THE INFLUENCE OF HYDROLOGY AND TIME ON ... - OhioLINK ETD

8 downloads 0 Views 2MB Size Report
Mitsch, W. J., C. J. Anderson, M. E. Hernandez, A. Altor, and L. Zhang. 2004. Net ...... Bioscience 29:344-352. ...... Anne Altor assisted with soil texture analyses.
THE INFLUENCE OF HYDROLOGY AND TIME ON PRODUCTIVITY AND SOIL DEVELOPMENT OF CREATED AND RESTORED WETLANDS

DISSERTATION Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University

By Christopher J. Anderson, M.S. ***** The Ohio State University 2005

Dissertation Committee: Approved by William J. Mitsch, Adviser Warren A. Dick P. Charles Goebel

Adviser School of Natural Resources

ABSTRACT In created and restored wetlands, hydrology (the depth, duration, and dynamics of water in wetlands) and time play an important role in regulating most ecological processes including productivity and soil development. The influence of hydrology on created and restored wetlands was examined using full-scale ecosystems and replicated mesocosm systems at the Olentangy River Wetland Research Park (ORWRP). In one study, twenty 540-liter tubs or ‘mesocosms’ were planted with either one of two wetland plants common to the region: narrow-leaved cattail (Typha angustifolia L.) or softstemmed bulrush (Schoenoplectus tabernaemontani C.C. (Gmel) Palla). For each species, half the mesocosms were pumped with river water based on a monthly pulsing regime while the other half was pumped on a steady-flow regime (an even amount of water was provided weekly). Overall, Typha wetlands were significantly more productive than Schoenoplectus wetlands; however no significant differences in productivity or morphology were observed between pulsed or steady-flow wetlands among species groups. No significant differences in nutrient concentrations, uptake or uptake efficiency were detected among species groups either; however hydrology did influence plant tissue N:P ratios (P154 m3 sec-1 over the preceding two years and basal area increment (BAI) and b) the number of days of river discharge >154 m3 sec-1 and BAI over 2yr periods from 1991-2004. Open symbols represent post-restoration years........77

4.1

The two experimental wetlands at The Olentangy River Wetland Research Park (ORWRP) at The Ohio State University in Columbus, Ohio pumping system and water control structures. ....................................................................107

4.2

The 10-m grid and locations used for soil sampling at the ORWRP experimental wetlands in 1993, 1995 and 2004. Shaded areas within the grid map represent approximate location of the deeper, open water (OW) zones. ...................................................................................................................108

4.3

Comparison of combined mean (±1 SE) for a) percent organic matter (n=32), b) total C (n=18), c) total P (n=18), d) available P (n=28), e) exchangeable Ca (n=28), f) exchangeable K (n=28), g) exchangeable Mg (n=28), and h) soil pH (n=28) at the ORWRP experimental wetlands in 1993, 1995 and 2004 at 0-8 and 8-16 cm depths. Letters denote differences between years and depths detected at p35) are in bold.......................79

3.3

Results of paired t-tests for mean (±1 SE) basal area increment (BAI) (% and cm2 yr-1) of canopy trees pre- and post-restoration. NS denotes nonsignificant p-value..................................................................................................80

3.4

Results of paired t-tests for mean (±1 SE) basal area increment (BAI) (% and cm2 yr-1) pre- and post-restoration for boxelder (Acer negundo L.) and Ohio buckeye (Aesculus glabra Willd.). NS denotes non-significant pvalue.......................................................................................................................81

4.1

Variogram characteristics for soil percent organic matter concentrations in Wetland 1 and 2 at 0-8 cm depth for 1993, 1995 and 2004.................................115

5.1

Mean (±1 SE) physiochemical conditions of sediment in the emergent vegetation (EM) and open water (OW) zones of the planted (Wetland 1) and naturally colonized (Wetland 2) wetlands at the Olentangy River Wetland Research Park in May 2004. .................................................................151

5.2

Range of mean annual accumulation rates of sediment and nutrients for Wetland 1 and 2 at the Olentangy River Wetland Research Park, 19942004......................................................................................................................152

A.1

Aboveground biomass, belowground biomass, total biomass and root:shoot (R:S) ratio estimated from the experimental Schoenoplectus (Sch.) and Typha (Typ.) mesocosms in August 2003.. .........................................................168 xviii

A.2

Mean (±1 SE) stem/ramet density for experimental mesocosm in 20022003......................................................................................................................169

A.3

Mean (±1 SE) number of inflorescence for experimental mesocosms in 2002-2003. ...........................................................................................................169

A.4

Mean (±1 SE) maximum stem height length based on the measured length of the five longest stems/ramets at each mesocosm plot .....................................170

A.5

Mean (±1 SE) stem height length based on the measured length of 12 randomly selected stems/ramets at each mesocosm plot .....................................170

A.6

Nutrient concentrations of plant tissue (live and senescent) collected from experimental Schoenoplectus (Sch.) and Typha (Typ.) mesocosms. All plant biomass were harvested in August 2003.....................................................171

A.7

Olentangy River water P and NO3-N concentration and input into wetland mesocosms during the 2003 experimental wet season.........................................172

B.1

Species, importance value, relative density, relative dominance and relative frequency of trees (>5cm dbh) observed in plots at the north section of the bottomland forest .................................................................................................174

B.2

Species, importance value, relative density, relative dominance and relative frequency of trees (>5cm dbh) observed in plots at the south section of the bottomland forest .................................................................................................175

B.3

Cumulative mean leaf litter biomass and monthly section mean (±1 SE) collected in the bottomland hardwood forest leaf traps between June 2003 and May 2004.......................................................................................................176

B.4

Cumulative mean reproductive material biomass and monthly section mean (±1 SE) collected in the bottomland hardwood forest leaf traps between June 2003 and May 2004 .....................................................................................177

B.5

Cumulative mean woody material biomass and section mean (±1 SE) collected in the bottomland hardwood forest leaf traps between June 2003 and May 2004.......................................................................................................178

B.6

Tree specific gravity (per Alden 1995 and U.S. Forest Products Laboratory 1974), dbh, tree height and estimated wood production for all trees >5cm dbh in the bottomland hardwood forest tree plots ...............................................179

xix

B.7

Surveyed elevations of plots corners (NW, NE, SE, and SW) and leaf traps (LT) for each bottomland hardwood forest tree plot. Elevation mean and variance based on all measured plot elevations ...................................................190

B.8

Mean trap and plot canopy cover for each plot in the bottomland hardwood forest ....................................................................................................................191

B.9

Annual tree basal growth increments and mean annual growth increment (±1 SE) for the pre-restoration years (1991-2000) and post-restoration years (2001-2004)..........................................................................................................193

C.1

Physiochemical soil characteristics at 0-8 and 8-16 cm depths in 2004. Percent organic C results in bold-type were lab-analyzed and those in regular-type were based on regression analysis with percent organic matter. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text).....................................................................................................201

C.2

Total C, total N, total P, and pH of experimental wetland soils at 0-8 and 816 cm depths in 2004. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text)........................................................................204

C.3

Available P, exchangeable cations (Ca, Mg and K) of the experimental wetland soils at 0-8 and 8-16 cm depths in 2004. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text) ...................................206

C.4

Percent and mean (±1 SE) textural classes of sediment in the experimental wetlands in 2004. Sample coordinates based on the 10x10 m grid system at the experimental wetlands; cover type consisted of emergent (EM) and open water (OW) zones; and sub-basin refers to OW zone in proximity to wetland inflow/outflow (see text) ........................................................................208

C.5

Micronutrient concentrations of sediment (0-8 depth) at the experimental wetlands in 2004. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text)........................................................................209

xx

CHAPTER 1

INTRODUCTION

In the last 35 years, interest in wetland ecology has increased substantially. This has come after wide-spread recognition that wetlands provide very valuable services to the landscape and society (e.g., flood attenuation, water quality improvement and habitat for wildlife). Perhaps the most important piece of legislation in the United States interpreted to protect wetlands was Section 404 of the Federal Water Pollution Control Act and other subsequent amendments [a.k.a. the Clean Water Act (CWA)]. Through a dredge-and-fill permitting process, Section 404 mandated that wetlands be protected for their ability to protect the quality of the nation’s waterways. Administered primarily by the U.S. Army Corps of Engineers and the U.S. Environmental Protection Agency, policy was implemented that wetland mitigation would be required for wetland impacts that were deemed unavoidable. Therefore starting in the late 1970s the United States entered the foray of creating compensatory wetlands. Originally there was very little science involved with the design of these wetlands and not surprising, many of the initial created wetlands were unsuccessful (Erwin 1991). 1

However, as the number of designers trained in wetland ecology grew, there was increasingly more created wetlands that appeared to be comparable to natural wetlands. However, the criteria used by regulatory agencies to assess successful wetland mitigation have been criticized as being too cursory and the timelines for monitoring too short (Mitsch and Wilson 1996, Zedler and Calloway 1999). Consequently there is still debate as to how much we are truly replacing through the policy of wetland mitigation and how long does it take. These questions were the motivation for much of this research. Studying at The Olentangy River Wetland Research Park (ORWRP) (Fig. 1.1), I used created and restored wetlands to evaluate the progression of wetland functions and attributes over time and under variable hydrologic conditions.

1.1 Research goals and objectives The goal of this dissertation research was to evaluate the role of hydrology and time on wetland productivity and soil development. Using the wetland facilities at the ORWRP, the specific objectives of this dissertation were the following:

1)

to evaluate the influence of a pulsing water regime on the productivity and nutrient use efficiency of wetland macrophytes (Chapter 2);

2)

to examine the effect of hydrologic restoration on the productivity of a bottomland hardwood forest (Chapter 3);

3)

to examine the change in soil physiochemical parameters over the ten-year history of two experimental wetlands and to examine the spatial changes in the concentration of soil organic matter (Chapter 4); and 2

4)

to evaluate the amount and spatial distribution of sediment and nutrient accumulation in the same two experimental wetlands (Chapter 5).

Establishing hydrology (the depth, duration, and dynamics of water in wetlands) in a created wetland or reintroducing it to a restored wetland generates conditions that are distinct from terrestrial systems. For instance, a flood-pulse hydrology (the periodic flooding from an adjacent river or lake) that was once common to riverine wetlands in this United States Midwest (Baker et al. 2004), has been shown to increase plant productivity (the production of biomass) and nutrient uptake in forested wetlands throughout the world (Mitsch and Ewel 1979, Junk et al. 1989, Tockner et al. 2000). There is the potential for this to occur in Ohio as well because many of the inflowing river waters are often laden with high nutrient loads that can have a “fertilizer effect” (Mitsch and Gosselink 2000). However, when inundation is prolonged, flooding can also cause adverse conditions that may impede plant growth and nutrient uptake (Mitsch and Rust 1994, Kozlowski 1997). This research included two studies that explored the relationship between flood pulsing and wetland/floodplain productivity. The first study used to 20 replicated mesocosm tubs at the ORWRP (Fig. 1.1) to evaluate pulsing effects on herbaceous wetlands. While pulsing has been documented to enhance forested wetland (Mitsch and Ewel 1979) and aquatic (Hein et al. 1999) productivity, few studies have evaluated the influence of a pulsing regime on herbaceous riverine wetlands. This study was designed to complement an ecosystem-scale, flood pulsing study conducted concurrently at the 1ha ORWRP experimental wetlands (Fig 1.1). 3

A second study evaluated the restoration of a pulsing regime to a 5.2-ha bottomland hardwood forest at the ORWRP (Fig. 1.1). The effect of hydrology on riparian forest productivity has been the subject to several studies (Mitsch and Ewel 1979, Taylor et al.1990, Tockner et al. 2000, Robertson et al. 2001) and many investigators have concluded that periodic flood pulses have an important and positive effect on the productivity of the ecosystem. This is consistent with Odum’s subsidy-stress model (Odum et al. 1979), in which flooding can be beneficial to the productivity of the system, depending upon the frequency, timing and duration of the flood events. Other studies however have different results. Brown and Peterson (1983) and Burke et al. (1999) found that permanently flooded forest zones rather than periodically flooded zones had higher productivity while Megonigal et al. (1997) found no difference between upland and periodically flooded forest productivity. Given the inconsistencies in bottomland responses to flooding, it has been suggested that more studies that evaluate an existing forest under a changing hydrologic regime are needed to elucidate the influence of hydrology on bottomland forests (Conner 1994, Megonigal et al. 1997). This is the scenario at the ORWRP bottomland where this research evaluated changes in productivity 4 years after a pulsing hydrology was restored. Changes in soil condition have been the least studied component of created wetlands. The anaerobic conditions of wetland soils induced by hydrology have several unique biogeochemical processes that contribute to their ability to absorb pollutants and improve water quality (Mitsch and Gosselink 2000). The organic matter stored in wetland soils is an important food source for invertebrates that become the food base for higher 4

organisms. The organic matter also is a valuable and permanent sink for atmospheric carbon that would otherwise contribute to climate change. Despite the critical importance of soils to overall wetland functions, they are rarely considered when mitigation wetlands are evaluated for success. Research studies that have been designed to compare the soils of created wetlands to natural reference wetlands (Bishel-Machung et al. 1996, Shaffer and Ernst 1999, Zedler and Callaway 1999; Nair et al. 2001, Campbell et al. 2002, Brooks et al. 2005) typically find some progression toward natural wetland soil conditions but with substantial deficiencies in many key characteristics (e.g., lower soil organic matter concentrations, coarser texture, and dissimilar nutrient concentrations and pH). As part of this dissertation, a survey of soil physiochemical parameters was evaluated at two 1-ha created marshes at the ORWRP and compared with conditions observed in 1993 and 1995. Using these data, changes were analyzed over space and time. Finally, a separate study using the same two 1-ha wetlands was conducted to evaluate sediment and nutrient accumulation. Studies that have evaluated conditions in open wetland systems have shown that newly created wetlands (Fennessy et al. 1994, Braskerud 2001, Harter and Mitsch 2003) typically report much higher sedimentation rates than those studying older created wetlands (Craft et al. 2003) or natural wetlands (Johnston 1991, Peterjohn and Correll 1994, Craft and Casey 2000). Other research has demonstrated that sediment and nutrient accumulation can vary considerably within wetlands depending upon preferential flow and proximity to inflows (Reddy et al. 1993, Mitsch et al. 1995). In a conceptual model developed by Craft (1997), sediment and P accumulation in created estuarine wetlands is suggested to occur rapidly in the first few 5

years but as it accumulates, the rate of retention eventually peaks and declines, eventually becoming comparable to natural systems.

1.2 Literature cited Baker, D. B., R. P. Richards, T. F. Loftus, and J. W. Kramer. 2004. A new flashiness index: characteristics and applications to Midwestern rivers and streams. Journal of the American Water Resources Association 40:503-522. Bishel-Machung, L., R.P. Brooks, S.S. Yates, and K.L. Hoover. 1996. Soil properties of reference wetlands and wetland creation projects in Pennsylvania. Wetlands 16:532541. Braskerud, B. C. 2001. The influence of vegetation on sedimentation and resuspension of soil particles in small constructed wetlands. J. Environ. Qual. 30:1447-1457. Brooks, R. P., D. H. Wardrop, C. A. Cole, and D. A. Campbell. 2005. Are we purveyors of wetland homogeneity? A model of degradation and restoration to improve wetland mitigation performance. Ecological Engineering 24:331-340. Brown S. and D.L. Peterson. 1983. Structural characteristics and biomass production of two Illinois bottomland forests. American Midland Naturalist 110:107-117. Burke, M. K., B. G. Lockaby, and W. H. Conner. 1999. Aboveground production and nutrient circulation along a flooding gradient in a South Carolina Coastal Plain forest. Can. J. For. Res. 29:1402-1418. Campbell, D.A., C.A. Cole, and R.P. Brooks. 2002. A comparison of created and natural wetlands in Pennsylvania, USA. Wetlands Ecology and Management 10:41-49. Conner, W. H. 1994. Effect of forest management practices on southern forested wetland productivity. Wetlands 14: 27-40. Craft, C.B. 1997. Dynamics of nitrogen and phosphorus retention during wetland ecosystem succession. Wetlands Ecology and Management 4:177-187. Craft, C. B., P. Megonigal, S. Broome, J. Stevenson, R. Freese, J. Cornell, L. Zheng and J. Sacco. 2003. The pace of ecosystem development of constructed Spartina alterniflora marshes. Ecological Applications 13:1417-1432. Craft, C. B. and W. P. Casey. 2000. Sediment and nutrient accumulation in floodplain and depressional freshwater wetlands of Georgia, USA. Wetlands 20: 323-332. 6

Erwin, K. L. 1991. An evaluation of wetland mitigation in the South Florida Water Management District, Vol. I. Final Report to the South Florida Water Management District, West Palm Beach, FL, USA. Fennessy, M. S., C. C. Brueske, and W. J. Mitsch. 1994. Sediment deposition patterns in restored freshwater wetlands using sediment traps. Ecological Engineering 3:409428. Harter, S. K. and W. J. Mitsch. 2003. Patterns of short-term sedimentation in a freshwater created marsh. J. Environ. Qual. 32:325-334. Hein, T., G. Heiler, D. Pennetzdorfer, P. Riedler, M. Schagerl, and F. Schiemer. 1999. The Danube Restoration Project: Functional aspects and planktonic productivity in the floodplain system. Regulated Rivers 15: 259-279. Johnston, C. A. 1991. Sediment and nutrient retention by freshwater wetlands: effects on surface water quality. Critical Reviews in Environmental Control 21:491-565. Junk, W. J., P. B. Bayley, and R. E. Sparks. 1989. The flood pulse concept in riverfloodplain systems. In D. P. Dodge, ed. Proceedings of the International Large River Symposium. Special Issue of Journal of Canadian Fisheries and Aquatic Sciences 106:11-127. Kozlowski, T. T. 1997. Responses of woody plants to flooding and salinity. Tree Physiology Monograph 1:1-17. Megonigal, J.P., W.H. Conner, S. Kroeger and R.R. Sharitz. 1997. Aboveground production in southeastern floodplain forests: a test of the subsidy-stress hypothesis. Ecology 78:370-384. Mitsch, W.J. and K.C. Ewel. 1979. Comparative biomass and growth of cypress in Florida wetlands. American Midland Naturalist 101:417-426. Mitsch, W. J. and W. G. Rust. 1984. Tree growth responses to flooding in a bottomland forest in northeastern Illinois. Forest Science 30: 499-510. Mitsch, W. J., J. K. Cronk, X. Wu, and R. W. Nairn. 1995. Phosphorus retention in constructed freshwater riparian marshes. Ecological Applications 5: 830-845. Mitsch, W. J. and R. F. Wilson. 1996. Improving the success of wetland creation and restoration with know-how, time, and self-deign. Ecological Applications 6:77-83. Mitsch, W.J. and J.G. Gosselink. 2000. Wetlands, Third Edition. John Wiley & Sons, Inc., New York, NY.

7

Nair, V.D., D.A. Graetz, K.R. Reddy, and O.G. Olila. 2001. Soil development in phosphate-mined created wetlands of Florida, USA. Wetlands 21:232-239. Odum, E.P., J.T. Finn and E.H. Franz. 1979. Perturbation theory and the subsidy-stress gradient. Bioscience 29:344-352. Peterjohn, W. T. and D. L. Correll. 1984. Nutrient dynamics in an agricultural watershed; observations on the role of a riparian forest. Ecol. 65: 1466-1475. Reddy, K. R., R. D. DeLaune, W. F. DeBusk and M. S. Koch. 1993. Long-term nutrient accumulation rates in the Everglades. Soil Sci. Soc. Am. J. 57:1147-1155. Robertson, A.I., P.Y. Bacon, and G. Heagney. 2001. The response of floodplain primary production to flood frequency and timing. Journal of Applied Ecology 38:126-136. Shaffer, P.W. and T.L. Ernest. 1999. Distribution of soil organic matter in freshwater emergent/open water wetlands in the Portland, Oregon metropolitan area. Wetlands 19:505-516. Taylor, J.R., M.A. Cardamone and W.J. Mitsch. 1990. Bottomland hardwood forests: their function and values. p. 14-34. In J.G. Gosselink, L.C. Lee and T.A. Muir (eds.) Ecological processes and cumulative impacts illustrated by bottomland hardwood wetland ecosystems. Lewis, Chelsea, MI, USA. Tockner, K., F. Malard and J.V. Ward. 2000. An extension of the flood pulse concept. Hydrological Process 14:2861-2883. Zedler, J.B. and J.C. Callaway. 1999. Tracking wetland restoration: do mitigation sites follow desired trajectories? Restoration Ecol. 7:69-73.

8

Fig. 1.1 The Olentangy River Wetland Research Park (ORWRP) at The Ohio State University in Columbus, Ohio.

9

CHAPTER 2

EFFECT OF PULSING ON MACROPHYTE PRODUCTIVITY AND NUTRIENT UPTAKE: A WETLAND MESOCOSM EXPERIMENT

2.1 Abstract A study was conducted to evaluate the effect of a pulsing hydrology on the productivity and nutrient uptake of an herbaceous, riverine wetland. Pulsing effects were evaluated using twenty 0.9-m2 wetland mesocosms: 10 planted with Schoenoplectus tabernaemontani (C.C. Gmel) Palla and the other 10 planted with Typha angustifolia L. For each species, half the mesocosms were subjected to a 3-month pulsing regime while the others were subjected to steady-flow conditions. Hydrology parameters were selected to approximate a pulsing experiment being carried out concurrently at two 1-ha wetlands at the research site. Typha wetlands were significantly more productive than Schoenoplectus wetlands; however no significant differences in productivity or morphology were observed between pulsed or steady-flow wetlands among species groups. No significant differences in nutrient concentrations, uptake or uptake efficiency were detected among species groups either, however hydrology did influence plant tissue 10

N:P ratios. For all wetland mesocosms, the mean (±1 SE) N:P ratio was 9.2 ±0.6 for steady flow wetlands and 11.7 ±0.5 for pulsed, suggesting that the steady flow wetlands were more N limited than pulsed wetlands. The potential applications and limitations of applying these results to the 1-ha wetlands study are discussed.

2.2 Introduction Wetland plant productivity and nutrient uptake can be influenced by hydrologic parameters including flood depth (Waters and Shay 1992, Newman et al. 1998, Kellogg et al. 2003), flood duration (Newman et al., 1998) and flooding frequency (Giovannini and Da Motta Marques 1998, Tanner 1999, Casanova and Brock 2000). Further, water level fluctuations associated with a pulsing hydrology may increase wetland productivity and nutrient uptake (Mitsch and Ewel 1979, Mitsch et al. 1979, Junk et al. 1989, Odum et al. 1995, Day et al. 2000). Inflowing flood pulses often contribute higher than normal concentrations of nutrients thus providing a fertilizing effect for plant growth (Spink et al. 1998). Increased nutrient availability can also occur in wetland sediments by the release of phosphorus during anaerobic conditions and the increase in nutrient mineralization caused by the fluctuation of wet-dry soil conditions. However, prolonged wetland inundation increases the potential for anoxic conditions that can be detrimental to macrophyte vegetation (van der Valk and Davis 1978). Pulsing water regimes are characteristic of riparian wetlands adjacent to flashy streams or rivers in the Midwest United States (Baker et al. 2004). The temporary nature of river flood pulses often leads to rapidly dropping water levels, thus decreasing the chance for soil anoxia while increasing nutrient mineralization and the potential for 11

greater productivity (Mitsch and Rust 1984). The influence of a pulsing hydrology on increased productivity has been well demonstrated for wetland forests (Mitsch et al. 1991, Brown 1981) and planktonic communities (Hein et al. 1999), but has been less predictable for herbaceous wetlands. Because of shallow rooting, herbaceous wetland vegetation may be more sensitive to different flood conditions (depth, duration, etc.) that may make a pulsing effect more difficult to detect. The purpose of this study was to determine if a pulsing hydrologic regime would elicit morphological and functional responses by two common wetland plants in experimental mesocosms. This project was conducted in conjunction with a multi-year experiment using two 1-ha created wetlands to examine the effects of a pulsing regime. We hypothesized that productivity and nutrient uptake would be higher in the pulsed water regime, and that pulsing river water into wetlands would result in greater N uptake by vegetation compared to a steady flow regime.

2.3 Methods 2.3.1 Experimental design The project was conducted at the 12-ha Olentangy River Wetland Research Park (ORWRP) at The Ohio State University in Columbus, Ohio, USA (latitude N40.021◦, longitude E83.017◦). A total of twenty mesocosm tubs (Fig. 2.1; area=0.9 m2, volume= 540 liters) were used. In each mesocosm a total of 8-10 cm of pea-size gravel was placed in the bottom to allow seepage. Approximately 30 cm of on-site alluvial, upland soil (Ross and Eldean series, consisting of silt loam, silt clays and clay loams; Mcloda and Parkinson 1980) was placed on top of the gravel. Water levels were controlled with a 12

French drain system established for each mesocosm using 5-cm diameter pvc pipes (Fig. 2.1). The control structure (top of the stand pipe) was established 5 cm above the soil layer. To provide gradual water subsidence, a bleed down orifice (2 mm hole) was drilled into the side of each pipe at the mesocosm soil elevation. To evaluate the effect of pulsing, mesocosms were randomly assigned to one of four treatments based on macrophyte species and hydrology (Table 2.1). In March 2002, a total of 30 rhizomes each of Typha angustifolia L. (Cooperrider et al. 2001) (narrowleaved cattail, hereafter Typha) and S. tabernaemontani (C.C. Gmel) Palla (Cooperrider et al. 2001) (soft-stemmed bulrush, a.k.a. Scirpus validus, hereafter Schoenoplectus) from a Midwestern nursery were trimmed to approximately 25 ±5 g in size and planted 3 cm below the soil surface. Three rhizomes (with no inter-species mixing) were planted in each mesocosm. To expedite growth of the planted rhizomes, wetlands were kept in moist condition throughout the summer and autumn 2002 with groundwater equally distributed to all mesocosms. At the end of August 2002, initial condition aboveground biomass was non-destructively estimated in each mesocosm by relating various morphological parameters to reference plants. For Schoenoplectus, stem density, mean stem height, number of flowers, and mean maximum stem height were recorded. For Typha, the number of leaves per ramet, ramet density, mean ramet height, number of flower spikes, and mean maximum ramet height were recorded. Over the winter of 20022003, the mesocosms were not watered; a layer of snow and ice covered them throughout most of the season. In March 2003 it was discovered that muskrats (Ondatra zibethicus) excavated 3 Schoenoplectus wetlands of nearly all its rhizomes during winter 2002-03. In April 2003, 13

sod containing Schoenoplectus rhizomes were collected from control wetlands used in another mesocosm experiment, and transplanted into the disturbed mesocosms. Each disturbed mesocosm received five pieces of Schoenoplectus sod that were approximately 15 x 20 cm wide and 4 cm thick. All measurements from these mesocosms were analyzed separately and designated as ‘Schoenoplectus (1-yr)’ wetlands, compared to ‘Schoenoplectus (2-yr)’ for the undisturbed wetlands (Table 2.1). During the experimental hydrology period (April through June 2003), water was pumped from the Olentangy River based on the pulsing schedule in Table 2.1. By design, all mesocosms received a near equivalent monthly hydrologic load of 160 cm month-1. Water was pumped directly from the Olentangy River to the mesocosms via a lowpressure pump and an elevated reservoir tank system. A garden hose was used for each water regime and extended along each mesocosm. At each nexus point, the hose was attached to the mesocosm rim and a volume-adjustable irrigation sprayer (Raindrip® R180C) was used to control the amount of water pumped into each mesocosm. In midMay 2003, the pumping schedule was postponed for 11 days to repair the pumping system. Pumping volumes and water depths were recorded for each water regime. Three mesocosms were omitted from the study in 2003 because of faulty drainage systems that developed and could not be repaired. During the experimental hydrology period, a 250mL sample of the river water was collected weekly and analyzed for NO3-N and total P using a Lachat QuikChem IV automated system and Lachat methods (U.S. EPA 1983). Total P was analyzed using the ascorbic acid and molybdate color reagent methods after digesting with 0.5 ml of 5.6N H2SO4 and 0.2 g NH3SO4 to 25ml of sample and exposing 14

the samples to a heated and pressurized environment for 30 minutes in an autoclave. Using the Lachat automated system, nitrate was analyzed using the cadmium reduction method. Nutrient concentrations were used with pumping rates to estimate weekly NO3-N and total P nutrient loads into the mesocosms (converted to g m-2 for comparison). Plant morphology was measured on 18 June, 23 July and 28 August 2003 using the same parameters measured in September 2002. After the August 2003 measurements, all aboveground vegetation in each mesocosm was harvested and the mesocoms were covered until belowground biomass was harvested. Harvested aboveground material was air-dried and a subsample was oven dried for 3 d (or until constant mass was achieved) at 80◦ C. Total aboveground biomass was calculated and converted to g m-2 for comparison. In September 2003, soil was extracted from each mesocosm and carefully washed from all root material. Root material was weighed for each mesocosm and a subsample was air dried for 5 d at 105◦ C or until constant mass was achieved. Total belowground biomass was estimated for each mesocosm and converted to g m-2 for comparison. Belowground biomass data was combined with aboveground data in August 2003 to calculate total biomass and the mean root:shoot ratio. To estimate tissue nutrient concentrations for each mesocosm, 10 mature stems were randomly selected and analyzed for nutrient content. Tissue specimens were air dried, ground to pass through a 2 mm sieve, and mixed to make a homogenous sample. In mesocosms where significant plant senescence occurred (>50% of the plant surface yellowed), 10 random senescent stems were analyzed separately from the living stems and because of the potential for nutrient translocation, the belowground tissue was analyzed separately by randomly selecting five 5-cm root/rhizome sections. All tissue 15

specimens were sent to Service Testing and Research (STAR) Laboratory, Ohio Agricultural Research and Development Center, Wooster, Ohio and analyzed for total N and P. Samples were digested with HClO4/HNO3 and analyzed for total P by inductively coupled plasma emission spectrometry (Isaac and Johnson 1985). Samples were analyzed for total N through combustion analysis (AOAC 1989). Using mean nutrient concentrations, the mean N:P ratio of the aboveground tissue was calculated for all treatment types and compared to thresholds developed by Koerselman and Meuleman (1996) to detect possible community level N or P limitations. As an indication of site fertility, tissue nutrient concentrations were used with peak biomass measurements in August 2003 to calculate total nutrient uptake for each wetland.

2.3.2 Statistical analyses For each species group [Typha, Schoenoplectus (2-yr) and Schoenoplectus (1-yr)], a two-way, repeated measure analysis of variance (ANOVA) was used to examine for differences in mean macrophyte morphological measurements between pulsed and steady-flow wetlands. An independent t-test was used to compare mean productivity, nutrient uptake and nutrient loads (mg total P and NO3-N wk-1) between the pulsed and steady-flow mesocosms. In the case of root:shoot ratios and leaf-tissue N and P concentrations, Schoenoplectus (2-yr) and (1-yr) data were comparable (P>0.05, t-test) and therefore pooled. A two-way ANOVA was conducted to compare the factors of hydrology (pulsed and steady flow wetlands), species, and species x hydrology interaction. Data were analyzed and transformed when necessary to meet assumptions for parametric statistics. For all comparative tests, P-values 5cm were identified by species, tagged and measured for dbh in April 2004 and April 2005 to determine 1-yr basal increase. Using tree data, species importance values were calculated in 2004 using the following equation:

Importance value = relative density + relative dominance + relative frequency

(1)

The increase in tree basal area (Ai) (cm2 yr-1) was calculated by the following equation (Newbould 1967):

Ai = π [r2-(r-i)2]

(2)

Where, r = radius of tree at breast height (cm), and i = radial increment per year (cm2 yr-1)

Tree heights were measured using a clinometer in May 2005 and the annual wood production per tree (Pi)(g yr-1) was calculated by the following parabolic volume equation (Whittaker and Woodwell 1968, Phipps 1979):

Pi = 0.5ρ Ai h

Where, ρ = wood specific gravity (g cm-3), and h = tree height (m)

44

(3)

Wood specific gravity values were obtained from the U. S. Forest Products Laboratory (1974) and Alden (1995). The plot wood production was calculated as the summation of all wood production per tree and converted to g m-2 yr-1. A total of 50 leaf litter traps (5 per plot) were installed in May 2004. Each plot was divided into 4 quadrants and a leaf trap was randomly placed in each quadrant with a fifth trap randomly placed near the center (Fig. 3.2). Leaf traps were 15 cm tall, 0.25 m2 in area, lined with 2-mm screen and installed approximately 1.0 m off the ground to avoid flooding and litter saturation. Litterfall was collected for one year starting in May 2004. Traps were emptied twice a month from June-December and once a month from January-May. After each collection, the contents were separated into leaves, reproductive material and woody material, air-dried at room temperature for 1 week and then at 105ûC for four days or until constant mass prior to being weighed. Leaf traps were averaged per plot and the summation of all fine litter production (leaf litter and reproductive materials) was calculated. Because of vandalism and flood/ice damage, several sampling periods had plots with less than the 5 traps available and were averaged only using the plots that were undamaged. Using litterfall and wood production data, aboveground net primary productivity (ANPP) (g m-2 yr-1) for each plot was estimated using the following equation (Whittaker and Woodwell 1968):

ANPP = plot wood production + litterfall production

45

(4)

3.3.4 Predicting ANPP, litterfall production and wood production Various environmental parameters known to influence forest productivity were selected to predict forest productivity in 2004 (ANPP, wood production and litterfall) through linear regression. The 2004 river hydrograph and observations of flooded conditions at different river stages were used to determine 1) the number of flood events that directly connected to each plot, and 2) the number of days that the river had a surface water connection to each plot. Flooding frequencies and durations in 2004 for each plot were estimated for the preceding year (October 2003-September 2004), preceding two years (October 2002Sepetmebr 2004) and the growing season (April-September 2004) and used for regression analyses. To assess the potential influence of tree plot elevation on ANPP, the corners of each plot and each random leaf litter trap within the plot quadrants (Fig. 3.2) were surveyed for elevation using a TOPCON RL-H3CTM rotating laser level and the mean plot elevation (m MSL) was calculated. To assess the potential influence of topographic variability on forest productivity, the variance of all elevation points at each plot was also calculated and used to predict forest productivity. Other data used as predictor variables included canopy cover and tree basal area. Canopy cover (%) was estimated for each plot in August 2004 using a convex spherical crown densitometer. Cover was measured at each trap facing the four cardinal directions and the mean of all measurements were calculated for the entire plot. Tree basal area (cm2 m-2) per plot was calculated based on the total basal area of all trees >5 cm dbh measured in April 2004.

