Você está na página 1de 24

Mitig Adapt Strateg Glob Change

DOI 10.1007/s11027-013-9464-0
ORIGINAL ARTICLE

Artificial neural networks models for predicting effective


drought index: Factoring effects of rainfall variability
Muthoni Masinde

Received: 9 February 2013 / Accepted: 26 March 2013


# Springer Science+Business Media Dordrecht 2013

Abstract Though most factors that trigger droughts cannot be prevented, accurate, relevant
and timely forecasts can be used to mitigate their impacts. Drought forecasts must define the
droughts severity, onset, cessation, duration and spatial distribution. Given the high probability of droughts occurrence in Kenya, her heavy reliance on rain-fed agriculture and lack of
effective drought mitigation strategies, the country is highly vulnerable to impacts of
droughts. Current drought forecasting approaches used in Kenya are not able to provide
short and long term forecasts and they fall short of providing the severity of the drought. In
this paper, a combination of Artificial Neural Networks and Effective Drought Index is
presented as a potential candidate for addressing these drawbacks. This is demonstrated
using forecasting models that were built using weather data for thirty years for four weather
stations (representing 3 agro-ecological zones) in Kenya. Experiments varying various
input/output combinations were carried out and drought forecasting network models were
implemented in Matrix Laboratorys (MATLAB) Neural Network Toolbox. The models
incorporate forecasted rainfall values in order to mitigate for unexpected extreme climate
variations. With accuracies as high as 98 %, the solution is a great enhancement to the
solutions currently in use in Kenya.
Keywords Drought Forecasts . Artificial Neural Networks(ANNs) . Effective Drought
Index(EDI) . Available Water Resource Index(AWRI) . Rainfall Variations . Kenya

1 Introduction
Droughts are among the most expensive disasters in the world and their negative impacts
span economic, social and environmental aspects of the affected society. It is for this reason
M. Masinde (*)
Department of Information Technology, Central University of Technology, Private Bag X20539,
Bloemfontein, 9300, South Africa
e-mail: emasinde@cut.ac.za
M. Masinde
e-mail: muthonimasinde@yahoo.com
URL: http://www.muthonimasinde.net

Mitig Adapt Strateg Glob Change

that research on drought has attracted a lot of interests from environmentalists, ecologists,
hydrologists, meteorologists, geologists, economists and agricultural scientists (Mishra and
Desai 2006). In Byun and Wilhite (1999), literature on droughts is classified under 4
categories; (1) causes of droughts (atmospheric science); (2) frequency and severity of
droughts in a given region; (3) impacts of droughts; and (4) response, mitigation and
preparedness for droughts; this paper focuses on some aspects of the latter. Natural disasters
triggered by critical climate extremes especially droughts continue to affect millions of
people in Africa. According to the World Disasters Report of 2011, Africa contributed over
50 % of the droughts that occurred in the world between 2001 and 2011 (Armstrong et al.
2011).
Kenya has consistently contributed the highest number of people affected by natural
disasters in the Continent (Deely et al. 2010). Droughts are heavily felt in Kenya because of
three factors: (1) regularity droughts are so regular in Kenya; they were for instance
experienced in 50 % of the years between 1980 and 2008; (2) like most other developing
countries of Africa, Kenyas rain-fed agriculture is the backbone of the economy. Over 80 %
of Kenyans rely on the agricultural sector that contributes 24 % of the Gross Domestic
Product (GDP) (Mwagore 2002). The sector is highly sensitive to droughts; and (3) inability
of Kenyas population to prepare and adapt to droughts. Being a developing country with
46 % of the rural population living below the poverty line, the governments priority list is
filled with items such as providing basic education, implementing democratic constitution,
peace initiatives, providing basic healthcare, and so on. This leaves no room for developing
drought mitigation strategies. Two examples of recent droughts that affected Kenya include:
the 19992002 drought in western and central Kenya that affected 23 million people (WHO
Collaborating Centre for Research on the Epidemiology of Disasters (CRED) 2012) and the
20052006 in the northern and eastern regions of Kenya that left over 70 % of the livestock
dead (CERF 2009).
Droughts are quantified using indices such as the Effective Drought Index (EDI). EDI is able
to quantify droughts in absolute terms and also provide answers to: (1) the when, (2) the how
long (onset to cessation) as well as (3) the severity of droughts/floods (Byun and Wilhite 1999;
Mishra and Singn 2010 ). EDI quantifies droughts in terms of droughts classes composed of
positive and negative real values; for example, 2.50 indicates extreme drought, +3.28 indicates
extreme floods and 0.98 indicates close to normal wetness. EDI further qualifies droughts by
providing Available Water Resource Index (AWRI) that can for example reliably inform a
farmer of the amount of the available soil moisture at any given day.
Though the well-developed drought indices such as EDI perform very well in mapping
droughts in spatial and timescale dimensions, they only detect the events already happening
(Masinde and Bagula 2011). In this paper, Artificial Neural Networks (ANNs) play the role
of forecasting/predicting future droughts. The numerical approaches that are used in Kenya
produce forecasts with low accuracies; 70 % at best (Mutua 2011). With such forecasts, it
becomes difficult to make sound policy decisions to mitigate for droughts.
The work presented in this paper is part of a larger project whose main objective was to
develop an effective, sustainable and relevant home-grown Drought Early Warning System
(DEWS) for the Sub-Saharan Africa by making use of Artificial Intelligence technologies to
integrate: (1) indigenous knowledge on droughts; (2) scientific weather forecasts; (3)
Wireless Sensor Networks (WSNs); and (4) mobile phones (Masinde and Bagula 2012).
The main component of this DEWS is Drought Monitoring and Prediction. While the
evaluation of the strengths of EDI as a tool for monitoring droughts in terms of their severity,
onset, cessation and probability of their occurrence has been described by Masinde and
Bagula (2012), this paper tackles the Drought Prediction sub-component of DEWS. It

Mitig Adapt Strateg Glob Change

describes drought-forecasting component that combines the use of Artificial Neural


Networks (ANNs) and Effective Drought Index (EDI) to forecast drought for periods
ranging from one day to four years with accuracies ranging from 98 % to 75 %. This is a
phenomenal improvement on the accuracy (most existing solutions accuracy is about 70 %)
of the forecasts.
The models have been tested using historical weather data from 4 weather stations in
Kenya. Though the integration of EDI and ANNs has been attempted elsewhere in the world
especially in South-East Asia, no research known to the author has attempted to model the
micro (daily) level of details like it has been done in this paper. In designing the long-term
category of network models, the correlation among EDI values of similar calendar months
was utilised. For instance, given EDI values for February 2008, February 2009, February
2010 and February 2011, one can forecast EDI value for February 2013. Forecasted rainfall
values are used as input into the ANNs models to take care of unexpected extreme rainfall
variations patterns.

