You are on page 1of 23

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/241592875

Debris Flow and Hyperconcentrated Flood-Flow


Deposits in an Alluvial Fan, Northwestern Part of
the Cretaceous Yongdong Basin, Central Korea
Article in The Journal of Geology January 1999
DOI: 10.1086/314334

CITATIONS

READS

138

1,984

3 authors, including:
Young Kwan Sohn

Chul Woo Rhee

Gyeongsang National University

Chungbuk National University

97 PUBLICATIONS 1,859 CITATIONS

19 PUBLICATIONS 448 CITATIONS

SEE PROFILE

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

AMS(anisotropy of magnetic susceptibility) of pumice in pyroclastic rocks, Baekdusan (Mt.) and


Ulleungdo View project
Is it possible an AMS approach on basaltic tuff?: a case study on the Holocene Songaksan tuff
ring (Phreatomagmatic erution), Jeju View project

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Young Kwan Sohn


Retrieved on: 25 October 2016

Debris Flow and Hyperconcentrated Flood-Flow Deposits in an


Alluvial Fan, Northwestern Part of the Cretaceous Yongdong
Basin, Central Korea
Young Kwan Sohn, Chul Woo Rhee,1 and Bok Chul Kim2
Department of Earth and Environmental Sciences, Research Institute of Natural Sciences,
Gyeongsang National University, Chinju 660-701, Korea
(e-mail: yksohn@nongae.gsnu.ac.kr)

ABSTRACT
The alluvial-fan deposits in the Cretaceous Yongdong Basin, Korea, consist of conglomerates with a muddy or sandy
matrix and sandstones with thick or thin laminations. The conglomerates and sandstones occur commonly in couplets,
constituting apparently single sedimentation units. Facies transitions in the deposits can be summed up in a tripartite
facies sequence: (1) a clast-supported conglomerate with a muddy or sandy matrix and parallel clast alignment (facies
A and B), (2) a matrix-supported coarse-tail normally graded conglomerate with random clast orientation (facies C)
and thickly stratified pebbly sandstone (facies D), and (3) laminated sandstone (facies E). The clast-supported conglomerate is interpreted as deposits of debris flows dominated by frictional grain interactions. Development of pervasive parallel clast alignment, lacking large floating clasts and inverse grading, suggests deposition via incremental
aggradation rather than en masse freezing. The matrix-supported conglomerate and thickly stratified sandstone are
interpreted as deposits of dense inertia layers or traction carpets developed beneath a high-concentration bipartite
flow. The laminated sandstone indicates traction sedimentation associated with dilute flows. The facies sequence
therefore suggests deposition from a composite sediment flow that comprises a preceding debris flow, a trailing watery
flow, and an intermediate flow between. The intermediate flow is regarded as a hyperconcentrated flow on the basis
of its bipartite nature because a hyperconcentrated suspension has a meager yield strength and is prone to be density
stratified. The measured section comprises three depositional sequences, decameters thick and separated by thick
mudstone beds, which could be interpreted in terms of fan evolution (progradation and retreat) under an influence
of changing sediment supply from a drainage basin. Close association of sediment type with constituent facies in the
three sequences suggests that composite sediment-flow deposits are favorably developed by sand-matrix debris flows
drained from large and less rugged catchments.

Introduction
Debris flows are important depositional processes
on alluvial fans, responsible for many coarsegrained deposits (Blackwelder 1928; Bull 1963;
Hooke 1967; Gloppen and Steel 1981; Nemec and
Steel 1984; Whipple and Dunne 1992; Blair and McPherson 1994). They are gravity-driven flows of
highly concentrated mixtures of sediment and water, resulting in massive, poorly sorted, and generally ungraded deposits with lobate fronts and latManuscript received August 12, 1997; accepted July 30, 1998.
1
Department of Earth and Environmental Sciences, College
of Natural Sciences, Chungbuk National University, Cheongju
361-763, Korea; e-mail: gloryees@cbucc.chungbuk.ac.kr.
2
Geological Research Division, Korea Institute of Geology,
Mining and Materials, Taejon 305-350, Korea; e-mail: kbc@
rock25t.kigam.re.kr.

eral levees (Sharp and Nobles 1953; Johnson 1970,


1984; Fisher 1971; Costa 1984; Major 1997; Sohn
et al. 1997). There is a wide spectrum of debris
flows, of which the mud-rich and clast-rich endmembers can be modeled as viscoplastic fluid
(Johnson 1970, 1984) or an inertial grain flow (Takahashi 1978, 1981). Both models assume that the
deposition from a debris flow occurs abruptly or en
masse as the driving shear stress drops below the
plastic yield strength (viscoplastic model) or as
grains lock up because of a decrease in the dispersive pressure (inertial grain-flow model). Most debris flows are, however, neither a single-phase homogeneous plastic substance nor an inertial grain
flow that lacks the effects of interstitial fluid. Instead, they are multiphase materials, in which a

[The Journal of Geology, 1999, volume 107, p. 111132] q 1999 by The University of Chicago. All rights reserved. 0022-1376/99/10701-0002$01.00

111

112

Y. K. SOHN ET AL.

number of momentum transport processes operate


involving both solid and fluid forces (Iverson and
Denlinger 1987; Iverson 1997). It is also recognized
that incremental aggradation by multiple waves or
surges is an important means of debris-flow emplacement (Wan and Wang 1994; Major 1997; Vallance and Scott 1997).
In contrast to debris flows, streamflows are dominated by fluid forces and deposit their sediment
loads on a grain-by-grain basis, resulting in wellsorted and stratified or cross-stratified deposits. Hyperconcentrated flood flows having characteristics
intermediate between debris flows and streamflows
have also been recognized as an important agent of
sedimentation both in alluvial fans and in volcanic
terrains (Bull 1963, 1977; Beverage and Culbertson
1964; Harrison and Fritz 1982; Nemec and Muszynski 1982; Waitt et al. 1983; Ballance 1984; Nemec and Steel 1984; Pierson and Scott 1985; Smith
1986, 1987a; Scott 1988; Todd 1989; Pierson et al.
1990; Smith and Lowe 1991; Best 1992; Wan and
Wang 1994). These flows develop from either debris
flows or torrential streamflows by incorporating
water or sediment along their flow paths. They produce ungraded or coarse-tail normally graded gravel
deposits and crudely or thickly stratified gravelly
sand deposits (Harrison and Fritz 1982; Pierson and
Scott 1985; Smith 1986; Smith and Lowe 1991; Best
1992), similar to those of high-concentration turbidity currents (Lowe 1982; Postma et al. 1988;
Chun and Chough 1992). In contrast to debris
flows, hyperconcentrated flood flows have meager
yield strengths and are prone to be density stratified
into a high-concentration clast-rich lower part and
a dilute finer-grained upper part (Smith 1986; Pierson and Costa 1987; Wan and Wang 1994; Coussot
and Meunier 1996). Recognition of such hybrid processes raises the possibility that couplets of conglomerate and stratified sandstone may represent
single depositional events rather than multiple
events of different flows.
Many sediment flows cannot, however, be categorized into one of these end-member types because individual flows may comprise more than
one flow type at an instant in time and are subject
to a series of flow transformation during transport
(Beaty 1963; Pierson 1980, 1986; Pierson and Scott
1985; Smith 1986, 1987b; Scott 1988; Pierson et al.
1990; Smith and Lowe 1991; Best 1992; Wan and
Wang 1994). Depositional events on alluvial fans
are therefore a complex mixture of different processes, such as debris flows, hyperconcentrated
flood flows, and streamflows. Understanding the
nature of these composite events, such as the evolution of flow properties with time and space and

