Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

The Fly River, Papua New Guinea: Environmental Studies in an Impacted Tropical River System
The Fly River, Papua New Guinea: Environmental Studies in an Impacted Tropical River System
The Fly River, Papua New Guinea: Environmental Studies in an Impacted Tropical River System
Ebook984 pages

The Fly River, Papua New Guinea: Environmental Studies in an Impacted Tropical River System

Rating: 0 out of 5 stars

()

Read preview

About this ebook

In 1984 the OK Tedi Mining Company Limited began mining copper and gold mineralization from Mt. Fubilan, which is located at the headwaters of the OK Tedi. Subsequent mining in the region followed in 1990. Since this time there has been intense monitoring of the environment undertaken by those in the field in order to better understand the possible impact of mining. This book assembles and summarizes research spanning two decades undertaken by leading experts with firsthand experience. Much of this research is contained in internal company reports, giving the reader rare insight and firsthand knowledge.

* documents physical and biologic change in a large tropical river system brought about largely by mining in an otherwise pristine environment.
* this book brings together a broad rand of disciplines to provide a comprehensive overview of change in a complex and dynamical tropical river system based largely on previously unpublished company reports.
* the book provides examples of state-of-the-art strategies and methodologies for monitoring environmental impact in a large river system.
LanguageEnglish
Release dateJan 9, 2009
ISBN9780080558837
The Fly River, Papua New Guinea: Environmental Studies in an Impacted Tropical River System

Related to The Fly River, Papua New Guinea

Titles in the series (6)

View More

Earth Sciences For You

View More

Reviews for The Fly River, Papua New Guinea

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    The Fly River, Papua New Guinea - Elsevier Science

    Australia

    Preface

    Barrie Bolton


    Environment Department, Ok Tedi Mining Limited

    Publisher Summary

    Ever since the discovery of rich copper and gold mineralization under Mount Fubilan in the rugged central high lands of western Papua New Guinea there have been concerns about the potential for environmental impacts on the nearby Fly River and in particular its tributary, the Ok Tedi. These concerns prompted a wide-ranging and comprehensive series of studies of the river system designed, in the first instance, to develop an understanding of the pre- mine environmental setting, and then following the initiation of mining in 1984, the nature and extent of mine-related impacts. Through the combined efforts of various universities, government departments, individuals and in particular, the management of Ok Tedi Mining Limited, the owner and operator of the Ok Tedi mine, this on-going attempt to monitor environmental impact has resulted in the assembly of an enormous body of work on this important river system.

    Ever since the discovery of rich copper and gold mineralization under Mount Fubilan in the rugged central highlands of western Papua New Guinea there have been concerns about the potential for environmental impacts on the nearby Fly River and in particular its tributary, the Ok Tedi. These concerns prompted a wide-ranging and comprehensive series of studies of the river system designed, in the first instance, to develop an understanding of the pre-mine environmental setting, and then following the initiation of mining in 1984, the nature and extent of mine-related impacts.

    Through the combined efforts of various universities, government departments, individuals and in particular, the management of Ok Tedi Mining Limited, the owner and operator of the Ok Tedi mine, this on-going attempt to monitor environmental impact has resulted in the assembly of an enormous body of work on this important river system. While much of the data acquired from these studies has been presented in workshops, conferences, and a wide range of specialist publications, there have been few attempts at bringing together the results of this work into a single volume. The 17 chapters comprising this volume are intended to summarize aspects of the natural history of the system as well as provide a multidisciplinary appraisal of the nature and extent of mine-related impacts. They provide in most instances both a synthesis of existing knowledge and a review of the status of current research and development in the respective fields addressed.

    The chapters have been grouped under five main themes. The first group includes six chapters that deal with aspects of the physical environment. The first chapter describes the geomorphology, hydrology, and climate of the river system. It also provides an account of the changes in the geomorphology that have occurred as a result of the release of overburden and tailings from the Ok Tedi mine. This is followed by a chapter that examines the changes that have occurred to the texture, geochemistry, and mineralogy of the sediments carried down the river system before and after the start of mining at Mount Fubilan. Another explores the impact of mine tailings and waste rock on the river floodplain and particularly the speed of these changes as evidenced by the concentrations of copper in floodplain sediments. The remaining chapters in this group examine the processes, sediments, and stratigraphy of the Fly River delta, the fate of mine-derived copper in the sediments of the shallow Gulf of Papua adjacent to the mouth of the Fly River and finally, the variable styles of sedimentation, including material from the Fly River, in the clinoform developing at the mouth of the river.

    The two chapters of the second group deal with predicting the impact of mining on the physical environment. The first presents the results of efforts to model the changes in the rate of river bed aggradation brought about mainly by the discharge of waste rock and tailings from the Ok Tedi mine, while the second focuses on the resultant increase in the incidence of floodplain inundation and how this is likely to change in the future.