2 Drought forecasting in Kenya


Droughts are the most common disasters in Kenya; they result from variations in weather/
climate, but there are occasional floods too. In the period 1999 to 2008, Kenya contributed a
whopping 32.85 % of people affected by natural disasters in the Africa (Collins et al. 2009).
In August 2011, Kenya was among the countries at the Horn of Africa that experienced a
devastating drought that was described as the worst drought in 60 years (http://
www.bbc.co.uk/news/world-africa-13944550)(lastly accessed on 10 October 2012). Such
disasters cannot be avoided but can be managed through effective early warning systems.
When they occur, droughts affect more that 25 % of Kenyas population not mentioning the
ripple effects such as inadequate hydro-electric power supply, increased commodity prices
and loss of jobs just to mention a few.
Currently, monitoring of climatic/weather variations in Kenya is the mandate of the
Kenya Meteorological Department (KMD). The Department is handicapped in the sense
that all they have are sparsely distributed weather stations. They run 3 main types of stations
that are currently managed by the Climatological Section of the Department (http://
www.meteo.go.ke/): (1) 700 rainfall stations, (2) 62 temperature stations and (3) 27 synoptic
stations. The Agrometeorological Section on the other hand manages 13 stations related to
agriculture; data is remitted from these stations every 10 days. Apart from the normal
meteorological observations, other observations by the Agrometeorological Section include:
soil temperature, sunshine duration, radiation, pan evaporation and potential evapotranspiration. All this data is stored in semi-automated formats at the Departments Head Quarters
in Nairobi. The data is available to interested stakeholders and on request. The data used in
this research was obtained from this Department. The Meteorological Department uses the
data collected to provide five main types of forecasts: (1) Daily forecast for main cities/
towns in Kenya; (2) Four-day forecast; (3) Seven-day forecast; (4) Monthly forecast; and (5)
Seasonal forecast. The latter are mainly issued two times a year a few weeks before each of
the two main rain seasons. The Country has two main rainy seasons: (1) October-NovemberDecember (OND); (2) March-April-May (MAM); MAM is the main season. Sometimes the
rainfall may occur in the period June-July-August (JJA). Among other factors (not part of
this research work), the amount of rain and moisture index in each of the above season is
used to classify Kenya under 14 climate and agro-ecological zones (http://
www.meteo.go.ke).

Mitig Adapt Strateg Glob Change

KMD uses statistical and dynamical models to generate seasonal rainfall forecasts for
each region in Kenya; these are then compared against the forecasts from neighbouring
countries, as well as against the forecasts from several global dynamical models, issued by
Meteorological agencies in Europe and the United States. These are then compiled to give
probabilistic forecast of rainfall under three categories: above-, near- or below-normal
rainfall. One of the shortcomings of KMDs approach is the fact that their forecasts provide
conceptual indications of droughts/floods without giving operational indicators. This makes
it difficult for key stakeholders to develop solid strategic plans. Conversely, Effective
Drought Index (EDI) is able to quantify droughts in absolute terms and also provide answers
to: (1) the when, (2) the how long (onset to termination) as well as (3) the severity of
droughts/floods. Combining EDI with Artificial Neural Networks(ANNs) leads to a more
effective drought forecasting solution described in this paper.

3 Drought severity indices and effective drought index


Drought is an insidious hazard of nature which according to Elsa et al. (2008) qualifies as a
hazard because it is a natural accident of unpredictable occurrence but of recognizable
recurrence. As a disaster, drought corresponds to the failure of the precipitation regime,
causing the disruption of the water supply to the natural and agricultural ecosystems. There
is no one universally accepted definition of drought yet. Palmer came to this conclusion as
early as 1965 when he stated Drought means various things to various people depending on
their specific interest. To the farmer drought means a shortage of moisture in the root zone of
his crops. To the hydrologist, it suggests below average water levels in the streams, lakes,
reservoirs, and the like. To the economist, it means a shortage which affects the established
economy (Palmer 1965). Since then, attempts have been made to define the term drought.
There are three main categories of drought: meteorological, hydrological and agricultural
(Wilhite and Glantz 1985; Dracup et al. 1980). Meteorological drought is defined as a period
of abnormally dry weather sufficiently prolonged for the lack of water/rainfall to cause
serious hydrologic imbalance in the affected area (Huschke 1970). As a matter of fact,
meteorological drought is precursor for the other categories of droughts, hence the most
critical.
According to Panu and Sharma (2002), the severity of a drought is a function of the
drought duration and probability distribution of the drought variable and its autocorrelation
structure. In meteorological drought, the severity is defined in the form of indices such as the
Palmer Drought Severity Index (PDSI) (Palmer 1965). PDSI is the best-known index;
historically, it has been the most commonly implemented. Its weaknesses however range
from its complexity to poor applicability (underlying computation is based on the climate of
the Midwestern United States). This has led to various variations of the Index such as: (1) the
Standardized Precipitation Index (SPI) that measures the deviation of the actual precipitation
from the average conditions in a given area (Mckee et al. 1993); (2) Palmer Hydrological
Drought Index (PHDI), which is used for water supply monitoring; (3) self-calibrated PDSI,
which provided a solution to the inappropriately higher frequency of extreme drought (Wells
et al 2004); (4) Normalized Difference Vegetation Index (NDVI) on the other hand takes
advantage of the reflective and absorptive characteristics of plants in the red and nearinfrared portions of the electromagnetic spectrum; and (5) AWRI that expresses the actual
amount of available water (Wilhite and Glantz 1985). Like the PDSI, SPI has its own share
of popularity; however its major undoing (among others) is the predication granularity,
which happens to be monthly scale. This way, it is not as accurate as required; it does not

Mitig Adapt Strateg Glob Change

reflect daily/weekly patterns. Researchers such as Byun and Wilhite (1999) developed the
EDI to address SPI shortcomings.
EDI is different from the rest of the indices in a number of ways; it was developed to
address weaknesses identified in the existing (at the time) drought indices. Some desirable
features of EDI are: (1) More accurately calculates current level of available water resources;
(2) It considers drought continuity, not just for a limited period; it can therefore diagnose
prolonged droughts that continue for several years; (3) It is computed using precipitation
alone; and (4) It considers daily water accumulation with a weighting function for time
passage. Calculation is therefore made with consideration of the fact that the quantity of
rainfall that can be used as a water resource drops gradually over time after the rain has
fallen. Effective Precipitations (EPs) are used to compute deficiency or surplus of water
resources for a particular date and place. EP here refers to the summed value of daily
precipitation with a time-dependant reduction function; it makes use of Eq. 1 below.
EPi

Xi
n1

 Pn

m1 Pm


1

where Pm is the precipitation of m days before and the index i represents the duration of
summation (DS) in days. Here i=365 is used, that is, summation for a year which is the most
dominant precipitation cycle worldwide. The 365 can then be a representative value of the
total water resources available or stored for a long time.
For instance, if i=2, then m varies from 1 to 2, EP2 becomes [P1+P1+P2)/2]. Once the
daily EP is computed, a series of indices can be calculated to highlight different characteristics of a stations water resources. These are: Mean Effective Precipitation (MEP),
Deviation of EP (DEP) and standardised value of DEP (EDI). EDI produced satisfactory
results in measuring drought severity and was adopted (and adapted) to analyse 200-year
drought climatology of Seoul, Korea (Kim et al. 2009) and to analyse droughts in Kenya
(Masinde and Bagula 2011)

