You are on page 1of 14

Geomorphology 136 (2012) 4558

Contents lists available at SciVerse ScienceDirect

Geomorphology
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / g e o m o r p h

Structural and morphometric irregularities of eroded Pliocene scoria cones at the


BakonyBalaton Highland Volcanic Field, Hungary
Gbor Kereszturi a, b, c,, Kroly Nmeth a
a
b
c

Volcanic Risk Solutions, Institute of Natural Resources, Massey University, PO Box 11 222, Palmerston North, New Zealand
Department of Geology and Mineral Deposits, University of Miskolc, Miskolc-Egyetemvros, H-3515, Hungary
Geological Institute of Hungary, Stefnia t 14, H-1143, Budapest, Hungary

a r t i c l e

i n f o

Article history:
Received 7 January 2010
Received in revised form 8 February 2011
Accepted 4 August 2011
Available online 16 August 2011
Keywords:
Morphometric age
Lava-spatter
Scoria cone
Slope angle
Digital Elevation Model
Slope decrease rate

a b s t r a c t
Scoria cones of the Mio-Pliocene BakonyBalaton Highland Volcanic Field (BBHVF) are built up by wide range
of volcanic rocks, including intercalated lava ows/dykes, pyroclastic breccias and scoriaceous lapilli with
various degrees of welding or agglutination. According to KAr and ArAr dating, ages of the fourteen scoria
cones within the eld span between 5.2 and 2.5 Ma. From these fourteen, seven cones were selected which
are suitable for morphometric analysis, i.e. visible in the eld and have identiable boundaries. The
morphometric data were either derived by manual measurement on topographic maps and by Digital
Elevation Model-based calculations. Using the same input contour line data from 1:10,000 maps, basic cone
parameters such as cone height, basal and crater width have been measured in order to calculate parameters
like Hco/Wco ratio and average slope angle. The results of these three-parameter-based manual calculations
have been compared to the DEM-based results in order to highlight the controls of degradation, pitfalls in
morphometric parameterization and the differences between these two calculation methods. Based on the
results, three main controlling conditions have been identied that are together responsible for the
preservation and erosion of the scoria cones located in the BBHVF: (1) age of the cone, (2) climate during the
degradation and the (3) inner architecture of the edice. In terms of morphometric dating, the traditional,
three-parameter-based method tends to give inaccurate results on (1) scattered and/or truncated cones and
(2) on the edices characterised by highly effusive behaviour during the emplacement.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Scoria (or cinder) cones are the most common volcanic landforms on
Earth (Wood, 1980a,b) as parasitic cones on larger polygenetic
volcanoes or as volcanic elds (Settle, 1979). Eruption styles associated
with scoria cone formation range from periodic magmatic fragmentation to minor intermittent phreatomagmatic phases (Martin and
Nmeth, 2006). Given dominantly a basaltic magma composition, the
majority of the eruptions are generally governed by the speed of the
rising magma, which is primarily determined by viscosity and gas
content and can produce predominantly Strombolian-style eruptions
(McGetchin et al., 1974; Partt and Wilson, 1995). However, typical
Strombolian-style eruptions at most of scoria cones are completed by
additional Hawaiian lava fountaining (Di Traglia et al., 2009), violent
Strombolian (Pioli et al., 2008) or phreatomagmatic eruptions (Gisbert
et al., 2009).
Previous work revealed that the main morphometric parameters of
scoria cones, such as cone height and average cone slope decrease during
Corresponding author at: Volcanic Risk Solutions, Institute of Natural Resources,
Massey University, PO Box 11 222, Palmerston North, New Zealand.
E-mail address: kereszturi_g@yahoo.com (G. Kereszturi).
0169-555X/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.geomorph.2011.08.005

degradation (Colton, 1967; Porter, 1972; Settle, 1979). Other morphometric parameters have been developed (e.g. cone/crater elongation,
breaching azimuth etc.) to characterise cone erosion (Dohrenwend et al.,
1986; Hooper and Sheridan, 1998), and tectonic setting (Corazzato and
Tibaldi, 2006). Morphometric characterization also aims at providing
basic information about the relative age of scoria cones (e.g. Sucipta et al.,
2006). However, as a consequence of the aforementioned scoria cone
diversity as well as of post-eruptive tectonic inuence and erosion, in
many cases morphometric age estimation is particularly difcult on older
(N1 Ma) scoria cones.
Our knowledge about the precision of morphometric dating is also
limited because most of the morphometric studies have focused on
relatively young (b1 Ma) scoria cones. Scoria cone dating based on their
morphometry benets from standard morphometric characteristics
such as ratio of cone height/basal width (Hco/Wco) and slope angle (Save,
Fig. 1). Both characteristics decrease with time (Dohrenwend et al.,
1986; Hooper and Sheridan, 1998). Since they are determined from
cone height, basal and crater width (if any), these three parameters are
essential to get age-representative values for the calculations.
According to Favalli et al. (2009), the Hco values of ank cones of
Mt. Etna (Italy) are largely controlled by the inclination of the substrate
as well as the lava burial effect, which reduce the cone height. Favalli

46

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

past decade (Martin and Nmeth, 2004; Auer et al., 2007). Of these
fourteen scoria cones, this study deals with the morphometry of seven
selected locations (Fig. 2), which have identiable/visible geologically
well-dened boundaries and available, reliable KAr and/or ArAr ages.
Each scoria cone of the BBHVF has a unique eruption history recorded in
their primary pyroclastic successions such as intercalated welded lava
spatters (Figs. 3AB), interbedded coherent lava units (Fig. 3C), and
various types of scoriaceous lapilli-dominated cone-building pyroclastic
successions (Fig. 3DF). These types of deposits are typical of
Strombolian-style explosive eruptions, Hawaiian-style lava fountaining
and lava effusion (Martin and Nmeth, 2004). The majority of these
processes have been inferred to take place during the cone-building
eruptive phases, except for post-eruption mass-movement, and are
together responsible for the complex cone-building events. However, the
relative roles of these processes are not yet fully understood in terms of
morphometry. A further aim of this study is to demonstrate the possible
pitfalls in automatic application of formulae-based morphometrical
dating generally used for younger, b1 Ma, scoria cone elds to establish
the relative morphological age of the cones.
2. Materials and methods
Fig. 1. Denitions of traditional morphometric parameters of a scoria cone.

et al. (2009) also suggest that a slightly decreasing trend of Hco/Wco ratio
with time can be observed among the ank cones in spite of the
relatively young age of the examined cones (6500 years BP).
Another morphometric parameter for scoria cone dating methods is
the slope angle, which has been partly studied by Favalli et al. (2009). It
raises some unsolved questions: how slope angles vary with different
methods of calculation, e.g. formulae-based or Digital Elevation Modelderived? How the eruption/erosion diversity of the cones is reected in
the slope angle? What are the precision and limitations of slope anglebased dating?
Here, we explore the complexity of scoria cone morphology
preserved in eroded, Pliocene scoria cones of western Hungary with
an aim of understanding internal and external factors that may have
been involved in the formations and preservation of the morphological
features of scoria cones. Scoria cones of the BakonyBalaton Highland
Volcanic Field (BBHVF) are older, 3.82.5 Ma (Balogh et al., 1986;
Wijbrans et al., 2007), than the cones which have been analysed by
morphometrically elsewhere (Wood, 1980b). At least ten scoria cone
remnants, with an addition of two deeply eroded cones (Tihany and
Kab-hegy) and a further two covered by thick Quaternary sediments
have been recognised at BBHVF through detailed investigations over the

The input data for both the manual and GIS-based parameterization were the Hungarian Military maps with scale of 1:10,000 with
5 m contour intervals. After georeferencing, Hungarian National Grid
(EOTR) projection and Hungarian Datum (1972), the contour lines of
these topographic maps were digitalized. The boundaries of lava
elds/ow (if any) and the scoria cones derived from 1:50,000
geological map of the BBHVF (Budai et al., 1999) updated by new eld
observations.
2.1. Manual parameterization
Hco (Fig. 1) is expressed as the arithmetic mean of the difference of
the basal height and maximum (Hco max) and minimum (Hco min)
elevation of the cone measured on topographic maps (Porter, 1972;
Settle, 1979). This parameter only gives valuable results in the case of
scoria cones located on gentle slopes (2.55), platform-type volcanic
elds (Favalli et al., 2009), and the BBHVF meets this requirement. Prior
to the volcanism at the BBHVF, the intra-Carpathian basins (especially
the later Pannonian Basin) were lled by a large volume of siliciclastic
sediments of Lake Pannon and associated river systems during the Late
Miocene through the Pliocene (Magyar et al., 1999) and therefore a
morphologically at landscape was in place before the volcanism.

