You are on page 1of 21

INTERNATIONAL JOURNAL OF CLIMATOLOGY

Int. J. Climatol. 26: 439–459 (2006)


Published online 25 January 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/joc.1249

RESPONSES OF LARGE VOLCANIC ERUPTIONS IN THE INSTRUMENTAL


AND DOCUMENTARY CLIMATIC DATA OVER CENTRAL EUROPE
JAN PÍSEKa, * and RUDOLF BRÁZDILb
a University of Toronto, Department of Geography and Program in Planning, 100 St George St, Room 5047 Toronto, Ontario, Canada
M5S 3G3, Canada
b Masaryk University, Institute of Geography, Kotlářská 2, 611 37 Brno, The Czech Republic

Received 23 February 2005


Revised 29 June 2005
Accepted 30 June 2005

ABSTRACT
Responses of large volcanic eruptions in selected long temperature series from Austria, the Czech Republic and Germany
as well as in three global radiation series in Central Europe are studied. In the example of seven large tropical eruptions
(Krakatau 1883; Pelée, Soufriére and Santa Marı́a 1902; Agung, 1963; El Chichón, 1982; Mt Pinatubo, 1991) it has
been demonstrated that volcanic signal in regional series is not so strongly expressed as in the hemispheric scale owing
to different local effects and circulation patterns. This is also valid in the case of two further discussed eruptions of
Tambora (1815) and Katmai (1912). The responses of eruptions in areas closer to Central Europe such as Iceland or Italy
are more important. In nine analysed cases with VEI = 4–5 with a single exception of the Hekla eruption (1917), cold
seasons were observed to follow the eruption. Responses to the Lakagı́gar eruption (1783) of Iceland with important
impacts are also discussed in detail. Moreover, correlation between temperatures (annual and winter half-year series) and
NAOI is prevailingly smaller for the period following eruptions than in the period preceding eruptions. The importance
of documentary evidence as a valuable source of the information about the impacts of volcanic eruptions is demonstrated.
Copyright  2006 Royal Meteorological Society.

KEY WORDS: large volcanic eruptions; air temperature; global radiation; documentary data; Central Europe

1. INTRODUCTION

A number of studies have been carried out to assess the climatic and environmental impacts of large volcanic
eruptions (see e.g. Bradley, 1988; Bradley and Jones, 1992; Robock and Mao, 1992, 1995; Camuffo and
Enzi, 1995; Briffa et al., 1998; Dunn, 2004). During eruptions, substantial quantities of sulphur and volatiles
can be injected into the upper atmosphere where contamination may persist for extended periods of time
(Robock, 2000). It has been argued that sulphate aerosols can perturb the radiation balance by warming the
stratosphere and cooling the Earth’s surface. Numerous difficulties are encountered in making assessments
such as concurrent atmospheric and oceanic variations or lacking statistical significance owing to a limited
number of available data. Nevertheless, post-eruption cooling of several tenths of degrees Celsius on a global
scale has been detected (e.g. Kelly et al., 1996; Parker et al., 1996). Many papers have dealt with some
outstanding events such as the historically largest known eruption of Tambora (Indonesia) in 1815 (see e.g.
Stommel and Stommel, 1983; Stothers, 1984; Harington, 1992; Vupputuri, 1992; Oppenheimer, 2003) or
the Laki (Lakagı́gar) eruption (Iceland) in 1783 with many consequences in Europe (see e.g. Wood, 1992;
Stothers, 1996; Demarée et al., 1998; Grattan and Pyatt, 1999; Grattan and Sadler, 1999; Demarée and Ogilvie,
2001; Brázdil et al., 2003). The greatest climatological attention from more recent eruptions was devoted to

* Correspondence to: Jan Pı́sek, University of Toronto, Department of Geography and Program in Planning, 100 St George St, Room
5047 Toronto, Ontario, Canada M5S 3G3, Canada; e-mail: jan.pisek@utoronto.ca

Copyright  2006 Royal Meteorological Society


440 J. PÍSEK AND R. BRÁZDIL

the St Helen eruption (USA) in 1980 (e.g. Robock, 1981; Mass and Robock, 1982; Robock and Mass, 1982),
El Chichón (Mexico) in 1982 (e.g. Parker and Brownscombe, 1983; McCracken and Luther, 1984; Vupputuri
and Blanchet, 1984; Angell, 1997a,b) or Mt Pinatubo (Philippines) in 1991 (e.g. Groisman, 1992; McCormick
et al., 1995; Parker et al., 1996; Kirchner et al., 1999); most of these studies have focused on the global or
hemispheric impacts of the eruption.
Significantly less attention has been paid to the climatic effects of large volcanic eruptions on a local or
regional scale. For example, Kyncl et al. (1990) analysed climatic and tree-ring responses to the Katmai
eruption (Alaska) in 1912 mainly on the Central European scale. Brůžek (1992) studied impacts of large
volcanic eruptions in the temperature series of Prague-Klementinum, the Czech Republic. Dawson et al.
(1997) discussed three clear peaks in storminess in series of gale frequency for Edinburgh, Scotland, that
occurred after the volcanic eruptions of Tambora (1815), Krakatau (1883) and El Chichón (1982). Recently,
Jones et al. (2003) used the region of Central Europe (45–55 ° N, 5–30 ° E), Fennoscandia and Central England
for a study of the response of long temperature records to explosive volcanic eruptions.
This paper focuses on the assessment of volcanic impacts on temperatures and global radiation over Central
Europe (with special attention to the Czech Republic) and compares them with the effects on the global and
hemispheric scales. Besides large volcanic eruptions, attention has also been paid to the effect of smaller
volcanic eruptions in the closest hot spots of volcanic activity, Iceland and Italy. In addition to instrumental
records, responses to volcanic activity recorded in the Czech documentary evidence are also mentioned,
mainly for the Lakagı́gar and Tambora eruptions.

2. DATA AND METHODS

To assess the climatic impacts of large volcanic eruptions in Central Europe, numerous long data sets
are available which make it possible to carry out the analysis for the selected area. As a basic series,
monthly air temperatures from station Prague-Klementinum in the Czech Republic (for location of all
meteorological stations and mentioned places in Central Europe, see Figure 1), extending from 1775 to
the present, were used (for homogeneity and more details regarding this station, see Brázdil and Budı́ková,
1999). Further, the homogenized monthly temperatures from Austrian climatological stations Vienna-Hohe
Warte and Kremsmünster, extending from 1775 and 1769, respectively, were available (Auer et al., 2001).
Finally, two long series of spatial averages were used for analysis: a composite series of monthly anomalies
(reference period 1961–1990) compiled from the stations in Germany (Rapp, 2000) since 1761, and another
one, calculated from weather stations in the Czech Republic and extending from year 1848 (Štěpánek, 2003).
Temperature fluctuations in Central Europe based on annual or monthly averages show very similar features
(Figure 2). They can be documented with the example of correlation coefficients between the station Prague-
Klementinum and other Central European series used (Table I). Correlations are highest during winter months
(in most cases higher than 0.95), while in summer months they sometimes drop below 0.90 (mainly in August).
All of the computed correlations are statistically significant at 0.05 level. It means that Central Europe is
climatically a relatively homogeneous region, and any conclusions derived from any of the examined series
may be valid for the region as a whole.

Table I. Correlation coefficients between the monthly temperature series of Prague-Klementinum with other Central
European series (calculated always for the longest overlapping period of observations)

Series J F M A M J J A S O N D Year

Czech Republic 0.960 0.969 0.954 0.941 0.921 0.918 0.906 0.895 0.933 0.891 0.905 0.933 0.908
Kremsmünster 0.948 0.958 0.959 0.949 0.950 0.899 0.928 0.895 0.940 0.930 0.951 0.957 0.947
Vienna-Hohe Warte 0.963 0.969 0.969 0.956 0.943 0.918 0.922 0.878 0.939 0.950 0.947 0.952 0.963
Germany 0.971 0.958 0.963 0.929 0.942 0.926 0.955 0.907 0.930 0.919 0.930 0.961 0.958

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 441

0 100 200 km

Potsdam

Germany Poland

3
2
7
Hradec 1
Prague Králové 6 4

Czech Republic 5
Skalnaté Pleso

Slovakia
Kremsmünster Vienna

Austria Hungary

Figure 1. Location of the meteorological stations used in Central Europe and places in the Czech Republic mentioned in the text
(1 – Kunvald, 2 – Mt Milešovka, 3 – Noviny pod Ralskem, 4 – Opava, 5 – Pı́sečné, 6 – Rýmařov, 7 – Žatec)

