You are on page 1of 184

1

Friends of the Pleistocene Field Trip


1996 South-Central Cell
April 19 - 21, 1996

Upland, Lowland, and In Between - Landscapes in the Lampasas Cut Plain

Co-Sponsored by

Texas A&M University


and
Baylor University

EXPLAN"nON
AthNiat lJn~5 Age Iy, e pI
FO Ford 400·plesenl TA Tlench
WA West Range 4200·600 T1 Landlorm
FH For1 Hood 8000-5000 ~ Rovally paleosol
OT GeOfgetown 11 ,OOO·BOOO I"'TIT'I Othel soil
JA Jackson 15,000
Table of Contents

Lampasas Cut Plain: Episodic Development of an Ancient and Complex


Regional Landscape, Central Texas
o. T. Hayward, Peter M. Allen, and David L. Amsbury . . . . . . . . . . . . . . . . . . . . . . . . 1-1 - 1-97

Soils and Geomorphology of Cowhouse Creek, Fort Hood, Texas


Lee Nordt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-1 - 246

Amino Acid Racemization Analysis of the Chronology and Integrity of


Archaeological Sites in Central Texas
Glenn A. Goodfriend. G. Lain Ellis. and James T. Abbott . . . . . . . . . . . . . . . . . . . . . . . 3-1 - 3-13

Archaeology at Fort Hood


David L. Carlson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-1 - 4-21
L C P 1 Day 96

Episodic Development
of an
Ancient and Complex Regional Landscape

Central Texas

O. T. Hayward, Peler M. Allen, David L. Amsbury

I
.I

FRIENDS OF THE PLEISTOCENE


South Central Cell
1996 Field Trip-Central Texas
April 19, 1996

I-I
Publlshed by Friends of the Pleistocene, David L Carlson, Department of
Anthropology, Texas A & M University, College Station, nc,
77843-4352

AUTHORS

O. T. Hayward, Department of Geology, Baylor University, Waco, TX., 76798-7354, (817)


755-2361; e-mail, O_T_Hayward@baylor.edu

Peter M. Allen, Department of Geology, Baylor University, Waco, TX., 76798-7354, (817)
755-2361; e-mail.Dr.Alien is illiterate.

David L. Amsbury, NASA Johnson Space Center, Code SN 15, Houston, TX., 77058, (713)-
483-5160; e-mail, amsbury@snjsc.nasa.gov

We solicit your Interest, comments, observations, condemnations, vituperation, whatever.

J..'2
THE CRETACEOUS PRAIRIES OF TEXAS

While the principal interest of this trip is in the geomorphic region of the Lampasas Cut Plain, it
is also important to appreciate that the Cut Plain is but a subprovince of the of the far larger
Grand Prairie of Texas, itself only a part of the Cretaceous Prairies of Texas (FIG. A-1, A-2).

To the earliest visitors, Texas was known as "...... a land of prairies .... : (Roemer, 1848),
grasslands and savannah extending from the coast to the most distant reaches then explored.
But among the most beautiful and productive of all of these seemingly-endless grasslands were
the Cretaceous Prairies of Texas, developed on the outcrop of Cretaceous Rocks, a widening
band of distinctive landscapes extending from the Rio Grande to the Red River (Fig A-1)

Two major subdivisions further divide the Cretaceous Prairies; on the west the Grand Prairie, on
the east the Black Prairie. On a smaller scale as north-south trending belts within the Grand and
Black Prairies are yet smaller subprovinces, each differing from those adjoining, each on
different geologic foundations.

THE BLACK AND GRAND PRAIRIES


The major determinants in the geomorphic differentiation of these Cretaceous Prairies from all
the other Texas prairies, and also for the existence of subprovinces within the Cretaceous
Prairies, are structural and lithologic differences in underlying bedrock. The Grand Prairie,
westernmost of the Cretaceous Prairies, is developed on Comanchean (Lower Cretaceous)
rocks, mostly near-horizontal beds of limestone with lesser sands at base and top, exposed on
the Central Texas Platform. The Black Prairie, the eastern subdivision of the Cretaceous Prairies,
exists on Gulfian (Upper Cretaceous) rocks, principally great thicknesses of near-homogeneous
mudstones with subordinate chalks, dipping gently eastward into the western margin of the East
Texas Basin (Fig. A-3, 0-8)

On a smaller scale, and within both the Grand and Black Prairies, lesser subdivisions (Fig A-4),
are distinguished one from another on the basis of landform, soils, and vegetation, each
reflecting differences in underlying stratigraphy and geologic structure.

Since the emphasis of this trip is on the Lampasas Cut Plain, a subprovince of the Grand Prairie,
the remainder of this introduction is devoted largely to the Grand Prairie Subprovince of the
Cretaceous Prairies of Texas. Of the Black Prairie subprovinces, the Eastern Crosstimbers, the
common border of Grand and Black Prairies, is given added consideration (Fig A-4, 0-4).

THE BLACK PRAIRIE The Black Prairie, at its widest some forty miles across, extends from
the Red River to San Antonio, continuing as a narrow band to the Rio Grande. In the area of its
greatest width, it consists of four distinctive subprovinces, each reflecting a lithologic difference
In the underlying bedrock. Of these Black Prairie subprovinces the only one of immediate
Interest to us on this trip is the westernmost, the Eastern Crosstlmbers (Fig. A-2, A-4).

I.he Eastern Crosstlmbers We encounter the southern terminus of the Eastern Cross
Timbers at Waco. At its southern end the dark, narrow band of Crosstimbers woodland
extending northward from Waco to the Red Rivers is developed on sands and clays of the
Quaternary alluvial deposits of the Brazos River. From Hill County northward, it is formed on
the outcrop of the sandy Woodbine Group (lower Gulfian). Throughout this length, it forms the
common border of Grand and Black prairies (Fig. A·4).

1-3
Figure A·1 The Cretaceous Prairies of t?
Texas are developed on the outcrop belt of 'k.""
...'Ii
Cretaceous rocks. Thus they comprise one <t
of the largest of Texas geomorphic ;:,'"
q,o
subprovinces, and in early day Texas . (J

perhaps the most distlnct/ve of Texa; 'q,


.tJ
landscapes. o . WACO
/
• TEMPLE,

i
,I
, I

IZI I '
e;-.'S.
ra (
I
'li
ctl<S' I
't> I
C:. \ ....
f!
<.!1 I
"WACO
I

I
I. TEMPLE
I
I
I
!

Figure A-Z The Cretaceous Prairies are


further divided into Black and Grand Prairies, the
Black Prairie on Gulfian rocks, principally muds;
and the Grand Prairie on Comanchean rocks,
prinCipally limestones. The Black Prairie is
further divided into five smaller subprovinces, and
the Grand Prairie into four, each subdivision
reflecting differences in the underlying bedrock.

\-4
I -< TRINITY SHELF
EAST TEXAS BASIN .
GRAND - PRAIRIE

BLACK PRAIRIE

WESTERN CRDSSTIMBERS
LAMPASAS CUT PlAIN '
WASHITA PRAIRIE

\. ~
TEXAS BLACKLANDS:

"'" PENNSYLVANIAN FORMA nONS


""

Figure A-3 The landscape of the Cretaceous Prairies is dependent upon composition and
structure of the Cretaceous bedrock. The Grand Prairie is on near-horizontal limestones and
sands of Comanchean age. The Black Prairie is on Gulfian muds and chalks dipping eastward into
the East Texas Basin. Where each differing lithic unit crops out, it forms a band of recognizable
landscape, unique to that unit and different from those to either side.
BLACK PRAIRIE
WICHITA PALEO PLAIN

• DALLAS

LLANO BASIN

50 miles

Figure A-4 Subdivisions of the Grand and Black Prairies reflect differences in the
underlying bedrock. From Waco northward, the Black and Grand Prairies are separated by the
Eastern Cross-timbers, on sands of the Woodbine Group, basal unit of the Gulfian section, and
hence westernmost of the Black Prairie subprovinces. From east to west within the Grand
Prairie, the subprovinces are (1 )the Washita Prairie on rocks of the Washita Group; (2) the
Lampasas Cut Plain on rocks of the Fredericksburg group; (3) the Glen Rose Prairie, a narrow
band along the westem margin of the Lampasas Cut Plain, and as windows within the Cut Plain;
on rocks of the upper Trinity Group, and (4) the Western Crosstimbers, on sands of the basal
Trinity Group.
\-6
To early visitors the densely wooded Crosstimbers was so different from the near-treeless open
grasslands on either side that it became the principal marker on all early maps and emigrant
guides (irving, 1835, Ikin, 1841; Kendall, 1844). It remains a conspicuous geomorphic element
to the present day.

The Eastern Crosstimbers separates the Grand Prairie, a subdued but angular landscape of
mixed-grass and savannah on Lower Cretaceous "hard lime-rock", from the Black Prairie, gently
rounded "jello-mold" landforms of treeless tall-grass prairie on soft Upper Cretaceous blocky
"joint clay".

Throughout its length the Crosstimbers is veneered by deep, peach-colored, sandy soil derived
from the sands and muds of the underlying Quaternary Brazos Alluvium and the Upper
Cretaceous Woodbine Group. It supports a dense post oak forest, with lesser amounts of
blackjack oak and other deciduous trees, ani:J an almost impenetrable deciduous undergrowth.
The result is a narrow wall-like band of woodland in the midst of the Cretaceous Prairies. For
those immigrants of the last century, mostly from the southeastern states, this woodland was a
momentary return to the forests of all their previous experience. It recalled home and fireside,
while the prairies to east and west were strange, inhospitable, woodless, waterless, threatening
worlds.

THE GRAND PRAIRIE


Abruptly abutting the Eastern Crosstimbers on the west is the Grand Prairie (Fig. A-2, A-4).
The common denominator for all Grand Prairie landscapes was the underlying lower Cretaceous
rock foundation. Yet within this common inheritance were sufficient lithologic variations to
create distinctive geomorphic subprovinces. .From east to west, the Grand Prairie is divided
Into four geomorphic subdivisions, (1) the Washita Prairie, (2) the Lampasas Cut Plain (the
landscape of our major interest), (3) the Glen Rose Prairie, and (4) the Western Crosstimbers
(Fig. A-3, A-4). Each of these subdivisions is briefly described in the following paragraphs

The Washita Prairie We see the Washita Prairie at Stops 1, 2, and 3 of the present trip.
Formerly it was in mixed-grass prairie, now it is in farm and grazing land. The Washita Prairie
consists of broad, gently rolling upland grasslands with streams in shallow but angular valleys
marked by gallery forests. Typically the prairies are also dotted with isolated trees, live oak
mottes, and small mesquite savannas, associations early described as the most beautiful of the
Texas Prairies. This landscape is developed on limestone and shale of the Washita Group
(Comanchean, Cretaceous), lower Washita limestone (Georgetown) cropping out mostly on the
western Washita Prairie, upper Washita shale (Grayson Marl) exposed mostly on the east.
Therefore from west to east the landscape varies in relief, and the soils vary in depth, in
fertility, and in water holding capacity, depending on whether their inheritance was from
limestone or shale.

The eastern part of the Washita Prairie was originally mixed-grass prairie dominated by Little
Bluestem and was almost devoid of trees (Kendall, 1846; Roemer, 1848; Poole, 1964). Though
tie richest parts were in farmland from early in the era of settlement, a great deal more was
broken to the plow during the First World War and into the early 1920s. It remained largely in
farmland until the late 'twenties and early 'thirties, to be abandoned as marginal land during the
farm programs of the pre-World War II New Deal. Much of it has since reverted to secondary
grassland and savannah. The best of the soils, those of the eastern Washita Prairie, on Grayson
Marl (Del Rio Clay), remain in farmland today.

The western Washita Prairie was also originally grassland, but the soils were generally thin and
rocky, formed from the limestones of the lower Georgetown Formation (Washita Group,
Comanchean), and the vegetation included juniper savannah and even dense juniper brakes

1-7
along rocky hllisiopes. Still, within this limestone-dominated environment are several significant
clay horizons, and on the outcrop bands of each of these, following the contour of the land, one
can often see bands of Little Bluestem marking these more hospitable soil-environments. Thus
over the western Washita Prairie the grass tended to be shorter and less abundant, savannah
was far more common, tree bands and mattes marked the poSitions of perched water tables,
and bands of Little Bluestem followed the outcrop lines of thin but persistent clays. Because of
limited productivity, the western Washita Prairie has tended to remain more as it was in the
beginning, though each year woodlands expand at the expense of prairie.

The Lampasas Cut Plain The Lampasas Cut Plain is the geomorphic subdivision to the west
of the Washita Prairie, and the central and largest subprovince of the Grand Prairie (Fig. A-3, A-
4, 0-4). It is also the landscape that will dominate the view at all the remaining Stops of this
trip, Stops 4 through 17 (NOTE For reasons later explained, Stops 9 through 15 have been
omitted from this trip). The Cut Plain is a far more rugged land than the Washita Prairie, of
buttes and mesas with distant and dramatic views of broad grassland valleys separated by
narrow wooded mesa-like divides, almost the whole carved in rocks of the Fredericksburg Group
(Comanchean, Cretaceous).

The term 'cut plain", coined by Robert T. Hill in 1900, describes landscapes everywhere that
have characteristics he saw so clearly demonstrated here in the Lampasas Cut Plain;

A cut plain (dissected plain) is a stratum plain of any kind which has been so dissected
Into remnants by erosion that the level of the overall stratum plain is still recognizable in
the summits of the dissected members (Hill, 1900, p. 6).

Within the Cut Plain, the dramatic views of diverse and distant landscapes were, from the
beginning of European exploration, a common experience and an attractant to settlement. This
subprovince clearly reflects the stratigraphy and structure of the underling Fredericksburg rock-
foundation (Fig. A-3, 0-2, 0-9). The flat-topped mesas and divides are defended by the near-
horizontal, thin but massive, highly erosion-resistant Edwards Limestone (uppermost
Fredericksburg Group). On this flat upland surface soils are rocky and thin, and vegetation is
dominated by juniper and postoak. Still, on top of larger divide lands, particularly near the
eastern margin of the Cut Plain, remnants of the Washita Prairie remain, veneered with
productive soils derived from the Kiamichi Clay (lowermost member of the Georgetown
Formation, Washita Group). Where these clay-derived soils are present even divide-crests, once
upland prairie, have long been in farmland.

Slopes from the Edwards caprock to valley-floors are carved In the white, chalky, nodular
limestones of the Comanche Peak Formation, next underlying the Edwards caprock (Fig 0-6) ..
Soils on the steep Comanche Peak slopes are stripped by erosion almost as soon as they are
formed, the slope-microclimate is arid, and juniper brakes typically surround almost every mesa
and divide.

The valley floors, far the largest part of the Cut Plain landscape, are formed on the Walnut Clay,
the formation that next underlies the Comanche Peak Limestone. While the Walnut Clay is not
all clay, it is far more clay than anything else, and that has had a marked effect on landform,
soils, native vegetation, and land use. In the eyes of the earliest beholders, both Indian and
European, the Cut Plain valleys provided an almost "Garden of Eden" aspect. These broad valleys
constitute an area of major interest on the present trip, and the locus of most towns, farms,
roads, barbecue joints, and coffee stops in the whole of the Grand Prairie.

The Glen Rose Prairie But not all Cut Plain valleys are formed entirely in Walnut Clay.
Centered in the broad basins of the western Cut Plain, occupying low, rounded in-basin divides,
and existing also as long southeastward-narrowing tongues within most the entrenched valleys

1-8
of major drainages (particularly in the northern and western parts of the Cut Plain) are rugged,
inner-valleys carved into Glen Rose Limestone (of the Upper Trinity Group, Fig. 0-9). These
landscapes on the outcrop of Glen Rose rocks constitute the Glen Rose Prairie, the smallest
subprovince of the Grand Prairie though in reality most of it is surrounded by Edwards-capped
buttes and divides within the region typically described as Lampasas Cut Plain.

This more rugged, rockier, somewhat stair-stepped landscape of basin-center and river-
entrenched valley systems was originally in mixed-grass, savannah, and thin oak and juniper
woodland, all developed on the thin soils and varied rock foundations of the Glen Rose
Limestone. Thin, rocky soils and highly dissected landscapes generally precluded agriculture
over most of the Glen Rose Prairie, and it has therefore tended to remain as grazing land. With
the control of fire, the area of prairie has diminished and savannah and woodlands have
encroached, but the broad view from uplands still shows a clearly delineated Glen Rose Prairie,
within the broad expanse of the Lampasas Cut Plain.

The Western Crosstlmbers The western boundary of the Glen Rose Prairie, and generally
also of the Lampasas Cut Plain, is the westernmost subprovince of the Grand Prairie, the
Western Crosstimbers (Fig A-4). This is a more extensive, and more diffuse counterpart of the
Eastern Crosstlmbers, gentle wooded landscapes of sandy soils, in marked contrast to the lands,
vegetation, and soils of the remainder of the Grand Prairie. The Western Crosstimbers is
developed almost entirely on Twin Mountains and Antlers sands of the lower Trinity Group (Fig
0-9) These earliest Cretaceous sand, silt and clay beds underlie the "hard lime-rock" formations
of the Upper Trinity, Fredericksburg, and Washita Groups of the rest of the Grand Prairie and are
the substrate for the sandy soils of the Crosstimbers.

In the years of exploration and settlement, the western Crosstimbers was more open than the
Eastern Crosstimbers. Because the underlying rocks contained substantial sections of clay and
sandy clay, there were numerous glade-like prairies within the belt of woodland and savannah.
These islands of prairie within the sea of woodland became early targets of immigrants who
appreciated the combination of wood, water, and easily tilled soil typical of these glades.

For those western-bound from the southeastern states, as most Central Texas immigrants were,
the Western Crosstimbers was a second reminder of a distant, eastern, wooded and watered
home. To those continuing on westward, it was also the westernmost and last of the
Cretaceous Prairies. The contrast was enormous. Beyond the Crosstimbers lay the Red Rolling
Plains; of sparse grassland, of brush land, and of streams with grand names but little water,
mostly gyp and mostly bad. And mile by mile to the west prospects became ever more grim.

From the western margin of the Western Crosstimbers almost to the Great Valley of California,
lands became more arid, wood less plentiful, soils less rich, water less abundant, and life more
difficult (Marcy, 1854, p. 85). The effect on westward-bound immigrants was profound. This
was noted particularly in letters from women, wives and daughters of westering families, who
remembered with longing the distant watered world so many would not see again. It would be
Interesting to know how many of those continuing westward ultimately turned 'round to return
to this favored land-this last reminder of the eastern world they had left so far behind
(Hayward, Dolliver, Amsbury, Yelderman, 1992, p. 8-9)

THE CUT PLAIN FIELD TRIP


The present field trip consists of ten stops extending from Waco through McLennan, Bosque,
Hamilton, Coryell, and Bell counties, to terminate at Lake Belton Dam, near Belton, Texas.

The stops are arranged in two major "packages". Stops 1 through 3 show the processes and
landform of the initial establishment of Cut Plain drainage as an inheritance from the

1-9
disappearing Washita Prairie. Streams on the Washita Prairie followed an apparently "normal"
and orderly evolution. from (1) valley Inception on the Washita Prairie (Stop 1). (2)
entrenchment through Georgetown (Washita) rocks into the Edwards Limestone
(Fredericksburg) to Impress an incipient valley-pattern onto a Lampasas Cut Plain yet-to-be
(Stop 2). (3) valley deepening and widening through entrenchment and slope retreat through
the Comanche Peak Limestone (Fredericksburg) (Stop 3) prior to (4) the greater valley
widening in the less erosion-resistant Walnut Clay (not yet encountered. but to be seen as Stop
4) that will typify the Lampasas Cut Plain. The processes of entrenchment and valley-widening
that resulted in these varied landscapes were originally perceived as continuous. and the
evidence of their activity should be visible on the time-scale of a human lifetime. This
perception is questioned in the later stops.

Stops 4 through 17 are examples of processes and landforms In a Cut Plain that fails to fit
easily as the next logical step in the orderly. simple sequence outlined above. The reality of Cut
Plain landform leads to a series of questions and speculative answers indicating a far more
complex and far longer history than that suggested by our first "package" of stops. and brings
Into question the traditional evolutionary sequence previously described ..

The sequence of landforms and the tentative chronology of landscape evolution resulting from
that sequence describes the grand design of the Lampasas Cut Plain. A particularly interesting
peripheral aspect Is that a number of landscape elements. and the tentative landscape
chronology developed from the Lampasas Cut Plain. appear to be represented far beyond the
boundaries of the Cut Plain. and across the full width of the Cretaceous Prairies. Speculatively.
the major elements of that landscape and chronology may well exist in landscapes of much of
the southern Midcontinent. and even into the eastern ,Basin-and-Range of Texas and northern
Mexico (Amsbury and Hayward. 1996) ,

1-10
THE LAMPASAS CUT PLAIN
NOTE: The following guide to the Lampasas Cut Plain is an adaptation of an
earlier Geological Society of America Guidebook (Hayward, Alien, Amsbury,
1990), designed for a two-day trip across the full width of the Cut Plain. This
present guide for a one-day trip across the eastern half of the Cut Plain we
believe "covers" most of the points emphasized In that earlier trip. In addition,
In the "Conclusions" section of the present guide we attempt to bring you up to
date on accomplishments (mostly speculative) that postdate the earlier volume.
When you read these latest conclusions, you will readily see why we encourage
your observations, your questions, and even your enthusiastic participation In a
geomorphic project that has vastly outgrown our own capabilities and
calendars.

AN INTRODUCTION
At the turn of the century, Robert T Hili, the first to work extensively in this unique region, and
the one who gave it its name, best described its landscape:

"The Lampasas Cut Plain is the modified northern extension of the great Edwards
Plateau. It is a greatly dissected dip plain, now recognized by the general level of its
many remnant summits, which dominate all of the country south of the Brazos, between
the Western Cross Timbers and the Balcones Fault Zone. These summits, which are
called mountains by the inhabitants, are numerous remnantal circular flat-topped buttes
crowning the divides between the drainage valleys. Life and industry are mostly found in
the valleys scored below the summit level." (Hill, 1901, p. 78)

Hili, the first to study the Cut Plain, had as his principal interest the Cretaceous stratigraphy, but
he never lost sight of the land. Few since Hill have appreciated the landscape of the Cut Plain
more than did Hill, who attributed his interest in geology largely to that landscape. No one
since Hill has thought quite so much about the evolution of that landscape, though most of Hili's
comments on geomorphology were reserved for his newspaper columns in his later years.

Hili worked without adequate maps or aerial photographs, and in his broad reconnaissance, time
and Inclination did not permit detailed assessment of landscapes, processes, and rates of
change. His interpretations were largely intuitive, but it was an intuition guided by immense
knowledge and a lifetime in the field, the most productive years of which were spent in the Cut
Plain. Later authors (proctor, 1969; Mikels, 19n) have generally agreed with Hill's model of
landscape evolution, and have thus confined their interests largely to problems of process in the
Cut Plain. Hill's concept of Cut Plain evolution-and the one generally accepted-can be
summarized under five major conclusions.

First, the Cut Plain is an intermediate stage in the evolution of the Grand Prairies (Fig.
0-4)

Second, the shape of the Cut Plain is dictated by the contrast in physical properties
between Edwards Limestone and underlying and overlying formations (Fig. 0-5).

Third, the processes of slope retreat and landscape evolution in the formation of the
Cut Plain involve the infiltration of precipitation through the Edwards, its emergence as
seepage at margins of Edwards uplands, the undercutting of Edwards by spring sapping

1-11
~~ Lld:~ 11'1411.,,,

1-
-;:;.
~

~
'. US .. - - . " . .
::;:. '->
. Evant '"----

Figure 0-1

ROUTE MAP
LAMPASAS CUT PLAIN
Central Texas

1996 ReId Trtp


Friends of the Pleistocene
South Central Cell
j
o 5 10 I 15 20

I.
Mil ••
Iii
«
w uu.uu !m"il'utl

'"


!
...i
w .~
...J
u:

'i
g

.;,
0
0
a:
Do

1=01=
«~(J
::E w
z
0
,.
"
c
t;
!/
,
0
'.
.'

i
''""
Ii:
::E Ul
«
a: ~
Cl
« 3
2i
I/II(UII /JII;U"

I-
Ul
W
:;:

1-13
>-
Ul
«
w

'"
,

J
'"

i/r
'I~I~"I'O

... J ~ ..
i •.
og~S
i~;s
-.
j: U J
""r
,~; ~:!
!:;iH~ ~ I
;;o:!1
ii tl *
i J
... _.
--~,.
E -;
IJ
e

D.
o
Iii 0.

W
..J
•!
Ii: •
I
0
a:
D.

~ !:!
>-01=
do «~(J
Z
0
~
,.""
-.
3-
.<

;o~
••
"1
!:
..
0
IE: :;; W u - <
:;; II) ::f ""

..
a:
(!J

Q g
<
~
,
.~ ;
i~
~
1

1-14
IV! tp/rn

D

Miles

--,
u.

_ Edwlrda·damlnl,ld Iind.clp ..

_ Antler.-domlnlted Iind.cap ..

Figure 0-4 The Lampasas Cut Plain lies between the Brazos and Colorado Rivers, and Is drained largety by the Brazos River
and its tributaries. Cut Plain landscape, dependent on the geomorphic effect of Edwards Limestone, is confined to the
outcrop belt of that formation. Similar Edwards-dominated landscapes include those of the Callahan Divide, and of the
Edwards Plateau (for which the Brady Mountains mark the northern margin). Other ancient Edwards-dominated landscapes
are buried beneath the Southern High Plains (from Hayward and Allen, 1987, 1988).
and subsequent failure by slumping, leading to almost continuous slope retreat and
valley widening (Fig. 0-5).

Fourth, the Cut Plain has migrated eastward through time, as the Grand Prairie
landscape moved slowly down dip from some distant westward margin of Edwards
deposition (Fig. 0-6). Northward, beyond the margin of Edwards reef-like deposition, the
Cut Plain merges with the Fort Worth Prairie (Hill, 1901, p. 77), a gentler, more rounded
landscape, but with numerous reminders of Cut Plain physiography

Fifth, stages in Cut Plain evolution include (a) initiation of dissection onto the Edwards
Limestone, (b) Initial entrenchment through the Edwards Limestone, (c) rapid
entrenchment and valley widening through the Comanche Peak Limestone and into the
Walnut Clay to create the typical profile of narrow divides and broad valleys, (d) the
ultimate stripping of Fredericksburg rocks and the consequence disappearance of the
Cut Plain as the Grand Prairie retreats eastward

These original conclusions have survived, largely Intact, again chiefly because they are intuitively
acceptable to one seeing the Cut Plain for the first time, or seeing Cut Plain evolution as an
aside in the progress of other work.

On this trip we will examine each of the widely accepted concepts of Cut Plain evolution to see
how they hold up to a more analytical view, with emphasis on landscape rather than
stratigraphy. More recent interpretations are based on observations and data unavailable to
earlier workers, or unnoticed because to casual view they are concealed by the far more obvious
elements of the landscape.

Toward that end, this trip is presented as a series of vignettes which together form a panorama
of episodic Cut Plain evolution. At each stop, the panorama widens, and at the end of the trip
there will be a summing-up describing what has been seen, and what It appears to mean in
terms of landscape evolution.

The trip is designed to emphasize five points; (1) the geological foundations of the Cut Plain;
(2) some details of modern processes of stream behavior and underground water, that must
also have contributed to the formation of the Cut Plain landscape; but that in significant ways
fail to explain all, (3) the formation of the Cut Plain from an earlier regional Washita Prairie
landscape; (4) a marked change in landscape formation midway in the orderly history of the Cut
Plain; and (5) a proposed sequence of stages, and a very tentative chronology for that
sequence In Cut Plain evolution, based on field relationships.

GEOLOGICAL FOUNDATIONS
perhaps nowhere in Texas are stratigraphy, structure, and landform more closely Interrelated
than they are In the Cut Plain (Fig. 0-2, 0-6, 0-9). So obvious are these relationships that for
many years they worked as an impediment to real understanding of Cut Plain evolution.
Appearances suggest that each landform element reflects absolutely the simple stratigraphy
and the near-horizontal structural attitude of specific stratigraphic units so that no other
explanation is needed or justified. Evidence presented on this trip Indicates a far more involved
history. To understand the Cut Plain we must first understand the general Interrelationships of
structure, stratigraphy and landscape.

Structure and landform


Geologic structure in the Lampasas Cut Plain is principally that of the Trinity Shelf, the most
stable part of the Central Texas Platform (Parrish, 1995; ; Hayward, 1988; Lemons, 1987).

1-16
WEST EAST
Cui Plain mlgrales E, down dip, as scarps Washita Prairie mlgrales E, down dip, as
rei real from 1 10 2 10 3. CuI Plain scarps relreal ~1 10 2 10 3. Washita Prairie
~/ \3 ........--- ~ _/_ 1. ____ _
55,!
---
1 ./ 2/r -- S.... I Ph - 1/_-;- 2,,_ 3 _/
'" (I
... '-- -' _/ c!?:r:
-
/ ,/
/' -' j ::
-' -'
-'
~
-.:.-- ~

. ,.: -.,~::'~ ' " Glan Ros,;cPralrla replaces CuI Plain as


::,::,,::,::.\~FredarICkSbUrg rocks ara slrlpped away RegIonal dip?

r weSlern Cross Timbers replaces Glen Rosa


Pralrla as landscape mlgrales easl-
ward, downdlp.
An exhumed aub-Crelaceous landscape I~
exposed 8S Crelaceous rocks ere slrlpped
FIgure 0-5 The mechanism and effect of the •... slow eastward migration of the Cut
Plain... of the earlier, traditional view. According to this model, the Cut Plain migrates
H

eastward, down dip, as erosion lowers the landscape. An observer (of enormous
patience) at any point on the eastern margin of the Washita Prairie would see a
away succession of landscapes pass in review, first the Washita Prairie, then the Lampasas Cut
Plain, the Glen Rose Prairie, the Western Cross Timbers, and eventually landscapes in
,
~ Paleozoic rocks (from Hayward , Allen, AmsbUlY, /990).
:::; Georgelown- L's

Edwards

Comanche Peak

Glen La ledges form benched valley


floors

FIgure 0-6 The generally accepted model of Cut Plain evolution, Note the close
correlation of stratigraphy and landform. In this model, the landscape is graded to
present drainage, slopes retreat by failure of the Edwards Limestone as seepage erosion
undermines this massive caprock, the divide-to-basin slope break occurs at the
Comanche Peak/Walnut contact, and slope retreat is a continuous process (from (from
Hayward, Allen, Amsbury, /990).
Near its eastern margin, the Cut Plain shares structural characteristics of the far western margin
of the East Texas Basin (Fig. 0-2, 0-7, 0-9).

From the western margin of the Cut Plain to the eastern divide of the Leon River Basin (our Stop
7, a distance of more than fifty miles, the dip increases from ten fUmi to twenty five fUmi.
Recent studies indicate that even in the areas of gentle dip there are small, step-like descents in
the regional near-homoclinal dip. Detailed studies of small areas also show minor crenulations in
structural contours (Amsbury, 1995; Parrish, 1995).

A major increase in dip from twenty five fUmi to fifty fUmi occurs in the transition zone
between Trinity Shelf and East Texas Basin, over a distance of less than twenty miles near the
eastern margin of the Cut Plain. The disappearance of the Cut Plain to the east is a product of
this increase in dip, as the Fredericksburg plunges beneath younger Comanchean and Gulfian
sections (Fig. 0-8).

The Cut Plain province is almost free of visible surface faulting, but there are several "belts' and
regions in which lineaments are conspicuous on Landsat imagery, aerial photos, and on drainage
patterns on topographic maps. Lineament zones appear to have been localized by older
structure in Paleozoic rocks, possibly reactivated by subsidence of the East Texas Basin. These
appear to have had significant effect on the orientation and position of major drainages and Cut
Plain divides, while having almost no influence on minor tributaries. Other lineaments, more
distant from the known fault zones but equally localized, show little relationship to recognized
structure or existing landform (Parrish, 1995).

Because the gentle eastward regional dip is essentially' equal to the mean eastward gradient of
the Cut Plain (measured from divide elevations), the Cut Plain is both very broad, and

remarkably uniform in landscape over the whole area of gentle dip. Near the east margin, where
the dip Increases, the Cut Plain disappears to be replaced by landscapes developed in younger
rocks as the dip continues to steepen into the East Texas Basin (Fig. 0-8), In succession, the
Washita Prairie, Eastern Cross Timbers, Black Prairie. and East Texas Timber Belt (Fig. A-3).

Stratigraphy and landform


The Grand Prairie is developed in rocks of the Trinity, Fredericksburg and Washita Groups (Fig.
A-3, A-4, 0-2, 0-9) This Comanchean section thins westward through depositional thinning and
erosional truncation, from about 1160 feet thick in Central Hill County where it disappears
beneath Gulfian rocks to 300 feet thick near Goldthwaite at the western margin of the Grand
Prairie (Fig. 0-8) (Lemons, 1987, p. 37).

The Trinity Group: Over most of the area of the Cut Plain, the Trinity Group (lowermost
Comanchean) consists of two formations, (1) the Twin Mountains Sand, and (2) the Glen Rose
Limestone. Along the northern and western margins of the Cut Plain where the Glen Rose
formation is missing through pinchout, Trinity rocks are represented by the Antlers Sand, of
combined Trinity and lower Fredericksburg deposition (Fig. 0-9). The outcrop belt of Trinity
rocks is represented by two geomorphic provinces; (1) the Western Cross Timbers, and (2) the
Glen Rose Prairie (Figs. 0-2, 0-3, 0-6, 0-9).

The Western Cross Timbers (Fig. A-4, 0-2, 0-4), is the broad belt of woodland and savannah, of
low relief, that is veneered in sandy soil derived from sands and sandy muds of the Twin
Mountains Sand or the Antlers Sand. Where the Glen Rose Limestone is present in the section
the Western Cross Timbers is separated from the Lampasas Cut Plain by a belt of Glen Rose
Prairie, developed on Glen Rose Limestone (Fig. 0-2, 0-9). Along the north and west margins of
the Cut Plain, where the Glen Rose is absent through pinchout, the Western Cross Timbers
adjoins the Cut Plain (Fig 0-8, 0-9).

1-18
The Glen Rose Prairie is generally included in the Lampasas Cut Plain, where it forms a
conspicuous dissected lowland landscape of prairie and savannah within the upper basins of
larger southeastward-flowing streams. These inner valleys of the northern and western Cut Plain
are cut into Glen Rose rocks, and veneered in shallow stony soil formed from the alternating
dense limestones, chalky limestones, and marls of the Glen Rose Formation.

The Fredericksburg Group: The Cut Plain is the remarkably conspicuous geomorphic
expression of the Fredericksburg Group, consisting of four formations, (1) Paluxy Sand, (2)
Walnut Clay, (3) Comanche Peak Limestone, and (4) Edwards Limestone (Fig 0-9). Buttes and
divides are capped by the thin but resistant Edwards Limestone. Upper slopes are formed in
Comanche Peak Limestone. Upper valley floors are developed in Walnut Clay-and in the
northern Cut Plain, in Paluxy Sand. Inner valley entrenchments are formed in Glen Rose
Limestone-and to the margins of the Cut Plain, in Twin Mountains sand. The lower three
formations of the Group thin to the west, so Cut Plain total relief also diminishes westward.
Larger areas of the westernmost basins are in Glen Rose and older formations, yet the Cut Plain
landforms are highly distinctive all the way to the western margin of the province (Figs. 0-2a, 0-
2b).

Washita Group: In the area of the Grand Prairies, the Washita Group, youngest of the
Comanchean Series, consists of the Georgetown Limestone and the Grayson Marl (Del Rio Clay)
(Fig. 0-9). The Washita Prairie, the geomorphic expression of the outcrop of the Washita
Group, is therefore subdivided into the Georgetown and Del Rio Prairies. Only the Georgetown
Prairie is significant to our study of the Cut Plain, since the Del Rio Prairie is largely confined to
the Eastern Washita Prairie, and therefore distant from the eastern margin of the Lampasas Cut
Plain. The Georgetown Prairie, formed on the alternating chalky limestones and marls of the
Georgetown Limestone, is typically gently rolling prairie, cut by shallow by angular, rock-walled
streamways. It was originally almost treeless prairie of generally thin and stony soils, capping
the divide areas of the eastern Cut Plain, and forming the foundation of the Washita Prairie
geomorphic province that bounds it to the east (Fig. 0-2)

1-19
Wlchlta-Arbuckle-Ouachlta Trend

Zone of gentle subsidence. Bay floor


frequently exposed, shoreline rapidly
~~
oscillating.
/'-'----
/
/1'. _-
/ S Fort Worth
r!
~'f'

Shelf st~ble and mostly '/.,~<c I generally submerged .


non-manne. / i/'«;I / /
I ~
Gj
r--t'J'-;I---,--;\-'L._.L----, ~ I '"
East
Brownwood 'lit

Llano "Island"
or "peninsula"
Miles
f 50

Figure 0-7 Map of the Trinity Sh elf at the time of deposition of Trinity rocks. The Shelf lies
to the west of the dashed line, the western margin of the East Texas Basin lies to the east. The
western margin of Trinity marine deposition Is deeply crenulate, a product of the rldge-and-
valley topography on the sub-Cretaceous surface at the time of Trinity deposition (From
Hayward, 1988).

1-20
WEST EAST

High Plains Callahan Divide Grand Prairie Black Prairie

--------
\
Gentle east dip of Trinity Shell, 2 to ~
fUmi. Dip increasing from 10 to 100
ftlmi into western margin East Texas
Basin

Callahan Divide·llke landscape buried.


beneath High Plains deposits

Figure O-S Diagrammatic profile and section, showing the gentle regional dIp and regional
thickening of Comanchean section, from the High Plains, along the Callahan Divide, across the
Cut Plain, and into the East Texas Basin. There are questIons about the correct depositional
model for Edwards Umestone (one would result in a very gentle southwestward initial dIp, the
other an essentially horizontal initial dip), but this disagreement would not significantly change
this picture. In either case, the gentle eastward dIp is not explained by deposition, but is a
product of post-Edwards structural warp (from Hayward, Allen, Amsbury, 1990),

w E

<
~

.,
i
<
~

EDWARDS .,• "

COMANCHE PEAK
""'II:
2",
Z
w:> <
ffiOl w
::
~, :~ ...'" u
z
....:.;;::-:;; , - t - P A L U X Y - ' - ' - ? _ •.. <
~
.?,::'. ::. U. GLEN AOSE 0
u
~NTLEA8-- _,OLE.N ROlE THORP PRINOS
::;-- =="'- -,-, >-
:~:". • OLEN ROSE i
',:' . TWIN MOUNT AINI
ii:
~

...;::.. .- .. .. .....
-' III!NH-TAIAIIIC ,

FIgure O-g Comanchean stratIgraphic nomenclature on the Trinity Shelf. The landscape of the
Cut Plain Is developed In rocks of the Fredericksburg Group. Rocks of the Washita Group are
confined to uplands in the eastern Cut Plain, and to the Washita Prairie, which bounds the Cut
Plain on the East. Trinity rocks are exposed in the inner valleys of the central and western Cut
Plain. Beyond the pinchout of the Glen Rose Limestone is the physiographic province of the
Western Cross Timbers, developed on Antlers Sand (Fig, 0-4) (from Hayward, 1988),

1-21
STAGES AND PROCESS IN EARLY CUT PLAIN FORMATION

STOP 1--BRAZOS-BOSQUE DIVIDE--The "Time Zero" Surface (31 0 41' 44" Nj


970 23'19" Wj Valley Mills Quadrangle, 1/24,000, 1979) (Fig. 1-1, 0-1, 0-2)

Along our line of traverse from the Brazos River to this point are occasional exposures of mid·
Cretaceous rocks (Georgetown Limestone, Washita Group (Fig. 0-9) in small roadcuts along the
highway, and as rubble in fields and pastures to the right and left of the road. Near the Brazos
River the Washita Prairie is seen in a degree of dissection encountered only near major streams.
Terrace veneers, common near the Brazos River, have been absent from the highlands the last
few miles to Stop #1.

Here on the divide between the Brazos and Bosque rivers (Fig. 1-2), where the gentle
landscape is typical of much of the Washita Prairie highest divide-lands, terrace lags are again
evident, but these are different from the better· represented terrace veneers along the Brazos.
Here on the highest divide is a record of river history which antedates the highest terraces of
the Brazos (see Fig 17-3), and which also antedates the Lampasas Cut Plain, the principal
emphasis of this trip. The landscape upon which these gravels were initially deposited by
streams unrelated to modern drainage has been called the "Time Zero Surface" (Montgomery,
1986) of Cut Plain evolution. It was the surface immediately antecedent to major incision into
Fredericksburg Rocks.

This is an important point.

The history of the Cut Plain clearly began before there was a Cut Plain. Drainage nets,
established on Washita Prairie (such as we see here) or even earlier landscapes, were the first to
be incised into Fredericksburg rocks. With that incision, the Cut Plain first became a distinctive
landscape. About eight miles north-northwest of this locality is the headwaters of Childress
Creek, which flows southeastward to join the Brazos just north of Waco (Fig. 1-2). its basin is
almost entirely in Georgetown rocks, while in its lower reaches the channel floor is just now
beginning to cut into Edwards Limestone, the first stage in formation of a Cut Plain landscape.

Of critical importance to the history of the Cut Plain is that major tributaries such as Childress
Creek tend to be dip-controlled, and to elongate rapidly up dip by processes which involve both
groundwater and overland flow, a point to be discussed later. Obsequent tributaries tend to be
short, giving larger basins an obvious asymmetry. Trunk streams tend to migrate down dip on
rock surfaces, and to a lesser degree, to be driven down dip by clastic input from dip-controlled,
steep-gradient tributaries. Valley lengthening by streams is thus accomplished both by major
headward migration of tributaries, and by lesser, lateral (generally down-dip) migration of trunk
streams.

The Brazos River is the major trunk stream toward which flow almost all Lampasas Cut Plain
steams In the region of today's trip (Fig. 0-4). Prior to Cut Plain incision, drainage was radically
different in both organization and bedload from present systems. But by the beginning of Cut
Plain inciSion, an ancestral Brazos In about the same pathway as the present river,

1-22
))

Figure 1·1 (Valley Mills Quadrangle, 1124,000; 1979) Location of Stop 1 on the high divide
between the Brazos and North Bosque rivers. This locality is in the Washita Prairie, of thin rocky
soils, and mixed grassland and savanna. It is on the outcrop belt of Georgetown Limestone, and
typical of many high divide localities in the Washita Prairie, the soil surface is covered by a thin
veneer of alien lag gravels, coarse, vel}' well rounded, siliceous, apparently remnant from the
"Time Zero Surface", an ancient surface from which was formed the landscape that now
surrounds us. Streamways we see present on this Washita Prairie surface will be impressed into
the rocks that will ultimately form the Lampasas Cut Plain, the next stage in the slowly-changing
landscape. of the Grand Prairie.

1-23
':Z.
) \. Washita ~
'1 ,/,
1,
'\.
t.... -.. . . .
STOP 1.... \.- . '. '
1
'.
)...
I"
-"'\ \
'I." \
Prairie Whitney

\ \ ·.... ,l'I.-______-tQuartzose pea gravei in auger


'to ~,,\ .\ '1 ¢. hoies all aiong divide
.,. '\ ~\ "~.

.....
,p,'
-':,

,;:,~,
I \
\'-\"
I,'
:1 _' ...
'0:
"' ........ &
; ·--_.'-.c:;"':.·"
\ ~ :~ "\ '. -'''''~~''
~" ...... '\'-..

Eastern margin \
\ '-'-"".~ .".
c;;.,
' ..... \ . , e""p.f.
Cut Plain <1>' -,,,,::::_, -, 7'-:."!I...._....j
.....ltI.'.. ' . .::::'-.1".
. ,- ~-
0- 5
Valley Miffs .,
- I
miles
I

Figure 1-2 Basin of Childress Creek in relation to the Brazos and Bosque Rivers, and the
eastern margin of the Cut Plain. Childress Creek flows generally down dip into the Brazos River.
Where the stream is on the Edwards Limestone, gradient and dip are equal. Exotic, quartzose
gravels on the divides are alien to Central Texas, and mark an earlier episode in landscape--the
"Time Zero" surface of Cut Plain evolution. The eastern margin of the Cut Plain, here marked by
the eastern divide of the North Bosque Basin, is everywhere abrupt. There is no transitional
landform (from Havward. Allen. Amsburv. 19901
WEST EAST

~"' "' "'


--.:..
Gi

-.
> Gi
e c
a; >

";;.,
'tI
ii:
=
CD

-."'.
a. CD

-.
a.
-
u
u a;
is
. ."' a. ... u
.'"
"'. -
c u
.c .! 0

.-
0
a. :c
.!!
a;
u. --------------==== 3 .
"'
.!!
:c
;::
!!

'7Z-- ---------..- == u
= _ ~-- c-
- .
-------~-----,

EAST TEXAS BASIN

1-24
apparently flowed at a level from two to four hundred feet higher than at the present time, on
Late Comanchean and Gulfian rocks which have since been partially stripped by erosion.

Initial Cut Plain dissection was by streams similar to the North Bosque River and Childress Creek
(Fig. 1-2), inherited from an earlier Washita Prairie landscape. Thus the factors controlling
drainage evolution on the Washita Prairie were important to the beginning Cut Plain
geomorphology. The surface upon which ancestral Washita Prairie drainage was established was
apparently this "Time Zero" surface, a remnant of which is yet present here at Stop 1. In
addition to the exotic gravel lag seen here, several shallow auger holes along the northern divide
of Childress Creek Basin have encountered quartzose pea-gravel lags in stony, caliche soils (Fig.
1-2), always at elevations far above the highest dated terrace of the Brazos. Similar quartzose
gravels have been described from major Cut Plain divides to the west and southwest (Smith,
1984); on the crest of the Callahan Divide between Colorado and Brazos headwaters drainages
(Byrd, 1971; Knapp, 1990); and eastward on high divides in the Blacklands and East Texas
Timber Belt (Knapp, 1990; Amsbury, 1990; Spencer, 1991) (Fig. 1-4).

Sampling has been too limited to establish clearly the topography of the original Time Zero
Surface (or surfaces) on which these fluvial sediments were deposited. The high gravels show
considerable diversity in size and petrology, and occur at various elevations above modern
drainage. The Time Zero Surface may have been an alluvial plain; but if so, there were also
favored trunk routes across that plain (Spencer, 1990), and each of these had local tributary
networks, reflected in differing ratios between locally-derived and exotic gravels. West-to-east
drainage was accomplished by a few larger streams. These originated far west, beyond the
present High Plains, and were served by local tributary networks that supplied abundant
Fredericksburg cherts, and farther east, abundant Woodbine orthoquartzites (Flowers, 1987;
Tharp, 1988; Montgomery, 1986). This surface, whatever its configuration, was the "Time
Zero" surface of our discussion--the landscape that immediately preceded the Lampasas Cut
Plain.

The Cut Plain was formed after reorganization of drainage originally on this ancient surface.
Drainage reorganization was In response to southward structural tilt during Tertiary episodes of
subsidence in the East Texas Basin and its northern borderlands, or to other factors as yet
unrecognized. The pattern of reorganized drainage, somewhat askew with present regional dip,
may in part be a key to older drainage yet visible in the Cut Plain.

The age of these highest gravel lags of Stop 1 is uncertain, but they are clearly older than the
highest terraces of the Brazos, and they reflect a time when the pattern of Central Texas
drainage was radically different than it is today. Attempts to date this beginning of Cut Plain
evolution are tenuous at best, and call for the correlation of these gravels with similar gravels in
other areas, but the problem is that we do not know whether we deal with one gravel or many.
There are gravels atop the Callahan Divide far to the west, and to the east detrital gravels on
high divides in the western margin of the East Texas Basin (Fig. 1-3). Both are quartzose,
highly mature with similar but undistinctive cobble petrology, and both have ferruginate
matrices. There the correlation ends.

The high gravels of the Callahan Divide are clearly much older than oldest Ogallala, for they are
encountered more than 200 feet above the feather edge of Ogallala deposits where the eastern
margin of the High Plains onlaps the toe of the Divide, at Roscoe (Nolan County).

When the Callahan gravels were deposited, the Callahan Divide, then the lowest part of the
landscape, and the Time Zero surface, was the site of deposition of coarse fluvial gravels from
far to the west (probably from the Southern Rocky Mountains). The present gradient from the

1-25
Ancestral Washita Prairie or alluvial plain
Time Zero Surface

NORTH
SOUTH
Time Zero Surface
_____ ?~__~Qaa9~O~9"Q~O~_\~_______ r'-------______~S?Q~~C?~a~e~CM~ ___
?-
Georgetown

Edwards I

Comanche Peak ;4 \ "

Figure 1·4 Block diagram and section across an ancestral Washita Prairie, showing prominent
west-east drainage avenues. The profile shows an almost-undissected surface, and the
suggestion of an alluvial veneer. Whether the streams of the pre-Cut Plain era flowed in gentle
valleys cut into the Time Zero surface, or on an alluvial plain is not yet determined. Stop 1 is on
a remnant of the Time Zero Surface, the landscape from which the Lampasas Cut Plain would
ultimately be formed (from Hayward, Allen, Amsbury, 1990).

1-26
Callahan Divide deposits to the Central Texas deposits is about 10 ft.lmile., a gradient
somewhat greater than that of the present High Plains surface ( ± 8 feeV mi.).