Fig. 2. Digital Elevation Model (DEM) of the central part of the BBHVF with the studied locations.

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

47

Fig. 3. Textural diversity of pyroclastic successions of scoria cones of BBHVF. (A) Agr-tet: lava spatter-rich blocks around the scoria cone remnant; (B) Kopasz-hegy: alternating
erosion-resistant, lava ow units (LR2) and intercalated tuff breccias (TB1); (C) Bondor: outcrop at the breaching site (TB1weakly agglutinated, scoriaceous breccias; LR1dyke
unit); (D) Kopcsi-hegy: ballistically transported, weakly agglutinated scoriaceous lapilli-dominated unit; (E) Badacsony: contact zone of the maar lake inlling lava lake (LR1) and
the capping (TB1) scoriaceous pyroclastic breccias; white arrows represent small-scale tumuli structures; (F) Boncsos-tet: alternating highly (TB1) and weakly (TB2) agglutinated,
scoriaceous pyroclastic breccia units.

Wco is the average of the maximum and minimum diameter of a cone


(Fig. 1). Based on the pioneering works of Porter (1972) and Wood
(1980a) the worldwide median value of Wco is ~800 m, whereas the
mean value is ~900 m (based on data measured from 910 scoria cones).
The average terrestrial and still intact scoria cone size is characterised by
the following equations between the cone height and width (Porter,
1972; Wood, 1980a):
Hco = 0:18Wco

Wcr = 0:40Wco

Crater diameter or width (Wcr) is calculated from the average of


the minimum and maximum diameter of the crater measured from
topographic maps.
From the parameters above, further age-representing parameters
such as Hco/Wco ratio or average slope angle can be calculated. Both
parameters decrease gradually with time, thus, correlate well with the
age of the volcanic edice (Porter, 1972; Dohrenwend et al., 1986;
Hooper, 1995; Hooper and Sheridan, 1998; Sucipta et al., 2006; Doniz
et al., 2008; Favalli et al., 2009).
The average slope angle (Save) derived from the three basic
parameters namely the Hco, Wco and Wcr. Formally written as (Hasenaka
and Carmichael, 1985):
1

Save = tan

2Hco = Wco Wcr 

Save = tan

2Hco = Wco 

where the Eq. (4) is for those scoria cones with lack of measurable
crater.
2.2. DEM-based parameterization
As parallel with the manual parameterization, DEM-based slope
angle calculations were performed for the study areas. The DEMs were
obtained by linear interpolation method (Gorte and Koolhoven, 1990).
This method rasterizes the input contour lines into the user-dened
horizontal grid cell size (Gorte and Koolhoven, 1990), which is
commonly referred to horizontal resolution. Extremely high resolution
DEMs therefore can be created from relatively low-scale topographic
input data. However, the proper DEM resolution is dependent on the
nature of the input data, i.e. contour lines or spot heights,and the
properties of the terrain modelled (Hengl, 2006; Jordan, 2007).Thus, the
resolution of DEMs was determined on the basis of input data properties
such as distance between neighbouring contour lines (Hengl, 2006;
Hengl and Evans, 2008). To nd the proper horizontal grid cell size,
neighbourhood operator was used to nd those rasterized contour line
pixels that lie immediately adjacent to each other on a 3 3 pixel matrix.
This special pixel, called touching pixels, may be a source of error. These
measurements showed that a 2 2 m grid cell size was small enough to
avoid touching pixels in the area of interest (outer anks). In order to

48

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

get a more generalised picture for slope angles, our nal DEMs
were smoothed by an average 3 3, i.e., 6 6 m, moving window. The
elevation values for each pixel have been calculated for 3 decimals.
An additional source of error in digital modelling is that at pixels
are mostly the result of an inadequate determination of horizontal
grid cell size (Garbrecht and Martz, 1997; Jordan, 2007). These pixels
have zero rst derivates, i.e. zero slope angles, thus can modify the
results of slope angle estimates. No at pixels have been found within
the area of interest. The overall accuracy of DEMs was characterised by
root mean square error, which is dened as (Fisher and Tate, 2006):

intends to approximate the average slope angle regardless the


small, local-scaled variance of the terrain modelled. In other words,
the linear, eight-point, unweighted (called Prewitt operator) lter
gives the best results because this lter (1) calculates a more
generalised slope estimates as compared to three- or four-point
methods (Jones, 1998), (2) shows high resistance for interpolation
errors (Raaaub and Collins, 2006) because of its high smoothing
effect due to the rst-order trend surface tted to the 3 3 pixel
matrix (Sharpnack and Akin, 1969). In this lter, the derivates dened
as:


2
RMSE = ZDEM Zref = n

f x = Z3 + Z6 + Z9Z1Z4Z7 = 6X

where the ZDEM is the pixel elevations of the DEM, the Zref is the
reference points (in this case spot heights from the topographic maps)
and the n is the number of reference points. The RMSE is always
positive. In terms of accuracy, the DEM studied in the present study
are characterised by 0.8 m up to 2.8 m.
The rst derivates, i.e. slope angle, of a DEM is formally written as:

f y = Z1 + Z2 + Z3Z7Z8Z9 = 6Y

SLOPE = arctanfx + fy

where the fx = z / x and fy = z / y, for bivariate function, z = f(x,y),


then the absolute value of the gradient vector. In a grid-based
environment, rst derivates can be calculated by various lters for
example three-, four- or eight-point methods (Sharpnack and
Akin, 1969; Zevenbergen and Thorne, 1987; Jones, 1998). In these
lters above, the numbers correspond to the number of pixels evolved
in the gradient vector calculation on a 3 3 pixel matrix. In the case of
morphometric parameterization for scoria cones, the user usually

where the Z1Z9 correspond to the pixel elevation reads from the top
left cornel to the bottom right position in a 33 pixel matrix. The X
and Y refer to the grid cell size along the two main directions.
In this study, four types of slope angle values such as Smean, Smed,
Smode and Smax were calculated only on the outer anks of the volcanic
cones. The Smean is the weighted average of the slope angles by the
total number of pixels related to a slope value, while the Smed, Smode
and Smax refer to the median, mode and maximum values of the slope
angles, respectively. The inner crater slopes, the breached side of the
cones, slopes dissected by large valleys as well as local at areas, e.g.
local maxima/peaks or local minima/depressions, due to interpolation
error were eliminated from the delimited areas (e.g. Figs. 4 and 5).
Nonetheless, the delimited cone slopes still contained 3686191,677
individual pixel values, which bring relevant information on the
present slope angle pattern (Table 1).

Fig. 4. Agr-tet: (A) The location of the scoria cone and the measured morphometric parameters such as maximum and minimum Wco as well as the measured outer slopes. (B) The
cross-section of the present cone and lava eld with the applied base height surface which were used in the volume calculations. (C) Slope angle histogram and cumulative (red
curve) slope angles of the outer cone's ank.

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

49

Fig. 5. Kopasz-hegy. For detailed information see Fig. 4.

Besides the slope angle estimates, the volume of the preserved


edices (Vcone) and related lava ow eld (Vlava) were calculated from
the high resolution DEMs by applying various dipping base heights
(see details in Figs. 410). The source of lava ows have been
established on the basis of (1) stratigraphical and topographical
relationship, (2) existence of the low, radial alignment of lava rock
covered ridges around the base of the cone, (3) existence of outcrops
as well as (4) present cone morphology.
3. Studied locations
In this section, seven scoria cones from the BBHVF have been
described, which have various geomorphic shapes and ages. We present
their locations, eruption ages, main volcanological and geomorpholog-

ical features as well as the main morphometric values, and address


whether or not they are suitable for traditional morphometric dating.
Furthermore, we also select which slopes represent the best the rough
age of the cone-formation (i.e. not disturbed by subsequent volcanological and/or erosional processes) and are therefore suitable for testing
(e.g. compare with the existing KAr and ArAr ages).
3.1. Agr-tet
Agr-tet Volcanic Complex is located at the northern, elevated part
of BBHVF (Figs. 2 and 4). The volcanic cone is surrounded by a lava
plateau erupted on Mesozoic carbonates (Budai et al., 1999). Based on
ArAr geochronology the scoria cone erupted about ~3.3 Ma ago and the
extended lava eld was emplaced around ~3.0 Ma ago (Wijbrans et al.,

Table 1
Morphometric parameters of the seven scoria cones from the BBHVF Estimated by both manual and DEM-based methods. Notes: 1Calculated by Eq. (3); 2Calculated by Eq. (4); 3The
weighted arithmetic mean (Smean), median (Smed), mode (Smode) and maximum (Smax) slope angle values. Values in bold were used to evaluate the cones' age based on their morphometry.
Cone

Measured from DEM

Calculated manually

Elevation Hco min Hco max Hco


a.s.l.

