Next Article in Journal
Mineralogical and Engineering Properties of Soils Derived from In Situ Weathering of Tuff in Central Java, Indonesia
Previous Article in Journal
Vegetation, Climate and Habitability in the Marseille Basin (SE France) circa 1 Ma
Previous Article in Special Issue
Quantifying Drought Characteristics in Complex Climate and Scarce Data Regions of Afghanistan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Neural Network-Based Climate Prediction for the 21st Century Using the Finnish Multi-Millennial Tree-Ring Chronology

1
Institute of North Industrial Ecology Problems, Kola Science Centre, Russian Academy of Sciences, 184209 Apatity, Russia
2
Natural Resources Institute (LUKE), 96200 Rovaniemi, Finland
*
Author to whom correspondence should be addressed.
Geosciences 2024, 14(8), 212; https://doi.org/10.3390/geosciences14080212
Submission received: 30 June 2024 / Revised: 25 July 2024 / Accepted: 6 August 2024 / Published: 8 August 2024
(This article belongs to the Special Issue Advanced Statistical Modelling in Climate Change)

Abstract

:
The sun’s activity role in climate change has become a topic of debate. According to data from the IPCC, the global average temperature has shown an increasing trend since 1850, with an average increase of 0.06 °C/decade. Our analysis of summer temperature records from five weather stations in northern Fennoscandia (65°–70.4° N) revealed an increasing trend, with a range of 0.09 °C/decade to 0.15 °C/decade. However, due to the short duration of instrumental records, it is not possible to accurately assess and predict climate changes on centennial and millennial timescales. In this study, we used the Finnish super-long (~7600 years) tree-ring chronology to create a climate prediction for the 21st century. We applied a method that combines a long short-term memory (LSTM) neural network with the continuous wavelet transform and wavelet filtering in order to make climate change predictions. This approach revealed a significant decrease in tree-ring growth over the near term (2063–2073). The predicted decrease in tree-ring growth (and regional temperature) is thought to be a result of a new grand solar minimum, which may lead to Little Ice Age-like climatic conditions. This result is significant for understanding current climate processes and assessing potential environmental and socio-economic risks on a global and regional level, including in the area of the Arctic shipping routes.