46

3.3.5 Tree-ring analysis For each forest canopy tree (>25 cm dbh and >15 m height) in the plots, two cores were extracted using a 5.15 mm inside increment borer. Seven supplemental trees located between tree plots (5 in the north section and 2 in the south section) were added. For comparison with trees not in the flood zone (the upland area between sections) a total 7 trees from species representative of the flooded sections were randomly selected for coring. For each tree, two cores were taken at 90° angles from each other to account for natural variation and were collected at least 12 cm into the tree to collect >15 years of increment growth. The cores were temporarily stored in straws, air-dried and then glued into grooved holders. Cores were sanded with a series of finer sandpaper grit (80-600) and polished with lamb’s wool. Tree cores were scanned and the image was analyzed for tree-ring widths (to the nearest 0.01 mm) using WinDENDRO TM (2002). Replicate tree-ring increments were compared for comparable growth patterns, verified with a stereoscope when necessary and averaged for each tree. Using the tree cores and tree diameter, basal area (Ai) increments (BAI) were calculated from 1991 to 2004. Years 1991-2000 were selected as representative pre-restoration growth and years 2001-2004 were analyzed as post-restoration years. Although most restoration work was conducted in June 2000, each cut was excavated further in early 2001. The first flood event to overflow into bottomland forest did not occur until April 2001. For comparison of trees between sections and species, the BAI (cm2 yr-1) from each tree were standardized to reflect percent basal increase relative to total tree basal area [BAI (%)], and were calculated using the following equation:

47

BAI(%) = [(Ai Year X – Ai Year X-1)/Ai Year X-1] * 100

(5)

3.3.6 Predicting basal growth The series of BAI data collected for 1991-2004 (both cm2 and %) were evaluated to determine if flood stage (based on river discharge) could be used to predict basal growth. A discharge curve prepared for river depth at this section of the Olentangy River (Mitsch 1995) was used to determine the number of bank-full flood days/events (221.2 mMSL or >70 m3 sec-1) and the number high-flood days/events (221.6 mMSL or >154 m3 sec-1) between March 1994 and September 2004. The high-flood discharge was selected because this is the discharge level that was likely required to directly flood both sections of the bottomland despite the presence of the levee. Daily river discharge data from an upstream United States Geological Survey (USGS) stream gauge (near Delaware, Ohio, Station No. 03225500) was used to estimate discharge rates at the study site between October 1990 and March 1994. A regression between known ORWRP and USGS discharge rates were used to estimate discharge on the days where no water level data was available at the study site (ORWRP= 1.43*USGS + 5.34, R2=0.92). Using daily river discharge data, the frequency and duration of flood events were determined for each year (from 1 October in the preceding year to 30 September) and growing season (1 April to 30 September). Frequency and duration were determined for bank-full and high-flood discharge events. For both thresholds, the number of days and events in which these rates occurred were counted for each applicable year and growing season.

48

In addition to conducting a regression analysis on the concurrent flood and BAI data for a given year, regressions were also conducted to evaluate the possibility of a lag in tree basal growth response to floods. A regression analysis was used to evaluate flood frequency and BAI data lumped into 2-yr increments. In addition to capturing potential lag effects, lumping BAI data in this manner has been suggested as an effective guard against potential errors due to false-rings or other measurement errors (Mitsch et al. 1991). A second regression analysis was conducted using two years of preceding river discharge data to predict the BAI for a given year.

3.3.7 Statistical analyses An independent t-test was conducted to compare the mean ANPP between the north and south sections of the bottomland. Because litterfall and wood production have been shown to respond independently to environmental factors, independent t-tests were also conducted to compare these parameters. Analyzing tree-ring data, paired t-tests and trend analyses were used to compare BAI (%) between pre- and post-restoration years for each section. Similarly, paired t-tests and trend analyses were used to compare BAI (%) for pre- and post restoration specimens of A. negundo and A. glabra. All pre- and post-restoration data were tested for normality using the Kolmogrov-Smirnov test, homogeneity of variances using Levene’s test, and transformed as needed to meet test assumptions. For all t-tests, p-values 75 cm dbh) had consistently low BAI (%) values that had an excessive influence on mean comparisons and trend analyses, and were therefore omitted. We presumed that these older trees had reached an age where a high proportion of gross production is used for maintenance metabolism (Kimmins 1987) and were unlikely to provide a growth response to changing moisture conditions. Evaluation of trend analyses showed that none of the sections had an abrupt shift in basal growth immediately after hydrologic restoration (Fig. 3.9). However, canopy trees in the south section showed increased radial growth in 2003 and 2004 compared to a trend of consistent decline in BAI(%) since 1994. Trees in the north and upland sections showed a slight increase of BAI(%) in 2004, however conditions in both sections during pre-restoration years were more variable making this shift difficult to assess. Two of the most dominant trees in the bottomland forest (A. negundo and A. glabra) were evaluated separately to see if responses between species were different. Because similar trends were detected in A. negundo between the north and south sections, these trees were pooled. No canopy-sized A. glabra occurred in the north section plots. Like trees in the north and south sections, A. negundo had a significantly lower mean BAI (%) after the restoration (P>0.05) but with no significant change in BAI (cm2 yr-1) (Table 3.4). Trend analyses indicated that unlike canopy specimens in the upland section, A. negundo trees in the flooded sections may have responded positively to the restoration based on the increased BAI (%) in 2003 and 2004 (Fig. 3.10a). A. negundo trees in the upland sections seemed to follow a basal growth trend that extended back to 1991. 54

The BAI (% and cm2 yr-1) of A. glabra canopy trees were not significantly different between pre- and post-restoration years (Table 3.4). Upland and flooded specimens had similar BAI (%) extending back to 1991 (Fig. 3.10b). After 2000, there was a separation between the upland and flooded trees, however BAI trajectories did not shift substantially in the post-restoration period.

3.4.6 Predicting basal growth A significant relationship was detected between the total number of days where the river discharged at high-flood stage (>154 m3sec-1) and BAI (cm2 yr-1) when analyzed using the preceding 2-yr river data (Fig. 3.11a). Similarly, a significant relationship was detected between the number of high-flood days over a 2-yr period and the corresponding 2-yr BAI (cm2 yr-1) (Fig. 3.11b). No significant relationships were detected between the number of days or events of discharge and the BAI for that corresponding single year. A significant relationship between the total number of high-flood discharge events and 2-yr preceding river data was also detected (R2=0.54, F=13.93, P=0.003), but the number of events was less predictive than the number of days. No other significant relationships between BAI and river discharge were detected. In all cases, BAI (%) data showed indications of autocorrelation and therefore were omitted from consideration in favor of BAI (cm2 yr-1).

3.5 Discussion 3.5.1 Bottomland productivity One of the most commonly cited benefits associated with the hydrologic restoration of a bottomland forest is the likely enhancement in productivity. Based on the results of this 55

study there is some evidence to suggest that after only four years, there was an increase in bottomland productivity. In terms of ANPP, we found no significant differences between the north (807 ±86 g m-2 yr-1) and south (869 ±56 g m-2 yr-1) sections. This was important because using productivity data from plots comparable to our study, Cochran (2001) found that ANPP in the north section (531-641 g m-2 yr-1) was significantly lower than the south (793-1033 g m-2 yr-1). This suggests that the north section has increased in productivity since the restoration activity occurred. The biggest difference between pre- and post-restoration productivity in the north section was in mean wood productivity which, in 2004 (346 g m-2 yr-1, this study) was nearly triple that estimated in 2000 (117 g m-2 yr-1, Cochran 2001). However, the change in wood productivity conflicts somewhat with our tree canopy ringanalysis data which saw relatively consistent basal area growth (cm2 yr-1) between pre- and post-restoration years in the north section. Given the high variability of wood production estimated for plots in the north section, plot location may have greatly influenced estimates in both studies. Cochran (2001) only used 2 plots (20 x 25m) in the north section directly affected by the levee, compared to 4 plots used in this study. Therefore we conclude only tentatively that ANPP has increased in the north section. Based on estimates by Cochran (2001), ANPP in the north section had clearly exceeded its pre-restoration range while in the south section ANPP was still within the pre-restoration range. Furthermore, the ANPP range seen at the bottomland forest was still below what has been recorded at other sections of the Olentangy River. At two other unrestricted bottomland hardwood forests upriver from the ORWRP (both within 12 km), forest ANPP was estimated at 1283 ±56 and 1297 ±302 g m-2 yr-1 (Cochran 2001). The ANPP range seen at the ORWRP bottomland also seems to be lower than what has been observed at most other bottomland 56

forests in the region. Mitsch et al. (1991) found ANPP at 1280 and 1334 g m-2 yr-1 in two hardwood bottomland forests along the Ohio River in western Kentucky. ANPP for a floodplain forest in Illinois was estimated at 1250 g m-2 yr-1 (Johnson and Bell 1976). However, Brown and Peterson (1983) found that ANPP at another bottomland forest in Illinois with stagnant water conditions was 960 g m-2 yr-1 while a seasonally flooded forest was at 668 g m-2 yr-1. It seems that in terms of long-term productivity, the ORWRP bottomland may still have an opportunity to increase. Although leaf productivity was significantly higher at the south section, it appeared that wood production was more the responsive component affecting ANPP based on the wide ranges observed at the ORWRP (Fig. 3.5). This in contrast to other studies (Burke et al. 1999) which found leaf production to be more variable. Part of the reason that litterfall was more consistent between plots may have been the frequent occurrence of paw paw (A. triloba) in north section plots. Although these plots had less canopy-tree cover and overall basal area, there was a large contribution of litterfall provided by subcanopy A. triloba which produced a dense cover of large leaves.

3.5.2 Relationship between bottomland productivity and flooding Although plot ANPP was predicted by the number of days each was flooded in 2004, the best relationships were found using flood data added from 2003 and 2004. The results of these analyses confirmed that surface water flooding was an important factor in determining forest productivity and also suggests that flood events may influence productivity beyond the year they occur. Similar patterns were revealed using river discharge to predict basal tree growth (see Section 3.5.4 below). It is possible that this delayed response represents the 57

time it takes for deposited nutrients to desorb from sediment and mineralize from matter and become available. The decomposition of organic matter, the desorption of nutrients from sediment and the alteration of soil chemistry are all factors that dictate nutrient availability in bottomland soils (Mitsch and Gosselink 2000). The rates of these processes are eventually dependent upon environmental conditions including hydrology and climate. Therefore if it takes several months for ecological processes to make nutrients available, nutrients from material deposited in the spring and early summer (when most flooding traditionally happens) may not become available to plants until the subsequent growing season. Using regression analyses, total ANPP and wood production were significantly influenced by total basal area and topographic variability (elevation variance). It was no surprise that existing basal area influenced productivity however elevation variance was one of the least considered predictor variables at the onset of this study. Floodplain bottomlands can have naturally diverse topographies consisting of repeated ridges, swales and meandering scrolls (Leopold et al. 1964). The influence of topography has been demonstrated on forest productivity in the southern Appalachian (Bolstad et al. 2001), on riparian plant diversity in Alaska (Pollock et al. 1998) and canopy gap regimes in a Texas bottomland forest (Almquist et al. 2002), however there is little information pertaining to its influence on bottomland tree productivity. A diverse topography such as that of a ridge-and-swale would perhaps allow the greatest interface between flood waters and trees on slightly elevated ground. In the case of the ORWRP bottomland, topographic variability was provided by swales and ridges in the south section, however in the north section it was provided by the old fill material from the remnant levee. The influence of topography on bottomland productivity is an interesting

58

result from this study and we would encourage future bottomland research to consider this component.

3.5.3 Forest basal growth before and after restoration Evaluating canopy tree cores, we did not find an occasion where radial tree growth made an immediate and clear response to the restored hydrology. Given that the north section was a more complete restoration (hydrology was only enhanced in the south section) we were expecting to see a positive response to the restored hydrology. However, compared to the other sections, only the south section showed a potential response. The change in BAI (%) seen at the south section in 2003 and 2004 was interpreted to be a more significant shift because it represented a clear break in a very consistent growth trend dating back to 1994. An increase was detected in the north section in 2004, however given the modest size of the increase, the more sporadic growth trend leading up to it, and that upland trees showed a similar increase; this change cannot be considered conclusive. It may be that because the south section trees were exposed to occasional flooding prior to the restoration work, trees in this section were better conditioned to the altered hydrology. Assuming that the increased flooding has been a stress to trees, when stressors are introduced more gradually, trees can generally make the physiological adjustments to protect themselves much more than if the stressor is introduced rapidly (Kozlowski and Pallardy 2002). The canopy tree response in the south section may have been in response to the high inflows that occurred in 2003 and 2004, or perhaps more likely, it may be a lag response to the new hydrology. This would not be unprecedented, as lags in forest response have been documented in the case of other habitat improvements. Jones and Thomas (2004) found that in Ontario forest stands 59

dominated by sugar maple (Acer saccharum Marsh), peak growth enhancement in response to canopy gaps did not occur until 3-5 years later. Given the shift in hydrology is even more substantial in the north section it may take longer for trees there to positively respond. Anaerobic conditions caused by flooding may have been exacerbated in this section where flooding was previously rare. A. negundo was the dominant tree in the south section and therefore its trend in BAI (%) over time (Fig. 3.10a) was similar to that seen for all south section trees (Fig. 3.9). However, A. negundo specimens tended to respond similarly in the north section as well. The response of trees in the flooded sections since the restoration is in contrast to upland specimens where BAI (%) maintained the same trend set before the restoration occurred. The physiology of A. negundo may make it well adapted to changing water conditions as it has been shown that its net photosynthesis can be resilient to seasonal changes in soil water potential (Foster 1992). A. glabra on the other hand did not show a substantial response although its BAI (%) has not declined during the post-restoration period as the upland specimens have. Nevertheless, this tree tends to occurs in moist soils and while it is considered resistant to saturation, it is a facultative upland species and might be less resilient to prolonged anaerobic conditions.

3.5.4 Basal growth in response to flooding Based on the results of this study, there is evidence that flooding may have a lagged effect on tree growth. In both scenarios where river discharges from the current and previous years were added, there was a significant relationship between the number of days with high-flood discharge and BAI (cm2 yr-1). Given the pre-restoration exclusion of bank-full flood waters it is not surprising a relationship was only detected using the high-flood events, and as 60

indicated in Figs. 3.11a and b, the bottomland forest was still responding to these high-flood occurrences during the post-restoration years. The evidence of a lagged response by bottomland canopy trees to flooding has been rarely documented however it isn’t unexpected given the amount of other circumstances where forests have shown a lagged response in growth. Factors such as climate (Fritts 1976, Camill and Clark 2000), newly formed canopy gaps (Jones and Thomas 2004) and the removal of shelterwoods (Holgen et al. 2003) have all been shown to induce a lagged response on tree basal growth. Significant regressions using current- and previous-year river discharge data indicated that basal tree growth occurred at an optimal number of high-water discharge days (~10) suggesting that trees are benefiting from a nutrient subsidy to a point. After about 10 highflood discharge days, the bottomland may no longer be nutrient limited and anaerobic conditions may have reduced productivity. It is important to point out that in the 2-yr periods where high-flood discharge exceeded 10 days, the decrease in basal growth was marginal compared to those years where floods events were scarcer. The general relationship seen in this case is not unprecedented. Golet et al. (1993) showed that the highest tree basal growth at red maple (Acer rubrum L.) swamps in Rhode Island occurred at intermediate annual water levels. The results from this study support findings such as these and demonstrate the pushpull influence that flooding has on forest productivity. The fact that flooding throughout entire years (and not just the growing seasons) was the best predictor of basal growth supports the idea that these trees were responding more to a nutrient subsidy and less to the anaerobic stress of flooding. If flooding stress was more important, we would have expected a relationship with BAI to manifest during the growing season. However, as seen in other studies, it is likely that the anaerobic stress caused by 61

flooding in the growing season was negated by a nutrient subsidy, and therefore a relationship between growing season flood occurrence and BAI was unapparent. Furthermore, it appears that trees are responding to sediment and nutrient deposition occurring year-round. Through the work of Zhang et al. (2005) and personal observations, it has been shown that these flood events can deposit significant amounts of material into the bottomland forest and the amount of material, sediment and nutrients available to trees may ultimately be dependent upon the frequency of major flood events in the preceding years.

3.6 Conclusions Hydrologic restoration of the ORWRP bottomland forest was conducted in 2000 and as a result, the north section has received direct surface flows from river floods and the south section has increased its surface flow and frequency. The two sections were similar in ANPP, but compared to previous estimates conducted before the restoration, the north section has increased its mean ANPP since the restoration occurred. No abrupt and clear changes in canopy tree basal growth has occurred since the restoration occurred, however since 2003, trees in the south section of the bottomland have shifted from a continuous trend of declining BAI (%) extending back about ten years. Evaluating BAI and river discharge data since 1991, these results suggest that for a two-yr period, optimal basal growth will occur when ~10 days of high-flood discharge occur during that period. These results also suggest that basal growth in response to flooding is lagged by at least one year as no relationships were detected between tree basal growth and concurrent flooding over one year. The lack of any significant relationship between tree basal growth and flooding in the growing season

62

suggests that sediment and nutrient deposition are likely more important to forest productivity than the stress caused through flooding.

3.7 Acknowledgements This study was partially funded through a contract with the Ohio Department of Transportation. Field assistance was provided by Jeremiah Miller. Editorial comments by Charles Goebel improved this chapter.

3.8 Literature cited Alden, H. A. 1995. Hardwoods of North America. U. S. Department of Agriculture, Forest Service, Madison, WI, USA. FPL-GTR-83. Almquist, E. B., S. B. Jack and M. G. Messina. 2002. Variation of the treefall gap regime in a bottomland hardwood forest: relationships with microtopograpy. Forest Ecology and Management 157:153-163. Baker, III, T. T., W. H. Conner, B. G. Lockaby, J. A. Stuart, and M. K. Burke. 2001. Fine root productivity and dynamics on a forested floodplain in South Carolina. Soil Science Society of America Journal 65:545-556. Bolstad, P. V., J. M. Vose, and S. G. McNulty. 2001. Forest productivity, leaf area, and terrain in southern Appalachian deciduous forests. Forest Science 47:419-427. Brown S. and D.L. Peterson. 1983. Structural characteristics and biomass production of two Illinois bottomland forests. American Midland Naturalist 110:107-117. Burke, M. K., B. G. Lockaby, and W. H. Conner. 1999. Aboveground production and nutrient circulation along a flooding gradient in a South Carolina Coastal Plain forest. Can. J. For. Res. 29:1402-1418. Camill, P. and J. S. Clark. 2000. Long-term perspectives on lagged ecosystem responses to climate change: permafrost in boreal peatlands and the grassland/woodland boundary. Ecosystems 3:534-544.

63

Cochran, M. 2001. Effect of Hydrology on bottomland hardwood forest productivity in central Ohio (USA). M.S. Thesis. The Ohio State University, Columbus, OH, USA. Conner, W. H. 1984. Effect of forest management practices on southern forested wetland productivity. Wetlands 14: 27-40. Conner, W. H. and J. W. Day, Jr. 1982. The ecology offorested wetlands in the southeastern United States. p. 69-87. In B. Gopal, R. E. Turner, R. G. Wetzel, and D. F. Whigham, (eds) Wetlands: ecology and management. National Institute of Ecology and International Scientific Publications, Jaipur, India. Dudek, D.M., J.R. McClenahen and W.J. Mitsch. 1998. Tree growth responses of populus deltoides and Juglans nigra to streamflow and climate in a bottomland hardwood forest in central Ohio. American Midland Naturalist 140:233-244. Foster, J. R. 1992. Photosynthesis and water relations of the floodplain tree, boxelder (Acer negundo L.). Tree Physiol. 11:133-149. Golet, F. C., A. J. K. Calhoun, W. R. DeRagon, D. J. Lowry, and A. J. Gold. 1993. Ecology of red maple swamps in the glaciated northeast: a community profile. Biological Report 12, U.S. Fish & Wildlife Service, Washington, DC. 151 pp. Holgen P., U. Soderberg, and B. Hanell. 2003. Diameter increment in Picea abies shelterwood stands in northern Sweden. Journal of Forest Research 18:163-167. Johnson, F. L. and D. T. Bell. 1976. Tree growth and mortality in the streamside forest. Castanea 41:34-41. Jones, T. A. and S. C. Thomas. 2004. The time course of diameter increment to selection harvests in Acer saccharum. Can. J. Res./Rev. Can. Rech. For. 34:1525-1533. Kimmins, J. P. 1987. Forest ecology. Macmillan Publishing Company. New York, NY, USA. Kozlowski, T. T. 1997. Responses of woody plants to flooding and salinity. Tree Physiology Monograph 1:1-17. Kozlowski, T. T. and S. G. Pallardy. 2002. Acclimation and adaptive responses of woody plants to environmental stresses. The Botanical Review 68:270-334. Leopold, L. B., M. G. Wolman and J. E. Miller. 1964. Fluvial processes in geomorphology. W. H. Freeman, San Francisco, CA, USA. Martens, D. M. 1993. Hydrologic inferences from tree-ring studies on the Hawkesbury River, Sydney, Australia. Geomorphology 8:147-164. 64

Megonigal, J.P., W.H. Conner, S. Kroeger and R.R. Sharitz. 1997. Aboveground production in southeastern floodplain forests: a test of the subsidy-stress hypothesis. Ecology 78:370-384. Mitsch, W. J. and K. C. Ewel. 1979. Comparative biomass and growth of cypress in Florida wetlands. American Midland Naturalist 101:417-426. Mitsch, W. J. and W. G. Rust. 1984. Tree growth responses to flooding in a bottomland forest in northeastern Illinois. Forest Science 30: 499-510. Mitsch, W.J., J.R. Taylor and K.B. Benson. 1991. Estimating primary productivity of forested wetland communities in different hydrologic landscapes. Landscape Ecology 5:75-92. Mitsch, W.J. and J.G. Gosselink. 2000. Wetlands, third edition. John Wiley & Sons, Inc., New York, NY, USA. Mitsch, W. J. and L. Zhang. 2004. Wetland monitoring of the bottomland hardwood forest at the Olentangy River Wetland Research Park (Year 3 – 2003). p.137-147. In W.J. Mitsch, L. Zhang and C. Tuttle (eds.) Olentangy River Wetland Research Park at The Ohio State University, Annual Report 2003. Columbus, OH, USA. Newbould, J. 1978. Methods for estimating the primary production of forests. Blackwell, Oxford, England. Odum, E.P., J.T. Finn and E.H. Franz. 1979. Perturbation theory and the subsidy-stress gradient. Bioscience 29:344-352. Phipps, R.L. 1979. Simulation of wetlands forest vegetation dynamics. Ecological Modelling 7:257-288. Pollock, M. M., R. J. Naiman, and T. A. Hanley. 1998. Plant species richness in riparian wetlands- a test of biodiversity theory. Ecology 79:94-105. Regent Instruments, Inc. 2002. WinDENDRO 2002 a,b. Regent Instruments, Inc., Quebec, Canada. Robertson, A.I., P.Y. Bacon, and G. Heagney. 2001. The response of floodplain primary production to flood frequency and timing. Journal of Applied Ecology 38:126-136. Taylor, J.R., M.A. Cardamone and W.J. Mitsch. 1990. Bottomland hardwood forests: their function and values. p. 14-34. In J.G. Gosselink, L.C. Lee and T.A. Muir (eds.) Ecological processes and cumulative impacts illustrated by bottomland hardwood wetland ecosystems. Lewis, Chelsea, MI, USA.

65

Teskey, R. O. and T. M. Hinckley. 1977a. Impact of water level changes on woody riparian and wetland communities, Vol. I: plant and soil response. U. S. Fish and Wildlife Service, Columbia, MO, USA. FWS/OBS-77/58. Teskey, R. O. and T. M. Hinckley. 1977b. Impact of water level changes on woody riparian and wetland communities, Vol. III: the central forest region. U. S. Fish and Wildlife Service, Columbia, MO, USA. FWS/OBS-77/60. Tockner, K., F. Malard and J.V. Ward. 2000. An extension of the flood pulse concept. Hydrological Process 14:2861-2883. U.S. Forests Products Laboratory. 1974. Wood handbook: wood as an engineering material. USDA Agriculture Handbook No. 72. Washington, D.C., USA. Whittaker, R.H. and G.M. Woodwell. 1968. Dimension and production relations of trees and shrubs in the Brookhaven Forest, New York. Journal of Ecology 57:155-174. Zhang, L., W. J. Mitsch, V. Bouchard and K. Hossler. 2005. Sediment chemistry in a hydrologically restored bottomland hardwood forest in Midwestern U.S. Program and abstracts, restoration and design of ecosystems, fifth annual meeting, American Ecological Engineering Society, The Ohio State University, Columbus, OH, USA.

66

Cut #1 Cut #2

Cut #3

Clinton Park weir staff gauge station

Levee

Existing Wetlands at the Olentangy River Wetland Research Park

Cut #4

LEGEND

Elevation (m MSL) < 221.0 221.0 - 221.3 221.3 - 221.6 221.6 - 221.9 > 221.9

N

Total

Area 0.5 ha 0.6 ha 1.3 ha 1.4 ha 1.4 ha 5.2 ha

Clinton Park weir gage elevations normal river pool 220.6 (m MSL) average river level 220.8 (m MSL) 0

50

100

meter source: Dodson - Lindblom survey (1987) drawing by: N. Wang, ORWRP, OSU modified by C. Anderson , ORWRP, OSU (2005)

Figure 3.1. Map of the bottomland forest at the Olentangy River Wetland Research Park (ORWRP) at The Ohio State University in Columbus, Ohio, USA indicating site topography and levee breeches (Mitsch and Zhang 2004). Hydrologic restoration was conducted by breaching a levee (Cuts #1-3) along the north section and breaching the river bank at the south section (Cut #4).

67

Figure 3.2. Experimental layout at the ORWRP bottomland hardwood forest indicating the location and dimensions of tree plots and litter traps. Each tree plot was divided into four quadrants (NW, NE, SW, and SE) for placement of random litter traps including a fifth trap near the plot center.

68

69

0

10

20

30

40

50

60

70

80

1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004

Actual Normal

69

Figure 3.3. Quarterly-annual normal and recorded precipitation totals for Columbus, Ohio based on data collected from the Ohio Agriculture and Development Center weather station (www.oardc.ohiostate.edu/centernet/weather.htm). Precipitation totals reported for January-March, April-June, JulySeptember and October-December of 1991-2004.

Precipitation (cm)

Olentangy River water level (m MSL)

222.0

221.6

221.2

220.8

220.4 Jan-00 Jul-00 Jan-01 Jul-01 Jan-02 Jul-02 Jan-03 Jul-03 Jan-04 Jul-04 Jan-05

Figure 3.4. Hydrograph of river water levels (m above MSL) for the Olentangy River for 2001-2004 based on data collected at the Olentangy River Wetland Research Park (Mitsch and Zhang 2004).

70

a) Litterfall Production

ANPP (g m-2 yr-1)

1200

Wood Production

1000 800 600 400 200 0 1

2

3

4

6

North

7

8

9

10

South

5 Upl

Plot Number and Section

b) Litterfall Production

1200

Wood Production ANPP (g m-2 yr-1)

1000 800 600 400 200 0 North

South

Figure 3.5. Aboveground net primary productivity (ANPP), including litter-fall and wood production for a) tree plots in the north, south and upland sections, and b) mean (±1 SE) for north and south section plots for 2004-05. Error bars for the section means represent standard error for ANPP.

71

-2

-1

Total ANPP (g m yr )

1100 1000 900 800 700 Total ANPP = 607 + 6*No. days flooded 2 R =0.48, P=0.04

600 500 0

20

40

60

80

Total no. of days flooded (2003-2004)

Figure 3.6. Linear relationship between the number of days flooded (2003-2004) and aboveground net primary productivity for experimental plots in 2004.

72

-1 -2

Wood production (g m yr )

600 500

a)

400 300 200 Wood prod. = 591 + 236 * Log (el. var.) 2 R =0.73, P154 m3 sec-1 over the preceding two years and basal area increment (BAI) and b) the number of days of river discharge >154 m3 sec-1 and BAI over 2-yr periods from 1991-2004. Open symbols represent post-restoration years.

77

Plot environmental parameters 2004 No. of floods (total)* No. of floods (growing season)* Days connected with the river (total)* Days connected with the river (growing season)* 2003 - 2004 No. of floods (total)* No. of floods (growing season)* Days connected with the river (total)* Days connected with the river (growing season)* 2001 - 2004 No. of floods (total)* No. of floods (growing season)* Days connected with the river (total)* Days connected with the river (growing season)* Mean plot elevation (m above MSL) Plot elevation variance Mean canopy cover (%) Total basal area (cm2 m-2)

Mean (±1 SE)

Range

4.4 ± 0.2 3.4 ± 0.2 21.1 ± 2.9

4–5 3–4 15 – 30

17.7 ± 2.7

7 – 26

11.1 ± 0.6 8.2 ± 0.6 37.8 ± 5.8

8 - 13 6 - 10 16 - 56

28.8 ± 4.9

11 - 44

16.2 ± 1.3 11.1 ± 0.9 54.0 ± 8.1

10 - 20 8 -14 22 - 81

38.3 ± 6.0

17 - 57

221.38 ± 0.07 1.15 ± 0.43

221.08 – 221.86 0.21 – 4.67

81.7 ± 1.3 39.0 ± 3.8

72.9 – 88.2 27.2 – 65.0

Note: Total year consists of 12 months (from preceding October-September). Growing season consists of 6 months (April-September) * Flood parameters do not include upland Plot #5 which was estimated to have flooded only once (in 2003) from 2001-2004.

Table 3.1. Synopsis of tree plot environmental variables used for regression analyses with forest productivity.

78

Importance Value Species (common name) North Sec. South Sec. Acer negundo L. (boxelder) 36.6 94.3 Acer saccharinum L. (silver maple) 15.0 8.9 Acer saccharum Marsh. (sugar maple) 7.1 9.2 Aesculus glabra Willd. (Ohio buckeye) 48.5 51.1 Asimina triloba (L.) Dunal (paw paw) -68.0 Celtis occidentalis Willd. (hackberry) 8.1 46.1 Fraxinus pennsylvanica Marsh. (green ash) 3.9 6.2 Gleditsia triacanthos L. (honey locust) -16.0 Juglans nigra L. (black walnut) 13.8 4.6 Lonicera maackii (Rupr.) Amur honeysuckle -7.0 Maclura pomifera (Raf.) (osage-orange) 3.8 -Morus alba L. (white mulberry) 8.8 7.1 Morus rubra L. (red mulberry) 8.7 6.8 Platanus occidentalis L. (sycamore) 18.9 20.4 Populus deltiodes Bartr. Ex (cottonwood) 11.2 41.3 Prunus serotina Ehrh. (black cherry) -5.8 Salix nigra L. (black willow) -7.5 Ulmus americana L. (American elm) __9.5__ __6.0__ Total 300.0 300.0

Table 3.2. Importance value (= rel. density + rel. dominance + rel. frequency) of all tree species identified in the north and south sections of the ORWRP bottomland forest. Dominant species (Impt.value >35) are in bold.

79

Mean basal area increment PrePostrestoration restoration (1991-2000) (2000-2004)

Section (n=)

BAI parameter

North (n=17)

% cm2 yr-1

4.3 ±0.6 33.5 ±4.6

South (n=25)

% cm2 yr-1

Upland (n=7)

% cm2 yr-1

Paired t-test T-value

P

3.3 ±0.6 30.8 ±4.0

2.99 0.90

0.009 NS

3.0 ±0.4 28.5 ±3.6

2.3 ±0.2 27.4 ±3.7

2.28 0.41

0.032 NS

3.0 ±0.5 24.8 ±3.6

3.8 ±0.6 24.6.0 ±4.9

2.30 0.04

NS NS

Table 3.3. Results of paired t-tests for mean (±1 SE) basal area increment (BAI) (% and cm2 yr-1) of canopy trees pre- and post-restoration. NS denotes non-significant p-value.

