENCES

149 149 151 155 167 172 174 175

9 FINITE ELEMENT MODELING 1.1 INTRODUCTION 1.2 MATHEMATICAL FORMULATION 1.3 SUMMARY 1.4 REFERENCES

177 177 182 194 195

10 DISTRIBUTED MODEL CALIBRATION 1.1 INTRODUCTION 1.2 CALIBRATION APPROACH 1.3 DISTRIBUTED MODEL CALIBRATION 1.4 AUTOMATIC CALIBRATION 1.5 SUMMARY 1.6 REFERENCES

197 197 199 201 208 214 214

11 DISTRIBUTED HYDROLOGIC MODELING 1.1 INTRODUCTION 1.2 CASE STUDIES 1.3 SUMMARY 1.4 REFERENCES

217 218 218 236 237

12 HYDROLOGIC ANALYSIS AND PREDICTION 1.1 INTRODUCTION

239 239

x

Distributed Hydrologic Modeling Using GIS 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9

VFLO™ EDITIONS VFLO™ FEATURES AND MODULES MODEL FEATURE SUMMARY VFLO™ REAL-TIME DATA REQUIREMENTS RELATIONSHIP TO OTHER MODELS SUMMARY REFERENCES

241 242 245 256 258 259 260 260

Glossary

263

Index

287

Preface

Distributed modeling is becoming a more commonplace approach to hydrology. During ten years serving with the USDA Soil Conservation Service (SCS), now known as the Natural Resources Conservation Service (NRCS), I became interested in how millions of dollars in construction contract monies were spent based on simplistic hydrologic models. As a project engineer in western Kansas, I was responsible for building flood control dams (authorized under Public Law 566) in the Wet Walnut River watershed. This watershed is within the Arkansas-Red River basin, as is the Illinois River basin referred to extensively in this book. After building nearly 18 of these structures, I became Assistant State Engineer in Michigan and, for a short time, State Engineer for NRCS. Again, we based our entire design and construction program on simplified relationships variously referred to as the SCS method. I recall announcing that I was going to pursue a doctoral degree and develop a new hydrologic model. One of my agency’s chief engineers remarked, “Oh no, not another model!” Since then, I hope that I have not built just another model but have significantly advanced the state of hydrologic modeling. This book sets out principles for modeling hydrologic processes distributed in space and time using the geographic information system (GIS), a spatial data management tool. Any hydrologic model is an abstract representation of a component of a natural process. The science and engineering aspects of hydrology have been long clouded by gross simplifications. Representation by lumping of parameters at the river basin scale such that a single value of slope or hydraulic roughness controls the basin response may have served well when computer resources were limited and spatial datasets of soils, topography, landuse, and precipitation did not

xii

Distributed Hydrologic Modeling Using GIS

exist. Shrugging off these assumptions in favor of more representative modeling will undoubtedly advance the science of hydrology. To advance from lumped to distributed representations requires reexamination of how we model for both engineering purposes and for scientific understanding. We could reasonably ask what laws govern the complexities of all the paths that water travels, from precipitation falling over a river basin to the flow in the river. We have no reason to believe that each unit of water mass is not guided by Newtonian mechanics, making conservation laws of momentum, mass, and energy applicable. It is my conviction that hydrologists charged with making predictions will opt for distributed representations if it can be shown that distributed models give better results. No real advance will be made if we continue to force lumped models based on empirical relationships to represent the complexity of distributed runoff. Once we embark on fully distributed representations of hydrologic processes, we have no other choice than to use conservation laws (termed “physics-based”) as governing equations. What was inconceivable a decade ago is now commonplace in terms of computational power and spatial data management systems that support detailed mathematical modeling of complex hydrologic processes. Technology has enabled the transformation of hydrologic modeling from lumped to distributed representations with the advent of new sensor systems such as radar and satellite, high performance computing, and orders-ofmagnitude increases in storage. Global digital datasets of elevation at thirty meters (or smaller) or soil moisture estimates from satellite and data assimilation offer tantalizing detail that could be of use in making better predictions or estimates of the extremes of weather, drought, and flooding. When confronted with the daunting task of modeling a natural process in uncontrolled non-laboratory conditions, the academic ranks are usually illequipped because neat disciplinary boundaries divide and subdivide the domain. In reality, water does not care whether it is flowing through the meteorologist’s domain or that of the soil scientist’s. Thus, any realistic treatment of hydrology necessarily taps the ingenuity and scientific understanding of a wide number of disciplines. Distributed hydrologic modeling requires disciplinary input from meteorology and electrical engineering in order to derive meaningful precipitation input from radar remote sensing of the atmosphere. Infiltration is controlled by soil properties and profile depth, which is the domain of the soil scientist, who most often is employed by an agricultural agency responsible for mapping soils. Managing spatial information using GIS requires aspects of geographic projections to map and overlay parameters and inputs needed in the model. Indeed, most land use/cover maps were not compiled for hydrologic purposes. An understanding of the origin and techniques used to map the

Distributed Hydrologic Modeling Using GIS

xiii

land use/cover is required in order to transform such datasets into useable hydrologic parameters. Computationally, numerical methods are used to solve the governing conservation equations. Finite difference and finite element methods applied to hydrology require data management tools such as GIS. If a GIS is used to supply parameters and input to these computational algorithms, then the interface between data structures of the spatial data and those in the numerical algorithm must be understood. Filling in the gaps between academic disciplines is necessary for a credible attempt at hydrologic modeling. Thus, the physical geographer who is involved in modeling river basin response to heavy rainfall for purposes of studying how floods impact society would likely benefit from seeing in this book how geographical analysis and datasets may be transformed from thematic maps into model input. A meteorologist who wishes to gain a clearer understanding of how terrestrial features transform rainfall into runoff from hillslope to river basin scale will gain a better appreciation for aspects of spatial and temporal scale, precision, and data format and their importance in using radar inputs to river basin models. Soil scientists who wish to map soils according to hydrologic performance rather than solely as aids to agricultural production would also likely benefit, especially from the chapters dealing with infiltration, model calibration, and the case studies. Several options exist for writing about GIS and hydrology. One choice would be to weight the book heavily in favor of GIS commands and techniques for specific software packages. Such books quickly become outdated as the software evolves or falls into disfavor with the user community. A more balanced choice is to focus on distributed hydrology with principles on how to implement a model of hydrologic processes using GIS. As the subject emerged during the writing of this book, it became clear that there were issues with GIS data formats, spatial interpolation, and resolution effects on information content and drainage network that could not be omitted. Included here are fewer examples of specific GIS commands or software operation. However, the focus is to illustrate how to represent adequately the spatially distributed data for hydrologic modeling along with the many pitfalls inherent in such an undertaking. Many of the details of how to accomplish the operations specific to various GIS software packages are left to other books. This book is not intended to be a survey of existing models or a GIS software manual, but rather a coherent treatment by a single author setting forth guiding principles on how to parameterize a distributed hydrologic model using GIS. Worldwide geospatial data has become readily available in GIS format. A modeling approach that can utilize this data for hydrology offers many possibilities. I expect those interested in smaller or larger scales or other hydrologic components will be able to apply many of the principles

xiv

Distributed Hydrologic Modeling Using GIS

presented herein. For this reason, I beg your indulgence for my narrow approach. It is my hope that this monograph benefits those hydrologists interested in distributed approaches to hydrologic modeling. Since the First Edition, software development and applications have created a richer set of examples, and a deeper understanding of how to perform distributed hydrologic analysis and prediction. This Second Edition is oriented towards a recent commercially available distributed model called Vflo™. The basic edition of this model is included on the enclosed CDROM. Baxter E. Vieux, Ph.D., P.E. Presidential Professor School of Civil Engineering and Environmental Science University of Oklahoma Norman, Oklahoma, USA

Foreword

‘Distributed Hydrologic Modeling Using GIS’ celebrates the beginning of a new era in hydrologic modeling. The debate surrounding the choice of either lumped or distributed parameter models in hydrology has been a long one. The increased availability of sufficiently detailed spatial data and faster, more powerful computers has leveled the playing field between these two basic approaches. The distributed parameter approach allows the hydrologist to develop models that make full use of such new datasets as radar rainfall and high-resolution digital elevation models (DEMs). The combination of this approach with Geographic Information Systems (GIS) software, has allowed for reduced computation times, increased data handling and analysis capability, and improved results and data display. 21st century hydrologists must be familiar with the distributed parameter approach as the spatial and temporal resolution of digital hydrologic data continues to improve. Additionally, a thorough understanding is required of how this data is handled, analyzed, and displayed at each step of hydrologic model development. It is in this manner that this book is unique. First, it addresses all of the latest technology in the area of hydrologic modeling, including Doppler radar, DEMs, GIS, and distributed hydrologic modeling. Second, it is written with the intention of arming the modeler with the knowledge required to apply these new technologies properly. In a clear and concise manner, it combines topics from different scientific disciplines into a unified approach aiming to guide the reader through the requirements, strengths, and pitfalls of distributed modeling. Chapters include excellent discussion of theory, data analysis, and application, along with several cross references for further review and useful conclusions.

xvi

Foreword

This book tackles some of the most pressing concerns of distributed hydrologic modeling such as: What are the hydrologic consequences of different interpolation methods? How does one choose the data resolution necessary to capture the spatial variability of your study area while maintaining feasibility and minimizing computation time? What is the effect of DEM grid resampling on the hydrologic response of the model? When is a parameter variation significant? What are the key aspects of the distributed model calibration process? In ‘Distributed Hydrologic Modeling Using GIS’, Dr. Vieux has distilled years of academic and professional experience in radar rainfall applications, GIS, numerical methods and hydrologic modeling into one single, comprehensive text. The reader will not only gain an appreciation for the changes brought about by recent technological advances in the hydrologic modeling arena, but will fully understand how to successfully apply these changes toward better hydrologic model generation. ‘Distributed Hydrologic Modeling Using GIS’ not only sets guiding principles to distributed hydrologic modeling, but also asks the reader to respond to new developments and calls for additional research in specific areas. All of the above make this a unique, invaluable book for the student, professor, or hydrologist seeking to acquire a thorough understanding of this area of hydrology. Philip B. Bedient Herman Brown Professor of Engineering Department of Civil and Environmental Engineering Rice University Houston, Texas, USA

Acknowledgments

I wish to thank my colleagues who contributed greatly to the writing of the First and Second Editions of this book. I am indebted to Professor Emeritus, Jacques W. Delleur, School of Civil Engineering, Purdue University, for his review; and to Philip B. Bedient, and his students, in the School of Civil and Environmental Engineering, Rice University, for their continued and helpful suggestions and insights, which improved this book substantially. I wish to thank my own students who have lent their time and energies to distributed hydrologic modeling using GIS. Over the course of many years, I have enjoyed collaborations with colleagues that have encouraged the development and application of distributed modeling. In particular, I am indebted to Bernard Cappelaere, Thierry Lebel, and others with l’Institut de Recherche pour le Développement (IRD), France. To my colleagues at the Disaster Prevention Research Institute, Kyoto Japan; Yasuto Tachikawa, Eichi Nakakita and others, I am indebted. During the writing of the First Edition, I enjoyed fruitful discussions and support from the NOAA-National Severe Storms Laboratory (NSSL), and the Cooperative Institute for Mesoscale Meteorological Studies (CIMMS). I wish to thank, Kenneth Howard and Jonathan J. Gourley with NSSL, and Professor Peter Lamb, School of Meteorology, Director of CIMMS, University of Oklahoma, who have helped promote the application of radar for hydrologic applications. Special thanks go to Ryan Hoes, Eddie Koehler of Vieux & Associates, Inc.; and especially to Jean E. Vieux, CEO/President, for her confidence, assistance, and support. The editing assistance of Carolyn Ahern and Daphne Summers improved the text immensely.

This page intentionally left blank

Chapter 1 DISTRIBUTED HYDROLOGIC MODELING

1.1

Introduction

An ongoing debate within the hydrology community, both practitioners and researchers, is how to construct a model that best represents the Earth’s hydrologic processes. Distributed models are becoming commonplace in a variety of applications. Through revision of existing models along with new model development, hydrology is striving to keep pace with the explosive growth of online geospatial data sources, remote sensing, and radar technology for measurement of precipitation. When geospatial data is used in hydrologic modeling, previously unfamiliar issues may arise. It is not surprising that Geographic Information Systems (GIS) have become an integral part of hydrologic studies considering the spatial character of parameters and precipitation controlling hydrologic processes. The primary motivation for this book is to bring together the key ingredients necessary to use GIS to model hydrologic processes, i.e., the spatial and temporal distribution of the inputs and parameters controlling surface runoff. GIS maps describing topography, land use and cover, soils, rainfall, and meteorological variables become model parameters or inputs in the simulation of hydrologic processes. Difficulties in managing and efficiently using spatial information have prompted hydrologists either to abandon it in favor of lumped models or to develop more sophisticated technology for managing geospatial data (Desconnets et al., 1996). As soon as we embark on the simulation of hydrologic processes using GIS, the issues that are the subject of this book must be addressed.

Chapter 1

2 1.2

Why Distributed Hydrologic Modeling?

Historical practice has been to use lumped representations because of computational limitations or because sufficient data was not available to populate a distributed model database. How one represents the process in the mathematical analogy and implements it in the hydrologic model determines the degree to which we classify a model as lumped or distributed. Several distinctions on the degree of lumping can be made in order to better characterize a mathematical model, the parameters/input, and the model implementation. Whether representation of hydrologically homogeneous areas can be justified depends on how uniform the spatially variable parameters are. For example, the City of Cherokee, Oklahoma suffers repeated flooding when storms having return intervals of approximately 2-year frequency occur on Cottonwood Creek (Figure 1-1). A lumped subbasin approach using HECHMS (HEC, 2000) is represented schematically in Figure 1-2. ‘Junction-2’ is located where the creek crosses Highway 64 on the northwestern outskirts of the city limits. Each subbasin must be assigned a set of parameters controlling the hydrologic response to rainfall input.

Figure 1-1. Contour map of the City of Cherokee in northwestern Oklahoma and Cottonwood Creek draining through town.

1. DISTRIBUTED HYDROLOGIC MODELING

3

Though contour lines are the traditional way of mapping topography, distributed hydrologic modeling requires a digital elevation model. The Cottonwood basin represented using a 60-m resolution digital elevation model is seen in Figure 1-3. Considerable variation in the topographic relief is evident in the upper portions of the watershed where relatively flat terrain breaks into steep areas; from there the terrain becomes flatter in the lower portions of the watershed near the town. A distributed approach to modeling this watershed would consist of a grid representation of topography, precipitation, soils, and land use/cover that accounts for the variability of all these parameters. Lumping even at the subbasin level would not be able to account for the change in slope and drainage network affecting the hydrologic response of the basin.

Figure 1-2. HEC-HMS subbasin definitions for the 125 km2 Cottonwood Creek.

422

Chapter 1

Figure 1-3. Hillshade digital elevation model and road network of the Cottonwood Creek watershed and the City of Cherokee (upper right).

Practitioners are beginning to profit from research and development of distributed hydrology (ASCE, 1999). As distributed hydrologic models become more widely used in practice, the need for scientific principles relating to spatial variability, temporal and spatial resolution, information content, and calibration become more apparent. Whether a model is lumped or distributed depends on whether the domain is subdivided. It is clear that this distinction is relative to the domain. If the watershed domain is to be distributed, the model must subdivide the watershed into smaller computational elements. This process often gives rise to lumped subbasin models that attempt to represent spatially variable parameters/conditions as a series of subbasins with average characteristics. In this manner, almost any lumped model can be turned into a semidistributed model. Most often, such lumping results in an empirically-based model, because conservation equations break down at the scale of the subbasin. Subbasin lumping is an outgrowth of the concept of hydrologically homogeneous subareas. This concept arises from overlaying areas of soil, land use/cover, and slope attributes producing subbasins of homogeneous parameters. Subbasins then could logically be lumped at this level. Drawbacks associated with subbasin lumping include: 1. The resulting model may not be physics-based

235

1. DISTRIBUTED HYDROLOGIC MODELING

2. Deriving parameters at the scale of subbasins is difficult because streamflow is not available at each outlet 3. Model performance may be affected by the number of subbasins 4. Parameter variability is not properly represented by lumping at the subbasin scale Subbasin lumping can cause unexpected parameter interaction and degraded model performance as the number of subbasins are changed. 1.3

Distributed Model Representation

It is useful to consider how physics-based distributed (PBD) models fit within the larger context of hydrologic modeling. Figure 1-4 shows a schematic for classifying a deterministic model of a river basin. Deterministic River Basin Model

PhysicsBased

Runoff Generation

Distributed Parameter

Conceptual

Runoff Routing

Runoff Generation

Runoff Routing

Lumped Parameter

Distributed Parameter

Lumped Parameter

Figure 1-4. Model classification according to distributed versus lumped treatment of parameters.

624

Chapter 1

Deterministic is distinguished from stochastic in that a deterministic river basin model estimates the response to an input using either a conceptual mathematical representation or a physics-based equation. Conceptual representations usually rely on some type of linear reservoir theory to delay and attenuate the routing of runoff generated. Runoff generation and routing are not closely linked and therefore do not interact. Physics-based models use equations of conservation of mass, momentum, and energy to represent both runoff generation and routing in a linked manner. Following the lefthand branch in the tree, the distinction between runoff generation and runoff routing is somewhat artificial, because they are intimately linked in most distributed model implementations. However, by making a distinction we can introduce the idea of lumped versus distributed parameterization for both overland flow and channel flow. A further distinction is whether overland flow or subsurface flow is modeled with lumped or distributed parameters. Routing flow through the channels using lumped or distributed parameters distinguishes whether uniform or spatially variable parameters are applied in a given stream segment. Hybrids between the branches in Figure 1-4 exist. For example, the model TOPMODEL (Beven and Kirkby, 1979) simulates flow through the range of hillslope parameters found in a watershed. The spatial arrangement is not taken into account, only the statistical distribution of index values, in order to develop a basin response function. It is a semi-distributed model since the statistics of the spatially variable parameters are operated on without regard to location. TOPMODEL falls somewhere between conceptual and distributed, though with some physical basis. Changing time steps of the model input amounts to lumping, can influence the PBD models significantly depending on the size of the basin. Unit-hydrograph approaches are based on rainfall accumulations and to a lesser degree on intensity. Temporal lumping occurs with aggregation over time of such phenomena as stream flow or rainfall accumulations at 5minute, hourly, daily, 10-day, monthly, or annual time series. Hydrologic models driven by intensities rather than accumulations can be more sensitive to temporal resolution. Scale is an issue where a small watershed may be sensitive to rainfall time series at 5-minute intervals, whereas a large river basin may be sensitive to only hourly or longer time steps. The spatial resolution used to represent spatially variable parameters is another form of lumping. Changing spatial resolution of datasets requires some scheme to aggregate parameter values at one resolution to another. Resampling is essentially a lumping process, which in the limit, results in a single value for the spatial domain. Resampling a parameter map involves taking the value at the center of the larger cell, averaging, or other operation.

1. DISTRIBUTED HYDROLOGIC MODELING

257

If the center of the larger cell happens to fall on low/high value, then a large cell area will have a low/high value. Resampling rainfall maps can produce erratic results as the resolution increases in size, as found by Vieux and Farajalla (1996). For the basin and storms tested, as the resolution exceeded 3 km, the simulated hydrograph became erratic because of the resampling effect. Farajalla and Vieux (1995) and Vieux and Farajalla (1994) applied information entropy to infiltration parameters and hydraulic roughness to discover the limiting resolution beyond which little more was added in terms of information. Over-sampling a parameter or input map at finer resolution may not add any more information, either because the map, or the physical feature, does not contain additional information. Of course, variations exist physically; however, these variations may not have an impact at the scale of the modeled domain. How to determine which resolution is adequate for capturing the essential information contained in a parameter map for simulating the hydrologic process is taken up in Chapter 4. Numerical solution of the governing equations in a physics-based model employs discrete elements. The three representative types are finite difference, finite element, and stream tubes. At the level of a computational element, a parameter is regarded as being representative of an average process. Thus, some average property is only valid over the computational element used to represent the runoff process. For example, porosity is a property of the soil medium, but it has little meaning at the level of the pore space itself. Thus, resolution also depends on how well a single value represents a grid cell. From a model perspective, a parameter should be representative of the surface or medium at the scale of the computational element used to solve the governing mathematical equations. This precept is often exaggerated as the modeler selects coarser grid cells, losing physical significance. In other words, runoff depth in a grid cell of 1-km resolution can only be taken as a generalization of the actual runoff process and may or may not produce physically realistic model results. Computational resources are easily exceeded when modeling large basins at fine resolution, motivating the need for coarser model resolution. At coarser resolution, the sub-grid scale processes take on more importance. One of the great questions facing operational use of distributed hydrologic models for large river basins is how to parameterize the sub-grid processes. At the scale of more than a few meters in resolution, runoff depth and velocity do not have strict physical significance. Depending on the areal extent of a river basin and the spatial variability inherent in each parameter, small variations may not be important while other variations may exercise a strong influence on model performance.

826 1.4

Chapter 1 Mathematical Analogy

Physics-based distributed (PBD) models solve governing equations derived from conservation of mass, momentum, and energy. Unlike empirically based models, differential equations are used to describe the flow of water over the land surface or through porous media, or energy balance in the exchange of water vapor through evapotranspiration. In most physics-based models, simplifications are made to the governing equations because certain gradients may not be important or accompanying parameters, boundary and initial conditions are not known. Linearization of the differential equations is also attractive because nonlinear equations may be difficult to solve. The resulting mathematical analogies are simplifications of the complete form. The full dynamic equations describing the flow of water over the land surface or in a channel may contain gradients that are negligible under certain conditions. In a mathematical analogy, we discard the terms in the equations that are orders of magnitudes less than the others are. Simplifications of the full dynamic governing equations give rise to zero inertial, kinematic, and diffusive wave analogies. Using simplified or full dynamic mathematical analogies to generate flow rates is a hydraulic approach to hydrology. Using such conservation laws provides the basis in physics for fully distributed models. If the physical character of the hydrologic process is not supported by a particular analogy, then errors result in the physical representation. Difficulties also arise from the simplifications because the terms discarded may have afforded a complete solution while their absence causes mathematical discontinuities. This is particularly true in the kinematic wave analogy, in which changes in parameter values can cause discontinuities, sometimes referred to as shock, in the equation solution. Special treatment is required to achieve solution to the kinematic wave analogy of runoff over a spatially variable surface. Vieux (1988), Vieux et al. (1990) and Vieux (1991) found such a solution using nodal values of parameters in a finite element solution. This method effectively treats changes in parameter values by interpolating their values across finite elements. The advantage of this approach is that the kinematic wave analogy can be applied to a spatially variable surface without numerical difficulty introduced by the shocks that would otherwise propagate through the system. Vieux and Gauer (1994) presented a distributed watershed model based on this nodal solution using finite elements to represent the drainage network called r.water.fea. Chapter 9 presents a detailed description of the finite element solution to the kinematic wave equations. This second edition presents a recent distributed hydrologic model called Vflo™ that uses finite elements in space and finite difference in time. The kinematic wave analogy is useful in

1. DISTRIBUTED HYDROLOGIC MODELING

279

watersheds where backwater is not important. Such watersheds are usually in the upper reaches of major river basins where topographic gradients dominate flow velocities. Vflo™ is a commercially available distributed hydrologic model (Vieux and Vieux, 2002). The diffusive wave analogy is necessary where backwater effects are important. This is usually in flatter watersheds or low-gradient river systems. CASC2D (Julien and Saghafian, 1991; Julien, et al., 1995) uses the diffusive wave analogy to simulate flow in a grid-cell (raster) representation of a watershed. The US Army Corps of Engineers Engineering Research and Development Center (USACE ERDC) has supported development of the Gridded Surface Subsurface Hydrologic Analysis model (GSSHA). The GSSHA model extends the applicability of the CASC2D model to handle surface-subsurface interactions associated with saturation excess runoff (non-Hortonian). These models solve the diffusive wave analogy using a finite difference grid corresponding to the grid-cell representation of the watershed. The diffusive wave analogy requires additional boundary conditions to obtain a numerical solution in the form of supplying a gradient term at boundaries or other locations. The models described herein, r.water.fea and Vflo™; use a less complex mathematical analogy, the Kinematic Wave Analogy, to represent hydraulic conditions in a watershed. The following sections outline the contents of each chapter as it relates to distributed modeling using GIS. 1.5

GIS Data Structures and Sources

New sources of geographic data, often in easily available global datasets, offer tantalizing detail if only they could be used in a hydrologic model designed to take advantage of the tremendous information content. Hydrologic models are now available that are designed to use geospatial data effectively. Once a particular spatial data source is considered for use in a hydrologic model, and then we must consider the data structure, file format, quantization (precision), and error propagation. GIS offers efficient algorithms for dealing with most of this geospatial data. However, the relevance of the particular geospatial data to hydrologic modeling is often not known without special studies to test whether a new data source provides advantages that merit its use. Chapter 2 deals with the major data types necessary for distributed hydrologic modeling. Depending on the particular watershed characteristics, many types of data may require processing before they can be used in a hydrologic model.