1. Introduction

The nature of modern and future climate change is currently a hot topic of discussion. According to the 6th Intergovernmental Panel on Climate Change (IPCC) Report [1], the global surface temperature over the past two decades has been 0.99 °C higher than it was between 1850 and 1900, mainly due to human activities that cause greenhouse gas emissions. However, greenhouse gas emissions are not the only cause of modern climate change. Indeed, according to some estimates [2,3], the contribution of solar activity to global warming in the past century was at least 50%. In addition, a strong correlation between solar activity and global surface temperature has been found during the instrumental period [4,5]. Furthermore, it is certain that glacial and interglacial periods coincide with changes in the geometry of Earth’s orbit, as predicted by the Milankovitch theory [6,7]. Proxy records have revealed sun–climate connections also on sub-Milankovitch timescales of hundreds to thousands of years [7,8,9,10,11,12,13,14]. The main solar cycles, such as the Suess (or de Vries) cycle (~200 years), the unnamed cycle (350–400 years), and the Eddy cycle (~900 years) were identified in the Finnish super-long (~7600 years) tree-ring chronology (FLTR) [7,8]. These cycles were also found in lake sediments in Alaska [9] and in records of Asian summer monsoons [13]. Bond et al. [14] identified a millennial-scale solar cycle in the North Atlantic drift-ice abundance. Zharkova [15] suggested that this cycle, with a duration of 350–400 years, manifests itself as a grand solar minimum of the Maunder minimum (1645–1715 AD) type. During this period, there were practically no sunspots, and the total solar irradiance (TSI) decreased by about 3 W/m2 [16]. This led to a decrease in the temperature of the northern hemisphere by 1.5 °C [17,18]. This period coincided with a period of more severe winters and frozen rivers (the Thames and Danube) in the United Kingdom and Europe [17,18]. Moreover, Zharkova [15] also demonstrated that the sun has entered a modern grand solar minimum (2020–2053) of Maunder minimum type with Little Ice Age climatic conditions.
These cycles differ from changes in the orbital forcing of the climate and are associated with other solar factors and mechanisms. Currently, variations in solar radiation at different wavelengths and galactic cosmic rays (GCR) are considered to be the main solar factors affecting the atmosphere and climate [16,19,20,21,22]. Variations in TSI cause direct heating of the Earth’s atmosphere [16], while solar ultraviolet radiation (UV) affects atmospheric chemistry, ozone abundance, and the thermal structure [19]. Additionally, the GCR, modulated by the solar magnetic field, affect the cloud cover in the low atmosphere, thereby changing the radiation balance [20,21] and the atmospheric circulation pattern [22]. In addition, cosmic dust fluxes of interplanetary and interstellar origins, modulated by magnetic and gravitational fields, can act as condensation centers in the Earth’s atmosphere and therefore affect the aerosol content and cloud cover [23,24].
Increasing global warming will lead to more intense climate-related risks and hazards [1]. Climate change can indeed become a security issue for certain countries and communities [1,25]. For instance, climate change and the associated sea-level rise may lead to population movements from islands and coastal areas [1,25,26]. Furthermore, for Arctic indigenous communities, climate change may restrict their traditional ways of life (hunting and fishing) [1,25,27]. For northern territories, climate change presents both risks and opportunities. The melting of icebergs increases the environmental risks, but at the same time, it provides the possibility of extracting natural resources and opens up the potential for Arctic shipping routes [28,29]. The estimation and prediction of shipping route transformation based on sea ice data has become a significant topic in modelling of climate change [28,29]. Most of these models indicate a declining trend in sea ice concentration and thickness [29]. Indeed, the Arctic sea ice extent in April 2024 was approximately 200,000 square kilometers below the 1991–2000 average. At the same time, it has shown an increasing trend since 2019 [30]. The decline of the sea ice extent will create new trade opportunities, especially for the Northern Sea Route along the Russian coasts [27,28,31].
According to the IPCC, the global surface temperature may rise by 1.4 °C to 4.4 °C due to greenhouse gas emissions in the 21st century [1]. These estimates are based on instrumental temperature records. However, these records are not extensive enough to accurately predict future climate change. The development and expansion of urban heat islands, or urbanization bias, has led to another issue with temperature trend estimates based on weather stations [4]. The proxy series, in particular the FLTR chronology, can extend the climatic record back over the past centuries and millennia [32]. This chronology is based on time series from Scots pine (Pinus sylvestris L.), composed of both modern and subfossil wood samples from northern Finnish Lapland (68°–70° N, 20°–30° E) [32].
Recently, artificial neural networks (ANN) have been used more and more for climate prediction and reconstruction [33,34,35,36,37,38,39,40,41,42,43]. For example, ANNs have been used to reconstruct summer temperatures in Europe [33]. Kalugin et al. [34] used ANNs in the reconstruction of temperature and precipitation using sediment proxies. The ANN approach has also been used to predict the temperature rise over Iran [35]. The recurrent neural network (RNN) is a type of ANN that is particularly useful for analyzing time series [36]. Additionally, the LSTM network has been developed, which is an improved version of the RNN [36,37]. Additionally, it has been found that ANN models, including more recent versions such as LSTM networks, are very suitable for analyzing the nonlinear relationship between climate and tree rings [38,39,40,41,42]. Khaleghi [38] has used an ANN model to evaluate climate change based on the correlation between tree rings, temperature, and precipitation. Molina et al. [39] also used LSTM modelling to investigate tree-ring-based precipitation reconstruction in Europe. Helama et al. [40] compared linear and ANN-based reconstructions of summer temperatures in northern Fennoscandia using the FLTR chronology. Other studies have also compared the performance of ANN models with linear models used in tree-ring-based climate reconstructions [41,42,43,44]. All these studies have confirmed that ANN models show better results than linear models, especially for the nonlinear climate–tree rings relationship [39,40,41,42,43,44]. In any case, the performance of different (linear and nonlinear) models for paleoclimate reconstructions can vary depending on various factors, such as the region, the nature of the data, and the timescales used [39,40,41,42].
To correctly interpret and model climate change for the 21st century, it is essential to take into account natural climate variability over timescales ranging from centuries to millennia [45]. In previous studies, the FLTR chronology has revealed a significant (but highly nonlinear) sun–climate relationship on sub-Milankovitch timescales from hundreds to thousands of years [7,8]. In light of the above context, this study aims to predict climate change for the 21st century based on the FLTR chronology using a new LSTM model coupled with the wavelet transform and filtering. The wavelet transform and wavelet filtering have been used to reveal a solar cycle of 350–400 years in the tree-ring record. This solar cycle may be responsible for the emergence of a new grand solar minimum of the Maunder minimum type in the 21st century [15].

2. Materials and Methods

2.1. Data Sets

To develop and train the ANN model, we used the Finnish super-long Scots pine (Pinus sylvestris Linnaeus) tree-ring chronology, covering the period from 5634 BC to 2004 AD [32]. This chronology includes tree-ring-width series from living and megafossil trees from the area of subarctic timberline in northern Fennoscandia (68°–70° N, 20°–30° E) (Figure 1). All tree-ring samples were processed using dendrochronological techniques and dated to calendar years using standard procedures. For more details, see [32,40,46]. The dendroclimatic potential of the FLTR chronology has been statistically confirmed [32,46]. The high level of the expressed population signal (EPS) of the chronology (EPS > 0.85) indicates its reliability and usefulness for paleoclimate reconstruction [47]. FLTR-based reconstructions of mid-summer temperatures for the region were created using linear [46] and ANN-based [40] models. A comparison of these different models revealed that ANN-based temperature reconstructions performed better [40].
The summer (JJA) temperature data from five meteorological stations in the GISTEMP database [48,49] were used to assess recent temperature changes on a regional scale. These stations are located throughout Northern Fennoscandia covering areas from northern Karelia (Kem, 65° N, 34.8° E) and Finnish Lapland (Sodankyla, 67.4° N, 26.6° E) to the Kola Peninsula (Murmansk, 69° N, 33.1° E; Teriberka, 69.2° N, 35.1° E) and northern Norway (Vardo, 70.4° N, 31.1° E) (Figure 1 and Table 1).

