Spaces:
Paused
Paused
The moisture trapped in the soil affects a lot more than the health of | |
crops and trees. Look at natural--color satellite images and it becomes | |
clear that most of the water on Earth (about 97 percent) is stored in | |
the oceans. Next you might notice some on the land: liquid water fills | |
lakes and rivers, while frozen water blankets the poles and | |
mountaintops. In the atmosphere, water is visible in the countless tiny | |
droplets that compose the clouds, though there is plenty of moisture | |
even in cloud-free skies. | |
Soil moisture has many expressions and influences in Earth's climate, | |
from evaporation to freezing and thawing ice to droughts and floods. | |
(Photos used under a Creative Commons license, courtesy of Guido | |
Appenzeller (top left), TREEAID (bottom left), and Mike Rosenberg (top | |
right). NASA Earth Observatory image (bottom right) by Joshua Stevens, | |
using Landsat data from the U.S. Geological Survey.) | |
Not immediately visible, however, is the water residing in the soil. | |
This water does not appear brilliantly blue or white, like the oceans or | |
ice. In fact, it is hard to spot in natural-color satellite images. | |
Compared to the amount of water stored elsewhere on the planet, the | |
amount in the soil is minuscule. But that small volume has great | |
significance. It can affect when, where, and what a farmer will plant. | |
It can influence the weather. And at high northern latitudes, soil | |
moisture has serious implications for global climate. | |
For all of these reasons, researchers have developed satellite | |
instruments to measure the water hidden between soil particles. The | |
instruments are either active or passive. | |
Active radar sensors transmit microwave radiation toward the ground and | |
measure the echoes. Depending on the moisture content, the reflected | |
signal will "look" different---information scientists then use to derive | |
the amount of soil moisture. (One such sensor flies on the Metop | |
satellites operated by the European Organization for the Exploitation of | |
Meteorological Satellites.) The radar approach allows scientists to | |
measure moisture in very specific areas (high resolution), but it is | |
less accurate than other approaches. | |
"Passive radiometers detect microwave wavelengths of light that are | |
naturally emitted by the soil. Because the signal varies with wetness, | |
scientists can use the information to estimate the amount of moisture in | |
the top few inches of soil. These measurements give better estimates of | |
the amount of water, but over a broader area (coarse spatial resolution) | |
than active radar. The European Space Agency has flown a passive | |
radiometer on the Soil Moisture and Ocean Salinity (SMOS) mission since | |
2009, and NASA put them on the Aqua (2002) and Aquarius (2011) | |
satellites. NASA's Soil Moisture Active Passive (SMAP) satellite was | |
launched in January 2015 and carries both a radiometer and radar. | |
However, the radar stopped transmitting data a few months after launch. | |
The Soil Moisture Active Passive (SMAP) satellite can observe global | |
soil moisture daily at a 36-kilometer resolution. (NASA Earth | |
Observatory map by Joshua Stevens, using data courtesy of JPL and the | |
SMAP science team.) | |
All of these platforms, combined with insights from ground-based | |
sensors, contribute to a growing record of global soil moisture. The | |
goal is to establish a standardized set of measurements for the entire | |
planet so that everyone from meteorologists to climate modelers can | |
track the movement of this small but vital reservoir of water. | |
The most obvious users of soil moisture data are farmers and ranchers. | |
There's more to it, however, than the simple fact that plants need water | |
to grow. Knowing something about the moisture in the soil is important | |
before, during, and after the growing season. For example, will mud | |
prevent a tractor from safely driving across the fields? How much water | |
will fruits, nuts, and vegetables have available at each stage of | |
growth, from germination through harvest? What is the forecast for crop | |
yields around the world? How will the amount of moisture and | |
agricultural output affect trade policy and food aid? | |
Ground-based sensors can monitor soil moisture over small areas, | |
typically less than one square meter. To find out what is happening over | |
larger areas, researchers in several U.S. states have patched together a | |
network of sensors. In Oklahoma, for example, a network to monitor | |
weather and climate parameters (including soil moisture) was conceived | |
after a disastrous flood struck Tulsa in 1984. Interest in this type of | |
network for agricultural purposes also arose in Stillwater. The result | |
was an environmental monitoring network called Mesonet, which in 1996 | |
started to include soil moisture sensors. | |
Today, more than 100 stations across Oklahoma are making measurements at | |
various soil depths down to 60 centimeters (24 inches). Sensors record | |
the temperature during and after a imparting a pulse of heat; the amount | |
of water in the soil can then be inferred from the temperature change. | |
(Other ground-based methods involve neutron scattering or soil coring.) | |
Data from these sensors, updated every 30 minutes, can help farmers | |
quickly figure out where there is inadequate moisture in their fields. | |
Mesonet is just one of 31 networks and 1,479 stations in North America. | |
But, in situ networks do not cover all areas of the United States, and | |
certainly not the planet. | |
Nearly 1,500 stations track soil moisture in locations across the United | |
States. As an example, the plot shows data from the Coastal Sage UCI | |
station (California). As the map shows, the network is relatively sparse | |
for the size of the nation. (NASA Earth Observatory map and chart by | |
Joshua Stevens, using data from the TAMU North American Soil Moisture | |
Database.) | |
To fill in the gaps, some scientists estimate global soil moisture by | |
running computer models loaded with precipitation, temperature, and | |
humidity data. Gathering all of the data to run such models can take as | |
long as two to three months, which makes real-time applications | |
impossible. | |
"What we really want is soil moisture information that can be used to | |
understand how plants are growing and what's going on in the atmosphere | |
right now," said Susan Moran, a hydrologist with USDA's Agricultural | |
Research Service and chair of the SMAP Applications Working Group. "We | |
have to get soil moisture information to the agriculture community, and | |
the only way to do that is from satellites." | |
With the recent satellite missions, Moran and colleagues have been | |
learning more about how soil moisture affects plant growth and | |
agricultural productivity, especially during conditions of water | |
shortage and drought. For instance, she notes there have been drier and | |
longer droughts than the one currently parching the western U.S., but | |
none that have been so hot. The combination of heat and the lack of | |
water is driving soil moisture to unprecedented deficits. | |
"Data from SMAP will make a giant difference for my work," she said. "We | |
have already looked at five years of data from SMOS; add SMAP onto that | |
and we begin to get a good time series of global soil moisture to help | |
us figure out where vegetation has a high risk of mortality." | |
Soil moisture has an obvious, visible effect on the landscape. The | |
high-profile examples are droughts and floods. But the water in the soil | |
has a more subtle, yet equally important role in day-to-day weather. | |
Soil moisture forms a vast, thin, and mostly out-of-sight reservoir of | |
water that accumulates in the root zone of plants. The water is released | |
to the atmosphere through evaporation and plant transpiration. Averaged | |
globally, this evapotranspiration contributes to more than 60 percent of | |
the precipitation that falls over land each year. | |
Today, satellites can measure soil moisture globally and quickly. | |
Saturated soils in the map above---measured by SMAP on October 5, | |
2015---were the result of intense rains that caused flooding in the | |
southeastern United States. (NASA Earth Observatory map by Joshua | |
Stevens, using soil moisture data courtesy of JPL and the SMAP science | |
team.) | |
"The first time we were struck by the importance of soil moisture for | |
weather forecasts was in July 1993," said Patricia de Rosnay, a | |
researcher at the European Centre for Medium-Range Weather Forecasts | |
(ECMWF). During the first six months of that year, extreme amounts of | |
rain and snow fell on the central United States. Yet the existing | |
weather models were not accounting for the storage and evaporation of | |
all of that water. They could not see how the water on the land was | |
feeding back into the weather patterns to make the deluge more extreme. | |
By July 1993, the Upper Mississippi River faced its worst flooding on | |
record. | |
July 1993 also happened to be the same month that ECMWF scientists began | |
testing a new weather forecast model. Their model accounted for soil | |
moisture in the root zone, allowing researchers to see how the soil | |
sustained a high level of evaporation and fed the extreme rainfall | |
event. The new model had produced a closer representation of reality. | |
The strength of the connection between soil moisture and the weather is | |
not the same everywhere. According to NASA scientist Randy Koster, there | |
are hot spots---about 10 percent of Earth's surface where the amount of | |
soil moisture plays a more critical role in the weather. | |
Water that evaporates from Earth's surface is linked to the formation of | |
clouds and rainfall. In dry areas, variations in the amount of | |
evaporation are too small to have much of an effect on the atmosphere. | |
In humid regions, particularly the tropics, changes in soil moisture do | |
not matter much for evaporation because it is limited by the amount of | |
water that the atmosphere can hold. | |
Landsat 8 acquired this image N'Djamena, Chad, on October 20, 2015. The | |
city sits along the Logone River and within the African Sahel. As | |
pictured here after the rainy season, the river's saturated banks are | |
surrounded by a dry, sandy landscape. (NASA Earth Observatory image by | |
Joshua Stevens, using Landsat data from the U.S. Geological Survey.) | |
The soil moisture hot spots are areas that are neither too dry nor too | |
wet. They are located in the transition zones between dry and wet | |
areas---places that have suitably high evaporation that is more | |
dependent on moisture on the ground than in the atmosphere. The | |
Midwestern United States is one of those hot spots. So, too, are | |
northern India and the African Sahel. | |
Better estimates of soil moisture in weather models will not necessarily | |
make for perfect long-range forecasts. Randomness in the variables that | |
cause weather will always hinder the accuracy beyond a few days. But | |
with better information on the thin reservoir in the soil, forecasters | |
can tip the scales further in favor of getting weather prediction right. | |
In the planet's highest northern latitudes, even the water in the soil | |
is locked away as ice, making it mostly inaccessible to plants. But just | |
a short distance to the south, in the boreal areas of Alaska, Canada, | |
Siberia, and Scandinavia, the landscape comes alive each year after the | |
spring thaw. | |
The transition is relatively rapid, occurring over just a few weeks, and | |
coincides with increasing sunlight and spring snowmelt. Rapid warming | |
releases liquid water. As liquid water becomes more readily available, | |
plant and animal activity are energized. The land greens up, and animals | |
return to graze. | |
"I'm always impressed by how rapidly northern landscapes transition from | |
frozen and dormant conditions in the winter to a rapid burst of life and | |
activity in the spring," said John Kimball, a scientist at the | |
University of Montana. | |
The transition between frozen and thawed land is something researchers | |
have observed for more than 30 years with satellites. The Nimbus-7 | |
Pathfinder, the Defense Meteorological Satellite Program (DMSP) | |
satellites, and Aqua all have carried passive microwave radiometers. | |
They detect the microwave energy coming from the Earth's surface, which | |
will have different characteristics depending on whether the soil is | |
frozen or thawed. When surface water and soil moisture is locked away as | |
ice, this frozen landscape looks like a desert to a microwave sensor. | |
Thawed landscapes look comparatively wet, so this large contrast is the | |
basis for something called a freeze-thaw measurement. | |
Across a year, the ice and frozen lands advance and retreat in the high | |
northern and southern latitudes. (NASA Earth Observatory image by Joshua | |
Stevens, using NASA's Blue Marble data.) | |
Kimball and colleagues have mined 30 years of freeze-thaw patterns from | |
the satellite record. In a paper published in 2012, the researchers | |
showed that soils in the Northern Hemisphere thawed for as many as 7.5 | |
days more in 2008 than they did in 1979. The change was primarily driven | |
by an earlier start to the spring thaw and coincided with measureable | |
warming in the region. | |
"This was a real eye-opener to me," Kimball said. "We found that the | |
earlier spring-thaw was driving widespread increases in northern growing | |
seasons." The start and the length of the growing season have | |
implications for how much carbon is exchanged between the land and | |
atmosphere. | |
Each year, half of all global carbon emissions are removed from the | |
atmosphere by natural processes on the surface. It is sequestered | |
somewhere on land, and a large amount of that carbon is stored at high | |
latitudes. According to scientists at the Woods Hole Research Center, | |
the boreal region covers about 15 percent of the global land surface, | |
yet holds more than 30 percent of all carbon contained on land. | |
A longer growing season in the north could make vegetation a more | |
important "sink," removing carbon dioxide from the atmosphere and | |
storing it in forest biomass, dead organic matter, and the soil. But | |
those boreal lands also could become a carbon source though burning, | |
decay, and thawing soil. Currently, the region is thought by many to be | |
a net sink, absorbing more carbon than it releases. But how might | |
thawing soils affect that balance? | |
The answer is complicated by the fact that the timing of the thaw can | |
vary dramatically over a small area. Sunlight sweeps over the landscape | |
at a low angle, so areas with even the slightest rolling topography can | |
be cast in either shade or sunlight. South-facing slopes thaw first. And | |
just a few extra weeks of thawing time can have a huge impact on plant | |
growth. | |
Another complication in the carbon equation is permafrost. Even where | |
the top layer of soil has thawed, there is often a long-term frozen | |
layer below. This frozen layer locks up carbon so that it cannot | |
decompose. But as seasonal thaws reach greater soil depths, more organic | |
matter can decompose and get flushed out into the atmosphere by streams | |
or degassing into the atmosphere. "There is debate as to how stable that | |
soil will be with continued global warming," Kimball says. | |
But progress is being made. Freeze-thaw monitoring, according to | |
Kimball, has made a major advance thanks to the development of | |
well-calibrated, long-term satellite soil moisture records. As those | |
observations continue, and as they encompass more of the planet, it | |
stands to reason that our understanding of the entire water cycle will | |
improve. | |
Soil moisture has an obvious, visible effect on the landscape. The | |
high-profile examples are droughts and floods. But the water in the soil | |
has a more subtle, yet equally important role in day-to-day weather. | |
Soil moisture forms a vast, thin, and mostly out-of-sight reservoir of | |
water that accumulates in the root zone of plants. The water is released | |
to the atmosphere through evaporation and plant transpiration. Averaged | |
globally, this evapotranspiration contributes to more than 60 percent of | |
the precipitation that falls over land each year. | |
"The first time we were struck by the importance of soil moisture for | |
weather forecasts was in July 1993," said Patricia de Rosnay, a | |
researcher at the European Centre for Medium-Range Weather Forecasts | |
(ECMWF). During the first six months of that year, extreme amounts of | |
rain and snow fell on the central United States. Yet the existing | |
weather models were not accounting for the storage and evaporation of | |
all of that water. They could not see how the water on the land was | |
feeding back into the weather patterns to make the deluge more extreme. | |
By July 1993, the Upper Mississippi River faced its worst flooding on | |
record. | |
July 1993 also happened to be the same month that ECMWF scientists began | |
testing a new weather forecast model. Their model accounted for soil | |
moisture in the root zone, allowing researchers to see how the soil | |
sustained a high level of evaporation and fed the extreme rainfall | |
event. The new model had produced a closer representation of reality. | |
The strength of the connection between soil moisture and the weather is | |
not the same everywhere. According to NASA scientist Randy Koster, there | |
are hot spots---about 10 percent of Earth's surface where the amount of | |
soil moisture plays a more critical role in the weather. | |
Water that evaporates from Earth's surface is linked to the formation of | |
clouds and rainfall. In dry areas, variations in the amount of | |
evaporation are too small to have much of an effect on the atmosphere. | |
In humid regions, particularly the tropics, changes in soil moisture do | |
not matter much for evaporation because it is limited by the amount of | |
water that the atmosphere can hold. | |
The soil moisture hot spots are areas that are neither too dry nor too | |
wet. They are located in the transition zones between dry and wet | |
areas---places that have suitably high evaporation that is more | |
dependent on moisture on the ground than in the atmosphere. The | |
Midwestern United States is one of those hot spots. So, too, are | |
northern India and the African Sahel. | |
Better estimates of soil moisture in weather models will not necessarily | |
make for perfect long-range forecasts. Randomness in the variables that | |
cause weather will always hinder the accuracy beyond a few days. But | |
with better information on the thin reservoir in the soil, forecasters | |
can tip the scales further in favor of getting weather prediction right. | |
Remotely sensed biomass carbon density maps are widely used for myriad | |
scientific and policy applications, but all remain limited in scope. | |
They often only represent a single vegetation | |
type and rarely account for carbon stocks in belowground biomass. To | |
date, no global product integrates these disparate estimates into an | |
all-encompassing map at a scale appropriate | |
for many modelling or decision-making applications. We developed an | |
approach for harmonizing vegetation-specific maps of both above and | |
belowground biomass into a single, comprehensive | |
representation of each. We overlaid input maps and allocated their | |
estimates in proportion to the relative spatial extent of each | |
vegetation type using ancillary maps of percent tree | |
cover and landcover, and a rule-based decision schema. The resulting | |
maps consistently and seamlessly report biomass carbon density estimates | |
across a wide range of vegetation types | |
in 2010 with quantified uncertainty. They do so for the globe at an | |
unprecedented 300-meter spatial resolution and can be used to more | |
holistically account for diverse vegetation carbon | |
stocks in global analyses and greenhouse gas inventories. Background & | |
Summary Terrestrial ecosystems store vast quantities of carbon (C) in | |
aboveground and belowground biomass1. At | |
any point in time, these stocks represent a dynamic balance between the | |
C gains of growth and C losses from death, decay and combustion. Maps of | |
biomass are routinely used for benchmarking | |
biophysical models2,3,4, estimating C cycle effects of disturbance5,6,7, | |
and assessing biogeographical patterns and ecosystem services8,9,10,11. | |
They are also critical for assessing | |
climate change drivers, impacts, and solutions, and factor prominently | |
in policies like Reducing Emissions from Deforestation and Forest | |
Degradation (REDD+) and C offset schemes12,13,14. | |
Numerous methods have been used to map biomass C stocks but their | |
derivatives often remain limited in either scope or extent12,15. There | |
thus remains a critical need for a globally | |
harmonized, integrative map that comprehensively reports biomass C | |
across a wide range of vegetation types. Most existing maps of | |
aboveground biomass (AGB) and the carbon it contains | |
(AGBC) are produced from statistical or data-driven methods relating | |
field-measured or field-estimated biomass densities and spaceborne | |
optical and/or radar imagery12,15,16. They largely | |
focus on the AGB of trees, particularly those in tropical landscapes | |
where forests store the majority of the region's biotic C in aboveground | |
plant matter. Land cover maps are often | |
used to isolate forests from other landcover types where the predictive | |
model may not be appropriate such that forest AGB maps intentionally | |
omit AGB stocks in non-forest vegetation | |
like shrublands, grasslands, and croplands, as well as the AGB of trees | |
located within the mapped extent of these excluded landcovers17. | |
Non-forest AGB has also been mapped to some | |
extent using similar approaches but these maps are also routinely masked | |
to the geographic extent of their focal landcover18,19,20,21. To date, | |
there has been no rigorous attempt to | |
harmonize and integrate these landcover-specific, remotely sensed | |
products into a single comprehensive and temporally consistent map of C | |
in all living biomass. Maps of belowground | |
biomass (BGB) and carbon density (BGBC) are far less common than those | |
of AGB because BGB cannot be readily observed from space or airborne | |
sensors. Consequently, BGB is often inferred | |
from taxa-, region-, and/or climate-specific "root-to-shoot" ratios that | |
relate the quantity of BGB to that of AGB22,23,24. These ratios can be | |
used to map BGB by spatially applying | |
them to AGB estimates using maps of their respective strata5. In recent | |
years, more sophisticated regression-based methods have been developed | |
to predict root-to-shoot ratios of some | |
landcover types based on covariance with other biophysical and/or | |
ecological factors25,26. When applied spatially, these methods can allow | |
for more continuous estimates of local BGB5,27. | |
Like AGBC, though, few attempts have been made to comprehensively map | |
BGBC for the globe. Despite the myriad of emerging mapping methods and | |
products, to date, the Intergovernmental | |
Panel on Climate Change (IPCC) Tier-1 maps by Ruesch and Gibbs28 remains | |
the primary source of global AGBC and BGBC estimates that transcend | |
individual landcover types. These maps, | |
which represents the year 2000, were produced prior to the relatively | |
recent explosion of satellite-based AGB maps and they therefore rely on | |
an alternative mapping technique called | |
"stratify and multiply"15, which assigns landcover-specific biomass | |
