Pas, Volume Z ee eee ASSe oe Water Resources ILLINOIS RIVER BLUFFS AREA ASSESSMENT “ILUNOIS STATE WATER SURVEY LIBRARY COPY IDNR ILLINOIS RIVER BLUFFS [ CTAP AREA ASSESSMENT. v.2 99062906 ISSUED To IDNR ILLINOIS RIVER BLUFFS CTAP AREA ASSESSMENT. v.2 99062906 DEMCO ILUINUIS STATE WATER SURVEY LIBRARY COPY JY 0799 ILLINOIS RIVER BLUFFS AREA ASSESSMENT VOLUME 2: WATER RESOURCES Illinois Department of Natural Resources Office of Scientific Research and Analysis Illinois State Water Survey 2204 Griffith Drive Champaign, Illinois 61820 (217) 244-5459 1998 Jim Edgar, Governor State of Illinois Brent Manning, Director Illinois Department of Natural Resources 524 South Second Springfield, Illinois 62701 300 Printed by the authority of the State of Illinois Other CTAP Publications The Changing Illinois Environment: Critical Trends, summary and 7-volume technical report Illinois Land Cover, An Atlas, plus CD-ROM Inventory of Ecologically Resource-Rich Areas in Illinois Rock River Area Assessment, 5-volume technical report The Rock River Country: An Inventory of the Region's Resources Cache River Area Assessment, 5-volume technical report The Cache River Basin: An Inventory of the Region's Resources Mackinaw River Area Assessment, 5-volume technical report The Mackinaw River Country: An Inventory of the Region's Resources The Illinois Headwaters: An Inventory of the Region’s Resources Headwaters Area Assessment, 5-volume technical report The Illinois Big Rivers: An Inventory of the Region's Resources Big Rivers Area Assessment, 5-volume technical report The Fox River Basin: An Inventory of the Region’s Resources Fox River Area Assessment, 5-volume technical report The Kankakee River Valley: An Inventory of the Region’s Resources Kankakee River Area Assessment, 5-volume technical report The Kishwaukee River Basin: An Inventory of the Region's Resources Kishwaukee River Area Assessment, 5-volume technical report Embarras River Area Assessment, 5-volume technical report Upper Des Plaines River Area Assessment, 5-volume technical report Annual Report 1997, Illinois EcoWatch Stream Monitoring Manual, Illinois RiverWatch Forest Monitoring Manual, Illinois ForestWatch Illinois Geographic Information System, CD-ROM of digital geospatial data All CTAP and Ecosystems Program documents are available from the DNR Clearinghouse at (217) 782-7498 or TDD (217) 782-9175. Selected publications are also available on the World Wide Web at http://dnr.state.il.us/ctap/ctaphome.htm, or . http://dnr.state.il.us/c2000/manage/partner.htm, as well as on the EcoForum Bulletin Board at 1 (800) 528-5486 or (217) 782-8447. For more information about CTAP, call (217) 524-0500 or e-mail at ctap2@dnrmail state.il.us; for information on the Ecosystems Program call (217) 782-7940 or e-mail at ecoprog@dnrmail state.il.us. About This Report The Illinois River Bluffs Area Assessment examines an area in west-central Illinois that includes parts of the upper and lower Illinois River watersheds from the vicinity of Hennepin southward to East Peoria. Because significant natural community and species diversity is found in the area, it has been designated a state Resource Rich Area.’ This report is part of a series of reports on areas of Illinois where a public-private partnership has been formed. These assessments provide information on the natural and human resources of the areas as a basis for managing and improving their ecosystems. The determination of resource rich areas and development of ecosystem-based information and management programs in Illinois are the result of three processes -- the Critical Trends Assessment Program, the Conservation Congress, and the Water Resources and Land Use Priorities Task Force. Background The Critical Trends Assessment Program (CTAP) documents changes in ecological conditions. In 1994, using existing information, the program provided a baseline of ecological conditions.* Three conclusions were drawn from the baseline investigation: 1. the emission and discharge of regulated pollutants over the past 20 years has declined, in some cases dramatically, 2. existing data suggest that the condition of natural ecosystems in Illinois is rapidly declining as a result of fragmentation and continued stress, and 3. data designed to monitor compliance with environmental regulations or the status of individual species are not sufficient to assess ecosystem health statewide. Based on these findings, CTAP has begun to develop methods to systematically monitor ecological conditions and provide information for ecosystem-based management. Five components make up this effort: 1. identify resource rich areas, 2. conduct regional assessments, 3. publish an atlas and inventory of Illinois landcover, 4. train volunteers to collect ecological indicator data, and 5. develop an educational science curriculum which incorporates data collection ' See Inventory of Resource Rich Areas in Illinois: An Evaluation of Ecological Resources ? See The Changing Illinois Environment: Critical Trends, summary report and volumes 1-7 iil At the same time that CTAP was publishing its baseline findings, the Illinois Conservation Congress and the Water Resources and Land Use Priorities Task Force were presenting their respective findings. These groups agreed with the CTAP conclusion that the state's ecosystems were declining. Better stewardship was needed, and they determined that a voluntary, incentive-based, grassroots approach would be the most appropriate, one that recognized the inter-relatedness of economic development and natural resource protection and enhancement. From the three initiatives was born Conservation 2000, a six-year program to begin reversing ecosystem degradation, primarily through the Ecosystems Program, a cooperative process of public-private partnerships that are intended to merge natural resource stewardship with economic and recreational development. To achieve this goal, the program will provide financial incentives and technical assistance to private landowners. The Rock River and Cache River were designated as the first Ecosystem Partnership areas. At the same time, CTAP identified 30 Resource Rich Areas (RRAs) throughout the state. In RRAs where Ecosystem Partnerships have been formed, CTAP is providing an assessment of the area, drawing from ecological and socio-economic databases to give an overview of the region's resources --.geologic, edaphic, hydrologic, biotic, and socio-economic. Although several of the analyses are somewhat restricted by spatial and/or temporal limitations of the data, they help to identify information gaps and additional opportunities and constraints to establishing long-term monitoring programs in the partnership areas. The Illinois River Bluffs Assessment The Illinois River Bluffs Assessment covers an area of about 560,871 acres in west central Illinois. It includes parts of the upper and lower Illinois River watersheds from the vicinity of Hennepin southward to East Peoria. Counties encompassed in this assessment include most of Marshall and Woodford counties as well as small portions of Stark, Bureau, La Salle, Tazewell, Putnam, and Peoria counties. In addition to containing a portion of the Illinois River Drainage basin (Illinois River upper and lower), this area also encompasses portions of the Crow Creek west, Sandy Creek, Senachwine Creek and Crow Creek east drainage basins as identified by the Illinois Environmental Protection Agency. Three of the sub-basins in this assessment area (Illinois River lower, Senachwine Creek, and Crow Creek east) were designated as “Resource Rich Areas” (a total of 277,847 acres) because they contain significant natural community diversity. The Illinois River Bluffs Ecosystem Partnership was subsequently formed around this core area of high quality ecological resources. This assessment is comprised of five volumes. In Volume 1, Geology discusses the geology, soils, and minerals in the assessment area. Volume 2, Water Resources, discusses the surface and groundwater resources and Volume 3, Living Resources, describes the natural vegetation communities and the fauna of the region. Volume 4 rs me PECATONICA/ \ i \ie oe \ SUGAR ~~ UPPER } & P\ RP in ae fir! | a ae a f UPPER | big x ef if Fg | : IROQUOIS | LA ( SALT FC FORK - SANGAMON as VERMILION » Fs WABASH) pee FORK - | Ta | SANGAMON! Scale 1:2700000 SSiin OF a Oe a | = 100 Mites SOUTH =F ae part ans NA —— ee “en, «=6BAY/ | >) o 160 Kiomaters LUSK 27 —E———— CACHE & ry ae, ~~ OHIO TRIBS\ ea ek Drainage basins from 1:24000 scale watershed boundaries as ae j aa delineated by the U.S.G.S. Water Resources Division. ry x4 So Major Drainage Basins of Illinois and Location of the Illinois River Bluffs Assessment Area Bradford STARK CO. __ | PEORIA CO. oS Pa v4 LZ Ilinols hae 7” Senachwine Cr. é Bureau — ss Ilinols R. lower Washington TAZEWELL CO. row ty. east —_——_—_— WOODFORD CO. crows Senechwine Lake Hennepin | Sid olf a, W ala < el< Tonica Washbum Br. Cro > JrranS SRS - fr PARASTC rf fh YL naw R AQ Mackin Ss =i Tol w Crk. \ S. Br. Crow Gr. ES \- sh | | | | | Rutland | | | | | \ | 5 a | ¢ | | r—s |S E] Paso 'UO |2 w ie S = aN | roe Scale 1:370000 25 Kilometers EE — EE — 0 15 Miles EE — Subbasins in the Illinois River Bluffs Assessment Area. Subbasin boundaries depicted are those determined by the Illinois Environmental Protection Agency. contains three parts: Part I, Socio-Economic Profile, discusses the demographics, infrastructure, and economy of the area, focusing on the three counties with the greatest amount of land in the area — Marshall, Peoria and Woodford; Part I], Environmental Quality, discusses air and water quality, and hazardous and toxic waste generation and management in the area; and Part III, Archaeological Resources, identifies and assesses the archaeological sites, ranging from the Paleoindian Prehistoric (B.C. 10,000) to the Historic (A.D. 1650), known in the assessment watershed. Volume 5, Early Accounts of the Ecology of the Illinois River Bluffs Area, describes the ecology of the area as recorded by historical writings of explorers, pioneers, early visitors and early historians. Vii Digitized by the Internet Archive in 2010 with funding from University of Illinois Urbana-Champaign http://www.archive.org/details/illinoisriverbluO2illi Contributors PARES COORGIN ALON . Qqeca yoniwV JT... ee mares cd A EET AR OSS 5 i“ 7 y fs A = - 5 a: 5 ridoll ekmoT anata ndot shill mee oss tee zit Wi 140 c : . wo Table of Contents PIITOGUICHION BI42. 30.5 Mev .c. bts Beste eee AEs. eats te eca RSI OGE Pas sake leanchigusesesedsevecusesoaneueaceecceetes | RIVETS, ATIC S Cr emtEMISHEE Se A Seek iced RA AMER EE 595 8S Ss basen cctuessuascseseaccusecdsene l TEAK CS): cS Rvevs the dead Bcesene Mhc ds beans enenacev Maca tetete sete nsadlsnnaseaa gi Ehlaasihcveséseseceusssesgsunscsseséce 4 DEL ATS oes ce cots aan ae ous RN po nese PAR LS acne ds suunaida div vsseuanasuentiinpaiyvaaessbedsiacsvs 5 PET VSUO STADIA pace acs as vbds dze cscs nstntenis sso tauasulteasa waka casts sudapaaed edactonsundassvaawas dass ccesedactciesdesss 10 MAAN a cit cat sui catenins Ba vcuo ot ope sae cu dat ecuvssa causes hatasatiiuetiasessbdesivesucsrassonatsbecicat: 11 Whimate ang Arenas rimgS lita ale esse ste secs cacscsusnnaesensusosesssseseenassesnandsbncs ssbsGTeese cd coccacees 13 PheMperatune css. ev sscctiweeas eta weeewes Sebecblcbsce INE tive dap Sanseeda dea Svauecate concealers Meacesccces: 13 EPG) po Mea Hite ieee eee tees hee ccc tcb tect SececBenowsddbseesnseatibendgisnstcadtienensennendeeReueatets.-s 17 BRecipitaniOn: MeniCits ANG EXCESSES sa..5c220:5cccsnaseseorostsbesvseusiiscstescessesepecsscsicanics 19 BROT VV CAL Che crce tccecccte reece ee saate css ciiics eave dsspecsusiscbsesgaSssasadnécaseedetnsiiessseeteeteatecczenvers 20 IMO TERACLOGS wenes cece eetccen es eee see cn cte sat sca aual yesesadcoedudces desstcaccessouse¥aisa tdesossonaval Obvedes 20 |S EY eo Ede ag fee nr RAS ene rere ac ee 21 siunderstorms: 5. «Ad Sess Siete: aercee edited ewes spsten, Avis ded Be es 21 STDIN) “eee asd cee ses Lantos oe ratte Seats cccsu save cease SRS eezecets detest) gepcdeacesnisccaséccovedees 22 IS Eire eATgE Es OW Pepe eens, Benes eee ee cae ame Rte 2d aR SI RUN ET gs 23 Shredmi Gagne RECOrdS s. 2is3 secs ete e-checeace a cacatt acon te otaaa sao ecngh cases eSasgeteosbbbasesacsieses 28 Human Impacts'on Streamflows)in the Basin (22.00.42. ...0.