Wco min Wco max Wco

Wcr Dcr

620
699
916
529
710
665
786

250

650
350
400

In metre
Agr-tet
Kopasz-hegy
Bondor
Kopcsi-hegy
Badacsony
Boncsos-tet
Gajos-tet

511
303
378
303
437
448
373

56
8
40
20
17
52
20

81
22
60
43
29
67
22

68.5
15.0
50.0
31.5
23.0
59.5
21.0

788
732
1614
742
876
1207
1226

704.0
715.5
1265.0
635.5
793.0
936.0
1006.0

bre
bre
bre
15.0
bre

Calculated from DEM


1

2
Save

Hco/Wco Eco

Save

Ratio

In degree

0.097
0.021
0.040
0.050
0.029
0.064
0.021

0.787
0.955
0.568
0.713
0.811
0.551
0.641

16.9

9.2
12.4
6.6

11.0
2.4
4.5
5.6
3.3
7.2
2.3

3
3
3
3
Smean
Smed
Smode
Smax
Std. Dev. Vcone

in km3
14.5
11.9
10.8
11.7
8.0
9.2
4.0

13.3
10.8
10.8
11.0
7.1
8.7
3.8

12.7
10.1
9.4
8.3
8.3
5.4
1.7

41.9
36.4
35.9
35.6
40.9
36.1
13.9

6.0
4.3
3.5
5.7
4.1
3.5
2.1

0.01632
0.00065
0.03185
0.00592
0.00592
0.01277
0.00796

50

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

Fig. 6. Bondor. For detailed information see Fig. 4.

2007). The relatively long-lasting eruption (~0.3 Ma) indicates a stable


magma supply beneath Agr-tet (Martin and Nmeth, 2004) resulting
in the accumulation of 0.365 km3 lava (Table 2).
Explosive volcanic activity of Agr-tet is characterised by Strombolian- and Hawaiian-styles that produced scoriaceous fall beds interbedded with agglutinated lava spatter (Martin and Nmeth, 2004;
Csillag et al., 2008). The inner structure of the eroded cone consists of
strongly to moderately agglutinated, lava-spatter dominated deposits
(Fig. 3A).
The original crater of Agr-tet is still recognisable; however, a large
breach has occurred at the NW side of the cone (Fig. 4). The preserved
geomorphology of the cone is characterised by the existence of low,
radial ridges around the eastern foot of the Agr-tet, which can be
interpreted as the remnants of individual lava ows (Csillag et al., 2008).
The size of scoria cone of Agr-tet is slightly smaller than an
average scoria cone after Wood (1980a), with ~ 704 m (Wco) and
~ 68 m (Hco) and a ~ 250 m wide crater as well as small cone volume
about 0.016 km 3 (Table 1). The Save calculated by traditional method
between 16.9 and 11 (Table 1) and with the DEM-based method
~ 14.5 (Smean). In addition, the Hco/Wco ratio is around 0.097.
Morphometric assessment: Agr-tet and its lava eld are not a classic
example of a scoria cone because the present cone mostly comprises
erosion resistant, lava spatter-dominated deposits as a result of Hawaiian
and Strombolian-type eruptions. Furthermore, the long-lasting eruption
with multiple lava effusion stages, according to Wijbrans et al. (2007),
likely inuenced the freshly deposited scoriaceous lapilli and ash beds
causing signicant welding and agglutination. However, the capping
scoria cone edice is suitable for all methods of morphometry-based
dating because its shape is preserved, except for the crater opening from
NNW. Consequently, both the traditional (including Eqs. (3) and (4))

and DEM-based methods yield valid results, which are in the same range
(1214; Table 1).

3.2. Kopasz-hegy
Kopasz-hegy Volcanic Complex (KVC; Figs. 2 and 5) consists of two
individual volcanic centres in the western boundary of the Kl Basin
(Fig. 2). According to KAr dating, the age of Kopasz-hegy spans
between 2.82 and 2.59 Ma (Balogh, K. pers. comm.). Based on
distribution, type and bedding characteristics of pyroclastic deposits,
initial phreatomagmatic activity at Kopasz-hegy was within a NS
aligned paleovalley (Kereszturi and Nmeth, in press). The northern
eruption centre is capped by at least four preserved, lava spatter and
scoriaceous breccia-dominated (Fig. 3B) mound-shaped hills, which
are ~515 m high and few hundreds of metres in diameter today.
These mounds consist of layers (usually a few dm thick) of greyish to
blackish, densely to non-welded, scoriaceous pyroclastic breccias
(Fig. 3B). Rootless (clastogenic) lava ows 1015 m long with a
volume of a few tens of m 3 were emplaced during the late stage
evolution of Kopasz-hegy. These erosion resistant rocks must have
had control over the erosion of the cone.
The distribution of the preserved scoria cone anks is limited due
to the quarrying and the volcanic/erosional processes. The crater
width is doubtful and its location is only inferred between the small,
preserved scoria mounds (Fig. 5). The direction of the breaching is
south. This has an obvious effect on morphometry; for example the
Wco is 715 m and the Hco is only 15 m and the volume is relatively low
0.0006 km 3 (Table 1). This results in a very low Hco/Wco ratio 0.021.
The slope angles strongly vary between 2.4 and 11.9 (Table 1).

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

51

Fig. 7. Kopcsi-hegy. For detailed information see Fig. 4.

Morphometric assessment: The present geomorphic shape, in spite of


its young age, indicates that the Kopasz-hegy was probably truncated by
volcanic (e.g. vent migration) or erosional processes, e.g. gully erosion.
The existence of a double cone system with a phreatomagmatic tuff ring
base, one partly destroying the other and the central crater, is supported
by the contrasting KAr ages measured (Balogh, K. pers. comm.). Thus,
the cone is not suitable for traditional, morphometric dating because its
crater has almost been removed: only the Save is calculated by the Eq. (4)
(Table 1). However, the DEM-based method for Smean gave a reasonable
value of slope angle (~11.8; Table 1).
3.3. Bondor
Bondor is a remnant of a complex volcano situated at the northern
part of the BBHVF (Fig. 2). Several KAr radiometric datings from the
coherent lava body of Bondor suggest an age range of 3.8 and 2.29 Ma
(Balogh et al., 1986; Balogh and Pcskay, 2001). The wide range of KAr
ages may result from long-lasting and intermittent volcanic activity
with individual eruption stages (Kereszturi et al., 2010). Based on the
depositional setting, the age of the upper scoria cone is inferred to be
between 2.9 and 2.29 Ma by KAr dating methods (Balogh et al., 1986;
Balogh and Pcskay, 2001; Kereszturi et al., 2010). This age interval
has been conrmed with the last unpublished dating of 2.52 Ma
(Balogh, K. pers. comm.).
Initial stages of the evolution of Bondor comprise a basal
phreatomagmatic tuff ring unit, which is ~ 1.5 km in diameter. This
basal tuff ring hosts a few lava ow units and the capping scoria cone
(Martin and Nmeth, 2004). A few outcrops at the eastern sector
expose the inner structure of the Bondor scoria cone. The scoria cone
is characterised by chaotically bedded, reddish, moderately vesiculated, partly welded scoriaceous lapilli and block layers and crosscutting dykes (Fig. 3C). The effusive activity, which is directly linked to