estimates or "defaults" (often derived from field measurements or | |
literature reviews) to the corresponding classified | |
grid cells of a chosen landcover map12. While this approach yields a | |
comprehensive wall-to-wall product, it can fail to capture finer-scale | |
spatial patterns often evident in the field | |
and in many satellite-based products12,15. The accuracy of these maps is | |
also tightly coupled to the quality and availability of field | |
measurements29 and the thematic accuracy and | |
discontinuity of the chosen landcover map. Given the wealth of | |
landcover-specific satellite based AGB maps, a new harmonization method | |
akin to "stratify and multiply" is needed to | |
merge the validated spatial detail of landcover-specific remotely sensed | |
biomass maps into a single, globally harmonized product. We developed | |
such an approach by which we (i) overlay | |
distinct satellite-based biomass maps and (ii) proportionately allocate | |
their estimates to each grid cell ("overlay and allocate"). | |
Specifically, we overlay continental-to-global scale | |
remotely sensed maps of landcover-specific biomass C density and then | |
allocate fractional contributions of each to a given grid cell using | |
additional maps of percent tree cover, thematic | |
landcover and a rule-based decision tree. We implement the new approach | |
here using temporally consistent maps of AGBC as well as matching | |
derived maps of BGBC to generate separate | |
harmonized maps of AGBC and BGBC densities. In addition, we generate | |
associated uncertainty layers by propagating the prediction error of | |
each input dataset. The resulting global maps | |
consistently represent biomass C and associated uncertainty across a | |
broad range of vegetation in the year 2010 at an unprecedented 300 meter | |
(m) spatial resolution. Our harmonization | |
approach (Fig. 1) relies on independent, landcover-specific biomass maps | |
and ancillary layers, which we compiled from the published literature | |
(Table 1). When published maps did not | |
represent our epoch of interest (i.e. grasslands and croplands) or did | |
not completely cover the necessary spatial extent (i.e. tundra | |
vegetation), we used the predictive model reported | |
with the respective map to generate an updated version that met our | |
spatial and temporal requirements. We then used landcover specific | |
root-to-shoot relationships to generate matching | |
BGBC maps for each of the input AGBC maps before implementing the | |
harmonization procedure. Below we describe, in detail, the methodologies | |
used for mapping AGBC and BGBC of each landcover | |
type and the procedure used to integrate them. Woody tree biomass Since | |
the first remotely sensed woody AGB maps were published in the early | |
1990s, the number of available products | |
has grown at an extraordinary pace16 and it can thus be challenging to | |
determine which product is best suited for a given application. For our | |
purposes, we relied on the GlobBiomass | |
AGB density map30 as our primary source of woody AGB estimates due to | |
its precision, timestamp, spatial resolution, and error quantification. | |
It was produced using a combination of | |
spaceborne optical and synthetic aperture radar (SAR) imagery and | |
represents the year 2010 at a 100 m spatial resolution -- making it the | |
most contemporary global woody AGB currently | |
available and the only such map available for that year. Moreover, | |
GlobBiomass aims to minimize prediction uncertainty to less than 30% and | |
a recent study suggests that it has high | |
fidelity for fine-scale applications31. The GlobBiomass product was | |
produced by first mapping the growing stock volume (GSV; i.e. stem | |
volume) of living trees, defined following Food | |
and Agriculture Organization (FAO) guidelines32 as those having a | |
diameter at breast height (DBH) greater than 10 centimeters (cm). AGB | |
density was then determined from GSV by applying | |
spatialized biomass expansion factors (BEFs) and wood density estimates. | |
These factors were mapped using machine learning methods trained from a | |
suite of plant morphological databases | |
that compile thousands of field measurements from around the globe33. | |
The resulting AGB estimates represent biomass in the living structures | |
(stems, branches, bark, twigs) of trees | |
with a DBH greater than 10 cm. This definition may thereby overlook AGB | |
of smaller trees and/or shrubs common to many global regions. Unlike | |
other maps, though, the GlobBiomass product | |
employs a subpixel masking procedure that retains AGB estimates in 100 m | |
grid cells in which any amount of tree cover was detected in finer | |
resolution (30 m) imagery34. This unique | |
procedure retains AGB estimates in tree-sparse regions like savannahs, | |
grasslands, croplands, and agroforestry systems where AGB is often | |
overlooked17, as well as in forest plantations. | |
The GlobBiomass product is the only global map that also includes a | |
dedicated uncertainty layer reporting the standard error of prediction. | |
We used this layer to propagate uncertainty | |
when converting AGB to AGBC density, modelling BGBC, and integrating | |
with C density estimates of other vegetation types. Bouvet et al.35 -- | |
some of whom were also participants of the | |
GlobBiomass project -- independently produced a separate AGB density map | |
for African savannahs, shrublands and dry woodlands circa 2010 at 25 m | |
spatial resolution35 (hereafter "Bouvet | |
map"), which we included in our harmonized product to begin to address | |
the GlobBiomass map's potential omission of small trees and shrubs that | |
do not meet the FAO definition of woody | |
AGB. This continental map of Africa is based on a predictive model that | |
directly relates spaceborne L-band SAR imagery -- an indirect measure of | |
vegetation structure that is sensitive | |
to low biomass densities36 -- with region-specific, field-measured AGB. | |
Field measurements (n = 144 sites) were compiled from 7 different | |
sampling campaigns -- each specifically seeking | |
training data for biomass remote sensing -- that encompassed 8 different | |
countries35. The resulting map is not constrained by the FAO tree | |
definition and is masked to exclude grid cells | |
in which predicted AGB exceeds 85 megagrams dry mater per hectare (Mg | |
ha−1) -- the threshold at which the SAR-biomass relationship saturates. | |
To avoid extraneous prediction, it further | |
excludes areas identified as "broadleaved evergreen closed-to-open | |
forest", "flooded forests", "urban areas" and "water bodies" by the | |
European Space Agency's Climate Change Initiative | |
(CCI) Landcover 2010 map37 and as "bare areas" in the Global Land Cover | |
(GLC) 2000 map38. While the Bouvet map is not natively accompanied | |
by an uncertainty layer, its authors provided | |
us with an analytic expression of its uncertainty (SE; standard error of | |
prediction) as a function of estimated AGB (Eq. 1) which we used to | |
generate an uncertainty layer for subsequent | |
error propagation. We combined the GlobBiomass and Bouvet products to | |
generate a single woody biomass map by first upscaling each map | |
separately to a matching 300 m spatial resolution | |
using an area-weighted average to aggregate grid cells, and then | |
assigning the Bouvet estimate to all overlapping grid cells, except | |
those identified by the CCI Landcover 2010 map | |
as closed or flooded forest types (Online-only Table 1) which were not | |
within the dryland domain of the Bouvet map. While more complex | |
harmonization procedures based on various averaging | |
techniques have been used by others39,40, their fidelity remains unclear | |
since they fail to explicitly identify and reconcile the underlying | |
source of the inputs' discrepancies41. | |
We thus opted to use a more transparent ruled-based approach when | |
combining these two woody biomass maps, which allows users to easily | |
identify the source of a grid cell's woody biomass | |
estimate. Given the local specificity of the training data used to | |
produce the Bouvet map, we chose to prioritize its predictions over | |
those of the GlobBiomass product when within | |
its domain. In areas of overlap, the Bouvet map values tend to be lower | |
in moist regions and higher in dryer regions (Fig. 2), though, where | |
used, these differences rarely exceed ±25 | |
megagrams C per hectare (MgC ha−1). We then converted all woody AGB | |
estimates to AGBC by mapping climate and phylogeny-specific biomass C | |
concentrations from Martin et al.42. Climate | |
zones were delineated by aggregating classes of the Köppen-Gieger | |
classification43 (Table 2) to match those of Martin et al.42. | |
Phylogenetic classes (angiosperm, gymnosperm and mixed/ambiguous) | |
were subsequently delineated within each of these zones using aggregated | |
classes of the CCI Landcover 2010 map (Online-only Table 1). Martin et | |
al.42 only report values for angiosperms | |
and gymnosperms so grid cells with a mixed or ambiguous phylogeny were | |
assigned the average of the angiosperm and gymnosperm values and the | |
standard error of this value was calculated | |
from their pooled variance. Due to residual classification error in the | |
aggregated phylogenetic classes, we weighted the phylogeny-specific C | |
concentration within each climate zone | |
by the binary probability of correctly mapping that phylogeny where, | |
within each climate zone, μc is the mean probability-weighted C | |
concentration of the most probable phylogeny, μm | |
is the mean C concentration of that phylogeny from Martin et al.42, pm | |
is the user's accuracy of that phylogeny's classification (Table 3), and | |
μn and μo are the mean C concentrations | |
of the remain phylogenetic classes from Martin et al.42. Standard error | |
estimates for these C concentrations were similarly weighted using | |
summation in quadrature where is the probability-weighted | |
standard error of the most probable phylogeny's C concentration and and | |
are the standard errors of the respective phylogeny-specific C | |
concentrations from Martin et al.42. Probability-weighted | |
C concentrations used are reported in Table. Mapped, | |
probability-weighted C estimates were then arithmetically applied to AGB | |
estimates. Uncertainty associated with this correction | |
was propagated using summation in quadrature of the general form (Eq. 4) | |
is the uncertainty of μf, and , are the respective uncertainty estimates | |
of the dependent parameters (standard | |
error unless otherwise noted). Here, μf, is the estimated AGBC of a | |
given grid cell, and is the product of its woody AGB estimate, and its | |
corresponding C concentration. Tundra vegetation | |
biomass The tundra and portions of the boreal biome are characterized by | |
sparse trees and dwarf woody shrubs as well as herbaceous cover that are | |
not included in the GlobBiomass definition | |
of biomass. AGB density of these classes has been collectively mapped by | |
Berner et al.18,45 for the North Slope of Alaska from annual Landsat | |
imagery composites of the normalized difference | |
vegetation index (NDVI) and a non-linear regression-based model trained | |
from field measurements of peak AGB that were collected from the | |
published literature (n = 28 sites). Berner | |
et al.18 note that while these field measurements did not constitute a | |
random or systematic sample, they did encompass a broad range of tundra | |
plant communities. In the absence of | |
a global map and due the sparsity of high quality Landsat imagery at | |
high latitudes, we extended this model to the pan-Arctic and | |
circumboreal regions using NDVI composites created | |
from daily 250 m MODIS Aqua and Terra surface reflectance images46,47 | |
that were cloud masked and numerically calibrated to Landsat ETM | |
reflectance -- upon which the tundra model is | |
based -- using globally derived conversion coefficients48. We generated | |
six separate 80th percentile NDVI composites circa 2010 -- one for each | |
of the MODIS missions (Aqua and Terra) | |
in 2009, 2010 and 2011 -- following Berner et al.18. We chose to use | |
three years of imagery (circa 2010) rather than just one (2010) to | |
account for the potential influence that cloud | |
masking may exert upon estimates of the 80th NDVI percentile in a single | |
year. We then applied the tundra AGB model to each composite, converted | |
AGB estimates to AGBC by assuming a | |
biomass C fraction of 49.2% (SE = 0.8%)42 and generated error layers for | |
each composite from the reported errors of the AGB regression | |
coefficients and the biomass C conversion factor | |
using summation in quadrature as generally described above (Eq. 4). A | |
single composite of tundra AGBC circa 2010 was then created as the | |
pixelwise mean of all six composites. We also | |
generated a complementary uncertainty layer representing the cumulative | |
standard error of prediction, calculated as the pixelwise root mean of | |
the squared error images in accordance | |
with summation in quadrature. Both maps were upscaled from their native | |
250 m spatial resolution to a 300 m spatial resolution using an area | |
weighted aggregation procedure, whereby | |
pixels of the 300 m biomass layer was calculated as the area weighted | |
average of contained 250 m grid cells, and the uncertainty layer was | |
calculated -- using summation in quadrature | |
-- as the root area-weighted average of the contained grid cells | |
squared. Grassland biomass Grassland AGBC density was modelled directly | |
from maximum annual NDVI composites using a | |
non-linear regression-based model developed by Xia et al.19 for mapping | |
at the global scale. This model was trained by relating maximum annual | |
NDVI as measured by the spaceborne Advanced | |
Very High-Resolution Radiometer (AVHRR) sensor to globally distributed | |
field measurements of grassland AGBC that were compiled from the | |
published literature (81 sites for a total of | |
158 site-years). Like the tundra biomass training data, these samples | |
did not constitute a random or systematic sample but do encompass a | |
comprehensive range of global grassland communities. | |
Given the inevitable co-occurrence of trees in the AVHRR sensor's 8 km | |
resolution pixels upon which the model is trained, it's predictions of | |
grassland AGBC are relatively insensitive | |
to the effects of co-occurring tree cover. We thereby assume that its | |
predictions for grid cells containing partial tree cover represent the | |
expected herbaceous AGBC density in the | |
absence of those trees. Maximum model predicted AGBC (NDVI = 1) is 2.3 | |
MgC ha−1 which is comparable to the upper quartile of herbaceous AGBC | |
estimates from global grasslands49 and | |
suggests that our assumption will not lead to an exaggerated estimation. | |
For partially wooded grid cells, we used modelled grassland AGBC density | |
to represent that associated with | |
the herbaceous fraction of the grid cell in a manner similar to Zomer et | |
al.17 as described below (See "Harmonizing Biomass Carbon Maps"). We | |
applied the grassland AGBC model to all | |
grid cells of maximum annual NDVI composites produced from finer | |
resolution 16-day (250 m) MODIS NDVI imagery composites circa 201050,51. | |
Here again, three years of imagery were used | |
to account for potential idiosyncrasies in a single year's NDVI | |
composites resulting from annual data availability and quality. As with | |
AGB of tundra vegetation, annual composites | |
(2009--2011) were constructed separately from cloud-masked imagery | |
collected by both MODIS missions (Aqua and Terra; n = 6) and then | |
numerically calibrated to AVHRR reflectance using | |
globally derived conversion coefficients specific to areas of herbaceous | |
cover52. We then applied the AGBC model to each of these composites and | |
estimated error for each composite | |
from both the AVHRR calibration (standard deviation approximated from | |
the 95% confidence interval of the calibration scalar) and the AGBC | |
model (relative RMSE) using summation in quadrature. | |
A single map of grassland AGBC circa 2010 was then created as the | |
pixelwise mean of all six composites and an associated error layer was | |
created as the pixelwise root mean of the squared | |
error images. Both maps were aggregated from their original 250 m | |
resolution to 300 m to facilitate harmonization using the area-weighted | |
procedure described previously for woody and | |
tundra vegetation (see section 1.2). Cropland biomass Prior to harvest, | |
cropland biomass can also represent a sizable terrestrial C stock. In | |
annually harvested cropping systems, the | |
maximum standing biomass of these crops can be inferred from annual net | |
primary productivity (ANPP). While spaceborne ANPP products exist, they | |
generally perform poorly in croplands53,54. | |
Instead, cropland ANPP is more commonly derived from crop | |
yields20,21,53. We used globally gridded, crop-specific yields of 70 | |
annually harvested herbaceous commodity crops circa 2000 | |
by Monfreda et al.20 -- the only year in which these data were | |
available. These maps were produced by spatially disaggregating | |
crop-yield statistics for thousands of globally distributed | |
administrative units throughout the full extent of a satellite-based | |
cropland map20. These maps were combined with crop-specific parameters | |
(Online-only Table 2) to globally map AGBC | |
as aboveground ANPP for each crop following the method of Wolf et al.21. | |
This method can be simplified as (Eq. 5) where y is the crop's yield (Mg | |
ha−1), ω is the dry matter fraction | |
of its harvested biomass, h is its harvest index (fraction of total AGB | |
collected at harvest) and c is the carbon content fraction of its | |
harvested dry mass. This simplification assumes, | |
following Wolf et al.21, that 2.5% of all harvested biomass is lost | |
between the field and farmgate and that unharvested residue and root | |
mass is 44% C. Total cropland AGBC density | |
was then calculated as the harvested-area-weighted average of all | |
crop-specific AGBC estimates within a given grid cell. Since multiple | |
harvests in a single year can confound inference | |
of maximum AGBC from ANPP, we further determined the harvest frequency | |
(f) of each grid cell by dividing a cell's total harvested area (sum of | |
the harvested area of each crop reported | |
within a given grid cell) by its absolute cropland extent as reported in | |
a complementary map by Ramankutty et al.55. If f was greater than one, | |
multiple harvests were assumed to have | |
occurred and AGBC was divided by f to ensure that AGBC estimates did not | |
exceed the maximum standing biomass density. Since the yields of many | |
crops and, by association, their biomass | |
have changed considerably since 200056,57, we calibrated our circa 2000 | |
AGBC estimates to the year 2010 using local rates of annual ANPP change | |
(MgC ha−1 yr−1) derived as the Theil-Sen | |
slope estimator -- a non-parametric estimator that is relatively | |
insensitive to outliers -- of the full MODIS Terra ANPP timeseries | |
(2000--2015)58. Total ANPP change between 2000 and | |
2010 for each grid cell was calculated as ten times this annual rate of | |
change. Since MODIS ANPP represents C gains in both AGB and BGB, we | |
proportionately allocated aboveground ANPP | |
to AGBC using the total root-to-shoot ratio derived from the circa 2000 | |
total crop AGBC and BGBC maps (described below). Since error estimates | |
were not available for the yield maps | |
or the crop-specific parameters used to generate the circa 2000 AGBC | |
map, estimated error of the circa 2010 crop AGBC map was exclusively | |
based on that of the 2000--2010 correction. | |
The error of this correction was calculated as the pixel-wise standard | |
deviation of bootstrapped simulations (n = 1000) in which a random | |
subset of years was omitted from the slope | |
estimator in each iteration. The 8 km resolution circa 2000 AGBC map and | |
error layer were resampled to 1 km to match the resolution of MODIS ANPP | |
using the bilinear method prior to | |
ANPP correction and then further resampled to 300 m to facilitate | |
harmonization. Woody crops like fruit, nut, and palm oil plantations | |
were not captured using the procedure just described | |
and their biomass was instead assumed to be captured by the previously | |
described woody biomass products which retained biomass estimates in all | |
pixels where any amount of tree cover | |
was detected at the sub-pixel level (see section 1.1). Belowground | |
biomass carbon maps Matching maps of BGBC and associated uncertainty | |
were subsequently produced for each of the landcover-specific | |
AGBC maps using published empirical relationships. With the exception of | |
savannah and shrubland areas, woody BGBC was modelled from AGBC using a | |
multiple regression model by Reich | |
et al.25 that considers the phylogeny, mean annual temperature (MAT), | |
and regenerative origin of each wooded grid cell and that was applied | |
spatially using maps of each covariate in | |
a fashion similar to other studies5,27. Tree phylogeny (angiosperm or | |
gymnosperm) was determined from aggregated classes of the CCI Landcover | |
2010 map37 (Online-only Table 1) with | |
phylogenetically mixed or ambiguous classes assumed to be composed of | |
50% of each. MAT was taken from version 2 of the WorldClim bioclimatic | |
variables dataset (1970--2000) at 1 km resolution59 | |
and resampled to 300 m using the bilinear method. Since there is not a | |
single global data product mapping forest management, we determined tree | |
origin -- whether naturally propagated | |
or planted -- by combining multiple data sources. These data included | |
(i) a global map of "Intact Forest Landscapes" (IFL) in the year 201360 | |
(a conservative proxy of primary, naturally | |
regenerating forests defined as large contiguous areas with minimal | |
human impact), (ii) a Spatial Database of Planted Trees (SDPT) with | |
partial global coverage61, (iii) national statistics | |
reported by the FAO Global Forest Resources Assessment (FRA) on the | |
extent of both naturally regenerating and planted forests and woodlands | |
within each country in the year 201062, | |
total area of natural and planted trees was equal to the corresponding | |
FRA estimates. If the FAOSTAT-reported area of tree crops exceeded | |
FRA-reported planted area, the difference | |
was added to FRA planted total. All areas mapped as IFL were assumed to | |
be of natural origin and BGB was modelled as such. Likewise, besides the | |
exceptions noted below, all tree plantations | |
mapped by the SDPT were assumed to be of planted origin. In countries | |
where the extent of the IFL or SDPT maps fell short of the FRA/FAOSTAT | |
reported areas of natural or planted forests, | |
respectively, we estimated BGBC in the remaining, unknown-origin forest | |
grid cells of that country (BGBCu), as the probability-weighted average | |
of the planted and natural origin estimates | |
using Eq. 6 and are the respective BGBC estimates for a grid cell | |
assuming entirely planted and natural origin, respectively, and and are | |
the respective differences between (i) the | |
FRA/FAOSTAT and (ii) mapped extent of planted and natural forest within | |
the given grid cell's country. While the mapped extent of IFL forests | |
within a given country never exceeded | |
that country's FRA reported natural forest extent, there were infrequent | |
cases (n = 22 of 257) in which the mapped extent of tree plantations | |
exceeded the corresponding FRA/FAOSTAT | |
estimate of planted forest area. In these cases, we down-weighted the | |