    The two papers of the third group discuss aspects of the mine's influence on the chemistry and biotoxicity of river waters. The first provides a review of the processes controlling the fate of copper, the main metal contaminate originating from the mine, in the river system while the second discusses the considerable research into the speciation, bioavailability, and toxicity of copper.

    The fourth group of three chapters deals with the fish population of the river and how it has been affected by mining. The first of these chapters’ focuses on the biology of barramundi in the river, one of the main food sources for the communities living along the river's edge. A second chapter examines the effects of mine-derived waste materials on fish habitat while the third and last explores the spatial and temporal changes in fish assemblage to assess the impact of the mine.

    The last group of four chapters deals with the fauna of the region, the riparian vegetation, and the development of a food web for the system.

    It is hoped that publication of these contributions will stimulate further discussion and research on the impact of mining on riverine and their adjacent near-coastal environments.

    Finally, the job of an editor of a volume such as this would be impossible without the help and support of a lot of people. First I wish to thank Keith Faulkner, a former Managing Director of Ok Tedi Mining Limited (OTML) and Jim Veness, current Manager of the OTML Environment Department, for their unfailing support and encouragement throughout this project. OTML Board members, Allan Roberts, Ross Garnaut, Jochen Tilk, Dick Zandee, and Simon Tosali are also thanked for their support and encouragement in the production of this volume. The financial support provided by Ok Tedi Mining Limited for the publication of this book is also gratefully acknowledged.

    I would also like to acknowledge the efforts of the following people who provided critical reviews and many editorial improvements to the contributions of this volume: Geoff Pickup, Len Murray, Barry Hart, Mead Allison, Simon Apte, Ted Edwards, David Heggie, Kent Hottle, Paul Humphries, Denis Mackey, Alan Orpin, Max Moulds, John Chapman, Wade Hadwen, Robert Hughes, Peter Teasdale, François Edwards, Scott Markich, Susan Adams, John Rinne, Katie Farnsworth, Robert Dalrymple, Scott Miller, Jim Veness, Jacky Croke, Ian Cowx, Alan Whitfield, and H. Gill.

    And finally, I wish to thank the support of the very patient staff at Elsevier for this project and specifically, Linda Versteeg-Buschman, Suja Narayana and Femke Wallien.

    Part 1. The Physical Environment

    Chapter 1 Geomorphology, Hydrology, and Climate of the Fly River System

    Geoff Pickup¹,*, Andrew R. Marshall²


    ¹ Consulting Geomorphologist, 1538 Sutton Road, Sutton, New South Wales 2620, Australia

    ² Andrew Marshall & Associates Pty Ltd., 43 Warrangarree Drive, Woronora Heights, New South Wales 2233, Australia

    * Corresponding author. Tel.: +61 26238 3427;

    E-mail address: pickup01@westnet.com.au

    Abstract

    The Fly River occupies a large humid tropical drainage basin in western Papua New Guinea. In its upper reaches, the river system occupies steep mountain country but further downstream, there is a 500-km floodplain reach that is backed up by the Strickland. This chapter examines the climate, hydrology, and geomorphology of the system. It also describes the changes that have occurred in response to disposal of tailings and waste rock from the Ok Tedi mine. These include extensive deposition, especially in the upper part of the system, and rising water levels that have had a major effect on the floodplain.

    1.1 Introduction

    The Fly River system drains an area of about 75,000 km² in western Papua New Guinea (Fig. 1.1). There are three major rivers in the system: the Ok Tedi, which drains the Hindenburg Ranges; the Upper Fly, which drains the southern part of the Victor Emanuel Range; and the Strickland, which drains the Victor Emanuel and Central Ranges. The Upper Fly and the Ok Tedi meet at D’Albertis Junction to form the Middle Fly, which meanders down a 400-km-long floodplain with extensive scroll bar complexes, cutoffs, and blocked valley lakes. The Middle Fly and the Strickland meet at Everill Junction before entering the Fly Delta. The delta covers about 10,000 km² and extends downstream for another 400 km before entering the ocean in the Gulf of Papua. The Ok Tedi receives waste rock and tailings from Ok Tedi while the Strickland receives sediment load from mining operations at Porgera.

    Figure 1.1 Location map. Mining operations at Tabubil and Porgera are shown by arrows.

    This chapter covers the climate, geomorphology, and hydrology of the Fly. It examines rainfall distribution and basin hydrology. It describes the river system from the mountains to the lower end of the floodplain, including linkages between the river channel and the many off-river water bodies (ORWBs) in tributary valleys and on the floodplain. It concludes with a summary of mine-related impacts on the river system.

    1.2 Climate

    The Fly basin has a humid tropical climate. Average rainfall varies with elevation (Moi et al., 2001). Falls in excess of 10,000 mm/year occur at the Ok Tedi mine site (1,500–2,000 m), declining to about 8,000 mm/year along the upper and middle Ok Tedi. Further decreases occur downstream with values of 5,250 mm/year being recorded at Kuambit just downstream of D’Albertis Junction, 3,869 mm/year at Manda about two-thirds of the way down the Middle Fly floodplain, and 1,847 mm/year at Obo near Everill Junction. Heavy rainfall occurs throughout the year in the mountains, but there is a slightly less wet period from September to November. Seasonal variations are also stronger over the southern part of the floodplain, where rainfalls are lower than in the mountains.

