0% found this document useful (0 votes)
57 views79 pages

Abney 2017

Recovery of U from seawater.

Uploaded by

Igor Dejanovic
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
57 views79 pages

Abney 2017

Recovery of U from seawater.

Uploaded by

Igor Dejanovic
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 79

Review

pubs.acs.org/CR

Materials for the Recovery of Uranium from Seawater


Carter W. Abney,* Richard T. Mayes, Tomonori Saito, and Sheng Dai
Chemical Sciences Division, Oak Ridge National Laboratory, One Bethel Valley Road, Oak Ridge, Tennessee 37831, United States

ABSTRACT: More than 1000× uranium exists in the oceans than exists in terrestrial
ores. With nuclear power generation expected to increase over the coming decades,
access to this unconventional reserve is a matter of energy security. With origins in the
mid-1950s, materials have been developed for the selective recovery of seawater uranium
for more than six decades, with a renewed interest in particular since 2010. This review
comprehensively surveys materials developed from 2000−2016 for recovery of seawater
uranium, in particular including recent developments in inorganic materials; polymer
adsorbents and related research pertaining to amidoxime; and nanostructured materials
such as metal−organic frameworks, porous-organic polymers, and mesoporous carbons.
Challenges of performing reliable and reproducible uranium adsorption studies are also
discussed, as well as the standardization of parameters necessary to ensure valid
comparisons between different adsorbents.

CONTENTS 5.3.2. Computationally-Guided Amidoxime


Development 13979
1. Introduction 13935 5.3.3. De Novo Design and Synthesis of Uranyl
2. Brief History of the Recovery of Seawater Chelating Ligands 13979
Uranium 13937 5.3.4. Determination of Binding Constants for
3. Challenges of Working with Seawater 13939 Uranium with Amidoxime and Deriva-
3.1. Conservative vs Nonconservative Metals 13940 tives 13980
3.2. Uranium Speciation 13940 6. Nanostructured Materials 13982
3.3. Salinity and pH Effects 13941 6.1. Porous Carbon 13982
3.4. Biofouling 13942 6.2. Mesoporous ATRP Initiators for Advanced
3.5. Temperature 13942 Adsorbents 13986
3.6. Analytical Considerations 13943 6.3. Metal−Organic Frameworks (MOFs) 13988
3.7. Summary of Challenges 13944 6.3.1. Orthogonally-Functionalized MOFs 13989
4. Inorganic Adsorbents 13944 6.3.2. Unfunctionalized MOFs 13992
4.1. Layered Metal Sulfides or Polysulfide-Inter- 6.3.3. Summary and Perspectives 13995
calated Layered Hydroxides 13944 6.4. Covalent Organic Frameworks (COFs) 13995
4.2. Chalcogel Sorbents 13945 6.5. Porous Organic Polymers (POPs) 13996
4.3. Metal Oxides and Phosphates 13945 6.6. Porous Silica 13999
4.4. Magnetic Adsorbents 13947 6.7. High Affinity Genetically-Engineered Pro-
4.5. Summary and Perspectives 13947 teins 14002
5. Synthetic Polymers 13947 7. Concluding Summary and Perspectives 14002
5.1. Poly(amidoxime) and Derivatives 13947 Author Information 14006
5.1.1. Traditional Preparation of Poly- Corresponding Author 14006
(Amidoxime) 13947 ORCID 14006
5.1.2. Identification of How Amidoxime Binds Author Contributions 14006
Uranyl 13948 Notes 14006
5.1.3. Cyclic Imidedioxime Functionality 13949 Biographies 14006
5.2. Polymerization Methods 13953 Acknowledgments 14006
5.2.1. Introduction 13953 References 14006
5.2.2. Polymer Adsorbents Prepared by Nei-
ther RIGP Nor ATRP 13953
5.2.3. Radiation-Induced Graft Polymerization 13961
5.2.4. Atom-Transfer Radical Polymerization 13974
5.3. Rational Design and Synthesis of Uranyl-
Chelating Ligands 13978
Received: June 21, 2017
5.3.1. Introduction 13978
Published: November 22, 2017

© 2017 American Chemical Society 13935 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

1. INTRODUCTION National Laboratory summarizes more recent international


efforts, organized by country.18
Limited resource availability and unchecked population growth
It is important to note the aforementioned reviews often
have motivated interest in identifying and utilizing unconven-
address, at least in part, many of the engineering and site
tional reserves for industrial metals and minerals. This is
identification challenges related to extracting uranium from
especially true when social, political, or environmental factors
seawater. These aspects were of particular importance during
could affect supply security, resulting in severe global economic
the early years at the height of debate over fixed vs fluidized
repercussions.1 The primary source of uranium is the ore
beds and whether active pumping or passive methods over
uraninite, U3O8, which is largely concentrated in several regions
inorganic adsorbents would provide sufficient extraction of
across the earth.2 Australia and Kazakhstan contain approx-
uranium to mitigate the energy requirements of the pumping
imately 33% of all economically recoverable uranium, with only
itself. Indeed, Kelmers from Oak Ridge National Laboratory
15 countries possessing a quantity adequate to facilitate nuclear
noted, “At 100% uranium recovery efficiency, an ocean stream
power generation. Uranium reserves are predicted to be equivalent to 25 times the annual Mississippi River flow would
sufficient for sustained power generation at the current rates have to be processed to recover 10,000 tons/year. This scale of
of consumption for the next 80−120 years,2,3 but global energy operation raises fundamental engineering questions...”19 While
requirements are expected to double by 2050 as emerging there remain essential challenges to surmount prior to large
economies drive annual growth4 and the global population scale deployment, the engineering questions are ultimately a
surpasses 13 billion.5 Despite per capita energy consumption function of the adsorbent to be deployed. Research with
gradually declining in the United States and developed nations, braided polymer adsorbents, the current state of the art
globally it has increased by more than 30% over the past two technology, was initiated in the 1990s, and the accompanying
decades.6 It is projected that nuclear power generation will engineering questions have not changed significantly in recent
double by 2040 to address this growing energy demand,7 years.
making uranium availability a matter of energy security.4 Several further clarifications in the scope of this review are
Uranium exists in seawater at a concentration of 3.3 μg L−1, also important to address at this juncture. First, the focus of the
forming a highly stable Ca2[UO2(CO3)3] complex.8 Though review is on materials which have been intentionally designed
dilute, this amounts to an estimated 4.5 billion tons of uranium, and developed. That is to say, this review does not cover the
which is approximately 1000× more than is available from adsorption of uranium on minerals, diatomaceous earth,
conventional sources such as terrestrial ores.9 Even considering pyrolyzed organic matter, or other such materials which
dramatic increases in nuclear power generation and ignoring occur naturally or have not been rationally synthesized. Second,
any waste reprocessing, the quantity of uranium in seawater is materials developed for the pre-concentration of uranium
sufficient to ensure sustained power generation for thousands specifically for analytical purposes are also not covered.
of years. Development of technology capable of economical Although related in several respects, the intent behind the
uranium extraction from seawater would afford a financial design of such materials and the methods for evaluating their
backstop, ensure resource accessibility to nations devoid of performance do not facilitate adequate comparison with
uranium reserves, and impede any dramatic fluctuations in the materials developed with environmental deployment for
uranium supply chain, thereby promoting investment in this selective uranium recovery as the end goal. Finally, the
mature, low-carbon source of base load power generation.3,10,11 application of the material is the recovery of uranium from
The purpose of this review is to summarize the current state environmental seawater, not simply aqueous solutions. As
of knowledge regarding materials developed for recovery of discussed below, seawater is a highly complex matrix, replete
uranium from seawater. Although select seminal articles with high ionic strength and competing ions and possessing an
predating the era of interest will be touched upon, this work incredibly low concentration of uranium. The ionic strength
is primarily focused on developments occurring from 2000− (specifically, the presence of the carbonate anion CO32−),
2016. Numerous other reviews have been published, especially concentration, and pH of the aqueous solution dictate the
pertaining to efforts in the late 1970s through 1996.9,12−14 speciation of uranium,8,20 which in turn dramatically influences
Government reports also provide an often untapped the adsorption properties. In general, materials are not included
information stream, with a two volume work prepared by the if they are only investigated for uranium uptake from aqueous
U.S. Department of Energy providing a truly exhaustive survey media, with no consideration for the concentration and salinity
of the field through 1979, including a list of all known materials challenges that accompany seawater deployment.
tested for the extraction of uranium from environmental The noteworthy exception to the preceding statement is
seawater and two separate bibliographies with references to all section 6, Nanostructured Materials, which includes research on
known related publications.15,16 Of particular interest, one porous carbon, metal−organic frameworks (MOFs), covalent
volume includes meeting notes from trips to Europe organic frameworks (COFs), porous organic polymers (POPs),
(specifically Italy, Germany, and Great Britain) and Asia porous silica, and genetically engineered proteins. In contrast to
(Japan and China), discussing the current effort by those more technologically mature materials, such as traditional metal
countries at the end of the 1970s.15 This document provides a oxides or graft-polymer adsorbents, most of the aforementioned
capstone for the early work on uranium extraction from nanostructured materials have not been developed to the point
seawater, as inorganic adsorbents had been investigated for where deployment in environmental seawater is practical, even
almost 30 years and may be where research on organo- on a small scale. As a result, uranium extraction studies are
functionalized chelating resins is mentioned for the first time. A often performed at lab-scale on systems which are far less
seminal report by Tamada was published in 2009 providing a complex than environmental seawater. Significant work has
high-level overview of significant Japanese contributions, been performed to develop these new classes of adsorbents, and
including two of their pilot-scale studies with polymer to the best of our awareness no comprehensive review exists
adsorbents.17 Finally, a 2010 report from Lawrence Berkeley summarizing their capabilities for seawater uranium recovery.
13936 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 1. Development of representative adsorbent technologies throughout history for the extraction of uranium from seawater. Only adsorbents
contacted with environmental seawater are displayed. The majority of samples are exposed to seawater for 20−60 days, with the shortest contact
times less than 1 day (denoted with *) and the longest contact time 240 days (denoted with †). References for displayed materials are provided in
brackets. Triangular markers denote open ocean deployments.

Such materials often display dramatic improvements in 2. BRIEF HISTORY OF THE RECOVERY OF SEAWATER
selectivity or uranium adsorption capacity as well, suggesting URANIUM
at high levels of success should technological development be The extraction of uranium from seawater has been a topic of
pursued further. Nanostructured materials represent the very active research by various state entities for more than six
cutting-edge of materials research for seawater uranium decades (Figure 1). The first mention of this research area in
recovery and are far behind adsorbent polymers with regards the open literature occurs in 1964 by Davies and colleagues,21
to material development and deployment. Nevertheless, in an as is often cited in recent articles. While there are no citations in
effort to provide a comprehensive review on the materials the aforementioned work to earlier publications explicitly
developed for seawater uranium recovery, they are included in regarding uranium extractions, even a casual reading makes
apparent that the article is in fact reviewing an extensive
the final section.
research effort. The genesis of this research dates back to
Starting with a brief overview of the history of seawater “Project Oyster”, initiated in the early 1950s in the United
mining of uranium and followed by a discussion of the Kingdom. In the wake of WWII and the initiation of both
challenges associated with performing this research, this review nuclear power and weapons programs, a sense of urgency
will categorically summarize developments in synthesized developed as it was realized sparse terrestrial reserves would
adsorbent materials. Inorganic sorbents were one of the earliest force reliance on potentially volatile international markets and
technologies developed, and though falling out of favor in the create vulnerability in both defense and power production. It is
1980s, recently renewed efforts are articulated. Synthetic noteworthy that similar concerns motivate the current research
polymers represent the category of adsorbent most capable of effort, as well as efforts to harness unconventional reserves of
other critical minerals, despite sparse economic opportunity.
large-scale deployment in the near term. Research pertaining to
Declassified documents available from the U.K. Archives
modifications of the trunk material, new developments in recount the pioneering investigations with early inorganic
polymer functionalization, tuning material properties through adsorbents, liquid- liquid extractions, and identification of
postsynthesis treatment, and the use of computer-guided design promising sites for establishing adsorbent beds and recovery
for the preparation of highly selective ligands are all covered in facilities.22,23
section 5. Nanostructured materials constitute the newest Efforts of these early decades focused on development of
avenue for potential uranium adsorbents, leveraging extremely inorganic adsorbents deployed either as fixed or floating
high surface areas and readily tailorable properties to achieve adsorbent beds. The performance of 81 different inorganic
unprecedented uranium capacities. Research performed in this materials are reported as of 1979, with lead sulfide, lead
naphthalene, and hydrous titanium dioxide all achieving
area includes porous carbon, MOFs, organic framework
uranium capacities exceeding 1 mg of uranium/g of metal in
materials (e.g., COFs and POPs), and mesoporous silica. adsorbent. Despite the ultimate consensus that hydrous
Finally, this review concludes with a research summary and the titanium dioxide was the best material, the number of lead-
authors’ perspectives regarding where ongoing efforts should be based inorganic adsorbents that were investigated is approx-
focused to achieve greatest impact. imately equal to the number of nonlead-based adsorbents.15,16
13937 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

In the early 1980s the technology for extracting uranium be subsequently transformed to poly(amidoxime). The
from seawater changed dramatically. While several organo- polyolefin trunk provided the necessary strength for the
functionalized polymers had been investigated late in the adsorbent, while the poly(amidoxime) graft chains were
preceding decade, unremarkable capacities and speculation accessible for uranium binding. Electron beam radiation of
regarding biofouling and limited stabilities in seawater made the trunk fiber under an inert atmosphere generateed free
them more of a novelty than a research emphasis. In 1979 and radicals in the polymer. The fibers could then be placed in
1980, Egawa and colleagues first reported the application of an solutions of AN to graft poly(AN) onto the sites of the free
amidoxime-functionalized polymeric adsorbent, formed by radicals.33 Comonomers, such as methacrylic acid, were
treating poly(AN-co-divinylbenzene) (AN = acrylonitrile) included to increase the hydrophilicity of the final polymer
beads with hydroxylamine. Using this resin-based amidoxime fiber. Finally, a treatment with KOH deprotonated the
adsorbent in a column system, more than 80% of uranium carboxylic acids and induced swelling of the polymer fiber for
could be recovered from a volume of seawater 4 orders of contact with seawater.17,32,33
magnitude larger than the volume of the adsorbent.24,25 A Large scale marine experiments were also performed by the
subsequent report articulated the recovery of 450 μg of Japanese, with the longest lasting from 1999−2001.34 Noting
uranium/g of adsorbent following 130 days of continuous the additional energy demand from pumping seawater across an
contact with seawater, as well as an average 82.9% uranium adsorbent bed, they sought to use passive diffusion of seawater
recovery over ten recycles of the adsorbent.26 In 1982, a through adsorbent stacks deployed in the ocean. Adsorbent
seminal work by Schwochau and co-workers articulated the key stacks were prepared from amidoxime-functionalized poly-
parameters for adsorbents for recovering uranium from ethylene fabric, which was then cut into smaller sheets. 144
seawater.27 Large-scale availability at low cost, stability under adsorbent stacks were placed into a 16 m2 adsorption bed, of
seawater conditions, physical resilience, ease of deployment, which three were suspended from an 8 m × 8 m floating frame
rapid binding kinetics, high loading capacity, and facile elution anchored to the seafloor. Figure 2 displays an example of the
were listed as essential requirements to merit further
investigation. These criteria were then used to assess the
results of a screening study of approximately 200 organo-
functionalized resins, with the authors determining cross-linked
poly(acrylamidoximes) to be the most promising candidate.27
This work was quickly followed by a more detailed
investigation using amidoxime and imidoxime functionalized
resins,28 as well as the first crystallographic investigation of how
amidoxime binds uranyl.29 Simultaneous with and independent
of Schwochau’s submission, an MIT report comparing
performance of hydrous titanium oxide and four experimental
ion exchange resins was submitted to the U.S. Department of
Energy. The authors indicate an amidoxime functionalized resin
matched the performance of the hydrous titanium oxide, and
displayed considerably superior mechanical properties.30 They
speculated improvements in performance upon optimization,
which they subsequently observed and published in a separate Figure 2. Adsorbent bed technology developed for extraction of
uranium from seawater. (Left) An illustration of an adsorbent stack.
report submitted later in the same year.31 The fact that
(Right) Photograph of the floating adsorbent frame during open ocean
amidoxime- functionalized adsorbents are still state of the art deployment studies. The frame contains three adsorbent beds, each
attests in part to the validity of the aforementioned composed of 144 stacks of adsorbent. Figure reprinted with
investigations, but should also hint at an area for potential permission from ref 34. Copyright 2003 Taylor & Francis Ltd.
advances in the field. After over 30 years of further research, it
is not unthinkable that materials eliminated by Schwochau due
to failing a key metric could be developed more economically adsorbent stack as well as a scheme of the floating frame and
and achieve superior performance than amidoxime. adsorbent beds. A crane ship extracted adsorbent beds every
From 1981 to 1988, an experimental plant was operated in 20−40 days for fractional elution with 0.5 M HCl. Over the two
Japan using a hydrous titanium dioxide sorbent. However, low year study, approximately 1 kg yellowcake was collected,
adsorption capacities of approximately 0.1 mg of uranium/g of affording an average of 0.5 mg uranium/g-adsorbent every 30
sorbent led to discontinuation of the project. Economic factors days.17
such as the poor mechanical resistance of the sorbent and the It was determined the most expensive part of the experiment
necessity of mechanical pumping of seawater across the with sorbent stacks was the floating frame. An innovative
adsorbent bed made the approach unsustainable. Forming the redesign involved preparation of a braided adsorbent that could
sorbent into pellets increased mechanical durability, but be anchored to the seafloor, creating a sorbent bed biomimetic
intraparticle diffusion dramatically decreased the sorption that of a kelp farm. Attaching a float to the top of the sorbent braid
was achieved.32 allowed their suspension in seawater, and upon wireless release
Major developments in preparation of amidoxime-function- from the anchor, facilitated their recovery with a fishing boat.
alized fibers were made in Japan over the 1980s and 90s.17 After 30 days of soaking, the polymer fiber recovered 1.5 mg
Discovering poly(amidoxime) to have poor mechanical uranium g−1 adsorbent.35 Figure 3 displays the proposed
strength, Tamada and co-workers used radiation-induced graft deployment system for the braided adsorbent.17
polymerization (RIGP) to functionalize a strong polymeric Noteworthy results from Korea from 1999 to 2000
trunk fiber, such as polyethylene, with poly(AN) which could articulated the postsynthetic installation of a cationic site
13938 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Since 2011, there has been renewed enthusiasm in the


United States for extracting uranium from seawater, stemming
from the formation of an interdisciplinary, multi-institution
working group founded by the Office of Nuclear Energy.
Starting with the goal of doubling the adsorption capacity of the
Japanese materials, multifaceted efforts were pursued leveraging
advances in high performance computing, advanced character-
ization instruments, and nanoscience to enable technical
breakthroughs. This multifaceted effort involved seven key
thrusts: (1) investigation of uranium coordination and
application of computation to ligand design; (2) thermody-
namic, kinetic, and structural characterization; (3) development
of advanced polymeric adsorbents through RIGP; (4)
preparation of novel nanoscale adsorbents; (5) establishment
of marine testing for performance assessment; (6) improving
durability and reusability; and (7) cost analysis and deployment
modeling. At the time of writing, these efforts are ongoing, with
major contributions including development of high surface-area
trunk materials for polymer adsorbents,48 two optimized
polymer formulations prepared by RIGP,49,50 application of
atom-transfer radical polymerization (ATRP) for adsorbent
preparation with record-setting performance,51,52 and the first
instance of a metal−organic framework (MOF) for uranium
recovery from seawater.53 Most recently there has been an
explosion of pertinent literature from China, with select
noteworthy reports including a promising organic−inorganic
composite adsorbent,54 a graphene-oxide/amidoxime hydro-
gel,55 amidoxime-functionalized on ultra high-molecular weight
polyethylene (UHMWPE),56 and electrospun polyamidoxime
with polyvinylidene fluoride (PVF) to enhance bulk mechanical
robustness.57
Over more than six decades, the extraction of uranium from
seawater has processed through several different material
Figure 3. Proposed deployment system for braided amidoxime- generations, with representative seawater capacities displayed
functionalized polymer sorbents, intended to simulate kelp fields. in Figure 1. From the initial studies in Project Oyster involving
fixed bed adsorbents of hydrous titanium dioxide, to organic
chelating resins of the 1980s, to the ocean-deployable polymer
within the adsorbent polymer, advantageous from the adsorbents in the 1990s, to the current generation of high
perspective of increasing hydrophilicity, facilitating mass surface area fibers, a steady multi-national effort has persisted
transport through Coulombic repulsion-induced swelling of throughout and is expected will continue for the foreseeable
graft chains, and promoting electrostatic concentration of the future.
anionic UO2(CO3)34− in seawater.36−38
Simultaneously, research from Turkey and Japan afforded a 3. CHALLENGES OF WORKING WITH SEAWATER
wealth of information further developing RIGP processes for The complexity of seawater constitutes one of the greatest
the design of polymeric adsorbents for uranium recovery from challenges in the development of effective adsorbents for
seawater. Systematic investigations were performed to thor- uranium extractions, as well as one of the greatest difficulties in
oughly studied poly(amidoxime-co-methacrylic acid) adsorb- analyzing or reproducing the performance of adsorbents
ents,39,40 while several new avenues were developed for the reported in the literature. Rigorous application of the scientific
development of amidoxime-functionalized polymers. Work method requires variation of only one parameter per experi-
initiated by Güven, Seko, and Tamada reported the preparation ment, and with respect to developing materials for extracting
of glycidyl methacrylate (GMA) grafted by RIGP to a uranium from seawater, this variable is routinely a difference in
polyolefin support.41 This approach allowed for ring-opening adsorbent or treatment process. However, the composition and
of the epoxide group on the GMA by a nucleophile, such as a acidity of seawater can be attributed to a wide range of
sterically unencumbered amine, enabling the postgraft environmental and biological aspects, in turn creating
installation of functional groups which are not amenable to numerous other parameters which must be adjusted or, at the
RIGP conditions or designer chelating moieties such as bis- very least, monitored and normalized. Such effort requires an
amidoximes.42 Emulsion polymerization was also investigated extensive investment in manpower and resources, ostensibly
for the first time for development of uranium adsorbents.43 beyond what is reasonably achievable by most independent
Literature from the mid- to late-2000s is dominated by India, research groups, despite the necessity for reproducible and
with studies reporting application of multidentate chelating relevant extraction data. While extraction of uranium from one
ligands grafted on resin,44 hydrogels composed of various batch of seawater would ensure matrix consistency, the
copolymers,45 amidoxime-based polypropylene sheets,46 and in- extremely low concentration (approximately 3.3 μg L−1)
field demonstration experiments in the Trombay Estuary.47 requires either extremely large volumes of water (at least one
13939 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

liter per mg of adsorbent, assuming a capacity not exceeding 3 Table 1. Concentration of Selected Elements in Seawater, as
mg of uranium/g of adsorbent) or the addition of a uranium Reported in ref 58
spike, increasing the concentration to something more readily
concentration in seawater
detected, but which is not actually representative of the
environmental conditions it is intended to depict. As the element mg kg−1 (ppm) mol L−1
binding of uranium by any chelating group is inherently an Cl 19 400 0.546
equilibrium reaction, increasing the concentration of uranium Na 10 800 0.468
in solution will necessarily increase the quantity of uranium Mg 1290 53 × 10−3
bound by the adsorbent material, resulting in exaggerated Ca 413 10.3 × 10−3
performance which is not representative of performance in K 400 10.2 × 10−3
actual seawater. Li 0.18 26 × 10−6
Preparation of simulated seawater could afford a reproducible Ni 0.005 8 × 10−9
surrogate for environmental seawater, but due to the low Fe 0.0034 0.5 × 10−9
concentration of uranium in seawater, similar challenges are U 0.0033 14 × 10−9
encountered as discussed above. Importantly, a representative V 0.00183 36 × 10−9
simulant must be of appropriate pH and composition, including Cu 0.001 3 × 10−9
total dissolved solids, organic content, and competing ions. S 0.0009 28 × 10−3
Perhaps the best experimental seawater simulant has been Pb 0.00003 0.01 × 10−9
reported very recently by the group of Wu and colleagues, TICa 0.00029 24.2 × 10−3
where the simulant was prepared from sea salt dissolved in DI DIPb 0.00071 2.3 × 10−6
a
water, further adjusted to the environmentally relevant salinity, Total Inorganic Carbon. bDissolved Inorganic Phosphorus.
pH, and metal concentration.56 At the very least, a salt solution
should be utilized which affords relevant concentrations of
carbonate, ionic strength, and pH. It is well-known that concentration is relatively homogeneous throughout the
adsorption of uranium is greatly enhanced at low pH, while oceans, uranium complexes interact only weakly with
speciation of uranium is known to vary as a function of both pH suspended particles, and the residence time of uranium exceeds
and concentration, and adsorbents have superior performance 105 years.58 In contrast, aluminum interacts strongly with
when deployed in matrixes of low ionic strength. While suspended particles and displays a short oceanic residence time,
awareness of these difficulties is necessary for the design and giving it a nonconservative scavenged-type distribution. Iron
and copper have distributions which are influenced by both
engineering of advanced adsorbents, manipulating these
recycling and scavenging processes, making them non-
established parameters to achieve the highest uranium uptake
conservative and displaying a hybrid distribution.
also appears to be a common practice. Some of the major
Perhaps the most important competing ion in extracting
challenges of working with seawater are summarized in this
uranium from seawater is vanadium, as increased concen-
section.
trations of vanadium are known to dramatically reduce uranium
3.1. Conservative vs Nonconservative Metals uptake.61 While it has been known for a long time that
As mentioned previously, uranium is dissolved in seawater at a competing ions have deleterious effects on uranium extrac-
relatively uniform 3.3 μg L−1. (In this review we refer to ppb tion,15,62 the extent to which these effects are realized is a
and μg L−1 synonymously even though the density of seawater function of the both the adsorbent and the local seawater
is not precisely 1 kg L−1. Minor differences in uranium composition.
distribution as well as experimental error intrinsic in analytical 3.2. Uranium Speciation
analysis of trace elements in seawater surpass any significant The speciation of uranium in seawater is an area that has been
difference between whether the quantity of seawater is 1 L or 1 extensively investigated and is reasonably well understood.
kg, making the discrimination largely semantic for our Early work reported on uranium extraction from seawater
purposes.) As displayed in Table 1, numerous other elements regarded uranium speciation as forming the anionic triscarbo-
spanning the periodic table are also dissolved at detectable nato-uranyl complex, UO2(CO3)34−,27 resulting in development
concentrations in seawater,9,58 many at higher concentrations, of adsorbents capable of extracting a highly anionic complex.63
creating the need for a material which selectively binds with However, the literature abounds with instances of
uranium. Importantly, different adsorbents will display different MnUO2(CO3)32(2‑n)‑ complex identification, particularly in
affinity for the various metals present in seawater. For instance, areas of geochemical research and environmental remedia-
the aforementioned amidoxime-functionalized polymers are tion. 6 4 − 6 7 The first work to report an aqueous
generally known to extract large amounts of vanadium, iron, Can[UO2(CO3)3]2(2‑n)‑ complex involved the application of
and copper, in addition to the alkali and alkali earth cations.59 time-resolved laser fluorescence to seepage waters of a mine
The relative concentration of competing ions in the seawater tailing pile. 6 8 The authors proposed the neutral
has the potential to dramatically affect the uranium capacity of Ca2[UO2(CO3) 3] as the dominant species, which was
the adsorbent material.60 independently supported shortly thereafter.69 X-ray absorption
While a rigorous discussion of marine geochemistry is fine structure spectroscopy (XAFS) has been applied at various
beyond the scope of this review, for the sake of discussion points, and while Can[UO2(CO3)3]2(2‑n)‑ complexes can be
metals dissolved in seawater can be loosely divided into the identified, precise quantification of the Ca coordination number
categories of conservative and nonconservative, based on the is impeded by spectral overlap in the extended XAFS (EXAFS)
extent to which it participates in biological, chemical, or region with the distal oxygen of the carbonate group,70,71 which
geological processes, relative to its original concentration. is further complicated by the presence of a
Uranium is a conservative metal, meaning the overall NamCan[UO2(CO3)3](2(2‑n)‑m)‑ complex.67 More recently, En-
13940 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