the scoria cone, produced lava ows that are only preserved at the
eastern part of Bondor (Fig. 6). Small topographic ridges at the SE
foot of the cone indicate that the lava ow came from the scoria cone
(Kereszturi et al., 2010). The rest of the lava plateau may be related to
other effusive activity during the evolution of Bondor, probably after
the initial maar-forming eruptions.
The Bondor is one of the largest scoria cones within the BBHVF in
Wco (~1265 m), Wcr (~650 m) and volume (V = 0.031 km 3; Table 1).
The cone height is small (~50 m). The Hco/Wco ratio is 0.04 due to the
large basal diameter of the cone (Table 1). The slope angle varies
between 4.5 (Save calculated by Eq. (4)) and 10.8 (Smean and Smed).
The Smode is 9.4.
Morphometric assessment: Bondor is one of the largest erosional
remnants of BBHVF built up by typical scoria cone deposits. Smallscale breaching occurs at the eastern slopes, but this morphological
irregularity does not signicantly disturb the morphometric parameterization process. As a result of this, the Bondor is suitable for both
dating methods (Fig. 6).
3.4. Kopcsi-hegy
The Kopcsi-hegy scoria cone (Figs. 2 and 7) has a well-preserved
cone with a still enclosed crater, located near the largest nested maarsystem of Fekete-hegy Maar Volcanic Complex in the central part of
BBHVF (Martin and Nmeth, 2004; Auer et al., 2007). The age of
Kopcsi-hegy is about 2.61 Ma (Wijbrans et al., 2007).
The basal part of the cone consists of massive lapilli tuff beds
interpreted as products of initial phreatomagmatic explosive eruptions (Nmeth and Martin, 1999). A NS elongated ridge can be
identied south of the present foot of the volcanic edice (Fig. 7),
which may have been formed by valley-lling deposits of the initial
phreatomagmatic eruption (Nmeth and Martin, 1999). The scoria

52

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

Fig. 8. Badacsony. For detailed information see Fig. 4.

cone grew over these basal phreatomagmatic rock units and is


characterised by chaotically bedded, reddish to brownish, partly to
weakly welded, highly vesicular lapilli and block-sized scoriaceous
deposits with abundant ballistically transported spindle bombs
hosting mantle-derived peridotite lherzolite nodules (Fig. 3D). No
lava ows are known to be associated with Kopcsi-hegy (Fig. 7).
The morphometry of the small Kopcsi-hegy scoria cone is simple
with Wco 635 m and Hco 31 m, as consequences of the partly closed
crater that has a diameter of 350 m and depth ~15 m (Table 1). The
cone volume is about 0.005 km 3 (Table 1). The lowest slope angle is
5.6 (Eq. (4)) and the highest is 12.4 (Eq. (3)). The Hco/Wco is around
0.05.
Morphometric assessment: The Kopcsi-hegy scoria cone, the only
cone of the BBHVF with an almost intact crater, was formed by
Strombolian activity following initial phreatomagmatic eruptions. No
late-stage lava ow is known to be derived from the Kopcsi-hegy
scoria cone. This cone is one of the most suitable for morphometric
parameterization both with the traditional and the DEM-based
methods. However, in the case of the latter method, we calculated
the slope angle only from the undisturbed slopes (Fig. 7). Here, Smean
gave an 11.7 relatively steep slope whereas formulae-based methods
gave 12.4 (Table 1).
3.5. Badacsony
Badacsony (Figs. 2 and 8) forms a butte in the southern part of the
Tapolca Basin (Fig. 2). The Badacsony volcano is about 3.45 Ma old on
the basis of KAr radiometric datings (Borsy et al., 1987). Badacsony is
an eroded tuff ring lled with thick capping lava and associated scoria
cone (Martin and Nmeth, 2004). The contact zone of the capping
scoria cone and the basal lava lake units are exposed due to quarrying

(Fig. 3E). The Badacsony volcanic complex consists of a present day


23 m high, intra-maar scoria cone with ~ 793 m basal width and
~400 m crater width, which correlates well with average world
examples (Wood, 1980a). The cone has been largely removed by the
erosional processes, thus the present volume of the cone is only
0.005 km 3 and the Hco/Wco is very low (0.029).
As a consequence of erosion, the crater of the scoria cone is not
clearly visible in the eld and it seems to have been breached from the
north. In addition, the original edice has been incised by a small gully
from the east. The present boundary of the scoria cone is slightly
undermined by 20th century quarrying. As a consequence of slope
angle pixels in rim position are higher due to the quarrying (Fig. 8).
Thus, we measured DEM-based slope angle parameter (Smean) only for
the S and SW slopes of the Badacsony (Fig. 8). The Save of the whole
volcanic edice is 6.6 and 3.3 calculated using the Eqs. 3 and 4 and a
slightly higher values (Smean = 8.0) is accounted for the southern
slopes of Badacsony (Fig. 8).
Morphometric assessment: The morphometry of the present scoria
cone has been modied by the quarrying as well as the marginal
situation of certain (e.g. eastern and western ones) slopes of the cone
(Fig. 8). This destruction highly disturbs the traditional, formulaebased results of slope angle. However, for the DEM-based slope angle
calculation, the southern slopes which are less eroded have been
found suitable (Fig. 8).
3.6. Boncsos-tet
Boncsos-tet (Figs. 2 and 9) is a large asymmetrical scoria cone
remnant in the central part of BBHVF and it lies at the margin of the
Fekete-hegy Maar Volcanic Complex (Auer et al., 2007). There are no
available KAr and ArAr radiometric ages directly from the cone, but

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

53

Fig. 9. Boncsos-tet. For detailed information see Fig. 4.

there is a KAr age date obtained (3.3 Ma) from the nearby lava
plateau of Fekete-hegy (Auer et al., 2007). In addition, morphometric
estimation and comparison to the nearby Gajos-tet scoria cone
conrmed partly a similar age range (Kereszturi, 2010).
The inner structure of Boncsos-tet scoria cone is exposed in a small,
abandoned quarry at the northern side of the remnant (Fig. 3F). In this
quarry, thickly bedded, reddish to greyish, alternating densely (TB1 in
Fig. 3F) and partly welded (TB2 in Fig. 3F) scoriaceous pyroclastic
breccias and lapilli tuffs with steeply dipping beds (3545) are in
agreement with its proximal location relative to the presumed original
crater (Csillag, 2004).
The position of Boncsos-tet (located on the rim of the lava
plateau of Fekete-hegy) favours the formation of large landslides due
to slope instability (Csillag, 2004). There is no morphological evidence
of any type of breaching because the location of the crater is also
doubtful.
The morphometry of Boncsos-tet is still youthful because Hco is
about 59 m and the Wco is 936 m. The volume is 0.012 km 3. The
average slope angle parameters vary from 7.2 up to 9.2 (Table 1).
Morphometric assessment: The present erosion remnant of Boncsostet is largely asymmetric due to degradation during the post-eruptive
period. As a result of landslides and the crater removal, Save is only
calculated by the Eq. (4) (Save = 7.2) due to the missing Wcr value.
Although only the eastern slopes of Boncsos-tet have been found
suitable for the slopes-angle calculation by DEM, it gives a similar result
Smean ~9.2 (Smean).
3.7. Gajos-tet
Gajos-tet (Figs. 2 and 10) is a poorly preserved, eroded scoria
cone from the Fekete-hegy Maar Volcanic Complex in the central part

of BBHVF (Fig. 2). The lava ow of Gajos-tet was penetrated by the


Kapolcs-1 drill hole (Jmbor, 1980; Kereszturi, 2009) and its age of
3.82 Ma is constrained by the upper lava ow unit (Balogh et al.,
1986).
The scoria cone and related lava ow are inferred to have lled a
previously formed maar crater and spilt over the crater rim (Auer et al.,
2007). The Gajos-tet scoria cone is made up of black to reddish-basaltic
scoriaceous pyroclastic rocks, but the exposure of the cone deposits is
very limited due to the vegetation cover and the gentle slope angle.
However, small, WE aligned lava ridges can be identied in the DEM
and conrmed in the eld, which are similar in size and location to the
ridges in the cases of Bondor and Agr-tet.
The advanced erosional stage is reected by the morphometry of
the volcanic cone (Fig. 10). Hco is 21 m, Wco is 1006 m, hence Hco/Wco
ratio is only 0.021 (Table 1). The preserved volume of the deeply
eroded scoria cone is as low as 0.007 km 3. The related lava ow
volume has also been calculated as Vlava = 0.017 km 3. Because the
crater is fully degraded, the Save calculation is only carried out by the
Eq. (4), giving a result of 2.3. At the same time, for DEM-based
calculation, we sampled the southern slopes, which are situated in a
sheltered position on the lava mesa of Fekete-hegy, and the result
shows a slightly higher slope angle value ~ 4.0 (Table 1).
Morphometric assessment: The Gajos-tet is a normal-sized scoria
cone, when compared with the world average of 800 m in width,
sitting on relatively undisturbed position upon the basalt mesa of
Fekete-hegy. The small-scale topographic ridges demonstrate that
effusive stage(s) led to the emplacement of a 1.5 km long lava ow
(Fig. 10). The preserved volcanic edice is suitable for the morphometric measurements (with the DEM-based method), but due to the
missing crater the calculation of Save values by the traditional,
formulae-based methods is impossible (Table 1).