BGB estimates of SDPT forests in a similar fashion such that the weight | |
of their planted estimate ( ) was equal | |
to the quotient of (i) the FRA/FAOSTAT planted area and (ii) the SDPT | |
extent within the country, and the weight of the natural origin estimate | |
applied to the SDPT extent ( ) was equal | |
to A BGBC error layer was then produced using summation in quadrature | |
from the standard error estimates of the model coefficients, the AGBC | |
error layer, the relative RMSE of MAT (27%), | |
and the derived global uncertainty of the phylogeny layer. Phylogeny | |
error was calculated as the Bernoulli standard deviation (δ) of the | |
binary probability (p) of correct classification | |
(i.e. "area weighted user's accuracy"44; Table 3) using Eq. 7. Since | |
savannahs and shrublands are underrepresented in the regression-based | |
model25, their BGBC was instead estimated | |
using static root-to-shoot ratios reported by Mokany et al.22, which are | |
somewhat conservative in comparison to the IPCC Tier-1 defaults23,24 put | |
favoured for consistency with methods | |
used for grasslands (see below). Error was subsequently mapped from that | |
of the AGBC estimates and the root-to-shoot ratios applied BGBC of | |
tundra vegetation was mapped from AGBC using | |
a univariate regression model derived by Wang et al.26 that predicts | |
root-to-shoot ratio as a function of MAT. We applied the model using the | |
WorldClim version 2 MAT map59 and propagated | |
error from the AGBC estimates, the relative RMSE of MAT and the standard | |
error of regression coefficients. Where tundra AGB exceeded 25 Mg ha−1 | |
-- the maximum field-measured shrub biomass | |
reported by Berner et al.18 -- vegetation was considered to include | |
trees and the Reich et al.25 method described earlier for woody | |
vegetation was used instead. In the absence of a | |
continuous predictor of grassland root-to-shoot ratios, we applied | |
climate specific root-to-shoot ratios from Mokany et al.22 to the | |
corresponding climate regions of the Köppen-Gieger | |
classification43 (Table 2). Here, again, these ratios vary slightly from | |
the IPCC Tier-1 defaults23,24 but were chosen for their greater sample | |
size and specificity. Grassland BGBC | |
error was mapped from the error of the AGBC estimates and the respective | |
root-to-shoot ratios. Cropland BGBC was again estimated from | |
crop-specific yields and morphological parameters | |
(Online-only Table 2) following Wolf et al.21 and Eq. 8 where y is the | |
crop's yield (Mg ha−1), r is the root-to-shoot ratio of the crop, and h | |
is its harvest index. Here again we assume | |
that 2.5% of all harvested biomass is lost between the field and | |
farmgate and that root biomass is 44% C, following Wolf et al.21. BGBC | |
error was mapped from the error of the 2000-to-2010 | |
ANPP correction for BGBC allocation as described above for cropland | |
AGBC. Harmonizing biomass carbon maps The AGBC and BGBC maps were | |
harmonized separately following the same general | |
schema (Fig. 3). Given that our harmonized woody biomass map contains | |
biomass estimates for grid cells in which any amount of tree cover was | |
detected at the subpixel level (see section | |
1.1), we conserved its estimates regardless of the landcover reported by | |
the 2010 CCI map in order to more fully account for woody biomass in | |
non-forested areas17. We then used the | |
MODIS continuous vegetation fields percent tree cover map for 201063 to | |
allocate additional biomass density associated with the most probable | |
herbaceous cover (grass or crop) to each | |
grid cell in quantities complementary to that of the grid cell's | |
fractional tree cover estimate (Eq. 9) where μT is the total biomass | |
estimate of a grid cell, μw is the woody biomass | |
estimate for the grid cell, μh is its herbaceous biomass estimate, and q | |
is the MODIS fractional tree cover of the grid cell. Since MODIS tree | |
cover estimates saturate at around 80%64, | |
we linearly stretched values such that 80% was treated as complete tree | |
cover (100%). Moreover, we acknowledge that percent cover can | |
realistically exceed 100% when understory cover | |
is considered but we were unable to reasonably determine the extent of | |
underlying cover from satellite imagery. As such, our approach may | |
underestimate the contribution of herbaceous | |
C stocks in densely forested grid cells. The most likely herbaceous | |
cover type was determined from the CCI Landcover 2010 map, which we | |
aggregated into two "likely herbaceous cover" | |
classes -- grass or crop -- based on the assumed likelihood of cropland | |
in each CCI class (Online-only Table 1). However, due to inherent | |
classification error in the native CCI Landcover | |
map, when determining the herbaceous biomass contribution we weighted | |
the relative allocation of crop and grass biomass to a given grid cell | |
based on the probability of correct classification | |
by the CCI map (i.e. "user's accuracy", Table 6) of the most probable | |
herbaceous class ( ) such that μh can be further expressed as (Eq. 10) | |
where μi is the predicted biomass of the | |
most probable herbaceous class, and μj is that of the less probable | |
class. The uncertainty of a grid cell's total AGBC or BGBC estimate ( ) | |
was determined and mapped from that of its | |
components ( ) by summation in quadrature which can be simplified as | |
(Eq. 11) is the error of the grid cell's estimated μw, is the error of | |
its estimated μh, and is the error of its | |
q. Here, can be further decomposed and expressed as Eq. 12 to account | |
for the accuracy weighted allocation procedure expressed previously | |
(Eq. 10) is the error of the estimated biomass | |
density of the most probable herbaceous class, is the estimated standard | |
deviation of that class's Bernoulli probability (p; Eq. 7), and is the | |
error of the estimated biomass density | |
of the less probable herbaceous subclass. Exceptions to the above schema | |
were made in the tundra and boreal biomes -- as delineated by the | |
RESOLVE Ecoregions 2017 biome polygons65 -- | |
where thematic overlap was likely between the woody and tundra plant | |
biomass maps. A separate set of decision rules (Fig. 3) was used to | |
determine whether grid cells in these biomes | |
were to be exclusively allocated the estimate of the tundra plant map or | |
that of the fractional allocation procedure described above. In general, | |
any land in these biomes identified | |
as sparse landcover by the CCI landcover map (Online-only Table 1) was | |
assigned the tundra vegetation estimate. In addition, lands north of 60° | |
latitude with less than 10% tree cover | |
or where the tundra AGBC estimate exceeded that of the woody AGBC | |
estimate were also exclusively assigned the tundra vegetation estimate. | |
Lands north of 60° latitude not meeting these | |
criteria were assigned the woody value with the additional contribution | |
of grass. Subtle numerical artefacts emerged from the divergent | |
methodologies employed north and south of 60°N | |
latitude. These were eliminated by distance weighting grid cells within | |
1° of 60°N based on their linear proximity to 60°N and then averaging | |
estimates such that values at or north | |
of 61°N were exclusively based on the northern methodology, those at | |
60°N were the arithmetic average of the two methodologies and those at | |
or south of 59°N were exclusively based | |
on the southern methodology. This produced a seamless, globally | |
harmonized product that integrates the best remotely sensed estimates of | |
landcover-specific C density. Water bodies | |
identified as class "210" of the CCI 2010 landcover map were then masked | |
from our final products. Data Records Data layers (n = 4, Table 7) for | |
the maps of AGBC and BGBC density (Fig. | |
4) as well as their associated uncertainty maps which represent the | |
combined standard error of prediction (Fig. 5) are available as | |
individual 16-bit integer rasters in GeoTiff format. | |
All layers are natively in a WGS84 Mercator projection with a spatial | |
resolution of approximately 300 m at the equator and match that of the | |
ESA CCI Landcover Maps37. Raster values | |
are in units megagrams C per hectare (MgC ha−1) and have been scaled by | |
a factor of ten to reduce file size. These data are accessible through | |
the Oak Ridge National Laboratory (ORNL) | |
DAAC data repository. In addition, updated and/or derived | |
vegetation-specific layers that were used to create our harmonized 2010 | |
maps are | |
available as supplemental data on figshare67. Technical Validation Our | |
harmonized products rely almost exclusively upon maps and models that | |
have been rigorously validated by their | |
original producers and were often accompanied by constrained uncertainty | |
estimates. Throughout our harmonization procedure, we strived to | |
conserve the validity of each of these products | |
by minimizing the introduction of additional error and by tracking any | |
introductions, as described above, such that the final error layers | |
represent the cumulative uncertainty of the | |
inputs used. Ground truth AGB and BGB data are almost always collected | |
for individual landcover types. Consequently, we are unable to directly | |
assess the validity of our integrated | |
estimates beyond their relationships to individual landcover-specific | |
estimates and the extents to which they were modified from their | |
original, previously-validated form prior to | |
and during our harmonization procedure. Modifications to independent | |
biomass layers Temporal and spatial updates made to existing | |
landcover-specific maps of non-tree AGB resulted in | |
relatively small changes to their predictions. For example, we used | |
numerically calibrated MODIS imagery to extend the Landsat-based tundra | |
plant AGB model beyond its native extent | |
(the North Slope of Alaska) to the pan-Arctic region since neither a | |
comparable model nor a consistent Landsat time series were available for | |
this extent. We assessed the effects of | |
these assumptions by comparing our predictions for the North Slope with | |
those of the original map18 (Fig. 6a). Both positive and negative | |
discrepancies exist between ours and the original, | |
though these rarely exceed ±2 MgC ha−1 and no discernibly systematic | |
bias was evident. Fig. 6 figure 6 Differences between landcover-specific | |
AGBC estimates from the original published | |
maps and the modified versions used as inputs to create the 2010 | |
harmonized global maps. Tundra vegetation AGBC (a) is compared to the | |
Landsat-based map of Berner et al.45 for the | |
north slope of Alaska after converting it to units MgC ha−1. Here, the | |
comparison map was subsequently aggregated to a 1 km resolution and | |
reprojected for visualization. Grassland | |
AGBC (b) is compared to the AVHRR-based map of Xia et al.19 which | |
represents the average estimate between 1982--2006. For visualization, | |
the map was aggregated to a 5 km resolution | |
and subsequently reprojected after being masked to MODIS IGBP grasslands | |
in the year 200685 following Xia et al.19. As such, this map does not | |
necessarily represent the spatial distribution | |
of grid cells in which grassland estimates were used. Cropland AGBC (c) | |
is compared to the original circa 2000 estimates to assess the effects | |
of the 2000-to-2010 correction. The map | |
is masked to the native extent of the combined yield maps and aggregated | |
to a 5 km resolution for visualization. For all maps, negative values | |
indicate that our circa 2010 estimates | |
are lower than those of the earlier maps while positive values indicate | |
higher estimates. Full size image Our updated map of grassland biomass | |
carbon in the year 2010 was similarly | |
made by applying the original AVHRR-based model to calibrated MODIS | |
imagery. This too resulted in only subtle changes to the original | |
biomass map (Fig. 6b) that were rarely in excess | |
of 0.5 MgC ha−1. In most areas, our estimates were higher than those of | |
Xia et al.19 who mapped the mean AGBC density between 1986 and 2006. | |
Most of these elevated estimates corresponded | |
with areas in which significant NDVI increases ("greening") have been | |
reported while notably lower estimates in the Argentine Monte and | |
Patagonian steppe biomes of southern South America, | |
likewise, correspond with areas of reported "browning"68,69. Both | |
greening and browning trends are well documented phenomena and have been | |
linked to climatic changes70. Moreover, we | |
further compared AGBC estimates from both the original Xia et al.19 map | |
and our 2010 update to AGBC field measurements coordinated by the | |
Nutrient Network that were collected from | |
48 sites around the world between 2007 and 200949. The RMSE (0.68 MgC | |
ha−1) of our updated map was 10% less that of the Xia et al. map for | |
sites with less than 40% tree cover. Likewise, | |
our 2010 estimates were virtually unbiased (bias = −0.01 MgC ha−1) in | |
comparison to the Xia map (bias = 0.25 MgC ha−1). While still noisy, | |
these results suggest that our temporal update | |
improved the overall accuracy of estimated grassland AGBC. Finally, | |
cropland biomass carbon maps were also updated from their native epoch | |
(2000) to 2010 using pixel-wise rates of | |
MODIS ANPP change over a ten-year period. While MODIS ANPP may be a poor | |
snapshot of crop biomass in a single year, we assumed that its relative | |
change over time reflects real physiological | |
shifts affecting the cropland C cycle. This correction also resulted in | |
only small differences that rarely exceeded ±2 MgC ha−1 and that, | |
spatially, correspond well with observed declines | |
in the yields of select crops that have been linked to climate | |
change71,72 (Fig. 6c). Nonetheless, updated global yield maps comparable | |
to those available for 2000 would greatly improve | |
our understanding of the interactions between climate change, crop | |
yields, and C dynamics. Belowground biomass estimates Belowground | |
biomass is notoriously difficult to measure, model, | |
and also to validate. We accounted for the reported uncertainty of | |
nearly every variable considered when estimating belowground biomass and | |
pixel-level uncertainty, but we were unable | |
to perform an independent validation of our harmonized estimates at the | |
pixel level due to a paucity of globally consistent field data. To | |
complete such a task, a globally orchestrated | |
effort to collect more BGB samples data across all vegetation types is | |
needed. Given this lack of data, we instead compared the estimated | |
uncertainty of our BGBC maps to that of our | |
AGBC estimates to infer the sources of any divergence (Fig. 5). As | |
expected, our cumulative BGBC uncertainty layer generally reveals | |
greater overall uncertainty than our AGBC estimates, | |
with BGBC uncertainty roughly twice that of AGBC throughout most of the | |
globe. The highest absolute uncertainty was found in biomass rich | |
forests. Arid woodlands, especially those | |
of the Sahel and eastern Namibia, generally had the greatest relative | |
BGBC uncertainty, though their absolute uncertainty was quite small | |
(generally less than 3 MgC ha−1). Here, biomass | |
estimates of sparse woody vegetation were primarily responsible for | |
heightened relative uncertainty. High relative and absolute BGBC | |
uncertainty were also associated with predictions | |
in select mountainous forests (e.g. east central Chile) as well as | |
forested areas in and around cities. These patterns were largely driven | |
by AGB uncertainty in the GlobBiomass product. | |
Biomass harmonization The GlobBiomass global woody AGB map produced by | |
Santoro et al.30 comprises the backbone of our integrated products and, | |
with few exceptions, remains largely | |
unchanged in our final AGBC map. The native version of the GlobBiomass | |
map is accompanied by an error layer describing the uncertainty of each | |
pixel's biomass estimate and this too | |
forms the core of our integrated uncertainty layers. In areas with tree | |
cover, the global average error of GlobBiomass estimates is 39 Mg ha−1 | |
or 50% with greater relative uncertainty | |
in densely forested areas, along the margins of forested expanses like | |
farm fields and cities, and in similar areas with sparse tree cover. | |
Adding additional grass or crop biomass | |
in complementary proportion to a grid cell's tree cover often did not | |
exceed the estimated error of the original GlobBiomass map (Fig. 7). | |
Grid cells exceeding GlobBiomass's native | |
uncertainty comprise less than 40% of its total extent. Exceptions were | |
primarily found in grassland and cropland dominated regions where tree | |
cover was generally sparse, and, consequently, | |
the herbaceous biomass contribution was relatively high. Even so, the | |
absolute magnitude of these additions remains somewhat small (less than | |
2.3 MgC ha−1 for grassland and 15 MgC | |
ha−1 for cropland). Fig. 7 figure 7 Differences between the final | |
harmonized AGBC map and GlobBiomass AGBC. GlobBiomass AGB was aggregated | |
to a 300 m spatial resolution and converted | |
to C density prior to comparison. Negative values indicate areas where | |
the new map reports lower values than GlobBiomass while positive value | |
denote higher estimates. Full size image | |
Larger deviations from GlobBiomass were also present in areas of both | |
dryland Africa and the Arctic tundra biome, where we used independent | |
layers to estimate woody biomass. In African | |
drylands, GlobBiomass likely underestimates woody biomass by adopting | |
the conservative FAO definition (DBH \> 10 cm), which implicitly omits | |
the relatively small trees and shrubs that | |
are common to the region. The Bouvet map of Africa that we used to | |
supplement these estimates is not bound by this constraint, was | |
developed from region-specific data, and predicts | |
substantially higher AGB density throughout much of its extent with | |
comparatively high accuracy (RMSE = 17.1 Mg ha−1)35. GlobBiomass also | |
included sporadic biomass estimates throughout | |
much of the Arctic tundra biome. Trees are generally scarce throughout | |
this biome, which is instead dominated by dwarf shrubs and herbaceous | |
forbs and graminoids, so given GlobBiomass's | |
adherence to FAO guidelines, its predictions here may be spurious. We | |
thus prioritized the estimates of the independent model developed | |
specifically to collectively predict biomass | |
of both woody and herbaceous tundra vegetation. These estimates were | |
generally higher than GlobBiomass but agreed well with independent | |
validation data from North America (RMSE = 2.9 | |
Mg ha−1)18. Comparison with the IPCC Tier-1 global biomass carbon map | |
While far from a perfect comparison, the only other map to | |
comprehensively report global biomass carbon density | |
for all landcover types is the IPCC Tier-1 map for the year 2000 by | |
Ruesch and Gibbs28. As previously described, this map was produced using | |
an entirely different method ("stratify | |
and multiply") and distinct data sources23 and represents an earlier | |
epoch. However, the map is widely used for myriad applications, and it | |
may thus be informative to assess notable | |
differences between it and our new products. Ruesch and Gibbs28 report | |
total living C stocks of 345 petagrams (PgC) in AGBC and 133 PgC in BGBC | |
for a total of 478 PgC, globally. Our | |
estimates are lower at 287 PgC and 122 PgC in global AGBC and BGBC, | |
respectively, for a total of 409 PgC in living global vegetation | |
biomass. Herbaceous biomass in our maps comprised | |
9.1 and 28.3 PgC of total AGBC and BGBC, respectively. Half of all | |
herbaceous AGBC (4.5 PgC) and roughly 6% of all herbaceous BGBC (1.7 | |
PgC) was found in croplands. Moreover, we mapped | |
22.3 and 6.1 PgC, respectively, in the AGB and BGB of trees located | |
within the cropland extent. These trees constituted roughly 7% of all | |
global biomass C and are likely overlooked | |
by both the Ruesch and Gibbs map28 and by remotely sensed forest C maps | |
that are masked to forested areas. Zomer et al.17 first highlighted this | |
potential discrepancy in the Ruesch | |
and Gibbs map28 when they produced a remarkably similar estimate of 34.2 | |
Pg of overlooked C in cropland trees using Tier-1 defaults. However, | |
their estimates were assumed to be in | |
addition to the 474 PgC originally mapped by Ruesch and Gibbs28. Here, | |
we suggest that the 28.4 PgC we mapped in cropland trees is already | |
factored into our 409 PgC total. Our AGBC | |
product predicts substantially less biomass C than Ruesch and Gibbs28 | |
throughout most of the pantropical region and, to a lesser extent, | |
southern temperate forests (Fig. 8a). This | |
pattern has been noted by others comparing the Ruesch and Gibbs map28 to | |
other satellite-based biomass maps73 and may suggest that the IPCC | |
default values used to create it23 are spatially | |
biased. In addition, well-defined areas of high disagreement emerge in | |
Africa that directly correspond with the FAO boundaries of the "tropical | |
moist deciduous forest" ecofloristic | |
zone and suggest that this area, in particular, may merit critical | |
review. Moreover, the opposite pattern is observed in this same | |
ecofloristic zone throughout South America. Our map | |
also predicts greater AGBC throughout much of the boreal forest as well | |
as in African shrublands and the steppes of South America. We observed | |
similar, though less pronounced discrepancies, | |
when comparing BGBC maps (Fig. 8b). Notably, our map predicts | |
substantially more BGBC throughout the tundra biome -- a previously | |
underappreciated C stock that has recently risen to | |
prominance74 -- the boreal forest, African shrublands and most of South | |
America and Australia. However, we predict less BGBC in nearly all | |
rainforests (Temperate and Tropical). These | |
differences and their distinct spatial patterns correspond with the | |
vegetation strata used to make the IPCC Tier-1 map28 and suggest that | |
the accuracy of the "stratify and multiply" | |
method depends heavily upon the quality of the referenced and spatial | |
data considered. Inaccuracies in these data may, in turn, lead to false | |
geographies. Integrating, continuous spatial | |
estimates that better capture local and regional variation, as we have | |
done, may thus greatly improve our understanding of global carbon | |
geographies and their role in the earth system. | |
Congruence with IPCC Tier-2 and Tier-3 nationally reported woody carbon | |
stocks The error and variance between our woody biomass estimates -- | |
when aggregated to the country level -- and | |
comparable totals reported in the FRA were less for comparisons made | |
against FRA estimates generated using higher tier IPCC methodologies | |
than for those based on Tier-1 approaches | |
(Fig. 9). Across the board for AGBC, BGBC, and total C comparisons, the | |
relative RMSE (RMSECV) of our estimates, when compared to estimates | |
generated using high tier methods, was roughly | |
half of that obtained from comparisons with Tier-1 estimates (Table 8). | |
Likewise, the coefficient of determination (R2) was greatest for | |
comparisons with Tier-3 estimates. For each | |
pool-specific comparison (AGBC, BGBC, and total C), the slopes of the | |
relationships between Tier-1, 2, and 3 estimates were neither | |
significantly different from a 1:1 relationship | |
nor from one another (p \> 0.05; ANCOVA). Combined, these results | |
suggest that our maps lead to C stock estimates congruent with those | |
attained from independent, higher-tier reporting | |