    Based on other stations in lowland Papua (McAlpine et al., 1975), annual evaporation is probably within the range 1,500–2,000 mm, which is less than annual rainfall in many areas.

    Rainfall not only varies between seasons. The region experiences both El Niño and La Niña events, including severe drought episodes in 1972, 1983, and 1997. There may also be longer-term variations in rainfall. For example, Moi et al. (2001) report that rainfalls for 1999–2000 exceeded long-term averages across the region, with the largest percentage increases occurring in lowland areas.

    The rainfall record for Tabubil, just below the mine site, shows some connection with the Southern Oscillation Index (SOI) (Fig. 1.2) but mainly during severe droughts. There is no obvious relationship at other times, making it difficult to use the SOI as a predictive tool.

    Figure 1.2 Monthly rainfalls for Tabubil compared with the Southern Oscillation Index. Both datasets have been smoothed using a five-point moving average.

    1.3 River System Geomorphology

    In its upper reaches, the Ok Tedi and its tributaries drain a heavily dissected ridge and ravine landscape. The ridges rise to over 2,000 m in the north, but most of the basin lies between 200 and 800 m. The eastern part of the basin is karst country, including the massive Hindenburg Wall escarpment, and contains large areas of landslide debris and old debris flow deposits. In the west, igneous rocks are exposed on high mountains, but much of the area consists of shales, limestones, and sandstones. Slopes are unstable in spite of a dense rain forest cover, and landslides and debris flows are common.

    Mine waste is disposed of into two tributaries of the Ok Tedi: Sulphide Creek (which joins the Ok Gilor, and then the Ok Mabiong in its lower section) and the Ok Mani (Fig. 1.3). These tributaries are referred as the Mine Area Creeks. These systems occupy steep, narrow valleys and, prior to mining, were cut down to bedrock for most of their length. Sediment deposits were restricted to local boulder chokes and short reaches with thin layers of armored cobbles. Sulphide Creek contains a waterfall, as does Harvey Creek, a small tributary of the Ok Mani that now receives waste rock from dumps on the southern side of the mine. These creeks are typical of the supply zone (Table 1.1) and were transporting virtually all delivered load before mining operations started.

    Figure 1.3 A SPOT image of the Ok Tedi mine and the Mine Area Creeks.

    Table 1.1 Classification scheme for river reaches on the Fly River System prior to mining

    Source: Adapted from Pickup (1984) and Higgins et al. (1987).

    The Upper Ok Tedi is a fast-flowing river capable of transporting large quantities of coarse sediment, including boulders. For much of its length, it occupies a bedrock gorge less than 200 m wide at its base, although there are wider sections with a few islands developing in the lower sections. Debris flow deposits flank the upper reaches of the gorge, and bed levels have fluctuated over time in response to massive landslides from the Hindenburg Wall and other areas, including one of 7 km³ about 8,800 BP (e.g., Pickup et al., 1979; Blong, 1991). Prior to mining, this reach of the river was classified as a supply zone (Table 1.1). However, unlike the Mine Area Creeks, the bed had a veneer of coarse sediment in transit, much of which probably came from the 1977 Hindenburg Wall landslide.

    Just upstream from Ningerum, the Lower Ok Tedi emerges from its gorge into a gradually widening valley filled with what are probably Pleistocene alluvial deposits. These include weathered and indurated sand and gravel deposits topped by distinctive red soils. Initially, these red beds form valley edges and confine the channel but, as the valley widens, form terraces or terrace remnants, and the Ok Tedi develops a currently active floodplain within them. A braided channel develops, and forested islands have formed on some of the more stable braid bars. Before mining increased sediment input to the river, this reach was classified as an armored zone merging into a gravel–sand transition zone (Table 1.1). Finer material, including a large volume of sand, is now being deposited in this area.

    Prior to mining, the armored zone ended upstream of the junction with the Ok Mart where braiding ceases, and there was a distinctive gravel front. Beyond this point, there was a gravel–sand transition zone (Table 1.1). Flow in the gravel–sand transition zone was (and still largely is) concentrated in a single channel, and meandering begins to develop. However, although there are several meander cutoffs and some channel traces, the floodplain is confined to a narrow belt, 1–4 km wide within the alluvial piedmont, and shows some traces of minor incision. The outer sections of some meanders also cut into Pleistocene terrace remnants, and lateral movement is restricted by indurated and highly weathered soils developed on former fluvial sediments.