28 10 1.6

Chapter 1 Surface Generation

Digital representation of terrain requires that a surface be modeled as a set of elevations or other terrain attributes at point locations. Much work has been done in the area of spatial statistics and the development of Kriging techniques to generate surfaces from point data. In fact, several methods for generating a two-dimensional surface from point data may be enumerated: • Linear interpolation • Local regression • Distance weighting • Moving average • Splines • Kriging The problem with all of these methods when applied to geophysical fields such as rainfall, ground water flow, wind, temperature, or soil properties is that the interpolation algorithm may violate some physical aspect. Gradients may be introduced that are a function of the sparseness of the data and/or the interpolation algorithm. Values may be interpolated across distinct zones where natural discontinuities exist. Suppose, for example, that several piezometric levels are measured over an area, and that we wish to generate a surface representative of the piezometric levels or elevations within the aquifer. Using an inverse distance-weighting scheme, we interpolate elevation in a raster array. We will almost certainly generate a surface that has artifacts of interpolation that violate physical characteristics, viz., gradients are introduced that would indicate flow in directions contrary to the known gradients or flow directions in the aquifer. In fact, a literal interpretation of the interpolated surface may indicate that, at each measured point, pressure decreases in a radial direction away from the well location, which is clearly not the case. None of the above methods of surface interpolation is entirely satisfactory when it comes to ensuring physical correctness in the interpolated surface. Depending on the sampling interval, spatial variability, physical characteristics of the measure, and the interpolation method, the contrariness of the surface to physical or constitutive laws may not be apparent until model results reveal intrinsic errors introduced by the surface generation algorithm. Chapter 3 deals with surface interpolation and hydrologic consequences of interpolation methods. 1.7

Spatial Resolution and Information Content

How resolution in space affects hydrologic modeling is of primary importance. The resolution that is necessary to capture the spatial variability

1. DISTRIBUTED HYDROLOGIC MODELING

29 11

is often not addressed in favor of simply using the finest resolution possible. It makes little sense, however, to waste computer resources when a coarser resolution would suffice. We wish to know the resolution that adequately samples the spatial variation in terms of the effects on the hydrologic model and at the scale of interest. This resolution may be coarser than that dictated by visual esthetics of the surface at fine resolution. The question of which resolution suffices for hydrologic purposes is answered in part by testing the quantity of information contained in a data set as a function of resolution. We can stop resampling at coarser resolution once the information content begins to decrease or be lost. Information entropy, originally developed by communication engineers, can test which resolution is adequate in capturing the spatial variability of the data (Vieux, 1993). We can relate the information content to model performance effects. For example, resampling rainfall at coarser resolution and inputting this into a distributed hydrologic model can produce erratic hydrologic model response (Vieux and Farajalla, 1996). Chapter 4 provides an overview of information theory with an application showing how information entropy is descriptive of spatial variability and its use as a statistical measure of resolution impacts on hydrologic parameters such as slope. 1.8

Runoff Processes

Two basic flow types can be recognized: overland flow, conceptualized as thin sheet flow before the runoff concentrates in recognized channels, and channel flow, conceptualized as occurring in recognized channels with hydraulic characteristics governing flow depth and velocity. Overland flow is the result of rainfall rates exceeding the infiltration rate of the soil. Depending on soil type, topography, and climatic factors, surface runoff may be generated either as infiltration excess, saturation excess, or as a combination. Infiltration is a major determinant of how much rainfall becomes runoff. Therefore, estimating infiltration parameters from soil maps and associated databases is important for quantifying infiltration at the watershed scale. 1.8.1

Infiltration Excess (Hortonian)

Infiltration excess first identified by Horton is typical in areas where the soils have low infiltration rates and/or the soil is bare. Raindrops striking bare soil surfaces break up soil aggregates, allowing fine particles to clog surface pores. A soil crust of low infiltration rate results particularly where vegetative cover has been removed due to urban construction, farming, or fire. Infiltration excess is generally conceptualized as flow over the surface

30 12

Chapter 1

in thin sheets. Model representation of overland flow uses this concept of uniform depth over a computational element though it differs from reality, where small rivulets and drainage swales convey runoff to the major stream channels. Figure 1-5 shows two zones, one where rainfall, R, exceeds infiltration I (R>I), the other where R < I. In the former, runoff occurs; in the latter, rainfall is infiltrated, and infiltration excess runoff does not occur. However, the amount of infiltrated rainfall may contribute to the watertable, subsurface conditions permitting. Figure 1-5 is a simplified representation. From hill slope to stream channel, there may be areas of infiltration excess which runs on to areas where the combination of rainfall and run-on from upslope does not exceed the infiltration rate, resulting in losses to the subsurface.

Figure 1-5. Schematic diagram of runoff produced by infiltration excess.

Simulation of infiltration excess requires soil properties and initial soil moisture conditions. Figure 1-6 shows two plots: rainfall intensity as impulses, and infiltration rate as smoothly decreasing with time. Simulation of infiltration over a watershed is complicated because it depends on the rainfall, soil properties, and antecedent soil moisture at every location or grid cell. The infiltration rate calculated from soil properties is a potential rate that depends on the initial degree of saturation. Richards’ equation fully describes this process using principles of conservation of mass and momentum. The Green and Ampt equation (Green and Ampt, 1911) is a simplification of Richards’ equation that assumes piston flow (no diffusion). Loague (1988) found that the spatial arrangement

31 13

1. DISTRIBUTED HYDROLOGIC MODELING

Rainfall Intensity (cm/hr)

6

9 8 7 6 5 4 3 2 1 0

5 4 3 2 1 0 0

1

2

3

4

5

Infiltration Rate (cm/hr)

of soil hydraulic properties at hillslope scales (< 100 m) was more important than rainfall variations. Order-of-magnitude variation in hydraulic conductivity at length scales on the order of 10 m controlled the runoff response. This would seem to say that infiltration rates at the river-basin scale is impossible to know unless very detailed spatial patterns of soil properties are measured. The other possible conclusion is that not all of this variability is important over large areas. Considering that detailed infiltration measurement and soil sampling is not economically feasible over large spatial extent, deriving infiltration rates from soil maps is an attractive alternative. Modeling infiltration excess at the watershed scale is more feasible if infiltration parameters can be estimated from mapped soil properties.

Rain Infil

6

Time (hr) Figure 1-6. Infiltration excess modeled using the Horton equation.

1.8.2

Saturation Excess (Dunne Type)

Saturation excess runoff is common in mountainous terrain or watersheds with highly porous surfaces (Dunne et al., 1975). Under these conditions, overland flow may not be observed. Runoff occurs by infiltrating to a shallow watertable. As the gradient of the watertable increases, runoff to stream channels also increases. As the watertable surface intersects the ground surface areas adjacent to the stream channel, the surface saturates. As

32 14

Chapter 1

the saturation zone grows in areal extent and rain falls on this area, more runoff occurs. Figure 1-7 shows the location of saturation excess next to a stream channel. Representing this type of runoff process requires information about the soil depth and hydraulic properties affecting the velocity of water moving through the subsurface. Infiltration modeling that relies on soil properties to derive the Green and Ampt equations is considered in Chapter 5. 1.9

Hydraulic Roughness

Accounting for overland and channel flow hydraulics over the watershed helps our ability to simulate hydrographs at the outlet. In rural and urban areas, hydraulics govern flow over artificial and natural surfaces. Frictional drag over the soil surface, standing vegetative material, crop residue, and rocks lying on the surface, raindrop impact, and other factors influence the hydraulic resistance experienced by runoff. Hydraulic roughness coefficients caused by each of these factors contribute to total hydraulic resistance.

Figure 1-7. Schematic diagram of runoff produced by saturation excess.

Detailed measurement of hydraulic roughness over any large spatial extent is generally impractical. Thus, reclassifying a GIS map of land use/cover into a map of hydraulic roughness parameters is attractive in spite of the errors present in such an operation. Considering that hydraulic roughness is a property that is characteristic of land use/cover classification,

1. DISTRIBUTED HYDROLOGIC MODELING

33 15

hydraulic roughness maps can be derived from a variety of sources. Aerial photography, land use/cover maps, and remote sensing of vegetative cover, become a source of spatially distributed hydraulic roughness. Each of these sources lets us establish hydraulic roughness over broad areas such as river basins or urban areas with both natural and artificial surfaces. The goal of reclassification of a landuse/cover map is to represent the location of hydraulically rough versus smooth land use types for watershed for simulation. Chapter 6 deals with the issue of how landuse/cover maps are reclassified into hydraulic roughness, and then used to control how fast runoff moves through the watershed. 1.10

Drainage Networks and Resolution

Drainage networks may be derived from digital elevation models (DEMs) by connecting each cell to its neighbor in the direction of principal slope. DEM resolution has a direct influence on the total drainage length and slope. Too coarse resolution causes an undersampling of the hillslopes and valleys where hilltops are cut off and valleys filled. Two principal effects of increasing the resolution coarseness are: 1. Drainage length is shortened 2. Slope is flattened The effect of drainage length shortening and slope flattening on hydrograph response may be compensating. That is, shorter drainage length accelerates arrival times at the outlet, while flatter slopes delay arrival times. The influence of DEM grid-cell resolution is discussed in Chapter 7. 1.11

Spatially Variable Precipitation

Besides satellite, one of the most important sources of spatially distributed rainfall data is radar. Spatial and temporal distribution of rainfall is the driving force for both infiltration and saturation excess. In the former case, comparing rainfall intensity with soil infiltration rates determines the rate and location of runoff. One of the most hydrologically significant radar systems is the WSR-88D (popularly known as NEXRAD) radar. Beginning in the early 1990’s, this system was deployed by the US National Weather Service (NWS) for surveillance and detection of severe weather. Understanding how this system may be used to produce accurate rainfall estimates provides a foundation for application to hydrologic models. Resolution in space and time, errors, quantizing (precision), and availability in real-time or post-analysis is taken up in Chapter 8. Without spatially distributed precipitation, distributed modeling cannot be accomplished with full efficiency making radar an important source of input.

34 16 1.12

Chapter 1 Distributed Hydrologic Model Formulation

Physics-based distributed hydrologic modeling relies on conservation equations to create a representation of surface runoff. The kinematic wave mathematical analogy may be solved using a network of finite elements connecting grid cells together. Flow direction in each grid cell is used to layout the finite elements. Solving the resulting system of equations defined by the connectivity of the finite elements provides the possibility of hydrograph simulation at any location in the drainage network. The linkage between GIS and the finite element and finite difference algorithms to solve the kinematic wave equations are examined in detail in Chapter 9. Assembly of finite elements representing the drainage network produces a system of equations solved in time. The resulting solution is the hydrograph at selected stream nodes, cumulative infiltration, and runoff depth in each grid cell. The model r.water.fea was described in the first edition of this book. Recent software development has resulted in another finite element model called Vflo™. In this edition, additional material is presented using this model, which is provided in the enclosed CD-ROM along with sample data sets, tutorials, and help files. 1.13

Distributed Model Calibration

Once the assembly of input and parameter maps for a distributed hydrologic model is completed, the model must usually be calibrated or adjusted. The argument that physics-based models do not need calibration presupposes perfect knowledge of the parameter values distributed throughout the watershed, and of the spatially/temporally variable rainfall input. This is clearly not the case. Besides the parameter and input uncertainty, there are resolution dependencies as presented by Vieux et al. (1996) and others. Hydrologists have argued that there are too many degrees of freedom in distributed modeling vis-à-vis the number of observations. This concern does not take into account that if we know the spatial pattern of a parameter, we can adjust its magnitude while preserving the spatial variation. This calibration procedure can be performed manually by applying scalar multipliers or additive constants to parameter maps until the desired match between simulated and observed is obtained. The ordered physicsbased parameter adjustment (OPPA) method described by Vieux and Moreda (2003) is adapted to the particular characteristics of physics-based models. Predictable parameter interaction and identifiable optimum values are hallmarks of the OPPA approach that can be used to produce physically realistic distributed parameter values.

1. DISTRIBUTED HYDROLOGIC MODELING

35 17

Automatic calibration of hydrologic models can be approached by methods that use generic optimization schemes. For example, the shuffled complex evolution (SCE) method described by Duan et al. (1992) has application to empirically-based hydrologic models. Physics-based models have the advantage that there are governing differential equations whose properties may be exploited. This fact may be taken advantage of with manual or automated calibration techniques. The adjoint method has enjoyed success in meteorology in retrieving initial conditions for atmospheric models. The adjoint method in the context of optimal control has application to distributed hydrologic model calibration. Chapter 10 covers both the manual and automatic calibration methods that exploit the properties of the governing equations of physics-based models. 1.14

Case Studies

The case studies presented in Chapter 11 illustrate several aspects of distributed hydrologic modeling using GIS. The case studies provide examples of using a distributed model in both urban and rural areas. Case I demonstrates application of Vflo™ in the 1200 km2 Blue River basin located in South Central Oklahoma. This watershed area is predominantly rural and was the subject of the Distributed Model Intercomparison Project (DMIP) organized by the US National Weather Service. Physics-based models use conservation of mass and momentum, referred to as a hydrodynamic or hydraulic approach to hydrology. As a result, channel hydraulics play an important role in predicting discharge using a PBD model. The benefit of using representative hydraulic cross-sections is demonstrated. Distributed model flood forecasting is described in Case II. In this case study, an example is offered of operational deployment of a physics-based distributed model configured for site-specific flood forecasts in an urban area, Houston Texas. The influence of radar rainfall input uncertainty is illustrated for five events. As with any measurement, uncertainty may be separated into random and systematic errors. Hydrologic prediction accuracy depends heavily on whether the systematic error (bias) in radar rainfall has been removed using rain gauges. Using rain gauge data removes bias in the radar rainfall input. Random error has less impact than systematic error on simulated hydrologic response. Because of this, real-time rain gauge data becomes an important factor affecting the hydrologic prediction accuracy of a distributed flood forecasting system. As illustrated by this case study, without accurate rainfall input, the full efficiency of the distributed model cannot be achieved.

36 18 1.15

Chapter 1 Hydrologic Analysis and Prediction

In this second edition, software development has resulted in availability of a fully distributed physics-based hydrologic model. An integrated network-based hydraulic approach to hydrologic prediction has advantages that make it possible to represent both local and main-stem flows with the same model setup and simultaneously. This integrated approach is used to make hydrologic forecasts for flood warning and water resources management. The model formulation supports prediction at scales from small upland catchments to large river basins. This model is designed to utilize multisensor inputs from radar, satellite, and rain gauge precipitation measurements. Continuous simulations including soil moisture support operational applications. Post-analysis of storm events allows calibration and hydrologic analysis using archived radar rainfall. Advances in modeling techniques; multisensor precipitation estimation; and secure client/server architecture in JAVA™, GIS and remotely sensed data have resulted in enhanced ability to make hydrologic predictions at any location. This model and the modeling approach described in this book represent a paradigm shift from traditional hydrologic modeling. Chapter 12 describes the Vflo™ model features and application for hydrologic prediction and analysis. The enclosed CD-ROM contains the Vflo™ software with Help and Tutorial files that are useful in understanding how distributed hydrologic modeling is performed. 1.16

Summary

While this book answers questions related to distributed modeling, it also raises others on how best to model distributed hydrologic processes using GIS. Depending on the reader’s interest, the techniques described should have wider application than just the subset of hydrologic processes that are addressed in the following chapters. The objective of this book is to present scientific principles of distributed hydrologic modeling. In an effort to make the book more general, techniques described may be applied using many different GIS packages. The material contained in the second edition has benefited from more experience with the application of distributed modeling in operational settings and from advances in software development. Advances in research have led to better understanding of calibration procedures and the sensitivity of PBD models to inputs and parameters. Discovery that an optimal parameter set exists is a major advance in hydrologic science. As with the first edition, the scientific principles contained herein relate to the spatial variability, temporal and spatial resolution, information content, calibration, and application of a fully

1. DISTRIBUTED HYDROLOGIC MODELING

37 19

distributed physics-based distributed model. The result of this approach is intended to guide hydrologists in the pursuit of more reliable hydrologic prediction. 1.17

References

ASCE, 1999, GIS Modules and Distributed Models of the Watershed, Report, ASCE Task Committee GIS Modules and Distributed Models of the Watershed, P.A. DeBarry, R.G. Quimpo, eds. American Society of Civil Engineers, Reston, VA., p. 120. Beven, K.J. and M.J. Kirkby, 1979, “A physically based variable contributing area model of basin hydrology.” Hydrologic Sciences Bulletin, 240(1):43-69. Duan, Q., Sorooshian, S.S., and Gupta, V.K, 1992, “Effective and efficient global optimization for conceptual rainfall runoff models.” Water Resour. Res., 28(4):1015-1031. Dunne, T., T.R. Moore, and C.H. Taylor, 1975, Recognition and prediction of runoffproducing zones in humid regions.” Hydrological Sciences Bulletin, 20(3): 305-327. Desconnets, J.-C., B.E. Vieux, and B. Cappelaere, F. Delclaux (1996), “A GIS for hydrologic modeling in the semi-arid, HAPEX-Sahel experiment area of Niger Africa.” Trans. in GIS, 1(2): 82-94. Farajalla, N.S. and B.E. Vieux, 1995, “Capturing the essential spatial variability in distributed hydrologic modeling: Infiltration parameters.” J. Hydrol. Process., 9(1):pp. 55-68. Green, W.H. and Ampt, G.A., 1911, “Studies in soil physics I: The flow of air and water through soils.” J.of Agricultural Science, 4:1-24. HEC, (2000), Hydrologic Modeling System: HEC-HMS, U.S. Army Corps of Engineers Hydrologic Engineering Center, Davis California. Julien, P.Y. and B. Saghafian, 1991, CASC2D User’s Manual. Civil Engineering Report, Dept. of Civil Engineering, Colorado State University, Fort Collins, Colorado. Julien, P.Y., B. Saghafian, and F.L. Ogden, 1995, “Raster-based hydrological modeling of spatially-varied surface runoff.” Water Resources Bulletin, AWRA, 31(3): 523-536. Loague, K.M., 1988, “Impact of rainfall and soil hydraulic property information on runoff predictions at the hillslope scale.” Water Resour. Res., 24(9):1501-1510. Vieux, B.E., 1988, Finite Element Analysis of Hydrologic Response Areas Using Geographic Information Systems. Department of Agricultural Engineering, Michigan State University. A dissertation submitted in partial fulfillment of the degree of Doctor of Philosophy. Vieux, B.E., V.F. Bralts, L.J. Segerlind and R.B. Wallace, 1990, “Finite element watershed modeling: One-dimensional elements.” J. of Water Resources Planning and Management, 116(6): 803-819. Vieux, B.E, 1991, “Geographic information systems and non-point source water quality modeling” J. Hydrol. Process, John Wiley & Sons, Ltd., Chichester, Sussex England, Jan., 5: 110-123. Invited paper for a special issue on digital terrain modeling and GIS. Vieux, B.E., 1993, “DEM aggregation and smoothing effects on surface runoff modeling.” ASCE J. of Computing in Civil Engineering, Special Issue on Geographic Information Analysis, 7(3): 310-338. Vieux, B.E., N.S. Farajalla and N.Gauer (1996), “Integrated GIS and Distributed storm Water Runoff Modeling”. In: GIS and Environmental Modeling: Progress and Research Issues. Edited by. Goodchild, M. F., Parks, B. O., and Steyaert, L. GIS World, Inc., Colorado, pp. 199-204. Vieux, B.E. and N.S. Farajalla, 1994, “Capturing the essential spatial variability in distributed hydrologic modeling: Hydraulic roughness.” J. Hydrol. Process, 8(3): 221-236.

38 20

Chapter 1

Vieux, B.E. and N.S. Farajalla, 1996, “Temporal and spatial aggregation of NEXRAD rainfall estimates on distributed hydrologic modeling.” Proceedings of Third International Conference on GIS and Environmental Modeling, NCGIA, Jan. 21-25, on CDROM and the Internet:(Last accessed, 30 January 2004): www.ncgia.ucsb.edu/conf/SANTA_FE_CD-ROM/sf_papers/vieux_baxter/ncgia96.html Vieux, B.E. and N. Gauer, 1994, “Finite element modeling of storm water runoff using GRASS GIS”, Microcomputers in Civil Engineering, 9(4):263-270. Vieux, B.E. and J.E. Vieux, 2002, Vflo™: A Real-time Distributed Hydrologic Model. Proceedings of the 2nd Federal Interagency Hydrologic Modeling Conference, July 28August 1, 2002, Las Vegas, Nevada. Abstract and paper on CD-ROM. Available on the Internet (Last accessed, 23 January 2004): http://www.vieuxinc.com/vflo.htm.