80

Species (n=)

BAI parameter

Mean basal area increment Pre-restoration Post-restoration (1991-2000) (2000-2004)

Paired t-test T-value

P

A. negundo (n=19)

% cm2 yr-1

3.2 ±0.3 29.1 ±3.8

2.3 ±0.2 27.7 ±3.5

2.75 1.38

0.013 NS

A. glabra (n=6)

% cm2 yr-1

1.9 ±0.4 14.1 ±4.5

1.8 ±0.3 14.8 ±3.9

0.25 -0.35

NS NS

Table 3.4. Results of paired t-tests for mean (±1 SE) basal area increment (BAI) (% and cm2 yr-1) pre- and post-restoration for boxelder (Acer negundo L.) and Ohio buckeye (Aesculus glabra Willd.). NS denotes non-significant p-value.

81

CHAPTER 4

TEMPORAL AND SPATIAL DEVELOPMENT OF SURFACE SOIL CONDITIONS AT TWO CREATED RIVERINE MARSHES

4.1 Abstract

The amount of time it takes for created wetlands to develop soils comparable to natural wetlands is relatively unknown. Surface soil changes over time were evaluated in two created wetlands (~1 ha each) at the Olentangy River Wetland Research Park in Columbus, Ohio. The two wetlands were constructed in 1993 to be identical in size and geomorphology, and maintained to have the same hydrology. The only initial difference between the wetlands was that one was planted with native macrophytes while the other was not. In May 2004, soil samples were collected (10 years and 2 months after the wetlands were flooded) and compared to samples collected in 1993 (after the wetlands were excavated but prior to flooding) and 1995 (18 months after the wetlands were flooded). In all three years, soils were split into surface (0-8 cm) and subsurface (8-16 cm) depths and analyzed for soil organic matter, total C, total P, available P, exchangeable cations and pH. Soils in the two wetlands have changed substantially 82

through sedimentation and organic accretion. Between 1993 and 1995, soils were most influenced by the deposition of senescent macroalgae, the mobilization of soluble nutrients and the precipitation of CaCO3. Between 1995 and 2004, soil parameters were influenced more by the deposition of organic matter from colonized macrophyte communities. Mean percent organic matter at the surface increased from 5.3 ±0.1% in 1993, 6.1 ±0.2% in 1995, to 9.5 ±0.2% in 2004. Mean total P increased from 493 ±18 μg g-1 in 1993, 600 ±23 μg g-1 in 1995, to 724 ±20 μg g-1 in 2004. Spatial analyses of percent organic matter (a commonly used indicator of hydric soil condition) at both wetlands in 1993, 1995 and 2004 showed that soil conditions have become increasingly more variable. High spatial structure (autocorrelation) between data points was detected in 1993 and 2004, with data in 2004 exhibiting a much higher overall variance and narrower range of spatial structure than in 1993.

4.2 Introduction Wetlands are constructed throughout the United States to provide landscape functions such as wildlife habitat, flood attenuation, and water quality enhancement (Mitsch and Gosselink 2000). Where regulatory requirements stipulate monitoring of wetland creation areas, hydrology and vegetation are usually used as indicators of wetland condition. Soils are often the least considered component of wetland systems despite their importance in providing the substrate for many of the biological and chemical processes that make them valuable components to the landscape (Vepraskas and Faulkner 2001, Collins and Kuehl 2001). There have been an increasing number of studies conducted to evaluate soil conditions in created wetlands. Many of the studies have been 83

designed to compare the soils of created wetlands to natural reference wetlands (BishelMachung et al. 1996, Shaffer and Ernst 1999, Zedler and Callaway 1999; Nair et al. 2001, Campbell et al. 2002) with most finding some progressions toward natural wetland soil conditions but with substantial differences in many key characteristics (e.g., lower soil organic matter concentrations, coarser texture, and dissimilar nutrient concentrations and pH). When terrestrial soils become flooded, there are several biogeochemical transformations that can occur over different time intervals. After only a few days of flooding, oxygen in the soil column becomes depleted and microbial activity will be dominated by facultative and strict anaerobes (Mitsch and Gosselink 2000). Soil colors will become darker as reduced Fe and Mn are transported out of the soil column during flooded conditions. Soils with chroma values < 2 are used to indicate the presence of hydric soil conditions (Tiner 1999). Flooding also influences soil P availability due to its release into the water column (Sanyal and De Datta 1991). Longer-term changes in soil condition are influenced by the buildup of soil organic matter at the surface caused by the reduced rate of decomposition. The accumulation of soil organic matter has been identified as an indication of soil maturity in created wetlands because of the time required for it to develop (Craft 2001, Nair et al. 2001). Several important biogeochemical processes associated with wetlands (e.g. denitrification) are dependent upon adequate soil carbon being present (Mitsch and Gosselink 2000). While several studies have evaluated temporal changes in the soil organic matter of created wetlands (Bishel-Machung et al. 1996, Nair et al. 2001, Anderson and Cowell 2004), few have evaluated the spatial patterns that occur over time. This is partially 84

because of the explicit sampling design that is required to evaluate spatial dynamics. Spatial patterns associated with natural wetland soil characteristics such as P enrichment in the Everglades (DeBusk et al. 2001) and P-sorption capabilities in North Carolina floodplains (Bruland and Richardson 2004) have been studied. Changes in how soil properties were distributed within a created wetland were observed after two years in response to flooding at the Des Plaines River wetlands near Chicago, Illinois (Fennessy and Mitsch 2001). They found that spatial variability of soil organic C and exchangeable nutrient concentrations as measured by the range of autocorrelation influence declined after two years of flooding. At the Olentangy River Wetland Research Park (ORWRP) in Columbus, Ohio, two 1ha experimental wetlands were constructed in 1993. One was planted with native macrophytes and the other was not. Extensive soil surveys were conducted at the two wetlands in August 1993 (after excavation but prior to flooding), September 1995 (18 months after flooding) and May 2004 (10 years and 2 months after flooding) to evaluate changes in response to flooding. Over the last ten years, several investigations have identified the rapid development of a sediment layer, and significant increases in soil organic C, Ca, Fe, P and total C (Nairn 1996, Liptak 2000, Harter and Mitsch 2003). Initial accumulations were attributed to the autochthonous production of dense algae mats (Wu and Mitsch 1998) and allochthonous import of sediment (Harter and Mitsch 2003). After the third year, both wetlands had developed significant cover by macrophyte communities, which are now considered the primary contributor to soil organic matter. This study represents the first to examine changes in soil condition at the two wetlands since its creation along with changes in spatial variability. Detailed descriptions of the 85

hydrologic, biogeochemical, and ecological patterns of these experimental wetlands are given by Metzker and Mitsch (1997), Mitsch et al. (1998, 2005a,b,c), Kang et al. (1998), Koreny et al. (1999), Nairn and Mitsch (2000), Spieles and Mitsch (2002a,b, 2003), Ahn and Mitsch (2002), Anderson et al. (2002), Selbo and Snow (2004), and Zhang and Mitsch (2005). The first objective of this study was to compare soil data collected in 1993, 1995 and 2004 to evaluate changes at the soil surface that have occurred as a result of the created riverine-wetland conditions. Given the high productivity and flooded conditions, we hypothesized that the wetland soil surface has substantially increased in its concentration of organic matter and nutrients associated with organic matter (organic C, N, P and exchangeable cations). Our second objective was to compare the spatial patterns of soil organic matter concentrations in samples collected in 1993, 1995 and 2004. Starting with the antecedent soil conditions (1993), we expected to see an increase in the concentration and spatial variability of organic matter.

4.3 Methods and materials 4.3.1 Study area The study was conducted at the Olentangy River Wetland Research Park (ORWRP) on The Ohio State University campus in Columbus, Ohio, USA (latitude N40.021◦, longitude E83.017◦). The ORWRP is a 10-ha facility located along the Olentangy River and was constructed on abandoned agricultural land. Underlying soils in this area are alluvial floodplain soils, comprised of the Ross and Eldean series (classified as a Cumlic Hapludoll), which include silt loams, silt clay, and clay loams (Mcloda and Parkinson 86

1980). Two 1-ha experimental marshes were excavated at the ORWRP in 1993 and flooded in March 1994 with pumped Olentangy River water. The two wetlands were built and managed identically with Olentangy River water being pumped at a similar rate (typically ~25 m yr-1) throughout their ten-year history. As part of a long-term study, the only difference between wetlands was that the western marsh (Wetland 1) was planted with native, wetland vegetation while the eastern marsh (Wetland 2) was left unplanted (Mitsch et al. 2004) (Fig. 4.1). Based on their topography, both wetlands have developed two distinct cover zones: a shallow, emergent vegetation (EM) zone and three deeper, open-water (OW) sub-basins spaced longitudinally along each wetland (Fig. 4.2). The EM zones were constructed approximately 0.3 m below natural grade and the OW basins were typically 0.6 m below grade. In the first three years, both wetlands were similar in form with large areas of open water gradually colonizing with macrophyte cover in the EM zones, predominantly Schoenoplectus tabernaemontani (C.C. Gmel) Palla. However, between the years of 1998 and 2001, Wetland 2 became dominated by dense stands of Typha spp. (mostly Typha angustifolia L., Selbo and Snow 2004) while Wetland 1 maintained a more mixed community assemblage. Because of their depth, the OW zones have only supported sparse amounts of emergent macrophytes, but have supported macroalgae and other aquatic vegetation (e.g., Ceratophyllum sp. and Lemna sp., Anderson and Mitsch 2003). Although both wetlands were excavated to be 1-ha in size, after ten years of peripheral shrub encroachment, the combined marsh area of Wetland 1 and 2 in 2004 was approximately 0.81 and 0.88 ha, respectively with the OW zone covering approximately 29 and 28% of each wetland, respectively.

87

4.3.2 Soil sampling design Soil sampling in 1993, 1995 and 2004 was conducted based on a 10-m grid system marked at each intersection point with a permanently installed 2-cm diameter PVC pole (Fig. 4.2). In 1993 (after wetland construction, but before flooding) and 1995 (18 months after flooding), soil samples were collected at the same 43 intersection points (Fig. 4.2) at a depth of 0-8 cm and 8-16 cm. In 2004, a total of 127 intersection points were used to collect samples at 0-8 cm and 8-16 cm depths (Fig. 4.2). Sampling methods described below are specific to the 2004 sampling period, but were designed to be consistent with methods used in 1993 and 1995 (Nairn 1996). Water depths were lowered to minimize standing water at each grid point and facilitate soil extraction. At each grid point, soils were collected 0.5 m east of the field marker. Soils were collected using a 10-cm diameter steel soil-corer, carefully removed, and split into 0-8 and 8-16 cm sections using a sharp knife. The 0-8 cm section was then halved length wise and stored in separate water-tight freezer bags. Because most 8-16 cm sections were typically dense clay, this section was split into quarters and two of the quarter-sections were placed in separate plastic freezer bags. Soil remnants were replaced into the sample hole. For each sample, the hue, value and chroma were determined using a Munsell Color Chart and other visual characteristics were noted. Because of the dense consistency of the antecedent soil surface, the development and boundary of the accreted sedimentlayer was usually apparent. When it was, the depth was measured to the nearest 0.5 cm. Each sample section was placed in a plastic freezer bag and kept in an ice-packed cooler 88

until being returned to the laboratory where they were refrigerated at 4 ûC until laboratory analysis.

4.3.3 Physical and chemical soil analyses One section of each soil sample was weighed and placed in a drying-oven at 105 ûC for five days or until constant mass occurred. Soil sections were reweighed to determine soil moisture content and bulk density. The second section of each soil sample was kept in its field-moist, natural condition and completely homogenized by hand. A 30-g subsample of each sample was air-dried at room temperature, ground using a pestle and mortar, and passed through a 2-mm sieve. Duplicate subsamples (approximately 10 g each) were placed in a crucible, oven-dried at 60 ûC overnight, weighed, and ignited in a muffle furnace at 550ûC for 1 hour. The post combustion material was reweighed and the duplicates averaged to estimate the percent organic matter of the soil. A second subsample from the field-moist section was used to characterize soils for various chemical parameters. For each year, samples (at 0-8 and 8-16 cm depths) were collected and analyzed from the same grid points. Samples were selected to analyze chemical conditions over an even spatial distribution and to be proportionate among the cover zones. A total of 168 samples [56 per year based on 2 samples (0-8 and 8-16 cm) collected at 28 grid points] were analyzed for available P by the Bray-P1 extraction (Kuo 1996), exchangeable K, Ca, and Mg by 1M ammonium acetate extraction (Warncke and Brown 1998), and pH (Thomas 1996). A total of 108 of these samples [36 per year based on 2 samples (0-8 and 8-16 cm) collected at 18 grid points] were further analyzed for total C by combustion (ISO, 1995; AOAC, 1989) and total P by digestion with 89

HClO4/HNO3 followed by inductively coupled plasma emission spectrometry (Sommers and Nelson 1972).

4.3.4 Temporal and geostatistical statistical analyses Because several of the soil parameters had unequal variances and could not be transformed to fit a normal distribution, mean comparison of each soil parameter in 1993, 1995 and 2004 was conducted using nonparametric Friedman Two-Way Analysis of Variance of repeated measure with post hoc comparison of years conducted using Wilcoxon Signed Ranks Test. The Friedman test was used to strictly evaluate changes over time at each repeatedly sampled grid point (no intra-annual comparisons were considered) therefore the potential ramifications of using non-independent data were minimized. All tests were conducted using Systat v.10.2 (Systat Software Co. 2002). For each statistical test, differences were considered significant at p8 >7 >6 >5 >3

0 0

88

16

Distance (10 m)

Figure 4.5 Spatial distribution maps of soil organic matter for Wetland 1 and 2 in a) 1993, b) 1995 and c) 2005. Maps for 1993 and 2004 were generated by ordinary point kriging using isotropic variogram models. Maps for 1995 were generated using inverse distance weighing method for spatial interpolation (see text).

112

19

Distance (10 m)

b) 1995

Percent Organic Matter > 13 > 10 >9 >8 >7 >6 >5 >3

0 0

8 8

16

Distance (10 m)

Figure 4.5 (continued)

113

19

Distance (10 m)

c) 2004

Percent Organic Matter > 13 > 10 >9 >8 >7 >6 >5 >3

0 0

88

16

Distance (10 m)

Figure 4.5 (continued)

114

Variogram characteristic nugget (C0)

1993 0.10

Wetland 1 1995 --†

2004 0.45

Wetland 2 1993 1995 8 cm deep. Each core was visually inspected, split vertically into two equally sized sections, and placed in separate watertight freezer bags. Collected sediment samples were kept in an iced cooler until being returned to the laboratory where they were stored at 4 ◦C. At the laboratory, each sediment sample was homogenized by hand. One section from each sampling point was oven-dried at 105ûC for determination of bulk density. The second section of each soil sample was kept in its field-moist, natural condition and a 122

30-g subsample of each sample was air-dried at room temperature, ground using a pestle and mortar, and passed through a 2-mm sieve. Duplicate subsamples (approximately 10 g each) were placed in a crucible, oven-dried at 60 ûC overnight, weighed, and ignited in a muffle furnace at 550ûC for 1 hour. Ten subsamples were further analyzed for textural conditions using the hydrometer method (Gee and Bauder 1986). Using the field-moist sections, 47 100-g subsamples were air-dried at room temperature, ground using a pestle and mortar, passed through a 2 mm sieve, and. analyzed for total N and C by combustion (AOAC 1989, ISO 1995). From this group, 21 samples were further analyzed for total P and Ca by digestion with HClO4/HNO3 followed by inductively coupled plasma emission spectrometry (Sommers and Nelson 1972). A total of 22 samples were analyzed for inorganic C (USEPA 2005). Percent organic C was calculated based on the difference of total C and inorganic C and a regression was used to estimate percent organic C for all samples based on its percent organic matter (Konen et al. 2002). For all chemical analyses, samples were selected to provide an even spatial distribution and proportionate sampling between cover zones. A weighted average of each nutrient concentration was calculated for both wetlands based on the percent surface area of each cover zone. The samples analyzed for nutrient concentrations were also used to characterize sediment conditions within each respective wetland cover zone. Mean sediment and nutrient content (g cm-3) were calculated for each individual wetland cover zone using sampled concentrations and bulk densities. Mean content was multiplied with sediment depth to estimate sediment and nutrient accumulation (converted to g m-2) at each grid point measured for depth. Mean accumulation for each wetland cover zone was calculated and a weighted average was 123

calculated for the each entire wetland. Mean sediment and nutrient accumulations for each wetland were used to estimate annual accumulation rates (kg yr-1) based on a wetland age of 10.2 years.

5.3.3 Statistical analyses The mean and standard error of sediment/nutrient concentrations and accumulations were determined for each wetland cover zone. Given the non-independent and potentially autocorrelated nature of the grid-sediment data, standard parametric statistical analyses were considered inappropriate and not used. To evaluate if wetland differences in 10-yr macrophyte diversity influenced sediment and organic C accumulation sample variance, Wetland 1 and 2 were compared using Levene’s Test for equal variances, with P0.05) in sample variance between wetlands. Examining the frequency distribution of sediment accumulation based on kriging analysis (Fig. 5.3b, see variograms results below), both wetlands had significant overlap, however Wetland 2 had a greater frequency of high range sediment accumulation (>60 kg m-2) because of the higher accumulation in its OW zones.

5.4.2 Spatial patterns of sediment accumulation In Wetland 1, no spatial structure associated with sediment accumulation was detected based on prepared isotropic variograms and correlograms. Strong spatial structure was detected at the 0◦ direction based on an anisotropic variogram (C0=61.2, C0+C=480.3, A0=5.1) and correlogram (Fig. 5.4a). Reviewing correlogram results, a Moran’s I value of 0.63 was determined for grid pairs at the first lag interval and steadily decreased over the 60 m lag distance. None of the other anisotropic directions provided evidence of spatial structure. In Wetland 2, evidence of moderate spatial structure was detected in the isotropic correlogram along with comparable anisotropic structure at 0◦ and 135◦ directions (Fig. 5.4b). However, the 135◦ variogram (C0=160.8, C0+C=616.0, A0=4.8) was the only one that corroborated with the correlogram results and therefore 127

was used for kriging analyses. In the 135◦ correlogram, a Moran’s I value of 0.30 was detected at the first interval, immediately dropped below 0 at the second lag interval, and stayed near 0 throughout the remaining lag distance. All anistotropic variograms and correlograms were re-analyzed after adjusting the direction at 5◦ intervals, however, none of the adjusted variograms improved upon the results reported. Kriging maps of Wetland 1 and 2 showed the range and locations of sediment accumulation throughout the wetlands (Fig. 5.5). Kriging analysis tends to suppress the range of interpolated values; therefore extreme high and low measurements tend to be less represented by this procedure. Nevertheless, in Wetland 2, there was a clear distinction of higher accumulation occurring within all three OW sub-basins compared to its surrounding EM zone (Fig. 5.5). In Wetland 1, only the middle sub-basin showed the same distinction. Both Wetland 1 and 2 showed highly variable conditions in the EM zones with accumulations ranging from 13.9 to 70.9 kg m-2.

5.4.3 Nutrient accumulation 5.4.3.1 Carbon No substantial differences in total C accumulation (Fig. 5.6a) were observed between wetlands, however there was 17% more in the OW zones of Wetland 2 (2.8 ±0.2 kg m-2) compared to Wetland 1 (2.4 ±0.2 kg m-2). This was primarily due to the accumulation of organic C which represented the majority of detected C accumulation and had similar proportions between wetlands and cover zones as did total C (Fig. 5.6b). Levene’s Test for equal variances indicated no significant difference (P>0.05) in organic C sample variance between wetlands. Both wetlands had significant amounts of inorganic C in the 128

OW zones sediment (Fig. 5.6c). Inorganic C represented 30 and 26% of the total C accumulated in the OW zones of Wetland 1 and 2, respectively. Based on the mean wetland accumulation total, annual rates of total, organic and inorganic C were prepared (Table 5.2). Wetland 2 had higher rates of all C constituents except for inorganic C.

5.4.3.2 Nitrogen, phosphorus and calcium Total N accumulation was nearly identical in Wetland 1 and 2, with slightly higher accumulations detected in the OW zones (Fig. 5.7a). Highest N accumulation was detected in the OW zone of Wetland 2 (228 ±14 g m-2). P accumulation was also highest in the OW zones of Wetlands 2 (58 ±4 g m-2) and was 41% higher than the OW zones of Wetland 1 (41 ±4 g m-2) (Fig. 5.7b). However, because P accumulation in the EM zone was higher in Wetland 1, differences in total wetland P accumulation were negligible. Total Ca accumulation was consistent between wetlands but showed extreme intrawetland variation (Fig. 5.7c). In Wetland 1 and 2, total Ca accumulation was 13 and 11 times higher, respectively, in the OW zones compared to respective EM zones (Fig. 5.7c; Appendix C, Table C.5). Annual accumulation rates and net accumulation of N, P and Ca were similar between wetlands (Table 5.2).

5.5 Discussion 5.5.1 Sediment accumulation and spatial patterns Based on whole-wetland estimates, there was no evidence that the higher 10-yr productivity in the naturally colonizing Wetland 2 resulted in a greater accumulation of sediment. We had expected that wetland to have more sediment accumulation primarily 129

through greater accretion of autochthonous organic matter, but accumulations of sediment and organic C were similar between the EM zones of both wetlands (Fig. 5.3a and 6b). However sediment accumulation was 19% higher in the OW zones of Wetland 2 compared to Wetland 1, and the sediment kriging maps clearly showed that the OW zones of Wetland 2 had accumulated more than Wetland 1. It is uncertain exactly why this difference occurred but it is possible that vegetation may have played a role. Between 1998 and 2001, Wetland 2 was dominated by dense stands of T. angustifolia that stood well above 2 m in height with mean shoot densities >32 m-2 (Selbo and Snow 2004). Vegetation communities in Wetland 1 tended to be more diverse but lower-growing and with less productive macrophytes. It is possible that the dense Typha stands surrounding the OW zones of Wetland 2 were more effectively sheltered these areas and increased sediment deposition while suppressing re-suspension. Other studies have found evidence that plant density and height can influence sediment deposition rates (Horppila and Nurminen 2001, Darke and Megonigal 2003). This would explain why accumulation rates were highest at the inflow sub-basin of Wetland 2 and were higher overall compared to Wetland 1. The highest sediment accumulation in the planted wetland occurred at the middle sub-basin after it had been transported a considerable distance from the inflow. Both wetlands demonstrated spatial structure associated with the accumulation of sediment. The 0◦ anisotropic results in the planted Wetland 1 suggest that spatial structure was occurring primarily in a north-to-south direction. This effect was caused in part by the gradual decrease in overall accumulation from inflow to outflow, but also the influence from the middle OW sub-basin which is elongated in a north-south direction (Fig. 5.2). This sub-basin was the only one that accumulated sediment in stark contrast to 130

the surrounding EM zones and its concentric nature within the wetland likely contributed to the 0◦ spatial structure detected (Fig. 5.5a). This was in contrast to the naturally colonized Wetland 2 which exhibited more moderate spatial structure but at isotropic and anisotropic (0◦ and 135◦) directions (Fig. 5.5b). This was caused by the same decrease in accumulation from inflow to outflow seen in Wetland 1, but with a greater influence by all three OW sub-basins. As indicated in the sediment kriging maps (Fig. 5.5), the greatest contrast between the EM and OW zones in Wetland 2 occurred at the inflow (north) section of the wetland which was analyzed most extensively using the 135◦ anisotropic analyses (Fig. 5.4). While this direction produced correlogram results similar to other directions, the 135◦ variogram was the only one that detected spatial structure. In both wetlands, the variograms and correlograms that detected spatial structure were using directional analyses that most extensively intercepted the OW sub-basin with the highest sediment accrual. Therefore, we conclude that these wetland features had a prevailing influence on sediment spatial structure.

5.5.2 Longitudinal patterns Both wetlands showed a general longitudinal decrease in sediment accumulation. This is a common occurrence in flow-through wetlands where sedimentation rates are highest near the inflow source and decrease as they continue through the system (Fennessy et al. 1994, Braskerud 2001). In the case of both Wetland 1 and 2, the southern half of each wetland had EM zones with sediment accumulation at 40 kg. Both wetlands also had some of the least sediment accumulation occur in the EM zones along the concave 131

mid-section of each wetland (Fig. 5.4). This area of both wetlands represents some of the shallowest sections and consists of a gradual elevation relief from the wetland edge to the mid OW sub-basin. The low sediment accumulation in these sections and high accumulation in the adjacent OW zone suggests that water is preferentially flowing through the open water area and by-passing this shallow emergent area. It was also noted that the OW zones in Wetland 2 also followed a decreasing accumulation gradient from inflow to outflow, unlike Wetland 1. Again, this may be the result of several years of dense Typha that may have promoted earlier sedimentation (closer to the inflow) and reduced re-suspension. Other factors also influenced sediment accumulation. Local topographic conditions were shown to have an effect as sediment accretion was often minimal near the crest of the OW sub-basins (Anderson et al. 2005). Without the benefit of rooted macrophyte cover, these areas appear to be highly susceptible to sediment transport either southward towards the outflow or sloughing downward into the deeper sections of the OW subbasin. Other factors influencing sediment distribution are more difficult to account for. Faunal species such as beaver (Castor canadensis) and muskrats (Ondatra zibethicus) have been active in the wetlands during portions of their history and have likely contributed to the patchiness of sediment seen.

5.5.3 Sedimentation rates When averaged for the entire wetland, sediment accumulation rates for Wetland 1 and 2 were between those reported for newly created wetlands and those seen in natural wetlands. An investigation of short-term sedimentation at Wetland 1 and 2 (between 132

1996 and 1997) using horizon markers showed an average sedimentation rate of 36 kg m2

yr-1 (Harter and Mitsch 2003). High sediment accumulation rates have also been

reported in other newly created wetlands. At the Des Plaines River Wetland Demonstration Project north of Chicago, Illinois, Fennessy et al. (1994) reported rates between 5.9 – 12.8 kg m-2 yr-1 during its first year and 1.2 – 4.2 kg m-2 yr-1 during its second year. Braskerud (2001) reported rates between 15 – 75 kg m-2 yr-1 for four small newly-created wetlands in an agricultural watershed in Norway, and Craft (1997) reported rates between 21 – 36 kg m-2 yr-1 for newly constructed estuarine marshes in North Carolina. Our study found overall annual sedimentation rates (4.5 – 4.9 kg m-2 yr-1) that were much lower than those found by the Harter and Mitsch (2003) horizon marker study suggesting that the wetlands are no longer retaining sediment at the pace it was in the first few years after it was constructed. This was supported by long term water quality data that was collected daily from the inflow and outflow of both wetlands since they were flooded (Mitsch and Zhang 2004). In the three years leading up to this sediment survey, both wetlands showed a decrease in turbidity abatement between inflow and outflow (an indicator of sediment retention), and by 2001, both wetlands had become sediment sources (Mitsch et al. 2004). This development has also influenced P retention and as of 2003 both wetlands have exported more P than retained (Mitsch et al. 2004).

5.5.4 Nutrient accumulation 5.5.4.1 Carbon

133

The lack of any substantial differences in C accumulation between Wetland 1 and 2 was somewhat unexpected given the higher 10-yr productivity in Wetland 2. However since 2001, the planted Wetland 1 has been the more productive wetland (mean net annual aboveground primary productivity after 2001 was 561 ±77 in Wetland 1 and 356 ±59 in Wetland 2) and the higher deposition of recent organic matter may explain why the two wetlands are now similar in total and organic C. Evidence of the higher proportion of recent organic deposition in Wetland 1 is also reflected by the higher concentrations of total N detected in Wetland 1 (Table 5.1). The higher accumulation of inorganic C in the OW zones of each wetland is an indication that CaCO3 precipitation has remained a critical process in sediment condition. Macrophyte shading has effectively precluded this process in the EM zones and therefore accumulation was substantially less. Higher accumulation of inorganic C in the EM zone of Wetland 1 (compared to Wetland 2) suggests that overall algal productivity may have also been higher. This would be expected given several years in which dense Typha stands in Wetland 2 likely precluded algae growth more so than vegetation communities in Wetland 1. C accumulation rates at the ORWRP wetlands were within the range seen in the literature. In an evaluation of four created estuarine marshes in North Carolina (1-15 yrs old), Craft (1997) found the accumulation of organic C at an average rate of 80 g m-2 yr-1. This rate was also determined for reference natural wetlands used for comparison. Along a Typha gradient in the anthropogenically influenced section of the Everglades, rates for C accumulation have been documented between 86 – 387 g m-2 yr-1 (Reddy et al. 1993) and soil organic matter at 492 – 1160 g m-2 yr-1 (Davis 1991). 134

5.5.4.2 Nitrogen, phosphorus and calcium Wetland removal/retention of N is typically dependent upon organic matter accumulation as an energy source for denitrifying bacteria and for retaining organically bound N. At the two experimental wetlands, N accumulated in proportions similar to organic C and minimal differences between wetlands were detected (16.2 and 16.6 g m-2 yr-1 for Wetland 1 and 2, respectively). These rates are above average for created wetlands but are comparable to other studies where high nutrient loads cause increased primary productivity and deposition of organic matter (Craft 1997). Armentano and Woodwell (1975) found N accumulation was 14 g m-2 yr-1 in a marsh on Long Island, NY and Reddy et al. (1993) found that Typha dominated areas of the Everglades most affected by anthropogenic nutrient loads had N accumulation rates of 11 – 24 g m-2 yr-1, compared to unaffected areas that only averaged 5 g m-2 yr-1. Nutrient loading and organic accretion likely explains the high N storage in the ORWRP sediment as these wetlands are pumped at a fairly high rate (>22 m yr-1 in 2003, Zhang and Mitsch 2004) with river water that is often very high in nitrate concentration (mean NO3 + NO2 inflow concentration for 2003 was 4.1 ±0.3 mg N L-1, Mitsch et al. 2004). As seen in other riverine wetlands, P accumulation in the sediment layer can be attributed to a combination of organic matter build-up, deposition of P bound sediment, and biochemical processes such as sorption and precipitation of dissolved P (Johnston 1991, Axt and Walbridge 1999). We expect that all these processes have occurred at the ORWRP wetlands. Water column productivity in both wetlands has remained high since the wetlands were flooded and the co-precipitation of CaCO3 and P has been found to be 135

a highly significant P-retention process in these wetlands (Liptak 2000). Like inorganic C, the accumulation of Ca was exceptionally high in the wetland OW zones compared to the EM zones because of the precipitation of CaCO3. A nearly perfect correlation (r2=0.99) was detected between Ca and inorganic C suggesting that carbonates are the primary form of sediment Ca fractions. Because P readily adsorbs to CaCO3, this process has likely contributed to significant P retention within both wetlands. Between 1994 and 1998, Liptak (2000) found precipitated calcite had accumulated at 0.45 kg m-2 yr-1and was one of the primary mechanisms for P sorption. Likewise, in a 10-yr old created and reference coastal marsh in North Carolina, Craft (1997) found that large amounts of Ca in the soil (159 - 516 g m-2) contributed to extensive P sorption (6.1 - 8.4 g m-2 yr-1). Calcite and other forms of CaCO3 have been attributed to the sorption of other elements beside P including NO3 (Jurinak and Griffin 1992) and K (Galvez-Cloutier and Dube, 1998). Total S concentrations were found to 3 to 4 times higher in the OW zones of the ORWRP wetlands compared to the EM zones (Appendix C; Table C.5). There has been minimal research regarding the potential sorption of S and CaCO3, however our evidence strongly suggests that this may be occurring. Although S constituents were not determined, Vepraskas and Faulkner (2001) reported that SO42- can be retained by the same adsorption processes that affect PO43-, although PO43- tends to displace it when both are available. The adsorptive capacity of CaCO3 may also contribute to the high N accumulations rates observed in both wetlands. Based on the conceptual model by Craft (1997), P removal is dominated by sedimentation and sorption/precipitation processes during its first three years but after ten years, removal declines significantly and becomes increasingly dependent on organic 136

matter accumulation. This pattern appears to be occurring at the ORWRP wetlands. P accumulation rates for the two ORWRP wetlands were generally less than those seen at newly created wetlands and were comparable to natural wetlands where rapid accretion occurs. Previous estimates of P retention at the ORWRP wetlands have been conducted by analyzing changes in surface water concentrations. Through water quality analyses, Nairn and Mitsch (2000) estimated that in the first two years, Wetland 1 and 2 retained 6.7 and 7.5 g m-2 yr-1, respectively although water quality analyses can sometimes overestimate retention compared to actual sedimentation rates (Mitsch et al. 1995). After ten years, P accumulation rates in Wetland 1 and 2 have decreased to 3.3 and 3.5 g m-2 yr1

, respectively. Other researchers have found comparable rates in natural open wetland

systems. DeLaune and Patrick (1980) found P accumulated at 2.3 g m-2 yr-1 in a Louisiana salt marsh and Cooper and Gilliam (1987) found P accumulation was 4.3 g m-2 yr-1 for riparian areas in North Carolina. Highly productive Typha invaded areas in the Everglades were found to have an elevated rate of P accumulation (0.54 – 1.14 g m-2 yr-1) compared to areas of still dominated by Cladium (0.11 – 0.25 g m-2 yr-1, Reddy et al. 1993 and 0.4 g m-2 yr-1, Richardson and Craft 1993).