the interaction between different parts of a flow and


its environment, is crucial for proper interpretation
of alluvial-fan facies and processes.
The alluvial-fan succession on the northwestern
margin of the Cretaceous Yongdong Basin, Korea
(fig. 1), consists of conglomerates with muddy or
sandy matrix and sandstones with thick or thin
laminations that are interpreted to be deposits of
debris flows, highly concentrated bipartite flows,
and dilute streamflows. The conglomerate and
sandstone units commonly occur in couplets, the
former gradationally overlain by the latter, apparently constituting single sedimentation units.
These facies transitions suggest sequential deposition from a composite sediment flow that comprises a preceding debris flow, a trailing dilute flow,
and probably with a hyperconcentrated flow between. This study documents the sedimentary
characteristics of the deposits and infers the characters and depositional processes of composite sediment flows on alluvial-fan surfaces.
Geologic Setting
In the southwestern Korean Peninsula, several Cretaceous basins were formed along a series of NESW trending strike-slip faults (fig. 1). The Yongdong
Basin (ca. 10 # 40 km2 in area) is an elongate halfgraben basin filled with siliciclastic, nonmarine deposits (ca. 68 km thick) (Lee and Paik 1989, 1990;
Chun et al. 1993). To the southeast, the deposits
are in fault contact with Precambrian gneiss and
schist (Shimamura 1924; Yun and Park 1968; Won
and Kim 1969; Kim and Hwang 1986). On the
northwestern margin, they unconformably overlie
phyllites and limestones of unknown age, Permian
metasandstones, and Permian to Jurassic granitic
rocks (Shimamura 1924; Won and Kim 1969; Kim
et al. 1978; Hong et al. 1980; Kim and Hwang 1986;
Kim and Lee 1986; Jin et al. 1993; Shin and Jin
1995). They are intruded by hypabyssal and volcanic rocks, such as quartz porphyry, rhyolite, and
trachytic andesite (fig. 1b). The occurrence of fossil
flora (coniferous gymnosperms and angiosperms)
and fauna (gastropods, estherids, ostracods, and
charophytes) suggests that the deposits were
formed in Neocomian to Aptian-Albian times under temperate climatic conditions (Chun et al.
1993; Choi et al. 1995).
In the southern part of the basin, the basin-fill
successions can be divided into two allostratigraphic units (alloformations 1 and 2) by a welldeveloped paleosol horizon, which is tens of meters
thick and extensively exposed across the basin (fig.
1b; Kim 1996). Each alloformation comprises var-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

113

Figure 1. a, Distribution of Cretaceous nonmarine basins and fault pattern in the Korean Peninsula (modified after
Korea Institute of Geology, Mining, and Materials 1995; Choi 1996). Numbers indicate Cretaceous basins. 1, Pungam;
2, Eumsung; 3, Kongju; 4, Puyeo; 5, Yongdong; 6, Jinan; 7, Neungju; 8, Kyongsang. b, Distribution of dominant
lithologic units in the southern part of the Yongdong Basin. I, Conglomerate; IIai, conglomerate/gravelly sandstone;
IIIa, b, gravelly sandstone; IVad, gravelly sandstone/reddish mudstone; V, sandstone; and VI, mudstone. Paleoflow
patterns measured from clast imbrications and attitude of channel walls or trough axes are indicated by arrows. The
measured section (marked by asterisk) is present in the unit IIg adjacent to the northern basin boundary.

114

Y. K. SOHN ET AL.

ious lithologic units of conglomerate, (gravelly)


sandstone, and mudstone, distinguished from one
another on the basis of dominant sedimentary facies and/or clast composition. Apparent paleoflow
directions (fig. 1b) and the overall distributional
pattern of lithologic units in each alloformation
(fig. 1b) indicate that discrete coarse-grained deposits (conglomerate and gravelly sandstone) along
the basin margin grade laterally into fine-grained
deposits (mudstone) in the basin center (Kim 1996).
Detailed facies analysis in the southern part of the
Yongdong Basin reveals that each alloformation
represents a depositional system comprising alluvial fans/fan deltas along the basin margin, downstream braided rivers, and lacustrine depositional
environments in the basin center (Kim 1996).
In this article, we describe in detail a road-cut
section in the lithologic unit IIg (fig. 1b), which
consists of conglomerates, gravelly sandstones, and
subordinate amounts of sandstones. The unit is interpreted to represent an alluvial-fan depositional
environment on the basis of its occurrence adjacent
to the Jurassic basement of granodiorite to the west,
the presence of debris-flow deposits, and the overall
sediment dispersal pattern in the study area (fig.
1b). Although the geometry and dimension of the
conglomerate body are unclear because of postdepositional deformation, the conglomerate body fits
into the widely used definition of an alluvial fan
as a body of detrital sediments built up at a mountain base by a mountain stream (Blissenbach 1954;
Nilsen 1982).

Facies Description and Interpretation


Facies A: Clast-Supported Conglomerate with Muddy
Matrix. The conglomerates of this facies are me-

dium to thick bedded (0.31 m thick) and are generally massive and ungraded (fig. 2). Unit boundaries are mostly parallel sided, but some units have
slightly scoured bases and undulatory upper contacts with positive primary relief. Rarely, units of
this facies pinch out abruptly (fig. 3; arrows). Clasts
are mostly pebble to cobble sized; boulder-size
clasts are less abundant; large mudstone chips are
abundant in some units; large protruding clasts are
not observed. Clasts are subangular to subrounded
and generally elongate. They are mostly in tight
contact and are aligned parallel to bedding planes
(fig. 4). This facies occurs mainly in the middle
sequence as thick amalgamated packets (fig. 2).
Thin discontinuous units of sandy deposits (facies
D or E) commonly overlie this facies. Sparingly, the
matrix of facies A conglomerates becomes sandier

upward, gradually passing into facies B


conglomerates.
General characteristics of this facies, such as the
massive and ungraded nature, poor sorting, muddy
matrix, and rare positive primary relief and tapered
margins, are suggestive of emplacement by debris
flows (Johnson 1970, 1984). The tight clast-supported fabric suggests that the debris flows had high
mechanical strengths imparted from grain friction.
Clast collision is interpreted to have been hampered by highly viscous muddy matrix, resulting in
the lack of inverse grading and clast imbrication.
The muddy matrix probably enhanced the mobility
of debris flows by lubricating frictional clast interactions and maintaining high pore fluid pressures,
rather than increasing the mechanical strengths of
debris. The pervasive parallel-to-bedding clast
alignment throughout the conglomerate units,
lacking large floating or protruding clasts and subvertical clast alignment, suggests that the debris
flows experienced full laminar shear before deposition (Fisher 1971; Enos 1977), without developing
a nondeforming plug at the upper part.
Facies B: Clast-Supported Conglomerate with Sandy
Matrix. Facies B conglomerates share many char-

acteristics in common with facies A conglomerates, having similar clast sizes and shape, fabric,
bed thickness, and bed geometry, except that the
matrix is composed of poorly sorted granular coarse
sand. Scoured bases are also more common. This
facies commonly passes upward into matrix-supported conglomerates or stratified sandstones (facies C or D; fig. 5). This facies is most common in
the upper sequence (fig. 2).
This facies is suggestive of cohesionless debris
flows dominated by frictional grain interactions
(Kim et al. 1995; Sohn et al. 1997). The lack of
inverse grading and imbrication of clasts is probably due to suppression of clast collision. The welldeveloped parallel clast alignment suggests frictional interaction among clasts along deposit margins (snout/lobe) (Major 1998).
Facies

C:

Sand-Matrix-Supported Conglomerate.

This facies refers to medium- to thick-bedded


(0.31 m thick), pebble-to-cobble conglomerates
whose clasts are either supported by sand matrix
or in loose contacts. Gravel clasts are poorly aligned
relative to those of facies A and B conglomerates.
This facies occurs either as basal portions of upward-fining units (figs. 6, 7) or as solitary units. In
the former, facies C conglomerates have scoured
bases and are overlain by stratified units of facies
D and E. They are coarse-tail normally graded and
laterally discontinuous. The upper contacts show
undulatory primary relief. In the latter, conglom-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

115

Figure 2.
Columnar log of the measured section (for location, see fig. 1). The section comprises three sequences
(lower, middle, and upper sequences) bounded by thick homogeneous mudstone units. Capital letters (AF) represent
facies codes. Numbers represent stratigraphic levels above the base of the measured section in meters. M 5
mudstone; S 5 sandstone; G 5 conglomerate.

erates have lenticular geometry with positive primary relief and abrupt lateral terminations. Grading is mostly absent. This facies is common in the
lower sequence (fig. 2).
The coarse-tail normally graded conglomerates

overlain by stratified sandy deposits are interpreted


as deposits of high-concentration dispersion of sand
and gravel that were developed beneath a turbulent
heavily sediment-laden flow (Lowe 1982; Todd
1989). Since gravel clasts are difficult to fully sup-

116

Y. K. SOHN ET AL.

Figure 2 (Continued)

port by turbulence, they are concentrated toward


the base of a turbulent flow to form a dense inertia
layer. Gravel clasts in the layer can be effectively
segregated according to their sizes under a hindered
settling condition (Druitt 1995). The coarse-tail
normal grading in this facies probably resulted from
abrupt immobilization of the size-graded inertia
layer rather than from prolonged suspension sedi-

mentation. The overall random clast orientation


and undulatory upper contacts support this interpretation. However, the solitary units with pronounced lenticular geometry resulted more likely
from cohesionless debris flows that are not necessarily associated with a turbulent flow.
Facies D: Stratified Pebbly Sandstone. This facies
includes well- to crudely stratified pebbly sand-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