Chapter 2 DATA SOURCES AND STRUCTURE Geospatial Data for Hydrology

1.1

Introduction

Once we decide to use GIS to manage the spatial data necessary for hydrologic modeling, we must address data characteristics in the context of GIS. Digital representation of topography, soils, land use/cover, and precipitation may be accomplished using widely available or special purpose GIS datasets. Each GIS data source has a characteristic data structure, which has implications for the hydrologic model. Two major types of data structure exist within the GIS domain: raster and vector. Raster data structures are characteristic of remotely sensed data with a single value representing a grid cell. Points, polygons, and lines are more often represented with vector data. Multiple parameters may be associated with the vector data. Even after considerable processing, hydrologic parameters can continue to have some vestige of the original data structure, which is termed an artifact. Some data sources capture characteristics of the data in terms of measurement scale or sample volume. Rain gauges measure rainfall at a point, whereas radar, satellites, and other remote sensing techniques typically average a surrogate measure over a volume or area. Source data structures can have important consequences on the derived parameter and, therefore, model performance. This chapter addresses issues of data structure, projection, scale, dimensionality, and sources of data for hydrologic applications. The geospatial data used to derive model parameters can come in a variety of data structures. Topography, for example, may be represented by a series of point elevations, contour lines, triangular facets composing a triangular irregular network (TIN), or elevations in a gridded, rectangular coordinate system. Rainfall may be represented by a point, polar/gridded

40 22

Chapter 2

array, or isohyetal lines (contours). Infiltration rates derived from soil maps are generalized over the polygon describing the soil-mapping unit. Resulting infiltration modeled using this data source will likely show artifacts of the original soil map polygons. Land use/cover may be used to develop evapotranspiration rates or estimates of hydraulic roughness from polygonal areas or from a raster array of remotely sensed surrogate measures. In both cases, the spatial variability of the parameters will be controlled by the source and data structure. These examples illustrate two important points. First, in most cases a data source may be either a direct measure of the physical characteristic or an indirect (surrogate) measure requiring conversion or interpretation. Second, because of how the data is measured, each source has a characteristic structure including spatial and temporal dimensions as well as geometric character (points, lines, polygons, rasters, or polar arrays of radar data). Because the model may not be expecting data in one form or another, transformation is often necessary from one data structure to another. This necessity often arises because hydrologic processes/parameters are not directly observable at the scale expected by the model, or because the spatial or temporal scale of the measured parameter differs from that of the model. This issue can require transformations from one projection to another, from one structure to another (e.g., contours to TINs), interpolation of point values, or surface generation. Because the GIS data forms the basis for numerical algorithms, hydrologic modeling requires more complex GIS analysis than simple geographical modeling using maps. Typically, distributed hydrologic modeling divides a watershed or region into computational elements. Given that the numerical algorithms used to solve conservation of mass and momentum equations in hydrology may divide the domain into discrete elements, a parameter may be assumed constant within the computational element. At the sub-grid scale, parameter variation is present and may be represented in the model as statistical distribution known as sub-grid parameterization. Hydraulic roughness may be measured at a point by relating flow depth and velocity. In a distributed model, the roughness is assigned to a grid cell based on the dominant landuse/cover classification. In most applications, the model computational element will not conform exactly to the measured parameter. Conformance of the data structure in the GIS map to the model representation of the process is a basic issue in GIS analysis for hydrology. Transformation from one data structure to another must be dealt with for effective use of GIS data in hydrology. Components of data structure are treated below.

2. DATA SOURCES AND STRUCTURE 1.2

41 23

Dimensionality

This chapter discusses the major types of data encountered in hydrologic modeling. Their arrangement does not follow precise Euclidean notions of 1, 2, and 3-dimensional data, or, with the addition of time, 4-dimensional data. Depending on how the data is represented, there may be mixtures of several types of Euclidean data. A stream network may be composed of vectors in 2dimensional space, yet the nodes and various points along the stream may be represented by 1-dimensional point data. The complexity of various data representations offers many possibilities for portraying hydrologic data. For instance, we can define the order of a stream by putting together the stream network composed of a series of vectors and the topology of how the branches of the network connect with, say, zero-order streams at the headwaters of the basin. Each stream segment of a particular order begins and ends with nodes. Along the stream, distance markers identifying contaminant sampling, measurement, or some other value may be associated at some distance along the stream. Distance along the stream is different from simply specifying another xy-point. Distance, in this definition, only has significance along the set of vectors identifying the stream segment. The term point data, as used here, refers to a representation of a quantity at a location. Meteorological stations measure many variables at a particular location. Rain gauge accumulations, wind speed, temperature, and insolation are examples of data that may not be measured precisely at the same location since instrumentation is often mounted on a tower at various heights. Nonetheless, the measured quantities are often represented at a single point in two-dimensional space. Except for measurements taken over some representative volume or area, e.g., radar and satellite measurements, nearly all measurements in hydrology are considered point estimates. 1.3

Map Scale and Spatial Detail

A map of topography can be shown at any scale within a GIS. Once data is digitized and represented electronically in a GIS, resolution finer than that at which it was compiled is lost. Subsequent resampling to a coarser or finer resolution obscures the inherent information content captured in the original map. A small-scale map is one in which features appear small, have few details, and cover large areas. An example of a small-scale map is one with a scale of 1:1000000. Conversely, large-scale maps have features that appear large and cover small areas. An example of a large-scale map is one with a scale of 1:2000. A map compiled at 1:1000000 can easily be displayed in a GIS at a 1:2000 scale, giving the false impression that the map contains more information than it really does.

42 24

Chapter 2

Because GIS provides the ability to easily display data at any scale, we must distinguish between the compilation or native scale and the userselected scale. The scale and resolution at which the data is collected or measured is termed the native scale or resolution. If the spot or point elevations are surveyed in the field on a grid of 100 m, this is its native resolution. Once contours are interpolated between the points and plotted on a paper map at a scale of, for example, 1:25000, we have introduced a scale to the data. Once the paper map is digitized, there will be little more information contained at a scale larger than 1:25000; enlarging to 1:1000 would make little sense. The importance to hydrology is that variations in landform, slope or topography may not be adequately captured at resolutions that are too coarse. Further, to claim that we have a 1:1000 scale map simply because we can set the scale in the GIS is misleading, because the small variations that may have been present were lost when the contours were compiled at 1:25000. The hydrologist must decide what scale will best represent hydrologic processes controlled by the topography. If microtopography at the scale of rills or small rivulets controls the rate of erosion and sediment transport, then small-scale maps (e.g., 1:1000000) will contain little information relevant to the modeling of the process. However, such a map may contain ample detail for modeling river basin hydrologic response. 1.4

Datum and Scale

In geodesy, datum refers to the geodetic or horizontal datum. The classical datum is defined by five elements, which give the position of the origin (two elements), the orientation of the network (one element), and the parameters of a reference ellipsoid (two elements). The World Geodetic System (WGS) is a geocentric system that provides a basic reference frame and geometric figure for the Earth, models the Earth gravimetrically, and provides the means for relating positions on various data to an Earthcentered, Earth-fixed coordinate system. Even if two maps are in the same coordinate system, discrepancies may still be apparent due to different datum or scale used to compile each map. This problem is common when a map is compiled with an older datum and then used with data compiled with an updated datum. The usual remedy is to adjust the older datum to bring it into alignment with the revised datum. Conversion routines exist to transform spatial data from one datum to another. Correction from one datum to another will not remove differences caused by compilation scale. If the aerial photography is collected at 1:25000 but the hydrography was compiled at a smaller scale, then the streams will not line up with the photography. Mis-registration is a common problem when

2. DATA SOURCES AND STRUCTURE

43 25

using generally available geospatial datasets that have been compiled at disparate scales. A similar effect of mis-registration may be observed when combining vector hydrography and raster elevation data because of differences in data structure. Digital aerial photography is available for many parts of the US either through government-sponsored acquisition or on a project basis. Correcting aerial photography so that it overlays properly with other geospatial data in a georeferenced coordinate system can provide a useful data source for obtaining hydrologic information. Orthophotography means that the photography has been rectified to a georeferenced coordinate system and is referred to as an orthophoto. As an example, Figure 2-1 shows two streams overlaid on top of a digital orthophoto. This orthophoto was compiled in North American Datum 83 (NAD83). The stream shown in black is compiled in a consistent datum (NAD83). The second stream (shown in white), displaced to the south and east, was compiled in an older datum NAD27. The NAD83 stream (shown in black) matches well with stream channel features in the photograph, whereas the NAD27 stream (shown in white) is inconsistent with the photograph.

Figure 2-1. Effect of mismatched datums on hydrographic features. (Black stream, NAD83; white stream, NAD27).

44 26 1.5

Chapter 2 Georeferenced Coordinate Systems

Georeferenced coordinate systems are developed to consistently map features on the Earth’s surface. Each point on the Earth’s surface may be located by a pair of latitude and longitude. Because the Earth is an oblate spheroid, the distance between two points on that surface depends on the assumed radius of the earth at the particular location. Both spherical and ellipsoidal definitions of the Earth exist. The terms spheroid and ellipsoid refer to the definition of the dimensions of the Earth in terms of the radii along the equator and along a line joining the poles. As better definitions of the spheroid are obtained, the spheroid has been historically updated. Clark in 1866 developed the Clark 66 spheroid. Geocentric spheroids have come into use with the advent of satellites. The GRS1980 and WGS84 are spheroids in which the geoid has been surveyed and updated with the aid of satellite techniques and measurements. Traditionally, GIS overlay of data has required that both datasets share the same coordinate system. As GIS software becomes more sophisticated, much of this process is becoming automated, but understanding the underlying process is essential to understanding problems that may arise. Projection/datum problems painfully present themselves when overlaying two different data layers that don’t line up. While small disparities such as those shown in Figure 2-1 are often due to differences in the data, major differences are likely caused by projection system differences. 1.6

Map Projections

A map projection transforms coordinates expressed by latitude and longitude to planar coordinates. Together with a geoid and datum definition, the equations are called a projection. Depending on the source of the GIS data, one may encounter a variety of map projections. When distances or areas are needed from geospatial data, the data is almost always projected from latitude and longitude into a 2-D plane. All projections introduce distortion, because the projection transforms positions located on a three-dimensional surface, i.e., spheroid, to a position located on a two-dimensional surface, called the projected surface. There are three main types of projections. Conformal projections are those that maintain local angles. If two lines intersect each other at an angle of 30º degrees on the spheroid, then in a conformal projection, the angle is maintained on the projected surface only if the projection is equal distance. The stereographic projection is conformal but not equal area or equal distance. Because hydrology is often concerned with distances and areas, map projections that preserve these quantities find broadest usage.

2. DATA SOURCES AND STRUCTURE

45 27

The usefulness of maps for navigation has made geographic projections an important part of human history. Figure 2-2 shows the countries of the world projected onto a plane tangent to the North Pole using the stereographic projection. The polar form was probably known by the Egyptians. The first Greek to use it was Hipparcus in the 2nd century. Francois d’Aiguillon was the first to name it stereographic in 1613.

Figure 2-2. Stereographic projection of countries together with lines of longitude and latitude.

While this projection has long been used for navigational purposes, it has been used more recently for hydrologic purposes. The US National Weather Service (NWS) uses it to map radar estimates of rainfall on a national grid called HRAP (Hydrologic Rainfall Analysis Project). The standard longitude is at 105ºW with a grid positioned such that HRAP coordinates of the North Pole are at (401,1601). The grid resolution varies with latitude but is 4.7625

46 28

Chapter 2

km at 60ºN latitude. A feature of this type of projection is that circles on the Earth’s surface remain circles in the projection. Radar circles remain as circles on the projected map. Figure 2-3 shows the basic idea of the stereographic projection. The choice of map plane latitude, termed reference latitude, is a projection parameter that depends on location and extent of features to be mapped. The distance between A’ and B’ is less on the map plane at 60ºN latitude than the distance between A” and B” at 90ºN latitude. Changes in distance on a given plane constitute a distortion.

A'' Map Planes

A'

B'' B'

A B

R

Figure 2-3. Stereographic projection method for transforming geographic location of points A and B to a map plane.

Parameters of a projection and the type of projection are important choices since the accuracy of mapped features depends on the selection. The spheroid may be represented with a single radius, for example the radius of 6371.2 km used in the HRAP projection. In choosing a projection, we seek to minimize the distortion in angles, areas or distances, depending on which aspect is more important. There is no projection that maintains all three characteristics. While we cannot have a projection that is simultaneously conformal, equal area, and equal distance, we can preserve two of the three quantities. For example, the Universal Transverse Mercator (UTM) projection is designed to be conformal and equal area. Figure 2-4 shows a transverse developable surface tangent along a meridian of longitude. To minimize the distortion in distance, the UTM projection divides the Earth’s surface into 60 zones. In mid-latitudes around

2. DATA SOURCES AND STRUCTURE

47 29

the world, identical projections are made in each zone of 6º longitude (360º/60). This means that the coordinates in the projected surface uniquely describe a point only within the zone. The projected coordinates are in meters with the x-coordinate (east west) of 500 000 m being assigned to the central meridian of each zone. It is not enough to say that a particular point is located at x = 500 000 and y = 2 000 000 m, because in the UTM projection we must also specify the zone.

Figure 2-4. Transverse cylinder developable surface used in the UTM projection.

Depending on the projection, distortions in either the east-west direction or the north-south direction may be minimized, but we cannot have both. If the geographic feature spans more distance in the north-south direction, a projection that minimizes distortion in this direction is often used. The direction of least distortion is determined by the developable surface used to transform the coordinates on the spheroid to a two-dimensional surface. Typical developable surfaces are cylinders or cones. The orientation of axis of the developable surface with respect to the Earth’s axis determines whether the projection is termed transverse or oblique. A transverse projection would orient a cylinder whose axis is at right angles to the Earth’s axis and is tangent to the Earth’s surface along some meridian. The cylinder

48 30

Chapter 2

is then “unwrapped” to produce a two-dimensional surface with Cartesian coordinates. In an oblique projection, the axis of the developable surface and the Earth’s axis form an oblique angle. Distortion in distance is minimized in the UTM projection by making this cylinder tangent at the central meridian of each zone and then unwrapping it to produce the projected map. Choosing the appropriate projection for the spatial extent of the hydrologic feature is important. Figure 2-5 shows the Arkansas-Red-White River basins and subbasins along with state boundaries. The graticule at 6ºintervals in longitude corresponds to the UTM zones. Zones 13, 14, and 15 are indicated at the bottom of Figure 2-5.

Figure 2-5. Arkansas-Red-White River basin and UTM zones 13, 14, and 15.

The watershed boundaries shown are in decimal degrees of latitude and longitude and not projected. A watershed that crosses a UTM zone cannot be mapped because the x-coordinate is non-unique. The origin starts over at 500 000 m at the center of each zone. The UTM projection is an acceptable projection for hydrologic analysis of limited spatial extent provided the area

2. DATA SOURCES AND STRUCTURE

49 31

of interest does not intersect a UTM zone boundary. Many of the subbasins shown in Figure 2-5 could effectively be mapped and modeled using UTM coordinates, whereas the entire Arkansas-Red-White basin could not because it intersects two zonal boundaries. Selecting the appropriate projection and datum for hydrologic analysis depends on the extent of the watershed or region. If the distance spanned is large, then projection issues take on added importance. In small watersheds, the issue of projection is not as important because errors or distortions are on the order of parts per million, and do not generally have a significant impact on the hydrologic simulation. In fact, for small watersheds we can use an assumed coordinate system that relates features to each other without regard to position on the Earth’s surface. The following section deals with general concepts of geographic modeling along with the major types of data representation with specific examples related to hydrology. 1.7

Data Representation

1.7.1

Metadata

Knowing the origin, lineage, and other aspects of the data you are using is essential to understanding its limitations and getting the most usefulness from the data. This information is commonly referred to as data about the data, or metadata. For example, the fact that an elevation dataset is recorded to the nearest meter could be discovered from the metadata. If the topography controlling the hydrologic process is controlled by surface features, e.g., rills and small drainage channels which physically have scales less than one meter in elevation or spatial extent, then digital representation of the surface may not be useful in modeling the hydrologic process. An example of metadata documentation for GIS data is provided by the Federal Geographic Data Committee (FGDC) with content standards for digital geospatial metadata (FGDC, 1998). The FGDC-compliant metadata files contain detailed descriptions of the data sets and narrative sections describing the procedures used to produce the data sets in digital form. Metadata can be indispensable for resolving problems with data. Understanding the origin of the data, the datum and scale at which it was originally compiled, and the projection parameters are prerequisites for effective GIS analysis. The following section describes the major methods for representing topography and its use in automatic delineating of watershed areas. Types of surface representation can be extended to other attributes besides elevation such as rainfall.

50 32 1.7.2

Chapter 2 Digital Representation of Topography

A digital elevation model (DEM) consists of an ordered array of numbers representing the spatial distribution of elevations above some arbitrary datum in a landscape. It may consist of elevations sampled at discrete points or the average elevation over a specified segment of the landscape, although in most cases it is the former. DEMs are a subset of digital terrain models (DTMs), which can be defined as ordered arrays of numbers that represent the spatial distribution of terrain attributes, not just elevation. Over 20 topographic attributes can be used to describe landform. Slope and slope steepness have always been important and widely used topographic attributes. Many land capability classification systems utilize slope as the primary means of ascribing class, together with other factors such as soil depth, soil drainability, and soil fertility. Other primary topographic attributes include specific catchment area and altitude. Surface geometry analysis plays an important role in the study of various landscape processes (Mitasova and Hofierka, 1993). Analytically derived compound topographic indices may include soil water content, surface saturation zones, annual precipitation, erosion processes, soil properties, and processes dependent on solar radiation. Reliable estimation of topographic parameters reflecting terrain geometry is necessary for geomorphological, hydrologic and ecological studies, because terrain controls runoff, erosion, and sedimentation. When choosing the particular method of representing the surface, it is important to consider the end use. The ideal structure for a DEM may be different if it is used as a structure for a distributed hydrologic model than if it is used to determine the topographic attributes of the landscape. There are three principal ways of structuring a network of elevation data for its acquisition and analysis: 1. Contour 2. Raster 3. Triangular irregular network With this introduction to the three basic types of surface representation, we now turn in more detail to each of the three. 1.7.3

Contour

A contour is an imaginary line on a surface showing the location of an equal level or value. For example, an isohyet is a line of equal rainfall accumulation. Representation of a surface using contours shows gradients and relative minima and maxima. The interval of the contour is important, particularly when deriving parameters. The difference between one level

2. DATA SOURCES AND STRUCTURE

51 33

contour and another is referred to as quantization. The hydrologic parameter may be quantized at different intervals depending on the variability of the process and the scale at which the hydrologic process is controlled. Contour-based methods of representing the landsurface elevations or other attributes have important advantages for hydrologic modeling because the structure of their elemental areas is based on the way in which water flows over the landsurface (Moore et al., 1991). Lines orthogonal to the contours are streamlines, so the equations describing the flow of water can be reduced to a series of coupled one-dimensional equations. Many DEMs are derived from topographic maps, so their accuracy can never be greater than the original source of data. For example, the most accurate DEMs produced by the United States Geologic Survey (USGS) are generated by linear interpolation of digitized contour maps and have a maximum root mean square error (RMSE) of one-half contour interval and an absolute error of no greater than two contour intervals in magnitude. Developing contours in hydrologic applications carries more significance than merely representing the topography. Under certain assumptions, it is reasonable to assume that the contours of the landsurface control the direction of flow. Thus, surface generation schemes that can extract contour lines and orthogonal streamlines at the same time from the elevation data have advantages of efficiency and consistency. 1.7.4

Raster

The raster data structure is perhaps one of the more familiar data structures in hydrology. Many types of data, especially remotely sensed information, is often measured and stored in raster format. The term raster derives from the technology developed for television in which an image is composed of an array of picture elements called pixels. This array or raster of pixels is also a useful format for representing geographical data, particularly remotely sensed data, which in its native format is a raster of pixels. Raster data is also referred to as grids. Because of the vast quantities of elevation data that is in raster format, it is commonly used for watershed delineation, deriving slope, and extracting drainage networks. Other surface attributes such as gradient and aspect may be derived from the DEM and stored in a digital terrain model (DTM). The term DTED stands for digital terrain elevation data to distinguish elevation data from other types of DTM attributes. Raster DEMs are one of the most widely used data structures because of the ease with which computer algorithms are implemented. However, they do have several disadvantages:

52 34 • •

Chapter 2

They cannot easily handle abrupt changes in elevation. The size of grid mesh affects the results obtained and the computational efficiency. • The computed upslope flow paths used in hydrologic analyses tend to zig-zag and are therefore somewhat unrealistic. • The definition of specific catchment areas may be imprecise in flat areas. Capturing surface elevation information in a digital form suitable to input into a computer involves sampling x, y, z (i.e, easting, northing and elevation) points from a model representing the surface, such as a contour map, stereo-correlated serial photographs or other images. DEMs may be sampled using a variety of techniques. Manual sampling of DEMs involves overlaying a grid onto a topographic map and manually coding the elevation values directly into each cell. However, this is a very tedious and timeconsuming operation suitable only for small areas. Alternatively, elevation data may be sampled by direct quantitative photogrammetric measurement from aerial photographs on an analytical stereo-plotter. More commonly, digital elevation data are sampled from contour maps using a digitizing table that translates the x, y and z data values into digital files. Equipment for automatically scanning line maps has also been developed based on either laser-driven line-following devices or a raster scanning device, such as the drum scanner. However, automatic systems still require an operator to nominate the elevation values for contour data caused by poor line work, the intrusion of non-contour lines across the contour line being automatically scanned, or other inconsistencies. DEMs may be derived from overlapping remotely-sensed digital data using automatic stereo-correlation techniques, thereby permitting the fast and accurate derivation of DEMs. With the increasing spatial accuracy of remotely sensed data, future DEMs will have increasingly higher accuracy. A recent US space shuttle mission launched by NASA mapped 80% of the populated Earth surface to 30 meter resolution. The major disadvantage of grid DEMs are the large amount of data redundancy in areas of uniform terrain and the subsequent inability to change grid sizes to reflect areas of different complexity of relief (Burrough, 1986). However, various techniques of data compaction have been proposed to reduce the severity of this problem, including quadtrees, freeman chaincodes, run-length codes, block codes and the raster data structure. The advantages of a regularly-gridded DEM is its easy integration with raster databases and remotely-sensed digital data, the smoother, more natural appearance of contour maps and derived terrain features maps, and the ability to change the scale of the grid cells rapidly.

2. DATA SOURCES AND STRUCTURE 1.7.5

53 35

Triangular Irregular Network

A triangular irregular network (TIN) is an irregular network of triangles representing a surface as a set of non-overlapping contiguous triangular facets of irregular sizes and shapes. TINs are more efficient at representing the surface than is the uniformly dense raster representation. TINs have become increasingly popular because of their efficiency in storing data and their simple data structure for accommodating irregularly spaced elevation data. Advantages have also been found when TIN models are used in inter-visibility analysis on topographic surfaces, extraction of hydrologic terrain features, and other applications. A TIN has several distinct advantages over contour and raster representations of surfaces. The primary advantage is that the size of each triangle may be varied such that broad flat areas are covered with a few large triangles, while highly variable or steeply sloping areas are covered with many smaller triangles. This provides some efficiency over raster data structures since the element may vary in size according to the variability of the surface. Given the advantages of TINs in representing data requiring variable resolution, we will examine the features of and methods for generating the TINs. A TIN approximates a terrain surface by a set of triangular facets. Each triangle is defined by three edges and each edge is bound by two vertices. Most TIN models assume planar triangular facets for the purpose of simpler interpolation or contouring. Vertices in TINs describe nodal terrain features, e.g., peaks, pits or passes, while edges depict linear terrain features, e.g., break, ridge or channel lines. Building TINs from grid DEMs therefore involves some procedures for efficiently selecting the locations of TIN vertices for nodal terrain features or TIN edges for linear terrain features. Grid DEMs are widely available at a relatively low cost. Because of this increasing availability, the need for an efficient method to extract critical elevation points from grid DEMs to form TINs has increased. Some caution should be exercised to ensure that critical features are not lost in the conversion process. Figure 2-6 is a TIN created from a grid DEM. Notice the larger triangles representing flat areas and the smaller, more numerous, triangles in areas where there is more topographic relief. Using triangles of various sizes, where needed, reduces computer storage compared with raster formats. Methods for constructing a TIN from a grid DEM was investigated by Lee (1991). In general, methods for creating TINs consist of selecting critical points from grid DEMs. Nonessential grid points are discarded in favor of representing the surface with fewer points linked by the triangular facets. These conversion methods may be classified into four categories: 1)

54 36

Chapter 2

skeleton, 2) filter, 3) hierarchy, and 4) heuristic method. Each of these methods has its advantages and disadvantages, differing in how they assess point importance, and in their stopping rules, i.e., when to stop eliminating elevation points. The common property of all the methods is that the solution sets depend on some pre-determined parameters; these are several tolerances in the skeleton method and either a tolerance or a prescribed number of output points in the other three methods.