Global and Northern Hemisphere monthly average temperature anomalies from the reference period
1961–1990 as compiled by the Climate Research Unit, University of East Anglia, Norwich (for details,
see Folland et al., 2001), were used to assess the volcanic effect on the global and hemispheric scale. The
decision to use these data sets was made with respect to the fact that available data extend from 1856 to
the present, which significantly increases the number of important volcanic eruptions for analysis. Secondly,
the data are of known reliability (though decreasing to the past), and, thus, the risk of possible flaws in the
results is reduced.
To evaluate changes in radiation, monthly totals of global radiation as measured in Potsdam (81 m),
Germany (http://www.klima-potsdam.de), Hradec Králové (278 m) in the Czech Republic and Skalnaté
Pleso (1778 m) in the Tatra Mts, Slovakia (Ostrožlı́k, 2002) were used. Also, series of the North
Atlantic Oscillation Index (NAOI; Lamb and Peppler, 1987; Hurrell, 1995) and the Central European
Zonal Index (CEZI; Jacobeit et al., 1998) were used to analyse eruption effects upon atmospheric circu-
lation and temperature data. The NAOI, which describes the character of airflow in Europe, is defined
as the standardized (1901–1980) difference between the sea-level pressure average of four grid points
on a 5 × 5 longitude–latitude grid over the Azores and over Iceland (for more details, see Luter-
bacher et al., 2002). CEZI characterizes airflow due to meridional change of air pressure in Central
Europe. It is calculated as the difference between the normalized values of surface pressure averaged
for grid points 35 ° N/0° , 35 ° N/20 ° E, 40 ° N/0° , 40 ° N/20 ° E and 60 ° N/0° , 60 ° N/20 ° E, 65 ° N/0° , 65 ° N/20 ° E,
respectively.
Large volcanic eruptions with Volcanic Explosivity Index (VEI) equal to 5 and higher were selected
(Table II). Using these criteria, the list contained all known large volcanic eruptions during the last 250 years
as summarized by Robock (2000). Furthermore, the original data set of selected volcanic eruptions has been
extended by smaller volcanic eruptions (with VEI = 4) that took place in Iceland and Italy (Table III). Despite
the smaller scale of these eruptions, their closeness to the area of interest does not completely exclude their
possible effect on the local climate. Tables II and III summarize the final list of volcanic eruptions utilized,
with information about their location, the year and month the eruption took place and the respective scale of
the eruption.
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
442 J. PÍSEK AND R. BRÁZDIL

2
Prague-klementinum
1

-1

-2
1750 1800 1850 1900 1950 2000

2
Czech Republic
1

-1

-2
1750 1800 1850 1900 1950 2000

Kremsünster
Temperature (°C)

-1

-2
1750 1800 1850 1900 1950 2000

2
Vienna-Hohe Warte
1

-1

-2
1750 1800 1850 1900 1950 2000

2
Germany
1

-1

-2
1750 1800 1850 1900 1950 2000

Figure 2. Fluctuations of annual mean temperature anomalies (reference period 1961–1990) for selected Central European series

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 443

Table II. List of large volcanic eruptions with VEI ≥ 5 considered in the analysis (source:
http://www.volcano.si.edu/gvp/world/largeeruptions.cfm). In the case of 1902, eruptions of
Pelée and Soufriére with VEI = 4 only were included

Month-year Name Location VEI

Jan–1800 St Helens Washington, USA 5


Apr–1815 Tambora Lesser Sunda Islands, Indonesia 7
Oct–1822 Galunggung Java, Indonesia 5
Jan–1835 Cosiguina Nicaragua 5
Dec–1853 Chikurachki Kurile Islands 5
Feb–1854 Shiveluch Kamchatka, Russia 5
Mar–1875 Askja Iceland 5
Aug–1883 Krakatau Indonesia 6
May–1902 Pelée West Indies 4
May–1902 Soufriére West Indies 4
Oct–1902 Santa Marı́a Guatemala 6
Mar–1907 Ksusach Kamchatka, Russia 5
June–1912 Novarupta Alaska, USA 6
Jan–1933 Kharimkotan Kurile Islands 5
Mar–1956 Bezyminanny Kamchatka, Russia 5
Mar–1963 Agung Lesser Sunda Islands, Indonesia 5
May–1980 St Helens Washington, USA 5
Apr–1982 El Chichón Chiapas, Mexico 5
Jun–1991 Pinatubo Luzon, Philippines 6

Table III. List of analysed volcanic eruptions with VEI ≥ 4


located in the European region (source: http://www.volcano.si.
edu/gvp/world/largeeruptions.cfm)

Month-year Name Location VEI

June–1783 Lakagı́gar Iceland 4


July–1787 Etna Italy 4
Sep–1845 Hekla Iceland 4
Jan–1873 Laki Iceland 4
Mar–1875 Askja Iceland 5
May–1903 Laki Iceland 4
Apr–1906 Vesuvius Italy 4
Oct–1918 Katla Iceland 4
Mar–1947 Hekla Iceland 4

While identifying the effects of this historic sample of eruptions, their impact on global and north hemi-
spheric monthly average temperatures was first considered. Superposed epoch analysis was used (Conrad and
Pollak, 1962), following the methodologies adopted in previous work (Sear et al., 1987; Kelly et al., 1996).
The month of the eruption was taken as month zero. The mean temperature for each month was calculated
using temperature data from the 5 years prior to the eruption. Each monthly average temperature for the
5 years before and after the eruption was then expressed as a departure from the calculated mean value.
This was done for all selected eruptions in relation to global radiation data, temperature data and circulation
indexes.
Robock and Mao (1992, 1995) argued that the effects of volcanic eruptions on climate vary according
to their location and that the largest effects should be observed following large sulphur-rich, low-latitude
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
444 J. PÍSEK AND R. BRÁZDIL

volcanic eruptions. They found that Northern Hemispheric winter warmth tended to occur in the first winter
after tropical eruptions, but in the second winter after higher-latitude eruptions. For the study of described
climatic effects usually, eruptions with the highest VEI values are selected (e.g. Kelly et al., 1996; Parker
et al., 1996; Jones et al., 2003).
It is necessary to mention that selection of volcanic eruptions according the VEI values for the study of
the climate impacts might be problematic because the total sulphate loading is the primary factor controlling
the radiative perturbation. For this reason, according to Ammann et al. (2003), existing volcanic data sets
include only limited information about seasonal or latitudinal distribution of volcanic aerosols to realistically
simulate the perturbations. They produced the new data set on the basis of the total amount of sulphate
released by each eruption, which contains the meridional spread and decay of volcanic aerosols at monthly
resolution, taking into account the seasonally changing stratospheric transport. With this data, they proved
greatest negative forcing from volcanic activity during the twentieth century prior to 1915 and after 1960.
On the basis of the above-mentioned papers (e.g. Kelly et al., 1996; Parker et al., 1996; Jones et al., 2003),
seven large tropical volcanic eruptions (VEI = 5 and higher) Krakatau (1883), Pelée, Soufriére and Santa
Marı́a (all in 1902), Agung (1963), El Chichón (1982), and Mt Pinatubo (1991), associated with the largest
estimated stratospheric aerosol optical depth (SAOD) (Sato et al., 1993) were selected for the following
analysis. Comparable with these large tropical eruptions would probably also be another unknown event of
1809 revealed from high-resolution analyses of ice cores from Antarctica and Greenland (Dai et al., 1991) and
later shown in tree-ring data (e.g. Briffa et al., 1998; for further discussion concerning dating of this event,
see Chenoweth, 2001; Mosley-Thompson et al., 2003), but owing to the amount of persisting uncertainties
about this event, it was not included in our analysis.
The selected seven large tropical eruptions were further averaged in order to assess and compare their
effect on global, hemispheric and local scales. This technique should have produced a composite profile,
highlighting features common to the post-eruption temperature series. Similar to the procedure undertaken by
Kelly et al. (1996), a Monte Carlo approach was used for testing the statistical significance of the composite
profiles. Owing to the limited number of samples, the procedure was repeated 5000 times by randomly
selecting eruption months for each of the composite profiles. The distribution of the resulting temperature
departures was used to define the 0.05 significance level; the test was one-tailed because of particular interest
in assessing the post-eruption cooling in the profiles. Finally, a low-pass Gaussian filter with smoothing
interval of 6 months was applied to all calculated profiles.

3. CLIMATIC EFFECTS OF LARGE VOLCANIC ERUPTIONS IN CENTRAL EUROPE

3.1. Large tropical volcanic eruptions


Radiative effect of volcanogenic pollution after eruptions should induce some responses in measured solar
radiation. Significant radiation responses were reported, e.g. after the Mt Pinatubo eruption. Blumthaler and
Ambach (1994) recorded an increase in diffuse sky total radiation and a concomitant decrease in global total
radiation (sun and sky) of 4% comparing May–June 1991 (before the aerosol cloud of the Mt Pinatubo
eruption) with the same months a year later for the high-Alpine station Jungfraujoch. Non-continuous solar
radiation measurements in Abisko (northern Sweden) between 1990 and 1994 showed after the Mt Pinatubo
eruption a decrease in the direct radiation varying between 18 and 30% (at sun elevations of 40° and 15° ,
respectively) (Adeyefa et al., 2000). In the global radiation, the corresponding changes were 1.4 and 8%
(for the same sun elevations). Diffuse radiation in many cases penetrates better into plant canopies, thus
potentially enhancing photosynthesis (Reichenau and Esser, 2003). Worldwide increased diffuse radiation
related to atmospheric aerosol loading after Mt Pinatubo eruption enhanced noontime photosynthesis of a
deciduous forest by 23% in 1992 and 8% in 1993 under cloudless conditions (Gu et al., 2003).
For Central Europe, only the variations of measured monthly sums of global radiation on the stations
Potsdam, Hradec Králové and Skalnaté Pleso around the time of large volcanic eruptions in the second half
of the twentieth century were analysed. The profiles created for Agung, El Chichón and Pinatubo eruptions
are shown in Figure 3. Global radiation anomalies were calculated as deviations from the appropriate monthly
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 445