drizzi and Rao used calcium ion selective electrode 3.3. Salinity and pH Effects
potentiometry and optical absorption spectrophotometry to In addition to speciation as a function of dissolved alkali and
provide precise equilibrium constants for Ca and Mg alkali earth cations, uranium concentration, ionic strength, and
complexation with UO2(CO3)34−. These data were then used pH are also known to affect speciation and even result in
to calculate the speciation of uranyl in environmental seawater, formation of polynuclear complexes.72,73 Work by Krestou and
with enthalpies of complexation determined for Ca- Panias reports the calculated speciation of uranium as a
[UO2(CO3)3]2− and Ca2[UO2(CO3)3].8 It was observed that function of concentration, pH, ionic strength, and carbonate.73
more than 90% of the overall uranium dissolved in seawater In all instances, the free uranyl cation (UO22+) is calculated at
could be accounted for in Ca2[UO2(CO3)3] and Mg- low pH < 4, while UO2(CO3)34− is obtained above pH 8.5.
[UO2(CO3)3]2− complexes, as displayed in the updated However, formation of polynuclear (UO)2CO3(OH)3− is
speciation diagram in Figure 4. The same team recently enhanced over pH 6−8 as uranium concentration is increased.
While conditions representative of seawater were not
investigated, a speciation diagram was calculated for a solution
similar to many uranium-spiked brine simulants reported in the
literature (ionic strength of 0.31 M, uranium concentration 1 ×
10−5 M). UO22+ dominates until approximately pH 4.5, where
UO2(OH)2 accumulates, peaking at pH 6. From pH 6.5−7.5,
the polynuclear complex (UO2)2CO3(OH)3− is most abundant,
replaced by UO2(CO3)34− from pH 7.5−10. Comparison of
this speciation, likely representative of many screening
solutions, with that calculated for environmental seawater20
underscores the necessity of realistic sorption conditions and
one of the major challenges in developing adsorbents for
extraction of uranium from seawater.
Salinity, the amount of salt dissolved in a body of water, is
another important factor governing uranium chemistry and
concentration, thereby affecting adsorbent performance.
Laboratory research has revealed that the presence of sodium
chloride, calcium, magnesium, and bicarbonate significantly
retards uranium adsorption kinetics to braided fiber adsorbents,
as revealed in Figure 5.74 Standardization of seawater simulant
Figure 4. Uranium(VI) speciation in seawater. 1. UO22+; 2. and normalization of uranium extraction data from environ-
(UO2)(CO3)(aq); 3. UO2OH+; 4. (UO2)(CO3)22−; 5. Ca2[(UO2)- mental seawater is essential for valid head-to-head comparisons
(CO3)3](aq); 6. Mg[(UO2)(CO3)3]2−; 7. Ca[(UO2)(CO3)3]2− 8.
of material performance. For example, in an effort to compare
(UO2)(CO3)34−. The dashed gray line corresponds to an approximate
seawater pH of 8.2. Figure reprinted with permission from ref 20. samples screened at different locations around the United
Copyright 2016 American Chemical Society. States, the DOE-NE Fuel Resources team chose to normalize
the seawater data to 35 practical salinity units, a universally
accepted value for seawater salinity around the world, roughly
equivalent to 35 parts-per-thousand.61 As uranium concen-
reviewed the efforts to determine uranium speciation in tration scales with salinity, it was prudent to normalize to this
value. Moreover, salinity normalization also provides a basis for
seawater, as well as investigated the solubility products for
comparison between locales where salinity differs.
several uranyl- containing minerals in seawater.20 The reader is The application of normalizing uranium uptake as a function
directed there for more comprehensive discussion. of salinity was first reported in the work of Tsouris and

Figure 5. Influence of sodium chloride “salt” (left) and sodium bicarbonate (right) on uranium adsorption kinetics by a braided polymer adsorbent.
Figure reprinted with permission from ref 74. Copyright 2016, American Chemical Society.

13941 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

colleagues,59 who observed a salinity concentration of 28.5


practical salinity units (psu) in the environmental seawater used
for their uranium uptake experiments. Previous work identified
uranium concentrations in seawater from the coast of Savannah
(Georgia, U.S.A.) and Charleston (South Carolina, U.S.A.), 3.6
and 3.2 ppb, respectively,75 which differed from the 2.85 ppb
uranium concentration in Sequim Bay (Washington, U.S.A.).
The same adsorbent displayed different performance when
deployed under these conditions, with uptake of 3.94, 3.36, and
2.7 mg of uranium/g achieved in Savannah, Charleston, and
Sequim Bay seawater. When the salinity and uranium
concentration of the Sequim Bay seawater was normalized to
35 psu and 3.3 ppb, consistent with open ocean conditions, the
uranium uptake was determined to be 3.3 mg of uranium/g. An
adsorbent provided by the JAEA was also tested simultaneously
for comparison, affording an uptake of 1.1 mg of uranium/g in
Sequim Bay seawater, which normalized to 1.3 mg of uranium/
g.59 This approach was replicated in subsequent publications to
normalize uranium uptake obtained from deployment in low
salinity seawater to values anticipated following deployment in
the open ocean.49,50,76−78
Figure 6. Growth of E. choli on Ag-nanoparticle-loaded polymer
3.4. Biofouling adsorbents. (i) Whatman filter disc. (ii) Control adsorbent without
Biofouling is the surface accumulation of microorganisms, algae, nanoparticles; (iii) 1.6 wt % Ag; (iv) 3.0 wt % Ag. Figure reprinted
plants, and eventually animals. Due to potential deployment with permission from ref 89. Copyright 2011 Elsevier.
under environmental conditions, the effects of biofouling also
constitute a challenge for the development and screening of
adsorbents for uranium extraction.79 Biofouling is considered to
occur in four stages,80,81 beginning only seconds after
immersion in seawater where surfaces are coated with a thin
organic film.82 The second stage is characterized by the settling
and colonization of bacteria and diatoms on the surface of the
material.83 These are followed by the development of microbial
films, constituting the third stage, which create rough surfaces
which trap additional particles and organisms. The final stage
occurs when macroorganisms, such as barnacles or mussles,
begin to grow from the fouled surface.84,85 Design features
commonly considered advantageous for adsorbents, such as
large porosities or nanotexturing, may promote the rate of
biofouling and ultimately render the materials unusable.80,86−88
Initial reports indicate biofouling on inorganic materials was Figure 7. Effect of biofouling on an amidoxime-functionalized polymer
not directly observed,15 but more advanced microscopy adsorbent as a function of deployment time. Figure reprinted with
permission from ref 79. Copyright 2016 American Chemical Society.
techniques applied to organic chelating resins reveal filling of
pore structures with unidentified debris after seawater contact,
consistent with biofouling.31 Indian literature articulates the Time and temperature are also known to affect the extent of
preparation of adsorbents containing silver nanoparticles which biofouling,90 influencing the length of any campaign, as well as
display marked improvements in biofouling resistance and are the location of adsorbent deployment.
amenable to deployment in bioaggressive environments (Figure 3.5. Temperature
6).89 Japanese literature reports the facile removal of biofouling The influence of temperature on adsorbent performance is truly
by simply immersing the adsorbent in fresh water prior to multifaceted, and ocean temperatures can vary dramatically
elution of uranium, with the change in ionic strength sufficient depending on location and season. Computational and
to dissolve the biological growth.34 Recent investigations have thermochemical titrations demonstrate chelation of uranyl by
been performed to investigate the effects of biofouling on amidoxime is an endothermic process. Accordingly, increasing
amidoxime-functionalized polymer adsorbents used for sea- solution temperature is expected to improve adsorbent
water extraction of uranium.79 The adsorbents were tested in performance.63,91 While such investigations cannot account
the presence or absence of light to simulate deployment in for the effect of the polymer support itself, complementary
shallow or deep marine environments, with continuous flow of work has been performed on seawater-contacted adsorbent
environmental seawater designed to represent natural ocean polymers, confirming the results obtained from small molecule
currents. After 42 days of exposure, uranium uptake was 30% studies,60,92,93 and putatively affording extension to nanoscale
lower for the fiber exposed in light, with numerous algal cells adsorbents as well. However, as discussed above, biofouling is
observed by microscopy (Figures 7 and 8). A major conclusion also known to correlate with temperature, which may
of this work is that deployment of the adsorbents below the dramatically reduce the performance of an adsorbent. Never-
photic zone may be necessary to mitigate biofouling effects.79 theless, achieving the right balance between promoting an
13942 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 8. Visible (top) and microscopic (bottom) evidence for biofouling on an amidoxime-functionalized polymer adsorbent. Figure reprinted with
permission from ref 79. Copyright 2016 American Chemical Society.

endothermic reaction and mitigating biofouling is not problems, requiring dilution (either of matrix or in plasma)
necessarily obvious. For example, in-field demonstration or sample preconcentration with seawater matrix elimination.
experiments performed at the intake and outfall canals of the For the latter, several publications have indicated metal
Tarapur Atomic Power Station in India’s Trombay Estuary recovery from seawater of approximately 67%, 80−107%, and
correlated uranium uptake with fouling, deployment time, above 90%, as analyzed by electrothermal atomic absorption
temperature, and other parameters. More than a 30% increase spectrometry,99 graphite-furnace atomic absorption spectrom-
in uranium uptake was observed at the outfall canal due to the etry,100 and inductively coupled plasma atomic emission
increased temperature, despite the increased biofouling that was spectrometry,101 respectively. In these approaches, seawater
observed.47 samples were preconcentrated by reductive precipitation with a
Temperature control can be readily achieved under mixture of sodium borohydride with iron and palladium. The
laboratory conditions, as well as by judicious choice of precipitate was collected and dissolved with nitric acid for
deployment location and date, making it one of the difficulties analysis.
most easily addressed. Furthermore, while there is some effect In contrast, simple dilution of the matrix can afford more
of temperature, the sensitivity is not overly remarkable, and lab- reproducible and accurate measurements, but requires use of an
based adsorption studies performed under ambient conditions inductively coupled plasma mass spectrometer (ICP-MS) to
will not vary dramatically based on modest changes in room achieve the necessary limits of detection. Recent work
temperature. Nevertheless, adsorbent design will be optimized articulates this approach, diluting seawater 20× in DI water
when uranium binding is exothermic, allowing for deployment and implementing standard addition calibration to match the
below the oceanic phototrophic layer or at reduced temper- matrix of the standards to the samples. When used to analyze
atures to suppress the rate of biofouling. standard reference material as a quality control, the uranium
recovery ranged from 93−99% for 15 replicates, while recovery
3.6. Analytical Considerations
of uranium-spiked seawater samples ranged from 93−109% for
Accurate, reproducible quantification of uranium concentra- seven replicates. The analytical approach is clearly articulated in
tions in seawater is significantly more difficult than in simplistic the aforementioned work, and constitutes a robust and
uranium-spiked brines. For the latter, various avenues for validated method for quantifying trace concentrations of
quantitation are available, including spectrometric detection by uranium in high TDS matrixes.59 Alternative methods are
chelation of free uranium using the Arsenazo III dye, allowing also available for using online preconcentrating chelating ion
facile analysis by UV−vis spectroscopy.94−98 Such an approach exchange resins have also been reported.61,76 Finally, a very
can afford a reasonably accurate measure of uranium remaining recent publication compares eight different methods for
in solution, but is also known to bind Ca2+, preventing its use in quantifying uranium in seawater, revealing the method of
environmental seawater or representative simulants. The standard addition calibration (discussed above) and offline
complex matrix effects resulting from a large amount of total preconcentration with an ion-exchange resin afford the most
dissolved solids (TDS) in seawater further compounds accurate and reproducible results.76 The previously discussed
13943 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 9. (a) PXRD data for pristine KMS-1 (red) and UO22+-exchanged KMS-1 (black) revealing differences in the crystalline lattice following ion
exchange. (b) Analysis of the Bragg peaks of the layered phases in the UO22+-exchanged KMS-1 supporting the assignment of the uranyl cation. (c)
Schematic of UO22+-intercalation parallel to the layers of the KMS-1 material. Figure reprinted with permission from ref 111. Copyright 2012
American Chemical Society.

reductive precipitation method was one of the least effective tion.21,102,103 The inorganic adsorbent research in that time is
methods and produced results consistent with direct analysis of mainly focused on hydroxides; oxides; and sulfides of metals
environmental seawater. such as titanium, aluminum, and magnesium. Among all of the
3.7. Summary of Challenges tested adsorbents, hydrous titania has given the best results.103
The highest uranium adsorption capacity for a hydrous titania
Performing realistic studies to assess material performance for reported was 1.2 mg of uranium/g of titanium (0.58 mg of
seawater uranium recovery is inherently difficult. Ideal studies uranium/g of adsorbent).104 However, due to the poor
will involve deployment in the open ocean or utilize
selectivity, durability, and deployment challenges, the metal
continuously replenished environmental seawater in order to
oxides were abandoned for more selective organic-based
achieve the most accurate data. However, access to such
sorbents such as chelation polymers. More recently, the
capabilities is rare, expensive, and generally unreasonable for
research on inorganic adsorbents has primarily focused on
independent research groups. Batch contact with seawater
porous silica as a support105−110 or as an organo-silica hybrid
provides a more feasible option, but the large volume required
where the silica is impregnated with organic chelation ligands.
to achieve an appreciable quantity of uranium presents a
different challenge, as does the analytical rigor necessary for The silica research will be discussed more within section 6.6
reproducibly accurate and precise quantification. Adequate due to the nanostructured porosity present; however, it is
seawater simulants are available in the literature, which afford important to note the instability of silicas under basic
environmentally relevant pH, ionic strength, uranium concen- conditions, which may mitigate large- scale use within the
tration, and competing trace metals. While alleviating some of oceanic environment. Nevertheless, several inorganic materials
the difficulties, the low uranium concentration nevertheless have recently been reported which display particular innovation
retains the analytical and volumetric challenges discussed and hold promise as potential “game-changing” technologies in
above. At the bare minimum, consideration of appropriate the field of uranyl extraction.
pH and ionic strength are essential for any degree of relevance. 4.1. Layered Metal Sulfides or Polysulfide-Intercalated
Temperature, salinity, and biofouling also affect adsorbent Layered Hydroxides
performance, often through both direct and indirect avenues, Recent work has been published by the Kanatzidis group
and while temperature can be easily controlled in laboratory regarding the use of sulfide-functionalized ion exchangers for
settings, the influence of biofouling cannot be readily separated the extraction of uranyl from seawater. A previously reported
from some form of marine testing. K2MnSn2S6 (KMS-1) was deployed in various water sources,
including environmental sources from Lake Michigan and the
4. INORGANIC ADSORBENTS Gulf of Mexico.111 Upon contact with uranyl-containing
Inorganic adsorbents were early extractants for uranium solutions, the K+ ions exchange for uranyl which locate in the
adsorption dating back to the 1960s, due to low cost, high interlayer space, based upon analysis of powder X-ray
surface area, tunable pore structure, and easy prepara- diffraction (PXRD) data. Based on the change in interlayer
13944 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 10. (Left) Schematic depicting the preparation of the ZIF-67 templated Mg−Co LDH. (Right) TEM image of Mg-CO LDH derived through
use of a sacrificial ZIF-67 template and cobalt source. Figure reprinted with permission from ref 113. Copyright 2017, Elsevier.

spacing and the difference in diameters between UO22+ and K+, greater than 850 mg of uranium g−1 achievable at high uranium
it was proposed the dehydrated uranyl cation intercalates concentrations. When contacted with a brine solution
parallel to the KMS-1 planes (Figure 9). Even in high containing an environmentally appropriate 3 ppb uranium
concentrations of competing ions such as Na+, Mg2+, Ca2+, concentration, an uptake of 0.006 mg of uranium g−1 was
and K+, in the environmental samples, KMS-1 was able to achieved. This uptake increased up to 0.4 mg of uranium g−1
extract between 75−99% of the uranyl in solution when added when contacted with a brine solution possessing 200 ppb
at a phase ratio of 10−20 g L−1 over 12 h contact time. A uranium concentration.113
saturation capacity of 380 mg g−1 was determined from a 4.2. Chalcogel Sorbents
sorption isotherm obtained in deionized water with no pH
A series of chalcogel sorbents were investigated for capture of
adjustment or measurement.111 While these preliminary results 99
Tc, 238U, and gaseous I2, as sorbents for environmental
are exciting and the materials merit further study, the phase
ratios reported are more than 3 orders of magnitude beyond remediation of radionuclides.114 These interesting materials
what is achievable in practical application. contained compositions of Co0.7Bi0.3MoS4, Co0.7Cr0.3MoS4,
A very recent report from the same group depicts use of Co0.5Ni0.5MoS4, PtGe2S5, and Sn2S3 and are prepared by
polysulfide/layered double hydroxide composite materials for chemically reacting chalcogenide clusters with interlinking
uranyl sequestration.112 When added at a reasonable phase ratio metal ions in solvent and then allowing gelation to occur
of 1 g L−1, the layered double hydroxide materials were able to over several weeks. Following gelation, the materials were aged
extract 96−99% of uranium from deionized water, achieving a 1−3 days in ethanolic solutions, before washing with
sorption capacity of 330 mg g−1. While significant variation in supercritical CO2 and mild heating.
initial pH values for isotherm solutions adds uncertainty the The sorbents were investigated for 99Tc and 238U sorption at
accuracy of this number, the results are impressive nevertheless. a phase ratio of 50−100 g L−1 in deionized water for 1 week,
Distribution coefficients (Kd) as high as 3.4 × 106 were with radionuclide removal ranging from 57−98% for 99Tc and
achieved, which are competitive with the best inorganic 68−99% for 238U. Sorption of iodine was measured by flowing
uranium sorbents reported in the literature. These results gaseous I2 over a sorbent bed at 125 °C. All chalcogel materials
were decreased by 1−2 orders of magnitude upon addition of bound >99% of all iodine, as determined by ICP-MS of a 0.1 M
Ca2+ and Na+ but remain among the best literature results for NaOH solution through which chalcogel-filtered gas was
bubbled. The ability to process these chalcogels into monolithic
uranyl extraction. The authors subsequently investigated these
materials makes them a potentially useful technology for uranyl
materials for extracting uranium from environmental seawater
extraction from seawater; however, the use of expensive, rare,
from near Tianjin City, China. The report that 78% of uranium
and toxic metals limits their applicability in current form.
was removed over 24 h would suggest rapid kinetics and great
Furthermore, external validation must be performed, as the
promise as an adsorbent material. However, a close reading
current report only discusses performance when deployed at
reveals adsorbents were added at a phase ratio of 15 g L−1 for
phase ratios which are untenable for environmental conditions
this experiment. Furthermore, control experiments reveal the
and large-scale deployment.114
concentration of uranium in this seawater exceeded 9 ppb,
more than 2.5× the concentration of uranium typically 4.3. Metal Oxides and Phosphates
anticipated under open ocean conditions.112 External validation Gamma-Al2O3 nanosheets were prepared by hydrothermal
of the performance for this material is essential, in particular synthesis in the presence of supercritical CO2 and tested for
when deployed at a reasonable phase ratio and in environ- their performance in extracting uranium from aqueous
mentally relevant uranium concentrations. solutions.115 Al(NO3)3 and urea were dissolved in deioinized
A Mg−Co layered double hydroxide (LDH) possessing water and subsequently pressurized with CO2 to 9 MPa,
hierarchical structuring was synthesized from a sacrificial beyond the CO2 critical pressure of 7.38 MPa. This system was
zeolite-imidizole framework (ZIF) template, which doubled as heated at 160 °C for 24 h and the resulting precipitate further
the Co source. Mg(NO3)2 was added to the ZIF-67 template in calcined at 500 °C for 4 h after thorough washing. The material
an aqueous/ethanolic system, promoting growth of nanosheets forms irregular nanosheets, similar in morphology to the
on the facets of the template and ultimately affording hollow boehmite precursor. Detailed time-dependent microscopy
rhombic dodecahedron structures (Figure 10). These adsorb- studies were performed to investigate the formation process,
ents were contacted with a brine screening solution at a phase with the authors proposing a mechanism including quick
ratio of 0.5 g L−1 for 3 h contact time. Isothermal investigations nucleation resulting from Al3+ hydrolysis followed by gradual
revealed pH 5 is optimal for uranium recovery, with saturation crystallization and agglomeration. The supercritical CO2
13945 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

promoted exfoliation of the disordered sheets, while the affording an uptake of 4.5 mg of uranium g−1. The material
calcination removed residual H2O and CO32− to form the saturation capacity was investigated through an isotherm, but a
hierarchically porous alumina. The synthetic process is modest maximum uptake of approximately 10 mg g−1 was
summarized in Figure 11. observed. Comparatively poor performance at acidic pH was
ascribed to competition of the uranium with the proton in
solution, suggesting an ion exchange mechanism and
anticipation of poor performance in solutions with high ionic
strength, such as seawater.115
An inorganic/organic composite, zirconyl-molybdopyrophos-
phate-tributyl phosphate (ZMPP-TBP), was prepared by
coprecipitation of sodium molybdate and potassium pyrophos-
phate with zirconyl oxychloride.54 While a maximum uptake
was observed at pH 6, there was minimal performance variation
across a wide pH range, and >99% of the uranium was
recovered from a 5 ppm aqueous solution when contacted at a
1 g L−1 phase ratio. Inclusion of various mono- or divalent
cations (Zn2+, Na+, Ca2+, and Mg2+) or carbonate effected a <
Figure 11. Schematic illustration of the proposed mechanism for 5% decrease in uranium uptake, suggesting promising perform-
formation of γ-Al2O3 nanosheets. Figure reprinted with permission
from ref 115. Copyright 2013 Royal Society of Chemistry.
ance in complex matrixes. The saturation isotherm was also
obtained, yielding a maximum capacity of 196 mg of uranium
g−1 and achieving equilibrium in 80 min. In contrast to most
Uranium uptake studies reveal optimal performance is adsorbents, variable temperature investigations with ZMPP-
achieved at pH 5 in a 25 ppm uranium solution, while TBP reveal uranium uptake is inversely related to temperature.
approximately 90% of the uranium is recovered at pH 8, A negative Gibbs free energy for uranyl complexation was

Table 2. Performance of Inorganic Adsorbents for Recovery of Uranium from Water and Seawatera
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
Layered Metal Sulfides or Polysulfide-Intercalated Layered Hydroxides
sulfide ion exchangers (KMS- 1) 380b 111
polysulfide layered double 332 1478 ppm uranium 0.007 24 h batch contact at 9 ppb 112
hydroxides uranium
Mg−Co layered double 850b pH 5 113
hydroxide
0.006 brine at 3 ppb uranium 113
0.4 brine at 200 ppb uranium 113
Chalcogels
CoBiMoS 0.022c 238 ppb uranium 114
CoCrMoS 0.016c 238 ppb uranium 114
CoNiMoS 0.021c 238 ppb uranium 114
PtGeS 0.024c 238 ppb uranium 114
SnS 0.024c 238 ppb uranium 114
Metal Oxides and Phosphates
γ-Al2O3 4.5 pH 8, 25 ppm uranium 115
10b pH 5 115
ZMPP-TBP 5.5 pH 3−9, 5 ppm uranium 54
196b pH 6 54
9.5d pH 6, 10 ppm uranium, with competing 54
ions
Magnetic Adsorbents
CoFe2O4/MWCNT 212.7 pH 6, 50 ppm uranium 116
140.4 pH 8, 50 ppm uranium 116
115.5e pH 6, 50 ppm uranium, with competing 116
ions
CoFe2O4 100 pH 6, 50 ppm uranium 116
MWCNT 95 pH 6, 50 ppm uranium 116
zero valent iron polyamidoxime 206b pH 5, 0.1 M NaCl solution, 117
zero valent iron control 45.8b pH 5, 0.1 M NaCl solution, 117
polyamidoxime control 188b pH 5, 0.1 M NaCl solution, 117

a
Unless otherwise stated, adsorption experiments were performed in DI water and pH was unadjusted. bMaximum capacity at saturation. cPrecise
mass not provided; asserted 0.1 g of adsorbent from provided mass range. dCompeting ion composition: 50 ppm of Zn2+, Na+, Ca2+, Mg2+, or CO32−.
e
Competing ion composition: 50 ppm of Na+, Ca2+, K+, or Mg2+.