2.2. Statistical and Wavelet Analysis

To assess the regional temperature trends, we used the Kendall–Theil robust line (KTRL) regression method. The KTRL, also known as the Sen’s slope estimator, is based on the KTRLine software [50] and is a nonparametric regression method that calculates medians instead of means [50]. Therefore, it is not necessary for the data to be normally distributed, and it is robust to outliers. The statistical significance of the trend was assessed using a nonparametric Mann–Kendall test [51] with the XLSTAT 2022 statistical software.
The wavelet filtering of tree-ring data was used to assess fluctuations in power within a frequency range from S1 to S2 [52]:
W ¯ 2 = δ i δ t D i = i 1 i 2 W ( S i ) 2 S i ,
where W denotes the continuous wavelet transform of tree-ring time series, W ¯ is the scale-averaged wavelet power, δt is the sampling interval, δi denotes the factor for scale averaging, and the parameter D is a constant for each mother wavelet function (D = 0.776 for the Morlet wavelet with w 0 = 6 ) [52]. The Morlet function is defined as a plane wave modulated by a Gaussian function [52]:
φ ( t ) = π 1 / 4 e i w 0 t / S e t 2 / ( 2 S ) 2 ,
where S denotes the wavelet scale and w 0 is the non-dimensional frequency, here taken to be 6 to provide a good balance between time and frequency localization [52]. This value of w 0 makes the wavelet scale almost identical to the corresponding Fourier period [52].
Tree-ring data were resampled to a 5-year time resolution and standardized (zero mean, unit standard deviation) before wavelet filtering. All statistical parameters and wavelet filtering were computed through the MATLAB R2022a software.

2.3. Development and Training of the LSTM Network

The basic RNN architecture consists of an input layer, a hidden layer, and an output layer. The hidden layer consists of hidden units, or neurons, which update their weights. Data inputs are first passed through the input layer and then transferred to the hidden layers before being sent to the output layer. Each layer of a neural network contains neurons that are connected to neurons in the next layer through weighted links. The training process involves adjusting the weights of these connections to minimize the error between the predicted output and the desired output. [53]. This process is known as backpropagation and is used to optimize the network’s performance. The LSTM model is an advanced version of an RNN with a more complex structure. It contains additional cells, or memory blocks, with several gates that can be activated or deactivated depending on the error value [53]. The hyperbolic tangent (tanh) function and the sigmoid function were used as the activation functions for the state and gate, respectively, as shown in Equations (3) and (4):
t a n h y = 2 1 + e 2 y 1 ,
σ y = 1 1 + e y ,
In this study, an LSTM network was developed and trained using a Levenberg–Marquardt algorithm [53]. Future values of an output series were predicted based on past values of the same series:
y(t) = f(y(t–1), y(t–2),…y(td)),
The mean square error (MSE) was used as a loss function:
M S E = 1 N 1 N ( T t Y t ) 2 ,
where Yt represents the ith observation, Tt is the ith prediction, and N is the number of samples.

2.4. Evaluation of the LSTM Performance

The quality of the model was evaluated using three metrics: mean absolute error (MAE), root-mean square error (RMSE), and determination coefficient ( R 2 ):
M A E = 1 N 1 N T t Y t ,
R M S E = 1 N ( T t Y t ) 2 N ,
R 2 = ( 1 N ( T t T ) ( Y t Y ) ) 2 1 N Y t Y 2 1 N ( T t T ) 2 ,
where Y is the average of the observations.
The LSTM model was developed using the Neural Network Toolbox of MATLAB R2022a.

3. Results

3.1. Temperature Trends in Northern Fennoscandia during the Instrumental Period

Considering the entire instrumental period, the summer temperatures at all five sites have increased at a significant level (p < 0.01, MK test) (Figure 2 and Table 1). The summer temperature showed an increasing trend at a rate of 0.09 °C/decade for Teriberka, 0.1 °C/decade for Murmansk, 0.12 °C/decade for Vardo, 0.13 °C/decade for Kem, and 0.15 °C/decade for Sodankyla (Table 1). Interestingly, the warming trends in the coastal areas (Murmansk, Teriberka, and Vardo) were lower than in the continental ones (Kem and Sodankyla).

3.2. LSTM Network Development and Time Series Prediction

A method using the LSTM network, coupled with a wavelet transform and wavelet filtering, was applied to predict tree-ring growth. The wavelet transform and wavelet filtering were used to highlight the 350–400-year solar cycle in the tree-ring record. The proposed model consists of two phases (Figure 3). First, a continuous wavelet transform (CWT) was applied to the tree-ring time series. This revealed a statistically significant (or nearly significant) high power in a band (300–400 years) (Figure 4b). The signal is weak (below the 95% confidence level) around 1800–1000 B.C. and 200–1000 A.D. (Figure 5b). Additionally, the CWT analysis reveals other periodicities associated with the main solar cycles at ~200 years (Suess or de Vries) and ~900 years (Eddy). Further, the input tree-ring-width data were filtered using a wavelet filter in a frequency band (300–400 years). Then, the filtered FLTR was used to train an LSTM model. The LSTM model consists of five layers: (1) the first layer is an input layer, which receives the wavelet-filtered tree-ring chronology; (2) the second layer is the LSTM layer with 128 neurons; (3) the third layer is a dropout layer with a probability of 0.5; (4) the fourth layer is a fully connected layer; and (5) the fifth and the final layer is the output layer, which produces the predicted time series.
A fully connected layer is used to multiply the input by a weight matrix [53]. In our model, we used a fully connected layer with an input dimension of 128 (the default) and an output dimension of 1. To avoid overfitting, we also added a dropout layer [54]. In the present model, the hidden layer consisted of 128 neurons, and a dropout of 50% was applied. Every iteration, 64 of these neurons were randomly eliminated. Therefore, the configuration of the LSTM network was 1-128-1-1-1 (number of input parameters (neurons) in the first layer—number of neurons in the hidden layer—dropout layer with probability of 0.5—fully connected layer—number of output parameters in the third layer) (Table 2).
In this study, we trained an LSTM network using the first 92% of a sequence and tested the model on the remaining 8%. We used the adaptive moment estimator (Adam) with a learning rate of 0.005 during training. After training and testing, the model demonstrated low error, a high degree of accuracy, and good performance (Table 2 and Figure 5). Therefore, this LSTM model can be used to predict climate change in the near future, at least for the 21st century.
Figure 4c shows the tree-ring sequence with predicted values using the developed LSTM model for the period 2005–3005 A.D. The curve shows a significant decrease in tree-ring growth in the near future (2063–2073 A.D). Interestingly, there was a similar decrease (1705–1715 A.D.) that coincides with the Maunder minimum of solar activity (1645–1715).