In East Texas no rocks of appropriate composition and texture, and predating Ogallala deposits
(late Miocene/Pliocene) of the High Plains have been recognized. Additionally, there is
ccnsiderable question whether the high gravels of the Cut Plain are truly correlative with those
of the Calahan Divide, though their elevations seem to be about right (Fig. 1-3). A pre-Late
Miocene, dissected "Calahan Divide" (or perhaps even mature Cut Plain) appears to be
preserved beneath the Ogallala Formation along the eastern margin of the High Plains (Cronin,
1969; Seni, 1980; Gustavson and Finley, 1985 Fallin, 1989)). suggesting that a mature Cut
Plain in that location is at least as old as mid-to-Iate Miocene.

To clarify a point that may be thoroughly ccnfusing by this time, you must remember that the
Ogallala of the Southern High Plains is an enormous fan-like deposit. It pinches out where it
onlaps the lower slopes of the Calahan Divide near Roscoe. At that point the Calahan divide
rises 200 feet above the surface of the High Plains. From that point the Ogallala thickens to
the north and west, and at Fluvana (about 50 miles northwest of Roscoe) it has completely
buried a Callahan Divide-like landscape, now exhumed by retreat of the Caprock Escarpment, the
present eastern margin of the High Plains.

In summary, the Time Zero surface is of uncertain age, though clearly older than late Miocene.
By late Miocene there existed a Cut Plain-like landscape In the area now occupied by the eastern
High Plains. It was later buried by the late Miocene/Pliocene Ogallala deposits. The Calahan
Divide jOins the Cut Plain north of Brownwood (Brown County) where marked similarities clearly
Indicate that at that point both are one, and both participated in the same pedimentation
episode that led to the formation of a pediplain (King, 1962. p. 161). the one that we have
called the Comanche Pediplain ..

Visualize this pediplain not as an areally extensive surface, but as a complex of local in-basin
pediments, each individual pediment confined to a single drainage basin, but all together forming
a stem-and-ieaf-like Joined-pediment surface. Each individual basin pediment was graded to a
Junction with the next higher-order stream. The common base level for all individual pediments
was the major trunk stream . At the time of pedimentation the ccmplex surface thus formed
was the Comanche Pediplain, at a base level then much above than that of today's drainage
network.

EARLY EVOLUTION
INTRODUCTION TO CUT PLAIN DEVELOPMENT

Stops 2, 3, and 4 (Fig. 2-1)


The next three stops (Stops 2. 3, and 4) illustrate stages in the earliest development of the
Cut Plain. Assuming non-competitive systems with uniform climate. lithologies and available
relief. such a sequence appears reasonable (Faulkner. 1974).

Two orders of major drainage are apparent in the Cut Plain.: (1) the main trunk streams. such as
the Brazos. Leon. Bosque, and (2) iower order tributaries to this main system. such as the
Middle Bosque River. Hog Creek. and Coryell Creek. among others. Streams of the first category
In the Cut Plain flow at angles to the regional dip (Fig. 1-4). These streams presumably
established their ccurses long before Cut Plain dissection began.

The streams of the second category align along the dip of the Edwards Limestone (Fig. 2-2).
Their patterns may have been determined by sapping. i.e. groundwater control of drainage (Fig.
2-3). A combination of weakening by solution. followed by removal during overland flow is

1-27
believed responsible (Bunting, 1961): (1) stratal dip and regional groundwater flow patterns
parallel stream flow (Yelderman, 1987); (2) the dominant direction of drainage dissection seen
on maps and in the field is up-dip and up-gradient; and (3) amphitheater-shaped areas with
associated seep zones are present at the headwaters of many streams (Fig. 2-3).

Valley cross-profiles
In the basins of Hog Creek (Stop 2) and the Middle Bosque River (Stop 3) the major
characteristic of the landscape is the dominance of divide areas over valley areas. Most of the
land is in interfluve on Washita Prairie. Only in the immediate vicinity of stream courses is the
deeply dissected, Lampasas Cut Plain landscape dominant. Thus, for these stream basins the
principal landscape is that of the Washita Prairie, in marked contrast to much of what we will see
in the remainder of the Cut Plain (Brown, 1988)

At Stops 2 and 3, valley width and side slope formation are clearly related to stream
downcutting. But at Stop 4 (Neils Creek), the side slopes have retreated far from the stream,
and the vastly greater distance between the Edwards capped hillslopes is apparent. This marks
a major break with the simple model suggested at the beginning of this guidebook.

Past theories of Cut Plain evolution explain valley widening solely by lithologic control; that is,
once the resistant Edwards Limestone is removed, the much-less resistant underlying rocks are
quickly eroded to form broad valleys. But the transition from narrow valleys (Stops 2 and 3) to
wide valleys (Stop 4) is everywhere abrupt. There are no intermediate examples. We believe
that the present landscape of the Cut Plain was formed in part, if not largely, during a major
period of side-slope retreat by pedimentation along streams that previously had cut through the
Edwards Limestone. Evidence for this conclusion will be presented throughout the field trip;
some evidence can be seen within the Cut Plain and evaluated in the field by the participants,
but other evidence comes from surrounding regions and can best be presented during
discussions.

A plot of valley width versus drainage areas (Fig. 2-4) shows two major populations of Cut Plain
streams. "Type I" streams were still flowing on the Georgetown formation (Washita Prairie)
during the period of slope retreat.. "Type II" are those streams which had breached the
Edwards prior to the period of slope retreat. Pedimentatlon in the valleys of Type I streams
would therefore be less apparent, owing to the behavior of the more easily eroded Georgetown
Formation.

A second comparison of basin parameters, valley width versus width-to-depth ratios of main
stem channel incision (Fig. 2-5) indicates: (I) a remarkable correlation of slope retreat with
degree of incision and (2) that the two groups (I and II, identified in the previous figure) are stili
apparent. These relationships (Brown, 1988) support the hypothesis that Cut Plain drainage
and landscape modification Is not simply the effect of the work of streams on lithologies of
varying resistance. The evolution of the Cut Plain is perhaps more appropriately tied to three
major factors: (I) major controls on local stream base levels that Influenced the degree of
Incision and stream extension, (2) the comparative level of incision of these streams with
respect to the Edwards Limestone during a period of equilibrium in stream grades and major
valley side-slope retreat and pediplaination, and (3) groundwater flow direction and volume, and
its Influence on stream incision and slope retreat (Fig. 2-3).

Phases of Cut Plain Evolution


The substitution of spatial variations for temporal ones has been a popular geomorphic strategy
to solve problems in the long term evolution of landforms. Patterns of landform evolution have
been inferred either directly or Indirectly on the assumption that time and space are
interchangeable under certain circumstances (Knighton, 1984).

1-28
V' \ v:

'I /~) [

I

/",,- . I ( ('" )

\.~ I 1(. K~~:;.


Iw·mdm.1I.
If
! l/ •,
J~. ""
/ ) L
,/
/-<
/'
1'-

;vl
';".
~
,I
/(
,.,-.

'.'1' l.
\
865>:.
\\

(
,,

,

\ ".
'.
"')\'
. .':
Figure 2-1 (Coryell [1955] Quadrangle, 1124,000) Stops 2 and 3, Hog Creek and the Middle Bosque River, two stages of ,a'
-;,
incision. antecedent to development of the Cut Plain. While these localities are in the Washita Prairie, the streams are
entrenched through the Edwards Limestone, These are progressively-later examples of Washita Prairie incision than that
shown at Stop 1.. While, locally, stream valleys have taken on some of the attributes of the Cut Plain, the regional landscape
",
I
N' )
I

I
is still Washita Prairie. and will remain so for a very long time. However, stream networks inherited from the stage of \ ) \'
Childress Creek, here amplified, are the first to be incised into the forming Cut Plain. Notice that while these two localities j'
\ apparently show significant dissection of the region, the Washita Prairie is still the dominant landscape, encompassing all but ~l I'
tile immediate areas of the valleys. '!P. /
\ \ ~ /-----~I.. /1' \/ !~!"rill / ) r;(,Jtrr '~\ --- ~ ~?~ I I
) \ I ( ",<> II
- •.
~
.~
$I,;. ~
Strike of Edwards

I
...~
f ... - .• -
., . I I
-"0 -"- ..
.,.~ _.' ,. \I
T__ I t-
-"" I
-;.~"
1/ I
l-
I
\

-··-r I
Major streams at angle to dip. Tributaries
flow downdip, probably controlled by
groundwater flow.

Figure 2·2 Comparison of directions of flow of trunk and tributary streams of the Lampasas
Cut Plain. Notice that while the tributaries flow in dip·directed valleys, trunk streams flow in
directions at significant angles to the regional dip. Tributary orientation is believed to be largely
a product of dip-contrOlled, groundwater flow (Fig. 2-3). Trunk stream direction was
apparently inherited from consequent flow at the time of establishment of modern drainage
(from Hayward, Allen, Amsbury, 1990).
Flow Paths Ordered reentrants mark spring exits
/11
I
I
'I I I I
I
I '
I I I" I
,
I
I ,
I,
I I
I I
I T I
I I '
I I
I 'I ;I I
I
I
I
I
I II I I I I '
I I I
, ' I I Time 1
I I I I I : II I : : II I : " Time 2
l"~! I I I I ' ,If

..\. '-. .t.ol.:;'


II

Edwards Free Fae


.lLU.
I I I,
.LU. ,t;-*."LtLl
Downdip flow perturbed by
permeability variations

- .... ~
Valley elongation by spring action

Figure 2·3 Valley entrenchment by spring sapping. Groundwater flow is generally down dip
(Time 1). However, locally subsurtace flow is perturbed by permeability differences. On
margins of divides, this creates a series of orderly minor reentrants along exposed Edwards
faces, with local more prominent, permeability-favored paths (Time 2). These favored pathways
divert flow, leading to slow headward (up dip) erosion along selected routes, ultimately
determining, at least in part, the pattern of later streams (Time 3) (from Hayward, Allen,
Amsbury, 1990)
..•••••
E
...
30 /~
.•-
~.----- ........ ........ , I

",
/ \
S I
I
. ,
!! \

-
c
c I z I
20
)(

-
:::.
/ :;
1/
I ~
/
.c: /'
\
.
0
'Ci 10 'i U //
0
~ \ , • /' ..."
..
G

>-
'--- ~--- -'" =. :
.!
OJ
> a 50 100
tI
o " >
:z:.m-~
,:;-..
150
. . ")
200
Basin Area (mi 2)

Figure 2-4 Plot of valley width versus basin area shows two families of Cut Plain valleys,
Types I and II. While Type I and II valleys have similar basin areas, they have radically different
valley widths, suggesting that basin area was not a significant factor in determining valley type
(from Brown. 1988).

300 I
-:::.
--- --- ---
u. Lampasas
Coryell
.c:
..
ii.
'D
200
--+"
Owl
• Neils

>-
.! I
OJ Valley wldlh
> I
r
vary~:]:?:6
100 Cave
M. Basque
Bluff deplh
Hog

a 10 20 30 40
Valley widlh (II X 1000)

Figure 2-5 Plot of Cut Plain valley width versus depth. Again valleys of both Type I and Type
II are clearly evident, and again they form two distinct populatlons--Type I with narrow valleys,
and Type II with wide valleys. While all the valleys are of the same general age, Type II valleys
went through a phase of valley widening in which Type I valleys did not participate (from Brown,
1988).

1-31
Following this logic, and based on flume studies and analysis of topographic maps (Glock, 1931;
Parker, 1976), stream networks are believed to go through several stages of development:

(1) initiation (of a skeletal pattern)


(2) elongation (by headwater tributary growth)
(3) elaboration (through addition of small tributaries)
(4) abstraction (as tributaries and relief are lost throughout time).

This process of drainage development, which apparently describes much of the reduction of
Edwards-capped uplands during the valley-widening phase of Cut Plain evolution, is also the
present sequence in which modern dissection is destroying the Comanche Pediplain (Brown,
1988).

STOP 2--HOG CREEK--The Cut Plain Is Born--(31 0 37' 22" N; 97 0 30' 44" W-
(Coryell [19551, and Mosheim [19791, quadrangles, 1124,000) (Fig. 2-1)

Stop 2, on Hog Creek (Fig. 2-1) illustrates a stage in which the area in Washita Prairie upland is
far greater than the area in entrenched valley. But throughout much of its upper basin, Hog
Creek is now incised through the Edwards and into the Comanche Peak Umestone, in a drainage
pattern inherited from the Washita Prairie stage of landscape evolution.

Recall that the earliest drainage network to be incised into the Edwards Umestone to form the
Cut Plain was that which already existed on the Washita Prairie. Once Washita Prairie drainage
had entrenched"into the Edwards, the stage was set for the formation of a Cut Plain, though the
ultimate mesa-and-valley landscape was still concealed in a distant future. Initial incision and
slope retreat probably were slow, so that the landscape (even adjacent to the deepening valleys
cutting into the Edwards) remained Washita Prairie for a very long time.

Stratigraphy: Initial incision through the Edwards Umestone; the Georgetown limestone has
been stripped from the immediate valley, but caps the interfluves (Figs. 2-6).

Relief: From top of the interfluve to stream level, relief is about 75-100 feet; the valley is
dominantly v-shaped in cross-valley profile (Figs. 2-6).

Drainage: Drainage density (miles of channellsq mi of basin) is about 3.0 mllmi2. (Fig 2-7);
drainage density is greater on the south side of the river than the north side; streams appear
to fall within "initiation and elongation" stage (Fig. 2-7).

Slope Retreat: Slope retreat is minimal; cross valley distance between the Edwards free-
faces is 200 feet or less (Fig. 2-6).

Soils: Flood plain soils are Frio Series silty clay loams; overlying the Edwards are terraces
capped by Krum and Searsville clayey soils.

Drainage History: We infer that the river flowed just above (not on) the resistant Edwards
Limestone prior to incision, based on the presence of extensive terraces,. Regional mapping
suggests that Hog Creek and the Middle Bosque River flowed eastward across the White Rock
Escarpment during that time to discharge into an ancestral Brazos River far below Waco (Fig. 4-
6). If so the elevated Austin Chalk would have provided a local base level. Hog Creek and the
Middle Bosque would thus have been more isolated from Brazos River Incision than the North
Bosque, which discharged into the Brazos at Waco, or which for an extended period may have
been the major pre-Brazos trunk stream at Waco.

1-32
If we visualize the valleys of Hog Creek and the Middle Bosque as early stages in Cut Plain evo-
lUtion, it is natural to think of them as somehow younger than the more typical broad Cut Plain
valleys we shall see on the rest of this trip. But Hog Creek and the Middle Bosque are "divide
streams", flowing along the "divides" between the more deeply incised North Bosque and Leon
rivers. Their patterns were established at the time of Cut Plain incision, and they were part of
an original stream network, draining at least the basins they now drain. Bifurcation ratios clearly
indicate that Hog Creek has lost basin to the encroaching North Bosque network (Proctor,
1969). The Middle Bosque may yet have much of its original basin area.

The lack of major incision in comparison to adjacent basins was apparently a product of lower
gradient, and of inadequate basetlow. Stream basin area apparently had no significant effect
(Fig. 2-4). For example, Hog Creek flows more than forty-five miles at an average gradient of
about 13 ftlmi to fall some 600 feet before it joins the North Bosque River at Waco. At the
time of formation of the Cut Plain, stream gradient between the headwaters divide and the
White rock at Waco was about 9 ftlmi, and stream length may have been as much as fifteen
miles greater than it is now. In addition, almost throughout the history of landscape evolution,
Hog Creek has lost surface basin area to the larger and growing North and Middle Bosque Rivers,
and consequentiy, surface water flow was reduced through time relative to those streams (Fig.
4-6), a point to be considered in greater detail at Stop 4.

Finally, the Middle Bosque was in a favored position to intercept groundwater flow directed
down dip. Groundwater, the source for basetlow, was apparently essential to both incision and
slope retreat in the developing Cut Plain. The reduction in baseflow to Hog Creek, resulting
from interception of groundwater by the up-dip Middle Bosque River, very early limited the
ability of Hog Creek to deepen or widen its valley. Similarly, baseflow to the Middle Bosque was
intercepted by the updip Coryell Creek, and base flow to Coryell Creek by the Leon River. This
factor may have had major effect, though the actual mechanics of basetlow solution-erosion
are only vaguely understood. The role of groundwater in Cut Plain evolution will be pursued
throughout this trip.

An increase in either gradient or volume could also radically change landscape and stream
organization. At one point near Stop 2, Hog Creek is within one mile of the valley wall of the
North Bosque basin, where it could fall one hundred fifty feet to the North Bosque drainage in a
distance of about two miles. Capture seems imminent, and under conditions in which power will
Increase greatly. Yet this situation has remained essentially unchanged for perhaps the past
half-million years. Capture under similar conditions during very active slope retreat, aided by
equally active groundwater flow may have been a principal mechanism of drainage
reorganization in the then-evolving Cut Plain.

1-33
Drainage reoriented, forms pattern for
initial Cui Plain incision

SOUTHWEST NORTHEAST

High gravels Time Zero profile Washita Prairie


"

~
I
",qe.=-.!~!*~'
.~-~
"'*.
___ ~-- -
:7
~=f
~~!
---;coo - . . .--'-
Georgetown

Edwards -'---T-<--r--'--,.- ) ---,


! !

Comanche Peak

High gravels

-----=~-~--J_
Washita Prairie profile f
Beginning Cut Plain

~-
--- Washita Prairie

Georgetown - .. - - - '- ---- .e!~~= •


~-

Edwards
Comanche Peak

~L
100 «

Figure 2-6 Block diagram. and section across Hog Creek valley at Stop 2. Notice That while
In the immediate vicinity of the creek the profile is somewhat reflective of the Cut Plain. the
region is still clearly Washita Prairie. However. once stream paths were clearly established on
the Washita Prairie. only major structural tilt. or major incision in adjacent valleys could induce
significant change in the principal stream-net (from Hayward. Allen. Amsbury. 1990)

1-34
_ _~

... \ \

'1~~i, ,,\uL~
.... ~_........
\ \
__J..
~

::!l!'-~
.- I I .. /.,:"\, 1 f \ :,
"
~
".{":' 't:)h\ "\'i\,\
I i
I (. I I I _----,
!! \
e\/i\,..O
-,

-
i. I
.... ' !

't-4'W\cS.e
....... - -

Miles
'-+-
,

Figure 2-7 Drainage map and block diagram showing the effect of headward elongation of
Hog Creek on tributary formation (see also Figure 2-3), Drainage density of a segment of Hog
Creek Basin (taken from a 1162,000 scale topographic map) is about 3 mi/mf!· The difference
in tributary length on north-facing and south-facing sides of the valley is a product of difference
In antecedent moisture and slope direction (and hence insolation), On the block diagram, the
unit cell is the minimum first-order basin area. notice that as the main stem elongates, the
tributary network also elongates, This illustrates the stages of initiation and elongation (from
Hayward, Allen, Amsbury, 1990)

\-35
STOP 3--MIDDLE BOSQUE RIVER--Drainage Elaboration And Extension, (31 0 35'
53"N; 97 0 33' 55"W; Coryell, 1955; Moseeim, 1979, quadrangles, 1/24,000)
(Fig. 3-1)

A later stage in dissection··though not an 'older" stream.. is represented by the Middle Bosque
River at Stop 3. Here drainage, locked into position when this was Washita Prairie, has not only
cut through the Edwards Limestone, but almost through the underlying Comanche Peak
Umestone. Valley widening has been extensive, almost entirely as a product of slope processes,
but the width·to·depth ratio is not greatly different than that of Hog Creek at Stop 2 (Fig. 2·5).
The landscape is still dominated by Washita Prairie, and entrenched streams, while far more
deeply entrenched than at Stop 2, are yet minor elements in the Washita Prairie landscape.

Stratigraphy: The Edwards has been breached, as has most of the Comanche Peak, and the
stream is flowing on the basal Comanche Peak Formation (Fig. 3·1).

Relief: It is about 120 feet from the top of the Edwards·capped interfluve to stream level
(Fig. 3·1).

Drainage: Drainage density overall is about 2.95 mi.!mi2.; drainage density on the south side
is less than on the north side (Fig. 3·3). This is interpreted as elaboration of drainage,
cessation of extension on the north side; on the south side the drainage continues to extend
(Figs. 2·2, 3·2).

Slope Retreat: With elongation and elaboration' of the drainage net, the Interfluves are being
dissected from the east and west resulting in inner·valley widening to the north and south along
the main channel. Total width of the main valley has now reached 1000 feet (Fig. 3·1).

Soils: Soils of the slopes consist of the thin Cranfils (formerly Bracket) Series, while the flood
plain/colluvial apron consists principally of Frio silty ciay loams.

Drainage Evolution: Stops 2 and 3 form a very convincing time·sequence ..and a thoroughly
acceptable model for the evolving Cut Plain, but their stream valleys are actually of about the
same age. Valley deepening apparently progressed as a product of interception of groundwater
flow and knickpoint migration from an entrenching Brazos River. Stream segments of greatest
base flow, and during floods of greatest power, entrenched readily, leading to rapid upstream
migration of knickpoints. Stream segments of lower baseflow and power were far slower to
entrench.

1·36
stili dominates landscape

Divides remain at original height

Entrenched through Edwards and deeply


through Comanche Peak
Washita Prairie

-- ---
L ___ -- -_.__ .... Hog Creek profile
............
,----/
;'

,r - - ---.I'-~~~:
Georgetown ..........

Edwards
---- I
.L. -"~

Comanche Peak

--'~
100 "

Figure 3-1 Block diagram and section across Middle Bosque valley at Stop 3. The valley
profile here is far more characteristic of the Cut Plain, but the broad divide regions are yet
dominated by Washita Prairie. Hog Creek and the Middle Bosque have similar basin areas (Fig. 2-
4), but the degree of entrenchment is markedly different. Basin area is not a significant factor
in valley entrenchment, nor was time of capture (?) by the South Bosque River, since Hog Creek
was the first encountered. The interception of groundwater flow by the Middle Bosque, up dip
from Hog Creek, may be an explanation. Hog Creek, having lost basin to both Middle and North
Bosque rivers, now has a vel}' long and narrow basin, also a factor (from Hayward, Allen,
Amsbul}', 1990)
t--.. . .........
''I at",·
'a~ I
J! 1 I \ 'f >-J
~~
1,)(/\(1
\'1'. 4\' 1. I,...,.~ I\
o Milo. '\ . i ( ~/; \
"'.1 y . Y,l
'"
'!i\UY1_ - .'
. . :I: 11/
:'
\\ 0\ _ - 17.!" j j. ! I
\\tf\,... _ -"---r" :!... ,,,'
""
-........,. /
'"
.... _.- i '
J
.,..,. .
..,. I
/" /"r _.'-i
/ ~._- . -,- i - 7io- I ~
A
1\ ,
\
l'/\:
' ••,. A , " , .......J...
..

'
\ { .." \ l'
' ..

, : , .. -'.
-

10
I~
I _;;
- . . , , ; - _ _ ..." ' J J . _}
\ /" I J . ,-_.... ::
, t' . . . . - . . . ~ .... \08 I.§
0'" -
'.....
--.....
.... - I
I

Figure 3-2 Drainage density of a segment of Middle Bosque Basin (taken from 1/62,000
topographic map) is again about 3 mi/mi2. This value appears to be a property of regional
climate and stratigraphy. Here elaboration is accomplished by increase in valley width, not by
increase in drainage density. Again the difference in tributary length on the north and south-
facing sides of the valley is at least in part the result of differences in antecedent moisture as a
product of slope direction. The block diagram shows the effect of elaboration on tributary
formation and elaboration by headward migration in a part of Middle Bosque Basin. The unit-cell
shown here is again the minimum first-order basin area, and has the same dimension as in the
stage of elongation. Notice that as the main stem entrenches, the tributary network elongates
and elaborates, with more complex tributary networks. This is clearly the stage of elaboration
(from Hayward, Allen, Amsbury, 1990).

INTERRUPTION IN THE SEQUENCE

STOP 4--NEILS CREEK--Elaboratlon And Abstractlon--The beginning of the true


Cut Plain--(31 0 42' 5" N; 97 0 37' 48" W; Hurst Spring [1956]. and Mosheim
[1979] quadrangles, 1/24,000) (Fig. 4-1)

As streams approach grade with constant discharge, down cutting Is reduced. Valley widening
continues independently, so width-to-depth ratios increase in an orderly sequence (defined by
an almost hyperbolic curve) from initial incision to ultimate stripping. But here at Stop 4 that
model fails. Immediately obvious is that while width-to-depth ratios have increased, it is not by
the predicted continuous series, but by an enormous jump. There are no intermediate stages
between that of Stop 3 (Middle Bosque River), and Stop 4 (Neil's Creek). Here the valley of
Neil's Creek is typical of most of the Cut Plain. It is not an exception (Fig. 4-2). Here we are
only about eight miles from Stops 1 and 2, yet we emphasize again that all these streams are
about the same age.

The slow reduction of Washita Prairie to form Cut Plain (noted at Stops 1, 2, and 3) has
suddenly become the total elimination of Washita Prairie. The landscape whiCh was dominated
by uplands has become a landscape totally dominated by valley floor; Washita Prairie is
confined to occasional small areas on narrow high divides, entirely in the eastern Cut Plain. Even
there those small areas are insignificant. The Washita Prairie is gone.

1-38
Stratigraphy: The main channel is now flowing in the Walnut Clay; the Comanche Peak
nodular limestone and Edwards limestone form a continuous slope, which rounds over at the top
of the Edwards (Fig. 4-2).

Relief: Maximum valley relief from the top of the Edwards to the valley floor is about 160 feet
(Fig. 4-2).

Drainage: Drainage density is about 3.5 milmi2.; drainage density on the south side of the
main channel is considerably higher than that on the north side. The lower density on the north
side is attributed to abstraction caused by reduced rates of downcutting and the presence of
the Walnut Clay which has a different hydrologic response (Fig. 4-3). The greater drainage
density on the south side is related to continued elaboration of the drainage in the Edwards
dominated uplands (Fig. 4-3, 4-5).

Slope Retreat: Two phases of interfluve dissection are evident here; (1) to the north, the
last remnants of the dissected interfluves are visible, representing the final phase of drainage
dissection and slope retreat; and (2) we are presently standing on a retreating interfluve,
similar to that seen at Stop 3. Total cross valley distance is now about 3,000 feet (Fig. 4-2, 4-
5)

Soils: Soils are largely residual except along slope bases where Cranfils soils mark thick
colluvial, caliche-dominated deposits, and on the elevated pediplain where Krum soils indicate old
terrace remnants related to an episode of reduced downcuttlng or erosional still-stand.

Drainage Evolution: What we have seen at ·Stops 2, 3, and 4 are examples of the two
families of stream valleys we have called valleys of Types I and II (Figs. 2-4, 2-5). They are
clearly differentiated on the basis of width-depth ratios (Fig. 2-5). Type I valleys have
width/depth ratios of 12 to 16. Type II valleys are characterized by width/depth ratios of 40
to 80 (Brown, 1988), and even higher (Dahl, 1987, oral communication.) (Fig. 2-5). The
difference in ratios appears to be a product of stream history at the time of incision into the
forming Cut Plain. Increased baseflow, and steeper gradient to major base-level, with its
attendant increase in power, appear to have been the dominant control. Basin area was not
critical, as indicated by field observations (Fig. 2-4, and Stop 8).

The broad valleys of Type II (Stop 4) have another common characteristic. Slopes leading down
from Edwards-capped divides are not graded to present entrenchment, but to a surface much
above present drainage and substantially below divide elevation (Fig. 4-2). Earlier this older
surface was given the name "Intermediate surface", because that term had no genetic
connotation, and we weren't sure what the surface represented. From habit, that name
continued to be used (Hayward and Allen, 1987, 1988). In 1990 (Hayward, Allen, Amsbury,
1990) we reviewed the history of the Cut Plain (and by implication, far wider regions of the
south central United States), and adopted as a newer name, "Comanche Pediplain", named for
the rock Series on which it is developed, and the shape which it exhibits (King, 1962, p. 162).

Finally, as is evident from the profile of the far divide to the north, gaps appear for the first time
as conspicuous elements in the landscape. The origins and Significance of these gaps to Cut
Plain evolution is a recurring question throughout the progress of this trip.

1-39
-:

\ j
.'

,-..r,

l /
I
i

,I

.,
'.

l .
.i") r.
'
/f_.__',
/.>~/' (
!/ /0
/ '\~(.
.'

Rgure 4·1 (Hurst Spring [1956}, and Mosheim [1979} quadrangles, 1/24,000) Neils Creek
Valley, the Lampasas Cut Plain. In contrast to Stops 2 and 3, here the valley dominates the
landscape, and the divides are reduced to remnants, Here for the first time we see the typical
Lampasas Cut Plain--Type II broad valleys with narrow Edwards capped divides. Here, for the
first time, we see an "intermediate surface" descending from the divides, but not graded to
present drainage. This is the most conspicuous difference between Type I and Type /I
landscapes, Hog Creek and Middle Bosque basins are entirely Type I valleys.
1-40
Plane of section

Cut Plain

Washita Prairie now gone

SOUTHWEST NORTHEAST
Streams entrench into pediment

Middle Bosque profile Comanche Pediplain


----1 ---- ---- -
Edwards:t:;;;:=1>.
) r-
Colluvium ,
\
... -""\" , I

-- ~

Walnut --

\i
------~lQ~Q~Q~ft~--~

Figure 4·2 Block diagram and section across Neils Creek valley at Stop 4. The valley profile
here is typical of the Cut Plain, and Washita Prairie is gone. The degree of entrenchment is
slightly greater than in Middle Bosque basin (Stop 3) but valley width is far greater (Fig 2-5).
Basin area (and hence flow volume) was apparently not a factor in valley-widening, for even very
small tributary basins adjacent to our present location show similar profiles (from Hayward,
Allen, Amsbury, 1990)

1-41
dissected pediment

Figure 4-3 Drainage density of a segment of Neils Creek Basin (taken from 1162,000
topographic map) is about 3.5 milmi2. The increase in density is a result of greater basin area
in shales of limited permeability, of the Walnut Formation exposed on the Comanche Pediplain.
This is the abstraction stage (note particularly the northern side of the valley) where increase in
relatively impermeable valley area has resulted in loss of more permeable divide lands. Block
diagram shows the effect of abstraction on tributary network as smaller tributary basins are lost
to other more effective stream systems. The main-stem entrenchment is not greatly different
from that of the Middle Bosque, but valley width is far greater, and the abandoned pediment is
the dominant element of the landscape (from Hayward, Allen, Amsbury, 1990).

Groundwater In Cut Plain evolution


In summary, evolution of the Cut Plain landscape was principally the result of groundwater
sapping and drainage integration. The sequencing and spatial variability of these processes has
been regulated by base level controls exerted on the system by the Brazos River and the range
In permeabilities of the local stratigraphic sequence.

This hypothesis is based on three lines of evidence: (1) recent work on hydrogeology of this
area (Cannata, 1988; Collins, 1989; and Myrick, 1989); (2) analysis of the evolution of streams
and landforms in the region (Epps. 1973; Montgomery. 1986; Brown. 1988; and Tharp. 1988);
(3) recent work on the Colorado Plateau (Laity and Malin. 1985).

Groundwater sapping is an erosional process that produces landforms with unique


characteristics (Laity and Malin. 1985) (Fig. 1-5. Tables I and II). Perhaps the most important
control on such processes is the dip direction of the beds relative to the direction of elongation
of the stream valleys. since this controls the occurrence and distribution of groundwater
seepage at the valley walls (Fig. 2-3) (Laity and Malin. 1985. p. 203).

The hypotheSized evolution of the drainage (Fig. 4-6a) involved three major steps; (1) a pre-
Whiterock Escarpment stage of normal consequent flow of the Brazos and its major tributaries
across the Cut Plain. in which the tributaries joined the Brazos below Waco; (2) a stage of
capture of the major Brazos tributaries by the short. shale-subsequent South Bosque River
along the northern face of the developing Whiterock Escarpment (Fig. 4-7b); and (3) the

1-42
subsequent "Isolation" of Hog Creek from groundwater supply through the interception of
groundwater flow by the Middle Bosque River, up dip from Hog Creek (Fig. 4-7c).

Among the explanations for the unusual properties of Hog Creek Basin, this appears the most
reasonable. Of the three streams (Hog Creek, The Middle Bosque River, and Coryell Creek) the
Middle Bosque River has the largest drainage area, and the best developed tributary network,
mostly on the up dip (western) side. Both Hog Creek and Coryell Creek flow in long narrow
valleys, with very simple, largely first-order tributary networks (Fig. 4-6). It appears that the
Middle Bosque River, near the up dip margin of the regional groundwater system was able to
capture more of the available groundwater flow, and was more successful in extending its
drainage for this reason. Coryell Creek, a stream which preserves much of the Cut Plain history,
was apparently cut off from groundwater acquisition by the rapid entrenchment of Leon River,
as the Leon cut through the Whiterock near Temple. Scarp retreat in the Leon Basin had
already cut away the up-dip section that earlier supplied groundwater to Coryell Creek.

Evidence In support of this model of drainage evolution includes six components; (1) shapes,
locations, and tributary networks of the three basins (Fig 4-6); (2) contrasting landforms of
their inner basins, particularly the significant differences between the landforms of Coryell
Creek, and those of Hog Creek and Middle Bosque (Figs. 2-6, 3-1, 4-2); (3) major springs
entering the Middle Bosque from the southwest (up-dip) side, flowing from a diffuse aquifer
(COllins, 1989) (Figs. 4-7, 4-8); (4) greater baseflow in the upper, more-dissected areas of the
basin, where the aquifer is cut by numerous tributaries (Myrick, 1989); (5) basetlow
contributions to the total flow than exceed storm runoff (Cannata 1988); (6) constant, main-
channel valley-width with increasing drainage area, indicating that It is more dependent upon
sapping than upon stream flow. Tables I and II summarize the attributes of basins of surface
water origin and basins of groundwater origin. <

In the middle Bosque Basin (and contrary to intuitive expectations) baseflow In 1987 was found
to comprise about 6.9 % of the total rainfall, whereas storm runoff accounted for only 5.3 %
(Cannata, 1988). This indicates a major potential for groundwater sapping under the proper
stratigraphic and geomorphic circumstances.

There Is no question that overland flow and related erosion is still a powerful mechanism in
landscape modification of interfluves (Flowers, 1989) and stream channels (Baker, 1977) in the
Central Texas region. But groundwater as a geomorphic agent in the Cut Plain has not
previously been considered a major agent in landscape modification. Yet both geomorphic and
hydrologic evidence seem to Indicate that it may well have been dominant during major
episodes of scarp retreat and landscape modification.

The organization of early drainage


The transition from Washita Prairie to Lampasas Cut Plain was originally thought to represent
simple entrenchment through the Edwards Limestone to a generally graded profile. This
entrenchment would be accompanied by rapid valley widening, until as incision diminished, the
"normal" Cut Plain profile developed. By this interpretation the degree of development of a
valley is one measure of its actual age (though other factors would, of course, be involved).

1-43
..,c
"c c
cOo
C Q= -
.!:! in ~ U
- c C) ~
~ QI.Q U;
- .CD .c
.E--w "---<1:-

Hog Cr. M. Basque Neils Cr.

Figure 4-4 Stages in drainage network evolution and changes in drainage density in Hog
Creek, Middle Bosque, and Neils Creek basins. Notice particularly the differences between north
and south sides of the valleys, differences caused by factors related to slope direction. The
slight Increase in drainage density during the abstraction phase is attributed to greater area of
basin in relatively impermeable shales (from Hayward, Allen, Amsbury, 1990).

c c
c .2 .2

-
o
0; .
iii U
e
'"E
C
II
;c .
o
.D iii

!-
.D
iii
W

f-- -' Coryell Ct. I , I


<I:

B
-. ...........
Owl Cr.

lZ~ff C!.:.
;e-
o; 8
INeils br. U~Lamp.
c Cave Cr.
Q" 4
M. Bosque R.
Hog Cr.

.'"
II

2
C
Q
.
c

0 10 ro ~ ~ ~ ~ ro ~ ~ 1~
Valley width/Basin width

Figure 4-5 Stages in drainage development, Lampasas Cut Plain, as indicated by the ratio of
drainage density versus valley width/basin width (valley width is distance between Edwards
faces on opposite sides of the valley, basin width is distance divide-to-divide). Notice that as
the basin matures and divides narrow, drainage density first increases and then decreases
through stages of elaboration and abstraction (after Brown, 1988).

1-44
Hog Cr.
M. Bosque
Leon


As'.
Brazos
.'
N. Bosque

Leon

,,
. - ... ;;
-"u ~
Brazos

N. Bosque ":~
0

~og Cr. a:
::"
M. Bosque .• , ., _ .... ' ....... ;: 3:
-", .... E .. ;:::..; . Leon
,....... CJ "'......
.:::i .....

,gl"
.,.

Figure 4-6 Postulated early connections of Washita Prairie streams with the Brazos River.
The South Bosque River, now an interceptor of Cut Plain drainage, may not then have existed,
nor did the interposing ridge of the Whilerock Cuesta, developed on Austin Chalk. North Bosque
River, Hog Creek, and the Middle Bosque River may have then flowed directly to successively
lower points on the Brazos. Interception of this drainage by the South Bosque (a shale
subsequent) affected first the North Bosque, and then In succession Hog Creek and the Middle
Bosque. In the sequence of diagrams: Time 1, drainage antedates the formation of the
Whiterock 'scarp; Time 2, at A the Brazos River cuts through the evolving 'scarp, at B the South
Bosque River extends headward; at C the Leon River entrenches, cutling regional groundwater
flow, and at 0 the favored Middle Bosque River elaborates; Time 3, at E, the Leon River
continues to downcut, Coryell Creek extends and elaborates headward, Hog Creek, cut off from
groundwater flow, extends without elaboration (Hayward, Allen, Amsbury, 1990).

1-45
TABLE 1

DRAINAGE EVOLUTION UNDER OVERLAND FLOW


1. Valley widths increase with distance from the valley head, and with stream order.
2. Valley heads are tapered, narrowest at the divides, widening downstream.
3. The longitudinal Profiles are usually concave.
4. Knickpoints in the profiles of tributaries match those of the trunk stream
5. A branched network of tributaries fills the drainage basin, evolving rapidly in the early
stages of development, and undergoing slow changes thereafter.
6. Direction of basin growth is related to surface landform, and fracture control.
7. Drainage lines may show prominent joint control where fractures are exposed at the
surface, but may show almost no structural control where such joints are deeply
mantled with soil or debris.

TABLE 2

DRAINAGE EVOLUTION UNDER GROUNDWATER FLOW


1. Valley widths tend to remain constant, dependent upon sapping rate, and not on stream
volume.
2. Valley heads are steep, rounded, cuspate theaters, not tapered "vee" notches.
3. Longitudinal profiles tend to be linear, with the direction and gradient of the dip.
4. Runoff from the uplands is sufficient to account for large scale headfalls by waterfall
erosion.
5. Valleys extend slowly headward by sapping, ultimately to define a basin. There is large scale
variation in space-filling characteristics from basin to basin, depending on the relative
ages of the systems.
6. Valleys grow in response to the hydraulic gradients of the groundwater flow. The
direction of growth may mimic that of the joint-controlled streams where surface
topography reflects subsurface flow.
7. Valley growth reflects the regional joint pattern even in debris-mantled basins. Joints at
depth provide pathways for internally moving groundwater.
8. Stream flow derived from groundwater discharge increases steadily from the headwall
to the midsection of the valley, and then increases only slightly to the outlet.

But recall that at the Washita Prairie (Stop 1, Hog Creek (Stop 2), Middle Bosque River (Stop 3),
and Neils Creek (Stop 4), the point was made that all valleys are of the same true age, they
flow in similar stratigraphic sections, and that basin-area is apparently not a major consideration
(Fig. 2-4). The differences in development are attributed to different histories of origin,
Involving base flow differences as a result of groundwater contributions, development of knick--
pOints, basin shape, and perhaps other factors as yet unrecognized. Thus rather than time
being the dominant control on Cut Plain valley configuration, process was dominant, and time
was secondary. This point was considered at each of the four stops.

Both Type I and Type II streams have similar ranges of basin area, but radically different valley
widths. (Fig. 2-4). Even minor tributary basins of Type II streams show similar wide-valley
profiles, with well-developed intermediate surfaces, and soil blankets similar to those of larger
Type" basins. Thus the history of Type II trunk streams was shared by the total tributary
networks of those streams. Type I streams remained largely in Washita rocks, and while they
may have shared in a valley widening phase, the evidence was not preserved. Type" streams
had already cut through the Edwards Umestone and participated in the valley widening episode,
always in Fredericksburg section, which preserved the evidence of that event. (Brown, 1988)
Washita Prairie
L.ampasas Cut I"ain

;~-~-~-:-~~~~~~tE~~
"Edwards
-----_ _ -".Sprlng
,/ .
,,
' ....

~--::------
Warer T.b'I""' .......... __
-----

Figure 4-7 Diagrammatic section showing transition between Washita Prairie and Lampasas
Cut Plain. Perennial springs discharge near the margin of the Cut Plain where steep-gradient
tributaries intersect the water table (from CoJlins, 1989). Runoff from the uplands and
seepage from the margins of the uplands are particularly important mechanisms of scarp
retreat, particularly on the downdip margins of the interfluves. Early in the history of slope
retreat when the uplands were wide, both runoff and seepage were at a maximum. Antecedent
moisture on the slopes was also at maximum Therefore, slope retreat was at a maximum. As
divides narrowed through slope retreat, runoff and seepage diminished, as did antecedent
moisture, and rates of slope retreat declined. Ultimately, as the divide narrowed to a width that
no longer provided significant runoff or seepage,antecedent moisture diminished dramatically,
and slope retreat then almost ended, sustained only by occasional intense rains falling directly
on the slope faces. It was at that stage that, except for those periods of intense and frequent
rainfall, landscape modification became almost dormant. This last stage is that of the present
Cut Plain.

Also emphasized, a second feature of importance in the evolution of the Cut Plain landscape is
the orientation of trunk (consequent?) drainage (Fig. 2-2). Two prominent stream orientations
exist. Major trunk streams, such as the Brazos River, North Bosque River, Leon River, Cowhouse
Creek, the Lampasas River, and a number of lesser tributaries flow southeastward, at a marked
angle to the dip. Hog Creek, the Middle Bosque River, Neils Creek, Meridian Creek, and a number
of other tributaries of trunk drainage flow generally eastward, down dip. This pattern of
drainage is repeated throughout Central Texas.

An imaginary surface contoured on the highest divides of the Cut Plain clearly reflects the dip
direction, and the direction of the tributary streams, but fails to explain the direction of the
trunk streams (Fig. 2-2).

If trunk drainage was initially consequent drainage, then the initial slope was to the southeast.
This consequent drainage may have resulted from (1) different geologic structure at the
inception of modern drainage, (2) from trunk drainage established on higher units, now
removed, in which a northward thickening wedge of overlying sediment provided the gradient
along which consequent drainage developed (Montgomery, 1986), or (3) effects of Tertiary
reactivation of deep-seated Paleozoic structural lineaments (Amsbury, 1990, oral
communication.).

The Miocene (Oakville) strike is essentially normal to this trunk drainage direction (Fig. 4-9) as
are the strikes of later units, leading to the speculation that perhaps initial slope was to the

1-47
southeast, and that this general drainage direction was established after eastward Eocene
drainage, and before Miocene Oakville deposition (Sam mel, 1978).

o
I
\ 5
I

Miles

Structural contours on Edwards

Figure 4-8 Topographically controlled groundwater flow in the Washita Prairie. Notice that
the groundwater flow-vectors generally follow the regional dip, but they are greatly modified by
topography. Typically, flow-vectors follow topographic divides, eventually to feed into
entrenched streams. Structural contours are on the Edwards Limestone over the area of Hog
Creek and Middle Bosque basins. The Cretaceous limestones of the aquifer system (Comanche
Peak, Edwards, and Georgetown) dip gently east-southeast at 20 to 40 ftlmi. DissectIon of the
landscape alters the groundwater flow from dip-controlled to topographic-controlled (Cannata,
1988, p. 94-95). Note particularly the marked modification of flow lines along the divide
between Hog Creek and the Middle Bosque, and along the divide between Coryell Creek and the
Middle Bosque River. This modification is a product of dissection through the aquifer section by
stream entrenchment. The spaCing of tributary valleys is related to the competition for
groundwater discharge, which is, In turn, related to the degree of dissection of the landscape,
and to the potential groundwater drainage area up-gradient from the site (simplified from
Cannata, 1988).

1-48
I
I Marked angle to early Tertiary contact

_ _ Creta ceous-Eocene contact


I

_ Normal to mid-Tertiary contact

(Early Miocene) contact

Figure 4-9 Relationship of (consequent 7) trunk drainage direction to strike of early and
middle Tertiary rocks, Texas. While there is evidence for ancestral Brazos drainage as early as
Eocene (Fisher and McGowen, 1969; Bammel, 1976), the dip-orlentatlon of main trunk drainage
normal to Oakville strike suggests (admittedly very tentatively) that present trunk stream
orientation dates from early Miocene Time and major subsidence in the Gulf Coast. While other
factors were also involved, slope was ultimately the dominant control on stream dIrection (from
Hayward and Allen, 1987, 1988).

STOP 5-TURKEY CREEK DIVIDE--Turning back the clock-(Sugarloaf Mountain


Quadrangle [1979). 1/24,000; 31 04S'27"N; 97° 39'22''W) (Fig. 5-1)

We are on the Comanche Pediplain, at the headwaters of Turkey Creek, a southeastward flowing
tributary of Neils creek. An unnamed tributary also flows northward from this divide Into
Meridian Creek. This Stop, has fou r pOints of emphasis: (1) the anomalous location and
configuration of the drainage divide; (2) the unusual cross-valley profile; (3) the equally unusual
soils of the divide; and (4) the relationship of these features to problems of Cut Plain history.

Within this part of the Cut Plain most divides (particularly between small drainages) are defined
by Edwards-capped highlands into which intermittent streams have incised narrowing, well
defined valleys. The divide here occurs within a valley, a windgap, on a broad gentle
Intermediate level well below the Edwards-capped uplands to east and west, but substantially
above present dissection to north and south (Fig. 5-2). The cross-valley profile here at Turkey
Creek Divide is in marked

1-49
IA\
Figure 5-1 (Sugarloaf Mountain Quadrangle, 1979, 1124,000;) Stop 5, Turkey Creek
headwaters divide is in the bol/om of the valley in which we stand. The surface around us,
veneered in Krum alluvial soil, is the original "intermediate surface", now called the Comanche
Pediplain. Here is a Type II broad valley with narrow Edwards-capped divides, almost unmodified
by modern drainage. We see the Pediplain graded from the divides, but clearly not to present
drainage. Here it is an almost unaltered remnant of a distant geomorphic past.

1-50
contrast to most divide notches in the Cut Plain. The breadth of this divide-valley cannot be
explained by present processes, for only a very few other such landforms exist in the region. It
cannot be attributed to lithologic control, for no unusually resistant units are known to exist in
the proper stratigraphic-topographic position. It is not developed in Walnut Clay, but in
Comanche Peak Umestone, an unusual circumstance in this region of the Cut Plain where most
such surfaces have been explained by the erosional properties of Walnut Clay. The profile of the
upper slopes to the crests of Norse Hill to the west and Bee Rock Flats to the east are markedly
more gentle and rounded than are equivalent slopes descending to Meridian Creek and the North
Bosque River (Fig. 5-3), both more clearly products of modern dissection.

Also of note here are the soils of the Krum series, derived from old stream terraces, and formed
originally "In valley areas over thick beds of unconsolidated clayey sediments" (Stringer, 1980,
p. 48). Here the soils are in agriculture, an unusual land-use for Cut Plain divide lands, but very
common for the pediplain. Perhaps most unusual is the presence of abundant sand and silt in
the soil. Sands are not known from the Edwards and Comanche Peak formations, above us in
the section (Amsbury, 1989, oral communication), so this soil-component was transported from
another, more distant source.

There is coarse silt in the lower Walnut Texigryphaea banks (Amsbury, 1995), but here these
beds are substantially below our present elevation of 860'. The Paluxy Sand (which is below the
Walnut Clay, Fig. 0-2, 0-9), has abundant sand similar to much of the sand of this soil, but the
nearest locality where Paluxy Sand occurs at an elevation above our present locality and high
enough to provide for an adequate delivery gradient, is almost thirty miles westward along
Meridian Creek valley from our present location, and across the divide into the Leon River Basin
(Stop 7).

While we recognize the possibility of a distant source for quartzose sand from the Paluxy
Formation, there yet remains a question. A characteristic of the soils of the pedlplain
everywhere is a surface component high in quartz silt, even in locations where there is no
possible access to Paluxy Sand. The silt so common to these soils may have been derived from
loess. for over much of the Cut Plain they appear more loess-like than Paluxy-like. Slit Is also
present in many of the soils of the Edwards uplands, and over most of the landscape identified
as "old". Since silts and sands do not occur in the Edwards and Comanche Peak limestones
(Amsbury, 1995), they are apparently not residual, yet to date, efforts to compare
the silts of the Comanche Pediplain with loess from the Mid-Continent, and from the High Plains
have been inconclusive (Bradley, 1988).

All of these factors combine to suggest that here at Turkey Creek Divide we are on an ancient
landform, unrelated to present topography. The state of preservation suggests that processes
active In this locality since the Initial isolation of this surface have done little to alter this relict
landscape. It has been suggested (Moore, 1970) that this was a former pathway of Meridian
Creek (Fig. 5-2), and that Turkey Creek is the beheaded remnant of that ancestral stream. This
Is supported to a degree by the Krum soil at this locality (Stringer, 1980), derived from ancient
alluvium If so, Meridian Creek then jOined Neils Creek, but at present that interpretation is
based largely on stream alignments and soil series. Testing this hypothesis for absolute
confirmation would require drilling a series of holes across this surface, but to date that has not
been attempted.