13946 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

experimentally determined, demonstrating adsorption to be a that engender favorable uptake. While chalcogels mitigate this
spontaneous reaction, while the inverse relationship with problem by forming comparatively large monoliths, the
temperature attests to a enthalpically driven process. Finally, necessity of toxic, expensive, and rare metals makes them
the uranium could be easily eluted with dilute HCl and the prohibitive from safety and cost perspectives. It is worth
material recycled up to four-times with negligible loss of recalling that inorganic materials, including metal sulfides, were
performance. Although not investigated directly for recovery of extensively investigated and subsequently abandoned after two
uranium from seawater or an appropriate brine, the perform- decades of research. Although the work discussed above reveal
ance is promising and may merit further investigation. some interesting materials, extensive continued effort in this
4.4. Magnetic Adsorbents direction is not anticipated to afford materials suitable for
practical implementation.
Several magnetic inorganic adsorbents have also been reported
of recent for uranium recovery from seawater. A magnetic
cobalt ferrite/multiwalled carbon nanotube (CoFe2O 4 / 5. SYNTHETIC POLYMERS
MWCNT) system was investigated for uranium extraction Polymeric adsorbents comprise the category of material which
from several aqueous solutions. When contacted at a phase is closest to feasible large-scale deployment, due in large part to
ratio of 0.4 g L−1 in a 50 ppm uranium solution, a strong pH pioneering work from Japan and Turkey in the 1980s and
dependence was observed, with a maximum uptake of 212.7 mg 1990s. As discussed previously, several in-field deployment
of uranium/g of adsorbent reported at pH 6. When the pH investigations of advanced polymer-based adsorbents have been
increased to 8, a 33% decrease in performance was reported. performed by research programs of various nations, further
Experiments performed in solutions containing alkaline and demonstrating proof of concept for such materials and
alkali earth cations reveal preferential uptake of uranium.116 affording them the highest technology readiness level of all
Later work articulated the use of zerovalent iron nano- proposed seawater uranium adsorbents. Traditional materials
composites, surface functionalized with AN and subsequently have been prepared by graft polymerization from a polyolefin
converted to amidoxime.117 When used to recover uranium support, such as polyethylene fibers or a polypropylene
from aqueous solutions, a maximum uptake was achieved at pH unwoven fiber mat. Advantages afforded by this approach
5. Increasing ionic strength up to 0.5 M NaCl results in a include scalability, economical material costs, versatility in graft
decrease of adsorption at lower pH, but no decrease from pure chain functionalities, and the diversity of form factors available
aqueous solutions were observed at pH 6 or 7. Investigation of for a support. However, fundamental challenges remain
the uranium-contacted material by X-ray photoelectron spec- regarding the direct investigation of graft chain molecular
troscopy (XPS) reveals a shoulder in the uranium 4f spectrum, structure and positive identification of the uranyl binding
indicative of partial reduction from U(VI) to U(IV), an environment. Moreover, hydrophilicity and salinity influence
oxidation state of uranium known to be much less soluble in graft chain structure, putatively resulting in interesting
water and part of the driving force for uranium binding. A morphological phenomena which remain largely uninvesti-
saturation capacity of 206 mg of uranium/g of adsorbent was gated.
achieved at pH 5 by application of an adsorption isotherm. This section begins by discussing the synthesis of amidoxime,
Nevertheless, while promising performance has been amidoxime-containing polymers, and derivatives of amidoxime,
observed for magnetic adsorbent materials, from the including recent investigations on how amidoxime binds uranyl.
perspective of practical uranium recovery from seawater they While other ligands have been proposed and investigated for
can be little more than an academic curiosity. Uncontrolled uranium recovery, amidoxime-derived adsorbents nevertheless
deployment at scales necessary for significant uranium recovery overwhelmingly predominate. Next, we discuss the polymer-
could not be accommodated biologically, or with regards to ization techniques and resulting materials that have been
subsequent collection from the open ocean, while integration in investigated for seawater uranium recovery. A section
some form of support or monolith defeats the purpose of being discussing the rational design and synthesis of uranyl chelating
magnetic. These materials are not anticipated to be of ligands follows, including alternative methods for incorporating
significant practical use for the recovery of uranium from these well-designed chelators onto polymer-based adsorbents.
seawater. Finally, this section concludes with a brief summary and
4.5. Summary and Perspectives outlook regarding polymer based adsorbents.
Renewed interest in development of inorganic adsorbents for 5.1. Poly(amidoxime) and Derivatives
extraction of uranium from seawater has resulted in several new 5.1.1. Traditional Preparation of Poly(Amidoxime).
materials which display promising uptake under idealized The amidoxime functionality shows favorable binding for the
conditions, with a summary of material performance provided uranyl ion, thus making it the preferred choice for the
in Table 2. Unfortunately, deployment concentrations for some development of uranium adsorbent materials. However, to
of these materials range from 10−100 g L−1, which are produce these materials on a scale capable of economical
unacceptably high phase ratios to achieve suitable performance extraction, this uranium binding moiety must be synthesized
in environmental matrixes. In contrast, advanced polymer through fast, efficient, and high yielding techniques. While there
adsorbents, vide infra, are can achieve comparable performance are a number of synthetic methods for the production of
when deployed at only 0.015 g L−1, a phase ratio approximately amidoximes (see reviews in refs 118 and 119 and references
3 orders of magnitude smaller than the best inorganic therein for broader synthetic scope), the approach most
adsorbent discussed previously. commonly utilized involves conversion of a nitrile into the
Deployment issues also present a challenge for the desired amidoxime product by treatment with hydroxylamine.
aforementioned inorganic materials. While unbound powders This method is generally performed at moderate temperatures
are suitable for lab-scale investigation, the processing necessary (60−80 °C) and in common solvents, such as water and
to enable deployment often undermines the very properties alcohols, for times ranging from a few hours to a few days. This
13947 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

approach to amidoxime formation is largely unchanged from nitrogen had occurred.29 Subsequent potentiometric studies
the pioneering work performed by Egawa and colleagues with both acetamidoxime120 and benzamidoxime121 revealed
around 1980.24−26 deprotonation of the oxime did occur prior to uranyl chelation,
Nucleophilic attack by the nitrogen lone pair of hydroxyl- forming charge-neutral bis-amidoximato uranyl complexes.
amine results in formation of an imino hydroxylamine tautomer Presumably influenced by the early crystallographic data,
which exists in dynamic equilibrium with the desired researchers proposed binding motifs involving four amidoxime
amidoxime (Figure 12). While in the imino tautomer, free ligands coordinating in a monodentate fashion through the
oxime oxygen, with charge balance achieved through
deprotonation of two ligands.46,122 Others have promoted a
chelating interaction between oxime oxygen and amide
nitrogen.123−126 These motifs are summarized in Figure 13.
In an effort to conclusively identify the binding motif for
uranyl-amidoximate complexes, Vukovic and colleagues utilized
Figure 12. Conversion of a nitrile to amidoxime. a complementary approach combining DFT calculations with
small molecule crystallographic studies and investigation of the
rotation of the hydroxylamine about the σ-bond will result in Cambridge Structural Database.127 In this seminal work, DFT
preferential adoption of either E- or Z-conformer upon calculations were performed to evaluate the geometries and
formation of the amidoxime as a function of the electronic relative stabilities of proposed bonding motifs for a series of
structure and relative steric bulk of the R-group. uranyl acetamidoxime (AO) complexes (Figure 13). The
5.1.2. Identification of How Amidoxime Binds Uranyl. authors report DFT calculations in structures I, II, and III
Witte and colleagues were the first to report small molecule achieved local minima, though attempts to optimize IV were
uranyl-amidoxime complexes, as investigated through X-ray unsuccessful, instead yielding III. The relative stabilities were
crystallography.29 To prepare uranyl acetamidoxime, uranyl compared for 1, 2, and 3 coordinating ligands, with the
nitrate was dissolved in an ethanolic solution at 50 °C, to which remainder of the equatorial coordination sphere filled by
10 equiv of acetamidoxime were added. Yellow crystals were individual aqua ligands. Representative geometries are displayed
precipitated by the addition of diethyl ether. Similarly, uranyl in Figure 14. For calculations performed both for solvent and
benzamidoxime was prepared by addition of 4 equiv of gas phase formalisms, the η2 motif, III, was observed to be the
benzamidoxime to an ethanolic solution of uranyl nitrate at most stable, albeit frequently by less than 3 kcal mol−1.
50 °C. Orange-yellow crystals were obtained by recrystallization More recently, high level coupled-cluster CCSD(T)/aug-cc-
from hot nitromethane.29 Both complexes revealed mono-
pVDZ calculations were used to investigate the potential
dentate coordination by four ligands through the oxime oxygen
binding motifs of uranyl complexes with formamidoxime
(Figure 13, top), influencing the proposed motif for uranyl
ligands, as this molecule is small enough that high-level ab
coordination for the next two decades.
initio calculations can be readily performed. Relative energies
While the crystal structures were solved, slight disorder of
obtained from CCSD(T) calculations were used as benchmarks
acetamidoxime and benzamidoxime residues did not accom-
against DFT and MP2 predicted energies. The most stable five-
modate location of H atoms. Though short bond lengths
and six-coordinate B3LYP geometries of the uranyl complexes
suggested an anionic ligand with resonance contributions, it
with 1−3 ligands involving η2-binding with the N−O bond and
was not definitive whether the oxime had been deprotonated or
chelation through the oxime oxygen and amine nitrogen donor
whether tautomerization via H atom transfer to the oxime
atoms are shown in Figure 15.128 Consistent with previous
DFT calculations and X-ray diffraction data,127,129,130 the η2-
binding motif is the most stable at the B3LYP and MP2 levels
for all amidoxime complexes; however, analysis of the relative
stabilities of uranyl complexes with amidoxime reveals that the
difference in the stability of the η2 and chelate binding motifs is
smaller than previously thought.127 After application of higher
order correlations estimated at the coupled-cluster theory,
differences of 0−2 kcal mol−1 were observed between motifs II
and III, suggesting both forms would coexist in thermodynamic
equilibrium in aqueous solution. Moreover, a five-coordinate
1:2 uranyl complex with mixed coordination of the two ligands
is slightly more stable than that with pure η2-coordination at the
CCSD(T) level (Figure 16).128 These two motifs are effectively
isoenergetic, suggesting the structure of the adsorbent and
orientation of functions upon uranyl binding are likely to be the
dominant factors in determining how uranyl is bound on an
adsorbent polymer.
Crystal structures obtained for
UO2(acetamidoxime)2(MeOH)2 and the related
Figure 13. (Top) Crystal structure of [UO2(acetamidoxime)4]2+.
(Bottom) Four possible uranyl binding motifs for the amidoximate UO2(benzamidoxime)2(MeOH)2 display the η2-coordination
anion. Representative crystal structures from the CCDC are displayed proposed in the preliminary study, while investigation of the
adjacent. Complexes II (chelate) and III (η2) are displayed as bound in Cambridge Structural Database reveals all known uranyl-
molybdenum complexes. Figure reprinted with permission from ref oximate complexes adopt an identical coordination motif.127
127. Copyright American Chemical Society, 2012. Reported almost simultaneously, a uranyl bound by an
13948 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 14. Representative structural isomers optimized at the B3LYP level. Figure reprinted with permission from ref 127. Copyright American
Chemical Society, 2012.

amidoxime-functionalized imidizolidine also displayed an η2- In contrast, the solid-state configurations of 1:2 uranyl
coordination motif.129 Recently, studies on related 2-pyridyl complexes with acetamidoxime and benzamidoxime attained
ketoximes have also been pursued in an effort to better the coordination number six by having two solvent molecules
understand the fundamentals of uranyl binding. Similar to the (CH3OH) in the equatorial plane.127 Theoretical calculations
aforementioned studies on acetamidoxime and benzamidoxime, (M06 and SMD) were used to determine the most stable
methyl pyridyl ketoxime binds in an η2-configuration as coordination environment in aqueous solution, with compar-
determined through crystallographic studies. However, a ison of relative stabilities between one and two solvent
dinuclear molecule displaying a κ2-binding mode and a peroxo molecules being ≤1 kcal mol−1 and indicating both forms
bridge, [(UO2)2(O2)(OAc)2(L)2] (where L = methyl 2-pyridyl likely coexist in thermodynamic equilibrium in solution.
ketoxime), is obtained when crystallized from methanol in 5.1.3. Cyclic Imidedioxime Functionality. Due to the
direct sunlight and layered with diethyl ether and n-hexane strong nucleophilic nature of hydroxylamine, several side
(Figure 17).131 The authors propose formation of the peroxide reactions may occur. Aldehydes and ketones can be converted
through a photochemical process whereby UO22+ is excited, to oximes, and carboxylic acids can react to form hydroxamic
reduced to UVO2+, and subsequently re-oxidized to UO22+ acid. Although the original carbonyl compound can be restored
forming H2O2 in the process. Similar multinuclear uranyl upon treatment with acid, functional group stability must
complexes were observed during crystallization of glutardiami- nevertheless be considered in adsorbent design.135 In addition
doxime with UO2(NO3)2, where the metal nitrate induced in to reacting with non-nitrile functional groups, important side
situ cyclization of the ligand to form a bridged dinuclear species reactions involving the amidoxime can also occur, most notably,
(Figure 17).132 While DFT and small molecule results are cyclization of adjacent amidoximes to form a cyclic
consistent with each other, it is worth noting they are imidedioxime.
performed under idealized conditions and may not be Astheimer and colleagues were the first to report the
representative of a complex and dynamic system, as is the generation of the cyclic imidedioxime functionality, determin-
case for a polymer adsorbent deployed in seawater. ing “uptake of uranium from natural seawater is closely related
To further understand the amidoxime ligand, other simple to the presence of cyclic imidoxime [sic] configurations” from
amidoxime species were also studied computationally. The- their investigation of cross-linked poly(amidoxime) resins.28
oretical calculations were performed to elucidate the structure Importantly, they indicate adsorbent elution as being a critical
and relative stability of uranyl complexes with acetamidoxime component of seawater uranium recovery and identify the acid-
and benzamidoxime. The most stable geometries of the uranyl instability of the cyclic imidedioxime functionality as respon-
complexes with one and two ligands obtained at the M06/SSC/ sible for the degradation of performance following multiple
6-311++G(d,p) level of theory are shown in Figure 18.133 recycles. Nevertheless, while providing this definitive statement,
Consistent with previous calculations and X-ray diffraction close inspection of the reported data reveal significant
data,127,129,130 the η2 binding with the N−O oximate bond to ambiguity and the following conclusion: “The comparative
uranyl is the energetically preferred coordination mode for all investigation of amidoxime and imide dioxime [sic] resins
amidoxime complexes. As was observed previously,134 the most identified the cyclic imidoxime [sic] structures as being
stable structural configuration for 1:1 uranyl complexes with responsible for the observed instability of poly-
mono and bidentate donor ligands was solely five-coordinate. (acrylamidoximes) in dilute acids, but allowed no discrim-
13949 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 15. Structures of uranyl complexes with formamidoxime (ao), [UO2(ao)x(H2O)y]2‑x, obtained after geometry optimization at the B3LYP/
SSC/6-311++G(d,p) level of theory. Figure reprinted with permission from ref 128. Copyright 2016, Royal Society of Chemistry.

and cyclic imidedioxime binding sites, proposed by Astheimer.


Glutaronitrile was contacted with hydroxylamine at different
temperatures and for different times to afford the cyclic
analogue, glutarimidedioxime (H2A),63 and the open-chain
analogue, glutardiamidoxime (H2B),136 as displayed in Figure
19. Following a thorough investigation of uranyl binding by
H2A, including computational studies, crystallography, as well
as both spectrophotometric and potentiometric titrations, they
determined not only was binding strongly favored but that the
H2A ligand when present at concentrations roughly analogous
to seawater-deployed polymers was competitive with carbonate
for uranium binding.63 While binding by the open-chain H2B
Figure 16. Structures of uranyl complexes with formamidoxime molecule was also determined to be thermodynamically
obtained after geometry optimization at the B3LYP/SSC/6-311+ favored, they concluded it was not capable of competing with
+G(d,p) level of theory. Relative stabilities for the complexes environmental carbonate for uranium recovery,136 nor was a
calculated with coupled cluster corrections at the CCSD(T) level partial hydrolysis product of H2A, glutarimidoxioxime.137
are displayed below. Results are consistent with those obtained by
MP2 optimized geometries. Figure reprinted with permission from ref Work by Hay and colleagues provided further insight into
128. Copyright 2016, Royal Society of Chemistry. cyclic imidedioximes, resolving three fundamental character-
istics involving (1) conditions that maximize conversion of
ination to be made between open-chain amidoxime and cyclic adjacent nitriles to cyclic imidedioxime, (2) whether base-
imidedoxime [sic] groups with respect to uranium uptake from conditioning results in cyclic imidedioxime formation, and (3)
natural seawater”.28 conditions under which cyclic imidedioxime functionalities
Rao and colleagues performed thermochemical titrations of would hydrolyze.138 They noted that the relative synthetic
uranium binding to small molecule analogues of the open-chain yields of open-chain vs cyclic functionalities was a condition of
13950 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 17. (Left) Crystal structure of the dinuclear [(UO2)2(O2)(OAc)2(L)2] (where L = methyl 2-pyridyl ketoxime). The unanticipated presence
of the peroxo bridge and the chelating pyridyl ketoxime ligands make this a particularly interesting uranyl complex affording potentially important
insight into how uranyl is bound by poly(oxime) adsorbents under environmental conditions. Figure reprinted with permission from ref 131.
Copyright 2016, Royal Society of Chemistry. (Right) Crystal structure of the dinuclear [(UO2)2(L)2(NO3)2] (where L = 2,6-diiminopiperidin-1-ol).
The uranyl-binding ligand is formed by an in situ cyclization reaction of glutardiamidoxime prior to crystallization. Figure reprinted with permission
from ref 132. Copyright 2016, Royal Society of Chemistry.

Figure 19. Formation of cyclic imidedioximes (path A, upper) and


open-chain amidoximes (path B, lower) following treatment of
glutaronitrile with hydroxylamine. The small molecule analogues of
these two binding sites are glutarimidedioxime (H 2 A) and
Figure 18. Structures of uranyl complexes with acetamidoxime glutardiamidoxime (H2B), respectively. Reprinted with permission
(acetam) and benzamidoxime (bzam) obtained after geometry from ref 63. Copyright 2012, Royal Society of Chemistry.
optimization at the M06/SSC/6-311++G(d,p) level of theory. Figure
reprinted with permission from ref 133. Copyright 2016, Elsevier.
imino carbon, thereby slowing the nucleophilic attack of water
and subsequent hydrolytic degradation. The addition of an
reaction temperature,139 reporting that the complete trans- aromatic backbone for the cyclic structure stabilized the
formation of glutaronitrile to H2A was achieved in just 3 h at resulting phthalimide dioxime molecule and imparted the
130 °C (Figure 20). Accordingly, treatment of poly(AN) with requisite chemical stability.138 A general summary of cyclic
hydroxylamine at elevated temperatures is the most straightfor- imidedioxime formation and degradation is provided in Figure
ward avenue for generating a maximum amount of cyclic 21.
imidedioxime sites. When glutarimidedioxime was investigated Recently these aforementioned studies were tested through
under basic conditions, glutarate was formed with no development of adsorbents possessing maximized cyclic
appreciable generation of H2A observed. Similarly, exposure imidedioxime functionalities. An adsorbent polymer was
of H2B to 1 M NaOD at various temperatures also afforded formed by graft copolymerization of AN and itaconic acid
glutarate as the sole reaction product (Figure 20). Finally, from a polyethylene trunk fiber, followed by conversion of the
investigation of H2A and H2B under acidic conditions revealed nitrile into amidoxime by treatment with hydroxylamine in a
that, while the open-chain H2B was stable in 1 M DCl for water/methanol solution at 80 °C.77 Adsorbents were then
several months at room temperature, H2A degraded rapidly heated at various temperatures (110−150 °C) in polar aprotic
through a glutarimide intermediate in less than 20 h. An acid- solvents to convert open-chain amidoxime groups into the
resistant form could be produced by lowering the charge on the desired cyclic imidedioxime functionalities. Chemical speciation
13951 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 20. NMR data revealing glutarimidedioxime (H2A) formation (left) in DMSO-d6 at 130 °C and degradation (right) in 1 M NaOD at 80 °C
as a function of time. (Left) Peaks denoted with a blue circle correspond to H2B while peaks with a red star correspond to H2A. (Right) Spectrum
(h) is H2A in D2O, provided for comparison. The circled region in spectrum (b) suggests only trace formation of H2A under basic conditions. Figure
reprinted with permission from ref 138. Copyright 2012, American Chemical Society.

Figure 21. Routes of conversion of poly(acrylamidoximes) into poly(acrylimidedioximes) and related degradation pathways and products. Reprinted
with permission from ref 138. Copyright 2012, American Chemical Society.

was investigated by 13C cross-polarization magic angle spinning regarding what amidoxime species is present on the adsorbent
NMR spectroscopy (13C CP-MAS NMR), which suggests and how uranyl is actually bound.
successful conversion of all adjacent amidoximes to the cyclic The adsorbents were investigated for uranium uptake from
imidedioxime,77 consistent with the aforementioned work of an 8 ppm uranyl nitrate brine solution, also containing 193 ppm
Hay.138 An important point is that, assuming the 13C CP-MAS sodium bicarbonate and 25 600 ppm sodium chloride. The
NMR peak characterizations are correct, treatment of poly- sample that was not heat-treated achieved an uptake of
(AN) with hydroxylamine at comparatively mild temperatures approximately 175 mg of uranium/g of adsorbent, while
and protic solvents results in formation of the cyclic maximum uptake for samples receiving post-amidoximation
heat-treatment in DMSO, DMF, and propylene carbonate
imidedioxime functionality as the predominant species on the
achieved uptake of approximately 225, 220, and 185 mg of
adsorbent, even prior to subsequent heat treatment in polar uranium/g of adsorbent, respectively. Further investigations
aprotic systems (Figure 22). It is essential to note that were performed using adsorbents heated in DMSO, which
previously reported 13C NMR analysis performed by Seko and displayed more rapid uranyl-binding kinetics compared to the
colleagues identified the upfield shoulder as attributable to the untreated material and improvement in recovery of uranium
cyclic imidedioxime, rather than the open-chain amidoxime following 56-days of contact with seawater from Sequim Bay,
(Figure 23),39 with degradation observed upon treatment with WA in flow through columns. A seawater uranium recovery
HCl, consistent with work reported by Hay and colleagues.138 capacity of approximately 4.2 mg of uranium/g of adsorbent
Such conflicting results continue to feed ongoing controversy was achieved for the DMSO-treated sample, while 3.5 mg of
13952 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 22. 13C CP-MAS NMR data of poly(AN-co-itaconic acid) (a) before treatment with hydroxylamine, (b) after treatment with hydroxylamine
in 50:50 methanol:water at 80 °C, and (c) after subsequent heating at 130 °C in DMSO. A slight shoulder is assigned to the 13C NMR signature of
the open-chain amidoxime which largely disappears following heating in polar aprotic solvents, suggesting the near-complete formation of the cyclic
imidedioxime functionality. Figure reprinted with permission from ref 77. Copyright 2016, Elsevier.

uranium/g of adsorbent were obtained for the untreated technologically mature, while ATRP is much more recent and
analogue. However, increases in uptake were also observed in has provided the best-performing synthetic polymer in the
the binding of numerous other metals present in seawater, in literature. Materials prepared by these two techniques are
particular Ca, Mg, Fe, V, Mn, and Cr.77 Taken comprehen- covered in their own individual sections. There are also
sively, it is clear the cyclic imidedioxime plays a critical role in numerous adsorbents prepared by neither RIGP nor ATRP
the performance of amidoxime-functionalized polymer adsorb- which constitute significant advances in the field and change the
ents, though achieving the necessary selectivity and chemical way polymer adsorbents are prepared or understood. Such
stability for repeated use remains an elusive and highly desirable materials are discussed in a third category, “Polymer
objective. Adsorbents Prepared by Neither RIGP Nor ATRP”, as they
5.2. Polymerization Methods do not possess a common polymerization technique. While it is
customary that such a category would follow sections on RIGP
5.2.1. Introduction. Polymer-based materials comprise and ATRP, it is also where the report of the first amidoxime-
perhaps the most important category of material for uranium functionalized polymers is categorized, and so it is the section
recovery from seawater, spanning a diverse range of in which this portion of the review begins.
compositions and forms, and the ensuing publications afford 5.2.2. Polymer Adsorbents Prepared by Neither RIGP
a wealth of knowledge to the disciplined reader. Nevertheless, Nor ATRP. 5.2.2.1. Suspension/Emulsion Polymerization.
identifying a coherent and logical format for the introduction The first amidoxime-functionalized polymer-based adsorbents
and discussion of this volume of information is challenging, as were prepared by thermally initiated suspension polymerization
many of the works discuss related concepts, while seemingly of (AN) and divinylbenzene (DVB) dissolved in toluene and
orthogonal synthetic techniques developed independently can dispersed in water (Figure 24).24−26 As discussed above,
dovetail with surprising ease and promising outcome. In an subsequent treatment of the poly(AN-co-DVB) with hydroxyl-
effort to systematically and coherently survey the pertinent amine generates the amidoxime (AO) functionality and the
literature that defines this research area, polymers are final poly(AO-co-DVB) adsorbent. An identical polymerization
categorized and presented based on the polymerization approach was implemented by Hirotsu, who investigated the
technique used for their synthesis. Within each category, the role of the hydrophilicity of the cross-linking comonomer on
report of new materials is arranged chronologically, but with adsorbent performance. Substituting tetraethylene glycol
complete development of each system presented before moving dimethacrylate (4EGDM) for divinylbenzene resulted in an
on to the next new material. Pertinent monomers are presented adsorbent with superior kinetics compared to poly(AO-co-
as figures at the beginning of each section. DVB). A 40-day test of the best performing poly(AO-co-DVB)
Many different polymerization approaches have been used afforded an uptake of approximately 0.2 mg of uranium/g of
for the preparation of uranium adsorbents; however, RIGP has adsorbent, while the poly(AO-co-4EGDM) achieved 0.4 mg of
afforded the majority of materials and is the most uranium/g in only 28 days.140 Follow up work with poly(AO)
13953 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

dimethacrylate, followed by postsynthetic hydrolysis of the


carbamate functionality. These amines could then be chemically
reacted to afford specific chelating ligands, such as iminodi-
acetic acid, aminophosphoric acid, thiourea, and dithiocarba-
mate. While capacities were not reported, X-ray microanalysis
revealed uranium was bound exclusively at the surface of the
beads, rather than homogeneously throughout.142,143
A St/AN copolymer was prepared by emulsion polymer-
ization with potassium persulfate radical initiator and a sodium
lauryl sulfate emulsifier.144 Following polymerization, the
nitriles were converted to amidoxime groups. Different St:AN
ratios were used to prepare materials (40:60, 30:70, 20:80),
which were subsequently screened for uranium uptake from a
45 ppm solution at pH 2.2, achieving capacities as high as 0.85
mg of uranium/g of adsorbent. Introduction of competing
metals, increasing ionic strength, or pre-exposure of the
adsorbent to solutions containing 1.4 ppm chlorine reduced
the performance. Nevertheless, when 35.8 L of filtered
environmental seawater was passed through 2 g of adsorbent,
an uptake of 0.0145 mg of uranium/g of adsorbent was realized,
equating to 25% the total quantity of uranium dissolved in the
seawater.
5.2.2.2. Anionic Polymerization. Kabay reported the
Figure 23. 13C NMR results displaying the change in spectrum application of anionic polymerization to adsorbent develop-
following treatment with HCl and subsequent regeneration with an ment in the mid-90s, where methacrylonitrile was polymerized
unspecified alkaline solution. Spectrum (a) displays a poly- with the anionic initiators BuLi or Et2Mg. The resulting
(acrylamidoxime-co-methacrylic acid) grafted poly(ethylene)/poly- polymers possessed high molecular weight, ranging from 105 to
(propylene) adsorbent following amidoximation by treatment with 106 g mol−1. Upon conversion of the nitrile group to AO,
hydroxylamine. Spectrum (b) displays the same adsorbent following pretreatment with 1 M NaOH, and subsequent neutralization,
contact with HCl, while spectrum (c) displays regeneration by the best performing material achieved an uptake of 0.176 mg of
treatment with an alkaline solution. Reprinted with permission from uranium/g of adsorbent/day deployment in 5 L of natural
ref 39. Copyright 2005, Taylor & Francis.
seawater.145
5.2.2.3. Chemically-Initiated Polymerization. Adsorbents
composed of poly(acrylamide-co-methylenebis(acrylamide))
(poly(AcA-co-MBA)) were synthesized using benzoyl peroxide
as a radical initiator to generate fully water insoluble polymers
(Figure 25).146 The amides were then converted to hydroxamic

Figure 24. Relevant monomers for suspension/emulsion polymer-


ization and corresponding abbreviations in the text.

adsorbents cross-linked with different ratios of DVB/4EGDM


revealed by X-ray microanalysis that uranium was more
homogeneously distributed throughout the adsorbent as more
4EGDM was incorporated, whereas binding in the poly(AO-co- Figure 25. Monomers and cross-linkers of relevance to chemically
DVB) was localized to the surface. While the DVB cross-linked initiated polymerization.
adsorbent displayed the lowest capacity and slowest kinetics,
the 100% 4EGDM adsorbent was only marginally better. The
materials with the best kinetics were composed of different acid through treatment with hydroxylamine. Different ratios of
ratios of DVB/4EGDM, indicating hydrophilicity of the AcA:MBA were investigated, including 95:5, 85:15, and 75:25.
adsorbent is not the only consideration for effective uranium Uranyl binding was investigated from pH 1−11, with maximum
uptake. The poly(AO-co-4EGDM) nevertheless displayed an performance observed at pH 5 and 50−60% maximum
impressive 4.5 mg of uranium/g uptake when deployed in a performance at pH 8. The best performing material contained
glass column and contacted with seawater for 180 days at 25 °C the smallest quantity of MBA, and higher moisture uptake was
and 250 mL min−1.141 observed to correlate with higher metal uptake. A modest
Morcellet and co-workers also used suspension polymer- uranium uptake of 0.018 mg of uranium/g of adsorbent is likely
ization to generate a series of porous polymers containing attributable to the intense hydrophobicity of the adsorbent.
vinylamine units through suspension copolymerization of N- Similar performance was reported by an insoluble cross-linked
vinyl t-butyl carbamate, styrene (St), and ethylene glycol r es i n f o r m ed f r o m p o l y ( N - ( 3 - d i m e t h y l a m i n o ) -
13954 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

propylmethacrylamide), formed by radical polymerization.147 105 mL g−1 was more than three times greater than that for V
While the aforementioned adsorbents are unlikely to be directly and more than 150× greater than the other competing ions.
applicable for seawater uranium recovery, the work does This selectivity was rationalized by the authors as attributable to
provide a good negative result corroborating the need for both intramolecular H-bonding promoting formation of the syn-
hydrophilicity and appropriate uranyl-binding receptors. hydroxyamino form, where metal-assisted hydroxyl deprotona-
Dipropionitrile acrylamide was polymerized through free tion would occur prior to bonding and be governed by hard/
radical polymerization by application of the chemical initiator soft acid/base theory.148
2,2′-azobis(2-methylpropionitrile), more commonly known as This adsorbent was investigated in seawater collected from
AIBN (Figure 26). Following postsynthetic conversion to the the coast of Mutsu Sekine-Hama in the Aomori Prefecture in a
recirculated column configuration at a flow rate of 6 mL min−1.
After 1 h, 1 day, and 1 week of contact, the adsorbents were
eluted with HCl and the uranium uptake determined, affording
values of 13.1, 28.1, and 27.1 mg of uranium/g of adsorbent.148
While these results are remarkably good, it is perplexing that
such technology has not been successfully commercialized. It is
suggested that the results be externally validated to confirm the
performance and ensure the results are not a typographical
error. In comparison to other adsorbents contacted under
similar conditions, results in the μg of uranium g−1 adsorbent
are more commonplace.
Compolymers of AA, AcA, and trimethylolpropane triacry-
late (TMPTA) were formed with an ammonium persulfate
initiator, with the resulting nonamidoxime-functionalized
adsorbents used to extract uranyl from a 5% NaCl solution.45
Performance was tested as a function of time, temperature, pH
Figure 26. Synthetic scheme depicting the preparation of the and ionic strength. Saponification of the polymer with NaOH
poly(N,N′-dipropionitrile acrylamide) and subsequent amidoximation afforded carboxylate groups, which are the putative uranyl-
through treatment with hydroxylamine. Reprinted with permission binding species in the adsorbent and also increase adsorbent
from ref 148. Copyright 2005, Springer. hydrophilicity. The swelling of the hydrogel increased upon
formation of carboxylic functionalities, though both swelling
amidoxime form, the adsorbent displayed a high capacity for and adsorbent performance decreased when contacted with
uranium and V over Cu2+, Co2+, and Ni2+ competing ions. electrolytes possessing high ionic strength, such as seawater.
Calculation of the Kd for each metal, i.e., the ratio of metal ions Swelling also decreased with increasing cross-linking, afforded
in the solution to the metal ions bound in the adsorbent, reveal by a greater molar quantity of TMPTA. A maximum uranium
strong selectivity for U. The Kd value for uranium of nearly 2 × uptake of 250 mg of uranium/g of adsorbent was achieved from

Figure 27. General scheme depicting synthesis of an IIP. Figure reprinted with permission from ref 150. Copyright 2013, Springer.