4. Discussion

In this study, the FLTR, filtered using a wavelet in the range of 300–400 years, was used to predict climate change in the current century. The use of this chronology for reconstructing paleotemperatures has already been shown [7,32,40]. FLTR-based mid-summer temperature reconstructions were created using both linear and ANN models [32,40]. A comparison of these models revealed that the ANN-based temperature reconstructions performed better [40]. The temperature sensitivity of the FLTR chronology is higher on multidecadal timescales. The FLTR-based temperature reconstruction explains more than 90% of the variability on these timescales during the instrumental period [7,40]. Additionally, the FLTR chronology showed a significant correlation with another regional P. sylvestris tree-ring chronology from the Kola Peninsula in northwestern Russia (68.6° N, 33.3° E) during the period of 1445–2005 A.D. in response to powerful (volcanic explosivity index, VEI > 5) volcanic eruptions [55]. All the above facts indicate a high dendroclimatic potential for this chronology and its usefulness for temperature reconstructions. Climate predictions were made using an LSTM neural network, coupled with a wavelet transform and wavelet filtering. This model showed good performance with low errors and high fitting accuracy (Table 2 and Figure 5), and, therefore, it can be used to predict climate change in the near future, at least in the 21st century. The wavelet analysis of the FLTR chronology revealed significant (above or close to 95% confidence level) periodicities that coincide with well-known solar cycles, such as the Suess (or de Vries) cycle (~200 years), an unnamed cycle (350–400 years), and the Eddy cycle (~900 years) (Figure 5b). Previously, all these periodicities were identified in solar–FLTR connections [8]. Zharkova [15] suggested that the 350–400-year solar cycle manifests itself in the form of a grand solar minima of Maunder minimum (A.D. 1645–1715) type. After applying a wavelet filter, we found a significant periodicity above the 95% confidence level in the 300–400-year range. This periodicity was not regular in time, suggesting a nonlinear nature of the signal. Given the nonlinearity of this periodicity, an LSTM model was used, based on a nonlinear approach. This approach revealed the presence of a significant decrease in tree-ring width in the near future (2063–2073). Therefore, considering the close relationship between tree growth and summer temperatures, a temperature minimum is expected in the near future (2063–2073). This is a consequence of the upcoming new GSM of Maunder minimum type with Little Ice Age climatic conditions [15]. Interestingly, a similar minimum (1705–1715), revealed in the FLTR, coincides with the Maunder minimum of solar activity.
At first glance, this result may seem to be in contradiction with regional temperature trends. In this study, we found that the summer (JJA) temperatures at five stations in northern Fennoscandia (between 65° and 70.4° N) increased during the instrumental period, at a rate ranging from 0.09 °C/decade to 0.15 °C/decade (Figure 2). This is consistent with the IPCC’s data, which show an increasing trend at an average rate of 0.06 °C/decade since 1850 [1]. According to the IPCC, the global surface temperature may rise by 1.4 °C to 4.4 °C due to human-induced greenhouse gas emissions in the 21st century [1]. These estimates were based on instrumental temperature data [1]. However, according to some studies, the contribution of solar activity to 20th century global warming may be similar to that of the anthropogenic component [2,3,4,5]. Our results and other studies [3,45] have shown that natural climate variability on decadal and millennial timescales is particularly significant for accurately predicting climate change in the 21st century.
Our findings approximately agree with and explain the results of Zharkova’s suggestion that a new grand solar minimum (GSM) of the Maunder minimum type may occur in the near future (2020–2053) [15]. In addition, there are other indications of the approaching new GSM with Little Ice Age-like climatic conditions (or global cooling) [56,57,58,59,60]. According to the Solar Radiation and Climate Experiment (SORCE) satellite observations, during the early 2000s the fluxes of solar spectral radiation (SSR) in the visible and near-infrared ranges increased, while the solar activity level (sunspot number) and TSI fluxes decreased [58,59]. It was hypothesized that the unusual behavior of SSR may be related to the upcoming new GSM [58,59,60]. Moreover, the solar activity–tree rings connections were most clearly evident during and around the GSM, including the Maunder minimum, when sunspots were practically absent [60].
The potential reduction in global temperature in the near future (2063–2073) due to the upcoming GSM could have significant implications for the environment and socio-economic development in the region, including Arctic shipping routes.

5. Conclusions

An LSTM neural network-based climate prediction for the 21st century was created using the Finnish super-long (~7600 years) tree-ring chronology. The chronology had been previously wavelet-filtered in the range of 300–400 years. This approach revealed a significant decrease in tree-ring growth in the near future (2063–2073). The predicted decrease in tree ring growth (and regional temperature) seems to be a result of the possible onset of a new grand solar minimum, which could lead to climatic conditions similar to those of the Little Ice Age.
Our findings have shown that natural climate variability on decadal and millennial timescales is particularly significant for accurately predicting climate change in the 21st century. Therefore, this research contributes to the understanding of current climatic processes and the assessment of potential environmental and socio-economic risks at a global and regional level, including in the area of Arctic shipping routes.