54

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

Fig. 10. Gajos-tet. For detailed information see Fig. 4.

4. Discussion
4.1. Controls on scoria cone degradation in the BBHVF
According to early papers (Breed, 1964; Colton, 1967; Dohrenwend
et al., 1986), two parameters (Save and Hco/Wco ratio) are the most
suitable for the morphometry-based dating of monogenetic scoria
cones. However, neither slope angle nor Hco/Wco ratio decreases at a

constant rate, thus the morphometric modications do not show a


linear trend (Dohrenwend et al., 1986). Because, the slopes of young
cones are more susceptible to rapid erosion, whereas degradation
especially where vegetation developsslows down with time (Wood,
1980b). Thus, the precision of the relative dating also decreases with
increasing geological age. Consequently, whereas general rules are
applicable for young and fresh cones, the older cones may follow these
trends to a lesser extent having larger errors in relative age estimation.

Table 2
Morphometric age vs. radiometric age. Note: 1DEM-based mean slope angle values of the studied cones; 2Hco/Wco ratio calculated from the Table 1; 3Volumetric parameters of
cone related lava elds; 4Geological ages based on KAr and ArAr radiometric determinations (Balogh et al., 1986; Balogh and Pcskay, 2001; Auer et al., 2007; Wijbrans et al.,
2007; Kereszturi, 2010).
Cone

Agr-tet

1
Smean

2
Hco/Wco

3
Vlava

Effusion scale

Expected age based


on morphometry

Real geological age


based on KAr and
ArAr dating (Ma)4

degree

Ratio

km

14.7

0.097

0.365

Large-scale

Absolute youngest

3.33.0

0.0001
0.000
0.056
0.084
0.102
0.017

Small-scale
Non
Medium-scale
Medium-scale
Medium-scale
Medium-scale

Youngest

2.52.8
2.6
2.32.9
3.3 and (2.83?)
3.4
3.8

0.084
0.000
0.056
0.102
0.000
0.017

Medium-scale
Small-scale
Medium-scale
Medium-scale
Small-scale
Medium-scale

Youngest

Smean based morphometrical age estimate


Kopasz-hegy
11.9
Kopcsi-hegy
11.7
Bondor
10.8
Boncsos-tet
9.2
Badacsony
8.0
Gajos-tet
4.0
Hco/Wco based morphometrical age estimate
Boncsos-tet
Kopcsi-hegy
Bondor
Badacsony
Kopasz-hegy
Gajos-tet

0.064
0.050
0.040
0.029
0.021
0.021

Oldest

Oldest

3.3 and (3.02.8?)


2.6
2.92.3
3.4
2.52.8
3.8

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

Previously, two main factors have been identied that determine


the major changes on the shape and morphometry of the scoria cones:
age of the edice and the climatic setting (Wood, 1980b). Based on
morphometry and volcanological settings of the scoria cones of BBHVF,
the effects of these main factors on degradation can be recognised.
However, we also found a third factor, namely, the textural characteristics of the cones (e.g. erosion-resistant components), which has also
played an important role on slope processes and thus morphometry.
4.1.1. Age control on degradation
Previous attempts that were targeted to date scoria cone elds used
various parameters to approximate edice ages (Sucipta et al., 2006). In
order to detect the variations of these age indicator parameters among
older cones, we systematically examined the behaviour of Smean and
Hco/Wco ratio (Table 2). The results are varied because Hco/Wco ratios do
not show a clear trend (for the detailed explanation see later; Table 2).
Only the DEM-based Smean values show a systematic decreasing trend
with time similar to other studied scoria cones worldwide (Dohrenwend
et al., 1986; Favalli et al., 2009). However, there is one exception, Agrtet, which does not t to the overall decreasing trend, because of its
having an old age but the highest slope angle values (Table 2). Based on
this trend, the eruptive age control on morphometric parameters is
conrmed in the case of older scoria cones as well.
4.1.2. Climatic control on degradation
The progressively more gentle slopes, i.e. the decreasing slopes
values with time due to degradation were calculated from the
obtained Smean and Save parameters (Fig. 11). The rates show a wide
range because of the differences in morphometric methods. The slope
angle rates derived from the traditional, formulae-based methods
(Eqs. (3) and (4)) are 3.94/Ma and 0.79/Ma, respectively. A higher
rate (5.03/Ma) was calculated using the DEM-based Smean values
(Fig. 11). In the case of formulae-based method (Eq. (3)), the lower

55

value of Save degradation rates are the result of the limited number of
suitable cones (only three cones were suitable for parameterization
due to the lack of crater), whereas the DEM-based estimation is based
on six cones. The Agr-tet scoria cone, which evolved in a different
way (e.g. dominant effusive activity, different deposits preserved in
the scoria cone successions), was systematically discarded from the
slope angle rate calculations in order to get more representative slope
decrease rate for the classical scoria cone examples. From the above
mentioned rates, we accepted the result of DEM-based method
because these result are in agreement with KAr and ArAr ages (see
details later).
The calculated rates of slope decrease at BBHVF (~5/Ma) are
comparable with other published rates, for example in the Cima
volcanic eld, Mojave Desert, California, where the average rate of
slope angle decrease (Save) is around 6/Ma (Dohrenwend et al.,
1986). In addition, the San Francisco Volcanic Field (Arizona) has
shown similar slope decrease over the last 2 Ma period of time, for
instance the Woodhouse age cones; ~ 8/Ma (Wood, 1980b).
However, we have to keep in mind the fact that these cones are
mostly characterised by erosion resistant inner structure, such as
welded or agglutinated deposits, thus the calculated Smean decrease
rate is at least minimum estimate. Interestingly, both of the analogue
elds for BBHVF are characterised by mostly low annual precipitation
(100500 mm/year) and dry climates which favour a slow degradation and relatively good preservation of the volcanic edice (Wood,
1980b). This slow rate of slope degradation may indicate a similar
semi-arid climate in the Carpathian Basin during the Pleistocene,
which is in accordance to other climate studies (Fbin et al., 2000;
Kovcs et al., 2011; Sebe et al., 2011).
4.1.3. Architectural control on degradation
Internal settings, e.g. age of the edice, and external factors, e.g.
climate, both control scoria cone degradation as we have seen

Fig. 11. Diagram of slope angle values and degradation rates of scoria cones of BBHVF calculated by various methods including traditional, formulae-based (Eqs. (3) and (4)) and
newly developed DEM-based (Smean) methods. Note the poor line tting for the slope angles calculated by traditional techniques and the limitation of this method on older and/or
truncated cones.