Figure 2-6 TIN elevation model derived from a grid DEM.

The objective of any method of converting DEMs to TINs is to extract a set of irregularly spaced elevation points which are few as possible while at the same time providing as much information as possible about topographic structures. Unfortunately, it is impossible to achieve both goals concurrently and difficult to balance the two effects. Surface geometry (especially breaklines) best represented by smaller TINs may be lost with larger TINs. When extracting TINs from grid DEMS, this trade-off would have to be dealt with first by finding the best middle ground. Since it would differ from one application to another and among different types of terrain surfaces, either trial and error or cross-validation experiments are usually needed to assess the fidelity of the topographic surface.

2. DATA SOURCES AND STRUCTURE

55 37

The importance of digital elevation data and derivative products for distributed hydrologic modeling cannot be overstated. Cell resolution effects are discussed in Chapter 4, which deals with information content and spatial variability. Chapter 6 addresses aspects of drainage networks derived from DEMs in detail. Methods for automatic delineation of watershed boundaries from DEMs are discussed in the following section. 1.8

Watershed Delineation

Raster or TIN DEMs are the primary data structures used in the delineation of watershed boundaries. Watersheds/basins and subwatersheds/subbasins are terms used interchangeably with catchment. Subwatersheds and the corresponding stream or drainage network is a type of data structure for organizing hydrologic computations. Lumped models must be applied based on delineated watersheds. Distributed models are usually organized by subbasins. Parallelization of distributed modeling codes is easiest if the computational work is organized according to watersheds. Defining the watershed and the drainage network forms the basic framework for applying both lumped and distributed hydrologic models. Moore et al. (1991) discussed the major data structures for watershed delineation ranging from grid, TIN, and contour methods. Figures 2-7 and 2-8 show the Illinois River Basin delineated to just below Lake Tenkiller in Eastern Oklahoma. The area is 4211 km2 delineated at a DEM resolution of 60 and 1080 m, respectively. At 60 m the number of cells is 1169811, whereas at 1080 m, the number of cells is only 3610. With larger resolution, computer storage is reduced; however, sampling errors increase. That is to say, the variability of the surface may not be adequately captured at coarse resolution, resulting in difficulties in automatic watershed delineation. GIS algorithms exist for forcing delineation when pits or depressions occur due to coarse resolution. For the watersheds delineated from the DEMs shown in Figure 2-7 and 2-8, the pits were filled and then the watershed draining to the outlet was delineated using the ArcView Hydro extension. It is common to find varying watershed areas and shapes depending on the DEM resolution. The two watersheds delineated in Figure 2-7 and 2-8 look similar in shape. Some differences in shape along the edges are introduced because coarser resolution may not accurately resolve flow directions in flat areas along the watershed divide. Depressions may be natural features or simply a result of sampling an irregular surface with a regular sampling interval, i.e., grid resolution. The hydrologic significance of depressions depends on the type of landscape represented by the DEM. In some areas, such as the prairie pothole region in the Upper Midwest (North Dakota) of the United States or in parts of the

56 38

Chapter 2

Sahel in Africa, surface depressions dominate and control the hydrologic processes. In areas with coordinated drainage, depressions are an artifact of the sampling and generation schemes used to produce the DEM. We should distinguish between real depressions and those that are artifacts introduced by the sampling scheme and data structure.

Figure 2-7. Digital elevation map of the Illinois River Basin with 60-m resolution.

1.8.1

Algorithms for Delineating Watersheds

O’Callaghan and Mark (1984) and Jenson (1987) proposed algorithms to produce depressionless DEMs from regularly spaced grid elevation data. Numerical filling of depressions, whether from artifacts or natural depressions, facilitates the automatic delineation of watersheds. Smoothing of the DEM can improve the success of delineation but can have deleterious effects on derived slopes (Vieux, 1993). If the depressions are hydrologically

2. DATA SOURCES AND STRUCTURE

57 39

significant their volume can be calculated. Jenson and Dominique (1988) used the depressionless DEM as a first step in assigning flow directions. Their procedure is based on the hydrologically realistic algorithm discussed by O’Callaghan and Mark (1984), but it is capable of determining flow paths iteratively where there is more than one possible receiving cell and where flow must be routed across flat areas.

Figure 2-8. Digital elevation map of the Illinois River Basin at 1080-m resolution.

Three main methods were examined by Skidmore (1990) for calculating ridge and gully position in the terrain. The main algorithms are: Peucker and Douglas algorithm

58 40

Chapter 2 Peucker and Douglas (1975) mapped ridges and valleys using a simple moving-window algorithm. The cell with the lowest elevation in a two-by-two moving window is flagged. Any unflagged cells remaining after the algorithm has passed over the DEM represent ridges. Similarly the highest cell in the window is flagged, with any unflagged cells in the DEM corresponding to valley lines. O’Callaghan and Mark algorithm O’Callaghan and Mark (1984) described an algorithm for extracting stream and ridge networks from a DEM. This algorithm forms a DEM based on quantifying the drainage accumulation at each cell in the DEM. Cells which had a drainage accumulation above a userspecific threshold were considered to be on a drainage channel. Ridges were defined as cells with no drainage accumulation. Band algorithm Band (1986) proposed a method for identifying streamlines from a DEM, enhancing the Peucker and Douglas algorithm by joining ‘broken’ stream lines. Ridges and streamlines, found in the same way as the algorithm, were thinned to one cell wide using the Rosenfeld and Kak (1982) thinning algorithm. The upstream and downstream nodes on each stream fragment were then flagged. Each downstream node was “drained” along the line of maximum descent until it connected with another streamline. The streams were again thinned to the final, one-cell wide, line representation of the stream network.

More recent research has improved upon these models, and many GIS modules exist for processing DEMs and delineating watersheds. Skidmore (1990) compared these three methods for mapping streams and ridges from the digital elevation model with a new algorithm that utilizes basic map delineation. Ridges and streamlines closely followed the original contour map and improved upon the results from the three alternative algorithms. Mid-slope positions are located from the stream and ridgelines by a measure of Euclidean distance. The algorithm proposed by Skidmore (1990) generated a satisfactory image of streams and ridges from a regularly gridded DEM. The Peucker and Douglas, and O’Callaghan and Mark algorithms produced images with broken streams in areas of flat gradient. The delineation produced by the Band algorithm was similar to that of the proposed algorithm, although it caused a larger number of streams to appear in the flatter parts of the study area. The Band algorithm was found to be more computationally expensive than the Skidmore algorithm. The decision

2. DATA SOURCES AND STRUCTURE

59 41

on which algorithm to use depends largely on the resolution of indeterminate flow directions caused by flat slopes. 1.8.2

Problems with Flat Areas

Truly flat landscapes, or zero slope, seldom occur in nature. Yet, when a landscape is represented by a DEM, areas of low relief can translate into flat surfaces. This flatness may also be a result of quantization (precision) of the elevation data. This occurs when the topographic variation is less than one meter yet the elevation data is reported with a precision to the nearest meter. Flat surfaces typically are the result of inadequate vertical DEM resolution, which can be further worsened by a lack of horizontal resolution. Such flat surfaces are also generated when depressions in the digital landscape are removed by raising the elevations within the depressions to the level of their lowest flow. A variety of methods have been proposed to address the problem of drainage analysis over flat surfaces. Methods range from simple DEM smoothing to arbitrary flow direction assignment. However, these methods have limitations. DEM smoothing introduces loss of information to the already approximate digital elevations, while arbitrary flow direction assignment can produce patterns that reflect the underlying assignment scheme, which are not necessarily realistic or topographically consistent. Given these limitations, the application of automated DEM processing is often restricted to landscapes with well-defined topographic features that can be resolved and represented by the DEM. Improved drainage identification is needed over flat surfaces to extend the capabilities and usefulness of automated DEM processing for drainage analysis. Garbrecht and Martz (1997) presented an approach to addressing this problem that modifies flat surfaces to produce more realistic and topographically consistent drainage patterns than those provided by earlier methods. The algorithm increments cell elevations of the flat surface to include information on the terrain configuration surrounding the flat surface. As a result, two independent gradients are imposed on the flat surface: one away from the higher terrain into the flat surface, and the other out of the flat surface towards lower terrain. The linear combination of both gradients, with localized corrections, is sufficient to identify the drainage pattern while at the same time satisfying all boundary conditions of the flat surface. Imposed gradients lead to more realistic and topographically consistent drainage over flat surfaces. The shape of the flat surface, the number of outlets on its edge, and the complexity of the surrounding topography apparently do not restrict the proposed approach. A comparison with the drainage pattern of an

60 42

Chapter 2

established method that displays the “parallel flow” problem shows significant improvements in producing realistic drainage patterns. One of the most satisfactory methods for assigning drainage directions on flat areas is that of Jenson and Dominique (1988). The Jenson and Domingue (JD) algorithm is useful over most of the DEM but does not produce satisfactory results in areas of drainage lines because it causes these lines to be parallel. The JD algorithm assigns drainage directions to flat areas in valleys and drainage lines such that flow is concentrated into single lines, and it uses the JD method over the rest of the DEM where less convergent flow is more realistic. Automated valley and drainage network delineation seeks to produce a fully connected drainage network of single cell width because this is what is required for applications such as hydrologic modeling. No automatically produced drainage network is likely to be very accurate in flat areas, because drainage directions across these areas are not assigned using information directly held in the DEM. One method for enforcing flow direction is to use an auxiliary map to restrict drainage direction where a mapped stream channel exists. In this case, a river or stream vector map may be used to burn in the elevations, forcing the drainage network to coincide with the vector map depicting the desired drainage network. By burning in, i.e., artificially lowering the DEM at the location of mapped streams, the correct location of the automatically delineated watershed and corresponding stream network is preserved. Watersheds and stream networks delineated from a DEM become a type of data structure for organizing lumped and distributed hydrologic model computations. The following section deals with another type of map representation useful in soil mapping or other thematic representations. 1.9

Soil Classification

Another important source of hydrologic modeling parameters is existing maps, such as soils for simulating infiltration. Mapping soils usually involves delineating soil types that have identifiable characteristics. The delineation is based on many factors germane to soil science, such as geomorphologic origin and conditions under which the soil formed, e.g., grassland or forest. Regardless of the purpose or method of delineation, there will be a range of soil properties within each mapping unit. This variation may stem from inclusions of other soil types too small to map or from natural variability. The primary hydrologic interest in soil maps is the modeling of infiltration as a function of soil properties. Adequate measurement of infiltration directly over an entire watershed is impractical. A soil-mapping unit is the smallest unit on a soil map that can be assigned a set of

2. DATA SOURCES AND STRUCTURE

61 43

representative properties. The soil properties are stated in terms of layers. At a particular location on the map, because the properties of the soil vary with depth, some infiltration scheme is adopted for representing an essentially one-dimensional (vertical) process. The infiltration model representation may not include all layers used in the soil classification. Soil maps and the associated soil properties form a major source of data for estimating infiltration. A map originally compiled for agricultural purposes may be reclassified into infiltration parameter maps for hydrologic modeling. As such, the parameter map takes on a data structure characteristic of the original soil map. The characteristic data structure associated with the original soil map carries forward into the hydrologic parameters used to simulate infiltration. Obtaining infiltration parameters from soil properties requires some type of reclassifying of the soil-mapping unit into a parameter meaningful to the hydrologic model and doing this for each model layer. Some adjustment of the number of soil layers is inevitable when fitting soil layers into a model scheme. Once the soil mapping unit delineated by a polygon is transformed to an infiltration parameter, the artifact of the original mapping unit will generally be present. A large area of uniform values from a single mapping unit may obscure the natural soil variability that is not mapped. The vector data structure of the soil map, whether converted to raster or not, will carry over into the hydrologic modeling. Efforts to estimate the hydraulic conductivity and other parameters from soil properties will be covered in more detail in Chapter 5, which deals with infiltration. Other readily available digital maps may be used to derive parameters as seen in the following section. 1.10

Land use/Cover Classification

It is well known that land use, vegetative cover, and urbanization affect the runoff characteristics of the land surface. The combination of the land use and the cover, termed land use/cover, is sometimes available as geospatial data derived from aerial photography or satellite imagery. To be useful, this land use/cover must be reclassified into parameters that are representative of the hydrologic processes. Examples of reclassification from a land use/cover map into hydrologic parameters include hydraulic roughness, surface roughness heights affecting evapotranspiration, and impervious areas that limit soil infiltration capacity. The data structure of the parent land use/cover map will carry into the model parameter map similar to how soil mapping units define the spatial variation of infiltration parameters.

62 44

Chapter 2

Maps derived from remote sensing of vegetative cover can be used to develop hydrologic parameters. Land use/cover classification schemes are devised for general purposes and not specifically for direct application to hydrology. An important concern is whether the land use/cover map classification and scale contains sufficient detail to be useful for hydrologic simulation. The hydraulic roughness is a parameter that controls the rate of runoff over the landsurface, and therefore affects the peak discharge and timing of the hydrograph in response to rainfall input. The hydraulic roughness parameter map takes on a data structure characteristic of the original land use/cover map. If derived from remotely sensed data, a raster data structure results rather than a polygonal structure. Figure 2-9 shows the polygonal data structure of a land use/cover map.

Figure 2-9. Land use/cover polygonal outlines for the Illinois River Basin near Lake Tenkiller, Oklahoma, USA.

This map is a special purpose land use/cover map derived from aerial photography at the same scale but more detailed than the Land Use and Land Cover (LULC) data files generally available in the US. The USGS provides these data sets and associated maps as a part of its National Mapping Program. The LULC mapping program is designed so that standard

2. DATA SOURCES AND STRUCTURE

63 45

topographic maps of a scale of 1:250000 can be used for compilation and organization of the LULC data. In the classification scheme shown in Figure 2-9, a distinction is made between pasture and cropland. The classification scheme used in the LULC maps makes no distinction between these two categories. In terms of runoff rate or evapotranspiration, these two categories may behave very differently. The primary (Level I) Anderson Classification codes used in LULC maps are listed in Table 2-1. Second-level categories exist within the major categories, but are not shown here. Table 2-1. Anderson land use/cover primary classification codes Anderson Classification Codes

1 Urban or Built-Up Land 2 Agricultural Land 3 Rangeland 4 Forest Land 5 Water

6 Wetland 7 Barren Land 8 Tundra 9 Perennial Snow and Ice

The level of detail in the LULC map determines the spatial variability of the derived hydrologic model parameter. The hydrologist must assign representative values of this parameter to each classification. Simple reclassification of a land use map into hydraulic roughness represents the deterministic variation described by the LULC map, but may ignore variability that is not mapped. On-site experience and published values are helpful in establishing the initial map for calibration of a distributed parameter model. Some judgment is required to assign hydrologically meaningful parameters to generalized land use/cover classification schemes. In Chapter 6 we will examine how to derive hydraulic roughness maps from such a classification scheme. 1.11

Summary

Effective use of GIS data in distributed hydrologic modeling requires understanding of type, structure, and scale of geospatial data used to represent the watershed and runoff processes. GIS data often lacks sufficient detail, space-time resolution, attributes, or differs in some fundamental way from how the model expects the character of the parameter and how the parameter is measured. Existing data sources are often surrogate measures that attempt to represent a particular category with direct or indirect relation to the parameter or physical characteristic. Generation of a surface from measurements taken at points, reclassification of generalized map categories into parameters, and extraction of terrain attributes from digital elevation data are important operations in the preparation of a distributed hydrologic

64 46

Chapter 2

model using GIS. Having transformed the original map into useable parameter maps, the parameter takes on the characteristic data structure of the original map. Thus, the data structure inherent in the original GIS map has lasting influence on the hydrologic process simulated using the derived parameters. In the following chapter we turn to the generation of surfaces from point data. 1.12

References

Band, L.E., 1986, “Topographic partition of watershed with Digital Elevation Models.” Water Resour. Res., 22 (1): 15-24. Burrough, P. A., 1986, Principles of Geographic Information Systems for Land Resources Assessment. Monographs on Soil and Resources Survey, No.12, Oxford Science Publications pp. 103-135. FGDC, 1998, Content Standard for Digital Geospatial Metadata. rev. June 1998 Metadata Ad Hoc Working Group, Federal Geographic Data Committee, Reston, Virginia, pp.1-78. Garbrecht, J. and Martz, L. W., 1997, “The assignment of drainage direction over flat surfaces in raster digital elevation models.” J. of Hydrol., 193 pp.204-213. Jenson, S. K., 1987, “Methods and applications in Surface depression analysis.” Proc. AutoCarto: 8, Baltimore, Maryland pp.137-144. Jenson, S. K. and Dominique, J. O, 1988, “Extracting Topographic Structure from Digital Elevation Data for Geographic Information System Analysis.” Photogramm. Eng. and Rem. S., 54 (11) pp.1593-1600. Lee, J., 1991, “Comparison of existing methods for building triangular irregular network models from grid digital elevation models.” Int. J. Geographical Information Systems, 5(3):267-285. Mitasova, H. and Hofierka, J., 1993, “Interpolation by Regularised Spline with Tension: II. application to Terrain and Surface Geometry Analysis.” Mathematical Geology, 25(6):657-669. Moore, I. D., Grayson, R. B., and Ladson, A. R., 1991, “Digital Terrian Modeling: A Review of Hydrological, Geomorphological and Biological Applications.” J. Hydrol. Process, 5:3-30. O’Callaghan, J. F. and Mark, D. M., 1984, “The extraction of drainage networks from digital elevation data.” Computer vision, graphics and image processing, 28:323-344. Peucker, T. K. and Douglas, D. H., 1975, “Detection of surface specific points by local parallel processing of discrete terrain elevation data.” Computer vision, graphics and image processing, 4:p.375. Rosenfeld, A. and Kak, A. C., 1982, Digital picture Processing. No. 2. Academic Press, New York. Skidmore, A.K., 1990, “Terrain position as mapped from a gridded digital elevation model.” Int. J. Geographical Information Systems, 4(1):33-49. Snyder, J. P., 1987, Map projections—A working manual: U.S. Geological Survey Professional Paper 1395, p.383. Vieux, B. E., 1993, “DEM Aggregation and Smoothing Effects on Surface Runoff Modeling.” ASCE J. of Computing in Civil Engineering, Special Issue on Geographic Information Analysis, 7(3): 310-338.

Chapter 3 SURFACE GENERATION Producing Spatial Data from Point Measurements

Figure 3-1. Hillslope shading of topography near the confluence of the Illinois River and Baron Fork Creek, Oklahoma.

66 48

Chapter 3

The shaded relief image shown above in Figure 3-1 is a portion of the Illinois River basin. This image reveals a myriad of ravines forming a drainage network. This DEM was derived from a 60-meter grid DEM produced by the USGS from contour lines drawn photogrammetrically using stereo-aerial photographs. Generating surfaces such as the DEM shown is the subject of the following sections. 1.1

Introduction

The importance of the raster data structure for hydrology cannot be overstated. Transformation from point to raster data structures is often accomplished through interpolation of a surface. This surface is generally developed by interpolating a grid-cell value from surrounding point values. Even if a surface is not required, point values may be required at locations other than at the measured point. This requires interpolation whether or not an entire grid is needed. Several surface generation utilities exist within general-purpose GIS packages such as ESRI ArcView, ArcGIS, or the public domain software called GRASS. It is important to understand how these surfaces are generated, the influence of the resulting surface on hydrologic modeling, and some of the pitfalls. Much work has been done in the area of spatial statistics, commonly referred to as geostatistics. This type of statistics is distinguished by the fact that the data samples invariably are autocorrelated and are therefore statistically dependent at some range or length scale. Temperature, rainfall, and topography, together with derived slope and aspect, and infiltration rates or soil properties are measured at a point but are required in the raster data structure for modeling purposes. When generating a surface from irregularly spaced data points, we must choose whether it will be done by 1) interpolating directly on the irregularly spaced data, or by 2) regularizing the data into a grid and then applying some interpolation or surface generation algorithm. The first method is accomplished using the TIN data structure, which has the advantage that real surface facets or breaklines can be preserved. The definition and characteristic data structure for TINs discussed in Chapter 2 have the advantage of minimizing computer storage while maintaining an accurate representation of the surface. Triangular irregular networks are surface representations and could logically be included here as a type of surface generation. TINs are useful because contours are easily interpolated along each facet. The second method has the advantage of producing a raster data structure directly with a smooth appearance. Smoothness, in the sense of differentiability, while desirable for many applications, may violate physical

3. SURFACE GENERATION

67 49

reality. Local gradients may not be particularly important at the river basin scale as long as the global mean of the slopes is not suppressed due to local smoothing. However, local gradients and associated slope may be particularly important for erosion modeling and sediment transport. Although the second method has the advantage that a smoother surface will result, this smoothness, though pleasing to the eye, may possibly obliterate real surface facets or breaklines. In the following sections, we examine some of the more popular surface generation methods and application to hydrology. 1.2

Surface Generators

Interpolation from scattered data is a well-developed area of scientific literature with obvious applications in hydrology. A number of new mathematical techniques for generating surfaces have been developed. Although the surfaces generated by sophisticated mathematical approaches do solve the interpolation problem, questions as to usefulness or practicality for various applications, such as hydrology, remain (Chaturvedi and Piegl, 1996). Surface generation is a more general case of interpolation in the sense that a value is needed in every grid cell of a raster array derived from point data. Among the many methods for generating a two-dimensional surface from point data, three will be addressed: 1. Inverse distance weighting (IDW) 2. Thin splines 3. Kriging. The problem with all of these methods when applied to digital elevation models or geophysical fields such as rainfall, ground water flow, wind, or soil properties is that some vital aspect of the physical characteristics becomes distorted or is simply wrong. The problems that arise when creating a surface from point data are: • Gradients are introduced that are artifacts, e.g., the tent pole effect. • Discontinuities and break lines, such as dikes or stream banks, may or may not be preserved. • Smoothing induced through regularized grids may flatten slopes locally and globally. • Interpolation across thematic zones such as soil type may violate physical reality. Further difficulties arise when applying surfacing algorithms to point measurements of geophysical measurements such as soil properties. Interpolation or surface generation may extend across physical boundaries, resulting in physically unrealistic results. Because of this, it may be

68 50

Chapter 3

necessary to interrupt the smooth interpolated surface by breaklines so that true linear features such as streams or edges of an embankment can be preserved. The quantitative evaluation of the amount of spatial distribution of precipitation is required for a number of applications in hydrology and water resources management. Many techniques have been proposed for mapping rainfall patterns and for evaluating the mean areal rainfall over a watershed by making use of data measured at points, e.g., rain gauges. Methods for precipitation interpolation from ground-based point data range from techniques based on Thiessen polygons and simple trend surface analysis, inverse distance weighting, multi-quadratic surface fitting, and Daulaney triangulations to more sophisticated statistical methods. Among statistical methods, the geostatistical interpolation techniques such as Kriging have been often applied to spatial analysis of precipitation. Kriging requires the development of a statistical model describing the variance as a function of separation distance. Once the underlying statistical model is developed, the surface is generated by weighting neighboring elevations according to the separation distance. This is perhaps the most critical and difficult task since the choice is somewhat arbitrary. The applicability or physical reality of the resulting model can hardly be assessed a priori. In following sections we will examine in more detail the application to hydrology of IDW, Kriging, and splines for generating surfaces composed of elevations, precipitation, or hydrologic parameter. 1.2.1

Inverse Distance Weighted Interpolation

Inverse Distance Weighted (IDW) interpolation is widely used with geospatial data. The generic equation for inverse distance weighted interpolation is, n

¦Z W i

Z x, y =

i

i =1 n

¦W

3.1

i

i =1

where, Zx,y is the point value located at coordinates x,y to be estimated; Zi represents the measured value (control point) at the ith sample point; and Wi is a weight that determines the relative importance of the individual control points in the interpolation procedure. Both Kriging and IDW depend on weighting neighboring data values in the estimation of Zx,y.