\.200252s265ssossesecesSecnsecousceee 25 Anal: Streamitlow dy ariabilit yy ee eee seek eee, 2 TS cs sccaseudsz 26 StatisticaleDrendy A tial ysis eset ccesezss, cots Mae ere eitonn seas tanesonsSetes tadstiaddinstenece tess 26 DailyandSeasonaliow Variability ascccccoscescsncsors 30 7.3 Source: Runge et al. (1969) Land Use Agriculture is a major land use in the eight counties (Bureau, LaSalle, Marshall, Peoria, Putnam, Stark, Tazewell, and Woodford) in the Illinois River Bluffs area. The Illinois Department of Agriculture, Illinois Agricultural Statistics (LAS) data indicate that in 1995 agriculture acreage accounted for approximately 31% of the total surface area in the Illinois River Bluffs assessment area and has increased only 4% from 341,650 acres in 1925 to 370,008 acres in 1995. Figure 5 shows the changes in the harvested acres of selected crops in the basin from 1925 to 1995. 450,000 400,000 350,000 300,000 250,000 200,000 ACRES 150,000 100,000 0 1320 1930 1940) 7 19505, 1960: W970) 1980 19,90" .2000 Figure 5. Acreages of Selected Crops in the Illinois River Bluffs Area Based on IAS Data In 1925 the dominant crops were grassy crops (wheat, oats, and hay) and corn, accounting for 99% of the agricultural crops grown in the basin (170,883 acres for corn and 167,275 for grassy crops). Corn acreage has remained fairly steady over time, increasing only slightly to levels above 250,000 acres in 1976, with a significant drop (approximately 30%) in 1983 to 167,192 acres. In 1925 soybeans were confined to little over 1000 acres; however, it steadily increased to 167,935 acres in 1995, the most acres harvested to date. The average grassy crop acreage from 1925 to 1950 was 140,000 and from this time steadily decreased to approximately 13,000 acres in 1995. The inverse relationship between soybean and grassy crop acreage is shown in figure 5, where the trends in acreage cross during 1964-66. In 1995 the dominant crops were corn and soybeans as opposed to corn and grassy crops in 1925. Ninety-six percent of crop acres harvested in the Dlinois River Bluffs area is corn and soybeans (356,803 acres). 12 Climate and Trends in Climate This chapter reviews climate trends in and around the Illinois River Bluffs area since the turn of the century. Climate parameters examined include annual mean temperature, the number of days with highs above or equal to 90°F, the number of days with lows below or equal to 32°F, the number of days with lows below or equal to 0°F, annual precipitation, the number of days with measurable precipitation, annual snowfall, and the number of days with measurable snowfall. Extreme weather events examined in this report are tornadoes, hail, and thunderstorms. The Illinois River Bluffs area in north-central Illinois occupies portions of Bureau, Putnam, La Salle, Stark, Marshall, Peoria, Woodford, and Tazewell Counties. The climate of this area is typically continental, as shown by its changeable weather and the wide range of temperature extremes. Summer maximum temperatures are generally in the 80s or 90s, with lows in the 60s or 70s, while daily high temperatures in winter are generally in the 20s or 30s, with lows in the teens or 20s. Based on the latest 30-year average (1961-1990), the average first occurrence of 32°F in the fall is October 17, and the average last occurrence in the spring is April 22. Precipitation is normally heaviest during the growing season and lightest in midwinter. Thunderstorms and associated heavy showers are the major source of growing season precipitation, and they can produce gusty winds, hail, and tornadoes. The months with the most snowfall are November, December, January, February, March, and April. However, snowfalls have occurred as early as October and as late as May. Heavy snowfalls rarely exceed 12 inches. The climate data used in the following discussions originate at Peoria, Illinois (Peoria County), which houses the National Weather Service (NWS) Coop site with the longest record (1901-1996) near the southern portion of the basin. Supportive data and analyses for nearby Illinois sites can be found in reports by the Illinois Department of Energy and Natural Resources (1994) and Changnon (1984). Temperature The mean January maximum temperature is 30°F and the minimum is 13°F, whereas the mean July maximum and minimum temperatures are 86°F and 65°F, respectively (Table 5). The mean annual temperature at Peoria is 50.7°F. The warmest year of record was 1901, with an average of 57.2°F, while the coldest was 1917, with an average of 47.8°F. Table 5. Temperature Summary for Peoria (Averages are from 1961-1990 and extremes are from 1901-1996. Temperatures are in °F) #ofdays #ofdays #of days Avg. Avg. Record Record with high =withlow — with low i high (year) —_ low (year) >90°F $32°F January 71 (1909) = -25 (1977) 0 28 February 34.9 17.7 74 (1932) -26 (1905) 0 25 ‘ March 48.1 29.8 87 (1907) = -11 (1943) 0 19 0.2 April 62.0 40.8 92 (1930) 14(1920) 0.1 5.8 0 May 72.8 50.9 104 (1934) 25 (1966) 1.0 0.4 0 June 82.2 60.7 105 (1934) 39 (1945) 5.6 0 0 July 85.7 65.4 113 (1936) 46(1911) 9.8 0 0 August 83.1 63.1 106 (1936) 41 (1910) V2 0 0 September 76.9 33:2 102 (1939) 24 (1942) 2.8 0.1 0 October 64.8 43.1 92 (1922) 7 (1925) 0.1 3.8 0 November 49.8 32:5 81 (1937) -2 (1977) 0 16 0.1 December 34.6 19.3 71 (1970) -24 (1924) 0 26 DD Although there is a great deal of year-to-year variability, mean annual temperatures at Peoria show a warming trend from 1901 to 1930, followed by a cooling trend until 1960, warming again through 1996 (Figure 6) 56 Mean Annual Temperature (F) - oa 164] oO oO oa oi o (=) - ine) Ww - oa nh foe) 47 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 6. Mean Annual Temperature for Peoria, 1901-1996 14 Examination of mean temperatures over time is one way to clarify trends. The NWS has adopted 30-year averages, ending at the beginning of the latest new decade, to represent climate "normals." These averages filter out some of the smaller scale features and yet retain the character of the longer term trends. Consecutive, overlapping "normals" for the last seven 30-year periods at Charleston are presented in Table 6. The consecutive means demonstrate the warming trend through the 1931-1960 period, followed by a cooling trend through the 1961-1990 period. Table 6. Average Annual Temperature during Consecutive 30-Year Periods Averaging Average period temperature (°F) 1901-1930 51.4 1911-1940 51.8 1921-1950 51.9 1931-1960 51.9 1941-1970 51.0 1951-1980 50.5 1961-1990 50.5 The frequency of extreme events sometimes conveys a clearer picture of trends than mean values. The annual number of days with temperatures equal to or above 90°F is shown in Figure 7. Not too surprisingly, this bears little resemblance to annual temperature (Figure 6), because the number of days with temperatures above 90°F represents only the high summer temperature extremes. Figure 7 shows an increase through 1938, followed by a slow decline through 1970, before returning to somewhat higher numbers from 1971 to 1996. = | sol ot AY fT SMALL TL oh WN UW LANL eI Ey! SER Se 0 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 7. Annual Number of Days with Maximum Temperatures Equal to or Above 90°F at Peoria, 1901-1996 15 Figure 8 shows the winter frequency of daily minimum temperatures equal to or below 32°F. The frequency of such temperatures shows no trends. Figure 9 shows the number of days per year when the minimum temperature was equal to or below O°F, beginning with the 1903-1904 winter. No long-term trends are evident. However, there is a large degree of variability from year to year. 160 150 = s= oO _ _ ib?) Ww Oo Oo L_——}—j— # of Days with Low <= 32F 80 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 8. Annual Number of Days with Minimum Temperatures Equal to or Below 32°F at Peoria, Winters 1903-1904 to 1995-1996 35 nm ao PVT Sie ve AT BM as VT ATH AAA AP PUMEAVINTVPWLA EV! VOT ee ae 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 # of Days with Low <= OF Figure 9. Annual Number of Days with Minimum Temperatures Equal to or Below O°F at Peoria, Winters 1903-1904 to 1995-1996 16 Precipitation Mean annual precipitation at Peoria is 36.25 inches, with more rainfall in the spring and summer than in fall and winter (Table 7). Late spring, summer, and early fall precipitation is primarily convective in nature, often associated with short thunderstorms (1-2 hours in duration). During the remainder of the year, precipitation is of longer duration and associated with synoptic-scale weather systems (cold fronts, occluded fronts, and low pressure systems). The wettest year of record was 1990 (55.35 inches). The driest year was 1988 (22.17 inches). Table 7. Precipitation Summary for Peoria (Averages are from 1961-1990 and extremes are from 1901-1996. Precipitation is in inches.) Largest one- # of Avg. Record Record day amount Snow- days w/ Month precip. high(year) low (year) (year) fall precip. January 15 8.11 (1965) 0.07(1919) 4.43(1965) 7.3 9 February 1.42 4.95 (1942) 0.14(1907) 2.83(1942) 5.9 8 March 2.91 6.95 (1973) 0.40(1958) 2.88(1944) 3.4 11 April CaM 8.66 (1947) 0.71(1971) 5.06 (1950) 12 12 May 3.70 11.49(1915) 0.47 (1934) 5.52(1927) 0 12 June 3.99 11.69(1974) 0.45(1936) 4.74(1911) 0 10 July 4.20 10.15 (1993) 0.33 (1988) 3.56(1953) 0 9 August 3.10 8.61 (1955) 0.25(1992) 4.32(1955) 0 9 September 3.87 13.09(1961) 0.03(1979) 4.11(1961) 0 9 October 2.65 10.53(1941) 0.03 (1964) 3.62(1969) 0.1 8 November 2.69 7.62 (1985) 0.07(1917) 4.26 (1990) 1.9 9 December 2.44 6.34 (1949) 0.29(1930) 2.52(1965) 64 9 Annual precipitation at Peoria is shown in Figure 10. No long-term trends are evident; however, the last 10 years of data have the highest degree of variability. The number of days per year with measurable precipitation (i.e., more than a trace) is shown in Figure 11. No trend is evident from 1901 to 1960. From 1961 to 1996, the variability in the number of days has increased dramatically. The much lower values in the first few years of the record may be due to a change in exposure, location, or observer. The annual precipitation (Figure 10) shows no such pattern, suggesting that the changes shown in Figure 11 mainly impact the very light precipitation events. Precipitation is more frequent during summer months than during winter months. 17 # of Days with Precipitation Precipitation (in) 45 .= oO Ee) ramen v0 I AIL EME NN i ae ake Ww oO Ww oO ‘i (ia ibe) oa 20 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 10. Annual Precipitation at Peoria, 1896-1995 ede: deo pe A "ti tr aa 80 70 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 1]. Annual Number of Days with Measurable Precipitation at Peoria, 1901-1996 Average winter snowfall in Peoria is 21.6 inches, with great year-to-year variability. The most snowfall during any one winter was 52.3 inches in 1977-1978, and the least was only 5.8 inches in 1916-1917. Figure 12 shows snowfall from winter 1903-1904 through winter 1995-1996. A similar upward trend was evident through the mid 1980s, followed by a slight decline through the winter of 1995-1996. 60 50 b oO Snowfall (in) 3 Lye] Oo fe caret) TNO VA UAL whe yen 10 e) 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 12. Annual Snowfall at Peoria, Winters 1903-1904 to 1995-1996 Figure 13 shows the number of days each winter with snowfall, from 1948-1949 through 1995-1996. The number of days with snow shows a somewhat different pattern than that for total snowfall with increases through 1966-1967, followed by decreases through 1995- 1996. A snowfall of more than 6 inches occurs about once a year. Snow cover is frequently experienced at Peoria, lasting from a few days at a time to three months. Precipitation Deficits and Excesses Following are the driest years in the Illinois River Bluffs area in terms of annual precipitation shortfall, starting with the driest: 1988, 1989, 1910, 1930, 1914, 1962, 1994, 1956, 1963, and 1901. Driest summer seasons (June, July, and August) in the basin include: 1988, 1936, 1910, 1922, 1930, 1991, 1912, 1920, 1914, and 1933. Significantly above average precipitation fell at Peoria in 1990, 1993, 1927, 1973, 1926, 1902, 1965, 1982, 1970, and 1985. No single decade dominated in terms of years with excessive precipitation. —— lah 3 25 i | Ht 2. AEE 5 0 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 Figure 13. Annual Number of Days with Measurable Snowfall at Peoria, Winters 1903-1904 to 1995-1996 Severe Weather Tornadoes Although tornadoes are not uncommon in []linois, most people do not expect to be affected directly by one, even if they live in the state for a lifetime. This is because tornadoes are generally only one-quarter mile in diameter, travel at roughly 30 miles per hour for only 15-20 minutes, and then dissipate, affecting a total area less than 2 square miles. Since Illinois observes an average of 28 tornadoes a year (though the actual number varies from fewer than ten to almost 100 during the last 35 years), the total area directly affected by tornadoes annually is only about 55 square miles, 0.1% the total area of the state. Even with 96 tornadoes reported in Illinois in 1974 (the greatest number reported in the last 30 years), the affected area was only about 0.3% the total area of the state. These numbers do not diminish the effect on those experiencing property damage, injury, or worse, but they demonstrate the extremely low probability of direct impact at any single location. The most recent study on tornadoes in [llinois examined events from 1955 to 1986 and found no apparent trend in tornado frequency or intensity (Wendland and Guinan 1988). On average, the Illinois River Bluffs area experiences about one tornado every three years. 