Author Contributions

Conceptualization, E.A.K., O.I.S. and M.T.; methodology, E.A.K., O.I.S. and M.T.; software, E.A.K. and M.T.; validation, E.A.K., O.I.S. and M.T.; formal analysis, E.A.K., O.I.S. and M.T.; investigation, E.A.K., O.I.S. and M.T.; resources, O.I.S. and M.T.; data curation, M.T.; writing—original draft preparation, E.A.K., O.I.S. and M.T.; writing—review and editing, E.A.K., O.I.S. and M.T.; visualization, E.A.K., O.I.S. and M.T.; supervision, E.A.K.; project administration, O.I.S.; funding acquisition, E.A.K. and O.I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was carried out as part of government contract with the Institute of North Industrial Ecology Problems, Kola Science Center RAS (no. FMEZ-2022-0010).

Data Availability Statement

Meteorological data are contained within the article, and all data sources are mentioned.

Acknowledgments

The authors are grateful to anonymous Reviewers for their valuable comments.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. IPCC. Summary for Policymakers. In Climate Change 2023: Synthesis Report. Contribution of Working Groups I, II and III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Core Writing Team, Lee, H., Romero, J., Eds.; IPCC: Geneva, Switzerland, 2023; pp. 1–34. [Google Scholar] [CrossRef]
  2. Scafetta, N.; West, B.J. Phenomenological solar contribution to the 1900–2000 global surface warming. Geophys. Res. Lett. 2006, 33, L05708. [Google Scholar] [CrossRef]
  3. Scafetta, N. Empirical assessment of the role of the Sun in climate change using balanced multi-proxy solar records. Geosci. Front. 2023, 14, 101650. [Google Scholar] [CrossRef]
  4. Soon, W.; Connoly, R.; Connolly, M. Re-evaluating the role of solar variability on Northern Hemisphere temperature trends since the 19th century. Earth Sci. Rev. 2015, 150, 409–452. [Google Scholar] [CrossRef]
  5. El-Borie, M.A.; Thabet, A.A.; El-Mallah, E.S.; Abd El-Zaher, M.; Bishara, A.A. The response of the atmosphere to solar variations. Indian J. Phys. 2020, 94, 737–752. [Google Scholar] [CrossRef]
  6. IPCC. Technical Summary. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change; Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L., Eds.; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2007; pp. 56–58. Available online: https://www.ipcc.ch/site/assets/uploads/2018/02/ar4-wg1-ts-1.pdf (accessed on 24 June 2024).
  7. Helama, S.; Fauria, M.M.; Mielikäinen, K.; Timonen, M.; Eronen, M. Sub-Milankovitch solar forcing of past climates: Mid and late Holocene perspectives. GSA Bull. 2010, 122, 1981–1988. [Google Scholar] [CrossRef]
  8. Kasatkina, E.A.; Shumilov, O.I.; Timonen, M.; Potorochin, E.O. Millennial-scale solar variability in tree rings of Northern Fennoscandia at the end of the Holocene. Tree-Ring Res. 2023, 79, 1–11. [Google Scholar] [CrossRef]
  9. Hu, F.S.; Kaufman, D.; Yoneji, S.; Nelson, D.; Shemesh, A.; Huang, Y.; Tian, J.; Bond, G.; Clegg, B.; Brown, T. Cyclic variation and solar forcing of Holocene climate in the Alaskan Subarctic. Science 2003, 301, 1890–1893. [Google Scholar] [CrossRef]
  10. Kern, A.K.; Harzhauser, M.; Piller, W.E.; Mandic, O.; Soliman, A. Strong evidence for the influence of solar cycles on a Late Miocene lake system revealed by biotic and abiotic proxies. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2012, 329–330, 124–136. [Google Scholar] [CrossRef]
  11. Steinhilber, F.; Abreu, J.A.; Beer, J.; Brunner, I.; Christl, M.; Fischer, A.; Heikkilla, U.; Kubik, P.W.; Mann, M.; McCracken, K.G.; et al. 9,400 years of cosmic radiation and solar activity from ice cores and tree rings. Proc. Natl. Acad. Sci. USA 2012, 109, 5967–5971. [Google Scholar] [CrossRef]
  12. Soon, W.; Velasco Herrera, V.M.; Selvaraj, K.; Traversi, R.; Usoskin, I.; Chen, C.-T.A.; Lou, J.-Y.; Kao, S.-J.; Carter, R.M.; Pipin, V. A review of Holocene solar-linked climatic variation on centennial to millennial timescales: Physical processes, interpretative frameworks an a new multiple cross-wavelet transform algorithm. Earth-Sci. Rev. 2014, 134, 1–15. [Google Scholar] [CrossRef]
  13. Sun, W.; Liu, J.; Wang, B.; Chen, D.; Wan, L.; Wang, J. Holocene multi-centennial variations of the Asian summer monsoon triggered by solar activity. Geophys. Res. Lett. 2022, 49, e2022GL098625. [Google Scholar] [CrossRef]
  14. Bond, G.; Kromer, B.; Beer, J.; Muscheler, R.; Evans, M.N.; Showers, W.; Hoffmann, S.; Lotti-Bond, R.; Hajdas, I.; Bonani, G. Persistent solar influence on North Atlantic climate during the Holocene. Science 2001, 294, 2130–2136. [Google Scholar] [CrossRef]