5.6 Conclusions After ten years, sediment accumulations within the two experimental marshes were highly variable, but mean accumulation rates were generally between those seen in newly created and natural open-wetlands. The decrease in sediment accumulation rates from those reported for 1996-1997 by Harter and Mitsch (2001) and water quality data from the last three years, indicates that these wetlands may no longer function as significant 137

sinks of sediment and P. Despite several years of markedly higher productivity in Wetland 2, this wetland showed no evidence of higher organic C accumulation. However, the dense Typha established during those years may have elicited greater sediment deposition and reduced re-suspension in the OW zones. Sediment accumulation in both wetlands showed anisotropic spatial structure that was caused in part because by the decrease in accumulation from inflow to outflow, but more so by the influence of high accumulation in OW sub-basins. Nutrient accumulation was consistent between wetlands, however high intra-wetland variation occurred for inorganic C, P and Ca indicating that the co-precipitation of CaCO3 in the OW zones is still a primary factor on sediment condition.

5.7 Acknowledgements Funding for this project was provided through an Ohio Agricultural Research & Development Center Graduate Research Enhancement Grant, the USDA (Grant No. 2002-35102-13518) and by support from the School of Natural Resources at The Ohio State University. Anne Altor assisted with soil texture analyses.

5.8 Literature Cited Anderson, C.J., W.J. Mitsch, and R.W. Nairn. 2005. Temporal and spatial development of surface soil conditions at two created riverine marshes. Journal of Environmental Quality 34(6): 2072-2081. Anderson, K.L., K.A. Wheeler, J.A. Robinson, and O.H. Tuovinen. 2002. Atrazine mineralization potential in two wetlands. Water Research 36: 4785-4794. AOAC Official Methods of Analysis. 1989. Method 990.03. Protein(crude) in Animal Feed Combustion Method(Dumas method). 17th edition 2002. Reference: JAOAC 72, 770 (1989). 138

Armentano, T. V. and G. M. Woodwell. 1975. Sedimentation rates in a Long Island marsh determined by 210Pb dating. Limnol. Oceanogr. 20:452-455. Axt, J. R. and M. R. Walbridge. 1999. Phosphate removal capacity of palustrine forested wetlands and adjacent uplands in Virginia. Soil Sci. Soc. Am. J. 63:1019-1031.

Braskerud, B. C. 2001. The influence of vegetation on sedimentation and resuspension of soil particles in small constructed wetlands. J. Environ. Qual. 30:1447-1457. Brenner, M., C. L. Schelske, and L. W. Keenan. 2001. Historical rates of sediment and nutrient accumulation in marshes of the Upper St. Johns River Basin, Florida, U.S.A. Journal of Paleolimnology 23:241-257. Brueske, C. C. and G. W. Barrett. 1994. Effects of vegetation and hydrologic load on sedimentation patterns in experimental wetland ecosystems. Ecological Engineering 3:429-447. Cooper, J. R. and J. W. Gilliam. 1987. Phosphorus redistribution from cultivated fields into riparian areas. Soil Sci. Soc. Am. J. 51:1600-1604. Craft, C. B. 1997. Dynamics of nitrogen and phosphorus retention during wetland ecosystem succession. Wetlands Ecology and Management 4:177-187. Craft, C. B. and W. P. Casey. 2000. Sediment and nutrient accumulation in floodplain and depressional freshwater wetlands of Georgia, USA. Wetlands 20: 323-332. Craft, C. B., S. Broome, and C. Carlton. 2002. Fifteen years of vegetation and soil development after brackish-water marsh creation. Restoration Ecology 10: 248-258. Craft, C. B., P. Megonigal, S. Broome, J. Stevenson, R. Freese, J. Cornell, L. Zheng and J. Sacco. 2003. The pace of ecosystem development of constructed Spartina alterniflora marshes. Ecological Applications 13:1417-1432. Darke, A. K. and J. P. Megonigal. 2003. Control of sediment deposition rates in two mid-Atlantic Coast tidal freshwater wetlands. Estuarine, Coastal and Shelf Science 57: 255-268. Davis, S. M. 1991. Growth, decomposition, and nutrient retention of Cladium jamaicense Crantz and Typha domingensis Pers in Florida Everglades. Aquat. Bot. 40:203-224.

139

DeLaune, R. D. and W. H. Patrick, Jr. 1980. Rate of sedimentation and its role in nutrient cycling in a Louisiana salt marsh. p. 401-412. In P. Hamilton and K. B. Macdonald (eds.) Estuarine and wetland Processes with Emphasis on Modeling. Plenum Publishing, New York, NY. Fennessy, M. S., C. C. Brueske, and W. J. Mitsch. 1994. Sediment deposition patterns in restored freshwater wetlands using sediment traps. Ecological Engineering 3:409428. Galvez-Cloutier, R. and J. S. Dube. 1998. An evaluation of freshwater sediments contamination: the Lachine Canal sediments case, Montreal, Canada. I. Quality assessment. Water, Air and Soil Poll. 102:259-279. Gee, G.W. and J.W. Bauder. 1986. Particle-size analysis. p. 383–411. In A. Klute (ed.) Methods of soil analysis. Part 1. 2nd ed. Agron. Monogr. 9. ASA and SSSA, Madison, WI. Harter, S. K. and W. J. Mitsch. 2003. Patterns of short-term sedimentation in a freshwater created marsh. J. Environ. Qual. 32:325-334. Horppila, J. and L. Nurminen. 2001. The effect of an emergent macrophyte (Typha angustifolia) on sediment resuspension in a shallow north temperate lake. Freshwater Biology 46:1447-1455. Hupp, C. R. and D. E. Bazemore. 1993. Temporal and spatial patterns of wetland sedimentation, West Tennessee. Journal of Hydrology 141:179-196. International Standard Organization, ISO. 1995. Soil quality – Determination of organic and total carbon after dry combustion (elementary analysis). International Organization for Standardization 10694:1995(E). Geneva, Switzerland. Isaaks, E.H. and R.M. Srivastava. 1989. An Introduction to Applied Geostatistics. Oxford University Press, Inc. New York, NY. Johnston, C. A. 1991. Sediment and nutrient retention by freshwater wetlands: effects on surface water quality. Critical Reviews in Environmental Control 21:491-565. Johnston, C. A., S. D. Bridgham, J. P. Schubauer-Berigan. 2001. Nutrient dynamics in relation to geomorphology of riverine wetlands. Soil Sci. Soc. Am. J. 65:557-577. Jurinak, J. J. and R. A. Griffin. 1992. Nitrate ion adsorption by calcium carbonate. Soil Sci. 113: 130-135.

140

Kang, H., C. Freeman, D. Lee, and W.J. Mitsch. 1998. Enzyme activities in constructed wetlands: Implication for water quality amelioration. Hydrobiologia 368:231-235. Konen, M. E., P. M. Jacobs, C. L. Burras, B. J. Talaga, and J. A. Mason. 2002. Equations for predicting soil organic carbon using loss-on-ignition for north central U.S. soils. Soil Sci. Soc. Am. J. 66:1878-1881. Koreny, J.S., W.J. Mitsch, E.S. Bair, and X. Wu. 1999. Regional and local hydrology of a created riparian wetland system. Wetlands 19:182-193. Liptak, M. 2000. Water column productivity, calcite precipitation, and phosphorus dynamics in freshwater marshes. PhD. Dissertation. The Ohio State University, Columbus, OH, USA. Mcloda, N. A. and R. J. Parkinson, 1980. Soil Survey of Franklin County, Ohio. USDASCS. US Government Printing Office, Washington, DC, USA. Mitsch, W. J., J. K. Cronk, X. Wu, and R. W. Nairn. 1995. Phosphorus retention in constructed freshwater riparian marshes. Ecological Applications 5: 830-845. Mitsch, W.J., X. Wu, R.W. Nairn, P.E. Weihe, N. Wang, R. Deal, and C.E. Boucher. 1998. Creating and restoring wetlands: A whole-ecosystem experiment in self-design. BioScience 48:1019-1030. Mitsch, W. J., C. J. Anderson, M. E. Hernandez, A. Altor and L. Zhang. 2004a. Net primary productivity of macrophyte communities after ten growing seasons in experimental marshes. p. 75-78. In W. J. Mitsch, L. Zhang and C. L. Tuttle (eds.) Olentangy River wetland Research Park at The Ohio State University Annual Report 2003. The Ohio State University, Columbus, OH, USA. Mitsch, W. J., L. Zhang, N. Dillon, and D. Fink. 2004b. Biogeochemical patterns of created riparian wetlands: tenth-year results (2003). p. 59-68. In W. J. Mitsch, L. Zhang and C. L. Tuttle (eds.) Olentangy River wetland Research Park at The Ohio State University Annual Report 2003. The Ohio State University, Columbus, OH, USA. Mitsch, W. J., N. Wang, L. Zhang, R. Deal, X. Wu, and A. Zuwerink. 2005. Using ecological indicators in a whole-ecosystem wetland experiment p. 211-236. In S. E. Jorgensen, R. Costanza and F. L. Xu (eds.) Handbook of Ecological Indicators for Assessment for Ecosystem Health. Lewis Publishers, CRC Press, Inc., Boca Raton, FL, USA. Mitsch, W.J., L. Zhang, C.J. Anderson, A. Altor, and M. Hernandez. Creating riverine wetlands: Ecological succession, nutrient retention, and pulsing effects. Ecological Engineering. (In Press). 141

Nairn, R. W., 1996. Biogeochemistry of newly created riparian wetlands: evaluation of water quality changes and soil development. PhD. Dissertation. The Ohio State University, Columbus, OH, USA. Nairn, R.W. and W.J. Mitsch. 2000. Phosphorus removal in created wetland ponds receiving river overflow. Ecological Engineering 14: 107-126. Olila, O. G., K. R. Reddy, and D. L. Stites. 1997. Influence of draining on soil phosphorus forms and distribution in a constructed wetland. Ecological Engineering 9:157-169. Pasternack, G. B. and G. S. Brush. 1996. Sedimentation cycles in a river-mouth tidal freshwater marsh. Estuaries 21:407-415. Perry, J. N., A. M. Liebhold, M. S. Rosenberg, J. Dungan, M. Miriti, A. Jakomulska and S. Citron-Pousty. 2002. Illustrations and guidelines for selecting statistical methods for quantifying spatial patterns in ecological data. Ecography 25: 578-600. Peterjohn, W. T. and D. L. Correll. 1984. Nutrient dynamics in an agricultural watershed; observations on the role of a riparian forest. Ecol. 65: 1466-1475. Reddy, K. R., R. D. DeLaune, W. F. DeBusk and M. S. Koch. 1993. Long-term nutrient accumulation rates in the Everglades. Soil Sci. Soc. Am. J. 57:1147-1155. Richardson, C. J. and C. B. Craft. 1993. Effective phosphorus retention in wetlands: fact or fiction? p. 271-282. In G. A. Moshiri (ed.) Constructed Wetlands for Water Quality Improvement. Lewis Publishers, CRC Press, Inc., Boca Raton, FL, USA. Selbo, S.M. and A.A. Snow. 2004. The potential for hybridization between Typha angustifolia and Typha latifolia in a constructed wetland. Aquatic Botany 78: 361369. Sommers, L. E. and D. W. Nelson. 1972. Determination of Total Phosphorus in Soils: A Rapid Perchloric Acid Digestion Procedure. Soil Science Society of America Proceedings. Vol. 36, 6: p.902-904. Madison, WI. Spieles, D.J. and W.J. Mitsch. 2000a. The effects of season and hydrologic and chemical loading on nitrate retention in constructed wetlands: A comparison of low and high nutrient riverine systems. Ecological Engineering 14: 77-91. Spieles, D.J. and W.J. Mitsch. 2000b. Macroinvertebrate community structure in highand low-nutrient constructed wetlands. Wetlands 20: 716-729. US EPA Test Methods for Evaluating Solid Wastes. 2005. Chapter 5 Miscellaneous Test Methods. Methods 9060A: Total Organic Carbon. SW-846. 3rd edition. 142

Vepraskas, M. J. and S. P. Faulkner. 2001. Redox chemistry of hydric soils. p. 85-105. In J. L. Richardson and M. J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes and Classification. Lewis Publishers, Boca Raton, FL, USA. Wetzel, R. G. 2001. Limnology: Lake and River Ecosystems, Third Edition. Academic Press, Inc., San Francisco, CA, USA. Wu, X. and W. J. Mitsch. 1998. Spatial and temporal patterns of algae in newly constructed freshwater wetlands. Wetlands 18: 9-20. Zhang, L. and W.J. Mitsch. 2005. Modelling hydrological processes in created wetlands: An integrated system approach. Environmental Modelling & Software 20: 935-946.

143

Figure 5.1 Pumping system and water control structures for the two experimental wetlands at The Olentangy River Wetland Research Park (ORWRP) at The Ohio State University in Columbus, Ohio.

144

Figure 5.2 The 10-m grid and locations used for measuring sediment depth and collecting samples at the ORWRP experimental wetlands in May 2004. Shaded areas within the grid map represent approximate location of the deeper, open water (OW) zones.

145

a)

Wetland 1

100 -2

Sediment (kg m )

Wetland 2 80 60 40 20 0 EM

OW

Total

b) Frequency (% of total area)

Wetland 1 0.4

Wetland 2

0.3 0.2 0.1 0 0

10

20

30

40

50

60

70

80

90

100

-2

Sediment accumulation (kg m )

Figure 5.3 a) Mean (±1 SE) sediment accumulation at the emergent vegetation (EM) zones, open water (OW) zones and total wetland and; b) frequency distribution based on spatially interpolated data from kriging analyses for Wetland 1 and 2. Total wetland sediment accumulation is derived based on weighted average. Kriging data for Wetland 1 and 2 were based on anisotropic analyses (0◦ and 135◦, respectively; see text for variogram results and further details).

146

110

19 0◦ 45◦ 90◦ Distance (10 m)

135◦

Sediment (kg m-2)

> 80.0 > 70.0 > 60.0 > 50.0 > 40.0 > 30.0 > 20.0 > 10.0

0 0

88

16

Distance (10 m)

Figure 5.4 Spatial distribution maps of sediment accumulations for Wetland 1 and 2. Maps were generated by ordinary point kriging using anisotropic variogram models (0◦ and 135◦, respectively). Degree bearings provided for reference.

147

a)

1

Moran's I

Isotropic Anisotropic (0 degree)

0

Anisotropic (45 degree) Anisotropic (135 degree)

-1 0

10

20

30

40

50

60

Mean distance (m)

b) 1

Moran's I

Isotropic Anisotropic (0 degree)

0

Anisotropic (45 degree) Anisotropic (135 degree)

-1 0

10

20

30

40

50

60

Mean distance (m)

Figure 5.5 Correlogram (Moran’s I over mean distance) for sediment accumulation in a) Wetland 1 and b) Wetland 2 for isotropic and anisotropic (0◦, 45◦ and 135◦) analyses.

148

a)

Wetland 1 Wetland 2

3500

-2

Total C (g m )

3000 2500 2000 1500 1000 500 0 EM

b)

OW

Wetland 1 Wetland 2

2500 -2

Organic C (g m )

Total

2000 1500 1000 500 0 EM

OW

Wetland 1 Wetland 2

1000 -2

Inorganic C (g m )

c)

Total

800 600 400 200 0 EM

OW

Total

Figure 5.6 Mean (±1 SE) accumulation of a) total C, b) organic C, and c) inorganic C for Wetland 1 and 2. Total wetland accumulation based on weighted average.

149

a)

Wetland 1 Wetland 2

300

-2

Total N (g m )

250 200 150 100 50 0 EM

b)

OW

Total

Wetland 1

70

Wetland 2

-2

Total P (g m )

60 50 40 30 20 10 0 EM

c)

OW

Wetland 1 Wetland 2

3000 2500 -2

Total Ca (g m )

Total

2000 1500 1000 500 0 EM

OW

Total

Figure 5.7 Mean (±1 SE) accumulation of a) total N, b) total P, and c) total Ca for Wetland 1 and 2. Total wetland accumulation based on weighted average.

150

151

OW Zone 0.53 ±0.02 (33) 3.9 ±0.1 (22) 3.5 ±0.1 (28) 0.48 ±0.08 (10) 0.36 ±0.01 (22) 722 ±29 (10) 14.3 ±1.7 (10)

706 ±27 (7) 5.0 ±0.5 (7)

Total

0.55 ±0.02 (22) 3.8 ±0.1 (17) 3.6 ±0.1 (18) 0.21 ±0.06 (9) 0.36 ±0.01 (17)

Wetland 1 EM Zone

34.4 ±6.4 (4)

806 ±28 (4)

0.48 ±0.04 (16) 3.8 ±0.4 (6) 2.9 ±0.1 (16) 0.97 ±0.19 (4) 0.31 ±0.03 (6)

OW Zone

5.4 ±0.5 (7)

670 ±37 (7)

0.50 ±0.03 (20) 3.9 ±0.2 (16) 3.9 ±0.2 (20) 0.07 ±0.02 (8) 0.32 ±0.02 (19)

Wetland 2 EM Zone

12.1 ±1.9 (11)

701 ±35 (11)

0.49 ±0.03 (36) 3.8 ±0.2 (24) 3.7 ±0.2 (36) 0.28 ±0.06 (12) 0.32 ±0.02 (25)

Total

151

Table 5.1 Mean (±1 SE) physiochemical conditions of sediment in the emergent vegetation (EM) and open water (OW) zones of the planted (Wetland 1) and naturally colonized (Wetland 2) wetlands at the Olentangy River Wetland Research Park in May 2004.

Total wetland values are weighted averages based on the relative proportion of OW and EM area at each wetland.

Values are means ±SE with sample size provided in parentheses.

Bulk density (g cm ) 0.44 ±0.03 (11) Total C (%) 4.3 ±0.1 (5) Organic C (%) 2.9 ±0.1 (10) Inorganic C (%) 1.40 ±0.14 (4) Total N (%) 0.36 ±0.01 (5) Total P (ug g-1) 782 ±36 (3) Total Ca (mg g-1) 47.2 ±6.1 (3)

-3

Sediment parameter

Mean annual accumulation rates

Sediment parameter Total sediment (kg m-2 yr-1) -2

4.5 - 4.9

-1

Total C (g m yr ) -2

180.9 - 192.9 -1

Organic C (g m yr ) -2

152.5 - 166.0

-1

Inorganic C (g m yr )

26.1 - 22.9

-2

-1

16.2 - 16.6

-2

-1

3.3 - 3.5

Total N (g m yr ) Total P (g m yr ) -2

-1

Total Ca (g m yr )

80.8 - 86.3

Table 5.2 Range of mean annual accumulation rates of sediment and nutrients for Wetland 1 and 2 at the Olentangy River Wetland Research Park, 1994-2004.

152

REFERENCES Ahn, C. and W.J. Mitsch. 2002. Scaling considerations of mesocosm wetlands in simulating large created freshwater marshes. Ecological Engineering 18: 327-342. Alden, H. A. 1995. Hardwoods of North America. U. S. Department of Agriculture, Forest Service, Madison, WI, USA. FPL-GTR-83. Almquist, E. B., S. B. Jack and M. G. Messina. 2002. Variation of the treefall gap regime in a bottomland hardwood forest: relationships with microtopograpy. Forest Ecology and Management 157:153-163. Anderson, C.J. and W.J. Mitsch. 2003. Open-water autotrophs: biomass and distribution in deepwater basins of two experimental wetlands. p. 41-44. In: W.J. Mitsch, L. Zhang and C.J. Anderson (eds.) Olentangy River wetland Research Park at The Ohio State University Annual Report 2002. The Ohio State University, Columbus, OH. Anderson, C.J. and B.C. Cowell. 2004. Mulching effects on the seasonally flooded zone of west-central Florida, USA wetlands. Wetlands 24:811-819. Anderson, C.J., W.J. Mitsch, and R.W. Nairn. 2005. Temporal and spatial development of surface soil conditions at two created riverine marshes. Journal of Environmental Quality 34(6): 2072-2081. Anderson, K.L., K.A. Wheeler, J.A. Robinson, and O.H. Tuovinen. 2002. Atrazine mineralizaation potential in two wetlands. Water Research 36: 4785-4794. AOAC Official Methods of Analysis. 1989. Method 990.03. Protein(crude) in Animal Feed Combustion Method (Dumas method). 17th edition 2002. Reference: JAOAC 72, 770 (1989). Armentano, T. V. and G. M. Woodwell. 1975. Sedimentation rates in a Long Island marsh determined by 210Pb dating. Limnol. Oceanogr. 20:452-455. 153

Axt, J. R. and M. R. Walbridge. 1999. Phosphate removal capacity of palustrine forested wetlands and adjacent uplands in Virginia. Soil Sci. Soc. Am. J. 63:1019-1031. Baker, D. B., R. P. Richards, T. F. Loftus, and J. W. Kramer. 2004. A new flashiness index: characteristics and applications to Midwestern rivers and streams. Journal of the American Water Resources Association 40:503-522. Baker, III, T. T., W. H. Conner, B. G. Lockaby, J. A. Stuart, and M. K. Burke. 2001. Fine root productivity and dynamics on a forested floodplain in South Carolina. Soil Science Society of America Journal 65:545-556. Bayley, S. E., J. Zoltek, Jr., A. J. Hermann, T. J. Dolan and L. Tortora. 1985. Experimental manipulation of nutrients and water in a freshwater marsh: effects on biomass, decomposition, and nutrient accumulation. Limnology and Oceanography 30:500-512. Bedford, B. L., M. R. Walbridge and A. Aldous. 1999. Patterns in nutrient availability and plant diversity of temperate North American wetlands. Ecology 80:2151-2169. Bishel-Machung, L., R.P. Brooks, S.S. Yates, and K.L. Hoover. 1996. Soil properties of reference wetlands and wetland creation projects in Pennsylvania. Wetlands 16:532541. Bolstad, P. V., J. M. Vose, and S. G. McNulty. 2001. Forest productivity, leaf area, and terrain in southern Appalachian deciduous forests. Forest Science 47:419-427. Braskerud, B. C. 2001. The influence of vegetation on sedimentation and resuspension of soil particles in small constructed wetlands. J. Environ. Qual. 30:1447-1457. Brenner, M., C. L. Schelske, and L. W. Keenan. 2001. Historical rates of sediment and nutrient accumulation in marshes of the Upper St. Johns River Basin, Florida, U.S.A. Journal of Paleolimnology 23:241-257. Brooks, R. P., D. H. Wardrop, C. A. Cole, and D. A. Campbell. 2005. Are we purveyors of wetland homogeneity? A model of degradation and restoration to improve wetland mitigation performance. Ecological Engineering 24:331-340. Brown, S. L. 1981. A comparison of the structure, primary productivity, and transpiration of cypress ecosystems in Florida. Ecological Monographs 51:405-415. Brown S. and D.L. Peterson. 1983. Structural characteristics and biomass production of two Illinois bottomland forests. American Midland Naturalist 110:107-117. 154

Brueske, C. C. and G. W. Barrett. 1994. Effects of vegetation and hydrologic load on sedimentation patterns in experimental wetland ecosystems. Ecological Engineering 3:429-447. Bruland, G.L. and C.J. Richardson. 2004. A spatially explicit investigation of phosphorus sorption and related soil properties in two riparian wetlands. J. Environ. Qual. 33:785-794. Burke, M. K., B. G. Lockaby, and W. H. Conner. 1999. Aboveground production and nutrient circulation along a flooding gradient in a South Carolina Coastal Plain forest. Can. J. For. Res. 29:1402-1418. Camill, P. and J. S. Clark. 2000. Long-term perspectives on lagged ecosystem responses to climate change: permafrost in boreal peatlands and the grassland/woodland boundary. Ecosystems 3:534-544. Campbell, D.A., C.A. Cole, and R.P. Brooks. 2002. A comparison of created and natural wetlands in Pennsylvania, USA. Wetlands Ecology and Management 10:41-49. Casanova,M. T. and M. A. Brock. 2000. How do depth, duration and frequency of flooding influence the establishment of wetland plant communities? Plant Ecology 147:237-250. Chadde, S. W. 2002. A Great Lakes Wetland Flora.- A complete guide to the aquatic and wetland plants of the upper Midwest. 2nd edition. PocketFlora Press, Laurium, MI, USA. Cochran, M. 2001. Effect of Hydrology on bottomland hardwood forest productivity in central Ohio (USA). M.S. Thesis. The Ohio State University, Columbus, OH, USA. Collins, M.E. and R.J. Kuehl. 2001. Organic matter accumulation and organic soils. In: J.L. Richardson and M.J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes and Classification. Lewis Publishers, Boca Raton, FL. Conner, W. H. 1994. Effect of forest management practices on southern forested wetland productivity. Wetlands 14: 27-40. Cooper, J. R. and J. W. Gilliam. 1987. Phosphorus redistribution from cultivated fields into riparian areas. Soil Sci. Soc. Am. J. 51:1600-1604. Cooperrider, T. S., A. W. Cusick and J. T. Kartesz (eds.). 2001. Seventh Catalog of the Vascular Plants of Ohio. Ohio State University Press. Columbus, Ohio. 155

Craft, C.B. 1997. Dynamics of nitrogen and phosphorus retention during wetland ecosystem succession. Wetlands Ecology and Management 4:177-187. Craft, C.B. 2001. Biology of wetland soils. In: J.L. Richardson and M.J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes and Classification. Lewis Publishers, Boca Raton, FL. Craft, C. B. and W. P. Casey. 2000. Sediment and nutrient accumulation in floodplain and depressional freshwater wetlands of Georgia, USA. Wetlands 20: 323-332. Craft, C. B., S. Broome, and C. Carlton. 2002. Fifteen years of vegetation and soil development after brackish-water marsh creation. Restoration Ecology 10: 248-258. Craft, C. B., P. Megonigal, S. Broome, J. Stevenson, R. Freese, J. Cornell, L. Zheng and J. Sacco. 2003. The pace of ecosystem development of constructed Spartina alterniflora marshes. Ecological Applications 13:1417-1432. Crawford, R. M. M. 1992. Oxygen availability as an ecological limit to plant distribution. Advances in Ecological Research 23:93-185. Cronk, J. K. and M. S. Fennessy. 2001. Wetland plants: biology and ecology. Lewis Publishers, Boca Raton, FL, USA. 462 p. Darke, A. K. and J. P. Megonigal. 2003. Control of sediment deposition rates in two mid-Atlantic Coast tidal freshwater wetlands. Estuarine, Coastal and Shelf Science 57: 255-268. Davis, S. M. 1991. Growth, decomposition, and nutrient retention of Cladium jamaicense Crantz and Typha domingensis Pers in Florida Everglades. Aquat. Bot. 40:203-224. DeBusk, W.F., S. Newman, and K.R. Reddy. 2001. Spatio-temporal patterns of soil phosphorus enrichment in Everglades Water Conservation Area 2A. J. Environ. Qual. 30:1438-1446. DeLaune, R. D. and W. H. Patrick, Jr. 1980. Rate of sedimentation and its role in nutrient cycling in a Louisiana salt marsh. p. 401-412. In P. Hamilton and K. B. Macdonald (eds.) Estuarine and wetland Processes with Emphasis on Modeling. Plenum Publishing, New York, NY. Dubbe, D. R, E. G. Graver, and D. C. Pratt. 1988. Production of cattail (Typha spp.) biomass in Minnesota, USA. Biomass 17:79-104. 156

Dudek, D.M., J.R. McClenahen and W.J. Mitsch. 1998. Tree growth responses of populus deltoides and Juglans nigra to streamflow and climate in a bottomland hardwood forest in central Ohio. American Midland Naturalist 140:233-244. Erwin, K. L. 1991. An evaluation of wetland mitigation in the South Florida Water Management District, Vol. I. Final Report to the South Florida Water Management District, West Palm Beach, FL, USA. Farnsworth, E. J. and L. A. Meyerson. 2003. Comparative ecophysiology of four wetland plant species along a continuum of invasiveness. Wetlands 23:750-762. Fennessy, M. S., C. C. Brueske, and W. J. Mitsch. 1994. Sediment deposition patterns in restored freshwater wetlands using sediment traps. Ecological Engineering 3:409428. Fennessy, M. S., J. K. Cronk and W. J. Mitsch. 1994. Macrophyte productivity and community development in created wetlands under experimental hydrological conditions. Ecological Engineering 3: 469-484. Fennessy, M.S. and W.J. Mitsch. 2001. Effects of hydrology and spatial patterns of soil development in created riparian wetlands. Wetlands Ecology and Management 9:103-120. Foster, J. R. 1992. Photosynthesis and water relations of the floodplain tree, boxelder (Acer negundo L.). Tree Physiol. 11:133-149. Galvez-Cloutier, R. and J. S. Dube. 1998. An evaluation of freshwater sediments contamination: the Lachine Canal sediments case, Montreal, Canada. I. Quality assessment. Water, Air and Soil Poll. 102:259-279. Gamma Design Software. 2004. GS+ User’s Guide Version 7: GeoStatistics for the Environmental Sciences. Gamma Design Software, LLC. Plainwell, MI. Garver, E. G., D. R. Dubbe, and D. C. Pratt. 1988. Seasonal patterns in accumulation and partitioning of biomass and macronutrients in Typha spp. Aquatic Botany 32:115127. Gaudet, C. L. and P. A. Keddy. 1995. Competitive performance and species distribution in shoreline plant communities: a competitive approach. Ecology 76:280-291. Gee, G.W. and J.W. Bauder. 1986. Particle-size analysis. p. 383–411. In A. Klute (ed.) Methods of soil analysis. Part 1. 2nd ed. Agron. Monogr. 9. ASA and SSSA, Madison, WI. 157

Giovannini, S. G. T. and D. M. L. Da Motta Marques. 1998. Establishment of three emergent macrophytes under different water regimes. Water Science & Technology 40:233-240. Grace, J. B. and J. S. Harrison. 1986. The biology of Canadian Weeds. 73. Typha latifolia L. , Typha angustifolia L., and Typha xglauca Godr. Canadian Journal of Plant Science 66:361-379. Grace, J.B. and R.G. Wetzel. 1981. Habitat partitioning and competitive displacement in cattails (Typha): experimental field studies. The American Naturalist 118:463-474. Harter, S. K. and W. J. Mitsch. 2003. Patterns of short-term sedimentation in a freshwater created marsh. J. Environ. Qual. 32:325-334. Hein, T., G. Heiler, D. Pennetzdorfer, P. Riedler, M. Schagerl, and F. Schiemer. 1999. The Danube Restoration Project: Functional aspects and planktonic productivity in the floodplain system. Regulated Rivers 15: 259-279. Hem, J.D. 1989. Study and Interpretation of the Chemical Characteristics of Natural Water, Third Edition. U.S. Geological Survey Water- Supply Paper 2254. Washington, DC. Horppila, J. and L. Nurminen. 2001. The effect of an emergent macrophyte (Typha angustifolia) on sediment resuspension in a shallow north temperate lake. Freshwater Biology 46:1447-1455. Hupp, C. R. and D. E. Bazemore. 1993. Temporal and spatial patterns of wetland sedimentation, West Tennessee. Journal of Hydrology 141:179-196. International Standard Organization, ISO. 1995. Soil quality – Determination of organic and total carbon after dry combustion (elementary analysis). International Organization for Standardization 10694:1995(E). Geneva, Switzerland. Isaac, R.A. and W.A. Johnson. 1985. Elemental Analysis of Plant Tissue by Plasma Emission Spectroscopy: Collaborative Study. J. Assoc. Off. Anal. Chem. 68:499. Isaaks, E.H. and R.M. Srivastava. 1989. An Introduction to Applied Geostatistics. Oxford University Press, Inc. New York, NY. Johnson, F. L. and D. T. Bell. 1976. Tree growth and mortality in the streamside forest. Castanea 41:34-41. 158

Johnston, C. A. 1991. Sediment and nutrient retention by freshwater wetlands: effects on surface water quality. Critical Reviews in Environmental Control 21:491-565. Johnston, C. A., S. D. Bridgham, and J. P. Schubauer-Berigan. 2001. Nutrient dynamics in relation to geomorphology of riverine wetlands. Soil Sci. Soc. Am. J. 65:557-577. Jones, T. A. and S. C. Thomas. 2004. The time course of diameter increment to selection harvests in Acer saccharum. Can. J. Res./Rev. Can. Rech. For. 34:15251533. Junk, W. J., P. B. Bayley, and R. E. Sparks. 1989. The flood pulse concept in riverfloodplain systems. In: D. P. Dodge (ed.). Proceedings of the International Large River Symposium. Special Issue of Journal of Canadian Fisheries and Aquatic Sciences 106:11-127. Jurinak, J. J. and R. A. Griffin. 1992. Nitrate ion adsorption by calcium carbonate. Soil Sci. 113: 130-135. Kang, H., C. Freeman, D. Lee, and W.J. Mitsch. 1998. Enzyme activities in constructed wetlands: Implication for water quality amelioration. Hydrobiologia 368:231-235. Kellogg, C. H., S. D. Bridgham, and S. A. Leicht. 2003. Effects of water level, shade and time on germination and growth of freshwater marsh plants along a simulated successional gradient. Journal of Ecology 91:274-282. Kimmins, J. P. 1987. Forest ecology. Macmillan Publishing Company. New York, NY, USA. Konen, M. E., P. M. Jacobs, C. L. Burras, B. J. Talaga, and J. A. Mason. 2002. Equations for predicting soil organic carbon using loss-on-ignition for north central U.S. soils. Soil Sci. Soc. Am. J. 66:1878-1881. Koerselman, W. and A. F. M. Meuleman. 1996. The vegetation N:P ratio: a new tool to detect the nature of nutrient limitation. The Journal of Applied Ecology 33:14411450. Koreny, J.S., W.J. Mitsch, E.S. Bair, and X. Wu. 1999. Regional and local hydrology of a created riparian wetland system. Wetlands 19:182-193. Kozlowski, T. T. 1997. Responses of woody plants to flooding and salinity. Tree Physiology Monograph 1:1-17.