13955 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

contact of 0.15 g of material with 50 mL of a 502 ppm uranyl Shamsipur and colleagues later reported preparation of IIPs
solution at pH 7. grafted on the surface of a silica gel adsorbent, with the IIPs
Ion-imprinted polymers (IIPs) have also been prepared polymerized from surface-bound azo groups.151 Polymerization
through chemical polymerization. This type of material is was performed in the presence of a MA-uranyl complex with
prepared by first coordinating uranyl with chelating monomers EGDMA, followed by elution of the uranium. Adsorption
and then polymerizing the complex followed by elution of the studies were performed by adding 100 mg of adsorbent to 10
bound uranium (Figure 27). For example, Rao and co-workers mL of an aqueous 5 ppm uranyl solution over pH 1.5−4. The
dissolved stoichiometric quantities of uranyl nitrate with select best uptake was achieved at pH 3, where 100% of the uranium
chelating ligands, including salicylaldoxime, catechol, succinic was extracted affording an uptake in excess of 0.5 mg of
acid, 5,7-dichloroquinoline-8-ol (DCQ), and 4-vinylpyridine uranium/g of adsorbent. Significant selectivity was displayed for
(VP), in 2-methoxy ethanol, followed by polymerization in the uranyl over a wide range of di- and trivalent cations, including
presence of 2-hydroxyethyl methacrylate (HEMA) and ethyl- Cu(II), Ni(II), Fe(III), and Mn(II), among others. The
ene glycol-dimethacrylate (EGDMA) with an AIBN initiator adsorbent also achieved efficient recovery of uranium from
(Figure 28).149 Following elution of the bound uranium, spiked environmental samples, though recovery of uranium
from tap water and Caspian Sea water were below limits of
quantification for their analytical method.
5.2.2.4. UV-Activated Polymerization. Pandey and co-
workers prepared AO-functionalized polymeric adsorbents
through both UV-activated and electron-beam radiation
induced graft polymerization (RIGP, discussed in the
subsequent section) of AN on polypropylene (PP) fibers and
microporous sheets (Figure 29). The UV initiator 2,2-

Figure 29. Comonomers pertinent to UV polymerization.


Figure 28. Relevant monomers, cross-linkers, and chelating groups
used in IIP synthesis for recovery of seawater uranium. dimethoxy-2-phenylacetophenone (DMPA) was dissolved in
the polymerizing solution containing the PP support, MBA
cross-linker, and AN, with the resulting mixture exposed to 365
adsorbents prepared with succinic acid, DCQ, and VP afforded nm UV light for 20 min. Only 120−130 wt % Dg could be
enrichment of uranium recovery in seawater simulant when achieved by either UV-initiated polymerization or RIGP, and
contacted at 0.02 g of resin per 1 L of seawater. While capacities adsorbents prepared by both routes displayed similar wt %
of up to 33 mg of uranium/g of adsorbent were achieved during water uptake for a common support and comparable Dg. While
screening in uranyl-spiked DI water, the nonimprinted controls a uranium uptake was not provided for the UV-prepared
also achieved up to 80% of this capacity. Good selectivity was adsorbents, the RIGP-prepared adsorbents with similar Dg and
reported for these materials, achieving >99% extraction of water uptake achieved a capacity of 380 mg of uranium g−1
uranium in the presence of Th(IV), Fe(III), Co(II), Ni(II), when contacted with a 10 ppm solution of 233U in seawater.46
Mn(II), Cu(II), and Zn(II), while adjustment of pH had no Biofouling-resistant polymer adsorbents were also prepared
effect on performance. When Na2CO3 was added to the by UV-initiated photopolymerization, where poly(ethylene
screening solution, recovery for the succinic acid/VP adsorbent glycol methacrylate phosphate) was polymerized in solution
was only 25%, while 99% recovery was still achieved by the with Ag nanoparticles and subsequently reinforced with a
DCQ/VP material. The DCQ/VP IIP was tested for uranium thermally bonded nonwoven PP fibrous sheet. As mentioned
extraction from Arabian Seawater, recovering 1.59 μg of earlier, a uniform distribution of Ag nanoparticles in the
uranium/L of seawater and an uptake value of 0.08 mg of adsorbent was observed, and when incubated with E. coli, no
uranium/g of adsorbent. growth was observed on the adsorbent. This suggests inclusion
Later work by Shen et al. reported formation of an IIP from of Ag nanoparticles imparts some significant resistance to
the uranyl complex with 2,4-dioxypentan-3-yl methacrylate, biofouling. When contacted with a 1788 ppm uranyl solution in
copolymerized with EGDMA in 1,4-dioxane with AIBN.150 2,4- seawater, a capacity of 550 mg of uranium/g of adsorbent was
Dioxypentan-3-yl methacrylate was synthesized to afford both a obtained, though large quantities of other cationic metals were
β-diketone structure and a methacrolyl group, incorporating the also extracted.89
ability for cross-linking with a uranyl receptor possessing a pKa 5.2.2.5. Chemical Modification or Functionalization of a
close to seawater pH. An uptake of approximately 24 mg of Pre-Synthesized Polymer. Suh and colleagues functionalized
uranium/g of adsorbent was achieved from an aqueous solution Sephadex G-25 resin with a metal-binding ligand composed of
of UO2(NO3)2, while 15.3 mg of uranium/g of adsorbent was 2,2′-dihydroxyazobenzene (DHAB) attached to poly-
recovered from an aqueous 200 ppm uranyl solution containing (ethylenimine) (PEI; Figure 30).38 This work followed
carbonate and at pH 9. The IIP was studied for the selective previous studies by the same group where DHAB was attached
adsorption of uranium from an aqueous solution with to a chloromethylated cross-linked poly(St) support,36
numerous monovalent cations of varying ionic radii, affording chloromethylated poly(St-co-DVB),37 and pure PEI.152 In the
87−140× selectivity for uranium. former references, the chloromethyl group could also be
13956 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

amounts to a 72% uranium recovery efficiency and an uptake of


0.0042 mg of uranium/g of adsorbent. Nevertheless, the
extremely high phase ratio of 500 mg L−1 is unreasonable for
realistic seawater deployment conditions.
Sellin and Alexandratos reported the application of a series of
polymer-supported primary amines for uranium extraction
(Figure 32).154 Poly(VBC-co-DVB) (VBC = vinylbenzyl
chloride) gel beads were synthesized as a support, followed
by postsynthetic functionalization with an array of amines
possessing different lengths, structures, and substituents. HCl
titrations revealed comparable quantities of amine sites with
capability for uranyl binding, while uranium uptake was
analyzed from a synthetic seawater spiked to 50 ppm with
uranyl. On a mass- uranium per mass-adsorbent basis, the
short-chain primary amine (pA) afforded the maximum uptake
of 14.8 mg of uranium/g of adsorbent. However, as noted by
the authors a mmol of uranium per mol of ligand affords a more
appropriate comparison in light of the added mass from more
Figure 30. Comonomers relevant to chemical modification of expansive amine functionalities. By this metric, the adsorbent
presynthesized polymers. functionalized with pTEPA achieved the best uptake of 17.7
mmol of uranium mol−1, followed closely by pTAEA (16.0
converted to a quaternary ammonium function by treating with mmol mol−1) and then by pDETA and pEDA (13.3 and 13.2
various tertiary amines.36,37 In this work, spectrophotometric mmol mol−1, respectively). These uptake values also correlate
titration experiments revealed 1:1 complexes of various metals, reasonably well with the number of amine sites in the
including uranium, with PEI-bound DHAB (Figure 31).38 After adsorbent, with values of 8.51, 8.43, 7.10, and 6.61 mmol of
N/g of polymer for pTEPA, pDETA, pTAEA, and pEDA,
respectively. A poly(DVB-co-VBC) functionalized with dia-
midoxime was also synthesized and investigated under the same
conditions as a control. A comparatively modest uranyl capacity
of 2.34 mg of uranium/g of adsorbent was achieved, while the
adsorbent has 10.4 mmol g−1 protonation sites, converting to
3.79 mmol of uranium/mol of binding site.
A particularly interesting follow-up work from the same
group involved formation and performance of bifunctional
Figure 31. Proposed uranyl complex formed by a DHAB ligand. amidoxime fibers synthesized by functionalization of commer-
cially available poly(AN) fibers (Figure 33).155 A series of
grafting to the Sephadex resin with cyanuric chloride as a adsorbents were prepared by first attaching various derivatives
coupling reagent, the adsorbent was investigated for recovery of of diethylenetriamine (DETA) to the poly(AN) by simple
metals from seawater. Following exposure to flowing seawater amination chemistry. To prepare bifunctional versions, the
for 20 h, 0.018 mg of uranium/g of adsorbent were extracted, residual nitrile species were converted to amidoxime by
compared to 1.3 mg of Fe/g of adsorbent and 0.47 μg of Zn/g treatment with hydroxylamine. To provide an assessment of
of adsorbent. As noted by the authors, the low uranium uptake relative performance, uranium was loaded onto the fibers by 1-
and high affinity for competing ions indicates the Sephadex- week contact with a synthetic seawater spiked with uranyl at
supported DHAB-PEI does not meet the criterion for concentrations ranging from 0.9 to 50 ppm, followed by a
economically feasible recovery of seawater uranium. second 1-week contact with a fresh uranyl-spiked synthetic
Ramkumar reported the chemical oxidation of aniline to seawater solution of the same concentration. Samples were also
generate polyaniline, followed by its application in recovery of investigated following 21-days contact with filtered environ-
uranium from seawater.153 Due to incorporation of anions, such mental seawater in a flow-through column at Pacific Northwest
as Cl− or SO42−, in the polymer during synthesis, it was National Laboratory (PNNL).61
expected anionic exchange with anionic uranyl complexes could While DETA- and DETA-derivative-functionalized poly(AN)
be achieved. Using ethylenediaminetetraacetic acid (EDTA) as adsorbents displayed some affinity to uranium, inclusion of
a metal complexant to form an anionic uranyl species, a modest amidoxime functionalities dramatically improved the uptake.
affinity for uranium was reported (40% in solution), which was Uranium was completely removed from 0.9 ppm uranyl
further suppressed by the presence of other competing metals. solutions, 97−99% removed from 6 ppm solutions, and
In contrast, substitution of EDTA for SCN− or CO32− resulted approximately 95% removed from the 50 ppm solution. In
in a marked improvement in performance. For SCN−, 85% of comparison, a pure poly(amidoxime) adsorbent prepared from
the uranium was extracted from solution (albeit, at pH 1), while the same poly(AN) precursor achieved similar performance at
95% of uranium was extracted in a carbonate solution. 0.9 and 6 ppm, but only extracted 82.5% of the uranium from
Furthermore, at pH 5.5 uranium could be selectively separated the 50 ppm solution. Seawater-contacted results also reveal
from solutions containing up to 100× excess of competing superior performance for the bifunctional adsorbents, with
metal ions. When investigated for recovery of uranium from AO−DETA extracting 0.54 mg of uranium/g of adsorbent and
seawater, 200 mL of seawater was contacted with 100 mg of AO-phon-DETA recovering 0.79 mg of uranium/g of
adsorbent for 4 h, with 0.42 μg of uranium recovered. This adsorbent. Pure poly(amidoxime) recovered 0.12 mg of
13957 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 32. Series of amines investigated on poly(DVB-co-VBC) supports for uranyl binding. Figure reprinted with permission from ref 154.
Copyright 2013, American Chemical Society.

Figure 33. DETA and derivatives used to functionalize poly(AN) adsorbents for recovery of seawater uranium. Figure reprinted with permission
from ref 155. Copyright 2016, American Chemical Society.

uranium/g of adsorbent, while the nonamidoximated phon- are needed to determine whether the enhanced performance is
DETA extracted a comparable 0.14 mg of uranium/g of attributable to true activation of the amidoxime group or is
adsorbent. This work demonstrates that affinity of amidoxime- primarily the sum of individual binding from multiple
functionalized adsorbents for uranyl can be enhanced by functionalities.155
incorporation of auxiliary DETA ligands. While detailed FTIR Chitin, a biopolymer, was also used as a chemically modified
investigations by the authors support activation of the polymer support for the recovery of uranium.156 Rogers and co-
amidoxime by the amine, they assert additional experiments workers started with shrimp shell waste, extracting chitin into
13958 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

an ionic liquid composed of 1-ethyl-3-methylimidazolium Fe3+, Pb3+, and Ni2+) reduced adsorption efficiency from 4−
acetate. The chitin was wet-jet spun into fibers which were 30% at pH 6, and a modest decrease in performance was
subsequently deacetylated with NaOH to form primary amine observed over five recycles.157
groups on the surface. Treatment with hydroxylamine Related work by the same group reported preparation of
appended hydroxamic acid functionalities for seawater uranium bifunctional poly(ionic liquid) microspheres by applying a
recovery. Approximately 2.5 mg of wet fibers were contacted similar process.158 Block copolymers of poly(N-vinylimida-
for 144 h with 1 mL of DI water spiked with a 233U radiotracer, zole)-block-poly(St) were prepared by RAFT. The poly(N-
affording a final uptake of approximately 0.28 mg of uranium/g vinylimidazole) could be transformed into a quaternary amine
of adsorbent.156 Although the performance is underwhelming, bromine by treatment with bromoalkylnitrile, thus forming the
and poor fiber integrity is noted by the authors of the poly(ionic liquid), and the nitrile subsequently transformed
publication, this work constitutes a particularly innovative into AO by treatment with hydroxylamine. Adsorbent micro-
approach to the preparation of novel uranium adsorbents, as spheres were then prepared by emulsion polymerization of St/
well as the effective utilization of a biological waste stream. DVB with the block copolymer. Varying the bromoalkylnitrile
Assuming adequate performance can be achieved, additional species changes the number of carbons between the amidoxime
benefits from this class of polymer include the expectation of and the quaternary amine, with adsorbents prepared with n = 1,
mild environmental impact, as well as access to a hydrophilic 3, and 6 carbons. Poly(AO) and poly(St) blocks could also be
trunk material capable of facilitating transport of water through finely controlled during RAFT polymerization, affording
the adsorbent. poly(AO) blocks ranging from 40−82 units and poly(St)
5.2.2.6. RAFT Polymerization. More recently, block blocks ranging from 9−11 units. A synthetic scheme is
copolymer adsorbent precursors of poly(AN)-block-poly(St) displayed in Figure 36.
have been prepared by reversible addition−fragmentation
transfer (RAFT) polymerization (Figure 34). The poly(AN)

Figure 34. Comonomers of relevance to adsorbents prepared by


RAFT polymerization.

block can then be converted to poly(AO) with hydroxylamine, Figure 36. Synthesis of poly(ionic liquid)-functionalized microspheres.
and the resulting poly(AO)-block-poly(St) can then be further St and DVB are as defined in the text. Reprinted in part with
polymerized into microspheres via emulsion polymerization of permission from ref 158. Copyright 2015, Royal Society of Chemistry.
the poly(St) block with additional St or DVB (Figure 35).157
This approach affords fine-tunability of the AO chain length
through controlling the poly(AN) and poly(St) blocks, while To investigate uranium uptake performance, 18 mg
the AO concentration can be modulated through varying the adsorbents were deployed in 200 mL of 12 ppm aqueous
concentration of the poly(AO)-block-poly(St) in the emulsion uranyl solution. For the adsorbent prepared from poly(AO)40-
polymerization step. block-poly(St)9 with a propyl spacer, an adsorption optimum of
Performance of the poly(AO)-block-poly(St)-co-poly(St/ approximately 65 mg of uranium/g of adsorbent was obtained a
DVB) was analyzed by deploying 15 mg particles in a dialysis pH 7, but a statistically insignificant decrease was observed for
bag, immersed in 200 mL of uranyl solution. A maximum pH 8. As the poly(AO) block was increased, leaving the carbon
capacity of 85 mg of uranium/g of adsorbent was obtained in a bridge between the zwitterion and AO function unchanged, an
6 ppm uranyl solution at pH 6.5, while the analogous uptake at uptake maximum of 170 mg of uranium/g of adsorbent was
the more seawater-relevant pH 8 was approximately 60 mg of obtained when n = 82. Additionally, while performance was
uranium/g of adsorbent. A saturation capacity of 247 mg of comparable for the carbon bridge of 1 and 3 atoms, distancing
uranium/g of adsorbent was obtained via an isotherm at pH the AO functionality from the quaternary ammonium bromide
6.5. Introduction of competing cations (Na+, Mg2+, Zn2+, Co2+, by 6 carbons resulted in a significant decrease in performance.

Figure 35. Synthesis of [poly(AO)-block-poly(St)-co-poly(St/DVB)] nanoparticle adsorbents for uranium adsorption. APS is the ammonium
persulfate initiator. PS-b-PAO is the poly(AO)-block-poly(St) nanoparticle. DVB and St are as defined in the text. Figure reprinted with permission
from ref 157. Copyright 2015, Springer.

13959 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 3. Performance of Adsorbents Prepared by neither RIGP nor ATRP for Recovery of Uranium from Water and Seawatera
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
Suspension/Emulsion Polymerization
poly(AO-co-DVB) 0.2 40 d column 140
poly(AO-co-4EGDM) 0.4 40 d column 140
4.5 180 d column 141
poly(St-co-AO) 0.85 45 ppm uranium, pH 2.2 0.0145 column; 35.8 L fixed 144
volume
0.2 45 ppm uranium, 0.598 M NaCl, pH 2.2 144
Anionic Polymerization
poly(methamidoxime) 0.176 1d batch; 5 L fixed 145
volume
Chemically-Initiated Polymerization
poly(AcA-co-MBA) 61.3 1% uranyl solution, pH 8 0.018 column; 50 L fixed 146
volume
amidoximated poly 13.1 1h column, recirculated 148
(dipropionitrile acrylamide) volume
28.1 1d column, recirculated 148
volume
27.1 7d column, recirculated 148
volume
poly(AA-co-AcA-co-TMPTA) 250 502 ppm uranium, pH 7 45
DCQ-VP IIP 0.5 10 ppb uranium, pH 7 0.08 batch, 1 L fixed 149
volume
0.42 10 ppb uranium, pH 8, with NaCl, KCl, 149
CaCl2, MgCl2, and transition metals
succinic acid-VP IIP 0.5 10 ppb uranium, pH 7 149
0.13 10 ppb uranium, pH 8, with NaCl, KCl, 149
CaCl2, MgCl2, and transition metals
sailcylaldoxime-VP IIP 0.35 10 ppb uranium, pH 7 149
catechol-VP IIP 0.32 10 ppb uranium, pH 7 149
2,4-dioxypentan-3-yl 15.3 200 ppm uranium, pH 9, with Na2CO3 150
methacrylate-co- EGDMA IIP
IIP-grafted silica gel 0.5 5 ppm uranium, pH 3 below analytical limit 151
of quantification
UV-Activated Polymerization
Ag NP@poly(ethylene glycol 550 1788 ppm uranium in seawater 89
methacrylate phosphate)
Chemical Modification/Functionalization of a Pre-Synthesized Polymer
DHAB/PEI@Sephade x G-25 0.018 20 h column 38
polyaniline >90% initial uranium concentration not provided in 0.0042 4h batch, 200 mL fixed 153
text volume
pA@Poly(VBC-co-DVB) 14.8b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
b
pEDA@Poly(VBC-co-DVB) 10.4 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pDETA@Poly(VBC-co-DVB) 8.89b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pTEPA@Poly(VBC-co-DVB) 7.17b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pTAEA@Poly(VBC-co-DVB) 6.78b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
b
pMA@Poly(VBC-co-DVB) 0.14 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pDMA@Poly(VBC-co-DVB) 0.09b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pDAP@Poly(VBC-co-DVB) 6.05b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pDAB@Poly(VBC-co-DVB) 6.62b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
pPip@Poly(VBC-co-DVB) 0.97b 50 ppm uranium, pH 8, in artificial seawater 154
(beads)
DETA@poly(AN) 0.02b 900 ppb uranium, pH 8, in artificial seawater 155
poly(AO) 0.09b 900 ppb uranium, pH 8, in artificial seawater 0.124 21 d column, continuously 155
replenished
phon- DETA@poly(AN) 0.07b 900 ppb uranium, pH 8, in artificial seawater 0.142 21 d column, continuously 155
replenished
phin- DETA@poly(AN) 0.017b 900 ppb uranium, pH 8, in artificial seawater 155

13960 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 3. continued
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
Chemical Modification/Functionalization of a Pre-Synthesized Polymer
AO−DETA@poly(AN) 0.09b 900 ppb uranium, pH 8, in artificial seawater 0.535 21 d column, continuously 155
replenished
AO-phon- DETA@poly(AN) 0.09b 900 ppb uranium, pH 8, in artificial seawater 0.789 21 d column, continuously 155
replenished
hydroxamic acid @ chitin 0.28 116 ppm uranium 156
(fibers)
Reversible Addition−Fragmentation-Transfer (RAFT) Polymerization
poly(AO)-block-poly(St)-co- 85 6 ppm uranium, pH 6.5 157
poly(St/DVB)
60 6 ppm uranium, pH 8 157
247c pH 6.5 157
45−80d 6 ppm uranium, pH 6.5, with competing 157
mono-, di-, and trivalent cations
poly(AO)x-block-poly(St)y
x = 40, y = 9, propyl bridge 65 13.5 ppm uranium, pH 7 158
x = 40, y = 9, propyl bridge 102c pH 7 158
x = 58, y = 9, propyl bridge 127c pH 7 158
x = 82, y = 9, propyl bridge 170c pH 7 158
a
Unless otherwise stated, adsorption experiments were performed in DI water and pH was unadjusted. bContacted in ASTM D1141 artificial
seawater; composition: NaCl (24.53 g L−1), MgCl2 (5.20 g L−1), Na2SO4 (4.09 g L−1), CaCl2 (1.16 g L−1), KCl (0.695 g L−1), and NaHCO3 (0.201
g L−1). cMaximum capacity at saturation. dCompeting cation composition: 2.7 × 10−4 M of Fe3+, Na+, Mg2+, Co2+, Ni2+, Zn2+, or Pb2+.

Figure 37. Generic scheme depicting RIGP of AN onto a PE trunk, followed by transformation to amidoxime by contact with hydroxylamine.
Reprinted from ref 92. Copyright 2012, American Chemical Society.