Agung 1963
15000
(a)
10000

5000

-5000

-10000

-15000
-60 -48 -36 -24 -12 0 12 24 36 48 60
El Chichón 1982
15000
Global radiation (J.cm-2)

(b)
10000

5000

-5000

-10000

-15000
-60 -48 -36 -24 -12 0 12 24 36 48 60
Pinatubo 1991
15000
(c)
10000

5000

-5000

-10000

-15000
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Potsdam Hradec Králové Skalnaté Pleso

Figure 3. Comparison of profiles of low-pass-filtered monthly sums of global radiation for Potsdam station in Germany, Hradec Králové
station in the Czech Republic and Skalnaté Pleso in Slovakia around the time of Agung eruption in March 1963 (a), El Chichón eruption
in April 1982 (b) and Mt Pinatubo eruption in June 1991 (c). The month of the corresponding eruption is taken as month zero

mean for the 5 years preceding the January of the eruption year. As can be seen, it is difficult to observe any
common features of the profiles. However, around month 14 after the moment of eruption, all profiles record
a similar rise in levels of measured global radiation. On the other hand, a significant drop in global radiation
is evident in the beginning of the third year after the Agung and El Chichón eruptions. Owing to the lack of
any other large volcanic events during the period, it remains difficult to determine if these features are a pure
coincidence or a typical sign of the effect of the eruptions in global radiation in the Central European region.
The composite profiles of the five large volcanic eruptions as demonstrated in Global, Northern Hemisphere
and composite Czech Republic series of monthly averaged temperature anomalies from the reference period
1961–1990 are shown in Figure 4. In agreement with previous studies, apparent cooling for 3 years after
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
446 J. PÍSEK AND R. BRÁZDIL

Tropical Eruptions
0.2
(a) Global
0.1

-0.1

-0.2
-60 -48 -36 -24 -12 0 12 24 36 48 60

0.2
(b)
Northern Hemisphere
Temperature (°C)

0.1

-0.1

-0.2

-0.3
-60 -48 -36 -24 -12 0 12 24 36 48 60

2
(c)
Czech Republic
1

-1

-2
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Figure 4. (a) Composite sequence of low-pass-filtered monthly global-mean surface air temperature anomalies from the reference
period 1961–1990 around the time of seven tropical volcanic eruptions (Krakatau, Pelée-Soufriére-Santa Marı́a (composite), Agung, El
Chichón, Pinatubo). Values are expressed as anomalies relative to the 5-year pre-eruption periods. (b) As (a) but for monthly Northern
Hemisphere temperature series. (c) As (a) but for the Czech Republic composite series. The dashed horizontal line indicates the 0.05
significance level. The month zero is taken as month of eruptions

the eruption can be observed in the Global and Northern Hemisphere series. This cooling tendency reaches
its maximum during the second year after the eruption and the cooling is statistically significant at the 0.05
level even during the third post-eruption year. In agreement with Robock (2000), it was possible to observe
warming during the post-eruption months 4–6 (Figure 4(b)). This warming is produced by an enhanced pole-
to-equator temperature gradient that results in stronger polar vortex, characterized by a stationary wave pattern
of tropospheric circulation. This consequently causes winter warming of the Northern Hemisphere continents
(Robock, 2000). However, the warming is not as significant as the cooling in the following seasons. Regarding
both winter warming and the subsequent cooling, both these events are more pronounced in the Northern
Hemisphere data set than on the global scale.
The first winter-warming tendency can be observed in the Czech composite series (Figure 4(c)); however,
it completely misses the cooling tendency during the second post-eruption year typical for the Global and
Northern Hemisphere series. The cooling reaches its maximum during month 25 after the eruption; however,
even then it does not drop below the 0.05 significance level. The most striking feature is the suspicious
pre-eruption cooling, which would require further research for explanation. The course of the profile for the
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 447

other series from Central Europe is essentially identical to the composite profile for described series from the
Czech Republic.
The Global and Northern Hemisphere temperature series do not cover the period of the strongest volcanic
eruption of Tambora (Lesser Sunda Islands, Indonesia) in April 1815 (Sigurdsson and Carey, 1992), classified
with VEI = 7. The subsequent year 1816 was characterized as ‘the year without a summer’ (see e.g. Stommel
and Stommel, 1983; Harington, 1992; Vupputuri, 1992; Oppenheimer, 2003). Cooler periods shown in
the first and second post-eruption years in the Central European temperature series were interrupted by
warming in the first winter after the eruption (Figure 5). However, modification of weather by the Tambora
eruption was not particularly extreme in comparison with other cold years in the data set. The summer
of 1816 was classified as the 18th coldest one in the long-term Prague measurements. This corresponds
with the work of Briffa and Jones (1992) who classified only the summer of 1816 as extreme in the
context of the 1810s in Europe. In contrast, according to documentary evidence, the period from June
1815 to December 1816 was characterized as extremely cold for Opava in Silesia (Kreuzinger, 1862).
Another source mentions a very poor harvest due to continuous wet weather and very bad vintage in
1815 in Bohemia (Bachmann, 1912). Cold rainy weather with a bad harvest in this year is reported also
by many other documentary sources in the Czech Lands. Another poor grain harvest in 1816 caused a
great dearth in Bohemia. It was partly due to May frosts which damaged flowering cereals and partly a
consequence of cold and rainy summer which delayed usual harvest time and decreased the amount of
grain (see e.g. Lenhart, 1840; Tille, 1905; Blau, 1908; Komárek, 1911; Trnka, 1912; Robek, 1974; Roubic,
1987).
Further attention was paid to the comparison of the most pronounced eruptions in the Northern Hemisphere
composite profile and the profiles from Central Europe. The effect of the large volcanic eruptions in Global and
Northern Hemisphere composite profiles has been previously assessed in other studies (see Kelly et al., 1996;
Parker et al., 1996). Since most of these eruptions lacked any more pronounced effect on the temperature in
the Central European series, only two of the eruptions (Santa Marı́a, 1902; Agung, 1963) with noteworthy
features in the Central European temperature series are presented below.
The comparison of temperature development for the Santa Marı́a eruption in 1902 (Williams and Self,
1983) for the Northern Hemisphere profile and for the Central European series can be seen in Figure 6.
For the Northern Hemisphere, winter warming can be discerned in the first post-eruption year, followed
by steep temperature drop in the following seasons; the maximum cooling agrees with first post-eruption
summer period. This summer was shown to be the coldest summer in Europe during the period 1500–2003
(Luterbacher et al., 2004). During the second post-eruption winter, it was possible to notice slight warming
tendency; however, even during the consequent months the temperature stayed well below pre-eruption

Tambora 1815
4
Temperature (°C)

-2

-4
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Prague Kremsmünster Vienna Germany

Figure 5. Comparison of monthly mean surface air temperature profiles for Prague-Klementinum, Kremsmünster, Vienna-Hohe Warte
and composite from German stations around the time of Tambora eruption in 1815. Values are expressed as anomalies relative to the
5-year pre-eruption period. The month zero – April 1815

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
448 J. PÍSEK AND R. BRÁZDIL

Pelée - Soufriére - Santa Maria 1902


0.4
(a) Northern Hemisphere
0.2

-0.2

-0.4
Temperature (°C)

-0.6
-60 -48 -36 -24 -12 0 12 24 36 48 60

4
(b)
2

-2

-4
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Prague Kremsmünster Vienna Germany

Figure 6. Comparison of monthly mean surface air temperature profiles (a) for Northern Hemisphere and (b) for Prague-Klementinum,
Kremsmünster, Vienna-Hohe Warte and composite from German stations around the time of Pelée-Soufriére-Santa Marı́a eruption in
1902. Values are expressed as anomalies relative to the 5-year pre-eruption period. The month zero – October 1902

monthly average temperatures. In the case of the Central European series, with the exception of the
immediate cooling after the eruption, it is not possible to notice any expected behaviour in the profile.
Temperatures mostly stay above the calculated pre-eruption month averages during the whole second and
third years, while in the Northern Hemisphere profile they drop significantly below pre-eruption temperature
levels.
In Figure 7, a similar comparison between the Central European and Northern Hemisphere series was
carried out for the Agung eruption in 1963 (Self and King, 1996). The Northern Hemisphere temperature
profile follows basically the same course as that of the Santa Marı́a eruption, marked by significant winter
warming in the first year, followed by a steep drop in temperature during the following years. The profile
for the Central European series shows significantly different tendencies during the post-eruption period.
The maximum temperature drop begins before the eruption and is followed by steep warming during the
2–8 post-eruption months. In contrast to the Northern Hemisphere profile, the first post-eruption winter is
significantly cooler, cooling appears at the end of the first post-eruption year and the temperature during the
most of the second year reaches above-average levels. Other signs of cooling can be observed during the
third post-eruption year.
From this analysis, it can be concluded that in the case of large tropical volcanic eruptions the temperature
follows a more or less similar course during the post-eruption period on global and Northern Hemisphere
scales. It is typical that a warming occurs during the first post-eruption winter and cooling occurs in the two
post-eruption years, which is statistically significant at the 0.05 level. In the case of the Central European
region, the profiles show a rather different course. Surprisingly, the most significant cooling can sometimes
be observed immediately before the time of the eruption. Although different cooler periods appear also during
post-eruptions years, it is rather difficult to observe any common features in the corresponding temperature
profiles.
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 449