    The Middle Fly extends for about 420 km from D’Albertis Junction to Everill Junction, where it meets the Strickland. There are two distinct reaches: a sand zone and a backwater zone (Table 1.1). Prior to mining, the sand zone had a riverbed consisting of well-sorted fine-to-medium sand and extended as far downstream as Manda. This reach has a steeper water surface slope than the backwater zone that extends from Manda downstream to Everill Junction. The pre-mining backwater zone had a bed composed of mainly fine sand but with an increasing proportion of silt clay particles further downstream until less than 20% sand was present. A localized body of medium sand occurred downstream from Bosset and is probably associated with the slightly steeper river long profile there. It may have originated from the catchment of the blocked valley lake of Bosset Lagoon.

    The evolution of the Middle Fly may be as follows. During the Pleistocene, the river repeatedly entrenched and dissected surrounding hills during periods of low sea level (Loffler, 1977). As sea level rose, reaching its current height around 5,000 BP, the river aggraded at about 1 mm/year (Pickup et al., 1979), declining to about 0.1 mm/year (Dietrich et al., 1999).

    The aggradation process created a number of distinctive landforms (Paijmans et al., 1971) (Figs. 1.4 and 1.5). First, there is a central meander belt made up of the river channel, a wide and sometimes discontinuous levee, scroll bar complexes, Pleistocene terrace remnants, and partly filled meander cutoffs, often connected to the main river by tie channels. This system forms an inner floodplain, accumulates most of the sediment from upstream that is deposited, and is perched above the surrounding outer floodplain. Second, the aggradation and levee buildup of the inner floodplain was sufficient to block off or back up most of the lowland tributary systems. This created a system of blocked valley lakes, backswamps, flood basins, and drowned tributary channel systems on each side of the central meander belt. These systems may be up to 7 m lower than the levees that border the main river channel. The elevation difference decreases downstream and is smallest in the backwater zone upstream of Everill Junction (Dietrich, 2000). However, it remains a significant feature in determining both floodplain behavior and landforms down as far as the Fly Delta.

    Figure 1.4 A LANDSAT TM image showing the floodplain systems of the upper Middle Fly during low-flow conditions. Flooded off-river water bodies are shown in black or blue. Dried-out areas that are normally flooded are pink or purple. Green areas are mainly rain forest. Yellow squares are gauging stations or sampling sites.

    Figure 1.5 A LANDSAT TM image showing the floodplain systems of the lower Middle Fly and lower Strickland during low flow conditions. Flooded off-river water bodies are shown in black or blue. Dried-out areas that are normally flooded are pink or purple. Green areas are mainly rain forest. Meander cutoffs, blocked valley lakes, and swamps behind the main channel levees are all clearly visible. Yellow squares are gauging stations or sampling sites.

    While the Middle Fly contains spectacular evidence of meandering in the form of cutoffs and scroll complexes, it has not experienced rapid lateral migration. Pickup et al. (1979) found that only three loops were cut off between 1899 and 1979. More extensive analyses by Barr Engineering (1995) and Dietrich et al. (1999) found that bank erosion and meander migration rates were highest close to D’Albertis Junction and decreased by a factor of three, downstream to the backwater-affected reach.

    The vegetation of the floodplain corresponds fairly closely with the steepness of the long profile. Rain forest extends from D’Albertis Junction to a point beyond Mabaduam, entering a transitional reach close to the junction with the Agu. The lower, backwater-affected reach, beyond Manda, largely consists swamp grass or riparian vegetation (Day et al., 1993). The low gradients down the Middle Fly and the small transverse differences in floodplain elevation make floodplain vegetation systems sensitive to changes in water level on the Fly if they are of sufficient duration.

    It is likely that the Fly has evolved only slowly since the sea level reached its present level. Before sediment load increased due to mining, the levee system was probably growing upward and extending downstream through deposition of sediment. However, given the limited sediment load reaching the Middle Fly, the rate of progress was very slow. For example, Pickup (1984) estimates that it would take somewhere between 1,400 and 27,000 years for the sand zone in the Middle Fly to extend as far as Everill Junction, provided that the bed level of the Strickland remained constant.

    The backswamps receive limited amounts of material from the main channel, apart from close in to the levees, and may grow upward more by accumulation of organic material than by deposition of sediment. Studies using copper as a marker show that most of the sediment leaving the main river channel is deposited on the levees, in off-channel water bodies with tie channels to the main river, and by backflow up tributaries (Day et al., 1993; Dietrich, 2000).

    The Strickland has a catchment area of 36,740 km² at Everill Junction and is a much larger river than the Fly (catchment area of 18,400 km²). It is also closer to the main sources of natural sediment load in the mountains. Like the pre-mining Ok Tedi and the Middle Fly, the Strickland has an armored zone, a gravel–sand transition zone, and a sand zone, but these extend much further down the system. The Strickland has a steeper gradient than the Fly, and its floodplain probably had more active channel shifting. The Strickland backs up some very large off-channel water bodies, including Lake Murray. It is separated from the Middle Fly floodplain by a 10- to 30-km-wide pre-Holocene former floodplain surface that forms an extensive set of terraces.