The effect of including the quaternary amine was also of RIGP by Omichi and colleagues in 1985,159 surpassed in
investigated by preparing a control sample without the AO importance perhaps only by the introduction of the amidoxime
functionality, as well as one without the cationic species. In functionality.24−26 The publication reports the preparation and
contrast to an uptake of 60 mg of uranium/g of adsorbent for performance of a fibrous polymer adsorbent functionalized by
the comparable poly(ionic liquid) microsphere with both AO RIGP with poly(AN) followed by amidoximation with
and quaternary amine, the charged adsorbent without the AO hydroxylamine (Figure 37). The work was motivated by the
achieved only 22 mg of uranium/g uptake, suggesting up to poor mechanical stability of other fibrous adsorbents where the
36% of the uranium uptake could be attributed to Coulombic nitrile precursor is homogeneously distributed through the
interactions. Additionally, while the microsphere without the adsorbent, such as poly(AN). Polymeric substrates, such as
quaternary amine displayed almost identical uptake perform- poly(ethylene) (PE) or poly(propylene) (PP) (Figure 38), can
ance (54 mg of uranium/g for the neutral adsorbent and 59 mg
of uranium/g for the cationic adsorbent), slower adsorption
kinetics were observed. This interesting adsorbent system may
show promise, especially if it can be efficiently translated to a
platform more amenable to large scale, open ocean deploy-
ment. Moreover, the inherent tunability of the adsorbent is
particularly noteworthy for fundamental investigations of the Figure 38. Common trunk materials used for preparation of
influence of soft matter properties on metal binding, as this adsorbents by RIGP.
work reports the ability to finely control both intermediate and
atomistic length scales by varying the size of the copolymer be irradiated with 60Co γ-rays or with an electron beam,
blocks as well as the distance between AO and cation.158 forming 1019 radicals per gram of polymer. These radicals
A summary of material performance for all polymer-derived induce uncontrolled graft polymerization upon immersion of
adsorbents prepared by neither RIGP nor ATRP is provided as the irradiated support in a solution containing appropriate
Table 3. monomers (Figure 39). In the work by Omichi, RIGP was used
5.2.3. Radiation-Induced Graft Polymerization. A truly to graft AN onto fibrous poly(tetrafluoroethylene-co-ethylene),
seminal contribution to the development of polymer-supported followed by conversion to amidoxime by treatment with
adsorbents for seawater uranium recovery was the introduction hydroxylamine. A second batch of material first involved
13961 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 39. Common comonomers used for preparation of adsorbents


by RIGP.

grafting of acrylic acid (AA) by RIGP to the polymeric trunk,


prior to a second RIGP of AN and subsequent amidoxima-
tion.159
When applied to recovering uranium from seawater, several
Figure 40. Uranium recovery by adsorbents possessing different
important observations were reported. First, uranium uptake surface areas. Note the y axis is intended to be a logarithmic scale.
was observed to be linearly correlated with the amount of Reprinted with permission from ref 161. Copyright 1987, Taylor &
amidoxime grafted to the trunk. Second, the molar ratio of Francis.
uranium to amidoxime groups was approximately 1:10−4,
revealing poor overall efficiency. Finally, inclusion of the AA
comonomer was observed to appreciably increase uranium 5.2.3.1. Modern Developments in RIGP. Although these
uptake over the pure amidoxime graft polymer by approx- foundational works established the fundamentals of developing
imately 20% following 7 d of batch contact with environmental polymer adsorbents by RIGP for uranium recovery, research
seawater. A maximum uranium uptake of slightly more than 0.5 continues to elucidate and optimize the complex relationships
mg of uranium/g of adsorbent was obtained over 50 d of that govern soft matter adsorbents. In 2000, Seko and
contact with 2 L of environmental seawater, which was colleagues reported the preparation of methacrylic acid (MA)
replenished every 2 d.159 copolymerized with AN onto PP fibers to improve uranium
The authors noted in a subsequent work that when adsorption kinetics by inclusion of a different hydrophilic
hydrophilic monomers, such as N,N-dimethyl acrylamide comonomer.162 In contrast to the aforementioned adsorbents,
(DMAAm) or AA were copolymerized with poly(AN), MA and AN were copolymerized from the same solution
significant increases in water uptake were observed.160 following one irradiation, rather than by two distinct RIGP
Nevertheless, while water uptake by the adsorbents was processes performed in series. Optimized performance was
poly(AO-co-DMAAm) > poly(AO-co-AA) > poly(AO), achieved for a AN/MA ratio of 80:20. Although similar to the
uranium uptake differed slightly: poly(AO-co-AA) > poly(AO- work reported by Omichi,160 water content was maximized for
co-DMAAm) > poly(AO). This work firmly established the a 60:40 ratio, suggesting competing factors in uranium
precedent for including hydrophilic comonomers to improve adsorption performance, such as hydrophilicity vs binding site
adsorbent performance, also demonstrating poly(AA) has no density. Inclusion of the hydrophilic MA resulted in a 4-fold
appreciable affinity for uranyl and is coordinatively innocent. enhancement in uranium binding kinetics, compared to
However, consideration of the difference between water uptake conventional fibers prepared exclusively from RIGP of AN,
and uranium uptake reveal that hydrophilicity is not the only and achieved a seawater uptake of 0.2 mg of uranium/g
consideration for achieving performance improvements,160 as following 1 d of contact with 20 L of natural seawater in a flow-
also observed by Hirotsu and colleagues in their studies on the through column.162
influence of hydrophilic vs hydrophobic cross-linkers on A report from the same group compared the performance of
uranium uptake.120,140,141 adsorbents prepared by RIGP of AN with different hydrophilic
The final foundational work reported by Omichi and comonomers from a PE trunk fiber, followed by amidoximation
colleagues articulates the first foray into the role of the shape with hydroxylamine.163 As in their prior work, MA was used as
of the polyolefin support on adsorbent performance. AN was one comonomer, while HEMA was the other. As expected,
polymerized via RIGP from PP fibrous supports possessing modulating the weight ratios of the comonomers afforded
either round or cross-shaped cross sections. When uranium adsorbents with differing hydrophilicities, and it was noted that
recovery performance was compared directly, the adsorbents the amidoxime density and water content needed to be
with cross-shaped supports afforded faster uranium binding balanced to achieve optimal uranium uptake. Water uptake by
kinetics and superior uranium uptake. This result was attributed the adsorbents (postconversion with hydroxylamine treatment)
to the increase in surface area afforded by the different trunk was maximized when AN was not included in the initial
material. Similarly, using small radius round fiber supports, a copolymer. For the MA copolymer, water uptake decreased
uranium uptake of 5 mg of uranium/g of adsorbent could be linearly with increasing amidoxime, while the HEMA
achieved following 140 d contact with natural seawater, in copolymer did not display an appreciable decrease in water
contrast to 0.5 mg of uranium/g of adsorbent for the larger uptake until amidoxime exceeded 60 wt %. Seawater uranium
round fiber support following 100 d contact. Careful recovery by the adsorbent materials was investigated both in
observation of these results led the researchers to propose a lab-based flow-through experiments, where 20 L of natural
logarithmic relationship between uranium recovery and trunk seawater was circulated through a column for 24 h, as well as in
surface area, as displayed in Figure 40.161 open-ocean deployments off the coast of Japan, where the
13962 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

materials were submerged at 15 m below the surface. flowers, gears, trilobal shapes, and others were investigated as
Submersion experiments subsequently validated by column supports for RIGP-prepared adsorbents (Figure 43). Similar to
studies revealed ratios of 60:40 for amidoxime:MA and 70:30 the round PE supports, increasing the surface area with shaped
amidoxime:HEMA were optimal (Figure 41). Uranium uptake, fibers resulted in an increase surface area and uranium uptake,
but without the loss of mechanical strength. Unlike the round
fibers of varying diameter, uranium uptake by the majority of
the shaped fibers displayed little variation, achieving uptake of
120−140 mg of uranium/g of adsorbent. The quasi-trilobal
material displayed poor performance, achieving uptake of only
20 mg of uranium/g of adsorbent, while the hollow gear
support achieved an uptake of 160 mg of uranium/g of
adsorbent. Screening capacities for all adsorbents are displayed
in Figure 44.
Several adsorbents on nonround PE supports were
investigated for recovery of uranium from filtered environ-
mental seawater at PNNL,61 compared against an adsorbent
sample obtained from the JAEA.48 Following 11−12 weeks of
contact in column-based flow through systems, the JAEA fiber
achieved an uptake of 1.1 mg of uranium/g of adsorbent, while
uptakes ranging from 2.3 to 3.1 mg of uranium/g of adsorbent
were obtained for the high surface area materials. As suggested
by lab-based screening, the hollow gear fiber was the best
performing. However, simultaneous investigation of six batches
Figure 41. Uranium adsorption as a function of weight percent AN in
the adsorbent, following 20-days deployment in the open ocean. of adsorbent prepared on the same hollow-gear trunk revealed
Circles correspond to poly(acrylamidoxime-co-MA)-graf t-PE, triangles variability in performance, with uranium uptake ranging from
correspond to poly(acrylamidoxime-co-HEMA)-graft-PE, diamonds 2.7 to 3.5 mg of uranium/g of adsorbent over 8 weeks of
denote the poly(amidoxime)-graft-PE adsorbent devoid of hydrophilic exposure. A JAEA sample investigated as a control achieved the
comonomer. Reprinted with permission from ref 163. Copyright 2000, same uptake as in the other column study, 1.1 mg of uranium/g
American Chemical Society. of adsorbent. Similar to the pioneering work by Omichi,161
these results demonstrate the potential for dramatic enhance-
measured at 10, 20, and 40 d, revealed superior kinetics for the ment of uranium uptake by modulation of trunk surface area.
MAA-containing adsorbent, outperforming the HEMA-con- Another study investigated functionalizing a PE support by
taining adsorbent by 1.7× and the pure poly(amidoxime) RIGP with AN and AA, followed by treatment with
material by 3×. The best performing material, a 60:40 hydroxylamine. Similar to aforementioned work involving
amidoxime:MA copolymer on a fibrous PE trunk, achieved an MA, water uptake and Dg increased with AA content, with a
uptake of 0.9 mg of uranium/g of adsorbent following 20 d maximum uptake afforded from a 50:50 amidoxime:AA ratio.
deployment in the open ocean.163 Uranium extraction was assessed from a 100 ppm solution over
More recently, work from Oyola and colleagues built off the 1-d contact, with poly(acrylamidoxime-co-AA) performing
work of Omichi161 and investigated a series of RIGP-prepared comparably to simple poly(amidoxime) and achieving uptake
adsorbents composed of PE-graf t-poly(amidoxime-co-MA), of 25 mg of uranium/g. This similarity in performance is almost
where various high surface area PE trunk fibers were used certainly attributable to the excessive uranium concentration in
instead of the traditional round fibers presented in previous the screening solution.165 A resin-based adsorbent was also
work.48 These PE supports were prepared by bicomponent prepared by RIGP of AN and AA. In this particular example, an
melt-spinning technology164 prior to RIGP, reducing the fiber 80:20 amidoxime:AA ratio was optimal, affording an uptake of
diameter, changing the cross-sectional shape, or by combina- 924 mg of uranium/g of adsorbent from a 100 ppm
tions of both. Fibers as small as 0.25 μm in diameter were UO2(NO3)2 solution.166 AN and MA grafted on PP by RIGP
prepared through an approach known as “islands-in-the-sea,” was also investigated, then analyzed for uranium uptake as a
where nanofibers are embedded in a dissolvable polymer function of pH. In contrast to aforementioned studies, water
matrix, such as poly(lactic acid) (PLA), resulting in a coaxial- uptake increased with additional amidoxime content. An uptake
like fiber. Dissolution of the PLA results in a fiber containing as of 45 mg of uranium/g of adsorbent was obtained from a 95
many as 156 000 nanofiber “islands.” High surface area PE ppm uranyl solution at pH 8 and 0.1 M ionic strength.40
fibers possessing round morphologies of varying diameter Seko and colleagues reported the preparation of an adsorbent
(Figure 42) were grafted with 70:30 AN:MA by RIGP, followed prepared on PE coated PP nonwoven fibers by RIGP of AN
by amidoximation and performance screening in an 8 ppm and MA followed by amidoximation through treatment with
uranium brine solution. The best performing adsorbent also hydroxylamine.39 Following irradiation in an inert environment,
had the smallest diameter, 0.24 mm, and in general adsorbents the PE-coated PP fibers were immersed in a comonomer
with smaller diameters outperformed those with larger mixture in degassed DMSO, the composition of which was
diameters. Uranium uptake ranged from 30 mg of uranium/g tuned to afford AN:MA ratios ranging from 60:40 to 80:20.
of adsorbent (d = 20 mm) to 140 mg of uranium/g of Performance was investigated at pH 8 in natural seawater,
adsorbent (d = 0.24 mm). contacted with a lab-based column flow-through system at a
The performance of nonround fibers as PE trunks was also velocity of 3 L min−1. A total of 2 days of contact afforded
investigated in the same work.48 Using bicomponent melt- 0.141, 0.145, and 0.135 mg of uranium/g of adsorbent for 60,
spinning technology, new shapes including solid or hollow 70, and 80 wt % amidoxime-containing materials, respectively.
13963 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 42. Summary of round PE fibrous supports investigated for seawater uranium recovery. Figure reprinted with permission from ref 48.
Copyright 2016, American Chemical Society.

Notably, the 70:30 material displayed the best performance, in Critical consideration of this suggestion finds it unlikely, but
contrast to the 60:40 ratio determined for purely a fibrous PE contingent upon the identity of the base treatment. Alkaline
support.163 Increasing seawater contact time for the optimized pretreatment before seawater deployment is a common
adsorbent to 7 days afforded a maximum uptake of 0.300 mg of practice, known to increase uranium uptake, though the precise
uranium/g of adsorbent.39 This nonwoven adsorbent was rational remains unresolved. Common bases used for this
subsequently used for a large scale open-ocean deployment. A purpose include KOH, NaOH, and Na2CO3.167,168 In light of
total of 52 000 sheets totaling 350 kg were packaged in 3 × 16 the acid stability reported for the open-chain amidoxime and
m2 adsorption cages, then deployed for 240 days at 20 m deep, the instability reported for the cyclic imidedioxime group, it is
7 km off the coast of Japan. 1 kg of uranium (yellowcake) was likely the decreasing peak upon HCl contact can be attributed
obtained, affording an uptake of 2.8 mg of uranium/g of to hydrolysis of cyclic imidedioxime groups. As reported by
adsorbent per 240 days.34 Hay and colleagues,138 glutarimidedioxime (H2A) degrades to
A significant decrease in performance was observed for the glutarate upon contact with HCl, suggesting formation of
adsorbents over five recycles, with uranium eluted with HCl. adjacent carboxylate species when part of a polymer. Treatment
of carboxylates with none of the aforementioned bases will
Treating the recycled adsorbent with an alkaline solution
result in regeneration of a cyclic imidedioxime, amidoxime, or
afforded 75% retention in original capacity, while no capacity
any related functionalities. However, in the event the alkaline
was achieved if the alkaline treatment was not performed. In
solution was composed of hydroxylamine, there is the
contrast, when tartaric acid, a weaker organic acid, was possibility of further graft-chain chemistry occurring. Further
substituted for HCl, > 80% performance during adsorbent work or additional information is necessary to understand and
recycle was achieved. 13C NMR was used to investigate the better interpret the changes in the 13C NMR spectra and the
degradation process. While peaks attributable to amidoxime implications for adsorbent recycle.
and cyclic imidedioxime were visible following treatment with A PE-coated PP support was used to attach bis-amidoxime
hydroxylamine, contact with HCl resulted in an increase in the groups through a glycidyl methacrylate (GMA) intermediate,
peak assigned to a carbonyl carbon, indicative of a hydrolysis polymerized by RIGP.41 The PE@PP-graf t-poly(GMA) was
degradation pathway (Figure 23).138 Upon treatment with an then reacted with 3,3′-iminodipropionitrile, ring-opening the
unspecified alkaline solution, a spectrum is obtained similar to epoxide group on the GMA functionality and installing two
the one following the initial treatment with hydroxylamine, nitriles. Subsequent treatment with hydroxylamine results in
prompting the suggestion from the authors that the uranium the desired bis-amidoxime-functionalized adsorbent (Figure
binding site is regenerated by this process.39 45). Although this adsorbent was not tested for uranium uptake
13964 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 43. Summary of shaped PE fibrous supports investigated for seawater uranium recovery. Figure reprinted with permission from ref 48.
Copyright 2016, American Chemical Society.

Figure 44. Performance for PE-graft-poly(amidoxime-co-MA) adsorbents on (left) round and (right) nonround PE trunk supports, following 24-h
contact with a 8 ppm uranium brine solution. Adsorbents on round fibers were prepared in monomer solutions containing either 25% or 50%
DMSO; adsorbents on nonround fibers were only prepared from monomer solutions containing 25% DMSO. Figures reprinted with permission
from ref 48. Copyright 2016, American Chemical Society.

from natural seawater, adsorption performed in a 100 ppb In 2008, Pandey and co-workers reported the application of
uranyl solution at pH 5 revealed rapid uptake (>80% in 30 RIGP to prepare PP membranes grafted with poly(AN).122
min). When contacted with a 1 ppm uranyl solution, an uptake Following irradiation in air to a 200 kGy dose, the PP
of 1.6 mg of uranium/g of adsorbent was obtained, though membranes were immersed in a DMF/AN solution at 50 °C,
competing ions V, Cu and Co all displayed significant affording a 125 wt % Dg. The poly(AN) functionalities were
adsorption, with values of 1.5, 0.9, and 0.9 mg/g, respectively.41 converted to amidoxime by treatment with hydroxylamine, and
Although the performance of the GMA-grafted adsorbents are the adsorbent was pretreated in an aqueous KOH solution.
not remarkable, these works comprise a particularly important Water uptake capacity for the adsorbent was 200 wt % in
contribution because they establish the use of an intermediate seawater, compared to 260 wt % in DI water, suggesting an
ligand to afford the rational design and inclusion of precisely influence of ionic strength on polymer hydrophilicity. Uranium
engineered uranyl binding sites. uptake was determined by scintillation counting using a 233U
13965 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 45. Preparation of a GMA-grafted adsorbent and subsequent chemical transformations to afford installation of bis-amidoxime functionalities.
Reprinted with permission from ref 41. Copyright 2007, Wiley Periodicals, Inc.

radiotracer, though no explicit concentration values were


provided. An uptake of 380 mg of uranium/g of adsorbent
was determined, affording a 1:4 uranium:amidoxime binding
mode assuming exclusive formation of the open-chain
amidoxime.
Using the RIGP-prepared poly(AN) adsorbent membranes
discussed above, in-field demonstration experiments were
performed, including deployment in intake/outfall canals of
the Tarapur Atomic Power Station.47 Adsorbents were
prepared by RIGP on PE, PP, and polyester supports, with
fibrous forms of the latter two polyolefins favored during lab
scale experiments and PP fibers in felt form used during oceanic
deployment. During lab-based screenings, the authors inves-
tigated the effect of various synthetic parameters on the
resulting material, including dose, duration, and solvent. As
observed in earlier work,43 graft yield increases as a function of
radiation dose with the most significant gains observed up to Figure 46. Diagram of the submergence facility at the Trombay
200 kGy, and improvements were still noticed as high as 250 Estuary, used for seawater- exposure studies. Reprinted with
kGy.47 Selection of the appropriate solvent was also determined permission from ref 169. Copyright 2010 American Chemical Society.
to be an important parameter, with DMF preventing phase
separation and facilitating polymerization. It was determined poly(amidoxime) by treatment with hydroxylamine. When
30% DMF in the AN monomer was ideal, as further dilution irradiated in air, the Dg increases with cumulative dose,
resulted in deleterious concentration effects. While polar protic reaching 80 wt % around 220 kGy. In contrast, when irradiated
solvents were investigated exclusively for treatment of the under an inert N2 environment, a plateau of 20 wt % is reached
poly(AN) with hydroxylamine, the authors noted that 1:1 around 180 kGy. Exposure of the irradiated sample to air for 15
mixtures of methanol or ethanol in water more than doubled min prior to contact with the monomer solution improved the
the amount of amidoxime generated on the final adsorbent grafting to 90 wt % Dg, while heating the polymerization
material. solution to 60 °C yielded further improvements and a Dg of
Initial studies performed in the Trombay Estuary (Figure 46) 130 wt %. These data are summarized in Figure 47.
suggested 12 d deployment to be the optimal time frame, Performance was analyzed by oceanic deployment at the
balancing uranium uptake against biofouling, which was Trombay estuary submergence facility, with the optimized
determined to be detrimental to uranium uptake. Subsequent adsorbent achieving 0.6 × 10−3 mg of uranium/g of adsorbent
experiments performed at the power station intake/outfall over 13 d of contact at the pilot scale.169
canals supported this result. Approximately 30% improvement Poly(vinyl alcohol) (PVA) was investigated as a trunk
in uranium uptake was observed for the adsorbents deployed by material for preparation of PVA-graf t-poly(AO) adsorbents.
the outfall canal over those deployed at the intake canal, Following preirradiation, the PVA trunk was immersed in the
putatively attributed to the elevated temperature and suggesting AN monomer solution, and a 43.5 wt % Dg was obtained.
endothermic thermodynamics for uranium binding. The elution Following conversion with hydroxylamine to form amidoximes,
of 1.8 and 4.85 ppm uranium was reported for the outfall canal the saturation capacity of the adsorbent was determined
(following 330 and 570 h deployment, respectively), while the through production of an isothermal study. A theoretical
comparative results for the intake canal were 1.3 and 3.92 saturation capacity of 69 mg of uranium/g of adsobent was
ppm.47 Insufficient data were provided to determine the uptake obtained by linearization of the Langmuir isotherm, while the
capacity of the adsorbent following seawater deployment. experimental maximum was 43 mg of uranium/g of adsorbent.
The same group reported a parametric study on RIGP of While the maximum adsorption in screening experiments was
adsorbents for recovery of heavy metals, investigating the effect obtained at pH 8.5, pertinent to seawater, isotherms and all
of cumulative dose, time exposed to air, atmosphere of reported values were obtained at pH 4, making relevant
radiation, and reaction temperature on final grafting yield. PP comparisons difficult.170
fiber sheets were irradiated and subsequently immersed in AN In an effort to address the poor mechanical properties
dissolved in DMF, followed by conversion of poly(AN) to exhibited by polymers following irradiation, as well as increase
13966 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 47. Degree of grafting of poly(AN) as a function of (left) cumulative dose, (center) grafting temperature, and (right) retention time in air
between irradiation and grafting. The center plot also displays the threshold limit values for AN vapors, measured at breathing levels in the
operational area. Figures reprinted with permission from ref 169. Copyright 2010, American Chemical Society.

Figure 48. Preparation of UHMWPE-graft-poly(GMA) and subsequent transformations to form amidoxime-functionalized adsorbent via addition of
AN and treatment with hydroxylamine. Figure reprinted with permission from ref 56. Copyright 2016, Elsevier.

the Dg by promoting postirradiation radical stability, Wu and column packed with glass beads which was then contacted with
co-workers applied RIGP to graft ultrahigh molecular weight filtered seawater from Sequim Bay, Washington, U.S.A.
polyethylene (UHMWPE).171 When investigated by electron Following 42 d of contact, an uptake of 2.3 mg of uranium/g
paramagnetic resonance, the radical lifetime on the UHMWPE of adsorbent was obtained.171 Subsequent studies on the same
was approximately 100× longer than on PE, suggesting superior adsorbent deployed in a 331 ppb uranium screening solution
potential for polymerization.172 Upon exposure to 50 kGy γ- contacted by batch and flow through techniques afforded
radiation from a 60Co source, UHMWPE fibers were immersed uranium uptake of 4.54 and 2.97 mg of uranium/g of
in a solution of AN and AA which had been purged with N2. adsorbent, respectively. When exposed to filtered seawater at
Ammonium iron(II) sulfate, more commonly known as Mohr’s PNNL,61 a capacity of 0.48 mg of uranium/g of adsorbent was
salt, was added to the polymerization solution to suppress obtained over 42 d. Finally, when deployed directly in the East
homopolymerization and promote the growth of the poly(AN- China Sea of Xiamen Island for 60 d, 0.25 mg of uranium/g of
co-AA) graft chain. The study reveals that the tensile strength of adsorbent was reported.172
the fiber decreases dramatically upon radiation, from 3 to 1.3 UHMWPE was subsequently used by the same group as the
GPa. While surface grafting improved tensile strength, fibrous polymer support for grafting of GMA and MA by RIGP,
subsequent treatment with hydroxylamine mitigated any gains again using a 60Co source to afford γ-irradiation.56 The Dg was
and a final tensile strength of 1.2 GPa was obtained. The analyzed in the material as a function of monomer
UHMWPE-graf t-poly(amidoxime-co-AA) adsorbent was concentration and dose. The max Dg obtained was 553 wt %,
screened in a 3 ppb solution for 11 d at 30 °C, affording a but a slow increase in Dg with dose in excess of 20 kGy
respectable uptake of 1.3 mg of uranium/g of adsorbent. When suggests a high amount of homopolymerization at higher dose
the uranyl concentration was increased to 5 ppm, the capacity and monomer concentration. 10 kGy and 10% monomer
improved to 21 mg of uranium/g of adsorbent at 30 °C and 28 concentration in the graft solution were deemed ideal.
mg of uranium/g of adsorbent at 40 °C, again indicative of an Following polymerization, the GMA was ring-opened by
endothermic uranium binding process. Lastly, the adsorbent reaction with EDA, with subsequent addition of AN to pendant
was investigated for recovery of uranium from natural seawater, amino groups, yielding nitrile functionalities which were
performed by PNNL.61 The adsorbent was suspended in a subsequently converted to amidoxime (Figure 48). Although
13967 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

affording the traditional amidoxime functionality on the


adsorbent, this approach also separates the chelating functions
from the graft chain, while adding amino groups to promote
hydrophilicity. Furthermore, this physical separation should
dramatically reduce the extent of cyclic imidedioxime
formation, as adjacent amidoximes are necessary to accom-
modate the generation of the 6-member ring, affording
exclusively the open-chain amidoxime on the adsorbent. For
these reasons, the synthetic avenue pursued in the development
of this adsorbent shows significant promise for the rational
design of model systems which would afford insight into
identification of adsorbent functional groups and metal binding
environments. The adsorbent was tested in a seawater simulant
under continuous flow conditions, affording an adsorbent
uptake of 1.97 mg of uranium/g of adsorbent. The adsorbent
also displayed much better uptake of uranium over vanadium, a
metal which traditionally is highly competitive for binding sites. Figure 49. Comonomers investigated as part of RIGP-prepared
The uptake of metals by the adsorbent was as follows: U > Cu adsorbents for extraction of uranium from seawater.
> Fe > Ca > Mg > Ni > Zn > Pb > V > Co.56
A later work reported the tuning of AN/AA feed ratios However, as uranium exists in seawater as the UO2(CO3)4−
during the RIGP preparation of UHMWPE adsorbents.173 The complex, a tetra-anion, a negatively charged polymer is not
series of materials possessed AN:AA ratios of 4:1, 1.3:1, 0.92:1, expected to facilitate concentration of the desired complex near
and 0.3:1, and were screened with a 300 ppb uranium solution the uranium binding amidoxime moieties. Moreover, such a
in seawater simulant, as well as deployment off the coast of material would presumably repel the alkali and alkaline earth
Xiamen Island at 3 m deep for 60 d. Tensile tests revealed a cations, ubiquitous in seawater. Hua and co-workers developed
50% reduction in strength upon irradiation, with further an adsorbent prepared with a cationic copolymer by grafting
reduction upon grafting and deployment in seawater. poly(1- vinylimidazole) onto PP via RIGP.176 Reaction with 4-
Furthermore, the authors noted higher ratios of AN:AA were bromobutyronitrile afforded a quaternary amine as well as the
not favorable for uranium extraction, consistent with previously necessary nitrile group for subsequent conversion to amidoxime
reported work,34,160,162,163 and proposed a potential synergism by treatment with hydroxylamine (Figure 50). Zeta-potential
between the carboxylic acid and amidoxime functionalities in measurements revealed the cationic polymer had a surface
achieving uranium binding.173 Uranium recovery from simulant charge of 145 mV, whereas poly(amidoxime) had a surface
was 5.49, 8.10, 7.01, and 3.62 mg-uranium/g-adsorbent, charge of −9.51 mV.
respectively. Similarly, adsorption upon oceanic deployment
afforded uptake of 0.29, 0.50, 0.77, and 0.33 mg-uranium/g-
adsorbent. From this work, the authors concluded that the
0.92:1 ratio of AN:AA afforded the best adsorbent, but also that
there was no definitive correlation between the quantity of
amidoxime groups and the adsorption capacity.173
5.2.3.2. Amidoxime-Functionalized Adsorbents with Dif-
ferent Comonomers. As can be readily gleaned from the
aforementioned publications, AA and MA constitute the two
archetypical comonomers used in the preparation of amidox-
ime-functionalized polymeric adsorbents (Figure 49).160,162
Other hydrophilic comonomers have also been used, with
DMAAm160 reported in some of the earliest literature applying
RIGP to the extraction of uranium from seawater. HEMA163 Figure 50. Schematic for synthesis of cationic amidoxime adsorbent.
was also investigated. In both instances, they were observed to Reprinted with permission from ref 176. Copyright 2015 American
afford less uranyl uptake than similar adsorbents prepared with Chemical Society.
AA or MA. Nevertheless, while hydrophilic comonomers are
known to be essential for adsorbent performance, the totality of Uranium uptake experiments were performed in a brine
their role in influencing uranium uptake remains unresolved, solution of seawater-relevant salinity and carbonate, with pH
and many adsorbents have been prepared with more exotic adjusted to 8. When contacted with brine spiked with 7.2 ppm
comonomers in an effort to improve seawater uranium uranium, capacities were correlated with polymer Dg. Inclusion
recovery. of the cationic species appeared to improve performance, as
Based on early work where water uptake and uranium isothermal studies afforded a saturation capacity of 100 mg of
binding were both investigated,120,140,141,160,162,163 one can uranium/g of adsorbent, compared to 20 mg of uranium/g of
conclude that improvement in adsorbent hydrophilicity is one adsorbent for the neutral poly(amidoxime). While clearly
role the comonomer does definitively play. Both AN and MA displaying a solid trend, significant differences in polymer
contain carboxylic acids which will deprotonate at seawater pH, degree of grafting prevent an ideal comparison. When the
encouraging the adsorbent to swell due to Coulombic cationic adsorbent was immersed in a 1.7 ppb solution, a
repulsion, while pKa values for both amidoxime small capacity of 0.04 mg of uranium/g of adsorbent was achieved
molecules133,174 and polymers175 reveal it will be neutral. following 8 days contact.176
13968 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Oyola and Dai reported the systematic investigation of