methodologies. Fig. 9 figure 9 Comparison of woody biomass density | |
estimates to corresponding estimates of the FAO's FRA and the USFS's | |
FIA. National woody AGBC totals derived from | |
the woody components of our harmonized maps are compared to national | |
totals reported in the 2015 FRA62 (a) in relation to the IPCC inventory | |
methodology used by each country. Likewise, | |
we derived woody AGBC totals for US states and compared them to the | |
corresponding totals reported by the 2014 FIA75 (b), a Tier-3 inventory. | |
We also show the additional effect of considering | |
non-woody C -- as is reported in our harmonized maps -- in light green. | |
Similar comparisons were made between our woody BGBC estimates and the | |
corresponding estimates of both the FRA | |
(c) and FIA (d). We further summed our woody AGBC and BGBC estimates and | |
compared them to the total woody C stocks reported by both the | |
FRA (e) and FIA (f). Full size image Table 8 | |
Statistical comparison of woody biomass carbon totals derived from the | |
2010 harmonized maps and those reported by the FRA in relation to the | |
IPCC inventory methodology used. Full size | |
table To explore this association at a finer regional scale, we also | |
compared our woody C estimates to the United States Forest Service's | |
Forest Inventory Analysis75 (FIA) and found | |
similarly strong congruence for AGBC and Total C stocks but subtle | |
overestimates for BGBC (Fig. 9). The FIA is a Tier-3 inventory of woody | |
forest biomass C stocks that is based on | |
extensive and statistically rigorous field sampling and subsequent | |
upscaling, We used data available at the state level for the year 2014 | |
-- again, the only year in which we could obtain | |
data partitioned by AGBC and BGBC. Like our FRA comparison, we found a | |
tight relationship between our woody AGBC totals and those reported by | |
the FIA (Fig. 9b; RMSECV = 25.7%, R2 = | |
0.960, slope = 1.10, n = 48). Our woody BGBC estimates, though, were | |
systematically greater than those reported by the FIA (Fig. 9d; RMSECV = | |
86.4%, R2 = 0.95, slope = 1.51, n = 48). | |
This trend has been noted by others27 and suggests that the global model | |
that we used to estimate woody BGBC may not be appropriate for some | |
finer scale applications as is foretold | |
by the elevated uncertainty reported in our corresponding uncertainty | |
layer (Fig. 5b). Our total woody C (AGBC + BGBC) estimates (Fig. 9f), | |
however, agreed well with the FIA (RMSECV | |
= 34.1%, R2 = 0.961, slope = 1.17, n = 48) and thus reflect the outsized | |
contribution of AGBC to the total woody C stock. When the contribution | |
of herbaceous C stocks is further added | |
to these comparisons, our stock estimates intuitively increase in rough | |
proportion to a state's proportional extent of herbaceous cover. The | |
effect of this addition is particularly | |
pronounced for BGBC estimates due to the large root-to-shoot ratios of | |
grassland vegetation. The relative congruence of our results with | |
higher-tier stock estimates suggests that our | |
maps could be used to facilitate broader adoption of higher-tier methods | |
among countries currently lacking the requisite data and those seeking | |
to better account for C in non-woody | |
biomass. This congruence spans a comprehensive range of biophysical | |
conditions and spatial scales ranging from small states to large | |
nations. Moreover, a recent study suggests that | |
the fidelity of the underlying GlobBiomass AGB map may extend to even | |
finer scales31. While our BGBC estimates may differ from some fine-scale | |
estimates (Fig. 9d), their tight agreement | |
with high tier BGBC totals at the national level (Fig. 9c) suggests that | |
they may still be well suited for many national-scale C inventories -- | |
especially for countries lacking requisite | |
high tier data. Use of our maps is unlikely to introduce error in excess | |
of that currently implicit in Tier-1 estimates. Credence, though, should | |
be given to the associated uncertainty | |
estimates. To facilitate wider adoption of higher-tier methodologies, | |
our maps could be used to derive new, region-specific default values for | |
use in Tier-2 frameworks76 or to either | |
represent or calibrate 2010 baseline conditions in Tier-3 frameworks. In | |
so doing, inventories and studies alike could more accurately account | |
for the nuanced global geographies of | |
biomass C. Usage Notes These maps are intended for global applications | |
in which continuous spatial estimates of live AGBC and/or BGBC density | |
are needed that span a broad range of | |
vegetation types and/or require estimates circa 2010. They are loosely | |
based upon and share the spatial resolution of the ESA CCI Landcover | |
2010 map37, which can be used to extract | |
landcover specific C totals. However, our products notably do not | |
account for C stored in non-living C pools like litter or coarse woody | |
debris, nor soil organic matter, though these | |
both represent large, additional ecosystem C stocks77,78,79. Our maps | |
are explicitly intended for global scale applications seeking to | |
consider C in the collective living biomass of | |
multiple vegetation types. For global scale applications focused | |
exclusively on the C stocks of a single vegetation type, we strongly | |
encourage users to instead use the respective | |
input map or model referenced in Table 1 to avoid potential errors that | |
may have been introduced by our harmonization procedure. For AGB | |
applications over smaller extents, users should | |
further consider whether locally specific products are available. If | |
such maps are not available and our maps are considered instead, | |
credence should be given to their pixel-level | |
uncertainty estimates. As mentioned above, the biomass of shrublands was | |
only explicitly accounted for in Africa and the Arctic tundra, since | |
neither broad-scale maps nor models generalizable | |
to other areas were available in the existing literature. As such, we | |
caution against the use of our maps outside of these areas when | |
shrubland biomass is of particular interest or | |
importance. Moreover, in contrast to the estimates for all other | |
vegetation types considered, which we upscaled to a 300 m resolution, | |
cropland C estimates were largely based on relatively | |
coarse 8 km resolution data that were downscaled using bilinear | |
resampling to achieve a 300 m spatial resolution. As such, these | |
estimates may not adequately capture the underlying | |
finer-scale spatial variation and should be interpreted with that in | |
mind. Likewise, we reiterate that some BGBC estimates may differ from | |
locally derived Tier-3 estimates, and attention | |
should thus be given to our reported pixel-level uncertainty for all | |
applications. Finally, our maps should not be used in comparison with | |
the IPCC Tier-1 map of Ruesch and Gibbs (2008) | |
to detect biomass change between the two study periods due to | |
significant methodological differences between these products. An | |
estimated 720 and 811 million people in the world faced | |
hunger in 2020, according to the United Nations (UN), and nearly one in | |
three people in the world (2.37 billion) did not have access to adequate | |
food in 2020. The vulnerabilities and | |
inadequacies of global food systems are expected to further intensify | |
over the coming years. The combination of NASA Earth science data with | |
socioeconomic data provides key information | |
for sustainable use of available resources. NASA's Socioeconomic Data | |
and Applications Center (SEDAC) is the home for NASA socioeconomic data | |
and is a gateway between the social sciences | |
and the Earth sciences. SEDAC provides numerous datasets and data | |
collections that may be useful for studies into agriculture and water | |
management. SEDAC also provides information | |
about the connections that support efforts to end hunger, achieve food | |
security and improved nutrition, and promote sustainable agriculture. | |
Women in the West African country of Senegal | |
take a break from crushing millet. The United Nations World Food Program | |
estimates that 46 percent of households in Senegal lack reliable access | |
to adequate amounts of food. Credit: | |
Molly Brown Women in the West African country of Senegal take a break | |
from crushing millet. The United Nations World Food Program estimates | |
that 46% of households in Senegal lack reliable | |
access to adequate amounts of food. Credit: Molly Brown. NASA helps | |
develop tools to address food security and works with decision-makers | |
and data users to tailor these tools to specific | |
locations and user needs. These efforts help address issues like water | |
management for irrigation, crop-type identification and land use, | |
coastal and lake water quality monitoring, | |
drought preparedness, and famine early warnings. Much of this work is | |
carried out and supported fully or in part by the agency's Applied | |
Sciences Program, which works with individuals | |
and institutions worldwide to inform decision-making, enhance quality of | |
life, and strengthen the economy. The Applied Sciences Program co-leads | |
the international Earth Observations | |
for Sustainable Development Goals initiative, which advances global | |
knowledge about effective ways that Earth observations and geospatial | |
information can support the SDGs. The NASA | |
datasets and resources listed below, coupled with other data and | |
resources in this Data Pathfinder, also help measure progress toward | |
meeting United Nations' Sustainable Development | |
Goals (SDGs), particularly SDG 2: Zero Hunger. These data can provide a | |
better overall view for monitoring the food insecurity of vulnerable | |
populations, tracking agricultural production | |
related to incomes of small-scale food producers, and monitoring | |
environmental impacts to soil, water, fertilizer, pesticide pollution, | |
and changes in biodiversity. More information | |
is available in the Connection of Sustainable Development Goals to | |
Agriculture and Water Management section on the main Pathfinder landing | |
page. Agriculture and Human Dimensions Agriculture | |
and Food Security theme landing page Global Agricultural Inputs, v1 The | |
five datasets in this data collection provide global gridded data and | |
maps on pesticide application, phosphorus | |
in manure and chemical fertilizers, and nitrogen in manure and chemical | |
fertilizers Global Pesticide Grids (PEST-CHEMGRIDS), v1.01 (2015, 2020, | |
2025) Global coverage; 5 arc-min spatial | |
resolution; GeoTIFF, netCDF-4 Web Map Service Layers Food Supply Effects | |
of Climate Change on Global Food Production from SRES Emissions and | |
Socioeconomic Scenarios, v1 (1970 -- 2080) | |
Global coverage; national resolution; .xlsx Web Map Service Layers Food | |
Insecurity Hotspots Data Set, v1 (2009 -- 2019) Global coverage; | |
national resolution; GeoTIFF, Shapefile Web | |
Map Service Layers Groundswell Spatial Population and Migration | |
Projections at One-Eighth Degree According to SSPs and RCPs, v1 (2010 -- | |
2050) Allows users to understand how slow-onset | |
climate change impacts on water availability and crop productivity, | |
coupled with sea-level rise and storm surge, may affect the future | |
population distribution and climate-related internal | |
migration in low to middle income countries Crop Production Twentieth | |
Century Crop Statistics, v1 (1900 -- 2017) Global coverage (selected | |
countries); national/sub-national resolution; | |
annual Global Population Projection Grid Data Groundswell Spatial | |
Population and Migration Projections at One-Eighth Degree According to | |
SSPs and RCPs, v1 (2010 -- 2050) Climate Change | |
Impact Effects of Climate Change on Global Food Production from SRES | |
Emissions and Socioeconomic Scenarios, v1 (1970 -- 2080) Global | |
coverage; national resolution; .xlsx Web Map Service | |
Layers Groundswell Spatial Population and Migration Projections at | |
One-Eighth Degree According to SSPs and RCPs, v1 (2010 -- 2050) | |
Environmental Performance 2022 Environmental Performance | |
Index Global coverage; national resolution; .xlsx, csv 15 static maps | |
Poverty-related Data Humidity is a measure of the amount of water vapor | |
present in the air. High humidity impairs | |
heat exchange efficiency by reducing the rate of moisture evaporation | |
from the skin and other surfaces. This can create challenges for | |
agricultural workers, as well as the crops they | |
grow. The Modern-Era Retrospective analysis for Research and | |
Applications, Version 2 (MERRA-2) provides data beginning in 1980. Due | |
to the amount of historical data available, MERRA-2 | |
data can be used to look for trends and patterns as well as anomalies. | |
There are several options available: 1-hourly, 3-hourly, 6-hourly, | |
daily, and monthly. These options provide | |
information on precipitation. The NASA Earth Exchange Global Daily | |
Downscaled Projections (NEX-GDDP) dataset is comprised of | |
high-resolution, bias-corrected global downscaled climate | |
projections derived from the General Circulation Model (GCM) runs | |
conducted under the Coupled Model Intercomparison Project Phase 6 | |
(CMIP6) and across all four "Tier 1" greenhouse | |
gas emissions scenarios known as Shared Socioeconomic Pathways (SSPs). | |
This dataset provides a set of global, high resolution, bias-corrected | |
climate change projections that can be | |
used to evaluate climate change impacts on processes that are sensitive | |
to finer-scale climate gradients and the effects of local topography on | |
climate conditions. Uses include: air | |
temperature, precipitation volume, humidity, stellar radiation, and | |
atmospheric wind speed. The atmosphere is a mixture of gases that | |
surrounds the Earth. It helps make life possible | |
by providing us with air to breathe, shielding us from harmful | |
ultraviolet (UV) radiation coming from the Sun, trapping heat to warm | |
the planet, and preventing extreme temperature | |
differences between day and night. Without the atmosphere, temperatures | |
would be well below freezing everywhere on Earth's surface. Instead, the | |
heat absorbed and trapped by our atmosphere | |
keeps our planet's average surface temperature at a balmy 15°C (59°F). | |
Some of the atmosphere's gases, like carbon dioxide, are particularly | |
good at absorbing and trapping radiation. | |
Changes in the amounts of these gases directly affect our climate. Gases | |
in Earth's Atmosphere Each of the planets in our solar system has an | |
atmosphere, but none of them have the | |
same ratio of gases or layered structure as Earth's atmosphere. Nitrogen | |
and oxygen are by far the most common gases in our atmosphere. Dry air | |
is composed of about 78% nitrogen (N2) | |
and about 21% oxygen (O2). The remaining less than 1% of the atmosphere | |
is a mixture of gases, including argon (Ar) and carbon dioxide (CO2). | |
The atmosphere also contains varying amounts | |
of water vapor, on average about 1%. There are also many, tiny, solid or | |
liquid particles, called aerosols, in the atmosphere. Aerosols can be | |
made of dust, spores and pollen, salt | |
from sea spray, volcanic ash, smoke, and pollutants introduced through | |
human activity. Earth's Atmosphere Has Layers The atmosphere becomes | |
thinner (less dense and lower in air pressure) | |
the further it extends from the Earth's surface. It gradually gives way | |
to the vacuum of space. There is no precise top of the atmosphere, but | |
the area between 100-120 km (62-75 miles) | |
above the Earth's surface is often considered the boundary between the | |
atmosphere and space because the air is so thin here. However, there are | |
measurable traces of atmospheric gases | |
beyond this boundary, detectable for hundreds of kilometers/miles from | |
Earth's surface. There are several unique layers in Earth's atmosphere. | |
Each has characteristic temperatures, | |
pressures, and phenomena. We live in the troposphere, the layer closest | |
to Earth's surface, where most clouds are found and almost all weather | |
occurs. Some jet aircraft fly in the | |
next layer, the stratosphere, which contains the jet streams and a | |
region called the ozone layer. The next layer, the mesosphere, is the | |
coldest because the there are almost no air | |
molecules there to absorb heat energy. There are so few molecules for | |
light to refract off of that the sky also changes from blue to black in | |
this layer. And farthest from the surface | |
we have the thermosphere, which absorbs much of the harmful radiation | |
that reaches Earth from the Sun, causing this layer to reach extremely | |
high temperatures. Beyond the thermosphere | |
is the exosphere, which represents the transition from Earth's | |
atmosphere to space. Planetary Atmospheres Earth is not the only world | |
with an atmosphere. Each of the planets - and | |
even a few moons - in our solar system have an atmosphere. Some planets | |
have active atmospheres with clouds, wind, rain and powerful storms. | |
Scientists use light spectroscopy to observe | |
the atmospheres of planets and moons in other solar systems . Each of | |
the planets in our solar system has a uniquely structured atmosphere. | |
The atmosphere of Mercury is extremely thin | |
and is not very different from the vacuum of space. The gas giant | |
planets in our solar system - Jupiter, Saturn, Uranus and Neptune - each | |
have a thick, deep atmosphere. The smaller, | |
rocky planets - Earth, Venus and Mars - each have thinner atmospheres, | |
hovering above their solid surfaces. The moons in our solar system | |
typically have thin atmospheres, with the | |
exception of Saturn's moon, Titan. Air pressure at the surface of Titan | |
is higher than on Earth! Of the five officially recognized dwarf | |
planets, Pluto has a thin atmosphere that expands | |
and collapses seasonally, and Ceres has an extremely thin and transient | |
atmosphere made of water vapor. But only Earth's atmosphere has the | |
layered structure that traps enough of the | |
Sun's energy for warmth while also blocking much of the harmful | |
radiation from the Sun. This important balance is necessary to maintain | |
life on Earth. Forests are one of the world's | |
largest banks of carbon-rich biomass. This is why when researchers | |
mapped biomass in the past, they typically focused on the world's | |
forests. But this approach leaves out considerable | |
amounts of biomass in grasslands, shrublands, croplands, and other | |
biomes. New maps, published at NASA's Oak Ridge National Laboratory | |
Distributed Active Archive Center (ORNL DAAC) | |
and described in Nature Scientific Data, combine remotely sensed biomass | |
data for different land cover types into harmonized global maps of above | |
and belowground biomass for the year | |
2010. People often conflate forest biomass with total biomass," said | |
Seth Spawn, lead author of the research and doctoral candidate at | |
the University of Wisconsin, Madison. "Researchers | |
have spent a lot of time developing nice remotely sensed maps of | |
aboveground forest biomass, but they intentionally omit other land | |
covers and the carbon stored below ground in plant | |
roots. We haven't had the whole picture." Spawn and his team combined | |
maps of forest biomass with other land cover specific biomass maps that | |
use remotely sensed data. They allocated | |
fractional contributions to a given grid cell using data on land cover, | |
percent tree cover, and the presence of secondary vegetation. These maps | |
show a sizable stock of biomass outside | |
of forested areas, especially in the trees located in savannas and farm | |
fields. "There's more carbon on croplands than I would have expected," | |
said Spawn. Trees on orchards and farms | |
practicing agroforestry have a carbon stock that is overlooked in | |
previous biomass maps. Globally, trees on croplands stored about 28 | |
metric gigatons of carbon in 2010, which is 7 | |
percent of the total stock of carbon in plants, according to these new | |
maps. The team produced maps of uncertainty with the data they used and | |
published them along with the dataset. | |
Published in March, the dataset is already being used for a number of | |
applications. Researchers are using the dataset in integrated assessment | |
models that incorporate economics and | |
the Earth system. It is also being used to model carbon emissions from | |
past and potential land use changes and the carbon impacts of bioenergy | |
transitions. Water is a key component | |
of the overall Earth system, cycling through each component, moving | |
within the atmosphere, the ocean, the cryosphere (including snow cover | |
and snowpack), surface water of rivers and | |
lakes, and subsurface water. Water availability is critical for human | |
consumption, agriculture and food security, industry, and energy | |
development. Assessing water availability, including | |
the amount and type of precipitation is critical to monitoring | |
agricultural practices and water resource availability and for providing | |
interventions when necessary. According to the | |
U.N., water use has been growing globally at twice the rate as the | |
global population is increasing. More and more areas are reaching the | |
limit at which water services can be sustainably | |
delivered, especially in arid regions. Groundwater, a major water | |
resource for maintaining cropland productivity, is declining through the | |
extensive use of water for agricultural irrigation, | |
where aquifer recharge cannot keep up with groundwater extraction. | |
Unfortunately, changes in terrestrial water storage, especially with | |
regard to groundwater, are poorly known and | |
sparsely sampled. Complicating matters further, global freshwater is not | |
only unevenly distributed, but sources of freshwater such as lakes and | |
rivers often cross geopolitical boundaries. | |
Integrating satellite data with land-based and other measurements, | |
geospatial data, and hydrologic models help to better understand | |
controls on global water resources and how changing | |
water resources impact social-environmental systems across geopolitical | |
boundaries. Earth Observation Data by Sensor According to the U.N., | |
water use has been growing globally at twice | |
the rate as the global population is increasing. More and more areas are | |
reaching the limit at which water services can be sustainably delivered, | |
especially in arid regions. Groundwater, | |
a major water resource for maintaining cropland productivity, is | |
declining through the extensive use of water for agricultural | |
irrigation, where aquifer recharge cannot keep up with | |
groundwater extraction. Unfortunately, changes in terrestrial water | |
storage, especially with regard to groundwater, are poorly known and | |
sparsely sampled. Complicating matters further, | |
global freshwater is not only unevenly distributed, but sources of | |
freshwater such as lakes and rivers often cross geopolitical boundaries. | |
Integrating satellite data with land-based | |
and other measurements, geospatial data, and hydrologic models help to | |
better understand controls on global water resources and how changing | |
water resources impact social-environmental | |
systems across geopolitical boundaries. Earth Observation Data by Sensor | |
GRACE, GRACE-FO Instruments aboard the joint NASA/German Space Agency | |
Gravity Recovery And Climate Experiment | |
(GRACE, operational 2002 to 2017) and GRACE Follow-On (GRACE-FO, | |
launched in 2018) satellites obtain measurements about changes in | |
Earth's gravity. Since water has mass, changes in | |
groundwater storage can be detected as changes in gravity. GRACE and | |