    Once the Strickland and the Middle Fly join, river characteristics change. Prior to mining, the Strickland transported perhaps 8–10 times the natural load of the Fly. Even now, it probably contributes more than 60% of the total load (Moi et al., 2001). Also, the bed material of the Strickland and the Lower Fly contain much more sand than the Middle Fly.

    The Fly Delta is a funnel-shaped system dominated by tidal action. The tidal range is about 3.5 m at the mouth and 5 m at the apex (Wolanski and Eagle, 1991). Channel bed material is mainly sand, and the majority of fine material is advected offshore (Harris et al., 1993).

    Terrace remnants in the delta are substantially higher than those surviving in the Middle Fly. The channel also seems to become more confined downstream and does not have the extensive floodplains and backswamps of the Middle Fly. This may suggest minor uplift due to tectonic activity or isostatic adjustment. The confined channel, increased river discharges, and tidal flushing all limit the deposition of mine-related sediment reaching the delta.

    1.4 Hydrology

    1.4.1 Runoff and Flooding

    Most of the runoff in the Fly River originates in the mountains. Runoff at Kuambit just below the junction of the Ok Tedi and the Upper Fly is about 6,600 mm/year or about 70–80% of the total rainfall. However, once runoff enters the floodplain reach of the Middle Fly, flow patterns change (Fig. 1.6).

    Figure 1.6 Water levels on the Middle Fly 1995–2000.

    At the upstream end of the floodplain, water levels vary rapidly as floods from the Ok Tedi and the Upper Fly enter the system. By the time flow reaches Manda, about 230 km downstream, much of the short-term variability has disappeared, and most of the changes in water level occur at a scale of weeks to months. The main cause of this flood wave attenuation is exchange of water between the main river and many off-river floodplain storages through tie channels, levee breaches, and, if water levels are high enough, directly across the floodplain.

    Dietrich (2000) describes the floodplain storages as follows:

    There are 38 oxbow lakes within the scroll complex of the study area, 25 of which have well maintained tie channels which transport water and sediment over a large range of main channel stage. A further 8 of the oxbows have smaller tie channels which transport water and sediment only over the upper ranges of main channel flow with the remainder (5 oxbows) having no apparent tie channel. There are at least 35 blocked valley lake systems within the backswamp areas of the floodplain ranging in size from <0.1 km² to > 50 km². The off-river water bodies (ORWB) (oxbow and blocked valley lakes) comprise 10% of the total area of the floodplain or approximately 350 km².

    The ORWBs can hold 2–3% of the annual flow of the Fly before general flooding occupies most of the floodplain. However, because there is continuous inflow and outflow between the floodplain storages and the main channel during floods, a much larger proportion of flow is cycled through them. Dietrich and Day (2004) has suggested that at least 20%, and possibly up to 40%, of the Fly River water passes through the floodplain. Model calculations (see Pickup, 2009) suggest that the value is closer to 10% but could increase to between 14 and 16% by 2050, depending on the level of bedload extraction by dredging from the Lower Ok Tedi.

    Most of the transfer between the main river channel and the floodplain is inflow and outflow to and from storages rather than flow down backplains between levee breaches. However, there is evidence on satellite imagery of backplain flow on the west side of the Fly near Bosset.

    Dietrich and Day (2004) have suggested that water may leave the middle reaches of the Middle Fly via tie channels and flow into the Agu River. This water then returns to the lower Middle Fly via the Agu junction below Manda. It is not clear how important this effect is. Three small tie channels connect ORWBs to both the Fly and the Agu, and it is possible for water to flow into the Agu across the floodplain at high water levels. At the same time, flood levels on the Fly propagate up the lower Agu very quickly, so the flow path from the Fly via the upper Agu would soon be restricted by rising water levels.

    1.4.2 Links between the River and its Floodplain – The Fly River Tie Channels

    The term tie channel was coined by Blake and Ollier (1971) to describe the narrow channels that link the main river channel and the oxbow lakes that occupy meander cutoffs on the Fly floodplain. These channels experience bidirectional flow, depending on differences in water level between the oxbow lake and the main river. Typically, water flows into the oxbow on the rising stage of floods in the main river and out of it as floodwater levels recede (Fig. 1.7). Hydraulic gradients therefore vary with time but may be steep, and flow velocities of 1–2 m/sec are quite common at times of rapidly changing water levels. This is enough to keep many tie channels clear of sediment and some have functioned for hundreds and perhaps thousands of years (Rowland et al., 2005).

    Figure 1.7 Some features of tie channels. Upper left shows rapid and turbulent flow through a narrow vegetation-lined tie channel. This is sufficient to move sandy bed material from the Middle Fly into the receiving water body. Upper right shows a tie channel when water levels on the main river and the connected water body are similar. Virtually no flow is occurring. Lower left shows a tie channel draining clean water from the ORWB into the Fly. Lower right shows what occurs when water levels in the river and the ORWB have equalized during a large flood. The levee breach tie channel is completely submerged and contains still, clear water.