Agung 1963
0.4
(a) Northern Hemisphere
0.2

-0.2

-0.4
Temperature (°C)

-0.6
-60 -48 -36 -24 -12 0 12 24 36 48 60

4
(b)
2

-2

-4

-6
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Prague Kremsmünster Vienna Germany

Figure 7. Comparison of monthly mean surface air temperature profiles (a) for Northern Hemisphere and (b) for Prague-Klementinum,
Kremsmünster, Vienna-Hohe Warte and composite from German stations around the time of Agung eruption in 1963. Values are
expressed as anomalies relative to the 5-year pre-eruption period. The month zero – March 1963

3.2. Northern hemisphere volcanic eruptions


More important climatic impacts in Central Europe were related to another large eruption with VEI = 6,
Katmai (Novarupta), Alaska, in June 1912. In this month, very weak traces of sunshine during clear days
were already recorded with a Campbell-Stokes sunshine recorder at the meteorological station Mt Milešovka
(Stöhr, 1912). As the author of this report comments in the entry for 28–29 June, the traces had a weak
brown colour. He hypothesized that weak impact of the sun’s rays together with the pale-blue colour of the
sky and remarkable twilight could be attributed to a volcanic eruption appearing, according to him, a short
time before in Mexico. Similar clear evidence of the Katmai eruption influence in sunshine records was found
by Hanzlı́k (1918) for the station Prague-Petřı́n. According to Prague-Klementinum observations, autumn of
1912, following immediately after the eruption, was the second coldest autumn (in the order of individual
months, August 1912 was the 4th coldest, September, absolutely the coldest and October, the 11th coldest),
and summer of 1913, the third coldest summer in the 226-year Prague series. Another extremely cold autumn
was recorded in 1915 (as the fifth coldest) (Figure 8). Moreover, the mean relative maximum density of the
1912 tree ring was the lowest in investigated tree-ring-density chronologies from the mountain spruce forests
of Central Europe (Kyncl et al., 1990). On a European scale, the autumn of 1912 was the coldest autumn
during the period 1500–2004 (Xoplaki et al., 2005). Regarding the Northern Hemisphere profile, expected
cooling appeared immediately in 1912 similar to that in Central Europe, but from the second post-eruption
year, temperature anomalies were only very positive (Figure 8).
Central Europe could be further influenced by effects of smaller volcanic eruptions from closer geographic
areas such as Iceland and Italy (see Table III). A total sample of nine eruptions was examined. Composite
profiles were not created this time, owing to the limited number of samples (only two samples from Italy)
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
450 J. PÍSEK AND R. BRÁZDIL

Katmai 1912
0.6
(a) Northern Hemisphere
0.4

0.2

-0.2
Temperature (˚C)

-0.4
-60 -48 -36 -24 -12 0 12 24 36 48 60

4
(b)

-2

-4
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Prague Kremsmünster Vienna Germany

Figure 8. Comparison of monthly mean surface air temperature profiles (a) for Northern Hemisphere and (b) for Prague-Klementinum,
Kremsmünster, Vienna-Hohe Warte and composite from German stations around the time of Katmai eruption in 1912. Values are
expressed as anomalies relative to the 5-year pre-eruption period. The month zero – June 1912

and the necessity of further discarding some eruptions in the case of Icelandic eruptions because of possible
bias, as was explained above. Eruptions were thus examined individually. Again, it was very difficult to track
any common features.
The most striking case concerns the so-called Lakagı́gar eruption in Iceland in 1783. On 8 June, a fissure
opened up along the Laki crater-row and until February 1784, it had been followed by the largest terrestrial
lava flow of the last millennium (Thordarson and Self, 1993, 2003; for atmospheric chemistry modelling,
see Stevenson et al., 2003). Moreover, Mt Asama (Japan) erupted at the beginning of August 1783 with
VEI = 4. No significant cooling is presented around the time of the eruption (Figure 9). The Lakagı́gar
eruption was followed by a slight warming during the immediately following months (summer of 1783);
however, during the first post-eruption winter, temperatures dropped significantly (1783/84 – the sixth coldest
in Prague-Klementinum). After a short return to the average temperatures of the pre-eruption period in the
second summer (1784), the temperature again decreased significantly during the spring of 1785, which was
the coldest spring for instrumental records since 1775 in Prague-Klementinum. The situation was similar
with the coldest measured autumn, 1786 (Brázdil et al., 2003), which followed a cool summer. Significant
cooling was thus observed during four consequent post-eruption years. The same course can be followed in
other series from Central Europe as well. The spring of 1785 was found to be the coldest spring in Europe
during the period 1500–2004 (Xoplaki et al., 2005). Unfortunately, data for the Northern Hemisphere are
not available from this period in order to find out if this phenomenon had only local or regional character
or could have been noticed on a larger scale as well. However, similar effects of the Iceland eruption
on temperature data were also detected in Vermont (Sigurdsson, 1982) and in Central England (Manley,
1974).
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 451

Lakagígar 1783
4

2
Temperature (°C)

-2

-4

-6
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Prague Kremsmünster Vienna Germany

Figure 9. Comparison of monthly mean surface air temperature profiles for Prague-Klementinum, Kremsmünster, Vienna-Hohe Warte
and composite from German stations around the time of Lakagı́gar eruption in 1783. Values are expressed as anomalies relative to the
5-year pre-eruption period. The month zero – June 1783

On the other hand, further climatological evidence was collected in Europe for the summer of 1783
immediately after the Lakagı́gar eruption (see e.g. Camuffo and Enzi, 1994; Stothers, 1996; Demarée
et al., 1998; Grattan and Pyatt, 1999; Sadler and Grattan, 1999; Demarée and Ogilvie, 2001; Brázdil et al.,
2003; Highwood and Stevenson, 2003). As in other European countries, in Central Europe, mainly dry fog,
remarkable optical phenomena and heavy thunderstorms immediately after the eruption were documented.
Dry fog after the Lakagı́gar eruption appeared at Prague first on 16 June (Strnadt, 1785). At Noviny pod
Ralskem, even fog smelling of sulphur from 18 June to 18 July was observed by the teacher Anton Lehmann
(Wiechowsky, 1928). An example of contemporary perception of this event can be taken from memoirs of
Jioı́ Vrbas from Pı́sečné (Paměti starých pı́smáků moravských, 1916): ‘In 1783 from John’s holiday [24 June]
to the harvest every day and night there were such dense fogs everywhere that no one could see anything but
a small piece of the world. Also the sun and the moon were changing. The sun rose every day blood red and
the moon was like a black sack.’ A similar report comes from the chronicle of Antonı́n Kodytek, a teacher
at Kunvald (Nezbeda and Šůla, 1970): ‘. . . in summer there was such a heat that if there were not for the
unusual fog which shaded the sun, perhaps everything would have been burnt by the sun’s heat. Because the
rising morning sun could not be seen due to the fires and then from six to nine o’clock the sun looked like a
red hot iron ball, then from nine to three or four o’clock it shone more intensely, but looked sad, which made
the people wonder. And people at that time much considered, what it would cause or what it might mean. . . ’
News of the Lakagı́gar eruption did not reach Europe until 1 September when it was brought to Copenhagen
from Iceland, carried by the vessels of the Danish trading monopoly. The connection between volcanic
eruptions and climate seems to have been noted first by Benjamin Franklin 2 years later (Demarée and
Ogilvie, 2001). Owing to these perceptions at that time about casual links, it is not surprising that a chronicler
Johann Josef Langer at Rýmařov related the outstanding summer weather of 1783 to the earthquake in
Messina (Italy) from 5 February of that year (Tutsch, 1914): ‘During the following summer, outstanding fogs
were observed in the European countries, soon here and soon there, also in the surroundings. The cause of fog
was adjudicated to this [Messina] earthquake.’
Grattan et al. (2003), studying the human health response to volcanogenic pollution and dry fog events,
found a significant increase in the national death rate in England coincident with the early phases of this
eruption. Witham and Oppenheimer (2005) speak, after the Laki eruption, about two distinct mortality crises
in August–September 1783 and January–February 1784, which together accounted for ∼19 700 extra deaths.
The East of England was the worst affected region. They speculate that fine acid aerosol and/or gases in the
volcanic haze may also have contributed, after a very hot July, to the first mortality wave. In the case of winter
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
452 J. PÍSEK AND R. BRÁZDIL

mortality, the severe cold of January 1784, after people had been weakened by the summer phenomena, might
be responsible. Grattan et al. (2005) showed that mortality crisis after eruption was also found in France and
in the Low Countries. Data for selected French parishes indicate an anomalous peak of deaths centered on
September and October 1783. They argue that this situation was similar to modern severe air pollution crises
with critical sulphur concentration (for possible implications of these effects, see Courtillot, 2005).
Another important eruption for climatic patterns in Central Europe was the Katla eruption in Iceland in
October 1918 (Figure 10). A cold spring and summer were already observed in the first post-eruption year
1919 (in the Prague-Klementinum chronology, the 18th and 9th coldest ones, respectively). Also, autumn of
1920 was very cold (the 11th coldest).
If for every season of Prague temperatures only 10% of the coldest seasons is always taken into
consideration, every Icelandic or Italian volcanic eruption (see Table III) was followed by such outstanding
season (with the exception of the Hekla eruption from 1947). Fifteen extremely cold seasons were found with
a predominant concentration in autumn (7 cases), followed by winter and spring (3 cases each), and lastly
summer (2 cases).