56

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

previously. In the BBHVF, however, an additional third main


controlling parameter such as volcanic architectural diversity can be
recognised in the BBHVF that affect the cone morphometry.
In general, the architecture of the scoria cones has two endmember types based on their explosive and effusive activity. On
the one hand, most explosive cones are built up mostly by loose lapilli
and ash as well as minor, e.g. locally distributed, lava spatter,
favouring rapid erosion. Alternatively, the more ash-dominated
cones can be generated by violent Strombolian type eruptions (Martin
and Nmeth, 2006; Valentine and Gregg, 2008). On the other hand,
more effusive scoria cones are made up of the lava spatter-dominated
beds, which in turn are more resistant than the lapilli-dominated
ones. Both types above are usually formed by Strombolian-style
eruptions. Based on eld observations, the scoria cones of BBHVF are
closer to the last end-member, i.e. lava spatter-dominated cones.
In order to understand the role of effusive activity on degradation,
we categorised the seven studied scoria cones of BBHVF by the total
volume of the emitted lava ow/eld (Table 2). The groups are the
followings: (A) large-scale (0.1 km 3); (B) medium-scale (~0.1
0.01 km 3) and (C) non-effusive or small-scale effusive activity (~0
0.0001 km 3).
(A) Large-scale effusive activity means a large volume (0.1 km3) of
lava emplaced during the course of the eruption, forming an
extended lava eld, e.g. in the case of Agr-tet (Fig. 4).
Nevertheless, at Agr-tet, the large volume of lava may be
interpreted as the effect of possible additional ssure vent(s)
during the emplacement of the lava eld. However, the existence
of small ridges around the cone seems to conrm that Agr-tet
emitted at least a certain portion of the measured lava plateau.
The eruption history of Agr-tet is the only example for which no
evidence exists for phreatomagmatic activity, the entire evolution
seems to have been governed by magmatic processes and not
phreatic and/or phreatomagmatic fragmentation (Kereszturi
et al., 2011). The large-scale effusive activity of Agr-tet makes
the whole inner part of the volcanic edice more erosion resistant
(Fig. 3A). This resistance largely controls the morphometric
parameters as well as slopes of the present volcanic cone and
hence the result of morphometric dating and relative age of the
cone.
(B) Medium-scale effusive activity (Table 4) ranges between 0.1 and
0.01 km3 of emitted lava, for instance Bondor, Boncsos-tet,
Badacsony and Gajos-tet. These cones are characterised by mild
Strombolian eruptions. In individual cases (e.g. Bondor, Boncsostet), we documented weakly to moderately agglutinate or
welded pyroclastics; however, their distribution is limited and
generally associated with the central crater area and along crosscutting dykes. These types of architectural irregularity have not
disturbed the morphometric parameters to any great extent nor
the age estimates.
(C) Non-effusive or small-scale effusive activity means that the total
volume of the lava is relatively small or not evident, e.g. Kopcsihegy, Kopasz-hegy. Although they have been formed around
2.6 Ma, these cones have not shown as young morphometric
parameters as the Agr-tet. This anomaly can be explained by a
rapid degradation period. As a consequence, the minor effusive
activity is favourable to rapid erosion of the cones.
Based on this classication and the preserved Smean values, the
impact of this architectural diversity may correlate with the total
volume of the lava ow eld related to the monogenetic edice. In
other words, the larger effusive activity is favourable to a higher
degree of welding or agglutination leading the formation of an
erosion-resistant inner core of the edice. Due to erosion, these inner
irregularities may be exposed and are able to distort the morphometric parameters of the cones, like in the case of Agr-tet.

4.2. Morphometric dating of the Pliocene scoria cones of BBHVF


4.2.1. Pitfalls of traditional morphometric dating method on older scoria
cones
The traditional morphometric dating (based on either Save or Hco/Wco
ratio) has a wide range of pitfalls, especially for older scoria cones or for
complex cones that underwent various post-eruptive tectonic/erosive
processes. Scoria cones are obviously more diverse volcanic landforms
than previously considered, and it is likely that the classical loose coarse
ash and lapilli dominated edices rather represent a small fraction of the
full spectrum.
The eruption diversity, both internally (e.g. eruption style
variations, magma ux changes) and externally (e.g. gravitational
collapse and/or cone over-growth) strongly affect the initial geometry
of the cone on which the syn and subsequent erosion takes place
(Nmeth et al., 2011). Thus, understanding and identifying any sign of
these processes could play a key role in interpreting morphometric
data. The initial effects of eruption diversity can be enhanced over
long-term (millions of years) erosion history, and likely create
unexpected morphological scenarios of the eroded cones such as
youthful morphology on an otherwise old cone.
However, not only are the traditional methods, i.e. Hco/Wco or
Save, affected by syn- and post-eruptive processes. The evidence
for these modications can clearly be seen in the large diversity of
slope angle histogram and the standard deviation of slope angles
(Figs. 410). As a result of various eruptive mechanisms and the
distribution of pyroclastics preserved in the cone edice, the scoria
cones in the BBHVF can be subdivided into two groups based on
their shapes.
The Type 1 cones have either a scattered/truncated shape, e.g.
Boncsos-tet and Kopasz-hegy and/or a multi-peaked slope angle
histogram such as the Agr-tet. The standard deviation of slope angles
is generally large in this case 4 that Type 1 cones (Figs. 4 and 9). These
asymmetric cones resulted from syn- and post-eruptive modication
and truncation of the edice, e.g. landslide, vent migration, gully erosion.
These two cones show the largest morphometric irregularity and
disturbing effects on traditionally measured Save and Hco/Wco (Table 1
and 2). In these cases, the Save is only calculated using the Eq. (3) due to
the removed crater.
In the case of Agr-tet, the multi-peaked and slightly asymmetric slope histogram can be the result of physical properties on
preserved pyroclastics on the western and eastern anks of the
volcanic edice (Fig. 4). In terms of morphometry, the Agr-tet
scoria cone also differs signicantly from the other cones studied due
to its different eruption mechanism (magmatic processes without
phreatomagmatic eruptions), architecture (lava spatter-dominance)
as well as its large-scale lava effusive activity (lava eld volume
0.365 km 3). The absolute age of Agr-tet is 3.3 Ma (Wijbrans
et al., 2007) which is similar to Boncsos-tet, 3.33.0 Ma (Auer et al.,
2007), but the primary morphometric age-indicators, Save (14.7 vs.
9.2) and the Hco/Wco ratio (0.097 vs. 0.064), differ considerably.
Consequently, the aforementioned differences together may be
responsible for deviance of Agr-tet, i.e. the very youthful
appearance in spite of its old age.
The Type 2 cones are mostly symmetrical in shape that favour
better the results of parameterization both by tradition and DEMbased methods. The cones, e.g. Kopcsi-hegy, Badacsony, Bondor and
Gajos-tet, show generally compact, one peaked slope histograms and
relatively narrow, i.e. under 4, standard deviation of slope angles
(Figs. 6 and 8). The only exception is the Kopcsi-hegy, which is
characterised by three mounds and saddles between each other
(Fig. 7). Around these saddles, the slope angles are lower than on the
anks, leading to multi-peaked slope histograms and a wider range of
standard deviations, i.e. 5.7. In terms of morphometry, however, the
Kopcsi-hegy is one of the most suitable cones for morphometric
dating because of the intact shape with a relatively closed crater. A