20 Hail Hail events are somewhat rare and typically affect a very small area (from a single farm field up to a few square miles). Unfortunately, very few NWS Coop sites measure hail. The combination of small, infrequent events being measured by a sparse climate network makes for very few reliable, long-term records of these events, particularly for large areas. Based on Changnon (1995), the Illinois River Bluffs area experiences two hail days per year, with the actual number varying greatly from year to year. The years with the most hail days were 1927, 1950, and 1954, each with seven. There are no indications of trends in hail days, based on these records. Thunderstorms On average, the Illinois River Bluffs area experiences about 40 days with thunderstorms each year. The annual number of days with thunder over the Illinois River Bluffs area since 1948 is shown in Figure 14, which is composed of data from Peoria (1948-1995). There is substantial year-to-year variation in thunderstorm days, ranging from as many as 56 in 1975 to as few as 23 in 1968. There is no significant trend in thunderstorm days. oa oO SS oa | PONUNEN sooaeon cnc IN a ie) ao Saal tg eco LU a a (a eros are tered 1940 1950 1960 1970 1980 1990 2000 # of Days with Thunder BS oO w Oo ie) oO Figure 14. Annual Number of Days with Thunderstorms at Peoria, 1948-1995 21 Summary Mean annual temperatures for Peoria show a warming trend through 1930, followed by a cooling trend until the early 1960s, before warming through 1996. The number of days with temperatures above or equal to 90°F shows an upward trend through 1938, followed by a slow decline through 1970, before returning to somewhat higher numbers from 1971 to 1996. The number of days with temperatures below or equal to 32°F shows no trends. The number of days with temperatures below or equal to 0°F shows no trends. For precipitation, there are no trends. There were no trends in the number of days with measurable precipitation. For snowfall and the number of days with snow, there was an upward trend through the 1980s, followed by a downward trend through 1996. Records extending back to 1901 show no clear trends in hail events. Similarly, there are no apparent trends in tornado events, although records date only to 1955. The number of days with thunderstorms has no significant trends since 1948. 22 Streamflow Surface water resources are an essential component of any ecosystem because they provide different types of habitats for aquatic and terrestrial biota. In addition to their natural functions, they are sources of water supply for domestic, industrial, and agricultural uses. Changes in natural and human factors, such as climate, land and water use, and hydrologic modifications, can greatly affect the quantity, quality, and distribution (both in space and time) of surface waters in a river basin. There are about 1,450 miles of rivers and streams in the [linois River Bluffs area. Their streamflow is monitored by stream gaging stations, which measure the flow of water over time, providing information on the amount and distribution of surface water passing the station. Since it is not feasible to monitor all streams in a basin, gaging stations are established at select locations, and the data collected are transferred to other parts of the watershed by applying hydrologic principles. Streamflow records are used to evaluate the impacts of changes in climate, land use, and other factors on the water resources of a river basin. The streams of the Illinois River Bluffs area consist of the Illinois River and a number of small- to medium-sized streams that drain the uplands and the bluffs. The variability of flows on the Illinois River is to a great degree influenced by large-scale rain events and climate influences from northeastern Illinois, which provides the major portion of the- river’s drainage area. Many of the tributary streams in the Illinois River Bluffs area are small, with flows rising and falling quickly in response to local climatic conditions. As a result, it is a fairly rare coincidence for the Illinois River and the local tributaries to be flooding at the same time or, in some cases, to be experiencing low flows at the same time. Stream Gaging Records Four stream gages in the Illinois River Bluffs area, presently or previously operated by the U.S. Geological Survey, have fifteen or more years of continuous daily flow data. These stations are listed in Table 8 and their locations are shown in Figure 15. Also listed in Table 8 is the Illinois River gage at Kingston Miles, located 22 miles downstream of the Illinois River Bluffs area, which provides a longer flow record for the Hlinois River. The Gimlet Creek gaging station is located along the bluff line, while the Crow Creek (West) gage near Henry is located in the alluvial valley just below the bluff. The Crow Creek gage near Washburn is located in the flatter upland portion of that stream. Stream profiles (elevation versus distance upstream) for both Crow Creeks were given earlier (Figure 2). 23 Cro w xs Y ee or —~ S, eh De 5 fn andy Cre s é ‘ert ite N Basin Boundary = 4 Lakes Streams A Gaging stations Figure 15. Stream Gaging Stations in the Illinois River Bluffs Table 8. USGS Stream Gaging Stations with Continuous Discharge Records Drainage Record Area Length . mi.) (years) USGS ID {Station name (s Period of record 05558300 {Illinois River at Henry 13543.0 15 198 l-present 05558500 |Crow Creek (West) near Henry 56.2 22 1949-7] 05559000 |Gimlet Creek at Sparland 57 | 1950-71 05559500 |Crow Creek near Washburn 115.0 27 1945-72 05568500 |Illinois River at Kingston Mines’ 15818.0 a7 1939-present Note: 'Located 22 miles downstream of the Illinois River Bluffs area Human Impacts on Streamflows in the Illinois River Bluffs Area The characteristics of streamflow in any moderately developed watershed will vary over time because of the cumulative effect of human activities in the region. Like most locations in Illinois, the Illinois River Bluffs area has experienced considerable land use modification since European settlement, including cultivation, drainage modification, removal of wetland areas, and deforestation. Most modifications began prior to the onset of streamgaging activities, and thus their impact cannot usually be detected in the gaging records. Climate variability has the greatest influence on streamflows from year to year and decade to decade. Its influence is usually large enough to help mask the impacts of the less obtrusive human modifications to flows, including that of land use modification. The major changes to streamflow during this century are assumed to occur from natural climatic variability, but it is possible that in the future they may be shown to have human influences. Other modifications to the watershed, such as the construction of reservoirs, point withdrawals from, and discharges to the streams have readily definable impacts on the stream flows. The most noticeable impact of this type comes from the diversion of Lake Michigan water to the Illinois River, for use in public water supply to most of the Chicago metropolitan area and for maintaining water levels in the Chicago Ship and Sanitary Canal. This diverted water accounts for over 20 percent of the total annual flow in the river and over 70 percent of the flow during drought conditions. ZS Annual Streamflow Variability Average streamflow varies greatly from year to year, and can also show sizable variation between decades. Figures 16a and 16b show the annual series of average streamflow for the Illinois River, and the tributaries in the Illinois River Bluffs area, respectively. For the Illinois River, the greatest and least annual runoffs occurred in 1993 and 1964, respectively. The long-term average flow for the Illinois River has been noticeably greater in the last 25 years since 1970. This can be attributed to coincident increases in annual precipitation and heavy rainfall events that have been observed in northeastern Illinois (Knapp, 1994; Kunkel, 1997). Streamgage records for the Illinois River at Henry indicate that the average annual flow for 1981-1996 has been 15,680 cubic feet per second (cfs); roughly 10% greater than the expected long-term average flow of 14,200 cfs. Of this flow amount, approximately 3,200 cfs originates from the Chicago diversion of Lake Michigan water into the Illinois River waterway. The remaining amount of flow is runoff from all portions of the watershed, and on average represents an equivalent runoff of 11 inches per year. The average flow for the tributaries in the Illinois River Bluffs area do not appear to have any trends. The average runoff of these tributaries over their periods of record ranges from 7 to 9 inches per years, and the long-term average runoff from these streams is expected to be about 9 inches. The greatest total annual flow on the tributaries occurred in 1970, with an annual runoff of over 20 inches. The least annual runoff, less than inch, was experienced in 1956. Statistical Trend Analysis Table 9 shows trend coefficients estimated for the annual flow record for individual stations. The trend analysis identifies a statistically significant increase in average flow for the Illinois River at Kingston Mines since 1939. On the other hand, the Illinois River at Henry (1981-1996) shows a significant decreasing trend over the last 15 years. This emphasizes the fact that trends in streamflow are dynamic and can vary significantly depending on the period of years being analyzed. Of additional interest is the season during which the flow increases have occurred. The trend statistics indicate that the average streamflows during the fall season have increased for all stations. The change in streamflows during other seasons are variable depending on location and period of record. 26 AVERAGE STREAM FLOW, inches AVERAGE STREAMFLOW, inches —6— Illinois River at Kingston Mines 1940 1950 1960 1970 1980 1990 —6— Gimlet Creek at Sparland —#-— Crow Creek near Washbum —@— Crow Creek (West) near Henry 1945 1950 1955 1960 1965 1970 Figure 16. Average Annual Streamflow for a) the Illinois River, and b) the Tributaries in the Illinois River Bluffs Area 27 2000 1975 Table 9. Trend Correlations for Annual and Seasonal Flows Kendall trend correlation Annual Fall Winter Spring Station and Period of Record Summer Crow Creek (West) near Henry (1949-71) Gimlet Creek at Sparland (1950-71) Crow Creek near Washburn (1946-72) Illinois River at Henry (1981-95) Illinois River at Kingston Mines (1940-95) -0.057 0.133 0.048 -0.076 -0.152 -0.060 0.066 -0.128 0.128 -0.202 -0.303 0.121 -0.030 -0.303 0.212 0.244 0.343 0.216 0.110 0.079 Daily and Seasonal Flow Variability Figure 17 plots the flow duration curves for the gages in the Illinois River Bluffs area. The flow duration curve provides an estimate of the frequency with which the given flows are exceeded. The shapes of the flow duration curves shown in Figure 17 display the differences that would be expected between small and large watersheds. The flows for the smaller tributaries tend to be highly variable; the peak flow rates measured at these gages are typically five to ten times greater than the maximum flow rate averaged over a one-day period, indicating a “flashy” nature with a quick rise and fall. Smaller streams will also typically dry up during the late summer and fall. As shown in Figure 17, Gimlet Creek is dry over one-third of the time. Larger streams will go dry during drought periods, perhaps only 10 percent of the time, as shown for Crow Creek near Washburn. The gage for Crow Creek (West) near Henry is located in an alluvial valley and shows a more sustained amount of low flow. : 100000 10000 1000 | | —@— Crow Creek(West) near Henry —t— Crow Creek near Washbum —*<— Gimlet Creek at Sparland 1 —6— Illinois River at Kingston Mines DISCHARGE, cfs 1 10 20 30 50 70 80 90 99 PERCENT CHANCE OF EXCEEDENCE Figure 17. Flow Duration Curves (Discharge Versus Probability) 28 Flows on the Illinois River are much more gradual, influenced by the great amount of water storage in its large watershed. The typical range of flows on the Illinois River is from 4,000 to 50,000 cfs, with a historical minimum and maximum of 2,100 and 108,000 cfs, respectively. As with all other locations in Illinois, streams in the Ilinois River Bluffs area display a well-defined seasonal cycle. Figure 18 shows the probability of flow rates on Crow Creek near Washburn for each month of the year. As shown, flows tend to be greatest during the spring and early summer months, March through June, dropping to their minimum values by late summer and autumn. 