4. DISCUSSION AND CONCLUSIONS

It should be noted that caution must be applied when interpreting the results of the analyses presented here.
This is true especially when evaluating series on a global or hemispheric scale, where the final course may
have been contaminated by the effects of other large-scale atmospheric and oceanic processes, such as the

Katla 1918
0.6
(a) Northern Hemisphere
0.4

0.2

-0.2
Temperature (°C)

-0.4
-60 -48 -36 -24 -12 0 12 24 36 48 60

4
(b)

-2

-4
-60 -48 -36 -24 -12 0 12 24 36 48 60
before eruption Month after eruption

Prague Kremsmünster Vienna Germany

Figure 10. Comparison of monthly mean surface air temperature profiles (a) for Northern Hemisphere and (b) for Prague-Klementinum,
Kremsmünster, Vienna-Hohe Warte and composite from German stations around the time of Katla eruption in 1918. Values are expressed
as anomalies relative to the 5-year pre-eruption period. The month zero – October 1918

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 453

El Niño-Southern Oscillation (ENSO). Moreover, Kirchner and Graf (1995) showed that episodes of high-
magnitude volcanism are often followed by well-defined El Niño events. Adams et al. (2003) demonstrated
a significant multi-year El Niño-like response to explosive tropical volcanic forcing from AD 1649 to the
present. Roughly, a doubling of the probability of an El Niño event was shown for the winter following the
volcanic eruption. ENSO and volcanic aerosols may also influence inter-annual fluctuation of atmospheric CO2
(Reichenau and Esser, 2003). Increasing photosynthetic carbon uptake due to the enhanced diffuse fraction
of the incoming radiation related to volcanic aerosols may sometimes cancel out or even over-compensate
for the carbon release caused by ENSO effects.
ENSO can result in global warming on a timescale similar to the effects of volcanic eruptions (Cress and
Schönwiese, 1992). ENSO-related effects could therefore significantly obliterate the volcanic signal in the
series. However, a possible removal of the ENSO signal is not straightforward and its execution could possibly
introduce even bigger flaws into the series (Kelly et al., 1996). Since the effect of ENSO on the climate in
the area of Central Europe is rather disputable (see e.g. Fraedrich, 1990, 1994; Fraedrich and Müller, 1992;
Brázdil and Bı́l, 2000), the removal of ENSO signal has not been executed, and all results were assessed
bearing this fact in mind.
Local and regional climatic series that tend to be very variable are strongly biased by different local
effects and circulation patterns. It means that in series such as those in Central Europe it is more difficult
to find any clear volcanic signal than it is in the case of global and hemispheric temperature series with
typical, statistically significant (0.05 level) warming during the first post-eruption winter and cooling in the
two post-eruption years. For example, for Central Europe, only 3 months of the eruption year (February,
August and December) indicated significant temperature departures during the post-eruptive sequences for
the eight tropical eruptions analysed by Jones et al. (2004). On the other hand, dynamically caused warm
anomalies in post-eruption winters occur throughout northern Eurasia besides anomalies in other parts of the
world (Shindell et al., 2004). But the standard deviation of the response is larger than the mean signal nearly
everywhere. It means that the anomaly following a single eruption is unlikely to be representative of the
mean.
Moreover, as discussed by Sadler and Grattan (1999), stratospheric and tropospheric warming after eruption
is also possible, which is in contradiction to generally expected cooling. This was, e.g. the case of the Hekla
eruption in Iceland on March 1947 which was not followed in the Czech Republic by any real cool season
but by a very warm summer and autumn in 1947 (each the 12th warmest in the Prague-Klementinum series),
and spring of 1948 (the 7th warmest one), instead. Similarly, the eruption of Etna (Italy) in July 1787 was
accompanied with only one cold winter 1788/89 (the 11th coldest), but with three very warm seasons: summer
of 1787 (the 6th warmest), autumn of 1787 (the 8th warmest) and summer of 1788 (the 8th warmest). Brázdil
et al. (2001) even found a statistically significant positive correlation between Czech Republic 1961–1995
temperatures in the months of June, July and August and the whole summer with series of SAOD for the
Northern Hemisphere (correlation coefficients were consequently 0.35, 0.34, 0.38 and 0.48, respectively). It
was mainly due to a surprising coincidence of the high Czech summer temperatures in 1982–1983 and 1992
with the highest values of SAOD after El Chichón (1982) and Mt Pinatubo (1991) eruptions. Although SAOD
characterizes atmospheric transmission for wavelength 550 nm only (Sato et al., 1993), the positive correlation
can hardly be connected with a radiation effect of volcanic eruptions. On the other hand, 1982–1983 was a
time of a very strong El Niño event (Diaz and Markgraf, 1992) and also the year 1992 corresponds to another
such case in 1991–1994 (Klein et al., 1999; Chiang and Sobel, 2002).
Another open question is the possible relation between circulation patterns and volcanic events (see e.g.
Graf et al., 1994; Stenchikov et al., 2002, 2004, Shindell et al., 2004). While summer temperature anomalies
related to large volcanic eruptions are largely radiative, the dynamical response to eruptions is most important
for two following Northern Hemisphere winters. In this case, anomalies strongly resemble AO/NAM (Arctic
Oscillation/Northern Annual Mode) and NAO in the Atlantic–Eurasian sector. A strengthening of polar
vortex forces a positive phase of the AO. As shown by Shindell et al. (2004), GCM simulations produce
enhanced westerlies, driven by the solar heating because of stratospheric volcanic aerosols, propagating from
the lowermost stratosphere down to the troposphere via interactions with planetary waves. Westerly QBO
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
454 J. PÍSEK AND R. BRÁZDIL

Table IV. Correlation coefficients between the monthly Prague-Klementinum temperatures and circulation indices (NAOI,
CEZI) during 5 years prior eruption (a) and after eruption (b). Correlations significant for 0.05 level are in bold. For the
list of eruptions used see Tables II and III

Year of NAOI
eruption
Year Winter half-year Summer half-year
(October–March) (April–September)
a b a b a b

Tropical eruptions
1815 0.520 0.537 0.317 0.684 0.768 0.352
1883 0.483 0.361 0.589 0.415 0.201 0.361
1902 0.278 0.408 0.346 0.495 0.193 0.323
1963 0.475 0.297 0.625 0.386 0.173 0.123
1982 0.094 0.444 0.099 0.683 0.105 0.226
1991 0.443 0.321 0.606 0.398 0.194 0.205
Katmai
1912 0.249 0.340 0.359 0.408 0.116 0.266
Icelandic and Italian eruptions
1783 0.632 0.367 0.735 0.512 0.568 −0.063
1787 0.362 0.387 0.544 0.346 −0.017 0.445
1845 0.321 0.315 0.091 0.379 0.637 0.150
1873 0.557 0.417 0.608 0.376 0.559 0.505
1875 0.457 0.450 0.538 0.496 0.538 0.413
1903 0.319 0.279 0.383 0.303 0.225 0.249
1906 0.378 0.134 0.430 0.236 0.345 0.009
1918 0.373 0.282 0.500 0.347 0.204 0.215
1947 0.428 0.287 0.620 0.275 0.111 0.282

Year of CEZI
eruption
Year Winter half-year Summer half-year
(October–March) (April–September)
a b a b a b

Tropical eruptions
1815 0.562 0.463 0.675 0.602 0.389 0.024
1883 0.438 0.369 0.525 0.466 0.213 0.239
1902 0.155 0.407 0.306 0.530 −0.160 0.248
1963 0.548 0.302 0.756 0.440 0.144 −0.015
1982 0.315 0.479 0.503 0.746 0.079 −0.007
1991 0.459 0.429 0.626 0.805 0.021 −0.129
Katmai
1912 0.260 0.338 0.598 0.471 0.003 0.040
Icelandic and Italian eruptions
1787 0.346 0.467 0.631 0.595 −0.245 0.221
1845 0.351 0.444 0.456 0.651 0.172 −0.109
1873 0.465 0.536 0.654 0.589 0.131 0.486
1875 0.514 0.247 0.575 0.310 0.446 0.197
1903 0.358 0.313 0.518 0.598 −0.088 0.318
1906 0.195 0.220 0.439 0.525 −0.223 0.037
1918 0.280 0.356 0.467 0.598 −0.068 0.084
1947 0.600 0.290 0.810 0.459 0.096 −0.008