4 Artificial neural networks in forecasting droughts


Artificial Intelligence (AI) refers to the computing paradigm that aims to develop solutions
that mimic human perception, learning and reasoning to solve problems. Two AI techniques
stand out when it comes to modelling environmental systems such as droughts; Intelligent
Agents and Artificial Neural Networks (ANNs). ANNs are composed of simple elements
(inspired by biological nervous systems) that mimic the brain neurons; they are designed to
operate in parallel to achieve a goal. The network function is determined by the connections
between elements. This function can be altered by varying the values of the connections
(weights) between these elements; this is referred to as training of the network.
Chang-Shian et al. (2010) described and compared AI techniques performance in
modelling environmental system; they singled out ANNs approach as being excellent in
solving data-intensive and multivariable problems with unclear mapping rules. ANNs are
highly interconnected systems that operate by mimicking the operation of the human brain.
ANNs perform seven categories of tasks: pattern classification, clustering, function approximation, prediction, optimisation, retrieval by content and process control.
For decades, pattern recognition techniques, regression models and autoregressive moving average (ARMA) and autoregressive integrated moving average (ARIMA) models for
statistical time series have been used for drought forecasting. ANNs provide one of the

Mitig Adapt Strateg Glob Change

nonlinear non-stationary alternative models for drought forecasting as explained in (Mishra


and Desai 2006). ANNs have several advantages to this end; they can be used to solve
problems with nonlinear/unknown multivariate and less controlled environments (Bodri and
Cermak 2001). The greatest strength of ANNs, though, is its ability to weave together
various mathematical components capable of tackling very complex physical systems such
as droughts. ANNs are also flexible and less assumption-dependent and this has seen them
find applications in modelling extremely complex hydrological domains such as rainfallrunoff, stream flows, water quality and precipitation estimation (American Society of Civil
Engineers (ASCE) Task Committee on Application of Artificial Neural Networks in
Hydrology 2000).
In most hydrological applications, multi-layer perceptron (MLP) neural network model is
adopted with feed-forward back propagation (BPN) as the training algorithm (Rumelhart et
al. 1986; Freitas and Billib 1997; Antoni et al. 2001; Bodri and Cermak 2001; Kung et al.
2006, and Chang-Shian et al. 2010). Antoni et al. (2001) used seven climatic variables
obtained from 127 weather stations in Croatia to develop empirical drought models using
BPN. Mishra and Desai (2006) observed that most ANNs implementations adopted the
multi-layer feed-forward neural network and were based on BPN.
BPN has been applied in modelling animal/plant population growth to predict
hydrological/meteorological phenomena such as droughts and floods. BPN has three general
layers: input, hidden and output, and has two processing types: learning/training and
recalling. At the input layer(s) data are fed to the system while the processing takes place
within the hidden layer(s) and finally, results are produced at the output layer(s) (Morid et al.
2007). Besides it being difficult to determine the representativeness of learning data, using
BPN in hydraulic modelling is faced with the problem that BPN is a non-linear black box
that does not consider/explain the underlying physical process of watershed (Chang-Shian et
al. 2010). This is the weakness that is addressed by Decision Group BPN (Chang-Shian et al.
2010).
Using a data (precipitation) set for 36 years (1965 to 2001), Mishra and Desai (2006)
compared the performance of BPN and ARIMA for different Standard Precipitation Index
(SPI) lead-times. Applying both the Recursive Multi-step Neural Network (RMSNN) and
the Direct Multi-step Neural Network (DMSNN) approaches, they found that RMSNN
performed better than both DMSNN and ARIMA for one-month lead-time and that
DMSNN performed better than the other two for a lead-time of four months. Other projects
in which ANNs have been used for predicting droughts/floods include drought prediction in
north-east Brazil(Freitas and Billib 1997) and summer flood forecasting in Moravia (Bodri
and Cermak 2001). Shin and Salas (2000) used ANNs to quantify the spatial and temporal
patterns of meteorological droughts for the south-western region of Colorado.
Though invented in 1943 by Mcculloch and Pitts (1943), it is the coming-on-stage of the
Back-Propagation training algorithm for feed-forward ANNs in 1986 that boosted the
application of ANNs in several real-life application areas such as drought forecasting
(Panu and Sharma 2002).
Maier and Dandy (2000) described five key considerations to successful implementation
of ANNs as:
(a)

ANNs performance evaluation criteria: how will the performance of the model be
judged? The relevant criterion for this study was accuracy of predicted results. This
also happens to be the most commonly used criterion (Demuth et al. 2009).
(b) Data set division criteria; deals with how to share the available dataset among the
training, validation and testing phases of ANNs. Most neural network tools such as

Mitig Adapt Strateg Glob Change

MATLAB Neural Network Toolkit adopt 70:15:15 percentages for training, validation
and testing respectively (Demuth et al. 2009). This data set division criterion was
adopted in this paper.
(c) Data pre-processing with the aim of ensuring some form of standardisation. In this
paper, all rainfall data was first subjected to EDI computation and hence, it was already
standardised/homogenisation. Further, the mapminmax function implemented in the
MATLAB Neural Network Toolkit was applied. The latter puts all the input values into
[1:1] range. This in turn ensured that all variables involved had equal representation
during the training.
(d) Determining model input, this is purely based on a priori knowledge of causal variables in conjunction with inspections of time series plots of potential inputs and
outputs. In this paper an array of weather parameters such as daily readings of
temperature, rainfall, relative humidity and wind direction were considered. From the
rainfall values, both daily and monthly EDI and AWRI were computed and extensively
used as inputs/outputs of the neural networks.
(e) Selecting a suitable network architecture which is made up of two aspects:
The type of connection (are there loops?) among the nodes; and how many hidden layers
and how many nodes in each of the hidden layers. The number at the input and output layers
is generally determined by the problem domain; this is referred to as the geometry of the
network.

5 Materials and methods


5.1 Case study with experimental design
Yin (1993) defines case study approach as an empirical enquiry that investigates a contemporary phenomenon within its real life context, when the boundaries between phenomenon
and context are not clearly evident and in which multiple sources of evidence are used. The
case study design was used to analyse historical daily weather data for over 30 years for four
weather stations in Kenya. First it was used to determine if EDI could be used to
qualify/quantify droughts and two, to investigate the suitability of ANNs in forecast
droughts. The design of the case studies followed a three-phase experiment made up of
pilot, exploratory and confirmatory experiments. This led to the variation of Case Study with
Experimental Design as shown in the Fig. 1. Each of the case studies started off with a PilotSingle Case Study of one weather station followed by recursive Exploratory Multiple Case
Studies of two weather stations. In the last phase of the case study design, detailed
Confirmatory Single Case Study using data from the fourth weather station was used.
5.2 Data sample
Droughts impacts are more pronounced in highly populated areas whose main economic
activity is agricultural production. In Kenya, these areas are found in western, central rift
valley and the highlands and in a narrow strip along the Indian Ocean coast. Four weather
stations were selected to represent four (excluding the rift valley) these regions. Data
composed of daily readings for temperature (highest and lowest), wind (direction and
speed), relative humidity (at 06Z and at 12Z), atmospheric pressure and rainfall was sourced
from KMD. These are: Dagoretti, Embu, Kakamega and Makindu. In the this paper, only

Mitig Adapt Strateg Glob Change

ANNs Design Experiments

Pilot Single-CaseStudy

Eploratory MultipleCase-Studies

Confirmatory SingleCase-Study

Data for 1 weather Station


(Embu)
Formulate theory

Data for 2 weather stations


(Dagoretti and Makindu)
Test theory

Data for 1 weather station


(Kakamega)
Confirm and extend theory

Fig. 1 The 3-steps of the case study with experimental design

daily precipitation data for years 1979 to 2009 from these stations was used; this translated
into 436531 records (Table 1).
5.3 EDI calculation
Two Phases: (1) Data cleaning; (2) Data analysis; were carried out as per the flow chart
shown in Fig. 2. A Java program was written to covert the original format to formats
acceptable by the EDI Fortan program: A Fortan program developed by Byun and Wilhite
(1999) and available at (http://atmos.pknu.ac.kr/~intra2/down_src.php (lastly accessed on 10
October 2012) was then used to compute EDI values. This computer program uses Eq. 1