159

Kozlowski, T. T. and S. G. Pallardy. 2002. Acclimation and adaptive responses of woody plants to environmental stresses. The Botanical Review 68:270-334. Kuo, S. 1996. Phosphorus. p. 894-895. In: Methods of Soil Analysis, Part 3—Chemical Methods. Soil Science Society of America, Madison, WI. Kvet, J. and S. Husak. 1978. Primary data on biomass and production estimates in typical stands offishpond littoral plant communities. p.211-216. In: Pond Littoral Ecosystems, D. Dykyjova and J. Kvet (eds.) Springer-Verlag, Berlin. Leopold, L. B., M. G. Wolman and J. E. Miller. 1964. Fluvial processes in geomorphology. W. H. Freeman, San Francisco, CA, USA. Liptak, M. 2000. Water column productivity, calcite precipitation, and phosphorus dynamics in freshwater marshes. PhD. Dissertation. The Ohio State University, Columbus, OH, USA. Martens, D. M. 1993. Hydrologic inferences from tree-ring studies on the Hawkesbury River, Sydney, Australia. Geomorphology 8:147-164. Martin, I and J. Fernandez. 1992. Nutrient dynamics and growth of a cattail crop (Typha latifolia L.) developed in an effluent with high eutrophic potential- application to wastewater purification systems. Bioresource Technology 42:7-12. Mason, C. F. and R. J. Bryant. 1975. Production, nutrient content and decomposition of Phragmites communis Trin. and Typha angustifolia L. Journal of Ecology 63:71-95. Mcloda, N. A. and R. J. Parkinson. 1980. Soil survey of Franklin County, Ohio. USDASCS. US Government Printing Office, Washington, DC, USA. Megonigal, J.P., W.H. Conner, S. Kroeger and R.R. Sharitz. 1997. Aboveground production in southeastern floodplain forests: a test of the subsidy-stress hypothesis. Ecology 78:370-384. Metzker, K. and W.J. Mitsch. 1997. Modelling self-design of the aquatic community in a newly created freshwater wetland. Ecological Modelling 100:61-86. Mitsch, W.J. and K.C. Ewel. 1979. Comparative biomass and growth of cypress in Florida wetlands. American Midland Naturalist 101:417-426. Mitsch, W. J. and W. G. Rust. 1984. Tree growth responses to flooding in a bottomland forest in northeastern Illinois. Forest Science 30: 499-510. 160

Mitsch, W. J., J. R. Taylor and K. B. Benson. 1991. Estimating primary productivity of forested wetland communities in different hydrologic landscapes. Landscape Ecology 5:75-92. Mitsch, W. J., J. K. Cronk, X. Wu, and R. W. Nairn. 1995. Phosphorus retention in constructed freshwater riparian marshes. Ecological Applications 5: 830-845. Mitsch, W. J. and R. F. Wilson. 1996. Improving the success of wetland creation and restoration with know-how, time, and self-deign. Ecological Applications 6:77-83. Mitsch, W.J., X. Wu, R.W. Nairn, P.E. Weihe, N. Wang, R. Deal, and C.E. Boucher. 1998. Creating and restoring wetlands: A whole-ecosystem experiment in self-design. BioScience 48:1019-1030. Mitsch, W.J. and J.G. Gosselink. 2000. Wetlands, Third Edition. John Wiley & Sons, Inc., New York, NY. Mitsch, W. J., C. J. Anderson, M. E. Hernandez and L. Zhang. 2003. Net primary productivity of macrophyte communities after nine growing seasons in experimental planted and unplanted marshes. p 31-36. In: W. J. Mitsch, L. Zhang and C. J. Anderson (eds.) The Olentangy River Wetland Research Park Annual Report 2002. The Ohio State University, OH, USA. Mitsch, W.J. and L. Zhang. 2004a. Plant community development after ten growing seasons in two experimental wetland basins. p. 69-73. In: W.J. Mitsch, L. Zhang and C.L. Tuttle (eds.) Olentangy River Wetland Research Park at The Ohio State University Annual Report 2003. The Ohio State University, Columbus, OH. Mitsch, W. J. and L. Zhang. 2004b. Wetland monitoring of the bottomland hardwood forest at the Olentangy River Wetland Research Park (Year 3 – 2003). p.137-147. In W.J. Mitsch, L. Zhang and C. Tuttle (eds.) Olentangy River Wetland Research Park at The Ohio State University, Annual Report 2003. Columbus, OH, USA. Mitsch, W. J., C. J. Anderson, M. E. Hernandez, A. Altor and L. Zhang. 2004c. Net primary productivity of macrophyte communities after ten growing seasons in experimental marshes. p. 75-78. In W. J. Mitsch, L. Zhang and C. L. Tuttle (eds.) Olentangy River wetland Research Park at The Ohio State University Annual Report 2003. The Ohio State University, Columbus, OH, USA. Mitsch, W. J., L. Zhang, N. Dillon, and D. Fink. 2004d. Biogeochemical patterns of created riparian wetlands: tenth-year results (2003). p. 59-68. In W. J. Mitsch, L. Zhang and C. L. Tuttle (eds.) Olentangy River wetland Research Park at The Ohio 161

State University Annual Report 2003. The Ohio State University, Columbus, OH, USA. Mitsch, W.J., J.W. Day, Jr., L. Zhang, and R. Lane. 2005a. Nitrate-nitrogen retention by wetlands in the Mississippi River Basin. Ecological Engineering 24: 267-278. Mitsch, W. J., N. Wang, L. Zhang, R. Deal, X. Wu, and A. Zuwerink. 2005b. Using ecological indicators in a whole-ecosystem wetland experiment p. 211-236. In S. E. Jorgensen, R. Costanza and F. L. Xu (eds.) Handbook of Ecological Indicators for Assessment for Ecosystem Health. Lewis Publishers, CRC Press, Inc., Boca Raton, FL, USA. Mitsch, W.J., L. Zhang, C.J. Anderson, A. Altor, and M. Hernandez. 2005c. Creating riverine wetlands: Ecological succession, nutrient retention, and pulsing effects. Ecological Engineering (in press). Motivans, K. and S. Apfelbaum. 1987. Element stewardship abstract for Typha spp., North American cattails. The Nature Conservancy, Arlington, VA, USA. Nair, V.D., D.A. Graetz, K.R. Reddy, and O.G. Olila. 2001. Soil development in phosphate-mined created wetlands of Florida, USA. Wetlands 21:232-239. Nairn, R. W., 1996. Biogeochemistry of newly created riparian wetlands: evaluation of water quality changes and soil development. PhD. Dissertation. The Ohio State University, Columbus, OH, USA. Nairn, R.W. and W.J. Mitsch. 2000. Phosphorus removal in created wetland ponds receiving river overflow. Ecological Engineering 14: 107-126. Newbould, J. 1978. Methods for estimating the primary production of forests. Blackwell, Oxford, England. Newman, S. and K. Pietro. 2001. Phosphorus storage and release in response to flooding: implications for Everglades stormwater treatment areas. Ecological Engineering 18:23-38. Newman, S., J. Schuette, J.B. Grace, K. Rutchey, T. Fontaine, K.R. Reddy, and M. Pietrucha. 1998. Factors influencing cattail abundance in the northern Everglades. Aquatic Botany 60:265-280. Odum, E.P., J.T. Finn and E.H. Franz. 1979. Perturbation theory and the subsidy-stress gradient. Bioscience 29:344-352. 162

Olila, O. G., K. R. Reddy, and D. L. Stites. 1997. Influence of draining on soil phosphorus forms and distribution in a constructed wetland. Ecological Engineering 9:157-169. Pasternack, G. B. and G. S. Brush. 1996. Sedimentation cycles in a river-mouth tidal freshwater marsh. Estuaries 21:407-415. Perry, J. N., A. M. Liebhold, M. S. Rosenberg, J. Dungan, M. Miriti, A. Jakomulska and S. Citron-Pousty. 2002. Illustrations and guidelines for selecting statistical methods for quantifying spatial patterns in ecological data. Ecography 25: 578-600. Peterjohn, W. T. and D. L. Correll. 1984. Nutrient dynamics in an agricultural watershed; observations on the role of a riparian forest. Ecol. 65: 1466-1475. Phipps, R.L. 1979. Simulation of wetlands forest vegetation dynamics. Ecological Modelling 7:257-288. Pollock, M. M., R. J. Naiman, and T. A. Hanley. 1998. Plant species richness in riparian wetlands- a test of biodiversity theory. Ecology 79:94-105. Ponnamperuma, F.N. 1972. The chemistry of submerged soils. Advances in Agronomy 24:29-96. Prentki, R. T., T. D. Gustafson, and M. S. Adams. 1978. Nutrient movements in lakeshore marshes, p. 169-194. In: R. E. Good, D. F. Whigham, and R. L. Simpson (eds.). Freshwater wetlands ecological processes and management potential. Academic Press, New York, NY, USA. Reddy, K. R., R. D. DeLaune, W. F. DeBusk and M. S. Koch. 1993. Long-term nutrient accumulation rates in the Everglades. Soil Sci. Soc. Am. J. 57:1147-1155. Regent Instruments, Inc. 2002. WinDENDRO 2002 a,b. Regent Instruments, Inc., Quebec, Canada. Richardson, C. J. and C. B. Craft. 1993. Effective phosphorus retention in wetlands: fact or fiction? p. 271-282. In G. A. Moshiri (ed.) Constructed Wetlands for Water Quality Improvement. Lewis Publishers, CRC Press, Inc., Boca Raton, FL, USA. Robertson, A.I., P.Y. Bacon, and G. Heagney. 2001. The response of floodplain primary production to flood frequency and timing. Journal of Applied Ecology 38:126-136.

163

Sanyal, L.K. and L.K. De Datta. 1991. Chemistry of phosphorus transformations in soil. In: B. A. Stewart (ed.) Advances in Soil Science, Volume 16, Springer-Verlag, New York, NY. Selbo, S.M. and A.A. Snow. 2004. The potential for hybridization between Typha angustifolia and Typha latifolia in a constructed wetland. Aquatic Botany 78: 361369. Shaffer, P.W. and T.L. Ernest. 1999. Distribution of soil organic matter in freshwater emergent/open water wetlands in the Portland, Oregon metropolitan area. Wetlands 19:505-516. Sommers, L. E. and D. W. Nelson. 1972. Determination of Total Phosphorus in Soils: A Rapid Perchloric Acid Digestion Procedure. Soil Science Society of America Proceedings. Vol. 36, 6: p.902-904. Madison, WI. Spieles, D.J. and W.J. Mitsch. 2000a. The effects of season and hydrologic and chemical loading on nitrate retention in constructed wetlands: A comparison of low and high nutrient riverine systems. Ecological Engineering 14: 77-91. Spieles, D.J. and W.J. Mitsch. 2000b. Macroinvertebrate community structure in highand low-nutrient constructed wetlands. Wetlands 20: 716-729. Spieles, D.J. and W.J. Mitsch. 2003. A model of macroinvertebrate trophic structure and oxygen demand in freshwater wetlands. Ecological Modelling 161: 183-194. Spink, A., R. E. Sparks, M. Van Oorschot and J. T. A. Verhoeven. 1998. Nutrient dynamics of large river floodplains. Regulated Rivers: Research & Management 14:203-216. Svengsouk, L.M. and W. J. Mitsch. 2001. Dynamics of mixtures of Typha latifolia and Schoenoplectus tabernaemontani in nutrient-enrichment wetland experiments. American Midland Naturalist 145:309-324. Svensgouk, L. 1998. First-year response of Typha latifolia L. and Schoenoplectus tabernaemontani (C.C. Gmel.) Palla to nitrogen and phosphorus additions in experimental mesocosms. M.S. Thesis, The Ohio State University, Columbus, OH, USA. 128pp. Systat Software, Inc. 2002. Systat for Windows. Version 10.2. Systat Software, Inc. Richmond, CA.

164

Tanner, C. C. 2001. Growth and nutrient dynamics of soft-stem bulrush in constructed wetlands treating nutrient-rich wastewaters. Wetlands Ecology and Management 9:49-73. Tanner, C. C., J. D’Eugenio, G. B. McBride, J. P. S. Sukias and K. Thompson. 1999. Effect of water level fluctuation on nitrogen removal from constructed wetland mesocosms. Ecological Engineering 12:67-92. Taylor, J.R., M.A. Cardamone and W.J. Mitsch. 1990. Bottomland hardwood forests: their function and values. p. 14-34. In: J.G. Gosselink, L.C. Lee and T.A. Muir (eds.) Ecological Processes and Cumulative Impacts Illustrated by Bottomland Hardwood Wetland Ecosystems. Lewis, Chelsea, MI, USA. Teskey, R. O. and T. M. Hinckley. 1977a. Impact of water level changes on woody riparian and wetland communities, Vol. I: plant and soil response. U. S. Fish and Wildlife Service, Columbia, MO, USA. FWS/OBS-77/58. Teskey, R. O. and T. M. Hinckley. 1977b. Impact of water level changes on woody riparian and wetland communities, Vol. III: the central forest region. U. S. Fish and Wildlife Service, Columbia, MO, USA. FWS/OBS-77/60. Thomas, G.W. 1996. Soil pH and soil acidity. p. 475-490. In: Methods of Soil Analysis, Part 3—Chemical Methods. Soil Science Society of America, Madison, WI. Tiner, R. W. 1999. Wetland Indicators- A Guide to Wetland Identification, Delineation, Classification, and Mapping. Lewis Publishers, Boca Raton, FL. Tockner, K., F. Malard and J.V. Ward. 2000. An extension of the flood pulse concept. Hydrological Process 14:2861-2883. U.S. Environmental Protection Agency. 1983. Methods for chemical analysis of water and wastes. 600/4-79-020, U.S. Environmental Protection Agency, Cincinnati, OH, USA. U.S. Environmental Protection 2005. Agency test methods for evaluating solid wastes. Chapter 5: miscellaneous test methods. Methods 9060A: Total Organic Carbon. SW846. 3rd edition. Washington, DC, USA. U.S. Forests Products Laboratory. 1974. Wood handbook: wood as an engineering material. USDA Agriculture Handbook No. 72. Washington, DC, USA. van der Valk, A. G. and C. B. Davis. 1978. The role of seed banks in the vegetation dynamics of prairie glacial marshes. Ecology 59: 322-335. 165

Vepraskas, M. J. and S. P. Faulkner. 2001. Redox chemistry of hydric soils. p. 85-105. In J. L. Richardson and M. J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes and Classification. Lewis Publishers, Boca Raton, FL, USA. Warncke, D. and J.R. Brown. 1998. Potassium and other basic cations, p. 31-33. In: Recommended Chemical Soil Test Procedures for the North Central Region. NCR Publication No. 221. Missouri Agricultural Experiment Station, Columbia, MO, USA. Waters, I. and J. M. Shay. 1992. Effect of water depth on population parameters of a Typha glauca stand. Canadian Journal of Botany 70:349-351. Wetzel, R.G. 2001. Limnology: lake and river ecosystems, third edition. Academic Press, Inc., San Francisco, CA. Whittaker, R.H. and G.M. Woodwell. 1968. Dimension and production relations of trees and shrubs in the Brookhaven Forest, New York. Journal of Ecology 57:155-174. Wilson, R. F. and W. J. Mitsch. 1996. Functional assessment of five wetlands constructed to mitigate wetland loss in Ohio. Wetlands 16:436-451. Wu, X. and W.J. Mitsch. 1998. Spatial and temporal patterns of algae in newly constructed freshwater wetlands. Wetlands 18: 9-20. Zhang, L. and W.J. Mitsch. 2005. Modelling hydrological processes in created wetlands: An integrated system approach. Environmental Modelling & Software 20: 935-946. Zhang, L., W. J. Mitsch, V. Bouchard and K. Hossler. 2005. Sediment chemistry in a hydrologically restored bottomland hardwood forest in Midwestern U.S. Program and abstracts, restoration and design of ecosystems, fifth annual meeting, American Ecological Engineering Society, The Ohio State University, Columbus, OH, USA. Zedler, J.B. and J.C. Callaway. 1999. Tracking wetland restoration: do mitigation sites follow desired trajectories? Restoration Ecol. 7:69-73.

166

APPENDIX A WETLAND MESOCOSM PLANT AND RIVER DATA (2002-2003)

167

168

Hydrology Pulsed Pulsed Steady-flow Steady-flow Pulsed Pulsed Steady-flow Pulsed Steady-flow Steady-flow Pulsed Steady-flow Pulsed Steady-flow Steady-flow Pulsed Steady-flow

Species Typ. Sch. Typ. Sch.* Sch.* Typ. Typ. Typ. Sch. Sch. Sch.* Sch. Typ. Typ. Typ. Typ. Sch.

Senescent 0 173 0 59 24 0 0 0 335 431 70 69 0 0 0 0 100

Live 1125 348 917 356 261 1085 994 999 448 270 331 334 905 1046 1170 919 414

Total

Biomass (g m ) Biomass (g m-2) R:S ratio 2010 3135 1.79 793 1314 1.52 1274 2191 1.39 720 1135 1.73 608 893 2.14 1296 2381 1.19 1583 2577 1.59 1727 2725 1.73 909 1693 1.16 1086 1787 1.55 400 801 1.00 715 1117 1.78 1616 2521 1.79 1101 2146 1.05 1591 2762 1.36 1106 2024 1.20 1149 1663 2.24

-2

Belowground

168

Table A.1. Aboveground biomass, belowground biomass, total biomass and root:shoot (R:S) ratio estimated from the experimental Schoenoplectus (Sch.) and Typha (Typ.) mesocosms in August 2003.

* Schoenoplectus (1-yr) mesocosms

Plot 1 2 3 5 6 7 9 10 11 12 13 14 15 16 18 19 20

Aboveground biomass (g m-2)

Species group Schoenoplectus (1-yr) Schoenoplectus (2-yr) Typha

Hydrology Pulsed Steady-flow Pulsed Steady-flow Pulsed Steady-flow

Mean stem/ramet height (cm) per plot (m-2) Sept 02 Jun 03 Jul 03 Aug 03 -72.8 ±5.9 84.5 ±9.7 87.3 ±4.3 -67.4 82.2 85.8 61.1 101.6 108.8 96.3 75.7 ±7.0 107.3 ±5.9 123.2 ±3.2 101.8 ±3.5 88.4 ±7.6 134.5 ±8.9 146.6 ±6.9 155.9 ±5.1 101.6 ±8.9 136.2 ±6.4 140.4 ±8.6 157.0 ±6.4

Table A.2. Mean (±1 SE) stem/ramet density for experimental mesocosm in 20022003.

Species group Schoenoplectus (1-yr) Schoenoplectus (2-yr) Typha

Hydrology Pulsed Steady-flow Pulsed Steady-flow Pulsed Steady-flow

Sept 02 --2 20 ±10 0 0

No. of inflorescence (m-2) Jun 03 Jul 03 105 ±11 146 ±17 126 224 143 208 209 ±25 206 ±18 7.1 ±2.6 7.6 ±2.7 6.4 ±2.8 6.4 ±2.8

Aug 03 77 ±3 158 97 82 ±10 6.9 ±2.7 6.4 ±2.8

Table A.3. Mean (±1 SE) number of inflorescences for experimental mesocosms in 2002-2003.

169

Species group Schoenoplectus (1-yr) Schoenoplectus (2-yr) Typha

Hydrology Pulsed Steady-flow Pulsed Steady-flow Pulsed Steady-flow

Maximum stem/ramet height (cm) per plot (m-2) Sept 02 Jun 03 Jul 03 Aug 03 -123.0 ±11.0 113.7 ±19.1 122.9 ±7.7 -117.0 126.4 122.8 105.0 141.0 135.4 139.2 109.7 ±4.9 145.0 ±3.3 141.6 ±4.6 136.2 ±2.4 126.0 ±2.3 178.2 ±4.5 189.0 ±2.2 187.8 ±2.1 128.2 ±2.6 175.0 ±5.4 188.1 ±2.0 188.4 ±1.9

Table A.4. Mean (±1 SE) maximum stem height length based on the measured length of the five longest stems/ramets at each mesocosm plot.

Species group Schoenoplectus (1-yr) Schoenoplectus (2-yr) Typha

Hydrology Pulsed Steady-flow Pulsed Steady-flow Pulsed Steady-flow

Mean stem/ramet height (cm) per plot (m-2) Sept 02 Jun 03 Jul 03 Aug 03 -72.8 ±5.9 84.5 ±9.7 87.3 ±4.3 -67.4 82.2 85.8 61.1 101.6 108.8 96.3 75.7 ±7.0 107.3 ±5.9 123.2 ±3.2 101.8 ±3.5 88.4 ±7.6 134.5 ±8.9 146.6 ±6.9 155.9 ±5.1 101.6 ±8.9 136.2 ±6.4 140.4 ±8.6 157.0 ±6.4

Table A.5. Mean (±1 SE) stem height length based on the measured length of 12 randomly selected stems/ramets at each mesocosm plot.

170

171

-1

(%) (μg g ) 1.488 1325 1.589 1308 1.057 901 1.539 1264 1.396 1826 1.296 1017 1.213 1032 0.966 824 1.891 1433 1.228 1324 1.246 947 1.225 1221 1.032 1050 1.329 1627 1.087 1051 1.539 1240 1.377 964 1.637 1437 1.014 813 0.994 1087 0.880 1016 1.058 1356 1.423 1355 1.239 1545 1.072 780

P -1

(μg g ) 14087 17254 9311 11343 18849 10290 17517 10367 14270 14883 11123 19635 15735 19911 14408 16198 9410 22107 11625 10651 13902 16515 15972 23997 14011

K -1

(μg g ) 9656 5089 6861 9389 6207 9252 5951 6738 10616 10577 11556 3519 5736 2948 4394 5162 8151 4322 5717 10947 8525 9072 9325 3408 6546

Ca (μg g ) 1088 553 504 972 783 698 771 735 992 943 1202 556 680 480 411 714 870 511 541 1001 911 947 903 616 825

-1

Mg -1

(μg g ) 14.8 58.4 132.2 28.0 49.2 47.1 23.7 158.7 81.0 31.9 28.5 39.1 49.1 11.5 32.4 186.7 105.9 20.8 58.4 23.1 17.3 30.1 21.6 50.9 63.7

Al -1

Cu -1

Fe -1

-1

Mn -1

Mo

Na -1

Zn -1

(μg g ) (μg g ) (μg g ) (μg g ) (μg g ) (μg g ) (μg g ) 12.7 2.8 26 256 0.8 2748 7.1 7.2 1.6 70 249 0.6 207 7.4 10.0 0.9 119 323 1.5 188 6.6 7.7 2.9 36 253 0.3 2900 7.4 7.4 1.3 77 523 3.6 102 8.9 12.2 0.6 67 433 7.1 155 6.4 5.8 1.2 38 302 1.7 195 6.9 9.1 1.3 153 341 1.7 214 8.8 11.3 2.7 67 270 0.3 970 8.1 8.0 2.9 37 241 0.6 3422 6.4 8.0 5.4 35 276 0.3 3877 10.0 9.1 2.1 48 203 0.4 223 5.8 7.4 1.3 73 242 0.7 709 6.1 6.6 1.8 27 203 0.3 148 6.5 7.7 1.2 45 221 1.0 303 5.8 8.4 1.2 150 344 1.0 171 6.2 10.5 0.9 139 392 3.5 399 7.8 7.0 1.2 32 244 0.6 229 6.4 10.5 0.8 72 252 0.7 413 6.1 7.8 2.6 30 227 0.3 4900 7.6 6.8 2.6 22 194 0.4 2965 6.0 7.2 3.9 34 223 0.9 3289 8.0 8.1 4.0 30 219 1.1 2665 6.6 6.0 2.0 56 184 0.4 333 7.3 10.9 0.9 91 265 0.7 409 6.1

B

171

Table A.6. Nutrient concentrations of plant tissue (live and senescent) collected from experimental Schoenoplectus (Sch.) and Typha (Typ.) mesocosms. All plant biomass were harvested in August 2003.

Plot Species Hydrology Biomass 1 Typ. Pulsed Live 2 Sch. Pulsed Live 2 Sch. Pulsed Senescent 3 Typ. Steady-flow Live 5 Sch.* Steady-flow Live 5 Sch.* Steady-flow Senescent 6 Sch.* Pulsed Live 6 Sch.* Pulsed Senescent 7 Typ. Pulsed Live 9 Typ. Steady-flow Live 10 Typ. Pulsed Live 11 Sch. Steady-flow Live 11 Sch. Steady-flow Senescent 12 Sch. Steady-flow Live 12 Sch. Steady-flow Senescent 13 Sch.* Pulsed Live 13 Sch.* Pulsed Senescent 14 Sch. Steady-flow Live 14 Sch. Steady-flow Senescent 15 Typ. Pulsed Live 16 Typ. Steady-flow Live 18 Typ. Steady-flow Live 19 Typ. Pulsed Live 20 Sch. Steady-flow Live 20 Sch. Steady-flow Senescent * Schoenoplectus (1-yr) mesocosms.

N

172

3/31-4/6 4/7-4/14 4/15-4/22 4/23-4/30 5/1-5/7 5/8-5/14 5/15-5/21 5/22-5/30 6/1-6/7 6/8-6/14 6/15-6-21 6/22-6/29 6/30-7/6 7/7-7/13 7/14-7/20 7/21-7/27 7/28-8/3 8/3-8/10 8/11-8/18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

-1

-1

Nutrient conc. (mg L ) P NO3-N nd 5.03 nd nd nd 4.43 0.131 3.30 0.241 4.08 0.316 6.58 nd 4.04 0.234 3.19 0.088 4.27 0.047 5.86 0.138 6.32 0.051 3.02 nd 0.34 0.351 3.81 0.069 3.85 0.259 2.81 0.046 2.64 0.053 1.79 0.063 1.86 Total Mean SE 19 week estimate (= Mean * 19)

Pumping rate (L wk ) Pulsed Steady-flow 878 356 208 443 216 413 185 401 924 379 159 333 68 68 0 0 0 0 223 193 182 386 965 337 178 428 193 413 197 390 284 284 284 284 284 284 284 284

-2

P input (mg P m ) Pulsed Steady-flow nd nd nd nd nd nd 27 58 245 100 55 116 nd nd 0 0 0 0 12 10 28 59 54 19 nd nd 75 159 15 30 81 81 14 14 17 17 20 20 641 682 53 57 19 14 1015 1079

172

-2

NO3-N input (mg N m ) Pulsed Steady-flow 4859 1969 nd nd 1051 2010 673 1456 4145 1699 1151 2411 303 303 0 0 0 0 1439 1244 1263 2684 3206 1119 67 160 809 1729 834 1651 877 877 824 824 559 559 581 581 22641 21277 1509 1418 349 187 28679 26950

Table A.7. Olentangy River water P and NO3-N concentration and input into wetland mesocosms during the 2003 experimental wet season.

nd = no data available

Dates

Week

APPENDIX B BOTTOMLAND HARDWOOD FOREST DATA (2004-2005)

173

Species Acer negundo L. (boxelder) Acer saccharinum L. (silver maple) Acer saccharum Marsh. (sugar maple) Aesculus glabra Willd. (Ohio buckeye) Asimina triloba (L.) Dunal (paw paw) Celtis occidentalis Willd. (hackberry) Fraxinus pennsylvanica Marsh. (green ash) Juglans nigra L. (black willow) Maclura pomifera (Raf.) (osage-orange) Morus alba L. (white mulberry) Morus rubra L. (red mulberry) Platanus occidentalis L. (sycamore) Populus deltiodes Bartr. Ex (cottonwood) Ulmus americana L. (American elm) Total

Impt Val. Rel. Den. Rel. Dom. Rel. Freq. 36.6 10.3 13.0 13.3 15.0 3.4 8.2 3.3 7.1 1.1 2.6 3.3 48.5 22.9 12.2 13.3 68.0 46.6 8.1 13.3 46.1 6.5 26.3 13.3 3.9 0.4 0.2 3.3 13.8 1.1 6.0 6.7 3.8 0.4 0.1 3.3 8.8 1.5 0.7 6.7 8.7 1.5 0.5 6.7 18.9 1.5 10.7 6.7 11.2 2.3 5.6 3.3 9.5 0.4 5.8 3.3 300.0 100.0 100.0 100.0

Table B.1. Species, importance value, relative density, relative dominance and relative frequency of trees (>5cm dbh) observed in plots at the north section of the bottomland forest.

174

Species Acer negundo L. (boxelder) Aesculus glabra Willd. (Ohio buckeye) Populus deltiodes Bartr. Ex (cottonwood) Platanus occidentalis L. (sycamore) Gleditsia triacanthos L. (honey locust) Acer saccharum Marsh. (sugar maple) Acer saccharinum L. (silver maple) Celtis occidentalis Willd. (hackberry) Salix nigra L. (black willow) Morus alba L. (white mulberry) Lonicera maackii (Rupr.) Amur honeysuckle Morus rubra L. (red mulberry) Fraxinus pennsylvanica Marsh. (green ash) Ulmus americana L. (American elm) Prunus serotina Ehrh. (black cherry) Juglans nigra L. (black walnut) Total

Impt Val. Rel. Den. Rel. Dom. Rel. Freq. 94.3 48.5 26.5 19.2 51.1 27.6 8.1 15.4 41.3 3.0 30.6 7.7 20.4 1.5 11.2 7.7 16.0 2.2 6.0 7.7 9.2 3.0 2.4 3.8 8.9 2.2 2.8 3.8 8.1 2.2 2.0 3.8 7.5 0.7 2.9 3.8 7.1 0.7 2.5 3.8 7.0 3.0 0.1 3.8 6.8 2.2 0.7 3.8 6.2 0.7 1.6 3.8 6.0 0.7 1.4 3.8 5.8 0.7 1.2 3.8 4.6 0.7 0.0 3.8 300.0 100.0 100.0 100.0

Table B.2. Species, importance value, relative density, relative dominance and relative frequency of trees (>5cm dbh) observed in plots at the south section of the bottomland forest.

175

176

22.3 ±3.1 25.3 ±2.7

9.5 ±2.3 14.0 ±1.9

North South

43.5 ±11.3 42.5 ±12.0

Aug 19.6 45.1 35.7 73.6 32.3 83.4 55.9 31.1 19.8 22.4

Oct 251.7 193.7 239.1 123.7 204.0 236.2 252.1 275.7 237.8 224.6

99.9 ±11.8 202.1 ±28.9 87.7 ±20.3 245.3 ±8.8

Sept 104.1 107.5 66.2 121.7 83.5 160.0 80.3 58.4 97.2 42.7 25.7 ±5.2 53.6 ±6.9

1.7 ±1.1 5.6 ±2.2

0.8 ±0.8 0.9 ±0.5

Feb 0.0 0.0 0.4 0.0 0.0 0.6 0.1 0.4 0.0 0.0 0.1 ±0.1 0.2 ±0.1

Mean cumulative leaf litter (g dry weight m-2) Nov Dec Jan 26.9 4.8 0.0 39.8 0.2 0.0 18.0 1.9 3.0 17.9 0.1 0.0 93.4 19.9 0.4 41.5 10.1 0.3 54.2 2.7 0.6 75.2 11.9 2.9 36.5 2.0 0.2 60.5 1.6 0.3 0.0 ±0.0 0.0 ±0.0

Mar 0.0 0.0 0.0 0.0 0.0 0.1 0.0 0.0 0.0 0.0

0.0 ±0.0 0.0 ±0.0

Apr 0.0 0.0 0.0 0.0 0.0 0.2 0.0 0.0 0.0 0.0

4.9 ±0.6 9.2 ±2.5

May 3.6 4.5 4.8 6.6 14.9 4.4 3.6 15.9 7.8 14.0

410.4 ±9.7 484.3 ±28.7

Total 434.0 415.5 403.9 388.1 472.4 571.6 498.3 508.6 432.5 410.8

176

Table B.3. Cumulative mean leaf litter biomass and monthly section mean (±1 SE) collected in the bottomland hardwood forest leaf traps between June 2003 and May 2004.

Jul 14.9 19.6 25.9 28.6 14.7 26.2 28.6 23.7 15.9 31.9

Jun 8.3 5.1 8.7 15.9 9.6 8.7 20.1 13.4 15.2 12.7

Plot/ Section 1 2 3 4 5 6 7 8 9 10

177

3.3 ±1.9 5.6 ±2.8

7.0 ±1.3 17.7 ±6.0

North South

4.1 ±2.5 1.8 ±0.9

Aug 0.7 11.5 3.5 0.7 4.1 2.6 0.2 5.1 1.1 0.0 3.1 ±1.9 2.5 ±1.1

Sept 0.0 3.8 0.1 8.2 0.8 0.9 6.6 3.2 1.5 0.5 4.6 ±3.8 3.8 ±2.2

2.3 ±1.2 2.8 ±1.1

2.8 ±2.4 2.1 ±0.8

0.8 ±0.7 1.8 ±0.9

0.1 ±0.0 1.1 ±0.4

Mean cumulative reprodtuvive material (g dry weight m-2) Oct Nov Dec Jan Feb 0.0 0.8 1.2 0.0 0.0 2.5 1.8 0.3 0.1 0.1 0.1 0.7 0.0 0.0 0.0 15.8 5.7 9.9 3.0 0.2 0.5 0.2 1.6 1.4 0.4 1.7 0.4 1.8 0.8 0.7 0.1 1.1 1.4 2.3 1.0 2.7 3.3 5.2 5.0 1.7 12.6 6.6 0.7 0.3 0.1 2.0 2.9 1.6 0.5 2.2 0.1 ±0.0 1.6 ±0.8

Mar 0.1 0.1 0.1 0.0 0.3 1.4 0.2 1.8 0.1 4.4

9.2 ±3.2 12.9 ±3.2

Apr 16.8 11.7 2.7 5.4 6.0 10.6 24.9 9.2 6.9 13.3

12.5 ±2.3 16.9 ±4.9

May 18.8 10.2 12.5 8.5 17.3 16.0 32.9 13.7 2.8 19.0

49.9 ±5.1 70.8 ±7.9

Total 47.8 49.2 38.8 63.6 71.3 77.5 75.1 77.3 39.9 84.2

177

Table B.4. Cumulative mean reproductive material biomass and monthly section mean (±1 SE) collected in the bottomland hardwood forest leaf traps between June 2003 and May 2004.