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 455

(quasi-biennial oscillation) phase in the lower stratosphere results in an enhancement of the aerosol effect on
the AO (Stenchikov et al., 2004).
In the European region, important circulation mode is expressed by NAO with significant influence on
temperature patterns during wintertime. Mainly in Northern Europe winters with positive NAO indices are
warmer and vice versa. Jones et al. (2003), despite of lack of significance in NAO related to volcanic eruptions,
found that positive departures of the NAO dominate in post-eruption years for tropical and higher-latitude
eruptions as well. For tropical eruptions, this clearly occurs during the first and fourth winters after eruptions
and in much lesser extent in the second one. Significance was reached in May of the eruption year and January
and August of the following year for tropical eruptions, in March, May and December of the eruption year
and June of the following year for higher-latitude eruptions.
It was shown for the Czech Republic that December–March temperatures are closely related to fluctuations
in NAOI (in some cases, better for CEZI) when in other months, this relation is much weaker or even not
statistically significant (Brázdil et al., 1999; Brázdil et al., 2001). If there is any significant change between
pre-eruption and post-eruption years, it should be reflected in relation to NAOI temperatures. For a study of
these possible links, the calculation of correlations between Prague-Klementinum temperature profiles and
profiles for NAOI and CEZI in 5 years before and 5 years after eruption has been carried out (Table IV). Their
relation is much stronger for annual and winter half-year data (with significant correlation coefficients) than for
the summer half-year months. It fits with a much stronger link between circulation patterns and temperatures
in the winter half-year because of the important role of airflow direction compared to the summer half-year
when temperatures are significantly influenced also by radiation.
From Table IV, it follows that for NAOI there is no clear difference between pre-eruption and post-eruption
years for large tropical eruptions. On the other hand, for Icelandic and Italian eruptions, it has been documented
that with the exception of the 1787 Etna eruption in all cases the link between temperatures and circulation
indices was stronger during the period prior to eruption than after it. The same is valid also for the months
of the winter half-year (with the exception of Hekla in 1845). This tendency is more weakly expressed in the
months of the summer half-year. For another studied circulation characteristic of CEZI, the relations are not so
strongly expressed as was the case of NAOI. It can therefore be concluded that Icelandic and Italian volcanic
eruptions do weaken the link between NAOI and temperatures in Central Europe. Physical mechanisms of
this relation have to be further investigated.
Finally, it is also necessary to stress the importance of documentary evidence for describing climatic
responses of volcanic eruptions in the early instrumental or pre-instrumental period. It has documented
mainly the example of the Lakagı́gar eruption in 1783, where this type of evidence contains useful relevant
information. For example, a report of burgesses of the Prague New Town already described the Etna eruption
in Italy in 1536 (Zilynskyj, 1984): ‘In the same year, Mount Etna in Sicily that burnt for many years, with its
fire extended as a sea and caused great damage, destroyed nearby towns, small towns, villages and fruits of the
Earth.’ Another source quotes even a raining sulphur as was the case in Prague and Žatec from 31 May 1550
according to the ‘Historical calendar’ by Lupáč (1584) or a chronicler Marek Bydžovský of Florentino (Kolár,
1987): ‘On the same year on May 31st, in Prague and Žatec large and small pieces of sulphur, from greater part
four-sided, rained from the sky. People picked up them and used them instead of other sulphur, but after setting
fire it stank very much.’ For this year, the eruption of Vulcano in Italy is described (www.volcano.si.edu/).
It is important from the point of view of perception of these volcanic phenomena, that documentary evidence
also includes explanation of their possible causes showing contemporary knowledge in this field (such as an
earthquake in 1783 – see e.g. Tutsch, 1914). More than hundred years later, in 1883, optical phenomena were
observed during November–December 1883 and in the beginning of January 1884, and they were already
connected to the Krakatau eruption. After sunset, above the western horizon appeared a nice red coloured sky
(‘Weltbrand’ – world fire) for half an hour every day, reminiscent of the aurora borealis. The author explained
it already that volcanic dust was injected into the atmosphere (see Tille, 1905).
As follows from discussion and actual results, the most significant climatic impacts of volcanic eruptions in
Central Europe on the regional as well as on the local scale can be observed after eruptions in geographically
closer areas such as Iceland or Italy. On the other hand, impacts of more distant volcanic eruptions such as in
Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
456 J. PÍSEK AND R. BRÁZDIL

the tropics, even with stronger volcanic explosivity, are biased by many local and regional effects, and they
seem to be, in many cases, negligible (see also Jones et al., 2004).

ACKNOWLEDGEMENTS

The authors would like to thank the Grant Agency of the Czech Republic for giving financial support, Grant
No. 205/01/1067, and the research project MŠM0021622412 (INCHEMBIOL); John P. Grattan, Aberystwyth,
for valuable corrections in the manuscript; Alan Robock, New Jersey, Gaston Demarée, Brussels, for valuable
discussions about climatic effects of volcanic eruptions; Karel Vanı́ček, Hradec Králové, for providing global
radiation data from the Czech Hydrometeorological Institute, station Hradec Králové; Marian Ostrožlı́k,
Bratislava, for providing global radiation data from Geophysical Institute of Slovak Academy of Sciences,
station Skalnaté Pleso; Jarmila Macková, Brno, for drawing Figure 1; Katherine Miller and Melissa Blackie,
Toronto, for English style corrections; two anonymous reviewers whose constructive comments helped
improve the original manuscript.

REFERENCES
Adams JB, Mann ME, Ammann CM. 2003. Proxy evidence for an El Niño-like response to volcanic forcing. Nature 426: 274–278.
DOI: 10.1038/nature02101.
Adeyefa ZD, Holmgren B, Adedokun JA. 2000. Spectral solar irradiance in northern Scandinavia before and after Pinatubo. Atmósfera
13: 133–146.
Ammann CM, Meehl GA, Washington WM, Zender CS. 2003. A monthly and latitudinally varying volcanic forcing dataset in
simulations of 20th century climate. Geophysical Research Letters 30: 59-1–59-4, DOI: 1029/2003GL016875.
Angell JK. 1997a. Estimated impact of Agung, El Chichón, and Pinatubo volcanic eruptions on global and regional total ozone after
adjustment for the QBO. Geophysical Research Letters 24: 647–650.
Angell JK. 1997b. Stratospheric warming due to Agung, El Chichón, and Pinatubo taking into account the quasi-biennial Oscillation.
Journal of Geophysical Research 102: 9479–9485.
Auer I, Böhm R, Schöner W. 2001. Austrian Long-term Climate 1767–2000 – Multiple Instrumental Climate Time Series from Central
Europe. Österreichische Beiträge zu Meteorologie und Geophysik 25: Wien.
Bachmann J. 1912. Aufzeichnungen eines Alt-Leitmeritzers. Mitteilungen des Vereines für Geschichte der Deutschen in Böhmen 50:
107–131, 254–281.
Blau J. 1908. Ein Kapitel vom Getreidepreis. Mitteilungen des Vereines für die Geschichte der Deutschen in Böhmen 47: 277–283.
Blumthaler M, Ambach W. 1994. Changes in solar radiation fluxes after the Pinatubo eruption. Tellus 46B: 76–78.
Bradley RS. 1988. The explosive volcanic eruption signal in northern hemisphere continental temperature records. Climatic Change 12:
221–243.
Bradley RS, Jones PD. 1992. Records of explosive volcanic eruptions over the last 500 years. In Climate Since AD 1500, Bradley RS,
Jones PD (eds). Routledge: London.
Brázdil R, Budı́ková M. 1999. An urban bias in air temperature fluctuations at the Klementinum, Prague, the Czech Republic.
Atmospheric Environment 33: 4211–4217.
Brázdil R, Bı́l M. 2000. ENSO and its effects in the mean fields of air pressure, air temperature and precipitation in Europe in the
twentieth century. In El Niño–Southern Oscillation and its Global Impacts, Nkemdirim LC (ed). A Publication of the Commission
on Climatology of the International Geographical Union: Bonn.
Brázdil R, Štěpánek P, Květoň V. 2001. Temperature series of the Czech Republic and its relation to Northern Hemisphere temperatures
in the period 1961–1999. In Detecting and Modelling Regional Climate Change, Brunet India M, López Bonillo D (eds). Springer:
Berlin, Heidelberg, New York, Barcelona, Hong Kong, London, Milan, Paris, Tokyo.
Brázdil R, Valášek H, Macková J. 2003. Climate in the Czech Lands during the 1780s in light of the daily weather records of parson
Karel Bernard Hein of Hodonice (south-western Moravia): Comparison of documentary and instrumental data. Climatic Change 60:
297–327.
Brázdil R, Štekl J, Budı́ková M, Dobrovolný P, Fišák J, Kolář M, Prošek P, Sokol Z, Stěpánek P, Štěpánková P, Zacharov P. 1999.
Klimatické poměry Milešovky, Academia: Praha.
Briffa KR, Jones PD. 1992. The climate of Europe during the 1810s with special reference to 1816. In The Year without a Summer?
World Climate in 1816, Harington CR (ed). Canadian Museum of Nature: Ottawa.
Briffa KR, Jones PD, Schweingruber FH, Osborn TJ. 1998. Influence of volcanic eruptions on Northern Hemisphere summer temperature
over the past 600 years. Nature 393: 450–455.
Brůžek V. 1992. Major volcanic eruptions in the nineteenth and twentieth centuries and temperatures in Central Europe. In The Year
without a Summer? World Climate in 1816, Harington CR (ed). Canadian Museum of Nature: Ottawa.
Camuffo D, Enzi S. 1994. Chronology of ‘Dry Fogs’ in Italy, 1374–1891. Theoretical and Applied Climatology 50: 31–33.
Camuffo D, Enzi S. 1995. Impact of the clouds of volcanic aerosols in Italy during the last 7 centuries. Natural Hazards 11: 135–161.
Chenoweth M. 2001. Two major volcanic cooling episodes derived from global marine air temperature, A.D. 1807–1827. Geophysical
Research Letters 28: 2963–2966.
Chiang JC, Sobel A. 2002. Tropical tropospheric temperature variations caused by ENSO and their influence on the remote tropical
climate. Journal of Climate 15: 2616–2631.
Conrad V, Pollak LW. 1962. Methods in Climatology. Harvard University Press: Cambridge.