GRACE-FO measurements help assess water storage changes in monthly, | |
total surface, and groundwater depth. These | |
data are available from 2002 to present; the data track total water | |
storage time-variations and anomalies (changes from the time-mean) at a | |
resolution of approximately 90,000 km2 and | |
larger. These measurements are unimpeded by clouds and track the entire | |
land water column from the surface down to deep aquifers. GRACE and | |
GRACE-FO data are uniquely valuable for | |
regional studies to determine general trends in land water storage as | |
well as for assessing basin-scale water budgets (e.g., the balance | |
between precipitation, evapotranspiration, | |
and runoff). GRACE and GRACE-FO Mascon Ocean, Ice, and Hydrology | |
Equivalent Water Height dataset provides gridded monthly global water | |
storage/height anomalies relative to a time-mean. | |
The data are processed at NASA's Jet Propulsion Laboratory (JPL) using | |
the mascon approach. Mass Concentration blocks (mascons) are a form of | |
gravity field basis functions to which | |
GRACE observations are optimally fit. For more information on this | |
approach, see the JPL Monthly Mass Grids webpage. Data are represented | |
as Water Equivalent Thickness (WET), representing | |
the total terrestrial water storage anomalies from soil moisture, snow, | |
surface water (including rivers, lakes, and reservoirs), as well as | |
groundwater and aquifers. Scientists at | |
NASA's Goddard Space Flight Center use GRACE-FO data to generate weekly | |
groundwater and soil moisture drought indicators. The drought indicators | |
describe current wet or dry conditions, | |
expressed as a percentile showing the probability of occurrence for a | |
specific location and time of year, with lower values (orange/red) | |
indicating drier than normal conditions and | |
higher values (blues) indicating wetter than normal conditions. The | |
drought model is also used to make forecasts of expected drought | |
conditions one, two, and three months into the | |
future. NASA, in collaboration with other agencies, has developed models | |
of groundwater that incorporate satellite information with ground-based | |
data (when ground-based data are available). | |
These models are part of the Land Data Assimilation System (LDAS), which | |
includes a global collection (GLDAS) and a North American collection | |
(NLDAS). NASA's Goddard Earth Sciences | |
Data and Information Services Center (GES DISC) optimally reorganized | |
some large hydrological datasets as time series (also known as data | |
rods) for a set of water cycle-related variables | |
from the NLDAS and GLDAS, the Land Parameter Parameter Model (LPRM), | |
TRMM, and GRACE data assimilation. These are available at GES DISC | |
Hydrology Data Rods. The Modern-Era Retrospective | |
analysis for Research and Applications, Version 2 (MERRA-2) provides | |
data beginning in 1980. Due to the amount of historical data available, | |
MERRA-2 data can be used to look for trends | |
and patterns, as well as anomalies. There are several options available: | |
hourly and monthly from 1980. Remote sensing is the acquiring of | |
information from a distance. NASA observes | |
Earth and other planetary bodies via remote sensors on satellites and | |
aircraft that detect and record reflected or emitted energy. Remote | |
sensors, which provide a global perspective | |
and a wealth of data about Earth systems, enable data-informed decision | |
making based on the current and future state of our planet. Satellites | |
can be placed in several types of orbits | |
around Earth. The three common classes of orbits are low-Earth orbit | |
(approximately 160 to 2,000 km above Earth), medium-Earth orbit | |
(approximately 2,000 to 35,500 km above Earth), | |
and high-Earth orbit (above 35,500 km above Earth). Satellites orbiting | |
at 35,786 km are at an altitude at which their orbital speed matches the | |
planet's rotation, and are in what | |
is called geosynchronous orbit (GSO). In addition, a satellite in GSO | |
directly over the equator will have a geostationary orbit. A | |
geostationary orbit enables a satellite to maintain | |
its position directly over the same place on Earth's surface. Aqua | |
satellite orbit illustrating polar orbital track. NASA's Aqua satellite | |
completes one orbit every 99 minutes and | |
passes within 10 degrees of each pole. This enables the Moderate | |
Resolution Imaging Spectroradiometer (MODIS) aboard Aqua to acquire full | |
global imagery every 1-2 days. Credit: NASA | |
Aqua. Low-Earth orbit is a commonly used orbit since satellites can | |
follow several orbital tracks around the planet. Polar-orbiting | |
satellites, for example, are inclined nearly 90 | |
degrees to the equatorial plane and travel from pole to pole as Earth | |
rotates. This enables sensors aboard the satellite to acquire data for | |
the entire globe rapidly, including the | |
polar regions. Many polar-orbiting satellites are considered | |
Sun-synchronous, meaning that the satellite passes over the same | |
location at the same solar time each cycle. One example | |
of a Sun-synchronous, polar-orbiting satellite is NASA's Aqua satellite, | |
which orbits approximately 705 km above Earth's surface. Non-polar | |
low-Earth orbit satellites, on the other | |
hand, do not provide global coverage but instead cover only a partial | |
range of latitudes. The joint NASA/Japan Aerospace Exploration Agency | |
Global Precipitation Measurement (GPM) Core | |
Observatory is an example of a non-Sun-synchronous low-Earth orbit | |
satellite. Its orbital track acquires data between 65 degrees north and | |
south latitude from 407 km above the planet. | |
A medium-Earth orbit satellite takes approximately 12 hours to complete | |
an orbit. In 24-hours, the satellite crosses over the same two spots on | |
the equator every day. This orbit is | |
consistent and highly predictable. As a result, this is an orbit used by | |
many telecommunications and GPS satellites. One example of a | |
medium-Earth orbit satellite constellation is | |
the European Space Agency's Galileo global navigation satellite system | |
(GNSS), which orbits 23,222 km above Earth. While both geosynchronous | |
and geostationary satellites orbit at 35,786 | |
km above Earth, geosynchronous satellites have orbits that can be tilted | |
above or below the equator. Geostationary satellites, on the other hand, | |
orbit Earth on the same plane as the | |
equator. These satellites capture identical views of Earth with each | |
observation and provide almost continuous coverage of one area. The | |
joint NASA/NOAA Geostationary Operational Environmental | |
Satellite (GOES) series of weather satellites are in geostationary | |
orbits above the equator. Observing with the Electromagnetic Spectrum | |
Electromagnetic energy, produced by the vibration | |
of charged particles, travels in the form of waves through the | |
atmosphere and the vacuum of space. These waves have different | |
wavelengths (the distance from wave crest to wave crest) | |
and frequencies; a shorter wavelength means a higher frequency. Some, | |
like radio, microwave, and infrared waves, have a longer wavelength, | |
while others, such as ultraviolet, x-rays, | |
and gamma rays, have a much shorter wavelength. Visible light sits in | |
the middle of that range of long to shortwave radiation. This small | |
portion of energy is all that the human eye | |
is able to detect. Instrumentation is needed to detect all other forms | |
of electromagnetic energy. NASA instrumentation utilizes the full range | |
of the spectrum to explore and understand | |
processes occurring here on Earth and on other planetary bodies. Some | |
waves are absorbed or reflected by atmospheric components, like water | |
vapor and carbon dioxide, while some wavelengths | |
allow for unimpeded movement through the atmosphere; visible light has | |
wavelengths that can be transmitted through the atmosphere. Microwave | |
energy has wavelengths that can pass through | |
clouds, an attribute utilized by many weather and communication | |
satellites. The primary source of the energy observed by satellites, is | |
the Sun. The amount of the Sun's energy reflected | |
depends on the roughness of the surface and its albedo, which is how | |
well a surface reflects light instead of absorbing it. Snow, for | |
example, has a very high albedo and reflects up | |
to 90% of incoming solar radiation. The ocean, on the other hand, | |
reflects only about 6% of incoming solar radiation and absorbs the rest. | |
Often, when energy is absorbed, it is re-emitted, | |
usually at longer wavelengths. For example, the energy absorbed by the | |
ocean gets re-emitted as infrared radiation. All things on Earth | |
reflect, absorb, or transmit energy, the amount | |
of which varies by wavelength. Just as your fingerprint is unique to | |
you, everything on Earth has a unique spectral fingerprint. Researchers | |
can use this information to identify different | |
Earth features as well as different rock and mineral types. The number | |
of spectral bands detected by a given instrument, its spectral | |
resolution, determines how much differentiation | |
a researcher can identify between materials. Drought, vegetation health, | |
and soil moisture all can be tracked remotely. This Data Pathfinder | |
provides links to NASA Earth observations, | |
tools, and other resources applicable to agricultural production and | |
water management. The planet NASA studies the most is Earth. NASA's | |
end-to-end Earth observations enable agricultural | |
producers to make informed decisions about global market conditions, | |
water management, in-season crop conditions, severe weather, and | |
sustainability. This Data Pathfinder will help | |
guide you through the process of selecting and using datasets applicable | |
to agriculture and water management, and provides links to specific data | |
sources. If you are new to remote | |
sensing, the What is Remote Sensing? Backgrounder provides a | |
comprehensive overview. In addition, NASA's Applied Remote Sensing | |
Training Program (ARSET) provides numerous training | |
What's big, covered in water, yet 100 times drier than the Sahara | |
Desert? It's not a riddle, it's the Moon! For centuries, astronomers | |
debated whether water exists on Earth's closest neighbor. In 2020, data | |
from NASA's SOFIA mission confirmed water exists in the sunlit area of | |
the lunar surface as molecules of H2O embedded within, or perhaps | |
sticking to the surface of, grains of lunar dust. Here is a brief | |
history of the discoveries leading up to the confirmation of water on | |
the Moon. | |
When early astronomers looked up at the Moon, they were struck by the | |
large, dark spots on its surface. In 1645, Dutch astronomer Michael van | |
Langren published the first-known map of the Moon referring to the dark | |
spots as "maria" -- the Latin word for "seas" -- and putting into | |
writing the widely-held view that the marks were oceans on the lunar | |
surface. Similar maps from Johannes Hevelius (1647), Giovanni Riccioli | |
and Francesco Grimaldi (1651) were published over the next few years. We | |
now know these spots to be plains of basalt created by early volcanic | |
eruptions, but the nomenclature of 'maria' (plural) or 'mare' (singular) | |
remains. | |
American astronomer William Pickering made measurements in the late | |
1800s that led him to conclude the Moon essentially has no atmosphere. | |
With no clouds and no atmosphere, scientists generally agreed that any | |
water on the lunar surface would evaporate immediately. Pickering's | |
measurements led to a widespread view that the Moon was devoid of water. | |
As scientists made headway in understanding the behavior of substances | |
that are prone to vaporize at relatively low temperatures -- called | |
volatiles -- theoretical physicist Kenneth Watson published a paper in | |
1961 describing how a substance like water could exist on the Moon. | |
Watson's paper first popularized the idea that water ice could stick to | |
the bottom of craters on the Moon that never receive light from the Sun, | |
while sunlit areas on the Moon would be so hot that water would | |
evaporate near-instantly. These lightless areas of the Moon are called | |
"permanently shadowed regions." | |
The Apollo era brought humans to the lunar surface for the very first | |
time, giving researchers the opportunity to directly look for signs of | |
water on the Moon. When tested, soil samples brought back by Apollo | |
astronauts revealed no sign of water. Scientists concluded that the | |
lunar surface must be completely dry, and the prospect of water wasn't | |
seriously considered again for decades. | |
NASA's Clementine mission launched in 1994 to orbit the Moon for two | |
months and collect information about its minerals. Clementine data | |
suggested there was ice in a permanently shadowed region of the Moon. | |
The Lunar Prospector Mission focused on permanently shadowed craters to | |
look deeper into the discovery and in 1998 found that the largest | |
concentrations of hydrogen exist in the areas of the lunar surface that | |
are never exposed to sunlight. The results indicated water ice at the | |
lunar poles. However, the images were low resolution so no strong | |
conclusions could be made. | |
Capitalizing on major advances in technology since the Apollo era, | |
researchers from Brown University revisited the Apollo samples. They | |
found hydrogen inside tiny beads of volcanic glass. Since no volcanoes | |
are erupting on the Moon today, the discovery presented evidence that | |
water had existed in the Moon when the volcanoes erupted in the Moon's | |
ancient past. Additionally, the preserved hydrogen provided clues to the | |
origins of lunar water: if it emerged from erupting volcanoes, it must | |
have come from within the Moon. The discovery suggested that water was a | |
part of the Moon since its early existence -- and perhaps since it first | |
formed. | |
A suite of spacecraft enabled exciting discoveries in 2009. None were | |
designed to look for water on the Moon, yet the Indian Space Research | |
Organization's Chandrayaan-1 and NASA's Cassini and Deep Impact missions | |
detected signs of hydrated minerals in the form of oxygen and hydrogen | |
molecules in sunlit areas of the Moon. Researchers couldn't determine | |
whether they were seeing hydration by hydroxyl (OH) or water (H2O). They | |
also debated whether the amount of hydration depended on the time of | |
day. | |
The Lunar Crater Observation and Sensing Satellite (LCROSS) spacecraft | |
and Lunar Reconnaissance Orbiter (LRO) launched together in 2009. Later | |
that year, LCROSS intentionally discharged a projectile into a crater | |
believed to contain water ice, and flew through the debris from the | |
projectile's impact. Four minutes later, LCROSS itself intentionally | |
impacted the Moon while LRO observed. The combined observations showed | |
grains of water ice in the ejected material. The LRO and LCROSS findings | |
added to a growing body of evidence that water exists on the Moon in the | |
form of ice within permanently shadowed regions. LRO continues to orbit | |
the Moon and provide data used to characterize and map lunar resources, | |
including hydrogen. | |
Data from Moon Mineralogy Mapper (M3), carried by ISRO's Chandrayaan-1, | |
provided scientists with the first high-resolution map of the minerals | |
that make up the lunar surface. The NASA instrument flew aboard India's | |
Chandrayaan-1 mission in 2009. An analysis of the full set of data from | |
M3, announced in 2018, revealed multiple confirmed locations of water | |
ice in permanently shadowed regions of the Moon. | |
In 2020, NASA announced the discovery of water on the sunlit surface of | |
the Moon. Data from the Strategic Observatory for Infrared Astronomy | |
(SOFIA), revealed that in Clavius crater, water exists in concentrations | |
roughly equivalent to a 12-ounce bottle of water within a cubic meter of | |
soil across the lunar surface. The discovery showed that water could be | |
distributed across the lunar surface, even on sunlit portions, and not | |
confined to cold, dark areas. | |
In 2023, a new map of water distribution on the Moon provided hints | |
about how water may be moving across the Moon's surface. The map, made | |
using SOFIA data, extends to the Moon's South Pole -- the intended | |
region of study for NASA's Artemis missions, including the water-hunting | |
rover, VIPER. | |
Researchers have confirmed that water exists both in the sunlit and | |
shadowed surfaces of the Moon, yet many questions remain. Lunar | |
scientists continue to investigate the origins of water and its | |
behavior. There is evidence that the water on the Moon comes from | |
ancient and current comet impacts, icy micrometeorites colliding on the | |
lunar surface, and lunar dust interactions with the solar wind. However, | |
more research is needed to understand the full history, present, and | |
future of water on the Moon. Writer: Allison Gasparini and Molly | |
Wasser`\nScience `{=tex}Advisors: Casey Honniball, Tim Livengood, NASA's | |
Goddard Space Flight Center In 2019, scientists discovered that water is | |
being released from the Moon during meteor showers. Water on the Moon | |
could come from a surprising source, our Sun. In 2020, NASA scientists | |
confirmed the presence of H2O on the Moon. | |
Viewed from space, one of the most striking features of our home planet | |
is the water, in both liquid and frozen forms, that covers approximately | |
75% of the Earth's surface. Geologic evidence suggests that large | |
amounts of water have likely flowed on Earth for the past 3.8 billion | |
years---most of its existence. Believed to have initially arrived on the | |
surface through the emissions of ancient volcanoes, water is a vital | |
substance that sets the Earth apart from the rest of the planets in our | |
solar system. In particular, water appears to be a necessary ingredient | |
for the development and nourishment of life. | |
Water is practically everywhere on Earth. Moreover, it is the only known | |
substance that can naturally exist as a gas, a liquid, and solid within | |
the relatively small range of air temperatures and pressures found at | |
the Earth's surface. In all, the Earth's water content is about 1.39 | |
billion cubic kilometers (331 million cubic miles), with the bulk of it, | |
about 96.5%, being in the global oceans. As for the rest, approximately | |
1.7% is stored in the polar icecaps, glaciers, and permanent snow, and | |
another 1.7% is stored in groundwater, lakes, rivers, streams, and soil. | |
Only a thousandth of 1% of the water on Earth exists as water vapor in | |
the atmosphere. | |
Despite its small amount, this water vapor has a huge influence on the | |
planet. Water vapor is a powerful greenhouse gas, and it is a major | |
driver of the Earth's weather and climate as it travels around the | |
globe, transporting latent heat with it. Latent heat is heat obtained by | |
water molecules as they transition from liquid or solid to vapor; the | |
heat is released when the molecules condense from vapor back to liquid | |
or solid form, creating cloud droplets and various forms of | |
precipitation. For human needs, the amount of freshwater on Earth---for | |
drinking and agriculture---is particularly important. Freshwater exists | |
in lakes, rivers, groundwater, and frozen as snow and ice. Estimates of | |
groundwater are particularly difficult to make, and they vary widely. | |
(The value in the above table is near the high end of the range.) | |
Groundwater may constitute anywhere from approximately 22 to 30% of | |
fresh water, with ice (including ice caps, glaciers, permanent snow, | |
ground ice, and permafrost) accounting for most of the remaining 78 to | |
70%. | |
The water, or hydrologic, cycle describes the pilgrimage of water as | |
water molecules make their way from the Earth's surface to the | |
atmosphere and back again, in some cases to below the surface. This | |
gigantic system, powered by energy from the Sun, is a continuous | |
exchange of moisture between the oceans, the atmosphere, and the land. | |
Studies have revealed that evaporation---the process by which water | |
changes from a liquid to a gas---from oceans, seas, and other bodies of | |
water (lakes, rivers, streams) provides nearly 90% of the moisture in | |
our atmosphere. Most of the remaining 10% found in the atmosphere is | |
released by plants through transpiration. Plants take in water through | |
their roots, then release it through small pores on the underside of | |
their leaves. In addition, a very small portion of water vapor enters | |
the atmosphere through sublimation, the process by which water changes | |
directly from a solid (ice or snow) to a gas. The gradual shrinking of | |
snow banks in cases when the temperature remains below freezing results | |
from sublimation. | |
Together, evaporation, transpiration, and sublimation, plus volcanic | |
emissions, account for almost all the water vapor in the atmosphere that | |
isn't inserted through human activities. While evaporation from the | |
oceans is the primary vehicle for driving the surface-to-atmosphere | |
portion of the hydrologic cycle, transpiration is also significant. For | |
example, a cornfield 1 acre in size can transpire as much as 4,000 | |
gallons of water every day. | |
After the water enters the lower atmosphere, rising air currents carry | |
it upward, often high into the atmosphere, where the air is cooler. In | |
the cool air, water vapor is more likely to condense from a gas to a | |
liquid to form cloud droplets. Cloud droplets can grow and produce | |
precipitation (including rain, snow, sleet, freezing rain, and hail), | |
which is the primary mechanism for transporting water from the | |
atmosphere back to the Earth's surface. | |
When precipitation falls over the land surface, it follows various | |
routes in its subsequent paths. Some of it evaporates, returning to the | |
atmosphere; some seeps into the ground as soil moisture or groundwater; | |
and some runs off into rivers and streams. Almost all of the water | |
eventually flows into the oceans or other bodies of water, where the | |
cycle continues. At different stages of the cycle, some of the water is | |
intercepted by humans or other life forms for drinking, washing, | |
irrigating, and a large variety of other uses. | |
Groundwater is found in two broadly defined layers of the soil, the | |
"zone of aeration," where gaps in the soil are filled with both air and | |
water, and, further down, the "zone of saturation," where the gaps are | |
completely filled with water. The boundary between these two zones is | |
known as the water table, which rises or falls as the amount of | |
groundwater changes. | |
The amount of water in the atmosphere at any moment in time is only | |
12,900 cubic kilometers, a minute fraction of Earth's total water | |
supply: if it were to completely rain out, atmospheric moisture would | |
cover the Earth's surface to a depth of only 2.5 centimeters. However, | |
far more water---in fact, some 495,000 cubic kilometers of it---are | |
cycled through the atmosphere every year. It is as if the entire amount | |
of water in the air were removed and replenished nearly 40 times a year. | |
Water continually evaporates, condenses, and precipitates, and on a | |
global basis, evaporation approximately equals precipitation. Because of | |
this equality, the total amount of water vapor in the atmosphere remains | |
approximately the same over time. However, over the continents, | |
precipitation routinely exceeds evaporation, and conversely, over the | |
oceans, evaporation exceeds precipitation. | |
In the case of the oceans, the continual excess of evaporation versus | |
precipitation would eventually leave the oceans empty if they were not | |
being replenished by additional means. Not only are they being | |
replenished, largely through runoff from the land areas, but over the | |
past 100 years, they have been over-replenished: sea level around the | |
globe has risen approximately 17 centimeters over the course of the | |
twentieth century. | |
Sea level has risen both because of warming of the oceans, causing water | |
to expand and increase in volume, and because more water has been | |
entering the ocean than the amount leaving it through evaporation or | |
other means. A primary cause for increased mass of water entering the | |
ocean is the calving or melting of land ice (ice sheets and glaciers). | |
Sea ice is already in the ocean, so increases or decreases in the annual | |
amount of sea ice do not significantly affect sea level. | |
Throughout the hydrologic cycle, there are many paths that a water | |
molecule might follow. Water at the bottom of Lake Superior may | |
eventually rise into the atmosphere and fall as rain in Massachusetts. | |
Runoff from the Massachusetts rain may drain into the Atlantic Ocean and | |
circulate northeastward toward Iceland, destined to become part of a | |
floe of sea ice, or, after evaporation to the atmosphere and | |
precipitation as snow, part of a glacier. | |
Water molecules can take an immense variety of routes and branching | |
trails that lead them again and again through the three phases of ice, | |
liquid water, and water vapor. For instance, the water molecules that | |
once fell 100 years ago as rain on your great- grandparents' farmhouse | |
in Iowa might now be falling as snow on your driveway in California. | |
Among the most serious Earth science and environmental policy issues | |
confronting society are the potential changes in the Earth's water cycle | |
due to climate change. The science community now generally agrees that | |
the Earth's climate is undergoing changes in response to natural | |
variability, including solar variability, and increasing concentrations | |
of greenhouse gases and aerosols. Furthermore, agreement is widespread | |
that these changes may profoundly affect atmospheric water vapor | |
concentrations, clouds, precipitation patterns, and runoff and stream | |
flow patterns. For example, as the lower atmosphere becomes warmer, | |
evaporation rates will increase, resulting in an increase in the amount | |
of moisture circulating throughout the troposphere (lower atmosphere). | |
An observed consequence of higher water vapor concentrations is the | |
increased frequency of intense precipitation events, mainly over land | |
areas. Furthermore, because of warmer temperatures, more precipitation | |
is falling as rain rather than snow. | |
In parts of the Northern Hemisphere, an earlier arrival of spring-like | |
conditions is leading to earlier peaks in snowmelt and resulting river | |
flows. As a consequence, seasons with the highest water demand, | |
typically summer and fall, are being impacted by a reduced availability | |
of fresh water. | |
Warmer temperatures have led to increased drying of the land surface in | |
some areas, with the effect of an increased incidence and severity of | |
drought. The Palmer Drought Severity Index, which is a measure of soil | |
moisture using precipitation measurements and rough estimates of changes | |
in evaporation, has shown that from 1900 to 2002, the Sahel region of | |
Africa has been experiencing harsher drought conditions. This same index | |
also indicates an opposite trend in southern South America and the south | |
central United States. | |
While the brief scenarios described above represent a small portion of | |
the observed changes in the water cycle, it should be noted that many | |
uncertainties remain in the prediction of future climate. These | |
uncertainties derive from the sheer complexity of the climate system, | |
insufficient and incomplete data sets, and inconsistent results given by | |
current climate models. However, state of the art (but still incomplete | |
and imperfect) climate models do consistently predict that precipitation | |
will become more variable, with increased risks of drought and floods at | |
different times and places. | |
Orbiting satellites are now collecting data relevant to all aspects of | |
the hydrologic cycle, including evaporation, transpiration, | |
condensation, precipitation, and runoff. NASA even has one satellite, | |
Aqua, named specifically for the information it is collecting about the | |
many components of the water cycle. | |
Aqua launched on May 4, 2002, with six Earth-observing instruments: the | |
Atmospheric Infrared Sounder (AIRS), the Advanced Microwave Sounding | |
Unit (AMSU), the Humidity Sounder for Brazil (HSB), the Advanced | |
Microwave Scanning Radiometer for the Earth Observing System (AMSR-E), | |
the Moderate Resolution Imaging Spectroradiometer (MODIS), and Clouds | |
and the Earth's Radiant Energy System (CERES). | |
Since water vapor is the Earth's primary greenhouse gas, and it | |
contributes significantly to uncertainties in projections of future | |
global warming, it is critical to understand how it varies in the Earth | |
system. In the first years of the Aqua mission, AIRS, AMSU, and HSB | |
provided space-based measurements of atmospheric temperature and water | |
vapor that were more accurate than any obtained before; the sensors also | |
made measurements from more altitudes than any previous sensor. The HSB | |
is no longer operational, but the AIRS/AMSU system continues to provide | |
high-quality atmospheric temperature and water vapor measurements. | |
More recent studies using AIRS data have demonstrated that most of the | |
warming caused by carbon dioxide does not come directly from carbon | |
dioxide, but rather from increased water vapor and other factors that | |
amplify the initial warming. Other studies have shown improved | |
estimation of the landfall of a hurricane in the Bay of Bengal by | |
incorporating AIRS temperature measurements, and improved understanding | |
of large-scale atmospheric patterns such as the Madden-Julian | |
Oscillation. | |
In addition to their importance to our weather, clouds play a major role | |
in regulating Earth's climate system. MODIS, CERES, and AIRS all collect | |
data relevant to the study of clouds. The cloud data include the height | |
and area of clouds, the liquid water they contain, and the sizes of | |
cloud droplets and ice particles. The size of cloud particles affects | |
how they reflect and absorb incoming sunlight, and the reflectivity | |
(albedo) of clouds plays a major role in Earth's energy balance. | |
One of the many variables AMSR-E monitors is global precipitation. The | |
sensor measures microwave energy, some of which passes through clouds, | |
and so the sensor can detect the rainfall even under the clouds. | |
Water in the atmosphere is hardly the only focus of the Aqua mission. | |
Among much else, AMSR-E and MODIS are being used to study sea ice. Sea | |
ice is important to the Earth system not just as an important element in | |
the habitat of polar bears, penguins, and some species of seals, but | |
also because it can insulate the underlying liquid water against heat | |
loss to the often frigid overlying polar atmosphere and because it | |
reflects sunlight that would otherwise be available to warm the ocean. | |
When it comes to sea ice, AMSR-E and MODIS provide complementary | |
information. AMSR-E doesn't record as much detail about ice features as | |
MODIS does, but it can distinguish ice versus open water even when it is | |
cloudy. The AMSR-E measurements continue, with improved resolution and | |
accuracy, a satellite record of changes in the extent of polar ice that | |
extends back to the 1970s. | |
AMSR-E and MODIS also provide monitoring of snow coverage over land, | |
another key indicator of climate change. As with sea ice, AMSR-E allows | |
routine monitoring of the snow, irrespective of cloud cover, but with | |
less spatial detail, while MODIS sees greater spatial detail, but only | |
under cloud-free conditions. | |
As for liquid water on land, AMSR-E provides information about soil | |
moisture, which is crucial for vegetation including agricultural crops. | |
AMSR-E's monitoring of soil moisture globally permits, for example, the | |
early identification of signs of drought. Aqua is the most comprehensive | |
of NASA's water cycle missions, but it isn't alone. In fact, the Terra | |
satellite also has MODIS and CERES instruments onboard, and several | |
other spacecraft have made or are making unique water-cycle | |
measurements. | |
The Ice, Cloud, and Land Elevation Satellite (ICESat) was launched in | |
January 2003, and it collected data on the topography of the Earth's ice | |
sheets, clouds, vegetation, and the thickness of sea ice off and on | |
until October 2009. A new ICESat mission, ICESat-2, is now under | |
development and is scheduled to launch in 2015. | |
The Gravity Recovery and Climate Experiment (GRACE) is a unique mission | |
that consists of two spacecraft orbiting one behind the other; changes | |
in the distance between the two provide information about the gravity | |
field on the Earth below. Because gravity depends on mass, some of the | |
changes in gravity over time signal a shift in water from one place on | |
Earth to another. Through measurements of changing gravity fields, GRACE | |
scientists are able to derive information about changes in the mass of | |
ice sheets and glaciers and even changes in groundwater around the | |
world. | |
CloudSat is advancing scientists' understanding of cloud abundance, | |
distribution, structure, and radiative properties (how they absorb and | |
emit energy, including thermal infrared energy escaping from Earth's | |
surface). Since 2006, CloudSat has flown the first satellite-based, | |
millimeter-wavelength cloud radar---an instrument that is 1000 times | |
more sensitive than existing weather radars on the ground. Unlike | |
ground-based weather radars that use centimeter wavelengths to detect | |
raindrop-sized particles, CloudSat's radar allows the detection of the | |
much smaller particles of liquid water and ice in the large cloud masses | |
that contribute significantly to our weather. | |
The joint NASA and French Cloud-Aerosol Lidar and Infrared Pathfinder | |
Satellite Observations (CALIPSO) is providing new insight into the role | |
that clouds and atmospheric aerosols (particles like dust and pollution) | |
play in regulating Earth's weather, climate, and air quality. CALIPSO | |
combines an active laser instrument with passive infrared and visible | |
imagers to probe the vertical structure and properties of thin clouds | |
and aerosols over the globe. | |
July through October fall within the dry season in the western and | |
northern Amazon rainforest, but a particularly acute lack of rain during | |
this period in 2023 has pushed the region into a severe drought. The OLI | |
(Operational Land Imager) instrument on Landsat 8 captured this image | |
(right) of the parched Rio Negro in the Brazilian province of Amazonas | |
near the city of Manaus on October 3, 2023. On that date, the level of | |
the river, the largest tributary of the Amazon River, had dropped to | |
15.14 meters (50.52 feet), according to data collected by the Port of | |
Manaus. | |
For comparison, the image on the left shows the same area on October 8, | |
2022, when the water level was 19.59 meters, a more typical level for | |
October. Rio Negro water levels continued to drop in the days after the | |
image was collected, reaching a record low of 13.49 meters on October | |
17, 2023. Some areas in the Amazon River's watershed have received less | |
rain between July and September than any year since 1980, Reuters | |
reported. The drought has been particularly severe in the Rio Negro | |
watershed in northern Amazonas, as well as parts of southern Venezuela | |
and southern Colombia. "Overall, this is a pretty unusual and extreme | |
situation," said René Garreaud, an atmospheric scientist at the | |
University of Chile. | |
"The primary culprit exacerbating the drought appears to be El Niño." | |
This cyclical warming of surface waters in the central-eastern Pacific | |
functions somewhat like a boulder in the middle of a stream, disrupting | |
atmospheric circulation patterns in ways that lead to wetter conditions | |
over the equatorial Pacific and drier conditions over the Amazon Basin. | |
According to news outlets, the low river water levels on the Rio Negro | |
and other nearby rivers have disrupted drinking water supplies in | |
hundreds of communities, slowed commercial navigation, and led to fish | |
and dolphin die-offs. | |
Manaus, the capital and largest city of the Brazilian state of Amazonas, | |
is the primary transportation hub for the upper Amazon, serving as an | |
important transit point for soap, beef, and animal hides. Other | |
industries with a presence in the city of two million people include | |
chemical, ship, and electrical equipment manufacturing. | |
After rapidly growing in volume just a few years earlier, northwest | |
Iran's Lake Urmia nearly dried out in autumn 2023. The largest lake in | |
the Middle East and one of the largest hypersaline lakes on Earth at its | |
greatest extent, Lake Urmia has for the most part transformed into a | |
vast, dry salt flat. | |
It stands in contrast to the image from three years earlier (left), | |
acquired by the OLI on Landsat 8 on September 8, 2020, when water filled | |
most of the basin and salt deposits were only visible around the | |
perimeter of the lake. | |
The replenishment followed a period of above-average precipitation that | |
sent a surge of freshwater into the basin, expanding its watery | |
footprint. Drier conditions have since brought levels back down. The | |
longer-term trend for Urmia has been one toward drying. In 1995, Lake | |
Urmia reached a high-water mark; then in the ensuing two decades, the | |
lake level dropped more than 7 meters (23 feet) and lost approximately | |
90 percent of its area. | |
Consecutive droughts, agricultural water use, and dam construction on | |
rivers feeding the lake have contributed to the decline. | |
A shrinking Lake Urmia has implications for ecological and human health. | |
The lake, its islands, and surrounding wetlands comprise valuable | |
habitat and are recognized as a UNESCO Biosphere Reserve, Ramsar site, | |
and national park. | |
The area provides breeding grounds for waterbirds such as flamingos, | |
white pelicans, and white-headed ducks, as well as a stopover for | |
migratory species. However, with low lake levels, what water remains | |
becomes more saline and taxes the populations of brine shrimp and other | |
food sources for larger animals. | |
A shrinking lake also increases the likelihood of dust from the exposed | |
lakebed becoming swept up by winds and degrading air quality. Recent | |
studies have linked the low water levels in Lake Urmia with respiratory | |
health impacts among the local population. The relative effects of | |
climate, water usage, and dams on Lake Urmia's water level is a topic of | |
debate. The lake did see some recovery during a 10-year restoration | |
program beginning in 2013. | |
However, the efficacy of that effort has been difficult to parse since | |
strong rains also fell during that period. Some research has concluded | |
that climatic factors were primarily responsible for the recovery. | |
The deep-blue sea is turning a touch greener. While that may not seem as | |
consequential as, say, record warm sea surface temperatures, the color | |
of the ocean surface is indicative of the ecosystem that lies beneath. | |
Communities of phytoplankton, microscopic photosynthesizing organisms, | |
abound in near-surface waters and are foundational to the aquatic food | |
web and carbon cycle. | |
This shift in the water's hue confirms a trend expected under climate | |
change and signals changes to ecosystems within the global ocean, which | |
covers 70 percent of Earth's surface. Researchers led by B. B. Cael, a | |
principal scientist at the U.K.'s National Oceanography Centre, revealed | |
that 56 percent of the global sea surface has undergone a significant | |
change in color in the past 20 years. | |
After analyzing ocean color data from the MODIS (Moderate Resolution | |
Imaging Spectroradiometer) instrument on NASA's Aqua satellite, they | |
found that much of the change stems from the ocean turning more green. | |
The map above highlights the areas where ocean surface color changed | |
between 2002 and 2022, with darker shades of green representing | |
more-significant differences (higher signal-to-noise ratio). By | |
extension, said Cael, "these are places we can detect a change in the | |
ocean ecosystem in the last 20 years." | |
The study focused on tropical and subtropical regions, excluding higher | |
latitudes, which are dark for part of the year, and coastal waters, | |
where the data are naturally very noisy. The black dots on the map | |
indicate the area, covering 12 percent of the ocean's surface, where | |
chlorophyll levels also changed over the study period. | |
Chlorophyll has been the go-to measurement for remote sensing scientists | |
to gauge phytoplankton abundance and productivity. However, those | |
estimates use only a few colors in the visible light spectrum. The | |
values shown in green are based on the whole gamut of colors and | |
therefore capture more information about the ecosystem as a whole. A | |
long time series from a single sensor is relatively rare in the remote | |
sensing world. As the Aqua satellite was celebrating its 20th year in | |
orbit in 2022---far exceeding its design life of 6 years---Cael wondered | |
what long term trends could be discovered in the data. In particular, he | |
was curious what might have been missed in all the ocean color | |
information it had collected. "There's more encoded in the data than we | |
actually make use of," he said. | |
By going big with the data, the team discerned an ocean color trend that | |
had been predicted by climate modeling, but one that was expected to | |
take 30-40 years of data to detect using satellite-based chlorophyll | |
estimates. That's because the natural variability in chlorophyll is high | |
relative to the climate change trend. The new method, incorporating all | |
visible light, was robust enough to confirm the trend in 20 years. At | |
this stage, it is difficult to say what exact ecological changes are | |
responsible for the new hues. However, the authors posit, they could | |
result from different assemblages of plankton, more detrital particles, | |
or other organisms such as zooplankton. | |
It is unlikely the color changes come from materials such as plastics or | |
other pollutants, said Cael, since they are not widespread enough to | |
register at large scales. "What we do know is that in the last 20 years, | |
the ocean has become more stratified," he said. Surface waters have | |
absorbed excess heat from the warming climate, and as a result, they are | |
less prone to mixing with deeper, more nutrient-rich layers. | |
This scenario would favor plankton adapted to a nutrient-poor | |
environment. The areas of ocean color change align well with where the | |
sea has become more stratified, said Cael, but there is no such overlap | |
with sea surface temperature changes. More insights into Earth's aquatic | |
ecosystems may soon be on the way. | |
NASA's PACE (Plankton, Aerosol, Cloud, ocean Ecosystem) satellite, set | |
to launch in 2024, will return observations in finer color resolution. | |
The new data will enable researchers to infer more information about | |
ocean ecology, such as the diversity of phytoplankton species and the | |
rates of phytoplankton growth. | |
On September 10, 2023, a low-pressure storm brought heavy rains to | |
northeastern Libya, causing deadly flooding and devastation in cities | |
along the Mediterranean coast. On the coast of Libya's Cyrenaica region, | |
Al Bayda recorded 414 millimeters (16 inches) of rain in one day. | |
Nearby, the port city of Derna received more than 100 millimeters (4 | |
inches) over the course of the storm---far exceeding the city's average | |
monthly rainfall for September of less than 1.5 millimeters (0.1 | |
inches). Derna lies at the end of a long, narrow valley, called a wadi, | |
which is dry for most of the year. | |
Floods triggered two dams along Wadi Derna to collapse, sending | |
floodwater and mud to the city. According to news reports, floodwater | |
swept away roads and entire neighborhoods. The images above show the | |
Cyrenaica region before and after the storm. They are false color, which | |
makes water (blue) stand out from the surroundings. The image on the | |
right, acquired on September 13, shows water filling low-lying areas and | |
wadis inland from the coast. | |
The image on the left shows the same area on September 7. Both images | |
were acquired with the Moderate Resolution Imaging Spectroradiometer | |
(MODIS) on NASA's Terra satellite. The flooding and damage in Derna is | |
difficult to see at this resolution, although sediment flowing into the | |
Mediterranean is visible in natural color images. | |
In the days prior to making landfall in Libya, the same low-pressure | |
storm (named Storm Daniel by the Hellenic National Meteorological | |
Service) swamped parts of Greece, Turkey, and Bulgaria. As the storm | |
approached Libya, it developed characteristics of a tropical-like | |
cyclone, or "medicane," with winds measuring around 70 to 80 kilometers | |
(43 to 50 miles) per hour. | |
The natural-color image above, acquired with MODIS on NASA's Terra | |
satellite, shows the storm on September 10 as it made landfall in | |
northeastern Libya. Only one or two medicanes typically develop in a | |
year, according to NOAA. | |
As of September 13, authorities were still conducting search and rescue | |
operations in the region. Derna was still largely inaccessible on that | |
day, making it difficult to assess the full impact of the flood. | |
Sea surface temperatures have a large influence on climate and weather. | |
For example, every 3 to 7 years a wide swath of the Pacific Ocean along | |
the equator warms by 2 to 3 degrees Celsius. | |
This warming is a hallmark of the climate pattern El Niño, which changes | |
rainfall patterns around the globe, causing heavy rainfall in the | |
southern United States and severe drought in Australia, Indonesia, and | |
southern Asia. | |
On a smaller scale, ocean temperatures influence the development of | |
tropical cyclones (hurricanes and typhoons), which draw energy from warm | |
ocean waters to form and intensify. | |
These sea surface temperature maps are based on observations by the | |
Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA's Aqua | |
satellite. The satellite measures the temperature of the top millimeter | |
of the ocean surface. In this map, the coolest waters appear in blue | |
(approximately -2 degrees Celsius), and the warmest temperatures appear | |
in pink-yellow (35 degrees Celsius). | |
Landmasses and the large area of sea ice around Antarctica appear in | |
shades of gray, indicating no data were collected. | |
The most obvious pattern shown in the time series is the year-round | |
difference in sea surface temperatures between equatorial regions and | |
the poles. | |
Various warm and cool currents stand out even in monthly averages of sea | |
surface temperature. A band of warm waters snakes up the East Coast of | |
the United States and veers across the North Atlanticâ€"the Gulf Stream. | |
Although short-lived weather events that influence ocean temperature are | |
often washed out in monthly averages, a few events show up. | |
For example, in December 2003, strong winds blew southwest from the Gulf | |
of Mexico over Central America toward the Pacific Ocean, driving surface | |
waters away from the coast, and allowing cold water from deeper in the | |
ocean to well up to the surface. These winds are a recurring phenomenon | |
in the area in the winter; they are known as Tehuano winds. | |
At the base of the ocean food web are single-celled algae and other | |
plant-like organisms known as phytoplankton. Like plants on land, | |
phytoplankton use chlorophyll and other light-harvesting pigments to | |
carry out photosynthesis, absorbing atmospheric carbon dioxide to | |
produce sugars for fuel. Chlorophyll in the water changes the way it | |
reflects and absorbs sunlight, allowing scientists to map the amount and | |
location of phytoplankton. These measurements give scientists valuable | |
insights into the health of the ocean environment, and help scientists | |
study the ocean carbon cycle. | |
These chlorophyll maps show milligrams of chlorophyll per cubic meter of | |
seawater each month. Places where chlorophyll amounts were very low, | |
indicating very low numbers of phytoplankton are blue. Places where | |
chlorophyll concentrations were high, meaning many phytoplankton were | |
growing, are dark green. The observations come from the Moderate | |
Resolution Imaging Spectroradiometer (MODIS) on NASA's Aqua satellite. | |
Land is dark gray, and places where MODIS could not collect data because | |
of sea ice, polar darkness, or clouds are light gray. | |
The highest chlorophyll concentrations, where tiny surface-dwelling | |
ocean plants are thriving, are in cold polar waters or in places where | |
ocean currents bring cold water to the surface, such as around the | |
equator and along the shores of continents. It is not the cold water | |
itself that stimulates the phytoplankton. Instead, the cool temperatures | |
are often a sign that the water has welled up to the surface from deeper | |
in the ocean, carrying nutrients that have built up over time. In polar | |
waters, nutrients accumulate in surface waters during the dark winter | |
months when plants can't grow. When sunlight returns in the spring and | |
summer, the plants flourish in high concentrations. | |
A band of cool, plant-rich waters circles the globe at the Equator, with | |
the strongest signal in the Atlantic Ocean and the open waters of the | |
Pacific Ocean. This zone of enhanced phytoplankton growth comes from the | |
frequent upwelling of cooler, deeper water as a result of the dominant | |
easterly trade winds blowing across the ocean surface. In many coastal | |
areas, the rising slope of the sea floor pushes cold water from the | |
lowest layers of the ocean to the surface. The rising, or upwelling | |
water carries iron and other nutrients from the ocean floor. Cold | |
coastal upwelling and subsequent phytoplankton growth are most evident | |
along the west coasts of North and South America and southern Africa. | |
In March and April 2023, some earth scientists began to point out that | |
average sea surface temperatures had surpassed the highest levels seen | |
in a key data record maintained by NOAA. Months later, they remain at | |
record levels, with global sea surface temperatures 0.99°C (1.78°F) | |
above average in July. That was the fourth consecutive month they were | |
at record levels. Scientists from NASA have taken a closer look at why. | |
"There are a lot of things that affect the world's sea surface | |
temperatures, but two main factors have pushed them to record heights," | |
said Josh Willis, an oceanographer at NASA's Jet Propulsion Laboratory | |
(JPL). "We have an El Niño developing in the Pacific, and that's on top | |
of long-term global warming that has been pushing ocean temperatures | |
steadily upward almost everywhere for a century." | |
The map above shows sea surface temperature anomalies on August 21, | |
2023, when many areas were more than 3°C (5.4°F) warmer than normal. On | |
that date, much of the central and eastern regions of the equatorial | |
Pacific were unusually warm, the signature of a developing El Niño. As | |
has been the case for weeks, large patches of warm water were also | |
present in the Northwest Pacific near Japan and the Northeast Pacific | |
near California and Oregon. Portions of the Indian, Southern, and Arctic | |
Oceans also showed unusual warmth. | |
The map is based on data from the Multiscale Ultrahigh Resolution Sea | |
Surface Temperature (MUR SST) project, a JPL effort that blends | |
measurements of sea surface temperatures from multiple NASA, NOAA, and | |
international satellites, as well as ship and buoy observations. Rather | |
than showing absolute temperature, the anomaly reflects the difference | |
between the sea surface temperature on August 21, 2023, and the | |
2003-2014 average for that day. The video below, also based on MUR SST | |
data, shows global sea surface temperatures since April 1, 2023, the | |
period when they have been at record-breaking levels. The warmest waters | |
appear dark red. | |
In March and April 2023, some earth scientists began to point out that | |
average sea surface temperatures had surpassed the highest levels seen | |
in a key data record maintained by NOAA. Months later, they remain at | |
record levels, with global sea surface temperatures 0.99°C (1.78°F) | |
above average in July. | |
That was the fourth consecutive month they were at record levels. | |
Scientists from NASA have taken a closer look at why. "There are a lot | |
of things that affect the world's sea surface temperatures, but two main | |
factors have pushed them to record heights," said Josh Willis, an | |
oceanographer at NASA's Jet Propulsion Laboratory (JPL). "We have an El | |
Niño developing in the Pacific, and that's on top of long-term global | |
warming that has been pushing ocean temperatures steadily upward almost | |
everywhere for a century." | |
The map above shows sea surface temperature anomalies on August 21, | |
2023, when many areas were more than 3°C (5.4°F) warmer than normal. On | |
that date, much of the central and eastern regions of the equatorial | |
Pacific were unusually warm, the signature of a developing El Niño. | |
As has been the case for weeks, large patches of warm water were also | |
present in the Northwest Pacific near Japan and the Northeast Pacific | |
near California and Oregon. Portions of the Indian, Southern, and Arctic | |
Oceans also showed unusual warmth. The map is based on data from the | |
Multiscale Ultrahigh Resolution Sea Surface Temperature (MUR SST) | |
project, a JPL effort that blends measurements of sea surface | |
temperatures from multiple NASA, NOAA, and international satellites, as | |
well as ship and buoy observations. Rather than showing absolute | |
temperature, the anomaly reflects the difference between the sea surface | |
temperature on August 21, 2023, and the 2003-2014 average for that day. | |
The video below, also based on MUR SST data, shows global sea surface | |
temperatures since April 1, 2023, the period when they have been at | |
record-breaking levels. The warmest waters appear dark red. | |
"Over the long term, we're seeing more heat and warmer sea surface | |
temperatures pretty much everywhere," said Gavin Schmidt, the director | |
of NASA's Goddard Institute for Space Studies. "That long-term trend is | |
almost entirely attributable to human forcing---the fact that we've put | |
such a huge amount of greenhouse gas in the atmosphere since the start | |
of the industrial era." Schmidt noted that other factors---such as | |
weather and wind patterns or the distribution of dust and | |
aerosols---have short-term effects on sea surface temperatures in | |
certain regions, but they generally have a minor effect on the | |
longer-term global mean. Previous research shows that as much as 90 | |
percent of the excess heat that has occurred in recent decades due to | |
increasing greenhouse gas emissions is absorbed by the ocean, with much | |
of that heat stored near the surface. The most important factor that | |
helped push sea surface temperatures into record territory in 2023 was | |
the evolving El Niño in the Pacific, according to Willis. He came to | |
that conclusion by analyzing the timing and intensity of sea surface | |
temperature anomalies in several regions and comparing them to the | |
global trend. "We had a big jump in global surface temperature at the | |
beginning of April---exactly when the Pacific temperatures jumped up and | |
also when sea levels in the eastern Pacific started to rise," Willis | |
said. "The heat waves in the Atlantic are important and will have | |
serious effects on marine life and weather in Europe in the coming | |
months. But it's the Pacific that has taken the global mean on a wild | |
ride this year." What happens in the Pacific tends to have a large | |
influence on the global sea surface temperatures partly because of its | |
size. | |
The Pacific represents about half of the world's ocean area. | |
Marine heat waves---defined as periods of persistent anomalously warm | |
ocean temperatures (warmer than 90 percent of the previous observations | |
for a given time of year)---have occurred recently in several areas. | |
One NOAA analysis showed that 48 percent of the global oceans were in | |
the midst of a marine heat wave in August---a larger area than for any | |
other month since the start of the record in 1991. | |
Particularly intense events have warmed the North Atlantic and parts of | |
the Caribbean in recent months. | |
Willis expects the heat in the equatorial Pacific to have more staying | |
power than many of the other marine heat waves simmering around the | |
world. "Many of the marine heat waves we're seeing are ephemeral and | |
'skin' deep, generally lasting on the order of weeks and driven by | |
atmospheric forces," explained Willis. | |
The unusually warm water in the equatorial Pacific associated with the | |
developing El Niño after three consecutive years of La Niña is expected | |
to weaken trade winds in ways that reinforce and amplify the warming of | |
surface waters, fueling the El Niño further. | |
Forecasters from NOAA say that there is a greater than 95 percent chance | |
that El Niño conditions will persist throughout the Northern Hemisphere | |
winter. | |
"What's happening in the Pacific with El Niño will influence global | |
weather patterns and sea surface temperatures well into the winter and | |
possibly even longer," Willis said. | |
To monitor sea surface temperatures, scientists at NOAA and NASA analyze | |
observations from sensors and buoys in the oceans, ships, and several | |
different polar-orbiting and geostationary satellites. Groups of | |
scientists with NOAA's Physical Sciences Laboratory, NOAA's Coral Reef | |
Watch, and NASA's Jet Propulsion Laboratory track marine heat waves and | |
sea surface temperature anomalies closely. | |
You can use NASA's State of the Ocean Tool on Worldview to monitor daily | |
sea surface temperature anomalies. | |
One of the wettest wet seasons in northern Australia transformed large | |
areas of the country's desert landscape over the course of many months | |
in 2023. A string of major rainfall events that dropped 690 millimeters | |
(27 inches) between October 2022 and April 2023 made it the | |
sixth-wettest season on record since 1900--1901. | |
This series of false-color images illustrates the rainfall's months-long | |
effects downstream in the Lake Eyre Basin. Water appears in shades of | |
blue, vegetation is green, and bare land is brown. The images were | |
acquired by the Moderate Resolution Imaging Spectroradiometer (MODIS) on | |
NASA's Terra satellite between January and July 2023. | |
In the January 22 image (left), water was coursing through seasonally | |
dry channels of the Georgina River and Eyre Creek following weeks of | |
heavy rains in northern Queensland. By April 21 (middle), floodwaters | |
had reached further downstream after another intense period of | |
precipitation in March. This scene shows that water had filled in some | |
of the north-northwest trending ridges that are part of a vast fossil | |
landscape of wind-formed dunes, while vegetation had emerged in wet soil | |
upstream. Then by July 26 (right), the riverbed had filled with even | |
more vegetation. | |
The Georgina River and Eyre Creek drain approximately 210,000 square | |
kilometers (81,000 square miles), nearly the area of the United Kingdom. | |
Visible in the lower part of the images, the lake gets refreshed about | |
every three years; when it reaches especially high levels, it may take | |
18 months to 2 years to dry up. Two smaller neighboring lakes flood | |
seasonally. These three lakes and surrounding floodplains support | |
hundreds of thousands of waterbirds and are designated as an Important | |
Bird Area. | |
Seasonal flooding is a regular occurrence in these desert river systems. | |
However, the events of the 2022-2023 rainy season stood out in several | |
ways. They occurred while La Niña conditions were in place over the | |
tropical Pacific Ocean. (The wettest seasons in northern Australia have | |
all occurred during La Niña years, according to Australia's Bureau of | |
Meteorology.) In addition, major rains occurring in succession, as was | |
the case with the January and March events, have the overall effect of | |
prolonging floods. That's because vegetation that grows after the first | |
event slows down the pulse of water that comes through in the next rain | |
event. | |
The high water has affected both local communities and ecosystems. | |
Floods have inundated cattle farms and isolated towns on temporary | |
islands. At the same time, they are a natural feature of the | |
"boom-and-bust" ecology of Channel Country, providing habitat and | |
nutrients that support biodiversity. | |
After three consecutive years of La Niña, spring 2023 saw the return of | |
El Niño---a natural climate phenomenon characterized by the presence of | |
warmer than normal sea surface temperatures (and higher sea levels) in | |
the central and eastern tropical Pacific Ocean. | |
El Niño is associated with the weakening of easterly trade winds and the | |
movement of warm water from the western Pacific toward the western coast | |
of the Americas. The phenomenon can have widespread effects, often | |
bringing cooler, wetter conditions to the U.S. Southwest and drought to | |
countries in the western Pacific, such as Indonesia and Australia. | |
Satellite- and ocean-based measurements of sea surface temperature are | |
one way to detect the arrival of El Niño. Its signature also shows up in | |
satellite measurements of sea surface height, which rises as ocean | |
temperatures warm up. That's because warmer water expands to fill more | |
volume, while cooler water contracts. | |
The map above depicts sea surface height anomalies across the central | |
and eastern Pacific Ocean as observed from June 1--10, 2023. Shades of | |
blue indicate sea levels that were lower than average; normal sea level | |
conditions appear white; and reds indicate areas where the ocean stood | |
higher than normal. | |
Data for the map were acquired by the Sentinel-6 Michael Freilich and | |
Sentinel-3B satellites and processed by scientists at NASA's Jet | |
Propulsion Laboratory (JPL). Note that signals related to seasonal | |
cycles and long-term trends have been removed to highlight sea level | |
anomalies associated with El Niño and other short-term natural | |
phenomena. | |
In a report released on June 8, 2023, the NOAA Climate Prediction Center | |
declared El Niño conditions were present. The report pointed to sea | |
surface temperatures in the Niño 3.4 region of the tropical Pacific | |
(from 170° to 120° West longitude) that in May 2023 were 0.8°C (1.4°F) | |
above the long-term average. | |
Forecasters expected El Niño conditions to gradually strengthen into the | |
2023--2024 Northern Hemisphere winter, by which time they called for a | |
84 percent chance of a moderate strength El Niño developing and a 56 | |
percent chance of a strong El Niño. | |
As of June 2023, however, El Niño was not as far along as past El Niño | |
events by the same time of year, according to Josh Willis, an | |
oceanographer and Sentinel-6 Michael Freilich project scientist at JPL. | |
"It's still a bit too early to say whether this will be a big one," | |
Willis said. "It will probably have some global impacts, but there's | |
still time for this El Niño to underwhelm." | |
As spring turned to summer, phytoplankton came to life in the shallow | |
waters of the North Sea. Sunlight and warm ocean temperatures in June | |
2023 enabled the microscopic plant-like organisms to rapidly multiply | |
and form a dazzling turquoise display visible to satellites. | |
Satellites observed hints of the bloom developing between Scotland and | |
Norway for about two weeks, but the view from above was mostly hidden by | |
clouds. Then, mostly clear skies on the afternoon of June 15, 2023, | |
allowed the Visible Infrared Imaging Radiometer Suite (VIIRS) on the | |
NOAA-20 satellite to acquire this natural-color image of the abundant | |
phytoplankton. | |
Phytoplankton are to the ocean what plants are to land: primary | |
producers, an essential food source for other life, and the main carbon | |
recycler for the marine environment. Diatoms, coccolithophores, algae, | |
and other forms of phytoplankton are floating, plant-like organisms that | |
soak up sunshine, carbon dioxide, and nutrients to create their own | |
energy. | |
This bloom might contain some diatoms---a type of phytoplankton with | |
silica shells and ample chlorophyll that color the surface waters green. | |
The color of the water, however, indicates that coccolithophores are | |
likely abundant. Coccolithophores have calcium carbonate shells that | |
make the water appear milky blue in satellite imagery, and they | |
typically peak in abundance at these latitudes around the summer | |
solstice. | |
Phytoplankton are typically most abundant in the North Sea in late | |
spring and early summer when high levels of nutrients are available in | |
the water. Melting sea ice and increased runoff from European rivers---a | |
product of melting snow and spring rains---carry a heavy load of | |
nutrients out to sea. Intense seasonal winds blowing over the relatively | |
shallow sea also cause a lot of mixing that brings nutrients to the | |
surface. | |
Researchers in Norway studied the patterns and timing of phytoplankton | |
blooms in the North Sea using data from multiple satellite sensors, | |
including VIIRS and NASA's Moderate Resolution Imaging Spectroradiometer | |
(MODIS). They found that between 2000 and 2020, blooms in this region of | |
the North Sea peaked in mid-to-late April. These blooms lasted, on | |
average, about 46 days. They also found that in the 21-year study | |
period, phytoplankton blooms in the region were starting later in the | |
year and lasting slightly longer. The cause of this delay, however, was | |
not immediately clear. | |
The composition of phytoplankton blooms near Norway may be changing over | |
time with warmer sea surface temperatures, the researchers noted, but it | |
is difficult to tell the species composition of blooms without taking | |
physical samples. However, a future NASA Plankton, Aerosol, Cloud, ocean | |
Ecosystem (PACE) satellite mission will enable researchers to infer more | |
information about ocean ecology, such as the species of phytoplankton | |
present in blooms and the rates of phytoplankton growth. | |
Sea ice in the Sea of Okhotsk put on a dazzling display in late May | |
2023, as the winter's ice pack thinned and broke up. The freely drifting | |
ice, subject to wind and currents, formed a series of spirals off the | |
coast of Russia. | |
The Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA's Aqua | |
satellite captured this image on May 28, 2023. More-intact ice is | |
visible on the north side of the P'yagina Peninsula (Poluostrov | |
P'yagina), at the top of the image, with smaller pieces breaking away | |
and drifting to the south and west. A group of islands---too small to | |
see clearly at this scale---off the eastern tip of the land may be | |
responsible for the small eddies in that area. Spirals such as these can | |
form downstream of a stationary object that obstructs fluid flow. | |
The Sea of Okhotsk, which is hemmed in by the Siberian coast and the | |
Kamchatka Peninsula, is the southernmost sea in the Northern Hemisphere | |
that freezes seasonally. An influx of frigid Siberian air, in addition | |
to inflows of freshwater from rivers that lower the salinity and raise | |
the freezing point of the water, create conducive conditions for ice to | |
form during the colder months. | |
During the 2022-2023 winter, the extent of sea ice in the Arctic was | |
below average. The end-of-winter minimum extent, reached on March 6, was | |
the sixth lowest in the satellite record, according to data maintained | |
by the National Snow & Ice Data Center (NSIDC). The NSIDC also noted | |
that seasonal ice decline picked up in the last several days of May, | |
when this image was captured. | |
In a recent study, researchers in Japan found that yearly differences in | |
ice extent are largely governed by regional cold air masses and | |
low-pressure systems, along with large-scale patterns associated with | |
the El Niño/Southern Oscillation (ENSO). Looking at longer term | |
climate-driven trends, they reported that ice extent in the Sea of | |
Okhotsk decreased by about 9 percent per decade between 1979 and 2010. | |
Every fall, millions of people flock to the Shuangtaizi Estuary (also | |
called the Liao River Estuary) in northeastern China to marvel at its | |
brilliant red coastal landscapes. They are drawn by the expanses of rare | |
salt-loving seepweed that thrive in the estuary's alkaline tidal | |
mudflats. | |
The small shrubby plants, Suaeda salsa (also called Suaeda heteroptera), | |
start out greenish-red in the spring but become a bright crimson in fall | |
as seasonal shifts in rainfall and tides expose seepweed to slightly | |
saltier, cooler conditions. This leads to the increased production and | |
accumulation of the red pigment betalain. | |
However, the estuary has changed significantly in recent decades due to | |
coastal development, raising questions about the long-term viability of | |
its colorful seepweed beaches and wetland habitat. The scale of change | |
is apparent in the pair of Landsat images shown above. | |
The image on the left, acquired by the Thematic Mapper on Landsat 5, | |
shows the estuary in 1986; the image on the right, from the Operational | |
Land Imager-2 (OLI-2) on Landsat 9, shows the same area in 2022. Both | |
images were acquired in September, around the time when seepweed reaches | |
its deepest red. The green areas along the river are dominated by | |
Phragmites australis, a type of reed. The yellow areas are rice fields. | |
The photo below shows seepweed (red) in the foreground transitioning | |
into Phragmites australis (green) in the background. Large new | |
aquaculture facilities---and a new port (lower right of image)---have | |
replaced tidal flats where seepweed once thrived. The construction of a | |
dyke and reservoir on the eastern bank of the river has also isolated | |
part of the estuary from tidal waters, making the area unsuitable for | |
seepweed. In wider views of the 1986 and 2022 images, notice how | |
seepweed was further constrained by urbanization and the expansion of | |
aquaculture to the east of the port, as well as the expansion of gas and | |
oil drilling on the western bank of the Shuangtaizi River. | |
Human activities have affected seepweed in other less direct ways in | |
recent decades. The construction of dams, bridges, and canals caused | |
