Iron fertilization

From Wikipedia, the free encyclopedia
(Redirected from Iron hypothesis)
An oceanic phytoplankton bloom in the South Atlantic Ocean, off the coast of Argentina covering an area about 300 by 50 miles (500 by 80 km)

Iron fertilization is the intentional introduction of iron-containing compounds (like iron sulfate) to iron-poor areas of the ocean surface to stimulate phytoplankton production. This is intended to enhance biological productivity and/or accelerate carbon dioxide (CO2) sequestration from the atmosphere. Iron is a trace element necessary for photosynthesis in plants. It is highly insoluble in sea water and in a variety of locations is the limiting nutrient for phytoplankton growth. Large algal blooms can be created by supplying iron to iron-deficient ocean waters. These blooms can nourish other organisms.

Ocean iron fertilization is an example of a geoengineering technique.[1] Iron fertilization[2] attempts to encourage phytoplankton growth, which removes carbon from the atmosphere for at least a period of time.[3][4] This technique is controversial because there is limited understanding of its complete effects on the marine ecosystem,[5] including side effects and possibly large deviations from expected behavior. Such effects potentially include release of nitrogen oxides,[6] and disruption of the ocean's nutrient balance.[1] Controversy remains over the effectiveness of atmospheric CO
2
sequestration and ecological effects.[7] Since 1990, 13 major large scale experiments have been carried out to evaluate efficiency and possible consequences of iron fertilization in ocean waters. A study in 2017 determined that the method is unproven; sequestering efficiency is low and sometimes no effect was seen and the amount of iron deposits that is needed to make a small cut in the carbon emissions is in the million tons per year.[8]

Approximately 25 per cent of the ocean surface has ample macronutrients, with little plant biomass (as defined by chlorophyll). The production in these high-nutrient low-chlorophyll (HNLC) waters is primarily limited by micronutrients, especially iron.[9] The cost of distributing iron over large ocean areas is large compared with the expected value of carbon credits.[10] Research in the early 2020s suggested that it could only permanently sequester a small amount of carbon.[11]

Process[edit]

Role of iron in carbon sequestration[edit]

Ocean iron fertilization is an example of a geoengineering technique that involves intentional introduction of iron-rich deposits into oceans, and is aimed to enhance biological productivity of organisms in ocean waters in order to increase carbon dioxide (CO2) uptake from the atmosphere, possibly resulting in mitigating its global warming effects.[12][13][14][15][16] Iron is a trace element in the ocean and its presence is vital for photosynthesis in plants, and in particular phytoplanktons, as it has been shown that iron deficiency can limit ocean productivity and phytoplankton growth.[17] For this reason, the "iron hypothesis" was put forward by Martin in late 1980s where he suggested that changes in iron supply in iron-deficient seawater can bloom plankton growth and have a significant effect on the concentration of atmospheric carbon dioxide by altering rates of carbon sequestration.[18][19] In fact, fertilization is an important process that occurs naturally in the ocean waters. For instance, upwellings of ocean currents can bring nutrient-rich sediments to the surface.[20] Another example is through transfer of iron-rich minerals, dust, and volcanic ash over long distances by rivers, glaciers, or wind.[21][22] Moreover, it has been suggested that whales can transfer iron-rich ocean dust to the surface, where planktons can take it up to grow. It has been shown that reduction in the number of sperm whales in the Southern Ocean has resulted in a 200,000 tonnes/yr decrease in the atmospheric carbon uptake, possibly due to limited phytoplankton growth.[23]

Carbon sequestration by phytoplankton[edit]

An oceanic phytoplankton bloom in the North Sea off the coast of eastern Scotland

Phytoplankton is photosynthetic: it needs sunlight and nutrients to grow, and takes up carbon dioxide in the process. Plankton can take up and sequester atmospheric carbon through generating calcium or silicon-carbonate skeletons. When these organisms die they sink to the ocean floor where their carbonate skeletons can form a major component of the carbon-rich deep sea precipitation, thousands of meters below plankton blooms, known as marine snow.[24][25][26] Nonetheless, based on the definition, carbon is only considered "sequestered" when it is deposited in the ocean floor where it can be retained for millions of years. However, most of the carbon-rich biomass generated from plankton is generally consumed by other organisms (small fish, zooplankton, etc.)[27][28] and substantial part of rest of the deposits that sink beneath plankton blooms may be re-dissolved in the water and gets transferred to the surface where it eventually returns to the atmosphere, thus, nullifying any possible intended effects regarding carbon sequestration.[29][30][31][32][33] Nevertheless, supporters of the idea of iron fertilization believe that carbon sequestration should be re-defined over much shorter time frames and claim that since the carbon is suspended in the deep ocean it is effectively isolated from the atmosphere for hundreds of years, and thus, carbon can be effectively sequestered.[34]

Efficiency and concerns[edit]

Assuming the ideal conditions, the upper estimates for possible effects of iron fertilization in slowing down global warming is about 0.3W/m2 of averaged negative forcing which can offset roughly 15–20% of the current anthropogenic CO2 emissions.[35][36][37] However, although this approach could be looked upon as an easy option to lower the concentration of CO2 in the atmosphere, ocean iron fertilization is still quite controversial and highly debated due to possible negative consequences on marine ecosystems.[30][38][39][40] Research on this area has suggested that fertilization through deposition of large quantities of iron-rich dust into the ocean floor can significantly disrupt the ocean's nutrient balance and cause major complications in the food chain for other marine organisms.[41][42][43][44][45][46][47]

Methods[edit]

There are two ways of performing artificial iron fertilization: ship based direct into the ocean and atmospheric deployment.[48]

Ship based deployment[edit]

Trials of ocean fertilization using iron sulphate added directly to the surface water from ships are described in detail in the experiment section below.

Atmospheric sourcing[edit]

Iron-rich dust rising into the atmosphere is a primary source of ocean iron fertilization.[49] For example, wind blown dust from the Sahara desert fertilizes the Atlantic Ocean[50] and the Amazon rainforest.[51] The naturally occurring iron oxide in atmospheric dust reacts with hydrogen chloride from sea spray to produce iron chloride, which degrades methane and other greenhouse gases, brightens clouds and eventually falls with the rain in low concentration across a wide area of the globe.[48] Unlike ship based deployment, no trials have been performed of increasing the natural level of atmospheric iron. Expanding this atmospheric source of iron could complement ship-based deployment.

One proposal is to boost the atmospheric iron level with iron salt aerosol.[48] Iron(III) chloride added to the troposphere could increase natural cooling effects including methane removal, cloud brightening and ocean fertilization, helping to prevent or reverse global warming.[48]

Experiments[edit]

Martin hypothesized that increasing phytoplankton photosynthesis could slow or even reverse global warming by sequestering CO
2
in the sea. He died shortly thereafter during preparations for Ironex I,[52] a proof of concept research voyage, which was successfully carried out near the Galapagos Islands in 1993 by his colleagues at Moss Landing Marine Laboratories.[53] Thereafter 12 international ocean studies examined the phenomenon:

  • Ironex II, 1995[54]
  • SOIREE (Southern Ocean Iron Release Experiment), 1999[55]
  • EisenEx (Iron Experiment), 2000[56]
  • SEEDS (Subarctic Pacific Iron Experiment for Ecosystem Dynamics Study), 2001[57]
  • SOFeX (Southern Ocean Iron Experiments - North & South), 2002[58][59]
  • SERIES (Subarctic Ecosystem Response to Iron Enrichment Study), 2002[60]
  • SEEDS-II, 2004[61]
  • EIFEX (European Iron Fertilization Experiment),[62] A successful experiment conducted in 2004 in a mesoscale ocean eddy in the South Atlantic resulted in a bloom of diatoms, a large portion of which died and sank to the ocean floor when fertilization ended. In contrast to the LOHAFEX experiment, also conducted in a mesoscale eddy, the ocean in the selected area contained enough dissolved silicon for the diatoms to flourish.[63][64][65]
  • CROZEX (CROZet natural iron bloom and Export experiment), 2005[66]
  • A pilot project planned by Planktos, a U.S. company, was cancelled in 2008 for lack of funding.[67] The company blamed environmental organizations for the failure.[68][69]
  • LOHAFEX (Indian and German Iron Fertilization Experiment), 2009[70][71][72] Despite widespread opposition to LOHAFEX, on 26 January 2009 the German Federal Ministry of Education and Research (BMBF) gave clearance. The experiment was carried out in waters low in silicic acid, an essential nutrient for diatom growth. This affected sequestration efficacy.[73] A 900 square kilometers (350 sq mi) portion of the southwest Atlantic was fertilized with iron sulfate. A large phytoplankton bloom was triggered. In the absence of diatoms, a relatively small amount of carbon was sequestered, because other phytoplankton are vulnerable to predation by zooplankton and do not sink rapidly upon death.[73] These poor sequestration results led to suggestions that fertilization is not an effective carbon mitigation strategy in general. However, prior ocean fertilization experiments in high silica locations revealed much higher carbon sequestration rates because of diatom growth. LOHAFEX confirmed sequestration potential depends strongly upon appropriate siting.[73]
  • Haida Salmon Restoration Corporation (HSRC), 2012 - funded by the Old Massett Haida band and managed by Russ George - dumped 100 tonnes of iron sulphate into the Pacific into an eddy 200 nautical miles (370 km) west of the islands of Haida Gwaii. This resulted in increased algae growth over 10,000 square miles (26,000 km2). Critics alleged George's actions violated the United Nations Convention on Biological Diversity (CBD) and the London convention on the dumping of wastes at sea which prohibited such geoengineering experiments.[74][75] On 15 July 2014, the resulting scientific data was made available to the public.[76]