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 457

Courtillot V. 2005. New evidence for massive pollution and mortality in 1783–1784 may have bearing on global change and mass
extinctions. Comptes rendus Geoscience 337: 635–637. DOI: 10.1016/j.crte.2005.03.001.
Cress A, Schönwiese C-D. 1992. Vulkanische Einflüsse auf die Bodennahe und Stratosphärische Lufttemperatur der Erde. Berichte des
Instituts für Meteorologie und Geophysik der Universität Frankfurt/Main 82: Frankfurt am Main.
Dai J, Mosley-Thompson E, Thompson LG. 1991. Ice core evidence for an explosive tropical volcanic eruption 6 years preceding
Tambora. Journal of Geophysical Research 96: 17361–17366.
Dawson AG, Hickey K, McKenna J, Foster IDL. 1997. A 200-year record of gale frequency, Edinburgh, Scotland: possible link with
high-magnitude volcanic eruptions. The Holocene 7: 337–341.
Demarée GR, Ogilvie AEJ. 2001. Bons Baisers d’Islande: Climatic, environmental, and human dimensions impacts of the Lakagı́gar
eruption (1783–1784) in Iceland. In History and Climate. Memories of the Future, Jones PD, Ogilvie AEJ, Davies TD, Briffa KR
(eds). Kluwer Academic/Plenum Publishers: New York, Boston, Dordrecht, London, Moscow.
Demarée GR, Ogilvie AEJ, Zhang D. 1998. Further documentary evidence of Northern Hemispheric coverage of the great dry fog of
1783. Climatic Change 39: 727–730.
Diaz HF, Markgraf V (eds). 1992. In El Niño. Historical and Paleoclimatic Aspects of the Southern Oscillation. Cambridge University
Press: Cambridge.
Dunn J. 2004. The influence of volcanic activity on large-scale atmospheric processes: a discussion. Weather 59: 46–49.
Folland CK, Rayner NA, Brown SJ, Smith TM, Shen SSP, Parker DE, Macadam I, Jones PD, Jones RN, Nicholls N, Sexton DMH.
2001. Global temperature change and its uncertainties since 1861. Geophysical Research Letters 28: 2621–2624.
Fraedrich K. 1990. European Grosswetter during the warm and cold extremes of the El Niño/Southern Oscillation. International Journal
of Climatology 10: 21–31.
Fraedrich K. 1994. An ENSO impact on Europe? a review. Tellus 46A: 541–552.
Fraedrich K, Müller K. 1992. Climate anomalies in Europe associated with ENSO extremes. International Journal of Climatology 12:
25–31.
Graf HF, Perlwitz J, Kirchner I. 1994. Northern hemisphere tropospheric mid-latitude circulation after violent volcanic eruptions.
Contribution to Atmospheric Physics 67: 3–13.
Grattan JP, Pyatt FB. 1999. Volcanic eruptions dry fogs and the European palaeoenvironmental record: localised phenomena or
hemispheric impacts? Global and Planetary Change 21: 173–179.
Grattan J, Sadler J. 1999. Regional warming of the lower atmosphere in the wake of volcanic eruptions: the role of the Laki fissure
eruption in the hot summer of 1783. In Volcanoes in the Quaternary, Firth CR, McGuire WG (eds). Geological Society: London.
Grattan JP, Durand M, Taylor S. 2003. Illness and elevated human mortality in Europe coincident with the Laki fissure eruption. In
Volcanic Degassing, Oppenheimer C, Pyle DM, Barclay J (eds). Geological Society: London.
Grattan J, Rabartin R, Self S, Thordarson T. 2005. Volcanic air pollution and mortality in France 1783–1784. Comptes rendus
Geoscience 337: 641–651. DOI: 10.1016/j.crte.2005.01.013.
Groisman P. 1992. Possible regional climate consequences of the Pinatubo eruption: an empirical approach. Geophysical Research
Letters 19: 1603–1606.
Gu L, Baldocchi DD, Wofsy SC, Munger JW, Michalsky JJ, Urbanski SP, Boden TA. 2003. Response of a deciduous forest to the
Mount Pinatubo eruption: Enhanced photosynthesis. Science 299: 2035–2038.
Hanzlı́k S. 1918. Die Schwankungen der atmosphärischen Dursichtigkeit nach den Sonnenscheinregistrierungen in Prag. Meteorologische
Zeitschrift 35: 52–53.
Harington CR (ed). 1992. In The Year Without a Summer? World Climate in 1816. Canadian Museum of Nature: Ottawa.
Highwood EJ, Stevenson DS. 2003. Atmospheric impact of the 1783–1784 Laki Eruption: Part II Climatic effect of sulphate aerosol.
Atmospheric Chemistry and Physics 3: 1177–1189.
Hurrell JW. 1995. Decadal trends in the North Atlantic Oscillation regional temperatures and precipitation. Science 269: 676–679.
Jacobeit J, Beck C, Philipp A. 1998. Annual and Decadal Variability in Climate in Europe. Würzburger Geographische Manuskripte 43,
Würzburg.
Jones PD, Moberg A, Osborn TJ, Briffa KR. 2003. Surface climate responses to explosive volcanic eruptions seen in long European
temperature records in mid-to-high latitude tree-ring density around the Northern Hemisphere. In Volcanism and the Earth’s
Atmosphere, Geophysical Monograph 139 . Robock A, Oppenheimer C (eds.) American Geophysical Union, Washington, D.C.
239–254, DOI: 10.1029/139GM15.
Kelly PM, Jones PD, Jia PQ. 1996. The spatial response of the climate system to explosive volcanic eruptions. International Journal
of Climatology 16: 537–550.
Kirchner I, Graf HF. 1995. Volcanoes and El Niño: signal separation in Northern Hemisphere winter. Climate Dynamics 11: 341–358.
Kirchner I, Stenchikov GL, Graf HF, Robock A, Antuna JC. 1999. Climate model simulation of winter warming and summer cooling
following the 1991 Mount Pinatubo volcanic eruption. Journal of Geophysical Research 104: 19039–19055.
Klein SA, Soden BJ, Lau N. 1999. Remote sea surface temperature variations during ENSO: evidence for a tropical atmospheric bridge.
Journal of Climate 12: 917–932.
Kolár J. 1987. Marek Bydžovský z Florentina, Svět za třı́ českých králů. Výbor z kronikářoských zápisů o letech 1526–1596. Svoboda:
Praha.
Komárek F. 1911. Paměti panstvı́ a farnı́ osady budenické. Nákladem vlastnı́m: Praha.
Kreuzinger E. 1862. Chronik der alten und neuern Zeit Troppau’s, oder Troppau und seine Merkwürdigkeiten. Im Selbstverlage des
Herausgebers: Troppau.
Kyncl J, Dobrý J, Munzar J, Sarajishvili KG. 1990. Tree-ring structure response of conifers in Europe to weather conditions in 1912
(with regard to the volcano Katmai eruption). In Climatic Change in the Historical and the Instrumental Periods, Brázdil R (ed).
Masaryk University: Brno.
Lamb PJ, Peppler RA. 1987. North Atlantic Oscillation: concept and an application. Bulletin of American Meteorological Society 68:
1218–1225.
Lenhart JJ. 1840. Carlsbads Memorabilien vom Jahre 1325 bis 1839. Gottlieb Haase Söhne: Prag.

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
458 J. PÍSEK AND R. BRÁZDIL