The elevated surface here at Turkey Creek Divide, the Comanche Pediplain, represents an
episode in landscape evolution in which entrenchment had almost ceased, and in which the
dominant landform modification was valley widening over an ex1ended time-span.

1·51
Bee Rock Flats

Farmlands on Comanche Pediplain

Figure 5-2 Diagram showing the relationships between the old high-divide surface of Norse
Hill/Bee Rock Flats and Turkey Creek headwaters divide; and between Turkey Creek Divide and
the recent entrenchment of Meridian Creek. Notice that Turkey Creek and the northward-
flowing tributary of Meridian Creek head within a valley, not on an Edwards-capped divide, and
that Turkey Creek Divide is apparently unaffected by the recent entrenchment of Meridian Creek
Valley (from Hayward, Allen, Amsbury, (990).

NORTH
SOUTH

aee Rock Flat

d:=-~-,._~/..........;;::::"""'!~~- - t - -_. . --.- T i


Conllnuous slopa to Pedlplaln, /"., " O. II lJ

. /; -/--~i:'~~' 2·L·fI~
~::::-":\'/ Pedlplaln ven~.red with elluvlal (Krum) soli ---_. •...
Meridian Creek tributary
cutting Into Pedlplaln
Turkey Creek extending Into Pedlplaln

Figure 5-3 Diagrammatic section through Turkey Creek headwaters divide, showing
relationships of the Comanche Pediplain to the Edwards-capped uplands, and to modern
drainage. The Pediplain is clearly graded from the divide, but not to modern drainage. Here the
pediplain is veneered by Krum soil, an old alluvial soil, suggesting that at one time the divide was
the a pathway to through-flowing drainage (from Hayward, Allen, Amsbury, (990).

1-52
STOP 6··MERIDIAN CREEK VALLEY··Dissection and preservation of the Comanche
Pedipiain (Cranfiis Gap Quadrangle [1979], 1/24,000; 31 0 47'36" N; 97 0 47'25"
W) (Fig. 6·1)

Here at Stop 6 we are again on the Comanche Pedipiain first identified in Neils Creek Valley
(Stop 4), then at Turkey Creek Divide (Stop 5). The view to the north Is of Meridian Creek
Valley and the northern divide that separates Meridian Creek from Spring Creek drainage. To
the south is Jenson Mountain, which marks the divide between Meridian and Neils Creek
drainage.

At this locality there are four points of emphasis, (1) the very wide cross·valley profile, with a
well preserved pediplain; (2) the obvious grading of the divide lands to the pediplain, not to
present Meridian Creek; (3) the relationship of this pediplain to that of Turkey Creek Divide, and
(4) again the presence of what appear to be anomalous soils of the pediplain.

The width of the cross-valley profile is clearly evident in the view north from Stop 6 (Fig. 6-2).
The most conspicuous landform elements in the view are the flat Edwards-capped divides to the
north and south. Second only to these is the well preserved pediplain, now in farmland. The
lowest point on this surface is about 940 feet, 70 feet above present Meridian Creek, 80 feet
above Stop 4 at Turkey Creek Divide. The straight-line distance to Turkey Creek Divide is about
nine miles, giving a pediplain gradient to the east-southeast along Meridian Creek of about 9
ft.lmi.. West of this point to the pediplain at the head of Meridian Creek (11 miles west-
northwest) the gradient is 25 fUmi" The present gradient of Meridian Creek over that same
reach is 31 fUml, a difference of 6 fVmi Recent entrenchment has sharply increased tributary
gradients, suggesting that the dissection of the Comanche Pedlplain is a product of baselevel
lowering.

A second feature, visible in this view (Fig. 6-2), is that of uplands graded to the pediplain, not
to present dissection, a point emphasized at Stops 4 and 5. The consistent curve of the profile
from the base of the Edwards-armored divide to the low point on the projected profile is that of
a pediplain, graded to a former drainage nearer the south valley wall. A gentie southward dip
component of 13 fVmi may explain this valley asymmetry. While Meridian Creek has cut below
the pediplain, that remnant surface is still clearly visible, as Is the relationship of this surface to
the Edwards-capped divide.

At an elevation of 960' on the northern "pediplain" surface are soils of the Slidell Series, which
·".formed in valley areas over thick beds of unconsolidated, calcareous clayey sediments"
(Stringer, 1980, p. 52). Slidell soils are not unusual in the Cut Plain, but they commonly occur
as valley-fills, principally in first-order drainage on the pediplain (Bingham, 1989). Here they
occur as soil variants on the near-flat pediplain where there is no low point to explain their
presence. These therefore appear to mark former drainways no longer parts of the present
networks.

One problem with "alien" soils, such as the Slidell Series, Is that they are often not greatiy
different from other Cut Plain soils which we do not consider "alien". Here, clay soils of the
Slidell series are surrounded by more-calcareous clay soils of the Purvis, Tarrant, and Denton
series. Slidell soils have been identified as soils developed on unconsolidated clayey sediments
(Stringer, 1980), but they lack the clear isolation from surrounding soils that would derive from
a peculiar mineralogy which clearly separates the two. They appear out of place. They seem to
relate to the history of the pediplain, but they do not provide the deeper satisfaction that would
be provided by more convincing evidence (Brown, 1988).

\-53
,.'
...····r··
I~'
, ."

I s n <

("~
f ...... - '/
Figure 6-1 (Cranfils Gap Quadrangle, 1124,000; 1979) Stop 6, on the dissected toe of the
Comanche Pediplain, looking northward across Meridian Creek Basin to the Pediplain and divide
on the north side. The view to the north is of a Type " broad valley with narrow Edwards-
capped divides. The valley is now being actively entrenched by modern drainage. Again the
Pediplain is graded from the divides, but cfearly not to present drainage.
1-54
SOUTH NORTH

MerIdIan Creek entrenched


~ -:;:~\h~e
Pedlplaln 60 " below Comanche Pedlplaln

-'u.
u. ~
,....
~
~
,~~
"
"'. STOP 6
Slidell soli
--- ----.l--:::~-~:.<--. ....."l...:
----~, -.... ~-~---
\r.;>
-' .
, / - - - ~- - ~ _, -...: \,...,rlP!!"'"".:"
Profile of recent dlssecllon - - . ............. Rock defended terrace

Figure 6-2 Diagrammatic cross-valley profile, showing the Comanche Pediplain graded to the divides, and now being
dissected by recent drainage of Meridian Creek and its tributaries. When the Pediplain was formed, the bed of Meridian Creek
was 60 ft. higher than it is today. Here Meridian Creek flows on Paluxy Sand. At the time of formation of the Pediplain,
Meridian Creek flowed on upper Walnut beds, and the Pediplain extended upward into the lower Comanche Peak formation at
the base of the divide. Much of the view from this stop still consists of the relict earlier landscape (from Hayward, Allen,
Amsbury, 1990)
When we first recognized a widespread "intermediate surface", we were uncertain of its
significance. It was, however, truly a surface midway below the divides and above present
drainage--and so the name. Later mapping in the Leon Basin (Brown, 1988) clearly
demonstrated that the Leon Basin had its own "Intermediate surface", graded to an ancestral
Leon River which flowed at a level several tens to more than a hundred feet above present
drainage (Brown, 1988). Reconnaissance studies of several other basins indicate that most Cut
Plain basins of today's section of the trip have intermediate surfaces. All are graded to a
common high terrace level on the Brazos River.

Because streams of the Comanche Pediplain event in the Cut Plain were controlled by an
ancestral Brazos, all such surfaces should have lowest elevations graded to this common base
level. More importantly, all such surfaces reflect the same episode of landscape formation, for
pediplain surfaces are preserved far up even minor tributaries, where again are found thick
accumulations of colluvial-alluvial soils in protected pockets far above present drainage.

This latter again emphasizes that basin area was not a consideration. Tributaries that had cut
through the Edwards before valley widening, were served by lesser tributaries that also
participated in the episode of valley widening (Brown, 1988).

Thick, caliche-cemented, colluvial wedges occur at the toes of slopes of almost all uplands in the
Cut Plain (Brotherton, 1978), clearly indicating that slope retreat is at present extremely slow
or inactive (Fig. 5-3). These colluvial wedges--called Cranfils "soils"--vary from thin veneers to
ten feet or more in thickness, and they join slope bases to the pediplain. Since their
emplacement the toe of the slope has remained fixed in position. The only alteration of slopes
has been "rounding-over" of the upper slopes, which supplied colluvium to Cranfils soils. Thus
the Cranfils soil represents the most recent episode of slope modification. While it is caliche-
cemented, it lacks the thick petrocalcic caliche characteristic of some of the ancient Cut Plain
soils, and thus may represent a fairly recent addition, though the actual age is as yet unknown.
(Flowers, 1987; Brown, 1988)

The evidence at Stop 6 therefore suggests that valley formation followed the pattern
established at Stops 1 through 5; (1) initial downcuttlng extended through the Edwards
limestone, and into the Comanche Peak (2) entrenchment ended as ancestral Meridian Creek
reached grade here at an elevation of about 940 feet; (3) the ancestral streams remained at
that level while slope retreat led to the formation of pediplains and attendant divides; and (4)
far more recently, base-level lowering on the BrazOS-Bosque system, in a number of episodes of
entrenchment and alluviation, created the present terraced inner valley of Meridian Creek (Fig.
6-2).

STOP 7--LANHAM ROAD CUT AND OVERLOOK--Fredericksburg stratigraphy, Fairy


Gap, and the Bosque-Leon Divide (31 0 45' 25 "N; 97 0 55' 55" W; Fairy and
Cutoff Mountain quadrangles, 1956, 1124,000) (Fig. 7-1)

At this Stop there are two points of emphasis, (1) the roadcut, the stratigraphy which has
shaped Cut Plain evolution; and (2) the view, the geomorphic significance of the valley
configuration of this very large basin.

THE ROAD CUT

Fredericksburg Rocks and the Cut Plain


We have referred to the Fredericksburg rocks simply as Paluxy, Walnut, Comanche Peak, and
Edwards Formations (Fig 0-9). Those terms are sufficient to understand the main features of
the Cut Plain landscape, but variations in rock type within each formation, and lateral gradations

I-56
between formations, affect the second order of landscape features (Fig. 7-2, 7-3, 7-4, 7-5).
Rocks are important to landscape development because of their relative resistance to erosion,
permeability, and susceptibility to various soil-forming processes. Lateral and vertical changes in
rock type within the Fredericksburg Division determine their properties as grist for the
weathering and erosion mill.

The Lampasas Cut Plain also was formed where Edwards Limestone forms less than half of the
Fredericksburg thickness. This is significant, because soft beds under the resistant Edwards
Limestone are just as important to the physiographic process as is the limestone cap itself. We
saw at the first few stops that streams that had incised into the Edwards, but not far below,
either did not participate in the valley widening process, or if they did, the widening occurred in
soft material above the Edwards and the diagnostic landscape has not been recognized.

Paleotopography Figure 7-2 (compare to Fig. 0-9) shows a slice through a very broad
depostional basin (vertical exaggeration; 1000:1), a shallow extension of the East Texas Basin,
overlying and expanding the earlier Trinity Shelf (Fig. 0-7). Fredericksburg rocks record the
episodic northeastward spread of shallow water carbonate sediment from the San Marcos Arch.
Rudistid debris, oolitic and skeletal grainstones, and peritidal dolomites are collectively named
"Edwards Formation". Figure 7-3 illustrates dominant lithofacies near the end of Edwards
Deposition. Earlier, terrigenous sand (Paluxy Formation) spread into the shallow basin from the
northeast and northwest (Atlee, 1962; Owen, 1978). and terrigenous clay (Walnut Formation)
was introduced from the northeast. The terrigenous material also came in episodically so that
the rocks record a complex interplay of exotic terrigenous sediment with /n situ carbonate
sediment. Fine grained carbonate muds became the nodular, white limestone beds of the
Walnut and Comanche Peak formations (Fig. 7-4).

Where Edwards Umestone is 200-300 feet thick, the physiographic expression is the Edwards
Plateau. Where the basal terrigenous section is thick, the result is the Western Cross Timbers
(Fig. 0-4). In between, where 150-30 feet of resistant Edwards rocks are preserved, we find
the Lampasas Cut Plain. Edwards lithologies thin northward; beyond the pinchout edge, the
Fredericksburg and Washita rocks together form the Fort Worth Prairie.

Not all Edwards rock types are resistant to erosion. Soft, pure, almost uncemented dolomite
was formed on or near tidal flats southwest of the late Fredericksburg grainstone barrier (Fig. 7-
3). We suspect that this material formerly occurred widely under Washita strata across the
San Marcos Arch and far to the west. But we do not know, because if it existed it has been
eroded away completely. Clearly recognizable beds and surfaces can be traced northward
through the middle part of the Tredericksburg (Fig. 7-5a). by tracing these surfaces we do
know that the top of the Edwards to the west is older than the top of the Edwards to the east,
thus we surmise that 50 to 100 feet of Fredericksburg material has been eroded away, and that
the material was soft, peritidal dolomite.

Edwards/Comanche Peak Lithologies Stop 7 provides an opportunity to examine coarse


rUdistid-fragment rock of the Edwards, overlain by deep soil of the Washita Prairie formed in
Klamichi and Duck Creek (lower Georgetown) shale and nodular limestone (Fig. 7-5). The road
cut below the Edwards is in material mapped as Comanche Peak Limestone. Nearly half of the
unit exposed is shale and marl, but the convention is to use the dominant lithology found in
normal outcrops. The unit exposed was termed "Pancake Member" by Amsbury (1988). When
the correlative stratigraphic position can be seen again to the south across Leon Valley at
Evant, most of the unit is represented by Edwards reef rock. A series of detailed measured
sections down the north side of the Leon Valley demonstrated that the shale, marl, and nodular
limestone go to Edwards rock bed-by-bed, that is, deposition was episodic and limestone
formation was punctuated by innumerable interruptions that left shale partings and beds.

I-57
\

X//.36
...\••
o
+ 0"45

...
&
~
"' .
" .

Figure 7-7 (Fairy Quadrangle [7956]7124,000) Stop 7, on the eastern, Edwards-capped


divide of the Leon River Basin, looking northwestward along the divide to Fairy Gap, headwaters
divide of Meridian Creek Basin. The Pediplain is especially well preserved here, and Meridian
Creek headwaters divide is in a valley, as it was at Stop 5 (Turkey Creek Divide), though
alluvial soil is not present Here again the Pediplain is graded from the divides, but clearly not
to present drainage,
1-58

MOFFATT
LENTIL GEORGETOWN FMN
CDUCK CREEK MBR)

--\
;s--
~
'-
r'\
<.,
)
T~a.
(!)
II: :J
----
(EDWARDS A)
GEORGETOWN FMN
PERSON FMNbCDUCK CREEK MBA!..
t
EDWARDS
FMN
I" ,
')10
EDWARDS FMN
KIAMICHHII~F~M~N!....:=;;j;:::
- 7f -r
GOODLAND (!)
T
~ o
~;a.ffi
:J KAINER FMN _ _ COMANCHE PEAK
"
::,::a. ffi
~ m (EDWARDS B) I FMN
-..1',
--<- '"
If)
r
COMANCHE
FREDERICKSBURG DIVISION
~ O:J If)
tu

~ fil
WALNUTJJ 5:J
a:~ PEAK
V 'VI
@
J

i'> ,FMN UPPER FMNJJ _0 ::;:


~(!) c( ___ _ MARL ... ..;.... 0:0:
v 0
MBR
.; ... " ... .., ........................... _ ... - W(!)O0
e "- M! :::
--------- ~ ~l
KEYSVALLE'I'---j- __ _ .......... _w ...........
"-
u. ftl ~ ---
'>-" MARL WALNUT PALUXY
FMN II:

11
r, CEDAR PARK LS FMN UPPER pAL
,
~ U.

1-
~
V>
-0 BEE CAVE MAR LOWER PALUXY
~ ~ -=
~ ___-!B~U~L~L~C~R!E!E!K~L!S--~~~
.,,)
-< GLEN ROSE FMN ,. 100 KILOMETERS-I GLEN ROSE FMN
~

~
~
<:'\
~
~

"- Figure 7-2 South to north cross-section along the eastern edge of the central Texas

0 ----
~ Fredericksburg outcrop from San Marcos to Fort Worth. The region of the field trip lies between
the Moffatt Lentil and the right-hand gap in the section. Stop 7 is on the thin Edwards
'i':';.... ~,

"-
Limestone plus most of the Comanche Peak Formation. near that gap. Note that the
Fredericksburg Group on the San Marcos Arch, where the Person Group built up above the
\"- Kiamichi Shale. is not equivalent to the Fredericksburg Group west of Fort Worth. where the tOJr
Glen Rose bounding surface is lost in Paluxy Sand. Throughout most of the Cut Plain. the
Fredericksburg Group (rock-stratigraphic) and Division (allostratigraphic) are coincident.
'"
"\'-
[Diagram based on detailed surface studies by Rose (1972). Moore (1961). 1964). Young
(1974a.b). Amsbury (1995). and Perkins (1961). Modified from Lozo (1959. Rg. 1) and Rose
(1972. Figs. 19.33).
~ 7
- --~
~l
.-. \'V'I '
I,
\.;
.•
'
'•.
•o
o
~

'0
~
:

.
~~'
-ID .'
'(j

'. · .-
··- • " .::
"
Q
'..
/
-'

--'
.~

.~

,,,
~'i

i.~
\~
... ......·..r::.
i'i
...• \
\.-

1-60
Figure 7-2 Generalized Facies Map of the Edwards Formation. The Edwards
. Formation within the field trip area represents 1/3 or less of Fredericksburg
deposition. Growth of rudist reefs and grainstone bodies began along the northeastern
flank of the San Marcos Arch, far to the south, and progressed episodically into the trip
~
area. A linear body of oolitic, skeletal, and pelletal grainstone formed at and
northwestward from Lake Belton, growing to 130 feet or so thick. Peritidal dolomite
'"
~

formed southwestward from the graInstone, while northeastward from it, open· marine
deposition of lime mudstone, marl, and calcareous clay conlinued until near the end of
Fredericksburg deposllion. The final stage was the extensive - and presumably rapid -
spread of rudist patch reels, associated lagoonal deposits, and grainstone-filled channels
throughout the area.

Klamlchl Shale of the overlying Washita Division pinches out southward onto the
grainstone body. Note that Impressive Incised meanders are preserved in the grainstone
body now hosting Lake Belton. Similar IncIsed meanders along the Leon River are Incised
In Glen Rose rocks between Gatesville and Hamilton, and are preserved In floodplain
deposits that probably date to latest Wisconsin/earliest Holocene throughout the Leon
Valley (Amsbury, 1984).
. ·~.I;'brvWI~ IIUC .....ItIlU . . PS(.WS~ .-

§,
........ _1Id1111_1U. UI&Il _ CIMI l1li111.1
'tPJ,...,. Ir--.&. ~ &I... '
"ttlh"'~, .. Ita 11_"'5 IfId cI-Wllllle "5.
'.'Lo;'F .•.

~:
• IT"'. ".17El.
12.8 I I
I I
i

~
I
jW"'-, "' ........,....._ ... 1'... 5 .. '" r........
i I
""~~', _\111111'"1",1111_ WS',
!

-
rNhlta.

~ .ss
I,
,
- ILn'...... ,1 .
us ~ ~':"'" MS,."
s. INII';-

,,' ,
.. !
-- I-
lo.~
c.-...
._--
~
'.... _'1-' ..... !"OI WS.
I
- II_PS.

!~PS.~ ~WHall~
• ~'I"'-I'IS. \

=;:,':.::::f
~~ -t-·1·":"I-;
Ii"~!;

.-i '''r
."',.T"'·
15:8 !"-"

I
us

- ----;.~
i

1"..:.;;-,
, ..
.~....
f-'--~ 'IS. ',.. i
_II .....
....,
I

,. ,......
U_WS/PS ..

,
,
, '5.
-
--If--+- .. -. T
!
.!
.----.. -.----.1
-z., I'I~.II""S.
I
.
• -
·Z..j~~;1 . _ - - ,
;
..WI" ~ . . . t.-Ilus.II~- tMll...,.lllla.
- 'j-- ---:
I

Figure 7-4 Measured section of the Edwards Formation and part of the Comanche Peak
formation (Pancake member of Amsbury, 1995) along a gravel road about five miles
I, ."-.J..

----
.;u/" i
:

southeastward from Stop 7, The top of the fairly resistant, white, nodular limestone (Cold
Spring Limestone of Amsbury, 1995) at the base of the section is about 60 feet above the top
of the Keys Valley Marl member of the Walnut Formation, and about 200 feet above the top of
the Glen Rose Formation, In the section, ms=mudstone, ps=packstone, ws=wackestone,
gs=grainstone, bs=boundstone (from Amsbury field notes, 1963),

1-62
GRAINSTONE

._------
--_ .. _--:------ ------ --- -- --- --- ... - ... -
- - - --
- - ---:..- ---------
---- ~OIEPEJUt------~
-- -- -- -~ - -- - -~ - :------
-- ~
-------~--

Figure 7-5a Northwest-Southeast Cross Section Along the Colorado/Brazos, Lampasas/Leon,


and Cowhouse Creek/Leon Divides. The cross-section cuts obliquely through the Edwards
grainstone body near Lake Belton, extending north westwards into the peritidal dolomite facies.
Scattered patch reefs, particularly at or near the local base of the Edwards Formation, occur
along this line. Correlations based on recognizable surfaces and key beds underneath Edwards
rocks are shown by light dashed lines. The actual amount of peritidal dolomite that once capped
the divide is impossible to determine, but by projection of the stratal top from the southeast it
is evident that a significant amount of Fredericksburg-equivalent material has been removed by
erosion (from Amsbury, 1995).

SW
S"rRATAL = EIlHl\RDS G __
.B..-s::: _______ _
CXIlOOI~~=§~EIl~RARD~~S~
~s
___ _

--.--
~o~",,~tE~~~
----'N~
-,..;:'00 1..
__
REEF
... ~.1
- - -- -
-~--
-- - -
.
---~~: ;,..'-=='---=~- ~ =:-.-==
- - - - - - - - - ' - - - - -- --- - -",'"

Figure 7-5b Southwest-Northeast Cross Section Across the Lampasas/Leon and Leon/Bosque
Divides. The stratal position of basal Edwards rocks to the southwest compared to that to the
northeast is shown by light, dashed correlation lines. There are no upper Fredericksburg
outcrops within the Leon Valley, of course, but the interpretation Is controlled by detailed
outcrop and core data near Lake Belton. The cross-sections have a common point at Shive, in
extreme northwestern Coryell County (from Amsbury, 1995)

1-63
Distinctive thin beds, surfaces scoured by submarine{?) action and then bored by rock-boring
clams, and clayey beds are most welcome to the stratigrapherl

Evidently conditions were favorable for molluscan growth during deposition of the Pancake
Member near Lanham. Ammonites, clams, marine snails, echinoids, marine oysters, calcareous
algae, and burrowing organisms are extremely abundant in many beds.

The relatively resistant white, nodular limestone bed at the base of the section was termed
"Cold Spring Limestone" by Amsbury (1988). This unit is a persistently traceable tongue of
Comanche Peak lithology that underlies the muddier Pancake Member, and overlies the Upper
Shale Member of the Walnut Formation (probably the Marysville Marl of Perkins [1961)).

Fredericksburg Stratigraphy and the Comanche Pedlplain • Along the lower slope of
the Leon Valley, but northeast of Highway 36, we see a persistent line of knolls at about the
same elevation along the valley. Slopes cut into the Upper Walnut Shale are graded either to the
tops of the knolls, or to lower elevations between them. Slopes below the crest of the knolls are
graded to a second, lower base level near high terraces along the Leon River. The base level of
the knolls defines a valley surface about 100 feet above the present flood plain of the Leon
River; in this reach of the Leon, the high base level defines an intermediate surface, the
Comanche Pedlplain.

The knolls along Leon Valley are formed in a relatively resistant, massive accumulation of
gryphaeid oyster shells at the top of the Keys Valley Marl Member of the Walnut Formation
(Moore, 1964; Flatt, 1976). Lower slopes are cut in Keys Valley marl and shale, and in the
Paluxy Sand. Here, just as in the Upper WalnuVComanche Peak/Edwards triplet, a resistant bed
over less-resistant beds provide the material for a distinctive landform.

The fact that the Comanche Pediplain in this part of the Cut Plain is coincident with a resistant
unit at mid-level within the valley caused us to wonder if in fact the "Intermediate Surface" was
a figment of our imagination. Perhaps the "Surface" is merely a set of rock-defended benches at
approximately the same position within the valley, and does not record a discrete period of
erosional history. Further work, though, led to the recognition of the Comanche Pediplain is on
lower and lower stratigraphic units westward, and higher ones eastward. Brown 1988) mapped
the pediplain throughout the Leon Valley. Flowers (1987) extended recognition along the upper
Colorado River and into the upper Concho basin. And Amsbury recognized the pediplain in
Blum's (1989) Terraces A and B in the upper Pedernales River during the 1989 Friends of the
Pleistocene trip. We think now that correlative landscapes occur along major streams draining
the southern Edwards Plateau, marked by Hill's "Uvalde Gravel" (Hill, 1900, p. 5) along Median
River, Seco Creek, Sabinal River, Nueces and West Nueces rivers, and the Rio Grande at Del Rio.
Similar surfaces perhaps formed in the same way and at the same time occur in the Pecos Valley
(Mescalero Plain, and including the surface that extends southwestward through Fort Stockton
Into the Davis Mountains), and the Rio Grande Valley above EI Paso (the La Mesa Surface)
(Amsbury and Hayward, 1996).

THE LANDSCAPE

The view northwest from the overlook at the top of the road cut (Stop 7) is of the divides and
pedlplain which separate Meridian Creek drainage (North Bosque River) from Egg Creek (Leon
River) drainage. The crossroads about one mile west of our viewpoint is the small community of
Lanham (a whole community of Joneses, descendants of the original Jones whose wagon broke
down here in the late 1870's). Visible farther to the west is the recent inner valley of the Leon
River that has greatly dissected the Comanche Pediplain, (though the pediplain is still visible in
the accordant summits southwest of the River). The far Edwards-capped divide of the Leon
Basin is 50 miles west, and not visible from this point (Fig. 7-6).

\-64
Probably the most impressive aspect here is the immensity of the view to the northwest. The
view is a product of two factors; (1) the major dissection caused by the Leon River, and (2) the
effect of structural dip. Because the dip is slightly greater than the mean gradient of the Leon
River system the Edwards-capped divides rise to the west- northwest, and the total relief
increases. Thus, entrenchment first encountered Edwards Limestone far to the northwest,
where it created the first entrenched valley. As general landscape lowering exposed Edwards
Limestone down dip to the east-southeast, entrenchment also progressed to the east
southeast, while upstream the older valley widened by slope retreat, eventually creating the
landforms we see here (Fig. 7-6).

The most recent entrenchment into the Edwards Limestone is at Belton, Bell County, 50 miles
to the south-southeast. The northernmost Edwards-capped western divide is Stop 9, in Brown
County, 50 miles to the west, at the end of today's traverse. The northernmost Edwards-
capped eastern divide is Long Point, visible on the skyline about eight miles northwest of Stop
7. Thus in map view, the Edwards-capped divides form a somewhat distorted V, 100 miles
"high," 50 miles wide, converging downstream to the southeast (Fig. 7-6).

Probably the greatest significance of this view, and of the broader picture which this view
emphasizes, is that ancestral Leon River entrenchment may have started in the northwest
before dissection first encountered the Edwards Limestone. Since that time, entrenchment has
lowered the course of the Leon River 840' below the northernmost divide elevation.

During that same period of entrenchment, slope retreat created a basin 50 miles wide at the
northern part of Edwards capped divide. Downstream at Belton, general surface lowering has
just now exposed Edwards Limestone to active erosion.

Throughout much of the length of both western and eastern divides, a pediplain graded to an
ancestral Leon River is yet preserved. The divide profiles suggest that almost all the slope
retreat, which now accounts for this very wide basin, was the process of formation of the
pediplain. This again suggests that the episode of slope retreat, introduced at Stops 4 to 6,
was the product of a major event (or events) of slope retreat, and that little has happened
since the time of that event (events?) in these old divide lands.

Along major, through-flowing, trunk streams, such as the Brazos and Leon Rivers the Time Zero
Surface first began to be destroyed (and the Cut Plain formed) in the far northwest of
Fredericksburg presence. With continued erosion, Cut Plain formation progressed
southeastward. In contrast. along tributaries such as the North Bosque. Hog Creek. and the
Middle Bosque the direction was reversed. as nick-point migration proceeded northwestward
from the incising trunk streams. Over most of the Cut Plain. entrenching streams had achieved
grade before the pediplaination phase. It was during their time-at-grade that the Comanche
Pediplain formed.

A point of frequent emphasis is the contrast between the widths of the ancient valleys and the
widths of inner entrenchments. Slope retreat in the inner valley of the Leon River rarely
exceeds one mile. The older. pediplain valley is as much as fifty miles wide where the Edwards
capped divides are yet preserved. The width of the inner entrenchment represents the effect of
slope retreat since entrenchment began in mid-Pleistocene time (Tharp. 1988). How much
time. or what rate of slope retreat. is represented by the very wide valleys of the ancient Cut
Plain? The answers which come first to mind are (1) that these wide valleys represent an
amount of time perhaps far greater than all of early Pleistocene. or (2) that the process of
valley widening in the "old" Cut Plain was radically different in mechanisms and rate than that 0
the recent entrenchments (Tharp, 1988).

1-65
Margin of Cretaceous rocks I
\'..- 50

'"
a
,-V ""''''''/r'\.. \ \..- -
t....-,
_./
" \
....
' - -
Miles
- - - - - '

''"''--, \
I
~

STOP 11 / ( ~E:.-_-STOP 7
"-\1'
,
\

"-
""'\
\: L.ampasas Cut Plain

Figure 7-6 Sketch map showing the downstream 'V' of the Leon Basin Edwards-capped
divides. Stop 7, where we now stand, is near the northern limit of the eastern divide. The point
marked Stop 11 (from an earlier trip) is at the end of the western divide. At Stop 17, the last
stop of today's trip, the Edwards Limestone plunges into the subsurface. Far the 'oldest"
landscape of the Leon Basin is that near the point shown as Stop 11 (from Hayward, Allen,
Amsbury, 1990), where total relief from divide to Colorado and Leon Rivers exceeds 800 feet.

1-66
Edwards-capped dIvIdes

~~Jl!~'ltt~

Hamilton ."~
l:'
~
cr·
~ :.>'"
i:')o ...--

'- ~
• Comenche
mc~
~ ~
'1fQ)
o, 10 20 30 40 50

Miles

. Figure 7-7 The Comanche Pediplain in the Leon River Basin at the final stage of its formation (after Brown, 1988). The
clearly pediment-like surface shown here was reconstructed from extensive Pediplain remnants preserved throughout the
Basin. Adjacent basins, even small ones, also preserve such remnants, each graded to an ancestral tributary in that individual
basin. The common base level for all Pediplain streams was the ancestral Brazos River. Using methods similar to those for the
Leon Basin, it should now be possible to reconstruct the entire Pediplain surface of the Lampasas Cut Plain as of the time of
its formation (from Brown, 1988).
Initial Cut Plain

f'(~~---~--
- - -"-'. '"'-= ......
y W"h'" P~""

=---
~-
.-- ------- ---'~--= II.

Expanding Cut Plain STAGE 1

~ ';,~..--_ R. W..h'"

-";. , ----=-'
---. - - ~"- .-: :.----
?-<
_ _ • .-..." ,''1

-~~
Prairie

~"ry,.'-!.b ~ ~ - - - -----
:/-~
~.)..l'
- '--~;:::?~~3£~~
~
'.-l.-).----
__ ~

~~
STAGE 2

STOP 11--=~

:;~~1:~~STOP 17
"I.co" II.

STAGE 3

Colorado R.

Figure 7-8 A series of block diagrams showing stages in the evolution of the Lampasas Cut
Plain in the Leon Basin. Initial dissection began in the north (see "Stop II", Fig. 7·6), and over
time extended southward. Following entrenchment, the river flowed near grade for an extended
period while slope retreat continued unabated, forming the vel}' wide Comanche Pediplain.
Renewed downcutting by the Leon River in response to climatic variations, or to base level
changes in the Blacklands to the east (near Belton, Stop 17) led to the formation of a "new"
inner valley. Renewed downcutting by the Leon River and its tributaries in Late Pleistocene-
Recent time has modified the landscape, leaving the Pediplain as an abandoned high surface
(from Hayward, Allen, Amsbul}', 1990)

1-68
Clearly the Leon River is an old system, from earliest Cut Plain formation established generally in
its present position. Near the northern limit of the Cut Plain at the time of first encounter with
the Edwards Limestone, it flowed at an elevation at least 840 feet above its present grade.
Since that time, the river entrenched first to the level of the pediplain, and apparently flowed at
grade at that elevation for a period of time adequate to form pediplain slopes as much as 30
miles wide. Then in more recent times, it entrenched again, dissecting the older surface in an
involved recent history that included a number of cycles of entrenchment and aggradation,
eventually to form the terraced inner valley of the present Leon River.

The northwesternmost Edwards-capped divide of the Leon River Basin is the southeastern-
most terminus of the Callahan Divide, the mesa-like divide which separates Colorado and Brazos
river drainage in the upper reaches of both stream systems. It also marks the boundary
between two structural provinces; (1) to the west a Callahan Divide province where the dip in

Cretaceous rock ranges from two-to-five ftlmi to the east, and (2) from Brownwood eastward,
the westernmost influence of the East Texas Basin, where the eastward dip increases to about
ten ftlmi, and continues to increase to the eastward, where Comanchean rocks disappear into
the East Texas Basin (Fig. 0-2).

There is a remnant pediplain along the Callahan Divide as well, equivalent to that of the Cut
Plain, indicating that the episode of pediplaination was late, involved a very wide region, and
that it affected both structural provinces at the same time.

This view therefore introduces a series of questions. Is the landscape we see a product of slow
evolution over a very long period, under processes not greatly different from those now active?
Or is it a product of rapid evolution over a much· shorter time-span, and under circumstances
radically different from those now active? The discussion of groundwater effects, and of
Fredericksburg stratigraphy suggest that it may have been much more rapid than present
processes might indicate. Only very recently have we began to appreciate the enormous
potential of groundwater as a geomorphic agent In the Cut Plain (Allen, 1990). Of equal
Importance, the similar landscapes cited in this guide (In Texas, the Callahan Divide, parts of the
Caprock Escarpment; in New Mexico, the Canadian River Canyon, Ute Creek; in western
Oklahoma; in southern and western Kansas; and perhaps in northern Mexico) have groundwater
and landform relationships similar to those postulated for the Cut Plain.

STOP So-CORYELL CREEK-Scarp retreat and pediplain formation (31 0 2S' 37"
N; 93 0 37' 50" W; Gatesville East and Oglesby quadrangle [1979], 1/24,000)
(FlgS-1)

The view west from this point is of the valley of Coryell Creek, a minor tributary of the Leon
River. A much smaller, unnamed tributary enters Coryell Creek from our right (north) in the
Immediate foreground. It is separated from the main stem of Coryell Creek by an Edwards
capped divide, and on the east facing margin of this divide are a number of scarp face
drain ways, in various stages of dissection.

This view encompasses the two points for our discussion; (1) the Comanche Pediplain in
tributary valleys as a key to pediplain evolution; and (2) the role of groundwater In scarp
recession, as a model for the formation of the Cut Plain.

1-69
lJ

-
.o:...

- -- --... '.- '.


Figure 8-1 (Gatesville East Quadrangle [1979] 1124,000) Stop 8, overlooking Coryell Creek Valley, and a small
tributary valley, both of which show well developed Comanche Pediplain remnants. That in the tributary valley is almost

\
Cundissected, indicating that during the pedimentation phase, even the smallest tributaries participated. This suggests that
that the baselevel remained remarkably stable for an extended period, or that during the pedimentation episode slope retreat
was unusually rapid
v
( 06J) /. .~ .. ;;.~ \ I( ~ )"\ ~ 2~\j "( ~0~\:l \,\%~
Pedlplain in tributary valleys
The cross-valley profile of Coryell Creek shows clearly the Edwards/Georgetown armored
uplands, the rounded slope to the pediplain, the pediplain, and the recent entrenchment of
Coryell Creek into the pediplain. More surprisingly, the unnamed tributary just to our right also
shows the same features, though the basin area is only about 0.25 sq mi, clearly indicating that
basin area was not a factor in pediplain formation (Fig. 2-4).

Pedimentation along Coryell Creek valley and into its most minor tributary valleys proceeded in
response to entrenchment and valley widening by pedimentation along the Leon. That in turn
was in response to entrenchment and valley widening by pedimentation along the Brazos. The
ultimate cause for entrenchment of the major trunk streams is not yet known, nor is the actual
time of this earliest dissection. But it took place over a sufficient period of time, and under
such circumstances of scarp retreat that most of even the minor tributary valleys took on the
"traditional" Cut Plain profile. Here again, landform configuration indicates that groundwater
availability was a major control (Table II, Stop 4), a point to be considered in some detail here at
this Stop 8.

To belabor a point, through-flowing trunk streams were local baselevels for the Comanche
Pediplain, as they are for much of the present Cut Plain. Because of their volumes of flow,
during initial phases of cut plain formation they rapidly cut downward. As nick points migrated
upstream, each upstream tributary in turn felt the effect of trunk stream entrenchment, and
responded by downcutting. As grade was reached and maintained for an extended period in the
I main stem, a steeper graded profile was established for each of the tributaries, and along all
channels pedimentation proceeded by slope retreat, the rates apparently determined largely by
groundwater availability.

I I Again we emphasize, the pedimentation phase persisted long enough that it eventually involved
all of the network that had achieved grade in Fredericksburg rocks, even the smallest
tributaries. The undulating, leaf-like landform which includes the combined pediments of all the
I Cut Plain basins is the Comanche Pediplain. It consists of a vast number of Individual basin
pediments, joined at common elevations only at stream junctions, and all ultimately graded to a
distant base level in younger rocks far east of the Cut Plain.

Groundwater and scarp recession


The role of groundwater in tributary formation, scarp recession, and formation of individual
basin-pediments has been discussed (Stops 2, 3, 4). Here at Stop 8 the effects can actually be
seen. In the Washita Prairie (the uplands to the north of the road) the Georgetown and Edwards
limestones form an unconfined aquifer with a water table 20 It below the surface. Two types of
porosity are present in the aquifer, (1) matrix, and (2) fracture. Fracture porosity is far the
largest contributor to effective porosity (Cannata, 1988, p. 64), with values ranging from .51%
to 1.49%, and an average value of 1%. Solutioning has greatly increased the effective porosity
in some areas, and zones of highly effective vertical solutioning along fractures have been noted
(Cannata, 1988, p. 64). Where fracture porosity has been increased by solUtion, permeabilities
of as much as 22 Darcys occur, equivalent to hydraulic conductivity of .0007 fVsec, or an
actual flow rate of as much as 60 fVday at existing head differences.

This is of critical importance, for the values cited here are (for the long history of Cut Plain
formation) minimum values. Any increase in precipitation, reduction in evapotranspiration, or
increase in permeability (through original presence of more permeable beds, since removed by
erosion) would increase groundwater flow, and hence scarp retreat. At the flow now existing,
little is happening in the Cut Plain. But once a critical threshold value of groundwater flow is
achieved, scarp retreat rate should increase markedly.

1-71
,..
.! EAST
WEST a;
>
~
co
..,-; Spring line
~ :;r-t:'- _ ...
t::::>

- -
. .. ,. . . . . __
EdwarGs..,.. --=~--:-.....~ ~
'......
~"
....-.~.""
',
~
.
.... ~
~ ~.~~~s::Jijjr-=
Camanche Peak \

Comanche Pediplain Comanche Pediplaln

Recent entrenchment and aliuviation


I

Figure 8-2 Diagrammatic section and profile of Coryell Creek Valley, showing the Edwards-
capped divides; the Comanche Pediplain in both Coryell Creek Valley and the small tributary
valley to the east; the inner entrenchment of Coryell Creek, and the spring line that marks the
slope break in the tributary valley. The presence of the Comanche Pediplain in the tributary
valley indicates that basin area was not a limiting condition in pedimentation (from Hayward,
Allen, Amsbury, 1990). .

WEST EAST
o, 2,
Miles

Local flow system

Figure 8-3 Section showing local, intermediate and regional flow systems, Middle Bosque River
basin. Notice that regional flow systems most responsible for stream incision are confined to
the Washita Prairie east of the entrenched valley of the North Bosque River (from Cannata,
1988).

1-72
W= wldlh of area contributing flow to pipe

--W-i
II I I
II I I
II I I
I I I I
I I I I
II I I
I I I I Flaw lines
I I I I
III I
¥ pip.
I
pip.
valley deposits

Gully A Gully B Gully C

Figure 8·4 Diagram showing relation between width of the contributing area of groundwater
discharge to a stream, and stream length (after Selby, 1982). Notice that there is a cascade
effect. As the stream lengthens, the contributing basin increases in area. As the basin
increases in area, the rate of elongation increases. This sequence, continuing to the divide,
appears to explain the evolution of dip·controlled drainage in the Cut Plain (from Hayward, Allen,
Amsbury, 1990).

3~----------~------------~------~~--r
/'
.//


o 1 2 3
Aquifer length
Figure 8·5 AnaJysis of a small sample (12) of first·order streams In the Cut Plain indicates
that stream length (SL= stream head to junction with next higher-order stream) is directly
related to aquifer length (AL= stream head to divide) above the stream head. The effect is to
form first order tributaries which head some distance (AL) from the divides. The 'undissected"
divide area is, none-the-less, a major contributing basin, accounting for headward migration of
tributaries (from Hayward, Allen, Amsbury, 1990).

1-73
Flow systems in the Washita Prairie and Cut Plain are controlled by structural dip and local
topography. For example, local, intermediate, and deep flow systems occur within the shallow
aquifer system of the Middle Bosque Basin (Fig. 8-3) (Cannata, 1988), and have been
recognized in diverse shallow Cretaceous aquifer systems throughout Central Texas (Barrett,
1989; Barquest 1989).

Based on the Middle Bosque Basin, reconstruction of the long term changes in the
groundwater system at a given locality in the Cut Plain can be reconstructed by comparing
groundwater behavior in less-dissected sections of the basin (Fig. 8-3), to its behavior in
increasingly dissected areas of the same basin, and assuming that these represent the sequence
of phases in drainage development as the landscape evolved. The result of these comparisons
is simply that with increasing dissection flow systems become ever more localized.

The dominant dip controls groundwater flow lines along major divides, but topography becomes
the dominant control in dissected areas (Fig. 8-4). Dissection of the Georgetown Limestone
results in localized reorientation of the groundwater flow net. The increased flow, directed by
lithology and structure, enhances local weathering and initiates the sapping process. This in
turn initiates headward elongation of the valley. While the tendency for the main stem of the
stream network to be directed down dip, many of the tributaries are oriented with local fracture
sets (N 20 0 to 40 0 E). As the main stem tributary continues to lengthen headward, its size and
length are continuously related to the increasing basin area of groundwater flow that it has
captured (Fig. 8-S). Within an individual basin, groundwater flow volume is related to (I) the
degree of dissection, (2) the saturated thickness of the aquifer system through which the
stream has cut, (3) the effective porosity and permeability of the aquifer, and (4) climate.
Within a given area these four factors remain relatively the same for all first order basins, and
dissection should reach an equilibrium dependent upon the local structure and stratigraphy.

Dissection acts to concentrate both groundwater and overland flow. Surface water acts to aid
in the removal of sapped debris, and the reduction of fall blocks by both mechanical breakdown
and chemical solution. Level of entrenchment is controlled by local base level, and the volume
of groundwater redirected into overland flow is controlled by degree of topographic dissection.

Assuming that this is a valid model for the systems that formed the Cut Plain, there should also
be a relationship between stream length and the area of captured aquifer (also expressed as
length of captured basin above the channel head). As a test of this hypothesis, a small sample
of small tributary basins in the larger Coryell Creek Basin were measured and compared (Fig. 8-
S), indicating a consistent relationship between channel length below the Edwards free face, and
contributing basin area above the Edwards free face. Of particular interest is the difference
between north and south flowing tributaries. North flowing (up dip) tributaries have shorter
lengths (and hence areally less extensive areas) of contributing aquifer than the south flowing
(down dip) tributaries, tending to confirm the greater water availability from such dip-controlled
aquifer systems

Note: Here we leap from stop 8 to stop 16, but there Is an explanation_ These
stops were origInally part of a two-day field trIp (Hayward, Allen, Amsbury,
1990). For the present trip that fIeld guide has here been reduced to a one-
day trip, and the Stops missing from the one-day trip (those Stops that took us
to the far western margin of the Cut Plain) have been deleted in favor of tIme,
wIth the convIction that one day in the Cut Plain is vastly more satisfying than
no days In the Cut Plain. In order to equate the stops of the earlier trip to
those of the present one we have chosen to maintain the original stop numbers_
We hope this is not too confusing_

1-74
STOP 16--LEON VALLEY OVERLOOK-High terraces of the river (310 18' 47" N;
970 29' 15" W; Leon Junction and Eagle Springs quadrangles [1979],
1/24,000) (Fig. 16·1)

Here at Stop 16, near the eastern margin of the Lampasas Cut Plain, the Leon River Is deeply
entrenched through Edwards and Comanche Peak rocks. The valley Is narrow and clearly
defined, and obviously a product of the stream which now occupies It (Fig. 16-1).

At this stop we consider only one topic, the history of the Leon River as we now Interpret It.

Here where we stand Is a bedload terrace of an ancestral Leon River. It rests on Edwards
Limestone 100 feet above the floodplain. The narrow, entrenched valley, so conspicuous here,
begins upstream near the junction with Coryell Creek about 9 miles northwest of our present
location. The valley remains entrenched almost to Belton, our last stop on this trip.

The entrenched valley Is widely meandering, in loops far larger than those of the present River.
It has long been believed that these valley meanders reflect on older, larger River (Lewand,
1969; Tharp, 1988), which once conducted much of the flow now conducted by the modern
Brazos. This ancestral Leon River, then conducting both the Clear Fork and Double Mountain
Fork, was beheaded above Eastland (Eastland County) leaving the beheaded river about as we
see it now.

The time of this diversion (Tharp, 1988) may have been Wlsconslnlan. The terrace gravels of
the Leon are composed almost entirely of quartzose components, not available from the
present basin, and similar to Ogallala deposits. l.:.imestones, so widely available In the modern
basin, are almost absent in the terraces, but they become dominant In the modern floodplain.
Thus the diversion must have taken place between the time of emplacement of the lowest
terrace, and the formation of the modern floodplain (Tharp, 1988).

Several other episodes of river history are also reflected in the landscape. On the highest divide
about four miles southwest of our present position (and on major divides of the Leon Basin
throughout the Cut Plain) are lags of large, rounded, exotic gravels, of varicolored quartzite,
quartz and chert, some to six inches in diameter. Two suites are present; (1) an exotic suite of
quartz, quartzite, and varicolored cherts (some containing fusilinids); and (2) a "local" suite of
Edwards derived chert. The exotic suite has petrology similar to Ogallala gravels far to the
northwest. The local suite, commonly more angular and abundant, was clearly derived from
Edwards limestone, which Is increasingly cherty to the south and to the west of our present
location .•

Below the high divides just south of our present position are other terrace lags similar In
composition, but at lower elevations. These also show the great abundance of siliceous gravels,
and do not show the dilution by limestone which might be expected if they represent reworking
of higher gravels (Fig. 16-2).

The distribution of these gravels between the highest divides and the third terrace suggests
that there may be several levels in that interval. There are few areas where these occur, and to
date information Is inadequate to define additional terrace levels. The area of their occurrence
Is similar in topography to much of the divide region downstream from Gatesville, and it is
possible that other localities will eventually help to provide answers to such questions (Tharp,
1988).Thls overlook Is on the third terrace level, and it is 200 feet below the nearest gravel
lags of the highest divide. A little less than five miles west of our present position an

1-75
- -C,'"'' \
~--~~~~--'" ",\~
_/

(}ff'~ ,: ' l .;
" ,
" I
,-
"
,
,
""
"
rEI' ,;
"
"
'.
s;; ,/ " "
~"
,.,-
(/-2~;';' -,?-~~~

Figure 16-1 (Eagle Springs Quadrangle, 1124,000; 1978) Stop 16, overlook of the Leon
Valley at Mother Neff Park, with emphasis on the third terrace (100' terrace) and its relation to
the Comanche Pediplain. The terraced inner valley of the Leon River is clearly visible, and the
relation of terraces to drainage history is evident. The third terrace, Stop 16, is the age of the
Comanche Pediplain, and antedates the incision of the valley before us, and of the Pediplain
proper.
l-76
Coarse siliceous gravel on hIghest dIvIdes

other terraces

c~7:e Pedlplaln
Georgetown 100 " terrace

;:;;
I ) i •
Edward~';'
; J'!
I I I
I
I
I
I
I
I
I
I
i .~
FQ
-- ---
I~ 60 " terrace
' l ( I 1 , 1 i i
,
-
.:.,
-..l
Comanche Peak
30 ")erraCe FloodplaIn

'le ., £dS
" : '-', . ,"
/• .. .:....

Figure 16-2 Diagrammatic profile and section, showing the relationships of terraces to landform and stratigraphy.
Terraces are best preserved where they are rock defended, and higher terraces are almost limited to rock defended localities.
Cross profiles of the Leon Valley here, and particularly in the Lake Belton area just downstream from our present locality,
strongly suggest that the entrenched valley was the product of rapid (?) knickpoint migration initiated by baselevel lowering
along the Brazos River (from Hayward, Allen, Amsbury, 1990).
abandoned meander, 100 feet above the present floodplain as is our present position, also
exposes similar gravel. Upstream for many miles, at an elevation of about 100 feet above the
present floodplain are found siliceous gravels and old alluvial soils of the Krum series (McCaleb,
1985) ..