3. SURFACE GENERATION

69 51

Interpolation across boundaries may not produce a surface or interpolation consistent with physical reality; the surface may only have validity locally within a thematic type, zone or classification. This is particularly true where soil properties have zonal variation within each mapped soil type. Confining the interpolation to use measured values from within a zone will ensure that properties from two different soil types are not averaged to produce some intermediate value that is inconsistent with both soil types. If such a constraint is imposed, discontinuities in the surface should be expected and will be consistent with the boundaries delineating thematic types or zones. The univariate IDW technique assumes that the surface between any two points is smooth, but not differentiable; that is, this interpolation technique does not allow for abrupt change in height. Bartier and Keller (1996) developed a multivariate IDW interpolation routine to support inclusion of additional independent variables. Control was also imposed on the abruptness of surface change across thematic boundaries. Conventional univariate interpolation methods are unable to incorporate the addition of secondary thematic variables. Bartier and Keller (1996) found that the influence on surface shape did not allow for inclusion of interpretive knowledge. The multivariate IDW method of interpolation, on the other hand, supports the inclusion of additional variables and allows for control over their relative importance. The interpolated elevation, Zxy is calculated by assigning weights to elevation values, Zp found within a given neighborhood of the kernel: r

¦Z Z x, y =

p

dp

p =1 r

¦d

−n

3.2 −n p

p =1

where Zp is the elevation at point p in the point neighborhood r; d is the distance from the kernel to point p; and exponent n is the friction distance ranging from 1.0 to 6.0 with the most commonly used value of 2.0 (Clarke, 1990). The negative sign of n implies that elevations closer to the interpolant are more important than those farther away. Size of r is an important consideration in preserving local features such as ridges, breaklines or stream banks. Figure 3-2 shows the IDW interpolation scheme where the grid-cell elevation, Zx,y is interpolated from its nearest neighbors. The closer the neighboring value Zp, the more weight it has in the interpolated elevation. Best results are from sufficiently dense samples. In the IDW method, friction distance and the number of points or radius are the only two parameters in

70 52

Chapter 3

Eq. (3.2) that the user can adjust to fit the surface to the data. Visual inspection and validation are the only means for judging the value of the parameters that produce an optimal surface.

Z2

d2

Z1

(1)

(2)

(3)

d1

(7)

Zxy

(8)

d3

Z3

(4)

d4

(5)

(6)

Z4

Figure 3-2. IDW weighting scheme for interpolating grid-cell values from nearest neighbors.

One artifact of the IDW method is the tent pole effect. That is, a local minimum or maximum results at the measured point location. When applied to rain gauge accumulations, this gives the impression that it rained most intensely where there were measurements, which is clearly nonsensical. Figure 3-3 shows a surface interpolated from point rain gauge accumulations using the IDW method. The surface was interpolated using the inverse distance squared method using all data points. The contours indicate that the rainfall occurred mainly around the gauges, which is physically unrealistic.

3. SURFACE GENERATION

71 53

The tent pole effect is problematic for most surfaces of geophysical quantities unless sufficient data points are obtained. Neither the IDW nor the other methods described below can insure physical correctness or fidelity in the generated surface. Whether the surface is suitable for hydrologic application may not become apparent until the surface is generated and intrinsic errors are revealed or cross-validation proves that artifacts exist. The choice of interpolation method depends on the data characteristics and degree of smoothness desired in the interpolated result.

Figure 3-3. Tent pole effect when generating a surface with too few data values using the IDW method.

1.2.2

Kriging

One of the most important advantages of geostatistical methods such as Kriging and objective analysis is their ability to quantify the uncertainty in the derived estimates. In these methods, statistical hypotheses are made to identify and evaluate the multidimensional spatial structure of hydrological processes. Kriging is an optimal spatial estimation method based on a model of spatially dependent variance. The model of spatial dependence is termed a variogram or semivariogram (may be used interchangeably). Surface interpolation by Kriging relies on the choice of a variogram model. Inaccuracies due to noisy data or anisotropy cause errors in the interpolated

72 54

Chapter 3

surface that may not be obvious or be readily identified except through validation with data withheld from the surfacing algorithm. Todini and Ferraresi (1996) found that uncertainty in the parameters of the variogram could lead to three major inconsistencies: 1. Bias in the point estimation of the multi-dimensional variable 2. Different spatial distribution of the measure of uncertainty 3. Wrong choice in selecting the correct variogram model. For the Kriging analysis, the spatial autocorrelation of data can be observed from a variogram structure. When the spatial variability is somewhat different in every direction, an anisotropy hypothesis is adopted. If no spatial dependence exists, Kriging uses a simple average of surrounding neighbors for the interpolant. This implies that there is no dependence of the variance on separation distance. The theory underlying Kriging is well-covered elsewhere, especially in Journel and Huijbregts (1978) and subsequent literature. For those less familiar with the method, we examine the basic principles of Kriging. In the theory of regionalized variables developed by Matheron (1965), there are two intrinsic hypotheses: 1. The mean is constant throughout the region. 2. Variance of differences is independent of position, but depends on separation. If the first hypothesis is not met then a trend exists and must be removed. The second hypothesis means that once a variogram has been fitted to the data, it may be used to estimate the variance throughout the region. Semivariance, Ȗ(h), is defined over the couplets of Z(xi+h) and Z(xi) lagged successively by distance h:

γ (h) = 2N1(h) ¦[Z(x ) − Z (x + h)]

2

i

i

3.3

where N(h) is the number of couplets at each successive lag. Figure 3-4 shows how grid data is offset by multiples of h, in this case, 2h.

3. SURFACE GENERATION

73 55

Figure 3-4. Lagging grid data by multiples of h producing semivariogram.

The variogram is calculated by sequentially lagging the data with itself by distance, h. The expected valued of the difference squared is plotted with each lagged distance. The term semi- in semivariogram refers to the ½ in Eq. (3.3). Fitting a semivariogram model to the data is analogous to fitting a probability density function to sample data for estimating the frequency of a particular event. Once we have such a model, we can make predictions of spatial dependence based on separation distance. Only certain functions may be used to form a model of variance. These allowable models, (linear or exponential among others) ensure that variance is positive. The linear and exponential semivariogram models in Figure 3-5 establish the relationship between variance and spatial separation.

56 74

Chapter 3

10

Ȗ(h)

8

C

6

4

Linear Exponential

2

C0

0 0

1

2

3

4

5

6

7

8

9

10

Lag (h)

Figure 3-5. Exponential and linear variogram models.

Beyond a distance of a-units there is no dependence, resulting in the value of Ȗ(’) = s2, which is the variance in the classical statistical sense. At h=0, i.e., no lag, the variation is termed the nugget variance. Because geostatistics had its start in mining and the estimation of gold ore, the term ‘nugget’ refers to the random occurrence of finding a gold nugget. In the more general geostatistical problem, this variance corresponds to a Gaussian process with no spatial dependence. Linear model—

­ °C + mh Ȗ (h ) = ® 0 C +C ° ¯ 0

0≤h≤a h≥a

3.4

Exponential model— ­ °C + C °° 0 Ȗ (h ) = ® ° °C0 + C °¯

ª3 h 1 h 3º « §¨ ·¸ − §¨ ·¸ » 0 ≤ h ≤ a «2 © a ¹ 2 © a ¹ » ¼ ¬ h≥a

3.5

3. SURFACE GENERATION

75 57

where C0 is the nugget variance; C is the variance contained within the range, a. The variance C defines the variance as a function of the separation distance, h. It is the dependence of the variance on separation distance used to weight the neighboring data values when Kriging. In Kriging, we use the semivariogram to assign weights to data values neighboring the interpolant. Thus, Kriging is a surface interpolation with known variance. The semivariogram provides the basis for weighting the neighboring data values. Unlike IDW, the semivariogram provides the Kriging estimator with some basis for assigning weights to neighboring values. Weights for generating a Kriged surface Zˆ (B ) at location B is determined using weights taken from the semivariogram. The simple Kriging estimator within a restricted neighborhood of the interpolant, Zˆ (B ) at location B from n neighboring data points: n

Zˆ (B ) = ¦ Ȝi Z i

3.6

i =1

where Ȝi is the weights assigned to each of the elevations, Zi. In Kriging, the weights are not simply a function of distance, as in the IDW method, but depend on the variogram model. Figure 3-6 shows how the interpolant is calculated given the distances and weights from the observed or measured values. In practice, interpolated surfaces are more or less in error because either the underlying assumptions do not hold, or the variances are not known accurately, or both. An interpolated surface should be validated in any rigorous application. In geostatistical practice, the usual method of validation is known as cross-validation. It involves eliminating each observation in turn, estimating the value at that site from the remaining observations, and comparing the two datasets. To be true cross-validation, the semivariogram should be recomputed and fitted each time an observation is removed. The main difficulty is that point data is usually so sparse that none can be omitted from the surface generation. Furthermore, refitting a semivariogram each time is cumbersome and time consuming. The difficulty with Kriging and all surface generation techniques lies in having enough data to fit an acceptable model to, and then validating the interpolated surface.

76 58

Chapter 3

Figure 3-6. Kriging interpolation based on weighted observations.

These shortcomings can be avoided by using a separate independent set of data for validation. Values are estimated at the sites in the second set, and the predicted and measured values are compared. This is a true validation as described by (Voltz and Webster, 1990). The mean error, ME is given by,

ME =

1 m ¦ {zˆ( xi ) − z( xi )} m i −1

3.7

where z ( xi ) is the measured value at xi , and zˆ( xi ) is the predicted value. The quantity measures the basis of the prediction, and it should be close to zero for unbiased methods. The mean square error, MSE, is defined as,

77 59

3. SURFACE GENERATION 2

1 m MSE = ¦ {zˆ( xi ) − z ( xi )} m i −1

3.8

MSE measures the precision of the prediction and, of course, it is preferred to be as small as possible. The third index is the mean square deviation ratio, R, defined by, 2

1 m R = ¦ {zˆ( xi ) − z ( xi )} / σ Ei2 m i −1 2

3.9

where σ Ei is the theoretical estimation variance for the prediction of z ( xi ) . Voltz and Webster (1990) investigated the performance of simple Kriging applied to soil properties using Eqs. (3.7)-(3.9). Poor performance was found where there are sharp soil boundaries imposed by soil classification and where within-class soil variation is gradual. By combining both spline and Kriging, the weakness of each method was avoided: within class, Kriging avoids the worst effects of local non-stationarity and provides an efficient predictor. However, this procedure still has a few limitations: it reduces, sometimes drastically, the number of points in the Kriging neighborhood and depends on the correctness of the soil classification. Cubic splines and bicubic splines were found to be somewhat unpredictable as interpolators for the soil properties investigated. For soil classification, in an area with discontinuities and strong departures from stationarity, the spline interpolation estimator may still be the most reliable when there are too few data from which to compute a specific variogram in each soil class. Kriging is often criticized because it demands too much data. To compute a variogram with reasonable confidence, at least 100 data points are required for isotropic data (Voltz and Webster, 1990). If variation is anisotropic, then 300 to 400 data points are needed. In the absence of a variogram, values can be predicted by other more empirical methods of interpolation, such as spline methods. Laslett et al. (1987) compared methods for mapping the pH of the soil and concluded that Kriging was the most reliable, with splines being only slightly less. Huang and Yang (1998) investigated the regionalization of streamflow data using Kriging of streamflow regression equations. Estimating statistical parameters from existing stream gauge data requires long records of discharge at any particular location. Because streamflow data may not be available at desired locations, regionalization of statistical parameters is needed. The basic premise is that the amount of streamflow in a basin is related to rainfall and basin characteristics. Numerous techniques have been developed to explore the relationships between precipitation and runoff.

78 60

Chapter 3

Kriging can be used to estimate ungauged streamflows corresponding to different exceedance probabilities. Regression analysis can be used to build a model to estimate streamflow at ungauged locations. Following the approach of Huang and Yang (1988) we can write discharge volume, Q as a function of basin area, A, such that,

Qi = c( Ai ) n

3.10

where Qi is the discharge volume for a given period at site i , Ai is the basin area above the i-th site; c and n coefficients. Qi / Ai (runoff depth per unit of drainage area). Considering Qi / Ai as a regionalized variable, this quantity becomes one realization of possible runoffdepth random functions. The mathematical expectation of this variable gives:

E[Qi / Ai ] = E{[c( Ai ) n −1 ]} = f ( Ai )

3.11

where n=1 and n≠1 correspond to stationary and nonstationary conditions, respectively. In the case of nonstationarity, the expected value, E[], is no longer a constant, which requires universal Kriging that considers drift. However, the use of universal Kriging can be troublesome because the variogram cannot be obtained directly from the data. In the case of stationarity (n=1), the mean runoff depth (E[Qi/Ai]) is considered constant throughout the basin. Where the hypothesis of stationarity is valid, ordinary Kriging is applicable. In practice, the experimental variogram is determined from the measurement data. The validity of the selected variogram model is proved through a cross-validation process. By solving the Kriging equations for estimates of the areal runoff depth throughout the basin, a contour map can be drawn. The variance of the estimation error can also be plotted to visualize the uncertainty of streamflow estimation. The contour maps serve as the basis for estimating the streamflow from ungauged watersheds of interest. Multiplying the runoff depth corresponding to the centroid of the area above the site by the area will yield the estimated streamflow quantity (Huang and Yang, 1988). In developing the experimental variogram, both the isotropic and anisotropic cases should be considered. The validity of the fitted variogram model should be verified in terms of the mean of standardized residual error

3. SURFACE GENERATION

79 61

(MRE) and variance of standardized residual error. We now turn to the spline method for generating a surface. 1.2.3

Spline

Splines are a class of functions useful for interpolating a point between measured values and for surface generation. Cubic splines are the most commonly used for interpolation. A one-dimensional cubic spline consists of a set of cubic curves, i.e., polynomial functions of degree three, joined smoothly end-to-end so that they and their first and second derivatives are continuous. They join at positions known as knots, which may be either at the data points, in which case the spline fits exactly, or elsewhere. In the latter case, the spline is fitted to minimize the sum of squares of deviations from the observed values and may be regarded as a smoothing spline. The number and positions of the knots affect the goodness of the fit and the shape of the interpolated curve, or surface in two dimensions. If the knots are widely spaced, local variations are filtered out. On the other hand, too many knots can cause the spline to fluctuate in a quite unwarranted fashion. To choose sensibly, the investigator must have some preconception of the kind of variation to expect, even though the technique demands fewer assumptions than other methods (Voltz and Webster, 1990). The thin plate spline (TPS) is defined by minimizing the roughness of the interpolated surface, subject to having a prescribed residual from the data. The TPS surface is derived by solving for the elevation, z(x), with continuity in the first and second order derivatives:

z ( x ) = g ( x ) + h( x )

3.12

where g(x) is function of the trend of the surface, which depends on the degree of continuity of the derivatives; and h(x) contains a tension term that adjusts the surface between the steel plate and membrane analogies. The mathematical details for solving this from a variational viewpoint are contained in Mitasova and Mitas (1993), and comparison of splines with Kriging, in Hutchinson and Gessler (1994). The TPS method is a deterministic interpolation method using a mathematical analogy based on the bending of a thin steel plate. TPS is attractive because it is able to give reasonable-looking surfaces without the difficulty of fitting a variogram. Moreover, comparative studies among different interpolators often find that deterministic techniques do as well as Kriging. In this context, it is interesting to consider the ties of Kriging to deterministic methods. In fact, some deterministic interpolators can be considered to have an implicit covariance structure embedded within them.

80 62

Chapter 3

Matheron (1981), for instance, elucidated the connections between spline surface fitting technique and Kriging and demonstrated that spline interpolation is identical in form to Kriging with fixed covariance and degree of polynomial trend. Borga and Vizzaccaro (1997) found that a formal equivalence could be established between multiquadratic paraboloid surface fitting and Kriging with a linear variogram model. Spline methods are based on the assumption that the approximating function should pass as closely as possible to the data points and be as smooth as possible. These two requirements can be combined into a single condition of minimizing the sum of the deviations from the measured points and the smoothness of the function. Multidimensional interpolation is also a valuable tool for incorporating the influence of additional variables into interpolation, e.g., for interpolating precipitation with the influence of topography, or concentration of chemicals with the influence of the environment in which they are disturbed. The data-smoothing aspects of TPS have also remained largely neglected in the geostatistical literature. The links between splines and Kriging suggest that data-smoothing methodology could be applied advantageously with Kriging. Smoothing with splines attempts to recover the spatially coherent signal by removing noise or short-range effects. Splines are not restricted to passing exactly through the data point. Data usually contains measurement error or uncertainty making this restriction unrealistic. The goal of surface interpolation is retrieval of the spatial trend in spite of the noise. Figure 3-7 shows a schematic diagram of a spline surface passing precisely through nine points Z1-Z9. A grid value, Zg may then be interpolated at any location under the surface. Depending on the spline algorithm, various degrees of continuity can be enforced at each point. The resulting surface may be more like a membrane than a thin sheet of steel, depending on the continuity in higher order derivatives. Often, difficulties in making meaningful estimates from sparse data suggest the use of deterministic techniques such as splines. In comparison with Kriging, splines are not computed with known variance. Voltz and Webster (1990) recommend a posteriori validation of the surface interpolation when using splines. Various phenomena like terrain, climatic variables, surface water depth, and concentration of chemicals can be modeled at a certain level of approximation by smooth functions. Hutchinson and Gessler (1994) found that splines give rise to smooth, easily interpretable surfaces, while Kriged surfaces may or may not be particularly smooth, depending on the differentiability of the variogram at the origin. Since the minimum error properties of Kriging depend on the accuracy of the fitted variogram model, it is not immediately apparent whether spline or Kriging results in interpolation that is more accurate. Situations exist in

81 63

3. SURFACE GENERATION

which the control over surface roughness/smoothness exhibited by splines can produce interpolated surfaces that are at least as accurate as the Kriged surfaces.

Z2

Z3

Z1

Zg

Z4

Z5

Z7

Z6 Z8 Z9

Figure 3-7. Surface generation and interpolation using splines.

1.2.4

Generalizations of Splines

Thin plate splines are capable of several linear and non-linear generalizations. Hutchinson and Gessler (1994) list three direct generalizations of the basic thin plate spline formulations: Partial thin plate spline. This version of splines incorporates linear submodels. The linear coefficients of the submodels are determined simultaneously with the solution of the spline. They may be solved using the same equation structure as for ordinary thin plate splines. Correlated error structure. Using simple parametric models for the covariance structure of the measurement error, it can be shown that the spline and Kriging estimation procedures give rise to solutions which are not consistent as data density increases. This is because highly correlated

82 64

Chapter 3 additional observations give essentially no further information. Nevertheless, spline and Kriging analyses can result in significant error reduction over the original data. Other roughness penalties. A weakness of the basic thin plate spline formulation is that the spline has reduced orders of continuity at the data points. This effect may be countered by increasing the order of the derivative. Alternatively, the structure of the roughness penalty may be altered.

Spline surfaces that impose data smoothing yield interpolation errors similar to those achieved by Kriging. Though the TPS method has not been used as widely in the geosciences as Kriging, it can be used with advantages where the variogram is difficult to obtain. There is no guidance as to how to adjust the parameters of the interpolation method, TPS or IDW, except by visual inspection or crossvalidation. Experience with TPS indicates that it gives good results for various applications. However, overshoots often appear due to the stiffness associated with the thin plate analog. Using a tension parameter in the TPS enables the character of the interpolation surface to be tuned from thin plate to membrane. Mitasova and Mitas (1993) evaluated the effect of choosing both the order of derivative controlling surface smoothness, and enforcement of continuity. The amount of data smoothing or tension is used to control the resulting surface appearance. A class of interpolation functions developed by Mitasova and Mitas (1993) is completely regularized splines (CRS). The application of CRS was compared with other methods and was found to have a high degree of accuracy. TPS derived in a variational approach by Mitasova and Hofierka (1993) has been incorporated into the GRASS program call s.surf.tps. This program interpolates the values to grid cells from x,y,z point data (digitized contours, climatic stations, monitoring wells, etc.) as input. If irregularly spaced elevations are input, the output raster file is a raster of elevations. As an option, simultaneously with interpolation, topographic parameters slope, aspect, profile curvature (measured in the direction of steepest slope), tangential curvature (measured in the direction of a tangent to contour line) or mean curvature are computed. The completely regularized spline has a tension parameter that tunes the character of the resulting surface from thin plate to membrane. Higher values of tension parameter reduce the overshoots that can appear in surfaces with rapid change of gradient. For noisy data, it is possible to define a smoothing parameter. With the smoothing parameter set to zero (smooth=0), the resulting surface passes exactly through the data points. Because the values of these parameters have little physical meaning, the surface must be judged

83 65

3. SURFACE GENERATION

by some independent means, such as cross-validation using the GCV method. Figure 3-8 shows an analysis wherein a tension parameter somewhere around 20 would minimize the cross-validation error. Note that smoothing does improve the surface for this particular data set.

120

Crossvalidation x 100

110 Smoothing=0 Smoothing=0.1 Smoothing=0.25 Smoothing=0.5

100

90

80

70

60 0

20

40

60

80

100

Tension

Figure 3-8. Thin Plate Spline cross-validation error associated with smoothing and tension parameters used in s.surf.tps from Mitasova and Mitas (1993).

The choice of surface generation method and corresponding parameters for a particular dataset is somewhat subjective unless cross validation is performed. It is not always clear which method or set of interpolation parameters yields the best results or most realistic surface. The resulting surface should be evaluated from a hydrologic perspective, which examines whether physical reality is violated, if spurious local gradients are introduced, or if important hydrologic or geomorphic features are preserved.