    It is likely that tie channels play a crucial role in the ecology of ORWBs. They provide a route for the transfer of water and sediment, transport contaminants, and allow the exchange of carbon. They may also serve as refugia and breeding sites for fish species, especially during drought.

    Water circulation through tie channels may have a role in maintaining water quality in many ORWBs, especially if they are small. Most ORWBs will be replenished by rainfall, and virtually all become connected with the main river during major floods. However, during periods of low-medium flow on the main river, flow through tie channels will be the main process of water circulation. This may be a factor limiting eutrophication in the smaller ORWBs, where tie channel flow could replace 5–10% of the total water volume in a day. In larger water bodies, input from rainfall, inflows from the surrounding catchment, and outflow to the main river would play a greater role in water circulation. However, inflow and outflow of main channel water is still likely to be the dominant process.

    If we define a tie channel as one that links the main river channel and an ORWB and experiences bidirectional flow, then there are at least seven types on the Lower Ok Tedi and the Middle Fly (see Fig. 1.8 for examples). Each type may have evolved in a different way, and there may be variations in the potential for silting up. The main types are:

    • Cutoff tie channels that connect oxbow lakes of meander cutoffs to the main river. The oxbows are the deepest ORWBs on the floodplain and usually hold water during drought, even if water levels fall below the elevation of the tie channel bed. They are therefore likely to be important refugia for a whole range of aquatic species. Several subtypes occur. The most common form drains through a sediment plug at one or both ends of the former meander channel. Several others exit closer to the middle of the cutoff through the swales between adjacent former point bars. A third variation occurs where the tie channel links with a tributary, and there are even cases where the tributary crosses the cutoff and functions as the cutoff tie channel as well as draining its own catchment.

    • Levee breach channels. These may vary in size and usually connect the river to areas of standing water in the backswamps behind the levee. These backswamp lakes vary in depth and area over time and may dry out completely when river water levels are low during drought. Levee breach channels are usually short and may develop when localized levee collapses occur and open a path for water to flow. There are shallower and probably more short lived than the cutoff tie channels. Many levee breach channels may have silted up, as levee deposition has accelerated in response to deposition of mine waste.

    • Tributary tie channels. Over the last 5,000–7,000 years, the central meander belt of the Fly has grown in height faster than the outer edges of the floodplain and the floors of tributary valleys. This has created a set of blocked valleys often containing permanent or ephemeral lakes that provide the largest areas of standing water on the floodplain. The blocked valley lakes cover a large area, and some of the larger ones may have water depths similar to the meander cutoff oxbows close to the main river. The larger lakes hold water during drought, and many are distant from the main river channel. The more distant lakes are largely unaffected by deposition of sediment from the main river. Tributary tie channels can be highly sinuous and often have well-developed levees. While no dating has been done, it is likely that most of these channels are hundreds to thousands of years old, especially in areas distant from the main Fly River channel.

    • Drowned tributary tie channels. The backwater zone of the Middle Fly extends from Manda to Everill Junction. Backwater from the Strickland keeps large areas of the floodplain inundated for long periods. The tie channels that originally linked ORWBs to the river are now flooded for most of the time, and only their levees may be visible. These channels may offer a faster path for water flowing from the river to the inundated area when water levels rise after a drought, but for most of the time, they behave as part of the lake. Again, they are likely to be hundreds to thousands of years old.

    • Meander neck channels. A few meanders on the Fly and the Lower Ok Tedi have narrow channels that cut across the meander neck. Typically, they occupy a swale between adjacent scroll bars. Some provide a shortcut for canoe and small-boat traffic. They may experience bidirectional flow if they provide access to large areas of intermittently flooded backswamps behind levees.

    • Fossil tie channels. The floodplain contains traces of tie channels of all types that are no longer functional or only operate intermittently. Many of the blocked valley lakes show sinuous channel traces that are only visible at low water. Many disappear well before they reach the main river. Others appear to have been blocked by long-term levee development on the main channel. There are also some abandoned tie channels connecting the main river with meander cutoff oxbows. In at least one case, the cutoff tie channel has been abandoned when meander migration on the main river has intersected the oxbow and opened up a new path for water flow (Rowland et al., 2005).

    • Composite forms consisting of more than one tie channel type are fairly common. Both cutoff and tributary tie channels often have levee breaches that drain adjacent backswamps. Cutoff and tributary channels also sometimes capture each other and intersect.

    Figure 1.8 Examples of tie channel types. (A) A blocked valley tie channel with typically sinuous form entering the Lower Ok Tedi through a levee breach. (B) A cutoff tie channel. (C) A levee breach into a flooded backswamp. (D) A levee breech into a flooded area in an old scroll bar/swale complex. (E) Submerged former channel levees in a blocked valley lake. (F) A complex tie channel system at D’Albertis Junction. The large tie channel in the foreground connects the upper Fly and the Ok Tedi along a meander trace. The smaller tie channel above it connects a meander cutoff.