uranium uptake as a function of different comonomers grafted
by RIGP on a PE fiber.177 Using a brine solution for screening
performance, they observed uranium uptake as a function of
comonomer was as follows: AA < vinyl sulfonic acid < MA <
itaconic acid (IA) < vinyl phosphonic acid (PA), which
achieved an uptake of approximately 160 mg of uranium/g of
adsorbent. When exposed to environmental seawater, the
performance changed: MA < AA (with Mohr’s salt) < vinyl
sulfonic acid < IA (with Mohr’s salt) < IA (without Mohr’s salt)
< PA. Several different formulations of the adsorbents were
investigated, affording seawater uranium uptake as high as 5 mg
of uranium/g of adsorbnet over 11 weeks or 4.4 mg of
uranium/g of adsorbent over 8 weeks. CHN analysis revealed a
low cografting of hydrophilic monomers (e.g., <10 wt % IA and
<3 wt % PA). An important observation made by the authors
Figure 51. Uptake by AI-series adsorbents from 8 ppm uranium brine
was that the increase in uranium adsorption capacity was solution, following 3 h of KOH conditioning and 24-h contact. Figure
correlated with grafting of AN, suggesting that one potential reprinted with permission from ref 50. Copyright 2016, American
role of the comonomer is simply to promote AN polymer- Chemical Society.
ization.
Building off this work, Das and colleagues reported the
optimization of two promising adsorbents identified during the a 3.21:1 ratio. In contrast to the screening studies, the
previous investigation: PE-graf t-poly(amidoxime-co-IA)49 and differences in uranium uptake were much less distinct, with
PE-graf t-poly(amidoxime-co-PA),50 denoted as the “AF” and most materials, including PE-graf t-poly(amidoxime), achieving
“AI” series, respectively. The ensuing parametric studies capacities ranging from 2.8−3.2 mg of uranium/g of adsorbent.
optimized Dg by varying the ratio of AN:comonomer, as well The best uptake of 3.4 mg of uranium/g of adsorbent was
as by testing the effect of 1 h vs 3 h KOH predeployment achieved following 3 h of KOH conditioning of the material
treatment. For both series, material performance was first with a 3.52:1 AN:PA ratio.
assessed by 24-h contact with a brine screening solution, Similar to previous work with AN and AA, these reports
possessing appropriate salinity, pH, and carbonate for environ- confirm the importance of the hydrophilic comonomer in
mental seawater, but with an 8 ppm uranium concentration. achieving improved adsorbent performance, though they also
Samples were subsequently analyzed at PNNL suspended in demonstrate these two new comonomers require less monomer
columns supported by glass beads and contacted with filtered composition to achieve optimal performance. Although the
natural seawater over a period of 56 d.61 precise rationale for this has not been investigated directly,
For the AF materials, varying the AN:IA ratio in the grafting subsequent work has suggested at the innocence of the
solution from 3.76:1 to 23.36:1 afforded Dg values of 154−354 comonomer in binding uranium directly, suggesting improve-
wt %. For the AI series, AN:PA ratios varied from 1.91:1 to ments may be attributable to increased steric bulk and the
7.39:1, resulting in Dgs of 110−300 wt %. When screened for influence on graft chain morphology.
uranium uptake in the brine solution, the AF series showed 5.2.3.3. Emulsion Graft Polymerization. Emulsion polymer-
consistent performance of approximately 170−190 mg of ization affords an interesting variation on traditional RIGP
uranium/g of adsorbent, regardless of Dg. A maximum uptake processes. Instead of immersing a preirradiated polymer
of 200 mg of uranium/g of adsorbent obtained for the material support in a homogeneous monomer solution, a surfactant is
with a 7.57:1 AN:IA ratio, conditioned in KOH for 3 h prior to used to create microemulsions possessing high monomer
contact. In contrast, the AI series displayed a dramatic change concentrations. Polymerization proceeds rapidly once initiated,
in performance in the brine solution as a function of while formation of uniform microemulsions ensures similar
comonomer ratio. While a PE-graf t-poly(amidoxime) adsorb- quantities of monomer are polymerized at each initiated site,
ent yielded an uptake of 165 mg of uranium/g of adsorbent, theoretically reducing the polydispersity of the resulting graft
progressive inclusion of PA afforded increasing uptake, chain.
maximized at a ratio of approximately 3.2−3.5:1 (Figure 51). Following the initial report of GMA as an intermediate
The maximum uptake was approximately 187 mg of uranium/g functionality for preparing uranium adsorbents,41 GMA was
of adsorbent, following 3 h of KOH conditioning. used to achieve the first adsorbent prepared by emulsion
For both AF and AI materials, trends observed during brine polymerization for the recovery of uranium from seawater.
screening loosely predicted performance upon deployment in GMA was grafted to a PE fiber adsorbent through RIGP with
seawater (Figure 52). For AF, the best material performance Tween-20 as the surfactant.43 Improvements in Dg were
was achieved by a 10.14:1 AN:IA ratio contacted for 1 h with observed to be dependent upon radiation dose, up to 40 kGy,
KOH, affording 3.9 mg of uranium/g of adsorbent. This is a where the rate of polymerization exceeds the rate of monomer
slightly higher monomer ratio than determined ideal by the diffusion to the trunk polymer. The highest Dg values were
screening study, underscoring the importance of actual seawater obtained at 40 kGy, 50 °C, and with 0.5% Tween-20 in the
tests in providing meaningful adsorption data and for analyzing monomer solution. The GMA-grafted fibers were then reacted
the effect of synthetic modifications to adsorbents. For AI with DETA, triethylenetetramine (TETA), and ethylenedi-
materials, the same AN:PA ratios as identified in screening amine (EDA), again by ring-opening of the GMA epoxide
afforded the optimal performance, though the material (Figure 53). Uranium uptake was assessed by contacting with a
composed of 3.52:1 AN:PA out-performed the material with 100 ppb aqueous solution for 1 d at pH 5. While a precise
13969 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 52. Seawater uranium recovery by (left) AF and (right) AI materials following 1 or 3 h of KOH conditioning at 80 °C. Figures reproduced
with permission from ref 49 and 50. Copyright 2016, American Chemical Society.

Figure 54. Density of different amine chelators as a function of Dg for


PE-graf t-poly(GMA) adsorbents prepared by emulsion polymer-
Figure 53. Comonomers and amines applied for epoxide ring opening ization. Figure reprinted with permission from ref 178. Copyright
of GMA and 4-HAGE to generate uranyl adsorbent polymers. 2012, Elsevier.

uptake cannot be calculated from the data provided, a glycidyl ether) (4-HAGE) by RIGP and emulsion polymer-
minimum value of 0.25 mg of uranium/g of adsorbent can be ization in a 5% graft ligand and 0.5% Tween-20 solution.179
calculated based on the reported value for the Kd.43 Similar to GMA, 4-HAGE possesses an epoxide group which
Wang and colleagues performed a closely related work, using can be opened through nucleophilic attack by sterically
preirradiation to graft emulsified GMA (with Tween-20 unencumbered amines, as well as an acryl functionality which
surfactant) onto PE nonwoven fabric.178 In addition to is amenable to RIGP. The primary difference is the length of
TETA, DETA, and EDA, tetraethylenepentamine was also the alkyl chain between the acryl and epoxide groups, with 4-
used to functionalize the poly(GMA)-graf t-PE material. While HAGE an additional six atoms farther away and putatively
Dg of up to 325 wt % could be obtained with a dose of 30 kGy affording faster adsorption and improved performance. As
and 60 min polymerization at 75 °C, the high Dg materials reported previously, following polymerization on a nonwoven
were observed to embrittle, and an optimal Dg of 195 wt % was PE coated PP fabric, the epoxide group was opened with EDA,
selected. Uranium uptake was determined from various DETA, TETA, and diethylamine (DEA) affording chelator
screening solutions with the best performing material densities ranging from 1.6 to 3 mmol g−1. Uranium adsorption
containing the simple EDA chelator which afforded the highest studies were performed revealing DETA to achieve fastest
molar density (Figure 54). When contacted with a 1 ppm adsorption kinetics, followed by DEA, TETA, and EDA, and a
uranyl solution in water, a capacity of 1.6 mg of uranium/g of maximum uranium uptake of 64.3 mg of uranium/g. While pH
adsorbent was achieved. Inclusion of carbonate reduced the had a negligible effect on the uranium uptake over the range of
uptake to 1.3 mg of uranium/g of adsorbent. Reducing the 4−8.5, deploying the adsorbents in uranyl solutions with 3.5%
uranium concentration to that of natural seawater afforded salinity, appropriate for a seawater simulant, resulted in a 55%
uptake of 0.006 and 0.003 mg of uranium/g of adsorbent, for loss of capacity. Furthermore, adsorption was optimal when
DI water and carbonate solutions, respectively.178 performed at 25 °C, outperforming studies performed at 35 °C
Related work was recently published articulating the and far superior to performance at lower temperature. These
preparation of PE@PP-graf t-poly(4-hydroxybutyl acrylate results not only prevent determination of binding enthalpy, but
13970 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

also suggest factors in addition to those occurring at the


atomistic scale influence the energetics of uranium binding in
this system.179
Wang, Li and colleagues reported the first instance of
emulsion polymerization of AN for development of adsorbents
for the recovery of uranium from seawater.92 PE nonwoven
fabric was preirradiated and contacted with AN emulsified with
Tween-20, followed by treatment with hydroxylamine to
generate PE-graf t-poly(amidoxime), as reported in a previous
work.180 Characterization of uranium uptake was performed at
different concentrations and temperatures in 3.5 L of uranyl
solutions. At the environmentally relevant concentration of 3
ppb and pH 7.5, the adsorbent achieved 0.024−0.036 mg of
uranium/g of adsorbent following 10 days contact at 10−30 °C.
Application of the Arrhenius equation afforded an activation
energy of 65.8 kJ mol−1, and the increase in uranium uptake as
a function of temperature is indicative of an endothermic
binding process.92 Figure 56. Swelling curves for hydrogels formed from poly-
5.2.3.4. Hydrogels Prepared through Radiation Induced (amidoxime-co-VP) for various comonomer ratios. Figure reprinted
with permission from ref 182. Copyright 2000, Taylor & Francis.
Polymerization. The majority of work pertaining to adsorbent
development by RIGP involves the propagation of a graft chain
from a polyolefin trunk, as discussed above. The trunk is uranyl uptake. While the poly(AN-co-VP) IPN afforded a
irradiated and subsequently immersed in a monomer- uranium uptake of 540 mg of uranium/g of adsorbent prior to
containing solution. However, direct irradiation of the treatment with hydroxylamine, performance improved to 750
monomer solution, without inclusion of any trunk or support, mg of uranium/g of adsorbent after formation of the
results in the formation of a hydrogel which can also be used amidoxime groups. While the importance of forming the
for the recovery of uranium from seawater. amidoxime species cannot be overstated, these uptake experi-
Pekel, Güven, and colleagues have been protagonists of this ments were performed at a concentration of 1400 ppm
field of material development, with early work reporting the use uranium, without any carbonate or consideration for ionic
of poly(AN-co-N-vinyl-2-pyrrolidone) (VP) hydrogels for strength. Accordingly, a high uptake is observed for even the
uranium recovery (Figure 55).181 The respective monomers nonamidoximated material, and little insight is afforded
regarding the materials potential for seawater uranium
recovery.183
Synthesis of terpolymeric hydrogel adsorbents was achieved
by polymerization in water initiated by a 60Co source. Aqueous
solutions containing acrylamide (AA), 2-hydroxy propylmethyl
acrylate (HPMA), and maleic acid (MaA) were irradiated in air
and at ambient temperature at 0.45 kGy h−1 to total doses of 2,
3, 4, 5, and 6 kGy. Hydrogels composed of AA, AA/HPMA,
and AA/HPMA/MaA were synthesized, characterized, and
investigated for uranyl uptake. Adsorbent swelling upon contact
with water was greatest for the poly(AA) system, followed by
Figure 55. Comonomers used for the preparation of hydrogels poly(AA-co-HPMA-co-MaA), then poly(AA-co-HPMA). Im-
investigated for the recovery of uranium from seawater. portantly, the poly(AA) system did not adsorb any uranium,
while the others displayed a distinct color change following
were irradiated by a γ-radiation from a 60Co source, with contact with an aqueous uranyl solution for 48 h at pH 5−6.
AN:VP ratios of 0.67:1, 1:1, 1.5:1, and 2:1. Hydrogel The negligible uptake by the AA hydrogel was rationalized by it
performance was assessed by contacting with aqueous solutions being devoid of any ionizable groups, preventing interaction
1000−1850 ppm in uranium, affording uptakes as high as 540 with uranyl. ΔG values of <20 kJ mol−1 were calculated,
mg of uranium/g of adsorbent for the best materials, albeit at indicative of physisorption of uranium onto the material, rather
pH 4. Similar to work on supported graft polymers, a mixture of than chemical binding. Information on the quantity of uranium
AN with the hydrophilic comonomer VP afforded the best adsorbed was not provided, but the aforementioned thermody-
performance, with 1:1 and 1.5:1 achieving the best uptake due namic information reveals the hydrogel would not perform
to the improved hydrophilicity and accessibility of uranium favorably for seawater uranium recovery.184
binding sites. Concurrent work revealed higher quantities of VP AA and acrylic acid (AcA) were copolymerized at molar
afforded greater water uptake and swelling (Figure 56).182 ratios of 15:85, 20:80, and 30:70, with the resulting hydrogels
Based on the stoichiometry of amidoxime:uranyl, the authors investigated for uranium recovery.185 Swelling behavior was
proposed a 1:4 UO22+:amidoxime binding mode.181 related to the quantity of AA, and with maximum swelling for
Later work from the same group reported the preparation of the hydrogel with the 15:85 molar ratio at pH 8. In contrast,
poly(AN-co-VP) interpenetrating polymer networks (IPNs), swelling with UO 2 (NO 3 ) 2 or UO 2 (OAc) 2 resulted in
synthesized by γ-irradiation of solutions of AN in VP. Nitriles significantly lower swelling percentages, attributed to metal
were converted to amidoxime by treatment with hydroxyl- binding resulting in the exclusion of water from the hydrogel.
amine, which was monitored by FTIR and correlated well with Uranium uptake ranged from 70−400 mg of uranium/g of
13971 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 4. Performance of RIGP-Prepared Adsorbents for Recovery of Uranium from Water and Seawater
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
poly(TFE-co-E)-graf t-poly(AO) 0.5 50 d semibatch; replenished every 2 d 159
poly(TFE-co-E)-graf t-poly(AO) 0.11 5d semibatch; replenished at 2 L 160
per day
poly(TFE-co-E)-graf t-poly(AO-co- 0.22 5d semibatch; replenished at 2 L 160
AAc) per day
poly(TFE-co-E)-graf t-poly(AO-co- 0.16 5d semibatch; replenished at 2 L 160
DMAAm) per day
poly(TFE-co-E)-graf t-poly(AO) 5 140 d column, continuously 161
(small radius fibers) replenished
poly(TFE-co-E)-graf t-poly(AO) 0.5 100 d column, continuously 161
(large radius fibers) replenished
poly(TFE-co-E)-graf t-poly(AO) 0.4 20 d column, continuously 161
(small radius fibers) replenished
poly(TFE-co-E)-graf t-poly(AO) 0.8 20 d column, continuously 161
(cross-shaped fibers) replenished
PP-graft-poly(AO-co-MA) (fibers) 0.2 1d column, 20 L fixed volume 162
PE-graf t-poly(AO-co-MA) (fibers) 0.9 20 d open ocean, (Sea of Japan) 163
PE-graf t-poly(AO-co-HEMA) 0.4 20 d open ocean, (Sea of Japan) 163
(fibers)
PE-graf t-poly(AO-co-MA) (high
surface area fibers)
UHMWPE 30 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
“islands in the sea” (37) 85 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
round sheath/core 58 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
14 μm round (Aspun) 45 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
16 μm round (Dowex) 45 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
16 μm round (Engage) 23 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
“islands in the sea” (330) 60 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
0.24 μm round (Aspun) 140 7.9 ppm uranium with 193 ppm of 2.9 77 d column, continuously 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3 replenished
16 μm round (Dow) 55 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
18 μm round (Aspun) 55 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
14 μm flower 130 7.9 ppm uranium with 193 ppm of 2.9 77 d column, continuously 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3 replenished
12−17 μm hollow gear 160 7.9 ppm uranium with 193 ppm of 3.0 ± 0.3b 56 d column, continuously 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3 replenished
3.1 77 d column, continuously 48
replenished
18 μm quasi-trilobal 20 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
17 μm solid gear 130 7.9 ppm uranium with 193 ppm of 2.6 77 d column, continuously 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3 replenished (Sequim,
WA, U.S.A.)
2−5 μm hollow snowflake 120 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
17−20 μm caterpillar 130 7.9 ppm uranium with 193 ppm of 2.3 77 d column, continuously 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3 replenished
30 μm hollow gear (Equistar) 135 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
30 μm hollow gear (Aspun) 145 7.9 ppm uranium, 193 ppm of Na2CO3, 48
25,600 ppm of NaCl, pH 8.3
30 μm solid gear 140 7.9 ppm uranium with 193 ppm of 48
Na2CO3, 25,600 ppm of NaCl, pH 8.3
PE-graf t-poly(AO-co-AA) 25 100 ppm uranium 165
poly(AO-co-AA) (Resin) 924 100 ppm uranium 166
PP-graft-poly(AO-co-MA) 45 95 ppm uranium, pH 8, 0.1 M ionic 40
strength
PE-coated PP-graf t-poly(AO-co- 0.3 7d column, continuously 39
MA) (nonwoven fibers) replenished

13972 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 4. continued
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
PE-coated PP-graf t-poly(AO-co- 2.8 240 d open ocean pilot-scale study, 34
MA) (nonwoven fibers) (Sea of Japan)
PE@PP-graft- poly 1.6 1 ppm uranium 41
(dipropioamidoxime)
(nonwoven fabric)
PP-graft-poly(AO) (membrane) 380 uranium concentration not provided 122
PP-graft-poly(AO) (fiber sheet) 6 × 10−4 13 d open ocean, (Trombay Esutary) 169
PVA-graf t-poly(AO) 69c pH 4 170
UHMWPE-graft-poly(AO-co-AA) 1.3 3 ppb uranyl, 11 d contact time 2.3 42 d column, continuously 171
replenished
UHMWPE-graft-poly(AO-co-AA) 4.54d 331 ppb uranium, seawater simulant, 0.48 42 d column, continuously 172
pH 8.1 (batch contact) replenished
2.97d 331 ppb uranium, seawater simulant, 0.25 60 d open ocean, (East China Sea) 172
pH 8.1 (column contact)
UHMWPE-graft-poly(GMA-co- 1.97d 3 ppb uranium, seawater simulant, pH 8 56
MA) (fibers)
UHMWPE-graft-poly(AO-co-AA)
(fibers)
AN:AA = 4:1 5.5d 331 ppb uranium, seawater simulant, 0.3 60 d open ocean, (East China Sea) 173
pH 8.1
AN:AA = 1.3:1 8.1d 331 ppb uranium, seawater simulant, 0.5 60 d open ocean, (East China Sea) 173
pH 8.1
d
AN:AA = 0.92:1 7 331 ppb uranium, seawater simulant, 0.8 60 d open ocean, (East China Sea) 173
pH 8.1
AN:AA = 0.3:1 3.6d 331 ppb uranium, seawater simulant, 0.3 60 d open ocean, (East China Sea) 173
pH 8.1
PP-graft-poly(1-vinylimidazole) 100c 0.438 M NaCl, 2.3 mM NaHCO3, pH 8 176
(fibers)
0.04 1.7 ppb uranium, 0.438 M NaCl, 2.3 mM 176
NaHCO3, pH 8
PE-graf t-poly(AO-co-AA) (fibers) 105 7.9 ppm uranium, 0.453 M NaCl, 2.3 177
mmol NaHCO3, pH 8
w/Mohr’s salt 152 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.4 77 d column, continuously 177
mmol NaHCO3, pH 8 replenished
PE-graf t-poly(AO-co-vinyl 127 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.4 77 d column, continuously 177
sulfonic acid) (fibers) mmol NaHCO3, pH 8 replenished
PE-graf t-poly(AO-co-MA) (fibers) 141 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.1 ± 0.05e 77 d column, continuously 177
mmol NaHCO3, pH 8 replenished
2.9 56 d column, continuously 177
replenished
w/Mohr’s salt 60 7.9 ppm uranium, 0.453 M NaCl, 2.3 177
mmol NaHCO3, pH 8
PE-graf t-poly(AO-co-IA) (fibers) 158 7.9 ppm uranium, 0.453 M NaCl, 2.3 4.7 ± 0.04e 77 d column, continuously 177
mmol NaHCO3, pH 8 replenished
4.6 56 d column, continuously 177
replenished
w/Mohr’s salt 157 7.9 ppm uranium, 0.453 M NaCl, 2.3 4 77 d column, continuously 177
mmol NaHCO3, pH 8 replenished
PE-graf t-poly(AO-co-PA) (fibers) 155 7.9 ppm uranium, 0.453 M NaCl, 2.3 5.0 ± 0.1e 77 d column, continuously 177
mmol NaHCO3, pH 8 replenished
4.4 56 d column, continuously 177
replenished
PE-graf t-poly(AO-co-IA) “AF”
(fibers)
AN:IA = 3.76:1 182 7.9 ppm uranium, 0.453 M NaCl, 2.3 2.8 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 4.66:1 179 7.9 ppm uranium, 0.453 M NaCl, 2.3 2.9 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 5.87:1 183 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.1 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 6.90:1 180 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.5 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 7.57:1 200 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.6 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 10.14:1 193 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.9 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 14.50:1 195 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.5 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished
AN:IA = 23.36:1 170 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.3 56 d column, continuously 49
mmol NaHCO3, pH 8 replenished

13973 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 4. continued
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
PE-graf t-poly(AO-co-PA) “AI”
(fibers)
AN:PA = 1.91:1 172 7.9 ppm uranium, 0.453 M NaCl, 2.3 2.9 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 2.33:1 178 7.9 ppm uranium, 0.453 M NaCl, 2.3 3 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 2.86:1 185 7.9 ppm uranium, 0.453 M NaCl, 2.3 3 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 3.21:1 188 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.2 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 3.52:1 188 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.4 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 4.41:1 184 7.9 ppm uranium, 0.453 M NaCl, 2.3 3 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 5.62:1 181 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.1 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
AN:PA = 7.39:1 174 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.2 56 d column, continuously 50
mmol NaHCO3, pH 8 replenished
PE-graf t-poly(AO-co-IA) “AF”
(fibers)
heated in DMSO 225f 7.9 ppm uranium, 0.453 M NaCl, 2.3 4.2 56 d column, continuously 77
mmol NaHCO3, pH 8 replenished
as prepared (no heat treatment) 175f 7.9 ppm uranium, 0.453 M NaCl, 2.3 3.5 56 d column, continuously 77
mmol NaHCO3, pH 8 replenished
PE-graf t-poly(GMA-DETA) ≥0.003 g
100 ppb uranium, pH 5 43
PE-graf t-poly(GMA-TETA) ≥0.01g 100 ppb uranium, pH 5 43
PE-graf t-poly(GMA-EDA) ≥0.25g 100 ppb uranium, pH 5 43
PE-graf t-poly(GMA-TETA) 0.003 3 ppb uranium, with NaHCO3, pH 8.3 178
(nonwoven fabric)
1.2 1 ppm uranium, with NaHCO3, pH 8.3 178
PE-graf t-poly(GMA-EDA) 0.003 3 ppb uranium, with NaHCO3, pH 8.3 178
(nonwoven fabric)
1.3 1 ppm uranium, with NaHCO3, pH 8.3 178
PE@PP-graft-poly(4-HAGE- 12 1 ppm uranium, pH 4−10 179
EDA) (nonwoven fabric)
PE@PP-graft-poly(4-HAGE- 12 1 ppm uranium, pH 4−10 179
TETA) (nonwoven fabric)
PE@PP-graft-poly(4-HAGE- 12 1 ppm uranium, pH 4−10 179
DEA) (nonwoven fabric)
PE@PP-graft-poly(4-HAGE- 12 1 ppm uranium, pH 4−10 179
DETA) (nonwoven fabric)
64.3 7 ppm uranium, pH 8 179
6.5 1 ppm uranium, pH 8, 3.5% salinity 179
PE-graf t-poly(AO) (nonwoven 0.036 3 ppb uranium, pH 7.5, 30 °C 92
fabric)
poly(AO-co-VP) (hydrogel)
AN:VP = 1.0:1.0 540c pH 4 181
AN:VP = 1.5:1.0 540c pH 4 181
poly(AN-co-VP) (IPN) 540 1400 ppm uranium 183
poly(AO-co-VP) (IPN) 750 1400 ppm uranium 183
poly(AA-co-AcA) 400 1000 ppm uranium, pH 7 185
a
Unless otherwise stated, adsorption experiments were performed in DI water and pH was unadjusted. bAverage of six replicates; error is the
standard deviation. cMaximum capacity at saturation. dSeawater simulant contains 175 g of sea salt and 0.2 g of Na2CO3 dissolved in 5 L of deionized
water with uranium and transition metals added to 100× environmental concentration. eAverage of two replicates; error is the standard deviation.
f
Salinity normalized to 35 psu. gData insufficient to calculate precise uranium uptake.