1000 2 Oo wi o ira < r re) 2 a Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Figure 18. Monthly Flow Probabilities for Crow Creek near Washburn Flooding and High Flows Ramamurthy et al. (1989) and Singh and Ramamurthy (1990) examined the increases in peak flows observed on the Illinois River for the period 1941-1985. These studies found that the annual peak flows showed a significant increase of about 50% over this period, and that the higher flows were caused by concurrent increases in precipitation amounts in the river’s watershed. Northeastern Illinois, in particular, has experienced a significant increase in the magnitude and frequency of heavy precipitation events (Kunkel et al., 1997). The following data provide an update of these previous trend studies, using data up through 1996, as well as information on flooding trends for tributaries in the Illinois River Bluffs area. 29 Figure 19a shows the annual series of peak flood discharges for the Illinois River at Kingston Mines and Henry. The two highest floods on record for the Illinois River at Kingston Mines occurred in 1982 and 1943. As indicated by the Kingston Mines series, there has been an gradual increase in flooding over the last 55 years; however, over the last 15 years there has been a downward trend in peak flood values, as seen in both the Henry and Kingston Mines records. There is no detectable trend in flooding at any of the tributary stations, as illustrated in Figure 19b. Statistical Trend Analysis Results of a statistical trend analysis of flood records are given in Table 10. The results show that the detection of flood trends is greatly impacted by the period of record being analyzed, with higher coefficients observed when the gaging record either starts during a drought period or ends with a major flood. Two general conclusions may be drawn from these coefficients: 1) there is an general increase in flooding for the [linois River from 1940 to the present, but there is also a downward trend since 1981; and 2) the smaller tributaries in the Illinois River Bluffs area have generally not experienced significant flood trends over their period of gaging, although the flood peaks for Crow Creek near Washburn show a reduction in flooding for the period 1946-1979. Table 10. Trend Correlations for Flood Volume and Peak Flow Kendall trend correlation Station name Crow Creek near Henry 1950-1971 0.117 o----- 1950-1976 ------ 0.014 Gimlet Creek at Sparland 1951-1971 -0.133 ------ 1950-1982 —-- -0.055 Crow Creek near Washbum 1946-1972 -0.117 ------ 1946-1979 ------ -0.202 Illinois River at Kingston Mines 1940-1996 0.157 0.129 1981-1996 -0.385 -0.317 Illinois River at Henry 1981-1996 -0.455 -0.250 30 DISCHARGE, cfs 1000000 DISCHARGE, cfs 100000 1945 1950 1960 1970 1980 1990 2000 —@— Crow Creek(West) near Henry —6— Gimlet at Sparland —f— Crow Creek near Washburn 1950 1955 1960 1965 1970 1975 1980 1985 Figure 19. Annual Peak Discharges for a) the Illinois River, and b) the Tributaries in the Illinois River Bluffs Area 31 Impact of Peoria Lake on Peak Flows An examination of Figure 19a also shows that, for all major flood events, the peak discharge on the Illinois River is significantly greater at Henry than at Kingston Mines. This occurs despite the fact that the drainage area at Kingston Mines is 20 percent greater than that at Henry, which causes the Kingston Mines location to have a significantly greater volume of flood waters. As illustrated in Figure 20, the peak discharges on the Illinois River are greatly reduced when flood waters pass through Peoria Lake. The lake, and other bottomland areas along the Illinois River, temporarily store much of the flow volume of these major flood events. Flood water is naturally released from Peoria Lake at a much more gradual rate, causing lower flood peaks. The outflow from Peoria Lake, as observed at Kingston Mines, may not surpass the inflow (at Henry) for well over a week after the peak flood flow has passed. Operation of the Peoria Lock and Dam has minimal impact on the flood storage provided by Peoria Lake and the adjacent bottomlands. 120000 —@ Illinois River at Henry 100000 —6- Illinois River at Kingston Mines 80000 2 o wi & —< 60000 po oO Q a 40000 20000 0 11/30/82 12/5/82 12/10/82 12/15/82 12/20/82 12/25/82 Figure 20. December 1982 Flood Hydrographs for the Illinois River at Henry and Kingston Mines Seasonal Distribution of Flood Events Table 11 presents the monthly distribution of the top 25 flood events for four gaging stations. For the Illinois River, major flooding occurs predominantly during spring, March through May. For the tributaries, a combination of locally-heavy rainfall and wet soil moisture conditions causes late spring and early summer flooding. 32 Table 11. Monthly Distribution of Top 25 Flood Events Station Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Crow Creek (West) near Henry 0 2 2 4 3 So 3 l 2 0 0 0 Gimlet Creek at Sparland 0: 0 l 3 3 § 5 l l 0 0 0 Crow Creek near Washburn Illinois River at Kingston Mines Drought and Low Flows The 7-day low flow (Q7) is used herein to describe the minimum streamflows expected during a drought or dry period. The Q7 is defined as the minimum average flow experienced during a seven-day period in that year. This minimum flow is useful for evaluating the effect of dry periods on river navigation. The 7-day, 10-year low flow is the lowest Q7 that would be expected to occur on average only once in ten years, and is commonly used for defining the minimum amount of dilution for streams receiving treatment effluents. Figure 21 presents the 7-day low flows computed for the Illinois River at Kingston Mines and the three tributary streams in the Illinois River Bluffs area. For the Illinois River, there is a significant increase in its low flows beginning in the late 1960s. This increase is generally proportional to and coincides with the increase in average streamflows, presented earlier. Low flows on the Illinois River are considerably greater than they were prior to 1900, resulting from the diversion of Lake Michigan water to the Illinois River basin. Many of the smaller tributaries in the Illinois River Bluffs area have zero flows during most summers or any extended dry period. Some of the largest tributaries have at least a small amount of flow throughout the entire year except during major droughts. The low flow records for the tributary streams in the Illinois River Bluffs area do not show any trends. 33 STREAMFLOW, cfs STREAMFLOW, cfs 9000 8000 7000 6000 5000 4000 3000 2000 —@— Illinois River at Kingston Mines | 1950 1960 1970 1980 1990 —@— Crow Creek near Henry —t— Crow Creek near Washbum 2000 Figure 21. Annual 7-Day Low Flows for a) the Illinois River, and b) the Tributaries in the Illinois River Bluffs Area 34 1975 Summary Since 1970 there has been a significant jump in the average annual flow in the Illinois River Bluffs area, a trend in many Illinois rivers. This increase in streamflow directly corresponds to a concurrent increase in average annual precipitation. There have been no observed trends in streamflows since the early 1970s, nor were there any observed trends in flow for the earlier period of record prior to 1970. There has also been a general increase in high flows and low flows related to the considerable jump in average streamflow amounts. However, the trend analysis indicates no overall increase in peak discharges. 35 Gj A>) 50 : 7 21 NG le das ar ee t ? ‘ tls Dee a one Me Rn We % Oey Sie ae Pe : yy me pee 3° | a 100 ; = 10/1/92 10/1/93 10/1/94 10/1/95 10/1/96 Date Figure 23. Variabilities of Flow Discharge and Instantaneous Suspended Sediment Concentration and Load for the Illinois River at Chillicothe 40 1500 Crow Creek | . 1000 ; _~ 1000 : : : > e G3 100 - a2} . . ° a: | | i= ° . = L oy nF : | r ’ 3 l . ‘ . ! 10/1/88 10/1/89 10/1/90 Date Figure 32. Variabilities of Flow Discharge and Instantaneous Suspended Sediment Concentration and Load for Dickison Run 49 T Qs (tons/day) = ° oO oO T TTT T TTT . ] ri = Os | , F “ 10/1/88 10/1/89 10/1/90 Date Figure 33. Variabilities of Flow Discharge and Instantaneous Suspended Sediment Concentration and Load for Farm Creek 50 To provide values in tons per day, sediment load was computed by multiplying the daily water discharge by the instantaneous sediment concentrations and applying the proper unit conversion factors. For stations with weekly sediment sampling, it was not possible to compute average daily and annual sediment loads. However, instantaneous sediment load provides a range of values to compare variability of sediment from year to year and from station to station. For the Illinois River at Chillicothe, the sediment load varied from 284 tons per day to 93,800 tons per day for the four water years monitored. It should be noted that sediment load depends on the size of the drainage area; therefore, a station with a larger drainage area will generally have a higher sediment load than one with a smaller drainage area under similar conditions. No annual sediment load can be calculated for the Illinois River at Chillicothe. This is because sediment load for each water year at this station was only for a period of five, nine, or ten months (see Table 14). Table 14. Annual Sediment Load for the Illinois River at Chillicothe Water year Water discharge (cfs) Sediment load (tons) 1993 8,933,370 677,950 | 1994 4,836,430 730,182 7 1995 5,631,210 824,982 ° 1996 4,909,800 943,638 ° Note: ' Represents a five-month total, * represents a nine-month total, and > represents a ten-month total. Sedimentation Sedimentation is the process by which eroded soil is deposited in stream channels, lakes, wetlands, and floodplains. In natural systems that have achieved dynamic equilibrium, the rates of erosion and sedimentation are in balance over a long period of time. This results in a stable system, at least until disruption by extreme events. However, in ecosystems where there are significant human activities, such as farming, construction, and hydraulic modifications, the dynamic equilibrium is disturbed, resulting in increased rates of erosion in some areas and a corresponding increased rate of sedimentation in other areas. Erosion rates are measured by estimating soil loss in upland areas and measuring streambank and bed erosion along drainageways. These measurements are generally not very accurate and thus are estimated indirectly, most often through evaluation of sediment transport rates based on instream sediment measurements and empirical equations. Similarly, measurement of sedimentation rates in stream channels is very difficult and expensive. 