Jul 1.1 1.9 9.0 1.5 16.1 16.0 0.9 6.6 2.0 2.5

Jun 8.2 5.1 10.1 4.7 22.5 24.6 3.6 19.7 5.1 35.3

Plot/ Section 1 2 3 4 5 6 7 8 9 10

178

13.9 ±1.2 7.7 ±2.5

13.6 ±0.3 9.0 ±1.7

North South

4.2 ±0.6 5.8 ±2.5

Aug 2.4 4.0 5.1 5.2 3.2 2.6 6.4 4.6 0.3 15.0 8.4 ±3.8 3.8 ±1.2

Sept 0.1 6.1 18.1 9.5 7.3 7.5 2.9 2.4 0.6 5.5 10.7 ±3.7 5.7 ±2.2

8.3 ±2.8 10.5 ±3.4

7.0 ±4.4 10.4 ±4.2

1.0 ±0.4 2.8 ±1.2

3.3 ±2.5 1.8 ±0.4

Mean cumulative woody material (g dry weight m-2) Oct Nov Dec Jan Feb 19.4 2.6 0.0 0.0 2.2 1.5 5.9 1.0 0.6 0.2 11.5 16.0 19.1 2.0 0.3 10.3 8.9 7.8 1.5 10.5 0.0 15.7 38.2 1.6 1.2 13.7 17.3 1.7 0.0 2.3 5.8 5.9 3.4 5.9 2.4 2.9 6.6 11.9 0.6 1.0 1.0 3.0 9.9 3.0 0.7 5.0 20.0 25.4 4.8 2.6 4.5 ±3.0 6.3 ±4.4

Mar 2.8 1.6 0.0 13.4 4.4 23.7 2.4 3.7 1.2 0.5

1.4 ±0.5 2.6 ±1.3

Apr 1.7 0.9 0.4 2.5 1.3 7.5 1.3 0.8 0.3 3.1

1.2 ±0.5 6.7 ±3.8

May 0.9 2.4 0.2 1.2 3.5 21.8 3.3 5.1 1.1 2.3

77.6 ±13.6 73.2 ±14.3

Total 58.2 50.2 103.3 98.5 90.0 111.0 70.8 59.6 29.0 95.3

178

Table B.5. Cumulative mean woody material biomass and section mean (±1 SE) collected in the bottomland hardwood forest leaf traps between June 2003 and May 2004.

Jul 12.3 12.9 17.4 13.2 8.9 6.5 16.2 9.9 2.3 3.5

Jun 13.8 13.1 13.3 14.4 4.6 6.4 15.0 10.2 5.7 7.7

Plot/ Section 1 2 3 4 5 6 7 8 9 10

Specific gravity (g cm-3) Section Plot Tree Quadrant Species Maclura pomifera (Raf.) North 1 1 NW 0.80 Platanus occidentalis L. North 1 2 NW 0.46 Morus rubra L. North 1 3 NW 0.59 Platanus occidentalis L. North 1 4 NW 0.46 Acer negundo L. North 1 5 NW 0.54 Platanus occidentalis L. North 1 6 NW 0.46 Populus deltiodes Bartr. ex North 1 7 NW 0.37 Acer negundo L. North 1 8 NW 0.54 Morus rubra L. North 1 9 NW 0.59 Celtis laevigata Willd. North 1 10 NW 0.49 Asimina triloba (L.) Dunal North 1 11 NW 0.47 Asimina triloba (L.) Dunal North 1 12 NW 0.47 Asimina triloba (L.) Dunal North 1 13 NW 0.47 Celtis laevigata Willd. North 1 14 NW 0.49 Fraxinus pennsylvanica Marsh. North 1 15 NW 0.53 Populus deltiodes Bartr. ex North 1 16 NW 0.37 Populus deltiodes Bartr. ex North 1 17 NW 0.37 Populus deltiodes Bartr. ex North 1 18 NW 0.37 Acer saccharinum L. North 1 19 NW 0.44 Acer saccharinum L. North 1 19 NW 0.44 Acer saccharinum L. North 1 19 NW 0.44 Acer saccharinum L. North 1 19 NW 0.44 Asimina triloba (L.) Dunal North 1 20 NE 0.47 Acer saccharinum L. North 1 21 NE 0.44 Celtis laevigata Willd. North 1 22 NE 0.49 Asimina triloba (L.) Dunal North 1 23 NE 0.47 Asimina triloba (L.) Dunal North 1 24 NE 0.47 Asimina triloba (L.) Dunal North 1 25 NE 0.47 Asimina triloba (L.) Dunal North 1 26 NE 0.47 Asimina triloba (L.) Dunal North 1 27 NE 0.47 Asimina triloba (L.) Dunal North 1 28 NE 0.47 Asimina triloba (L.) Dunal North 1 29 NE 0.47 Aesculus glabra Willd. North 1 30 NE 0.33 Aesculus glabra Willd. North 1 31 NE 0.33 Asimina triloba (L.) Dunal North 1 32 NE 0.47 Aesculus glabra Willd. North 1 33 NE 0.33 Acer negundo L. North 1 34 NE 0.54 Aesculus glabra Willd. North 1 35 NE 0.33 Acer negundo L. North 1 36 NE 0.54

dbh (cm) Apr04 9.45 27.25 11.15 6.20 5.50 9.15 21.85 9.70 7.35 33.90 6.55 12.00 5.15 6.40 14.15 17.15 38.50 34.70 6.80 10.90 11.40 11.25 12.70 24.90 25.75 6.80 5.70 7.85 6.10 8.05 7.70 9.05 5.90 5.25 7.50 5.70 15.40 13.90 38.05

Mar05 9.80 27.75 11.50 6.35 5.80 9.20 22.20 9.90 7.40 34.45 6.55 12.10 5.20 6.60 14.60 17.65 39.50 35.70 6.80 11.05 11.40 11.65 13.70 25.70 26.40 6.90 5.85 8.00 6.15 8.05 7.90 9.10 6.75 5.80 7.90 5.80 15.40 14.50 38.40

Tree height

Wood prod.

(m) 9 20 8 5 5 13 18 9 2 14 8 10 4 5 14 13 25 18 11 11 11 11 11 15 13 8 8 12 8 9 10 12 9 3 2 9 10 10 15

(g yr-1) 1848 9925 1407 174 369 218 4004 743 34 10345 0 454 37 237 3653 3164 28330 18180 0 616 0 1715 5601 10421 8438 210 241 527 87 0 596 201 1245 251 227 130 0 2151 8593

Continued Table B.6. Tree specific gravity (per Alden 1995 and U.S. Forest Products Laboratory 1974), dbh, tree height and estimated wood production for all trees >5cm dbh in the bottomland hardwood forest tree plots. 179

Table B.6. continued Specific gravity Section North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North

Plot 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

Tree Quadrant 37 NE 38 NE 38 NE 39 S 40 S 41 S 42 S 42 S 42 S 43 S 44 S 45 S 45 S 46 S 46 S 47 S 1 NW 2 NW 3 NW 4 NW 5 NW 6 NW 7 NW 8 NW 9 NW 10 NW 11 NW 12 NW 13 NW 14 NW 15 NW 16 NW 17 NW 18 NW 19 NW 20 NW 21 NW 22 NW 23 NW 24 NE

Species Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer saccharinum L. Acer saccharinum L. Acer saccharinum L. Morus rubra L. Ulmus americana L. Acer negundo L. Acer negundo L. Populus deltiodes Bartr. ex Populus deltiodes Bartr. ex Acer saccharinum L. Celtis laevigata Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Celtis laevigata Willd. Asimina triloba (L.) Dunal Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Juglans nigra L. Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Celtis laevigata Willd. Asimina triloba (L.) Dunal Celtis laevigata Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal

-3

DBH (cm)

(g cm ) Apr04 0.54 8.00 0.54 12.30 0.54 24.70 0.54 6.65 0.54 27.40 0.54 25.30 0.44 15.50 0.44 49.40 0.44 61.90 0.59 14.90 0.46 74.30 0.54 8.45 0.54 16.30 0.37 18.55 0.37 39.75 0.44 11.90 0.49 19.05 0.33 5.15 0.33 7.60 0.33 15.20 0.47 11.90 0.33 13.15 0.33 5.00 0.49 31.85 0.47 8.50 0.54 53.80 0.33 7.00 0.33 15.85 0.47 6.20 0.51 66.80 0.47 4.90 0.33 21.60 0.47 5.15 0.49 14.45 0.47 9.70 0.49 12.45 0.47 14.45 0.33 7.35 0.47 11.15 0.47 10.80

Mar05 9.00 12.30 24.70 6.80 27.60 25.30 15.85 50.20 63.30 15.35 74.50 9.10 16.50 18.60 40.25 12.40 19.15 5.30 7.60 15.20 12.10 13.35 5.10 32.10 8.70 53.80 7.30 16.10 6.35 67.15 n/a 21.80 5.30 14.70 9.70 12.60 14.60 dead 11.15 11.00

Tree height

Wood prod.

(m) 9 6 9 7 13 8 10 23 25 7 18 4 9 15 20 9 15 5 5 10 11 9 4 15 7 9 7 11 9 28 -11 12 11 7 12 11 -11 15

(g yr ) 3198 0 0 317 3052 0 1886 31418 74847 2183 9894 1043 1200 417 11595 1833 1087 100 0 0 954 652 54 4583 422 0 368 1113 311 26423 -1234 361 1558 0 836 871 -0 1194

-1

Continued

180

Table B.6. continued Specific gravity Section North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North

Plot 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

Tree Quadrant 25 NE 26 NE 27 NE 28 NE 29 NE 30 NE 31 NE 32 NE 33 NE 34 NE 35 NE 36 NE 37 NE 38 NE 39 NE 40 NE 41 SE 42 SE 43 SE 44 SE 45 SE 46 SE 47 SE 48 SE 49 SE 50 SE 51 SE 52 SE 53 SE 54 SE 55 SW 56 SW 57 SW 58 SW 59 SW 60 SW 61 SW 62 SW 63 SW 64 SW 65 SW 66 SW

Species Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Celtis laevigata Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Celtis laevigata Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal

(g cm-3) 0.47 0.33 0.47 0.47 0.47 0.49 0.47 0.47 0.49 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.33 0.33 0.47 0.33 0.47 0.47 0.47 0.47 0.47 0.47 0.54 0.33 0.33 0.33 0.47 0.47 0.47 0.33 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47

DBH (cm) Apr04 7.10 8.25 6.40 9.15 8.95 16.15 7.65 5.30 15.10 5.40 5.90 8.05 9.35 13.50 10.00 10.85 7.05 4.80 12.00 10.60 14.00 11.60 6.20 7.70 7.15 4.80 72.25 5.60 4.70 7.50 11.55 12.40 6.25 8.60 6.70 11.65 5.85 8.70 7.75 9.80 8.15 7.55

Mar05 7.20 8.40 6.45 9.50 9.10 16.70 7.70 5.40 15.90 5.60 6.10 8.05 9.50 13.60 10.25 11.05 7.20 dead 12.30 10.75 14.15 11.85 6.25 7.80 7.15 5.15 57.50 5.85 n/a 7.80 11.90 12.60 6.35 8.90 6.80 12.00 dead 8.80 8.00 9.95 8.45 7.90

Tree height

Wood prod.

(m) 15 6 10 13 13 13 7 4 15 9 10 8 8 10 10 10 6 -10 9 13 10 5 8 8 6 23 5 -7 9 9 9 7 8 11 -6 7 10 11 10

(g yr-1) 392 190 114 1561 648 4494 98 71 7260 365 430 0 401 479 894 835 164 -1389 369 978 1051 54 218 0 399 203 -394 1435 874 220 451 199 1625 -179 512 565 1041 1018

Continued 181

Table B.6. continued Specific gravity Section Plot Tree Quadrant North 2 67 SW North 2 68 SW North 2 69 SW North 2 70 SW North 2 71 SW North 2 72 SW North 2 73 SW North 2 74 SW North 2 75 SW North 2 76 SW North 2 77 SW North 2 78 SW North 2 79 SW North 2 80 SW North 2 81 SW North 3 1 NW North 3 2 NW North 3 3 NW North 3 4 NW North 3 5 NW North 3 6 NW North 3 7 NW North 3 8 NW North 3 9 NW North 3 10 NW North 3 10 NW North 3 11 NW North 3 12 NE North 3 13 NE North 3 14 NE North 3 15 NE North 3 16 NE North 3 17 NE North 3 18 NE North 3 19 NE North 3 20 NE North 3 21 NE North 3 22 NE North 3 23 NE North 3 24 NE

Species Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Acer negundo L. Acer negundo L. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Acer negundo L. Asimina triloba (L.) Dunal Morus rubra L. Acer saccharum Marsh. Acer saccharum Marsh. Morus alba L. Asimina triloba (L.) Dunal Celtis laevigata Willd. Acer negundo L. Aesculus glabra Willd. Platanus occidentalis L. Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal

(g cm-3) 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.33 0.47 0.54 0.54 0.47 0.47 0.47 0.47 0.54 0.47 0.59 0.56 0.56 0.59 0.47 0.49 0.54 0.33 0.46 0.33 0.33 0.47 0.47 0.47 0.33 0.33 0.47

DBH (cm) Apr04 6.20 7.20 7.95 6.10 5.00 7.00 14.80 12.25 7.50 8.60 6.70 7.40 5.55 6.90 6.20 37.20 5.30 7.40 6.90 5.60 5.20 43.50 5.35 9.40 27.25 41.60 7.00 7.65 7.30 13.15 12.95 97.10 22.00 10.25 6.50 7.15 6.80 19.30 5.35 5.00

Mar05 6.35 7.25 8.30 6.10 5.05 7.15 14.95 12.25 7.65 8.85 6.80 7.50 5.60 7.20 6.30 37.35 5.55 7.90 7.45 5.75 5.50 43.70 5.75 9.60 27.80 42.45 7.00 8.30 7.60 13.60 13.00 97.35 22.40 10.25 6.60 7.40 7.05 19.70 5.75 5.00

Tree height

Wood prod.

(m) 8 8 9 8 7 8 9 10 12 9 8 9 7 6 8 10 7 8 8 7 6 20 7 6 20 20 6 8 7 12 8 47 13 8 8 9 8 10 6 6

(g yr-1) 284 110 920 0 66 320 780 0 500 694 191 248 76 324 179 2419 381 1129 1096 235 330 7273 564 546 13094 30897 0 1603 617 3016 135 41626 3066 0 188 614 531 2048 333 0

Continued 182

Table B.6. continued Specific gravity Section Plot Tree Quadrant North 3 25 NE North 3 26 NE North 3 27 NE North 3 28 NE North 3 29 NE North 3 30 NE North 3 31 NE North 3 32 NE North 3 33 NE North 3 34 NE North 3 35 SW North 3 36 SW North 3 37 SW North 3 38 SW North 3 39 SW North 3 40 SW North 3 41 SE North 3 42 SE North 3 43 SE North 3 44 SE North 3 45 SE North 3 46 SE North 3 47 SE North 3 48 SE North 3 49 SE North 3 50 SE North 3 51 SE North 3 52 SE North 3 53 SE North 3 54 SE North 3 55 SE North 3 56 SE North 3 57 SE North 3 58 SE North 3 59 SE North 3 60 SE North 3 61 SE North 3 62 SE North 3 63 SE North 3 64 SE

Species Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Acer saccharum Marsh. Morus alba L. Asimina triloba (L.) Dunal Acer negundo L. Acer negundo L. Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal

(g cm-3) 0.47 0.33 0.33 0.47 0.47 0.47 0.47 0.47 0.47 0.47 0.56 0.59 0.47 0.54 0.54 0.47 0.33 0.47 0.33 0.33 0.47 0.47 0.33 0.47 0.33 0.47 0.47 0.47 0.47 0.33 0.47 0.47 0.33 0.47 0.33 0.47 0.33 0.33 0.47 0.47

DBH (cm) Apr04 8.55 12.50 5.40 7.60 6.15 6.50 6.85 6.85 6.80 4.95 7.50 12.05 8.40 11.30 9.40 5.95 5.70 8.15 5.70 6.20 5.10 5.05 10.40 6.55 5.20 8.30 11.90 7.10 5.65 12.85 5.55 5.50 5.05 10.10 6.45 9.60 7.35 5.70 9.35 9.90

Mar05 8.55 12.65 5.50 7.85 6.40 6.70 7.00 6.95 6.90 5.10 7.90 12.15 8.50 11.75 9.60 6.10 6.05 8.40 6.10 6.40 5.50 5.10 10.70 6.75 5.20 8.40 12.10 7.30 5.80 13.20 5.70 5.55 5.15 10.35 6.50 9.85 7.35 5.70 9.45 10.10

Tree height

Wood prod.

(m) 9 8 5 11 5 8 6 7 8 5 6 12 7 8 7 8 4 9 5 5 5 6 8 4 6 12 10 7 8 8 7 6 3 9 5 9 5 6 9 9

(g yr-1) 0 406 76 750 294 402 222 171 203 148 869 701 213 1797 590 255 226 688 302 158 354 58 687 209 0 358 860 386 252 940 215 63 42 883 46 802 0 0 297 688

Continued 183

Table B.6. continued Specific gravity Section North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North North

Plot 3 3 3 3 3 3 3 3 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4

Tree Quadrant 65 SE 66 SE 67 SE 68 SE 69 SE 70 SE 71 SE 72 SE 73 SE 1 NW 2 NW 3 NW 4 NW 5 NW 6 NW 7 NW 8 NW 9 NW 10 NW 11 NW 12 NW 13 NW 14 NW 15 NW 16 NW 17 NE 18 NE 19 NE 20 NE 21 NE 22 NE 23 NE 24 NE 25 NE 26 NE 27 NE 28 NE 28 NE 29 NE 30 NE 31 SW 32 SW

Species Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Celtis laevigata Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Celtis laevigata Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Celtis laevigata Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Celtis laevigata Willd. Aesculus glabra Willd. Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Celtis laevigata Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Celtis laevigata Willd. Asimina triloba (L.) Dunal Aesculus glabra Willd.

-3

DBH (cm)

(g cm ) Apr04 0.47 7.80 0.47 5.50 0.49 33.45 0.47 7.75 0.47 5.70 0.47 5.70 0.49 16.60 0.47 5.35 0.47 6.40 0.33 9.50 0.49 96.80 0.33 21.50 0.33 6.20 0.33 22.10 0.33 12.35 0.33 6.45 0.47 6.50 0.47 6.45 0.47 5.30 0.47 5.55 0.47 5.75 0.33 14.15 0.33 9.65 0.47 5.20 0.47 5.95 0.33 45.55 0.33 56.05 0.33 5.00 0.54 23.90 0.54 6.30 0.49 97.10 0.33 9.85 0.54 23.75 0.33 6.70 0.33 7.10 0.49 18.00 0.33 5.25 0.33 6.90 0.33 6.85 0.49 16.90 0.47 5.55 0.33 5.65

Mar05 7.90 5.55 33.75 7.95 5.85 5.90 16.90 5.50 6.60 9.55 97.20 21.85 6.35 22.10 12.75 6.55 6.50 6.55 5.40 5.80 5.90 14.20 9.85 5.35 6.35 46.10 56.50 5.25 24.75 6.65 97.40 9.95 24.60 6.80 7.20 19.00 5.25 7.05 7.00 17.20 5.70 5.65

Tree height

Wood prod.

(m) 9 5 24 7 8 6 12 8 8 6 41 12 6 12 8 7 13 13 9 9 7 7 10 7 8 17 18 5 15 6 44 7 14 5 5 8 6 6 6 14 6 5

(g yr ) 257 53 9220 429 241 242 2339 229 365 76 61351 2337 157 0 1026 119 0 300 184 454 226 129 491 212 706 10831 12079 164 13559 617 49599 180 11824 94 99 5467 0 167 167 2831 176 0

-1

Continued 184

Table B.6. continued Specific gravity Section Plot Tree Quadrant North 4 33 SW North 4 34 SW North 4 35 SW North 4 36 SW North 4 37 SW North 4 38 SW North 4 39 SE North 4 39 SW North 4 40 SE North 4 40 SE North 4 41 SE North 4 42 SE North 4 43 SE North 4 44 SE North 4 45 SE North 4 46 SE North 4 47 SE North 4 48 SE North 4 49 SE North 4 50 SE Upland 5 1 NW Upland 5 2 NW Upland 5 3 NW Upland 5 4 NW Upland 5 5 NW Upland 5 6 NW Upland 5 7 NW Upland 5 8 NE Upland 5 9 NE Upland 5 9 NE Upland 5 10 NE Upland 5 10 NE Upland 5 11 NE Upland 5 12 SW Upland 5 13 SW Upland 5 14 SW Upland 5 15 SW Upland 5 16 SW Upland 5 17 SW Upland 5 18 SW Upland 5 18 SW

Species Aesculus glabra Willd. Celtis laevigata Willd. Acer negundo L. Aesculus glabra Willd. Asimina triloba (L.) Dunal Morus alba L. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Morus alba L. Juglans nigra L. Acer negundo L. Juglans nigra L. Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Asimina triloba (L.) Dunal Ulmus americana L. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Celtis laevigata Willd. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Aesculus glabra Willd. Maclura pomifera (Raf.) Maclura pomifera (Raf.) Aesculus glabra Willd. Aesculus glabra Willd. unknown Celtis laevigata Willd. Celtis laevigata Willd. Celtis laevigata Willd.

-3

DBH (cm)

(g cm ) Apr04 0.33 7.15 0.49 6.80 0.54 7.25 0.33 9.30 0.47 5.20 0.59 5.85 0.33 14.85 0.33 16.45 0.54 9.60 0.54 11.40 0.59 20.00 0.51 20.05 0.54 13.95 0.51 30.05 0.47 7.80 0.47 6.05 0.47 9.90 0.47 5.10 0.47 5.35 0.47 8.40 0.46 12.95 0.33 8.05 0.54 28.00 0.54 6.45 0.54 35.80 0.54 7.65 0.54 12.25 0.49 10.90 0.54 8.60 0.54 18.40 0.54 6.75 0.54 9.40 0.33 6.10 0.80 64.70 0.80 38.80 0.33 10.10 0.33 9.10 17.30 0.49 19.55 0.49 43.85 0.49 83.65

Mar05 7.55 7.05 7.50 9.65 5.70 5.85 15.05 16.80 9.75 11.60 20.40 20.50 14.10 30.45 7.95 6.20 10.25 5.30 5.50 8.60 12.95 8.05 28.85 dead dead dead 12.40 11.55 8.65 18.95 7.05 10.00 7.30 dead 39.30 10.15 9.15 dead 19.60 44.35 83.95

Tree height

Wood prod.

(m) 9 6 8 6 8 5 9 9 7 8 10 20 11 19 5 8 12 8 4 8 7 5 13 ---6 9 8 10 8 8 2 -13 6 5 -11 13 37

(g yr ) 673 413 655 530 783 0 677 1317 455 759 3636 7324 992 9378 197 260 1575 312 122 495 0 0 13256 ---503 2609 155 4383 668 2016 417 -15354 73 60 -412 11334 35477

-1

Continued 185

Table B.6. continued Specific gravity Section Plot Tree Quadrant Upland 5 19 SW Upland 5 20 SE Upland 5 21 SE Upland 5 22 SE Upland 5 22 SE Upland 5 23 SE Upland 5 23 SE Upland 5 24 SE Upland 5 25 SE Upland 5 26 SE Upland 5 27 SE Upland 5 28 SE South 6 1 NW South 6 2 NW South 6 3 NW South 6 4 NW South 6 5 NW South 6 6 NW South 6 7 NE South 6 7 NE South 6 8 NE South 6 9 NE South 6 10 NE South 6 11 NE South 6 12 NE South 6 13 NE South 6 14 NE South 6 15 NE South 6 16 NE South 6 17 SW South 6 18 SW South 6 19 SW South 6 20 SW South 6 21 SW South 6 22 SW South 6 23 SW South 6 24 SW South 6 25 SW South 6 26 SW

Species Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Maclura pomifera (Raf.) Maclura pomifera (Raf.) Maclura pomifera (Raf.) Aesculus glabra Willd. Acer negundo L. Acer negundo L. Acer negundo L. Populus deltiodes Bartr. ex Acer negundo L. Acer negundo L. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Acer saccharum Marsh. Acer saccharum Marsh. Aesculus glabra Willd. Acer negundo L. Aesculus glabra Willd. Acer saccharum Marsh. Acer saccharum Marsh. Acer saccharum Marsh. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Populus deltiodes Bartr. ex Acer negundo L. Ulmus americana L. Acer negundo L.

-3

DBH (cm)

(g cm ) Apr04 Mar05 0.54 5.00 5.30 0.33 7.80 8.00 0.33 7.05 7.25 0.54 11.90 12.30 0.54 46.10 46.40 0.80 9.45 9.60 0.80 72.85 73.20 0.80 45.45 45.70 0.33 22.50 22.55 0.54 6.55 6.55 0.54 5.50 5.70 0.54 10.60 11.25 0.37 113.40 113.50 0.54 13.85 14.10 0.54 8.40 8.40 0.33 10.75 10.80 0.54 14.70 14.75 0.54 39.35 39.65 0.56 7.95 7.95 0.56 21.40 21.70 0.33 4.75 n/a 0.54 37.80 38.60 0.33 12.90 13.20 0.56 44.85 45.35 0.56 20.10 20.20 0.56 4.70 n/a 0.33 8.50 8.60 0.33 23.15 23.85 0.54 29.20 29.80 0.54 22.45 22.50 0.54 6.80 6.80 0.54 8.55 8.95 0.54 8.15 8.20 0.54 5.45 5.45 0.54 6.30 6.65 0.37 118.20 118.50 0.54 10.50 10.60 0.46 41.25 41.65 0.54 37.00 37.60

Tree height

Wood prod.

(m) 4 5 4 6 21 9 25 12 10 7 8 12 51 7 6 5 8 14 7 11 -15 8 27 13 -6 17 17 11 7 8 8 7 7 51 7 23 20

(g yr ) 292 213 151 1188 12439 768 40756 8684 279 0 361 3622 16837 1075 0 71 243 6818 0 3160 -19765 808 26462 1184 -138 7129 12500 517 0 1135 133 0 664 52694 314 13562 18733

-1

Continued 186

Table B.6. continued Specific gravity Section South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South

Plot 6 6 6 6 6 6 6 6 6 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7 7

Tree Quadrant 27 SW 28 SE 29 SE 30 SE 31 SE 32 SE 33 SE 34 SE 35 SE 1 NW 2 NW 3 NW 4 NW 5 NW 6 NW 6 NW 7 NW 8 NW 9 NW 9 NW 10 NW 11 NW 12 NE 13 NE 13 NE 14 NE 14 NE 15 NE 16 SW 17 SW 18 SW 19 SW 20 SW 21 SW 22 SW 23 SW 24 SW 25 SW 26 SE 27 SE 28 SE 29 SE 29 SE 30 SW

Species Acer negundo L. Acer negundo L. Acer negundo L. Juglans nigra L. Prunus serotina Ehrh. Gleditsia triacanthos L. Gleditsia triacanthos L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Celtis laevigata Willd. Morus alba L. Aesculus glabra Willd. Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Celtis laevigata Willd.

-3

DBH (cm)

(g cm ) Apr04 Mar05 0.54 15.50 15.60 0.54 6.60 6.95 0.54 5.40 5.70 0.51 7.60 7.75 0.47 38.00 38.75 0.60 5.05 5.50 0.60 4.80 5.05 0.54 12.90 dead 0.54 9.15 9.40 0.54 9.00 9.15 0.54 5.55 5.60 0.54 8.30 dead 0.54 11.10 11.55 0.54 7.15 7.90 0.54 7.80 8.00 0.54 11.25 11.50 0.54 6.60 7.25 0.54 10.20 10.95 0.49 8.35 8.65 0.59 55.60 56.20 0.33 14.25 14.40 0.54 30.35 30.35 0.33 16.40 16.50 0.33 5.55 5.70 0.33 6.70 6.70 0.33 8.50 9.05 0.33 13.90 14.50 0.54 11.45 11.90 0.33 6.40 6.50 0.33 8.35 8.50 0.33 7.10 7.30 0.33 4.70 n/a 0.54 27.60 27.90 0.54 6.80 7.30 0.54 7.35 7.55 0.33 15.10 15.40 0.33 21.70 22.15 0.33 28.70 29.10 0.33 7.80 7.90 0.54 23.15 23.45 0.33 25.50 25.65 0.33 6.00 6.05 0.33 6.60 6.60 0.49 15.90 16.70

Tree height

Wood prod.

(m) 11 7 4 11 20 5 5 2 9 7 8 2 11 8 6 7 11 11 7 27 10 9 2 4 4 8 9 6 4 6 7 -15 7 9 10 13 12 6 15 12 4 5 13

(g yr ) 714 688 312 529 21489 590 279 -886 423 90 -2432 1888 392 875 2055 3777 712 41610 541 0 85 82 0 958 1986 1430 75 192 247 -5347 1096 552 1226 3371 3624 126 4334 1147 33 0 6281

-1

Continued 187

Table B.6. continued Specific gravity Section South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South South

Plot 7 7 7 7 8 8 8 8 8 8 8 8 8 8 8 8 8 8 8 8 8 8 8 9 9 9 9 9 9 9 9 9 9 9 9 9 9 9 9 9

-3

DBH (cm)

(g cm ) Apr04 Tree Quadrant Species Aesculus glabra Willd. 31 SE 0.33 5.15 Acer negundo L. 32 SE 0.54 23.65 Acer negundo L. 33 SE 0.54 6.30 Platanus occidentalis L. 34 SE 0.46 80.85 Acer negundo L. 1 NW 0.54 43.50 Fraxinus pennsylvanica Marsh. 0.53 2 NW 44.40 Gleditsia triacanthos L. 3 NW 0.60 86.40 Aesculus glabra Willd. 4 NE 0.33 6.55 Aesculus glabra Willd. 4 NE 0.33 8.10 Acer negundo L. 5 NE 0.54 32.30 Lonicera maacki Rupr. 6 NE 0.45 4.95 Aesculus glabra Willd. 7 SW 0.33 6.70 Acer negundo L. 8 SW 0.54 10.10 Acer negundo L. 9 SW 0.54 9.30 Celtis laevigata Willd. 10 SW 0.49 46.65 Acer negundo L. 11 SW 0.54 34.05 Aesculus glabra Willd. 12 SW 0.33 8.85 Acer negundo L. 13 SE 0.54 34.20 Lonicera maacki Rupr. 14 SE 0.45 6.00 Platanus occidentalis L. 15 SE 0.46 86.00 Acer negundo L. 16 SE 0.54 23.15 Lonicera maacki Rupr. 17 SE 0.45 7.20 Lonicera maacki Rupr. 17 SE 0.45 7.85 Aesculus glabra Willd. 1 NW 0.33 57.85 Aesculus glabra Willd. 2 NW 0.33 7.80 Acer negundo L. 3 NW 0.54 23.55 Acer negundo L. 4 NW 0.54 5.80 Acer negundo L. 4 NW 0.54 59.10 Acer negundo L. 5 NW 0.54 25.50 Aesculus glabra Willd. 6 NW 0.33 9.25 Acer negundo L. 7 NW 0.54 7.20 Acer negundo L. 7 NW 0.54 9.20 Acer negundo L. 7 NW 0.54 60.90 Acer negundo L. 8 NE 0.54 6.55 Acer negundo L. 8 NE 0.54 8.50 Acer negundo L. 8 NE 0.54 35.60 Acer negundo L. 9 NE 0.54 5.25 Acer negundo L. 10 NE 0.54 7.70 Aesculus glabra Willd. 11 SW 0.33 5.20 Aesculus glabra Willd. 12 SW 0.33 33.55

Mar05 5.45 24.20 6.40 81.05 43.80 45.20 86.60 6.55 8.15 32.75 5.10 6.80 10.50 9.30 47.25 34.65 8.85 34.60 6.20 86.15 23.30 8.00 8.00 58.20 8.05 24.00 6.10 59.70 25.90 9.60 7.30 9.65 61.40 6.60 8.60 35.70 5.50 7.75 5.35 33.85

Tree height

Wood prod.