  15. Zharkova, V. Modern Grand Solar Minimum will lead to terrestrial cooling. Temperature 2020, 7, 217–222. [Google Scholar] [CrossRef]
  16. Lean, J.R.; Beer, J.; Bradley, R. Reconstruction of solar irradiance since 1610, Implications for climate change. Geophys. Res. Lett. 1995, 22, 3195–3198. [Google Scholar] [CrossRef]
  17. Shindell, D.T.; Schmidt, G.A.; Mann, M.E.; Rind, D.; Waple, A. Solar forcing of regional climate change during the Maunder minimum. Science 2001, 294, 2149–2152. [Google Scholar] [CrossRef] [PubMed]
  18. Lockwood, M.; Harrison, R.G.; Woolings, T.; Solanki, S.K. Are cold winters in Europe associated with low solar activity? Environ. Res. Lett. 2010, 5, 024001. [Google Scholar] [CrossRef]
  19. Haigh, J.D. The impact of solar variability on climate. Science 1996, 272, 981–984. [Google Scholar] [CrossRef] [PubMed]
  20. Marsh, N.D.; Svensmark, H. Low cloud properties influenced by cosmic rays. Phys. Rev. Lett. 2000, 85, 5004–5008. [Google Scholar] [CrossRef]
  21. Palle, E.; Butler, J. Sunshine records from Ireland, cloud factors and possible links to solar activity and cosmic rays. Int. J. Climatol. 2001, 21, 709–729. [Google Scholar] [CrossRef]
  22. Tinsley, B.A.; Deen, G.W. Apparent tropospheric response to MeV-GeV particle flux variations: A connection via electrofreezing of supercooled water in high-level clouds? J. Geophys. Res. 1991, 96, 22283–22296. [Google Scholar] [CrossRef]
  23. Kasatkina, E.A.; Shumilov, O.I.; Lukina, N.V.; Krapiec, M.; Jacoby, G. Stardust component in tree rings. Dendrochronologia 2007, 24, 131–135. [Google Scholar] [CrossRef]
  24. Ermakov, V.I.; Okhlopkov, V.P.; Stozhkov, Y.I. Influence of cosmic rays and cosmic dust on the atmosphere and Earth’s climate. Bull. Russ. Acad. Sci. Phys. 2009, 73, 416–418. [Google Scholar] [CrossRef]
  25. Barnett, J. Security and climate change. Glob. Environ. Chang. 2003, 13, 7–17. [Google Scholar] [CrossRef]
  26. Vaidya, H.N.; Breininger, R.D.; Madrid, M.; Lazarus, S.; Kachouie, N.N. Generalized additive models for predicting sea level rise in coastal Florida. Geosciences 2023, 13, 310. [Google Scholar] [CrossRef]
  27. Sharapov, D. Northern Sea route and climate change. E3S Web Conf. 2023, 460, 09019. [Google Scholar] [CrossRef]
  28. Lee, T.; Kim, H.J. Barriers of voyaging on the Northern Sea Route: A perspective from shipping companies. Mar. Policy 2015, 62, 264–270. [Google Scholar] [CrossRef]
  29. Zhang, Y.; Sun, X.; Zha, Y.; Wang, K.; Chen, C. Changing Arctic Northern Sea Route and Transpolar Sea Route: A prediction of route changes and navigation potential before Mid-21st century. J. Mar. Sci. Eng. 2023, 11, 2340. [Google Scholar] [CrossRef]
  30. NOAA National Centers for Environmental Information, Monthly Global Snow and Ice Report for April 2024. Available online: https://www.ncei.noaa.gov/access/monitoring/monthly-report/global-snow/202404 (accessed on 10 May 2024).
  31. Moe, A. A new Russian policy for the Northern sea route? State interests, key stakeholders and economic opportunities in changing times. Polar J. 2020, 10, 209–227. [Google Scholar] [CrossRef]
  32. Helama, S.; Mielikäinen, M.; Timonen, M.; Eronen, M. Finnish supra-long tree-ring chronology extended to 5634 BC. Nor. J. Geogr. 2008, 62, 271–277. [Google Scholar] [CrossRef]
  33. Guiot, J.; Nicault, A.; Rathgeber, C.; Edouard, J.L.; Guibal, F.; Pichard, G.; Till, C. Last-millennium summer-temperature variations in western Europe based on proxy data. Holocene 2005, 15, 489–500. [Google Scholar] [CrossRef]
  34. Kalugin, I.; Daryin, A.; Smolyaninova, L.; Andreev, A.; Diekmann, B.; Khlystov, O. 800-yr-long records of annual air temperature and precipitation over southern Siberia inferred from Teletskoye Lake sediments. Quat. Res. 2007, 67, 400–410. [Google Scholar] [CrossRef]
  35. Ashrafi, K.; Shafiepour, M.; Ghasemi, M.; Najar Araabi, B. Prediction of climate change induced temperature rise in regional scale using neural network. Int. J. Environ. Res. 2012, 6, 677–688. [Google Scholar]
  36. Chen, L.; Han, B.; Wang, X.; Zhao, J.; Yang, W.; Yang, Z. Machine learning methods in weather and climate applications: A survey. Appl. Sci. 2023, 13, 12019. [Google Scholar] [CrossRef]
  37. Hamdan, A.; Al-Salaymeh, A.; AlHamad, I.M.; Ikemba, S.; Ewin, D.R.E. Predicting future global temperature and greenhouse gas emissions via LSTM model. Sustain. Energy Res. 2023, 10, 21. [Google Scholar] [CrossRef]
  38. Khaleghi, M.R. Application of dendroclimatology in evaluation of climatic changes. J. For. Sci. 2018, 64, 139–147. [Google Scholar] [CrossRef]
  39. Molina, A.A.R.; Bezak, N.; Tootle, G.; Wang, C.; Gong, J. Machine-learning-based reconstructions: A study on Slovenia’s Sava River Basin. Hydrology 2023, 10, 207. [Google Scholar] [CrossRef]