It is to this third terrace level that the Comanche Pediplain of the Leon Basin was graded. The
Leon River controlled the evolution of the early Cut Plain with its extensive pediplain. Even then
the Leon River was a stream, at grade, in about its present position, widely meandering on an
alluvial veneer (Fig. 17·7).

The inner entrenched valley into which we look was apparently the product of knickpoint
migration along a meandering valley extending northwest from near Belton. Knickpoints were
apparently generated by base level changes in the Brazos River system in the Blacklands to the
east. Even this last incision was a product of a number of events reflected in a sequence of
terraces in the inner valley.

Dissection by knickpoint migration appears to have been the mechanism of valley deepening
below the third terrace, because the entrenched meanders of the Leon, particularly in the reach
between Gatesville and Belton, are generally symmetrical in cross section. While meander necks
in softer rocks upstream, such as Twin Mountains Sand, give the clear impression of Slip off
slopes, in the lower river where the valley is walled in harder rock most meander necks exhibit
near vertical faces and cross valley profiles are generally symmetrical. Furthermore, valley
meander geometry is that of free meanders of a much larger river, and none of the preserved
meanders exhibit the exaggerated meander shape so often encountered in deeply ingrown
meanders. The mechanism of earliest entrenchment (that which initiated the development of
the Cut Plain) is of course unknown. The presence of possible terraces between the divide
gravels and the third terrace suggests that this entrenchment may have been the product of
staged incision as well (Tharp, 1988).

STOP H--BEL TON DAMSITE--The eastern margin of the Cut Plaln--(31° OS' 43"
N; 97 0 28' 17" W; Moffat and Belton quadrangles [1974], 1/24,000) (Fig. 17-
1)

Here at stop 17, we stand at the eastern margin of the Cut Plain, overlooking the Texas
Blacklands to the East. The Edwards Limestone plunges steeply eastward into the subsurface at
the western margin of the East Texas Basin (marked by the Balcones Fault Zone). At Stop 17
the top of the Edwards Limestone is at an elevation of 640 feet where it forms the top of the
bluffs along the river valley. Four miles south, where Nolan Creek crosses 1-35, Edwards
Limestone forms the floor of the creek, at an elevation of about 470', giving a dip component
of about 40 ftJmi.. Upstream to Stop 15, a distance of 15 miles, the elevation difference is 80
feet, and the dip component is a little more than 5 ft.lmi.. Thus, between here and 1-35 at
Belton, dip increases by eight times, and the rollover is clearly visible. East southeast of our
present position, on the skyline about seven miles away, the most prominent feature is Scott
and White Hospital. It is at an elevation of 740 feet, on the crest of the Whiterock Escarpment,
near the top of the Austin Chalk. At that point, the Edwards limestone is at an elevation of
about 200 feet, about 540 feet below our present position (Fig. 17-2).

Therefore, at Stop 17 we are on the eastern margin of the Central Texas Platform, and on the
western margin of the East Texas Basin proper. Both are structural provinces with great
significance in the evolution of the Lampasas Cut Plain. Here we consider (1) the relationship
between the Cut Plain and the Blacklands to the East, and (2) what we have seen on this trip,
and what it all may mean in the long history of the Lampasas Cut Plain.

1-78
/,:..--- / l
Figure 17-1 (Belton Quadrangle, 1124,000; 1974) Stop 17, overlook of the Leon Valley at
Belton Dam. This Is the last stop of the trip--a view of the eastern edge of the Cut Plain, the
western margin of the East Texas Basin, and of the Texas Blacklands which bound the Grand
Prairies to the east. This is also the end of the deeply entrenched inner valley of the Leon River
1-79
WEST EAST

~
<i
Comanche Pedlpleln . Approximate prollle of Comanche Pedlplaln ~
o
'"!!!
:;;
~
STOP 15 STOP 17
STO\ 16 "
C

-'-
G

ou
Ul

---
- --
---

-
-~ :
-
--
-- --- --.......- ..... ,
- -- AIJStJl) C'"
c::::.::; ... /k
~

-
Comanchean
.....................
~. ........
GUlf/an

Central TaX8S Platform


---
-ComaOCh.ao'GU",.o cOo/,c/

Balcones Faull Zone EdMlard3

Comanchean

East Te xas Basin

Figure 17-2 West-east diagrammatic profile and section from Payne Gap (Stop 15, Hayward, AI/en, Amsbury, 1990) to
the Whiterock Cuesta, showing topographic-structural relationships of the Grand Prairie (of which the Cut Plain is a part) to
the Black Prairie; and from the Central Texas Platform to the western margin of the East Texas Basin. Notice that the
stratigraphic position of the Comanche Pediplain climbs section from Walnut Clay in the west, through Comanche Peak
limestone, to the top of the Edwards Limestone at Stop 16. We are uncertain of its position at Stop 17. Eastward in the
Blacklands there are "Uvalde gravels· which undoubtedly correlate with one or another of the high gravels of the Cut Plain.
They occur at elevations much above the third terrace and Comanche Pediplain (Hayward, AI/en, Amsbury, 1990), though
earlier we suggested such a correlation (Hayward and AI/en, 1987)
The crest of the valley wall of the Leon River which we saw at Stop 16 we interpreted as the
base level to which Leon Valley pediments were originally graded. The inner valley we attributed
to knickpoint migration, controlled by base level change in the Blacklands to the east.

The history of the Cut Plain has involved several episodes:

First was the formation of a "Time Zero" surface, a stripped structural plain on Edwards
Limestone, extending northward from the present Edwards Plateau to the area of the
Callahan Divide and an unknown distance farther to the north. This was a northern
extension of the Edwards Plateau. This surface was locally or regionally veneered with
alluvium containing abundant coarse quartzose gravel, derived from far to the west. The
age of this surface antedates the Ogallala, for Ogallala deposits onlap only the lower
slopes of the Callahan Divide at Big Spring, Sweetwater, and Roscoe. Farther to the
northwest, near Post, the Callahan Divide is buried by High Plains deposits. The divide
beneath the High Plains appears to be a mature landscape (Cronin, 1969; Seni, 1980;
Gustavson and Finley, 1985; Fallin, 1989), which

existed in its present form at the time of burial. Thus the age of the ancient surface
exceeded that of the oldest Ogallala, and by a considerable time.

By mid-Miocene Time there was already a mature Cut Plain-like landscape along the
Callahan Divide. The western Cut Plain is continuous with the eastern limit of the
Callahan Divide, and appears to be of the same age. Thus, there may have been a long
period in which the older Cut Plain landscape existed before the Yarmouthian terraces
were finally emplaced on the pediplain along the axes of various basins (Dahl, 1987).

Second, a pattern similar to modern drainage was established on this Time Zero
surface. This included an ancestral Brazos and an ancestral Colorado in the region west
of the western limits of this trip. These streams entrenched through the old Edwards
Plateau to the level of the pediplain, where they achieved grade, a profile they
maintained for an extended period. This was apparently controlled by events in the
Blacklands to the east, for entrenchment on the upper rivers was dependent on local
base levels established by trunk drainage. During the period at which these streams
flowed at grade siliceous gravels and sands were principal components of trunk stream
loads, and they remain as the third terrace, tentatively dated as Yarmouthian (Stop 16)
(Tharp, 1988). So the old Cut Plain with its pediplain still existed in Yarmouth Time.

Inner valleys in that ancient surface were probably products of erosional stripping of the
Blacklands to the east (Montgomery, 1986) beginning in mid-Pleistocene time. This too
was episodic. The inner valley, post-Yarmouth entrenchment was intermittent. The
upper terrace represents the first alluviation event, followed by dissection and alluviation
through three rock defended terraces, and a floodplain.

The terrace chronologies of the Trinity, Brazos, Leon and Colorado Rivers have been compared
(Fig. 17-3) (Tharp. 1988). Most of these age assignments are based on assumptions that are
plausible to many, though not all, students of the problem, so here we consider some of the
evidence on which they are based. These major streams appear to have similar terrace
geometrys; in that each stream system has an alluvial floodplain of Wisconsin to Recent age; a
first terrace system at about 20 to 30 feet above the floodplain (the 30 foot terrace), of
Wisconsin age; a second terrace system at about 50 to 60 feet above the floodplain (the 60
foot terrace), of Sangamon age; and a third terrace system at 90 to 110 feet above the
floodplain (the 100 foot terrace), of Yarmouth age. Above the third terrace are numerous
gravel lags, and a few actual terrace sections, but generally too few to permit mapping of yet
higher terraces above the third terrace and below the lags (and occasional

1-81
RIVER • J.CT1JAL DATED EVEIlTS

,...
T'RrNrN RIVER BRAZosANeR l..,Mlp A$AS R lYE R CXl.CP.AlXJ RNeR w:NIWER
Shul ... 19~
S10vIII AIId UcAnuity
10.,
Crool!. and Haml 1157
BroN.U9" ISiSO
F". and L«II'IW lR5.2
Slncldin
'00'
B.mlld. 1'1c. 1M2
O'tM!um anct .sa....;hI..,
....Ul ... 1R42
u... _
Loo"" IOn
..
SId'\IPtotI and Co&. 1S137

,, 'IIlD
L. .."d
ThaI'D IU7
..
Adluna and NICIl

,,
SLaughll' IHO, M. Wi\lan ,M2 and a.u.r ISl77
NJE n, 17 anct RKm-.. s.nw et. aI. 1110 au. anct PtnlMd~

".,
,...
Lund.aus ,nt Or~.'Sln
a.,natd and lt8Llnc EDI:III 1Sl1:I
Nonn 1817
Wlllimo" tOn
Collins 1012
Farn"; 11111
ftcx:::oPLAIN •
VaLI::ISro and 0 ..... FLCXX>Pl,AIN •
LATE
1870 D. &33 L.....",d l!i116S1 D. 18
FLOCOP\.AIH • FLDODP LA>< • P..IX:CP\.AI< Eddy 1S11:1 ~ 2D
IClJXa.E FIlM; 1R811 p. 80t Blml&l'd ... .,. lV70
Pr..-t 1S17. p. ,.to
S1lugl'lfI' 1017 p. 11 P. SotlS
SaUl and Plntl.oo.
EARlY
...............
Nann 1017 pel'W)l'\&J
er,ILana 1077
p. ""

. TERRACE 1 PCTJ .
LATe TERRACE 1 P1l TE RRACE , (3CIl • Urt)anec lH3 p. eo
8rcnIU;t'I lV~ P. 22 OI..rum.-.:t sz.u;h\et BaNI, and Plntl.oo. TERRACE 1 QCTJ
WIS:ONSII EDI:III IIn '.20
p. 11e1 P. "1$ Orl'lllna tSin

EARlY
TERRACE 1 ~
Willi,"," IBM Po 4U
• ..""
Feml'l9 Hlee Po 801 TERRAC£ 2' (6CJ'1
SlaugMW lU7 D. SI

TERRACE 2' (110'1 .


Shw., lQ35 p. '2
StCHall &lid IoIcAII'Il1ty TERRACE 2' (601 • TERRACE 2 (.., TERRACe :3 (loa) TERRACE 2' (80') •
~ 1;., p. 2" Lund ...... 11" P. 241 Adkins and Aller.
CtOOll; II'Id H.",. 1;30 g. 87·SI
I1S7 p. Zl
SlaU;h1w 1117 g. I

(WNDISAN

YJIR,OOUlH teRRACE 3 (100'1 TERRACE 3 (100")

_SAN TERRACE 3 (100'1

.IFlt:N1OH

NEIPAS<AN

Figure 17-3 Table showing terrace correlation for Central Texas streams. The uniformity
from basin to basin strongly suggests that terrace events were products of regional climate
that effected aI/ Central Texas streams equal/y. Since climate is regional, this further suggests
that the terrace history of Central Texas may extend far beyond the confines of the region of
Central Texas, perhaps involving much of the south central United states (Tharp, 1988).

1·82
terrace sections) probably erroneously designated as Uvalde (300 or more feet above the
floodplain) (Tharp, 1988)

Valley widening since entrenchment has been almost insignificant in comparison to the great
valley width of the pediplain, leading to the conclusion that this earlier episode of scarp retreat
was probably more rapid, and probably of far greater duration than the episodes that created
the inner valleys ..
CONCLUSIONS
And now a return to the original assumptions which were stated in the introduction to this trip.

On casual glance the earlier assumptions appear valid, yet recent work directed specifically at
Cut Plain evolution has cast doubt on most or all of them. The present revision of these earlier
conclusions might read as follows:

First, the Cut Plain is indeed an intermediate stage in the evolution of the Grand Prairie,
but it is a complex Intermediate stage. Evidence suggests that several episodes of Cut
Plain modification, some of very long duration, were interposed between the initial stage
of Cut Plain formation and the Cut Plain we see today. Evidence of at least five of those
stages is yet widely visible and was seen on this trip (Figs. 17-4 to 17-8).

Second: no one will deny the overwhelming landscape control exerted by Edwards
Limestone, but within the Cut Plain almost of equal importance are the effects of
lithologic facies, groundwater, stream gradient, and time. Of particular importance are
thresholds of process initiation, for there were apparently times of rapid slope retreat,
almost province wide, followed by periods when the landscape remained essentially static,
apparently for long periods. Even when thresholds were exceeded, not all Cut Plain
valleys participated in valley widening, for groundwater fed baseflow and stream gradient
appear to have had major influence (Fig. 2-2, 2-3).

Third: the mechanism of slope retreat originally described appears to be more Imagined
than real. The slump blocks which should everywhere mark retreating slopes are rare.
They occur on only a very small fraction of upland margins. They should be most
abundant where spring lines are most conspicuous, but instead they often appear to have
no readily explained pattern. The profiles of Cut Plain divides, which should show the
slope breaks at the base of the Edward, show instead a continuous slope from the top of
the Edwards. Edwards blocky rubble, which should be a major component of colluvial
veneer, is commonly a minor fraction. While groundwater appears to have been perhaps
the major control on slope retreat, the actual mechanism appears to have involved the
entire slope, from the top of the Edwards down. If slump blocks were a common product
of retreat, they have since been weathered away over most of the Cut Plain.

On the larger scale, slope retreat appears to have involved processes of tributary
formation, elaboration, abstraction, and finally divide destruction by processes that are as
much chemical as physical and which originate on the slopes (Stops 2, 3, 4, 8). Most
models of slope retreat predict outliers in abundance. Most result in highly crenulated
divide margins. Yet the common characteristic of much of the Cut Plain is the near
absence of outliers, and the existence of very long straight divides, particularly straight
on up dip margins. The formation of such straight segments is not well understood, but it
appears to have involved mechanisms universally active on the slopes themselves, and
almost certainly involved groundwater effects. The sequence of drainage net initiation,
elongation, and abstraction (Fig. 4-5) clearly applies to stream network evolution, but
there were apparently other highly significant aspects not confined to channels, and

1-83
active over the whole surface of the slope. Of these latter, we have well-grounded
speculations, but they are yet unproven (Stops 2, 3, 4, 8).

Fourth; the "migrating" Cut Plain appears to have remained remarkably stationary for a
very long while. It exists now over much of the area which it originally covered. Parts of
it may have existed since late Miocene time. Even the earliest remnants of the Callahan
Divide, buried under Ogallala deposits, appear to have been of mature landscapes.
Modification since that time has changed the Cut Plain, but far less than intuition
suggests that it should.

Fifth; in an overall sense, the pattern of stages in Cut Plain evolution is as earlier
outlined, but a more detailed assessment introduces complexity. Stages in Cut Plain
evolution apparently began with a "time zero" surface upon which topography and
drainage were radically different than those of today. Into this surface initial dissection
began the process of Cut Plain formation. Eventually streams approached grade, and
those which had achieved grade maintained graded profiles for long periods during which
slopes retreated leaving the broad Comanche Pediplain. Later, streams again entrenched
in a history of dissection and alluvlation to the present time, and left a record in valleys
and terraces below the old pediplain surface.

Sixth--in this discussion of the Cut Plain, we have only briefly considered the terrace
histories of the rivers which drain the region. This constitutes another problem of
considerable complexity. Our justification for this omission is that clearly defined terraces
occur almost entirely within the inner valleys which have cut below the Comanche
Pediplain, and they record "minor" events in a largely modern drainage history. Our
discussion, ranging in time from Eocene(?) to Recent, has concentrated on the major
chronology of the larger elements of this landscape.

Seventh--the evolutionary sequence In the Cut Plain has implications far beyond the
limits of the Cut Plain. Major elements of Cut Plain landscapes appear to reflect long
duration climatic events, which also left their impress on landscapes of much of the South
Central United StatesFeatures similar to the Cut Plain pediplain have been
recognized over much of south central North America, along the eastern margin of the
High Plains, along the valley of the Canadian River and its tributaries in northeastern New
Mexico, in the Edwards Plateau of southwest Texas, and in southern Kansas, and over
much of the southwest. They may not be related, but they appear to reflect similar
histories, which suggests the possibility of equivalence.

Dating of the older events is yet highly speCUlative, but the sequence of events appears to
be decipherable.

1-84
Figure 17-4 The next five illustrations constitute a recapitulation. Figure 7-4 pictures the
oldest geomorphic event recorded in the landscape, the Time Zero surface. On the west this
was an Edwards Plateau-like landscape, on the east a Washita Prairie-like landscape, perhaps
with an alluvial veneer, watered by west-to-east flowing drainage, originating far to the west.
. The age of this surface Is uncertain, but some evidence suggests that it is older than Late
Miocene, for a Cut Plain-like landscape is apparently buried under late Miocene deposits of the
Southern High Plains. Relict upland gravels of the ancient Washita Prairie streams apparently
represent a minor fraction of the total bedload of mixed-load streams. The gravels of this high .
surface came from far west or northwest of the present High Plains. Other than that, little is
known about these earliest of systems (from Hayward, Allen, Amsbury, 1990).

1-85
Figure 17-5 While the Time Zero surface was stili in existence, and during most active
subsidence of the Gulf Coast, drainage was reoriented from the apparent west-to-east direction
on the Time Zero surface, to the northwest-to-southeast direction of the present pal/ern. The
time of this reorlentalion Is uncertain, but it is suggested that it took place in early Miocene,
since an apparently mature Cut Plain landscape is buried beneath the late Miocene deposits of
the High Plains. Initial entrenchment into an Edwards Plateau-like surface, shown here, began in
the northwest, beyond the limits of our trip. By late Miocene, both the Callahan Divide and the
Lampasas Cut Plain had been formed, at least to a substantial degree. Thick Ogallala deposits
(Late Miocene-Pliocene) bury a Cut Plain-like landscape near Post, and the feather edge of the
thinning Ogallala onlaps the foot of the Callahan Divide west of Abilene (from Hayward, Allen,
Amsbury, 1990) .

1-86
Figure 17-6 Entrenchment of reoriented drainage into Fredericksburg rocks to form the Cut
Plain began in the northwest and progressed southeastward, somewhat down dip, along major
trunk streams and their tributaries. The cause of this initial incision Is uncertain. We have long
considered local mechanisms, but since similar landscapes, of similar age apparently exist over
much of the south central United States, the mechanism was probably climatic. The absence of
clear evidence of intermediate stillstands in the episode of dissection suggests that this initial
incision to the level of the Comanche Pediplain may have been a single event of entrenchment,
apparently followed by a stillstand of long duration, in which valley widening was the dominant
activity (from Hayward, Allen, Amsbury, 1990).

1-81
Figure 17-7 The time of slope retreat was apparently a long one, for the expanding pediplain
entered even the smallest of tributary valleys along those trunk streams which had participated
in the valley-widening event. Before the end of this episode, valley-widening along trunk
streams was measured in tens of miles. The ultimate landform was the Lampasas Cut Plain of
very wide valleys and narrow divides, in which more than nine tenths of the original cover of
Edwards Limestone was stripped away. The processes and rates of slope retreat of the valley-
widening phase are unknown, but groundwater appears to have played a major role. The
presence of a loess·like silt veneer over much of the older Cut Plain, and the preservation of
petrocalcic soils suggests that climates varied from humid to arid, In a cycle common to much
of the unglaciated south central United States. In mid-Pleistocene time, toward the end of the
pedlmentation phase, major trunk streams flowed in wide meanders over valley alluvial plains.
Alluvial soils preserved as high terraces are remnants of that once extensive depositional
surface. Very large entrenched meanders mimic the original free meanders of the ancestral
streams. Meander geometry suggests that through-flowing trunk streams were greatly
augmented, probably by contributions from other sources since lost (from Hayward. Allen.
Amsbury. 1990).

1·88
Figure 17-8 During mid-Pleistocene time, renewed incision began on all trunk streams
transecting the Cut Plain, creating deeply entrenched inner valleys within the old Comanche
Pedlplaln. Here again the action was widespread, Involving the whole Cut Plain, adjacent
geomorphic provinces, and perhaps much of the south central United States. This last major
Incision, however, was a complex one, involving at least four episodes of entrenchment and
alluviation. This Included (1) the last alluvial phase of the Comanche Ped/plain, preseNed as the
"100 foot" terrace of mid-Pleistocene age, (2) entrenchment to the level of the "60 foot"
terrace, of early late-Pleistocene age: (3) entrenchment to the "3~ foot" terrace of Wisconsin
age; and (4) incision to the present floodplain, of late Wisconsin to Recent age. During the
period of inner entrenchment there have been widespread and significant fluctuations In climate,
but slope retreat within the Inner valleys has been modest. By implication, the same has been
true of the slopes of the older Comanche Pedlplain, suggesting that either there was a distinct
difference in the duration of slope retreat episodes in the era of the Comanche Pediplain, and
the post-pediplaln era of entrenchment and slope retreat, or there there was marked difference
in the mechanisms and rates of slope modification between the time of formation of the
Comanche Pediplain and the more recent inner valleys (from Hayward, A/len, Amsbury, 1990).

1-89
SOME OBSERVATIONS AND SUGGESTIONS FOR THE FUTURE

In the five years since the body of this guide was written, reconnaissance in other areas has
continued the search for landscapes having similar histories and similar chronologies. Our
purpose is to test the thesis that what we see in the Lampasas Cut Plain is truly the geomorphic
response to a regional climatic history, that is, a sequence of different climates. If it is,
landscapes over a wide area should have participated in that climatic history. Landforms in
other areas might be different depending on the effects of the initial landscape from which
today's descendent was formed, details of stratigraphy, structure, erosional agencies, and
climatic variations within the overall climatic zone. but landforms in all the areas under a similar
climatic sequence should have gone through similar histories of climate-induced change.

For example, the major events in the history of the Lampasas Cut Plain that should be reflected
in some form in time-equivalent landscapes beyond the Cut Plain inclUde:

I) --the existence of an earlier regional landscape from which the present landscape has
been derived. In the case of the Cut Plain, that ancient surface was apparently a
stripped structural plain on rocks of varying resistance to erosion, forming a surface
later veneered with coarse, supermature, siliceous gravels. The age of this surface is
uncertain, but evidence suggests that it could be as old as early-to-middle Miocene.

2) --the dissection of that ancient surface through the incision of trunk streams and
their major tributary networks. In the case of the Cut Plain, that incision (pOSSibly
staged) cut through a substantial thickness of rock, as much as 700 feet, ultimately to
establish a graded profile that became the base-level control for pediplaination on a
major regional scale.

3) --an episode (or episodes) of valley widening on a grand scale by processes of slope
retreat and pedimentation, and the establishment of widely meandering trunk streams in
alluvial bottoms, at grade, within these broad pediment valleys. In the case of the Cut
Plain there are valleys of this event that exceed fifty miles in width and 700 feet in
depth, that are even now in a remarkable state of geomorphic preservation. The time of
beginning of this pedimentation episode is uncertain, though there is evidence
suggestive of late Miocene landforms appropriate to this event. There is far firmer
evidence that the pediment-phase of landscape development was still the active
landscape until mid-Pleistocene time.

4) --the penultimate stage, and the last major event in the active developmental history
of this old pediplain landscape was a period of massive caliche formation largely along
the pediplain valley-centers. This was followed by a period, perhaps short, when
pulverulent caliche was altered to a highly-distinctive petrocalcic caliche, a distinctive
marker still existing over thousands of acres of pediplain surface.. In the case of the
Cut Plain, this petro calcic caliche is unique to the pediplain surface. It occurs nowhere
else in the landscape.

5) --the final stage in landscape modification has been episodic entrenchment of trunk
stream and tributary networks into the ancient pediplain surface. In smaller tributary
valleys it is a process that yet continues. In the case of the Cut Plain, there appears to
have been three major episodes of alluviation and entrenchment, evidence of each
alluvial event now preserved in a distinctive terrace level. Comparison of terrace
histories of major streams in Central Texas indicates that trunk streams share a common
or near-common terrace sequence. In the Cut Plain the oldest such terrace level is that
of the pediplain toe, the 100 foot terrace, late Kansan or Yarmouth in age. Below this is
the "second terrace", the 60 foot terrace, of Sangamon age. Yet lower, at an elevation

1-90
of about 30 feet above the floodplain, lies the third terrace, of Wisconsin age. Lowest
of all are the alluvial floors of the rivers, of Wisconsin to Recent age. And in this incision
of major trunk stream valleys across the Cut Plain, the wide meander loops of the
ancient pediplain streams have been remarkably preserved as great valley meanders.

In projecting these episodes to other regions, a wide variety of landforms might be involved, but
a general chronology should relate them as common participants in a history of regional
landscape development. Still, in our search for common landscape histories, our first need was
to identify landscapes, the form of which seemed to offer strong evidence of similar origins.

We found them.

First, we extended the search westward along the Brazos and Colorado Rivers to the Caprock
Escarpment and beyond the High Plains to the Pecos River. At the same time reconnaissance
was extended to the east into the Black Prairie, and on eastward to the Trinity River. Additional
search extended northward, first into the Fort Worth Prairie and then to the Red River.
Additional landscape-hunting trips sought 'familiar" landforms in western Oklahoma, southern
Kansas, south Texas, and northern Mexico.

As an aside, this was not unpleasant 'work". The nice thing about research in regional
geomorphic history is that it Is probably among the last of the 'golden age" reconnaissance
level geological problems still open to today's investigator. And these problems can be done
while on family vacation trips, public relations speaking tours, and bedrock geology field trips;
or while having coffee in a country cafe, enjoying the view from a high point, and while alone,
with companions, or with wife and kids ..
I I
Out of this most enjoyable of investigations there has come to us a strong indication that Cut
Plain history may be representative of much of the southern midcontinent, and may well extend
Into west Texas, Coahuila, and Chihuahua. And we feel that great region Is probably a minimum.
I
We now realize that we have bitten off far more than we can effectively chew.

I In each area we have found landscapes that appear to us to justify further investigation, but we
also recognize the need for Independent eyes, not already committed to a pre-approved
outcome. An effective effort will require the combined input of far-better Informed colleagues
In each area. Therefore we propose a multi-authored regional study of the evolution of late
Tertiary to Recent landscapes of the southern midcontinent. Our experience has led to the
generalization that in this case the mystical "scientific method' simply boils down to attempts
to answer the questions: ·What was going on across this great region for all of that time, and
how did it function?"

We need your help.

We appreciate any suggestion you might offer, we welcome you enthusiastically to this regional
effort, and we remind you that its greatest appeal is not only In the grand dimensions of the
effort in space and time, but that it takes place In the great open expanses of the most
beautiful parts of southern North America, In a region where memorable coffee, mouhwatering
pie, and barbecue of superior quality are the norm.

1-91
BIBLIOGRAPHY

Adkins, W. S., and Arick, W. P. (1930) The Geology of Bell County, Texas; University of Texas,
Bureau of Economic Geology, Bull. 3016, p. 92

Allen, P. M. (1990) oral communication

-=-_:_-;----:-;---;-;-,(1989) oral communication, also in Soil Survey of Limestone County, U S


Department of Agriculture, Soil Conservation Service, in press.

Amsbury, David L, (1995) Stratigraphy of Fredericksburg Rocks (Middle/upper Albion


Cretaceous) of north-central Texas; unpublisheahanuscript, Ferdinand Roemer Geological
Ubrary, Waco, TX.; p. 70.

Amsbury, D. L., and Hayward, O. T. (1996) Comanche Pediplain, a major element of the
southwestern landscape (Abs.) Geological Society of America, South Central Section, Abstracts
with programs, March 11-12, 1996, p. 2.

Allee, W. A., 1962, The Lower Cretaceous Paluxy Sand in Central Texas: Baylor Geological
Studies, Bulletin no. 2, p. 26.

Bain, J. S. (1973) The nature of the Cretaceous-Pre Cretaceous Contact, North-Central Texas;
Baylor Geological Studies, Bull. 25; Baylor Univ., Waco, TX.; p. 44.

Baker, V. (1977) Stream channel response to floods with examples from Central Texas; Geol.
Soc. Am., Bull., vol. 88, p.1057-1171. .

Bammel, Bobby (1979) Stratigraphy of the Simmsboro Formation, east·central Texas; Baylor
Geological Studies, Bull., No. 37p. 40.

Barquest, Bradley (1989) A hydrologic assessment of the Austin Chalk outcrop belt, Central
Texas, Unpublished Master's Thesis, Baylor University, Waco, TX., p. 174..

Barrett, Daniel (1989) A hydrologic assessment of the Ozan Formation, Central, Texas;
Unpublished Master's Thesis, Baylor University, Waco, TX., p. 141.

Beede, J. W., and Bentley, W. P. (1918) The Geology of Coke County, Texas; University of
Texas, Bureau of Economic Geology; Bull. 1850; p.81.

Bingham, N. L. (1989) Anomalous soils of the Lampasas Cut Plain, nature, distribution, and
significance; unpublished student paper, Geology 5346, Baylor University, p. 69.

Bradley, Robert G. (1988) Anomalous soils of the Lampasas Cut Plain--nature, distribution, and
significance; Unpublished student paper, Geology 5346, Baylor University, p. 69.

Brown, Thomas E. (1988) Relationships between basin parameters and landform, Lampasas
Cut Plain, Central Texas; Unpublished master's thesis, Baylor UniverSity; p. 166.

Brotherton, M. A. (1978) The geomorphic evolution of the North Bosque River; Unpublished
bachelor's thesis, Baylor Univ.; Waco, TX.; p.131.

Bunting, B. T. (1961) The role of seepage moisture in soil formation, slope development, and
stream initiation; Am. Journ ... Sci., vol. 259, p. 503-518.

1-92
Byrd, C. L. (1971) Origin and history of the Uvalde Gravel, Central Texas; Baylor Geological
Studies, Bull. 20; Baylor Univ.; Waco, TX.; p.48.

Cannata, Stan L. (1988) Hydrogeology of a portion of the Washita Prairie Edwards Aquifer:
Central Texas; Unpublished M. S. thesis, Baylor University, Waco, TX.; p. 205.

Collins, Andrew D. (1989) Geochemistry and flow characteristics of springs discharging from the
Edwards Aquifer: Washita Prairie, Central Texas; Unpublished M.S. thesis, Baylor University,
Waco, TX.; p.137.

Cronin, J. G. (1969) Ground Water In the Ogallala Formation of Texas and New Mexico; U. S.
Geological Survey, Hydrologic Atlas 330, p.9.

Dahl, Suzanne (1987) The Intermediate surface across the Brazos River, I\Iertlral Texas,
and its relationship to Lampasas Cut Plain evolution; Unpublished student paper, Geology
5346, Baylor University, p. 50 (approx.)

Epps, L. W. (1973) The geologic history of the Brazos River; Baylor Geological Studies, Bull. 24,
Baylor Univ., Waco, TX.; p.44.

Fallin, J. A. (1989) Hydrogeology of Lower Cretaceous strata under the Southern High Plains of
Texas and New Mexico; Tex. Water Development. Board, Report 314, Austin, TX.; p.39.

Faulkner, H. (1974) An allometric growth model for competitive gullies; Zeitschrlft for
geomorphologie; Supplement 21, p. 76-87.

Fisher, W. L. and McGowen, J. H. (1969) DepOSitional systems in the Wilcox Group of Texas and
their relationship to the occurrence of oil and gas; American Association of Petroleum Geologists
Bull. vol. 53, p. 30-54.

Flatt, C. D., 1976, Origin and significance of the oyster banks in the Walnut Clay Formation,
Central Texas: Baylor University Geological Studies, Bulletin no. 30,47 p.

Flowers, Beth A. (1987) Geomorphic evolution of the upper Colorado River basin; student
paper, Geology 5346; Baylor University; p. 66.

(1989) Sediment budget analysis of the Coryell Creek drainage Basin, Coryell
County, TX., Unpublished M. S. thesis, Baylor University, Waco, TX., p. 106.

Glock, W. S. (1931) The development of drainage systems: a synoptic view; Geographical


Review, Vol. 21, p. 475-82.

Gupta, A. (1983) High magnitude floods and stream channel response, In Spec. Pubs.
International. Assoc. Sedimentology; No.6, p. 219-227.

Gustavson, T. C., and Finley, R. J. (1985) Late Cenozoic geomorphic evolution of the Texas
Panhandle and northeastern New Mexico; Univ. of Texas, Bureau of Economic Geology, Report of
Investigations No. 148; p. 42.

Hayward, O. T. (1988) The Comanchean section of the Trinity Shelf, central Texas, In
Southcentral Section of the Geological Society of America, Centennial field guide, vol. 4 (0. T.
Hayward, ed.) Geological Society of America; p. 323-328.

\-93
Hayward, O. T., and Allen, P. M. (1987) The Lampasas Cut Plain; Form, Process, Evolution:
Guidebook, South Central Section, Geological Society of America; Baylor University, Waco, TX.,
p.84.

Hayward, O. T., Allen, P. M. (1988) The Lampasas Cut Plain; Form, Process, Evolution (2nd
ed.): Guidebook, Southwestern Association of Student Geological Societies, Baylor University,
Waco, TX., p. 119.

Hayward, O. T. Allen, P. M. and Amsbury, D. L. (1990) The Lampasas Cut Plain-Evidence for
the cyclic evolution of a regional landscape, Central Texas, Field Trip #19. Geological Society of
America 1990 National Meeting, Dallas, TX. p.128.

Hayward, O. T.; Dolliver, P. N.;, Amsbury, D. L.; and Yelderman, J. C. (1992) The Grand Prairie of
Texas: land, history, culture; Program for Regional Studies, Baylor University, Waco TX .• p. 98.

Hill, R. T., (1899) A preliminary annotated check list of the Cretaceous invertebrate fossils of
Texas. accompanied by a short description of the lithology and stratigraphy of the system:
Texas Geological Survey Bulletin 4, 57 p.

Hill. R. T. (1900) Physical geography of the Texas Region; United States Geological Survey,
Topographic Atlas of the United States, Folio 3; p. 22.

=-__ (1901) Geography and Geology of the Grand and Black Prairies, Texas; U.S.G.S.,
Twenty First Annual Report, Part VII; p.666.

Huckabee, J.; Thompson, D.; Wyrick, J.; Pavlat, E. '(1977) Soil Survey of Bell County Texas; U.S.
Department of Agriculture. Soil Conservation Service; p. 75.

Ikin. Arthur (1841) Texas: its history. topography. agriculture, commerce and general statistics;
reprint, Texian Press, Waco, TX, 1964, p.100.

Irving, Washington (1835) A tour of the prairies; reprint, University of Oklahoma Press. Norman
OK. p. 216.

Kendall, George W. (1844) Narrative of the Texas Santa Fe expedition ........... ; In two volumes,
reprint of the first edition, Steck Company, Austin, TX. (1936) Vol. 1, p. 405. Vol. 2, p. 406.

King, L. C. (1962) Morphology of the Earth; Haffner. N. Y.; p. 699.

Knapp, M. (1990) oral communications -- work In progress, nature, origin. and significance of
the high gravels. Callahan Divide, Texas).

Knighton, D. (1984) Fluvial forms and processes; Arnold, London; p.218.

Laity, Julie E. and Malin, Michael C. (1985) Sapping processes and the development of theater-
headed valley networks on the Colorado Plateau. Geological St).oiaYca:>fBulietin, v.
96, p. 203-217.

LeFevre. Stephen B. (1989) Geologic controls on the main channel of the Middle Bosque River,
Central Texas-a bedrock controlled stream system; Unpublished master's thesis; Baylor Univer-
sity, Waco. Tx.; 117 p.

Lemons, D. R. (1987) Structural evolution of the Lower Cretaceous (Comanchean) Trinity Shelf;
Unpublished master's thesis. Baylor Univ.; Waco. TX.; p.300.

\-94
Lewand, Raymond L. (1969) The geomorphic evolution of the Leon River system; Baylor
Geological Studies, Bull. 17; Baylor Univ.; Waco, TX.; p.27.

Lozo, F. E., 1959, Stratigraphic relations of the Edwards Limestone and associated formations
in North-central Texas: in Lozo, F. E., Nelson, H. F., Young, Keith, Shelburne, O. B., and Sandidge,
J. R., Symposium on Edwards Limestone in Cehblillersi:ljfarof Texas Bureau of
Economic Geology Pub. No. 5905, p. 1-19.

-=-_" and Stricklin, F. L., Jr., 1956, Stratigraphic notes on the outcrop basal Cretaceous,
Central Texas: Transactions, Gulf Coast Association of Geological Societies, vol. VI, p. 67-78.

Marcy, Randolph B. (1854) Exploration of the Red River of Louisiana in the year 1852; Executive
Document, 33d Congress, 1st Session, House of Representatives; A. O. P. Nicholson, Public
Printer, Washington D. C.; p. 286, 20 plates.

McCaleb, N. (1985) Soli survey of Coryell County, Texas; U.S. Department of Agriculture, Soil
Conservation Service; p. 129.

Mikels, J. K. (1977) Geomorphology of the Lampasas Cut Plain, Central Texas; Unpublished
bachelor's thesis, Baylor University, Waco, TX.; p.42.

Montgomery, James A. (1986) The geomorphic evolution of the Taylor Black Prairie between
the Trinity and Colorado Rivers, Central Texas; Unpublished master's thesis, Baylor University, p.
148.

Moore, C. H., Jr., 1961, Stratigraphy of the Walnut Formation, South-central Texas: Texas
Journal of Science, vol. 13, p. 17-40.

-;-::;-:-" 1964, Stratigraphy of the Fredericksburg Division, South-central TexaShe University


of Texas at Austin Bureau of Economic Geology Report of Investigations no. 52, 48 p.

Moore, T. H. (1970) Water geochemistry, Hog Creek Basin, Central Texas: Baylor Geological
Studies, Bull. 18; Baylor Univ.; Waco, TX.; p.44 .

.,,-___ (1989) Aquifer-stream interactions and their hydrologic implications in nonkarstic


limestones; Washita Prairie, Central Texas; Unpublished M. S. thesis, Baylor University, Waco, Tx.,
p.180.

North American Comm. on Stratigraphic Nomenclature (1983) North American Stratigraphic


Code: American Association of Petroleum Geologists Bulletin,
v. 67, no. 5, p. 841-875.

Owen, M. T., 1979, The Paluxy Sand in North-central Texas: Baylor University Geological Studies,
Bulletin no. 36, 36 p.

Parish, B. R. (1989) Geomorphic response to regional structure, Lampasas Cut Plain, Central
Texas; Baylor Geological Studies, No. 55, Baylor University, Waco, TX., pp. 28.

Parker, R. S. (1976) Experimental study of drainage basin evolution and its hydrologic
implications; Colo. State Univ., Ft. Collins, Colo.; Hydrologic Papers 90; p. 58.

Perkins, B. F. (1961) Biostratigraphic studies in the Comanche (Cretaceous) Series of Northern


Mexico and Texas: Geological Society of America Memoir 83, 138 p.

1-95
Poole, W. C. (1964) Bosque Territory; Chaparral Press, Kyle, TX., p.206.

Proctor, C. V. (1969) The North Bosque Watershed, Inventory of a drainage basin; Baylor
Geological Studies, Bull. 16; Baylor University; Waco, TX.; p. 40.

Roemer, Ferdinand (1848) Texas, with particular reference to German Immigration and the
physical appearance of the country,; translated by Oswald Mueller, (1935); reprint, German
Texan Heritage Society (1983); p. 308.

Rose, P. R., 1972, Edwards Group, surface and subsurface, CetltrBAerSiEyi:asxf The
Texas at Austin Bureau of Economic Geology Report of Investigations no. 74, 198 p.

Segrest, Charles C. (1977) Geomorphic evolution of the Middle Bosque River, Central Texas;
Unpublished student paper; Baylor University, Waco, Tx.; p. 107.

Selby, M. J. (1982) Hillslope materials and processes;Oxford University Press, p. 264.

Seni, S. J. (1980) Sand body geometry and depositional systems, Ogallala Formation, Texas;
University of Texas, Bureau of Economic Geology, Report of Investigations 105, p.36.

Slaughter, Bob H. (1966) The Moore Pit local fauna--Pleistocene of Texas; Journ. Paleon-
tology, v. 40; p.78-91.

Smith, Scott E. (1984) Evidence of ancient cross·divide drainage, western Bosque Basin, Central
Texas; Unpublished bachelor's thesis, Baylor University; p.65.

Spencer, Kevin (1991) Drainage evolution across the White Rock Cuesta, Central Texas;
unpublished Master's Thesis, Baylor University, Waco, TX, p. 177.

Stringer, B. (1980) Soil survey of Bosque County, TX; U. S. Department of Agriculture, Soil
Conservation Service; p. 102.

Tharp, Tommy Lee (1988) Aspects of Leon River drainage history with implications to other
Central Texas streams; Unpublished master's thesis, Baylor University, p. 260.

Thuma, Jeffrey J. (1987) The original topography of the Edwards Limestone across the Llano
Region, and its relationship to Colorado River History; Unpublished student paper, Geology
5346; Baylor University, p. 50 (approx).

Vaughan, T. W. (1900) Folio No. 64, Uvalde, Texas; United States Geological Survey,
Washington D. C.; p. 7, 3 maps.

Walker, J. R. (1978) Geomorphic evolution of the Southern High Plains; Baylor Geological
Studies, Bull. 35; Baylor Univ.; Waco, TX.; p.32.

Yelderman, J. (1987) oral communication, Baylor Univ., Waco, TX.

Young, Keith, 1967, Comanche Series (Cretaceous), south central Texas: in Leo Hendricks, ed.,
Comanchean (Lower Cretaceous) stratigraphy and paleontology of Texas: Permian Basin
Section, Society of Economic Paleontologists and Mineralogists Publication 67-8, p. 9- 29.

\-96
_ _, 1974a, Lower Albian and Aptian (Cretaceous) ammonites of Texas: in B. F. Perkins, ed.,
Aspects of Trinity Geology, Geoscience and Man, vol. VIII, Louisiana State University Press, p.
175-228.

_ , 1974b, Edwards Plateau ammonites: in P. R. Rose, ed., Stratigraphy of Edwards Group


and equivalents, Eastern Edwards Plateau, Texas: South Texas Geological Society Guidebook for
AAPG-SEPM Field Trip 2, March 30-31,1974, p. 59-75.

, 1986, Cretaceous, marine inundations of the San Marcos Platform, Texas: Cretaceous
Research, vol. 7, p.117-140.
Friends of the Pleistocene
April 19 - 21,1996
Trip Log

April 19, 1996,7:30 am. Lampasas Cut Plain.

0.0 Turn right (west) onto the frontage road beside La Quinta (next to Popeye's) and move
into the far left lane. Turn left under Hwy 190 and left again onto the frontage road
(east). Enter H wy 190 East

18.7 Hwy 190 curves left (north) and joins Interstate 35.

62.5 Exit #335 at Baylor (Fort Fisher) Exit. Meet in the parking lot to begin. We will
follow the trip described in your guidebook stops 1-8 and 16-17.

At the end of the trip take loop 121 to Highway 190 west and return to La Quinta.

April 20, 1996, 8:00 am. Fort Hood, Cowhouse Creek.

0.0 Turn right (north) onto Fort Hood SI. at the front of La Quinta.

1.3 Turn left onto base (Tank Destroyer Blvd) at T-intersection.

2.2 Turn right on 31st (stay in the left lane). Cross Central Ave and Park Ave.

2.9 Turn left (west) on North Ave. North Ave is a divided road but both sides have two
way traffic.

3.8 Turn right cross the division between the North Aves and then turn left. You are now
on the north side of North.

3.9 Turn right on Rod and Gun Club road. At the back of the parking lot, park near "Area
Acess. "

4.1 After checking in with area access turn left (south) on 53rd.

4.2 Turn right (west) on North Ave.

7.4 Curve right (north) onto West Range Road.

8.8 Stay right on West Range Road (to "North Fort Hood")
10.5 Cross Cowhouse Bridge. We will turn left after crossing Cowhouse Creek. Watch the
lead vehicle since our exact route may change depending on off road conditions. Stop
I. Cutbank exposure.

10.8 Turn right (south) on West Range Road. Cross Cowhouse Creek and watch lead
vehicle. Stops 2 and 3 are on the right after crossing the creek.

12.0 (approximate mileage). Turn right (south) on West Range Road.

13.8 Turn right (west) on Elijah Road.

16.4 Turn right on Old Georgetown Road.

20.3 Cross Cowhouse Creek. Turn right off Old Georgetown Road and follow the lead
vehicle to stop 4. Lunch.

24.3 (approximate mileage). Turn left (south) on Old Georgetown Road, cross Cow house
Creek and turn right following the lead vehicle.

25.2 (approximate mileage). Turn left (south) on Old Georgetown Road.

28.6 Turn left on Elijah road.

31.2 Fork right onto West Range Road (south).

34.5 Curve onto (heading east) North Ave.

35.6 Turn right, cross median, turn left onto North Ave.

36.6 Turn right (south) on 31st Ave.

37.5 Turn left (east) on Tank Destroyer.

38.5 Turn right (south) on Fort Hood.

39.8 Turn left into La Quinta

April 20, 1996, 6:30 pm. Trip to Frank's Lakeview and Anchor Club.

0.0 Turn right (west) onto the frontage road beside La Quinta (next to Popeye's) and move
into the far left lane. Turn left under Hwy 190 and left again onto the frontage road
(east). Enter Hwy 190 East

15.0 (mileage approximate). Exit at Loop 121. Turn left and head north. At 439 turn left
(west) and follow the signs to Frank's.

2
April 21, 1996, 8:00 am. Eastern Training Area archaeological sites: rockshelters, burned rock
mounds, and lithic procurement areas.

0.0 Turn right (north) onto Fort Hood St. at the front of La Quinta.

1.3 Turn right (east) on 439 (Rancier).

2.0 (Second traffic light). Turn left next to the Dollar Store. Proceed I block to Yield
sign, turn right and then immediately head straight/left onto Garth. You wiII be
heading north. Continue straight at the four-way stop. After this point continue
straight (the road name wiII change to Martin) until the road stops at Central.

2.7 At Central turn right and then left (back onto Martin).

3.4 At stop sign next to heliport, continue straight. Note: this is not a four-way stop. You
are now on East Range Road.

11.3 Cross Cow house Creek.

11.9 Turn right (east) on gravel road. Note results of recent fire on your left.

17.0 The road wiII curve to the left (north) and climb the escarpment.

17.5 Turn right onto dirt/gravel road (east).

IS.5 Turn right to lithic procurement site (watch lead vehicle). Stop I. 4IBL60S.

IS.7 Return to dirt/gravel road and turn right (east).

19.9 Turn left onto pipeline road (north).

22.3 Turn right onto dirt road (east).

22.4 Turn right. Follow lead vehicle. Stop 2, 41BL233. After Stop 2, turn around. At the
dirt road turn right.

22.7 Stop to walk down to the rockshelters. Stop 3, 4lBL671 and 41BL670. Final stop.
Turn vehicles around and return to the pipeline road.

23. I Turn left (south) on the pipeline road.

25.5 Turn right (west) on dirt/gravel road.

27.9 Turn left (sonth) and head down the escarpment. Notice view of Belton Reservoir.

33.5 Turn left (south) on East Range Road.

3
42.0 At stop sign proceed straight. You are now on Martin Road.

42.7 At Central curve right and make an immediate left turn to stay on Martin Road.
Continue straight on Martin/Garth until you reach Rancier.

43.4 Turn right on Rancier.

44.1 Turn left on 439 (Fort Hood)

45.4 La Quinta/Highway 190.

4
SOILS AND GEOMORPHOLOGY OF
COWHOUSE CREEK, FORT HOOD, TEXAS

by

Lee Nordt

2-1
Introduction

During the past decade geoarchaeological investigations in Texas have provided a


unique opportunity for reconstructing alluvial stream histories, studying rates of pedogenesis,
and making paleoenvironmental interpretations. Since 1989, with monies appropriated from
the environmental staff at the Fort Hood Military Reservation, the alluvial histories of eight
streams at Fort Hood, Texas have been reconstructed with the support of nearly 70
radiocarbon ages, and over 100 cutbank and 92 trench descriptions (Nordt, 1992). Cowhouse
Creek, the largest stream within the Fort Hood Military Reservation, has a rich archaeological
record that has yielded nnmerous radiocarbon datable samples. The focus of this report is the
late Quaternary alluvial history of Cowhouse Creek. Specific objectives were to:
1) identify the major stratigraphic units, interpret their environments of deposition and
establish a nnmerical stratigraphic chronology, 2) study soil chronosequences and
chronofunctions within the stratigraphic framework, and 3) assess presentation potentials for
the archaeological record.