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

possible explanation for the symmetrical shape may be the presence


of a homogenous distribution and properties of the cone-building
materials such as lapilli and breccias-dominated tuffs derived from a
series of Strombolian-type eruptions. Only locally do these cones
contain pyroclastic beds that are more resistance to erosion, i.e. lavaspatter beds, welded or agglutinated pyroclastics. The presence of
these erosion-resistant pyroclastic beds, however, has little apparent
inuence on the morphometric parameterizations of the cones
regardless the type of method that has been used.
To sum up, the possible pitfalls in the traditional, formulae-based
scoria cone age-estimation method are (1) highly scattered cones, e.g.
Type 1 cones: Kopasz-hegy and Boncsos-tet; (2) truncated cones and
(3) cones associated with signicant effusive activity, e.g. Agr-tet.
Any of these conditions may modify the results of relative cone dating
(Table 2).
4.2.2. Validity of DEM-based slope angle estimation and its implications
to older scoria cones
Our most signicant result is that the DEM-based morphometric
dating results has been found appropriate with regard to the KAr and
ArAr radiometric ages. In order to assess which of the morphometric
parameters (Smean and Hco/Wco) are more suitable to characterise the
age of the eroded cones of BBHVF, we carried out morphometric
estimation using both methods (Table 2). The Agr-tet was not
included in this test because of its different large-scale effusive activity,
which is suspected of causing the apparently young, morphometric age.
In the case of the Smean-based age estimation, the cones have
progressively, more gentle slopes in accordance with their age spectrum
from 2.4 to 3.8 Ma (Table 2). Effusive activity does not signicantly
distort the age obtained, because all the cones examined emitted
0.1 km 3 lava (medium-, small-scale or non-effusive activity).
At the same time, the Hco/Wco ratio shows a slightly distorted
trend. According to this ratio, the youngest cone is the Boncsos-tet
and one of the oldest is the Kopasz-hegy (Table 2) which signicantly
conicts with the KAr ages. In addition, these two examples from the
BBHVF have been truncated by various eruptive and/or erosionalrelated processes. In these cases, the DEM-based method provides a
more precise age estimate.
5. Conclusions
(1). Scoria cones are built up of wide range of eruption material,
such as lava rocks, densely to non-welded breccias and lapilli,
rarely ash. This textural and structural diversity of simple
scoria cones is common. At BBHVF, the present shapes of
studied scoria cones are inuenced by diverse primary
eruption-related (e.g. Hawaiian-type eruption, large to smallscale effusive activity) as well as secondary syn- and/or posteruptive erosion-related (e.g., scoria cone breaching, large scale
slope failure and anthropogenic activity) processes.
(2). Comparison of DEM-based and traditional methods shows that a
number of examples, e.g. Kopasz-hegy, Boncsos-tet and Gajostet, are unsuitable for the traditional, formulae-based morphometric parameterization (Eq. (4)) due to the erosion of the crater.
Specically, Hco/Wco ratio does not show a strong correlation
with the radiometric age of the scoria cones (Table 2) owing to
the diverse nature of cone materials as well as the highly
degraded and breached present morphology.
(3). The DEM-based method gave more reliable and acceptable
results compared to the radiometric ages (Table 2) determining
undisturbed outer slope angle values for the studied cones.
The DEM-based slope angle parameterization method is also
useful for highly truncated/asymmetric scoria cones, which
have only scattered remnants (e.g. Type 1 cones). This is
because certain parts of the outer cone slopes are still age
representatives after nearly 4 Ma of degradation. At the same

57

time, the Hco/Wco ratio does not show strong correlation with
the real, geological age of the older scoria cones (Table 2).
(4). Assessing the precision of the measurements, the DEM-based
slope angle estimation, based on thousands of pixels representing the cone surfaces, is more precise and robust. The
traditional formulae-based method is dependent on three
parameters (Hco, Wco and Wcr), hence imprecise parameter
measuring or taking highly disturbed, eroded zones of the cone
could yield unreliable results (see Kopasz-hegy or Boncsostet). A similar tendency between data obtaining methods has
been documented for the ank cones of Mt. Etna by Favalli et al.
(2009). However, the DEM-based slope angle can also show
scattered ranges of the slope histogram leading multi-peaked
histograms and high standard deviations. If the interpolation
noise is low, the individual peaks on the slope angle histogram
represent either differences in the architecture, i.e. eruption
diversity, and/or in the behaviour of erosional processes that
shaped the anks of the scoria cones.
(5). The accuracy of relative dating decreases with increasing
edice age because the inner structure of the cone (probably
more erosion resistant) gradually becomes more exposed with
time. The properties and the distribution of these exposed
resistant layers govern the precision of morphometric dating.
(6). In the BBHVF, three difference controlling factors have been
identied namely the age of the edice, the climate during the
degradation and the inner architecture of the edice, i.e.
erosion resistance of pyroclastic rocks. The last factor is more
likely to have played a major role in the formation of erosionresistant collars of scoria cones in the BBHVF.
Acknowledgements
This research was supported by Department of Geology and Mineral
Deposits, University of Miskolc, Hungary and PhD Research Fellowship of
the Volcanic Risk Solutions, Massey University, New Zealand (GK). The
authors would like to thank to K. Balogh (Institute of Nuclear Research,
Debrecen) for the help in interpreting the existing KAr radiometric data,
to G. Jordn (Geological Institute of Hungary, Budapest) for the helpful
discussions about digital modelling and to G. Csillag (Geological Institute
of Hungary, Budapest). Constructive comments by J. Procter, K. Arentsen,
D. Kartson, F. J. Dniz-Pez, J.-T. Thouret and A. Harvey signicantly
elevated the quality of the manuscript.
References
Auer, A., Martin, U., Nmeth, K., 2007. The Fekete-hegy (Balaton Highland Hungary) softsubstrate and hard-substrate maar volcanoes in an aligned volcanic complex
implications for vent geometry, subsurface stratigraphy and the paleoenvironmental
setting. Journal of Volcanology and Geothermal Research 159, 225245.
Balogh, K., Pcskay, Z., 2001. K/Ar and Ar/Ar geochronological studies in the Pannonian
CarpathiansDinarides (PANCARDI) region. Acta Geologica Hungarica 44, 281299.
Balogh, K., rva-Ss, E., Pcskay, Z., Ravasz-Baranyai, L., 1986. K/Ar dating of postsarmatian alkali basaltic rocks in Hungary. Acta Mineralogica-Petrographica 27,
7593.
Borsy, Z., Balogh, K., Kozk, M., Pcskay, Z., 1987. jabb adatok a Tapolcai-medence
fejldstrtnethez [Contribution to the evolution of the Tapolca Basin, Hungary].
Kzlemnyek a Debreceni Kossuth Lajos Tudomnyegyetem Fldrajzi Intzetbl
23, 79104 (in Hungarian with English abstract).
Breed, W.-J., 1964. Morphology and lineation of cinder cones in the San Franciscan
Volcanic Field. Museum of Northern Arizona Bulletin 40, 6571.
Budai, T., Csillag, G., Dudko, A., Koloszr, L., 1999. Geological map of Balaton Highland
(1:50,000). In: Budai, T., Csillag, G. (Eds.), Geology of the Balaton Highland
Explanation of the Geological Map of the Balaton Highland, 1:50,000. Geological
Institute of Hungary, Budapest, Hungary.
Colton, H.S., 1967. The basaltic cinder cones and lava ows of the San Francisco Mountain
Volcanic Field. Museum of Northern Arizona Bulletin 10 (revised edition), 158.
Corazzato, C., Tibaldi, A., 2006. Fracture control on type, morphology and distribution of
parasitic volcanic cones: an example from Mt. Etna, Italy. Journal of Volcanology
and Geothermal Research 158, 177194.
Csillag, G., 2004. Geomorphologic levels of the Kl Basin and its vicinity. A Magyar
llami Fldtani Intzet vi Jelentse 2004-rl, pp. 95110 (in Hungarian with
English abstact).