84 66 1.3

Chapter 3 Surface Generation Application

Generating a topographic surface from surveyed data points for the purposes of distributed hydrologic modeling requires some choice of surface generation method. This section demonstrates the effects of smoothing and tension parameters on a surface generated using the thin plate spline algorithm, s.surf.tps of Mitasova and Mitas (1993). The Wankama catchment is located in Niger, in the Sahel of West Africa. This region is the subject of a large-scale energy and hydrologic campaign, called HAPEX, conducted by many organizations under the leadership of the French agency, l’Institut de Recherche pour le Développement (IRD), formerly known as ORSTOM (Goutorbe et al., 1994). Because the Wankama catchment is typical of the region, characterizing one catchment has implications for understanding how and to what degree rainfall is transformed into runoff that collects in the valley bottoms called kori. Small lakes (mares) located in the kori are the major source of water in the region and are responsible for much of the recharge to a shallow ground water table. Elevation data was obtained by surveying the landsurface of the Wankama catchment, recording elevation and location in a relative coordinate system, and then adjusting vertically to mean sea level and horizontally to the projected coordinate system, UTM. This small endoreic catchment (2 km2) shown in Figure 3-9 is representative of the discontinuous drainage network in this part of the Sahel. On the left (West), a large number of contours pass between the data points, resulting in a very rough surface. These are artifacts of the surfacing algorithm responsible for much of the spatial variability. This is a tent-pole effect where the surface drapes like fabric supported with poles located over the data points. The distributed hydrologic model, r.water.fea was then applied in the catchment using the derived DEMs (Cappelaere et al. 2003). Even though the surface passes through the data points, its physical reality is reduced. The black features that nearly cover the western portion of the watershed are closely spaced between surveyed elevation points. The tension parameter affects the smoothness of the resulting surface. As the tension parameter increases, the surface takes on a uniform appearance. However, this smoothing may remove from the surface actual landforms such as ravines. A tension parameter value should be chosen such that artifacts are removed while important landscape features are preserved.

3. SURFACE GENERATION

85 67

Figure 3-9. Wankama catchment aerial extent draining from a plateaux in the west to a sandclogged valley bottom (kori and mare) in the east. (Data courtesy of l’Institut de Recherche pour le Développement, laboratoire d’Hydrologie Montpellier, France, now IRD).

The difficulty with surface generation is deciding which parameter value to choose because a range of surfaces results from seemingly arbitrary choices. Low smoothing and tension generates artifacts of the algorithm with spatial variability that may not be physically realistic. Too much smoothing and tension eliminate spatial variability, leading to a smooth surface that is visually pleasing, but with many terrain features eliminated. Whether these features are important to the ultimate goal of hydrologic modeling depends on the scale of the terrain features and the purpose of the modeling effort. If the purpose is to model erosion rates using the generated surface, local gradients due to the tent-pole artifact may produce increased runoff velocities and higher erosion rates than are reasonable. Figure 3-10 shows a surface generated using a range of tension parameters. Differences are made evident by the spacing of the contour lines. The effect of the tension parameter on surface generation is illustrated by the Wankama catchment, which is the subject of distributed hydrologic modeling reported by Séguis et al. (2003), Peugeot et al. (2003) and Cappelaere et al. (2003).

86 68

Chapter 3

Tension 20

Tension 60

Tension 80

Figure 3-10. Effect of the tension parameter (20, 60, 80) on elevation contours (Data courtesy of l’Institut de Recherche pour le Développement, laboratoire d’Hydrologie Montpellier, France, now IRD).

3. SURFACE GENERATION

87 69

The calibration of the distributed-parameter model for this basin relies on differential infiltration rates, higher rates for the ravines and lower values for crusted overland flow areas. The terrain was represented using a raster DEM derived from point elevation measurements. A topographic surface was then interpolated using the TPS (s.surf.tps) algorithm. Very different surfaces resulted from the interpolation using s.surf.tps depending on the choice made for tension and smoothness parameters. The resulting DEM was used to derived the drainage network and slope for hydrologic modeling of rainfall runoff in this catchment. Closer examination through visual interpretation can reveal the adequacy of the generated surface. Too high tension causes smoothing of fine-detailed topography, whereas too low tension causes tent-pole artifacts to appear. Figure 3-11 shows the contours near the plateau in the upper part of the Wankama catchment (western) resulting from a tension of 40 and 80.

Figure 3-11. Contours generated using s.surf.tps using tension parameters set to 40 (light grey) and 80 (black) for the Wankama Upper basin area.

It is readily seen that the lower tension parameter equal to 40 (shown in gray) preserves some landscape features such as the ravine better than with the higher tension=80 (shown in black). A more detailed view of the plateau area to the western portion of the watershed reveals the differences in the

88 70

Chapter 3

surfaces generated with the 40 and 80 tension parameters. More ravine detail is seen in the contours produced using a tension parameter of 40 than with 80. Large differences result in surfaces generated with the various parameters of tension. Besides heuristics and cross-validation, no a priori guidance is available to help us choose correct values. Erosion and sedimentation simulation may require much more detailed topography than is generally available from common digital data sets. The best surface is often chosen based on how well it preserves important topographic features. 1.4

Summary

Simulation of hydrologic processes requires extension of point measurements to a surface representing the spatial distribution of the parameter or input. With many surface generation techniques, unintended artifacts may be introduced. Gradients that are not physically significant may result along edges or where data are sparse. When using the IDW, the tent pole effect may be overcome by artificially making the number of points more dense or by obtaining more measurements. The IDW has no fundamental means of control except to limit the neighborhood, in terms of either a radius or number of points used, and the exponent (friction) used to interpolate the grid. Kriging, on the other hand, interpolates with weights assigned to neighboring values, depending on a spatially dependent model of variance. Control of the resulting surface is possible with the various spline algorithms. TPS permits adjustment of smoothing and tension for more representative surfaces. Withholding measurements for validation of the surface permits some independent means of assessing whether the surface is representative of the physical feature. This is often difficult because all data is usually needed to create a representative surface. Visual inspection may be the only alternative to cross-validation in assessing whether a reasonable reproduction of the surface resulted from the surface generation algorithm. The ultimate goal is to generate a surface that reproduces the hydrologic character of the actual landsurface. This may entail creating several DEMs, deriving slope and other parameters, and then running a hydrologic model to observe the suitability of the surface for hydrologic simulation. In the next chapter, we will examine means of assessing which resolution is sufficient for capturing spatial detail from an information content viewpoint.

3. SURFACE GENERATION 1.5

89 71

References

Bartier, P. M. and Keller, P., 1996, “Multivariate interpolation to incorporate thematic surface data using inverse distance weighting (IDW).” Computers & Geosciences, 22(7):795-799. Borga, M. and Vizzaccaro, A., 1997, “On the interpolation of hydrologic variables: formal equivalence of multiquadratic surface fitting and Kriging.” J. of Hydrol.,195:160-171. Cappelaere, B., B.E. Vieux, C. Peugeot, A. Maia, L. Seguis, 2003. “Hydrologic process simulation of a semi-arid, endoreic catchment in Sahelian West Niger, Africa: II. Model calibration and uncertainty characterization.” J. of Hydrol., 279(1-4) pp. 244-261. Chaturvedi, A.K. and Piegl, L.A., 1996, “Procedural method for terrain surface interpolation.” Computer & Graphics, 20(4):541-566. Clarke, K. C, 1990, Analytical and computer cartography. Prentice-Hall, Englewood, New Jersey, 290p. Franke, R., 1982, “Smooth interpolation of scattered data by local thin plate splines”. Comput. Math. Appl., 8:273-281. Goutorbe J-P., T. Lebel, A. Tinga, P. Bessemoulin, J. Bouwer, A.J. Dolman, E.T. Wngman, J.H.C. Gash, M. Hoepffner, P. Kabat, Y.H. Kerr, B. Monteny, S.D. Prince, F. Saïd, P. Sellers and J.S. Wallace, 1994, “Hapex-Sahel: a large scale study of land-surface interactions in the semi-arid tropics.” Ann. Geophysicae, 12: 53-64 Huang, Wen-Cheng and Yang, F.-T., 1998, “Streamflow estimation using Kriging.” Water Resour. Res., 34(6):1599-1608. Hutchinson, M. F. and Gessler, P. E., 1994, “ Splines-more than just a smooth interpolator.” Geoderma, 62:45-67. Journel, A. G. and Huijbregts, C. J., 1978, Mining Geostatistics. Academic Press, New York. Laslett, G. M., McBratney, A. B., Pahl, P. H., and Hutchinson, M. F, 1987, “Comparison of several spatial prediction methods for soil pH.” J. of Soil Sci., 37:617-639. Matheron, G., 1965, Les Variables Régionalisées et leur Estimation, Masson, Paris. Matheron, G., 1981, “Splines and Kriging: their formal equivalence.” In: Down to Earth Statistics: Solutions Looking for Geological Problems. Syracuse University Mitasova, H. and Hofierka, J., 1993, “Interpolation by Regularised Spline with Tension: II. application to Terrain Modeling and Surface Geometry Analysis.” Mathematical Geology, 25(6):657-669. Mitasova, H. and Mitas, L., 1993, “Interpolation by Regularized Spline with Tension I. Theory and Implementation.” Mathematical Geology, 25(6):641-655. Peugeot, C., B. Cappelaere, B.E. Vieux, L. Seguis, A. Maia, 2003, “Hydrologic process simulation of a semi-arid, endoreic catchment in Sahelian West Niger, Africa: I. Modelaided data analysis and screening.” J. of Hydrol., 279(1-4) pp. 244-261. Séguis, L., B. Cappelaere, C. Peugeot, B.E. Vieux, 2002, “Impact on Sahelian runoff of stochastic and elevation-induced spatial distributions of soil parameters.” J. of Hyd. Proc., 16(2): 313-332. Todini, E and Ferraresi, M., 1996,”Influence of parameter estimation uncertainty in Kriging”. J. of Hydrol., 175: 555-566. Voltz, M. and Webster, R., 1990, “A comparison of kriging, cubic splines and classification for predicting soil properties from sample information.” J. of Soil Science, 41:473-490.

This page intentionally left blank

Chapter 4 SPATIAL VARIABILITY Measuring Information Content

Rainfall Runon Runon Runoff Runon

Infiltration

Figure 4-1. Fully distributed grid-based runoff schematic for a 3x3 kernel in the Blue River Basin. The inset in the upper right shows spatially distributed runoff generated using a 270meter resolution grid.

When building a fully distributed grid-based distributed hydrologic model from geospatial data, a fundamental question is what resolution is adequate for capturing the spatial variability of a parameter or input. Figure 4-1 above shows the schematic for computing runoff in a 3x3 kernel. The model computational elements, whether finite element or difference, require

92 74

Chapter 4

parameter values that are representative of the grid cell. In the composite for the model should then be representative of the spatial variation found in the watershed. When solving conservation of mass, momentum, or energy, a representative parameter value is required for that element size. The correspondence of computational element size and the sampled resolution of the digital map affect how well the model will represent the spatial variation of parameters controlling the process. Within each grid cell, the conservation of mass and momentum is controlled by the parameters taken from special purpose maps whose values are hydraulic roughness, rainfall intensity, slope, and infiltration rate for each computational element. The grid cell in the GIS map supplies this value for use in the model. To take advantage of distributed parameter hydrologic models, the spatial distribution of inputs (e.g., rainfall) and parameters (e.g., hydraulic roughness) should be sampled at a sufficiently fine resolution to capture the spatial variability. 1.1

Introduction

Hydrologic simulations and model performance are affected by the seemingly arbitrary choice of resolution. Additionally, in the processing of DEMs, smoothing or other filters are applied that may have deleterious effects on slope and other derived terrain attributes. This chapter deals with the resolution necessary or sufficient to capture spatial information that controls hydrologic response. Information content can be measured using a statistical technique developed in communications called informational entropy. Having a statistic that is predictive of the hydrologic impacts brought about by smoothing and resampling to another grid-cell resolution, especially DEMs, is quite useful. Fully distributed grid-based hydrologic models require parameter values and precipitation input in every grid cell. The size of the grid cell determines how much spatial variability will be built into the model. Another way to view resolution is as a sampling frequency. Depending on the sampling frequency used and the spatial variability of the surface, the information may be under- or over-sampled. The goal is to capture the information content that is contained in a map of slope, soil infiltration parameters, surface hydraulic roughness, or rainfall. Each map and derivative parameter map is likely to have different resolutions that are necessary for capturing the spatial variability. Another consideration is that there may be more variability present in the actual surface than is mapped. The measurement of spatial variability at specific sampling frequency (resolution) is useful for deciding which resolution should be used for each map. Practical considerations involve finding the smallest grid-cell size and then using that for the other maps even if this involves over-sampling less-variable parameter maps.

4. SPATIAL VARIABILITY

93 75

The resolution that is sufficient to capture the variability of a parameter depends on how variable the values are spatially in the map. If the parameter is actually constant, then any resolution is adequate for sampling the surface. However, the more likely case is that the parameter or input is spatially variable, requiring a decision as to the most efficient choice of resolution. Opting for the smallest resolution possible wastes storage space and computational efficiency. Depending on the extent or size of the river basin, the resolution that adequately represents the spatial variability important to the hydrologic process may range from tens or hundreds of meters for hillslopes to hundreds or thousands of meters for river basins. For very large basins, even larger resolutions may be appropriate. The real question is whether we can use larger resolutions to minimize the use of computational resources and still represent the spatial variability controlling the response. Spatial data is widely recognized as having a high degree of autocorrelation. That is to say, adjacent cells tend to have values similar in magnitude. The correlation distance is the length scale that separates whether sampled values appear to be correlated or independent. This same concept was dealt with in Chapter 3, where the range of spatial dependence in the Kriging method determined the weights for neighboring data values. Sampling at resolutions greater than this length will produce variates that are independent, showing no high degree of autocorrelation. Sampling at resolutions smaller than this length will produce variates with a high degree of autocorrelation. Deriving maps of parameters from soil, land use, or other areal classification maps produces parameter maps that appear to be constant within the polygon delineating the class. More variation undoubtedly exists than is mapped. The within-class variation may or not be important to the simulation of the hydrologic process. Variation within the class may be assumed to follow some type of probability distribution or assumed to be a mean value. Many of the perceived problems with physics-based distributed models are difficulties associated with parameterization, validation and representation of within-grid processes. However, a strength of physicsbased models is the ability to incorporate the spatial variability of parameters controlling the hydrologic process. However, as Beven (1985) pointed out, the incorporation of spatial variability in physics-based models is possible only “...if the nature of that variability were known.” The “nature of variability” is a qualitative term describing the kind of variability, in contrast to quantitative terms such as “variance” or “correlated length,” which describe the amount of variability. Specifically, the nature of spatial variability may be represented as deterministic and/or stochastic (Smith and Hebbert, 1979; Philip, 1980; and Rao and Wagenet, 1985). The total

94 76

Chapter 4

variability of a given parameter is a composite of the deterministic and stochastic components. Within-class variation is the stochastic component, whereas between-class variation is deterministic. An important special case is when homogeneity, or no variability, is assumed. In reality, spatial variability is rarely entirely deterministic or stochastic. Within any deterministic trend or distribution of parameter values, there is invariably some degree of uncertainty or stochastic component. Similarly, stochastic variation may be nonstationary; containing systematic trends, or is composed of nested, deterministic variability. A number of studies have demonstrated that stochastic variability can have a large impact on hydrologic response. In particular, those studies have shown that processes such as runoff and infiltration, which result from hypothetical random distribution of parameters, are often not well represented by homogeneous effective parameters. This has led to stochastic modeling approaches and uncertainty prediction. In addition to work on stochastic variability, Smith and Hebbert (1979) have demonstrated that deterministic variability may have important impacts on surface runoff. Another view is to recognize some limiting size (Wood et al., 1990) of a subbasin or computational element called a representative elementary area (REA) to establish a “fundamental building block for catchment modeling.” Smith and Hebbert (1979) noted, however, that the deterministic length scale depends on the scale of interest, the processes involved, and the local ecosystem characteristics. Introducing new sources of deterministic variability related to topography, soils, climate, or land use/cover might be accomplished by mapping at finer resolution and scale. Seyfried and Wilcox (1995) examined how the nature of spatial variability affects hydrologic response over a range of scales, using five field studies as examples. The nature of variability was characterized as either stochastic, when random, or deterministic, when due to known nonrandom sources. In each example, there was a deterministic length scale, over which the hydrologic response was strongly dependent upon the specific, locationdependent ecosystem properties. Smaller-scale variability was represented as either stochastic or homogenous with nonspatial data. In addition, changes in scale or location sometimes resulted in the introduction of larger-scale sources of variability that subsumed smaller-scale sources. Thus, recognition of the nature and sources of variability was seen to reduce data requirements by focusing on important sources of variability and using nonspatial data to characterize variability at scales smaller than the deterministic length scale. The existence of a deterministic length scale implies upper and lower scale bounds. At present, this is a major limitation to incorporating scaledependent deterministic variability into physics- based models. Ideally, the deterministic length scale would be established by experimentation over a

4. SPATIAL VARIABILITY

95 77

range of scales. This, however, will probably prove to be impractical in many cases. In general, the lower boundary will be easier to estimate. Although it is difficult to define the nature of spatial variability with respect to sources of variability, it is important to recognize that these definitions are effectively made during model design and application. The determination to include certain kinds of elevation data or use a certain grid spacing, for example, implies specific treatments, whether intended or not, of spatial variability. These aspects of model design are more likely to be driven by computation time and data availability than by spatial variability. Even when driven by spatial variability, these considerations are not usually stated up front as part of the model. This makes interpreting model results and transferring the model to other locations difficult. Recognition of the nature of variability has other important implications for physics-based modeling. Although physics-based models are generally comprehensive in their consideration of the basic hydrologic processes, yet may fail to account for sources of variability that largely control these processes. This failure may be due to grid sizes that are too large or to lack of appreciation of impact of different variability sources. Dunne et al. (1991), for example, noted the lack of appreciation of shrub influences on snow drifting in alpine hydrologic modeling. Tabler et al. (1990) made similar observations concerning snow drifting. With regard to soil freezing, Leavesley (1989) noted that frozen soil was identified as a major problem area, but none of the 11 models that were included in the study had the capability to simulate frozen soils. The important advantage of physicsbased models over other approaches cannot be expected to accrue unless the effects of critical sources of variability are taken into account. The degree to which a model takes into account critical spatial variability has important effects on model data requirements, grid scale, and measurement scales. Whether a soil is frozen at a particular location and time or the location of shrubs in a landscape verges on unknowability, exemplifies the difficulty of distributed modeling. While the following considerations enumerated by Leavesley (1989) pertain to specific hydrologic applications, they should be taken into account when developing or applying a distributed hydrologic model as follows. Data requirements: The effects of scale on spatial variability suggest at least two reasons why virtually unlimited data (Freeze and Harlan, 1969) and computer demands may not have to be met to simulate hydrologic processes on a physical basis. First, increasing the scale of interest introduces additional sources of variability whose effects on hydrologic response may subsume those from smaller scales. In this context, greater data quality may not mean greater accuracy but simply mean more information.

96 78

Chapter 4 Grid scale: Another important consideration is the model grid size. Physics-based models generally account for spatial heterogeneity with inputs at nodes of a grid system. Grid point parameters are generally assumed to represent homogeneous grid, which results in empirically effective parameters. This approach is necessary for computational reasons and allows for the incorporation of a variety of data inputs via geographical information systems. Grid size should be determined within the context of the nature of the variability. The grid must be small enough to describe important deterministic variability. Measurement Scale: Many of the problems associated with spatial variability and scale have been related to scale of measurement. It has been suggested that the development of large-scale measurement techniques, by remote sensing, for example, will provide model input and verification, which will largely eliminate problems related to spatial variability (Bathurst and O’Connell, 1992).

If a particular data set is under-sampled, important variation in space will be missed, causing error in the model results. Over-sampling at too fine a resolution wastes computer storage and causes the model to run more slowly. The ideal is to find the resolution that adequately samples the data for the purpose of the simulation yet is not so fine that computational inefficiency results. We will assume that the GIS system used to perform the hydrologic analysis of input data has the capability for maintaining each map with its own resolution. The topography and the derived parameters such as slope may be at a relatively small resolution dictated by the available DEM, say on the order of 30 m, whereas precipitation estimates derived from radar or satellite may be at 4 km or larger. Whether the precipitation is adequately sampled for the size of the watershed depends on relative scales. Precipitation at 4 km may be adequate for simulating runoff produced in a 1200 km2 watershed. Yet such a coarse resolution would not be adequate for a smaller watershed of only a one or two km2. 1.2

Information Content

The goal of spatial information content measurement is to have some measure that indicates whether the resolution of a dataset over-samples, under-samples or is simply adequate to capture the spatial variability. We wish to be able to tell whether the resolution is sufficient to capture the information contained in a map of slope, soil infiltration parameters, surface hydraulic roughness, or rainfall.