51 Lake sedimentation surveys provide the most reliable sedimentation measurements. Since lakes are typically created by constructing dams across rivers, creating a stagnant or slow-moving body of water, they trap most of the sediment that flows into them. The continuous accumulation of eroded soils in lake beds provides a good measure of how much soil has been eroded in the watershed upstream of the lake. In the Illinois River Bluffs area, surveys have been conducted for 3 lakes (Table 15). The sedimentation rates (in percent per year) for these lakes are high in comparison to most Illinois lakes, primarily because they involve extensive watershed areas draining into relatively small lakes. There are no sedimentation surveys for constructed reservoirs in the Illinois River Bluffs area. Records for the water depth of Peoria Lake have been collected and analyzed for sedimentation rates for the years 1903, 1965, 1976, and 1985. The sedimentation analyses for these surveys were analyzed by Demissie and Bhowmik (1986). An additional survey by the Corps of Engineers in 1988 is not used in this analysis due to the limited record length between the 1985 Water Survey study and the 1988 survey. These analyses are presented in the Table 15. Comparison of the pre-1965 to the post- 1965 rates and volumes should be made with caution. The rate for the period 1903 to 1965 includes the influence of several significant alterations to the watershed and river systems. These include 1) the early flows of the Illinois waterway from the Sanitary and Ship Canal and the manipulation of these flows to meet court-ordered withdrawal rates from Lake Michigan, 2) the construction of the Peoria Lock and Dam structures, 3) development of agricultural levee and drainage systems in the Illinois River Valley, and 4) agricultural drainage systems in the Peoria Lake area that bypassed the shoreline wetlands around the Lake. Table 15. Lake Sedimentation Rates in the Illinois River Bluffs Area (Volumes in acre-feet) Year Volume Average depth Average depth loss Lake name surveyed acre-feet feet feet per year Upper Peoria Lake 1903 96,000 7.6 1965 55,200 4.4 0.05 1976 42,200 3.4 0.09 1985 11,800 53 0.07 Lower Peoria Lake 1903 24,000 9.8 1965 17,700 12 0.04 1976 14,400 3.9 0.12 1985 11,800 Sp) 0.07 Peoria Lake 1903 120,000 8.0 1965 72,900 4.8 0.05 1976 56,600 3.8 0.09 1985 38,300 2.6 0.13 52 Water Use and Availability Statewide, water use has increased a modest 27% since 1965 (Illinois Department of Energy and Natural Resources, 1994). Most of that increase is in power generation. Water use for PWS has risen only about 7% during that time, less than the concurrent percentage increase in population. The number of public ground-water supply facilities in Illinois has risen significantly during that time, yet the total amount supplied by ground water remains near 25%. A dependable, adequate source of water is essential to meeting existing and potential population demands and industrial uses in Illinois. Modifications to and practical management of both surface and ground-water use have helped make Illinois’ water resources reliable. As individual facilities experience increases in water use, innovative alternative approaches to developing adequate water supplies must be developed, such as use of both surface and ground waters. Major metropolitan centers such as the Chicago area, Peoria, and Decatur, as well as smaller communities, have already developed surface and ground-water sources to meet their development needs and to sustain growth. The construction of impounding reservoirs has become and will remain economically and environmentally expensive, making it a less common approach. Proper management of water resources is necessary to ensure a reliable, high quality supply for the population. Water conservation practices will become increasingly important to reduce demand and to avoid exceeding available supplies. Both our ground- water resources and surface reservoir storage must be preserved to maintain reliable sources for future generations. Ground-Water Resources Ground water provides approximately one-third of Illinois’ population with drinking water. The sources of this water can be broken down into three major units: 1) sand and gravel, 2) shallow bedrock, and 3) deep bedrock. Most ground-water resources are centered in the northern two-thirds of Illinois. Sand-and-gravel aquifers are found along many of the major rivers and streams across the state and also in “buried bedrock valley” systems created by complex glacial and interglacial episodes of surface erosion. There are also many instances of thin sand-and- gravel deposits in the unconsolidated materials above bedrock. These thin deposits are used throughout Illinois to meet the water needs of small towns. Shallow bedrock units are more commonly used in the northern third of Illinois, whereas deep bedrock units are most widely used in the northeastern quarter (in and around the Chicago area). The variety of uses and the volume of water used vary widely throughout the state. This report describes ground-water availability and use in the Illinois River Bluffs area. 53 Data Sources Private Well Information The Illinois State Water Survey (ISWS) has maintained well construction reports since the late 1890s. Selected information from these documents has been computerized and is maintained in the Private Well Database. These data are easily queried and summarized for specific needs and form the basis of well distribution studies in the area. Public Well Information Public Water Supply (PWS) well information has been maintained at the ISWS since the late 1890s. Municipal well books (or files) have been created for virtually all of the reported surface and ground-water PWS facilities in Illinois. Details from these files are assembled in the Public-Industrial-Commercial Database, which was created to house water well and water use information collected by the ISWS. Ground-Water Use Information The water use data given in this report come from the records compiled by the ISWS’ Illinois Water Inventory Program (IWIP). This program was developed to document and facilitate planning and management of existing water resources in Illinois. Information for the program is collected through an annual water use summary mailed directly to each PWS facility. Data Limitations Several limitations must be taken into consideration when interpreting these data: 1. Information is reported by drillers and each PWS facility. 2. Data measuring devices are generally not very accurate. 3. Participation in the IWIP is voluntary. Information assembled from well construction reports and from the IWIP is considered “reported” information. This means that the data are as accurate as the reliability of the individual reporting or as mechanical devices dictate. The quality of the reported information depends upon the skill or budget of the driller or facility, respectively. Moreover, the ISWS estimates that only one-third to one-half of the wells in the state are on file at the Survey, mainly due to the lack of reporting regulations prior to 1976. In general, water use measuring devices, such as the meters used by PWS facilities, are not very accurate. In fact, errors of as much as 10% are not uncommon. Much of the information reported in the IWIP is estimated by the water operator or by program staff. 54 Participation in the program is not required by the State of Illinois, and each facility voluntarily reports its information through a yearly survey. However, not all facilities know of or respond to the water use questionnaire. After several mail and telephone attempts have been made to gather this information, estimates are made using various techniques. To help reduce errors associated with the program, reported water use information is checked against usage from previous years to identify any large-scale reporting errors. Ground-Water Availability The Illinois River Bluffs area encompasses portions of 9 counties: Bureau, LaSalle, Lee, Marshall, Peoria, Putnam, Stark, Tazewell, and Woodford. The portion of each county in the watershed varies from less than 1% (Stark County) to 96% (Marshall County). This section summarizes ground-water availability in the area, taking into consideration only those portions of each county that are actually in the watershed. Domestic and Farm Wells Available regional information indicates that ground water for domestic and farm use in the area is mostly obtained from two types of wells finished in the till (Salkregg, Kempton, 1958). Table 16 summarizes the number of reported private wells in the watershed by county and depth. Table 16. Number of Reported Private Wells in the Illinois River Bluffs Area (Source: ISWS Private Well Database) Depth range, feet 51-100 101-150 151-200 201-250 251-300 301-350 351-400 400+ Bureau 4 4 5 12 5 4 LaSalle 10 14 12 1 3 Lee 1 Marshall 209 72) 109 33 99 27 10 3 9 Peoria 246 483 172 98 56 15 17 14 13 Putnam 90 90 66 40 44 15 1 1 5 Stark 3 4 1 Tazewell Woodford Public Water Supply Wells Information from the ISWS’ Public-Industrial-Commercial Database indicates that most ground water for PWS use in the area comes from wells finished in the unconsolidated materials, generally the Sankoty sand, which supplies about 96% of the groundwater withdrawn. The Cambrian-Ordovician systems supply the remaining 4%. 55 Unconsolidated wells range in depth from 23 to 408 feet, while bedrock wells range in depth from 320 to 2,000 feet. A total of 40 public water supplies withdraw 13.50 million gallons per day (mgd), servicing a reported 199,872 residents at an average per capita daily water use of 71.7 gallons per day (gpd). 1995 Ground-Water Use Ground water constitutes a substantial portion of the total water used in the basin. Total ground-water use in the basin during 1995 is estimated to be 17.04 mgd, with 13.50 mgd for PWS facilities, 2.14 mgd for self-supplied industries (SSI), 0.84 mgd for rural/domestic uses, and 0.56 mgd for livestock watering. Public Water Supply In 1995, municipal residential use for 40 communities using ground water was reported to be 11.70 mgd, serving a combined population of 199,872. The average per capita use from these municipalities is 71.7 gpd. The facilities also delivered 1.80 mgd for industrial and commercial use. Self-Supplied Industry Self-supplied industries are defined as those facilities that meet all or a portion of their water needs from their own sources. In the Illinois River Bluffs area, 12 SSI facilities reported total ground-water pumpage of 2.14 mgd during 1995. Rural/Domestic There is no direct method for determining rural/domestic water use in the basin. To geta rough estimate for the area, several assumptions were made using existing information. The population served and number of services reported by PWS facilities were used to calculate an average population per service for all PWS facilities in the area. This number was used as an estimate of population per reported domestic well. The average PWS per capita use was then used as a multiplier to determine the total rural/domestic water use from each well. Based on information from the ISWS Private Well Database, which shows 3,646 reported wells in the Illinois River Bluffs area, an average of 3.2 people per service (well), and an average of 71.7 gpd per person, total rural/domestic water use was estimated to be 0.84 mgd. Livestock Watering Water withdrawals for livestock use in 1995 were estimated to be 0.56 mgd. Water use estimates for livestock are based on a fixed amount of water use per head for each type of animal. Percentages of the total animal population (Illinois Department of Agriculture, 1995) for the major livestock (cattle and hogs) in the counties were calculated based upon the percentage of county acres in the Illinois River Bluffs area. Daily consumption rates (beef cattle = 12 gpd, all other cattle = 35 gpd, and hogs = 4 gpd) provided the basis for these calculations. 