Table 1 Geo-Data of the weather stations studied


Name

Dagoretti

Embu

Kakamega

Makindu

WMO number

63741

63720

63687

63766

ICAO

HKNC

HKEM

HKKG

HKMU

Year opened

1954

1975

1957

1904

Latitude

01 18S

00 30S

00 17 N

2 17S

Longitude

36 45E

37 27E

34 47E

37 50E

Elevation

1,798 m

1,493 m

2,133 m

1,000 m

Annual average rainfall

1,023,mm

1,250 mm

1,913 mm

593 mm

Ecological-zone
Geographical region

11
Highland (Nairobi)

11
Central

13
Western

10
Coast

ICAO (International Civil Aviation Organisation), WMO (World Meteorological Organisation)


The values of the annual average rainfall in this table were calculated (by the author) from daily rainfall values
for each station for years 1979 to 2009 in Kenya.

Mitig Adapt Strateg Glob Change


Fig. 2 Flow chart showing data
cleaning and analysis phases for
the Kenyas case study

Phase I: Data Cleaning

Weather Data
in text files;
Source: KMD

Extract
rainfall data
for 4 stations

Using a Java Program, reformat


the data to Day, Month, Year and
Station

Identify and remove


data gaps

Formatted Data

EDI
Program

Computed
Monthly &
Daily EDI

Excel Charts
Generator

Phase II: Data


Analysis
Weather

EDI Web-Based
Decision Support

Database

System

(MySQL)

(presented in Section 3) to compute daily/monthly EDIs and outputs them into text files. The
raw data that was acquired from the Kenya KMD was first cleaned and formatted to suit the
pre-defined format (generating an input file to the EDI program) for each of the 4 weather
stations. For each input file, an output file containing the EDI and AWRI values was
generated. These were then exported and stored in a My Structured Query Language
(MySQL) database from where various data analysis reports were generated. Some of these
can be found in (Masinde and Bagula 2011).

Mitig Adapt Strateg Glob Change

5.4 Artificial neural networks models development


5.4.1 Overview
The ANNs network models were developed using the output from the EDI computation
described above. The ANNs described in this work utilised only EDI, AWRI and precipitation values for 30 years (1980 to 2009); an attempt to include temperature (and other
parameters) led to large errors because among other reasons, precipitation (used to derive
EDI and AWRI) is in itself a function of these parameters. Further, precipitation is itself a
function of so many other linear and nonlinear functions (Weichert and Burger 1998). The
network models were created by combing various Inputs/Targets; for each station, the data
set contained daily or monthly readings for precipitation, EDI and AWRI. The format of the
daily EDI/AWRI file is as shown in Table 2 while the monthly one had the format:
[Month/Year|Precipitation|AWRI|EDI]. To normalize the input values, default data preprocessing function, mapminmax that maps the range of input values to the range [1 1]
was used.
5.4.2 Network architecture
One hidden layer with 20 neurons was selected. During the network building, some
experiments of varying the number of neurons were carried out but the value 20
outperformed them; hence the decision to use the default value 20. For example, during
the training of the ANNs for predicting EDI value for Dagoretti for 7-day lead-time, the
number of hidden layers was varied between 10 and 40 at an interval of 5. The Mean Square
Error (MSE) and Regression(R) values were used to rank them; it emerged that the network
with 20 hidden layers had the best performance.
5.4.3 Data division criteria and network training
For each data set, the ratio of 70 %:15 %:15 % for training, validation and testing
respectively was used. Splitting of the data during training was done randomly using
dividerand() function. Training was carried out using Levenberg-Marquardt
backpropagation (trainlm) algorithm. This was selected because it is the fastest
backpropagation algorithm in the MATLAN ANNs toolbox that was used in this reasearch
(Demuth et al. 2009)

Table 2 Format of daily EDI


output

Date

Total Precipitation

AWRI

EDI

1980-01-01

0.0

117.0

0.66

1981-01-31

0.0

208.1

0.96

2009-12-31

26.6

146.7

0.20

Mitig Adapt Strateg Glob Change

5.4.4 Network geometry


The networks were first designed to predict future EDI and/or AWRI values for a given
combination of EDI (E), precipitation (P) and AWRI (W) for 6 previous days/months. The
networks were categorized as follows:

&
&
&

7 Networks with 2(EDI and AWRI) outputs


7 Networks with EDI as the output
7 Networks with AWRI as the output

A simple Java program was then created to convert input files in the format on Table 2 to
the following inputs-targets mapping.
(i)

Homogeneous inputs, homogeneous output example:

&
(ii)

Network 8: En+1 =f(En, En1, En2, En3, En4, En5):Given EDI values for the last 6 days, forecast the EDI value for the next day.
Heterogeneous inputs, homogeneous output example:

&

(iii)

Network 14: En+1 =f((Pn, Pn1, Pn2, Pn3, Pn4, Pn5 ) (En, En1, En2, En3, En4, En5)
(Wn, Wn1, Wn2, Wn3, Wn4, En5)):Given Precipitation, EDI and AWRI values for the last 6 days, forecast the EDI
value for the next day.
Homogeneous inputs, heterogeneous output example:

&

(iv)

Network 2: En+1,Wn+1 =f(En, En1, En2, En3, En4, En5):Given EDI values for the last 6 days, forecast the EDI and AWRI values for the
next day.
Heterogeneous inputs, heterogeneous output example:

&

Network7: En+1,Wn+1 =f((En, En1, En2, En3, En4, En5) (Wn, Wn1, Wn2, Wn3,
Wn4, Wn5) ((Pn, Pn1, Pn2, Pn3, Pn4, Pn5)):Given EDI, AWRI and Precipitation values for the last 6 days, forecast the EDI
and AWRI values the next day.

5.4.5 ANNs performance analysis


Both R and MSE values were used to select the networks models with the best performance.
Root Mean Square Error (RMSE) and Percentage RMSE were then used to determine the
implications of the errors on the resulting forecasts. For Regression Analysis, the higher the
value, the higher the rank; it is the measure of the correlation between the inputs and the outputs.
In the case of MSE: the smaller the value the better the network; it refers to the average squared
difference between outputs and targets. However, the MSE value is bound to be directly
proportional to the size of the outputs values. For example, the outputs/targets for Networks
15 to 21 are those of AWRI whose mean for Embu weather station is 194.58 while for Networks
7 to 14 are for EDI values that ranges between 2.28 to 4.32 (an absolute mean of 0.76521) for
the same station. In order to get the actual implications of the networks, RMSE and the
percentage of the RMSE were computed.