Study Area

Geology of Fort Hood

Fort Hood is situated on the dissected eastern margin of the Edwards Plateau in the
Grand Prairie Land Resource Area of Texas CD.S.D.A., 1981). Hill (1901) named the
southern portion of the Grand Prairie the "Lampasas Cut Plain." Differential erosion of
resistant and nonresistant Cretaceous rocks has produced a distinctive landscape configuration
whereby nnmerous prominent flat-topped ridges are dissected by narrow and deeply
entrenched stream valleys. These landscape features bear a close relationship with mapped
geomorphic surfaces and soil distributions.

Major rock units that crop out in the Fort Hood area are Lower Cretaceous, which
parallel ancient north-south trending marine shorelines (Bames, 1979) (Fig. 1). Sediments
from Lower Cretaceous formations were deposited as part of a series of marine transgressions
and regressions within an overall transgressive trend (Adkins, 1978; Rose, 1979). Marine
transgression proceeded from the east along a broad shallow shelf as a result of subsidence of
the East Texas Basin. Deposition of sediments in this low energy environment resulted in
formation of limestones, marls, shales, and clays (Davis, 1974; Moore, 1969; Stricklin, 1971;
and Nelson, 1973). Prograding fluvial and strandline sands, although not as abundant, were
also deposited (Owen, 1979; Anderson, 1989) during relatively brief periods of terrigenous
sediment influx, primarily from North Texas, Oklahoma, and Arkansas. Couplets of resistant
limestones and nonresistant limestones, sands, sandstones, and shales produce a micro-stepped
landscape on backwearing upland slopes in the Fort Hood area.

2-2
L.EON RIVER

E 3 WAL.NUT CLA.Y
EZEJ
I
~ EDWARDS, DUCX CREEK, AND COMANCHE PEAK LiMESTONES

GL.EN ROSE FORMATION


_ . - FORT HOOD MlL.lTARY RESERVATION BOUNDARY

REESE CREE)(
Ok.

Figure I. Major Rock units and stream patterns of Fort Hood Military Reservation.

2-3
The Cretaceous fonnations, from lower to upper, that are principal contributors of
Quaternary alluvium in the Fort Hood area are summarized below (Barnes, 1979) (Figs. 1 and
2):
(1) Glen Rose Fonnation - fme-grained, chalky to hard limestone; interbedded dark
gray clay and marl; partly arenaceous.
(2) Paluxy Sand - light gray to red, very fme grained quartz sands; locally
interbedded shales and limestones.
(3) Walnut Clay - nodular and chalky clay, limestone, and shale.
(4) Comanche Peak Limestone - hard and nodular, gray to white limestone; numerous
shale partings.
(5) Edwards Limestone/Kiarnichi Clay undifferentiated - massive, rudistid limestone,
abundant chert; nodular clay, limestone, and shale.

The Glen Rose Formation forms the basement rock on which Cowhouse Creek flows,
and along with the overlying Walnut, is the main contributor of sediment for fluvial transport
and deposition. Therefore, alluvial sediments are abundant in sand-sized carbonates, large
carbonate clasts derived from local limestones, and chert nodules from the Edwards
Limestone.

Landscape Evolution of Fort Hood Area

Scattered water-worn siliceous channel lag on flat-topped ridges provide some of the
earliest evidence of post-marine landscape history in the Lampasas Cut Plain (Hayward et al.,
1990). These gravels were deposited by ancestral Central Texas rivers as they migrated
across the Edwards Formation, thus forming what are now remnants of the "High" surface of
Hayward et al. and the Manning surface of this study (Fig. 2). Hayward et al. (1990)
correlate this surface with the Callahan Divide of North-Central Texas, which is
stratigraphically 60 to 70 m above the southeastward projection of the Southern High Plains
surface of Texas. This indicates that the Manning surface pre-dates deposition of the Ogallala
Formation and development of the High Plains. By inference, because the Ogallala was
deposited sometime between late Miocene and late Pliocene (Walker, 1978; Byrd, 1971), the
Manning surface may be at least Miocene-age (Hayward et al., 1990).

Approximately 25 to 35 m below the Manning surface, resting primarily on the


Walnut Clay and partly on the Comanche Peak Limestone, lies the broad and gently rolling
"Intermediate" surface of Hayward et al. (1990) and the Killeen surface of this study (Fig. 2).
Sometime during the late Tertiary or early Quaternary, widespread valley entrenchment
resulted in abandonment of the Manning surface. The Killeen surface then formed as a result
of long-term landscape stability and pedimentation as the Manning surface and Edwards rocks
retreated laterally. Tributaries that sculptured the Killeen surface graded down to large base-
level controlling trunk streams, such as the Brazos and Leon Rivers, when they flowed some
25 to 30 m above their modem counterparts. The Killeen surface is now left as a relict
drainage network containing erosional and depositional counterparts.

2-4
Manning Surface

'" Killeen Surface Modem Drainage

"" T2 !
T1 TO- - - 1 - - -

Cretaceous Units Quaternary Un~


ITJJ Kiamichi Clay IEdwards limestone [::=:::J Alluvium
~ Comanche Peak limestone
E-m--I Walnut Clay
~ Glen Rose Formation
...
t;'

Figure 2. Schematic geologic and landscape cross-section of the Lampasas Cut Plain,
Central Texas (from Nordt, 1992)(no scale).
Colluvial soils with petrocalcic horizons on footslopes connecting the Manning and
Killeen surfaces, provide strong evidence that lateral retreat of the Manning surface has been
inactive for some time. The Killeen surface lies just above the highest terraces, of the Brazos
and Leon Rivers, and as a result may be as old as early to middle Pleistocene (Hayward et
al., 1990).

The Killeen surface was probably abandoned sometime after the middle Pleistocene
during the last major episode of valley entrenclunent (Hayward et al., 1990). Since its
initiation several cut and fill episodes have followed fonning prominent terraces, particularly
in the larger stream valleys (Fig. 2). These terraces and the modem stream valleys are
Pleistocene to Holocene in age. Modem channels are commonly entrenched 25 to 30 m
below the Killeen surface.

Soils of Fort Hood

Soils at Fort Hood can be grouped into six associations that reflect the influence of
surface geology and landscape age (Fig. 3) (McCaleb, 1985). Family classification of the
soils are given in Table 1. The Eckrant -Real rock outcrop association consists of shallow
rocky soils mapped on the Edwards Formation of the Manning surface. Eckrant and Real
soils are thin Haplustolls and Calciustolls, respectively, and have silty clay to clay loam
surfaces that are sometimes gravelly. Given that the Manning surface is pre-Quaternary in
age, the soils mapped on this surface are poorly developed even considering that they may
have formed from hard limestone. Locally however, petrocalcic horizons do occur and have
also been identified on similar surfaces and formations in the western Edwards Plateau Land
Resource Area (Rabenhorst, 1983).

Several soil associations are mapped on the broad and gently rolling Killeen surface.
The Nuff-Cho association is found mainly on the Walnut Clay portion of the Killeen surface.
They have thin clay loam to silty clay loam surfaces that are locally gravelly to very stony.
Both the Nuff and Cho are Calciustolls and exhibit more pedogenesis than many soils on the
older Manning surface probably a result of fonning in clays and marls rather than in the more
resistant limestones. Cho soils are on more stable remnants of the Killeen surface and exhibit
Stage 4 (Oile et al., 1966) carbonate morphology in the form of well developed petroca1cic
horizons.

The soils of the Slidell-Topsey-Brackett association are Pellusterts, Calciustolls, and


Ustochrepts, respectively, and are mapped on the Walnut Clay Formation of the Killeen
surface. Slidell soils are mapped in thick clayey relict valley fills, while the Topsey and
Brackett soils formed in convex positions and steeper side slopes, respectively. Slidell soils
have appreciable carbonate accumulation deep in many profiles, which suggests that some
carbonates may have been present in the solum and now have been partially leached.
Soils of the Doss-Real-Krum association are mapped principally where the Olen Rose
Formation outcrops on erosional margins separating the Killeen surface and younger stream
terraces. This portion of the landscape is younger than the Killeen surface and in places
grades to high terraces of modem streams. Doss soils (shallow Calciustolls) form on

2-6
3 km
EXPLANATION

5 Bosque-FrIO-Lewisvi I Ie
1 Eckrant-Reai Rock
6 Bastl!-Mlnwells
2 Nuff-Cho
3 S Ilde II-T<.:lpse!:l-Bracket t
4 Ooss-Real-Krum

Figure 3. General soils map of Fort Hood, Texas. (McCaleb, 1985).

2-7
Table 1. Common Soils of the Southern Portion of the Grand Prairie and their U.S.D.A.
Classification.

Series Classification

Bolar Typic Calciustolls; fme-loamy, carbonatic, thermic

Bosque Cumulic Haplustolls; fine-loamy, mixed, thermic

Brackett Typic Ustochrepts; fine-loamy, carbonatic, thermic, shallow

Cho Petrocalcic Calciustolls; loamy, carbonatic, thermic, shallow

Crawford Udic Chrornusterts; fme, rnontrnorillonitic, thermic

Denton Typic Calciustolls; fme-silty, carbonatic, thermic

Doss Typic Calciustolls; loamy, carbonatic, thermic

Eckrant Lithic Haplustolls; clayey-skeletal, rnontrnorillonitic, thermic

Frio Cumulic Haplustolls; fme, rnontrnorillonitic, thermic

Krum Vertic Haplustolls; fme, rnontrnorillonitic, thermic

Lewisville Typic Calciustolls; fme-silty, mixed, thermic

Nuff Typic Calciustolls; fme-silty, carbonatic, thermic

Purves Lithic Calciustolls; clayey, rnontrnorillonitic, thermic

Real Typic Calciustolls; loamy-skeletal, carbonatic, thermic, shallow

Slidell Udic Pellustert; fme, rnontrnorillonitic, thermic

Topsey Typic Calciustolls; fme-loamy, carbonatic, thermic

2-8
erosional hilIslopes while Krum soils (Vertic Haplustolls) develop in a variety of alluvial and
colluvial landscape positions such as relict valley fills associated with the Killeen surface, or
on high strath terraces in upper drainage basins where the Walnut Clay was beveled by lateral
stream migration.

The Bastil-Minwells association formed from siliceous and mixed mineralogy alluvial
sediments deposited by the Leon River. These soils, unlike the carbonatic parent materials
derived from local sources, are acid PaleustaIfs with reddish Bt horizons and form mainly on
terraces.

The Bosque-Frio-Lewisville association consists of soils formed from alluvial


sediments deposited by modern streams in the form of terraces and floodplains. Bosque and
Frio are deep Curnulic Haplustolls mapped in rarely to frequently flooded floodplain
positions. Lewisville, a Typic Calciustoll, is mapped on higher stream terraces. The Fort
Hood portion of this field trip focuses on a soil chronosequence developed from soils
occurring on these alluvial landforms, specifically, those within Cowhouse Creek.

Methods

This investigation began with a review of literature pertinent to alluvial stream


histories and paleoclimates of Central Texas. Analyses of topographic, soil survey, and aerial
photographic maps of Cowhouse Creek were undertaken to assess geomorphic and soil
relations. Field reconnaissance proceeded with three principal goals in mind: 1) to identify
the major alluvial landforms for assistance in construction of the geomorphic maps, 2) to
locate cutbank exposures for identifying major stratigraphic units and establishing their
relative temporal order, 3) to locate trenching sites for testing the accuracy of landform maps
and lateral extensiveness and outcrop nature of the stratigraphic units, and 4) to collect
radiocarbon samples from critical stratigraphic localities.

Construction of the geomorphic map of Cowhouse Creek relied on 1:24,000 U.S.


Geological Survey and 1:4,800 U.S. Corps of Engineer topographic maps, 1:24,000 U.S.
Department of Agriculture aerial photographs, and field observations. Landform designations
were adapted from Brackenridge (1988) where TO represents the most frequently flooded
surfaces and n, T2 and so on, represent higher terraces that are infrequently flooded or not
flooded.

Soil-stratigraphic descriptions, including texture, color, horizonation, reaction to HCl,


structure, consistence, boundary, and identification of special features, were written from over
100 cutbank exposures and 92 trench sites following standard procedures outlined by the Soil
Survey Staff (1981). Soils were classified according to Soil Survey Staff (1990).

Subsurface sediments were grouped informally into stratigraphic units and designated
from oldest to youngest beginning with A and progressing through the alphabet toward
younger sediments. These stratigraphic units are similar to Allostratigraphic units in that they
represent packages of sediment that mayor may not have similar internal character and that

2-9
are bound by laterally traceable disconformities (North American Commission on
Stratigraphic Nomenclature, 1983). The upper boundary may bea geomorphic surface or
buried soil.

Complete soil characterization (Hallmark et ai., 1986) was performed on samples


collected from four pedons on Cowhouse Creek and are the focus of this field trip. Analyses
considered important for determining chronosequences included determination of particle size
distribution, organic carbon, calcium carbonate equivalent, dithionite-citrate extractable iron
(FeJ content, pH, bulk density, clay mineralogy, and thin section analysis. These properties
can be used as chronological indicators in the absence of numerical dating techniques and for
the detection of stratigraphic discontinuities and reconstruction analysis.

Standard radiocarbon assays were performed by Beta Analytic Laboratories and


Balcones Laboratory (University of Texas), and accelerator assays by Geochron Laboratories.
All ages were determined from charcoal, wood, or bulk humate samples, presented in years
before present (B.P.), and corrected for variations in carbon-13.

Stable carbon and oxygen isotope analyses were performed on bulk carbonates and
organic carbon of alluvial sediments and soils by Dr. Thomas Boutton, Department of
Rangeland Ecology and Management, Texas A&M University.

Stable Carbon Isotope Theory

An important indicator of soil age is the accumulation of secondary calcium carbonate.


Carbonate accumulations can be described quantitatively (Birkeland, 1984; Machette, 1985;
McFadden and Tinsley, 1985) or descriptively as stages (Gile et al., 1966). In soils
containing large quantities of lithogenic carbonate, morphological distinction between
pedogenic and lithogenic carbonate forms is not always possible (Rabenhorst, 1983; West,
1986). In recent years, however, stable carbon isotope analysis has been used to successfully
differentiate between pedogenic and lithogenic carbonate forms in soils formed in calcareous
sediments (Salomons and Mook, 1976; Magaritz and Arniel, 1980; Rabenhorst et al., 1984;
Pendall and Amundson, 1990) and to identify the kind and density of vegetation biomass
growing during pedogenesis (Ceding, 1984; Amundson et ai., 1988; Quade et al., 1989;
Pendall and Amundson, 1990). The following equation, derived by Salomons and Mook
(1976), expresses the relationship of the isotopic composition among bulk soil, pedogenic, and
parent material carbonate. This equation can then be used to calculate the percentage of
pedogenic carbonate present in a soil horizon that contains lithogenic components:

% pedogenic carbonate = 1113C (bulk soil) - 1113C (P.m.) X 100 (1)


1113C (pedogenic) ~ 1113C (p.m.)

where:
1113C (bulk soil) = bulk soil carbonate
1113C (pedogenic) = soil organic matter plus a fractionation factor of 14%0
11 13 C (p.m.) = parent material carbonate

2-10
Delta 13C values are ratios of 13C/12C relative to the Pee Dee belemnite standard expressed in
parts per thousand (%0 or per mil) (Craig, 1957).

Marine limestones commonly have /l13C values near zero (Hoefs, 1987) while that of
bulk soil carbonate, which typically contains both pedogenic and lithogenic carbonate forms,
has values that are somewhat less. For Equation (1), both the parent carbonate and bulk
carbonate values can be determined directly from the soil profile in question. The most
difficult estimation in Equation (1) is that of the /ll3C of pedogenic carbonate because it is
dependent on the isotopic composition of soil organic carbon, molecular diffusion of soil
respired CO2, and isotopic fractionation during chemical phase changes in the CO2(g)-
CaC03(s) system (Deines, 1980). The derivation of the pedogenic carbonate end member
value involves stable carbon isotope theory and is discussed below.

Dissolution and reprecipitation of carbonates in soil systems can be expressed as:

CO2(g) + H,D + Caco3 = Ca·2 + 2HCOf (2)

In humid regions where soil respiration is high, and there is little mixing of atmospheric CO2,
the source for soil respired CO 2 at depth is derived mainly from decaying organic matter and
root respiration (Ceding, 1984; Amundson et aI., 1988; Quade et aI., 1989). The isotopic
composition of soil organic matter is determined by the proportion of C3 and C4 species
growing on the soil. C3 plants, which includes all tree species and grasses that experience a
cool growing season, and C4 plants, which includes mostly warm season grasses, have
different photosynthetic pathways resulting in different 13C/12C fractionations (Deines, 1980).
Both plant types discriminate against the heavier 13C isotope, but by differing amounts, so that
their /I 13C values are always negative. Average /I 13C values for C3 plants and C4 plants are
-27%0 and -13%0 (Deines, 1980), respectively, and approximates that of the soil respired CO2,
Additionally, because the diffusional coefficient of 12C02 is greater than 13C02, CO 2 that
remains in the soil and occupies pore space is about 40/00 heavier than that respiring from the
soil (Ceding, 1984). Therefore, a pure C3 and C4 plant community should yield soil CO2
/ll3C values about -23%0 and -9°/00, respectively.

An "open system" is assumed during equilibrium-based CO 2(g)-C0 2(aq) -HC03" -


Caco3 carbon phase changes. In this system there is complete isotopic exchange between the
CO2 and HC03" carbon species in the presence of an unlimited reservoir of soil CO 2 before
carbonate precipitation (Magaritz and Amiel, 1980; Rabenhorst, 1983; Ceding, 1984; Quade
et al., 1989). Equilibrium is maintained during the phase changes of Esuation (2) because
carbonate precipitation proceeds slowly relative to the flux of soil respired CO2 (Ceding,
1984). As a result, carbon in the precipitated carbonate species is derived from soil CO2 with
little, if any, contribution from lithogenic carbonate sources. Due to isotopic discrimination
during physical chemical fractionation of the CO2(g)-C02(aq)- HC03"-CaC03 carbon phases,
precipitated carbonate will be enriched in 13C by about 10.2%0 relative to the gaseous phase
at 20· C (Emrich et al., 1970). Because virtually all CO2 consumed during the dissolution
and reprecipitation of carbonate is derived from soil CO2, the isotopic composition of
pedogenic carbonate will be equal to that of soil organic matter and the sum of the diffusional

2"1l
and chemical equilibria fractionations. Knowing this combined value and that of the parent
material (lithogenic) and bulk soil carbonate, the percentage of pedogenically derived
carbonate in calcareous soil systems can be ascertained using Equation (1).

For this study c3 H C values were determined on organic carbon. Considering


diffusional fractionation, a factor of 4% was added to this value to arrive at an approximation
for the c3 H C value of soil CO2 , To this value was added the sum of carbonate equilibria
fractionation factors equalling about 10%0. The total value added to the original c3 H C value
of organic carbon is therefore 14°Alo, which should approximate the c3 H C value for pedogenic
carbonate.

The next section is a discussion of the alluvial stratigraphy of Cowhouse Creek taken
from Nordt (1992).

Alluvial Stratigraphy of Cowhouse Creek

Terrace 2 (TI)

The oldest terrace bordering Cowhouse Creek, TI, is situated approximately 10 to 15


m below the Killeen surface, 16 to 18 m above the modern channel, and 5 to 7 m above T1
(Figs. 2, 4 and 5). Depositional remnants of TI are preserved on both sides of the modem
stream and suggest a relict valley width of up to 2 km. This relatively flat and featureless
surface supports soils mapped mainly as part of the Lewisville series (Table 1)(McCaleb,
1985). Pit and cutbank exposures show that the alluvial fill beneath TI, named the Jackson
alluvium (unit A), rests unconformably on the Glen Rose Formation. Thickness of the
Jackson alluvium ranges from 2.5 to 4 m. Trenches and several partial cutbank and pit
exposures reveal two alluvial facies. The lower facies consists of laterally accreted channel
gravels and sands (sometimes cemented) that exhibit trough cross stratification and minor
channel fills. The upper facies consists 1 to 2.5 m of laterally and vertically accreted, yellow
to gray, bioturbated loam from which the Lewisville soils formed. A bulk humate
radiocarbon age of 15,230 B.P. was obtained from a fine-grained channel fill about 2.8 m
below the modern TI surface (Fig. 4, cutbank 10; Fig. 5). Charcoal-humate radiocarbon pairs
in Holocene sediments demonstrate that humate ages are consistently 300 to 1700 years older
than charcoal ages. The 15,000 year age for the Jackson alluvium is therefore an
approximation of the time at which soil formation began for Pedon 4. Blum and Valastro
(1989) recognize several Pleistocene terraces from the Pedemales River of the Central
Edwards Plateau, the lowest terrace of which yielded a bulk humate radiocarbon age from
fine-grained sediments of about 18,000 B.P. This may suggest synchroneity of latest
Pleistocene deposition across the Plateau.

Terrace 1 (T1)

Terrace 1 (Tl) constitutes the broad, featureless paired surface immediately bordering
TO and the channel of Cowhouse Creek (Figs. 4 and 5). Three major stratigraphic units, the
Georgetown alluvium (B), Fort Hood alluvium (C), and West Range alluvium (0), are

2·12
EXPLANATION

c:::=J TO - MODERN CHANNEL A~D FLOODPLAINS


C3 TI - HOLOCENE TERRACE

~ T2 - PLEISTOCENE TERRACE

TRENCH LOCALITIES
• CUTBACK LOCALITIES

-t- UTM COORDINATES

OEOFlOETO'lll'N FlO

~a-a

+
CO.'HO'JSE CREEK

~1-13
+

K"

W(IT II"-NIIIIO.

Figure 4. Geomorphic surface map of upper Cowhouse Creek.

2-13
~

6o.0±16o. (TX87OO)
~ 15,270±26o. _ )
190±90 (8oIa371lOO)
,650±15o. (TX8701)
}

2720±11o. (451)
&,
~ T2 300±100 [IX"'7) 15o.0±6o. (801838"4)
<:>.. 370±18o. [IX....) 1690±90 _,<50) .
30.1 0±11 0. [lX1J703) 5210±23o. 1""""452)
390±6o. 1"""""'77)
<
T1
Pedon3

Ped~n 1 ,'------..--....:,-"1
"
TO
,",
C
./' 01

~I?\ E / 02, /
!
C B

Glen Rosa limeston,


/./{ '\. A 6:::: :::: :546O±701 ...
.,.,IltI
5740±300-(GX15602
\" 8830±7o.l ....3:lOn 6850±90\BOi03'618)
Z 395D:t290:/T'x~~bl)1 .:
23B0±15o. ~6~02tgg:4170±10o.~
2860±5Q<1..;.,i"h)

EXPLANATION

Alluvial Units
A -Jackson • Radiocarbon age (B.P.)
B - Georgetown
TO - Landform
C - Fort Hood
01 -lower West Range Royalty Paleosol
TTTi['Tl'
I&n
02 - upper West Range rrrn Other soil
E - Ford

Figure 5. Generalized composite geologic cross section of Cowhouse Creek showing the recognized stratigraphic units and trench localities from which the
soil chronosequence was developed. Charcoal ages are shown in bold and bulk humate ages in plain text (from Nordl, 1992). Valley width is approximately
1 km. .
associated with this terrace. In most areas the Fort Hood and West Range alluvium crop out
to the same surface elevation and therefore cannot be differentiated by geomorphic mapping,
thus making Tl a diachronous terrace. The Georgetown alluvium is the oldest unit beneath
Tl. It is always deeply buried, rests unconformably on a scoured bedrock valley floor just
above the thalweg of the modern channel, and is unconformably overlain by the Fort Hood
alluvium. The Georgetown alluvium consists of 1 to 2 m of fme, well sorted, channel gravels
overlain by 1 to 3 m of massive and bioturbated, yellow to gray, very fme sands to sandy
clay loarns. The upper portion of the Georgetown is recognized as a truncated paleosol
named the Royalty. This soil is represented by a brown Bk horizon having encrusted stage I
carbonate morphology. Modern analogues show that encrusted filaments and threads of
calcium carbonate form by the downward gravitational flow of water carrying Ca+2 and HCO',
ions and the subsequent precipitation of Caco, on ped faces with desiccation (Sobecki and
Wilding, 1983). Groundwater carbonates tend to be disseminated throughout the fme-earth
matrix because of the preferential upward flow of water by capillary forces. It is possible,
however, that the Royalty paleosol contains Caco, formed by both processes.

The bulk of sediment associated with Tl is the Fort Hood and West Range alluvium
(Figs. 4 and 5). The Fort Hood alluvium overlies the Georgetown alluvium unconformably
and comprises about one-half of the Holocene valley fill. Cutbank exposures show that the
Fort Hood alluvium channel regime completely removed the Georgetown alluvium in
numerous localities, in which case the total alluvial fill is up to 10 m thick. Well sorted, fine
to medium basal gravels, commonly contained in beds 1 to 3 m thick, are typically overlain
by thick, fme-grained, yellowish brown floodplain deposits, which in some cases enclose fme-
grained channel fills. Radiocarbon ages obtained from a basal channel fill and from upper
flood basin deposits, respectively, show that deposition of the Fort Hood alluvium had begun
by 6850 B.P. and was still ongoing at 5200 B.P., but to no later than 4170 B.P. (Fig. 5).
Blum and Valastro (1989) recognize early Holocene stratigraphic units similar to the Fort
Hood alluvium on the Pedernales and Colorado/Concho Rivers that range in age from about
10,000 to 5000 B.P.

The West Range alluvium crops out immediately adjacent to the modern, entrenched
Cowhouse Creek floodplain (TO) (Figs. 4 and 5). It is laterally inset to the Fort Hood
alluvium on the inside of modern meander bends, and in some places partly truncates and
overlies the Fort Hood alluvium on the outside of modern meander bends. The West Range
channel system trenched down to nearly the same depth as the Georgetown and Fort Hood
alluvium before aggrading. Aggradation occurred in two episodes between 4200 and 600
B.P., which was punctuated by a brief erosional period 2000 to 3000 B.P. The lower member
spans the early part of this episode and is called the lower West Range alluvium (D1), while
the upper member spans the latter part and is called the upper West Range alluvium (D2).
Both the Fort Hood and West Range alluvium accreted to the same surface elevation in much
of the valley. However, the upper West Range alluvium in the upper Cowhouse Creek basin
is laterally inset to the lower member, but at a slightly lower elevation (1m). Downstream,
the upper member overlaps and buries the lower member and the truncated Fort Hood
alluvium. Because both members of the West Range alluvium crop out beneath Tl, soil age
can vary considerably. Sedimentological character is similar for both

2·15
West Range units. Basal deposits typically consist of several meters of relatively coarse
gravels and sands contained in large sets of laterally accreted point bars. Brown and dark
brown loamy overbank deposits either directly overlie channel gravels and sands, or laterally
truncate and partly overlap, the Fort Hood alluvium.

Degree of soil development in the Fort Hood and West Range alluvium reflects
differences in soil age. However, both these soils have Stage I carbonate morphology of
encrusted fIlaments and threads in Bk horizons that meet the criteria for a calcic horiwn.
Soils associated with both alluvial units are mapped as part of the Lewisville and Bosque
series.

Modem Floodplain (TO)

The modern Cowhouse Creek is a meandering stream that is entrenched by about 9 to


10 m relative to the T1 surface (Figs. 4 and 5). Acute channel bends suggest that there may
be partial structural control from the underlying limestone bedrock. Cowhouse Creek is an
intermittent perermial stream with numerous periods of no flow. However, the surrounding
limestone uplands produce steep slopes, sparse vegetative cover, and shallow soils. Coupled
with high intensity convectional rainstorms, peak-flood discharges of Central Texas streams
are extremely large and episodic. Baker (1975) has shown by magnitude-duration flood
curves that Central Texas has rainfall runoff events that rank among the worlds greatest.

Modern Cow house Creek sediments are named the Ford alluvium (unit E) and consist
of several facies: channel lag, laterally accreted point bar sands and gravels, and overbank
sands and clays. Modern point bars have accreted to several meters above the low-flow
channel as a result of wide variations in stream discharge. Most overbank deposits are
located on the innermost portions of large meanders and are inset to, and lower than T1 by
about 2 to 3 m (Figs. 4 and 5). Along straight channel segments vertically accreted deposits
as much as 7 m thick may occur on both sides of the channel. Overbank sediments of the
Ford alluvium are highly stratified with differences in texture indicating coarse-grained
deposition during initial overbank flooding and fine-grained deposition during flood recession.
A suite of radiocarbon ages, all determined from dispersed and culturally-derived charcoal,
demonstrate that modern channel trenching occurred sometime between 600 to 400 B.P. (Fig.
5) and that deposition has been ongoing since. Numerous alluvial investigations in Central
and North-Central Texas demonstrate that a widespread episode of valley trenching occurred
sometime around 1000 B.P. (Hall, 1988, 1990; Mandel, n.d.; Blum and Valastro, 1989,
Ferring, 1990) and again, events at Cowhouse Creek appear to generally concur with the
regional alluvial record. The soils developed in Ford alluvium are also mapped as part of the
Bosque series.

Stratigraphic Summary and Paleoclimatic Inferences

Radiocarbon ages from around the Military Reservation suggest that widespread
channel trenching was proceeding sometime between 15,000 and 10,000 B.P. This event
abandoned the 1'2 floodplain and initiated scouring of the Holocene valley. Bryant and

2-16
Holloway (1985) have shown by pollen analysis that climatic conditions changed from
relatively cool and wet to wann and dry in Central Texas between 14,000 and 10,000 B.P.
Nordt et al. (1994) also show that cool season C3 species constituted as much as 60% of the
biomass production in the region approximately 15,000 B.P.. Blum and Valastro (1989)
believe that this change to drier conditions near the end of the Holocene induced widespread
valley trenching of the Pederna1es River. These climatic conditions may also account for
channel degradation of Cowhouse Creek at this time.

Widespread valley alluviation of the Georgetown alluvium began sometime after


valley entrenchment between 15,000 and 10,000 B.P. and had ended by 8500 to 8000 B.P.
This channel regime apparently carried relatively uniform, nonepisodic base flow with local
high water tables, as reflected in the fme-grained and well sorted sedimentological character
and mottled sediment colors. In addition, the vertical relief between the inferred low water
channel and the overbank sediments of the Georgetown alluvium is several meters less than
units deposited later in the Holocene. This suggests that the younger Holocene units were
deposited under more flashy stream discharge conditions than the Georgetown. Given that
climatic conditions were sufficiently moist to support an expansive upland vegetative cover
during the early Holocene, rapid upland knickpoint migration, in response to the lowered base
level of the trunk drainage network, may not have occurred. This would have limited the
source of Georgetown alluvium to reworking of Jackson sediment. The Royalty paleosol
contains carbonatic filaments and threads on ped faces that indicates some period of landscape
stability or slow cumulic deposition. Because this soil was truncated at about 7000 B.P.,
perhaps as much as 500 to 1000 years of soil formation occurred cumulicly.

Between 7000 and 4500 B.P., much of Central to Northwest Texas was undergoing a
warming trend (Hall, 1988; Holliday, 1989, Nordt et al. (1994) that was accompanied by slow
fluvial deposition and soil formation (Ferring, 1990), eolian sedimentation (Holliday, 1989;
Caran and Baumgardner, 1990), and fan deposition (Blum, 1989; Mandel, n.d.). This climatic
shift to warmer and drier conditions may have been the catalyst for brief channel trenching or
stream metamorphosis as low frequency, high intensity rainstorms increased. Following this
model, deposition of the Fort Hood alluvium proceeded as the surrounding uplands became
progressively depleted of vegetation and resulted in increased runoff and sediment yields.
This sediment was most likely derived from thick, fme-grained upland soils that formed
during the Pleistocene as the newly entrenched tributary network migrated into the Killeen
surface. The reddish-yellow hue of the Fort Hood sediments suggested that the eroded upland
soils were well oxidized.

Between 4800 and 4200 B.P. the channel network throughout Fort Hood adjusted to an .
increase in coarse-grained sediment load. This shift to coarse-grained sedimentation, which
resulted in deposition of the West Range alluvium between 4200 and 600 B.P., may be
explained by one of two working models. One postulates that as Holocene drying continued,
and many of the deep fme-grained soils of entrenched sections of the Killeen surface were
eroded, larger proportions of parent rock were eroded and delivered to the trunk streams. The
increased abundance of bedload sediment and subsequent imbalance between sediment load
and discharge may have induced channel metamorphosis by decreasing sinuosities and

2-17
increasing width-depth ratios, even in the absence of a significant shift in stream discharge.
Previous investigators have demonstrated this important relationship, which mayor may not
be climatically induced, between drainage basin soil sediment supply and fluvial erosional and
depositional cycles (Bull, 1991; Tonkin and Basher, 1990).

An altemative mechanism for delivering coarse-grained sediment to the trunk channel


network after 4200 B.P. is an increase in discharge in response to wetter and possibly cooler
climatic condition. Such a shift in climate at this time has been interpreted by Nordt et al.
(1994) and Holliday (1985a,b). Wetter conditions could have stabilized upland hillslopes by
increasing vegetative cover, reducing sheet erosion, and effectively slowing the rate of
knickpoint migration into the Killeen surface. Continued downcutting within the previously
entrenched tributary network would have provided significant quantities of parent rock for
delivery into trunk streams. The relatively higher stream discharges at this time would
presumably have had the competency to carry the newly introduced coarse-grained sediment.
In support of this model, Blum and Valastro (1989) attribute coarse-grained sediment
transport by the Pedemales River between 4000 and 1000 B.P. to moist climatic conditions
that produced higher effective discharges and competency levels.

Nordt et al. (1994) show that climate conditions because dryer for a brief period
between 3000 and 2000 B.P. This also coincides with the age of an erosional conformity
separating the lower and upper West Range alluvium. Radiocarbon ages from Cowhouse
Creek also demonstrate that stream metamorphosis, and brief channel trenching, occurred
between 600 and 400 B.P. However, the channel facies of both the West Range and Ford
alluvium are at nearly the same bedrock basal level indicating minimal channel trenching.
Rapid fluvial adjustments in response to a continued coarsening of the sediment load or shift
in discharge, without significant channel incision, must be considered as a possible cause of
stream metamorphosis at this time.

Soil Chronosequences

Development of the alluvial stratigraphy of Cowhouse Creek is nearly complete. Soil


interpretations, however, are in their preliminary stages. The following is an attempt to
express our current ideas and indicate approaches we are taking in efforts to establish
relationships between the observed soil properties and soil age.

The traditionally recognized factors influencing soil formation are climate, biota,
parent material, topography, and time (Jenny, 1941). If the first four factors are held
constant, with only time varying, a soil chronosequence, and possibly chronofunctions, can be
established. The chronological alluvial sequence of Cowhouse Creek was established in the
previous section. Based on the stratigraphy, assumed soil ages of the four soil pedons under
consideration on this field trip, are listed in Table 2.

2-18
Table 2. Stratigraphic unit ages and estimated soil ages of the study pedons.

Unit No. (Terrace) Unit Age Range (B.P.) Pedon Soil Age (B.P.)

Ford (E) TO 600 to present 1 modern


Upper West Range (D2) T1 4200 to 600 2 2000
Fort Hood (C) T1 7000 to 5000 3 5000
Jackson (A) T2 15,000 4 15000

The age of Pedon 1 is essentially modern (TO) because of active flooding and
deposition. The minimum age of Pedon 2 is 600 B. P. because channel trenching,
abandonment of T1, and termination of deposition of the West Range alluvium (D2) occurred
at about that time. The actual age of Pedon 2 can be deciphered from Figure 6, which
displays a schematic cross section of an alluvial exposure near Pedon 2 (Fig. 4; Trench 1). A
channel fill within the upper West Range alluvium (02) is shown beveling and overlying the
lower West Range (D1), Fort Hood alluvium (C), and Georgetown alluvium (B). This
demonstrates that active lateral channel migration was ensuing around 2380 B.P. Pedon 2
parent materials can be traced in beds dripping toward the buried channel that dates to 2380
B.P. This suggests a soil age of about 2000 years for Pedon 2.

Radiocarbon ages from around Fort Hood suggest that where the Fort Hood alluvium
was not truncated by West Range erosion, deposition terminated between 4800 and 5200 B.
P. (Fig. 5).. As a result, the soil age of Pedon 3 (associated with TI) is assumed to be 5000
years. The soil age of Pedon 4 (associated with T2) is more uncertain because it is based on
a single humate age. However, an age of 15,000 years provides an approximation until
further ages are obtained (Fig. 5).

Pedon 1 - Trench 5

Pedon I-Trench 5 is excavated in the Ford alluvium associated with TO, which is inset
to T1 by about 2 m. This area represents the modern floodplain that is actively aggrading
(Figs. 4, 5). Terrace 0 is mapped as part of the Bosque rarely flooded map unit (McCabeb,
1985). A radiocarbon age of 650 B.P. was determined on hearth charcoal from a depth of
383 cm beneath this surface and demonstrates that Pedon 1 is less than 650 years old. In
other localities, ages as young as 190 B.P. show the youthfulness of the Ford alluvium. A
second age of 2100 B.P. was obtained on hurnates from a depth of about 7 m at the cutbank
near pedon 1 and indicates that the upper West Range alluvium partly underlies the Ford
alluvium and TO in this position. Pedon 1 formed in stratified sediments typical of the Ford
alluvium as can be seen by depth distributions of the percentages of clay-free sand and silt,
and organic carbon content (see pedon 1, Appendix). Carbonate values show that at least
50% of the matrix is comprised of calcite, virtually all of which is lithogenic. Two faintly
expressed buried A horizons are evident within this unit indicating that deposition was
occasionally puntuated by brief periods of landscape stability and soil formation. Few to

2-l9
-v
"-
t ,,\.<,
(\of
,
T1 \
j; J

02 (upper)
c

• 238D±150

883D±70
Glen Rose limestone. B

'"~
""'" Surface soil in upper West Range alluvium (02) (Pachic Calciustolls)

TTT Truncated and buried soil in lower West Range alluvium (01)
iii i Truncated and buried soil in Fort Hood alluvium (C)

o Stratigraphic units
• Radiocarbon ages (charcoal)

Figure 6. Schematic cross section of alluvial and soil stratigraphic units beneath T1 of Cowhouse Creek (cutbanks
1 and 2 near West Range Road in Figure 4) (no scale-total alluvial thickness approximately 9 to 10m).
'7 rDf1
many, fme patchy carbonate pseudomycelia are on ped surfaces of loamy Bk horizons.
Because of the presence of an ochric epipedon, absence of a cambic endopedon, and an
irregular distribution of organic carbon with depth, this pedon classifies as a Typic
Ustifluvent; coarse-loamy, carbona tic, thermic.

Pedon 2 - Trench 1

Pedon 2-Trench 1 was excavated in sediments of the upper West Range alluvium (02)
associated with Tl, and like TO, is mapped as part of the rarely flooded Bosque map unit
(Figs. 4, 5, 6). Depth distributions of clay-free sand and silt percentages and organic matter
content (pedon 2, Appendix) show slight stratification in the original soil parent material,
although it is not as great as in Pedon 1 of the Ford alluvium. The bulk soil of Pedon 2
contains slightly fewer total carbonates than Pedon 1 as values range between 40 and 50%.
The Bk horizons of Pedon 2 described below a depth of 85 cm, contains between 5 and 15%
n
carbonate filaments and threads (Stage in ped interiors and on exteriors. These Bk
horizons meet the requirements of a calcic horizon because of greater than 15% calcium
carbonate equivalent and more than 5% by volume of identifiable secondary forms
(disseminated films and threads). Because of the calcic horizon and "overthickened" mollic
epipedon, this pedon classifies as a Pachic Calciustoll; fme-Ioamy, carbonatic, thermic.

Dithionite-extractable Fe (Fe.J values in the solum of Pedon 2 are slightly higher than
those of the parent materials for the Ford alluvium. Assuming that Ford and West Range
alluvium parent material values approximate each other, some accumulation of iron oxides
has occurred in Pedon 2 during 2000 years of pedogenesis.

Pedon 3 - Trench 19

Pedon 3-Trench 19 is located in the Fort Hood alluvium on the outer margin of Tl
near the Cretaceous valley wall (Figs. 4 and 5). The parent material that Pedon 3 developed
in is somewhat different than Pedons 1 and 2, in that Pedon 3 has 1) greater clay and silt
contents, 2) less total carbonate, and 3) higher Fed contents. As a result, caution must be
taken in differentiating pedogenic differences that are related to parent material and those
related to in situ weathering and the passage of time. The increase in Fed content in the mid-
profile position of Pedon 3 does suggest, however, some increase in the release of pedogenic
Fed with time. Significant clay translocation has not been initiated because of the absence of
clay films and the absence of an increase in the fme clay-total clay ratio with depth.

Carbonate morphology of Pedon 3 is also Stage I; however, filaments and threads are
thicker and more deeply encrusted into ped surfaces than in Bk horizons of Pedon 2. Greater
than 5% by volume of carbonate forms, coupled with more than 15% total carbonate in the
Bk horizons, substantiate the presence of a calcic horizon. As with Pedon 2, this soil has a
"overthickened" mollic epipedon. This soil has sufficient COLE values and cracking when
dry to be vertic; therefore, because vertic character has precedence over pachic, this soil is
classified in the Vertic Calciustolls; fme, montmorillonitic, thermic family.

2-21
Pedon 4 - Trench 13

Pedon 4-Trench 13 was opened in the Jackson alluvium associated with T2 (Figs. 4
and 5). Reconstruction and rnicromorphic techniques were performed on this pedon to make
inferences about the amount and kind of changes that soil properties have undergone since
pedogenesis began (Brewer, 1976). Applying this methodology assumes unifonnity in parent
material; however, particle size distribution expressed on a carbonate-free and clay-free basis
for Pedon 4, along with Zr and Ti contents, indicates that there is some stratification within
the solum (see Pedon 4, Appendix). Nevertheless, this analysis proceeds under the
assumption that parent material is sufficiently uniform to estimate volume changes of marker
constituents with time.

Reconstruction analysis shows a net loss of 69 and 50 g per cm2 of sand and silt-sized
carbonates in Pedon 4, respectively, relative to the assumed parent material (see Pedon 4,
Appendix). These losses are occurring from dissolution of carbonates in the sand and silt
fractions, which make up about 50% of the parent material. Most of this loss is in the upper
98 cm where carbonates have either been dissolved and moved to lower portions of the pedon
or completely leached from the solum. Although there is a net increase in sand, silt, and total
carbonate in the 150-250 cm section (Bk3 through Bk5 horizons), there is substantial loss of
lithogenic carbonate and substantial gain in pedogenic carbonate (see next section). Overall,
these data indicate a total loss of carbonates from Pedon 4 of 133 gjcm', or a rate loss of 8.9
gjcm'll000 years. Furthermore, as weathering proceeds, occluded clays are released from the
carbonates and left as residue, thus accounting for a higher proportion of the total mass of the
soils than in the parent material. Further support for net carbonate removal is exhibited in the
pitted nature of lithogenic carbonate clasts contained in the upper BAk and Bkss horizons.
Thin section analysis shows that these clasts are cretaceous renmants and not of pedogenic
origin because they contain Cretaceous fossils, do not engulf matrix particles, do not show
stress features around the nodules that would indicate incremental growth, do not have iron
manganese stains, and do have quartz rinds resulting from dissolution. Delta 13C values on
three of these nodules from different depths were all near + 1.0, which is in range of values
from Glen Rose limestone. These data indicate that the pits in the nodules are a result of
dissolution and leaching of carbonate. Dissolved carbonates may subsequently reprecipitate
deeper in the solum, or be removed from the solum via percolating waters.

As a result of carbonate dissolution, reconstruction analysis also shows a net gain of


19 g of fme clay per cm' colunm for Pedon 4. Only the Al horizon shows a fme clay loss
while the zone from 14 to 120 cm exhibits substantial fme clay gain. Patchy clay films in the
lower solum and fme clay-total clay depth distributions indicate that clay translocation has
occurred. However, because clay films constitute less than 1 percent of the surface area in
thin section and the amount of clay increase is inadequate, the minimum requirements for an
argillic horizon are not met. Because of high clay content and COLE values, and evidence of
shrink-swell in the form of slickensides in Bkss horizons, this pedon classifies as a Udic
Chromustert; fme, montrnorillonitic, thermic. Also note, from a chronological viewpoint, that
relative to Pedons 1, 2, and 3, Pedon 4 has redder hues and lower pH values in the upper
solum.

2-22
Pedogenic Carbonate Chronosequences

Bulk carbonate and organic carbon samples from selected A and Bk horizons of
Pedons 2, 3, and 4 were subjected to stable carbon isotopic analysis in an attempt to
differentiate pedogenic from lithogenic carbonates and to quantify pedogenic carbonate
accumulations (for theory see Methods section). Vertically superposed horizons containing
similar field estimated quantities of disseminated filaments and threads were grouped together
for analysis.

The &13C values for the organic carbon were all taken as -19.00Alo for all pedons based
on reconstructed vegetation history of the area (Nordt et al., 1994). In other words, during
pedogenesis for the last 15,000 years, -19.00/00 gives a reasonable representation of the
proportion of C3 and C4 plants growing in the area (Table 3). Given a 13C enrichment factor
of 14%0 in the organic matter-C02W -CaC03 system, pedogenic carbonate in the soils should
have &BC values around -5%0.

Samples from the Glen Rose limestone, considered to be the predominant source of
soil lithogenic carbonate in the alluvium, were analyzed to determine their BC composition.
Values of +1.5 and +2.8%0 were obtained from two Samples (Table 3). The lower value
(+ 1.5) was used as the parent material value in Equation (I) for calculating percent pedogenic
carbonate because the higher value did not fall within the ranges published by West et al.
(1988) for limestone-derived parent materials of Central Texas. All carbonate accumulation
in this area is assumed to be derived from in situ carbonates with little contribution from
airborne influx based upon monitoring of dust in the region (Rabenhorst, 1983; West, 1986).

Figure 7 displays total carbonate and calculated pedogenic carbonate contents of


Pedons 2, 3, and 4. Note that the percentage of pedogenic carbonate does not increase
systematically with time. Furthermore, pedogenic carbonates are somewhat uniformly
distributed throughout the profiles indicating a wide range of effective wetting depths.
Although a parent material Caco3 content is difficult to determine because of deep
weathering, all pedons have probably lost carbonate, most noticeably in Pedon 4.

A more revealing indicator of secondary carbonate accumulation in soils is the whole-proftle


index (Machette, 1985). This index is used to measure the cumulative total of grams of pedogenic
carbonate in a soil column I cm2 in diameter. It is calculated from bulk density, percent pedogenic
carbonate, and horizon thickness. We used moist bulk density values to eliminate volume changes
associated with shrink-swell soils. In addition, as revealed by the adjacent cutbank exposure, the
sola of Pedon 2 is 25 cm deeper than originally described. This thickness was added to the last
horizon for calculating the index. We also extended the pedogenic carbonate content of the last
horizon of Pedon 4 down to the same depth as Pedon 3. It is obvious from Figure 7 that pedogenic
carbonate exists in horizons above and below those horizons from which we have data. Our
calculations therefore slightly under estimate pedogenic carbonate accumulations.

The whole-profile index shows that Pedons 2, 3, and 4 have accumulated about 22, 69, and
41 grams of secondary carbonate per cm2 column (Fig. 8). For Pedon 2, pedogenic carbonate has

2-23
Table 3. Stable Carbon Isotope Analysis.

6 13C

Pedon No. (Unit) Depth Horizon Bulk Carbonate Organic Carbon

(cm)

2 Upper West Range


Unit 02 -19.0
85-155 Bk3&4 0.2
155-166 BkS 0.0
166-215 Bk6&7 0.1
215-268 Bk8 0.2
268-347 Bk9 0.2
3 Fort Hood
UnitC -19.0
59-94 BW . -1.6
94-122 Bkl -1.9
122-153 Bk2 -1.7
153-189 Bk3 -1.7
189-252 Bk4 -1.7
252-320 BkS -1.8
320-370 Bk6 -1.4
4 Jackson
Unit A -19.0
38-98 Bkss2 -2.1
98-120 Bkl -1.2
120-150 Bk2 -0.4
150-190 Bk3 0.2
190-250 Bk4&5 -0.1
250-270 BC 0.2
270-350 CB 0.4
Limestone Sample 1 1.5
Limestone Sample 2 2.8

2-24
%CaC03 20%caC03 %CaC03
0 20 40 60 0
0

~
Peden 2 Peden 3 Peden 4
2000 B.P. 5000 B.P. BAk 15,000 B.P.
:1 I AB
50
t-
Bk2

100
~

E
:2- 1-150
a.
'"
0
'"N
'-"
200

250

300

350
C=I Total calcium carbonate equivalent
';';;::::«:::";'~~'J
'%w!l'~. Total pedogenic carbonate

Figure 7 . Mean pedogenic carbonate and total calcium carbonate contents for Pedons 2, 3, and 4.
been accumulating at a rate of about 11 g/cm2/1000 years and for Pedon 3 at a rate of about
14 g/cm2/1000 years. These rates are comparable and suggest a relatively uniform carbonate
accumulation rate over the last 5000 years. If the rate of accumulation of Pedon 2 is
projected over the last 15,000 years, Pedon 4 should contain about 150 grams of pedogenic
carbonate. Because Pedon 4 has only 41 grams, as much as 11 0 grams or about 73 percent,
of the previously accumulated pedogenic carbonate has been removed from the soil solum.
Net carbonate loss by these calculations support our conclusions based on reconstruction
analyses. Holocene alluvial soils have secondary carbonate accumulations rates much greater
than those listed for the arid Southwest where rates are normally less than 1 g/cm2/1000 years
(Machette, 1985). Even though the calculated pedogenic carbonate percentages approximate
our field estimations, it is possible that parent materials contain small quantities of pedogenic
carbonate inherited from erosion of soils from the surrounding uplands. Additional parent
material samples will be analyzed in the future to make this determination.