  40. Helama, S.; Makarenko, N.G.; Karimova, L.M.; Kruglun, O.A.; Timonen, M.; Holopainen, J.; Meriläinen, J.; Eronen, M. Dendroclimatic transfer functions revisited: Little Ice Age and Medieval Warm Period summer temperatures reconstructed using artificial neural networks and linear algorithms. Ann. Geophys. 2009, 27, 1097–1111. [Google Scholar] [CrossRef]
  41. Jevšenak, J.; Levanič, T. Should artificial neural networks replace linear models in tree ring based climate reconstructions? Dendrochronologia 2016, 40, 102–109. [Google Scholar] [CrossRef]
  42. Jevšenak, J.; Džeroski, S.; Levanič, T. Predicting the vessel lumen area tree-ring parameter of with linear and nonlinear machine learning algorithms. Geochronometria 2018, 45, 211–222. [Google Scholar] [CrossRef]
  43. Fang, M.; Li, X. An artificial neural networks-based tree ring width proxy system model for paleoclimate data assimilation. J. Adv. Model. Earth Syst. 2019, 11, 892–904. [Google Scholar] [CrossRef]
  44. Salehnia, N.; Ahn, J. Modelling and reconstructing tree ring growth index with climate variables through artificial intelligence and statistical methods. Ecol. Indic. 2022, 134, 108496. [Google Scholar] [CrossRef]
  45. Scafetta, N. Impacts and risks of “realistic” global warming projections for the 21st century. Geosci. Front. 2024, 15, 101774. [Google Scholar] [CrossRef]
  46. Helama, S.; Timonen, M.; Holopainen, J.; Ogurtsov, M.G.; Mielikäinen, K.; Eronen, M.; Lindholm, M.; Meriläinen, J. Summer temperature variations in Lapland during the Medieval Warm Period and the Little Ice Age relative to natural instability of thermohaline circulation on multi-decadal and multi-centennial scales. J. Quatern. Sci. 2009, 24, 450–456. [Google Scholar] [CrossRef]
  47. Wigley, T.M.L.; Briffa, K.R.; Jones, P.D. On the average value of correlated time series, with applications in dendrochronology and hydrometeorology. J. Appl. Meteorol. Clim. 1984, 23, 201–213. [Google Scholar] [CrossRef]
  48. GISTEMP Team. GISS Surface Temperature Analysis (GISTEMP), Version 4; NASA Goddard Institute for Space Studies: New York, NY, USA, 2024. Available online: https://data.giss.nasa.gov/gistemp (accessed on 9 May 2024).
  49. Lenssen, N.; Schmidt, G.; Hansen, J.; Menne, M.; Persin, A.; Rudy, R.; Zyss, D. Improvements in the GISTEMP uncertainly model. J. Geophys. Res. Atmos. 2019, 124, 6307–6326. [Google Scholar] [CrossRef]
  50. Granato, C.E. Kendall-Theil Robust Line (KTRLine version 1.0)—A visual basic program for calculating and graphing robust nonparametric estimates of linear regression coefficients between two continuous variables. In Techniques and Methods of the U.S. Geological Survey, Book 4; U.S. Geological Survey: Reston, VA, USA, 2006; Chapter A7; 31p. [Google Scholar] [CrossRef]
  51. Salas, J.D. Analysis and modelling of hydrologic time series. In Handbook of Hydrology; Maidment, D.R., Ed.; McGraw-Hill: New York, NY, USA, 1993; Chapter 19; pp. 19.1–19.72. [Google Scholar]
  52. Torrence, C.; Compo, G.P. A practical guide to wavelet analysis. Bull. Am. Meteorol. Soc. 1998, 79, 61–78. [Google Scholar] [CrossRef]
  53. Aggarwal, C.C. Neural Networks and Deep Learning, 2nd ed.; Springer Nature Switzerland AG: Cham, Switzerland, 2023. [Google Scholar]
  54. Srivastava, N.; Hinton, G.; Krizhevsky, A.; Sutskever, I.; Salakhutdinov, R. Dropout: A simple way to prevent neural networks from overfitting. J. Mach. Learn. Res. 2014, 15, 1929–1958. [Google Scholar]
  55. Shumilov, O.I.; Kasatkina, E.A.; Mielikäinen, K.; Timonen, M.; Kanatjev, A.G. Palaeovolcanos, solar activity and pine tree-rings from the Kola Peninsula (northwestern Russia) over the last 560 years. Int. J. Environ. Res. 2011, 5, 855–864. [Google Scholar]
  56. Abdussamatov, H.I. Grand minimum of the solar irradiance leads to the Little ice age. J. Geol. Geosci. 2013, 2, 113. [Google Scholar] [CrossRef]
  57. Morner, N.-A. The approaching new Grand solar minimum and Little ice age climatic conditions. Nat. Sci. 2015, 7, 510–518. [Google Scholar] [CrossRef]
  58. Harder, J.W.; Fontenla, J.M.; Pilewskie, P.; Richard, E.C.; Woods, T.N. Trends in solar spectral irradiance variability in the visible and infrared. Geophys. Res. Lett. 2009, 36, L07801. [Google Scholar] [CrossRef]
  59. Haigh, J.D.; Winning, A.R.; Toumi, R.; Harder, J.W. An influence of solar spectral variations on radiative forcing of climate. Nature 2010, 467, 696–699. [Google Scholar] [CrossRef] [PubMed]
  60. Kasatkina, E.A.; Shumilov, O.I.; Timonen, M. Solar activity imprints in tree-ring data from northwestern Russia. J. Atmos. Sol. Terr. Phys. 2019, 193, 105075. [Google Scholar] [CrossRef]