adsorbent, though the yellow coloration and facile desorption affords precise control over the graft chain synthesis, permitting
of uranyl ions upon contact with DI water supports modulation of various important polymer parameters. For
physisorption as the mode of action.185 example, graft chain molecular weight, composition, and
Performance for uranium uptake for adsorbents prepared by morphologies can be tuned, including installation of various
RIGP are summarized in Table 4. block copolymer architectures. Furthermore, precise control
5.2.4. Atom-Transfer Radical Polymerization. Since the over the chain length can be achieved, allowing determination
initial discovery in the 1950s,186,187 controlled or “living” of the precise Dg to achieve optimal performance. Recent
polymerization has contributed tremendously to the advance- developments in controlled radical polymerization have also
ment of polymer science. In contrast to conventional free greatly expanded its applicability to a wide range of
radical polymerization, such as RIGP, living polymerization monomers,188,189 allowing the incorporation of monomers
13974 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

which are not suitable for RIGP due to radical quenching and fiber, and a hollow-gear-shaped PE fiber (shaped fibers,
making controlled radical polymerization an exciting avenue for including representative pictures, are discussed above in section
the rational development of new adsorbents for the recovery of 4.2.2). ATRP of AN and tBA from the halide-functionalized
uranium from seawater. ATRP is especially well-suited for trunk fibers was successfully achieved, with Dg as high as 2570
controlling AN polymerization, particularly in light of other wt % obtained. Amidoximation was achieved by the usual
techniques, such as anionic polymerization, which are not as approach. An overview of the synthesis of this new adsorbent is
well controlled.190 In addition to the intensely studied provided in Figure 59. The maximum adsorption capacity from
polymerization mechanism of AN via ATRP, the functionaliza- a 6 ppm uranium screening brine solution was 146.6 mg of
tion of trunk fibers to possess uniformly distributed ATRP uranium/g of adsorbent, approximately doubling an adsorbent
initiation sites is feasible, with some commercially available provided by the JAEA as a control (75.1 mg of uranium/g of
trunk polymers amenable to direct functionalization by ATRP. adsorbent).
Thus, several recent studies investigated the use of ATRP for a The same group simultaneously reported graft-polymerizing
graft polymerization on fibers and other substrates, as are from a commercially available poly(vinyl chloride)-co-chlori-
summarized in the following section. nated poly(vinyl chloride) (PVC-co-CPVC) fiber through
Saito and colleagues reported the first use of ATRP for the application of ATRP (Figure 60).78 This direct approach
preparation of adsorbents for uranium recovery from sea- eliminates any modification of the trunk material, simplifying
water.52 A high surface area PE trunk was first grafted by RIGP the process and reducing the adsorbent cost. The ratio of AN
with poly(vinylbenzyl chloride, VBC) (Figure 57) to install to tBA was systematically varied to identify the optimal ratio of
hydrophilicity and uranyl binding amidoxime functionalities,
derived from the AN by treatment with hydroxylamine. The
best performing adsorbent possessed a tBA:AN ratio of 0.356
and a Dg of 1390 wt %. When deployed in flow-through
column studies and contacted with filtered natural seawater,
remarkable uptake was obtained. Following 42 days exposure,
the material uptake ranged from 2.42−3.24 mg of uranium/g of
adsorbent, while 5.22 mg of uranium/g of adsorbent was
recovered following 49 d of exposure. To date, this remains the
highest uranium uptake reported in the open literature.
While displaying remarkable promise, questions remain
Figure 57. Trunk materials and comonomers used in preparation of unresolved regarding interbatch reproducibility for the PVC-
polymeric adsorbents by ATRP. co-CPVC-graf t-poly(AO-co-tBA) material. As reported in the
publication, three different batches of adsorbent were analyzed
ATRP initiation sites. AN and tert-butyl acrylate (tBA) were for recovery of uranium from filtered natural seawater, each
then copolymerzied by ATRP from the PE-graft-poly(VBC) affording different uptake kinetics and calculated saturation
trunk material, creating a “brush-on-brush” structure (Figure capacities. While the best performing batch reached a half-
58). This material was then converted to the amidoxime form saturation around 14 days contact, the worst performing batch
by reaction with hydroxylamine. Implementation of ATRP required nearly twice as long (Figure 61). Accompanying
afforded a much higher Dg than RIGP alone, with values calculated saturation uptake ranged from 3.94 to 6.91 mg of
ranging from 585−2818 wt % being achieved. Polymer uranium/g of adsorbent. Comparison against JAEA reference
composition was varied by tuning the ratio of AN:tBA, fibers deployed and measured simultaneously also displayed
observing that both amidoxime density and hydrophilicity significant fluctuations, though the changes in uptake did not
were essential for optimal adsorbent performance. correlate with those observed for the ATRP-prepared material.
This first generation of ATRP adsorbents were screened in a The authors attributed slight compositional differences to be
6 ppm uranium brine solution under laboratory conditions, the driving force in the dramatic changes in performance.
achieving uptake of 141−179 mg of uranium/g of adsorbent. Further investigation is necessary to achieve sufficiently precise
However, when deployed in filtered natural seawater in a flow- control over the polymer composition and validate the
through column setup, uptake was lower than that of other influence on adsorbent performance. If this hypothesis is
high-surface-area adsorbents,49,50 attributed to slower adsorp- correct, it implies extreme sensitivity of the adsorbent to subtle
tion kinetics resulting from suppressed mass transfer. The next changes in synthetic approach, and causes one to question
iteration on the adsorbent included the installation of a whether such control is achievable at industrial scale.
hydrophilic block of poly(tBA) on the end of the graft chain ATRP has also been applied to the synthesis of mesoporous
terminus, resulting in the doubling of the seawater adsorption polymer adsorbents, functionalized with poly(amidoxime). As
capacity from 1.56 mg of uranium/g of adsorbent to over 3 mg first reported by Yue and co-workers, a copolymer monolith
of uranium/g of adsorbent following 56 d of exposure. This support was synthesized by conventional free radical polymer-
investigation revealed the importance of polymer chain ization of VBC and DVB in the presence of AIBN.191 A surface
conformation and optimization of hydrophilic comonomers in area of 572 m2 g−1 was achieved by optimizing the ratio of the
the development of advanced adsorbents for uranium recovery two comonomers. Subsequent ATRP of AN was initiated from
from seawater. the chloride in the poly(VBC-co-DVB), which was subse-
Brown and colleagues recently reported the complete quently converted to amidoxime by treatment with hydroxyl-
elimination of the RIGP step to further advance the preparation amine. Figure 62 displays a representative schematic of the
of adsorbents via ATRP.51 In order to afford ATRP initiation synthesis process.
sites, a new class of trunk materials were prepared by When screened for uranium adsorption in a 6.8 ppm uranyl
chlorination of PP round fibers, a hollow-gear-shaped PP brine solution at pH 8, the material achieved an uptake of 80
13975 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 58. Preparation of the PE-graft-poly(VBC)-graf t-poly(amidoxime-co-tBA) adsorbent by combination RIGP and ATRP, affording a brush-on-
brush structure. Installation of a highly hydrophilic block at the end of the graft chain afforded improved uranium uptake, also demonstrating the
importance of polymer morphology and hydrophilicity in adsorbent performance. Figure reprinted with permission from ref 52. Copyright 2014,
Royal Society of Chemistry.

Figure 59. Chlorination of shaped polyolefin trunks and subsequent graft-chain polymerization by ATRP to afford adsorbents for seawater uranium
recovery. Figure reprinted with permission from ref 51. Copyright 2016, American Chemical Society.

mg of uranium/g of adsorbent, outperforming several conversion to PAF-graf t-poly(AO) with hydroxylamine.


commercially available adsorbent materials. Furthermore, Uranium uptake was assessed in laboratory screening tests
when deployed in 5 gallons (18.92 L) of environmental using a 6 ppm uranium brine solution, where the adsorbent
seawater for 27 d, an uptake of 2.0 mg of uranium/g of outperformed several commercially available adsorbents as well
adsorbent was achieved.191 as a reference fiber obtained from the JAEA. Uptakes of 64.1
The same group applied ATRP to the functionalization of a and 65.2 mg of uranium/g of adsorbent were obtained for the
porous aromatic framework (PAF) with poly(AN), followed by “as prepared” sample, as well as following a KOH-conditioning
13976 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 60. Synthesis of PVC-co-CPVC-graft-poly(AO-co-tBA) for recovery of uranium from seawater. This material currently displays the best
performance for extraction of uranium from natural seawater, surpassing 5 mg of uranium/g of adsorbent over 49 d of deployment. Figure reprinted
with permission from ref 78. Copyright 2016, American Chemical Society.

step.192 The statistically insignificant difference in uranium


uptake indicates the alkaline preconditioning normally required
for polymer adsorbents is not necessary for this particular
material. While easy to overlook this important detail, the
identity of the trunk is the primary difference that differentiates
this polymer adsorbent from the many different adsorbent
materials discussed above, and further research is warranted to
elucidate the relationship between trunk material and need for
KOH conditioning prior to achieving promising performance.
The adsorbent was also tested via batch-contact in filtered
environmental seawater spiked to an 80 ppb uranium
concentration. Following 42 d of contact, an adsorption
capacity of 4.81 mg of uranium/g of adsorbent was achieved.
Furthermore, kinetics analysis revealed approximately 80% of
Figure 61. Uranium recovery from filtered environmental seawater by
all uranium was adsorbed after 2 weeks of contact, suggesting a
the PVC-co-CPVC-graf t-poly(amidoxime-co-tBA) material. Calculated reduced deployment time would be plausible. Nevertheless, no
values for saturation and half-saturation are afforded based on the appreciable uranium uptake was detected by the authors when
single site binding model. JAEA reference fibers deployed and the material was deployed in unspiked seawater, though the low
collected simultaneously with the PVC-co-CPVC-graf t-poly- concentration of uranium was near the limits of their analytical
(amidoxime-co-tBA) material are provided as open symbols, with quantification.192
colors coordinating with the appropriate series. Figure reprinted with Chi and co-workers further developed the above work,
permission from ref 78. Copyright 2016, American Chemical Society.
varying the ratio of DVB:VBC and polymerizing AN by
ATRP.193 As in the work by Yue and colleagues, the surface

Figure 62. Generic process for preparation of poly(DVB-co-VBC) support, formation of poly(AN) graft by ATRP, and formation of amidoxime
species by treatment with hydroxylamine. Figure reprinted with permission from ref 193. Copyright 2016, Springer.

13977 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 5. Performance of ATRP-Prepared Adsorbents for Recovery of Uranium from Water and Seawater
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
PE-graf t-poly(VBC)-graf t-
poly(AO-co-tBA)
AN:tBA = 1000:0 179a 6 ppm uranium, with NaCl and 1.52 42 d column, continuously replenished 52
NaHCO3, pH 8 (Sequim, WA, U.S.A.)
AN:tBA = 1000:100 174.1a 6 ppm uranium, with NaCl and 1.56 42 d column, continuously replenished 52
NaHCO3, pH 8 (Sequim, WA, U.S.A.)
AN:tBA = 1000:738 161.3a 6 ppm uranium, with NaCl and 1.08 42 d column, continuously replenished 52
NaHCO3, pH 8 (Sequim, WA, U.S.A.)
1.18 56 d column, continuously replenished 52
(Sequim, WA, U.S.A.)
AN:tBA = 1000:950 6 ppm uranium, with NaCl and 0.99 55 d column, continuously replenished 52
NaHCO3, pH 8 (Sequim, WA, U.S.A.)
PE-graf t-poly(VBC)-graf t-
poly(AO-co-tBA)-graf t-
poly(tBA)
%Dg poly(tBA) = 0 80a 6 ppm uranium, with NaCl and 1.56 42 d column, continuously replenished 52
NaHCO3, pH 8 (Sequim, WA, U.S.A.)
a
%Dg poly(tBA) = 13.7 79.4 6 ppm uranium, with NaCl and 3.02 56 d column, continuously replenished 52
NaHCO3, pH 8 (Sequim, WA, U.S.A.)
PP-graf t-poly(AO-co-tBA) (round 147a 6 ppm uranium, with NaCl and 51
fibers) NaHCO3, pH 8
PE-graf t-poly(AO-co-tBA) (hollow 81a 6 ppm uranium, with NaCl and 51
gear fibers) NaHCO3, pH 8
PVC-co-CPVC-graf t- poly(AO-co- 174.7a 6 ppm uranium, with NaCl and 5.2 49 d column, continuously replenished 78
tBA) (round fibers) NaHCO3, pH 8 (Sequim, WA, U.S.A.)
3 49 d column, continuously replenished 78
(Sequim, WA, U.S.A.)
2.6 42 d column, continuously replenished 78
(Sequim, WA, U.S.A.)
poly(VBC-co-DVB)-graf t- 80a 6 ppm uranium, with NaCl and 2 27 d batch, 18.9 L fixed volume 191
poly(AO) NaHCO3, pH 8 (h)
PAF-graf t-poly(AO) 65.1a 6 ppm uranium, with NaCl and 14 d batch, 1 L fixed volume; below 192
NaHCO3, pH 8 (h) analytical limit of detection
4.8 Environmental seawater, 192
spiked to 90 ppb uranium
poly(VBC-co-DVB)-graf t- 75 100 ppm uranium, pH 8.3 193
poly(AO)
2 10 ppb uranium, pH 8.3 193
a
Solution composition: 10123 ppm of Na+, 15529 ppm of Cl−, 140 ppm of HCO3, 750 mL of solution

Figure 63. A poly(St)-supported uranyl receptor, as reported by Rebek and colleagues. Figure reprinted with permission from ref 194. Copyright
2013, Royal Society of Chemistry.

area and pore architectures of the poly(DVB:VBC) could be achieved. When the concentration was reduced to 10 ppb, an
tuned as a function of the DVB:VBC ratio. Furthermore, the uptake of 2 mg of uranium/g of adsorbent was reported.
authors noted that poly(AN) length varied as the structure of The uranium uptake performance of adsorbents prepared by
the poly(DVB-co-VBC) support changed. While the highest Dg ATRP are summarized in Table 5.
was obtained for the material formed on the 1:1 DVB:VBC
5.3. Rational Design and Synthesis of Uranyl-Chelating
support, the best performing adsorbent was prepared on 2:1 Ligands
DVB:VBC, suggesting the structure of the support may
supersede Dg in determining adsorbent performance. When 5.3.1. Introduction. While promising development has
contacted with a 100 ppm aqueous uranyl solution at pH 8.3, a been reported through preparation of polymer adsorbents, a
uranium uptake of 75 mg of uranium/g of adsorbent was second school of thought involves the rational design and
13978 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

synthesis of engineered uranyl receptors. For example, Rebek


and colleagues reported an ion exchange polymer where a
ligand derived from Kemp’s triacid oriented three carboxylates
to bind uranyl in the equatorial plane, while an N−H bond is
oriented toward the axial oxygen (Figure 63). This receptor
could be immobilized on PEG and subsequently supported on
a minimally cross-linked poly(St) substrate at 0.22 mmol g−1.
Uranium recovery from seawater spiked to 400 ppb with
uranium at pH 8.4 afforded recovery of 332 ng of U and was
able to be recycled over 15 times with negligible degradation in
performance.194
Although the aforementioned work is not on a support
capable of deployment in seawater, the general strategy
provides an important point for consideration. Rather than
forming long graft chains of poly(amidoxime-co-hydrophilic
monomer), better adsorbent performance may be achieved
through application of supramolecular chemistry principles with Figure 64. Plot of log(K/K0) vs Hammett σpara constants for the
rational design of highly selective chelators. Moreover, as adjacent substituents on oxime- derived ligands binding to uranyl. The
receptors are necessarily well-defined, such an approach is red line is the linear regression, with R2 = 0.9773. The point
inherently amenable to application of high-performance corresponding to the dimethyl amine substituent (N(Me)2) was not
computational design. This section summarizes recent work included due to sterically induced distortion in orbital conjugation.
Inset: the scope of oximate-derived ligands. Reprinted with permission
on the design of such engineered uranyl receptors, with from ref 195. Copyright 2013, American Chemical Society.
particular emphasis on application of computation to the design
and improvement of uranyl-binding species.
5.3.2. Computationally-Guided Amidoxime Develop- kcal mol−1. The noteworthy exception to this trend was
ment. Early improvements to the amidoxime functionality dimethylamidoxime, which possessed the strongest σpara value
were proposed by Abney et al., who performed DFT of −0.83, yet bound uranyl with a ΔG of −54.63 kcal mol−1.
calculations to identify critical orbitals participating in The authors rationalized this discrepancy by observing steric
uranium-amidoxime bonding.195 Natural bond orbital (NBO) interactions from the one methyl group resulted in disruption
analysis of [UO2(AO)(H2O)3]+ revealed bonds between of the conjugated π-orbitals, reducing the effective electron
uranium and hybridized orbitals from O and N were more donation.
than 85% localized on the light elements, and the authors Continuing this line of reasoning, the authors suggested an
observed that varying the energy levels of the bonding orbitals imidazole-derived oxime would display enhanced bond strength
would result in differing bond strength. By increasing electron with uranyl. Computationally this was observed, as an
donation, the energy level of the bonding molecular orbitals [UO2(imidazole oxime)(H2O)3]+ complex achieved a bond
should increase to afford greater contribution from uranyl and a strength of −60.92 kcal mol−1. However, the authors noted
more thermodynamically stable bond. When η2-bound to instability of this complex was reported in previous literature,
uranyl, a series of conjugated π-orbitals connect the oxime and subsequently suggested a reduced imidazolidine-oximate
functionality with the adjacent amine, affording a convenient ligand which would display enhanced stability and bond
route for varying the electron density on the oxime and strength of −58.06 kcal mol−1.
modulating the energy of the electronic orbitals. 5.3.3. De Novo Design and Synthesis of Uranyl
To investigate this hypothesis, the authors performed DFT Chelating Ligands. Design of host structures capable of
calculations on a series of oximate-derived ligands, coordinated selective guest binding is a central area of research in the
in an η2-fashion to uranyl. Substituents adjacent to the oxime subdiscipline of supramolecular chemistry. Such chemical
were selected on the basis of their electron donating character recognition is achieved when three criteria are met. First, the
via resonance, with Gibbs Free Energies (ΔG) for uranyl host structure must contain at least two binding sites that
binding calculated by DFT and used to determine the interact favorably with the intended guest. Second, the host
equilibrium constant K, assuming a temperature of 298 K. must provide a “complementary array” of these binding sites for
The unfunctionalized acetoxime ligand, with hydrogen in the interacting with the guest. That is, the binding species must be
adjacent position, was used as a reference and the resulting K oriented in such a fashion as to simultaneously bind the guest.
value defined as K0. The log of K/K0 for each ligand was plotted Finally, preorganization of the complementary array is
against the corresponding Hammett σpara value, as obtained necessary or at least a minimal number of stable conformations
from the literature.196 The Hammett σ value is a measure of a should be achievable for the host, with the ideal conformation
substituent’s ability to influence the rate of a reaction due to low in energy relative to the alternatives. The application of
electron donation or withdrawing. such structure-based approaches can compensate for modest
The plot of log(K/K0) revealed a direct correlation between affinity, achieving strong uranium binding through cooperative
uranyl binding strength and the electron donating properties of interactions. This was clearly demonstrated through develop-
the adjacent substituent (Figure 64). Functionalized with a ment of engineered proteins, where weakly coordinating atoms
nitro group, a strongly electron withdrawing substituent with were oriented to form a uranyl binding pocket, achieving
σpara value of 0.78, nitrooxime bound uranyl with a ΔG of femtomolar affinity for uranium.197
−27.34 kcal mol−1. In contrast, when functionalized with one Upon identification of the η2-motif as a valid configuration
methyl group and possessing a σ para value of −0.75, for uranyl binding, rational design of a selective chelating ligand
monomethylamidoxime bound uranyl with a ΔG of −56.45 could be pursued. Binding sites with uranyl were identified as
13979 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

the equatorial plane, and 2−3 AO ligands could readily modate some flexibility within the proposed complexes,
constitute the complementary array of binding moieties. consisting of ±20° for in-plane angle variation of the AO
However, identification of the host architecture necessary to groups, ±20° for the twist of the AO group, and ±15° for
preorganize the AO functionalities remained a challenge. The rotation about the N−O bond.
software HostDesigner was initially developed by Hay and HostDesigner constructed and scored 180 million geometries
colleagues198 and subsequently adapted for the design of in under 20 min to achieve the initial ranking. Subsequent
anionic hosts199,200 and the design of high-symmetry molecular molecular mechanics calculations determined ligand strain
polyhedra.201,202 Starting from an appropriate array of binding upon metal complexation for the top 1000 candidates,
sites located at optimized positions around the intended guest, reordering as appropriate. Conformational searches performed
the software identifies host architectures by assembling on the top 200 candidates, resulted in 87 candidates with ΔGrel
structures from user-defined input fragments to achieve the ≤ 10 kcal mol−1. The top 10 candidates all exhibit ΔGrel ≤ 5
appropriate preorganized arrangement for optimally binding kcal mol−1, with the top scoring candidate 2.7 kcal mol−1.
host molecule. Once HostDesigner had identified the best These fragments are displayed in Figure 66.
candidates from a geometrical perspective, subsequent
molecular mechanics analyses refined the top contenders.
HostDesigner was used to identify promising uranyl
receptors by the aforementioned approach.203 Assuming uranyl
forms a tris-carbonate [UO2(CO3)3]4− in seawater, displace-
ment of two carbonates by AO ligands was determined to
afford a good model for designing a complementary host.
Geometries of [UO2(AO)2(CO3)]2− were optimized by
quantum mechanical calculations, affording three potential
configurations. While not isoenergetic, the most thermody-
namically stable complex was achieved when the AO C−H
groups are facing carbonate, requiring an unreasonably long
linking molecule. By rotating the AO so the C−H groups face
each other, reasonable configurations were able to be obtained,
but at a mild energy penalty of 1.56 kcal mol−1. This
identification process is depicted graphically in Figure 65. Figure 66. Top hits identified by HostDesigner screening and ranked
Structural degrees of freedom were also applied to accom- by ΔGrel. Reprinted with permission from ref 203. Copyright 2013,
American Chemical Society.

Qualitative comparison of the top ranked candidates reveals


four sp3 or four to five sp2 bridging carbon atoms are necessary
to minimize reorganization energy and create a high-scoring
host. However, a lack of synthesizability hinders this computer-
guided approach. Inclusion of chiral centers demand the costly
and laborious preparation of enatiopure precursors. Alkene or
diene-comprised hosts are chemically reactive. Further filtering
of the results based on synthetic conditions resulted in a series
of three candidates for preparation and experimental testing,
displayed in Figure 67. However, these ligands have not been
successfully synthesized or grafted onto polymers for
subsequent testing as of the time of publication.
5.3.4. Determination of Binding Constants for
Uranium with Amidoxime and Derivatives. Understand-
ing the speciation of uranium ions in aqueous solution is the
basis for predicting and controlling the behavior of uranium
binding and extraction process.204−206 This understanding also
provides the essential foundation for the rational design of
ligands with enhanced uranium ion affinity and selectivity in
seawater. Central to this endeavor, an important and
fundamental characterization of ligand reactivity involves the
determination of equilibrium constants for the formation of
simple 1:1 metal:ligand complexes with a series of metal ions.
Such values, which are referred to as stability constants or
formation constants, are determined most frequently in
Figure 65. Relative free energies in water of optimized geometries for aqueous solution. In practice, stability constants are normally
uranyl complexes containing two AO ligands and one carbonate. The defined as the ratio of concentrations rather than activities. In
orange arrows denote the location where a connecting fragment will be the case of 1:1 complexes, the aqueous stability constant is
placed by subsequent HostDesigner analysis. Reprinted with defined as the concentration of the metal complex, ML, divided
permission from ref 203. Copyright 2013, American Chemical Society. by the concentrations of the metal ion, M, and the ligand, L, as
13980 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

compared with experimental data. These results were then


plotted for 13 distinct complexes, spanning almost 17 orders of
magnitude. Though calculated log K1 values were systemically
overestimated by up to 39 log units, a plot of experimental log
K1 values versus calculated log K1 values afford linear
correlations with R2 ≥ 0.950. These results are displayed in
Figure 68.134
The computationally directed preparation of improved
amidoxime functionalities and de novo designed uranyl
chelators leverage a powerful tool for the preparation of
adsorbents with great promise for enhanced selectivity and
uranyl uptake. Furthermore, application of computational
screening of uranyl binding provides a powerful tool for
identifying small molecules with high affinity for uranyl
chelation. Judicious application of these approaches holds
great potential for high-throughput computational screening
investigations, reducing waste and expediting discovery.
Ongoing efforts should be devoted to completing the synthetic
work necessary to synthesize N-methylamidoxime and N,N-
dimethylamidoxime functionalized polymer adsorbents, as well
adsorbents grafted with the uranyl receptors identified by
HostDesigner. While exciting results have been obtained from
Figure 67. Viable candidates identified by HostDesigner for
the first round of HostDesigner-based investigation, it is
subsequent synthesis and uranium extraction studies. Figure reprinted
with permission from ref 203. Copyright 2013, American Chemical peculiar that only two ligands were considered sufficient to
Society. form an array of binding sites. Recent reports indicate
structurally based adsorbents can display high affinity for
uranyl through cooperative adsorbent interactions, despite
shown by eq 1. Because they may span many orders of possessing individual log K1 values far below those listed in
magnitude, stability constants are usually reported in Figure 68.53,197 Using HostDesigner or similar software to
logarithmic form, log K1. Individual log K1 values quantify the design adsorbents with 3-D structure capable of fully occupying
strength of the metal−ligand interaction and the difference the equatorial coordination sphere should be pursued fervently.
between log K1 values for two metal ions measures the degree One major downside of computationally guided efforts is the
of selectivity. challenge and cost of subsequent synthetic confirmation. The
[ML] fact that, despite a list of 87 candidates, only three ligands were
M + L ⇌ ML K1 = deemed synthetically plausible is a glaring condemnation.
[M][L] (1)
Though appropriate from a perspective of fundamental science
Applying necessary corrections for using pure H2O as a and publication, identification of host molecules with facile
reference in thermochemical calculations, log K1 values were synthetic avenues is vital for implementation of these results to
calculated for a series of 18 donor ligands binding uranyl and real-world applications.

Figure 68. (Left) Plots of experimental log K1 versus calculated log K1, as obtained in aqueous solution. Darkened circles are for monoanions, while
open circles are dianions. Linear correlations had R2 values of 0.967 and 0.950 for mono- and dianions and were represented by the equations expt
log K1 = 0.309 × calc log K1 − 2.047 and expt log K1 = 0.473 × calc log K1 − 8.108, respectively. (Right) Plot of experimental log K1 versus
calculated log K1 for select dianionic uranyl chelating ligands. The linear correlation indicates computational screening can afford a powerful tool for
predictive determination of experimental uranyl affinity. Figures reprinted with permission from ref 134. Copyright 2015, American Chemical
Society.