Iron Oxide and Organic Carbon Chronosequences

Our data show that Fed increases linearly with time (Fig. 8). McFadden (1988) has
shown that in semiarid to arid regions of the Southwestern U.S., late Pleistocene and middle
Holocene soils have accumulated 0.5 to 2.5 percent Fed in the clay fraction. Furthermore,
Miles (1981) calculated a Fed percentage of 1.5 percent from a late Pleistocene terrace soil of
the Brazos River in North-Central Texas. Nordt (unpublished) calculated an Fed percentage
of 1.2 from a late Pleistocene Leon River terrace soil in the Fort Hood area.

Organic carbon distributions in the upper 50 cm show that similar quantities have
accumulated in the three pedons under study (Fig. 8). In addition, accumulation commonly
reaches steady state within 2000 years and may decline slightly with increasing age up to
15,000 years. Birkeland (1984) shows data generally supporting these results.

Soil Order Chronosequences

Even though currently aggrading, the surface horizon of Pedon 1 of the Ford alluvium
qualifies as a mollic epipedon except for light color values. TItis demonstrates that even
though surface age is constantly being rejuvenated, minimal time is needed to form a mollic
epipedon and Mollisol. The other three pedons have well expressed mollic epipedons with
similar quantities of organic carbon in the upper 50 cm.

It has been shown in this study that calcic horizons form within 2000 years because
rainfall is sufficient to rapidly dissolve and translocate lithogenic carbonates, but not so great
as to remove all of the dissolved carbonate by-products from the solum before secondary
carbonate precipitation occurs. Pedogenic carbonate accumulation appears to have been
continuous through time. However, sometime between 5000 and 15,000 B.P. appreciable
pedogenic carbonate depletion also becomes apparent. Petrocalcic and calcic horizons in soils
of the Killeen surface are poorly expressed and seem to be confmed more to areas where
limestone is situated in the near surface, which may contribute to petrocaicic

2-26
200
5000 B.P.

~
160 J
2000 B.P.
n 1-2.0 c
E
::J
c 0
()
E C\I
::J
15,000 B.P.
o 120 1.5 E
~
()

C\I ~

E
c
~ 0
.0

-
~
~

CD 80 1.0 ~
<ll
,
N
c ()
N
-J 0
.0
'c
<ll
~

<ll Modern 0)
~

0
40
• .5
0
'0
c
<ll
'0
Ql
U.

0
0 Pedon 1 Pedon 2 Pedon 3 Pedon 4

Carbonate

Total
Pedogenic
lIB Fed
- Organic carbon

Figure S. Pedogenic carbonate, dithionite-citrate extractable iron oxide (Fed» and organic carbon in g/cm2 column for
Pedons 2, 3, and 4.
horizon fonnation by in situ carbonate dissolution and reprecipitation. On the other hand,
these portions of the Killeen surface may have been eroded during the Holocene, thus
removing upper soil horizons and bringing the petrocaicics into near-surface positions. In
thick clayey soils on the Killeen surface, most of the carbonates have either been transported
to deep portions of the profIle or completely removed from the sola. Both processes are
consistent with the model of net carbonate loss coupled with apparent short -tenn pedogenic
carbonate accumulation shown for the alluvial soils.

Argillic horizons have not been identified in Fort Hood except in areas on the Leon
River terraces where mixed and siliceous parent material mineral assemblages occur. Clay
fIlms in the solum of Pedon 4 of Cow house Creek indicate that clay movement does occur
and becomes observable in 5000 to 15,000 years, but it is not the primary pedogenic process.
Interestingly, clay quantities increase with time as occluded clays are released during sand
and silt carbonate dissolution. As a result, vertic properties become more prevalent with time,
thus destroying many illuviation features.

Holliday (l985a, 1985b) has evidence demonstrating rapid rates of pedogenesis in the
Southern High Plains of Texas. He found that mollic, calcic, and argillic horizons could fonn
withing 2000 years. Rapid fonnation of calcic and argillic horizons was attributed in large
part to carbonate eolian influx. Within 450 years· all soils in his study developed into
Inceptisols, Mollisols or Alfisols. In the Fort Hood area, Entisols (Ustifluvents) develop
rapidly into Inceptisols and Mollisols (Hapustolls and Calciustolls), in part because of the
preweathered nature of the soil parent materials. The vertic subgroup of Pedon 3 may arise
from the fme-grained nature of the parent material; however, reconstruction analysis of Pedon
4 indicates that occluded clays are released from sand-sized and silt-sized carbonates and
increase in quantity with time. The development of a Vertisol (Chromusterts) in Pedon 4
from loamy carbonatic sediments within 15,000 years supports this conclusion. These
relationships are illustrated in Table 4.

Table 4. Summary of fonnation of diagnostic features with time for the alluvial soils of
Cow house Creek, Fort Hood, Texas.

Soil Age (Years Before Present)


Modern 2,000 5,000 15,000
Diagnostic Ochric mollic mollic mollic
Horizons (mollic l) calcic calcic calcic
Great Group Ustifluvents Calciustolls Calciustolls Chromusterts
(Haplustolls l)

lMollic epipedon may fonn quickly, or in ideal situation, be inherited from parent material.

2-28
ACKNOWLEDGEMENTS

I thank Kimbal Smith and Dr. Jack Jackson, archaeologist and environmental scientists
at the Fort Hood Military Reservation, for encouraging and obtaining permission for this
excursion. They have done so on numerous other occasions as well. Bill Roberts was in
charge of trenching operations. I also thank Glenda Kurten for patiently typing the
manuscript, John Jacob for coordinating soil characterization analysis of the four pedons, and
Drs. Larry Wilding and Richard Drees and their Agronomy 606 class for the reconstruction
analysis of soil pit 4. Special thanks to Dr. Tom Hallmark and Dr. Michael Waters for
contributing to various parts of this investigation.

2-29
· APPENDIX

Pedons 1-4

2-30
PEDON 1
SOIL SERIES: BOSQUE VARIANT PEDON: S90TX-099-OQ4 COUNTY: CORYELL

PEcaN CLASSIFICATION: TYPIC USTIFLUVENTS: COARSE-LOAMY. CARBONATIC. THERMIC


tOeA HON: FT. HOOD, TEXAS; CHST31TRf.

LANOFORM: FLOODPLAIN ELEVATION (M): SLOPE: 0-1~ SLOPE ASPECT:

PARENT ~ATERIALS: ALLUVIUM FORMATION: HOLOCENE ALLUVIUM (UNIT E)


TOPOGRAPHY: NEARLY LEVEL QRAINAGE: WELL DRAINED LANOUSE: NATIVE

COLLECTORS: JACOB ANO KORGEL DATE:

HORIZON DEPTH (C~) SOIL DESCRIPTION (COLORS FOR MOIST SOIL UNLESS STATED)

A 0-25 BROWN (10YR 4/3) LOAM. GRAYISH BROWN (fOVR 5/2) DRY; MODERATE COARSE AND
VERY COARSE ANGULAR BLOCKY STRUCTURE; HARD; FEW FINE PORES; COMMON FINE
ANa MEDIUM ROOTS: MANY VERY FINE CARBONATE CLASTS; MOOE'RATELY ALKALINE:
STRONGLY EFFERVESCENT; CLEAR SMOOTH BOUNDARY,
2BI<1 25-57 BROWN (fOYR 4/3) FINE SANOY LOAM, BROWN (10YR 5/3) ORv; WEAK COARSE SUBANGULAR
BLOCK V STRUCTURE; VERY HARD; FEW MEDIUM AND COARSE ROOTS: FEW SHELL FRAGMENTS;
~ANY WORM CASTS: FEW MVCELIAL CARBONATES: MANY FINE CARBONATE CLASTS: MODERATELY
ALKAL[NE; STRONGLV EFFERVESCENT: CLEAR SMOOTH BOUNDARY,
:lBK2 57-75 BROWN (fOYR 4/3) FINE SANOY LOAM, YELLOWISH BROWN (10YR 5/4) DRY: WEAK MEDIUM
SUBANGULAR BLOCKY STRUCTURE; HARO: FEW MEOIUM AND COARSE ROOTS: FEw SHELL
FRAGMENTS: COMMON TO MANY MVCELIAL CARBONATES: MANY FINE CARBONATE CLASTS;
FEW LARGE LIMESTONE FRAGMENTS AND PEBBLES: MOOERATELV ALKALINE; STRONGLY
EFFERVESCENT; CLEAR SMOOTH BOUNOARV,
3C 75-100 YELLOWISH BROWN (fOYR 5/4) FINE SANOY LOAM, VERY PALE BROWN (IOVR 7/3) DRY;
STRUCTURE LESS MASSIVE: SLIGHTLV HARO; FEW MEDIUM AND COARSE ROOTS: COMMON
SANO LENSES: SANDS ARE OOMINANTLV CARBONATE; COMMON DISCONTINUOUS LOAMY
STRATA WITHIN HORIZON: LOAMY STRATA HAS COMMON MYCELIA CARBONATES: MODERATELY
ALKALINE: STRONGLV EFFERVESCENT: CLEAR SMOOTH BOUNOARY,
4AKB1 100-114 BROWN (10YR 4/3) LOA~, yELLOWISH BROWN (IOYR 5/4) ORY: MOOERATE MEOIUM ANO
COARSE SUBANGULAR BLOCKV STRUCTURE: SLIGHTLY HARD: FEW MEDIUM AND COARSE
ROOTS: COMMON SCATTER EO CHARCOAL FRAGMENTS: COMMON MYCELIAL CARBONATES:
~OOERATELV ALKALINE: STRONGLV EFFERVESCENT; CLEAR SMOOTH BOUNOARV,

SBWS1 114-122 VELLOWISH BROWN (fOYR 5/4) LOAM, VERY PALE BROWN (fOYR 7/3) DRY: WEAK MEDIUM
SUBANGULAR BLOCKY STRUCTURE; SOFT; FEW MEDIUM AND COARSE ROOTS; SAND GRAINS
ARE DOMINANTLY CARBONATES: NOTABLE SAND LENSES; MODERATELY ALKALINE: STRONGLY
EFFERVESCENT: GRAOUAL SMOOTH BOUNOARY,
6AKB2 122-143 OARK BROWN (IOYR 3/3) LOAM, BROWN (IOYR 4/3) DRY: WEAK MEOIUM SUBANGULAR
BLOCKY STRUCTURE: HARO: FEW MEDIUM AND COARSE ROOTS; FEW SHELL FRAGMENTS:
COMMON MYCELIAL CARBONATES; MODERATELY ALKALINE: STRONGLY EFFERVESCENT;
CLEAR SMOOTH BOUNOARY,
7C1B2 143-171 OARK BROWN (10VR 3/3) FINE SANOY LOAM, BROWN (tOYR 4/3) ORY: WEAK COARSE
PRISMATIC STRUCTURE; SLIGHTLY HARO: FEW MEDIUM ROOTS: ~NO GRAINS ARE OOMINANTLY
CARBONATES: COMMON PATCHY AREAS OF ~YCELIAL CARBONATES; COMMON SCATTEREO
CHARCOAL FRAGMENTS: ~O~ OF HORIZON AS fOYR 5/3 (10YR 7/3. ORY) OISCONTINUDUS
SAND LENSES: MOOERATELY ALKALINE: STRONGLY EFFERVESCENT: CLEAR SMOOTH BOUNOARY,
aC2B2 171-220 OARK BROWN (fOYR 3/3) LOAM, BROWN (fOYR 4/3) ORY: WEAK COARSE pRISMATIC
PARTING TO WEAK MEOIU~ SUBANGULAR BLOCKY STRUCTURE: HARO; FEW MEOIUM ROOTS:
COMMON SANO LENSES; COMMON SCATTER EO CHARCOAL FRAGMENTS: COMMON TO MANY
MYCELIAL CARBONATES: MOOERATELY ALKALINE; STRONGLY EFFERVESCENT; CLEAR SMOOTH
BOUNOARY.
9C3B2 220-237 YELLOWISH BROWN (IOVR 5/4) VERY GRAVELLY SANOY LOAM, PALE BROWN (IOVR 6/3)
ORV: STRUCTURELESS MASSIVE: FEW MEOIUM ROOTS: GRAVEL AND COARSE SAND LENSES;
GRAVELS ANO SANOS ARE CARBONATES: MOOERATELY ALKALINE: STRONGLV EFFERVESCENT:
CLEAR SMOOTH BOUNDARY,

IOC4B2 237-266 YELLOWISH BROWN (IOYR 5/4) SANDV LOAM, PALE BROWN (IOVR 6/3) ORY; STRUCTURELESS
MASSIVE; SLIGHTLV HARD: NO ROOTS: COARSE SANO LENSES: ON TOP OF HORIZON
IS A LOAMY STRATA WITH MVCELIAL CARBONATES: MOOERATELY ALKALINE; STRONGLY
EFFERVESCENT.

2-31
Pedon 1
SOIL CHAR~CTERIZ~TICN lABORATORY
SOIL ANO CROP SCIENCES OEPT .. THE TEXAS AGRICULTURAL EXPERIMENT STATION
SOIL SERIES: 80SQUE VARIANT
SOIL FAMILY: TYPIC UST{FLUVENTS: COARSE-LOAMY, CARBONATIC. THERMIC
LOCATION: CORYELL COUNTY

PARTICLE SIZE OISTRIBUTION (M".)


-----------------SANO----------------
VC C M , v, TOTAL -----SlLT----- -----CLAY-----
FINE TOTAL FINE TOTAL COARSE
LAB (2.0- ( 1 .0- (0.5- 10.2'5- (0.10- (.2 .0- {a.02- (0.05- (< (< TEXTURE FRAG-
NO DEPTH HORIZON l.O) 0.5) O.lS} 0.10) 0.05) a.OS} 0.002) 0.002) 0.0002) 0.00:2) CLASS IoIENTS

40:14
40:15
4036
(eM)
0- 25
2'- '7
.7 - 7'
·A
26KI
2BK2
----------------------------------~----------------------------------
0.5
0.2
0.0
a .•
L,
,. ,
0.'
5
12. 7• 20.9
29.6
29.2
19.9
20.5
27 .5
48.0
64.8
GO.G
19.:3
12.2
12.3
32.6
20.1
2:1.4
12.1

to.2
.. , 19.4
14.5
16.0
L
FSL
FSL
a
a
a
%

..,
'.G
.t037 75-100 ,c o. , lS.5 :14.5 21.4 74.G B.7 12.6 7.G 1.2 . a 'SL a
4038 100-114 4AKBt o. , '.0 16.0 17.1 14.0 50.2 18.0 30.6 11.2 19 . .2 L a
4039 114- 122 5BWB1 O.G 11.3 37.0 22.8 7.' 79.G B. , 11.1 G. , LS 0
4040 122-143 6AKB2 o. , a .• '.0 15.9 18.3 .43.2 20.0 32.2 14.9 24.7 L a
4041
4042
4043
4044
143- 17 f
171-220
220-237
237-266
7CIB2
aC2B2
9C3B2
IOC4B2
o. ,
0.'
'.7
LA
1.2
1.7
15.7
'.5
13.6
10.1
25.9
24.5
33.5
19.5
11. f
19.3
.. ,
14.9
16.5

10.5
63.3
48. t
72.7
64.2
14.0
20.1
11.0
13.3
21.1
31.0
14.0
21.2
I
•••
1.6
7.'
B.8
15.6
20.9
13.3
14.6
'SL
L
SL
SL
a
a
A'
a
LAB ORGN PH -------NH40AC EXTR BASES------ KCL EXTR NAOAC BASE CAL- OoLO- CAC03 GYP
NO

4034
i!
1.09 ..
C (H2O)
1:1
~
CA MG NA K TOTAL AL
----------------------MEQ/lOOG---------------------
43.1 2. , o. , 1.0 413.3 113.2
CEC Ecec SAT ESP
----%---
'00
,,,
SAR CITE

43.3 0.0
"'ITe
-----------%-----------
43.3
EO SUM

4035
4036
4037
4038
0.63
0.54
0.34
0.58
B.'

..•••,
B.5
B.'
:16.3
37.1
33.6
36.6
L6
2.0
L6
2.'
o. ,
0.1
o. ,
o. ,
0.7
0.7
A.'
0.6
:18.7
40.0
35.8
39.6
.. ,
10.1
11.0

14 .2
'00
'00
'00
100 ,,
49.4
42.4
55.0
58.7
0.0
0.0
0.0
0.0
49.4
42.4
5S.0
58.7
4039
4040
0.30
0.64 B.2
31.2
51.0
LA
2.7
0.1
0.1
0.2
O.A
32.9
54.2 •••
16.6
'00
'00
,, 2 76.0
43.9
0.0
0.0
76.0
43.9
4041 0.32 8.A 43.8 1.7 0.1 0.2 45.9. '.0 '00 62.3 0.0 62.3
4042
4043
0.66
0.40
B.A
8.2
49.3
42.9
2.'
1.8
0.2
0.2
O.A
0.2
52.4
45.1
12.6
7.7
(00
'00 , 2 56.3
12.0
0.0
0.0
56.3
72.0
4044 0.57 '.2 44.3 1.. 0.2 0.' 46.7 B.' '00 2 136.6 0.0 66.6

SATURATEO PASTE EXTRACT BULK OEN WATER CONTENT


LAB ELEC H20 0.33 AIR 0.10 0.33 15 CO
NO CDNO CONT Col MG NA K C03 He03 CL S04 BAR ORY COLE BAR BAR BAR FE

4034
"'MHOS/CM X ---------------------MEQ/L-·------------------ ---G/cc-- CIot/CM ------WTX------ %
0.2
4035 1.29 1.580.070 30.0 0.2
4036 0.2
4037 0.2
4038 0.2

.•
4039 0.2
4040
4041
0.'
0.2
4042 I. 16 1.38 0.060 0.2
A""
4044 1.01 1. 23 0.068 50.9
0.2
0.2

LAB PARTICLE Sl2E DISTRIBUTION (CLAY-FREE BASIS)


NO
VCS C , V, TOTAL C ,
- - -.- - - -.--- -SANO- - - - - - - - - - - -- -----SlLT-----
• TOTAL -----------RATIOS-----------
----------------------%---------------------- s/st FSI/CSI VFs/FS FCITC CEclCLAY
4034
4035
O.G 1.' 7.2 25.9
0.2 2.1 14.9 34.6 24.0 75.8
24.7 59.6 16.5
•••
23.9 40.4
14.3 24.2
1.. LA
1.' ,. ,
1.0
0.7
O.G
0.6
0.a3
0.70
4036 0.0 0.4 4.3 34.8 32.7 72. t 13.3 14.6 27.9 2.G (., a .• 0.6 0.69
4037 O. ( 3.6 17 .8 39.6 24.5 85.6 4.' 10.0 14.'3 5.' 2.2 0.6 0.6 0.135
4038 0.1 '.7 19.8 21.2 t7 .3 62.1 15.6 22.3 37.9 (,6 (.4 0.8 0.6 0.74
4039 0.7 12.5 40.8 25.1 8.7 87.8 3.3 B.' 12.2 7.2 2.7 0.3 0.7 0.64
4040 0.1 1.2 10.6 2 t. 1 24.3 57.4 16.2 26.6 42.8 1.3 1.6 (,2 0.6 0.67
4041 O. , (.A 16.1 39.7 t7.7 75.0 B.4 16.6 25.0 3.0 2.0 0.4 0.6 0.58
4042 0.4 2.2 12.8 24.7 20.9 60.8 13.8 25.4 39.2 1.6 1.8 0.' 0.6 0.60
4043 6.6 lB.l 29.9 19.7 '.6 83.9 3.' 12.7 113.1 '.2 3.7 0.' 0.' 0.58
4044 (.6 10.0 28.7 22.6 12.3 75.2 '.2 15.6 24.8 3.0 1.7 a .• 0.6 0.61

2-32
PEDON 2
SOIL SERIES: BOSQUE VARIANT PEDaN: 59QrX-099-009 COUNTY = CORYELL

PEDaN CLASSIFICATION: PACHIC CALCIUSTOlLS: FINE-LOAMY, CARBONATIC. THERMIC


LOCATION: FT. HOOD. rEXAS: 10 M N OF CQWHOUSe CREEK; 300 M W OF WEST RANGE AOAD.

LANDFORM: STREAM TERRACE ELEVATION (104): SLOPE: 0-11. SLOPE ASPECT:

PARENT MATERIALS: ALLUVIUM FORMATION: HOLOCENE ALLUVIUM (UNIT 0)

TOPOGRAPHY: NEARLY LEVEL DRAINAGE: WELL DRAINED LANOUSE: NATIVE

COLLECTORS: NOROT ANO HALLMARK DATE: 07/03/90

HORIZON OEPTH (eM) SOIL DESCRIPTION (COLORS FOR MOIST SOIL UNLESS STATED)

• I 0-15 VERY DARK GRAYISH BROWN (IOVR 3/2) LOAM, GRAYISH BROWN (fOVR 5/2) DRY: MODERATE
COARSE SUBANGULAR BLOCKY PARTING TO MODERATE MEDIUM SUBANGULAR BLOCKY STRUCTURE'
VERY HARD: MANY FINE AND MEDIUM ROOTS; MODERATELY ALKALINE; STRONGLY EFFERVESCENT'
CLEAR SMOOTH BOUNDARY, '

'2 15-34 VERY DARK GRAYISH BROWN (IOYR 3/2) LOAM, VERY OARK GRAYISH BROWN (IOYR 3/2)
DRY; WEAK COARSE PRISMATIC PARTING TO WEAK MEOIUM AND COARSE SUBANGULAR
BLOCKY STRUCTURE: HARD; COMMON FINE ROOTS; COMMON FAUNAL CASTS; MODERATELY
ALKALINE; STRONGLY EFFERVESCENT; CLEAR WAVY BOUNDARY,

aKI 34-53 VERY DARK GRAYISH BROWN (tOYR 3/2) LOAM, OARK GRAYISH BROWN (IOYR 4/2) ORY~
WEAK COARSE PRISMATIC PARTING TO MODERATE MEOIUM AND COARSE SUBANGULAR BLOCKY
STRUCTURE: HARD; COMMON FINE ROOTS; FEW FAUNAL CASTS: MANY VERY THIN CARBONATE
MYCELIA ON VERTICAL PEO FACES OF PRISMS; SECONDARY CARBONATE IS <11. OF HORIZON
VOLUME: MODERATELY ALKALINE: STRONGLY EFFERVESCENT: GRA~UAL SMOOTH BOUNDARY.

aK2 53-85 VERY DARK GRAYISH BROWN (tOYR 3/2) LOAM, GRAYISH BROWN (10YR 5/2) DRY; WEAK
COARSE PRISMATIC PARTING TO MODERATE MEDIUM SUB ANGULAR BLOCKY STRUCTURE;
HARD: COMMON FINE ROOTS: FEW FAUNAL CASTS: MANY vERY THIN CARBONATE MYCELIA
ON VERTICAL PEO FACES OF PRISMS; SECONDARY CARBONATE IS 110 OF HORIZON VOLUME;
MODERATELY ALKALINE: STRONGLY EFFERVESCENT; GRADUAL WAVY BOUNOARY,

aKa 85-118 DARK GRAYISH BROWN (fOYR 4/2) LOAM, GRAYISH BROWN (IOYR 5/2) DRY; WEAK COARSE
PRISMATIC PARTING TO MOOERATE MEDIUM SUBANGULAR BLOCKY STRUCTURE; HARD:
FEW FINE AND MEDIUM ROOTS: FEW FAUNAL CASTS: COMMON (1510) FINE CARBONATE
FILAMENTS AND THREADS ON SURFACES OF PEDS; COMMON (5%) FINE CARBONATE FILAMENTS
AND THREADS IN INTERIORS OF PEDS: MODERATELY ALKALINE; VIOLENTLY EFFERVESCENT;
GRADUAL WAVY BOUNDARY.

BK4 118-155 BROWN (10YR 4/3) LOAM, BROWN (fOYR 5/3) DRY; WEAK COARSE PRISMATIC PARTING
TO MODERATE MEDIUM SUBANGULAR BLOCKY STRUCTURE: HARD; FEW FINE AND MEDIUM
ROOTS; FEW FAUNAL CASTS; FEW SHELL FRAGMENTS; COMMON (15%) FINE CARBONATE
FILAMENTS AND THREADS ON PEO SURFACES; COMMON (510) FINE CARBONATE FILAMENTS
AND THREADS IN INTERIOR OF PEDS: MODERATELY ALKALINE: VIOLENTLY EFFERVESCENT;
GRADUAL WAVY BOUNDARY.

aKe 155-166 BROWN (tOYR 4/3) lOAM, BROWN (tOYR 5/3) DRY: WEAK COARSE PRISMATIC PARTING
TO MODERATE MEDIUM SUBANGULAR BLOCKY STRUCTURE; HARD; VERY FEW FINE ROOTS:
FEW FAUNAL CASTS; FEW SHELL FRAGMENTS: COMMON (1010) FINE CARBONATE FILAMENTS
AND THREADS ON PED SURFACES: COMMON (5%) FINE CARBONATE FILAMENTS AND THREADS
IN INTERIOR OF PEDS: MODERATELY ALKALINE: VIOLENTLY EFFERVESCENT; CLEAR
SMOOTH BOUNDARY,

aK6 166-188 BROWN (10YR 4/3) LOAM, BROWN (10YR 5/3) DRY; MODERATE MEDIUM PRISMATIC PARTING
TO MODERATE MEDIUM SUBANGULAR BLOCKY STRUCTURE; HARD: VERY FEW FINE ROOTS:
FEW FAUNAL CASTS: FEW SHELL FRAGMENTS: MANY (20%) FINE CARBONATE FILAMENTS
AND THREADS ON PED SURFACES; COMMON (10%) FINE CARBONATE FILAMENTS ANO THREADS
IN PED INTERIORS: MODERATELY ALKALINE: VIOLENTLY EFFERVESCENT; GRAOUAL SNOOTH
BOUNDARY,

BK7 _!a8-2IS BROWN (10YR 4/3) LOAM, BROWN (IOYR 5/3) DRY: MODERATE MEDIUM PRISMATIC PARTING
TO MODERATE MEDIUM SUBANGULAR BLOCKY STRUCTURE; HARD: VERY FEW FINE ROOTS;
FEW FAUNAL CASTS: FEW SHELL FRAGMENTS: MANY (2010) FINE CARBONATE FILAMENTS
AND THREADS ON PED SURFACES: COMMON (10%) FINE CARBONATE FILAMENTS AND THREADS
IN PED INTERIORS: MOOERATELY.ALKALINE: VIOLENTLY EFFERVESCENT.

BKe 215-230 BROWN (10YR 4/3) LOAM. BROWN (10YR 5/3) ORY; WEAK COARSE PRISMATIC PARTING'
TO WEAK MEDIUM SUBANGULAR BLOCKY STRUCTURE; HARD: VERY FEW FINE ROOTS: FEW
FAUNAL CASTS; COMMON (10%) FINE CARBONATE FILAMENTS AND THREADS ON PEO FACES;
COMMON (10%) FINE CARBONATE FILAMENTS AND THREADS IN PED INTERIORS: MODERATELY
ALKALINE; VIOLENTLY EFFERVESCENT

REMARKS: BKS AND BK7 MAY HAVE BEEN A BURIED A. HORIZON BUT HAS CHARACTER OF B
HORIZON TODAY,

2-33
Pedon 2
SOIL CHARACTERIZATION LABORATORY
SOIL AND CROP SCIENCES QEPT .. THE TEXAS AGRICULTURAL eXPERIMENT STATION
SOIL SERIES: BOSQUe VARIANT PEDON NUMBER: S90TX-099-009
SOIL FAMILY: PACHIC CALC{USTOLlS: FINE-LOA~Y. CARBONAT[C, THERMIC
LOCATION: CORVELL COUNTY

PARTICLE SIZE DISTRIBUTION (~Ml


-----------------SAND---------------- -----SILT----- -----CLAY-----
VC eMF VF TOTAL FINE rOTAL FINE rOTAL COARse
LAB {2.a- (f.O- (0.5- (0.25- (0.10- (2.0- {O.02- (0.05- « (< TeXTURE FRAG-
NO DEPTH HOR I ZON 1.0) 0.5) 0.251 0.10) 0.05) 0.05) 0.002) 0.002) 0.0002) 0.002) CLASS MENTS

4165
(CIon
0- 15 Af
----------------------------------%----------------------------------
0.4 0.6 7.0 23.0 18.0 49.0 18.1 31.2 10.5 19.8 L
%
o
4166 15- 34 A2 0.4 1.8 7.4 21.3 15.5 46,4 21.5 34.3 9.0 19.3 L o
4167 34- 53 81(1 0.10.64.523.619.748.5 21.1 .31.4 10.8 20.1 L o
416B 53- 84 8K2 0.0 0.7 5.8 26.4 19.8 52.7 20.6 28.6 11.0 18.7 FSL o
4169 85-118 aK3 0.1 0.4 3.2 22 . .3 2L3 47 . .3 22.7 .31.8 12.6 20.9 L o
4170 118-155 8K4 0.2 0 . .3 4.0 20.1 21.2 45.8 19.6 31.7 13.4 22.S L o
4171 155-166 eK5 0.00.22.017.522.642.3 19.4 .34.6 14.0 23.1 L o
4172 166-188 eK6 0.1 0.2 1.2 15.3 21.8 38.6 21.9 36.5 14.8 24.9 L o
4173 188-215 8K7 0.1 0.4 3.0 16.4 17.4 37.3 23.4 37.0 15.6 25.7 L o
4174 215-230 aK8 0.00.55.421.117.044.0 20.6 33.1 13.8 22.9 L o
LAB ORCiN PH -------NH40AC EXTR 8ASES------ KCL EXTR NAOAC 8ASE CAL- . DOLO- CAC03 GVP
NO C (H2O) CA "G NA K TOTAL AL CEC fCEC 5A T ESP SA' CITE MITE EO SUM
r. 1: f ----------------------MEO/lOOG--------------------- ----y.--- -----------y.-----------
4165 1.17 S. I 47.2 ,. I 0.1 0.4 49.8 14.6 100 46. I 0.0 46. I
4166
4167
I. 89
0.97
S.O
S.,
47.0
47.0 ,. ,
, .0 0.1 0.7
0.1 0.'
49.9
49.8
16.3
13.2
100
100
46.4
46.2
0.0
0.0
46.4
46.2
4168 0.94 S.J 45.5 '.J 0.1 O.J 48.2 10.7 100 49.7 0.0 49.7
4169 0.70 S.J 47.8 '.5 0.1 O.J 50.7 1f.5 100 47.5 0.0 47.5
4170 0.69 S.' 47.9 '.B O. I O.J 51.1 t 1. 7 100 46.0 0.0 46.0
4171 0.72 B.' 47.9 J.J O. I O.J 51.5 12.4 100 42.6' 0.0 42.6
4172 0.62 B.' 50.1 J.' O. I 0.4 54.5 13.1 100 42. f 0.0 42.1
4173 0.73 S.J 50.8 4.' O. I 0.4 55.5 .!4. I 100 44.4 0.0 44.4
4174 0.66 S.J 45.5 4. I O. I O.J 49.9 12.2 100 49.1 0.0 49. f

SATURATED PASTE EXTRACT 8ULK OEN WATER CONTENT


LAB fLEC H20 0.33 AIR 0.10 0.33 15 co
NO COND CONT CA MG NA K COJ HC03 CL S04 BAR OR'" COLE BAR BAR BAR FE

4165
"''''Hos/c''' " ---------------------MEO/L-------------------- ---o/cc-- CM/c,,",
1.38 1.570.044
1.31 1.500.046
------WT%---- .. -
26.6
30.4

O.J
O.J
4166
4167 1.22 1.450.059 35.1 0.3
4168 1.21 1.440.060 35.7 o.J
4f69 1.26 1.500.060 33.2 o.J
4170 1.27 1.58 0.076 31.8 O.J
4171 1.291.460.042
1.30 1.4B 0.044
32.5
32.5
0.'
o.J
4172
4173 1.29 1.50 0.052 33.2 0.'
4174 1.28 1.470.047 33.9 0.'

LAB PARTICLE SIZE DISTRIBUTION (CLAY-FREE BASIS)


NO ------------SAND------------- -----SIlT-----
VCS C N F VF TOrAl C F TOTAL -----------RATIOS-----------
51st FSI/CSI VFS/FS Fc/TC CEC/ClAY
4165
----------------------%----------------------
0.5 0.7 8.728.722.461.1 15.623.338.9 1.6 1.5 0.8, 0.5 0.74
4166 0.5 2.2 9.2 26.4 19.257.5 15.926.642.5 1.4 1.7 0.7 0.5 0.84
O. I 0.8 5.629.524.760.7 12.926.439.3 1.5 2.0 0.8 0.5 0.66
4167
0.0 0.9 7.1 32.5 24.4 64.8 9.925.335.2 1.8 2.6 0.8 0.6 O.!H
4168
0.1 0.5 4.0 28.2 26.9 59.B It.5 2B.7 40.2 1.5 2.5 1.0 0.6 0.55
4169
0.3 0.4 5.225.927.459.1 15.625.340.9 1.4 1.6 1.1 0.6 0.52
04170 0.54
0.0 0.3 2.6 22.8 29.4 55.0 19.8 25.2 45.0 1.2 1.3 1.3 0.6
04'71 0.53
O. I 0.3 1.6 20.4 29.0 51.4 19.4 29.2 48.6 1. f 1.5 1.4 0.6
4172 0.55
4173 O. f 0.5 4.0 22.1 23.4 50.2 IB.3 31.S 49.8 1.0 1.7 1.1 0.6
0.0 0.6 7.0 27.4 22.1 57.1 16.226.742.9 1.3 1.6 O.B 0.6 0.53
4174

2-34
PEDON 3
SOIL SERIES: LEWISVILLE VARIANT PEDON: S90TX-Q99-002 COUNTY: CaRVELL

PEDON CLASSIFICATION: VERTIC CALCIUSTOLLS; FINE. MONTMORILlONITIC. THERMIC


LOCATION: FT. HOOD, TEXAS. CHSi8TRf.

LANOFOR~: STREA~ TERRACE ELEVATION (M): SLOPE: 0- 1% SLOPE ASPECT:


PARENT MATERIALS: ALLUvIUM FORMATION: HOLOCENE ALLUVIUM (UNIT C)

TOPOGRAPHY: NEARLY LEVEL DRAINAGE: WELL DRAINED LANOUSE: NATIVE

COLLECTORS: NOROT AND HALLMARK OA TE: 12/20/89

HORIZON DEPTH (eM) SOIL DESCRIPTION (COLORS FDA MOrST SOIL UNLESS STATED)

At 0-16 VERY OARK BROWN (IOVR 2/2) SILTV CLAY LOAM, VERY DARK GRAYISH BROWN (IOVR
3/2) CAY; MODERATE COARSE SUB ANGULAR BLOCKY PARTING TO MOOERATE FINE SUBANGULAR
BLOCKY STRUCTURE VERY HARO: COMMON FINE ANO MEOIUM ROOTS; NANY WORN CASTS:
FEW SANO SIZED CARBONATE GRAINS: NOOERATELY ALKALINE; STRONGLY EFFERVESCENT:
CLEAR SMOOTH BOUNOARY.

16-32 VERY OARK BRO~N (IOVR 2/2) SILTY CLAY. VERY OARK GRAyISH BROWN (IOVR 3/2)
ORY: ~EAK COARSE SUB ANGULAR BLOCKY PARTING TO MOOERATE FINE SUB ANGULAR BLOCKY
STRUCTURE: HARD: COMMON FINE ANO NEO{UN ROOTS; CONMON WORM CASTS: FEW SHELL
FRAGMENTS: FEW SANO SIZEO CARBONATE GRAINS: MOOERATELY ALKALINE: STRONGLY
EFFERVESCENT; GRADUAL SMOOTH BOUNOARV.

AB 32-59 VERY OARK GRAVISH BROWN (lOVR 3/2) SILTY CLAY, VERY OARK GRAVISH BROWN (lOYR
3/2) ORV; NOOERATE COARSE SUBANGULAR BLOCKV PARTING TO MODERATE MEOIUM SUBANGULAR
BLOCKV STRUCTURE: HARD: FEW FINE ANO MEOIUN ROOTS: COMMON WORM CASTS; FEW
SHELL FRAGMENTS: FEW SANO SIZEO CARBONATE GRAINS; MOOERATELV ALkALINE; STRONGLV
EFFERVESCENT; GRAOUAL ~AVV BOUNOARV.

BW 59-94 DARk BROWN (lOVR 3/3) SILTV CLAY, OARK BROWN (laVR 3/3) ORV; WEAK COARSE
PRISNATIC PARTING TO MODERATE NEOIUN ANGULAR BLOCkY STRUCTURE; VERY HARD:
FEW FINE ANO MEDIUM ROOTS: COMMON WORM CASTS: FEW SHELL FRAGNENTS; FEW SAND
SIZED CARBONATE GRAINS: OCCASIONAL MEDIUM (1 eM) HARD IRREGULAR CARBONATE
NODULES THAT ARE SONEWHAT PITTED: MOOERATELY ALkALINE; STRONGLV EFFERVESCENT;
CLEAR WAVY BOUNDARY.

BKl 94-122 DARk VELLOWISH BROWN (lOVR 4/4) SILTV CLAY LOAN, DARK YELLOWISH BROWN (laVR
4/4) ORV: WEAk NEOIUN PRISNATIC PARTING TO MOOERATE MEDIUM ANGULAR BLOCKV
STRUCTURE: VERy HARD: VERY FEW FINE ROOTS; FEW WORM CASTS; FEW SHELL FRAGMENTS;
FEW SAND SIZED CARBONATE GRAINS: COMMON (3~) THREADS ANO FILAMENTS OF CARBONATE
CONCENTRATED ON PED SURFACES BUT ALSO IN SOME PEO INTERIORS; SOME SMALL
TO MEOIUM WEDGE-SHAPED PEDS WITH DISTINCT PRESSURE FACES; MOOERATELV ALKALINE:
STRONGLV EFFERVESCENT; GRADUAL WAVY BOUNDARY.

BK2 1:22-153 DARk YELLOWISH BROWN (lOVR 4/4) SILTV CLAY LOAM, DARk YELLOWISH BROWN (IOVR
4/4) ORV: MODERATE MEDIUM ANGULAR BLOCKY STRUCTURE: VERY HARD; VERY FEW
FINE ROOTS: FEW WORM CASTS; FEW SHELL FRAGMENTS: FEW SAND SIZED CARBONATE
GRAINS: COMMON (5%) THREAOS AND FILAMENTS OF CARBONATE BOTH ON PED SURFACES
AND IN PEDS: SOME SMALL TO MEDIUM WEDGE-SHAPED PEOS WITH DISTINCT PRESSURE
FACES; MOOERATELY ALKALINE: STRONGLV EFFERVESCENT; GRAOUAL WAVY BOUNOARy.

BK' 153-189 DARK VELLOWISH BROWN (lOVR 4/4) SILTV CLAy LOAM; MOOERATE MEDIUM PRISMATIC
PARTING TO MODERATE MEDIUM ANGULAR BLOCKV STRUCTURE: VERY HARD; VERY FEW
FINE ROOTS; FEW SHELL FRAGMENTS: FEW SAND SIZED CARBONATE GRAINS: COMMON
(7%) THREADS AND FILAMENTS OF CARBONATE BOTH ON PED SURFACES AND IN PEDS:
MOOERATELV ALKALINE: STRONGLY EFFERVESCENT: GRADUAL WAVY BOUNOARV.

BK4 189-252 DARK YELLOWISH BROWN (IOYR 4/4) SILTV CLAY LOAM: MOOERATE MEDIUM PRISMATIC
PARTING TO ~OOERATE MEDIUM ANGULAR BLOCkY STRUCTURE: VERY HARD: NO ROOTS:
FEW SHELL FRAGMENTS: FEW SAND SIZED CARBONATE GRAINS; COMMON (9%) THREADS
AND FILAMENTS OF CARBONATE BOTH ON PED SURFACES AND IN PEDS: MODERATELY
ALKALINE: STRONGLY EFFERVESCENT.

Be 330-350 YELLOWISH BROWN (10YR 5/4) SILTV CLAY; HARD: NO ROOTS: FEW SOFT SEGREGATIONS
OF CARBONATE ABOUT 1-2 CM IN DIANETER; MODERATELV ALKALINE: STRONGLV EFFERVESCENT.

REfIIARKS: THE BC HORIZON WAS SAMPLED AND DESCRIBED BY HANO AUGER.

2-35
Pedon 3
SOIL CHARACTERIZATION LABORATORY
SOIL AND CROP SCIENCES DEPT .• THE TEXAS. AGRICULTURAL EXPERIMENT STATtON

SOIL SERIES: LEWIsvILLE VARIANT PEDaN NUMBER: 590TX-099-002


SOIL FAMILY: VERTIC CALCtUSTOLL5: FINE. MONTMQRIlLONITIC. THERWIC
LOCATION: CQIlYELL COUNn

PARTICLE SIZE DISTRIBUTION !101M)


-----------------SANO----------------
VC C F VF
-----SlLT----- -----CtAY-----
LAB
NO
" TOTAL
(2.0- ( j .0- (0.5- (0.25- (0.10- (:Z.o-
FINE
(0.01-
TOTAL
(0.05-
fINE
C<
TOTAL
C< TEXTURE
COARSE
FRAG-
OEPTH HORI ZON La I 0.5) 0,25) O. to) 0.05) O.OS) 0.002) 0.002) 0.0002) 0.002) CLASS IoIENTS

4018
{C"'~
0- IG A I
----------------------------------X----------------------------------
0.2 O .• 0.5 2.G 7 .• , 1. I 36.8 51.8 17.6 37. f SteL
%
0
4019
4020
4021
,,-
16- J> A2
59 AB
59- 9' BW
0.3
0.4
0.4
O .•
0.3
0.7
O.G
0.7
0.9
2.9
2.7
'.5
7. I
7.2
8.7
11.3
If.3
14.2
l6.9
36.4
34.0
47.7
45.7
43.7
22.6
23.4
23.1
41.0
4l.0
42. I
SIC
SIC
SIC
0
0
0
4022
4023
4024
94-122 BKI
122-153 BK2
153-189 BK3
4025 189-252
4026 330-350 C
,..
0.2
0.2
0.1
0.2
0.1
0.5
0.4
0.4
0.'
0.2
1.0
O.G
0.4
0.'
'.'
'.5
'.5
'.8
2.0
If. 1
12.0
13.3
12.1
B.G
17.2
17 .1
18.7
16.8
11.2
31.9
32. I
lO.5
lO.8
29.5
....
43.5
44.5
.3
.B
.G
21.8
21.5
22. t
22.5
25.l
39.3
37.8
37.0
38.4
41.2
SICL
SICL
SICL
SICL
0
0
0
0

LAB
0.'

ORGN PH -------NH40AC EXTR 8ASES------ KCl EXTR NAOAC BASE


" CAL- OOLO-
SIC

CAC03
0

GYP
NO C {H20} CA 0.1:0 NA K TOTAL AL CEC ECEC SAT ESP SA. CITE ~ITE EO SUM
~ 1: t ----------------------MEO/IOOG--------------------- ----1,--- -----------/.-----------
4018 1.817.9 51.8 2.4 0.0 t.l 55.5 26.8 100 0 0 36.5 0.0 36.5
4019 1.37 8.0 47.7 2.2 0.1 0.9 50.9 27.1 100 0 0 37. I 0.0 37. I
4020 1.04 8.1 49.5 2.3 0.1 0.7 52.6 28.5 100 0 0 38.6 0.0 38.6
4021 0.798.3 47.3 2.4 0.1 0.6 50.4 26.2 100 0 0 39.6 0.0 39.S
4022 0.58 8.3 48.1 3.0 0.1 0.5 52.3 2t.9 100 0 0 39.4 0.0 39.4
4023 0.44 8.3 47.6 3.3 0.1 0.5 51.5 25.3 100 0 0 39.6 0.0 39.6
4024 0-.34 8.4 45.3 3.9 0.1 0.6 50.0 21.1 100 0 31.3 0.0 31.3
4025 0.31 8.3 43.6 4.8 0.1 0.6 49.2 21.5 100 0 35.5 0.0 35.5
4026 0.l4 8.4 44.0 8.8 0.1 0.6 53.6 25.4 100 0 33.6 0.0 3l.6

SATURATEO PASTE EXTRACT BULl( OEN WATER CONTENT


LAB ELEC H20 0.33 AlR 0.10 0.33 15 CO
NO CONO CONT CA IotO NA I( C03 He03 CL S04 BAR DRY COLE BAR BAR BAR FE
M~HOs/cM % ---------------------MEO/L-------------------- u-o/cc-- C~/CM ------WT%-- ---- %
4018 0.5 61 4.4 0.3 0.2 0.3 0.0 4.4 0.2 0.3 1.23 1.54 0.078 32.8 0.'
4019 0.5 61 4.4 0.3 0.2 O. t 0.0 4.5 0.0 0.3 1.25 1.51 0.079 32.5 0.5
4020 0.4 61 3.5 0.3 0.2 O. I 0.0 3.9 0.0 0.1 t .31 1.61 0.084 34. I 0.5
4021 0.3 51 2.1 0.2 0.2 0.0 0.0 2.6 0.0 0.3 1.43 1. 74 0.06B 21. I 0.5
4022 0.2 56 1.1 0.2 0.2 0.0 0.0 1.9 0.2 0.1 1.39 1.13 0.016 29.1 0.5
4023 0.2 59 1.5 0.2 0.2 0.0 0.0 1.8 0.2 0.2 1.41 1.16 0.017 30.5 0.5
4024
4025
1.421.76
1.35 1.18
0.014
0.097
29.3
35.2
0.'
0.5
4026 0.4

LAB PARTICLE SIZE DISTRIBUTION (CLAY-FREE BASIS)


NO ------------SANO------------- -----SILT-----
ves e F VF TOTAL e F TOTAL -----------RATtOS-----------

4018
"
----------------------%----------------------
0.3 O.G O.B 4. I 11.8 17 .6 23.9 58.5 82.4
S/St FsIICSI Vfs/FS FCtTC
0.2 2.4 2.' 0.5
CEC/CLAY
0.72
4019 0.5 0.7 1.0 4.' 12.0 19.2 18.3 62.5 80.8 0.2 3.4 2.4 O.G 0.66
4020 0.7 0.5 1.2 '.7 12.6 19.8 16.3 63.9 80.2 0.2 3.' 2.7 0.5 0.66
4021 0.7 1.2 1.6 G.O 15.0 24.5 16.8 58.7 75.5 0.3 3.5 2.5 0.5 0.62
4022 0.3 O.B I.G 7.2 18.3 28.3 19. I 52.6 71.1 0.' 2.B 2.S 0.6 0.56
4023 0.3 O.G 1.0 7.2 19.3 28.5 19.9 51.6 71.5 0.4 2.6 2.7 0.6 0.67
4024 0.2 0.6 0.6 7. I 21. t 29.7 21.9 48.4 10.3 0.4 2.2 3.0 O.G 0.57
4025 0.3 0.5 0.7 G.2 19.6 21.3 22.7 50.0 12.7 0.4 2.2 3.2 O.G 0.56
4026 0.2 0.3 0.5 3.' 14 .6 19.0 .30.8 50.2 81.0 0.2 I.G 4.3 0.6 0.62

2-36
PEDON 4
SDIL SERIES: LEWISVILLE VARIANT PECCN: S90TX-099-001 coUNn: CORHLL
PEDON CLASSIFICATION: UOIe CHRO~STERTS; FINE. MONTWORILLONtT[C, THERMIC
LOCATION: FT. HOOD, TEXAS. STRATIGRAPHIC UNIT A.

LANOFORM: STREAM TERRACE ELEVATION (N): SLOPE: 1% SLOPE ASPECT:


PARENT MATERIALS: ALLUVIUN FORMATION: PLEISTOCENE ALLuvIUN (UNIT A)
TOPOGRAPHY: GENTLY SLOPING ORAINAGE: WELL DRAINED LANOU5E: NATIVE
COLLECTORS: NeRDT. HALLMARK, WILDING, DREES, .lGRD 606 CLASS CATE: 01/20/90

HORIZ.ON OEPYH (eM) SOIL DESCRIPTION (COLORS FOR ~IST SOIL UNLESS STATED)

o·a VERY DARK BROWN (IOVA 2/2) CLAY LOAM, VERY DARK BROWN {IOVR a/a} DRY: ~OOERATE
FINE AND MEDIUM SUB ANGULAR BLOCkY AND WEAK FINE GRANULAR STRUCTURE: VERY
HARD: FEW FINE PORES: CO~MON FJNE ANO MEOIUM ROOTS: FEW SANO-SIZEO CARBONATE
GRAINS; NONCALCAREOUS: CLEAR S~OTH BOUNDARY.