John Martin, director of the Moss Landing Marine Laboratories, hypothesized that the low levels of phytoplankton in these regions are due to a lack of iron. In 1989 he tested this hypothesis (known as the Iron Hypothesis) by an experiment using samples of clean water from Antarctica.[77] Iron was added to some of these samples. After several days the phytoplankton in the samples with iron fertilization grew much more than in the untreated samples. This led Martin to speculate that increased iron concentrations in the oceans could partly explain past ice ages.[78]

IRONEX I[edit]

This experiment was followed by a larger field experiment (IRONEX I) where 445 kg of iron was added to a patch of ocean near the Galápagos Islands. The levels of phytoplankton increased three times in the experimental area.[79] The success of this experiment and others led to proposals to use this technique to remove carbon dioxide from the atmosphere.[80]

EisenEx[edit]

In 2000 and 2004, iron sulfate was discharged from the EisenEx. 10 to 20 percent of the resulting algal bloom died and sank to the sea floor.[81]

Commercial projects[edit]

Planktos was a US company that abandoned its plans to conduct 6 iron fertilization cruises from 2007 to 2009, each of which would have dissolved up to 100 tons of iron over a 10,000 km2 area of ocean. Their ship Weatherbird II was refused entry to the port of Las Palmas in the Canary Islands where it was to take on provisions and scientific equipment.[82]

In 2007 commercial companies such as Climos and GreenSea Ventures and the Australian-based Ocean Nourishment Corporation, planned to engage in fertilization projects. These companies invited green co-sponsors to finance their activities in return for provision of carbon credits to offset investors' CO2 emissions.[83]

LOHAFEX[edit]

LOHAFEX was an experiment initiated by the German Federal Ministry of Research and carried out by the German Alfred Wegener Institute (AWI) in 2009 to study fertilization in the South Atlantic. India was also involved.[84]

As part of the experiment, the German research vessel Polarstern deposited 6 tons of ferrous sulfate in an area of 300 square kilometers. It was expected that the material would distribute through the upper 15 metres (49 ft) of water and trigger an algal bloom. A significant part of the carbon dioxide dissolved in sea water would then be bound by the emerging bloom and sink to the ocean floor.

The Federal Environment Ministry called for the experiment to halt, partly because environmentalists predicted damage to marine plants. Others predicted long-term effects that would not be detectable during short-term observation[85][unreliable source?] or that this would encourage large-scale ecosystem manipulation.[86][unreliable source?][87]

2012[edit]

A 2012 study deposited iron fertilizer in an eddy near Antarctica. The resulting algal bloom sent a significant amount of carbon into the deep ocean, where it was expected to remain for centuries to millennia. The eddy was chosen because it offered a largely self-contained test system.[88]

As of day 24, nutrients, including nitrogen, phosphorus and silicic acid that diatoms use to construct their shells, declined. Dissolved inorganic carbon concentrations were reduced below equilibrium with atmospheric CO
2
. In surface water, particulate organic matter (algal remains) including silica and chlorophyll increased.[88]

After day 24, however, the particulate matter fell to between 100 metres (330 ft) to the ocean floor. Each iron atom converted at least 13,000 carbon atoms into algae. At least half of the organic matter sank below, 1,000 metres (3,300 ft).[88]

Haida Gwaii project[edit]

In July 2012, the Haida Salmon Restoration Corporation dispersed 100 short tons (91 t) of iron sulphate dust into the Pacific Ocean several hundred miles west of the islands of Haida Gwaii. The Old Massett Village Council financed the action as a salmon enhancement project with $2.5 million in village funds.[89] The concept was that the formerly iron-deficient waters would produce more phytoplankton that would in turn serve as a "pasture" to feed salmon. Then-CEO Russ George hoped to sell carbon offsets to recover the costs. The project was accompanied by charges of unscientific procedures and recklessness. George contended that 100 tons was negligible compared to what naturally enters the ocean.[90]

Some environmentalists called the dumping a "blatant violation" of two international moratoria.[89][91] George said that the Old Massett Village Council and its lawyers approved the effort and at least seven Canadian agencies were aware of it.[90]

According to George, the 2013 salmon runs increased from 50 million to 226 million fish.[92] However, many experts contend that changes in fishery stocks since 2012 cannot necessarily be attributed to the 2012 iron fertilization; many factors contribute to predictive models, and most data from the experiment are considered to be of questionable scientific value.[93]

On 15 July 2014, the data gathered during the project were made publicly available under the ODbL license.[94]

Experiments with iron-coated rice husks in Arabian Sea[edit]

In 2022, a UK/India research team plans to place iron-coated rice husks in the Arabian Sea, to test whether increasing time at the surface can stimulate a bloom using less iron. The iron will be confined within a plastic bag reaching from the surface several kilometers down to the sea bottom.[95][96] The Centre for Climate Repair at the University of Cambridge, along with India's Institute of Maritime Studies assessed the impact of iron seeding in another experiment. They spread iron-coated rice husks across an area of the Arabian Sea. Iron is a limiting nutrient in many ocean waters. They hoped that the iron would fertilize algae, which would bolster the bottom of the marine food chain and sequester carbon as uneaten algae died. The experiment was demolished by a storm, leaving inconclusive results.[97]

Science[edit]

The maximum possible result from iron fertilization, assuming the most favourable conditions and disregarding practical considerations, is 0.29 W/m2 of globally averaged negative forcing,[98] offsetting 1/6 of current levels of anthropogenic CO
2
emissions. These benefits have been called into question by research suggesting that fertilization with iron may deplete other essential nutrients in the seawater causing reduced phytoplankton growth elsewhere — in other words, that iron concentrations limit growth more locally than they do on a global scale.[99][100]

Ocean fertilization occurs naturally when upwellings bring nutrient-rich water to the surface, as occurs when ocean currents meet an ocean bank or a sea mount. This form of fertilization produces the world's largest marine habitats. Fertilization can also occur when weather carries wind blown dust long distances over the ocean, or iron-rich minerals are carried into the ocean by glaciers,[101] rivers and icebergs.[102]

Role of iron[edit]

About 70% of the world's surface is covered in oceans. The part of these where light can penetrate is inhabited by algae (and other marine life). In some oceans, algae growth and reproduction is limited by the amount of iron. Iron is a vital micronutrient for phytoplankton growth and photosynthesis that has historically been delivered to the pelagic sea by dust storms from arid lands. This Aeolian dust contains 3–5% iron and its deposition has fallen nearly 25% in recent decades.[103]

The Redfield ratio describes the relative atomic concentrations of critical nutrients in plankton biomass and is conventionally written "106 C: 16 N: 1 P." This expresses the fact that one atom of phosphorus and 16 of nitrogen are required to "fix" 106 carbon atoms (or 106 molecules of CO
2
). Research expanded this constant to "106 C: 16 N: 1 P: .001 Fe" signifying that in iron deficient conditions each atom of iron can fix 106,000 atoms of carbon,[104] or on a mass basis, each kilogram of iron can fix 83,000 kg of carbon dioxide. The 2004 EIFEX experiment reported a carbon dioxide to iron export ratio of nearly 3000 to 1. The atomic ratio would be approximately: "3000 C: 58,000 N: 3,600 P: 1 Fe".[105]

Therefore, small amounts of iron (measured by mass parts per trillion) in HNLC zones can trigger large phytoplankton blooms on the order of 100,000 kilograms of plankton per kilogram of iron. The size of the iron particles is critical. Particles of 0.5–1 micrometer or less seem to be ideal both in terms of sink rate and bioavailability. Particles this small are easier for cyanobacteria and other phytoplankton to incorporate and the churning of surface waters keeps them in the euphotic or sunlit biologically active depths without sinking for long periods. One way to add small amounts of iron to HNLC zones would be Atmospheric Methane Removal.

Atmospheric deposition is an important iron source. Satellite images and data (such as PODLER, MODIS, MSIR)[106][107][108] combined with back-trajectory analyses identified natural sources of iron–containing dust. Iron-bearing dusts erode from soil and are transported by wind. Although most dust sources are situated in the Northern Hemisphere, the largest dust sources are located in northern and southern Africa, North America, central Asia and Australia.[109]

Heterogeneous chemical reactions in the atmosphere modify the speciation of iron in dust and may affect the bioavailability of deposited iron. The soluble form of iron is much higher in aerosols than in soil (~0.5%).[109][110][111] Several photo-chemical interactions with dissolved organic acids increase iron solubility in aerosols.[112][113] Among these, photochemical reduction of oxalate-bound Fe(III) from iron-containing minerals is important. The organic ligand forms a surface complex with the Fe (III) metal center of an iron-containing mineral (such as hematite or goethite). On exposure to solar radiation the complex is converted to an excited energy state in which the ligand, acting as bridge and an electron donor, supplies an electron to Fe(III) producing soluble Fe(II).[114][115][116] Consistent with this, studies documented a distinct diel variation in the concentrations of Fe (II) and Fe(III) in which daytime Fe(II) concentrations exceed those of Fe(III).[117][118][119][120]

Volcanic ash as an iron source[edit]

Volcanic ash has a significant role in supplying the world's oceans with iron.[121] Volcanic ash is composed of glass shards, pyrogenic minerals, lithic particles and other forms of ash that release nutrients at different rates depending on structure and the type of reaction caused by contact with water.[122]

Increases of biogenic opal in the sediment record are associated with increased iron accumulation over the last million years.[123] In August 2008, an eruption in the Aleutian Islands deposited ash in the nutrient-limited Northeast Pacific. This ash and iron deposition resulted in one of the largest phytoplankton blooms observed in the subarctic.[124]

Carbon sequestration[edit]

Previous instances of biological carbon sequestration triggered major climatic changes, lowering the temperature of the planet, such as the Azolla event. Plankton that generate calcium or silicon carbonate skeletons, such as diatoms, coccolithophores and foraminifera, account for most direct sequestration.[citation needed] When these organisms die their carbonate skeletons sink relatively quickly and form a major component of the carbon-rich deep sea precipitation known as marine snow. Marine snow also includes fish fecal pellets and other organic detritus, and steadily falls thousands of meters below active plankton blooms.[125]

Of the carbon-rich biomass generated by plankton blooms, half (or more) is generally consumed by grazing organisms (zooplankton, krill, small fish, etc.) but 20 to 30% sinks below 200 meters (660 ft) into the colder water strata below the thermocline.[126] Much of this fixed carbon continues into the abyss, but a substantial percentage is redissolved and remineralized. At this depth, however, this carbon is now suspended in deep currents and effectively isolated from the atmosphere for centuries.

Analysis and quantification[edit]

Evaluation of the biological effects and verification of the amount of carbon actually sequestered by any particular bloom involves a variety of measurements, combining ship-borne and remote sampling, submarine filtration traps, tracking buoy spectroscopy and satellite telemetry. Unpredictable ocean currents can remove experimental iron patches from the pelagic zone, invalidating the experiment.

The potential of fertilization to tackle global warming is illustrated by the following figures. If phytoplankton converted all the nitrate and phosphate present in the surface mixed layer across the entire Antarctic circumpolar current into organic carbon, the resulting carbon dioxide deficit could be compensated by uptake from the atmosphere amounting to about 0.8 to 1.4 gigatonnes of carbon per year.[127] This quantity is comparable in magnitude to annual anthropogenic fossil fuels combustion of approximately 6 gigatonnes. The Antarctic circumpolar current region is one of several in which iron fertilization could be conducted—the Galapagos islands area another potentially suitable location.

Dimethyl sulfide and clouds[edit]

Schematic diagram of the CLAW hypothesis (Charlson et al., 1987)[128]

Some species of plankton produce dimethyl sulfide (DMS), a portion of which enters the atmosphere where it is oxidized by hydroxyl radicals (OH), atomic chlorine (Cl) and bromine monoxide (BrO) to form sulfate particles, and potentially increase cloud cover. This may increase the albedo of the planet and so cause cooling—this proposed mechanism is central to the CLAW hypothesis.[128] This is one of the examples used by James Lovelock to illustrate his Gaia hypothesis.[129]

During SOFeX, DMS concentrations increased by a factor of four inside the fertilized patch. Widescale iron fertilization of the Southern Ocean could lead to significant sulfur-triggered cooling in addition to that due to the CO
2
uptake and that due to the ocean's albedo increase, however the amount of cooling by this particular effect is very uncertain.[130]

Financial opportunities[edit]

Beginning with the Kyoto Protocol, several countries and the European Union established carbon offset markets which trade certified emission reduction credits (CERs) and other types of carbon credit instruments. In 2007 CERs sold for approximately €15–20/ton COe
2
.[131] Iron fertilization is relatively inexpensive compared to scrubbing, direct injection and other industrial approaches, and can theoretically sequester for less than €5/ton CO
2
, creating a substantial return.[132] In August, 2010, Russia established a minimum price of €10/ton for offsets to reduce uncertainty for offset providers.[133] Scientists have reported a 6–12% decline in global plankton production since 1980.[103][134] A full-scale plankton restoration program could regenerate approximately 3–5 billion tons of sequestration capacity worth €50-100 billion in carbon offset value. However, a 2013 study indicates the cost versus benefits of iron fertilization puts it behind carbon capture and storage and carbon taxes.[135]

Debate[edit]

While ocean iron fertilization could represent a potent means to slow global warming, there is a current debate surrounding the efficacy of this strategy and the potential adverse effects of this.

Precautionary principle[edit]

The precautionary principle is a proposed guideline regarding environmental conservation. According to an article published in 2021, the precautionary principle (PP) is a concept that states, "The PP means that when it is scientifically plausible that human activities may lead to morally unacceptable harm, actions shall be taken to avoid or diminish that harm: uncertainty should not be an excuse to delay action."[136] Based on this principle, and because there is little data quantifying the effects of iron fertilization, it is the responsibility of leaders in this field to avoid the harmful effects of this procedure. This school of thought is one argument against using iron fertilization on a wide scale, at least until more data is available to analyze the repercussions of this.

Ecological issues[edit]

A "red tide" off the coast of La Jolla, San Diego, California.

Critics are concerned that fertilization will create harmful algal blooms (HAB) as many toxic algae are often favored when iron is deposited into the marine ecosystem. A 2010 study of iron fertilization in an oceanic high-nitrate, low-chlorophyll environment, however, found that fertilized Pseudo-nitzschia diatom spp., which are generally nontoxic in the open ocean, began producing toxic levels of domoic acid. Even short-lived blooms containing such toxins could have detrimental effects on marine food webs.[137] Most species of phytoplankton are harmless or beneficial, given that they constitute the base of the marine food chain. Fertilization increases phytoplankton only in the open oceans (far from shore) where iron deficiency is substantial. Most coastal waters are replete with iron and adding more has no useful effect.[138] Further, it has been shown that there are often higher mineralization rates with iron fertilization, leading to a turn over in the plankton masses that are produced. This results in no beneficial effects and actually causes an increase in CO2.[139]

Finally, a 2010 study showed that iron enrichment stimulates toxic diatom production in high-nitrate, low-chlorophyll areas[140] which, the authors argue, raises "serious concerns over the net benefit and sustainability of large-scale iron fertilizations". Nitrogen released by cetaceans and iron chelate are a significant benefit to the marine food chain in addition to sequestering carbon for long periods of time.[141]

Ocean acidification[edit]

A 2009 study tested the potential of iron fertilization to reduce both atmospheric CO2 and ocean acidity using a global ocean carbon model. The study found that, "Our simulations show that ocean iron fertilization, even in the extreme scenario by depleting global surface macronutrient concentration to zero at all time, has a minor effect on mitigating CO2-induced acidification at the surface ocean."[142] Unfortunately, the impact on ocean acidification would likely not change due to the low effects that iron fertilization has on CO2 levels.[139]

History[edit]

Consideration of iron's importance to phytoplankton growth and photosynthesis dates to the 1930s when Dr Thomas John Hart, a British marine biologist based on the RRS Discovery II in the Southern Ocean speculated - in "On the phytoplankton of the South-West Atlantic and Bellingshausen Sea, 1929-31" - that great "desolate zones" (areas apparently rich in nutrients, but lacking in phytoplankton activity or other sea life) might be iron-deficient.[53] Hart returned to this issue in a 1942 paper entitled "Phytoplankton periodicity in Antarctic surface waters", but little other scientific discussion was recorded until the 1980s, when oceanographer John Martin of the Moss Landing Marine Laboratories renewed controversy on the topic with his marine water nutrient analyses. His studies supported Hart's hypothesis. These "desolate" regions came to be called "high-nutrient, low-chlorophyll regions" (HNLC).[53]

John Gribbin was the first scientist to publicly suggest that climate change could be reduced by adding large amounts of soluble iron to the oceans.[143] Martin's 1988 quip four months later at Woods Hole Oceanographic Institution, "Give me a half a tanker of iron and I will give you an ice age,"[53][144][145] drove a decade of research.

The findings suggested that iron deficiency was limiting ocean productivity and offered an approach to mitigating climate change as well. Perhaps the most dramatic support for Martin's hypothesis came with the 1991 eruption of Mount Pinatubo in the Philippines. Environmental scientist Andrew Watson analyzed global data from that eruption and calculated that it deposited approximately 40,000 tons of iron dust into oceans worldwide. This single fertilization event preceded an easily observed global decline in atmospheric CO
2
and a parallel pulsed increase in oxygen levels.[146]

The parties to the London Dumping Convention adopted a non-binding resolution in 2008 on fertilization (labeled LC-LP.1(2008)). The resolution states that ocean fertilization activities, other than legitimate scientific research, "should be considered as contrary to the aims of the Convention and Protocol and do not currently qualify for any exemption from the definition of dumping".[147] An Assessment Framework for Scientific Research Involving Ocean Fertilization, regulating the dumping of wastes at sea (labeled LC-LP.2(2010)) was adopted by the Contracting Parties to the Convention in October 2010 (LC 32/LP 5).[148]

Multiple ocean labs, scientists and businesses have explored fertilization. Beginning in 1993, thirteen research teams completed ocean trials demonstrating that phytoplankton blooms can be stimulated by iron augmentation.[139] Controversy remains over the effectiveness of atmospheric CO
2
sequestration and ecological effects.[7] Ocean trials of ocean iron fertilization took place in 2009 in the South Atlantic by project LOHAFEX, and in July 2012 in the North Pacific off the coast of British Columbia, Canada, by the Haida Salmon Restoration Corporation (HSRC).[149]

See also[edit]

References[edit]

  1. ^ a b Traufetter, Gerald (January 2, 2009). "Cold Carbon Sink: Slowing Global Warming with Antarctic Iron". Spiegel Online. Archived from the original on April 13, 2017. Retrieved May 9, 2010.
  2. ^ Jin, X.; Gruber, N.; Frenzel1, H.; Doney, S.C.; McWilliams, J.C. (2008). "The impact on atmospheric CO
    2
    of iron fertilization induced changes in the ocean's biological pump"
    . Biogeosciences. 5 (2): 385–406. Bibcode:2008BGeo....5..385J. doi:10.5194/bg-5-385-2008. hdl:1912/2129. Archived from the original on October 16, 2009. Retrieved May 9, 2010.
    {{cite journal}}: CS1 maint: numeric names: authors list (link)
  3. ^ Monastersky, Richard (September 30, 1995). "Iron versus the Greenhouse - Oceanographers cautiously explore a global warming therapy". Science News. Archived from the original on August 20, 2010. Retrieved May 9, 2010.
  4. ^ Monastersky, Richard (September 30, 1995). "Iron versus the Greenhouse: Oceanographers cautiously explore a global warming therapy". Science News. 148 (14): 220–222. doi:10.2307/4018225. JSTOR 4018225.
  5. ^ "WWF condemns Planktos Inc. iron-seeding plan in the Galapagos". Geoengineering Monitor. June 27, 2007. Archived from the original on January 15, 2016. Retrieved August 21, 2015.
  6. ^ Fogarty, David (December 15, 2008). "Scientists urge caution in ocean-CO
    2
    capture schemes"
    . Alertnet.org. Archived from the original on August 3, 2009. Retrieved May 9, 2010.
  7. ^ a b Buesseler, K.O.; Doney, SC; Karl, DM; Boyd, PW; Caldeira, K; Chai, F; Coale, KH; De Baar, HJ; Falkowski, PG; Johnson, KS; Lampitt, R. S.; Michaels, A. F.; Naqvi, S. W. A.; Smetacek, V.; Takeda, S.; Watson, A. J.; et al. (2008). "Environment: Ocean Iron Fertilization—Moving Forward in a Sea of Uncertainty" (PDF). Science. 319 (5860): 162. doi:10.1126/science.1154305. PMID 18187642. S2CID 206511143. Archived (PDF) from the original on 2012-03-05. Retrieved 2009-03-27.
  8. ^ Tollefson, Jeff (2017-05-23). "Iron-dumping ocean experiment sparks controversy". Nature. 545 (7655): 393–394. Bibcode:2017Natur.545..393T. doi:10.1038/545393a. ISSN 0028-0836. PMID 28541342. S2CID 4464713.
  9. ^ Lampitt, R. S.; Achterberg, E. P.; Anderson, T. R.; Hughes, J. A.; Iglesias-Rodriguez, M. D.; Kelly-Gerreyn, B. A.; Lucas, M.; Popova, E. E.; Sanders, R. (2008-11-13). "Ocean fertilization: a potential means of geoengineering?". Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. 366 (1882): 3919–3945. Bibcode:2008RSPTA.366.3919L. doi:10.1098/rsta.2008.0139. ISSN 1364-503X. PMID 18757282.
  10. ^ Harrison, Daniel P. (2013). "A method for estimating the cost to sequester carbon dioxide by delivering iron to the ocean". International Journal of Global Warming. 5 (3): 231. doi:10.1504/ijgw.2013.055360.
  11. ^ "Cloud spraying and hurricane slaying: how ocean geoengineering became the frontier of the climate crisis". The Guardian. 2021-06-23. Archived from the original on 23 June 2021. Retrieved 2021-06-23.
  12. ^ Traufetter, Gerald (2009-01-02). "Cold Carbon Sink: Slowing Global Warming with Antarctic Iron". Spiegel Online. Retrieved 2018-11-18.
  13. ^ Jin, X.; Gruber, N.; Frenzel, H.; Doney, S. C.; McWilliams, J. C. (2008-03-18). "The impact on atmospheric CO2 of iron fertilization induced changes in the ocean's biological pump". Biogeosciences. 5 (2): 385–406. Bibcode:2008BGeo....5..385J. doi:10.5194/bg-5-385-2008. hdl:1912/2129. ISSN 1726-4170.
  14. ^ Monastersky, Richard (September 30, 1995). "Iron versus the Greenhouse: Oceanographers cautiously explore a global warming therapy". Science News. 148: 220. doi:10.2307/4018225. JSTOR 4018225.
  15. ^ Martínez-García, Alfredo; Sigman, Daniel M.; Ren, Haojia; Anderson, Robert F.; Straub, Marietta; Hodell, David A.; Jaccard, Samuel L.; Eglinton, Timothy I.; Haug, Gerald H. (2014-03-21). "Iron Fertilization of the Subantarctic Ocean During the Last Ice Age". Science. 343 (6177): 1347–1350. Bibcode:2014Sci...343.1347M. doi:10.1126/science.1246848. ISSN 0036-8075. PMID 24653031. S2CID 206552831.
  16. ^ Pasquier, Benoît; Holzer, Mark (2018-08-16). "Iron fertilization efficiency and the number of past and future regenerations of iron in the ocean". Biogeosciences Discussions. 15 (23): 7177–7203. Bibcode:2018AGUFMGC23G1277P. doi:10.5194/bg-2018-379. ISSN 1726-4170. S2CID 133851021.
  17. ^ Boyd, Philip W.; Watson, Andrew J.; Law, Cliff S.; Abraham, Edward R.; Trull, Thomas; Murdoch, Rob; Bakker, Dorothee C. E.; Bowie, Andrew R.; Buesseler, K. O. (October 2000). "A mesoscale phytoplankton bloom in the polar Southern Ocean stimulated by iron fertilization". Nature. 407 (6805): 695–702. Bibcode:2000Natur.407..695B. doi:10.1038/35037500. ISSN 0028-0836. PMID 11048709. S2CID 4368261.
  18. ^ Boyd, P. W.; Jickells, T.; Law, C. S.; Blain, S.; Boyle, E. A.; Buesseler, K. O.; Coale, K. H.; Cullen, J. J.; Baar, H. J. W. de (2007-02-02). "Mesoscale Iron Enrichment Experiments 1993-2005: Synthesis and Future Directions". Science. 315 (5812): 612–617. Bibcode:2007Sci...315..612B. doi:10.1126/science.1131669. ISSN 0036-8075. PMID 17272712. S2CID 2476669.
  19. ^ "John Martin". earthobservatory.nasa.gov. 2001-07-10. Retrieved 2018-11-19.
  20. ^ Ian, Salter; Ralf, Schiebel; Patrizia, Ziveri; Aurore, Movellan; S., Lampitt, Richard; A., Wolff, George (2015-02-23). "Carbonate counter pump stimulated by natural iron fertilization in the Southern Ocean". epic.awi.de (in German). Retrieved 2018-11-19.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  21. ^ "Text taken from a draft review, "The role of grazing in structuring Southern Ocean pelagic ecosystems and biogeochemical cycles"" (PDF). Tyndall Centre. 2007-11-29. Archived (PDF) from the original on 2007-11-29. Retrieved 2018-11-19.
  22. ^ Hodson, Andy; Nowak, Aga; Sabacka, Marie; Jungblut, Anne; Navarro, Francisco; Pearce, David; Ávila-Jiménez, María Luisa; Convey, Peter; Vieira, Gonçalo (2017-02-15). "Climatically sensitive transfer of iron to maritime Antarctic ecosystems by surface runoff". Nature Communications. 8: 14499. Bibcode:2017NatCo...814499H. doi:10.1038/ncomms14499. ISSN 2041-1723. PMC 5316877. PMID 28198359.
  23. ^ Lavery, Trish J.; Roudnew, Ben; Gill, Peter; Seymour, Justin; Seuront, Laurent; Johnson, Genevieve; Mitchell, James G.; Smetacek, Victor (2010-11-22). "Iron defecation by sperm whales stimulates carbon export in the Southern Ocean". Proceedings of the Royal Society of London B: Biological Sciences. 277 (1699): 3527–3531. doi:10.1098/rspb.2010.0863. ISSN 0962-8452. PMC 2982231. PMID 20554546.
  24. ^ J., Brooks; K., Shamberger; B., Roark, E.; K., Miller; A., Baco-Taylor (February 2016). "Seawater Carbonate Chemistry of Deep-sea Coral Beds off the Northwestern Hawaiian Islands". American Geophysical Union, Ocean Sciences Meeting. 2016: AH23A–03. Bibcode:2016AGUOSAH23A..03B.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  25. ^ Laurenceau-Cornec, Emmanuel C.; Trull, Thomas W.; Davies, Diana M.; Rocha, Christina L. De La; Blain, Stéphane (2015-02-03). "Phytoplankton morphology controls on marine snow sinking velocity". Marine Ecology Progress Series. 520: 35–56. Bibcode:2015MEPS..520...35L. doi:10.3354/meps11116. ISSN 0171-8630.
  26. ^ Prairie, Jennifer C.; Ziervogel, Kai; Camassa, Roberto; McLaughlin, Richard M.; White, Brian L.; Dewald, Carolin; Arnosti, Carol (2015-10-20). "Delayed settling of marine snow: Effects of density gradient and particle properties and implications for carbon cycling". Marine Chemistry. 175: 28–38. Bibcode:2015MarCh.175...28P. doi:10.1016/j.marchem.2015.04.006. ISSN 0304-4203.
  27. ^ Steinberg, Deborah K.; Landry, Michael R. (2017-01-03). "Zooplankton and the Ocean Carbon Cycle". Annual Review of Marine Science. 9 (1): 413–444. Bibcode:2017ARMS....9..413S. doi:10.1146/annurev-marine-010814-015924. ISSN 1941-1405. PMID 27814033.
  28. ^ Cavan, Emma L.; Henson, Stephanie A.; Belcher, Anna; Sanders, Richard (2017-01-12). "Role of zooplankton in determining the efficiency of the biological carbon pump". Biogeosciences. 14 (1): 177–186. Bibcode:2017BGeo...14..177C. doi:10.5194/bg-14-177-2017. ISSN 1726-4189.
  29. ^ Robinson, J.; Popova, E. E.; Yool, A.; Srokosz, M.; Lampitt, R. S.; Blundell, J. R. (2014-04-11). "How deep is deep enough? Ocean iron fertilization and carbon sequestration in the Southern Ocean" (PDF). Geophysical Research Letters. 41 (7): 2489–2495. Bibcode:2014GeoRL..41.2489R. doi:10.1002/2013gl058799. ISSN 0094-8276. S2CID 53389222.
  30. ^ a b Hauck, Judith; Köhler, Peter; Wolf-Gladrow, Dieter; Völker, Christoph (2016). "Iron fertilisation and century-scale effects of open ocean dissolution of olivine in a simulated CO 2 removal experiment". Environmental Research Letters. 11 (2): 024007. Bibcode:2016ERL....11b4007H. doi:10.1088/1748-9326/11/2/024007. ISSN 1748-9326.
  31. ^ Tremblay, Luc; Caparros, Jocelyne; Leblanc, Karine; Obernosterer, Ingrid (2014). "Origin and fate of particulate and dissolved organic matter in a naturally iron-fertilized region of the Southern Ocean". Biogeosciences. 12 (2): 607. Bibcode:2015BGeo...12..607T. doi:10.5194/bg-12-607-2015. S2CID 9333764.
  32. ^ Arrhenius, Gustaf; Mojzsis, Stephen; Atkinson, A.; Fielding, S.; Venables, H. J.; Waluda, C. M.; Achterberg, E. P. (2016-10-10). "Zooplankton Gut Passage Mobilizes Lithogenic Iron for Ocean Productivity" (PDF). Current Biology. 26 (19): 2667–2673. doi:10.1016/j.cub.2016.07.058. ISSN 0960-9822. PMID 27641768. S2CID 3970146.
  33. ^ Vinay, Subhas, Adam (2017). Chemical Controls on the Dissolution Kinetics of Calcite in Seawater (phd). California Institute of Technology. doi:10.7907/z93x84p3.{{cite thesis}}: CS1 maint: multiple names: authors list (link)
  34. ^ Jackson, R. B.; Canadell, J. G.; Fuss, S.; Milne, J.; Nakicenovic, N.; Tavoni, M. (2017). "Focus on negative emissions". Environmental Research Letters. 12 (11): 110201. Bibcode:2017ERL....12k0201J. doi:10.1088/1748-9326/aa94ff. ISSN 1748-9326.
  35. ^ Lenton, T. M.; Vaughan, N. E. (2009-01-28). "The radiative forcing potential of different climate geoengineering options" (PDF). Atmospheric Chemistry and Physics Discussions. 9 (1): 2559–2608. doi:10.5194/acpd-9-2559-2009. ISSN 1680-7375.
  36. ^ US 20180217119, "Process and method for the enhancement of sequestering atmospheric carbon through ocean iron fertilization, and method for calculating net carbon capture from said process and method", issued 2016-07-28 
  37. ^ Gattuso, J.-P.; Magnan, A.; Billé, R.; Cheung, W. W. L.; Howes, E. L.; Joos, F.; Allemand, D.; Bopp, L.; Cooley, S. R. (2015-07-03). "Contrasting futures for ocean and society from different anthropogenic CO2 emissions scenarios" (PDF). Science. 349 (6243): aac4722. doi:10.1126/science.aac4722. ISSN 0036-8075. PMID 26138982. S2CID 206639157.
  38. ^ El-Jendoubi, Hamdi; Vázquez, Saúl; Calatayud, Ángeles; Vavpetič, Primož; Vogel-Mikuš, Katarina; Pelicon, Primoz; Abadía, Javier; Abadía, Anunciación; Morales, Fermín (2014). "The effects of foliar fertilization with iron sulfate in chlorotic leaves are limited to the treated area. A study with peach trees (Prunus persica L. Batsch) grown in the field and sugar beet (Beta vulgaris L.) grown in hydroponics". Frontiers in Plant Science. 5: 2. doi:10.3389/fpls.2014.00002. ISSN 1664-462X. PMC 3895801. PMID 24478782.
  39. ^ Yoon, Joo-Eun; Yoo, Kyu-Cheul; Macdonald, Alison M.; Yoon, Ho-Il; Park, Ki-Tae; Yang, Eun Jin; Kim, Hyun-Cheol; Lee, Jae Il; Lee, Min Kyung (2018-10-05). "Reviews and syntheses: Ocean iron fertilization experiments – past, present, and future looking to a future Korean Iron Fertilization Experiment in the Southern Ocean (KIFES) project". Biogeosciences. 15 (19): 5847–5889. Bibcode:2018BGeo...15.5847Y. doi:10.5194/bg-15-5847-2018. ISSN 1726-4189.
  40. ^ Gim, Byeong-Mo; Hong, Seongjin; Lee, Jung-Suk; Kim, Nam-Hyun; Kwon, Eun-Mi; Gil, Joon-Woo; Lim, Hyun-Hwa; Jeon, Eui-Chan; Khim, Jong Seong (2018-10-01). "Potential ecotoxicological effects of elevated bicarbonate ion concentrations on marine organisms". Environmental Pollution. 241: 194–199. doi:10.1016/j.envpol.2018.05.057. ISSN 0269-7491. PMID 29807279. S2CID 44160652.
  41. ^ Traufetter, Gerald (2009-01-02). "Cold Carbon Sink: Slowing Global Warming with Antarctic Iron". Spiegel Online. Retrieved 2018-11-19.
  42. ^ "Reuters AlertNet - RPT-FEATURE-Scientists urge caution in ocean-CO2 capture schemes". 2009-08-03. Archived from the original on 2009-08-03. Retrieved 2018-11-19.
  43. ^ "WWF condemns Planktos Inc. iron-seeding plan in the Galapagos". Geoengineering Monitor. 2007-06-27. Retrieved 2018-11-19.
  44. ^ Glibert, Patricia; Anderson, Donald; Gentien, Patrick; Granéli, Edna; Sellner, Kevin (June 2005). "The Global, Complex Phenomena of Harmful Algal Blooms | Oceanography". Oceanography. 18 (2): 136–147. doi:10.5670/oceanog.2005.49. hdl:1912/2790. Retrieved 2018-11-19.
  45. ^ Moore, J.Keith; Doney, Scott C; Glover, David M; Fung, Inez Y (2001). "Iron cycling and nutrient-limitation patterns in surface waters of the World Ocean". Deep Sea Research Part II: Topical Studies in Oceanography. 49 (1–3): 463–507. Bibcode:2001DSRII..49..463M. CiteSeerX 10.1.1.210.1108. doi:10.1016/S0967-0645(01)00109-6. ISSN 0967-0645.
  46. ^ Trick, Charles G.; Bill, Brian D.; Cochlan, William P.; Wells, Mark L.; Trainer, Vera L.; Pickell, Lisa D. (2010-03-30). "Iron enrichment stimulates toxic diatom production in high-nitrate, low-chlorophyll areas". Proceedings of the National Academy of Sciences. 107 (13): 5887–5892. Bibcode:2010PNAS..107.5887T. doi:10.1073/pnas.0910579107. ISSN 0027-8424. PMC 2851856. PMID 20231473.
  47. ^ Fripiat, F.; Elskens, M.; Trull, T. W.; Blain, S.; Cavagna, A. -J.; Fernandez, C.; Fonseca-Batista, D.; Planchon, F.; Raimbault, P. (November 2015). "Significant mixed layer nitrification in a natural iron-fertilized bloom of the Southern Ocean". Global Biogeochemical Cycles. 29 (11): 1929–1943. Bibcode:2015GBioC..29.1929F. doi:10.1002/2014gb005051. ISSN 0886-6236.
  48. ^ a b c d Franz Dietrich Oeste; Renaud de Richter; Tingzhen Ming; Sylvain Caillol (13 January 2017). "Climate engineering by mimicking natural dust climate control: the iron salt aerosol method". Earth System Dynamics. 8 (1): 1–54. Bibcode:2017ESD.....8....1O. doi:10.5194/esd-8-1-2017.
  49. ^ Gary Shaffer; Fabrice Lambert (27 February 2018). "In and out of glacial extremes by way of dust−climate feedbacks". Proceedings of the National Academy of Sciences of the United States of America. 115 (9): 2026–2031. Bibcode:2018PNAS..115.2026S. doi:10.1073/pnas.1708174115. PMC 5834668. PMID 29440407.
  50. ^ Tim Radford (July 16, 2014). "Desert Dust Feeds Deep Ocean Life". Scientific American. Archived from the original on March 31, 2019. Retrieved March 30, 2019.
  51. ^ Richard Lovett (August 9, 2010). "African dust keeps Amazon blooming". Nature. Archived from the original on March 3, 2019. Retrieved March 30, 2019.
  52. ^ "Ironex (Iron Experiment) I". Archived from the original on 2004-04-08.
  53. ^ a b c d Weier, John (2001-07-10). "John Martin (1935-1993)". On the Shoulders of Giants. NASA Earth Observatory. Archived from the original on 2012-10-04. Retrieved 2012-08-27.
  54. ^ Ironex II Archived 2005-12-25 at the Wayback Machine, 1995
  55. ^ SOIREE (Southern Ocean Iron Release Experiment) Archived 2008-10-24 at the Wayback Machine, 1999
  56. ^ EisenEx (Iron Experiment) Archived 2007-09-27 at the Wayback Machine, 2000
  57. ^ SEEDS (Subarctic Pacific Iron Experiment for Ecosystem Dynamics Study) Archived 2006-02-14 at the Wayback Machine, 2001
  58. ^ SOFeX (Southern Ocean Iron Experiments - North & South) Archived 2018-08-27 at the Wayback Machine, 2002
  59. ^ "Effects of Ocean Fertilization with Iron To Remove Carbon Dioxide from the Atmosphere Reported" (Press release). Archived from the original on 2006-12-31. Retrieved 2007-03-31.
  60. ^ SERIES (Subarctic Ecosystem Response to Iron Enrichment Study) Archived 2007-09-29 at the Wayback Machine, 2002
  61. ^ SEEDS-II Archived 2006-05-16 at the Wayback Machine, 2004
  62. ^ EIFEX (European Iron Fertilization Experiment) Archived 2006-09-25 at the Wayback Machine, 2004
  63. ^ Smetacek, Victor; Christine Klaas; Volker H. Strass; Philipp Assmy; Marina Montresor; Boris Cisewski; Nicolas Savoye; Adrian Webb; Francesco d’Ovidio; Jesús M. Arrieta; Ulrich Bathmann; Richard Bellerby; Gry Mine Berg; Peter Croot; Santiago Gonzalez; Joachim Henjes; Gerhard J. Herndl; Linn J. Hoffmann; Harry Leach; Martin Losch; Matthew M. Mills; Craig Neill; Ilka Peeken; Rüdiger Röttgers; Oliver Sachs; et al. (18 July 2012). "Deep carbon export from a Southern Ocean iron-fertilized diatom bloom". Nature. 487 (7407): 313–319. Bibcode:2012Natur.487..313S. doi:10.1038/nature11229. PMID 22810695. S2CID 4304972.
  64. ^ David Biello (July 18, 2012). "Controversial Spewed Iron Experiment Succeeds as Carbon Sink". Scientific American. Archived from the original on August 8, 2012. Retrieved July 19, 2012.
  65. ^ Field test stashes climate-warming carbon in deep ocean; Strategically dumping metal puts greenhouse gas away, possibly for good Archived 2012-07-26 at the Wayback Machine July 18th, 2012 Science News
  66. ^ CROZEX (CROZet natural iron bloom and Export experiment) Archived 2011-06-13 at the Wayback Machine, 2005
  67. ^ Scientists to fight global warming with plankton Archived 2007-09-27 at the Wayback Machine ecoearth.info 2007-05-21
  68. ^ Planktos kills iron fertilization project due to environmental opposition Archived 2009-07-13 at the Portuguese Web Archive mongabay.com 2008-02-19
  69. ^ Venture to Use Sea to Fight Warming Runs Out of Cash Archived 2017-01-19 at the Wayback Machine New York Times 2008-02-14
  70. ^ "LOHAFEX: An Indo-German iron fertilization experiment". Eurekalert.org. Archived from the original on 2012-04-10. Retrieved 2012-04-17.
  71. ^ Bhattacharya, Amit (2009-01-06). "Tossing iron powder into ocean to fight global warming". The Times Of India. Archived from the original on 2009-01-11. Retrieved 2009-01-13.
  72. ^ "'Climate fix' ship sets sail with plan to dump iron - environment - 09 January 2009". New Scientist. Archived from the original on 2012-10-19. Retrieved 2012-04-17.
  73. ^ a b c "Lohafex provides new insights on plankton ecology". Eurekalert.org. Archived from the original on 2012-09-19. Retrieved 2012-04-17.
  74. ^ Martin Lukacs (October 15, 2012). "World's biggest geoengineering experiment 'violates' UN rules: Controversial US businessman's iron fertilisation off west coast of Canada contravenes two UN conventions". The Guardian. Archived from the original on October 12, 2013. Retrieved October 16, 2012.
  75. ^ Henry Fountain (October 18, 2012). "A Rogue Climate Experiment Outrages Scientists". The New York Times. Archived from the original on October 18, 2012. Retrieved October 19, 2012.
  76. ^ "Home: OCB Ocean Fertilization". Woods Hole Oceanographic Institution. Archived from the original on 2015-03-12.
  77. ^ "The Iron Hypothesis". homepages.ed.ac.uk. Archived from the original on 5 November 2020. Retrieved 2020-07-26.
  78. ^ John Weier (2001-07-10). "The Iron Hypothesis". John Martin (1935–1993). Archived from the original on 4 October 2012. Retrieved 27 August 2012.
  79. ^ John Weier (2001-07-10). "Following the vision". John Martin (1935–1993). Archived from the original on 15 October 2012. Retrieved 27 August 2012.
  80. ^ Richtel, Matt (2007-05-01). "Recruiting Plankton to Fight Global Warming". The New York Times. Archived from the original on 8 April 2017. Retrieved 2017-06-03.
  81. ^ Bakker, Dorothee C. E.; Bozec, Yann; Nightingale, Philip D.; Goldson, Laura; Messias, Marie-José; de Baar, Hein J. W.; Liddicoat, Malcolm; Skjelvan, Ingunn; Strass, Volker; Watson, Andrew J. (2005-06-01). "Iron and mixing affect biological carbon uptake in SOIREE and EisenEx, two Southern Ocean iron fertilisation experiments". Deep Sea Research Part I: Oceanographic Research Papers. 52 (6): 1001–1019. Bibcode:2005DSRI...52.1001B. doi:10.1016/j.dsr.2004.11.015. ISSN 0967-0637.
  82. ^ "Planktos Shareholder Update". Business wire. 2007-12-19. Archived from the original on 2008-06-25.
  83. ^ Salleh, Anna (2007). "Urea 'climate solution' may backfire". ABC Science Online. Archived from the original on 2008-11-18.
  84. ^ "Alfred-Wegener-Institut für Polar- und Meeresforschung (AWI) ANT-XXV/3". Archived from the original on 8 October 2012. Retrieved 9 August 2012.
  85. ^ "LOHAFEX über sich selbst". 2009-01-14. Archived from the original on 15 February 2009. Retrieved 9 August 2012.
  86. ^ Helfrich, Silke (2009-01-12). "Polarsternreise zur Manipulation der Erde". CommonsBlog. Archived from the original on 14 October 2016. Retrieved 2017-06-03.
  87. ^ John, Paull (2009). "Geo-Engineering in the Southern Ocean". Journal of Bio-Dynamics Tasmania (93). Archived from the original on 6 November 2018. Retrieved 2017-06-03.
  88. ^ a b c "Could Fertilizing the Oceans Reduce Global Warming?". Live Science. Archived from the original on 27 November 2016. Retrieved 2017-06-02.
  89. ^ a b Lucas, Martin (15 October 2012). "World's biggest geoengineering experiment 'violates' UN rules". The Guardian. Archived from the original on 4 February 2014. Retrieved 17 October 2012.
  90. ^ a b Fountain, Henry (18 October 2012). "A Rogue Climate Experiment Outrages Scientists". The New York Times. Archived from the original on 18 October 2012. Retrieved 18 October 2012.
  91. ^ "Environment Canada launches probe into massive iron sulfate dump off Haida Gwaii coast". APTN National News. 16 October 2012. Archived from the original on 2 February 2014. Retrieved 17 October 2012.
  92. ^ Zubrin, Robert (2014-04-22). "The Pacific's Salmon Are Back – Thank Human Ingenuity". Nationalreview.com. Archived from the original on 23 April 2014. Retrieved 2014-04-23.
  93. ^ "Controversial Haida Gwaii ocean fertilizing experiment pitched to Chile". CBC News. Archived from the original on 2 November 2017. Retrieved 2017-11-09.
  94. ^ "WELCOME TO THE OCB OCEAN FERTILIZATION WEBSITE". Archived from the original on 12 March 2015. Retrieved 22 October 2014.
  95. ^ "Scarcity of iron, puffing hot iron in the Arabian Sea and geoengineering climate change".
  96. ^ Chinni, Venkatesh; Singh, Sunil Kumar (2022). "Dissolved iron cycling in the Arabian Sea and sub-tropical gyre region of the Indian Ocean". Geochimica et Cosmochimica Acta. 317: 325–348. Bibcode:2022GeCoA.317..325C. doi:10.1016/j.gca.2021.10.026. S2CID 240313905.
  97. ^ Voosen, Paul (16 December 2022). "Ocean geoengineering scheme aces its first field test". www.science.org. Retrieved 2022-12-19.
  98. ^ Lenton, T. M., Vaughan, N. E. (2009). "The radiative forcing potential of different climate geoengineering options" (PDF). Atmos. Chem. Phys. Discuss. 9: 2559–2608. doi:10.5194/acpd-9-2559-2009. Archived (PDF) from the original on 2021-11-23. Retrieved 2019-10-01.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  99. ^ "Seeding iron in the Pacific may not pull carbon from air as thought". Phys.org. March 3, 2016. Archived from the original on August 4, 2017. Retrieved August 4, 2017.
  100. ^ K. M. Costa, J. F. McManus, R. F. Anderson, H. Ren, D. M. Sigman, G. Winckler, M. Q. Fleisher, F. Marcantonio, A. C. Ravelo (2016). "No iron fertilization in the equatorial Pacific Ocean during the last ice age". Nature. 529 (7587): 519–522. Bibcode:2016Natur.529..519C. doi:10.1038/nature16453. PMID 26819045. S2CID 205247036.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  101. ^ Smetacek, Victor. "Ocean fertilization" (PDF). Archived from the original (PDF) on 29 November 2007.
  102. ^ Traufetter, Gerald (2008-12-18). "Cold Carbon Sink: Slowing Global Warming with Antarctic Iron - Spiegel Online". Spiegel Online. Spiegel.de. Archived from the original on 2017-04-13. Retrieved 2012-04-17.
  103. ^ a b Ocean Plant Life Slows Down and Absorbs less Carbon Archived 2007-08-02 at the Wayback Machine NASA Earth Observatory
  104. ^ Sunda, W. G.; S. A. Huntsman (1995). "Iron uptake and growth limitation in oceanic and coastal phytoplankton". Mar. Chem. 50 (1–4): 189–206. Bibcode:1995MarCh..50..189S. doi:10.1016/0304-4203(95)00035-P. Archived from the original on 2020-02-06. Retrieved 2020-02-06.
  105. ^ de Baar H . J. W., Gerringa, L. J. A., Laan, P., Timmermans, K. R (2008). "Efficiency of carbon removal per added iron in ocean iron fertilization". Mar Ecol Prog Ser. 364: 269–282. Bibcode:2008MEPS..364..269D. doi:10.3354/meps07548.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  106. ^ Barnaba, F.; G. P. Gobbi (2004). "Aerosol seasonal variability over the Mediterranean region and relative impact of maritime, continental and Saharan dust particles over the basin from MODIS data in the year 2001". Atmos. Chem. Phys. Discuss. 4 (4): 4285–4337. doi:10.5194/acpd-4-4285-2004. Archived from the original on 2019-09-23. Retrieved 2019-09-23.
  107. ^ Ginoux, P.; O. Torres (2003). "Empirical TOMS index for dust aerosol: Applications to model validation and source characterization". J. Geophys. Res. 108 (D17): 4534. Bibcode:2003JGRD..108.4534G. CiteSeerX 10.1.1.143.9618. doi:10.1029/2003jd003470.
  108. ^ Kaufman, Y., I. Koren, L. A. Remer, D. Tanre, P. Ginoux, and S. Fan (2005). "Dust transport and deposition observed from the Terra-MODIS spacecraft over the Atlantic Ocean". J. Geophys. Res. 110 (D10): D10S12. Bibcode:2005JGRD..11010S12K. CiteSeerX 10.1.1.143.7305. doi:10.1029/2003jd004436.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  109. ^ a b Mahowald, Natalie M.; et al. (2005). "Atmospheric global dust cycle and iron inputs to the ocean" (PDF). Global Biogeochemical Cycles. 19 (4): GB4025. Bibcode:2005GBioC..19.4025M. doi:10.1029/2004GB002402. hdl:11511/68526. Archived (PDF) from the original on 2020-02-06. Retrieved 2020-02-06.
  110. ^ Fung, I. Y., S. K. Meyn, I. Tegen, S. C. Doney, J. G. John, and J. K. B. Bishop (2000). "Iron supply and demand in the upper ocean". Global Biogeochem. Cycles. 14 (2): 697–700. Bibcode:2000GBioC..14..697F. doi:10.1029/2000gb900001.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  111. ^ Hand, J. L., N. Mahowald, Y. Chen, R. Siefert, C. Luo, A. Subramaniam, and I. Fung (2004). "Estimates of soluble iron from observations and a global mineral aerosol model: Biogeochemical implications". J. Geophys. Res. 109 (D17): D17205. Bibcode:2004JGRD..10917205H. doi:10.1029/2004jd004574.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  112. ^ Siefert, Ronald L.; et al. (1994). "Iron photochemistry of aqueous suspensions of ambient aerosol with added organic acids". Geochimica et Cosmochimica Acta. 58 (15): 3271–3279. Bibcode:1994GeCoA..58.3271S. doi:10.1016/0016-7037(94)90055-8.
  113. ^ Yuegang Zuo; Juerg Hoigne (1992). "Formation of hydrogen peroxide and depletion of oxalic acid in atmospheric water by photolysis of iron (iii)-oxalato complexes". Environmental Science & Technology. 26 (5): 1014–1022. Bibcode:1992EnST...26.1014Z. doi:10.1021/es00029a022.