4. SPATIAL VARIABILITY

97 79

Which resolution suffices for hydrologic purposes is answered in part by testing the quantity of information contained in a data set as a function of resolution. Once the information content ceases to increase as we resample the original data at finer resolution, we can say that we have captured the information content of the mapped parameter and finer resolution is not merited or supported. Borrowing from communication theory, information content is a statistic that may be computed at various resolutions (Vieux, 1993). Such a measure can test which resolution is adequate in capturing the spatial variability of the data. Smoothing and resampling reduce the spatial variability of elevation and the derived slope data. Smoothing and resampling to coarser resolution are essentially data filters that reduce information content. The application of information theory to determining bandwidth needed in a communication channel was first introduced in the landmark theory by Shannon and Weaver (1964). Application of information content to hydrologic modeling was found to be useful by Vieux (1993), Farajalla and Vieux (1995) and Brasington and Richards (1998) among others. Information content helps in understanding the effects of data filters on hydrologic process simulation. Informational content or entropy, I, in this context is defined as, β

I = −¦ Pi log10 Pi i =1

4.1

where β is the number of bins or discrete intervals of the variate and Pi is the probability of the variate occurring within the discrete interval. A negative sign in front of the summation is by convention such that increasing information content results in positively increasing informational entropy. Base 10 logarithm is used in this application, yielding units of Hartleys. Information content in communication theory commonly uses base 2 when operating on bits of information yielding binits of information. For a complete description of informational entropy in communication-theory context, the reader is directed to Papoulis (1984). Informational entropy becomes a measure of spatial variability when applied to topographic surfaces defined by a raster DEM. As the variance increases, so does informational entropy. Conversely, as variance decreases, so does informational entropy. In the limit, if the topographic surface is a plane with a constant elevation, the probability is 1.0, resulting in zero informational entropy, zero uncertainty, and zero information content. Maximum informational entropy occurs when all classes or histogram bins are equally probable. With all bins in the histogram filled, every elevation is equally probable in the DEM. Measuring informational entropy at increasing

98 80

Chapter 4

resolutions provides an estimate of the rate of information loss due to resampling to larger cell sizes. 1.3

Fractal Interpretation

The rate of information loss with respect to cell size can be put into terms of a non-integer or fractal scaling law. As the sampling interval increases with increasing cell size, the information loss is greater for surfaces with higher fractal dimension. In turn, the information loss propagates errors in the hydrologic model output. Thus, information content is an indication of how much spatial variability is lost due to filtering by either smoothing or aggregation. The rate of information loss would vary in any natural topography. High variance at a local scale demands smaller cell sizes to capture the information content. When variance is low (a surface with constant elevation), any number of cells is sufficient to capture the information content including just one cell. For example, a plane surface will have a fractal dimension equal to the Euclidian dimension of 2. A completely flat or uniform slope will not lose information when smoothed or resampled because there is no spatial variability to be lost. On the other hand, topographic surfaces possessing high variability with a fractal dimension greater than 2.0 will experience a higher degree of information loss as smoothing and resampling progress. The information dimension is the fractal dimension computed by determining the information content as measured by informational entropy at differing scales. Several definitions of dimension, including the information dimension, for dynamic systems exhibiting chaotic attractors, are presented by Farmer et al. (1983). The fractal dimension is a measure common to many fields of application. Information content measured at different scales is not a self-similar fractal, because a self-similar fractal exhibits identical scaling in each dimension. In terms of a box dimension, a rectangular coordinate system may be scaled down to smaller grid cells of side ε . Following the arguments of Mandelbrot (1988), a self-affine function (such as informational entropy) follows the nonuniform scaling law where the number of grid cells is scaled by ε H . Writing the scaling law in terms of a proportionality, the number of grid cells, N(ε ) , of side ε is:

N(ε ) ∝ ε H

4.2

where ε is the grid dimension and H is the Hurst coefficient. Note that the fractal dimension d I must exceed the Euclidian dimension and that 0 ≤ H ≤ 1 , therefore,

4. SPATIAL VARIABILITY

99 81

d I = E +1 - H

4.3

H has special significance (Saupe, 1988): at H=0.5 the surface is analogous to ordinary Brownian motion, H < 0.5 there is negative correlation between the scaled functions, and H > 0.5 there is positive correlation between the scaled functions. The information dimension described by Farmer et al. (1983) is based on the probability of the variate within each cube in three dimensions or within a grid cell in two dimensions covering the set of data. As a test case, we examine the dimension of a plane. The concept is extended to the more general topographic surface: the study watershed basin in following sections. The plane surface is subdivided into cells of ε = 1/5th, 1/125th, and 1/625th. The plane surface is of uniform slope such that each elevation has equal probability and I(ε ) = log(N(ε )) . If a plane surface of uniform slope in the direction of one side of the cells is divided into a five by five set of cells, it will have five unique rows of equally probable elevations. There are five categories of elevation in the DEM and five bins. The number of bins is β. Each bin will have a probability of 1/5 and an informational entropy of 0.69897 which is simply log(B) . If the plane is divided into successively smaller grid cells, the rate at which the informational entropy changes should be equal to one for a plane surface. Table 4-1 presents the calculations for the plane of uniform slope and one cell wide. Under the condition of equal probability among classes and exactly one occurrence per class or interval, the number of bins, B equals the number of cells, N for a unit-width plane. The Hurst coefficient, H , is the proportionality factor relating the number of grid cells, N(ε) to the size of the cells (ε). We expect that informational entropy scales with the size of the cell of side ε covering the set according to Eq. (4.1). Taking logarithms of both sides of Eq. (4.1), replacing the proportionality with a difference relation, and recognizing that for equal probabilities, log( N(ε )) = I(ε ) , we find that H is the proportionality constant that relates the rate at which informational entropy changes with grid-cell size. H is related to the fractal dimension, d I by: H=

∆I( ε ) ∆ log(1/ε )

4.4

where I(ε ) is the information content computed for grid cells of side ε . Applying Eq.(4.4) to Table 1, we find that,

H = (2.79588 - 2.0969)/(log(625) - log(125)) = 1.0000

4.5

100 82

Chapter 4

Thus for a two-dimensional surface E=2 and H=1, and by Eq. (4.3), d I =2, as would be expected for a plane surface. Furthermore, H =1.0 which is greater than 0.5, which indicates that elevations are positively correlated. That is, similar elevations are found close together, a feature that is typical of real terrain. The information dimension is a measure of the rate at which information changes with resolution or by some other filter such as smoothing. Table 4-1 demonstrates that the information dimension measures the rate at which information changes for a plane surface. H=1 and dI=E. It also shows that any resolution is sufficient to capture the spatial variability. Table 4-1. Information content of a plane surface with uniform slope. I(ε) ε β 1/625 625 2.7958 1/125 125 2.0969 1/25 25 1.3979 1/5 5 0.6990

We have demonstrated how resolution affects the information content sampled by a range of resolutions for an idealized surface. Next, we turn our attention from a theoretical plane surface with uniform slope to the effect of resolution on an actual DEM that is used to derive slope. 1.4

Resolution Effects on DEMs

Resampling a DEM at larger resolutions means that we take the original resolution and increase the resolution by selecting the cell closest to the center of the new larger cell size. If we increase by odd multiples of the original resolution, say 3x3 or 5x5 windows, the center cell of the window is used as the elevation for the new larger cell. As resampling progresses, the topographic data becomes a data set of larger and larger cells. Resampling is a means of investigating the effects of using larger cell sizes in hydrologic modeling. The appropriate cell size or resolution is the largest one that still preserves essential characteristics of the digital data and its spatial characteristics or distribution. To gain an appreciation for how information content is related to spatial variability, we start with a plane surface. Informational entropy applied to spatially variable parameters/input maps is useful in assessing the resolution that captures the information content. Special considerations for different types of data are addressed in the following applications. The choice of resolution has an important impact on the hydrologic simulation. Choosing a coarse resolution DEM, deriving slope, and using this in a surface runoff model has two principal effects. One is to shorten the

4. SPATIAL VARIABILITY

101 83

drainage length, because many of the natural meanders or crookedness of the drainage network is short-circuited by connecting raster grid cells together by way of the principal slope. The effect of grid-cell resolution on the drainage network is addressed in Chapter 7. The other effect is a flattening of the slope due to a sampling of the hills and valleys at too coarse resolution. We can think of it as cutting off the hilltops and filling in the valleys. These effects together may have compensating effects on the resulting hydrograph response. A shortened drainage length decreases the time taken by runoff from the point of generation to the outlet. A flattened slope will increase the time. Brasington and Richards (1998) examined the effects of cell resolution on TOPMODEL and found that information content predicted a break in the relation between model response and resolution. Sensitivity analyses revealed that model predictions were consequently grid-size dependent, although this effect could be modulated by recalibrating the saturated hydraulic conductivity parameter of the model as grid size changed. A significant change in the model response to scale was identified for grid sizes between 100 and 200 m. This change in grid size was also marked by rapid deterioration of the topographic information contained in the DEM, measured in terms of the information content. The authors suggested that this break in the scaling relationship corresponds to typical hillslope lengths in the dissected terrain and that this scale marks a fundamental natural threshold for DEM-based application. Quinn et al. (1991) presents the application of TOPMODEL which models subsurface flow at the hillslope scale. Model sensitivity to flow path direction derived from a DEM was investigated. The application by Quinn et al. (1991) used a 50-meter grid-cell resolution, which is the default value of the United Kingdom database. Resampling at larger grid-cell resolutions was found to have significant effects on soil moisture modeling due to aggregation. The drainage network extracted from a DEM is affected by DEM resolution. Tarboton et al. (1991) investigated stream network extraction from DEMs at various scale resolutions. They found the drainage network density and configuration to be highly dependent on smoothing of elevations during the pit removal stage of network extraction. In fact, if smoothing was not applied to the DEM prior to extraction of the stream channel network, the result did not resemble a network. While smoothing may be expedient, it can have deleterious effects, viz., undesirable variations in delineated watershed area. Relating the quantitative effects of data filters to hydrologic simulation was reported by Vieux (1993). The effects on hydrologic modeling produced are measured using information content affected by two types of filters:

102 84

Chapter 4

smoothing and cell aggregation. As mentioned above, cell size selection is important in capturing the spatial variability of the DEM. Smoothing is often necessary before automatic delineation of the watershed and stream network to reduce the number of spurious high or low points, referred to as pits and spikes. Both smoothing and resampling to larger resolutions have the effect of flattening the slope derived from such a filtered DEM. A nearly linear relationship was found between delayed hydrograph response and the relative informational entropy loss due to both smoothing and aggregation. The error introduced into the simulation was not constant for all rainfall intensities. Higher intensity events revealed less change in response to filtering than did less intense events. Because the more intense events achieved equilibrium more quickly, the changes in spatial variability have less effect. Though rare in nature, an equilibrium hydrograph means that input equals output. When input (rainfall excess) equals output (discharge at the outlet), effects of spatial variability in the output are no longer present. Resolution effects on information content of elevation and slope can influence modeled response that relies on drainage length and slope. A decrease in the average slope of the Illinois River basin, as shown in Figure 4-2, results with increasingly coarse resolution.

Mean Basin Slope (%)

4

3.5

3

2.5

2 0

200

400

600

800

1000

1200

Resolution (m)

Figure 4-2. Slope flattening due to increased DEM resolution.

Using DEMs of different resolutions may produce different hydrographs due to flattening of the slope. In Figure 4-3, we see that information content decreases as well when resampling the DEM from 30 m to 960 m, then deriving slope maps, steadily decreases the mean basin slope. At 960-m resolution, there is essentially one slope class around 2.5% and therefore

103 85

4. SPATIAL VARIABILITY

zero information entropy. Little or no loss of information content was found for elevation during the resampling. This means that at these resolutions, equivalent distributions of elevations result.

Relative Change in Mean Basin Slope (%)

1.40 1.20

y = 0.477x R 2 = 0.996

1.00 0.80 0.60 0.40 0.20 0.00 0.00

0.50

1.00

1.50

2.00

2.50

3.00

Relative Change in Information Entropy (%)

Figure 4-3. Information entropy of slope lost due to resampling of DEM.

Resampling a DEM to a larger resolution decreases the slope derived from it. Resampling a DEM from 30 to 60 m and deriving the slope map shows the mean slope decreasing from 3.34 to 3.10%. This may be an acceptable degree of flattening, given that computer storage decreases to 302/602 or ¼ of that required to store the 30-meter DEM. Besides reduced storage, subsequent computational effort will be more efficient at larger resolution. A physical interpretation is that the hillslopes are not sampled adequately, resulting in a cutting of the hilltops and filling of the valleys. As slope is derived from coarser-resolution DEMs, the steeper slopes decrease in areal extent and are reflected in the decrease in mean slope. Figure 4-4 shows four maps produced from resampled DEMs. The darker areas are the steeper slopes (> 10%). Flatter slopes in the class 0-1% also increase, as is evident particularly in the upper reaches of the basin. Information content is reduced primarily because the bins in the histogram are not filled. This same effect of resolution-induced flattening can be seen in Figure 45, which reveals how steeper slope classes have fewer counts than flatter classes. As steeper slopes drop out of the distribution, they do not contribute to the information content statistic. At the same time, average watershed slope becomes flatter with coarser resolution.

104 86

Chapter 4

Figure 4-4. Slope maps derived from 30, 240, 480, and 960 m DEMs.

Figure 4-5. Histograms of slope derived from DEMS at 30, 240, 480, and 960-m.

87 105

4. SPATIAL VARIABILITY

The information content loss decreases on a relative basis compared to the finest resolution, as well as the mean slope. Figure 4-6 shows a nearly linear relationship (r2 = 0.96) between the relative loss in mean slope and information content.

1.40

Basin Mean Slope

1.20 1.00 0.80 0.60 0.40 0.20 0.00 0.00

0.50

1.00

1.50

2.00

2.50

3.00

Information Entropy

Figure 4-6. Relative loss in slope in relation to loss in information content.

Relative loss means the amount lost in relation to the information entropy or slope at 30-meter resolution. The decrease in mean slope is about two times the loss in information content. By resampling, if the information content decreases by 100%, then a 50% decrease in mean slope should be expected. The rate of decline for other regions will depend on the initial starting resolution and the length scales of the topography in relation to resolution.

Information content is useful for characterizing the spatial variability of parameters affecting distributed hydrologic modeling. Other methods for characterizing the variation of spatial data have been developed. In any case, the method should provide a robust statistic for characterizing effects of resolution and spatial filtering that is often necessary in preparation of parameters from geospatial data.

106 88

Chapter 4

Other approaches involving the measurement of spatial autocorrelation have been introduced such as the Moran Index. Moran introduced the first measure of spatial autocorrelation in order to study stochastic phenomena, which are distributed in space in two or more dimensions (Moran, 1950). The index is essentially a correlation coefficient evaluated for a group of adjacent or closely spaced data values. The value of Moran’s I ranges from +1 meaning a strong positive spatial autocorrelation, to -1 indicating strong negative spatial autocorrelation. A value of 0 indicates a random pattern without spatial autocorrelation. This statistic is designed for application to gridded ordinal, interval or ratio data. The Moran’s I is given as,

4.6 where Wij is the binary weight matrix of the general cross-product statistic, such that Wij.=1 if locations i and j are adjacent; and Wii=0 for all cells, points or regions which are not adjacent. Spatial data used in distributed hydrologic modeling are generally autocorrelated in space. Resampling at different resolutions will affect the mean value of the parameter. Completely random spatial fields, without autocorrelation, can be sampled at almost any resolution without affecting the mean value. Therefore, using an index, whether based on the Moran I, information content or other similar measure is useful for understanding the effects of resolution and filtering on slope and other parameters. 1.5

Summary

Changing cell resolution and smoothing of DEMs can have important effects on hydrologic simulation revealing dependence on grid-cell size. We have developed a measure of spatial variability called information content adapted from communication theory. It is easily applied to a map of parameters to gauge the effects of resolution or other filters. If information content is not lost as we increase cell size, we may be able to use coarser resolution maps, saving computer storage and making simulations more computationally efficient. Application of information content as a measure of spatial variability is demonstrated where derived slope maps become flatter. Selecting a too coarse resolution can be thought of as cutting of the hilltops and filling of the valleys. Flatter slopes derived from coarse DEMs will produce more delayed and attenuated hydrographs. Care should be taken

4. SPATIAL VARIABILITY

107 89

when using coarse resolution DEMs to recognize that the derived slope maps will have important consequences on hydrologic simulations. Information content may be used to assess the spatial variability of parameters and input maps used to simulate hydrologic processes. Distributed parameter models are dependent on the grid-cell size. When scaling up from one cell size to another, this dependence should be recognized. Having seen the impact on slope due to the choice of DEM resolution, we next examine the effect of resolution on drainage network length. 1.6

References

Bathurst, J. C. and O’Connell, P. E., 1992, “Future of distributed modeling - The Systemhydrologique-Europeen.” J. Hydrol. Process, 6(3):265-277. Beven, K., 1985, “Distributed Models.” In: Hydrological Forecasting. Edited by. Anderson, M. G. and Burt, T. P. John Wiley, New York pp.405-435. Brasington, J. and Richards, K., 1998, “Interactions between Model predictions, Parameters and DTM Scales for Topmodel.” Computers & Geosciences, 24(4):299-314. Dunne, T., Zhang, X. C., and Aubry, B. F., 1991, “Effects of rainfall, vegetation and microtopography on infiltration and runoff.” Water Resour. Res., 27 pp.2271-2285. Farajalla, N. S. and Vieux, B. E., 1995, “Capturing the Essential Spatial Variability in Distributed Hydrologic Modeling : Infiltration Parameters.”J. Hydrol. Process, 8(1):55-68. Farmer, J. D., Ott, E, and Yorke, J. A, 1983, “The Dimension of Chaotic Attractors.” Physica D: Nonlinear Phenomena,7(1-3), North-Holland Publishing Co., Amsterdam pp.265-278. Freeze, R. A. and Harlan, R. L., 1969, “Blueprint of a Physically-based digitally-simulated Hydrologic Response Model.” J. of Hydrol., 9:237-258. Leavesley, G. H, 1989, “Problems of snowmelt runoff modeling for a variety of physiographic climatic conditions.” Hydrol.Sci., 34:617-634. Mandelbrot, B. B., 1988, The Science of Fractal Images. Edited by. Heinz-Otto, Peitgin and Saupe, Deitmar. Springer-Verlag, New York pp.2-21. Moran, P.A.P., 1950, Notes on continuous stochastic phenomena. Biometrika 37: 17-23. Papoulis, A., 1984, Probability, Random Variables and Stochastic Processes. 2nd edition McGraw-Hill, New York pp.500-567. Philip, J. R., 1980, “Field heterogeneity.” Water Resour. Res., 16(2) pp.443-448. Quinn, P., Beven, K., Chevallier, P., and Planchon, O., 1991, “The Prediction of Hillslope Flow Paths for Distributed Hydrological Modeling using Digital Terrain Models.” In: Terrain Analysis and Distributed Modeling in Hydrology. Edited by. Beven, K.J. and Moore, I. D. John Wiley and Sons, Chichester, U.K. pp.63-83. Rao, P. S. C. and Wagenet, R. J., 1985, “Spatial variability of pesticides in field soils: Methods for data analysis and consequences.” Water Science, 33:18-24. Saupe, Deitmar, 1988, The Science of Fractal Images. Edited by. Heinz-Otto, Peitgin and Saupe, Deitmar. Springer-Verlag, New York pp.82-84. Seyfried, M. S. and Wilcox, B. P., 1995, “Scale and nature of spatial variability: Field examples having implications for hydrologic modeling.” Water Resour. Res., 31(1):173184. Shannon, C. E. and Weaver, W., 1964, “The Mathematical Theory of Communication.” The Bell System Technical Journal, University of Illinois Press, Urbana . Smith, R. E. and Hebbert, R. H. B., 1979, “A Monte Carlo analysis of the hydrologic effects of spatial variability of infiltration.” Water Resour. Res., 15:419-429.

108 90

Chapter 4

Tabler, R. D., Pomeroy, J.W., and Santana, B. W., 1990, “Drifting Snow.” In: Cold Regions Hydrology and Hydraulics. Edited by. Ryan, W. C., Crissman, R. D., Tabler, R. D., Pomeroy, J.W., and Santana, B. W. pp.95-145. Tarboton, D. G., Bras, R. L., and Iturbe, I. R., 1991, “On the Extraction of Channel Networks from Digital Elevation Data.” In: Terrain Analysis and Distributed Modeling in Hydrology. Edited by. Beven, K. J. and Moore, I. D. John Wiley, New York pp.85-104. Vieux, B. E., 1993, “DEM Resampling and Smoothing Effects on Surface Runoff Modeling.” ASCE, J. of Comput. in Civil Eng., Special Issue on Geographic Information Analysis, 7(3):310-338. Wood, E. F., Sivapalan, M., and Beven, K., 1990, “Similarity and Scale in Catchment Storm Response.” Review of Geophysics, 28(1):1-18.

Chapter 5 INFILTRATION MODELING

Figure 5-1.Intense rainfall caused severe erosion on sandy soil in Kalkaskia County, Michigan. (Photo by B.E. Vieux, 1985).

110 92

Chapter 5

Rainfall-runoff modeling depends on the estimation of infiltration extended over large areas covering experimental plots, river basins, or regions. Figure 5-1 shows an eroded ski slope in Kalkaskia County, Michigan. The Kalkaskia soil is of glacial origin with permeability in excess of 24 cm/hr. Surface runoff and subsequent erosion was possible only because surface vegetation had been removed during construction, and raindrop impact caused surface sealing, and thus produced runoff. Because the rainfall rate exceeded the infiltration rate of the soils, runoff was generated and large amounts of sediment were transported into a nearby trout stream. This chapter deals with how soil properties and anthropogenic changes affect runoff. Infiltration and saturation excess are two important mechanisms responsible for runoff generation. In the following sections, we will examine how infiltration rates may be derived from soil properties. 1.1

Introduction

It is difficult to consider modeling of infiltration independently from rainfall and antecedent soil moisture. Consider that before ponding of water on the soil surface, that the infiltration rate is equivalent to the rainfall rate. After ponding occurs, the infiltration rate is controlled by soil properties, soil depth, degree and depth of saturation, and antecedent soil moisture. Methods for estimating infiltration rates can range from the full partial differential equation, known as Richards’ equation, to the more simplified Green and Ampt equation. The basic idea is to estimate the infiltration rate as controlled by the interplay between rainfall rate and the landsurface. This may involve estimating potential infiltration rates or parameters from soil properties or by regionalizing the point estimate obtained by some type of measurement device. In either case, a parameter known at a point must be extrapolated to large areas to be useful in runoff simulations. It would be impractical to take sufficient measurements at each point in the watershed. Thus, we often rely on regionalized variable theory together with some form of soil map to model infiltration at the river basin scale. We will not cover the full range of soil and infiltration models but will provide in this chapter a view of how to model infiltration at a point. We will then generalize this technique using a map of soil properties to model infiltration over a river basin or region. There are many issues surrounding the derivation of the governing equation for unsteady unsaturated flow in a porous media, as first presented by Richards (1931). Green and Ampt (1911) derived a method that predated the Richards’ equation known as the Green-Ampt equation. Essentially, the Green-Ampt equation ignores diffusion of soil moisture through a range of saturation, considering only an abrupt wetting front. Richards’ equation is

5. INFILTRATION MODELING

111 93

usually applied where soil properties are known through extensive testing. At the river basin scale, detailed soil testing is not feasible. It is possible to estimate infiltration over large areas using soils maps in GIS format and soil property databases. The accuracy of these infiltration estimates may vary depending on how well the soil maps represent the soil and the hydrologic conditions controlling the process. This chapter reviews efforts to estimate infiltration from soil properties. An application follows showing the derivation of Green-Ampt parameters from soil properties. 1.2

Infiltration Process

To compute rainfall-runoff it is necessary to have a sub-model of infiltration that interacts with the rainfall input. We will consider the case of Hortonian runoff, where rainfall rates exceeding infiltration produces runoff. An infiltration model in turn relies on characterizing the soil and the wetting front moving through the soil. Representing the full physics of this subprocess is often prevented by numerical or data restrictions. Data describing the soil-water interface may not be well known except at a few point locations where detailed testing has been performed. Practically, to model the infiltration rate, we must construct a model, an idealized conception, of how the soil transmits water through the soil profile. The soil model represents the interplay between the soil and the infiltrating water under both saturated and unsaturated conditions. The importance of parameters controlling the infiltration process depends on the modeling purpose. For example, a flash flood may occur after a recent rainfall has saturated the soil; the saturated hydraulic conductivity may control the infiltration process, leaving the soil suction with only a small role in such conditions. On the other hand, soil moisture modeling is more concerned with the amount of water that infiltrates than that which runs off, making soil suction under unsaturated conditions more important. Which one of these two parameters controls has an impact on our choice of calibration strategy? See Chapter 10 for further discussion of calibration as it relates to hydraulic conductivity. 1.3

Approaches to Infiltration Modeling

Accurate infiltration components are essential for physics-based hydrologic modeling. Many current hydrologic models use some form of the Green-Ampt equation to partition rainfall between runoff and infiltration components. While decades of use have confirmed the validity of this equation, as with all models, accurate parameter estimates are required to obtain reliable results. To apply models to ungauged areas, we must develop