Figure 1. Map showing sample collection sites with subfossil pines [32] (triangles) and weather stations (black circles): 1—Vardo, 2—Teriberka, 3—Murmansk, 4—Sodankyla, 5—Kem. The blue dashed line indicates the Arctic circle.
Figure 1. Map showing sample collection sites with subfossil pines [32] (triangles) and weather stations (black circles): 1—Vardo, 2—Teriberka, 3—Murmansk, 4—Sodankyla, 5—Kem. The blue dashed line indicates the Arctic circle.
Geosciences 14 00212 g001
Figure 2. Mean summer (JJA) temperatures across northern Fennoscandia: (a) Murmansk, (b) Teriberka, (c) Kem, (d) Sodankyla, (e) Vardo. Red lines denote trends calculated using the nonparametric Kendall–Theil robust line regression method [50]. The numbers indicate the increasing rate (°C/decade), with the 95% confidence interval in square brackets.
Figure 2. Mean summer (JJA) temperatures across northern Fennoscandia: (a) Murmansk, (b) Teriberka, (c) Kem, (d) Sodankyla, (e) Vardo. Red lines denote trends calculated using the nonparametric Kendall–Theil robust line regression method [50]. The numbers indicate the increasing rate (°C/decade), with the 95% confidence interval in square brackets.
Geosciences 14 00212 g002
Figure 3. A block diagram of the developed LSTM network for climate change prediction.
Figure 3. A block diagram of the developed LSTM network for climate change prediction.
Geosciences 14 00212 g003
Figure 4. (a) Finnish super-long tree-ring chronology (FLTR) [32], (b) corresponding continuous wavelet transform (CWT), and (c) wavelet-filtered chronology over the 300–400-year band (blue) with predicted values using the LSTM (red). The 95% confidence level against red noise is shown as a black contour.
Figure 4. (a) Finnish super-long tree-ring chronology (FLTR) [32], (b) corresponding continuous wavelet transform (CWT), and (c) wavelet-filtered chronology over the 300–400-year band (blue) with predicted values using the LSTM (red). The 95% confidence level against red noise is shown as a black contour.
Geosciences 14 00212 g004
Figure 5. Comparison of the measured (blue) and predicted by the LSTM (red) time series of the FLTR over the testing period (1398–2003 A.D.) (a) and the difference between predicted and measured FLTR values (b).
Figure 5. Comparison of the measured (blue) and predicted by the LSTM (red) time series of the FLTR over the testing period (1398–2003 A.D.) (a) and the difference between predicted and measured FLTR values (b).
Geosciences 14 00212 g005
Table 1. Coordinates and summer temperature trends of five stations in northern Fennoscandia.
Table 1. Coordinates and summer temperature trends of five stations in northern Fennoscandia.
Station (Coordinates)Period (Years)Rate (°C/Decade) 1
Vardo (70.4° N, 31.1° E)1870–20230.12 [0.085 0.158] 2
Teriberka (69.2° N, 35.1° E)1893–20230.09 [0.034 0.15]
Murmansk (69° N, 33.1° E1919–20230.1 [0.027 0.176]
Sodankyla (67.4° N, 26.6° E)1908–20230.15 [0.088 0.202]
Kem (65° N, 34.8° E)1891–20230.13 [0.077 0.175]
1 Increasing rate was calculated according to the Kendall–Theil robust line regression method [50]. All trends are significant (Mann–Kendall test [51]; p < 0.01). 2 The 95% confidence interval of each increasing rate is given in square brackets.
Table 2. Parameters and performance validation (testing period) of the LSTM model.
Table 2. Parameters and performance validation (testing period) of the LSTM model.
LSTM Parameters
Number of LSTM layers (number of neurons)1 (128)
Number of fully connected layers1
Number of dropout layers (probability)1 (0.5)
Types of activation functiontanh (state); σ (gate)
OptimizerAdam
Learning rate0.005
Loss functionmean square error (MSE)
Number of epochs700
Performance validation
R 2 0.99
MAE0.008
RMSE0.0113
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kasatkina, E.A.; Shumilov, O.I.; Timonen, M. Neural Network-Based Climate Prediction for the 21st Century Using the Finnish Multi-Millennial Tree-Ring Chronology. Geosciences 2024, 14, 212. https://doi.org/10.3390/geosciences14080212

AMA Style

Kasatkina EA, Shumilov OI, Timonen M. Neural Network-Based Climate Prediction for the 21st Century Using the Finnish Multi-Millennial Tree-Ring Chronology. Geosciences. 2024; 14(8):212. https://doi.org/10.3390/geosciences14080212

Chicago/Turabian Style

Kasatkina, Elena A., Oleg I. Shumilov, and Mauri Timonen. 2024. "Neural Network-Based Climate Prediction for the 21st Century Using the Finnish Multi-Millennial Tree-Ring Chronology" Geosciences 14, no. 8: 212. https://doi.org/10.3390/geosciences14080212

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop