scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Evaluating the performance of infectious disease forecasts: A comparison of climate-driven and seasonal dengue forecasts for Mexico

TL;DR: A framework to assess and compare dengue forecasts produced from different types of models and evaluated the performance of seasonal autoregressive models with and without climate variables for forecasting d Dengue incidence in Mexico found short-term and seasonal autocorrelation were key to improving short- term and long-term forecasts.
Abstract: Dengue viruses, which infect millions of people per year worldwide, cause large epidemics that strain healthcare systems. Despite diverse efforts to develop forecasting tools including autoregressive time series, climate-driven statistical, and mechanistic biological models, little work has been done to understand the contribution of different components to improved prediction. We developed a framework to assess and compare dengue forecasts produced from different types of models and evaluated the performance of seasonal autoregressive models with and without climate variables for forecasting dengue incidence in Mexico. Climate data did not significantly improve the predictive power of seasonal autoregressive models. Short-term and seasonal autocorrelation were key to improving short-term and long-term forecasts, respectively. Seasonal autoregressive models captured a substantial amount of dengue variability, but better models are needed to improve dengue forecasting. This framework contributes to the sparse literature of infectious disease prediction model evaluation, using state-of-the-art validation techniques such as out-of-sample testing and comparison to an appropriate reference model.

Content maybe subject to copyright    Report

Evaluating the performance of infectious disease
forecasts: A comparison of climate-driven and
seasonal dengue forecasts for Mexico
Citation
Johansson, Michael A., Nicholas G. Reich, Aditi Hota, John S. Brownstein, and Mauricio
Santillana. 2016. “Evaluating the performance of infectious disease forecasts: A comparison
of climate-driven and seasonal dengue forecasts for Mexico.” Scientific Reports 6 (1): 33707.
doi:10.1038/srep33707. http://dx.doi.org/10.1038/srep33707.
Published Version
doi:10.1038/srep33707
Permanent link
http://nrs.harvard.edu/urn-3:HUL.InstRepos:29407787
Terms of Use
This article was downloaded from Harvard University’s DASH repository, and is made available
under the terms and conditions applicable to Other Posted Material, as set forth at
http://
nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of-use#LAA
Share Your Story
The Harvard community has made this article openly available.
Please share how this access benefits you.
Submit a story .
Accessibility

1
Scientific RepoRts | 6:33707 | DOI: 10.1038/srep33707
www.nature.com/scientificreports
Evaluating the performance of
infectious disease forecasts: A
comparison of climate-driven and
seasonal dengue forecasts for
Mexico
Michael A. Johansson
1,2
, Nicholas G. Reich
3
, Aditi Hota
4
, John S. Brownstein
4,5
&
Mauricio Santillana
4,5,6
Dengue viruses, which infect millions of people per year worldwide, cause large epidemics that strain
healthcare systems. Despite diverse eorts to develop forecasting tools including autoregressive
time series, climate-driven statistical, and mechanistic biological models, little work has been done
to understand the contribution of dierent components to improved prediction. We developed a
framework to assess and compare dengue forecasts produced from dierent types of models and
evaluated the performance of seasonal autoregressive models with and without climate variables
for forecasting dengue incidence in Mexico. Climate data did not signicantly improve the predictive
power of seasonal autoregressive models. Short-term and seasonal autocorrelation were key to
improving short-term and long-term forecasts, respectively. Seasonal autoregressive models captured a
substantial amount of dengue variability, but better models are needed to improve dengue forecasting.
This framework contributes to the sparse literature of infectious disease prediction model evaluation,
using state-of-the-art validation techniques such as out-of-sample testing and comparison to an
appropriate reference model.
Dengue is a substantial public health problem in most of the tropical and subtropical regions of the world
1
. In
most of these areas, dengue is endemic with cases occurring year-round, yet there is marked variation in inci-
dence of dengue both within and between years. In Mexico, for example, yearly reported incidence over the
last few decades has varied from several thousand cases to over 100,000 cases in 2009
2
. Even understanding the
current burden of dengue can be challenging. ere is oen an extended delay between symptom onset and o-
cial reports reecting care-seeking behavior or conrmed clinical reports of illness. In some settings, complex,
multi-tiered reporting systems may contribute to delays. Because of the vastly greater burden of disease in larger
epidemic years and diculty in understanding current and future needs, much eort has been placed on devel-
oping early warning systems to predict or detect large epidemics as early as possible with the hope of being able
to control epidemics in their early stages
3,4
.
Numerous mechanistic models have been developed to use detailed knowledge of dengue virus transmission
biology to predict the evolution of dengue epidemics
5,6
. However diculties emerge when attempting to use
mechanistic models for forecasting, oen the key mechanistic assumptions are unclear and the data needed to
parameterize and feed the model are oen dicult or impossible to obtain.
1
Dengue Branch, Division of Vector-Borne Diseases, Centers for Disease Control and Prevention, San Juan, Puerto Rico.
2
Center for Communicable Disease Dynamics, Harvard T. H. Chan School of Public Health, Boston, Massachusetts,
USA.
3
Department of Biostatistics and Epidemiology, University of Massachusetts, Amherst, Massachusetts, USA.
4
Computational Health Informatics Program, Boston Children’s Hospital, Boston, Massachusetts, USA.
5
Department
of Pediatrics, Harvard Medical School, Boston, Massachusetts, USA.
6
J.A. Paulson School of Engineering and Applied
Sciences, Harvard University, Cambridge, MA, USA. Correspondence and requests for materials should be addressed
to M.A.J. (email: eyq9@cdc.gov) or M.S. (email: msantill@g.harvard.edu)
Received: 17 March 2016
Accepted: 24 August 2016
Published: 26 September 2016
OPEN

www.nature.com/scientificreports/
2
Scientific RepoRts | 6:33707 | DOI: 10.1038/srep33707
Models for large-scale dengue early warning systems have therefore mostly focused on two components, temporal
autocorrelation and an association with weather or climate. Temporal autocorrelation results from the infectious
nature of dengue viruses; cases are more likely in the near future when the current prevalence of infection
is high. e inuence of weather is due to the mosquito vectors of dengue viruses, Aedes aegypti and Ae. albopictus.
Temperature, humidity, and precipitation are important determinants of mosquito reproduction and longevity
7–9
,
and temperature has a strong inuence on the ability of the mosquitoes to transmit dengue viruses
10
.
ese two components, autocorrelation and weather or climate, form the basis for a wide variety of eorts to
predict dengue incidence in countries such as Australia
11
, Bangladesh
12
, Barbados
13
, Brazil
14–17
, Colombia
18–20
,
Costa Rica
21
, China
22
, Ecuador
23
, Guadeloupe
24
, India
25
, Indonesia
26,27
, Mexico
28
, New Caledonia
29
, Peru
30
,
Philippines
31,32
, Puerto Rico
33
, Singapore
34–38
, Sri Lanka
39
, Taiwan
40,41
, ailand
42–46
, and Vietnam
47
.
In this manuscript, we focus on developing prediction models for dengue incidence in Mexico based on
observed dengue incidence and weather. We focus explicitly on building models that directly predict dengue
incidence rather than those designed to classify future transmission, for example as low or high
29,30,34,42
. Although
classication may be more useful for public health decision-making, the classication process is subjective
48
,
making estimation of uncertainty a less straightforward process compared to direct prediction of incidence.
Autoregressive integrated moving average (ARIMA) models
49
have been used extensively in dengue predic-
tion eorts
14,18,19,35,37,40,45
, oen incorporating a seasonal component (SARIMA)
11,14,15,24,25,27,43,46,47
. We used the
SARIMA framework to dene a dynamic suite of prediction models for dengue in Mexico that is at once exible
and wide-ranging, while remaining manageable in size for the purposes of careful evaluation and comparison.
While the fundamental biology leading to autocorrelation and associations with weather and dengue incidence
is clear, the specic contributions of these two components to dengue prediction remains less so. In other words,
we wanted to answer the question, to what extent does incorporating climate data into models improve den-
gue forecasts at dierent spatial scales? erefore, we assessed three specic features of these dengue prediction
models: (i) the most important autoregressive and climatological components for predicting dengue incidence;
(ii) the variability in importance of these components across dierent geographical areas; and (iii) the limits of
prediction accuracy across models at dierent time horizons. Another important motivation of this study was
to establish a forecast assessment framework that can serve as a reference for any infectious disease forecasting
problem, including comparison to a non-naïve baseline model and validation on completely out-of-sample data.
Materials and Methods
Data. e number of dengue and dengue hemorrhagic fever cases reported for each month from January 1985
to December 2012 was obtained from the Mexican Health Secretariat
2
. To model these counts as a linear process,
we log-transformed the highly skewed monthly dengue cases aer adding one to the observed count. is elimi-
nates computational problems associated with taking the log of zero, and can be thought of as accounting for the
potential entrance of new cases
50
. Weather data from January 1985 to December 2012 were obtained from the
National Oceanic and Atmospheric Administration North American Regional Reanalysis dataset (www.esrl.noaa.
gov/psd/ accessed on May 1
st
, 2013). For each month and state, we extracted the average temperature (°C), daily
precipitation (mm), and relative humidity (%).
Models. We analyzed three dierent types of temporal models: linear models, autoregressive (AR) models,
and seasonal autoregressive models (SAR). We label these models as (p,d,q)(P,D,Q)
s
+ covar
L
. For the ARIMA
component, (p,d,q), p indicates the autoregressive order, d indicates dierencing, and q is the order of the moving
average
49
. For the seasonal component, (P,D,Q)
s
, s indicates the season length (s = 12 months for the monthly data
presented here), P indicates the autoregressive order, D indicates dierencing, and Q is the order of the moving
average. e nal component, + covar
L
, indicates a particular covariate at a lag of L months included as a linear
regression term.
We used a systematic procedure to select SARIMA models to t to the data and evaluate for predictive per-
formance. Each time series was assessed for stationarity using the Augmented Dickey-Fuller test
51
. We used
domain-specic knowledge about infectious disease models to create a limited model space that we could explore
fully and justify scientically. Dening the model space in this way mitigated the risk of overtting the model in
the training period. Specically, we chose to focus on models that (a) included lagged observations of up to three
months (p = 1, 2, or 3) and three years (P = 0, 1, 2, or 3), (b) included dierencing terms of up to order 1 (i.e. d or
D = 0 or 1), and (c) excluded all consideration of moving averages (i.e. q and Q both xed at 0).
Prediction models that include a dierencing term (i.e. d = 1 or D = 1) have a term added to the model for-
mula that captures the dierence between the lag-1 and lag-2 case counts. ese models therefore incorporate
information about the most recently observed slope of dengue incidence: e.g. are reported cases increasing or
decreasing? In some formulations, this can be considered a crude approximation to the reproductive rate of the
disease, an important parameter for judging the trajectory of the outbreak
52
. As an example of a formula for a
model including dierencing, a (1,0,0)(2,1,0)
12
multiplicative SARIMA model can be expressed as:
αφ αφ
φαφ
=+ −+ −−
+−−−+
−−−−−−
−− −−
XX XX XX XX
XX XX
()()()
()()
(1)
tt tt tt tt
tt tt t
12 11 13 11224111325
22436122537
where
X
t
are the observed numbers of cases of dengue at time
t
;
t
are the residual error terms, assumed to be
normally distributed; and the coecients
α
1
,
φ
1
, and
are determined to minimize the residual squared error.
To evaluate the added predictive value of weather variables, we examined the best SARIMA models from the
model space dened above by adding individual weather variables as covariates, with lags of 1, 2, and 3 months
and the month with the maximum Pearson correlation with dengue incidence (labeled with the sub-script MX
in the gures), and assessing the change in predictive performance. Our aim was to single out the eect of each

www.nature.com/scientificreports/
3
Scientific RepoRts | 6:33707 | DOI: 10.1038/srep33707
variable (precipitation, relative humidity, and temperature). Additional models including multiple weather vari-
ables at once were considered. Ultimately, the models presented represent a given SARIMA model plus the most
highly correlated lagged weather covariate. All models were tted in R
53
, using the arima function from the stats
package.
Model evaluation. e data was separated into three subsets. e rst ve years of data (1985-1989) was
reserved for model training. Models trained on that data were used to dynamically predict dengue incidence
over an 18-year model evaluation period (1990–2007). e nal ve-year dataset (2008–2012) was reserved for a
complete out-of-sample validation aer model selection.
At the end of the training period, predictions were made for the next 1–6 months (e.g., predictions are made
for January–June based on data through December of the preceding year). en the model was retted with data
from the next month (i.e., January) and predictions were made for the following 1–6 months (February–July).
is process was repeated over all months of the evaluation period. Predictions were evaluated by comparing
observed incidence with model-predicted incidence using two metrics: mean absolute error (MAE) and the coef-
cient of determination (R
2
). For comparisons of predictions across dierent locations and predictions hori-
zons, we calculated the change in MAE relative to the best prediction among all models for a given location and
horizon:
=mm MrelMAE() MAE( )/min(MAE( ))
(2)
ii
where m
i
is a single model in the set of models M for a given time and location. us the model with the best point
prediction has relMAE = 1 and predictions further from the observations will have increasing relMAE values.
To compare two specic models (m
1
and m
2
), we calculated the relMAE:
=mm mmrelMAE(, )MAE()/MAE()
(3)
12 12
State-level models. We systematically assessed the predictive performance of a wide range of models at
dierent spatial scales, using data aggregated at the national level and at the level of states. e primary interest
of this manuscript is to assess models for forecasting dengue incidence in endemic locations, so we restricted the
analysis to the 17 states reporting cases in at least half of the months in the training and development time period
(1985–2007). Some of the other states, such as Sonora, Coahuila, and San Luis Potosi, had high median annual
incidence over this period due to sporadic outbreaks but reported cases in less than 50% of the months.
Using the selection algorithm described briey above in the “Models” subsection and in more detail below
in the “National dengue forecasts” subsection, we tted and evaluated 39 models at each spatial scale. e same
selection process was repeated for each location separately. We assessed the predictions of all models across each
location using the average relative MAE and R
2
for each prediction horizon. For each state we selected an optimal
model for short-term forecasts by identifying the model with minimum average relative MAE across the 1–3
month prediction horizons, for each state we refer to this model as the “local” model. Assessing model perfor-
mance across all spatial scales in the training period, we identied a single “common” model that performed well
across all spatial scales. During the prospective out-of-sample forecast phase, aer model selection, we compared
the performance between the common and local models.
Results
National dengue forecasts. During the 5-year training period (1985–1989) and 18-year evaluation period
(1990–2007) the monthly dengue incidence in Mexico showed strong seasonality (Fig.1A). Incidence also varied
substantially between years, with less than 2,000 cases reported in 2000 and more than 50,000 reported in 1997.
We first compared two naïve models for predicting national dengue incidence several months into the
future, one estimating that future incidence follows the historical average incidence and another assuming that
future incidence for a specic month will be the historical average for that particular month of the calendar year
(Fig.1B). ese models were implemented at the beginning of the evaluation period, with predictions made for
months 1 through 6 of that period. An additional month of data was then acquired (month 1), the model was
retted, and predictions for the next 6 months (i.e., months 2–7) were made, ensuring that all forecasts were made
on out-of-sample data. As predictions were made at each month throughout the 18-year evaluation period, a total
of 216 predictions were analyzed for each prediction horizon and each model. For all models, we used the same
dynamically increasing training set strategy to calculate the respective out-of-sample predictions, strictly avoiding
the use of forward-looking information. is allowed the use of predictive metrics to compare models directly.
Across all of these out-of-sample predictions, the model using the month of the year outperforms the
long-term mean model by both MAE (lower error) and R
2
(higher correlation). Despite being relatively naïve,
predictions from the seasonal model have an R
2
of approximately 0.24.
We then assessed twelve linear models including data on average temperature, daily precipitation, and relative
humidity in previous months. e maximum cross-correlation between each of these weather variables and den-
gue incidence was 1 month for relative humidity, 3 months for precipitation, and 4 months for temperature. We
assessed lags of 1–3 months for each variable and also a 4-month lag for temperature. e 4-month lag tempera-
ture model performed best, but did not outperform the simple monthly model by either metric.
Next we assessed sixteen models with only autocorrelation and dierencing terms: eight models including
only short-term autocorrelation and eight models including both short-term and seasonal autocorrelation. e
rst-order AR model (1,0,0)(0,0,0)
12
, substantially improved the 1-month predictions compared to all previous
models. At 2-months, the predictions were less accurate than those from the monthly model and some of the
covariate models. Increasing the order of the AR model slightly improved the predictions, while dierencing

www.nature.com/scientificreports/
4
Scientific RepoRts | 6:33707 | DOI: 10.1038/srep33707
decreased their accuracy. Adding a seasonal component markedly improved the predictions; the (1,0,0)(1,0,0)
12
SAR model outperformed all of the previous models with lower MAE and higher R
2
at all prediction times.
Increasing the short-term AR component slightly decreased accuracy and 1-month dierencing decreased accu-
racy substantially. Adding a seasonal dierencing term however, improved accuracy by both metrics for longer
prediction horizons (e.g. (1,0,0)(1,1,0)
12
). Increasing the order of the seasonal AR term to (1,0,0)(3,1,0)
12
further
improved predictions.
Finally, we assessed nine AR and SAR models with weather covariates. At all prediction horizons, models
containing rst order autoregressive terms and covariates consistently outperformed models containing only the
covariate information (e.g. (1,0,0)(0,0,0)
12
+ temp
MX
versus temp
MX
) and simple AR models (e.g. (1,0,0)(0,0,0)
12
+
temp
MX
versus (1,0,0)(0,0,0)
12
). e SAR models with covariates have very similar accuracy to the SAR models
without covariates at 1–4 month prediction horizons. For 5- and 6-month horizons, the covariate models could
not make predictions as they extended beyond the 4-month lag for temperature. Additional analyses showed that
once about 12 years of training data were used, the predictive performance of the models reached a stable average
value (between 0.85 and 0.9 correlation), although year-specic values showed some variation around the average
(data not shown). Finally, in a non-exhaustive search, we noticed that using all weather variables at once as input
in a given model did not yield noticeable improvements.
Overall, the best performing models were (1,0,0)(3,1,0)
12
for horizons of 1–2 months, (1,0,0)(3,1,0)
12
+ temp
MX
for the 3-month horizon, (1,0,0)(1,1,0)
12
for the 4-month horizon, and (2,0,0)(1,1,0)
12
for 5- and 6-month hori-
zons (Fig.1C). e best model across all horizons was (1,0,0)(3,1,0)
12
, followed by (1,0,0)(1,1,0)
12
and (1,0,0)
(2,1,0)
12
.
State-level dengue forecasts. Dengue incidence varies substantially between states within Mexico
(Fig.2). Similar to the national-level results, the rst-order AR model (1,0,0)(0,0,0)
12
outperformed the monthly
model at shorter prediction horizons of 1–3 months (Fig.3, Supplementary Figure 1). However across all hori-
zons and states, the most accurate model was the (1,0,0)(2,1,0)
12
model, which we now refer to as the “common
model. e coecients for the (1,0,0)(2,1,0)
12
model in each state varied, but the pattern was consistent across all
Figure 1. National-level forecast metrics. National-level incidence is shown during the training period
(1985–1989) and evaluation period (1990–2007). (A) For each of 39 models considered, the MAE (B) and R
2
(C) values for prospective forecasts over the entire evaluation period are shown for each prediction horizon
(dark red to yellow, corresponds to 1–6 months). For models including lagged weather covariates, forecasts
were not possible at prediction horizons beyond the lag and are not shown. An equivalent plot for each state
is shown in Figure S1.

Citations
More filters
Posted Content
TL;DR: From smart grids to disaster management, high impact problems where existing gaps can be filled by ML are identified, in collaboration with other fields, to join the global effort against climate change.
Abstract: Climate change is one of the greatest challenges facing humanity, and we, as machine learning experts, may wonder how we can help. Here we describe how machine learning can be a powerful tool in reducing greenhouse gas emissions and helping society adapt to a changing climate. From smart grids to disaster management, we identify high impact problems where existing gaps can be filled by machine learning, in collaboration with other fields. Our recommendations encompass exciting research questions as well as promising business opportunities. We call on the machine learning community to join the global effort against climate change.

441 citations


Cites background from "Evaluating the performance of infec..."

  • ...Disease surveillance and outbreak forecasting systems can be built from web data and specially-designed apps, in addition to traditional surveys [626–628]....

    [...]

  • ...[628] Michael A Johansson, Nicholas G Reich, Aditi Hota, John S Brownstein, and Mauricio Santillana....

    [...]

Journal ArticleDOI
TL;DR: In this paper, the authors adopt permutation entropy as a model independent measure of predictability and identify a fundamental entropy barrier for disease time series forecasting, which is often beyond the time scale of single outbreaks, implying prediction is likely to succeed.
Abstract: Infectious disease outbreaks recapitulate biology: they emerge from the multi-level interaction of hosts, pathogens, and environment. Therefore, outbreak forecasting requires an integrative approach to modeling. While specific components of outbreaks are predictable, it remains unclear whether fundamental limits to outbreak prediction exist. Here, adopting permutation entropy as a model independent measure of predictability, we study the predictability of a diverse collection of outbreaks and identify a fundamental entropy barrier for disease time series forecasting. However, this barrier is often beyond the time scale of single outbreaks, implying prediction is likely to succeed. We show that forecast horizons vary by disease and that both shifting model structures and social network heterogeneity are likely mechanisms for differences in predictability. Our results highlight the importance of embracing dynamic modeling approaches, suggest challenges for performing model selection across long time series, and may relate more broadly to the predictability of complex adaptive systems.

161 citations

Journal ArticleDOI
TL;DR: Internet-based data streams can be used as timely and complementary ways to assess the dynamics of the outbreak of Zika virus and show the predictive power of these data and a dynamic multivariable approach.
Abstract: Background Over 400,000 people across the Americas are thought to have been infected with Zika virus as a consequence of the 2015–2016 Latin American outbreak. Official government-led case count data in Latin America are typically delayed by several weeks, making it difficult to track the disease in a timely manner. Thus, timely disease tracking systems are needed to design and assess interventions to mitigate disease transmission. Methodology/Principal Findings We combined information from Zika-related Google searches, Twitter microblogs, and the HealthMap digital surveillance system with historical Zika suspected case counts to track and predict estimates of suspected weekly Zika cases during the 2015–2016 Latin American outbreak, up to three weeks ahead of the publication of official case data. We evaluated the predictive power of these data and used a dynamic multivariable approach to retrospectively produce predictions of weekly suspected cases for five countries: Colombia, El Salvador, Honduras, Venezuela, and Martinique. Models that combined Google (and Twitter data where available) with autoregressive information showed the best out-of-sample predictive accuracy for 1-week ahead predictions, whereas models that used only Google and Twitter typically performed best for 2- and 3-week ahead predictions. Significance Given the significant delay in the release of official government-reported Zika case counts, we show that these Internet-based data streams can be used as timely and complementary ways to assess the dynamics of the outbreak.

151 citations

Journal ArticleDOI
TL;DR: This study compares machine learning-based prediction models to commonly used regression models for prediction of undiagnosed T2DM and shows no clinically relevant improvement when more sophisticated prediction models were used.
Abstract: Most screening tests for T2DM in use today were developed using multivariate regression methods that are often further simplified to allow transformation into a scoring formula. The increasing volume of electronically collected data opened the opportunity to develop more complex, accurate prediction models that can be continuously updated using machine learning approaches. This study compares machine learning-based prediction models (i.e. Glmnet, RF, XGBoost, LightGBM) to commonly used regression models for prediction of undiagnosed T2DM. The performance in prediction of fasting plasma glucose level was measured using 100 bootstrap iterations in different subsets of data simulating new incoming data in 6-month batches. With 6 months of data available, simple regression model performed with the lowest average RMSE of 0.838, followed by RF (0.842), LightGBM (0.846), Glmnet (0.859) and XGBoost (0.881). When more data were added, Glmnet improved with the highest rate (+ 3.4%). The highest level of variable selection stability over time was observed with LightGBM models. Our results show no clinically relevant improvement when more sophisticated prediction models were used. Since higher stability of selected variables over time contributes to simpler interpretation of the models, interpretability and model calibration should also be considered in development of clinical prediction models.

134 citations

Journal ArticleDOI
Michael A. Johansson1, Michael A. Johansson2, Karyn M. Apfeldorf, Scott Dobson, Jason Devita, Anna L. Buczak3, Benjamin Baugher3, Linda J. Moniz3, Thomas Bagley3, Steven M. Babin3, Erhan Guven3, Teresa K. Yamana4, Jeffrey Shaman4, Terry Moschou5, Nick Lothian5, Aaron Lane5, Grant Osborne5, Gao Jiang6, Logan C. Brooks6, David C. Farrow6, Sangwon Hyun6, Ryan J. Tibshirani6, Roni Rosenfeld6, Justin Lessler7, Nicholas G. Reich8, Derek A. T. Cummings9, Stephen A. Lauer8, Sean M. Moore10, Hannah E. Clapham11, Rachel Lowe12, Trevor C. Bailey13, Markel García-Díez, Marilia Sá Carvalho14, Xavier Rodó, Tridip Sardar15, Richard Paul15, Evan L. Ray16, Krzysztof Sakrejda8, Alexandria C. Brown8, Xi Meng8, Osonde A. Osoba17, Raffaele Vardavas17, David Manheim, Melinda Moore17, Dhananjai M. Rao18, Travis C. Porco19, Sarah F Ackley19, Fengchen Liu19, Lee Worden19, Matteo Convertino20, Yang Liu21, Abraham Reddy21, Eloy Ortiz, Jorge Rivero, Humberto Brito22, Alicia Juarrero23, Leah R. Johnson24, Robert B. Gramacy25, Jeremy M. Cohen25, Erin A. Mordecai26, Courtney C. Murdock27, Jason R. Rohr10, Sadie J. Ryan9, Sadie J. Ryan28, Anna M. Stewart-Ibarra29, Daniel P. Weikel30, Antarpreet Jutla31, Rakibul Khan31, Marissa Poultney31, Rita R. Colwell32, Brenda Rivera-Garcia33, Christopher M. Barker33, Jesse E. Bell34, Matthew Biggerstaff1, David L. Swerdlow1, Luis Mier-y-Teran-Romero7, Luis Mier-y-Teran-Romero1, Brett M. Forshey, Juli Trtanj35, Jason Asher35, Matt Clay35, Harold S. Margolis1, Andrew M. Hebbeler36, Dylan B. George37, Jean Paul Chretien38, Jean Paul Chretien37 
TL;DR: An open collaborative forecasting challenge to assess probabilistic forecasts for seasonal epidemics of dengue, a major global public health problem, revealed that average forecast skill was lower for models including biologically meaningful data and mechanisms and that both multimodel and multiteam ensemble forecasts consistently outperformed individual model forecasts.
Abstract: A wide range of research has promised new tools for forecasting infectious disease dynamics, but little of that research is currently being applied in practice, because tools do not address key public health needs, do not produce probabilistic forecasts, have not been evaluated on external data, or do not provide sufficient forecast skill to be useful. We developed an open collaborative forecasting challenge to assess probabilistic forecasts for seasonal epidemics of dengue, a major global public health problem. Sixteen teams used a variety of methods and data to generate forecasts for 3 epidemiological targets (peak incidence, the week of the peak, and total incidence) over 8 dengue seasons in Iquitos, Peru and San Juan, Puerto Rico. Forecast skill was highly variable across teams and targets. While numerous forecasts showed high skill for midseason situational awareness, early season skill was low, and skill was generally lowest for high incidence seasons, those for which forecasts would be most valuable. A comparison of modeling approaches revealed that average forecast skill was lower for models including biologically meaningful data and mechanisms and that both multimodel and multiteam ensemble forecasts consistently outperformed individual model forecasts. Leveraging these insights, data, and the forecasting framework will be critical to improve forecast skill and the application of forecasts in real time for epidemic preparedness and response. Moreover, key components of this project-integration with public health needs, a common forecasting framework, shared and standardized data, and open participation-can help advance infectious disease forecasting beyond dengue.

132 citations

References
More filters
Journal ArticleDOI
TL;DR: Time series analysis san francisco state university, 6 4 introduction to time series analysis, box and jenkins time seriesAnalysis forecasting and, th15 weeks citation classic eugene garfield, proc arima references 9 3 sas support, time series Analysis forecasting and control pambudi, timeseries analysis forecasting and Control george e.
Abstract: time series analysis san francisco state university, 6 4 introduction to time series analysis, box and jenkins time series analysis forecasting and, th15 weeks citation classic eugene garfield, proc arima references 9 3 sas support, time series analysis forecasting and control pambudi, time series analysis forecasting and control george e, time series analysis forecasting and control ebook, time series analysis forecasting and control 5th edition, time series analysis forecasting and control fourth, time series analysis forecasting and control amazon, wiley time series analysis forecasting and control 5th, time series analysis forecasting and control edition 5, time series analysis forecasting and control 5th edition, time series analysis forecasting and control abebooks, time series analysis for business forecasting, time series analysis forecasting and control wiley, time series analysis forecasting and control book 1976, time series analysis forecasting and control researchgate, time series analysis forecasting and control edition 4, time series analysis forecasting amp control forecasting, george box publications department of statistics, time series analysis forecasting and control london, time series analysis forecasting and control an, time series analysis forecasting and control amazon it, box g e p and jenkins g m 1976 time series, time series analysis forecasting and control pdf slideshare, time series analysis forecasting and control researchgate, time series analysis forecasting and control 5th edition, time series analysis forecasting and control 5th edition, time series wikipedia, time series analysis forecasting and control abebooks, time series analysis forecasting and control, forecasting and time series analysis using the sca system, time series analysis forecasting and control by george e, time series analysis forecasting and control 5th edition, time series analysis forecasting and control 5th edition, box and jenkins time series analysis forecasting and control, time series analysis forecasting and control ebook, time series analysis forecasting and control, time series analysis and forecasting cengage, 6 7 references itl nist gov, time series analysis forecasting and control george e, time series analysis and forecasting statgraphics, time series analysis forecasting and control fourth edition, time series analysis forecasting and control, time series analysis forecasting and control wiley, time series analysis forecasting and control in

10,118 citations

Book
24 Aug 2012
TL;DR: This textbook offers a comprehensive and self-contained introduction to the field of machine learning, based on a unified, probabilistic approach, and is suitable for upper-level undergraduates with an introductory-level college math background and beginning graduate students.
Abstract: Today's Web-enabled deluge of electronic data calls for automated methods of data analysis. Machine learning provides these, developing methods that can automatically detect patterns in data and then use the uncovered patterns to predict future data. This textbook offers a comprehensive and self-contained introduction to the field of machine learning, based on a unified, probabilistic approach. The coverage combines breadth and depth, offering necessary background material on such topics as probability, optimization, and linear algebra as well as discussion of recent developments in the field, including conditional random fields, L1 regularization, and deep learning. The book is written in an informal, accessible style, complete with pseudo-code for the most important algorithms. All topics are copiously illustrated with color images and worked examples drawn from such application domains as biology, text processing, computer vision, and robotics. Rather than providing a cookbook of different heuristic methods, the book stresses a principled model-based approach, often using the language of graphical models to specify models in a concise and intuitive way. Almost all the models described have been implemented in a MATLAB software package--PMTK (probabilistic modeling toolkit)--that is freely available online. The book is suitable for upper-level undergraduates with an introductory-level college math background and beginning graduate students.

8,059 citations

Journal ArticleDOI
01 May 1971

7,355 citations

Journal ArticleDOI
19 Feb 2009-Nature
TL;DR: A method of analysing large numbers of Google search queries to track influenza-like illness in a population and accurately estimate the current level of weekly influenza activity in each region of the United States with a reporting lag of about one day is presented.
Abstract: This paper - first published on-line in November 2008 - draws on data from an early version of the Google Flu Trends search engine to estimate the levels of flu in a population. It introduces a computational model that converts raw search query data into a region-by-region real-time surveillance system that accurately estimates influenza activity with a lag of about one day - one to two weeks faster than the conventional reports published by the Centers for Disease Prevention and Control. This report introduces a computational model based on internet search queries for real-time surveillance of influenza-like illness (ILI), which reproduces the patterns observed in ILI data from the Centers for Disease Control and Prevention. Seasonal influenza epidemics are a major public health concern, causing tens of millions of respiratory illnesses and 250,000 to 500,000 deaths worldwide each year1. In addition to seasonal influenza, a new strain of influenza virus against which no previous immunity exists and that demonstrates human-to-human transmission could result in a pandemic with millions of fatalities2. Early detection of disease activity, when followed by a rapid response, can reduce the impact of both seasonal and pandemic influenza3,4. One way to improve early detection is to monitor health-seeking behaviour in the form of queries to online search engines, which are submitted by millions of users around the world each day. Here we present a method of analysing large numbers of Google search queries to track influenza-like illness in a population. Because the relative frequency of certain queries is highly correlated with the percentage of physician visits in which a patient presents with influenza-like symptoms, we can accurately estimate the current level of weekly influenza activity in each region of the United States, with a reporting lag of about one day. This approach may make it possible to use search queries to detect influenza epidemics in areas with a large population of web search users.

3,984 citations