spikes in the amount of sediment carried by the river and deposited on | |
tidal flats downstream. The extra accumulations made it difficult for | |
new seepweed plants to germinate in some areas. Researchers have also | |
found evidence indicating that the construction of boardwalks and the | |
rising number of tourists has harmed seepweed by scaring away waterbirds | |
that feed on crabs, leading to higher numbers of crabs grazing on | |
seepweed in certain areas. | |
Overall, hundreds of square kilometers of wetlands have been lost since | |
the 1980s, according to one analysis that spans three decades of Landsat | |
observations. The amount of land with seepweed dropped by roughly 25 | |
percent during that time, though certain parts of the estuary have seen | |
seepweed areas expand or grow more concentrated. | |
Seepweed-seeking tourists are not the only group affected by the loss of | |
tidal flats and wetlands in this area. The estuary provides habitat for | |
more than 100 water birds, including the critically endangered Siberian | |
Crane (Leucogeranus leucogeranus), the endangered Oriental stork | |
(Ciconia boyciana), and the red-crowned Crane (Grus japonensis). The | |
estuary was named a national nature reserve in 1998 and a Ramsar site in | |
2005. | |
For the past few decades, scientists have been observing natural ocean | |
fertilization events---episodes when plumes of volcanic ash, glacial | |
flour, wildfire soot, and desert dust blow out onto the sea surface and | |
spur massive blooms of phytoplankton. But beyond these extreme events, | |
there is a steady, long-distance rain of dust particles onto the ocean | |
that promotes phytoplankton growth just about all year and in nearly | |
every basin. | |
In a new study published May 5 in the journal Science, a team of | |
researchers from Oregon State University, the University of Maryland | |
Baltimore County, and NASA combined satellite observations with an | |
advanced computer model to home in on how mineral dust from land | |
fertilizes the growth of phytoplankton in the ocean. Phytoplankton are | |
microscopic, plant-like organisms that form the center of the marine | |
food web. | |
Phytoplankton float near the ocean surface primarily subsisting on | |
sunlight and mineral nutrients that well up from the depths or float out | |
to sea in coastal runoff. But mineral-rich desert dust---borne by strong | |
winds and deposited in the ocean---also plays an important role in the | |
health and abundance of phytoplankton. | |
This image, acquired on April 8, 2011, by the Moderate Resolution | |
Imaging Spectroradiometer (MODIS) on NASA's Terra satellite, shows | |
Saharan dust over the Bay of Biscay. A phytoplankton bloom in the bay | |
makes the water appear bright green and blue. Sediment is likely | |
contributing to some of the color, especially in areas closer to the | |
shore. | |
According to the new study, dust deposition onto the ocean supports | |
about 4.5 percent of yearly global export production---a measure of how | |
much of the carbon phytoplankton take up during photosynthesis sinks | |
into the deep ocean. However, this contribution approaches 20 percent to | |
40 percent in some ocean regions at middle and higher latitudes. | |
Phytoplankton play a large role in Earth's climate and carbon cycle. | |
Like land plants, they contain chlorophyll and derive energy from | |
sunlight through photosynthesis. They produce oxygen and sequester a | |
tremendous amount of carbon dioxide in the process, potentially on a | |
scale comparable to rainforests. And they are at the bottom of an | |
ocean-wide food pecking order that ranges from tiny zooplankton to fish | |
to whales. | |
Dust particles can travel thousands of miles before falling into the | |
ocean, where they nourish phytoplankton long distances from the dust | |
source, said study coauthor Lorraine Remer, a research professor at the | |
University of Maryland Baltimore County. "We knew that atmospheric | |
transport of desert dust is part of what makes the ocean 'click,' but we | |
didn't know how to find it," she said. | |
Seasonal allergy sufferers be warned: this story may have you reaching | |
for the antihistamines. Researchers have determined that "slicks" on the | |
surface of the Baltic Sea, visible in satellite images, are made up of | |
pine pollen. | |
Pollen slicks are visible in these images of the Baltic Sea, acquired on | |
May 16, 2018, with the MultiSpectral Instrument (MSI) on the European | |
Space Agency's Sentinel-2A satellite. The images are false-color (bands | |
8A, 3, and 2) and have been enhanced to increase the visibility of the | |
pollen. The patterns are caused by wind-driven currents and waves moving | |
the pollen around on the surface of the water. | |
The composition of slicks in this region was previously unclear. Other | |
types of floating material, such as cyanobacteria and marine debris, | |
have been known to appear in satellite imagery. But by combining | |
experimental results, ground-based observations, and satellite image | |
processing, the researchers could confidently attribute the material in | |
the eddies to pine (Pinus sylvestris) pollen. | |
The impetus for investigating this phenomenon came from a different | |
marine event, said Chuanmin Hu, an ocean optics expert at the University | |
of South Florida who led the research. "This work is inspired by a | |
recent sea snot event in the Marmara Sea that created a huge problem for | |
Türkiye and its coastal regions," he said. Sea snot, which is caused by | |
phytoplankton releasing a gooey substance, coated large swaths of the | |
sea in May 2021 and caught Hu's attention when it was detected by | |
satellites. | |
That led him to wonder if anything comparable was occurring on other | |
large bodies of water nearby. As it turned out, satellite images of the | |
Baltic Sea from that time looked similar to the satellite images of sea | |
snot in the Marmara Sea (to human eyes, at least). But Hu found it | |
strange that there were no reports of disruptive slime from the large, | |
heavily trafficked sea. | |
To identify potential slicks, Hu and colleagues inspected | |
medium-resolution satellite images from sensors such as the Moderate | |
Resolution Imaging Spectroradiometer on NASA's Terra and Aqua | |
satellites. When his team analyzed other satellite data for the spectral | |
signature of the mystery Baltic Sea substance, they realized it was | |
distinct from sea snot and other floating matter. The spectral shape had | |
a characteristically sharp increase between wavelengths of 400 and 500 | |
nanometers. | |
Given the timing of the slicks and the prevalence of pine trees in the | |
nine countries surrounding the sea, they suspected pollen as a possible | |
culprit. Collaborators in Poland had photographs of pollen on the | |
surface of the water, acquired during fieldwork in May 2013 (below). To | |
dig deeper, the U.S. and Polish groups conducted laboratory and field | |
experiments to measure the spectral reflectance of pollen. Indeed, the | |
results matched what was captured by satellites. | |
The researchers then looked back at springtime images of the Baltic Sea | |
from 2000 to 2021 and saw similar slick patterns in 14 of those years. | |
Notably, the pollen's footprint on the sea in the second half of the | |
study period was markedly larger than in the first half. In recent | |
years, slicks often cover some portion of the sea in parts of May and | |
June. | |
This observation aligns with trends toward longer pollen seasons and | |
more pollen production that have been documented in other areas of the | |
world. For example, one recent study found that pollen season in North | |
America starts nearly three weeks sooner and lasts about a week longer | |
than it did in 1990, driven by warming temperatures. In addition, more | |
carbon dioxide in the atmosphere fueling photosynthesis may increase | |
plants' potential to produce more pollen. | |
The profusion of pollen may have larger impacts beyond making people | |
sneeze. Though not well studied, pollen grains can affect aquatic | |
ecosystems by supplying carbon to the sea. Much like leaf litter | |
supports food webs in lakes and streams, pollen grains may be an | |
important source of nutrients for insect larvae, crustaceans, and other | |
invertebrates in coastal Baltic Sea waters. | |
Having cracked the code of distinguishing pollen in satellite imagery, | |
Hu thinks the imagery may lead to several new insights. "If we can track | |
pollen aggregation in different places, this may provide useful data for | |
fisheries studies," he said. Even more, the technique could complement | |
land-based air quality sensors to monitor allergens---all the more | |
relevant as human health impacts from allergies intensify. | |
For several weeks in April 2023, swirls of green and turquoise grew more | |
vibrant in the waters off the Mid-Atlantic coast of the United States. | |
Some of the color is due to an abundance of phytoplankton. Though each | |
of these floating plant-like organisms is microscopic, large groups of | |
them are visible to satellites. | |
A phytoplankton bloom was under way on April 20, 2023, when the Moderate | |
Resolution Imaging Spectroradiometer (MODIS) on NASA's Aqua satellite | |
acquired this image (top). The detailed image below was acquired the | |
same day with the Operational Land Imager-2 (OLI-2) on Landsat 9. | |
Phytoplankton are responsible for nearly half of Earth's primary | |
production. They turn carbon dioxide, sunlight, and nutrients into the | |
food that feeds almost all other life in the sea, from zooplankton to | |
finfish to whales. | |
The type of phytoplankton present in the bloom cannot be definitively | |
identified based on these natural-color images. But assessments of past | |
blooms in the area have turned up a mix of diatoms and coccolithophores. | |
Diatoms, a microscopic form of algae, have silica shells and plenty of | |
chlorophyll that can make the water appear green. Coccolithophores have | |
chalky calcium carbonate plates (coccoliths) that reflect light and can | |
make the water appear bright blue. | |
Color can also come from other sources, such as sediment or colored | |
dissolved organic matter (CDOM) that have mixed in the water. Discharge | |
from the Delaware River delivers sediment and CDOM to the coastal waters | |
in this region. It can also supply nutrients---contained in the runoff | |
from farms and urban and suburban areas---that help to fuel large | |
blooms. | |
Similar blooms have occurred in recent years, in both 2021 and 2022. | |
Those blooms, however, developed their most striking colors almost one | |
month later, around mid-May. | |
In February 2023, Tropical Cyclone Gabrielle churned south across the | |
Coral Sea and passed over the Bellona Plateau---a shallow area 600 | |
kilometers (400 miles) west of Grande Terre, the principal island of New | |
Caledonia. Once a sizable island during the Pleistocene ice ages, the | |
plateau is now submerged under 25-50 meters of water. It hosts reefs | |
that teem with corals, coralline algae, mollusks, foraminifera, and many | |
other types of marine life with calcium carbonate skeletons or shells. | |
Signs of underwater reefs and carbonate platforms are often subtle in | |
satellite imagery. But Gabrielle's winds were fierce enough that the | |
storm left a clear sign of the carbonate ecosystem below the water. The | |
passing storm stirred up enough carbonate sediment to temporarily | |
discolor more than 13,000 square kilometers of water, an area about the | |
size of Puerto Rico. Resuspension events of this size are rare at | |
Bellona Plateau, with this being only the second time it has happened at | |
this scale since the launch of the MODIS sensor on the Terra satellite | |
in 1999. | |
Gabrielle was passing over the area on February 9, 2023, when the | |
Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA's Terra | |
satellite acquired the first image above. After the storm clouds | |
cleared, the satellite observed carbonate sediment that had become | |
suspended in the water (second image) on February 11, 2023. The sediment | |
drifted in ocean currents over the span of a week, with water over the | |
Bellona plateau returning to its normal color by February 20, 2023. | |
The Operational Land Imager-2 (OLI-2) on Landsat 9 captured the detailed | |
images (below) showing sediment swirling in eddies around the plateau on | |
February 12, 2023. The sediment was likely fine-grained carbonate mud | |
with some larger carbonate sand mixed in. It likely formed due to the | |
erosion and accumulation of bits of coral skeletons, coralline algae, | |
and the hard shells of marine organisms that live on the plateau. | |
"With the right water chemistry and amount of light, plateaus like this | |
become major calcium carbonate factories," explained James Acker, an | |
oceanographer with ADNET Systems at the Goddard Earth Sciences Data and | |
Information Services Center (GES DISC). Previous estimates suggest that | |
although shallow coastal areas cover just 7 percent of the ocean's area, | |
they generate about half of the world's marine carbonate sediment. | |
Acker has been using satellites to observe carbonate resuspension events | |
since the launch of the Sea-viewing Wide Field-of-view Sensor (SeaWiFS) | |
in 1997. The goal is to develop better estimates for how much carbonate | |
sediment from shallow areas ends up getting pushed into deeper waters by | |
winds, currents, or other processes. | |
"Deep ocean water dissolves carbonate muds and sands when they sink," | |
explained Acker. "That can help counter the ongoing ocean acidification | |
we're seeing that is caused by the rising levels of carbon dioxide in | |
the atmosphere." In some cases, depending on the chemistry of the water, | |
carbonates dissolve at depths as shallow as 500 meters. In others, they | |
dissolve at depths closer to 4.5 kilometers. | |
Estimates suggest that oceans absorb about 30 percent of the carbon | |
dioxide that humans release into the atmosphere. Some of that carbon | |
gets incorporated into shells and sediment and eventually stored as | |
calcium carbonate in limestone and other sedimentary rocks, making | |
carbonate platforms and marine sedimentary rock an important carbon | |
sink. | |
However, the estimates for how much carbonate these shallow carbonate | |
reefs and plateaus produce, export, and store vary significantly. And | |
there is considerable uncertainty about how the ocean's ability to store | |
carbon will change as the acidity of the ocean changes. More acidic | |
ocean waters make it harder for many marine organisms to build calcium | |
carbonate shells and thrive, so acidification may reduce how much carbon | |
ends up stored in sedimentary rocks. | |
The first step to investigating how climate change might be changing the | |
marine carbon cycle is to simply understand and document how much | |
carbonate sediment is cycling between shallow and deep water, explained | |
Acker. That has led Acker and sedimentologist Jude Wilber to examine | |
decades of satellite data to find out if storm and wind events play an | |
important role in this cycling from shallow to deep. | |
At the American Geophysical Union's Oceans meeting in 2022, Acker and | |
colleagues presented an analysis of a previous resuspension event that | |
followed Tropical Cyclone Wati hitting the Bellona Plateau in 2006. That | |
event occurred after the Category 4 storm stalled over the plateau for | |
two days and battered it with winds that exceeded 209 kilometers (135 | |
miles) per hour. | |
However, the estimates for how much carbonate these shallow carbonate | |
reefs and plateaus produce, export, and store vary significantly. And | |
there is considerable uncertainty about how the ocean's ability to store | |
carbon will change as the acidity of the ocean changes. More acidic | |
ocean waters make it harder for many marine organisms to build calcium | |
carbonate shells and thrive, so acidification may reduce how much carbon | |
ends up stored in sedimentary rocks. | |
The first step to investigating how climate change might be changing the | |
marine carbon cycle is to simply understand and document how much | |
carbonate sediment is cycling between shallow and deep water, explained | |
Acker. That has led Acker and sedimentologist Jude Wilber to examine | |
decades of satellite data to find out if storm and wind events play an | |
important role in this cycling from shallow to deep. | |
At the American Geophysical Union's Oceans meeting in 2022, Acker and | |
colleagues presented an analysis of a previous resuspension event that | |
followed Tropical Cyclone Wati hitting the Bellona Plateau in 2006. That | |
event occurred after the Category 4 storm stalled over the plateau for | |
two days and battered it with winds that exceeded 209 kilometers (135 | |
miles) per hour. | |
"Due to decades of satellite observations---and dramatic examples like | |
this---we can say confidently that tropical cyclones play a very | |
important role," said Acker. "Nothing else exports the volume of | |
sediment into deeper water that they do. The next step is to demonstrate | |
that in a more systematic and rigorous way by analyzing the entire | |
satellite record with machine learning techniques and getting teams out | |
in the field to better understand the dynamics of transport events." | |
Water from recent winter storms is needed by farmers, wildlife, and | |
residents in the region, where precipitation and lake levels in recent | |
years have been among the lowest since the 1970s. However, scientists | |
caution that similar large precipitation events in the past have not | |
been enough to reverse the longer-term depletion of groundwater---a | |
reserve of water that supplements surface sources used for irrigation | |
and other purposes. | |
"The abundant water is expected to recharge the groundwater in the next | |
few months, as we have seen during similar events in 2011 and 2017," | |
said Pang-Wei Liu, a scientist at NASA's Goddard Space Flight Center. | |
"However, if the climate pattern is the same as before---dry and hot in | |
summer followed by low precipitation---and the water demands are still | |
high, then we expect the groundwater drawdown will continue." | |
The chart above, produced with data provided by Liu, shows a downward | |
trend in California's terrestrial water storage (dark blue line) | |
spanning nearly two decades. This includes surface and groundwater, and | |
water held within the soil and in snow. The rest of the lines show why | |
this is happening; amid some variability in all types of stored water, | |
it is groundwater (light blue line) that is sharply decreasing. | |
Liu and colleagues used data from the Gravity Recovery and Climate | |
Experiment (GRACE) and GRACE Follow-On satellite missions to show that | |
the depletion of groundwater in California's Central Valley has been | |
accelerating since 2003. Their results were published December 2022 in | |
Nature Communications. | |
"Even the wettest wet seasons are simply never enough to make up for the | |
far greater amount of groundwater that California extracts each year," | |
said Jay Famiglietti, a global futures professor at Arizona State | |
University and a co-author of the paper. "Hopefully California's | |
Sustainable Groundwater Management Act can slow what will otherwise be a | |
speedy trip to the bottom." | |
Earth's average surface temperature in 2022 tied with 2015 as the fifth | |
warmest on record, according to an analysis by NASA. Continuing the | |
planet's long-term warming trend, global temperatures in 2022 were 0.89 | |
degrees Celsius (1.6 degrees Fahrenheit) above the average for NASA's | |
baseline period (1951--1980), according to scientists at NASA's Goddard | |
Institute for Space Studies (GISS). | |
The past nine years have been the warmest years since modern | |
recordkeeping began in 1880. This means Earth in 2022 was about 1.11°C | |
(2°F) warmer than the late 19th century average. | |
The map above depicts global temperature anomalies in 2022. It does not | |
show absolute temperatures; instead, it shows how much warmer or cooler | |
each region of Earth was compared to the average from 1951 to 1980. The | |
bar chart below shows 2022 in context with temperature anomalies since | |
1880. The values represent surface temperatures averaged over the entire | |
globe for the year. | |
"The reason for the warming trend is that human activities continue to | |
pump enormous amounts of greenhouse gases into the atmosphere, and the | |
long-term planetary impacts will also continue," said Gavin Schmidt, | |
director of GISS, NASA's leading center for climate modeling. | |
Human-driven greenhouse gas emissions have rebounded following a | |
short-lived dip in 2020 due to the COVID-19 pandemic. Recently, NASA | |
scientists, as well as international scientists, determined carbon | |
dioxide emissions were the highest on record in 2022. NASA also | |
identified some super-emitters of methane---another powerful greenhouse | |
gas---using the Earth Surface Mineral Dust Source Investigation (EMIT) | |
instrument that launched to the International Space Station last year. | |
The Arctic region continues to experience the strongest warming | |
trends---close to four times the global average---according to GISS | |
research presented at the 2022 annual meeting of the American | |
Geophysical Union, as well as a separate study. | |
Communities around the world are experiencing impacts scientists see as | |
connected to the warming atmosphere and ocean. Climate change has | |
intensified rainfall and tropical storms, deepened the severity of | |
droughts, and increased the impact of storm surges. Last year brought | |
torrential monsoon rains that devastated Pakistan and a persistent | |
megadrought in the U.S. Southwest. In September, Hurricane Ian became | |
one of the strongest and costliest hurricanes to strike the continental | |
U.S. | |
NASA's global temperature analysis is drawn from data collected by | |
weather stations and Antarctic research stations, as well as instruments | |
mounted on ships and ocean buoys. NASA scientists analyze these | |
measurements to account for uncertainties in the data and to maintain | |
consistent methods for calculating global average surface temperature | |
differences for every year. These ground-based measurements of surface | |
temperature are consistent with satellite data collected since 2002 by | |
the Atmospheric Infrared Sounder on NASA's Aqua satellite and with other | |
estimates. | |
NASA uses the period from 1951--1980 as a baseline to understand how | |
global temperatures change over time. That baseline includes climate | |
patterns such as La Niña and El Niño, as well as unusually hot or cold | |
years due to other factors, ensuring it encompasses natural variations | |
in Earth's temperature. | |
Many factors can affect the average temperature in any given year. For | |
example, 2022 was one of the warmest on record despite a third | |
consecutive year of La Niña conditions in the tropical Pacific Ocean. | |
NASA scientists estimate that La Niña's cooling influence may have | |
lowered global temperatures slightly (about 0.06°C or 0.11°F) from what | |
the average would have been under more typical ocean conditions. | |
A separate, independent analysis by the National Oceanic and Atmospheric | |
Administration (NOAA) concluded that the global surface temperature for | |
2022 was the sixth highest since 1880. NOAA scientists use much of the | |
same raw temperature data in their analysis and have a different | |
baseline period (1901--2000) and methodology. Although rankings for | |
specific years can differ slightly between the records, they are in | |
broad agreement and both reflect ongoing long-term warming. | |