Mitig Adapt Strateg Glob Change

6 Results and discussion


6.1 Rainfall and temperature variability in Kenya
As shown in Figs. 3 and 4, data analysis point towards more frequent and pronounced
rainfall variations as well as a rise in temperatures in Kenya. There is evidence of extreme
within-season variability especially of the rains on-set leading to unseasonal rainfall patterns.
Considering that the design and operation of ANNs is heavily dependent on learning
historical patterns and using these to predict the future, the rainfall variability being
witnessed in Kenya and elsewhere in the world would lead to larger errors in the predicted
drought values. Forecasted (by KMD) rainfall values were incorporated into the ANNs
models to address such situations.
6.2 Pilot phase - Experiments results
In order to select the best network to use, data for one (Embu) weather station was used;
results of the runs are tabulated in Table 3. An ideal Network would be the one that predicts
EDI and/or AWRI given precipitation only, however, these (1, 8 and 15) had the worst
performance. Secondly, by selecting a network that uses all the 3 inputs (P, E and W) and
gives both EDI and AWRI as outputs is ideal because it ensures that only one network
is needed to solve the current problem. This scenario is represented in Networks 7
which had the best performance in Category 1 of the networks. What about creating
two different networks that use all the three inputs (P, E and W) to predict both EDI
and AWRI respectively? This is achieved in Network 14 and 21. Though Network14
had the best performances in Category 2, Network 21 was the third after Network 19
and 17 in Category 3.
The networks with average (Rank 1 to 3) performance and that have Precipitation as an
input were selected for further investigation. These are Networks 7, 11, 14, 20 and 21.
Finally, in order to exhaustively investigate the best networks for building bi-network (two
network; for EDI and AWRI respectively), the two best performing networks in each of
these categories were selected.
Networks 1, 8 and 15 that have precipitation alone as input had the worst performance.
An attempt to use either EDI to compute AWRI and vice versa did not yield desirable results.
This is an indication that there is no any form of relationship between the two. This made
Networks 2, 3, 10 and 16 give poor performance. Finally, combining EDI and Precipitation
to predict AWRI or AWRI with Precipitation to predict EDI did not work either. This made
Networks 4, 6, 13, and 18 not perform well.
6.3 Exploratory phase I - Experiments results
From the results in Table 3 and the analysis in above, 10 Neural Networks models (5, 7, 9,
11, 12, 14, 17, 19, 20 and 21) were identified for further analysis during the Exploratory
Experiments. Using data for Embu, Dagoretti and Makindu weather stations, the 10 Neural
Networks were trained and their performances analysed. For each weather station, a table
similar to Table 4 was generated.
For Dagoretti, Network 9 had the lowest value (0.01887) of MSE in the category with
EDI (Networks 9, 11, 12 and 14) as output. For the R value, both Networks 9 and 11 had
equal values (0.99029). For AWRI category, the one-input Network 17 had the best
performance. For Embu, in the category of Networks for computing EDI(Networks 9, 11,

Mitig Adapt Strateg Glob Change


Fig. 3 Temperature variability in
selected weather stations in
Kenya

Mitig Adapt Strateg Glob Change


Fig. 4 Rainfall variability in
selected weather stations in
Kenya

Mitig Adapt Strateg Glob Change


Table 3 Performance of the pilot phase neural networks
Network

Inputs (I)

Targets (T)

Mean Square Error (MSE)

Tr

Vl

Ts

Regression (R)
Rn

Tr

Vl

Ts

Rn

Category 1: Two outputs (EDI and AWRI)


Network1

2,340.90

2,358.28

2,358.05

0.58

0.68

0.62

Network2

1,169.10

1,365.17

1,287.50

0.94

0.94

0.95

Network3

41.24

40.19

47.63

0.76

0.77

0.75

Network4

795.01

861.62

885.25

0.96

0.96

0.96

Network5

43.00

32.74

42.15

0.97

0.98

0.97

Network6
Network7

Y
Y

N
Y

Y
Y

Y
Y

Y
Y

39.45
35.66

51.13
59.27

50.33
45.75

6
1

0.76
0.97

0.76
0.97

0.74
0.97

4
1

Category 2: Output EDI only


Network8

0.84

0.82

0.85

0.36

0.32

0.32

Network9

0.02

0.03

0.02

0.99

0.99

0.99

Network10

0.38

0.35

0.38

0.78

0.78

0.78

Network11

0.02

0.02

0.02

0.99

1.00

0.99

Network12

0.02

0.02

0.02

0.99

0.99

0.99

Network13
Network14

Y
Y

N
Y

Y
Y

Y
Y

N
N

0.31
0.02

0.31
0.02

0.33
0.02

5
1

0.82
0.99

0.82
0.99

0.82
0.99

5
1

Category 3: Output AWRI only


Network15

4,672.09

4,847.33

4,743.63

0.50

0.50

0.46

Network16

2,447.19

2,566.56

2,669.63

0.78

0.77

0.77

Network17

87.57

77.81

80.48

0.99

0.99

0.99

Network18

1,655.01

1,834.05

1,703.69

0.85

0.85

0.85

Network19

84.57

91.60

72.05

0.99

0.99

0.99

Network20
Network21

Y
Y

N
Y

Y
Y

N
N

Y
Y

77.17
80.67

78.17
79.87

118.44
97.39

4
3

0.99
0.99

1.00
0.99

0.99
0.99

1
1

P precipitation, E Effective Drought Index, W Available Water Resource Index (AWRI), Tr Training, Vl
Validation, Ts Testing, Y yes (included in the input/targets), N No (not included in the inputs/targets), Rn Rank
(the Rank of the Network by performance (MSE and R), for example, Network7 has the lowest error in the
Category 1 while Network 14 and Network 19 have the lowest error in Category 2 and 3 respectively)
Though the ranking of the networks was based on the average of the performance (MSE and R) of the three
data sets (Tr, Vl and Ts); the value of V1 is preferred because it is the measure of the networks ability to carry
out the future forecasting (classify dataset it has not encountered before).

12 and 14), Network 9 had the lowest MSE (0.0179) as well as the strongest R(0.99068).
Similarly, in the category of networks that compute AWRI, Network 17 had the lowest
MSE(64.9521) and highest R (0.99412). While for Makindu, Network 9 performed better
among the networks with EDI as output but unlike the case of Embu and Dagoretti, Network
21(3 inputs) had the least MSE and Network 19(has EDI as the second input) had the highest
value of R. Since the improvements (51.4371 to 47.7839 (7 %) for MSE and 0.99307 to
0.99361(0.05 %)) were too low considering that extra input(s) values are needed, Network
17 was selected to represent this category.
Networks 5 and 7 have both EDI and AWRI as output and therefore difficult to calculate
the implications of MSE. Besides, these networks had weaker Regression values for all the

Mitig Adapt Strateg Glob Change


Table 4 Performance of the exploratory phases neural networks - Dagoretti
Network