    Very little work has been done on tie channel behavior. The most detailed information in the literature comes from Rowland et al. (2005) and Rowland and Dietrich (2006). These papers describe the forms, sedimentology, and evolution of meander cutoff tie channels, together with deposition rates in associated oxbow lakes.

    Rowland describes tie channel evolution as follows. Sediment-laden water from the main river enters the cutoff as a jet that extends a small delta outward into the cutoff water body and advects sediment laterally to its margins. This allows a set of submerged levees to build up. Over time, the levees extend upward and outward, gradually filling up the river end of the cutoff with sediment. Flow is usually sufficient to keep the bed of the tie channels scoured, and finer sediment deposited on the banks when flow velocities are low tends to collapse into the channel during more active periods. Flow from the cutoff water body back to the river, as water levels recede after a flood, also plays a role in keeping the tie channel open since it scours sediment that may be deposited when water levels are similar in both the main river and the cutoff, and tie channel flow velocities are low. An example of tie channel evolution from the Fly floodplain is shown in a series of satellite images in Fig. 1.9.

    Figure 1.9 Evolution of a cutoff on the Middle Fly. Infilling and tie channel development begins very quickly once the cutoff has occurred. Initially, both ends of the cutoff remain open to water from the river, but one arm of the former meander develops a tie channel. Eventually, a tie channel begins to develop in the other meander arm. Imagery includes LANDSAT TM, PACRIM 2000 SAR, and ENVISAT SAR.

    Several factors govern the rate of tie channel extension into the cutoff oxbow lake and whether a tie channel is likely to be maintained. First, there is the amount of water flowing into and out of the oxbow. This is a function of lake size and water level variations on the main river channel. More water will flow into larger oxbows. Also, inflow and outflow will occur more frequently when river levels are more variable. Second, the sediment concentration in river water will affect delta growth rates since most of it will settle out as the jet enters the still water of the lake. Sediment concentration will also influence tie channel levee growth and backfilling of areas behind these levees.

    Tie channels not only grow outward into meander cutoffs but may also extend toward the river if meander migration moves the channel away from the original tie channel mouth (Rowland et al., 2005). Many of the long, straight tie channel sections on the Middle Fly floodplain have developed in this way.

    Rowland et al. (2005) have dated two active tie channels connecting the Fly with meander cutoffs. One was 347±44 years old while the other proved to be at least 928±128 years old. This indicates how persistent tie channel features can be. It also suggests caution in interpreting blocked tie channels. They may, in fact, have been blocked for hundreds of years and long before mine sediments were added to the river.

    Rowland's model can be applied to levee breach tie channels as well as those connecting meander cutoffs to the river. However, the tie channel levee complex and the delta will not develop to any great extent. This occurs because the floor of the levee breach remains close to the elevation of the backswamp behind the main channel levee. This limits the water surface slope that can develop between the ORWB and the river and keeps flow velocities lower than in cutoff tie channels. It also prevents much of the sandy main riverbed material from getting into tie channel flow. Most of the sediment reaching the ORWB will therefore be fine main river wash load that settles out slowly and disperses more widely in the ORWB.

    Blocked valley tributary tie channels may also evolve differently from cutoff tie channels. Many of these systems have long reaches with well-developed levees that have built up over thousands of years in response to bidirectional flow. However, because these channels are much longer than the cutoff tie channels, they do not tend to develop the steep water surface slopes that occur in cutoff tie channels during rising and falling flood stages. Deposition therefore tends to be dispersed over a longer distance, and the jet-deposited fans of cutoff oxbow lakes are less likely to occur. Many of the blocked valley tributaries also drain and fill very large ORWBs and have sufficient flow to remain open.

    1.4.3 Backwater Effects from the Strickland

    Backwater effects from the Strickland partly determine water levels in the lower reaches of the Middle Fly. The backwater effect extends upstream for about 200 km, and the gauging record at Manda (FLY16) closely follows that of Obo (FLY15) (Fig. 1.6).

    Lack of data makes it difficult to determine the precise influence of the Strickland. Gauging stations on the Lower Fly at Ogwa (FLY22) and the lower Strickland (STR01 and more recently SG4) have only operated for short periods, and there are many missing records. Obo (FLY15) has a longer record, but gaps in the data frequently coincide with periods when the other gauges were operating.

    There is a common dataset for Obo (FLY15) and the lower Strickland (STR01) with few missing records for 1989–1992 (Fig. 1.10). STR01 is 73 km upstream of Everill Junction and below Massey Baker Junction, where the flows from Lake Murray join the Strickland. There are no major inflow points between Station STR01 and Everill Junction, so the STR01 record should give a reasonable indication of flows in the Strickland at Everill Junction.

    Figure 1.10 Water levels on the lower Middle Fly and the lower Strickland. Strickland height datum is arbitrary.