(m) 4 14 6 39 19 24 40 6 7 7 6 4 7 8 23 13 6 17 4 43 8 5 5 21 6 19 6 18 16 6 6 7 24 4 4 14 5 5 4 14

(g yr ) 160 7571 158 22575 10688 35590 32437 0 79 4201 173 69 1287 0 24968 11185 0 9818 169 20055 1228 995 194 11016 301 8462 443 26679 6969 527 186 1313 30934 49 141 2115 284 82 83 3676

-1

Continued 188

Table B.6. continued Specific gravity Section South South South South South South South South South South South South South South South South South South South South South South South

Plot Tree Quadrant 9 13 SW 9 14 SW 9 15 SW 9 16 SW 9 17 SW 9 18 SW 9 19 SE 9 20 SE 9 21 SE 9 22 SE 10 1 NW 10 2 NW 10 3 NW 10 4 NW 10 5 NW 10 6 NE 10 7 NE 10 8 SW 10 9 SW 10 10 SW 10 11 SW 10 12 SE 10 13 SE

Species Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Aesculus glabra Willd. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer negundo L. Acer saccharinum L. Acer saccharinum L. Acer saccharinum L. Salix nigra Marsh. Acer negundo L. Populus deltiodes Bartr. ex Populus deltiodes Bartr. ex Morus rubra L. Acer negundo L. Morus rubra L. Morus rubra L.

189

-3

DBH (cm)

(g cm ) Apr04 0.33 12.65 0.33 9.25 0.33 20.15 0.33 11.30 0.33 15.65 0.54 8.25 0.54 42.10 0.54 23.70 0.54 21.75 0.54 33.25 0.54 9.45 0.54 46.70 0.44 37.80 0.44 24.05 0.44 38.70 0.36 60.50 0.54 7.10 0.37 81.70 0.37 68.10 0.59 7.65 0.54 10.20 0.59 25.05 0.59 13.10

Mar05 12.90 9.25 20.15 11.30 15.65 8.65 42.70 23.70 21.75 33.85 9.95 47.00 38.95 25.25 39.60 61.30 7.80 81.90 68.30 7.95 10.70 25.25 13.40

Tree height

Wood prod.

(m) 7 5 11 9 9 7 22 4 5 14 6 14 23 21 23 25 8 29 28 9 9 13 10

(g yr ) 572 0 0 0 0 940 23496 0 0 11762 1233 8112 35604 21787 28427 34139 1838 13949 21782 947 2074 3015 2385

-1

Plot 1

2

3

4

5

Plot corner/ Elevation Mean elev. Elevation leaf trap (m MSL) (m MSL) Variance NW 222.66 221.36 0.43 NE 221.74 SE 220.87 SW 221.06 LT 1 222.13 LT 2 221.07 LT 3 220.75 LT 4 221.03 LT 5 220.96 NW 221.47 221.45 0.02 NE 221.60 SE 221.45 SW 221.35 LT 1 221.53 LT 2 221.16 LT 3 221.61 LT 4 221.38 LT 5 221.51 NW 221.49 221.37 0.04 NE 221.23 SE 221.25 SW 221.42 LT 1 221.62 LT 2 220.98 LT 3 221.56 LT 4 221.48 LT 5 221.27 NW 222.13 221.56 0.09 NE 221.34 SE 221.51 SW 221.27 LT 1 222.02 LT 2 221.50 LT 3 221.42 LT 4 221.35 LT 5 221.52 NW 221.81 221.92 0.02 NE 222.16 SE 221.77 SW 221.80 LT 1 221.97 LT 2 222.03 LT 3 222.07 LT 4 221.77 LT 5 221.87

Plot 6

7

8

9

10

Plot corner/ Elevation Mean elev. Elevation leaf trap (m MSL) (m MSL) Variance NW 221.06 221.50 0.21 NE 221.67 SE 222.12 SW 221.29 LT 1 221.28 LT 2 221.73 LT 3 221.29 LT 4 220.88 LT 5 222.21 NW 221.77 221.31 0.08 NE 221.54 SE 221.20 SW 220.91 LT 1 221.00 LT 2 221.55 LT 3 221.39 LT 4 221.17 LT 5 221.21 NW 221.07 221.49 0.11 NE 221.80 SE 221.44 SW 221.69 LT 1 220.85 LT 2 221.45 LT 3 221.64 LT 4 221.67 LT 5 221.81 NW 221.00 221.13 0.04 NE 221.43 SE 221.21 SW 220.75 LT 1 221.00 LT 2 221.21 LT 3 221.30 LT 4 221.08 LT 5 221.19 NW 221.46 221.14 0.04 NE 221.03 SE 220.87 SW 220.84 LT 1 221.09 LT 2 221.16 LT 3 221.21 LT 4 221.28 LT 5 221.32

Table B.7. Surveyed elevations of plots corners (NW, NE, SE, and SW) and leaf traps (LT) for each bottomland hardwood forest tree plot. Elevation mean and variance based on all measured plot elevations. 190

Plot 1

2

3

4

5

Leaf Trap 1 2 3 4 5 1 2 3 4 5 1 2 3 4 5 1 2 3 4 5 1 2 3 4 5

N 91 nd 87 74 84 81 93 87 94 88 80 84 91 83 89 69 69 77 72 69 77 84 82 78 77

Canopy cover (%) S E 83 85 nd nd 80 75 74 74 78 74 87 86 83 80 87 90 90 94 91 93 68 83 81 82 73 90 75 84 81 85 80 77 64 69 71 72 80 77 67 74 82 80 75 92 77 79 77 93 88 86

W 86 nd 80 87 79 82 85 91 91 90 78 76 86 86 89 76 67 68 80 79 90 81 83 85 81

Trap mean Plot Mean 86.3 80.7 nd 80.5 77.3 78.8 84.0 88.2 85.3 88.8 92.3 90.5 77.3 82.2 80.8 85.0 82.0 86.0 75.5 72.9 67.3 72.0 77.3 72.3 82.3 82.4 83.0 80.3 83.3 83.0

‘nd’: no data available

Continued Table B.8. Mean trap and plot canopy cover for each plot in the bottomland hardwood forest.

191

Table B.8. continued

Plot 6

7

8

9

10

Leaf Trap 1 2 3 4 5 1 2 3 4 5 1 2 3 4 5 1 2 3 4 5 1 2 3 4 5

N 83 82 88 91 82 88 86 78 82 72 84 90 90 83 90 78 85 92 82 87 76 80 79 79 76

Canopy cover (%) S E 77 90 83 91 85 87 90 91 86 78 86 79 86 81 79 80 82 79 74 77 95 83 85 85 86 91 88 86 89 85 82 80 81 84 73 87 79 77 71 88 77 77 73 77 82 80 78 91 79 71

‘nd’ denotes no data available

192

W 72 83 75 85 85 90 77 84 88 81 87 80 66 82 84 83 91 89 80 73 62 83 77 78 76

Trap mean Plot Mean 80.5 84.2 84.8 83.8 89.3 82.8 85.8 81.5 82.5 80.3 82.8 76.0 87.3 85.5 85.0 83.3 84.8 87.0 80.8 82.1 85.3 85.3 79.5 79.8 73.0 77.6 78.3 79.5 81.5 75.5

193

North North

North

1-44 1-44 Mean

1-46

1

1 1

1 1

1 1

1 1

1 1

P. deltiodes

U. americana U. americana

A. saccharinum A. saccharinum

A. saccharinum A. saccharinum

A. negundo A. negundo

A. negundo A. negundo

C. laevigata C. laevigata

Species P. occidentalis

1

1 2

1 2

1 2

1 2

1 2

1 2

7.46

1.85 3.28 2.57

4.21 2.11 3.16

2.36 1.57 1.96

0.54 1.67 1.11

3.62 5.01 4.32

2.53 4.17 3.35

7.28

5.04 6.06 5.55

8.51 6.31 7.41

1.77 2.40 2.08

0.33 0.54 0.44

4.45 3.36 3.91

3.00 3.53 3.26

5.66

3.54 5.13 4.34

4.29 5.30 4.79

2.42 0.65 1.54

1.26 0.84 1.05

3.62 4.46 4.04

3.29 4.66 3.98

3.94

4.87 3.96 4.41

4.97 3.62 4.30

3.61 1.63 2.62

1.13 1.81 1.47

4.38 3.28 3.83

2.59 2.92 2.75

4.48

3.95 5.13 4.54

8.26 3.27 5.76

1.51 1.88 1.69

1.80 1.59 1.70

2.82 1.26 2.04

3.47 2.80 3.13

5.47

3.45 5.63 4.54

2.51 1.01 1.76

1.35 1.53 1.44

1.22 1.42 1.32

3.11 1.85 2.48

2.45 4.05 3.25

4.60

2.69 3.88 3.28

1.76 0.92 1.34

1.42 1.73 1.58

1.04 0.42 0.73

1.81 3.46 2.63

2.70 2.99 2.84

4.93

3.11 5.81 4.46

2.69 2.01 2.35

2.51 0.97 1.74

0.84 0.83 0.83

2.56 1.85 2.21

3.72 4.32 4.02

4.92

2.01 3.62 2.82

0.51 2.35 1.43

1.31 0.81 1.06

0.75 1.42 1.09

3.18 1.68 2.43

2.50 4.23 3.36

3.60

2.53 5.14 3.83

1.67 4.56 3.11

2.34 2.21 2.28

1.60 1.93 1.77

2.22 2.69 2.45

3.19 5.85 4.52

3.12

2.34 4.46 3.40

1.52 3.57 2.54

1.76 0.99 1.38

1.68 1.76 1.72

1.93 1.90 1.92

3.88 4.53 4.20

2.81

1.51 4.80 3.16

1.34 3.20 2.27

1.39 1.02 1.20

1.31 1.51 1.41

1.30 1.50 1.40

3.25 3.91 3.58

0.98

0.92 1.94 1.43

4.58 1.32 2.95

2.90 3.32 3.11

1.51 1.67 1.59

1.26 1.25 1.25

5.49 4.80 5.14

3.31

1.18 2.94 2.06

9.50 6.48 7.99

2.54 5.87 4.20

0.97 1.00 0.98

1.35 2.36 1.85

3.19 3.25 3.22

Annual tree-ring growth increments (mm) Sample 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2 5.98 3.95 4.54 5.18 6.12 6.20 2.27 2.95 4.68 6.89 3.02 2.51

5.23

3.30 4.76 4.03

3.94 3.14 3.54

2.06 1.54 1.80

1.05 1.25 1.15

3.18 2.89 3.03

2.94 3.95 3.45

0.41

0.35 0.31 0.29

0.86 0.57 0.63

0.23 0.18 0.14

0.14 0.17 0.13

0.27 0.39 0.28

0.14 0.30 0.18

1991-2000 Mean SE 4.65 0.53

0.53

0.31 0.67 0.46

1.91 1.07 1.36

0.35 1.16 0.72

0.15 0.17 0.16

0.16 0.24 0.16

0.53 0.35 0.42

Continued

2.55

1.49 3.53 2.51

4.23 3.64 3.94

2.15 2.80 2.47

1.37 1.48 1.43

1.46 1.75 1.61

3.95 4.12 4.04

2001-2004 Mean SE 4.28 0.99

193

Table B.9. Annual tree basal growth increments and mean annual growth increment (±SE) for the pre-restoration years (19912000) and post-restoration years (2001-2004).

North North

1-42(63) 1-42(63) Mean

North North

1-40 1-40 Mean

North North

North North

1-36 1-36 Mean

1-42(49) 1-42(49) Mean

North North

1-22 1-22 Mean

1 1

Section Plot North 1

Tree 1-2

194

North North

North North

North North

North North

North

North North

North

2-14 2-14 Mean

3-1 3-1 Mean

3-10 3-10 Mean

3-16 3-16 Mean

P. del.~3-16

N_sup1 N_sup1 Mean

4-2

4

3 3

3

3 3

3 3

3 3

2 2

Section Plot North 1 North 1

Tree Perm1 Perm1 Mean

Table B.9. continued

C. laevigata

A. negundo A. negundo

P. deltoides

P. occidentalis P. occidentalis

A. saccharinum A. saccharinum

A. negundo A. negundo

J. nigra J. nigra

Species A. negundo A. negundo

2

1 2

2

1 2

1 2

1 2

1 2

5.82

2.27 3.70 2.99

1.68

1.65 2.35 2.00

4.26 3.95 4.11

3.61 1.25 2.43

1.21 2.19 1.70

4.68

4.64 2.18 3.41

1.77

1.38 1.58 1.48

3.20 2.99 3.09

3.03 1.45 2.24

1.74 1.42 1.58

4.62

3.61 2.30 2.95

2.02

1.99 2.25 2.12

3.81 2.76 3.28

2.02 0.93 1.47

2.37 1.18 1.78

4.40

4.56 3.19 3.88

1.18

1.43 1.55 1.49

2.52 0.88 1.70

2.86 3.00 2.93

2.96 3.13 3.04

3.80

3.94 2.98 3.46

2.02

2.44 3.54 2.99

1.44 1.09 1.26

2.36 2.19 2.28

2.37 2.51 2.44

194

4.17

3.25 2.65 2.95

1.26

1.96 2.00 1.98

1.03 1.75 1.39

1.09 3.11 2.10

1.58 1.96 1.77

3.08

1.31 1.34 1.33

2.44

2.59 2.61 2.60

2.22 2.51 2.37

2.35 2.74 2.54

2.02 1.84 1.93

3.92

0.72 1.21 0.97

2.60

2.57 1.53 2.05

1.77 1.39 1.58

1.01 2.88 1.94

1.37 2.45 1.91

2.69

1.24 1.17 1.21

3.03

2.47 2.20 2.33

4.60 3.96 4.28

0.76 0.82 0.79

1.61 2.11 1.86

3.81

0.58 1.25 0.91

1.94

2.27 2.54 2.40

3.68 3.37 3.52

1.60 2.68 2.14

1.94 1.84 1.89

Annual tree-ring growth increments (mm) Sample 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 1 8.38 4.20 2.41 5.34 6.01 3.78 3.94 4.97 2.19 3.83 2 3.06 2.72 1.17 3.49 8.38 4.20 2.41 5.34 6.01 3.78 3.50 3.85 1.68 3.66

4.32

1.36 0.93 1.14

2.32

1.93 2.13 2.03

1.81 1.56 1.68

1.01 1.85 1.43

1.74 2.10 1.92

2001 2.94 3.49 3.22

4.58

1.16 1.25 1.21

1.56

1.24 2.18 1.71

1.31 2.02 1.67

0.59 1.06 0.82

1.29 1.50 1.39

2002 2.52 3.81 3.17

5.57

1.10 0.80 0.95

0.84

1.69 2.00 1.85

1.27 1.81 1.54

1.10 1.74 1.42

1.41 1.25 1.33

2003 3.07 2.99 3.03

4.29

1.07 1.01 1.04

1.26

2.25 2.60 2.42

0.57 0.51 0.54

0.84 1.66 1.25

1.90 1.75 1.82

2004 2.85 3.20 3.02

4.10

2.61 2.20 2.40

1.99

2.07 2.22 2.14

2.85 2.47 2.66

2.07 2.10 2.09

1.92 2.06 1.99

0.28

0.50 0.29 0.37

0.18

0.15 0.19 0.15

0.39 0.36 0.36

0.30 0.29 0.19

0.17 0.18 0.14

1991-2000 Mean SE 4.50 0.57 2.61 0.51 4.28 0.60

0.30

0.06 0.10 0.06

0.31

0.21 0.13 0.15

0.26 0.34 0.28

0.11 0.18 0.14

0.14 0.18 0.15

Continued

4.69

1.17 1.00 1.08

1.50

1.78 2.23 2.00

1.24 1.47 1.36

0.88 1.58 1.23

1.58 1.65 1.62

2001-2004 Mean SE 2.84 0.12 3.37 0.18 3.11 0.05

195

South South

South South

6-6 6-6 Mean

6-9 6-9 Mean

Upland Upland

5-18(83) 5-18(83) Mean

Upland Upland

North North

Perm6 Perm6 Mean

5-22 5-22 Mean

North North

4-44 4-44 Mean

6 6

6 6

5 5

5 5

4 4

4 4

Section Plot North 4 North 4

Tree 4-22 4-22 Mean

Table B.9. continued

A. negundo A. negundo

A. negundo A. negundo

A. negundo A. negundo

C. laevigata C. laevigata

J. nigra J. nigra

J. nigra J. nigra

Species C. laevigata C. laevigata

1 2

1 2

1 2

1 2

1 2

1 2

Sample 1 2

1.69 1.53 1.61

4.42 4.00 4.21

3.16 1.79 2.47

1.69 2.33 2.01

3.95 7.95 5.95

4.50 7.73 6.11

1.51 0.91 1.21

2.95 5.20 4.07

3.83 2.21 3.02

3.45 3.37 3.41

3.11 1.81 2.46

1.50 2.39 1.95

1.43 1.34 1.39

3.87 6.89 5.38

5.70 4.67 5.18

2.95 3.51 3.23

5.55 5.05 5.30

7.00 4.34 5.67

1.10 1.58 1.34

3.45 7.04 5.25

4.25 2.62 3.44

2.69 3.15 2.92

1.43 3.27 2.35

3.50 1.96 2.73

195

1.18 0.92 1.05

0.55 3.74 2.14

2.56 1.88 2.22

2.86 2.75 2.81

2.66 2.44 2.55

4.00 1.46 2.73

0.59 0.98 0.79

0.63 2.14 1.39

1.72 1.90 1.81

3.11 3.25 3.18

7.79 5.44 6.61

8.50 7.56 8.03

0.93 0.50 0.71

2.20 3.11 2.66

2.10 2.58 2.34

3.71 2.45 3.08

9.08 2.82 5.95

4.20 5.84 5.02

0.84 2.18 1.51

1.85 2.82 2.34

2.76 4.00 3.38

1.77 3.04 2.41

8.38 1.17 4.78

3.20 5.20 4.20

0.76 0.92 0.84

0.88 1.43 1.16

1.64 1.90 1.77

2.78 2.44 2.61

4.39 3.75 4.07

3.70 5.19 4.45

2.70 3.86 3.28

0.46 2.26 1.36

1.64 1.74 1.69

3.71 3.08 3.40

2.18 1.72 1.95

3.75 5.52 4.63

Annual tree-ring growth increments (mm) 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2.53 2.78 4.02 1.91 3.94 4.54 3.89 1.54 2.29 2.85 1.35 3.68 3.60 2.41 2.84 4.55 3.79 3.20 2.95 3.03 1.94 3.23 3.81 2.16 3.39 4.54 3.84 2.37 2.62 2.94

1.43 1.60 1.51

0.67 0.92 0.80

1.14 0.87 1.01

2.44 3.01 2.73

5.34 2.93 4.13

4.10 6.66 5.38

1.52 2.52 2.02

0.84 1.25 1.05

1.67 1.32 1.50

2.86 3.40 3.13

3.83 1.50 2.67

2.69 4.31 3.50

3.61 3.30 3.46

1.14 0.75 0.94

1.39 0.68 1.03

3.36 3.48 3.42

2.89 1.34 2.12

3.60 4.48 4.04

3.79 4.13 3.96

1.31 0.83 1.07

2.33 1.05 1.69

3.12 2.50 2.81

5.97 5.28 5.62

1.57 2.25 1.91

2001 2002 2003 2004 2.39 2.03 3.00 2.04 3.54 2.02 2.87 2.00 2.96 2.03 2.93 2.02

1.27 1.47 1.37

2.13 3.86 3.00

2.94 2.53 2.73

2.87 2.94 2.90

4.85 3.54 4.20

4.38 4.72 4.55

2.59 2.89 2.74

0.99 0.94 0.96

1.63 0.98 1.31

2.95 3.10 3.02

4.51 2.76 3.63

2.99 4.43 3.71

0.64 0.54 0.58

0.14 0.11 0.06

0.26 0.14 0.17

0.20 0.22 0.16

0.70 0.91 0.79

0.56 0.90 0.72

2001-2004 Mean SE 2.37 0.23 2.60 0.37 2.49 0.27

Continued

0.19 0.30 0.23

0.47 0.62 0.51

0.42 0.32 0.34

0.22 0.13 0.14

0.86 0.66 0.55

0.63 0.69 0.57

1991-2000 Mean SE 3.03 0.32 3.14 0.27 3.08 0.26

196

South South

South South

South South

South

S_sup1 S_sup1 Mean

7-20 7-20 Mean

7-25

South South

6-25 6-25 Mean

6-26 6-26 Mean

South South

6-23 6-23 Mean

7

7 7

6 6

6 6

6 6

6 6

A. glabra

A. negundo A. negundo

A. glabra A. glabra

A. negundo A. negundo

U. americana U. americana

P. deltoides P. deltoides

A. negundo A. negundo

South South

6-16 6-16 Mean

6 6

Section Plot Species South 6 A. saccharinum South 6 A. saccharinum

Tree 6-11 6-11 Mean

Table B.9. continued

2

1 2

1 2

1 2

1 2

1 2

1 2

0.83

2.36 1.30 1.83

1.77 2.23 2.00

1.62 1.67 1.65

3.79 4.66 4.23

1.94 1.47 1.71

2.53 0.93 1.73

0.58

1.85 2.32 2.08

2.36 4.68 3.52

0.88 0.92 0.90

2.77 3.50 3.13

1.68 1.26 1.47

1.67 2.02 1.85

Sample 1991 1992 1 2.08 1.76 2 1.41 1.93 1.75 1.85

0.91

1.50 2.10 1.80

2.02 1.98 2.00

4.67 2.75 3.71

2.61 3.47 3.04

2.36 1.01 1.68

1.43 2.48 1.96

1993 2.17 2.32 2.25

1.09

2.70 2.82 2.76

3.70 2.99 3.35

4.70 1.83 3.26

4.54 2.58 3.56

2.19 2.19 2.19

1.69 2.91 2.30

1994 2.10 4.78 3.44

0.67

1.43 2.40 1.92

2.61 3.50 3.05

3.20 0.59 1.90

2.69 2.49 2.59

2.57 2.61 2.59

1.77 0.38 1.07

196

0.92

2.93 0.97 1.95

2.86 4.26 3.56

3.87 0.66 2.27

2.25 2.41 2.33

2.85 2.06 2.46

3.04 0.76 1.90

0.85

2.01 1.05 1.53

1.09 1.98 1.54

3.44 1.16 2.30

2.24 2.24 2.24

2.60 1.60 2.10

2.22 2.14 2.18

1.09

3.07 1.10 2.09

2.19 0.83 1.51

2.67 0.58 1.63

3.83 3.17 3.50

2.10 3.07 2.59

1.85 1.26 1.55

1.01

1.33 1.18 1.25

1.18 0.68 0.93

2.27 1.42 1.85

2.34 1.72 2.03

2.11 2.06 2.08

2.19 1.27 1.73

1.25

1.09 0.97 1.03

2.35 1.18 1.77

3.37 2.49 2.93

3.81 4.29 4.05

1.93 2.82 2.37

1.43 0.51 0.97

Annual tree-ring growth increments (mm) 1995 1996 1997 1998 1999 2000 1.81 0.91 2.78 1.26 1.49 0.90 6.52 2.73 1.66 1.85 1.00 2.02 4.16 1.82 2.22 1.55 1.25 1.46

1.44

1.01 0.71 0.86

1.77 0.88 1.32

2.53 1.01 1.77

3.69 3.67 3.68

2.60 2.49 2.54

2.26 1.10 1.68

2001 1.08 2.95 2.01

1.92

1.35 0.89 1.12

2.44 0.80 1.62

2.02 0.75 1.39

3.09 2.19 2.64

1.92 2.02 1.97

2.06 0.51 1.28

2002 0.42 1.85 1.13

1.26

1.52 0.97 1.24

2.80 1.21 2.01

3.45 2.77 3.11

1.50 2.00 1.75

2.69 2.82 2.76

3.53 0.67 2.10

2003 1.92 2.53 2.22

1.93

1.35 0.80 1.07

1.67 0.93 1.30

3.29 2.15 2.72

2.00 2.26 2.13

1.76 1.64 1.70

5.05 0.92 2.99

2004 2.34 2.70 2.52

0.92

2.03 1.62 1.82

0.06

0.22 0.22 0.15

0.24 0.44 0.30

0.27

2.24 2.21 2.43 2.32

0.39

0.26 0.30 0.24

0.11 0.22 0.12

0.16 0.28 0.14

3.07

3.09 3.05 3.07

2.23 2.02 2.12

1.98 1.46 1.72

1991-2000 Mean SE 1.73 0.19 2.62 0.54 2.17 0.29

0.17

0.11 0.05 0.08

0.27 0.09 0.16

0.40

0.33

0.50 0.38 0.42

0.24 0.26 0.25

0.69 0.13 0.37

Continued

1.64

1.30 0.84 1.07

2.17 0.95 1.56

2.25

2.82

2.57 2.53 2.55

2.24 2.24 2.24

3.23 0.80 2.01

2001-2004 Mean SE 1.44 0.43 2.50 0.23 1.97 0.30

197

South South

South South

South South

9-1 9-1 Mean

9-12 9-12 Mean

8-11 8-11 Mean

South South

8-1 8-1 Mean

South South

South South

S_sup2 S_sup2 Mean

8-10 8-10 Mean

South South

7-34 7-34 Mean

8 8

9 9

9 9

8 8

8 8

7 7

7 7

Section Plot South 7 South 7

Tree 7-28 7-28 Mean

Table B.9. continued

A. negundo A. negundo

A. glabra A. glabra

A. glabra A. glabra

C. laevigata C. laevigata

A. negundo A. negundo

A. glabra A. glabra

P. occidentalis P. occidentalis

Species A. glabra A. glabra

1 2

1 2

1 2

1 2

1 2

1 1

1 2

2.87 3.79 3.33

0.52 1.00 0.76

1.49 2.73 2.11

3.35 2.74 3.04

2.00 1.01 1.50

0.79 1.01 0.90

1.58 1.76 1.67

3.01 5.33 4.17

0.75 1.52 1.13

2.46 2.86 2.66

3.35 2.46 2.91

2.50 1.51 2.01

0.71 1.85 1.28

1.35 1.79 1.57

Sample 1991 1992 1 0.80 0.93 2 1.01 0.58 0.91 0.76

4.00 4.59 4.30

0.52 1.81 1.17

0.91 1.52 1.21

2.94 2.71 2.83

2.50 1.01 1.76

1.18 1.09 1.13

2.06 1.84 1.95

1993 0.89 1.01 0.95

2.00 3.49 2.75

0.37 1.04 0.70

1.10 1.31 1.20

4.48 2.38 3.43

3.37 3.03 3.20

0.63 1.01 0.82

0.76 1.44 1.10

1994 1.49 1.43 1.46

1.89 2.32 2.10

0.55 0.54 0.54

2.16 1.98 2.07

3.53 2.95 3.24

0.67 1.74 1.21

0.50 0.75 0.63

2.36 2.30 2.33

197

1.42 2.61 2.01

0.56 0.54 0.55

2.69 2.74 2.72

3.44 2.72 3.08

2.52 2.67 2.60

1.39 0.76 1.07

2.57 1.67 2.12

1.98 2.86 2.42

0.41 0.41 0.41

4.27 1.98 3.13

4.21 4.50 4.36

0.93 1.42 1.17

0.88 1.01 0.95

1.81 2.05 1.93

3.44 3.78 3.61

0.48 0.79 0.63

2.44 2.11 2.27

4.30 4.69 4.50

1.09 1.18 1.14

0.88 0.67 0.78

1.09 2.02 1.55

2.35 2.10 2.23

0.36 0.96 0.66

1.35 1.77 1.56

5.97 4.61 5.29

0.93 1.33 1.13

0.93 0.84 0.88

0.72 2.14 1.43

1.80 1.90 1.85

0.27 0.50 0.38

1.25 1.43 1.34

8.00 6.24 7.12

0.76 0.75 0.76

0.93 1.01 0.97

1.34 2.40 1.87

Annual tree-ring growth increments (mm) 1995 1996 1997 1998 1999 2000 1.69 1.33 1.20 1.22 0.86 0.84 2.01 1.60 0.92 1.67 2.25 2.10 1.85 1.46 1.06 1.44 1.55 1.47

1.22 1.39 1.30

0.32 0.75 0.53

1.65 1.60 1.62

7.08 5.23 6.15

0.76 0.92 0.84

1.52 1.60 1.56

0.76 1.35 1.06

2001 1.05 1.50 1.28

1.51 0.92 1.22

0.33 0.63 0.48

1.81 1.69 1.75

5.73 4.76 5.24

1.43 0.92 1.18

0.93 1.60 1.26

1.73 1.49 1.61

2002 1.00 0.84 0.92

2.27 1.60 1.94

0.38 0.54 0.46

2.64 1.93 2.29

5.00 5.34 5.17

2.44 1.18 1.81

0.93 2.44 1.68

0.67 1.55 1.11

2003 1.06 1.59 1.32

3.67 2.60 3.13

0.70 0.51 0.60

1.80 1.52 1.66

3.00 3.37 3.19

1.85 1.00 1.42

1.77 2.19 1.98

0.59 1.27 0.93

2004 0.34 1.43 0.89

2.48 3.28 2.88

0.48 0.91 0.69

2.01 2.04 2.03

4.36 3.60 3.98

1.73 1.57 1.65

0.88 1.00 0.94

1.52 1.94 1.75

0.26 0.35 0.29

0.04 0.15 0.08

0.32 0.18 0.22

0.49 0.42 0.43

0.30 0.23 0.24

0.08 0.10 0.06

0.22 0.09 0.11

0.55 0.35 0.44

0.09 0.05 0.03

0.23 0.09 0.15

0.85 0.45 0.63

0.35 0.06 0.20

0.21 0.21 0.15

0.04 0.06 0.15

Continued

2.17 1.63 1.90

0.43 0.61 0.52

1.97 1.68 1.83

5.20 4.68 4.94

1.62 1.01 1.31

1.28 1.96 1.62

0.63 1.42 1.18

0.17 0.11

1.34 1.10

1.46 1.29

0.18 0.11

2001-2004 Mean SE

1991-2000 Mean SE

198

South South

South

South

South

South South

South

South

South South

Upland n/a Upland n/a

9-5 9-5 Mean

9-7

9-8

9-19

9-22 9-22 Mean

10-2

10-4

10-6 10-6 Mean

Upl 1 Upl 1 Mean

10 10

10

10

9 9

9

9

9

9 9

Section Plot South 9

Tree 9-4

Table B.9. continued

A. glabra A. glabra

S. nigra S. nigra

A. saccharinum

A. negundo

A. negundo A. negundo

A. negundo

A. negundo

A. negundo

A. negundo A. negundo

Species A. negundo

1 2

1 2

2

1

1 2

1

2

1

1 2

1.09 1.61 1.35

3.78 3.72 3.75

2.70

4.23

2.94 3.41 3.17

1.77

6.82

3.41

3.37 3.41 3.39

0.84 1.49 1.17

1.42 4.15 2.78

3.37

6.23

2.66 5.68 4.17

2.36

2.77

1.81

3.11 3.33 3.22

0.93 0.92 0.92

3.83 4.93 4.38

2.86

6.41

3.14 4.92 4.03

3.12

3.54

1.55

2.14 2.48 2.31

0.76 1.76 1.26

2.94 4.29 3.62

3.35

2.66

2.61 5.38 4.00

3.03

5.02

1.80

0.76 3.41 2.08

1.26 2.27 1.76

2.70 3.89 3.29

5.14

1.92

2.93 2.54 2.74

3.12

6.06

1.58

1.01 1.77 1.39

198

1.18 0.93 1.05

3.78 4.82 4.30

5.13

2.41

3.06 5.23 4.14

1.94

4.62

2.33

0.63 2.14 1.39

1.18 1.34 1.26

2.28 3.64 2.96

5.69

2.30

2.49 2.46 2.48

1.42

7.72

2.66

1.10 0.80 0.95

1.01 1.26 1.14

2.01 2.77 2.39

3.02

3.97

1.90 3.27 2.59

1.09

7.46

2.44

1.05 1.80 1.42

1.01 1.26 1.14

2.61 3.82 3.22

4.80

2.74

4.48 2.42 3.45

1.26

7.24

3.28

1.05 1.05 1.05

1.18 1.76 1.47

3.79 1.95 2.87

1.85

2.70

4.25 3.45 3.85

2.19

0.67

2.86

1.81 2.10 1.96

1.27 1.43 1.35

3.19 3.27 3.23

2.02

2.94

3.76 3.46 3.61

1.43

0.68

3.01

2.53 2.51 2.52

0.42 1.24 0.83

1.77 4.57 3.17

1.18

3.90

3.84 2.65 3.24

1.85

0.84

2.36

2.00 2.31 2.15

0.49 0.92 0.71

2.86 3.96 3.41

2.78

3.07

3.52 3.22 3.37

1.77

1.10

5.37

2.20 2.76 2.48

0.59 1.02 0.80

1.85 6.33 4.09

4.60

1.31

4.05 3.41 3.73

2.52

0.59

2.27

2.86 2.18 2.52

Annual tree-ring growth increments (mm) Sample 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 1 5.20 4.97 3.45 3.51 4.43 5.87 6.40 4.88 6.75 3.87 2.33 2.78 4.42 3.34