58

G. Kereszturi, K. Nmeth / Geomorphology 136 (2012) 4558

Csillag, G., Nmeth, K., Sebe, K., 2008. Paleofelsznek s vulkni szerkezetek kapcsolata a
Balaton-felvidk s a Bakony terletn [Relationship between paleosurfaces and
volcanic structures in the BakonyBalaton Highland Volcanic Fields, Hungary], IV.
Magyar Fldrajzi Konferencia, Debrecen, Hungary, pp. 8490 (in Hungarian with
English abstact).
Di Traglia, F., Cimarelli, C., de Rita, D., Gimeno Torrente, D., 2009. Changing eruptive
styles in basaltic explosive volcanism: examples from Croscat complex scoria cone,
Garrotxa Volcanic Field (NE Iberain Peninsula). Journal of Volcanology and
Geothermal Research 180, 89109.
Dohrenwend, J.C., Wells, S.G., Turrin, B.D., 1986. Degradation of Quaternary cinder
cones in the Cima volcanic eld, Mojave Desert, California. Geological Society of
America Bulletin 97, 421427.
Doniz, J., Romero, C., Coello, E., Guillen, C., Sanchez, N., Garcia-Cacho, L., Garcia, A., 2008.
Morphological and statistical characterisation of recent mac volcanism on Tenerife
(Canary Islands, Spain). Journal of Volcanology and Geothermal Research 173,
185195.
Fbin, S.., Kovcs, J., Varga, G., 2000. jabb szempontok haznk periglacilis
klmjhoz [New data on the periglacial climate in Hungary]. Fldrajzi rtest
49, 189204 (in Hungarian with English abstact).
Favalli, M., Kartson, D., Mazzarini, F., Pareschi, M.T., Boschi, E., 2009. Morphometry of
scoria cones located on a volcano ank: a case study from Mt. Etna (Italy), based on
high-resolution LiDAR data. Journal of Volcanology and Geothermal Research 186,
320330.
Fisher, P.F., Tate, N.J., 2006. Causes and consequences of error in digital elevation
models. Progress in Physical Geography 30, 467489.
Garbrecht, J., Martz, L.W., 1997. The assignment of drainage direction over at surfaces
in raster Digital Elevation Models. Journal of Hydrology 193, 204213.
Gisbert, G., Gimeno, D., Fernandez-Turiel, J.-L., 2009. Eruptive mechanisms of the Puig
De La Garrinada volcano (Olot, Garrotxa volcanic eld, Northeastern Spain): a
methodological study based on proximal pyroclastic deposits. Journal of Volcanology and Geothermal Research 180, 259276.
Gorte, B.G.H., Koolhoven, W., 1990. Interpolation between isolines based on the
Borgefors distance transform. ITC Journals 19903, 245247.
Hasenaka, T., Carmichael, I.S.E., 1985. A compilation of location, size, and geomophological parameters of volcanoes of the MichoacanGuanajuato volcanic eld,
central Mexico. Geosica Internacional 24, 577607.
Hengl, T., 2006. Finding the right pixel size. Computers & Geosciences 32, 12831298.
Hengl, T., Evans, I.S., 2008. Mathematical and digital models of the land surface. In:
Hengl, T., Reuter, H. (Eds.), Geomorphometry: Concepts, Software, Applications.
Elsevier B.V, Amsterdam, pp. 3163.
Hooper, D.M., 1995. Computer-simulation models of scoria cone degradation in the
Colima and MichoacnGuanajuato volcanic elds, Mexico. Geosica Internacional
34, 321340.
Hooper, D.M., Sheridan, M.F., 1998. Computer-simulation models of scoria cone
degradation. Journal of Volcanology and Geothermal Research 83, 241267.
Jmbor, ., 1980. Pannonian in the Transdanubian Central Mountains. A Magyar llami
Fldtani Intzet vknyve, 57. Magyar llami Fldtani Intzet, Budapest, Hungary.
Jones, K.H., 1998. A comparison of algorithms used to compute hill slope as a property
of the DEM. Computers & Geosciences 24, 315323.
Jordan, G., 2007. Digital terrain analysis in a GIS environment. Concepts and development.
In: Peckham, R.J., Jordan, G. (Eds.), Digital Terrain Modelling. Development and
Applications in a Policy Support Environment, Springer, Berlin, pp. 143.
Kereszturi, G., 2009. Morfometrikus mrsek a BakonyBalaton-felvidk vulkni terlet
pontosabb sfldrajzi viszonyainak megismershez [New data on paleogeographical
conditions of BakonyBalaton Highland Volcanic Field on the basis of morphometric
measurements]. MSc thesis Thesis, University of Miskolc, Miskolc, Hungary.
Kereszturi, G., 2010. A fekete-hegyi salakkp-roncsok morfometrija (BakonyBalatonfelvidk vulkni terlet) [Morphometry of eroded scoria cones of Fekete-hegy
(Balaton Highland, Hungary)]. Fldrajzi Kzlemnyek 133, 255267 (in Hungarian
with English abstact).
Kereszturi, G., Nmeth, K., in press. Shallow-seated controls on the evolution of the Upper
Pliocene Kopasz-hegy nested monogenetic volcanic chain in the Western Pannonian
Basin, Hungary. Geologica Carpathica. doi:10.2478/v10096-011-0038-3.

Kereszturi, G., Csillag, G., Nmeth, K., Sebe, K., Balogh, K., Jger, V., 2010. Volcanic
architecture, eruption mechanism and landform evolution of a Pliocene intracontinental basaltic polycyclic monogenetic volcano from the BakonyBalaton
Highland Volcanic Field, Hungary. Central European Journal of Geosciences 2,
362384.
Kereszturi, G., Nmeth, K., Csillag, G., Balogh, K., Kovcs, J., 2011. The role of external
environmental factors in changing eruption styles of monogenetic volcanoes in a
Mio/Pleistocene continental volcanic eld in western Hungary. Journal of
Volcanology and Geothermal Research 201, 227240.
Kovcs, J., Fbin, S.., Varga, G., jvri, G., Varga, G., Dezso, J., 2011. Plio-Pleistocene red
clay deposits in the Pannonian basin: a review. Quaternary International 240, 3543.
Magyar, I., Geary, D.H., Mller, P., 1999. Paleogeographic evolution of the Late Miocene
Lake Pannon in Central Europe. Palaeogeography Palaeoclimatology Palaeoecology
147, 151167.
Martin, U., Nmeth, K., 2004. Mio/Pliocene phreatomagmatic volcanism in the Western
Pannonian Basin. Geologica Hungarica Series Geologica 26 (Budapest).
Martin, U., Nmeth, K., 2006. How Strombolian is a Strombolian scoria cone? Some
irregularities in scoria cone architecture from the Transmexican Volcanic Belt, near
Volcn Ceboruco (Mexico), and Al Haruj (Libya). Journal of Volcanology and
Geothermal Research 155, 104118.
McGetchin, T.R., Settle, M., Chouet, B.A., 1974. Cinder cone growth modeled after Northeast
Crater, Mount Etna, Sicily. Journal of Geophysical Research 79, 32573272.
Nmeth, K., Martin, U., 1999. Small-volume volcaniclastic ow deposits related to
phreatomagmatic explosive eruptive centres near Szentbkklla, BakonyBalaton
Highland Volcanic Field, Hungary: pyroclastic ow or hydroclastic ow? Fldtani
Kzlny 129, 393417.
Nmeth, K., Risso, C., Nullo, F., Kereszturi, G., 2011. The role of collapsing and rafting of
scoria cones on eruption style changes and nal cone morphology: Los Morados
scoria cone, Mendoza, Argentina. Central European Journal of Geosciences 3,
102118.
Partt, E.A., Wilson, L., 1995. Explosive volcanic eruptionsIX. The transition between
Hawaiian-style lava fountaining and Strombolian explosive activity. Geophysical
Journal International 121, 226232.
Pioli, L., Erlund, E., Johnson, E., Cashman, N.K., Wallace, P., Rosi, M., Delgado Granados, H.,
2008. Explosive dynamics of violent Strombolian eruptions: the eruption of Parcutin
Volcano 19431952 (Mexico). Earth and Planetary Science Letters 271, 359368.
Porter, S.C., 1972. Distribution, Morphology, and Size Frequency of Cinder Cones on
Mauna Kea Volcano, Hawaii. Geological Society of America Bulletin 83, 36073612.
Raaaub, L.D., Collins, M.J., 2006. The effect of error in gridded digital elevation models
on the estimation of topographic parameters. Environmental Modelling & Software
21, 710732.
Sebe, K., Csillag, G., Ruszkiczay-Rdiger, Z., Fodor, L., Tham-Bozs, E., Mller, P.,
Braucher, R., 2011. Wind erosion under cold climate: a Pleistocene periglacial
mega-yardang system in Central Europe (Western Pannonian Basin, Hungary).
Geomorphology 134, 470482.
Settle, M., 1979. The structure and emplacement of cinder cone elds. American Journal
of Science 279, 10891107.
Sharpnack, D.A., Akin, G., 1969. An algorithm for computing slope and aspect from
elevation. Photogrammetric Survey 35, 247248.
Sucipta, I.G.B.E., Takashima, I., Muraoka, H., 2006. Morphometric age and petrological
characteristic of volcanic rocks from the Bajawa cinder cone complex, Flores,
Indonesia. Journal of Mineralogical and Petrological Sciences 101, 4868.
Valentine, G.A., Gregg, T.K.P., 2008. Continental basaltic volcanoesprocesses and
problems. Journal of Volcanology and Geothermal Research 177, 857873.
Wijbrans, J., Nmeth, K., Martin, U., Balogh, K., 2007. 40Ar/39Ar geochronology of
Neogene phreatomagmatic volcanism in the western Pannonian Basin, Hungary.
Journal of Volcanology and Geothermal Research 164, 193204.
Wood, C.A., 1980a. Morphometric evolution of cinder cones. Journal of Volcanology and
Geothermal Research 7, 387413.
Wood, C.A., 1980b. Morphometric analysis of cinder cone degradation. Journal of
Volcanology and Geothermal Research 8, 137160.
Zevenbergen, L.W., Thorne, C.R., 1987. Quantitative analysis of land surface topography.
Earth Surface Processes and Landforms 12, 4756.

You might also like