Mean Square Error

Regression

Tr

Vl

Ts

Average

Tr

Vl

Ts

Average

Network5

34.4754

27.7476

38.8559

33.6930

0.98315

0.98450

0.98408

0.98391

Network7

35.1539

35.1539

33.3021

34.5366

0.97997

0.99816

0.98864

0.98892

Network9

0.0188

0.0178

0.0200

0.01887

0.99018

0.99068

0.99001

0.99029

Network11

0.0165

0.0254

0.0149

0.01899

0.99126

0.98751

0.99211

0.99029

Network12

0.0166

0.0199

0.0221

0.0195

0.99138

0.98955

0.98865

0.98986

Network14

0.0174

0.0151

0.0240

0.0188

0.99110

0.99159

0.98793

0.99021

Network17
Network19

68.5308
68.4546

59.6619
81.1977

72.7853
53.5809

66.9927
67.7444

0.99378
0.99373

0.99442
0.99271

0.99334
0.99503

0.99385
0.99382

Network20

64.7377

75.1657

78.6352

72.8462

0.99399

0.99311

0.99332

0.99347

Network21

68.8056

75.3141

60.2966

68.1388

0.99382

0.99237

0.99457

0.99359

weather stations and were therefore not considered in building the final prediction models.
Further, including extra inputs to calculate either EDI or AWRI values did not have
significant improvements on the Networks performance, as such; only the two single-input
Networks (Network 9 for EDI and Network 17 for AWRI) were considered in the next
phases. With MSE values ranging from 3 % to 7 % on EDI, Network 9 generated values that
are within acceptable error-range. For example, an EDI value 4.32 (Extreme Floods) with an
error of 3 % will yield 4.320.13. Similarly, Network 17 has MSE values between 4 % and
8 % and an AWRI value of 194.6 would yield 194.6 8.1
6.4 Exploratory phase II - Forecast models results
Using the results above, neural networks were created to forecast future values of EDI
(Network 9) and AWRI (Network 17) for the short-term, medium-term and long-term timescales as follows:
6.4.1 Short-Term forecasts
(a)

D-Days-Lead-Time Forecast

The lead-time (in days) considered included: 1, 2, 3, 4, 5, 6, 7, 12 and 14. Given a specific
day n, to forecast EDI or AWRI value for d-days from n, the following expression was used:
Forecast[En+d]=f(En, En1, En2, En3, En4, En5) where En is the EDI value for day n;
the latter ranges from 1 to 6. That is, in order to forecast future values, 6 past values are input
into the neural network. In case of AWRI, E is replaced with W. For example to forecast the
value of EDI 5-days Lead-Time from 21st January (26th January), the following expression
applies:
ForecastE215  f E21 ; E20 ; E19 ; E18 ; E17 ; E16
For the purpose of training and testing the neural networks, the daily EDI/AWRI values
for years 1980 to 2009 were used. For each of the 3 weather stations, input files for each of
the neural networks (representing forecast durations above) were created. In all the networks
except 12b and 14b, 6 number of values (EDI or AWRI) were included as inputs to the

Mitig Adapt Strateg Glob Change

neural network models. In an attempt to find out if increasing the number of inputs would
significantly improve the network performance, the number of inputs were increased to 12
and 14 in 12b and 14b respectively. The results of this network model are shown in Figs. 5
and 6; the results indicate that the performance of the network declines as the number of days
being forecast increases; for instance, forecasting 3 days ahead gives a better performance
that forecasting 14 days head.
Increasing the number of inputs (previous days values) did not have a significant
improvement in network performance. For example, increasing number inputs from 6 to
14 resulted in an increase from 13.61 to 13.55 %, 15.21 to 15.29 % and 9.256.52 % for
Dagoretti, Embu and Makindu respectively.
(b)

D-Days-Lead-Time Forecast with Precipitation

In the era of climate change, extreme climate variations may trigger precipitation either
during periods that it does not normally rain or lead to precipitation amounts that are below
or above the expected normal for the area/region. Conventionally, ANNs-based forecasting
solutions learn from past events/patterns and use this knowledge to forecast future trends.
This means that faced with extreme climate variations, ANNs-based drought forecast
solution would result in poor forecast skill. In order to take care of such events, the DDays-Lead-Time Forecasting neural networks described in (a) above were modified to
include forecasted precipitation values for the lead-time considered. Example, to forecast
the drought (EDI and AWRI) values for the 5-days-Lead-Time counting from 21st January
(26th January), the anticipated daily precipitation (as forecasted by professional weather
forecasting institutions such as the Kenya Meteorological Department) for the 5 days are
included as inputs to the networks:
ForecastE215  f E21 ; E20 ; E19 ; E18 ; E17 ; E16 ; P22 ; P23 ; P24 ; P25 ; P26
Where Pi represent the forecasted (approximated) precipitation values
As shown in Figs. 7 and 8, with accuracies of 94.13 to 97.03 %, 93.57 to 97.62 % and
96.25 to 99.09 % for Dagoretti, Embu and Makindu respectively, the networks in the EDI
category (Network 9) displayed excellent performance. Further, the networks had a

Fig. 5 MSE graph for D-Days-Lead-Time for selected weather stations in Kenya

Mitig Adapt Strateg Glob Change

Fig. 6 Regression graph for D-Days-Lead-Time for selected weather stations in Kenya

correlation coefficient values between 0.98 and 0.99. It must be noted that the impressive
performance of the D-Day-Lead-Time Forecast were achieved because the actual historical
precipitation values were used as input. The real networks are expected to use values
forecasted by meteorological institutions. These values have accuracies as low as 70 %
and therefore this will slightly affect the neural networks performance.
6.4.2 Medium and long term forecasts
Similar to the daily EDI, monthly EDI/AWRI values are computed using monthly precipitation totals. This is the data that was used for the following long-term and short-term
forecasts.

Fig. 7 MSE graph for D-Days-Lead-Time (with precipitation) for selected weather stations in Kenya

Mitig Adapt Strateg Glob Change

Fig. 8 Regression graph for D-Days-Lead-Time(with precipitation) for selected weather stations in Kenya

(a)

M-Months-Lead-Time Forecast

Six inputs (previous months) were selected as the standard input in training neural
networks to forecast monthly EDI/AWRI values. The lead-time units considered were 2,
3,4, 5,6,7 and 12 months.
Forecast[En+m]=f(En, En1, En2, En3, En4, En5) where En is the EDI value for month
n; the latter ranges from 1 to 6.
For example to forecast the value of EDI 5-months Lead-Time from January 2012 (that is
forecast for July 2012), the following expression applies:


ForecastEJun2012  f EJan2012 ; EDec2011 ; ENov2011 ; EOct2011 ; ESep2011 ; EAug2011
As shown in Table 5, the performance of the neural networks decreases as the length of
the forecast duration increases. The error rates for majority of the networks are above 30 %
resulting in forecasts that with below 70 % accuracy.
(a)

YYears-Lead-Time Forecasting

Using the expression below, forecast for Y(1,2,3 and 4) years were developed and tested
for the 3 weather stations.
Network 9 : EMonthM;Year y1 f EMonthM;Year y5 ; EMonthM;Year y4 ; EMonthM;Year y3 ;
EMonthM;Year y2 ; EMonthM;Year y1 ; EMonthM;Year y ; PYear y5 ; PYear y4 ; PYear y3 ;
PYear y2 ; PYear y1 ; PYear y ; PYear Yy1
Network 17 : WMonthM; Year n1 f WMonthM;Year n5 ; WMonthM; Year n4 ;
WMonthM; Year n3 ; WMonthM; Year n2 ; WMonthM; Year n1 ; WMonthM; Year n ; PYear n5 ;
PYear n4 ; PYear n3 ; PYear n2 ; PYear n1 ; PYear n ; PYear n1
Illustration: to forecast EDI/AWRI values for August 1986, the neural network requires the
EDI/AWRI values for August 80, 81, 82, 83, 84 and 85. It also requires the annual precipitation
totals for these years as well as an approximate annual precipitation value for year 1986

Mitig Adapt Strateg Glob Change


Table 5 Network performance for the M-Months-Lead-Time Forecast Network 9
LT (D)