8~1" VERY DARK BROVN (101'R 2/2) CLAY LOAM, VERY DARK BRowN (IOYR '2/2) ORY: MODERATE
FINE AND MEDIUM SUB ANGULAR BLOCKY STRUCTURE: VERY HARD: COMMON FINE AND
MEOIUN ROOTS; FEW SAND-SIZED CARBONATE GRAINS: NONCALCAREOUS: GRAOUAL SNOOTH
BOUNOARY.
.A. 14-38 VERY OARK BROWN (IOYR 2/2) CLAI', VERY DARK BROWN (IOYR l/2) DRY: ~OOERATE
MEDWM PRISMATIC PARTING TO ."I00ERATE COARSE ANGULAR BLOCKY STRUCTURE: VERY
HARO; CONMON FINE ANO MEDIUM ROOTS; COMMON BROWN (7.5I'R 4/4) VORM CASTS;
FEW TO COMMON (:IX) MEorUM (f CM) HARO HIGHLY PITTEO CARBONATE NOOULES; SLIGHTLY
CALCAREOUS; GRADUAL WAVI' BOUNOARY.
BKSS f 3B-66 DARK BROWN (7.5YR 3/4) CLAY, DARK BROWN (7.5YR 3/4) ORY; MODERATE COARSE
PRISMATIC PARTING TO MODERATE COARSE ANGULAR BLOCKY STRUCTURE; VERY HARD:
COMMON FINE ANO MEDIUM PORES: FEV FINE ROOTS; COMNON VERY OARK BROWN (10YA
2/21 CHANNEL FILLINGS ON VERTICAL PEO FACES; CO"~"'ON BROWN (7.5YR 4/4) WORM
CASTS IN MATRIX: FEW sANQ-SIZEO CARBONATE GRAINS; FEW TO COMMON (1%) MEOIUI4
(I CM) HJGHLI' PITTED CARBONATE NOOULES: FEW PRESSURE FACES ANO SLiCKENSIOEs
IN LOWER PART OF HORIZON; ROOTS FLATTENEO AGAINST PEO FACES: STRONGll' EFFERVESCENT;
GRAOUAL WAVY BOUNDARY.
BKSS2 66-98 STRONG BROWN (7.5YR 4/6) CLAY, BROVN (7.5YR 4/4) DRY: VERY HARO: COMMON
FINE ANO MEDIUM PORES: FEW FINE ROOTS; FEW VERY DARK BROwN (IOYR 1/1) CHANNEL
FILLINGS ON VERTICAL PEO FACES; FEW BRO'ol"N (1.SYR 4/4) WORM CASTS IN !04ATRIX;
FEW SANO-sIZEO CARBONATE GRAINS: FEW (1)(,) MEOIUM (1-2 CI4) HIGHLY PITTEO
CARBONATE NOOULES; COM~ON INTERSECTING SLIGKENStOES AND PARALLELIPIPEOS;
FEW MEOIUM 1.5YR 5/6 POCKETS OF CARBONATE-RICH MATERIAL SIMILAR TO HORIZON

.. ,
BELOw: FEW FINE THIN CARBONATE FILAMENTS ON SOI4E PEa FACES; ROOTS FLATTENEO
AGAINST PEO FACES: STRONGLY EFFERVESCENT; CLEAR WAVY BOUNDARY.

98-120 STRONQ BROWN (7.5YR 4/6) CLAY LOA~. STRONG BROWN {7.5YR 5/6} DRY: WEAK COARSE
SUBANGULAR BLOCKY PARTINO TO WEAK FINE AND MEDIUM SUB ANGULAR BLOCKY STRUCTURE:
HARD; FEW FINE PORES; VERY FEW FINE ROOTS: COMMON BROWN (7.5~R 4/4) WOR~
CASTS IN MATRIX; ABOUT ~O~ OF HORIZON COMPOSEO OF 7.5YR 6/6 MATERIAL OF
SIMILAR TEXTURE BUT APPEARS TO BE HIGHER IN CARBONATE CONTENT: COMMON (15~)
MEDIUM (2-3 CM) HARO CARBONATE NODULES: CO~NON (15-20~) CARBONATE FILAMENTS
ANO THREAOS THROUGHOUT; HARD NOOULES ARE NOT AS PITTED AS ABOVE: STRONGLY
EFFERVESCENT: CLEAR WAVY BOUNOARY.

.'2 120-'50 REODISH YELLOw (7.5YR 6/6) CLAY LOAM: NANY COARSE FAINT REDDISH YELLOW (1.5YR
7/6) ANO I4ANY COARSE FAINT REDDISH YELLOW (7.5YR 6/B) MOTTLES; WEAK COARSE
SUBANGULAR BLOCKY PARTING TO 'oI"EAK FINE ANO MEDIUM SUBANGULAR BLOCKY STRUCTURE:
HARO: VERY FEW FINE ROOTS: COKNON BROWN (7.5YR 4/4) WORM CASTS IN MATRIX:
COMMON (5-10%) MEDIUM (1-2 CN) HARD CARBONATE NOOULES; COMMON (10%) CARBONATE
FILAMENTS ANO THREADS: STRONGLY EFFERVESCENT; GRAOUAL WAVY BOUNOARY.

.'3 1~-190 STRONG BROWN (7,5YR 5/6) SANOY CLAY LOAM: MANY COARSE FAINT REDDISH YELLOW
(7.5YR 6/6) ANO CO~N COARSE DISTINCT YELLOwISH REO (SYR 4/6) MOTTLES:
WEAK COARSE SUBANCULAR BLOCKY PARTING TO WEAK MEDIUM SU8ANGULAR BLOCKY STRUCTURE;
VERY HARD: VERY FEW FINE ROOTS: COMMON BROWN (7.SYR 4/4) WORN CASTS IN MATAIX;
FEw TO COfolMON (1%) MEOIUM (1-2 CN) KlRO CARBONAT[ NOOULES: COIolM(]N (3%) CARBONATE
FILAMENTS ANO THREADS: COMHON '-2 MI4 ~OOT CHANNELS: 30-40% OF HORIZON ~EWORKED
BY BIOTIC ACTIVITY: STRONGLY EFFERVESCENT; GRAOUAL wAVY BOUNOARY.

••• 1&0-213 YELLOWISH B~OWN (IOYA 5/6) SANOY CLAY LOAM:' MANY COARSE FAINT YELLOWISH
BAOWN (IOYR 5/4) ANO CONNON COARSE OISTINCT BROWN {7.9YA 4/4} MOTTLES; WEAK
COARSE SUBANQULAR BLOCKY PARTINO TO WEAK MEOIUM SUB ANGULAR BLOCKY STRUCTURE;
VEIIY HARD: VERY FEW FINE ROOTS': CONNON BROWN {7.SYR 4/4} WORM CASTS IN MATRIX:
'E~ THIN PATCHY CLAY FtL~s: VERY FEw ~EoIU~ HARD CARBONATE NOOULES; FEW
(IX) CARBONATE FILAMENTS ANO THREAOS; COfolMON 1-2 NM ROOT CHANNELS: 30-40~ .
OF HQIIIZON REWORKEO BY BIOTIC ACTIVITY; STRONGLY EFFEIIVESCENT: GRAOUAL WAVY BOUNOARY.

••• 213-250 YELLOWtSH BROWN (IOYA 5/6) SANOY CLAY LOA~; MANY COARSE FAINT YELLOWISH
BAOWN (IOYR 5/4) AND BROWN (7.5YR 4/4) MOTTLES; WEAK COARSE SUSANGULAR BLOCKY
PAATING TO WEAK MEDIUM SUSANGULAR 8LOCKY STRUCTURE: VERY HARO; VERY FEW
FtNE AND MEOIUM ROOTS; MANY BROWN (7.5YR 4/4) WORM CASTS IN MATRIX; FEW
THIN PATCHY BROWN (7.5YR 4/4) CLAY FILMS: VERY FEW MEDIUM HARD CARBONATE
NODULES; FEw (I~) CARBONATE FILAMENTS AND THREAOS: COMMON 1-2 ~ ROOT CHANNELS:
30 TO 40~ OF HORIZON REWORKEO BY BIOTIC ACTIVITY: STRONOLY EFFERVESCENT;
GRAOUAL WAVY BOUNOARY.

.e 250-270 REDDISH YELLOW (7.5YR 6/6) LOAM; CONYON M~IUM FAINT BROWNISH YELLOW (IOYR
6/8) ~TTLES: STRUCTURELESS MASSIVE: HARO; VERY FEW FINE AND MEDIUM ROOTS:
COMMON BROWN (7.5YR 4/4) WORM CASTS IN ~ATRIX: FEw THIN pATCHY BROWN (7.5YR
4/4) CLAY FILMS; CO~N BIOCHANNEL ANO BIOPORES I MM OIAMETER: LARGE BIOCASTS
wiTH WHITE FUNGAL GRowTH; FEW LINESTONE CLASTS 0.5 TO I CM OIAMETER: STRONGLY
EFFERVESCENT.

2-37
Pedon 4
SOll CHARACTERIZATION lABORATORV
SOIL ANO CROP SCIENCES DEPT., THE TEXAS AGRICULTURAL EXPERIMENT STATION
SOIL SERIES: LEWISVILLE VARIANT PEDON NUMBER: S90rX-099-001
SOIL FAMILY: UDIC CHROMUSTERTS: FINE. MONTMQRtLlONtT[C. THERMIC
LOCATION: CORVELL COUNTY

PARTICLE SIZE OISTRIBUTION o~ ... )


VC C N , V, TOTAL
-----------------SANO---------------- -----SILT-----
FINE TOTAL
-----CLAV-----
FINE
,<
TOTAL COARSE
LAB .0- f 1 .Q- {O.S- fO.2S- (0.10-
{:2 (2.0- {O.Ol- (0.05- ,< TeXTURE FRAG-
NO CEPTH HORIZON 1.0) 0.5) 0.25) 0.10) 0.05) 0.05) 0.002.) 0.002) 0.00021 0.002) CLASS MENTS
!C"'! ----------------------------------~---------------------------------- Yo
.,
e
....
J995 D- AI O. , 0 3 0.4 7.0 12.6 20.1 22.2 39.6 15.9 39.1 CL 0
3996
3997
3998
3999
4000
a-
14-
3B-
66-
14
3B BAK
66 BKSSI
'B BKSS2
98-120 BKl
0.2
0.1
0.4
0.6
1. 1
0.3
0.'
0.4
0.7
1.4
0.3
0.'
O.B
1.3
3.'
6.'
6
6
a .2
12. 3
13.3
13.8
IJ.6
13.8
14 .8
20.6
'21.6
22. f
24.6
JJ.l
22.0
18.6
20.1
20.4
22.4
34.5
31.1
33.0
32.2
33.1
23.5
28:7
27.2
26.5
20.0
44..•
44.
46 .7

43. 2
33. 2
C
C
C
C
CL
0
0
0
0
B
4001
'002
110-150 BK2
150-190 BK3
0.6
2.2
2 1
3.0
7.0
7.B
17 .7
18.1
18.6
19.3
46.0
51.0
20.3
16.9
30.3
29.7
14.7
ILl
23.7
19.3
L
L
•6
n.9
,,
4003 190-213 BK4 1.. 2.' '.6 20.:3 52.6 IS.8 29.5 10. I 17.9 FSL 2
4004 213-250 BK' 1.B 2.' 7.1 16. , 18.5 46.0 20. :2 35.0 10. J 19.0 L
400' 250-270 BC 0.' 1.7 6.' 15.2 19.3 44.0 20.7 36.5 10. f 19.5 L

LAB ORGN PH -------NH4QAC EXTR BASES------ KCL EXTR NACAC BASE CAL- OOlO- CAeOl GYP
NO C (H2O) CA "G NA K TOTAL AL CEC ECEC SAT ESP SA' CITE MITE EO SUN
% 1:1 ----------------------NEOllOOG--------------------- ----y.--- -----------1.-----------
3995 2.25 7.6 56.7 2.0 0.0 2.2 60.9 33.9 100 0 0 4.2 0.0 4.2
3996 1. 13 7.6 58.5 1.7 0.0 1.B 62.0 34.8 100 0 0 3.0 0.0 3.0
:3997 '·.30 7.6 62.7 1.6 0.0 1.2 65.4 32.8 100 0 0 4.3 0.0 4.3
3998 0.97 7.7 62.9 1.2 0.2 0.6 GS.a 29.8 100 1 11.4 0.0 11.4
3999
4000
0.12 7.B
0.35 7.8
65.8
54.8
1.3
0.'
0.1 o.s
0.2 0.'
67.8
56.4
21. ,
18.8
100
100
0
, 20.0
38.6
0.0
0.0
20.0
38.6
4001 0.26 7.' 54.9 (.0 O. , 0.3 56.3 11.5 100 1 48. I 0.0 48.1
4002
4003
0.42 7.'
0.23 B.O
52.6
52.0
0.9
1.0
0.0 0.3
0.1 O.J
53.8
53.4
•••
B.a
100
100
0
1
51.1
52.8
0.0
0.0
51.1
52.8
'004 0.00 B.O 51.7 0.' O.d 0.3 52.9 '.2 100 0 51.3 0.0 51.3
'oos 0.06 B.O 52.2 1.1 0.0 0.3 53.§ 9.0 '00 0 49.8 0.0 49.8

SATURATED PASTE EXTRACT BULK OEN WATER CONTENT


LAB ELEC. H20 0.33 AIR 0.10 0.33 15 CO
CONO CO NT CA NA C03 HC03 CL S04 BAR DRY COLE BAR BAR BAR
Nt
MMHOS/CN r.
MG 1<
----~-------------~--MEO/L-------------------- ~--G/CC-- CM/CN ------WT%------ "
%
3995 0.5 73 4.8 0 . .2 0.1 0,3 0.0 4,1 0.0 0.0 1.12 1.56 0.117 47.4 O.B
3996 0.4 79 4.2 0.2 O. I 0.2 0.0 2.9 0.0 0.0 1.25 1.68 0.104 40.0 O.B
3997 0.4 76 3.9 0.2 0.1 0, I 0-,0 3.0 0.0 0.0 1.30 1.71 0.096 34.9 0.'
3998 1.33 1.75 0.096 33.7 0.'
3999 1.40 1.78 0.083 31.3 0.'
4000 1.43 1.56 0.029 25.3 0.7
'00' 1.47 f .61 0.031 29.2 0.4
'002 1.47 1.62 0.033 24.8 0.3
'003 O.J
4004 1,41 1.59 0.041 27.5 O.J
400' 1.55 1.70 0.031 22,8 O.J

LAB PARTICLE SIZE DISTRIBUTION (CLAY-FREE BAStS)


NO --------~---SANO--------~---- -----SlLT-----
YCS C J4 F YF TOTAL C F TOTAL -----------RATIOS-----------
S/SI FSI/cSI YfS/FS Fc/TC CEC/CLAY
---------------------~%---------------------- 0.85
3995 0.7 0.5 0.7 11.6 20.9 34.3 28.9 36.8 65.7 0.5 1.3 1.8 0.4
3996 0.4 0.5 0.5 11.824.1 37.4 22.739.962.6 0.6 1.8 2.0 0.5 0.78
3997 0.2 0.8 0.8 12.9 2S.9 40.5 24.6 34.9 59.5 0.7 1.4 2.0 0.6 0.70
3998 0.7 0.7 1.512.524.740.1 23.436.559.9 0.7 1.6 2.0 0.6 0.66
3999 1.1 1.2 2.314.424.343.3 20.835.956.7 0.8 1.7 1.7 0.6 0.63
4000 1.6 2.1 5.2 18.422.249.6 17.0 33.5 50.5 1.0 2.0 1.2 0.6 0,57
'001 0.8 2.8 9.223.224.460.3 13.126.639.7 1.5 2.0 1.1 0.6 0.48
'002 2.7 3.7 9.723.223,963.2 IS.9 20.9 36,8 1. 7 1.3 1.0 0.6 0.51
2.3 3.5 11.724.721.864.1 IS.4 20.5 35,9 1.8 1.3 0.9 0.6 0.49
'003
'004 2.2 3.1 8.8 19.922.856.8 18.324.943.2 1.3 1.4 I. I 0.5 0.48
1.1 2.1 8.618.924.0 54.7 19.6 2!L7 45.3 1.2 1.3 1.3 0.5 0.46
'oos

2-38
CaIcoIaICd Carbonate-Free
Partlcle Sim Distribution of Pedon 4

Clay-free basis
Horizon Depth Sand Silt Clay Sand Silt

em %

Al 0-8 19.6 44.6 35.8 305 69.5


A2 8-14 18.4 39.1 42.5 32.0 68.0
BAk 14-38 19.8 36.3 43.9 35.3 64.7
Bkssl 38-66 20.2 36.2 43.6 35.8 64.2
Bkss2 66-98 21.3 32.2 46.5 39.8 60.2
Bkl 98-120 24.9 30.4 44.7 45.0 55.0
Bk2 120-150 30.5 27.7 41.8 52.4 47.6
N
.:., 150-190 41.1 27.3 31.6 60.1 39.9
Bk3
'"
Bk4 190-213 43.1 29.3 27.6 59.5 40.5
BkS 213-250 30.7 33.4 35.9 47.9 52.1
BC 250-270 32.5 31.7 35.8 50.6 49.4
Elemental Analysis of
CaIbonate-free Sands and SillS of Pedon 4

Horizon Depth Sand (2-.05 mm) Silt (.05-.002 mm)

Ti K Fe Zr Ti K Fe Zr

em %

Al 0-8 0.083 0.71 0.37 0.029 0.48 1.86 1.44 0.074


A2 8-14 0.077 0.62 029 0.029 0.49 1.77 1.30 0.078
BAle 14-38 0.083 0.68 0.39 0.035 0.49 1.71 1.23 0.062
Bless1 38-66 0.059 0.62 0.30 0.021 0.49 1.62 1.25 0.063
Bless2 66-98 0.058 0.60 0.43 0.017 0.54 1.48 1.26 0.053

'"0
1- Btl 98-120 0.104 : 0.69 0.77 0.023 0.49 1.28 1.11 0.075
Bk2 120-150 0.109 0.75 1.09 0.012 0.54 1.31 1.23 0.097
Bk3 150-190 0.101 0.73 1.04 0.013 0.48 1.22 1.22 0.083
Bk4 190-213 0.094 0.67 0.97 0.009 0.48 1.28 1.26 0.089
BkS 213-250 0.149 0.86 1.11 0.028 0.45 1.01 1.06 0.109
Be 250-270 0.120 0.76 0.92 0.021 0.47 1.27 1.08 0.074
ReoollStmctIon ADaJ,ysfs of Pedon 4

Gain-Lnss of Constituents Gain-Lnss of Carbonat~


Coarse Fine
Horizon Sand Silt Clay Clay Total Sand Silt Clay Total

g/cm'

Al -9.2 -5.0 .().4 -2.0 -16.5 -6.8 -5.2 -1.1 -13.1


A2 -7.0 -3.9 .().1 0.3 -10.7 -5.3 -4.0 '().9 -10.2
BAk -26.5 -15.5 -1.3 3.7 -39.6 -20.5 -15.7 -2.9 -39.1
Bkss1 -26.8 -14.4 '().5 4.7 -36.9 -20.9 -14.9 -1.7 -37.5
Bkss2 -26.0 -12.9 0.2 6.0 -32.7 -20.3 -11.7 -1.5 -33.5
Bkl -5.6 -2.1 0.9 2.7 -4.1 -4.3 -1.7 '().1 -6.1
Bk2 '().4 '().8 '().2 1.7 0.3 0.2 0.7 -1.7 '().8
't
- Bk3 7.2 '().6 '().2 1.6 7.9 4.0 0.7 0.0 4.7
Bk4 4.7 '().4 '().6 0.3 4.0 2.8 0.3 '().2 2.9
Bk5 0.7 0.4 -1.0 0.1 0.1 1.9 1.1 -2.4 0.6
BC '().1 0.0 .().1 '().2 '().5 0.1 0.6 -1.4 .().7

Sum (:t) -89.12 -55.08 -3.41 18.76 -128.85 -69.05 -49.70 -13.97 -132.72
LITERATURE CITED AND REFERENCES

Adkins, W. S. 1978. The Mesozoic system in Texas. In The geology of Texas, Vo!' I:
Stratigraphy. Bur. of Econ. Geo!. Bull. No. 3232. Univ. of Texas, Austin.

Amundson, R. G., O. A. Chadwick, 1. M. Sowers, H. E. Doner. 1988. Relationship between


climate and vegetation and the stable carbon isotope chemistry of soils in the eastern
Mojave, Nevada. Quat. Res. 29:245-254.

Anderson, L. M. 1989. Stratigraphy of the Fredericksburg Group, East Texas Basin. Baylor
Geo!. Studies Bull. No. 47. Baylor Univ., Waco.

Baker, V. R. 1975. Flood hazards along the Balcones Escarpment in Central Texas:
Alternative approaches to their recognition, mapping, and management. Geo!. Circ.
75-5. Bur. Econ. Geo!., Univ. of Texas, Austin. p. 22.

Barnes, V. E. 1979. Geologic atlas of Texas: Waco sheet. Bur. of Econ. Geo!., Univ. of
Texas, Austin.

Birkeland, P. W. 1984. Soils and Geomorphology. Oxford University Press. New York.
371 p.

Blum, M. D. 1989. Geoarcheological investigations. In Boyd, D. K., M. D. Freeman, M. D.


Blum, E. R. Prewitt, and J. M. Quigg. Phase I cultural resources investigations at
Justiceburg Reservoir on the Double Mountain Fork of the Brazos River, Garza and
Kent Counties, Texas. Rep. of Invest. 66. pp. 81-106. Prewitt and Associates
Consulting Archaeologists, Inc., Austin.

Blum, M. D. and S. Valastro, Jr. 1989. Response of the Pederna1es River of Central Texas
to Late Holocene climatic change. Annals-Assoc. Am. Geog. 79:435-456.

Brackenridge, R. G. 1988. River flood regime and floodplain stratigraphy. In Baker, V. R.


and P. C. Patton (eds.). Flood geomorphology, pp. 139-155.

Brewer, R. 1976. Fabric and mineral analysis of soils. Krieger Publishing Company. New
York.

Bryant, V. M. and R. G. Holloway. 1985. The late Quaternary paleoenvironmental record of


Texas. In Bryant, V. M. and R. G. Holloway (eds.) Pollen records of late Quaternary
North American Sediments. Am. Assoc. Strat. Paly. pp. 39-70.

Byrd, C. L. 1971. Origin and history of the Uvalde gravel of Central Texas. Baylor Geo!.
Studies Bull. No. 20. Baylor Univ., Waco.

2-42
Bull, W. B. 1991. Geomorphic responses to climatic change. Oxford Univ. Press. 326 p.

Caran, S. C. and R. W. Baumgardner, Jr. 1990. Quaternary stratigraphy and


paleoenvironments of the Texas Rolling Plains. Geol. Soc. Am. Bull. 102:768-785.

Cerling, T. E. 1984. The stable isotopic composition of modern soil carbonate and its
relationship to climate. Earth Planet Sci. Let. 71:229-240.

Craig, H. 1957. Isotopic standards for carbon and oxygen and correction factors for mass
spectrometric analysis of carbon dioxide. Geochim. Cosmochim. Acta 12:133-149.

Davis, K. W. 1974. Stratigraphy and depositional environments of the Glen Rose Formation,
North-Central Texas. Baylor Geol. Studies Bull. No. 26. Baylor Univ., Waco.

Deines, P. 1980. The isotopic composition of reduced organic carbon. In Fritz, P. and J.
Fontes, (eds.) Handbook of environmental isotope geochemistry. Vol. 1. The
terrestrial environment: Amsterdam, The Netherlands. Elsevier. pp. 329406.

Dougherty, J. P. 1980. Streamflow and reservoir - content records of Texas. Texas Dept.
Water Res. Rep. 244 Vol. 2. p. 111.

Emrich, K. D., H. Ehalt, and J. C. Vogel. 1970. Carbon isotope fractionation during the
precipitation of calcium carbonate. Earth Planet Sci. Let. 8:363-371.

Ferring, C. R. 1990. Archeological geology of the Southern High Plains. In Lasca, N. P.


and Donahue, J. (eds.) Archeological Geology of North America. Geol. Soc. Am.
Centennial Spec. Vol. 4. pp. 253-266.

Gile, L. H., F. F. Peterson, and R. B. G. Grossman. 1966. Morphology and genetic


sequences of carbonate accumulation in desert soils. Soil Sci. 101:347-360.

Hall, S. A. 1988. Environment and archaeology of the Central Osage Plains. Plains
Anthropologist 33:203-218.

Hall, S. A. 1990. Channel trenching and climate change in the Southern U.S. Great Plains.
Geology. 18:342-345.

Hallmark, C. T., L. T. West, L. P. Wilding, and L. R. Drees. 1986. Characterization data for
selected Texas Soils. Texas Agric. Exp. Sta. Misc. Pub. No. 1583. 239 p.

Hayward, O. T., P. M. Allen, and D. L. Amsbury. 1990. Lampasas Cut Plain-Cyclic


Evolution of a regional landscape, Central Texas Geol. Soc. Am. 122 p.

2-43
Hill, R. T. 1901. Geography and geology of the Black and Grand Prairies, Texas. U. S.
Geo!. Surv. 21st Ann. Rep. 666 p.

Hoefs, J. 1987. Stable isotope chemistry. 3rd edition Springer-Verlag. New York.

Holliday, V. T. 1985a. Morphology of Late Holocene soils at the Lubbock Lake


Archaeological Site, Texas. Soil Sci. Soc. Am. J. 49:938-946.

Holliday, V. T. 1985b. Early and middle Holocene soils at the Lubbock Lake
Archaeological Site, Texas. Catena 12:61-78.

Holliday, V. T. 1989. Middle Holocene drought on the Southern High Plains. Quaternary
Research 31:74-82.

Jenny, H. 1941. Factors of soil formation. McGraw-Hill Book Co., Inc. New York.

Machette, M. N. 1985. Calcic soils of the Southwestern United States. In Weide, D. L.


(ed.) Soils and Quaternary Geology of the Southwestern United States. Geo!. Soc.
Am. Special Paper 203. pp. 1-21.

Mandel, R. n.d. Geomorphology of the South Bend Area. In Saunders, J. and C. S.


Mueller-Wille. (eds.) An archaeological survey of the proposed South Bend
Reservoir Area, Young, Stevens, and Throckmorton Counties, Texas. Archaeological
Research Lab Archaeological Surveys No.6, Texas A&M University, College Station.

Magaritz, M. and A. J. Amie!. 1980. Calcium carbonate in a calcareous soil from the Jordan
Valley, Israel: Its origin as revealed by stable carbon isotope method. Soil Sci. Soc.
Am. 44:1059-1062.

McCaleb, N. L. 1985. Soil Survey of Coryell County, Texas. U.S.D.A.-Soil Conservation


Service, Texas Agric. Exp. Stn., and U. S. Dept. of the Army-Fort Hood, Texas. U. S.
Gov. Print. Office, Washington, D.C.

McFadden, L. D. 1988. Climatic influences on rates and processes of soil development in


Quaternary deposits of Southern California. Geo!. Soc. Am. Special Paper 216. pp.
153-177.

McFadden, L. D. and J. C. Tinsley. 1985. Climatic influences on rates and processes of soil
. development in Quaternary deposits of Southern California. In Weide, D. L. (ed.)
Quaternary geology of the Southwestern United States. Geo!. Soc. Am. Special Paper
203. pp. 23-41.

2-44
Miles, R. I. 1981. Development of soils on terraces associated with the Brazos River in
Young and Throckmorton Counties, Texas. Ph.D. diss. Texas A&M Univ., College
Station.

Moore, C. H., Ir. 1969. Stratigraphic framework, Lower Cretaceous, West-Central, Texas.
In C. H. Moore (ed.) Depositional environments and depositional history - Lower
Cretaceous shallow shelf carbonate sequence, West-Central Texas. Am. Assoc. Petro
Geo!. Guidebook. Dallas Geo!. Soc., Dallas, Texas.

Nelson, H. F. 1973. The Edwards (Lower Cretaceous) reef complex and associated
sediments in Central Texas. In H. F. Nelson (ed.) The Edwards reef complex and
associated sedimentation in Texas. Geo!. Soc. Am. Guidebook 15.

Nordt, L. C. 1992. Archeological geology of the Fort Hood Military Reservation. Delivery
Order No.8, Contract DACA 63-87-R-1055. U. S. Army, Fort Hood, Texas.

Nordt, L. C., T. W. Boutton, C. T. Hallmark, and M. R. Waters. 1994. Late Quaternary


vegetation and climate changes in central Texas based on the isotopic composition of
organic carbon. Quat. Res. 41:109-120.

North American Commission on Stratigraphic Nomenclature. 1983. North American


stratigraphic code. Am. Assoc. Petro Geo!. Bull. Vo!' 67, No.5. pp. 841-875.

Owen, M. T. 1979. The Paluxy Sand in North-Central Texas. Baylor Geo!. Studies Bull.
No. 36. Baylor Univ., Waco.

Pendall, E. and R. Amundson. 1990. The stable isotope chemistry of pedogenic carbonate in
an alluvial soil from the Punjab, Pakistan. Soil Sci. 149:199-211.

Quade, I., T. E. Cerling, and 1. R. Bowman. 1989. Systematic variations in the carbon and
oxygen isotopic composition of pedogenic carbonate along elevation transects in the
Southern Great Basin, United States. Geo!. Soc. Am. Bull. 101:464-475.

Rabenhorst, M. C. 1983. Genesis of soils and carbonate enriched horizons in a c1imo-


sequence developed over Cretaceous limestone in Central and West Texas. Ph.D.
Diss., Texas A&M Univ., College Station.

Rabenhorst, M. C., L. P. Wilding, and L. T. West.. 1984. Identification of pedogenic and


lithogenic carbonates using stable carbon isotope and micro fabric analyses. Soil Sci.
Soc. Am. 1. 48:125-132.

Rose, P. R. 1979. Edwards group, surface and subsurface, Central Texas. Bur. of Econ.
Geo!. Rep. Invest. No. 74. Univ. of Texas, Austin.

2-45
Salomons, W. and W. G. Mock. 1976. Isotope geochemistry of carbonate dissolution and
reprecipitation in soils. Soil Sci. 122: 15-24.

Sobecki, T. M. and L. P. Wilding. 1983. Formation of calcic and argillic horizons in


selected soils of the Texas Coast Prairie. Soil Sci. Soc. Am. J. 47:701-715.

Soil Survey Staff. 1981. Soil survey manual. U.S.D.A.-Soil Conservation Service Agric.
Handbook 18. U. S. Gov. Print. Office, Washington, D. C.

Soil Survey Staff. 1990. Keys to soil taxonomy. 4th Edition. Soil Manage. Supp. Servo
Techn. Mongr. No.6. Cornell Univ., Ithaca, NY.

Stricklin, F. L., Jr., C. I. Smith, and F. E. Lozo. 1971. Stratigraphy of Lower Cretaceous
Trinity deposits of Central Texas. Bur. Econ. Geo!. Rep. Invest. No. 71. University
of Texas, Austin.

Tonkin, P. J. and L. R. Brasher. 1990. Soil-stratigraphic study of soil and landform


evolution across the Southern Alps, New Zealand. In Knuepfer, P. L. K. and L. D.
McFadden (eds.) Soils and landscape evolution. Geomorphology. Vo!' 3, Nos. 3/4.
pp. 547-555.

U. S. Department of Agricultural (1981). Land resource regions and major land resource
areas of the United States. Soil Conservation Service Agriculture Handbook 296. U.
S. Government Printing Office, Washington, D. C.

Walker, 1. R. 1978. Geomorphic evolution of the Southern High Plains. Baylor Geol.
Studies Bull. No. 35. Baylor Univ., Waco.

West, L. T. 1986. Genesis of soils and carbonate enriched horizons associated with soft
limestones in Central Texas. Ph.D. Diss., Texas A&M Univ., College Station.

West, L. T., L. R. Drees, L. P. Wilding, M. C. Rabenhorst. 1988. Differentiation of


pedogenic and lithogenic carbonate forms in Texas. Geoderma. 43:271-287.
Amino Acid Racemization
Analysis of the Chronology
and Integrity of Archeological
Sites in Central Texas

GlennA. Goodfriend
G. Lain Ellis
James T. Abbott

Abstract

Amino acid racemization analysis (D-alloiso-Ieucine /L-isoleucine, or


.A/I, values) of shells of the land snails Rabdoills and OligJ)ra were used
to carry out detailed chronostratigraphic studies of archeological sites
at Fort Hood and San Angelo, in Central Texas. Because large numbers
of racemization analyses can be performed at a relatively low cost, the
method lends itself to the assessment of age uniformity within and
among archeological strata. Racemization analyses of a series of
individual radiocarbon-dated shells show that A/I values have good
age-predictive ability, especially for middle to late Holocene samples.

A series of examples is discussed in which a range of patterns of


distribution of A/I values is found. Some proveniences show uniform
A/I values, indicating uniformity of age and high site integrity.
However, other proveniences show nonuniform A/I values, which
indicates that reworking, slow burial, or disturbance has compromised
site integrity. Heating from fire produces anomalously high A/I
values, and fire sometimes can be identified as a probable source of
nonuniform A/I values.

[Adapted from poster presented at


the Annual Meetings of the Geological
Society of America, 7 November 1996,
New Orleans1

3-1
Introduction

Site Integrity Temporal Diagnostics: A conventional index for assessing


integrity.
Problem: Before we invest in archeological data recovery, we need to
know how confident we can be about the integrity of the data we Diagnostics of different ages frequently are recovered from the same
recover. provenience. Typically, this is interpreted as assemblage mixture.

Inferences from artifacts to human behavior are only as reliable as


judgments that artifacts in a given provenience are or are not
behaviorally related. Such judgments themselves are reliable only insofar
as:

(a) the assemblage was buried rapidly enough to prevent inclusion of


behaviorally unrelated artifacts in the same provenience; and
...,
.:., (b) disturbance, redeposition, or other processes did not introduce
beha viorally unrelated artifacts. Perdiz Castroville Pedernales
A.D. 1200-1500 800-400 B.C. 2000-1200 B.C.
Integrity assessment: evaluating the extent to which artifacts from a Diagnostics in Hypothetical Column A
given provenience can be known to have been buried

i" ~ ~ 6666
contemporaneously.

i· 7!;-~lf6tilS7S-
Problem: Is the assemblage mixed, or were Perdiz users avid projectile
point collectors? Co-occurrence of non-contemporaneous artifacts does
not mean that artifacts were not deposited and buried
contemporaneously.
Introduction (cont'd)

Deposition Rate: Another conventional index for assessing Amino Acid Racemization: A new index for assessing
integrity. integrity.

Deposition rate typically is used as ... more dates, another pattern Amino acid racemization dating is an economical tool that provides a
an index of integrity. This assumes might emerge. Rilpid deposition useful complement to other indices for assessing integrity.
that deposition rate is a direct and high integrity, although
measure of integrity. However, if usually related, are not the same
we run ... thing. This study applies racemization
dating of land snail shells
to integrity assessment
Hypothetical Column S, Hypothetical Column S, of archeological sites in
One Date per level Three Dates per level
, Central Texas.
------,
-3 ~ I 31 ~ Mixed assemblaga /C:f1trw T6);&:> \

~ - - -1 !~~~i~rn:i~- - - - - - - - -~ - - - - - - - - - - - l. *Fort \1
'-' .e
c 4 ~
_______ r.!:!!.o~r~..!!~ __ 41
~

___ ;;~ ~~~as:e:b~ge__ _


• San
tAngelo Hood /
(.,
re~ 5 ~
_________
"~. .' Inl.."y
Falilst!!~lilitIO~
5 r 1"+1
____ _ ':!'1________ _
Unmlxod assemblage
"",
..._---, ,J
w 6 ~ ~ ~h.'~I", Y 61 Mixed assemblage 1"+11"+1
§ ~w!!"$'P.2S~OIJ! ________ _
o lower mtGgnty ------------
Mixed a.ssembla go
~

~
~7 ~ 7 ~

I
, I - ,-- I
~

o 1000 2000 3000 0 1000 2000 3000


Here, we:
Radiocarbon Years B.P. Radiocarbon Years B.P.

• review the principles of amino acid racemization dating;


It takes at least two measurements of absolute or relative age per
provenience to provide an index of the contemporaneity of burial of • show that amino acid racemization can be calibrated to provide
the artifacts in the assemblage. The more measurements, the greater absolute and, hence, relative dates; and
one's confidence in the index. Obviously, integrity assessment can get
expensive quickly, especially if one has several columns to evaluate. • provide a series of cases studies with examples of high-integrity
deposits, compromised integrity, influences from fire, and
multicausal interpretation of integrity.
Amino Acid Racemization Dating

Principles of Amino Acid Racemization Principles of Integrity Assessment


Dating
• Variation among A/l values of shells from a
• Mollusk shells contain small amounts of protein given provenience is an index of variation of the
(ca. 0.03% by weight in land snails) which are relative ages of the shells.
preserved through time.
• Variation of the relative ages of the shells is an
• The proteins are constructed of amino acids, index of the potential for variation of the
which are initially all in the L- configuration. relative ages of artifacts in the same
provenience.
• Over time, the L- form converts (racemizes) to the
D- form, up to an equilibrium ratio of D:L. In • If A;1 values are tightly clustered within
this study, the racemization of L-isoleucine to analytical equality, the pattern reflects
D-isoleucine is analyzed. contemporaneity of burial of shells. Artifacts in
the same provenience also are likely to be
• The D/L value of the amino acids (here, the contemporaneously buried.
D-alloisoleucine/L-isoleucine, or A;1, value)
can be used as a measure of relative age. • If A/I values are not tightly clustered, the A/I
values and other evidence are examined to
• The rate of racemization depends on evaluate the likely cause(s) of A/I variability. In
temperature and varies among different taxa. many cases, evaluation is ambiguolls, but
Thus, estimation of absolute ages from A/l values specific reasons for doubting integrity can be
requires a rate calibration. This is usually identified.
accomplished through radiocarbon analysis of
selected samples. • In this study, most A/I values are obtained
from shells of Rabdotus mooreanus. For one site,
shells of Oligyra orbiculata also are used.
Random and Systematic Error

• Analytical (measurement) error is random. Why Use Racemization Dating?


Analytical error is a percentage of A;1 (typically
ca. 5%). In this study, analytical error is ca. 4%. • Although use of bone for racemization dating
has been discredited, mollusk shells have bee/!
• Temperature variation introduces systematic shown to yield reliable racemization dates when
error. properly calibrated.

• Calibrate to control for climatic temperature • Racemization analyses are performed with off-
history. Supplement with other chronometric the-shelf technologies, and are relatively
analyses under conditions of divergent slope, inexpensive.
aspect, and depth.
• Because racemization analyses are relatively
• Fire can induce racemization. This effect inexpensive, more dates can be budgeted at a
sometimes can be identified during integrity given funding level.
assessment.
• The ability to use more dates improves one's
• Amino acids can be leached. Interpret with ability to assess integrity by providing a more
caution when samples come from contexts robust data set.
with evidence of long-term saturation.

3-4
Calibration

Calibration of A/Ion Radiocarbon Age Calibration of All for Last 5000 BP Cali bra tion of All on Calendric Age

-+
0.25 -f1-'--'-.L....l-'--'-.L....ll.....a....L-..L...Jl.....a..-L-............... 0.15 -f'~~<-1.-'-~<-1.-'-~-'--'-"'-'-'--'-"'-'-+ 0.15 t' ~:'::->.-:-,-""",--,--,--,-:--,--,--,-""",-,-""",--,-+
AJ1 = 0.008 + 2.28. < 10"5 (RYBP) AI~ = -0.004 + 2.96 x 10"5 (RYBP) AI~ = 0.0122 + 2.67 x 10"5 (cal B.P.)
R2 = 0.906 A = 0.978 R = 0.966
• •
0.20
High R2
Points >5000 BP
• 0.10 Very high R2
0.10 Very high R2
scattered Points well
:ii :ii
Points in middle are 0.05
• distributed around
regression line
0.05
Calibration predicts calendric
age for last 5000 years
0.15 -I above regression line: / . ••
linear model not best fit • • Intercept = -0.004:

• •
belowmeasured modem Intercept = expected

~J 0.00 Iii I I l
v~lu,e <p, Oj12 0.0 O
I modern value
iii i'
G. o 1000 2000 3000 4000 5000 a 1000 2000 3000 4000 5000
Radiocarbon Years B.P. Calendric Years B.P .

w • For last 5,000 years, linear relationship is very Calibrating radiocarbon ages with correction

--
("
0.05
strong. However, intercept and radiocarbon for 445 year average age anomaly predicts
Points <5000 BP show measurements on modem shell indicate measured modem value in regression of A/I
tr:iJ8 strong linear trend
on calendric age.
radiocarbon age anomaly for Rabdotus.
0.00 -fl-,--.--,-,-..,.-,rr-r-r-r-.-r-r-r-TO-r-+
o 2000 4000 6000 BOOO
Radiocarbon Years B.P.

Regression of A/Ion radiocarbon age shows Results:


strong relationship. However, linear model may
not be best as result of climate change and/ or • Calibration provides calendric dates for last
other systematic error" 5000 years from All values.

• Strong age predictive value supports use of


A/I values for relative dating.

• All values as relative dates provide an index of


contemporaneity of burial and, hence, integrity.
High Integrity

Theoretical Signature of High Integrity A/I Values from Test Pit 2, 41CV1200, Fort Hood

-- . All values tightly clustered, 500


Calendric Years B.P.
1500 2500 3500

-- .
many within analytical error -90 -,
of youngest shell She All values clustered
within analytical error with
two older shells

-
Excavation error or intrusive 'E-100 >-+-0
.c ... - -E--- shells: material from newer -8. Clustered racemization dates

-
i5. • deposits included .c consistent with radiocarbon dates

.
~
Ql
i5.
Q
Ql / 0 0 0 ........ • •

-
-110
.~
Q
Three Two
~ shells t<:=:>I shells ......... He, Charcoal
~ 14C,ShoU

~ I 0,. R.acomlUroIiDfl

--
.£ Rodent ca. 250
'"0, ~ midden-
-120 yea",

•Redeposited shells stand out


-- •

Increasing All
as isolated higer values among
clustered lower values

:>
Interpretation:

Clustered A/I values imply that the assemblage was rapidly buried with
no detectable introduction of materials that post-date burial.
Duration of burial is ca. 250 years or less for 10 em of deposits.

Clusters of analytically equal AlI values are very strong evidence of Presence of scattered older shells without scattered newer shells
rapid burial and low levels of vertical mixing. Samples from rodent implies that redeposition-not disturbance-is source of older materials.
middens should be avoided since they reflect rodent behavior, not If the sediment source contains small artifacts, they could have been
conditions of burial. Presence of a few high values among well transported to tlte site. Small artifacts deposited at the site may have
clustered values is likely to reflect low levels of vertical displacement been rearranged during burial.
or redeposition of shells eroded from older contexts.
Overall, integrity of the artifact assemblage is probably high for any
problem requiring temporal resolution of 250 years or less per 10 em
level.
High Integrity (cont'd)

A/I Values from Test Pit 6, 41BL532, Fort Hood A/I Values from Test Pit 1, 41BL513, Fort Hood

Calendric Years B.P. Calendric Years B.P.


2000 3000 4000 1000 2000 3000
-60 -. -50 -,

Seven /VI values clustered Five- analytically


E -70
within analytical error
E -60 equal shells
.3- .3-
.c:
a. •
Intrusive or
o 0000
I
0 .c:
a.
o O<lX> •••

Q)
Two shells Q) 1=1 Three older
o -80 excavation o -70 ca. 325 shells
error 1< >1 years
ca. 430 ~: :>1
years ca. 735
u.>
-90 -80 years
.:..,
Interpretation: Interpretation:

All values clustered at analytical equality imply that burial occurred All values clustered at analytical equality imply burial over a period
over a period of ca. 430 years or less for a 10 em level. Youngest shell no longer than ca. 325 years. However, older shells may reflect relatively
either is intrusive or was introduced by excavation error. slow burial, systematic error, and/or redeposition.

Absence of older shells implies that detectably older materials were not If slow burial accounts for older dates, artifacts may be in situ, but
introduced by redeposition or disturbance. temporal resolution (10 em over ca. 735 years) is fairly low. If older
shells are redeposited, temporal resolution improves, but not reliability:
Overall, integrity of the artifact assemblage is probably moderately artifacts occur upslope at this site. If systenzatic error accounts for
high. For problems requiring relatively low temporal resolution older dates, limits of temporal resolution are unknown.
« ca. 430 years per 10 em level), integrity can be regarded as high.
Integrity is ambiguous, but there are good reasons to doubt that artifact
associations are reliable.

Whereas a pattern of highly clustered A/I values is very strong evidence of rapid burial and an
unmixed assemblage, less clustered patterns are ambiguous. However, even ambiguous patterns can
point toward specific reasons for doubting integrity.
Compromised Integrity

Theoretical Signature of Compromised Integrity A/I Values from Test Pit 2, 41BL754, Fort Hood

••••• All

----
.....
Distribution
randomized by
0.00 0.05 0.10 0.15 0.20
-20 r,- - ' - -. . .--r--y-~r--,---"--,

o
L::
C.
Q)

·w'"
c:
--_.. disturbance

E
~
.c.
15.
-30
Youngest shells are
effectively modem
...........
"- Very wide

....,, •••••• '"
0-40 Total of four range of
'~"
00
"
.EO ._ ...
••••• • Redeposited,
or buried
together -50
analytically
equal shells
All values

t •••• • •
. ...
in situ?

_• •
•• •• Interpretation:

. Poorly clustered A/I values imply assemblage mixture. This site


••••• • is a rock shelter with internally derived sediments as the
predominant fill. Redeposition of shells is unlikely.
Increasing All ~
Lowest A/I values are effectively modem (ca. 0.5% net
racemization). Presence of very young shells 30-40 cm deep
Disturbance, redeposition, and slow burial have similar signatures implies disturbance, but there was no evidence of vandalism or
since all involve mixing of shells of different ages. Inability to natural agents.
distinguish between these may not be important: all three
processes lead to low confidence in assemblage integrity. Other Sediments were powdery, and unit walls collapsed repeatedly
kinds of evidence are necessary to reduce ambiguity. during excavation. Excavation error may have mixed old and
new materials. Integrity of the shelter is unknown, but the
excavated assemblage is severely compromised regardless of the
cause.
Compromised Integrity (cont'd)

A/I Values from Test Pit 2, 41BL168, Fort Hood A/I Values from Test Pit 5, 4181.532, Fort Hood

All
0.00 0.05 0.10 0.15 0.20
All
0.00 0.05 0.10 0.15 0.20 0.25 0.30
Orl--~--'---~--r---~-'--~--,
40.

Wide range of values Heated or


E -10 Best case redeposited
~ f: H E -50.
f=i
so
15.
00 • • •
I
• ~
so «IN' • ••
shell?

o" -20 Two analytically 15. V
equal shells o" -60 Two analytically
equal shells
~: >1
w
, -30 Worst case
-70
'"
Interpretation: Interpretation:

Poorly c/ustl?red A/I values imply assemblage mixture. .. Poorly c/ustl?red A/I values imply assemblage mixture. Test pit is
on a colluvial bench, and the spread of values is consistent with
This site is a rock shelter with externally derived sediments as redeposition of shells-and artifacts-from the ancient upslope
the predominant fill, with no evidence of disturbance by vandals. surface. If systematic error does not affect values, then best-case
Redeposition of sheils is likely to account for A/I variation. interpretation is compromised integrihj.

A lithic production locale is on a mesa top 3-4 meters above the However, the highest value could be anomalously high as a
shelter opening, and artifacts are scattered on the slope from the result offire. If it is, then all A/I values could be affeded, and
mesa to the sheltl?r. The likelihood of redeposited shells is the range of A/I values would under-estimate integrity.
accompanied by a high likelihood of redeposited artifacts.
Confidence in assemblage integrity can be regarded as low. Upshot: Integrity is .mknown_
Effects of Fire

Theoretical Signature of Influence by Fire Extreme Values Reflecting Influence of Fire

-- As yet, there is no chemical or other test that can unambiguously


identify effects offire.

-- \
Range of values will be
·stretched- as a result of
differential distance
from fire
A robust calibration is the best currently available means for identifying
A/I values that are far too high to be plausible candidates for
redeposition. Scatter at the old end of the Fort Hood calibration prevents
s:
li
May have
\
\

...
• • • • • • confidently using it to determine A/I values that pre-date human

--
\- •
an apparent occupation.
0" stratigraphic ,
.,.'"c: reversal if all
shells were
Some values may be
Mextreme- (l.e., may have At 4ITG307 in San Angelo, three shells with identical radiocarbon ages

--
nI
affected Implausibly old apparent age) had widely divergent A/I values. Shells were recovered near a rock
~
,
w
"
.!: hearth and scattered burned rocks. Higher A/I values surely reflect
0

t
influence offire.

--
- Increasing N1 ----7
0..
CD
~
8900

nI

"
>- All values widely divergent
§ 8700
-enI
......... All surely / •
Combination of apparent stratigraphic reversal, "stretched" 14Cdates [ •
values, and "extreme" values is very strong evidence for "
.2
"C
identical
influence by fire. If other evidence (e.g., oxidized sediments, '"
II: increased
burned rock) is available, it reinforces this conclusion. by fire
8500 ------~----~~----~~----~
LI
0.05 0.10 0.15 0.20 0.25
N1
Effects of Fire (cant'd)

Examples: Shells Recovered from Levels Near Hearths Prospecting for Fire-Oriented Features

A/I Values from Test Pit 2, 4ITG307, San Angelo


A/I Values from Test Pit 6,
Nt 4ITG307, San Angelo
= '" 0.00 0.05 0.10 0.15 0.20 0.25


--
-1001 110
Shells can be as far as
, this from fire in Feature:

J .:. § -120 .
~ ~ and 'stretched" •
pattem probably
• • •
..
11
0.0
-110 I
02 0.4
Nt
0.6 0.8 1.0 1.2
111

r[
12,...
"4 .. ,
,
... Scattered.
- . bwned ,
Excavation rock. • Nine shells
;;
Ci.
~ -130
show effects of
differential dIStance
from flr8

• •
Feature 2
elevation
~
13 -
m

'" ro.
:[,'20 sIlt'S
Four Extreme values and ·stretched" pattem
probably show effects of fire

." / •
Um~s \ _ recovered from -140 Shells from same levels as
hearth and scattered rock
14 ~ -130 \ These values are 13
Two especially high
'f
~
-150 15 shells
_140 L ...J -14
Interpretation: o Analytically Equal Values

A/I values probably are affected by fire. Integrity


Backhoe Trench cannot be confidently assessed with these values. No evidence of fire was visible in Test
Pit 6. However, pattern of values and
A/I Values from Test Pit 8, 4ITG307, San Angelo the especially high values of two
-" Nt shells imply that fire may have been
'" 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 nearby. The area near Test Pit 6 may
-40 I 14
Excavation
be a good place to look for a hearth or
Feature 5
limits similar feature.