56 Ground-Water Use Trends Ground-water use in the Illinois River Bluffs area has remained relatively constant over the last six years. During this period, total ground-water use has averaged 13.84 mgd and ranged from 12.58 to 15.64 mgd; PWS use has averaged 11.85 mgd and ranged from 10.15 to 13.78 mgd; and SSI use has averaged 1.99 mgd and ranged from 1.33 to 2.55 mgd. Table 17 shows the individual totals per year since 1990. No significant trends are evident in terms of water withdrawals in the basin. Table 17. Ground-Water Use Trends in the Illinois River Bluffs Area (in million gallons per day, mgd) Surface Water Resources The rivers, streams, and lakes of the Illinois River Bluffs area serve a wide variety of purposes, including uses for public water supply, recreation (boating, fishing, and swimming), and habitat for aquatic life. The primary focus of this section is on water withdrawn from streams for public water supply and the surface water resources available for such use. Water supply systems generally obtain surface water in one of three manners: 1) direct withdrawal from a stream, 2) impoundment of a stream to create a storage reservoir, and 3) creation of an off-channel (side-channel) storage reservoir into which stream water is pumped. As described below, there is substantial potential for direct withdrawals from the Illinois River for water supply, and several locations along the river bluffs for potential impounding reservoirs. The potential for side-channel storage also exists along most streams. Water Use and Availability The only major user of surface water for water supply in the Illinois River Bluffs area is the city of Peoria, which withdraws water from the Illinois River. Over the six year period, 1990-1995, the average amount of water withdrawal from the river has been 8.35 million gallons per day (mgd), or roughly 46 percent of that used for the city’s public water supply. The only other use of surface water in the area is a small industrial supply which reports pumping 0.03 mgd. =f) Most of the small communities and industries in the Illinois River Bluffs area discharge their treated wastewater into the Illinois River. However, the total amount of these effluents is fairly small, totaling less than 4 mgd. The city of Peoria discharges their treated wastewater downstream of Peoria Lake. There are a few small discharges into tributary streams, but these are not sufficient to significantly alter the flow characteristics of such streams. Potential for Development of Surface Water Supplies Direct Withdrawals from Streams The Illinois River is the only reach of stream in the area that is able to support a direct withdrawal for water use. There are no practical limitations on the amount of water use that could be supported by the river, and no anticipated negative impacts to its potential use. Impounding Reservoirs The tributaries in the Ilinois River Bluffs area provide a number of possible reservoir sites, primarily because of their valley slopes. Figure 34 shows the locations of 19 potential reservoir sites in the region, as given in Dawes and Terstriep (1966, 1967). Many of the potential reservoir sites could support a safe yield in excess of 2 mgd. In general, the construction of impounding reservoirs has become a less common option for developing a water supply, primarily because of costs and environmental concerns. As a result, the proximity of alternative sources should be considered in their proposed development. Since the Illinois River provides an ample supply of water, the reservoir sites that are farther from the river are the ones of greatest interest. Side-Channel Reservoirs There are no side-channel reservoirs in the Illinois River Bluffs area. The construction of side-channel reservoirs is generally not limited by local topography and could be a viable water supply option for a small water supply along most of the tributary streams in the basin. The amount of water supply that off-channel storage can provide is limited primarily by the temporal distribution of flow in the stream and the size of the storage reservoir. 58 Dry Richlan qd Cree, x05. cot Scale 1:410000 N Basin Boundary #) Potential reservoirs /\/ Streams Figure 34. Potential Reservoirs in the Illinois River Bluffs wv Pa ee ae : ‘ ue oie rile oF jw! »>ats ion ree Be { ‘ is Vay blo vio’ hs lo) afl of thee nieve legs es , 4a Giactrres Ci? “ rears rif Tan kL: " are a te @uall ditcbarges Ing up Teer oan Sieh Lo alin! Cafe hee Fo w Oia i ‘ 7 “ “wh : 7 , , -_ by) det hele Ha neva ok Seg gg oor om a" Be } j we we . eesthiin iar Siemens OC -_ not - ans § . \ ; ; . —. ot Fe 7 ‘ a) al: iJ y a mip a in Wie oe) A G : ' be = reds wakrn wii There cs 4 “a he seul Warr l ah Mee War ealrias (‘Mars Syrpypieniest fo A), a iD pilte —Pe, ee ws ‘a ¢ La — oy e —_ 3. —_ \ \" q y » fi ms . 7 f ~ Oe a | yo# mA. 3 i gi 7 fag: faik ith ; % ‘ 7 sting the Eich Soo UH ag desc s “I ¥ bee: Wine © 7 Sy Bites , wig 34 Ww Ay tad WO lye ees as giver, ap ae fodpy foe yee >a) wor Pes > ANT ars 3 aie 1 « > a ial a LJ \itng f tr we | a a ie $e inet ;* nag ar terval wi a wart kopply. pone ly boone OF Cote Be exrvip Ass nf sip iy ol dternw) weeds io equi : teveikeenent. Sinan pie Tings River pyaeee el ’ gn dw Aver ue panes OF ide Clue! Besar tc There st 10. Sele ehingl oervenm in rr gide-channel reser s.irs le puleraily pop fisted wales apply ontivg tors, viml! wade: Rpgy bade. ‘Tix. acieeot of sree rely due primarity by Une nce it Sider aurt tn if rae we wine 7 aT Ground-Water Quality This section examines ground-water quality records to determine temporal trends and to provide baseline water quality parameters in the Illinois River Bluffs area. Increasingly, ground-water contamination is discussed in the news media, and it may seem that the entire ground-water resource has been affected. However, these contamination events are often localized and may not represent widespread degradation of the ground-water resource. By examining the temporal trends in ground-water quality in the area, it may be possible to determine whether large-scale degradation of the ground-water resource has occurred. The general term “ground-water quality” refers to the chemical composition of ground water. Ground water originates as precipitation that filters into the ground. As the water infiltrates the soil, it begins to change chemically due to reactions with air in the soil and with the earth materials through which it flows. Human-induced chemical changes can also occur. In fact, contamination of ground water is generally the result of human- induced chemical changes and not naturally occurring processes. As a general rule, local ground-water quality tends to remain nearly constant under natural conditions because of long ground-water travel times. Therefore, significant changes in ground-water quality can indicate degradation of the ground-water resource. Data Sources The ground-water quality data that are used in this report come from two sources: private wells and municipal wells. The private well water quality data are compiled by the Chemistry Division of the Illinois State Water Survey (ISWS) as part of its water testing program and are maintained by the Office of Ground-Water Information in a water quality database. The municipal well data come from ISWS analyses and from the Illinois Environmental Protection Agency (IEPA) laboratories. The combined database now contains more than 50,000 records of chemical analyses from samples analyzed at the ISWS and IEPA laboratories. Some of these analyses date to the early part of the century, but most are from 1970 to the present. Before 1987, most analyses addressed inorganic compounds and physical parameters. Since then, many organic analyses have been added to the database from the IEPA Safe Drinking Water Act compliance monitoring program. This report presents information for only a portion of the chemical parameters in the ISWS database. 61 Data Limitations Several limitations must be understood before meaningful interpretation of the water quality data can begin: Representativeness of the sample Location information Data quality (checked by charge balance) Extrapolation to larger areas WnhN Generally, private well samples are not representative of regional ground-water quality. In most cases, private well owners submit samples for analysis only when they believe there may be a problem such as high iron or an odd odor or taste. However, while one or more constituents may not be representative, in general the remainder of the chemical information will be accurate and useful. As a result, the composite data may be skewed toward analyses with higher than normal concentrations. On the other hand, private well information probably gives a better picture of the spatial distribution of chemical ground-water quality than municipal well information because of the larger number of samples spread over a large area. Recent IEPA data from municipal wells will not be skewed because each well is sampled and analyzed on a regular basis. While this produces a much more representative sample overall, samples are generally limited to specific areas where municipalities are located. Therefore, these data may not be good indicators of regional ground-water quality. Much of the location information for the private wells is based solely on the location provided by the driller at the time the well was constructed. Generally, locations are given to the nearest 10-acre plot of land. For this discussion, that degree of resolution is adequate. However, it is not uncommon for a given location to be in error by as much as 6 miles. To circumvent possible location errors, this report presents results on a watershed basis. The validity of water quality data was not checked for this report. However, previous charge balance checking of these data was conducted for a similar statewide project (Illinois Department of Energy and Natural Resources, 1994). Charge balance is a simple measure of the accuracy of a water quality analysis. It measures the deviation from the constraint of electrical neutrality of the water by comparing total cations (positively charged ions) with total anions (negatively charged ions). Because many of the early analyses were performed for specific chemical constituents, a complete chemical analysis is not always available from which to calculate a charge balance. The statewide study searched the water quality database for analyses with sufficient chemical constituents to perform an ion balance. The charge balance checking of those data found that more than 98% of the analyses produced acceptable mass balance, which suggests that the chemical analyses are accurate in the database. Using that assumption 62 for this report, we feel confident that most of the analyses used are accurate and give representative water quality parameters for the Illinois River Bluffs area. However, this may be true only for large samples, a factor that should be considered when reviewing the results, as this report presents data from ten decades and a wide range of sample sizes. Extrapolating a point value (a well water sample) to a regional description of ground- water quality is difficult theoretically and beyond the scope of this report. However, none of the data provide a uniform spatial coverage. Therefore, it seems best to summarize the data on a watershed basis to ensure an adequate number of values. The private well analyses are more numerous and will likely provide better spatial coverage than the municipal well data, which are concentrated in isolated locations. Chemical Components Selected for Trend Analysis In many cases, ground-water contamination involves the introduction into ground water of industrial or agricultural chemicals such as organic solvents, heavy metals, fertilizers, and pesticides. However, recent evidence suggests that many of these contamination occurrences are localized and form finite plumes that extend down gradient from the source. Much of this information is relatively recent, dating back a few decades, and long-term records at any one site are rare. As mentioned earlier, changes in the concentrations of naturally occurring chemical elements such as chloride, sulfate, or nitrate also can indicate contamination. For ; instance, increasing chloride concentrations may indicate contamination from road salt or oil field brine, while increasing sulfate concentrations may be from acid wastes such as metal pickling, and increasing nitrate concentrations may result from fertilizer application, feed-lot runoff, or leaking septic tanks. These naturally occurring substances are the major components of mineral quality in ground water and are routinely included in ground-water quality analyses. Fortunately, the ISWS has maintained records of routine water quality analyses of private and commercial wells that extend as far back as the 1890s. After examination of these records, six chemical constituents were chosen for trend analyses based on the large number of available analyses and because they may be indicators of human-induced degradation of ground-water quality. These components are iron (Fe), total dissolved solids (TDS), sulfate (SOq), nitrate (NO3), chloride (Cl), and hardness (as CaCQ3). Aquifer Unit Analysis Ground water occurs in many types of geological materials and at various depths below the land surface. This variability results in significant differences of natural ground-water quality from one part of Illinois to another and from one aquifer to the next even at the same location. For the purpose of this trend analysis, wells that were finished in 63 unconsolidated sand and gravel units were grouped together, as were wells finished in bedrock units. Unconsolidated units are by far the most frequently used in the Illinois River Bluffs area. Out of the more than 3,646 private wells reported in the