MSE

RMSE

%RMSE

0.552

0.556

0.514

0.743

0.746

0.717

26.75

29.01

24.37

0.6868

0.6713

0.7432

0.750

0.758

0.625

0.866

0.871

0.791

31.17

33.86

28.59

0.5483

0.4733

0.6317

0.903

0.947

0.845

0.950

0.973

0.919

34.19

37.86

36.74

0.3288

0.3907

0.4862

0.893

0.658

0.715

0.945

0.811

0.846

34.00

31.55

31.97

0.6583

0.5045

0.6686

0.995

0.730

0.801

0.998

0.855

0.895

35.90

33.24

35.15

0.3551

0.4204

0.5255

0.862

0.997

0.868

0.928

0.999

0.932

33.42

38.84

37.60

0.3359

0.1837

0.2477

12

0.809

1.087

1.009

0.900

1.043

1.004

27.15

34.93

36.27

0.3970

0.2593

0.4995

LT(D) Lead-Time in Days; D Dagoretti Station; E Embu station; M Makindu station; MSE Mean Square Error;
RMSE Root Mean Square Error computed by calculating the square root of the MSE; %RMSE Percentage
Root Mean Square Error calculated by computing the percentage of the actual mean values (EDI or AWRI)
the RMSE represents
Values 2, 3. ,12 represent the lead-time of the forecast in months. For example, entry 5 implies that the
forecast is for the 5th month given the previous 6 months. For example, given the values for June, July,
August, September, October and November 2011, the network forecasts the value for May 2012

6.5 Confirmatory phase Results of evaluating forecast models


From the exploratory phase, the following networks were identified for this confirmation: DDays-Lead-Time Forecasting, D-Days-Lead-Time Forecasting with Precipitation and Y-YearsLead-Time Forecast. Data set for Kakamega weather station with same dimensions as the data
for Dagoretti, Embu and Makindu was used to evaluate the performance of the ANNs models.
For each forecast category, the already trained networks were retrieved and used to evaluate the
respective networks input-output combinations created using the Kakamega data set.
In all the three categories of forecasts, the results showed consistency in performance. For
instance, the D-Days Lead-Time Forecast had accuracies of 88 % to 94 % for EDI and 88 %
to 95 % for AWRI. Just like it was the case in Exploratory Phase, the D-Days Lead-Time
Forecast with precipitation had much higher accuracies of up to 98 %. The fact that the
ANNs models gave these impressive results on data not encountered before helped to reaffirm that the ANNs can indeed be used to enhance the accuracy of drought forecasts.
7 Evaluation and conclusion
7.1 Evaluation
Artificial neural network as implemented in MATLABs Neural Network Toolbox was used
to develop drought prediction models discussed in this paper. The ANNs models developed
accepts as input a combination of EDI/AWRI and precipitation values and generate a set of
EDI/AWRI values representing predicted drought. For example, a 2-Years Lead-Time
(2 years in advance) Forecasting model could be used to forecast the 1998 floods a follows


Forecast EJanuary;1998 f EJanuary;1996; EJanuary;1995 ; ; EJanuary;1994 ; ; EJanuary;1993 ; EJanuary;1992 ;
EJanuary;1991 ; P1991 ; P1992 ; P1993 ; P1994 ; ; P1995 ; ; P1996 ; ; P1997 ; ; P1998

Mitig Adapt Strateg Glob Change

To forecast EDI values for January 1998, the neural network requires the EDI values for
January 91, 92, 93, 94, 95 and 96. It also requires the annual precipitation totals for these
years as well as an approximate annual precipitation value for year 1997 and 1998. Values
for other calendar months are computed in a similar version. That is, given 6 EDI and
precipitation values for 6 years and predicted precipitation values for the lead period (1997
and 1998), the ANNs model for 2-years lead-time outputs the EDI values (Fi) for 1998. The
accuracy of this value (Fi) was determined by its closeness to the actual drought value (Ai)
for this period (1998). Figure 9 is a sample plot of such relationships as generated by the
ANNs model for 2-years lead-time for the period October 1997 to September 1999 using
data for Makindu weather station.
MSE was then used to determine the accuracy of the forecasts. The best performance
was retrieved from the MATLAB Artificial Neural Network Toolkit as illustrated in Fig. 10.
In the case of 2-Years lead-time EDI forecast model for Makindu, the best performance
(MSE=0.37097) was attained after 16 epochs; epoch here refers to the presentation of the
entire training set to the neural network.
Throughout the experiments, validation error was used o evaluate the performance of our
network models. For instance, the data for Makindu for Y-Years Lead Time category (see
Table 6) had MSE values of 0.205, 0.233, 0.267 and 0.270 for 1, 2, 3 and 4 years lead times
respectively. The AWRI models for Makindu for the same network category gave values
ranging from 3,037 to 4,661. In order to put the MSE values into perspective, further he
Percentage RMSE which was easier to understand than the MSE were computerd. A value of
2.89 % implied that the network had an accuracy of 97.11 %. This analysis was performed
for all models and networks with accuracies of over 70 % were considered acceptable.
Networks with correlation values of 0.9 and above were considered acceptable.
7.2 Conclusion
Most of the drought indices applied in forecasting droughts in Kenya fall short of providing
the severity of the drought. In addressing this drawback, a combination of ANNs and EDI
was adopted. This was evaluated by developing forecasting models for time-scales ranging
from one day to four years. The models were then tested using weather data for over 30 years
for 4 weather stations in Kenya. By including forecasted precipitation values as inputs to the
ANNs models, our solution took care of the effects of unprecedented climate variations. It
also exploited the correlation among EDI values of calendar months to forecast annual
droughts. With accuracies raging from 75 % to 98 %, the solution is a great enhancement to

Fig. 9 Predicted versus actual values sample graph

Mitig Adapt Strateg Glob Change

Fig. 10 ANNs evaluation Best MSE sample graph

the solutions currently in use in Kenya. In developing the ANNs models, three phases were
followed:
Phase I: - using 30-years (19792009) data for Embu Station, 21 neural networks were
created using different combinations of Rainfall (Precipitation), EDI and AWRI values.
From these, 10 best performing networks were selected for further investigation during the
Phase II.
Phase II: - Data for three weather stations (Dagoretti, Embu and Makindu) for 30 (1979
2009) years was used to further evaluate the 10 networks selected above. After thorough
analysis, 2 networks (for forecasting EDI and AWRI respectively) were identified and used
to run the following forecasts: (1) Next-Day-Forecast; (2) D-Days-Lead-Time Forecast; (3)
D-Days-Lead-Time With Precipitation Forecast; (4) Next-Month Forecast; (5) M-MonthsLead-Time Forecast (6) Next-Year Forecast; and (7) Y-Years-Lead-Time Forecast. Overall,
the D-Days-Lead-Time With Precipitation Forecast (for up to 7 days Lead-Time) had
accuracies levels of over 94 % while the Y-Years-Lead-Time Forecast (for up to 4 years
Lead-Time) had over 75 % accuracies. The Next-Month Forecast had poorer performance;
Dagoretti and Embu had accuracies of between 7687 % and 72 %80 % for EDI and AWRI
network categories respectively. However, monthly forecasts could still be achieved using
Table 6 Network Performance for the Y-Years-Lead-Time Forecast Network 9
MSE