'~t
Rock. Hearth Extreme values and

H
·stretched" pattern Feature 5
probably show eHecis of elevation r-
:[-60 differential distance 6~
m from fire
10-20 em vertical m.
Nine Shells Cl distance from
recovered from this ·70 Feature 5 to shells 7
1 xl mtestpit
••0 L
• • • •• • !.U •
Shells can be as far as
from fire in Feature 5
Interpretation:

A/I values probably are affected by fire. Integrity


cannot be confidently assessed with these values.
Backhoe Trench
Multicausal Interpretation
Profile of Burned Rock Midden at 41B1598 Racemization and Radiocarbon Data from
Bumed rtldI: Test Pit 1, 41B1598
c:::::=::J Excavation em A/I Values
7 Level 0

a-<lj> 0.. (I) oo>-d Patterns lor Rabdotus 10 Rabdolusl


3 0> b OC 0
o and Oligym ara similar Q Oligyra I
\- - ---------
4 Modem
shells
13 10 5 o 9--000 0- f
Very hIgh NI
maiMS
All for e and f ara widely divergent
'0
• Test Pit 1 excavated adjacent to the central 6
hearth.
~o o-j Patterns for Rabdotus
• Levels 3 and S in burned rock midden 7 • h ,
and O/igyra are similar
~o

matrix. , 0
,
0.0 0.1 0.2 0.3 0.4 0.5

• Level 7 in underlying midden. Radiocarbon Dates


Problem: How extensive was assemblage ~
Pattern of radiocarbon
mixture during formation of the burned rock '">
,;
.~
b~
dates for a-d is similar to
midden? ~
~
,~
pattern of All values
d~

"'>
,;
~

Radiocarbon dates for e and f e lOt


~
~
are Identical, unlike NI values I 0

~ g~

"ii Pattern of radiocarbon l~


~ I Ralxiotus
>--0-< I dates for g.j Is similar to ho
. .--. ChM:oaI. pattern of All values J~

o 2000 4000 6000 6000 10000


Ca!endric Years B.P.

Interpretation:

• Level 7 has a mixed assemblage judging from A/I values and


radiocarbon dates. Mixture apparently results from slow burial
and/or reworking. Absence of newer shells shows that deposits
were sealed ca. 6000 cal BP.

• Level 7 deposits are reliable for problems requiring "Early


Holocene" assemblages, but unreliable for problems requiring
finer temporal resolution.

• Rabdotus in levelS have signature of slowly buried assemblage.


But, the extreme (!) value of Oligyra implies that the AlI values
under-estimate integrity due to heat from the nearby hearth. A/I
-radiocarbon pairs "e" and ':f' support this. Absence of lower
A/I values implies that dated charcoal filtered down. Integrity
is probably moderately high.

• Level 3 has signature of a mixed assemblage. Radiocarbon dates


and absence of extreme A/I values imply that the pattern was not
induced by fire. Although most radiocarbon dates are ca. 1000-
2000 cal BP, A/I values show that younger and older materials are
mixed in. Integrity of level 3 is low.

3-12
Conclusions
• Amino acid racemiZiltion analysis has allowed us to identify sites where we can have high
confidence in the integrity of archeological deposits.

• Because it is relatively inexpensive, amino acid racemiZiltion analysis of mollusk shells can
be an important addition to the tools we already use. In conjunction with other lines of
evidence, it is possible to provide robust interpretations of the likely conditions of burial
and preservation of archeological assemblages.

• Although some cases of ambiguitycould not be resolved, other assessments allowed us to


detennine that there were good reasons to doubt integrity. The ability to know whether
we can be confident about the integrity of archeological data is an important aspect of the
general problem of determining whether our data are sllitable for the analyses we want to
perfonn.

References
Ellis, G. L., G. A. Goodfriend, J. T. Abbott, P. E. Hare, and D. W. Von Endt (in press).
Assessment of integrity and geochronology of archeological sites using amino acid
racemization in land snail shells: Examples from central Texas. Geoarchaeology.
Goodfriend, G. A. (1987a). Radiocarbon age anomalies in shell carbonate of land snails from
semi-arid areas. Radiocarbon 29:159-167.
_____ (1987b). Chronostratigraphic studies of sediments in the Negev Desert, using
amino acid epimerization analysis of land snail shells. Quaternary Research 28:374-392.
_ _-,-__ (1987c). Evaluation of amino-acid racemization/ epimerization dating using
radiocarbon-dated fossil land snails. Geology 15:698-700.
_____ (1989). Complementary use of amino acid epimerization and radiocarbon
analysis for dating of mixed-age fossil assemblages. Radiocarbon 31:1041-1047.
_-:--::-__ (1991). Patterns of racemization and epimerization of amino acids in land snail
shells over the course of the Holocene. Geochimica et Cosmochimica Acta 55:293-302.
_ _ _ _ (1992). The use of land snail shells in paleoenvironmental reconstruction.
Quaternary Science Reviews 11:665-685.
Goodfriend, G. A., and V. R. Meyer (1991). A comparative study of amino acid
racemization/epimerization kinetics in fossil and modem mollusk shells. Geoc"imica et
Cosmochimica Acta 55:3355-3367.
Goodfriend, G. A., and J. J. Stipp (1983). Limestone and the problem of radiocarbon dating
of land-snail shell carbonate. Geology 11:575-577.
Johnson, B. J., and G. H. Miller (in press). Archaeological applications of amino acid
racemization dating: A review. Archaeometry.

3-13
ARCHAEOLOGY AT FORT HOOD
by David L. Carlson

Introduction

Cultural resources investigations began at Fon Hood in 1978 (Skinner, et. al. 1981). From 1978 to 1991
surface surveys of 95 percent of the surveyable area of the post are complete. Unsurveyable areas include the
anillery impact areas, the permanently dudded area, and the cantonment. These areas are either heavily disturbed
or too dangerous to survey. The total number of sites recorded by post archaeologists and contractors is 2,150.
Since 1992, systematic shovel testing and geomorphic investigations have been conducted at 571 prehistoric sites
and National Register eligibility testing has been conducted at approximately 100 sites.

Prehistory

The prehistory of Fon Hood may be broadly divided into six periods spanning the last 12,500 years:

Paleoindian (10,550-6,550 B.C.). The Paleo indian period includes the earliest inhabitants of the Texas.
During the beginning of this period, central Texas was cooler and moister than todays climate. Extinct animals
such as manunoth, mastodon, horse, and camel occupied the area. The earliest people in the area made a
distinctive projectile point called the Clovis point. The sites occupied by these people often contain extinct fauna.
Later Paleo indian people where hunters and gatherers who adapted to changing environmental conditions as the
last Ice Age ended about 10,000 B.C. This period is poorly studied because sites of this age are very rare.
Prewitt (1983: 210) lists only three sites in central Texas with radiocarbon dates falling into this period.

Early Archaic (6,550-3,050 B.C.). The Early Archaic represents as shift in projectile point styles from
the lanceolate styles of the Paleoindian to a variety of barbed and notched styles. Population is assumed to have
been low, but many sites have undoubtedly been destroyed over the years by geomorphic processes. This period
spans the Altithermal, a climatic period during which the environment was apparently warmer and drier than
today. The specific details of environmental change during the Altithermal are unknown. This period is only
slightly better understood than the Paleo indian. Prewitt (1983: 209-210) lists only six sites in central Texas with
radiocarbon dates falling into this period.

Middle Archaic (3,050-650 B.C.). While the initial appearance of burned rock features begins in the
Early Archaic, they are abundant and extensive during the Middle Archaic. These large accumulations of burned
rock called burned rock mounds or middens appear throughout central Texas. The burned rock was apparently
used repeated over a considerable period of time. Whether the burned rock mounds represent the cooking of root
foods, the processing of acorns, or some other purpose is still unknown. Some archaeologists believe that the
Middle Archaic represents a peak in the population of the hunter-gatherers who occupied central Texas (Weir
1976). Plant gathering may have become more imponant during this period. At Fon Hood there are more
Middle Archaic than Early Archaic sites, but there are considerable more Terminal Archaic sites. Bison which
had been absent during the Altithermal move back into central and even southern Texas. Eight sites in central
Texas are reponed by Prewitt to have radiocarbon ages within this period (1983: 206-208).

Late Archaic (650 B.C.-A.D. 200). Burned rock accumulations are fewer in number. Bison reinvade
Texas from the southern Plains and were hunted although their imponance may not have been that great in the
Fon Hood area. The Late Archaic is slightly better studied than the previous periods, but only nine sites in
central Texas have radiocarbon ages within this period (Prewitt 1983: 207-208).

Terminal or Transitional Archaic (A.D. 200-700). This period has the highest site density at Fon
Hood. Rockshelters often contain evidence of Terminal Archaic occupations, which may indicate a change in
settlement pattern. Bison are absent from central Texas during this period and foodgathering probably shifted

4-1
toward dispersed animals such as deer and heavy dependence on plant foods. Fourteen sites in central Texas have
radiocarbon dates falling in this range (Prewitt 1983: 206-207).

Late Prehistoric (A.D. 700-1780). The Late Prehistoric (called the Neoarchaic) is associated with the
introduction of the bow and arrow into the area. There may be linle or no change in hunting and gathering
practices. At Fort Hood and other parts of central Texas this period contains substantially fewer sites which may
reflect a population shift out of the area. Alternatively it could reflect a shift in settlement location. Small
rockshelters often have Late Prehistoric occupations, but these sites rarely have artifacts on the surface which
would allow their assignment to a period. Much of the Late Prehistoric occupation may be hidden in small
rockshelters and in the now flooded floodplain of the Leon River. The Late Prehistoric is divided into the Austin
and Toyah phases. The Toyah phase may represent a migration of Bison hunters from the southern Plains of
Oklahoma and the Texas panhandle into the area. The introduction of ceramic into the area indicates interaction
and trade with the Caddo to the east. Twenty sites with radiocarbon ages falling into this period are known from
central Texas (Prewitt 1983: 203-206).

History

Table 1 identifies the major events in the history of Bell and Coryell counties). The following summary
is based on the Fort Hood Historic Preservation Plan:

Pioneer Era (1830-1850). European senlement of the Fort Hood area did not begin until Rohert
Leftwich was awarded an impresario contract from the Mexican government in 1825. When Leftwich died in
1830, the administration of the colony passed to Sterling Robertson. In 1834 six families were granted title to
land in what is now Bell County. Robertson lost control of the colony in 1835 because he was unable to establish
100 families in the area. Control of the colony then passed io Austin and Williams. In 1849 the U.S. Army built
Fort Gates on the Leon River to protect settlers. Bell County was established in 1850.

RanchIng Era (1850-1880). Population grew steadily in the area except for the Interruption of the Civil
War. The early settlers in the area were largely self-sufficient. Based on the population density of Bell and
Coryell Counties we can estimate that 936 people lived within the boundaries of what is now Fort Hood in 1860.
That number nearly doubled to 1715 in 1870 and doubled again to 3819 in 1880 (Mueller-Wille and Carlson
1990b: 40). Domestic structures were often constructed of logs. At least seventeen small communties were
established withIn the confines of Fort Hood during this time. Very rew historic sites at Fort Hood can be
conclusively assigned to this period. Only three sites have components dating to 1860, 24 to 1870, and 27 to 1880
(Mueller-Wille and Carlson 1990a: 32).

Railroad Era (1880-1915). In 1880, the Gulf, Colorado, and Santa Fe railroad built a line through Bell
county. In 1882, the Missouri, Kansas, and Texas railway crossed tracks with the Santa Fe at Temple which then
became the major town in the area. Many communities foundered when the railroad passed them by. The
railroad connected Bell and Coryell counties to markets allowing them to obtain lumber, barbed wire, and other
goods more economically. Production for cash rather than home consumption grew as some ralsed sheep for
wool and others raised cattle. Rural population growth slowed and peaked as the estimated population of Fort
Hood grew from 3819 in 1880 to 6985 in 1900 and then declined to 6868 in 1910 (Mueller-Wille and Carlson
1990b: 40).

Automoblle Era (1915-1942). After a brief boom in conan prices, the area entered a post World War I
agricultural depression. The automobile facilitated travel and accelerated the shift from rural to urban areas.
During this time the estimated Fort Hood population dropped from 6868 in 1910 to 5409 in 1940. The towns and
communities of Fort Hood declined to the point that they were simply rural neighborhoods which depended on
towns like Gatesville, Copperas Cove, and Killeen for most services.

4-2
Table I. Summary of Bell County and Coryell County History
(from Anonymous [1893], Newcomb [1961]. Scott [1965]. and Tyler [1936]).

Date Summary
1687 Henri Joulel recorded Tonkawa and Mayeye Indians in Central Texas.
1698 Missions were established in northeast Mexico for the Ervipiame.
1801 Phillip Nolan went on hunting expedition in Brazos Falls region.
1825 Robert Leftwich granted empresario contract by Mexico.
1830 Leftwich's contract passed to Sterling Robertson; HamIet of Tenoxtitlan became first settlement in
Robertson's Colony.
1835 Nashville-on-the-Brazos founded; James Coryell given a headright grant in the Nashville Colony in
present-day Coryell County.
1836 Bell County residents fled eastward in "Runaway Scrape"; Milam County created out of the Milam Land
District; Coryell County was later created out of Milam County.
1841 Governor Sam Houston pacified Indian problems for settlers in Bell County.
1849 Fort Gates established as last garrison along the frontier line from Fort Duncan, near Eagle's Pass, to
Coffee's Station on Red River.
1850 Bell County officially organized; "Nolandsville n (renamed "Belton n in 1852) designated as county seat.
1852 Fort Gates was abandoned.
1853 Fort Gates was temporarily used as a quartermaster depot.
1854 Coryell County created; GatesvilJe later designated county seat.
1859 Belton (pop. 300) the only town of significance.in Bell County; Governor Houston gives direct aid to
settlers to repulse Indians~ First cattle drive out of Coryell County to Shreveport, Louisiana.
1866 Cattle business developed in Texas and trails to northern mark~ts passed through Bell County.
1870s Wends settle The Grove.
1880 Gulf, Colorado and Sanra Fe railroad passed through Bell County.
1882 Missouri, Kansas, and Texas railway passed through Temple; Missouri Pacific CKaty") branch passed
through Belton; Texas and St. Louis Railway Company completed tracks to Gatesville; Gulf, Colorado,
and Santa Fe Railway Company reach southwestern Coryell County from Galveston.
1890s Wends settle Copperas Cove; Cotton and wheat prices declined as the availability of manufactured
goods increased.
1893 Panic began and lasted until 1899.
1904 Boll weevil reached Bell County and destroyed crops.
1907 Stephenville North and South Texas Railway Company laid tracks from Stephenville to Hamilton.
1911 Stephenville North and South Texas Railway Company extended lines to both Comanche and Gatesville.
1913 Bond issue passed in Bell County for construction of better roads.
1914 Farm prices dropped with onset of World War I followed by a war-inflated boom.
1920 Period of deflation in Ben County.
1923 Federal aid for highway construction granted to Coryell County.
1930 Community Natural Gas Company provided service for 500 customers.
1935 Community Public Service provided electricity for 783 customers.
1936 Rural Electrical Association available in Bartlett region of BeII County.
1942 Camp Hood activated as a tank destroyer training center.
1951 Camp Hood renamed Fort Hood.

4-3
Military History

In 1942 the Army purchased most of the land which is now Fort Hood to establish a tank destroyer
training center called Camp Hood. In 1943 the initial construction was completed with 5,630 buildings and 35
fIring ranges on 160,127 acres. Between 1942 and 1945 the peak wartime troop strength of Camp Hood reached
130,000. At the conclusion of World War II troop strength dropped to 14,000. In 1950 the Army designated
Camp Hood as a permanent installation and changed the name to Fort Hood. In 1952 additional land was
purchased around Lake Belton and added to the Government reservation bringing it up to 216,915 acres.

Research Topics

While a great many research topics have been investigated using the data from Fort Hood, they can
generally be summarized under three broad areas: broad patterns in site distribution through time, settlement
patterns, lithic procurement and technology, and burned rock midden/mounds. During the FOP fIeld trip we will
have the opportunity to visit sites representing all of these areas of research.

Table 2 summarize the the distribution of prehistoric site components as of 1990. Two small surveys
since then have not altered the overall picture. The sites on the fort can be divided into those within 10 km (a
days foraging radius) from the Leon River, those within 10 km from the Lampasas River, and those more than 10
km from either river. Only small portion of Fort Hood lies within 10 km of the Lampasas River (West Fort Hood
located below Highway 290). The low site density in this area relative to the other two areas is probably at least
partly a reflection of this fact. Site density is highest near the Leon River. For most periods site density in the
area more than 10 km away from the Leon River is half to two thirds the density for sites within 10 km of the
Leon. The other striking feature of the site distribution is the dramatic decrease in site component density after
the Terminal Archaic. This pattern has been noted by Prewitt (1983) for other areas of Central Texas. He
suggested the the striking change in site density could be explained in terms of a population decrease in the area.
Subsequent research suggests that other factors, particularly a change in mobility strategies, may account for
much or all of the change. The Terminal Archaic is partly created by the short timespan estimated for the use of
several dart point forms such as Ensor. If these point types were used longer, the Terminal Archaic peak would
not be as large. Secondly, the pattern seems to hold only for upland surveys. Most of the reservoir surveys were
conducted many years ago and did not include a systematic examination of the entire surface and are not published
in the detail that modem surveys are. Late Prehistoric sites do seem to be common in the floodplains, however,
and most sites tested in the floodplain contained Late Prehistoric components. Third, two site types appear to
have been occupied during the Late Prehistoric, but this use is rarely detected during surface surveys. Burned
rock mounds regularly produce radiocarbon age estimates in the Late Prehistoric range, but often do not contain
Late Prehistoric arrow points which are the only way to identify sites of this period from the surface.
Rockshelters also regularly produce evidence of Late Prehistoric occupation, but arrow points are rarely found on
the surface. This means that upland surveys will systematically underestimate the number of sites occupied during
the Late Prehistoric.

Table 3 shows the distribution of historic components at Fort Hood. Figure 2 shows that site densities
increased rapidly between 1880 and 1910 where it peaks and that site density began to drop after 1910.
Interestingly, the relative densities for sites near the Leon, near the Lampasas, and away from either river is
similar to the prehistoric pattern. Comparison with historic population fIgures indicates that the site density
underestimates the nineteenth century population. That is, there should be more sites occupied between 1860 and
1890 than we have found. While rural population in Bell and Coryell Counties did drop off after 1920, the site
density drop is much more rapid.

4-4
Table 2. Prehistoric Chronological Components
(Source: Mueller-Wille and Carlson, 1990a)

Period Age Components


PALEOINDIAN
Early 12,500 - 10,000 BP 32
Late 10,000 - 8,500 BP 37
ARCHAIC
Early 8,500 - 5,000 BP 147
Early/Middle 4
Middle 5,000 - 2,600 BP 237
Middle/Late 7
Late 2,600 - 1,750 BP 141
Terminal 1,750 - 1,250 BP 227
Unknown Archaic 94
LATE PREHISTORIC
Austin 1,250 - 650 BP 48
Toyah 650 - 200 BP 23
Unknown Late Prehistoric 59

120 r-------------------------------------------------~
E
~
>10krn
CJ
(f) 100 < 10 krn Leon
Terminal Ardlaic /1,
o
o ••••••• < 10 km Lampasas ," ,
~ I

,,,
'-
Q} 80 I
a.. I

E ,
I
.2
ffi 60

~
'-
,,
40 late Archaic,'
Q}
a.. ,, ,
,,
-(f)
C
Q} Middle Archa~ , , ,
I
I
--.
......
c
o
a.
20
-- ..... . .••. Tnvah
-,

o t~E;"~'Y~~~~~~'e~~~~~"Y~M~~~~-::-~-:-=-::::=::.~
. :.~. ~..~.~. ~.~. ~. ~.~..~_A~V~S"~.n_.~.~J
E Paleoindian Pal~~~n.. __ .. .. .. .. . ., .•... , ...

o
o ------ ........... . .....
12,000 10,000 8,000 6,000 4,000 2,000 o
Years B. P.
Figure 1. Chronological and Geographic Distribution of Prehistoric Components (after Mueller-Wille and
Carlson 1990a).

4-5
Table 3. Historic Chronological Components
(Source: Mueller-Wille and Carlson. 1990b)

Decade Estimated Population Components


1860 930 3
1870 1703 25
1880 3793 28
1890 5395 161
1900 6939 139
1910 6822 385
1920 6048 325
1930 6056 275
1940 5373 143

80 r--------------------------------------------------,
-->10km
I,
I ,
• - - - < 10 km leon
·1
I ' ,,
••••••• < 10 km lampasas
E
• 60
I
, I ,
...... '\
::.:: I \

& \
\
(f) \
o \
\
040 \
~
\
~
\
Q)
Il..
ffJ
Q)
.t:
(f) 20
, ,
,,
'.
I • • •••••
I " ........................ .

,----'.... ...... .
oL-__~~--~·~.. ·~·~__~____L -_ _~_ _~
1840 1860 1880 1900 1920 1940 1960
Years A.D.
Figure 2. Chronological and Geographic Distribution of Historic Sites (after Mueller-Wille and Carlson
1990a).

4·6
Settlement pattern studies have focused on mobility strategies of hunter-gatherers. Archaeologists
distinquish between residential mobility where the whole band moves from site to site and exploits the resources in
the immediate vicinity of the site from tactical mobility where the band stays in one place and smaller task groups
collect foodstuffs and bring them back to the base camp. The first involves moving the mouths to the food, while
the second involves moving the food to the mouths. Tactical mobility allows greater settlement stability. This in
turn permits a group to have more possessions and to have possessions which are difficult to move. The greater
sedentism permitted by tactical mobility is often considered a prerequisite for the domestication of plants, the
construction of substantial housing, and the use of pottery. For these reasons, archaeologists are particularly
interested in documenting changes from residential to tactical mobility. The issue is complicated by the fact that
there are no straightforward diagnostic criteria of these mobility strategies and by the fact that real hunter-
gatherers can combine both strategies. One immediate consequence of mobility strategies is that the same number
of people can generate very different numbers of sites. A pure tactical mobility strategy would generate a large
base camp with relatively high visibility in the archaeological record and a number of specialized camps which
might be very difficult to fmd. A pure residential strategy will generate numerous sites throughout the year that
are all comparable in their visibility in the archaeological record. One way to interpret the dramatic decrease in
site density during the Late Prehistoric is to propose that mobility shifted from residential camps scattered in the
upland and floodplain environments (Archaic pattern) to base camps in the floodplains and much smaller (and
harder to detect) specialized camps in the uplands. During 1990 and 1991 The Texas A&M field school
investigated five rockshelters to search for data that might help to confirm or disprove this proposition. All of the
rockshelters contained evidence of Late Prehistoric occupation (although we did not know this initially). Four of
the rockshelters are quite small and could have housed only a single family or perhaps two. We thought that this
might suggest that the rockshelters were occupied by specialized task groups, but analysis of the lithic tools does
not support this expectation. In fact the range of tool types is quite diverse suggesting that the rockshelters were
used by one or two families. Furthermore, one of the small rockshelters contained a deciduous tooth from a child
of about 7 or 8. This evidence suggests that Late Prehistoric bands split into small groups (micro-bands) in
narrow protected canyons during some time of the year and iIIen aggregated into larger groups in the floodplain
during a different season. Further investigations at Fort Hood and elsewhere in central Texas might test this
model by careful comparison of floodplain and upland sites.

Fort Hood is an ideal place to study the way in which stone was selected, manufactured into chipped and
ground stone tools, worn out or broken, and discarded. The fort contains numerous lithic procurement areas
where chert is readily available on the surface. The abundance of this critical raw material probably helps to
explain why site density and size is so high on the post. Counting these large lithic exposures, which contain
abundant evidence of prehistoric tool manufacturing, 7.2 % of the post has some evidence of prehistoric activity.
[n addition, the chert is quite variable and there are some varieties that are found in fairly restricted parts of the
post. Research by Dickens (1995) and by Frederick and Ringstaff (in Trierweiler, 1994) has defmed 17 different
chert types for the fort. The spatial distribution of these types provides some indication of how far raw materials
moved across the fort and to what degree chert selection was affected by quality and availability. Examination of
chert diversity through time and by site type may help to clarify changes in mobility patterns over time, but these
comparisons have not yet been made and the samples from dated contexts are only just beginning to be large
enough to conduct such studies.

Burned rock mounds/middens are the enigmatic features of central Texas. Despite excavations of a large
number of these features, we still do not really know what they represent. The classic burned rock mound is a
circular or oval concentration of burned rock from 0.5 to about 2 m high. The mound often contains animal bone,
chert flakes and tools, charcoal, and carbonized plant remains. Some have a central depression. Burned rock
concentrations containing artifacts and ecofacts, but laCking distinct boundaries and are often buried in alluvial or
colluvial contexts are also found. [t is not clear if they represent three, or more, different uses of burned rock for
processing foodstuffs or a single use in different depositional environments. We do not know what they were used
for although numerous suggestions can be found in the literature from roasting pits for meat, to leaching and
cooking acorns, to cooking root foods. The features were apparently used repeatedly and may well have been
used for several purposes. Understanding what they were used for would help to clarify mobility strategies since
large-scale bulk processing of a foodstuff for storage should be part of a tactical mobility strategy, but not a
residential one. Research at Fort Hood to date has focused on chronological issues regarding these features.
Prewitt (1981) suggests that burned rock middens and mounds were constructed between about 5,000 and 2,250

4-7
years ago (the Middle Archaic). This inference is based on the fact that these features often contain dan points
that have been dated to this interval from other sites, especially stratified midden sites. Recently archaeologists
have begno to obtain multiple radiocarbon age determinations on these features with the result that many of them
contain evidence of Late Prehistoric use. This discrepancy between the age estimate based on the anifacts and the
age estimate based on the charcoal has not been explained. That the features represent multiple episodes of use
and reuse is indicated by the fact that stratigraphic reversals in radiocarbon age estimates are common. One
possibility is that the features are Late Prehistoric and the Archaic-age points were incorporated by Late
Prehistoric people either accidenrally or deliberately (Steve Black, personal communication). Another possibility
is that muliple re-use of the fearute oxidizes and breaks up the older charcoal so that the possibility of recovering
a piece of charcoal from the earlier use of the feature is very small.

While this brief summary does not exhaust the research potential of sites at Fort Hood, it should provide
some idea of how culrutal resources investigations at the fort are contributing to our understanding of prehistoric
lifeways.

Sites

The following are brief descriptions of the results of investigations undertaken at six sites on the fort. All
or only some of these sites will be visited depending on weather conditions and the availability of time.

Site 41CV960 consists of two burned rock concentrations or middens that were tested by TRC Mariah
and Associates in 1993 (Abbott and Trierweiler 1995: 483-498). The site straddles the confluence of Cowhouse
Creek with an unnamed tributary and includes portions of a Holocene terrace of Cowhouse Creek and the
flooplain of the unnamed tributary (Figure 3). The burned rock middens are located on either side of the unnamed
tributary. Test excavations consisted of six Ixl m test pits and six backhoe trenches. Artifacts were largely
confmed to the upper meter of the deposit. Despite the overall similarity in the appearance of the burned rock
fearutes on either side of the creek, radiocarbon age determinations indicate the the eastern midden is about 3200
years old (3200 ± 60 BP, Beta b-7(038) and the western midden is about 1700 years old (1730 ± 60 BP, Beta b-
70039 and 1690 ± 60, Beta b-7(037). Projectile points recovered from this midden include Late and Terminal
Archaic types (Ellis, Edgewood, Godley, Costroville).

4-8
!
I o
N .",,,,
" "
I CQ<'IIOU.-ed 01 1 l,Ielet Inle''tQAs.
(fe..al;o", !:lOUd 0t'I ossumed datum.

Test Pit
= Backhoe Trench
Site Boundary
Subarea Boundary
2- track Rood

,,
IMC,,),.... •,~"
,, ' '
,, ' '
, , ' "-
..... , , . .al-~
,, ' ,
A', ," ,
,, ' ,
,, ' ,
, ,
" '

Mopped Ronaid Corroll Texos. 9-29-9J

Figure 3. Site plan for 41CV96O (Trierweiler 1994).

Site 4IBL670 is a rockshelter that was tested by the 1990 TAMU Archaeological Field School (Carlson
1993a). It is located along Bull Branch and lies only 1.5 m above the floodplain. The shelter is 8 m wide by 10
m deep (Figure 4). Because it is triangular in plan section and because the ceiling is very low, the usable space is
very small. The shelter was tested with a 1x2 m test pit excavated to 90 cm. Flake density was modest varying
from 2 to 68 flakes per 10 cm level. Diagnostic artifacts recovered from the unit included one Seal/om (20-30
cm), one Bonham (30-40 cm), one Perdiz (40-50 cm), and two unidentifiable arrow points (30-40 cm).
Radiocarbon age estimates were obtained on charcoal at 20-30 cm (1080 ± 80 BP) and 60-70 cm (2020 ± 110
BP). Fauna recovered from the shelter include deer, rabbit, opossum, and bear.

4-9
,/ ".

...
DETAIL OF ROCKSHEL TER

41 BL 670
-==-'=- ..


Figure 4. Site plan for rockshelter at 41BL670 (Carlson 1993a).

Site 4lBL671 is a rockshelter that was tested by the 1990 TAMU Archaeological Field School (Carlson
1993a). The site has been heavily vandalized with obvious potholes covering about 25 % of the site (Figure 5).
The surface of the shelter is 3.5 m above an unnamed tributary of Bull Branch. The shelter is about 40 m long
and reaches a maximum depth of 4 m in the middle. Five different test pits were excavated into the site to
determine what periods of occupaton were represented and how badly disturbed the site was. Very little of the
site appears to be intact except for discontinuous patches near the bedrock. Human skeletal remains were found in
several units and a complete burial was excavated in units 2 and 5. The burial appears to be an adult female
between 22 and 29 years of age and about 5'2" tall. Portions of at least 5 other individuals were also recovered.
Some 15,132 flakes were recovered in the excavations. Dart points include Darl, Ensor, Marcos, and Marshall.
Arrow points include Perdiz and Scallom. The points indicate Late Arcliaic through Late Prehistoric use of the
shelter, but the there is no stratigraphic order to the materials. For this reason, no radiocarbon age estimates were
obtained for the shelter. Use of the shelter by non-human animals is indicated by numerous snake and rodent
bones as well as probably prey species such as deer and rabbit.

4-lO
\
'\
'V
\
--~---~---- ---

Figure 5. Site plan for rockshelter at 41BL671 (Carlson 1993a).

Site 4IBL233 includes two burned rock mounds that were tested by TRC Mariah Associates in 1993
(Trierweiler 1994: 207-213). The site contains at least five recognizable burned rock mounds. Two of these were
tested. Feature 1 is a 6 by 8 m mound consisting of burned limestone rock with associated flakes and charcoal. A
single lxl m test pit was excavated 50 cm into the mound adjacent to a backhoe trench. One Uvalde point (Early
Archaic) was recovered from level 3 and seven radiocarbon age estimates were obtained (Figure 6). Four
charcoal ages are statistically homogeneous and provide a pooled average age of 763 ± 43 BP. Feature 5 is a 7
by 7.5 m mound consisting of burned limestone rock with associated flakes and charcoal. A single Ixl m test pit
was excavated 70 cm into the mound adjacent to a backhoe trench. Seven radiocarbon age estimates were
obtained (Figure 7). Five charcoal age estimates form three distinct clusters: 2863 ± 70 BP. 1742 ± 62 BP. and
630 ± 70 BP. No diagnostic artifact types were recovered. but the dates suggest Late Archaic through Late
Prehistoric use. Both mounds contained modern. unburned seeds at depths of 40-5Ocm indicating movement of
small materials through the deposit is a continuous process.

4-11
Feature 1 Plan Map

BHT Backhoe Trench


t
H

I
TP Test Pit (1x1m)
~ Treeline

o 2 meters

Backhoe Trench Profile

• 5 J 2 o
midden mlrtru:; eJay Ioem; YtK'y stony; ~ subengo.IIar ~ 7Bdin9 into massive; hatd ay: st1clty....-eot atu'IdanI: buned, lX'Ibune<Ilimestone clasts;
aI'Ju"Idant fine f1JQtS ill ~ 20 ern; common WOOCy nxu (Me) ~ Yf¥.'( daOI. tJrown (10YR 211 o;nd-ng doWn into appro:lImately 7.5YR 211)

_" lmestone stabs w/day Joem in intentices (dismegn6ng beQ-odo;) stabs _ 15 ern ttlic:Io;. up to 0_5 m long. ~ dat1I; btown clay loam iI1 intentices
(m~sive, hatd dry; stic:Jo;.y wet. YtIry daf'k btown (7,5 YR 211 apprnximately)

West Wall Tn
cmbs
O.-__ ~~~ ____ ~
17Ot508
00 0 860 t70C

20-
'" 0' q <c. 870 t 70C
o limestone rock

C> DD '6 670 t 70 121" t "'" H


650 t 70C 15-4 t 1,2% S
level from which date
was obtained
Zone containing mkfden
bo<>od< mabix
S·SHd B -Bone
C'C~I H , Hunate

Figure 6. Plan and Profile of Feature I at 41 BL233 (AbbatE and Trierweiler 1995),

4-12
"- 1&.410
rn.ad
ilGC8SS

"-
"-
"-
"-
\
-~
~.....-.
Feature 5 Plan Map

BHT Backhoe Trench


I
N
I
ST / TI' Test p< (1xlm)
/ ST Shovel Test J
I
~
/ Treeline a 2 meteo
/

Backhoe Trench Profile


cmbs

--
meters

Zooo , middeomalrix: storry clay loam; ~sl.baogutl!ll bIodI.y Sb'\..JctU'9 9fiI'd.no:;J lI'ItomaS$Ne; clast-supported buT'.ed l,mestone rnodoenrIe 9fll.SS roots inwater
20 em; law 'oIrOOCy I1XlU ~; fitI.e to mednxn limu!OI'le clasts· _ailly mbnc.aled - except on c.omlll', wr«e $Iao$ up to -35 em diame1&r at8
sli'tlnQty ~bric3ted. slopinQ 10 S:SW:SE at IS.· 25>, Yf!!ry daf1c. ~sh tItOw!'Il0YR 312}

stony clay loam; ma.uive: hIt'd ay: sUdl.y wet aI:Iu"Idant medUn to l2Ige limestone coObIes: slMls (Il'10$1 ~): vf!!ry da1I. 9'2)'ish brown {tOVR
Zone "
312).ry: v«y daf1t tItOw!'I (10YR 212) _t: common fll"\e.oot1y roou: some lan;>e honzontal slabs at hmestone' sl.lCstrate: h¥d Ifmestone

cmbs East Walt TP1

Zone containing midden


matItx"
Limestone rock
Level from which date
t850 t: S(I C was obtained
21340 t: 70 C
630t:70C //h Unexcavated
117 t: 0,9'10 S
H -Hl$Tli!le
2880 t 70 C C -Chan;oal
60 - 1759 t 49 H
S-Seed

Figure 7. Plan and Profile of Feature 5 at 41BL233 (Abbott and Trierweiler 1995).

4-lJ
Site 41BL495 is a rockshelter that was tested by the 1991 TAMU Archaeological Field School (Carlson
1993b). The shelter faces southeast and is about 24 m long (Figure 8). At its deepest it is 6 m, but this is •
deceptive since the deepest ponion is a narrow hole in the limestone that one would have to crawl into. Seven
different test units were placed into the shelter and excavated to depths of 40 to 60 cm before reaching bedrock.
Debitage counts ranged from 1 to 361 flakes with the highest density occuring in the top 20 cm. The shelter was
occupied during the Late Prehistoric period as indicated by 8 Sea/lorn and 7 Perdiz points recovered from the
excavations. Fauna recovered from the shelter included a variety of freshwater mussels, deer, rabbits, raccoon
and other taxa.

\
\
\
'. '-.~.

-.

Figure 8. Site plan for rockshelter at 41BL495 (Carlson 1993b).

4·14
Site 41BL497 is a rockshelter that was tested by the 1991 TAMU Archaeological Field School (Carlson
1993b). The shelter is located high on the valley wall of Spicewood Creek near the top of the escarpment (Figure
9). It is very small and could have provided shelter for only three or four people. Four test units were excavated
to obtain information on the age and integrity of the cultural deposits. The occupation seems to center around
Excavations Units 1 and 2 and is buried by 10-30 cm of relatively sterile fill. Seven Seal/om points were
recovered from these two units. Debitage counts ranged from I to 235 flakes per level with peak densities
occuring at 20-50 cm in Unit I and 30-60 cm in Unit 2. Faunal remains included freshwater mussels, deer,
opossum and rabbit. Radiocarbon age estimates from Units I and 2 are statistically homogeneous and give a
pooled age estimate of 1224 ± 74 BP. A single human fibula shaft fragment was also recovered from the shelter
which may indicate burials that were not encountered by the test excavations.

1 ,_ "0" 10.

I 10 hOO_ 0', c •• (

41BL497
,.

Figure 9. Site plan for rockshelter at 41BlA97 (Carlson 1993b).

4-15
REFERENCES
CULTURAL RESOURCES AT FORT HOOD, TX

Abbott, James T. and W. Nicholas Trierweiler, editors


1995 NRHP Significance Testing of 57 Prehistoric Archaeological Sites on Fon Hood, Texas. United
States Army, Fort Hood Archaeological Resource Management Series, Research Report Number
34.
Anonymous
1893 Memorial and Biographical History of McLennan, Falls, Bell, and Coryell Counties, Texas.
Lewis Publishing Co., Chicago.
Briuer, Frederick L. and George B. Thomas, editors
1986 Standard Operating Procedure for Field Surveys. Revision No. 11. United States Army, Fort
Hood Archaeological Resource Management Series, Research Report Number 13.
Carlson, David L., editor
1993a Archaeological Investigations in Bull Branch: Results of the 1990 Summer Archaeological Field
School. United States Army, Fort Hood Archaeological Resource Management Series, Research
Report Number 19.
1993b Archaeological Investigations in Spicewood Creek: Results of the 1991 Summer Archaeological
Field School. United States Army, Fort Hood Archaeological Resource Management Series,
Research Report Number 22.
1993c Archaeological Site Tesling and Evaluation on the Henson Mountain Helicopter Range AWSS
Projeci Area, Fon Hood, Texas. United States Army, Fort Hood Archaeological Resource
Management Series, Research Report Number 26.
1996 Archaeological1nvesligations Along Owl Creek: Results of Ihe 1992 Summer Archaeological
Field School. United States Army, Fort Hood Archaeological Resource Management Series,
Research Report Number 29.
Carlson, David L. and Frederick L. Briuer
1986 Analysis of Mililary Training 1mpacls on Prolectd Archaeological Siles al Wesl Fon Hood,
Texas. United States Army, Fort Hood Archaeological Resource Management Series, Research
Report Number 9.
Carlson, David L., Frederick L. Briuer, and Henry Bruno
1983 Selecting a Slatistically Represenlative Sample of Archaeological Siles al WeSI Fon Hood, Texas.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 8.
Carlson, David L., Shawn Bonath Carlson, Frederick L. Briuer, Erwin Roemer, Jr., and William E. Moore
1986 Archaeological Survey al Fon Hood, Texas, Fiscal Year 1983, The Easlem Training Area.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number II.
Carlson, David L., John E. Dockall, and Ben Olive
1994 Archaeological Survey at Fon Hood, Texas: Fiscal Year 1990: The Nonheaslem Perimeler
Survey. United States Army, Fort Hood Archaeological Resource Management Series, Research
Report Number 24.
Carlson, Shawn Bonath
1984 Ethnoarchaeological Studies al a 20ih Century Farmslead in Central Texas: The W. Jarvis
Henderson Sile (41BL273). Fort Hood Archaeological Resource Management Series. Research
Report No. 12.
1986 A Preliminary Assessment of Environmental and Cultural Determinants of Settlement in Central
Texas During the Nineteenth Century. Bulletin oflhe Texas Archeological Sociely 57: 123-142.
1990 The Persistence of Nineteenth Century Lifeways in Central Texas. HislOrical Archaeology 24:
50-59.
Carlson, Shawn Bonath, H. Blaine Ensor, David L. Carlson, Elizabeth A. Miller, and Diane E. Young
1987 Archaeological Survey at Fon Hood, Texas, Fiscal Year 1984. United States Army, Fort Hood
Archaeological Resource Management Series, Research Report Number 14.
Carlson, Shawn Bonath, David L. Carlson, H. Blaine Ensor, Elizabeth A. Miller, and Diane E. Young

4-16
1988 Archaeological Survey at Fon Hood, Teras, Fiscal Year 1985, The Nonhwestern Training Area.
Archaeological Resource Management Series, Research Report Number 15.
Dibble, David S. and Frederick L. Briuer
1985 Archaeological Survey, Fiscal Year 1980 (Spring). United States Army, Fort Hood
Archaeological Resource Management Series, Research Report Number 3.
1983 Archaeological Survey, Fiscal Year 1980 (Fall). United States Army, Fort Hood Archaeological
Resource Management Series, Research Report Number 3.
Dickens, William A.
1995 Identification and Prehistoric Exploitation of Chert from Fort Hood, Bell and Coryell Counties,
Texas. Unpublished Master's Thesis, Department of Anthropology, Texas A&M University,
College Station.
Ellis, G. Lain, Christopher Lintz, W. Nicholas Trierweiler, and Jack M. Jackson
1994 Significance Standards for Prehistoric Culrural Resources: A Case Study From Fon Hood,
Texas. USACERL Technical Report CRC-94/04.
Ensor, H. Blaine
1991 Archaeological Survey at Fon Hood, Texas, Fiscal Year 1987, The MCA Range Construction,
Pidcoke Land Exchange, and Phantom Range Projects. United States Army, Fort Hood
Archaeological Resource Management Series, Research Report No. 23.
Frederick, Charles D., Michael D. Glascock, Hector Neff, Christopher M. Stevenson
1994 Evaluation of Chen Patination as a Dating Technique: A Case Study from Fon Hood, Texas.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 32.
Goran, William D., William E. Dvorak, Lloyd Van Warren, and Ronald D. Webster
1983 Fon Hood Geographic Information System: Pilot System Development and User Instructions.
Construction Engineering Research Laboratory. Technical Report N-154.
Jack, Jackson
1990 Building an Historic Settlement Database in GlS. In Interpreting Space: GIS and Archaeology,
edited by Kathleen M. S. Allen, Stanton W. Green, and Ezra B. W. Zubrow, pp 274-283.
Taylor and Francis, London.
Jackson, Jack M. and Frederick L. Briuer
1989 Historical Research and Remote Sensing: Applications for Archaeological Resource
Management at Fon Hood, Texas, Fiscal Year 1981. United States Army, Fort Hood
Archaeological Resource Management Series, Research Report Numbers 5, 6, and 7.
Koch, Joan K., and C. S. Mueller-Wille
1988 Archaeological Survey at Fort Hood, Texas, Fiscal Year 1985, The Nonhwestern Perimeter.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 16.
1989 Archaeological Survey at Fon Hood, Texas, Fiscal Year 1985, The Southwestern Training Area.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 17.
1989 Archaeological Survey at Fon Hood, Texas, Fiscal Year 1986, The Nonhern Training Area.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 18.
Korgel, Randy
1990 National Register Eligibility Assessment of 41CV514, Fon Hood, Coryell County, Texas. United
States Army, Fort Hood Historic Preservation Technical Compliance Report Number 2.
Moore, William E. and George B. Thomas
1987 Results of a Monitoring Project of 112 Stratified Random Sample Quadrats at Fon Hood, Texas.
Report submitted by Texas A&M University to the U.S. Army, Fort Hood.
Mueller-Wille, C.S.
1988 Archaeological Monitoring at Fon Hood, Texas, Fiscal Year 1986. Report submitted by Texas
A&M University to the U. S. Army, Fort Hood.
Mueller-Wille, C.S. and David L. Carlson

4-17
1990a Archeological Survey at Fon Hood, Texas, Fiscal Year 1986, The Shoal Creek Watershed.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 20.
1990b Archeological Survey at Fon Hood, Texas, Fiscal Year 1986, Other Training Areas. United
States Army, Fort Hood Archaeological Resource Management Series, Research Report Number
21.
Newcomb, W. W., Jr.
1961 The Indians of Texas. University of Texas at Austin.
Nordt, Lee
1992 Archaeological Geology of the Fon Hood Military Reservation, Fon Hood, Texas. United States
Army, Fort Hood Archaeological Resource Management Series, Research Report Number 25.
1993 Additional GeoarchaeologicalInvestigations at the Fon Hood Military Reservation, Fon Hood,
Texas. United States Army, Fort Hood Archaeological Resource Management Series, Research
Report Number 28.
Prewitt, Elton R.
1981 Cultural Chronology in Central Texas. Bulletin of the Texas Archeological Society 52:65-90.
1983 From Circleville to Toyah: Comments on Central Texas Chronology. Bulletin of the Texas
Archeological Society 54:201-238.
Quigg, Michael J., Charles D. Frederick, and Dorothy Lippert
1996 Archaeology and Native American Religion at the Leon River Medicine Wheel. United States
Army, Fort Hood Archaeological Resource Management Series, Research Report Number 33.
Roemer, Erwin, Jr., Shawn Bonath Carlson, David L. Carlson, and Frederick L. Briuer
1985 Archaeological Survey at Fort Hood, Texas, Fiscal Year 1982, The Range Construction Projects.
United States Army, Fort Hood Archaeological Resource Management Series, Research Report
Number 10.
Scott, Zelma
1965 A History of Coryell County, Texas. Texas State Historical Association, Austin.
Sedlak, Michael A. And Sammy Brown
1992 Manuever Activity Damage Assessment Model: Applications to Predicting the Effects of the
Restationing of the 5th Infantry Division (Mech) to Fort Hood, Texas. Report submitted by the
Texas A&M University Research Foundation to the U. S. Army Installation Fort Hood, Texas.
Skinner, S. AIan, F. L. Briuer, W. C. Meiszner, and I. Show
1984 Archaeological Survey at Fon Hood, Texas, Fiscal Year 1979. United States Army, Fort Hood
Archaeological Resource Management Series, Research Report Number 2.
Skinner, S. AIan, F. L. Briuer, G. B. Thomas, and I. Show
1981 Initial Archaeological Survey at Fon Hood, Texas, Fiscal Year 1978. United States Army, Fort
Hood Archaeological Resource Management Series, Research Report Number I.
Thoms, Alston V., editor
1993 Archaeological Survey at Fon Hood, Texas: Fiscal Years 1991 and 1992: The Cantonment and
Belton Lake Periphery Areas. United States Army, Fort Hood Archaeological Resource
Management Series, Research Report Number 1.
Turpin, Jeff
1983 Archaeological Site Monitoring at West Fort Hood: Final Report. Report submitted by Science
Applications, Inc. to the United States Army, Fort Hood, Texas.
Trierweiler, W. Nicholas, editor
1994 Archaeological Investigations on 571 Prehistoric Sites at Fon Hood, Bell and Coryell Counties,
Texas. United States Army, Fort Hood Archaeological Resource Management Series, Research
Report Number 31.
Tyler, George W.
1936 The History of Bell County. Edited by Charles W. Ramsdell. The Naylor Company, San
Antonio.
Turner, Ellen Sue and Thomas R. Hester
1993 A Field Guide to Stone Anifacts of Texas Indians. Second Edition. Gulf Publishing, Houston.

4-18
U. S. Army Engineer District, Fon Worth
1992 Final Environmentallrnpact Statement. Base Realignment and Closure: Realignment of the 5th
Infantry Division (Mechanized) from Fort Polk, Louisiana to Fort Hood, Texas. Report
submitted by the U.S. Army Corps of Engineers, Fon Worth District to Depanment of the
Army.
Weir, Frank
1976 The Central Texas Archaic. Unpublished Ph.D. Dissertation, Washington State University,
Pullman.
Williams, Ishmael, W. Frederick Limp, and Frederick L. Briuer
1990 Using Geographic Information Systems and Exploratory Data Arta1ysis for Archaeological Site
Classification and Analysis. In Interpreting Space: GIS and Archaeology, edited by Kathleen M.
S. Allen, Stanton W. Green, and Ezra B. W. Zubrow, pp 239·273. Taylor and Francis,
London.

4·19
Appendix
Pictoral Glossary of Fon Hood Projectile Point Types Mentioned in the Text

Uvalde (Early Archaic, Abbott 1994) Marslwll (Late Archaic, Carlson 1993a)

Castroville (Late Archaic, Carlson, Dockall, Olive Ellis (Late-Terminal Archaic, Turner and Hester
1994) 1993)

Marcos (Late Archaic, Carlson, Dockall, Olive Godley (Late-Terminal Archaic, Abbott and
1994) Trierweiler 1995)

4-20
Darl (TerminallTransitional Archaic, Abhott and Bonham (Late Prehistoric, Carlson 1993a)
Trierweiler 1995)

Edgewood (Terminal/Transitional Archaic, Abbott Perdiz (Late Prehistoric, Carlson 1993a)


and Trierweiler 1995)

Seal/om (Late Prehistoric, Carlson 1993a)

Ensor (Terminal/Transitional Archaic, Abbott and


Trierweiler 1995)

4-21

You might also like