117

Figure 2 (Continued)

stones and sandstones. They occur mostly as thin,


discontinuous interbeds (0.10.3 m thick) between
other facies units, gradually or abruptly overlying
units of facies AC (fig. 3). Stratification is planeparallel or low-angle inclined relative to bedding

planes. Internal truncation surfaces are rare. Meterthick units occur in the lower sequence, where the
stratification is defined by repetition of centimeters-thick, generally ungraded, coarse- to finegrained sand layers. Pebble clasts are either sparsely

118

Y. K. SOHN ET AL.

Figure 3. Sketch of part of the middle sequence. Stacked conglomerate beds (facies A and B) are either continuous
(lower left part) or abruptly terminating (open arrows) and are draped by (pebbly) sandstone layers (facies D and E).
Conglomerate beds overlying a topographic relief (facies B, bracketed) compensate for the relief and are laterally
offset, as shown by inclined layers of mudstone chips. A bouldery conglomerate bed in the upper part is rapidly
pinching out and suggests selective migration of large clasts to the lateral margins of a debris flow. For stratigraphic
position, see figure 2.

present or form discontinuous trains. Their long


axes are aligned parallel to bedding planes (fig. 7).
The centimeters-thick stratification, lacking
angle-of-repose cross-bedding and internal truncation surfaces, has been observed in many catastropic flood-flow deposits and attributed to rapid
accumulation under a swift, heavily sedimentladen turbulent flow (Harrison and Fritz 1982; Pierson and Scott 1985; Smith 1986, 1987a; Blair 1987;
Smith and Lowe 1991; Best 1992). Best (1992) and
Smith (1986, 1987a) interpreted the stratification
to be due to migration of low-relief bedforms
formed during the transition from dunes to upperstage plane beds (Bridge and Best 1988) on the basis

of lenticularity and the interdigitating nature of the


stratification. The centimeters-thick stratification
in this facies, which lacks these features, may however be a result of repetitive deposition from thin,
short-lived traction carpets developed beneath a
turbulent flow that carried abundant sediment load
and fluctuated in the suspended-load fallout rate
(Hiscott 1994; Sohn 1997).
Facies E: Thinly Stratified Sandstone. This facies
includes thinly stratified or horizontally laminated,
coarse- to fine-grained sandstones, centimeters to
decimeters thick (figs. 7, 8). They mostly overlie
units of facies B and D gradationally and only rarely
occur as solitary beds. They commonly show fin-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

119

Figure 4. Photograph showing amalgamated units of facies A conglomerates with clasts aligned parallel to bedding
planes. Note that large clasts are concentrated in the lower part (near scale). For stratigraphic position, see figure 2.
Hammer is 30 cm long.

ing-up trends and are devoid of gravel clasts. Thin


units are generally discontinuous, laterally extending less than a few meters, whereas thick units are
laterally persistent. Thick units are abundant in the
lower sequence.
Deposits similar to this facies are common in
alluvial fans and have been attributed to waningstage sheetfloods or water flows, dewatering of debris flows, or reworking of surficial deposits by
sheetwash during storms (Gloppen and Steel 1981;
Ballance 1984; Wells 1984). Common gradational
contacts with the underlying units of facies B or D
suggest that this facies resulted mostly from water
flows attendant to preceding sediment flows. The
horizontal lamination was most likely formed under upper-plane-bed conditions. The lack of angleof-repose cross-bedding and channel structures indicates sporadic shallow sheet flows (Ballance
1984).
Facies

F:

Fine-Grained

Homogeneous Deposits.

This facies includes all fine-grained deposits intercalated between gravelly and sandy beds. Except for
units along scoured contacts, they are bounded by
flat bases and tops. They are generally homogeneous and purple but appear to be stratified when
diffuse bands of sand are intercalated. They range
in grain size from clay to muddy sand and, in some
cases, contain scattered granule to fine pebble
clasts and thin sheetlike layers of massive or strat-

ified gravelly sandstones. This facies occurs either


as decimeter-thick discontinuous interbeds or as
meter-thick, laterally persistent beds along the sequence boundaries (2224-m and 4648-m intervals
in the measured section; fig. 2).
This facies most probably formed on inactive segments of alluvial-fan surfaces that were inundated
only by occasional floods. Thick units forming the
sequence boundaries indicate a long period of fan
abandonment and slow suspension sedimentation.
Thin discontinuous gravelly sandstone sheets
bounded by flat base and top are indicative of sheetflood deposits on distal fan surfaces.
Depositional Processes of Composite
Sediment Flows
Facies Relationships.
Individual sedimentation
units in the measured section consist of one to several facies units. The most commonly observed
type is graded and stratified conglomerate/sandstone units, in which conglomerates (facies AC)
are either sharply overlain by sandstones (facies D
and E) or gradually decrease in clast size and content toward the sandstone divisions (figs. 58). The
conglomerate divisions are generally thicker than
the sandstone divisions. Three representative bed
types (fig. 9) were obtained by summing up facies
transitions (fig. 10) and by comparison with real

120

Y. K. SOHN ET AL.

Figure 5. Photograph showing gradual facies transition from clast-supported conglomerate (facies B) to matrixsupported conglomerate (facies C) and crudely stratified pebbly sandstone (facies D). The facies C unit is wedge
shaped. For stratigraphic position, see figure 2. Hammer is 30 cm long.

occurrences in the outcrops. The relationship between these bed types and individual facies could
be determined by vertical facies transitions in several beds that are composed of more than four facies
units (figs. 7, 8) as well as by the summation of all
facies transitions in the measured section (fig. 10).
The facies sequence is tripartite, consisting of massive and clast-supported conglomerates with
muddy or sandy matrix in the lower part (facies A
or B), sand-matrix-rich and normally graded conglomerates and stratified gravelly sandstones in the
middle part (facies C and D), and thinly stratified
or laminated sandstones in the upper part (facies E)
(fig. 11). The facies sequence can be regarded as an
idealized depositional sequence produced by a composite sediment flow that comprises a debris flow,
a streamflow, and an intermediate flow condition
between. The facies sequence depicted in figure 10
is only a vertical expression of the facies relationships that may exist more commonly as proximaldistal relationships in the field rather than as vertical successions in a single locality.
Organization of Flow Types. The facies relationships suggest that the responsible sediment flows
comprised different flow types, either across the

flow depth (vertically) or along the flow length (longitudinally). We surmise that the flows were essentially longitudinally segregated, comprising a
preceding debris flow and a trailing dilute flow (fig.
12c). Such a flow segregation was documented by
a number of workers (Beaty 1963; Okuda et al. 1980;
Pierson 1980, 1986; Suwa and Okuda 1983; Suwa
1988; Best 1992; Wan and Wang 1994). They note
that the flow head or the preceding debris flow has
a steep front and contains the densest slurry and
the coarsest particles. The flow head consumes
much kinetic energy and hence moves slower than
the main body, occasionally slowing the flow and
allowing the slurry behind to pile up to great depths
(Pierson 1980, 1986; Wan and Wang 1994). The flow
body is more fluidal than the head and changes into
a progressively more dilute tail that accounts for
the recessional limb of a slurry-flood wave. The
body, having intermediate characteristics between
the preceding debris flow and the trailing water
flow, may be vertically segregated into a dense inertia layer and a dilute overlying flow as a result
of the meager yield strength of interstitial fluid (discussed later). Such a vertical segregation seems un-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

Figure 6. Photograph of matrix-supported conglomerates (facies C) gradationally overlain by stratified sandstones (facies D or E). Note the erosional lower contact
of the facies C conglomerate. For stratigraphic position,
see figure 2. Scale is 10 cm long.

stable and transient, being maintained as long as


the overlying flow is contained inside the highrelief front and lateral levees of the preceding debris
flow. Because there is a marked difference in mobility between the basal layer with high mechanical
strengths and the overlying fluid layer, the overlying fluid layer may spill over the debris-flow front
and levees and proceed in front of the underlying
basal layer, once the relief of the debris-flow front
and lateral levees is subdued via lateral spreading
and deposition. The facies sequence in figure 11
can thus be produced by a sediment flow with a
full gamut of flow types segregated longitudinally
along the flow length and partly across the flow
depth (fig. 12c).
There are several examples of longitudinally segregated flows, of which the organization of flow
types is opposite to that implied by the facies sequence. These flows were mostly initiated by tur-

121

bulent water floods and then transformed into debris flows (lahars) by incorporating sediments. The
debris flows transformed farther downcurrent into
hyperconcentrated flood flows (lahar-runout) by
mixing of the leading edges of the debris flows with
perennial streamflows (Pierson and Scott 1985;
Smith 1986; Scott 1988; Pierson et al. 1990; Best
1992; Scott et al. 1995). These flows therefore comprised a peak flow of hyperconcentrated flood flow
followed by a recessional debris flow (fig. 12a). In
the resultant deposits, hyperconcentrated floodflow deposits thicken downcurrent, whereas the
overlying debris flow deposits thin and disappear
downcurrent, producing an inversely graded transitional facies between (figs. 12b and 13). The opposite organization of flow types seems to have resulted from the difference in depositional setting.
The debris flows documented in the above studies
were mainly initiated by catastrophic events and
had large volumes. They mostly flowed and deposited their sediments along a river valley where they
were subject to continuous mixing with ambient
water. The facies sequence in this study is, however, applicable to noncatastrophic flows that had
already passed through river valleys in catchment
areas. The flows spread out laterally and moved on
unconfined alluvial-fan surfaces, where abundant
water for flow dilution is not available. Therefore,
these flows had only a small chance to be diluted
at the leading edges, thus resulting in different
organization of flow types and depositional
sequences.
Deposition from Debris Flows. Interpretation of
debris-flow deposits has relied mainly on the viscoplastic flow model (Johnson 1970, 1984) and the
inertial grain-flow model (Takahashi 1978, 1981).
Both models assume that the emplacement of a
debris flow occurs via en masse freezing as the driving shear stress drops below the yield strength of
the viscoplastic substance or as the dispersive pressure drops and grains lock up frictionally. Although
the applicability of these models has been questioned recently (Iverson 1997), many features of debris-flow deposits can be explained by these models, including lobate geometry, basal or overall
inverse grading, large floating or protruding clasts,
high-angle or subvertical clast imbrication, vertical
variation in clast alignment, semicircular channel
structures, and the presence of large friable sediment blocks and clast clumps (Johnson 1970, 1984;
Naylor 1980; Nemec et al. 1980; Postma et al. 1983;
Nemec and Steel 1984; Hiscott and James 1985; Cas
and Landis 1987; Nemec 1990; Sohn and Chough
1992; Kim et al. 1995).
Facies A and B conglomerates, however, show

122

Y. K. SOHN ET AL.

Figure 7. Photograph showing a fining-upward unit consisting of facies A (not shown), C, D, and E. Note dispersed
elongate clasts in the facies D unit, which are aligned parallel to the gradational facies boundaries. For stratigraphic
position, see figure 2. The pen is 13 cm long.

few of these features but have generally thin sheetlike geometry and parallel-aligned gravel clasts
from base to top of units. These features can be
explained if deposition has occurred incrementally
from a debris flow. Recent experiments and field
observations show that a debris flow usually consists of a number of surges or roll waves, and its
deposition occurs not by abrupt freezing of the entire flow but by incremental aggradation of individual surges (Li and Yuan 1983; Davies 1986, 1990;
Wan and Wang 1994; Major 1997). Each surge
shoulders aside deposits of earlier surges, spreads
out, and leaves a deposit that is thinner and longer
than the surge itself. However, the cumulative deposit thickness produced by a series of surges may
greatly exceed the average flow thickness. Deposits
of earlier surges do not consolidate before the arrival of later surges, resulting in amalgamation of
many surge layers into a single massive layer (Major 1997). A similar process of incremental deposition has also been observed in muddy debris flows
(Li and Yuan 1983; Wan and Wang 1994), suggesting
that such a process is common in both cohesive
and cohesionless debris flows.
We interpret the facies A and B conglomerates to
be mainly incrementally aggraded materials, rather

than frozen masses of debris flows, and thus have


resulted in several peculiar features. The welldeveloped parallel alignment of clasts is probably
related to high rates of laminar shear straining of
individual surges (Fisher 1971; Enos 1977) associated with spreading or extensional deformation
during emplacement and subsequent smearing by
later surges before the deposits of earlier surges consolidate. Combined with the incremental aggradation, successive shouldering aside of earlier deposits by later surges prevents the protrusion of
large clasts. Large clasts are therefore expected to
be found mostly along the front and lateral margins
of a deposit rather than on the top. An abruptly
pinching-out bed composed of large boulders (fig.
3) may be an example of clast concentration along
the flow margin. The lack of basal or overall inverse
grading may also be related to incremental deposition because vertical distribution of clast sizes
throughout a bed is determined by the grain size
distribution in individual surges. Vallance and
Scott (1997) showed, for example, that normal grading can be produced by incremental deposition of
a longitudinally segregated cohesive debris flow.
We suggest that the development of pervasive parallel-to-bedding clast alignment and the lack of

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

123

itary or unified debris flow. Otherwise, the following flow may modify the deposits of the preceding
debris flow to produce thinner and broader deposits.
It is suggestive that debris flow deposits associated
with composite sediment-flow events show thin
sheetlike geometry (e.g., Ballance 1984; Wells
1984).
Deposition from Hyperconcentrated Flood Flows.

Figure 8. Photograph of a fining-upward bed composed


of clast-supported conglomerate with muddy matrix (facies A; base not shown) and sand matrix (facies B) overlain by thickly stratified pebbly sandstone with gravel
bands (facies D) and laminated sandstone (facies E). This
outcrop is located 10 m apart from the measured section
and correlated to the 2830-m interval of the columnar
section of figure 2. Pencil is 13 cm long.

large floating clasts and inverse grading, in spite of


tight clast support, are possible evidence for incremental deposition.
The process of incremental aggradation seems especially accentuated in composite sediment flows
because different parts of a flow with different rheologies are not likely to be deposited all at once in
time and space but to form a composite bed through
several phases of deposition under different conditions. The generation of thin sheetlike geometry
in debris-flow deposits seems also improved by the
composite nature of a sediment flow. Researchers
have observed that the more fluid and mobile flow
body pushes the head of less mobile debris flow
forward (Pierson 1980, 1986; Wan and Wang 1994).
This behavior may help the debris flow spread its
materials over broader areas compared with a sol-

Recognition of hyperconcentrated flow deposits


from ancient sequences is hampered by the difficulties in properly estimating the flow properties,
such as sediment concentration and rheology, on
the basis of deposit characteristics. It is therefore
necessary to infer the hyperconcentrated flow processes on an approximate and indirect basis. There
have been many attempts to define hyperconcentrated flood flows, using various criteria such as
sediment concentration (2060 vol %; Beverage and
Culbertson 1964), rheology (Pierson and Costa
1987), transport and depositional processes (Smith
1986), and a combination of energy slope and sediment concentration (Wan and Wang 1994). In general, these studies show that (1) a hyperconcentrated flood flow is a two-phase or multiphase flow,
in which solid particles (sand and gravel) and interstitial fluid (water, silt, and clay) behave independently (Wan and Wang 1994; Coussot and Meunier 1996); (2) interstitial fluid possesses only a
meager yield strength and acts as a transporting
medium (Pierson and Costa 1987); and (3) the role
of turbulence can be important as a grain-supporting mechanism (Smith 1986). Therefore, grains of
different sizes can behave differently in a hyperconcentrated flood flow, leading to density stratification or bipartite division of the flow into a dense
and coarse-grained lower part and a dilute and finergrained upper part. Druitt (1995) demonstrated experimentally that a bidispersion, having a solid concentration between 20 vol % and 44 vol %,
segregates into clusters of different grains, which
then separate by gravitational convection.
Recognition of hyperconcentrated flow deposits
from ancient sequences can thus be achieved by
finding evidence for bipartite division of a flow (e.g.,
Todd 1989). This approach is indirect but may be
more practical in identifying supposed hyperconcentrated flood-flow deposits because criteria for
hyperconcentrations, such as sediment concentrations and rheologies (viscosity and yield strength),
can hardly be inferred from deposit characteristics.
Furthermore, a single rheology or sediment concentration is not sufficient to describe or characterize a hyperconcentrated flow that is prone to be
bipartite. One has to remember, however, that a
bipartite flow may not necessarily be in a hyper-

124

Y. K. SOHN ET AL.

Figure 9. Three representative bed types common in the middle sequence (bed type I), the upper sequence (bed
type II), and the lower sequence (bed type III). They are composed of relatively thick conglomerate divisions overlain
by thin sandstone divisions.

concentrated condition sensu stricto (Beverage and


Culbertson 1964) because a bipartite flow can be
generated by an excess of sediment supply toward
the flow base relative to the deposition rate, irrespective of the sediment concentration or rheology
of the flow (Sohn 1997).
The dense basal layer has a similar rheology and
similar momentum transport processes to those of
a debris flow (Lowe 1982; Smith 1986; Smith and
Lowe 1991), leading some researchers (Shanmugam
1996, 1997) to identify the basal layer with a debris
flow. There are some differences, however, such as
in the source of shear stress, direction of granular
temperature dissipation, shape of velocity profile,
vertical distribution of shear stress, and sediment
concentration profile (Sohn 1997). It therefore
seems rash to identify the dense basal part of a
bipartite flow with a unified debris flow simply on
the basis that a single flow should not be considered
to represent multiple rheologies.
Some of facies C and D deposits in the Yongdong
alluvial fan could be interpreted with some certainty to be deposits of intermediate or hyperconcentrated flows because they are sandwiched between debris-flow deposits (facies A and B) and
streamflow deposits (facies E) within apparently
single sedimentation units (e.g., figs. 7, 8). If these
deposits were to occur solitarily, it would not be
easy to interpret them as hyperconcentrated floodflow deposits because similar deposits can also be
produced by high-magnitude streamflows. Facies C
and D deposits are also interpreted to have resulted

from the dense basal part or traction carpet of a


bipartite flow. It is not possible to infer the overall
sediment concentration or rheology of the flow, but
the flow is estimated to have had a sediment concentration much larger than that of a normal
streamflow.
Depositional Setting of the Composite
Sediment Flows
Description of Sequences. The measured section
can be divided into three sequences by meter-thick
reddish mudstone beds (facies F): lower (022-m interval of the measured section), middle (2446 m),
and upper (4855 m) sequences (fig. 2). The three
sequences are comparable on scale to the sequence of Heward (1978). Each sequence is distinct in facies composition, clast size, and bedding
pattern, as described below.
The lower sequence is characterized by the
smallest content and size of gravel clasts compared
with the overlying sequences. It consists of alternations of disorganized conglomerates, decimeters
to a meter thick, and stratified (gravelly) sandstones, many of them forming graded and stratified
units. The conglomerates are generally sand matrixsupported, rest mostly on erosional surfaces,
and show marked lateral variations in unit thickness. Gravel clasts are mostly granitic and the
coarse sand- to granule-sized fraction of the matrix
appears to have been derived from the same granitic
rocks. This sequence contains two coarsening-to-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

Figure 10.

125

Number of facies transitions in the measured section

fining upward cycles (813 m and 1822 m) demarcated by decimeter-thick purple mudstones.
Such cycles are suggestive of sporadic emplacement of coarse-grained sediments, probably by
shifting debris-flow lobes across fan surface.
The middle sequence contains the most abundant and coarsest gravel clasts and is dominated by
amalgamated disorganized conglomerates, 23 m
thick, with a muddy or sandy matrix (fig. 2). Conglomerate beds are bounded by slightly erosional
bases and relatively flat tops. Each depositional
unit can be discerned by the contrast in texture or
by the presence of mudstone chips and lenticular
sandstone interbeds. Pebble- to boulder-sized clasts
are tightly packed, with elongate ones aligned parallel to bedding planes. Clasts consist of granite
(48%), slate (35%), and metasandstone (17%) fragments. Interbedded mudstones are rare and thin
and tend to drape underlying conglomerate beds
with positive relief (near 43 m; fig. 3). Mudstone
chips are relatively abundant and range in size from
granule to boulder. They are in some cases crossbedded (fig. 3).
The upper sequence consists of stacked, sheetlike conglomerates separated by (pebbly) sandstone
wedges. It is characterized by fining-up units, 5070
cm thick, composed of clast-supported conglomerates with sandy matrix and overlying stratified
conglomerates or pebbly sandstones (fig. 2). The
stratified deposits compensate for local relief produced by underlying conglomerates and are commonly scoured by overlying conglomerates. The
conglomerates occur mostly as discrete beds, compared with the amalgamated beds of the middle

sequence. Clasts are better sorted and finer grained


(cobble sized) than those of the middle sequence.
Development of Sequences. Contrasting depositional sequences demarcated by thick intervening
mudstone beds suggest several phases of aggradation of coarse-grained deposits during alluvial-fan
progradation interrupted by migration (switching)
and abandonment of an alluvial-fan segment (Heward 1978). The vertical variations in sequence
characteristics, such as overall clast size, bed thickness, and stacked nature, can be principally attributed to the downfan and down-flow variations of
alluvial-fan processes (Heward 1978). It is generally
reported that hyperconcentrated and dilute flows
become more common toward the distal part of an
alluvial fan (e.g., Meyer and Wells 1995). Such a
trend has been generally attributed to the differences in mobility between solitary debris flows and
water flows but can also be produced by differential
deposition from a composite sediment flow. It is
supposed that the frontal debris-flow part of a composite sediment flow is deposited near the proximal
part of an alluvial fan, via either forming lateral
levees or paving its materials along the flow path,
whereas the more fluid rear part can continue to
move farther downfan.
Facies occurrences in the three sequences suggest that the coarsest sequence (the middle sequence) dominated by mud-matrix debris-flow
deposits (facies A) represents the proximal alluvial-fan facies, whereas the fine-grained lower sequence dominated by hyperconcentrated floodflow deposits (facies C and D) represents a more
distal facies. The coarsening- and thickening-up-

126

Y. K. SOHN ET AL.

Figure 11. Facies sequence obtained by the summation of facies transitions within individual sedimentation units
(fig. 10) and comparison with real occurrences (e.g., figs. 7, 8), with brief descriptions and interpretations on the right
columns. The facies sequence represents an idealized sequence of facies that can be produced by a composite debris
flowhyperconcentrated flood flowstreamflow event.

ward trend from the lower to the middle sequences is therefore interpreted to represent progradation of a proximal fan segment over distal
fan deposits. The abrupt changes in clast composition from granitic (lower sequence) to polymictic (middle sequence) are, however, interpreted
to have been caused by the changes in drainage
basin characteristicsespecially the source-rock
lithology. The transition from the middle to the
upper sequences is characterized by fining of
clast sizes and thinning of bed thicknesses but
most notably by the changes in the matrix type
from muddy to sandy. There are only a few conglomerate units with sandy matrices in the middle sequence, whereas the upper sequence totally
lacks conglomerates with a muddy matrix. The
fining- and thinning-upward trend from the middle to the upper sequences is probably attributable to the retreat of an active fan segment, but
the abrupt changes in the matrix type are more
likely caused by the changes in drainage basin
characteristics.
Traditionally, alluvial fans have been divided

into fluvial dominated (or wet) and debris flow


dominated (or dry), depending on prevalent depositional processes (Bull 1972, 1977; Schumm 1977;
Gloppen and Steel 1981; Kostaschuk et al. 1986).
The alluvial-fan sequences in the Yongdong Basin,
comprising deposits of both debris flows and dilute
flows, have characteristics of both wet and dry alluvial fans. This does not, however, mean that the
Yongdong alluvial fan experienced drastic changes
in climatic or tectonic regimes. It is recognized that
both types of alluvial fan can occur simultaneously
within a single geographical area because depositional processes on an alluvial fan are primarily
determined by local sediment types and flood hydrographs of associated drainage basins (Kostaschuk et al. 1986; Wells and Harvey 1987). An alluvial-fan succession may also show considerable
temporal variations in constituent facies as its contributing drainage basin evolves. In this respect, significant stratigraphic variations in dominant facies
as well as in clast composition and matrix type can
be caused by changes in the drainage basin characteristicsespecially the source rock lithology, re-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

127

Figure 12. Comparison of two contrasting styles of debris flowhyperconcentrated flood flowstreamflow events.
a, A hyperconcentrated flood flow is generated by the dilution at the leading edge of a debris flow that entered a
river valley. The resultant sediment flow therefore comprises a preceding hyperconcentrated flood flow, followed by
a recessional debris flow. As the dilution front propagates behind a debris flow, the entire flow may transform into
a hyperconcentrated flow (Pierson and Scott 1985; Smith 1986, 1987b; Scott 1988). b, The resultant deposits from
this sediment flow consist of downstream-thickening (and eventually thinning) hyperconcentrated flood-flow deposits
overlain by upstream-thickening debris-flow deposits. Four columns, showing representative facies in proximal to
distal parts, are given. c, This study suggests a different organization of flow types with a debris flow preceding in
advance of a hyperconcentrated flood flow and streamflow. d, The resultant deposits are envisaged to comprise debrisflow deposits in the proximal part, which are successively overlain and replaced by hyperconcentrated flood flow and
streamflow deposits toward the downstream direction. Four representative columns are given. The transition facies
(col. 2) is markedly different from that of b. The two models may differ greatly in time scale and dimension and
should not be compared on a footing of equality.

lief, and drainage density. Changes in these factors


result in different flood hydrographs (e.g., flashy vs.
sluggish) and different transport modes of (coarsegrained) sediments from the catchment (Ritter et
al. 1995). The changes in clast composition and matrix type in the three sequences of the measured
section were most likely produced by successive
exposure of different source rocks, caused by retreat
of an alluvial fan and lowering of a drainage basin

following repeated tectonic uplift, which then led


to changes in the nature of sediment supply with
time. Sediments of the lower and upper sequences,
mostly lacking a mud fraction, were probably
drained from large and less rugged catchments,
whereas those of the middle sequence dominated
by mud-rich debris-flow deposits were from small
and steep (rugged) catchments (Kostaschuk et al.
1986; Allen and Hovius 1998).

128

Y. K. SOHN ET AL.

Figure 13. Photograph of 1980 Mount St. Helens lahar deposit exposed along the North Fork Turtle River, Washington. The lower part of the deposit is finer-grained and inversely graded and erosionally overlies pre-1980 mud
deposits (below the dotted line). This part is interpreted as being composed of deposits of a hyperconcentrated flood
flow (lahar runout) that proceeded a debris flow. The gravelly debris-flow deposit gradually overlies the hyperconcentrated flood-flow deposits, forming an overall inversely graded transition facies. The pencil is 17 cm long.
Implications of the Composite Sediment Flow Deposits. There is a close association between the ma-

trix type of conglomerates and the depositional processes in the measured section. The conglomerates
in the lower and upper sequences, mostly having
sandy matrices, are commonly intercalated with
hyperconcentrated flood-flow deposits (facies C and
D), whereas the conglomerates in the middle sequence, which comprise muddy matrices, generally
lack interbeds of dilute-flow deposits. This association suggests that sand-matrix debris flows are
more prone to transform into hyperconcentrated
flows. In general, noncohesive flows tend to transform easily into dilute flows via dilution, whereas
cohesive flows remain debris flows to their termini
(Pierson 1986; Scott 1988; Vallance and Scott 1997).
The availability of silt-clay fractions in the drainage
area therefore seems important in determining
types and spacial distribution of facies.
Flow behavior of debris flows from a drainage
basin to its associated fan is largely governed by
longitudinal profiles of tributaries because the
channel slope regulates transportation forces (trac-

tive force and the slope-direction component of sediment weight) that affect runout distances and geomorphic processes (erosion and deposition) of the
debris flows (Ohmori and Shimazu 1994; Cenderelli
and Kite 1998). In channels of an exponential function type, large curvature may significantly decrease transportation forces downstream, and thus
upstream debris flows cannot runout to the gentle
downstream segment (alluvial-fan surface) unless
they are transformed into fluid, turbidity flows
(Ohmori and Shimazu 1994). According to Ohmori
and Shimazu (1994), tributaries of an exponential
function type develop in a less rugged drainage basin. We infer that the facies variations among the
three sequences reflect the changes in relief development and thus the changes in the longitudinal
profiles of tributaries in the catchment area with
time.
According to recent experiments on cohesionless
debris flows (Major 1997), temporally distinct debris flows as well as contemporaneous ones produce mostly amalgamated beds without distinct in-

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

ternal boundaries that demarcate flow waves/


events even though their deposits are emplaced by
incremental sedimentation rather than en masse
freezing. Similar amalgamated conglomerates (facies A and B) are contained in the middle sequence
of the measured section. They, however, commonly
show textural variations in vertical section (figs. 2,
3). Relatively distinct internal contacts and prevalent clast alignments of the debris-flow deposits
might have resulted from prolonged incremental
sedimentation of a number of longitudinally wellsorted debris flows (Major 1997). That is, they may
result from discrete waves of debris flows rather
than from closely spaced surges of a single flow
event. Such a distinction of debris flows is circumstantially supported by the facies successions (figs.
9, 11) obtained from the measured section, representing a composite debris flowhyperconcentrated
flood flowstreamflow event. In this respect, presence or absence of the facies succession can be an
indirect measure in determining whether flow
units of debris-flow deposits can be correctly
recognized.
Conclusions
Alluvial-fan deposits in the Cretaceous Yongdong
Basin, Korea, consist of conglomerates and sandstones deposited by debris flows, intermediate
flows, and dilute flows. Some of the conglomerates
and sandstones constitute single sedimentation
units, suggestive of composite sediment-flow processes. The composite sediment flows are inferred
to have been longitudinally segregated into a preceding debris flow and a trailing dilute flow, with
an intermediate flow between. The deposits from
the preceding debris-flow part of the composite
flow are distinguished from the deposits of solitary
debris flows in that they have pervasive parallel
alignment of clasts and thin sheetlike geometry but
lack basal inverse grading, large floating clasts, and
other features of classical debris-flow deposits.
These differences were most likely produced by in-

129

cremental aggradation of debris flows and modification of debris-flow deposits by later surges of
composite sediment flows. The deposits of intermediate flows show evidence for deposition from
bipartite flows that comprise dense inertia layers
or traction carpets beneath dilute and turbulent
flows. We suggest that the bipartite flow can be
approximately equated with a hyperconcentrated
flow based on the fact that a hyperconcentrated
suspension has a meager yield strength and is prone
to be density stratified. The relationship between
density stratification and rheology (viscosity, yield
strength, and sediment concentration) of a flow
needs to be established.
As far as can be judged from exposures of the
Yongdong alluvial-fan succession, deposition from
the composite sediment flows produces spatiotemporal variations in facies composition both as a result of an alluvial-fan evolution and as a result of
flow transformation over travel distances: the
lower sequence dominated by sandy-to-gravelly hyperconcentrated flow deposits mostly with granitic
clasts, the middle sequence dominated by coarsegrained debris-flow deposits with mud matrix and
polymictic gravel clasts, and the upper sequence
dominated by sand-matrix debris-flow deposits.
The coarsening to fining of the sequences reflects
progradation and retreat of an alluvial fan, but the
changes in clast composition and matrix type more
likely resulted from the changes in the characters
of supplied sediments as different source rocks
were exposed in the drainage area.
ACKNOWLEDGMENTS

This research was supported through grants to Y.


K. Sohn (project BSRI-96-5402) and C. W. Rhee
(BSRI-96-5417) by the Basic Science Research Institute Program, Ministry of Education, Korea. This
article benefited from helpful review and comments by an anonymous reviewer, as well as comments by and discussions with S. B. Kim and J.
Major.

REFERENCES CITED

Allen, P. A., and Hovius, N. 1998. Sediment supply from


landslide-dominated catchments: implications for basin-margin fans. Basin Res. 10:1935.
Ballance, P. F. 1984. Sheet-flow-dominated gravel fans of
the non-marine Middle Cenozoic Simmler Formation,
Central California. Sed. Geol. 38:337359.

Beaty, C. B. 1963. Origin of alluvial fans, White Mountains, California and Nevada. Assoc. Am. Geog. Annals 53:516535.
Best, J. L. 1992. Sedimentology and vent timing of a catastrophic volcaniclastic mass flow, Volcan Hudson,
Southern Chile. Bull. Volcanol. 54:299318.

130

Y. K. SOHN ET AL.

Beverage, J. P., and Culbertson, J. K. 1964. Hyperconcentrations of suspended sediment. J. Hydraulic Div.,
ASCE 90:117128.
Blackwelder, E. 1928. Mudflow as a geological agent in
semiarid mountains. Geol. Soc. Am. Bull. 39:465484.
Blair, T. C. 1987. Sedimentary processes, vertical stratification sequences, and geomorphology of the Roaring River alluvial fan, Rocky Mountain National Park,
Colorado. J. Sed. Petrol. 57:118.
Blair, T. C., and McPherson, J. G. 1994. Alluvial fans and
their natural distinction from rivers based on morphology, hydraulic processes, sedimentary processes,
and facies assemblages. J. Sed. Res. 64A:450489.
Blissenbach, E. 1954. Geology of alluvial fans in semiarid
regions. Geol. Soc. Am. Bull. 65:175190.
Bridge, J. S., and Best, J. L. 1988. Flow, sediment transport
and bedform dynamics over the transition from dunes
to upper-stage plane beds: implications for the formation of planar lamination. Sedimentology 35:
753764.
Bull, W. B. 1963. Alluvial-fan deposits in Western Fresno
County, California. J. Geol. 71:243251.
. 1972. Recognition of alluvial fan deposits in the
stratigraphic record. In Rigby, J. K., and Hamblin, W.
K., eds. Recognition of ancient sedimentary environments. Soc. Econ. Paleont. Mineralogists Spec. Publ.
16:6383.
. 1977. The alluvial-fan environment. Prog. Phys.
Geogr. 1:222270.
Cas, R. A. F., and Landis, C. A. 1987. A debris flow deposit
with multiple plug-flow channels and associated side
accretion deposits. Sedimentology 34:901910.
Cenderelli, D. A., and Kite, J. S. 1998. Geomorphic effects
of large debris flows on channel morphology at North
Fork Mountain, Eastern West Virginia, USA. Earth
Surf. Proc. Landforms 23:119.
Choi, S. J.; Kim, Y. B.; and Kim, B. C. 1995. Stratigraphy
and paleontology of the Cretaceous sedimentary strata
in the Yeongdong Basin. Taejon, Korea Inst. Geol.,
Mining and Materials, 118 p.
Choi, Y. S. 1996. Structural evolution of the Cretaceous
Eumseong Basin, Korea. Unpub. Ph.D. dissertation,
Seoul National University, Korea.
Chun, H. Y.; Um, S. H.; Choi, S. J.; Kim, Y. B.; Kim, B.
C.; and Choi, Y. S. 1993. Fossil floral and faunal assemblage and paleoenvironmental modelling study on
the Cretaceous sedimentary basins scattered in/near
the Ogcheon Belt (I). Taejon, Korea Inst. Geol., Mining
and Materials, 122 p.
Chun, S. S., and Chough, S. K. 1992. Depositional sequences from high-concentration turbidity currents,
Cretaceous Uhangri Formation (SW Korea). Sed. Geol.
77:225233.
Costa, J. E. 1984. Physical geomorphology of debris flows.
In Costa, J. E., and Fleischer, P. J., eds. Developments
and applications of geomorphology. Berlin, Springer,
p. 268317.
Coussot, P., and Meunier, M. 1996. Recognition, classification and mechanical description of debris flows.
Earth-Sci. Rev. 40:209227.

Davies, T. R. H. 1986. Large debris flows: a macro-viscous


phenomenon. Acta Mechanica 63:161178.
. 1990. Debris-flow surgesexperimental simulation. J. Hydrol. 29:1846.
Druitt, T. H. 1995. Settling behaviour of concentrated
dispersions and some volcanological applications. J.
Volcanol. Geotherm. Res. 65:2739.
Enos, P. 1977. Flow regimes in debris flow. Sedimentology 24:133142.
Fisher, R. V. 1971. Features of coarse-grained, high-concentration fluids and their deposits. J. Sed. Petrol. 41:
916927.
Gloppen, T. G., and Steel, R. J. 1981. The deposits, internal structure and geometry in six alluvial fanfan
delta bodies (Devonian-Norway)a study in the significance of bedding sequences in conglomerates. In
Ethridge, F. G., and Flores, R. M., eds. Recent and
ancient nonmarine depositional environments: models for exploration. Tulsa, Okla., Soc. Econ. Paleont.
Mineral., Spec. Publ. 31:4969.
Harrison, S., and Fritz, W. J. 1982. Depositional features
of March 1982 Mount St. Helens sediment flows. Nature 299:720722.
Heward, A. P. 1978. Alluvial fan sequence and megasequence models: with examples from Westphalian DStephanian B coalfields, northern Spain. In Miall, A.
D., ed. Fluvial sedimentology. Can. Soc. Petrol. Geol.,
Mem. 5:669702.
Hiscott, R. N. 1994. Traction-carpet stratification in turbiditesfact or fiction? J. Sed. Petrol. 64A:204208.
Hiscott, R. N., and James, N. P. 1985. Carbonate debris
flows, Cow Head Group, western Newfoundland. J.
Sed. Petrol. 55:735745.
Hong, S. H.; Lee, B. J.; and Kim, W. Y. 1980. Explanatory
text of the geological map of Muju sheet. Seoul, Korea
Inst. Energy and Resources, 28 p.
Hooke, R. J. 1967. Processes on arid-region alluvial fans.
J. Geol. 75:438460.
Iverson, R. M. 1997. The physics of debris flows. Rev.
Geophys. 35:245296.
Iverson, R. M., and Denlinger, R.P. 1987. The physics of
debris flowsa conceptual assessment. In Beschta, R.
L.; Blinn, T.; Grant, G. E.; Swanson, F. J.; and Ice, G.
G., eds. Erosion and sedimentation in the Pacific Rim.
Int. Assoc. Hydraul. Sci., Corvallis Symp., Proc.,
p. 155165.
Jin, M. S.; Shin, S. C.; Kim, S. J.; Choo, S. H.; and Chi,
S. J. 1993. Radiometric ages from granites and hypabyssal rocks and igneous activity in the mid-western
part of the Ogcheon Fold Belt. Taejon, Korea Inst.
Geol., Mining and Materials, 54 p.
Johnson, A. M. 1970. Physical processes in geology. San
Francisco, Freeman Copper, 577 p.
. 1984. Debris flow. In Brunsden, D., and Prior, D.
B., eds. Slope instability. Chichester, Wiley, p.
257361.
Kim, B. C., 1996, Sequential development of depositional
systems in a strike-slip basin: southern part of the
Cretaceous Yongdong Basin, Korea. Unpub. Ph.D. dissertation, Yonsei University, Korea.

Journal of Geology

DEBRIS AND FLOOD-FLOW DEPOSITS

Kim, D. H.; Chang, T. W.; Kim, W. Y.; and Hwang, J. H.


1978. Explanatory text of the geological map of Ogcheon sheet. Seoul, Korea Inst. Energy and Resources,
21 p.
Kim, D. H., and Lee, B.J. 1986. Geological report of the
Chongsan sheet. Seoul, Korea Inst. Energy and Resources, 20 p.
Kim, K. B., and Hwang, J. H. 1986. Geological report of
the Yongdong sheet. Seoul, Korea Inst. Energy and Resources, 24 p.
Kim, S. B.; Chough, S. K.; and Chun, S. S. 1995. Bouldery
deposits in the lowermost part of the Cretaceous
Kyokpori Formation, SW Korea: cohesionless debris
flows and debris falls on a steep-gradient delta slope.
Sed. Geol. 98:97119.
Korea Institute of Geology, Mining and Materials. 1995.
Geologic map of Korea, scale 1 : 1,000,000.
Kostaschuk, R. A.; Macdonald, G. M.; and Putnam, P. E.
1986. Depositional process and alluvial fan-drainage
basin morphometric relationships near Banff, Alberta,
Canada. Earth Surface Process and Landforms 11:
471484.
Lee, D. W., and Paik, K. H. 1989. Sedimentological characteristics along Yongdong fault zone in Cretaceous
Yongdong Basin, Korea. J. Geol. Soc. Korea 25:
259272.
. 1990. Evolution of strike-slip fault-controlled
Cretaceous Yongdon Basin, South Korea: signs of
strike-slip tectonics during infilling. J. Geol. Soc. Korea 26:257276.
Li, J., and Yuan, J. 1983. The main features of the mudflows in Jiang-Jia Ravine. Z. Geomorphol. 27:326341.
Lowe, D. R. 1982. Sediment gravity flows. II. Depositional models with special reference to the deposits
of high-density turbidity currents. J. Sed. Petrol. 52:
279297.
Major, J. J., 1997, Depositional processes in large-scale
debris-flow experiments. J. Geol. 105:345366.
. 1998. Pebble orientation on large, experimental
debris-flow deposits. Sed. Geol. 117:151164.
Meyer, G. A., and Wells, S. G. 1995. Modern fire- and
flood-related sedimentation and facies distribution in
Holocene alluvial fans, Yellowstone National Park. In
Blair, T. C., and McPherson, J. G., eds. Alluvial
fansprocesses, forms, controls, facies models, and
use in basin analysis. SEPM Int. Res. Conf., Proc. and
Abstr., Death Valley, Calif., Oct. 1721, 1995, p. 66.
Naylor, M. A. 1980. The origin of inverse grading in
muddy debris flow depositsa review. J. Sed. Petrol.
50:11111116.
Nemec, W. 1990. Aspects of sediment movement on
steep delta slopes. In Colella, A., and Prior, D. B., eds.
Coarse-grained deltas. Spec. Publ. Int. Assoc. Sed. 10:
2973.
Nemec, W., and Muszynski, A. 1982. Volcaniclastic alluvial aprons in the Tertiary of Sofia District (Bulgaria). Ann. Soc. Geol. Poloniae 52:239303.
Nemec, W.; Porebski, S. J.; and Steel, R. J. 1980. Texture
and structure of resedimented conglomerates: exam-

131

ples from Ksiaz Formation (Famennian-Tournaisian),


southwestern Poland. Sedimentology 27:519786.
Nemec, W., and Steel, R. J. 1984. Alluvial and coastal
conglomerates: their significant features and some
comments on gravelly mass-flow deposits. In Koster,
E. H., and Steel, R. J., eds. Sedimentology of gravels
and conglomerates. Can. Soc. Petrol. Geol., Mem. 10:
131.
Nilsen, T. H. 1982. Alluvial fan deposits. In Scholle, P.
A., and Spearing, D., eds. Sandstone depositional environments. Tulsa, Okla., Am. Assoc. Petrol. Geol.,
Mem. 31:4986.
Ohmori, H., and Shimazu, H. 1994. Distribution of hazard types in a drainage basin and its relation to geomorphological setting. Geomorphology 10:95106.
Okuda, S.; Suwa, H.; Okunishi, K.; Yokoyama, K.; and
Nakano, M. 1980. Observations on the motion of a
debris flow and its geomorphology effects. Z. Geomorphol. 35:142163.
Pierson, T. C. 1980. Erosion and deposition by debris
flows at Mt. Thomas, North Canterbury, New Zealand. Earth Surf. Proc. 5:227247.
. 1986. Flow behavior of channelized debris flows,
Mount St. Helens, Washington. In Abrahams, A. D.,
ed. Hillslope processes. Boston, Allen & Unwin, p.
269296.
Pierson, T. C., and Costa, J. E. 1987. A rheologic classification of subaerial sediment-water flows. In Costa,
J. E., and Wieczorek, G. F., eds. Debris flows/avalanches: processes, recognition, and mitigation. Geol.
Soc. Am., Rev. Eng. Geol. 7:112.
Pierson, T. C.; Janda, R. J.; Thouret, J. C.; and Borrero,
C. A. 1990. Perturbation and melting of snow and ice
by the 13 November 1985 eruption of Nevado del
Ruiz, Colombia, and consequent mobilization, flow
and deposition. J. Volcanol. Geotherm. Res. 41:1766.
Pierson, T. C., and Scott, K. M. 1985. Downstream dilution of a lahar: transition from debris flow to hyperconcentrated streamflow. Water Resources Res. 21:
15111524.
Postma, G.; Nemec, W.; and Kleinspehn, K. 1988. Large
floating clasts in turbidites: a mechanism for their emplacement. Sed. Geol. 58:4761.
Postma, G.; Roep, T. B.; and Ruegg, G. H. J. 1983. Sandygravelly mass-flow deposits in an ice-marginal lake
(Saalian, Leuvenumsche Beek Valley, Veluwe, The
Netherlands), with emphasis on plug-flow deposits.
Sed. Geol. 34:5982.
Ritter, D.F.; Kochel, R.C.; and Miller, J.R. 1995. Process
geomorphology (3d ed.). Brown, Dubuque, Iowa,
538 p.
Schumm, S. A. 1977. The fluvial system. New York, Wiley, 338 p.
Scott, K. M. 1988. Origins, behavior, and sedimentology
of lahars and lahar-runout flows in the Toutle-Cowlitz
River System. U.S. Geol. Surv. Prof. Paper 1447-A:
A1A74.
Scott, K. M.; Vallance, J. W.; and Pringle, P. T. 1995. Sedimentology, behavior, and hazards of debris flows at

132

Y. K. SOHN ET AL.

Mount Rainier, Washington. U.S. Geol. Surv. Prof. Paper 1547:156.


Shanmugam, G. 1996. High-density turbidity currents:
are they sandy debris flows? J. Sed. Res. 66:210.
. 1997. The Bouma Sequence and the turbidite
mind set. Earth-Sci. Rev. 42:201229.
Sharp, R. P., and Nobles, L. H. 1953. Mudflow of 1941
at Wrightwood, southern California. Geol. Soc. Am.
Bull. 64:547560.
Shimamura, S. 1924. Geologic map of Chosen, Yeongdong and Cheongsan Sheet. Seoul, Geol. Surv. Chosen.
Shin, S. C., and Jin, M. S. 1995. Radiometric age maps
of volcanic, plutonic and metamorphic rocks and ore
deposits in the Korean Peninsula. Taejon, Korea Inst.
Geol., Mining and Materials, 43 p.
Smith, G. A. 1986. Coarse-grained nonmarine volcaniclastic sediment: terminology and depositional process. Geol. Soc. Am. Bull. 97:110.
. 1987a. The influence of explosive volcanism on
fluvial sedimentation: the Deschutes Formation (Neogene) in central Oregon. J. Sed. Petrol. 57:613629.
. 1987b. Sedimentology of volcanism-induced aggradation in fluvial basins: examples from the Pacific
Northwest, U.S.A. In Ethridge, F. G.; Flores, R. M.;
and Harvey, M. D., eds. Recent developments in fluvial sedimentology. Tulsa, Okla., Soc. Econ. Paleont.
Mineral., Spec. Publ. 39:217228.
Smith, G. A., and Lowe, D. R. 1991. Lahars: volcanohydrologic events and deposition in the debris flowhyperconcentrated flow continuum. In Fisher, R. V.,
and Smith, G. A., eds. Sedimentation in volcanic settings. Tulsa, Okla., Soc. for Sed. Geol. (SEPM), Spec.
Publ. 45:5970.
Sohn, Y. K. 1997. On traction-carpet sedimentation. J.
Sed. Res. 67:502509.
Sohn, Y. K., and Chough, S. K. 1992. The Ilchulbong tuff
cone, Cheju Island, South Korea: depositional processes and evolution of an emergent, Surtseyan-type
tuff cone. Sedimentology 39:523544.
Sohn, Y. K.; Kim, S. B.; Hwang, I. G.; Bahk, J. J.; Choe,
M. Y.; and Chough, S. K. 1997. Characteristics and
depositional processes of large-scale gravely gilberttype forests in the Miocence Doumsan fan delta, Pohang Basin, SE Korea. J. Sed. Res. 67:130141.
Suwa, H. 1988. Focusing mechanism of large boulders to

a debris-flow fronts. Trans. Jpn. Geomorphol. Union


9:151178.
Suwa, H., and Okuda, S. 1983. Deposition of debris flows
on a fan surface, Mt. Ykedake, Japan. Z. Geomorphol.,
N.F., Suppl. Band 46:79101.
Takahashi, T. 1978. Mechanical characteristics of debris
flow. J. Hydraul. Div., ASCE 104:11531169.
. 1981. Debris flow. Annu. Rev. Fluid Mech. 13:
5777.
Todd, S. P. 1989. Stream-driven, high-density gravelly
traction carpets: possible deposits in the Trabeg Conglomerate Formations, SW Ireland and theoretical
considerations of their origin. Sedimentology 36:
513530.
Vallance, J. W., and Scott, K. M. 1997. The Osceola Mudflow from Mount Rainier: sedimentology and hazard
implications of a huge clay-rich debris flow. Geol. Soc.
Am. Bull. 109:143163.
Waitt, R. B., Jr.; Pierson, T. C.; MacLeod, N. S.; Janda, R.
J.; Voight, B.; and Holcomb, R. T. 1983. Eruption-triggered avalanche, flood, and lahar at Mount St. Helenseffects of winter snowpack. Science 221:
13941396.
Wan, Z., and Wang, Z. 1994. Hyperconcentrated flow.
International Association of Hydraulic Research Monograph Series. Rotterdam, A. A. Balkema, 290 p.
Wells, N. A. 1984. Sheet debris flow and sheetflood conglomerates in Cretaceous cool-maritime alluvial fans,
south Orkney Islands, Antarctica. In Koster, E. H., and
Steel, R. J., eds. Sedimentology of gravels and conglomerates. Can. Soc. Petrol. Geol., Memoir 10:
133145.
Wells, S. G., and Harvey, A. M. 1987. Sedimentologic and
geomorphic variations in storm-generated alluvial
fans, Howgill Fells, northwest England. Geol. Soc.
Am. Bull. 98:182198.
Whipple, K. X., and Dunne, T. 1992. The influence of
debris-flow rheology on fan morphology, Owens
Valley, California. Geol. Soc. Am. Bull. 104:887900.
Won, C. G., and Kim, K. T. 1969. Explanatory text of the
geological map of Sangju sheet (1 : 50,000). Seoul,
Geol. Surv. Korea, 34 p.
Yun, S. K., and Park, B. K. 1968. Explanatory text of the
geological map of Seolcheon sheet. Seoul, Geol. Surv.
Korea, 15 p.

You might also like