RMSE

%RMSE

0.224

0.218

0.205

0.473

0.467

0.453

17.03

18.17

11.81

0.8521

0.8761

0.8516

0.305

0.305

0.233

0.552

0.552

0.482

19.87

21.47

13.02

0.8389

0.8618

0.9107

0.294

0.294

0.267

0.542

0.542

0.516

19.51

21.10

14.46

0.7953

0.8504

0.8682

0.285

0.315

0.270

0.533

0.561

0.520

19.20

21.83

14.60

0.8487

0.8167

0.8527

Mitig Adapt Strateg Glob Change

the Y-Years-Lead-Time Forecast. ANNs are known to rely on historical trends to forecast
future values but in the case of droughts, phenomena such as climate change may lead to
erroneous forecasts due to unexpected increase/decrease of precipitation. In this paper, the
ANNs forecasting models have addressed this by incorporating forecasted precipitation
values from issued at professional meteorological institutions such as the KMD. The fact
that the latter forecasts can have accuracies of as low as 70 % has been factored in to the
design and was found that it does not significantly affect accuracies of the ANNs outputs.
Phase III : this is the validation of the results achieved above; data set from Kakamega
weather station (not used before in the previous two phases) was utilised. The resulting
accuracies were similar to those achieved in Phase II.
The work presented in this paper is a part of an on-going research and one of the
immediate tasks is to put the ANNs models described here on to a web-based interface to
allow end users (such the meteorologists and staff at meteorological departments) to access
and use the system. This way, it will be possible to use the ANNs models together with the
seasonal climate forecasts carried out in Kenya and beyond.
Acknowledgement Special thanks to the Kenya Meteorological Department for providing the data set used
in this research.

References
Antoni O, Josip K, Antun M, Dragan B (2001) Ecol Model 138:255263
Armstrong S, Mark C, Randolph K, Dan M (2011) World Disasters Report Focus on Hunger and
Malnutrition. International Federation of Red Cross and Red Crescent Societies. International Federation
of Red Cross and Red Crescent Societies, Geneva
ASCE Task Committee On Application Of Artificial Neural Networks In Hydrology (2000) Artificial Neural
Networks in Hydrology: II. Hydrologic Applications. J Hydrol Eng 5:115123
Bodri L, Cermak V (2001) Prediction of extreme precipitation using a neural network: application to summer
flood occurrence in Moravia. Journal of Environmental Informatics 45:155167
Byun H, Wilhite DA (1999) Objective quantification of drought severity and duration. J Clim 12:27472756
CERF (Central Emergency Response Fund) (2009) CERF funds jump start emergency aid operations in
Kenya. United Nations Economic Commission for Africa. Addis Ababa, Ethiopia
Chang-Shian C, Boris P, Chen F, Chou N, Chao-Chung Y (2010) Development and application of a decision
group Back-Propagation Neural Network for flood forecasting. J Hydrol 385:173182
Collins A, Nick M, Michele M, Anne M, Maarten VA (2009) World Disasters Report Focus on
Early Warning, Action, 1st edn. International Federation of Red Cross and Red Crescent Societies,
Geneva
Deely S, David D, Jorgelina H, Cassidy J (2010) World Disasters Report Focus on Urban Risk, 1st edn.
International Federation of Red Cross and Red Crescent Societies, Geneva
Demuth, H., Mark, B. and Martin, H. 2009. MATLAB Neural Network Toolbox Users Guide. 6
Dracup JA, KS L, KS L, Paulson EG Jr (1980) On the definition of droughts. Water Resour Res 16:289296
Elsa E, Moreira C, Coelho A, Paulo AA (2008) SPI-Based Drought Category Predication Using Loglinear
Models. J Hydrol 354:116130
Freitas, M.A.S. and Billib, M.H.A. 1997. Drought prediction and characteristic analysis in semiarid Cear,
northeast Brazil. Sustainability of Water Resources under Increasing Uncertainty, Anonymous IAHS, No.
240, , 105.
Huschke, R. E., 1970: Glossary of Meteorology. Amer. Meteor. Soc., 638 pp.
Kim D, Hi-Ryong B, Ki-Seon C (2009) Evaluation, modification, and application of the Effective Drought
Index to 200-Year drought climatology of Seoul, Korea. J Hydrol 378:112
Kung H, Hua J, Chen C (2006) Drought forecast model and framework using wireless sensor networks. J Inf
Sci Eng 22:751769
Maier HR, Dandy GC (2000) Neural networks for the prediction and forecasting of water resources variables:
a review of modelling issues and applications. Environ Model Softw 15(1):101124

Mitig Adapt Strateg Glob Change


Masinde, M. and Bagula, A., 2011. The Role of ICTs in Quantifying the Severity and Duration of Climatic
Variations Kenyas Case, Proceedings of ITU Kaleidoscope 2011: The Fully Networked Human? Innovations for Future Networks and Services (K-2011), 1214 December 2011, IEEE Xplore, pp. 18
Masinde M, Bagula A (2012) ITIKI: bridge between African indigenous knowledge and modern science of
drought prediction. Knowl Manag Dev J 7(03):274290
Mcculloch WS, Pitts W (1943) A logical calculus of the ideas imminent in nervous activity. Bull Math
Biophys 5:115133
Mckee TB, Doesken NJ, Kleist J (1993) The relationship of drought frequency and duration to time scales. In:
8th Conference on Applied Climatology, January. Anonymous American Meteorological Society, Boston,
Massachusetts, pp 1723
Mishra AK, Desai VR (2006) Drought Forecasting Using Fee-Forward Recursive Neural Network. Ecol
Model 198:128138
Mishra AK, Singn VP (2010) A Review of Drought Concepts. J Hydrol 391:202216
Morid S, Smakhtin V, Bagherzadeh K (2007) Drought forecasting using artificial neural networks and time
series of drought indices. Int J Climatol 27:21032111
Mutua, S., 2011. Strengthening Drought Early Warning At The Community and District Levels: Analysis Of
Traditional Community Warning Systems In Wajir & Turkana
Mwagore D (2002) Land use in Kenya: The case for a national land use policy. Printfast Kenya Ltd., Nairobi
Palmer, W.C., 1965. Meteorological Drought. Research Paper 45. US Department of Commerce, Weather
Bureau, Washington, DC, pp. 158.
Panu, U.S. and Sharma, T.C. 2002. Challenges in drought research: some perspectives and future directions.
Hydrological Sciences-Journal .
Rumelhart DE, Hinton GE, Williams RJ (1986) Learning internal representations by error propagation.
Parallel distributed processing, Vol. 1. MIT Press, Cambridge, pp 318362
Shin HS, Salas JD (2000) Regional drought analysis based on neural networks. J Hydrol Eng 5(2):145155
Weichert A, Burger G (1998) Linear versus nonlinear techniques in downscaling. Clim Res 10:8393
Wells N, Goddard S, Hayes MJ (2004) A self-calibrating palmer drought severity index. J Clim 17:23352351
WHO Collaborating Centre For Research On The Epidemiology of Disasters (CRED), 2012-last update,
Emergency Events Database (EM-DAT) [Homepage of Government of Belgium, for WHO], [Online].
Available: http://www.emdat.be/ [May/11, 2012].
Wilhite DA, Glantz MH (1985) Understanding the drought phenomenon: the role of definitions. Water
International 10:111120
Yin RK (1993) Applications of Case Study Research. Sage, Newbury Park

Você também pode gostar