  114. ^ Siffert, Christophe; Barbara Sulzberger (1991). "Light-induced dissolution of hematite in the presence of oxalate. A case study". Langmuir. 7 (8): 1627–1634. doi:10.1021/la00056a014.
  115. ^ Banwart, Steven, Simon Davies, and Werner Stumm (1989). "The role of oxalate in accelerating the reductive dissolution of hematite (α-Fe 2 O 3) by ascorbate". Colloids and Surfaces. 39 (2): 303–309. doi:10.1016/0166-6622(89)80281-1.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  116. ^ Sulzberger, Barbara; Hansulrich Laubscher (1995). "Reactivity of various types of iron (III)(hydr) oxides towards light-induced dissolution". Marine Chemistry. 50.1 (1–4): 103–115. Bibcode:1995MarCh..50..103S. doi:10.1016/0304-4203(95)00030-u.
  117. ^ Kieber, R., Skrabal, S., Smith, B., and Willey (2005). "Organic complexation of Fe (II) and its impact on the redox cycling of iron in rain". Environmental Science & Technology. 39 (6): 1576–1583. Bibcode:2005EnST...39.1576K. doi:10.1021/es040439h. PMID 15819212.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  118. ^ Kieber, R. J., Peake, B., Willey, J. D., and Jacobs, B (2001b). "Iron speciation and hydrogen peroxide concentrations in New Zealand rainwater". Atmospheric Environment. 35 (34): 6041–6048. Bibcode:2001AtmEn..35.6041K. doi:10.1016/s1352-2310(01)00199-6.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  119. ^ Kieber, R. J., Willey, J. D., and Avery, G. B. (2003). "Temporal variability of rainwater iron speciation at the Bermuda Atlantic Time Series Station". Journal of Geophysical Research: Oceans. 108 (C8): 1978–2012. Bibcode:2003JGRC..108.3277K. doi:10.1029/2001jc001031.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  120. ^ Willey, J. D., Kieber, R. J., Seaton, P. J., and Miller, C. (2008). "Rainwater as a source of Fe (II)-stabilizing ligands to seawater". Limnology and Oceanography. 53 (4): 1678–1684. Bibcode:2008LimOc..53.1678W. doi:10.4319/lo.2008.53.4.1678.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  121. ^ Duggen S.; et al. (2007). "Subduction zone volcanic ash can fertilize the surface ocean and stimulate phytoplankton growth: Evidence from biogeochemical experiments and satellite data" (PDF). Geophysical Research Letters. 34 (1): L01612. Bibcode:2007GeoRL..34.1612D. doi:10.1029/2006gl027522. S2CID 44686878. Archived (PDF) from the original on 2021-08-10. Retrieved 2020-02-06.
  122. ^ Olgun N.; et al. (2011). "Surface Ocean Iron Fertilization: The role of airborne volcanic ash from subduction zone and hot spot volcanoes and related iron fluxes into the Pacific Ocean" (PDF). Global Biogeochemical Cycles. 25 (4): n/a. Bibcode:2011GBioC..25.4001O. doi:10.1029/2009gb003761. Archived (PDF) from the original on 2021-08-10. Retrieved 2020-02-06.
  123. ^ Murray Richard W., Leinen Margaret, Knowlton Christopher W. (2012). "Links between iron input and opal deposition in the Pleistocene equatorial Pacific Ocean". Nature Geoscience. 5 (4): 270–274. Bibcode:2012NatGe...5..270M. doi:10.1038/ngeo1422.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  124. ^ Hemme R.; et al. (2010). "Volcanic ash fuels anomalous plankton bloom in subarctic northeast Pacific". Geophysical Research Letters. 37 (19): n/a. Bibcode:2010GeoRL..3719604H. doi:10.1029/2010gl044629.
  125. ^ "Video of extremely heavy amounts of "marine snow" in the Charlie-Gibbs Fracture Zone in the Mid-Atlantic Ridge. Michael Vecchione, NOAA Fisheries Systematics Lab". Archived from the original on 2006-09-08.
  126. ^ "The Biological Productivity of the Ocean | Learn Science at Scitable". Archived from the original on 2021-05-03. Retrieved 2021-04-22.
  127. ^ Schiermeier Q (January 2003). "Climate change: The oresmen". Nature. 421 (6919): 109–10. Bibcode:2003Natur.421..109S. doi:10.1038/421109a. PMID 12520274. S2CID 4384209.
  128. ^ a b Charlson, R. J.; Lovelock, J. E.; Andreae, M. O.; Warren, S. G. (1987). "Oceanic phytoplankton, atmospheric sulphur, cloud albedo and climate". Nature. 326 (6114): 655–661. Bibcode:1987Natur.326..655C. doi:10.1038/326655a0. S2CID 4321239.
  129. ^ Lovelock, J.E. (2000) [1979]. Gaia: A New Look at Life on Earth (3rd ed.). Oxford University Press. ISBN 978-0-19-286218-1.
  130. ^ Wingenter, Oliver W.; Karl B. Haase; Peter Strutton; Gernot Friederich; Simone Meinardi; Donald R. Blake; F. Sherwood Rowland (2004-06-08). "Changing concentrations of CO, CH4, C5H8, CH3Br, CH3I, and dimethyl sulfide during the Southern Ocean Iron Enrichment Experiments". Proceedings of the National Academy of Sciences. 101 (23): 8537–8541. Bibcode:2004PNAS..101.8537W. doi:10.1073/pnas.0402744101. PMC 423229. PMID 15173582.
  131. ^ "February 2007 Carbon Update" (PDF). CO2 Australia Limited. Archived from the original (PDF) on 30 August 2007.
  132. ^ "Greening Up". Scienceline.[permanent dead link]
  133. ^ "Russia sets minimum carbon offset price Envirotech Online". www.envirotech-online.com. Archived from the original on 2011-07-10. Retrieved 2010-11-19.
  134. ^ Plankton Found to Absorb Less Carbon Dioxide Archived 2006-09-06 at the Wayback Machine BBC, 8/30/06
  135. ^ Iron fertilisation sunk as an ocean carbon storage solution Archived 2013-04-13 at the Wayback Machine University of Sydney press release 12 December 2012 and Harrison, D P IJGW (2013)
  136. ^ Drivdal, Laura; van der Sluijs, Jeroen P (August 2021). "Pollinator conservation requires a stronger and broader application of the precautionary principle". Current Opinion in Insect Science. 46: 95–105. doi:10.1016/j.cois.2021.04.005. ISSN 2214-5745. PMID 33930597. S2CID 233470544.
  137. ^ Tricka, Charles G., Brian D. Bill, William P. Cochlan, Mark L. Wells, Vera L. Trainer, and Lisa D. Pickell (2010). "Iron enrichment stimulates toxic diatom production in high-nitrate, low-chlorophyll areas". PNAS. 107 (13): 5887–5892. Bibcode:2010PNAS..107.5887T. doi:10.1073/pnas.0910579107. PMC 2851856. PMID 20231473.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  138. ^ JK, Moore; SC, Doney; DM, Glover; IY, Fung (2002-01-19). "Iron cycling and nutrient-limitation patterns in surface waters of the world ocean". Deep-Sea Research Part II: Topical Studies in Oceanography. 49 (1–3): 463–507. Bibcode:2001DSRII..49..463M. doi:10.1016/S0967-0645(01)00109-6. ISSN 0967-0645. Archived from the original on 2017-10-01. Retrieved 2017-09-30.
  139. ^ a b c Boyd, P.W.; Jickells, T; Law, CS; Blain, S; Boyle, EA; Buesseler, KO; Coale, KH; Cullen, JJ; De Baar, HJ; Follows, M; Harvey, M.; Lancelot, C.; Levasseur, M.; Owens, N. P. J.; Pollard, R.; Rivkin, R. B.; Sarmiento, J.; Schoemann, V.; Smetacek, V.; Takeda, S.; Tsuda, A.; Turner, S.; Watson, A. J.; et al. (2007). "Mesoscale Iron Enrichment Experiments 1993-2005: Synthesis and Future Directions" (PDF). Science. 315 (5812): 612–7. Bibcode:2007Sci...315..612B. doi:10.1126/science.1131669. PMID 17272712. S2CID 2476669. Archived (PDF) from the original on 2012-03-05. Retrieved 2009-03-27.
  140. ^ Trick, Charles G.; Brian D. Bill; William P. Cochlan; Mark L. Wells; Vera L. Trainer; Lisa D. Pickell (2010). "Iron enrichment stimulates toxic diatom production in high-nitrate, low-chlorophyll areas". Proceedings of the National Academy of Sciences of the United States of America. 107 (13): 5887–5892. Bibcode:2010PNAS..107.5887T. doi:10.1073/pnas.0910579107. PMC 2851856. PMID 20231473.
  141. ^ Brown, Joshua E. (12 Oct 2010). "Whale poop pumps up ocean health". Science Daily. Archived from the original on 2 September 2019. Retrieved 18 August 2014.
  142. ^ Cao, Long; Caldeira, Ken (2010). "Can ocean iron fertilization mitigate ocean acidification?". Climatic Change. 99 (1–2): 303–311. Bibcode:2010ClCh...99..303C. doi:10.1007/s10584-010-9799-4. S2CID 153613458.
  143. ^ Gribbin, John (1988). "Any old iron?". Nature. 331 (6157): 570. Bibcode:1988Natur.331..570G. doi:10.1038/331570c0. PMID 3340209. S2CID 4281828.
  144. ^ "Fertilizing the Ocean with Iron". Woods Hole Oceanographic Institution. Archived from the original on 2021-01-18. Retrieved 2021-02-25.
  145. ^ "Ocean Iron Fertilization – Why Dump Iron into the Ocean". Café Thorium. Woods Hole Oceanographic Institution. Archived from the original on 2007-02-10. Retrieved 2007-03-31.
  146. ^ Watson, A.J. (1997-02-13). "Volcanic iron, CO2, ocean productivity and climate". Nature. 385 (6617): 587–588. Bibcode:1997Natur.385R.587W. doi:10.1038/385587b0. S2CID 4316845.
  147. ^ Resolution LC-LP.1 (2008) On the Regulation of Ocean Fertilization (PDF). London Dumping Convention. 31 October 2008. Archived (PDF) from the original on 28 August 2013. Retrieved 9 August 2012.
  148. ^ "Assessment Framework for scientific research involving ocean fertilization agreed". International Maritime Organization. October 20, 2010. Archived from the original on 8 November 2012. Retrieved 9 August 2012.
  149. ^ Tollefson, Jeff (2012-10-25). "Ocean-fertilization project off Canada sparks furore". Nature. 490 (7421): 458–459. Bibcode:2012Natur.490..458T. doi:10.1038/490458a. PMID 23099379.