watershed, 3,095 indicate penetration into unconsolidated units. From the water quality analyses in the ISWS water quality database, 833 of 940 wells indicated that the water for the sample came from the unconsolidated units. In this report, unconsolidated and bedrock aquifers are treated separately in the descriptions of each chemical constituent. Discussion and Results Temporal trends in the six chemical constituents from unconsolidated and bedrock materials are summarized in this section. Tables 18 and 19 present the results of each decade’s analyses, including the maximum, minimum, mean, and median for each of the six chemical constituents for unconsolidated and bedrock materials, respectively. Median values are given in the tables by decade, beginning with 1900-1909 (Decade 0), 1910-1919 (Decade 1), and so on through the 1990s (Decade 9). Each decade covers the corresponding ten-year period, except for the partial decade of the 1990s. Median concentrations are given per decade so that temporal trends can be identified in the data set. Median values are the midpoints of a set of data, above which lie half the data points and below which is found the remaining half. These values are used to look at the central tendency of the data set. Although the arithmetic mean would also look at this statistic, it incorporates all data points into its analysis, which can move the mean value in one direction or another based upon maximum or minimum values. Outliers occur in many data sets. These are extreme values that tend to stand alone from the central values of the data set. They may lead to a false interpretation of the data set, whereas the median values are true values that are central to the data set. By looking at the median we can determine trends in the central portions of the data. However, for data sets with a small number of samples, the median may not necessarily be representative of the water quality in the area. It is important to recognize that the values included in these tables are reported values. While every attempt to verify the values was made, the validity of each value with regard to method error, etc. is not known. For this reason, the tables include every analysis in the database and all analysis results regardless of whether a value seems excessive and regardless of the sample size in the decade. Table 18. Chemical Constituents Selected for Trend Analysis, Unconsolidated Systems Chemical constituent Iron (Fe) Decade Sample size (N) 36 3 1 67 Minimum (mg/L) 0.0 0.0 0.1 0.0 Maximum (mg/L) 32 3.0 0.1 28.0 Mean (mg/L) 0.7 1.0 0.1 0.8 Median (mg/L) 0.2 0.1 0.1 0.2 TDS Sample size (N) 39 6 1 68 Minimum (mg/l) 350.0 417.0 557.0 313.0 Maximum (mg/l) 1476.0 794.0 557.0 1964.0 Mean (mg/l) 506.5 554.2 557.0 648.5 Median (mg/1) 45505295" 1557.0) S145 Sulfate (SO) Sample size (N) 26 4 1 67 Minimum (mg/1) 0.0 45.0 113.0 0.0 Maximum (mg/l) 38510" *110:0" “113:0"* 592:0 Mean (mg/l) 85.5 US 2" YA 1'13:09 1262 Median (mg/l) 60.5 73.0 113.0 80.0 Nitrate (NO;) Sample size (N) 29 2 1 66 Minimum (mg/1) 0.0 1.0 31.9 0.0 Maximum (mg/1) 73.0 61.9 31.9 64.0 Mean (mg/1) 13.2 18.3 31.9 13.0 Median (mg/l) 0.0 Sel 31.9 Me] Chloride (Cl) Sample size (N) 40 a 1 68 Minimum (mg/1) 3.0 2.0 29.0 2.0 Maximum (mg/l) 21305) 72:0) +29:0% 7070 Mean (mg/1) 25.0 40.4 29.0 66.0 Median (mg/1) 18.0 50.0 29.0 15.5 Hardness (as CaCQ;) Sample size (N) 36 1 1 68 Minimum (mg/l) 232.0 153.0 407.0 36.0 Maximum (mg/1) 590.0 153.0 407.0 735.0 Mean (mg/l) 379.2 153.0 407.0 423.7 Median (mg/l) 374.0 153.0 407.0 398.5 *Note: Decade 0=1900-1909, Decade 1=1910-1919, 65 169 19 55 210 0.0 0.0 0.0 0.0 16.2 4.7 8.7 14.0 1e7 0.9 1.5 1.4 0.7 0.2 0.7 0.3 188 19 49 207 306.0 366.0 308.0 248.0 1722.0 1088.0 2891.0 1179.0 517.0 562.1 590.8 475.5 452.5 464.0 454.0 449.0 179 5 8 145 0.0 57.0 0.0 0.0 559.0 447.0 235.0 370.0 70:7 "i206.850 1101.0 ¢5ie3 44.0) 45930455595 45:0 66 17 34 173 0.0 0.0 0.0 0.0 96.0 266.2 107.0 96.6 6.9 34.9 21.1 10.1 4.2 5.9 5.2 22 187 19 51 206 1.0 4.0 1.0 0.0 575.0 142.0 375.0 280.0 2TAGRE 2547 321 16.2 15.0 9.0 12.0 12.0 191 19 51 191 40 336.0 202.0 148.0 1224.0 1044.0 1810.0 650.0 379.4 462.1 414.8 358.4 352.0 404.0 342.0 348.0 and so on. 183 0.0 16.3 1.5 0.6 181 270.0 1510.0 463.2 431.0 52 125.0 1226.0 434.2 431.0 52 10.0 Table 19. Chemical Constituents Selected for Trend Analysis, Bedrock Aquifer Systems Decade Chemical constituent Iron (Fe) Sample size (N) 3 0) 1 14 8 21 12 38 20 3 Minimum (mg/L) 0.3 0.0 0.4 0.0 0.0 0.2 0.1 0.0 0.0 0.4 Maximum (mg/L) 4.0 0.0 0.4 6.0 21.9 22.0 14.0 17.0 1.6 0.9 Mean (mg/L) 1.6 0.0 0.4 1.6 5.2 4.4 1.9 1.5 0.6 0.7 Median (mg/L) 0.4 0.0 0.4 0.3 2.4 1.2 0.8 0.6 0.5 0.7 TDS Sample size (N) 4 1 l 14 8 24 12 37 20 3 Minimum (mg/1) 1260.0 1461.0 1454.0 356.0 374.0 1323.0 465.0 330.0 359.0 1110.0 Maximum (mg/1) 3154.0 1461.0 1454.0 3301.0 4186.0 3688.0 1510.0 6764.0 3428.0 2190.0 Mean (mg/l) 1916.0 1461.0 1454.0 1671.6 1571.2 1757.5 1049.9 1316.8 1581.0 1650.0 Median (mg/1) 1625.0 1461.0 1454.0 1446.5 1545.0 1554.5 1318.5 1306.0 1452.5 1650.0 Sulfate (SO) Sample size (N) 4 1 1 14 i) 10 5 31 19 3 Minimum (mg/l) 107:0 <181.0.92176:005 0.0 27:0: + 110:0's5 0:0 51:0. 59:0), -208:0 Maximum (mg/1) 225.0 181.0 176.0 555.0 515.0 280.0 232.0 390.0 422.0 396.0 Mean (mg/l) 176.8 181.0 176.0 190.3 215.6 224.8 184.4 216.6 212.8 284.7 Median (mg/l) 187.5 181.0 176.0 174.0 183.0 233.0 230.0 220.0 218.0 250.0 Nitrate (NO3) Sample size (N) 4 1 1 14 1 2 4 28 3 0 Minimum (mg/l) 0.0 0.7 1.0 0.8 0.8 0.6 0.3 0.0 0.3 0.0 Maximum (mg/l) 0.9 0.7 1.0 14.2 0.8 0.7 S16 2)k8 2.5 0.0 Mean (mg/l) 0.2 0.7 1.0 3.8 0.8 0.6 15.1 DES 1.1 0.0 Median (mg/1) 0.0 0.7 1.0 17 0.8 0.6 4.2 0.4 0.5 0.0 Chloride (Cl) Sample size (N) 5 2 1 14 8 25 12 37 20 3 Minimum (mg/l) 360.0 580.0 458.0 2.0 9.0 420.0 1.0 3.0 12.0 200.0 Maximum (mg/l) 1725.0 725.0 458.0 1683.0 2050.0 1900.0 560.0 3750.0 1720.0 956.0 Mean (mg/l) 722.0 652.5 458.0 633.7 545.9 645.4 267.6 449.3 610.1 588.0 Median (mg/1) 485.0 652.5 458.0 505.5 382.5 540.0 300.0 450.0 534.5 608.0 Hardness (as CaCQ;) Sample size (N) 3 0 0 14 8 25 12 32 14 3 Minimum (mg/l) 244.0 0.0 0.0 26:0 9200:0'5 42:0... 15210), 24.0. 42370 80.0 Maximum (mg/1) 430.0 0.0 0.0 1088.0 397.0 348.0 1090.0 508.0 416.0 325.0 Mean (mg/l) 321.3 0.0 0:0’ 278.1 ~285:8 223.1 ~265:2° 269:6° 25/6 1877 Median (mg/l) 290.0 0.0 0:0" "233.0, 251-5 © 220:0) 214:0) §255:0' -210'5;5 S8t0 *Note: Decade 0=1900-1909, Decade 1=1910-1919, and so on. 66 Iron (Fe) Iron in ground water occurs naturally in the soluble (ferrous) state. However, when exposed to air, iron becomes oxidized into the ferric state and forms fine to fluffy reddish-brown particles that will settle to the bottom of a container if allowed to sit long enough. The presence of iron in quantities much greater than 0.1 to 0.3 milligrams per liter (mg/l) usually causes reddish-brown stains on porcelain fixtures and laundry. The drinking water standards recommend a maximum limit of 0.3 mg/I iron to avoid staining (Gibb, 1973). Unconsolidated Systems Iron concentrations for unconsolidated systems in the watershed are given for each decade in Table 18. Minimum and maximum concentrations for all ten decades are 0.0 and 28.0 mg/l, respectively. These values clearly indicate a great deal of spatial variability in iron in the watershed. The median values range from 0.1 to 0.7 mg/I for all ten decades. While these median values show relatively high concentrations that could cause staining of porcelain fixtures (greater than 0.3 mg/l), they generally pose no threat to human health. In addition, the median values are all well above the Class I potable ground-water supply standard of 0.5 mg/l. Table 18 suggests no significant trend in iron concentrations in the area. Bedrock Aquifer Systems Iron concentrations for bedrock aquifer systems in the watershed are given for each decade in Table 19. Minimum and maximum concentrations for all ten decades are 0.0 and 22.0 mg/l, respectively. These values clearly indicate a great deal of spatial variability in iron in the watershed. The median values range from 0.3 to 2.4 mg/I for all ten decades. While these median values show relatively high concentrations that could cause staining of porcelain fixtures (greater than 0.3 mg/l), they generally pose no threat to human health. Table 19 suggests no significant trend in iron concentrations in the area. Total Dissolved Solids (TDS) The TDS content of ground water is a measure of the mineral solutes in the water. Water with a high mineral content may taste salty or brackish depending on the types of minerals in solution and their concentrations. In general, water containing more than 500 mg/l] TDS will taste slightly mineralized. However, the general public can become accustomed to the taste of water with concentrations of up to 2,000 mg/]. Water containing more than 3,000 mg/l TDS generally is not acceptable for domestic use, and at 5,000 to 6,000 mg/I, livestock should not drink the water. Because TDS concentration is a lumped measure of the total amount of dissolved chemical constituents in the water, it will not be a sensitive indicator of trace-level contamination. However, it is a good indicator of major inputs of ions or cations to ground water. 67 Unconsolidated Systems TDS concentrations in the unconsolidated systems in the watershed are given for each decade in Table 18. Minimum and maximum concentrations for all ten decades are 125.0 and 2,891.0 mg/l, respectively. Median values range from 431.0 to 557.0 mg/I for all ten decades. Generally, there are no significant trends in TDS concentrations in these aquifer systems in the watershed. Bedrock Aquifer Systems TDS concentrations for bedrock aquifer systems in the watershed are reported for each decade in Table 19. Minimum and maximum concentrations for all ten decades are 330.0 and 6,764.0 mg/l, respectively. Median values range from 1,306.0 to 1,650.0 mg/] for all ten decades. Generally, there are no significant trends in TDS concentrations in bedrock aquifer systems in the watershed. Any fluctuations from one decade to the next are more likely related to data limitations than to any inherent changes in ground-water quality. Sulfate (SO4) Water with high sulfate concentrations often has a medicinal taste and a pronounced laxative effect on those not accustomed to it. Sulfates generally are present in aquifer systems in one of three forms: magnesium sulfate (sometimes called Epsom salt), sodium sulfate (Glauber’s salt), or calcium sulfate (gypsum). They also occur in earth materials in a soluble form that is the source for natural concentrations of this compound. Human sources similar to those for chloride also can contribute locally to sulfate concentrations. Coal mining operations particularly are a common source of sulfate pollution, as are industrial wastes. Drinking water standards recommend an upper limit of 250 mg/I for sulfates. Upward trends in sulfate concentrations can suggest potential ground-water pollution. Unconsolidated Systems Sulfate concentrations for unconsolidated systems in the watershed are reported for each decade in Table 18. Minimum and maximum concentrations for all ten decades are 0.0 and 868.0 mg/l, respectively. Median values are all well below the drinking water standard, and range from 37.0 to 113.0 mg/I for all ten decades. Fluctuations from one decade to the next are more likely related to data limitations than to any inherent changes in ground-water quality. Bedrock Aquifer Systems Sulfate concentrations for bedrock aquifer systems in the watershed are reported for each decade in Table 19. Minimum and maximum concentrations for all ten decades are 0.0 and 555.0 mg/l, respectively. Median values are all well below the drinking water standard, and range from 174.0 to 250.0 mg/l for all ten decades. Table 19 indicates variability, but no significant trends in sulfate concentrations in the watershed. Fluctuations from one decade to the next are more likely related to data limitations than to any inherent changes in ground-water quality. 