Lupáč P. 1584. Rerum Boemicarum Ephemeris sive Kalendarium historicum, Ex reconditis veterum annalium monumentis crutum.
Authore M. Procopio Lupacio [de] Hlavaczov aeo, Pragensi. Opus nunc primum in lucem editum, una cum coronide ac locuplere
personarum et rerum memorabilium indice. In idem kalendarium. Eruta dum patriae monumenta Lupacius edit, et bonus est civis,
doctus et historicus. Suscitat, amplificat, manifestat, promovet, ornat, maiores, patriam, tempora, gesta, duces. M. Bern. Sturmii.
Pragae, Anno M.D.LXXXIIII [1584].
Luterbacher J, Dietrich D, Xoplaki E, Grosjean M, Wanner H. 2004. European seasonal and annual temperature variability, trends and
extremes since 1500. Science 303: 1499–1503.
Luterbacher J, Xoplaki E, Dietrich D, Jones PD, Davies TD, Portis D, Gonzalez-Rouco JF, von Storch H, Gyalistras D, Casty C,
Wanner H. 2002. Extending North Atlantic Oscillation reconstructions back to 1500. Atmospheric Science Letters 2: 114–124.
Manley G. 1974. Central England temperatures: monthly means 1659 to 1973. Quarterly Journal of the Royal Meteorological Society
100: 389–405.
Mass C, Robock A. 1982. The short-term influence of the Mount St. Helens volcanic eruption on surface temperature in the northwest
United States. Monthly Weather Review 110: 614–622.
McCormick MP, Thomason LW, Trepte CR. 1995. Atmospheric effects of the Mt. Pinatubo eruption. Nature 373: 399–404.
McCracken MC, Luther FM. 1984. Preliminary estimate of the radiative and climatic effects of the El Chichon eruption. Geofisica
Internacional 23: 385–401.
Mosley-Thompson E, Mashiotta TA, Thompson LG. 2003. Ice core records of late Holocene volcanism: Current and future
contributions from the Greenland PARCA cores. In Volcanism and the Earth’s Atmosphere, Geophysical Monograph 139 . Robock A,
Oppenheimer C (eds.) American Geophysical Union, Washington, D.C. 153–164, DOI: 10.1029/139GM09.
Nezbeda V, Šůla J. 1970. Kunvaldská kronika Antonı́na Kodytka 1740–1786. Nákladem Orlického muzea: Choceň.
Oppenheimer C. 2003. Climatic, environmental and human consequences of the largest known historic eruption: Tambora volcano
(Indonesia) 1815. Progress in Physical Geography 27: 230–259.
Ostrožlı́k M. 2002. Time variability of global solar radiation in high-mountain regions. Contrib. Geophys. Geodesy 32: 277–289.
Paměti starých pı́smáků moravských. 1916. Nákladem Selského archivu: Velké Meziřı́čı́.
Parker DE, Brownscombe JKL. 1983. Stratospheric warming following the El Chichon volcanic eruption. Nature 301: 406–408.
Parker DE, Wilson H, Jones PD, Christy JR, Folland CK. 1996. The impact of Mount Pinatubo on worldwide temperatures. International
Journal of Climatology 16: 487–497.
Rapp J. 2000. Konzeption, Problematik und Ergebnisse klimatologischer Trendanalysen für Europa und Deutschland. Berichte des
Deutschen Wetterdienstes 212: Offenbach am Main.
Reichenau TG, Esser G. 2003. Is interannual fluctuation of atmospheric CO2 dominated by combined effects of ENSO and volcanic
aerosols? Global Biogeochemical Cycles 17: 5-1–5-9. DOI: 10.1029/2002 GB002025.
Robek A. 1974. Lidové kronikářstvı́ na Kralupsku a Mělnicku. Edice lidových kronikářkých textů. Ústav pro etnografii a folkloristiku
ČSAV: Praha.
Robock A. 1981. The Mount St. Helens volcanic eruption on 18 May 1980: minimal climatic effect. Science 212: 1383–1384.
Robock A. 1984. Climate model simulations of the effects of the El Chichon eruption. Geofisica Internacional 23: 403–414.
Robock A. 2000. Volcanic eruptions and climate. Reviews of Geophysics 38: 191–219.
Robock A, Mao J. 1992. Winter warming from large volcanic eruptions. Geophysical Research Letters 19: 2405–2408.
Robock A, Mao J. 1995. The volcanic signal in surface temperature observations. Journal of Climate 8: 1086–1103.
Robock A, Mass C. 1982. The Mount St. Helens volcanic eruption of 18 May 1980: large short-term surface temperature effects. Science
216: 628–630.
Roubic A. 1987. Kronika rychtářů Urbanı́ka a Hořı́nka z Velké Bystřice z let 1789–1848. Okresnı́ archı́v v Olomouci 1987: 213–238.
Sadler JP, Grattan JP. 1999. Volcanoes as agents of past environmental change. Global and Planetary Change 21: 181–196.
Sato M, Hansen JE, McCormick MP, Pollack JB. 1993. Stratospheric aerosol optical depth, 1850–1990. Journal of Geophysical
Research 98: 22,987–22,994.
Sear CB, Kelly PM, Jones PD, Goodess CM. 1987. Global surface-temperature responses to major volcanic eruptions. Nature 330:
365–367.
Self S, King AJ. 1996. Petrology and sulfur and chlorine emissions of the 1963 eruption of Gunung Agung, Bali, Indonesia. Bulletin
of Volcanology 58: 263–285.
Shindell DT, Schmidt GA, Mann ME, Faluvegi G. 2004. Dynamic winter climate response to large tropical volcanic eruptions since
1600. Journal of Geophysical Research 109: D05104. DOI:10.1029/2003JD004151.
Sigurdsson H. 1982. Volcanic pollution and climate – the 1783 Laki eruption. American Geophysical Union, EOS Transactions 10:
601–602.
Sigurdsson HS, Carey S. 1992. The eruption of Tambora in 1815: Environmental effects and eruption dynamics. In The Year without a
Summer? World Climate in 1816, Harington CR (ed). Canadian Museum of Nature: Ottawa.
Stenchikov G, Hamilton K, Robock A, Ramaswamy V, Schwarzkopf MD. 2004. Arctic Oscillation response to the 1991 Pinatubo
eruption in the SKYHI general circulation model with a realistic quasi-biennial oscillation. Journal of Geophysical Research 109:
D03112. DOI:10.1029/2003JD003699.
Stenchikov G, Robock A, Ramaswamy V, Schwarzkopf MD, Hamilton K, Ramachandran S. 2002. Arctic oscillation response to the
1991 Mount Pinatubo eruption: effects of volcanic aerosols and ozone depletion. Journal of Geophysical Research 107: 4803.
DOI:10.1029/2002JD002090.
Stevenson DS, Johnson CE, Highwood EJ, Gauci V, Collins WJ, Derwent RG. 2003. Atmospheric impact of the 1783–1784 Laki
eruption: Part I Chemistry modeling. Atmospheric Chemistry and Physics 3: 487–507.
Stöhr A. 1912. Geringere Wirkung der Sonnenstrahlen. Das Wetter 29: 212–213.
Stommel H, Stommel E. 1983. Volcano Weather: The Story of 1816, The Year without a Summer. Seven Seas Press: Newport RI.
Stothers RB. 1984. The great Tambora eruption in 1815 and its aftermath. Science 224: 1191–1198.
Stothers RB. 1996. The great dry fog of 1783. Climatic Change 32: 79–89.
Štěpánek P. 2003. Temperature fluctuations in the Czech republic during the instrumental period. PhD Thesis, Masaryk University:
Brno.

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)
RESPONSES OF LARGE VOLCANIC ERUPTIONS OVER CENTRAL EUROPE 459

Strnadt A. 1785. Betrachtungen über einige meteorologische Gegenstände; besonders über die Ebbe und Fluth in der Luft, als auch über
die mehrmaligen täglichen Beobachtungen. Abhandlungen der böhmischen Gesellschaft der Wissenschaften 1: 300–342.
Thordarson T, Self S. 1993. The Laki (Skaftár Fires) and Grı́msvötn eruptions in 1783–1785. Bulletin of Volcanology 55: 233–263.
Thordarson T, Self S. 2003. Atmospheric and environmental effects of the 1783–1784 Laki eruption: a review and reassessment. Journal
of Geophysical Research 108: AAC7-1–AAC7-29.
Tille J. 1905. Geschichte der Stadt Niemes und ihrer nächsten Umgebung. Druck und Verlag von A. Bienert: Niemes.
Trnka F. 1912. Kroniky novoměstské. Část II. Nákladem a tiskem A. Veselého: Nové Město na Moravě.
Tutsch F. 1914. Die älteste Chronik Römerstadts und ihr Verfasser. Jahresbericht der Landesoberrealschule zu Römerstadt 16: 4–47.
Vupputuri RKR. 1992. The Tambora eruption in 1815 provides a test on possible global climatic and chemical perturbations in the past.
Natural Hazards 5: 1–16.
Vupputuri RKR, Blanchet JP. 1984. The possible effects of the El Chichon eruption on atmospheric thermal and chemical structure and
surface climate. Geofisica Internacional 23: 433–447.
Wiechowsky A. 1928. Aus den Aufzeichnungen des Lehrers Anton Lehmann aus Neuland bei Niemes. Mitteilungen des Nordböhm.
Vereines für Heimatforschung und Wanderpflege 51: 28–36, 110–115.
Williams SN, Self S. 1983. The October, 1902 plinian eruption of Santa Marı́a Volcano, Guatemala. Journal of Volcanology and
Geothermal Research 16: 33–56.
Witham CS, Oppenheimer C. 2005. Mortality in England during the 1783–4 Laki Craters eruption. Bulletin of Volcanology 67: 15–26.
DOI: 10.1007/s00445-004-0357-7.
Wood CA. 1992. Climatic effects of the 1783 Laki eruption. In The Year without a Summer? World Climate in 1816, Harington CR
(ed). Canadian Museum of Nature: Ottawa.
Xoplaki E, Luterbacher J, Paeth H, Dietrich D, Steiner N, Grosjean M, Wanner H. 2005. European spring & autumn temperature,
variability and change of extremes over the last half millennium. Geophysical Research Letters 32: L15713. DOI: 10.1029/2005G-
L023424.
Zilynskyj B. 1984. Letopis měšt’ana Nového Města pražského z let 1492 až 1539. Pražský sbornı́k historický 17: 52–89.

Copyright  2006 Royal Meteorological Society Int. J. Climatol. 26: 439–459 (2006)

You might also like