1.04 1.46 1.25

2.92 3.80 3.36

3.79

3.56

3.05 3.88 3.46

2.13

5.19

2.37

1.60 2.23 1.92

0.05 0.13 0.07

0.27 0.28 0.21

0.41

0.51

0.25 0.41 0.21

0.24

0.73

0.22

0.31 0.30 0.27

1991-2000 Mean SE 4.93 0.37

0.19 0.12 0.14

0.36 0.66 0.21

0.73

0.54

0.11 0.18 0.11

0.23

0.11

0.72

0.19 0.13 0.09

Continued

0.69 1.15 0.92

2.42 4.53 3.48

2.64

2.80

3.79 3.18 3.49

1.89

0.80

3.25

2.40 2.44 2.42

2001-2004 Mean SE 3.22 0.45

199

Section Plot Upland n/a Upland n/a

Upland n/a Upland n/a

Upland n/a Upland n/a

Upland n/a Upland n/a

Upland n/a Upland n/a

Upland n/a Upland n/a

Tree Upl 2 Upl 2 Mean

Upl 3 Upl 3 Mean

Upl 4 Upl 4 Mean

Upl 5 Upl 5 Mean

Upl 6 Upl 6 Mean

Upl 7 Upl 7 Mean

Table B.9. continued

A. glabra A. glabra

A. negundo A. negundo

P. occidentalis P. occidentalis

J. nigra J. nigra

C. laevigata C. laevigata

Species M. rubra M. rubra

1 2

1 2

1 2

1 2

1 2

1.56 2.34 1.95

1.50 1.94 1.72

2.18 2.18

4.29 2.27 3.28

1.68 2.36 2.02

2.51 1.57 2.04

1.25 3.53 2.39

2.44 2.44

4.16 2.44 3.30

2.20 1.68 1.94

Sample 1991 1992 1 5.64 4.69 2 3.53 3.41 4.58 4.05

1.76 1.10 1.43

1.51 4.14 2.83

1.51 1.51

3.18 4.24 3.71

0.93 0.67 0.80

1993 2.86 4.50 3.68

2.02 1.09 1.56

1.01 5.04 3.02

1.60 2.69 2.15

3.44 2.65 3.05

0.92 0.93 0.93

1994 3.20 5.13 4.16

1.26 1.60 1.43

2.00 4.45 3.22

1.17 1.76 1.47

3.80 4.05 3.92

1.43 0.83 1.13

199

1.43 0.92 1.17

1.05 5.21 3.13

1.75 2.20 1.97

3.17 3.71 3.44

1.18 0.66 0.92

1.26 0.75 2.93 2.72 2.09 1.74

2.05 1.80 4.61 5.79 3.33 3.79

1.01 2.50 2.40 0.59 1.70 1.55

4.41 5.31 5.23 5.15 4.82 5.23

1.60 1.09 1.17 1.42 1.39 1.25

0.75 2.71 1.73

1.41 4.30 2.85

2.67 0.99 1.83

5.48 4.78 5.13

1.51 1.26 1.39

2001 2.85 4.77 3.81

0.67 1.07 3.01 1.38 1.84 1.23

1.22 0.75 5.40 5.17 3.31 2.96

1.59 1.75 0.78 0.56 1.19 1.16

5.75 5.18 6.22 5.88 5.99 5.53

1.01 0.84 1.26 1.26 1.13 1.05

Annual tree-ring growth increments (mm) 1995 1996 1997 1998 1999 2000 3.29 3.53 3.79 2.53 3.11 2.44 5.22 5.05 4.14 3.38 4.03 4.47 4.25 4.29 3.96 2.95 3.57 3.45

1.33 1.60 1.46

1.78 4.45 3.11

2.70 0.67 1.69

5.34 5.60 5.47

1.35 2.70 2.02

2002 2.69 2.70 2.70

2004 3.95 3.03 3.49

0.66 1.24 1.84 1.43 1.25 1.33

2.33 1.17 4.08 3.27 3.20 2.22

0.95 0.81 0.43 1.18 0.69 0.99

3.82 4.80 4.13 3.70 3.98 4.25

0.68 2.25 1.77 2.11 1.23 2.18

2003 1.83 3.45 2.64

1.40 2.00 1.70

1.48 4.44 2.96

1.76 1.75 1.80

4.30 4.07 4.19

1.35 1.22 1.29

0.19 0.26 0.09

0.12 0.35 0.18

0.24 0.24 0.12

0.30 0.42 0.32

0.13 0.16 0.13

1991-2000 Mean SE 3.51 0.31 4.29 0.22 3.90 0.15

1.08 0.15 1.56 0.10 1.32 0.05

1.51 0.35 4.24 0.39 2.87 0.22

1.55 0.43 0.71 0.16 1.13 0.21

4.78 0.34 4.83 0.54 4.80 0.40

1.28 0.35 1.96 0.30 1.62 0.28

2001-2004 Mean SE 2.83 0.43 3.49 0.45 3.16 0.29

APPENDIX C EXPERIMENTAL WETLANDS SOIL AND SEDIMENT DATA (2004)

200

Cover Coord. Wetland Type x y 1 EM 1 4 1 EM 1 5 1 EM 1 7 1 OW 1 8 1 OW 1 9 1 OW 1 10 1 EM 1 11 1 EM 1 13 1 EM 1 14 1 EM 2 3 1 EM 2 4 1 EM 2 5 1 EM 2 6 1 OW 2 7 1 OW 2 8 1 OW 2 11 1 EM 2 12 1 EM 2 14 1 EM 2 15 1 EM 3 3 1 EM 3 5 1 EM 3 6 1 EM 3 7 1 EM 3 8 1 EM 3 9 1 EM 3 10 1 EM 3 11 1 EM 3 12 1 EM 3 13 1 EM 3 14 1 EM 3 16 1 EM 4 2 1 EM 4 5 1 EM 4 6 1 EM 4 7 1 EM 4 8 1 EM 4 9 1 EM 4 11 1 EM 4 12 1 EM 4 13 1 EM 4 14 1 OW 4 15 1 OW 4 16 1 EM 4 17

Sediment

Bulk density

Moisture

Organic

Organic

depth (cm) 5.0 7.0 8.0 9.5 19.0 21.5 9.0 8.0 5.0 13.0 11.0 9.0 9.0 20.5 14.0 16.0 10.0 10.0 8.0 5.0 9.0 8.0 6.5 3.5 1.5 4.0 3.5 2.0 7.0 8.0 2.0 5.0 5.0 10.0 13.0 8.0 8.5 5.0 -8.0 4.0 9.0 ---

(g cm-3) [0-8 cm] [8-16 cm] 0.65 1.37 0.51 1.08 0.48 1.23 0.39 0.65 0.66 -0.37 -0.61 1.25 0.69 1.33 0.6 1.3 0.49 1.05 0.47 1.36 0.46 1.03 0.49 -0.49 -0.38 -0.25 -0.6 0.93 0.59 1.19 0.80 1.34 0.73 1.50 0.45 1.28 0.56 1.23 0.57 1.40 1.03 1.34 1.48 1.67 1.19 1.66 0.87 1.53 0.89 1.27 0.76 1.37 0.85 1.62 0.96 1.51 0.81 1.27 0.98 1.41 0.54 1.14 0.46 0.77 0.54 1.36 0.50 1.31 0.62 -0.66 1.17 0.42 1.07 --0.38 1.08 -0.37 0.78 1.44

content (%) [0-8 cm] [8-16 cm] 49.7 30.3 60.9 29.7 56.7 31.5 64.2 49.0 50.2 59.9 65.6 68.4 54.0 28.5 48.1 29.1 50.1 27.5 60.8 44.7 60.8 33.9 63.2 31.1 58.3 31.8 61.7 66.7 66.7 40.8 72.9 70.3 47.7 39.5 57.1 33.7 44.0 27.5 48.6 22.2 62.4 30.9 52.7 25.8 53.0 25.5 31.5 25.5 23.5 22.9 29.7 22.3 33.5 24.4 34.7 27.5 50.5 26.4 43.5 25.0 34.9 24.0 41.3 26.8 40.2 25.3 57.8 33.4 61.8 52.4 58.6 30.0 64.5 35.7 52.0 53.4 31.0 60.7 37.8 --63.4 36.0 50.8 51.5 45.5 24.1

matter (%) [0-8 cm] [8-16 cm] 11.8 5.3 10.9 5.6 8.3 5.5 8.7 5.4 6.6 7.6 8.5 7.9 7.8 4.6 7.5 4.8 14.0 4.9 11.2 6.4 9.8 3.9 8.9 5.0 7.7 4.1 8.8 7.5 8.5 4.5 7.3 5.6 6.3 5.5 10.3 5.0 8.4 4.6 6.2 3.6 10.2 4.6 7.3 4.3 7.9 4.5 4.8 4.0 4.2 4.2 4.6 3.7 4.2 3.5 4.6 4.0 10.0 4.7 7.9 4.1 5.2 4.2 6.7 4.3 5.4 4.1 10.8 5.0 11.8 6.8 10.4 4.7 10.8 5.5 10.8 6.2 9.3 5.2 9.7 6.0 -5.4 8.1 -7.1 6.7 9.8 4.0

C (%) [0-8 cm] 4.02 3.78 3.12 3.06 2.70 3.17 2.99 2.92 4.58 4.05 3.50 3.26 2.96 2.57 3.18 2.83 2.60 3.62 2.86 2.59 3.59 2.87 3.03 2.22 2.08 2.19 2.07 2.17 3.57 3.02 2.33 2.72 3.15 3.75 4.01 3.64 3.76 3.75 3.37 4.15 -3.02 2.82 3.49

Continued Table C.1. Physiochemical soil characteristics at 0-8 and 8-16 cm depths in 2004. Percent organic C results in bold-type were lab-analyzed and those in regular-type were based on regression analysis with percent organic matter. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text). 201

Table C.1. continued

Cover Coord. Wetland Type x y 1 EM 5 2 1 OW 5 3 1 OW 5 4 1 EM 5 5 1 EM 5 6 1 EM 5 7 1 EM 5 8 1 EM 5 12 1 EM 5 13 1 EM 5 14 1 OW 5 15 1 EM 5 17 1 EM 6 3 1 EM 6 4 1 EM 6 6 1 EM 6 14 1 OW 6 16 1 EM 6 17 1 EM 7 15 1 EM 7 16 2 EM 8 5 2 EM 9 3 2 OW 9 4 2 EM 9 5 2 EM 9 6 2 EM 9 7 2 EM 9 16 2 EM 9 17 2 EM 9 18 2 EM 10 3 2 OW 10 4 2 OW 10 5 2 OW 10 6 2 EM 10 7 2 EM 10 8 2 EM 10 9 2 EM 10 15 2 EM 10 16 2 OW 10 17 2 EM 10 19 2 EM 11 3 2 OW 11 4 2 OW 11 6 2 EM 11 9 2 EM 11 10 2 EM 11 11 2 EM 11 12 2 EM 11 13 2 EM 11 16 2 OW 11 17 2 OW 11 18 2 EM 12 4 2 EM 12 6

Sediment

Bulk density

Moisture

Organic

Organic

depth (cm) 12.0 10.0 17.0 11.0 9.0 9.0 10.0 -5.0 -12.0 -5.0 2.0 --8.0 5.5 ---12.0 17.0 13.0 9.0 --7.0 5.5 15.0 12.0 22.0 18.0 8.0 10.0 -4.0 6.5 16.0 8.0 9.0 20.0 15.0 3.0 11.0 8.0 8.0 9.0 5.0 12.0 9.0 11.0 7.0

(g cm-3) [0-8 cm] [8-16 cm] 0.49 1.24 0.48 -0.50 -0.55 1.13 -0.93 0.62 0.98 0.43 1.21 0.81 1.43 0.65 1.29 0.74 -0.49 0.64 1.03 1.54 0.51 1.35 0.68 1.09 0.82 1.15 0.62 1.09 0.47 -0.74 1.74 0.53 -0.80 1.39 0.81 1.37 0.48 0.50 0.62 0.68 -1.59 0.35 -0.86 1.43 0.88 1.25 0.59 1.40 0.63 1.55 0.48 0.36 0.48 1.22 0.43 0.52 0.52 0.51 0.57 -0.54 1.37 0.99 1.41 0.91 1.44 0.54 1.48 0.46 0.64 0.55 1.29 0.51 1.40 0.32 -0.50 0.73 0.92 -0.38 1.41 0.58 1.39 0.73 1.35 0.49 0.86 0.82 1.49 0.81 1.27 0.85 --0.74 0.48 1.58

content (%) [0-8 cm] [8-16 cm] 58.2 34.4 62.3 54.7 59.1 56.0 57.1 36.8 -36.8 56.8 40.1 66.1 37.7 38.7 27.6 51.5 31.1 45.3 54.9 -55.9 37.3 25.1 59.5 25.6 53.9 35.6 41.6 32.7 53.1 34.7 58.1 -43.7 24.5 49.5 31.2 44.6 27.8 37.0 25.6 61.6 61.8 54.8 54.2 29.8 23.6 67.7 -37.4 27.9 36.7 27.5 58.8 27.5 43.2 25.2 58.3 70.4 59.9 39.0 61.1 50.3 62.0 57.6 56.2 26.7 57.0 29.8 32.9 26.3 38.6 28.0 57.0 27.1 65.0 52.6 57.0 27.7 65.0 26.4 65.1 67.6 62.0 54.3 38.5 -63.4 31.8 58.9 28.9 53.7 32.1 56.1 44.9 38.6 26.3 50.0 32.4 56.0 36.8 -54.3 57.0 25.8

matter (%) [0-8 cm] [8-16 cm] 9.8 4.5 8.5 6.8 7.2 7.0 9.6 5.3 -5.0 10.7 5.2 15.5 5.9 10.2 4.9 11.2 5.6 6.5 7.1 7.4 6.8 7.0 5.2 10.8 4.3 11.7 5.5 9.9 5.9 9.4 5.2 6.8 -7.3 5.3 10.7 5.3 8.0 5.1 9.4 4.5 13.5 9.3 7.3 7.9 5.0 3.7 12.4 -10.5 5.8 11.5 5.6 -4.8 5.3 4.7 12.3 12.1 8.6 5.6 8.3 6.4 8.0 7.4 8.5 4.0 13.5 4.9 8.8 5.0 6.9 4.8 8.3 5.0 8.2 6.2 9.3 5.2 12.3 4.1 10.2 9.7 7.7 6.3 6.2 -11.2 -9.1 5.1 8.7 5.4 8.5 6.4 5.6 4.9 6.3 5.2 7.5 5.3 13.3 5.3 8.5 4.1

C (%) [0-8 cm] 3.70 3.17 2.84 3.66 -3.73 4.96 3.61 3.85 2.49 2.90 2.78 3.93 3.99 3.53 3.41 2.73 2.72 3.74 3.05 3.41 4.45 2.86 2.28 4.16 3.68 3.93 -2.37 4.13 3.20 3.13 3.05 3.71 4.45 3.24 2.77 3.11 3.09 3.37 4.24 3.61 2.81 2.01 3.85 3.33 3.23 3.17 2.44 2.39 2.92 4.40 3.17

Continued 202

Table C.1. continued

Cover Coord. Wetland Type x y 2 EM 12 8 2 EM 12 9 2 EM 12 10 2 EM 12 11 2 EM 12 14 2 EM 12 16 2 EM 12 17 2 EM 12 18 2 EM 13 5 2 EM 13 6 2 EM 13 7 2 EM 13 8 2 OW 13 9 2 OW 13 10 2 OW 13 11 2 OW 13 12 2 OW 13 13 2 EM 13 14 2 EM 13 15 2 EM 13 17 2 EM 14 6 2 EM 14 7 2 EM 14 8 2 EM 14 9 2 OW 14 11 2 OW 14 12 2 OW 14 13 2 EM 14 15 2 EM 14 16 2 EM 15 8 2 EM 15 9 2 EM 15 11 2 EM 15 12 2 EM 15 13 2 EM 15 14

Sediment

Bulk density

Moisture

Organic

Organic

depth (cm) -5.0 6.0 6.5 5.0 6.0 5.0 4.0 9.0 10.0 7.0 7.0 21.5 11.0 10.0 14.0 15.0 -6.0 8.0 9.0 1.5 11.0 7.5 13.0 19.0 17.0 8.0 19.0 --6.0 8.0 8.0 --

(g cm-3) [0-8 cm] [8-16 cm] --0.78 1.63 0.60 1.47 0.68 1.43 0.83 1.01 0.86 1.16 0.66 1.28 0.64 1 0.38 -0.44 1.2 0.69 1.51 0.46 1.13 0.32 -0.3 1.07 0.45 1.16 0.37 0.65 0.33 0.73 -1.56 0.60 1.34 0.49 0.75 0.50 -1.09 -0.30 1.02 0.55 -0.61 0.58 --0.33 -0.45 1.46 0.85 0.91 1.31 -1.17 1.34 0.64 1.50 0.55 1.42 0.37 1.23 1.18 1.45

content (%) [0-8 cm] [8-16 cm] -62.0 45.4 24.3 47.6 24.6 45.0 27.3 41.6 29.0 44.6 31.6 48.5 28.7 44.9 37.4 67.8 -63.2 34.0 48.8 26.5 64.3 35.1 67.7 64.3 68.5 35.8 61.3 32.1 64.3 51.9 65.3 66.9 -25.6 51.1 28.0 60.8 44.7 58.0 -33.9 37.7 74.5 39.6 58.2 -56.5 53.4 68.3 -69.9 64.0 58.6 29.8 41.8 41.6 22.4 26.2 30.2 27.0 49.1 24.9 58.3 28.5 63.8 32.2 29.5 26.5

matter (%) [0-8 cm] [8-16 cm] 12.8 8.7 7.5 4.2 6.6 4.1 6.4 4.4 5.3 4.8 8.0 5.0 8.3 4.9 7.5 5.3 16.6 -12.1 4.8 9.1 3.5 11.3 5.3 8.8 7.8 6.8 5.1 8.3 5.0 7.7 5.7 7.9 7.6 -4.8 8.5 5.1 11.5 7.2 12.4 -4.5 4.4 18.9 5.3 10.4 -6.1 5.9 8.3 -8.1 7.8 10.6 4.9 6.7 6.4 5.6 5.6 5.5 -8.2 4.2 -4.5 19.1 4.9 7.2 5.1

C (%) [0-8 cm] 4.26 2.92 2.70 2.64 2.36 3.03 3.12 3.39 5.13 4.08 3.33 3.89 3.18 2.74 3.12 2.98 1.78 -3.60 4.31 4.15 2.16 5.82 3.66 2.55 3.12 3.06 3.71 2.73 2.43 2.40 3.09 -5.58 2.84

203

Wetland 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Cover Coord. Type x y EM 1 5 OW 1 8 OW 1 10 EM 1 14 EM 2 3 EM 2 4 EM 2 5 EM 2 6 OW 2 7 OW 2 11 EM 2 12 EM 2 14 EM 2 15 EM 3 3 EM 3 6 EM 3 8 EM 3 10 EM 3 11 EM 3 12 EM 3 13 EM 3 14 EM 4 2 EM 4 5 EM 4 7 EM 4 9 EM 4 11 EM 4 12 EM 4 13 OW 4 15 EM 4 17 EM 5 2 OW 5 3 EM 5 5 EM 5 6 EM 5 12 EM 5 13 EM 5 14 OW 5 15 EM 6 3 EM 6 4 EM 6 14

Total C

Total N

Total P

(%) [0-8 cm] [8-16 cm] --4.33 -----4.07 -4.39 1.71 ----4.06 4.46 4.59 4.09 --3.90 -2.87 1.59 --------------4.03 ---3.64 ---4.38 1.61 3.48 ---4.39 1.87 4.12 1.63 3.41 1.47 3.96 -4.48 -4.14 1.84 4.19 -----2.70 ---4.07 -4.64 ----

(%) [0-8 cm] [8-16 cm] --0.38 -----0.37 -0.40 0.16 ----0.32 0.39 0.37 0.31 --0.38 -0.29 0.16 --------------0.38 ---0.31 ---0.43 0.17 0.31 ---0.40 0.19 0.38 0.18 0.30 0.14 0.35 -0.36 -0.39 0.18 0.41 -----0.28 ---0.35 -0.44 ---

(μg g ) [0-8 cm] ----793 ---851 764 --679 ----------813 704 654 --731 ---------681 ---

-1

[0-8 cm] -7.49 7.48 6.82 7.02 7.47 7.64 7.57 7.44 7.63 -7.84 7.12 7.54 7.55 7.84 7.40 7.38 7.40 6.95 7.08 -7.52 6.39 6.15 6.85 6.51 7.48 7.55 6.74 7.29 7.62 7.72 6.79 6.86 6.84 7.78 7.66 6.86 7.25 6.60

pH [8-16 cm] 6.43 7.35 7.49 7.23 6.99 7.34 7.24 7.50 7.59 7.55 7.50 6.63 7.42 7.21 7.43 7.86 7.30 7.01 6.97 7.48 -7.48 7.09 7.21 6.72 7.19 7.07 7.72 7.36 7.38 7.32 7.75 6.73 6.90 6.98 6.09 7.83 7.58 6.48 7.27 6.74

Continued Table C.2. Total C, total N, total P, and pH of experimental wetland soils at 0-8 and 816 cm depths in 2004. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text). 204

Table C.2. continued

Wetland 1 1 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

Cover Coord. Type x y EM 6 17 EM 7 15 EM 9 3 EM 9 5 EM 9 7 EM 9 16 EM 9 17 OW 10 4 OW 10 5 EM 10 7 EM 10 8 EM 10 9 EM 10 15 EM 10 16 EM 10 19 EM 11 3 OW 11 6 EM 11 9 EM 11 11 EM 11 12 EM 11 13 OW 11 17 OW 11 18 EM 12 4 EM 12 8 EM 12 9 EM 12 10 EM 12 18 EM 13 5 EM 13 6 EM 13 7 OW 13 9 OW 13 11 OW 13 13 EM 13 14 EM 13 15 EM 13 17 EM 14 8 OW 14 12 EM 14 16 EM 15 9 EM 15 11 EM 15 13

Total C

Total N

Total P

(%) [0-8 cm] [8-16 cm] 2.76 1.65 ----1.23 1.07 ----3.50 -4.64 ---3.95 -----3.04 1.64 3.21 -3.58 -4.26 1.00 4.20 3.51 2.05 -3.56 1.45 --3.63 3.07 2.15 ----------3.43 -5.15 -4.45 -4.14 0.95 4.39 4.17 4.37 -2.39 1.96 --3.69 1.69 4.33 2.32 ----2.36 -2.08 1.43 --5.59 --

(%) [0-8 cm] [8-16 cm] 0.28 0.17 ----0.14 0.13 ----0.35 -0.33 ---0.34 -----0.26 0.17 0.31 -0.30 -0.35 0.11 0.31 0.27 0.19 -0.35 0.15 --0.36 -0.27 0.23 ----------0.27 -0.48 -0.41 -0.38 0.11 0.40 0.36 0.35 -0.21 0.19 --0.33 0.16 0.42 0.21 ----0.23 -0.18 0.13 --0.39 --

(μg g ) [0-8 cm] 620 ----------------650 715 -688 772 ------705 --863 844 747 --831 ---528 -575

205

-1

[0-8 cm] 7.07 6.70 6.76 7.01 7.12 -6.97 7.50 7.52 7.35 6.87 7.27 6.07 7.29 6.59 6.51 7.51 6.65 7.34 6.96 6.40 7.65 7.43 6.52 7.52 7.12 7.28 6.50 6.95 6.97 7.52 7.59 7.66 7.66 -7.48 6.12 5.59 7.55 7.26 7.06 6.26 6.71

pH [8-16 cm] 6.32 6.68 6.65 6.82 7.31 6.96 6.37 7.39 -6.38 6.44 -6.17 6.25 6.59 6.72 7.52 -6.73 6.96 7.29 7.29 7.41 6.60 7.72 6.70 -7.65 6.94 -7.45 7.62 -7.52 7.70 6.50 6.23 6.30 -6.89 6.28 -5.80

Cover Wetland

Type

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

EM OW OW EM EM EM EM EM OW OW EM EM EM EM EM EM EM EM EM EM EM EM EM EM EM EM EM EM OW EM EM OW EM EM EM EM EM OW EM EM EM EM EM

Coord. x y 1 5 1 8 1 10 1 14 2 3 2 4 2 5 2 6 2 7 2 11 2 12 2 14 2 15 3 3 3 6 3 8 3 10 3 11 3 12 3 13 3 14 4 2 4 5 4 7 4 9 4 11 4 12 4 13 4 15 4 17 5 2 5 3 5 5 5 6 5 12 5 13 5 14 5 15 6 3 6 4 6 14 6 17 7 15

Avail. P

Exch. Ca

Exch. Mg

Exch. K

(μg g-1) [0-8 cm] [8-16 cm] -3 1 5 1 1 14 4 14 9 9 6 9 7 2 10 2 2 1 1 -3 8 5 11 6 7 4 7 3 12 5 4 3 8 6 3 2 12 5 14 --9 5 7 13 8 5 4 9 6 9 4 7 8 2 3 13 4 11 13 2 2 5 4 12 10 3 2 10 2 8 8 1 2 7 5 21 12 7 3 6 2 8 4

(μg g-1) [0-8 cm] [8-16 cm] -2243 4271 2952 4050 4312 3466 2108 2811 2125 3198 1827 3621 2449 3343 2756 4470 4017 4059 4383 -2072 3395 2009 2582 2110 2472 1504 2942 1922 2273 1971 2240 1766 2547 1665 2019 1820 2455 2157 2924 --1998 3474 2076 2749 2417 2595 2078 2518 2196 2614 2170 3519 2706 4253 2275 2291 1865 2964 2373 4003 3592 3601 2190 2898 2734 2070 2047 2821 1967 3055 3072 3672 3955 2885 1885 3225 2418 2446 2148 2869 2105 2211 2018

(μg g-1) [0-8 cm] [8-16 cm] -363 469 365 465 415 657 373 474 360 375 232 465 318 380 328 392 426 465 419 -320 370 323 478 382 251 222 330 270 346 409 321 326 313 285 322 336 375 337 531 --369 417 386 472 396 428 342 440 391 437 391 372 317 440 336 424 427 481 369 434 376 440 328 552 488 415 404 491 348 392 378 443 418 488 328 615 422 470 429 492 380 414 411

(μg g-1) [0-8 cm] [8-16 cm] -78 206 156 211 204 184 129 128 101 165 71 184 80 164 105 177 208 208 189 -126 166 122 138 111 90 48 114 60 105 101 121 120 124 116 104 98 128 92 167 --102 150 97 217 144 186 90 148 138 172 93 195 109 221 165 169 105 176 130 200 202 210 103 179 168 119 81 158 72 142 148 224 197 155 105 201 137 152 113 139 113 176 107

Continued Table C.3. Available P, exchangeable cations (Ca, Mg and K) of the experimental wetland soils at 0-8 and 8-16 cm depths in 2004. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text). 206

Table C.3. continued Cover Wetland

Type

2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

EM EM EM EM EM OW OW EM EM EM EM EM EM EM OW EM EM EM EM OW OW EM EM EM EM EM EM EM EM OW OW OW EM EM EM EM OW EM EM EM EM

Coord. x y 9 3 9 5 9 7 9 16 9 17 10 4 10 5 10 7 10 8 10 9 10 15 10 16 10 19 11 3 11 6 11 9 11 11 11 12 11 13 11 17 11 18 12 4 12 8 12 9 12 10 12 18 13 5 13 6 13 7 13 9 13 11 13 13 13 14 13 15 13 17 14 8 14 12 14 16 15 9 15 11 15 13

Avail. P

Exch. Ca

Exch. Mg

Exch. K

(μg g-1) [0-8 cm] [8-16 cm] 12 8 9 5 12 6 -3 6 2 1 5 2 -8 4 7 7 9 -7 4 9 6 6 4 12 6 3 3 7 -14 10 12 7 8 14 2 5 4 4 6 10 11 5 6 6 6 -8 4 21 8 20 -14 12 3 2 2 -2 4 -5 8 10 11 8 13 11 2 -10 9 7 6 6 -8 5

(μg g-1) [0-8 cm] [8-16 cm] 2689 2061 2080 1802 2883 2346 -2198 2988 1997 3590 2777 3682 -3321 1797 2641 1934 3071 -2498 2055 1911 2908 3000 2105 2595 1805 3665 4087 2277 -2943 2056 3146 2207 3330 2647 3407 2488 2825 2600 3038 1889 3532 3404 2322 1975 2961 -2431 3027 2842 1980 2986 -3105 1790 4137 4854 3702 -3117 2844 -1967 3261 2156 2890 2270 2600 1811 4278 -2281 2220 2019 1688 1960 -2447 1502

(μg g-1) [0-8 cm] [8-16 cm] 449 379 306 331 509 425 -440 538 344 400 301 363 -405 254 467 360 500 -432 335 344 373 376 319 462 351 377 411 274 -491 298 383 278 427 353 357 357 383 348 570 368 383 269 243 254 298 -303 226 508 370 514 -445 318 419 430 377 -330 338 -388 357 301 448 336 538 333 428 -363 360 359 327 285 -434 298

(μg g-1) [0-8 cm] [8-16 cm] 149 154 164 137 181 138 -108 172 125 193 163 186 -161 65 152 137 152 -138 123 151 147 150 88 210 106 175 199 90 -189 129 160 123 165 149 192 163 167 164 240 119 201 111 69 54 140 -122 84 172 91 165 -140 109 228 233 214 -160 163 -72 159 133 200 139 187 131 211 -136 134 130 111 138 -127 117

207

Wetland Sample 1 4,7 1 5,3 1 2,7 1 6,16 1 4,15 2 13,6 2 11,13 2 13,17 2 13,11 2 11,17 Mean SE

Cover zone EM OW OW OW OW EM EM EM OW OW

Subbasin Out In Mid Out Out In Mid Out Mid Out

Sediment composition % Sand % Silt % Clay 51.5 41.7 6.8 27.7 55.2 17.1 22.7 61.5 15.8 40.9 40.3 18.8 36.0 47.2 16.8 40.3 52.4 7.3 48.6 46.9 4.5 28.0 54.9 17.1 38.7 47.7 13.6 34.0 50.1 15.9 36.9 2.9

49.8 2.0

13.4 1.6

Table C.4. Percent and mean (±1 SE) textural classes of sediment in the experimental wetlands in 2004. Sample coordinates based on the 10x10 m grid system at the experimental wetlands; cover type consisted of emergent (EM) and open water (OW) zones; and sub-basin refers to OW zone in proximity to wetland inflow/outflow (see text).

208

209

Type EM OW OW EM EM EM EM OW EM EM EM EM EM OW EM OW OW OW EM EM EM

x 2 2 2 2 4 4 4 4 6 6 11 11 11 11 13 13 13 13 13 15 15

y 3 7 11 15 7 9 11 15 3 17 9 11 13 17 5 9 11 13 17 9 13

(μg g-1) (μg g-1) (μg g-1) (μg g-1) (μg g-1) (μg g-1) (μg g-1) (μg g-1) (μg g-1) 38564 42.90 4582 27.34 24533 9477 4131 161.90 8.22 45950 38.79 46254 30.86 27939 11221 6458 242.28 8.95 44038 38.64 58255 29.27 26543 10842 5864 244.78 8.10 45768 36.80 4285 28.74 27018 10775 4621 162.36 9.32 44069 36.72 4767 29.87 24457 10756 4635 166.88 8.71 42547 35.94 4351 28.09 26004 10230 4346 148.92 9.96 38880 35.88 4414 24.66 25906 9588 4191 163.39 9.29 46170 37.62 37207 29.42 27054 11187 5475 196.36 6.94 39073 31.27 7940 29.15 25192 9558 4362 160.11 9.78 40115 31.38 4602 25.17 28500 9617 4104 173.02 8.87 40963 40.45 4152 26.41 30962 10287 3981 235.81 9.31 41592 33.62 6486 28.94 24524 10443 4534 160.89 8.52 40888 27.74 6434 27.37 24998 10338 4242 228.88 8.18 41811 28.97 24912 27.00 26716 10675 4451 272.98 8.01 49572 41.31 4862 34.16 23801 12214 4898 144.57 9.12 46906 33.74 39942 32.39 26409 11968 5814 219.78 7.92 42439 40.87 50012 29.38 25381 10909 5211 244.68 9.43 38463 27.10 22857 25.00 27019 9848 4121 311.96 8.08 51012 48.11 4813 32.96 29444 12543 4909 192.63 9.88 36114 35.46 3401 23.00 30897 9214 3632 245.26 9.40 36149 37.42 4235 25.26 21729 8921 3675 147.79 9.43

(μg g-1) 446.00 529.05 551.94 476.86 487.29 467.17 435.26 544.83 461.99 445.09 425.69 454.49 451.33 464.55 502.90 535.34 523.30 433.38 520.54 387.49 417.51

(μg g-1) 1021.44 5909.58 6267.34 741.23 1365.80 1240.96 865.70 3533.63 2005.91 915.78 568.13 1189.28 1329.46 3116.29 1262.34 5642.61 5151.37 2368.82 1443.85 343.66 1100.50

(μg g-1) 121.66 143.61 131.96 136.19 141.76 127.80 115.92 131.35 132.17 112.74 119.50 137.74 131.42 125.70 161.37 154.49 141.16 117.02 150.59 110.52 101.58

209

Table C.5. Micronutrient concentrations of sediment (0-8 depth) at the experimental wetlands in 2004. Coordinates based on the 10x10 m grid system at the experimental wetlands and cover type consisted of emergent (EM) and open water (OW) zones (see text).

Wetland 1 1 1 1 1 1 1 1 1 1 2 2 2 2 2 2 2 2 2 2 2

Cover Coord. Total Al Total B Total Ca Total Cu Total Fe Total K Total MgTotal MnTotal Mo Total Na Total S Total Zn