68 Nitrate (NO3) Nitrates are considered harmful to fetuses and children under the age of one when concentrations in drinking water supplies exceed 45 mg/l (as NOs), or the approximate equivalent of 10 mg/I nitrogen (N). Excessive nitrate concentrations in water may cause “blue baby” syndrome (methmoglobinemia) when such water is used in the preparation of infant feeding formulas. Inorganic nitrogen fertilizer has proven to be a source of nitrate pollution in some shallow aquifers, and may become an even more significant source in the future as ever increasing quantities are applied to Illinois farmlands. Upward trends in concentrations of nitrate may be a good indication that farm practices in the area are affecting the ground-water environment. Unconsolidated Systems Nitrate concentrations for unconsolidated systems in the watershed are reported for each decade in Table 18. Minimum and maximum concentrations for all ten decades are 0.0 and 266.2 mg/l, respectively. Maximum concentrations should be viewed as an outlier of the dataset, and not as representative of the water quality in the area. The majority of the median values are well below the drinking water standards, and range from 0.0 to 31.9 mg/l] for all ten decades. However, the ISWS has documented numerous cases of elevated nitrate levels associated with rural private wells in and around this area (Wilson et al., 1992). The evidence suggests that rural well contamination is associated more with farmstead contamination of the local ground water or well than with regional contamination of major portions of an aquifer from land application of fertilizers. This topic is actively being studied. Bedrock Aquifer Systems Nitrate concentrations for bedrock aquifer systems in the watershed are reported for each decade in Table 19. Minimum and maximum concentrations for all ten decades are 0.0 and 51.6 mg/l, respectively. Median values are well below the drinking water standards, and range from 0.0 to 4.2 mg/l for all ten decades. Chloride (Cl) Chloride is generally present in aquifer systems as sodium chloride or calcium chloride. Concentrations greater than about 250 mg/l usually cause the water to taste salty. Chloride occurs in earth materials in a soluble form that is the source for normal concentrations of this mineral in water. Of the constituents examined in this report, chloride is one of the most likely to indicate the impacts of anthropogenic activity on ground water. Upward trends in chloride concentrations may indicate contamination from road salt or oil field brine. The drinking water standards recommend an upper limit of 250 mg/l for chloride. In sand and gravel aquifers throughout most of the state, chloride concentrations are usually less than 10 mg/l. 69 Unconsolidated Systems Chloride concentrations for unconsolidated systems in the watershed are reported for each decade in Table 18. Minimum and maximum concentrations for all ten decades are 0.0 and 707.0 mg/l, respectively. Median values are well below the drinking water standard, and range from 9.0 to 50.0 mg/l for all ten decades. Table 18 indicates no significant trends in chloride concentrations in the watershed. Bedrock Aquifer Systems Chloride concentrations for bedrock aquifer systems in the watershed are reported for each decade in Table 19. Minimum and maximum concentrations for all ten decades are 1.0 and 3,750.0 mg/l, respectively. Median yalues range from 300.0 to 652.5 mg/l for all ten decades. Table 19 indicates no significant trends in median chloride concentrations in the watershed. Fluctuations from one decade to the next are more likely related to data limitations than to any inherent changes in ground-water quality. Hardness (as CaCQ3) Hardness in water is caused by calcium and magnesium. These hardness-forming minerals generally are of major importance to users since they affect the consumption of soap and soap products and produce scale in water heaters, pipes, and other parts of the water system. The drinking water standards do not recommend an upper limit for hardness. The distinction between hard and soft water is relative, depending on the type of water a person is accustomed to. The ISWS categorizes water from 0 to 75 mg/l as soft, 75 to 125 mg/1 as fairly soft, 125 to 250 mg/I] as moderately hard, 250 to 400 mg/l as hard, and over 400 mg/l as very hard. Unconsolidated Systems Hardness concentrations for unconsolidated systems in the watershed are reported for each decade in Table 18. Minimum and maximum concentrations for all ten decades are 4.0 and 1.810.0 mg/l, respectively. Median values range from 153.0 to 407.0 mg/I for all ten decades, indicating moderately hard to hard water in this area. Bedrock Aquifer Systems Hardness concentrations for bedrock aquifer systems in the watershed are reported for each decade in Table 19. Minimum and maximum concentrations for all ten decades are 23.0 and 1,090.0 mg/l, respectively. Median values range from 158.0 to 290.0 mg/I for all ten decades. The water is considered moderately hard in this area. No trends are observed in hardness concentrations from the bedrock in this area. 70 Summary This work was undertaken to examine long-term temporal trends in ground-water quality in the Ilinois River Bluffs area. Data from private and municipal wells were the primary sources of information used to show the trends in six chemical constituents of ground water in the area. These data demonstrate that on a watershed scale, ground water has not been degraded with respect to the six chemicals examined. Fluctuations from one decade to the next are more likely related to data limitations than to any inherent changes in ground-water quality. It is also evident that the sample size in each decade can play a role in trend analysis. Much of the contamination of Illinois’ ground water is localized. Nonetheless, this contamination can render a private or municipal ground-water supply unusable. Once contaminated, ground water is very difficult and expensive to clean, and clean-up may take many years to complete. Clearly it is in the best interests of the people of Illinois to protect their ground-water resource through prevention of contamination. Although no significant trends in water quality for these six constituents are apparent, the information provides baseline water quality for the watershed. This information can be used in future studies of the area as a reference to determine whether the local ground- water quality is degrading. 71 : ia ACnneTHR : t i §£ rl ie east 's r Dee Al, cae Pa ie THche sa Al Ds 4430) Aa .. atu uit ea’ hoe Aire ga! Stet! ge odingid VAS Cor jt» 4 " aa AEE Vy tp hg ip An a LM * Mi wei aes aera Roa 1s. dd fgh ‘ai otbatly {7 deen, Ba apis sae yous elke cea sccuh aethand Tea cn) overgant ot) ena alee: j set gurtoeqeen thre hobargsh asad ml iti writs ai itt gemma ash oF kan jor viral siar vie ah One 9 Lip Flu Seip ct Jeni aherntay sia sha 4 LAA Belle 19h y/ DAagTE, j : | vole wey 4 <9g] . RC wy ate F hens ; ‘ | | ; Nie Shel) aloe, lier saw oo i re | eit % . ae ane eal ay Mey Neer hf ata? AT eon eee rae - ott F See Cir Oa A610 RY BR 4, Dap ANY - ‘Bey LEO spies MAR ae Wy, Fab ae aati, havovy bianca. vadltlh to Aico at! te eosin 1220 od di 2c 36 oneal. sepslqeias ob emg ns a ale Nap pe etre PATE toh a8] Pouch S7illOev? 1S’ bm 11g ried) joars " : Tt) Je ot pny 9 a th ae eat bag Jona pee sit Ath ip a“ “5? 3 ei olf Sat Sig a Os wet iv rat a at bie ip). ne to), Wiel thik gh Felis ‘tuare Ost) 6 as ak tee me nay ; 7 owe Ay : men eee ‘ eu j 70a ss 1 aw | 5. iy oe oe seer y fe Rr pai r! * - &cype on — 7 ? to > 2k iin! (unt ett...” \¢ = ia Dea tee, 4 ae) and quar OD engl es eeny test. Uormiagl/ Aahedl S > coms: Oncsd eoncetnaom for ui iealidated eystenre sche watered any pas ems ¢ vtech doculé in Varte (6, bY ain ind fra xinit cConeenttlons toy al Ware PL WIO2 au). pepectivaly. Mediu ealact Gage Tg (Sowa mghe ten Cacgden. vadieaing Moderuitly her oo nard wher’ Hitivemee. m Rare Se ewZrons teerertreionn for betiradk agullarsypteme wo i oh a a eth Gecats ‘a Ta 16 Miner, eG cnx naios concerntronares tat emi n St apa O80 Omphreqpendvily. Madiun vale elige Ene IEG ees glk fer mecedes, ‘Tre worer Sv eondidened mideamainyy liad i Si Mowat Gewervead ln Lerdhes? coieesurptings Hila ce meh ; References Introduction Illinois Department of Natural Resources. 1996. Illinois Land Cover, An Atlas. IDNR- 96/05. Springfield, IL. Illinois Department of Natural Resources. 1996. Illinois Land Cover, An Atlas. Compact Disc. Springfield, IL. Leighton, M.M., G.E. Ekblaw, and L. Horberg. 1948. Physiographic Divisions of Illinois. Dlinois State Geological Survey Report of Investigation 129. Champaign, IL. Mattingly, R.L., and E.E. Herricks. 1991. Channelization of Streams and Rivers in Illinois: Procedural Review and Selected Case Studies. Illinois Department of Natural Resources Report ILENR/RE-WR-91/01. Suloway, L., and M. Hubbell. 1994. Wetland Resources of Illinois: An Analysis and Atlas. Illinois Natural History Survey Special Publication 15. Champaign, IL. Climate and Trends in Climate Changnon, S.A., Jr. 1984. Climate Fluctuations in Illinois: 1901-1980. Illinois State Water Survey Bulletin 68. Champaign, IL. Changnon, S.A., Jr. 1995. Temporal Fluctuations of Hail in Dlinois. [linois State Water Survey Miscellaneous Publication 167. Champaign, IL. Illinois Department of Energy and Natural Resources. 1994. The Changing Illinois Environment: Critical Trends. Volume 1: Air Resources. ILENR/RE-EA-94/05(1). Springfield, IL. Streamflow Knapp, H.V. 1994. Hydrologic Trends in the Upper Mississippi River Basin. Water International 19 (4): 199-206. Kunkel, K.E., K. Andsager, and D.R. Easterling. 1997. Trends in Heavy Precipitation Events over the Continental U.S. Manuscript. Ramamurthy, G.S., K.P. Singh, and M.L. Terstriep. 1989. Increased Duration of High Flows along the Illinois and Mississippi Rivers: Trends and Agricultural Impacts. Illinois State Water Survey Contract Report 478. Erosion and Sedimentation Demissie, M., L. Keefer, R. Xia. 1992. Erosion and Sedimentation in the Illinois River Basin. ILENR/RE-WR-92/04, Springfield, IL. 73 Water Use and Availability Dawes, J.H., and M.L. Terstriep. 1966. Potential Surface Water Resources of North- Central Illinois. Illinois State Water Survey Report of Investigation 56. Dawes, J.H., and M.L. Terstriep. 1967. Potential Surface Water Resources of North- Central Hlinois. Hlinois State Water Survey Report of Investigation 58. Illinois Department of Energy and Natural Resources. 1994. The Changing Illinois Environment: Critical Trends. Volume 2: Water Resources. ILENR/RE-EA- 94/05(2). Springfield, IL. Selkregg, L.F., and Kempton, J.P., 1958. Groundwater Geology in East-Central Illinois, Illinois State Geological Survey Circular 248. Ground-Water Quality Gibb, J.P. 1973. Water Quality and Treatment of Domestic Groundwater Supplies. Illinois State Water Survey Circular 118. Champaign, IL. Illinois Department of Energy and Natural Resources. 1994. The Changing Illinois Environment: Critical Trends. Volume 2: Water Resources ILENR/RE-EA- 94/05(2). Springfield, IL. Wilson, S.D., K.J. Hlinka, J.M. Shafer, J.R. Karny, and K.A. Panczak. 1992. “Agricultural chemical contamination of shallow-bored and dug wells.” In Research on Agricultural Chemicals in Illinois Groundwater: Status and Future Directions II. Proceedings of the second annual conference, Ilinois Groundwater Consortium, Southern Illinois University, Carbondale, IL. 74 (EE —————————— The Illinois Department of Natural Resources receives federal financial assistance and therefore must comply with federal anti-discrimination laws. In compliance with the Illinois Human Rights Act, the Illinois Constitution, Title VI of the 1964 Civil Rights Act, Section 504 of the Rehabilitation Act of 1973 as amended, and the U.S. Constitution, the Illinois Department of Natural Resources does not discriminate on the basis of race, color, sex, national origin, age or disability. If you believe you have been discriminated against in any program, activity or facility please contact the Equal Employment Opportunity Officer, Department of Natural Resources, 524 S. Second St., Springfield, IL 62701-1787, (217) 782-7616, or the Office of Human Rights, U.S. Fish & Wildlife Service, Washington, D.C. 20240. All public meetings conducted by the Department of Natural Resources will be accessible to handicapped individuals in compliance with Executive Order No. 5 and pertinent state and federal laws, upon notification of the anticipated attendance. Handicapped persons planning to attend and needing special accommodations should inform the Department of Natural Resources at least five days prior to the meeting by telephoning or writing the Equal Employment Opportunity Officer, Department of Natural Resources, 524 S. Second St., Springfield, IL 62701-1787, phone (217) 782-7616. Department of Natural Resources information is available to the hearing impaired by calling DNR’s Telecommunications Device for the Deaf: (217) 782-9175. The Ameritech Relay Number is (800) 526- 0844.