112 94

Chapter 5

a procedure for estimating the key parameters of the model. Estimation is based on theoretical considerations, soil properties, or through calibration to measured data. Calibration is often required for most current hydrologic models to account for spatial variations not represented in the model formulation, model error/inadequacies, and functional dependencies between model parameters (Risse et al., 1995a). Calibration is usually necessary to generalize the laboratory measurement of parameters to field conditions. Order or magnitude differences in parameters can easily be found between laboratory and field measurements. For single-event models, measured parameters can be used; however, continuous simulation models often require both an accurate initial estimate as well as a method to adjust these parameters over the course of the simulation. These adjustments are intended to account for natural changes in soil structure, such as consolidation and crusting, as well as the effects of human-induced changes, such as those associated with tillage (Risse et al., 1995b). The Water Erosion Prediction Project (WEPP) was initiated in 1985 by the USDA to develop a new generation of prediction technology. Infiltration in WEPP is calculated using a solution of the Green-Ampt equation developed by Chu (1978) for unsteady rainfall. It is essentially a two-stage process under steady rainfall. Initially, infiltration rate is equal to the rainfall application rate until ponding occurs, after which infiltration rate is controlled by the soil properties. Parameterization of the soil layers can range from a simple, single-layer soil model to the more complex WEPP formulation that allows the user to input up to ten soil layers. These layers are used in the water balance component of the model; the infiltration routine uses a single-layer approach. The harmonic mean of the soil properties in the upper 200 mm is used in this model to represent the effects of multi-layer soils. Effective porosity, soil water content, and wetting front capillary potential are all calculated based on the mean of these soil properties for each layer. Sensitivity analysis on the hydrologic component of WEPP indicates that predicted runoff amounts are most sensitive to rainfall rates and hydraulic conductivity (Nearing et al., 1990). Other factors besides the soil properties affect hydraulic conductivity of the soil matrix responding to changes in the surrounding environment such as crusting. Crusting results from raindrop impact disaggregating the soil into constituent particles. The dislodged and disaggregated sand, silt and clay particles settle into the soil pores or are transported by the runoff downstream. The soil particles that settle into the surface form the crust. Therefore, in continuous simulation models, surface hydraulic conductivity should change for each storm event. Soil crusting is one of the most influential processes in reducing infiltration on a bare soil. When using

5. INFILTRATION MODELING

113 95

infiltration models, there may be a need to quantify the interactive effects of crusting and tillage on infiltration parameters. Without such recognition, the predicted variance will be larger, because the parameters do not represent the actual surface conditions. Risse et al. (1995) sought to improve WEPP model estimates of runoff using over 220 plot years of natural runoff data from 11 locations. By optimizing the effective Green-Ampt hydraulic conductivity for each event within the simulation, a method of correlating hydraulic conductivity on any given day to many other parameters was established. The most significant factors affecting hydraulic conductivity fell into three distinct categories: 1. Soil crusting and tillage 2. Storm event size and intensity 3. Antecedent moisture conditions. Equations were developed to represent the temporal variability of hydraulic conductivity for each group. Equations describing adjustments to account for event size and antecedent moisture conditions were derived based on the optimized data. Each adjustment was incorporated into the WEPP model, and runoff predictions were compared with the measured data. The tillage adjustment improved the overall average model efficiency. The adjustments for rainfall and antecedent moisture also improved the model efficiency, although they also increased the bias of under-predicting runoff for larger events. It has to be noted that these results may not be applicable to conditions outside the realm of this database or for models that implement the Green-Ampt equations in a different manner. The WEPP model description is meant to show the range of considerations encountered when implementing an infiltration model. A calibration algorithm was required to determine the optimum baseline hydraulic conductivity, Kb based on the curve number predicted runoff. For each soil, curve number predictions were made based on the 20-year climate file. The WEPP model was then run iteratively until the value of Kb was found that made the average annual runoff predicted by WEPP equal to that predicted by the curve number method. The reason for this procedure was to make the widely used curve number approach consistent with the more recently adopted Green-Ampt method. Having established that the Kb values calibrated to the curve number runoff for these 11 sites seemed reasonable, the values for the 32 additional soils were determined. Regression analysis was then applied to determine which soil properties could be used to estimate these values of Kb. The soil properties investigated were limited to sand, clay, silt, very fine sand, field capacity, wilting point, organic matter, CEC, and rock fragments, as these properties were thought to be easily obtainable or relatively easy to measure. In addition, several transformations and interactions were tested. The

114 96

Chapter 5

variables that exhibited the highest correlation were percent sand, percent silt, and CEC. While developing an equation to predict Kb, it was evident that a few soils with very high clay contents and low values of Kb were being overpredicted. Therefore, these soils were separated and two equations were developed to improve the prediction for these soils. While several equations provided nearly equivalent values of r2, the following equations were selected based on their simplicity and the low standard error of the estimates. Four soils with greater than 40% clay had a baseline hydraulic conductivity, Kb (mm/hr) equal to:

§ 244 · K b = 0.0066 exp¨¨ ¸¸ © % clay ¹

5.1

For 39 soils with clay less than or equal to 40%, the baseline hydraulic conductivity was found to be:

K b = −0.265 + 0.0086% sand 1.8 + 11.46CEC −0.75

5.2

where CEC has units of meq/100g. Overall, for the 43 soils, this combination of equations had an R2 of 0.78. From Eq. (5.2), it is apparent that the sandy soils have higher baseline conductivity. Since the value of the estimated Kb is dependent on both the clay content and the exchange capacity of the clay, conductivity is affected by the amount of clay as well as differences in clay mineralogy. Eqs. (5.1) and (5.2) were sufficiently robust to predict realistic values of Kb for most naturally occurring soils, so no limits were established. Estimated values of Kb from this equation compared favorably with measured values and values calibrated to measured natural runoff plot data. WEPP predictions of runoff using both optimized and estimated values of Kb were compared to curve number predictions of runoff and the measured values. The WEPP predictions using the optimized values of Kb were the best in terms of both average error and model efficiency. WEPP predictions using estimated values of Kb were superior to predictions obtained from the SCS runoff curve number method. The runoff predictions all tended to be biased high for small events and low for larger events when compared to the measured data. Confidence intervals for runoff predictions on both an annual and event basis were also developed for the WEPP model. Another method of computing the Green-Ampt parameters from data available in the standard United States Department of Agriculture (USDA) soil surveys (Rawls et al., 1983a and b) makes the application of the Green–

5. INFILTRATION MODELING

115 97

Ampt equation possible wherever soil surveys have been made. The GreenAmpt infiltration model has been used successfully to predict infiltration and runoff volumes with laboratory and small plot data. Rawls and Brakensiek (1986) applied a Green-Ampt model to predict runoff volumes from natural rainfalls on a single land use area of up to 10 acres. For the model to be fully useful there must be methods of easily handling the wide variety of soil and cover conditions found in working-sized watersheds. A method to estimate the Green-Ampt parameters for areas where soil surveys have not been completed is also desirable. These methods have been developed by Van Mullem (1989). Prior to this study, peak discharge estimates based on Green-Ampt infiltration had not been reported in the literature. Van Mullem (1991) used the Green-Ampt infiltration model to predict runoff from 12 rangeland and cropland watersheds in the US states of Montana and Wyoming. Soil parameters derived from data in standard USDA soil surveys were used, and 99 rainfall events were modeled. The runoff distributions obtained from the model were then used with a hydrograph model to predict the peak discharge from the watershed. The model was applied to areas of up to 140 km2 (54 mi2) with a wide variety of soil and cover conditions. The runoff volumes and peak discharges were compared with the measured values and with those predicted by the Soil Conservation Service curve number procedure. The Green-Ampt model predicted both the runoff volume and peak discharges better than did the curve-number model. The standard error of estimate was found to be less for the Green-Ampt in 9 of 12 watersheds for runoff volumes, and in 1 of 12 watersheds for the range of peak discharge studied by Van Mullem (1991). Corradini et al. (1994) have extended an earlier model developed by Smith et al. (1993) in order to describe a wide variety of real situations. Specifically, the model addresses a range of complex cases characterized by arbitrary series of infiltration-redistribution cycles. The model reliability was analyzed by comparing its results with those obtained by numerical solutions of Richards’ equation with reasonable results. Smith et al. (1993) presented a simplified physics-based model of discontinuous rainfall infiltration suitable for use as a modular part of complex watershed models. The basic model considered the problem of point infiltration during a storm consisting of two parts separated by a rainfall hiatus, with surface saturation and runoff occurring in each part. The model employed a two-part profile for simulating the actual soil profile. When the surface flux was not at capacity, it used a slightly modified version of the Parlange et al. (1985) model for description of increases in the surface water content and the Smith et al. (1993) redistribution equation for decreases. Criteria for the development of compound profiles and for their reduction to single profiles were also incorporated. The extended model was tested by comparison with numerical

116 98

Chapter 5

solutions of Richards’ equation, carried out for a variety of experiments upon two contrasting soils. The model applications yielded very accurate results and supported its use as a part of a watershed hydrologic model. Infiltration properties have been shown to vary both spatially and temporally because of soil property variation, cultural influences of man, activities of flora and fauna, and the interaction of the soil surface with kinetic energy of rainfall. Variability in a physical property can often result from investigation at a scale inappropriate for the problem at hand. Sisson and Wieranga (1981) were able to show that the variability of the infiltration properties tended to decrease with increased sampling area and volume. In infiltration studies the most commonly used methods, such as sprinkling or ring infiltrometers, make measurements over an area of a square meter or less. If the area sampled by these methods is large enough to encompass the representative elemental volume (REV) or the repetitive unit, all pore sizes, fabric structures and textures will be included. A substantial proportion of spatial variability occurs because these traditional samples are too small to encompass the representative elemental volume. Spatial heterogeneities and resulting infiltrating parameters depend on the scale of measurement. Williams and Bonell (1988) compared Philip equation parameters for large field plots with those for ponded infiltration rings. Rainfall and overland flow were recorded as functions of time at the experimental site, a Eucalypt woodland on massive oxic soils in central north Queensland. Oxic soils are typically well drained, seldom waterlogged or lacking in oxygen. From the beginning of rainfall to the time of runoff commencement, these infiltration processes were well described by the twoparameter Philip infiltration equation (Philip, 1957), which yields parameter estimates of the sorptivity, S, and asymptotic infiltration, A. The former was small and was found to have little influence, while the latter approached the field saturated hydraulic conductivity K*. For relatively large elements of the natural landscape, A and K* were estimated from plot runoff and compared with similar estimates made using Talsma infiltration rings. The Philip infiltration parameters estimated at the scale of these plots were smaller but of the same order as estimates using Talsma infiltration rings located in bare soil between the tussock vegetation. The ring estimates for soils associated with the tussock vegetation were very much larger than the estimates from the plots. The Talsma ring estimates of S, A and K* parameters for both the tussock and bare soil areas were much more variable in both space and time than those estimated from the surface hydrology of the large field plots. The spatial variability of these large plots was shown to be quite small. Temporal variability, however, was significant. The authors suggested that many of the apparently intractable problems of field spatial variability, characteristically associated with soil infiltration

5. INFILTRATION MODELING

117 99

properties, were due to a scale of measurement in space and time that is inappropriate to the model structure. Such problems could perhaps be resolved by careful analysis of surface hydrology of large plots or of small catchments that are sufficiently large relative to characteristic length of the repetitive unit for the soil surface. Two aspects must be accounted for when describing spatial variations of infiltration parameters. Deterministic variability is defined by soil types of known or estimated properties. Stochastic variability recognizes the variance of properties within mapped soil types. Gupta et al. (1994) describes spatial patterns of three selected infiltration parameters using a stochastic approach. The parameters selected for consideration were: 1. Saturated hydraulic conductivity (K) 2. Sorptivity (S) 3. Steady-state infiltration rate (F). Hydrological data series are neither purely deterministic nor purely stochastic in nature but are a combination of both. Therefore, we can analyze them by decomposing them into their separately deterministic and stochastic components. These components can then be described using a variety of mathematical and statistical models. The experimental data analyzed using Philip’s infiltration equation resulted in hydraulic conductivity, sorptivity and infiltration rates estimated for each soil. The results revealed that the K, S and I values are not constant but vary in magnitude from one point of measurement to the other. The range and variance also indicated that hydraulic conductivity exhibited the maximum degree of variation when compared to the sorptivity and infiltration rate of the soil. A model involving deterministic and stochastically dependent and independent sub-components can thus describe the spatial patterns of each infiltration parameter. The deterministic component of each model can be expressed by a two-harmonic Fourier Series with the dependent stochastic sub-component described by a second-order autoregression model and a residual sub-component. Application of the model for each soil requires recalibrating the models and obtaining a new set of Fourier coefficients and model parameters (Gupta et al., 1994). The standard assumption implicit in the Green-Ampt formulation is that the moisture front infiltrating into a semi-infinite, homogeneous soil at uniform initial volumetric water content occurs without diffusion. Salvucci and Entekhabi (1994) have derived explicit expressions for Green-Ampt (delta diffusivity) infiltration rate and cumulative storage. Through simple integration, the wetting front model of infiltration yields an exact solution relating the infiltration rate, cumulative infiltration, and time. An accurate expression for the infiltration rate determined by the Green-Ampt equation can be expressed in the form of a rapidly converging time series. Truncating

118 100

Chapter 5

the series at four terms yields a sufficiently accurate expression for infiltration rate with less than 2% error at all times. Further, the infiltration rate is readily integrated with time to yield the cumulative infiltration. The proposed expression is useful in infiltration/runoff calculations because it avoids the nonlinearity of the Green-Ampt equation. Efforts to characterize infiltration have focused mainly on cropland, to estimate erosion, or in arid areas, to manage precious water resources. Though investigated less often, infiltration in tropical climates is also subject to large spatial variability. A nested sampling scheme was used by Jetten et al. (1993) in three major physiographic units of Central Guyana, South America, termed ‘White Sands’, ‘Brown Sands’ and ‘Laterite’. Cluster analysis yielded three sample groups that reflected the sharp landscape boundaries between the units. Multiple regression analysis showed that each unit had a different combination of soil properties that satisfactorily explained the variance in final infiltration rate and sorptivity. Nested analysis of the spatial variance indicated that clear spatial patterns on the order of several hundred meters existed for White Sands and Laterite. Infiltration rate in Brown Sands and sorptivity in all units have large short-distance variability and high noise levels. The correlated independent variables behave accordingly. For the majority of the soil properties, sampling at distances of 100 to 200 m results in variance levels of more than 80% of the total variance, which indicates that only a detailed investigation over short length scales can assess spatial variation in soil hydrological behavior. The use of simple soil properties to predict infiltration is only possible in a very general sense and with the acceptance of high variance levels. Other factors affecting spatial variability of infiltration found by Jetten et al. (1993) were abundance of roots and activity of soil fauna. Also, abiotic soil characteristics differ considerably from place to place and strongly influence the soil structure. Crucial processes such as growth and development of seedlings may be adversely affected by spatial variations in soil moisture content. Soil moisture content and its temporal and spatial changes are a result of various processes in the soil-plant-atmosphere system, among which infiltration of water in the soil plays an important role. Representing the infiltration process over a river basin can be approached directly by infiltration measurements or indirectly by using a model that relies on parameter estimates based on soil properties. Both approaches contain errors caused by spatial variation of soil properties. Infiltration models show a variance that also originates from spatial variance of the soil properties. However, in a model approach, this original spatial variance is altered through the combination and processing of the parameters in the algorithms used in the model. The spatial variation not captured in a soil

5. INFILTRATION MODELING

119 101

map or detected through measurement is one of the major sources of variance encountered in infiltration simulation at the watershed scale. As discussed in Chapter 3, Kriging has been applied to estimating soil properties across some spatial extent. The regionalized variable theory (Burrough, 1986; Webster and Oliver, 1990; Davis, 1973) assumes that a spatial variation of a variable z(x) can be expressed as the sum of three components:

z ( x ) = m( x ) + ε ' ( x ) + ε "

5.3

where x is the position given in two or three coordinates; m( x ) = structural component, i.e., a trend or a mean of an area; ε ' ( x ) = stochastic component, i.e., spatially correlated random variation; ε " = residual error component (noise), i.e., spatially noncorrelated random variation. Spatial analysis of a variable involves the detection and subsequent removal of trends m(x) in the data set, after which the spatially correlated random variation can be described. This is a departure from the approach of using a mapping unit and associated soil properties to estimate Green-Ampt infiltration. To apply the regionalized variable theory, we must have point estimates of the variate at sufficient density and spatial extent. This method suffers from the same obstacle as does infiltration measurement: sufficient measurements are difficult to obtain over large regions. This makes direct measurement and regionalized variable approaches difficult from lack of data over large areas such as river basins. The following describes the Green-Ampt theory and the variables controlling the process. Once the variables are defined, then regression equations can be formulated to predict values based on soil properties mapped over the river basin. 1.4

Green-Ampt theory

The rate form of the Green-Ampt equation for the one-stage case of initially ponded conditions and assuming a shallow ponded water depth is:

ª Ψ f ∆θ º f (t ) = K e «1 + » F (t ) ¼ ¬

5.4

where f (t ) = dF (t ) / dt = infiltration rate (L/T); K e = effective saturated conductivity (L/T), usually taken as ½ of (see Eq. (5.12) below) predicted by soil properties or measured in the laboratory accounting for the reduction due to entrapped air; Ψ f = average capillary potential (wetting front soil

120 102

Chapter 5

suction head) (L); ∆θ = moisture deficit (L/L); and F (t ) = cumulative infiltration depth (L). The soil-moisture deficit can be computed as:

∆θ = φtotal − θ i

5.5

where φtotal is the total porosity(L/L); and θ i is initial volumetric water content (L/L). Eq. (5.4) is a differential equation, which may be reformulated for ease of solution as:

ª F (t ) º K e t = F (t ) − Ψ f ∆θ ln «1 + » «¬ Ψ f ∆θ »¼

5.6

where t is time and the other terms are as defined above. Eq. (5.6) can be solved for cumulative infiltration depth, F for successive increments of time using a Newton-Raphson iteration together with Eq. (5.4) to obtain the instantaneous infiltration rate. For the case of constant rainfall, infiltration is a two-stage process. The first stage occurs when the rainfall is less than the potential infiltration rate and the second when the rainfall rate is greater than the potential infiltration rate. Mein and Larson (1973) modified the Green-Ampt equation for this two-stage case by computing a time to ponding, tp, as:

tp =

K e Ψ f ∆θ

5.7

i (i − K e )

where i = rainfall rate. The actual infiltration rate after ponding is obtained by constructing a curve of potential infiltration beginning at time t0 such that the cumulative infiltration and the infiltration rate at tp are equal to those observed under rainfall beginning at time 0. The time t is computed as:

t = t p − t0

5.8

The time to ponding, tp depends on soil properties and initial degree of saturation expressed as a soil moisture deficit, ∆θ and is computed as:

F − Ψ f ∆θ ln(1 + tp =

Ke

F ) Ψ f ∆θ

5.9

5. INFILTRATION MODELING

121 103

where F = cumulative infiltration depth (L) at the time to ponding, which, for the case of constant rainfall, is equal to the cumulative rainfall depth at the time to ponding. Before ponding, all rainfall is considered to be equal to the infiltrate. The corrected time as computed by Eq. (5.9) is used in Eq. (5.6), which is solved as with the one-stage case. Computing cumulative and instantaneous infiltration requires soil parameters, which in turn must be estimated from mapped soil properties. 1.5

Estimation of Green-Ampt Parameters

In this section, we briefly review 1) the Green-Ampt equation and the parameters used to estimate infiltration; 2) estimation of the Green-Ampt parameters from soil properties; and 3) an application using two soil maps, one with a high degree of detail (MIADS), and the other more generalized (STATSGO). MIADS is a soils database compiled by the USDA-Natural Resources Conservation Service (NRCS) for the State of Oklahoma from county-level soil surveys. Though in GIS format, the soil maps were compiled using non-rectified aerial photography. These soil maps together with a database of soil properties, known within the NRCS as the soils-5 database, comprise the most detailed soils information available over any large spatial extent in Oklahoma. These maps were rasterized at a grid-cell resolution of 200 m from soil maps that are generally at a scale of 1:20000. STATSGO is a generalized soils database and soil maps in GIS format compiled for each state in the US. It is generalized in the sense that the STATSGO map units are aggregated from soils of similar origin and characteristics. Thus, much of the detail present in county-level soil maps is lost. The composition of each STATSGO mapping unit was determined by transecting or sampling areas on the more detailed soil maps and expanding the data statistically to characterize the whole map unit. Each map unit has corresponding attribute data giving the proportionate extent of the component soils and the properties for each soil. The database contains both estimated and measured data on the physical and chemical soil properties and soil interpretations for engineering, water management, recreation, agronomic, woodland, range and wildlife uses of the soil. These soil map units were compiled at a scale of 1:250000 corresponding to USGS topographic quadrangles. From these mapped soil properties Green-Ampt parameters may be estimated for regions of large spatial extent such as river basins. In estimating the soil model parameters from soil properties, the infiltration parameters are effective porosity, θe; wetting front suction head, ψf; and saturated hydraulic conductivity, Ksat. These parameters are based on the following equations. The effective porosity, θe is:

122 104

Chapter 5

θ e = φ total − θ r

5.10

where φtotal is total porosity, or 1 – (bulk density)/2.65; and θr is residual soil moisture content (cm3/cm3). The effective porosity, θe is considered a soil property and is independent of soil moisture at any particular time, as is the wetting front suction head. Using the Brooks and Corey soil-water retention parameters to simulate the soil water content as a function of suction head, the wetting front suction head, Ψ f (cm), is given as:

Ψf =

2 + 3λ Ψb 1 + 3λ 2

5.11

where ψb is bubbling pressure (cm) and λ is the pore size distribution index. Saturated hydraulic conductivity, Ks in cm/hr, is estimated from the effective porosity, bubbling pressure and pore size distribution index as:

(θ e / Ψb ) 2 (λ ) 2 K s = 21.0 * (λ + 1)(λ + 2)

5.12

Eqs. (5.10) to (5.12) allow the expression of the Green-Ampt parameters, wetting front suction, and hydraulic conductivity in terms of the soil-water retention parameters, pore size index, bubbling pressure, and effective porosity (Brooks and Corey, 1964). Soil-water retention parameters have been estimated from soil properties using regression techniques by Rawls et al. (1983a and b) as follows: Bubbling Pressure ψb = exp[5.3396738 + 0.1845038(C) – 2.48394546 (φtotal) – 0.00213853(C)2–0.04356349(S)(φtotal)– 0.61745089(C)(φtotal)+0.00143598(S)2(φtotal)2– 0.00855375(C)2(φtotal)2–0.00001282(S)2(C)+ 0.00895359(C)2(φtotal)–0.00072472(S)2(φtotal)+ 0.0000054(C)2(S) + 0.50028060(φtotal)2(C)]

5.13

123 105

5. INFILTRATION MODELING Pore Size Distribution Index λ = exp[–0.7842831+0.0177544(S)–1.062498(φtotal) – 0.00005304(S)2– 0.00273493(C)2+1.11134946(φtotal)2– 0.03088295(S)(φtotal)+0.00026587(S)2(φtotal)2– 0.00610522(C)2(φtotal)2–0.00000235(S)2(C)+ 0.00798746(C)2(φtotal) – 0.00674491(φtotal)2(C)]

5.14

Residual Soil Moisture Content θr = –0.0182482+0.00087269(S)+0.00513488(C)+ 0.02939286(φtotal)–0.00015395(C)2– 0.0010827(S)(φtotal)– 0.00018233(C)2(φtotal)2+0.00030703(C)2(φtotal)– 0.0023584(φtotal)2(C)

5.15

where C is percent clay in the range 5%