    The datasets show both differences and similarities among stations. Some of the low water levels at Obo (FLY15) and on the lower Strickland (STR01) coincide, but this is not always the case. There is also some matching of flood peaks but, again, not in every case. A very crude multiple-regression analysis for a short period shows that, after adjusting for lag, Stations FLY10 (Kuambit) and STR01 have a similar degree of influence on water levels at Obo (FLY15). However, these results do not cover a wide enough range of water levels to be definitive. They show that water surface elevations on the lower Middle Fly are a product of flows coming down both the Fly and the Strickland.

    1.4.4 Tidal Effects

    Very few data are available to show the influence of tidal cycles on Middle Fly water levels. A tide gauge operated at Daru, southwest of the Fly Delta for short periods in mid-1983, but no water level data are available in the OTML database for the lower Middle Fly. The JTides model (http://www.arachnoid.com/JTides/) gives a good reproduction of the limited Daru Roads dataset. This model may be used to calculate tidal variation for other periods, including those when lower river system gauges were operating.

    The longest gauging record for the lower river system is for Station FLY15 at Obo. However, this station is affected by backwater from the Strickland as well as weak tidal effects. A Lower Fly station (FLY22) operated at Ogwa, just downstream of Everill Junction (Fig. 1.1) in late 1996 and early 1997. This station measures water levels for the combined flows from the Middle Fly and the Strickland and gives a clearer picture of tidal influences.

    Ogwa (FLY22) water levels are compared with the modeled tidal variation in Fig. 1.11. Daily tidal variation can be seen on the lower graph. By the time tidal influence reaches Ogwa, there is only about 0.1-m variation at low water and less during medium flows. At high flow, daily tidal variation is almost indistinguishable. Longer-term tidal variation (upper graph) is also difficult to identify, but during periods of steady flow, it may possibly reach 0.3–0.5 m at Ogwa. However, some of this variation could be due to small floods coming downriver rather than tidal effects.

    Figure 1.11 Modeled tidal variation at Daru compared with water levels at Station FLY22 (Ogwa) on the Lower Fly just downstream of Everill Junction. The upper graph shows medium-term tidal variation. The lower graph compares FLY22 water levels with individual tidal cycles over a period of low-medium river flow. The tide gauge datum is arbitrary.

    1.5 Geomorphic Changes Associated with Mining

    The Ok Tedi mine is located close to the headwaters of the Ok Tedi on what was formerly Mount Fubilan, although much of the mountain has now been removed (Fig. 1.12). The operation is open cut, producing an average of 83,000 tonnes/day of ore and a peak of 152,000 tonnes/day of overburden. About 80 million tonnes of waste rock and tailings are discharged into the Fly River system each year. The original mine design incorporated a tailings dam, and construction of a tailings dam began. However, the foundations collapsed in a landslide in 1984. An Interim Tailings Scheme contained tailings at start-up. Since this storage area is filled, tailings have been discharged directly to the river system (Ok Tedi Mining Ltd., 1999).

    Figure 1.12 The Ok Tedi mine and the Ok Mani in early 2004 (photo: Andrew Marshall).

    The discharge of waste rock and tailings has greatly increased the sediment load of the river. While much of the waste rock consists of coarse material, about 60% of it breaks down during transport into finer material in the sand–silt size range. While the gravels are not transported beyond the middle reaches of the Ok Tedi, the finer material can pass into the Fly. Prior to mining, the natural load of the Ok Tedi at D’Albertis Junction was 3–5 Mt/year with a similar amount coming from the Upper Fly (Higgins et al., 1987). Over the period 1985–2000, the load of the Ok Tedi has increased to an average about 45 Mt/year. Some reaches of the river have experienced extensive deposition, raising the riverbed, adding material to levees and floodplains, and increasing water levels.

    1.5.1 Mine Area Creeks

    Waste rock is disposed of in failing dumps on the north and south sides of the mine. These dumps feed a range of sediment sizes, from boulders to clay, into Sulphide Creek (which becomes the Ok Gilor, and then the Ok Mabiong in its lower section) and the Ok Mani. Mill tailings with a median size of about 75 μm are discharged directly into the Ok Mani downstream of the dump inputs. A summary of inputs is presented in Table 1.2.

    Table 1.2 Major sediment inputs to mine area creeks during mining operations

    The Sulphide Creek system, on the north side of the mine, receives material from the northern dumps. This material accumulates in a steep fan at the dump toe, filling the whole valley of Sulphide Creek. The creek then enters a steep and narrow slot with waterfalls. All sediment delivered from upstream passes through this reach. Beyond the slot, a second valley-wide fan develops, extending down to the junction with the Ok Tedi. This system experienced a major landslide in 1989, adding about 125 Mt of hillslope material to the system. This is equivalent to about 80% of the waste rock dumped into Sulphide Creek between 1985 and

    Enjoying the preview?
    Page 1 of 1