13981 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Finally, it is worth commenting that all of the computation- was observed by EXAFS. In contrast to research proposing
ally guided efforts are underpinned by several critical single functionalities could bind uranyl, EXAFS fits revealed a
assumptions, foremost of which is η2 binding mode between coordination environment with at least two binding ligands.
amidoxime and uranyl. As noted in Section 5.1.2, binding of Importantly, the uranyl coordination environment is distinctly
formamidoxime to uranyl is effectively isoenergetic for η2 and different depending on whether adsorbents were contacted with
chelating motifs, while acetamidoxime and benzamidoxime seawater simulant or environmental seawater, with environ-
prefer η2 mode by only 3 kcal mol−1. This relatively small, mental seawater spectra possessing features indicative of an
computationally determined energy difference should be placed adjacent metal forming an oxo-bridge with uranium. This work
in context. An 11 kcal mol−1 barrier was deemed experimentally casts doubt on the assumption that small molecule and
feasible for vanadium binding, and the ring flip of cyclohexane, computationally guided investigations are representative of
which occurs readily at room temperature, proceeds through a complex polymeric systems under environmental conditions.
10 kcal mol−1 half-chair intermediate. Free rotation of ethylene
around the σ bond, with the eclipsed conformer 2.9 kcal mol−1 6. NANOSTRUCTURED MATERIALS
higher in energy than staggered, is similar in energy to the Nanostructured materials represent a new category of
calculated difference between η2 and chelating motifs. adsorbents which were not available during initial development
While crystal structures confirm η2-binding to be possible, of materials for seawater uranium recovery. As such, they are
molecular packing interactions can strongly dictate which technologically far less mature than the more established
crystal phases are obtained. Furthermore, while there is adsorbent materials discussed in preceding sections and
apparent agreement between computation and crystallography frequently have not been tested under environmental
in terms of bonding motif, there is no consensus regarding the conditions. With the development of advanced analytical and
appropriate number of equatorially coordinating aqua ligands. processing techniques, the ability to synthesize, modify, and
Justification of computational work should not hold a double- functionalize materials on the nanoscale regime allows an
standard, claiming the bonding motif is confirmed while unprecedented ability to engineer materials for specific
ignoring diverging results in coordination number. applications. A much larger proportion of nanomaterials are
Importantly, these results are all obtained under extensively surface-accessible than a comparable quantity of bulk material.
controlled, highly idealized circumstances, which may not be This intrinsic characteristic engenders nanomaterials and
representative of complex, dynamic conditions present for nanostructured materials with several key properties, including
amorphous polymers deployed in environmental seawater. high surface areas, high binding site densities, and rapid mass
Indeed, recent work using X-ray absorption fine structure transport properties. These features are expected to facilitate
(XAFS) spectroscopy attempted to validate the use of small uranyl extraction selectivity, kinetics, and capacity, perhaps even
molecule standards to predict uranyl bonding on an to the extent that one day they may replace polymer-based
amidoxime-functionalized polymer.128,207 However, as shown adsorbents as the benchmark technology for this application.
in Figure 69, the extended XAFS (EXAFS) spectra reveal Despite the vast promise afforded by accessing nanoscale
distinct differences between the uranium coordination environ- regimes, such materials are not technologically mature and
ment for small molecule standards and adsorbent-bound uranyl. significant additional research emphasis is necessary prior to
While previous small molecule and computational research their meaningful application as uranium adsorbents.
suggested cyclized amidoxime groups are responsible for uranyl 6.1. Porous Carbon
binding,63,138,208,209 no contribution from this binding motif Activated carbons provide an exciting platform for the
development of new uranyl adsorbents. Prior to functionaliza-
tion, they display minimal background sorption210 and are
sufficiently stable for deployment in environmental seawater
without risk of degradation. Due to preparation by a variety of
routes, pore architectures can be tuned and systematically
investigated. Often possessing surface areas from 250−3000 m2
g−1 and capable of processing into monolithic materials or
sheets, they show significant promise as effective sorbents for
extracting uranium from seawater.
The first instance of functionalized carbon for extraction of
uranium was by Starvin et al. in 2004.211 An activated carbon
compound was functionalized with diarylazobisphenol through
treatment in an aqueous solution, similar to previously reported
functionalization of single-wall carbon nanotubes through
chemical modification with a diazonium compound.212 No
information is provided regarding surface area coverage or even
Figure 69. Comparison of EXAFS spectra for seawater-contacted characterization of the carbon material. It was packed into a
adsorbent, simulant-contacted adsorbent, UO2(glutarimidedioxime)2 column and investigated for preconcentration of uranium in the
small molecule standard, and UO2(benzamidoxime)2(MeOH)2 small
molecule standard. Elution refers to a 1 h soak in dilute HCl, eluting
pH range of 4−5, saturating at approximately 18 mg of
weakly bound transition metals. These spectra show the uranium uranium/g of adsorbent.211
coordination environment on adsorbents to be distinctly different Unfunctionalized activated carbon was also investigated for
from both motifs displayed by small molecule standards. Figure batch adsorption. Varying contact time, pH, concentration of
reprinted with permission from ref 128. Copyright 2016, Royal Society uranium, and temperature, a maximum uptake of 28 mg of
of Chemistry. uranium/g of adsorbent was obtained at pH 3 and 20 °C with 4
13982 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

h contact time.210 Uranium sorption was observed to decrease seawater cations, such as sodium and calcium (Figure 71). In
significantly below pH 2 and above pH 6. It was suggested contrast to traditional electrosorptive approaches, HW-ACE
optimal saturation occurs at pH 3 because this is the acidity at uses an electric field to guide the migration of the uranyl ions,
which “free” uranium is maximized, meaning uranyl speciates to preconcentrating the solution surrounding the electrode. The
a penta−aqua complex and there is minimal formation of amidoxime functionalities then readily bind the uranium,
dimers or tetramers. Sorption data were not collected above pH demonstrating enhancements in selectivity, kinetics, and
7, but this study reveals negligible sorption would be expected uptake.
by unfunctionalized activated carbon under seawater con- As displayed in Figure 71, adsorption by a HW-ACE method
ditions.210 can be conceptualized as a sequence of five repeating steps,
Electrosorption using activated carbon slurries prepared with which correspond with the cycling of the bias on the
polyvinylidene fluoride (PVDF) has also been studied.213 This electrode.214 Step I shows randomized ions in a seawater
technique utilizes the conductivity of the carbon to extract ions solution, prior to application of the bias, while step II depicts
through isolation within the electrical double layer formed the application of a negative bias, inducing the migration of the
when a bias potential is applied to the electrodes. Through ions and formation of an electrical double layer. Notably,
optimizing the applied potential at +0.4 V (Figure 70), uranyl application of a negative bias requires the adsorption of cations,
UO22+, rather than the anionic UO2(CO3)34− reported in
previous work.213 Step III reveals the binding and deposition of
uranium on the surface of the electrode. Due to the presence of
the amidoxime functionality on the surface of the carbon
electrode, only uranium remains bound upon removal of the
bias, while other ions redisperse in solution. Step V reveals
further growth of uranyl as a function of more cycles.
The authors first tested the amidoxime-functionalized carbon
electrode and HW-ACE method in seawater spiked with
uranium concentrations ranging from 150 ppb to 2000 ppm,
revealing the formation of (UO2)O2·2H2O crystals on the
surface of the electrode. In a head-to-head comparison against
pure physicochemical adsorption, performance improvements
through application of the HW-ACE approach were obvious.
Near 100% recovery of uranium was achieved by HW-ACE
across all concentrations, while physicochemical adsorption
varied from approximately 10% to 80%, as displayed in Figure
72. The difference in performance was also observed by EDX
Figure 70. Electrosorption of uranium from seawater as a function of using a poly(amidoxime)-coated electrode patterned with
applied potential (vs Ag/AgCl). Figure reprinted with permission from parallel Pt lines on an insulating quartz substrate. After
ref 213. Copyright 2015, Elsevier. extraction of uranium, EDX mapping clearly revealed uranium
binding to map directly on the conduction Pt pattern on the
can be selectively separated from seawater. A capacity after 300 electrode (Figure 72), confirming that the application of the
min of applied potential was reported as 3.4 mg of uranium/g bias and resulting uranium concentration resulted in the
of adsorbent, comparable to high surface area adsorbents improved uranium uptake. When subsequently tested in
recently developed at ORNL. The authors speculate that no filtered, unspiked, environmental seawater, the HW-ACE
cationic species is adsorbed; therefore, the triscarbonatouranyl- approach achieved a uranium uptake of 1.62 μg, compared to
(VI) anion must be the electrosorbed species. However, the 0.55 μg for physicochemical adsorption.
calcium and magnesium response, two cations reported to have Benzoylthiourea-functionalized activated carbons were pre-
strong binding affinity to the triscarbonatouranyl(VI) anion, do pared through in situ generation of diazonium cations.215 These
not appear to increase with the uranium uptake. While this materials were investigated for uranyl sorption from pH 1−5,
work is promising, more work is required due to the large with differing speciation of uranyl again providing the rationale
decrease in surface area from the use of 50 wt % binder for why sorption was not continued to environmental pH. A
(PVDF) to stabilize the activated carbon electrode. Further saturation capacity of 82 mg of uranium/g of adsorbent was
work is required to understand the electrosorption process and obtained, which surpassed the unmodified activated carbon by
further characterize the ion extraction as no high valent metals three times. The binding motif for the benzoylthiourea-
are discussed outside of uranium. The high oxidation state functionalized carbon was proposed to be chelation of one
transition metals should be electrosorbed as well and will uranyl by two cooperating benzoylthiourea functionalities
impact the reported economics of the process. through the carboxylic oxygen and thiourea sulfur, while the
Very recently, Cui and co-workers reported the application of imide and amine nitrogen were proposed to be Lewis bases
an amidoxime-functionalized conducting carbon electrode in a donating electron pairs to other uranyl molecules.215
half-wave rectified alternating current electrochemical (HW- CMK-3, a nanoporous ordered carbon, was functionalized
ACE) method for the extraction of seawater uranium.214 As with a carboxymethylated polyethylenimine substrate. Upon
noted by the authors, conventional physicochemical adsorption functionalization, the surface area of the material decreased by
of uranyl is limited due to slow diffusion of uranyl ions due to approximately 50%, while degree of functionalization was
low seawater uranium concentration, poor binding site estimated at 24% by total organic carbon analysis. Sorption
accessibility due to Coulombinc repulsion from adjacent isotherms displayed good sorption from pH 3−4, with
bound metals, and strong competition from other ubiquitous saturation capacities of 119 and 222 mg g−1, respectively. It
13983 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 71. Depiction of difficulties in traditional physicochemical adsorption of uranium by receptor-functionalized adsorbents, and mechanism of
the HW-ACE extraction approach. Panel (a) displays uranyl repulsed by Coulombic forces of other bound ions, and panel (b) displays the
competition between uranyl and other ubiquitous seawater cations. Panel (c) depicts the HW-ACE method, where a voltage applied to the
amidoxime-functionalized carbon electrode alternates between a negative value and zero, resulting in concentration and selective binding of uranium.
Figure reprinted with permission from ref 214. Copyright 2017, Nature Publishing Group.

Figure 72. Uptake of uranium from spiked seawater using the HW-ACE method (orange) and traditional physicochemical adsorption (purple).
Performance was analyzed for initial uranium concentrations of (a) 150 ppb, (b) 1.5 ppm, (c) 15 ppm, (d) 400 ppm, (e) 1000 ppm, and (f) 2000
ppm. The bottom right plot shows a schematic of the patterned electrode, with accompanying SEM image (top right) and EDX maps for Pt (bottom
left) and uranium (bottom right) following uranium adsorption by the HW-ACE method. Figure reprinted with permission from ref 214. Copyright
2017, Nature Publishing Group.

was proposed further deprotonation of the imine and carboxylic 7.5. Interestingly, the pristine CMK-3 also displayed a
acids resulted in the improved sorption capacity at higher maximum sorption capacity of 120 mg g−1 at pH 6, decreasing
pH.216 CMK-3 was also functionalized by oxidation to generate to 90 mg g−1 at pH 7.5.217
a surface carboxylic acid. A maximum saturation capacity of 190 A related activated carbon material, CMK-5, was function-
mg g−1 was observed at pH 5.5, decreasing to 160 mg g−1 at pH alized with 4-acetophenone oxime by thermally induced
13984 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

diazonitization with isoamyl nitrite (Figure 73).218 Batch Competitive adsorption of uranyl from simulated nuclear
sorption experiments were performed to investigate uranyl processing effluent revealed remarkable selectivity for uranyl,
with a Kd of approximately 9.5 × 104 mL g−1.219
In an effort to provide a more direct basis for comparison,
Carboni et al. prepared a library of organic functional groups
and grafted them to commercially available mesoporous carbon
through formation of in situ diazonium cations (Figure 74).95
Containing many functional groups investigated on MS
materials by Vivero-Escoto,96 a similar result was obtained: a
phosphylated mesoporous carbon material possessed the
highest sorption capacity, at both pH 4 and 8.2. Saturation
Figure 73. Schematic of the diazonium-based functionalization of capacities of 85 and 50 mg g−1 were obtained under these
CMK-5 ordered mesoporous carbon with aminoacetophenone oxime. conditions, respectively. Not surprisingly, sorption of uranyl
Figure reprinted with permission from ref 218. Copyright 2011,
was pH-dependent, with a dramatic increase observed above
Elsevier.
pH 3.5 attributed to the first deprotonation of the surface
phosphoryl groups.95
sorption from simulated effluent from nuclear processing. The
simulant contained multiple competing di- and trivalent Sonochemical grafting of AN into mesoporous carbon was
transition metals as well as several lanthanides. While poor also pursued for development of uranyl-extracting materials.220
sorption was observed below pH 3, a maximum sorption The authors report an increase in degree of grafting through
capacity around 65 mg g−1 was obtained at pH 4.5. This was sonochemical approaches compared to efforts with free radical
rationalized by the extent to which the oxime could be polymerization. They propose pore blockage occurs during free
deprotonated, forming a proposed uranyl complex of four radical polymerization, but penetration of the monomer is
equatorial monodentate-coordinating oxime ligands.218 encouraged through sonication. The authors reported that a
A salicylaldeneimine-functionalized hydrothermal carbon was high intensity ultrasonication, using an ultrasonic horn as the
prepared by functionalizing an activated hydrothermal carbon source, provided more conformal polymer coatings within the
with ethylenediamine.219 Condensation with salicylaldehyde pores compared to the standard lab bath ultrasound. This was
resulted in the final material, which was investigated for observed through the nitrogen sorption isotherms where a
extraction of uranyl over the pH range of <1−4.5. Uptake hysteresis of the lab bath ultrasound was extended instead of
increased with increasing pH, with a maximum uptake of 261 sharply closing, as in the case of the high intensity ultrasound, a
mg g−1 obtained at pH 4.3, while kinetic studies revealed more phenomenon attributed to pore constrictions (Figure 75).
than 80% of the total adsorption occurred in the first minute. Homogeneous polymerization of poly(AN) throughout the

Figure 74. Substrate scope investigated by Carboni et al. for uranium extraction on a common mesoporous carbon support. Figure reprinted with
permission from ref 95. Copyright 2013, American Chemical Society.

13985 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

imately 0.8 nm equatorially and 0.6 nm axially. This represents


a bulky anion that can significantly block microporosity thereby
negating the rest of the polymer in the micropores. It was also
determined that the larger pores are required as the
mesoporosity did not significantly contribute to the overall
capacity.
Mayes and colleagues subsequently reported a systematic
study investigating the impact of pore size on uranyl adsorption
by mesoporous carbon materials.221 Applying the aforemen-
tioned sonochemical polymerization technique,220 poly(AN)
(later modified to form amidoxime sites) was polymerized from
carbon supports which had been rationally modified to exclude
mesoporosity, microporosity, or macroporosity.221 The authors
observed N2-adsorption isotherm surface area measurements
were not representative of uranium-binding performance nor
was the polymer Dg. However, pore size and the accompanying
Figure 75. Nitrogen sorption isotherms for sonication-assisted curvature of the pore wall were directly correlated with uranium
polymerization within mesoporous carbon. The oa-C-p-25.6 sample extraction. While affording large specific surface areas, micro-
represents high intensity ultrasonication while the oa-C-b-24.0 porosity was detrimental to uranium capacity, putatively due to
represents a standard lab bath ultrasound. Figure reprinted with pore-filling or the inability of uranyl to access buried amidoxime
permission from ref 220. Copyright 2013, Royal Society of Chemistry. binding sites. In contrast, mesoporosity and macroporosity
were both advantageous to achieving good performance.
Carbon nanotubes have also been investigated for uranium
mesoporous material was coupled with nitric acid surface sorption.222−224 First reports involve acid treatment for
oxidation and CO2 pore activation to achieve enhanced sorbent enhancing colloidal stability, with sorption capacities of
performance. Modest sorption capacities of <1−4 mg g−1 were approximately 45 mg g−1 achieved at pH 5.222 Similar results
obtained, but these experiments were performed in a seawater were obtained by oxidation of multiwalled carbon nanotubes,
simulant at pH 8.2 with carbonate sources, which typically
affording a sorption capacity of 36 mg g−1.223 Finally,
reduces sorption dramatically.220
carboxymethyl cellulose was also attached to the surface of
To overcome the low capacities observed by mesoporous
multiwalled carbon nanotubes by plasma induced grafting. After
carbons, a combination of templating techniques were utilized
exposure to an N2 plasma, the nanotubes were combined with
to generate hierarchical porosity.221 Hard templating with
carboxymethyl cellulose and heated to 80 °C for 1 week with
uniform silica spheres was used to generate large mesopores of
stirring. Sorption isotherms obtained at pH 5 yield a sorption
35 and 50 nm average diameters along with 85 nm average
capacity of approximately 110 mg g−1.224
diameter macropores. This was coupled with soft templating to
Very recently, a graphene oxide (GO)/amidoxime hydrogel
generate an average 10 nm mesopore. The carbons were then
adsorbent was developed for the removal of uranium from
further activated with flowing carbon dioxide (CO2) at high
aqueous solution, including seawater simulant.55 Amino acids,
temperature to induce microporosity. This stepwise synthesis
allows for tailoring the porosity. Upon increasing the largest polyamines, and metformin possess multiple nitrogen basic
pore size from 10 to 35 nm, a significant increase in polymer moieties which bind with carboxylic or alcohol groups on GO
grafting and subsequent uranium capacity was observed. sheets, forming hydrogels. Iminodiacetonitirle (IDAN) was
Increasing the pore size to 50 nm enhanced the increase in treated with hydroxylamine to form the amidoxime derivative,
capacity. When 85 nm pores were introduced with the 10 nm AO-IDAN, which was then used as a cross-linking agent in the
mesopore and microporosity, the polymer grafting percentage GO dispersion. Uranium adsorption experiments reveal pH
decreased with an increase in the uranium capacity. This is affects performance significantly, with maximum uptake
believed to be due to better accessibility resulting in greater achieved from pH 5−7. A slight decrease was observed at pH
utilization of the polymer graft allowed by the larger pores. The 8, followed by a precipitous drop at higher pH. When contacted
stepwise approach also facilitated the understanding of the pore with a 3.7 ppb uranium seawater simulant at a phase ratio of 0.5
regime, i.e., micro-, meso-, or macroporosity, on the uranium mg L−1, the hydrogel adsorbed 88% of the uranium in solution,
sorption capacity. It was observed (Table 6) that the affording an uptake of 0.006 mg of uranium/g of adsorbent
microporosity, while good at providing anchor points for the over 6 h. An isotherm performed at pH 6 yields a saturation
polymer within the nanostructure, did not contribute capacity of 398 mg of uranium/g of adsorbent.
significantly to the overall capacity of the sorbent. This is Table 7 summarizes the uranium adsorption performance by
most likely due to the size of the uranyl ion. The nanostructured carbon-based adsorbents.
triscarbonatouranyl(VI) anion has a diameter that is approx- 6.2. Mesoporous ATRP Initiators for Advanced Adsorbents
An exciting development in the preparation of nanostructured
Table 6. Uranium Capacity as a Function of Pore Size materials for uranyl extraction is the preparation of mesoporous
heterogeneous supports from copolymerization of divinylben-
sample pore size (nm) uranium adsorption cap. (g of U/g of ads)
zene and vinylbenzyl chloride.191 These materials previously
85a-AO 2/10/85 20.9 had been developed for other separations processes and
85b-AO 2/85 13.4 catalysis, and varying the ratio of comonomers is known to
85c-AO 85 33.0 afford final materials with controllable porosity and high surface
85d-AO 10/85 28.9 areas without the use of porogens. In addition to providing a
13986 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 7. Performance of Nanostructured Carbons for Recovery of Uranium from Water and Seawatera
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
diarylazobisphenol-activated carbon 18 28 ppb uranium 211
activated carbon 28b pH 3 210
PVDF electrosorption 3.4 5h batch electrosorption 213
HW-ACE on AO-functionalized 1932 seawater spiked to 2000 ppm uranium 0.002 4 L; mass of electrode 214
electrode not provided
benzoylthiourea-activated carbon 82b pH 4.5 215
67c pH 4.5, with competing ions, 227 ppm 215
uranium
polyethylenimine@CMK-3 222b pH 4 216
oxidized CMK-3 190b pH 5.5 217
160b pH 7.5 217
CMK-3 120b pH 6 217
90b pH 7.5 217
4-acetophenone oxime@CMK-5 65d pH 4.5, with competing ions, 181 ppm 218
uranium
salicylaldeneimine@hydrothermal 261b pH 4.3 219
carbon
MC-Ph-AO 3 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-CA 13 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-PhCOOH 15 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-Ph−CH2−COOH 5 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-PhOH 3 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC−COOH 1 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-u-PO(OEt)2 5 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-u-PO(OH)2 4 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-bis(PO(OH)2) 2 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-O-PO(OH)2 50 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-Ph-O-PO(OH)2 5 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-Ph-PO(OEt)2 2 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-Ph-PO(OH)2 9 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-CA/PhCOOH 5 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
MC-Ph-AO/PhCOOH 4 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 95
NaHCO3, pH 8
AO@mesoporous carbon 4 7.9 ppm uranium, 0.453 M NaCl, 2.3 220
(sonochemical) mmol NaHCO3, pH 8
pore regime: 2/10/85 21 7.9 ppm uranium, 0.453 M NaCl, 2.3 221
mmol NaHCO3, pH 8
pore regime: 2/85 13 7.9 ppm uranium, 0.453 M NaCl, 2.3 221
mmol NaHCO3, pH 8
pore regime: 85 33 7.9 ppm uranium, 0.453 M NaCl, 2.3 221
mmol NaHCO3, pH 8
pore regime: 10/85 29 7.9 ppm uranium, 0.453 M NaCl, 2.3 221
mmol NaHCO3, pH 8
carbon nanotubes
acid treated 45b pH 5 222
oxidized 36 pH 6.3, 54 ppm uranium, 0.01 M 223
NaClO4
119 pH 6.9, 54 ppm uranium, 0.01 M 223
NaClO4
cellulose-functionalized MWCNTs 110b pH 5 224
AO-IDAN-GO hydrogel 0.006e 4.2 ppb uranium, seawater simulant 55
398b pH 6 55

13987 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 7. continued

a
Unless otherwise stated, adsorption experiments were performed in DI water and pH was unadjusted. bMaximum capacity at saturation.
c
Competing ion composition includes: 5 mM Na+, 1 mM for Co2+, La3+, Sr2+, and Cs+ dCompeting ion composition includes: 162 ppm of Na+, 113
ppm of Co2+, 124 ppm of Ni2+, 114 ppm Mn2+, 15 ppm Zn2+, 194 ppm of Sr2+, 254 ppm of Ce3+, 86 ppm of Cr3+, 274 ppm La3+ eSeawater simulant
composition: 33.3 g sea salts dissolved in 1 L distilled water

Table 8. Performance of Graft Polymers Generated by ATRP on Porous Carbon Supports


uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
poly(VBC-co-DVB)-graf t- 80a 6 ppm uranium, with NaCl and NaHCO3, 2 27 d batch, 18.9 L fixed 191
poly(AO) pH 8 volume
AO-co-AA-co-DVB 63a 6 ppm uranium, with NaCl and NaHCO3, 0.6 28 d batch, 18.9 L fixed 225
pH 8 volume
a
Composition of solution: 10123 ppm of Na+, 15529 ppm of Cl−, 140 ppm of HCO3−, 250 mL of solution

desirable support, the readily accessible and homogeneously 6000 m2 g−1) and porosities (up to 90% free volume), this new
distributed chlorides were used as initiators for atom-transfer class of materials has been of great interest in recent years.
radical polymerization of poly(AN). Treatment with hydroxyl- Unlike traditional coordination polymers, MOFs possess
amine resulted in conversion to amidoxime functionalities for extensive long-range ordering and can be readily characterized
uranyl sequestration.191 by crystallography, facilitating the study of structure−activity
Applying this process yielded supports with degree of relationships. The ability to extend the length of the organic
grafting ranging from 280−509 wt %, displaying a linear struts while retaining the overall material topology, known as
correlation with density of initiator sites. The authors do note isoreticular synthesis, enables the tuning of pore and channel
extensive grafting may ultimately fill or block mesopores,191 a structure to target desired applications. Finally, orthogonal
phenomenon discussed previously regarding functionalization functionalities can be included through direct synthesis or
of mesoporous carbon.220 Deployment of the mesoporous postsynthetic modification (PSM) to engender desirable
copolymer sorbents in simulated seawater at environmental pH properties to MOFs which would otherwise be chemically inert.
resulted in adsorption capacities as high as 80 mg g−1, While these three fundamental characteristics have made
surpassing that of four commercially available adsorbent MOFs interesting for a wide range of potential applications,
materials (12−50 mg g−1) and that of nonwoven amidoxime- they also suffer from several significant drawbacks, limiting their
functionalized polyethylene fabric (0.75 mg g−1).191 Batch application as solid-phase adsorbent materials. Most notably,
experiments involving 5 gallons of seawater were performed the relatively weak and labile metal−ligand coordination bonds
using the ATRP-grafted mesoporous polymer. A uptake make chemical stability a routine challenge, and displacement
behavior of the mesoporous polymer sorbent, 0.33 mg of of organic bridging ligands from MOF templates has even been
uranium/g of adsorbent/day, was observed with a moderate harnessed as a way to prepare porous inorganic materials under
capacity of 1.99 mg of uranium/g of adsorbent. This behavior is mild conditions.226,227 While design of complex and intricate
similar to that of an poly(amidoxime-co-MA) sorbent produced organic struts can be easily proposed, the synthetic challenge of
by ORNL (0.30 mg of uranium/g of adsorbent/day). their preparation must not be neglected, and thus far the
AN/AA comonomers were also copolymerized with DVB on industrial-scale synthesis of MOFs has only been pursued with
a mesoporous carbon framework, initiated thermally with an simple, commercially available ligands. Moreover, incorporation
AIBN activator.225 The ratio of AN:AA:DVB was maintained at of orthogonal functionalities capable of strongly chelating
7:2:1 (w/w) while the carbon:monomer ratios were varied metals is extremely difficult, due to the solvothermal synthetic
from 1:3 to 1:25. Screening of uranium uptake was performed routes commonly implemented to prepare MOFs. Direct
following conversion of the nitrile groups to amidoximes via synthesis will result in the occupation of the metal-chelating site
treatment with hydroxylamine. Approximately 15 mg of by the metal intended to form the MOF node (also referred to
adsorbents were suspended in 250 mL of a 700 ppb uranium as the Secondary Building Unit, or SBU). Besides preventing
brine solution for 24 h. The highest screening capacity of 63 mg the chelation of subsequent metals, this also reduces the
of uranium/g of adsorbent was achieved for a 1:12 carbon:- efficiency of the MOF preparation, if not prohibiting it outright.
monomer ratio, though performance by 1:15, 1:20, and 1:25 Finally, MOFs are not generally amenable to environmental
were similar. The material possessing a 1:12 carbon:monomer deployment due to small particle size. A method for
ratio was investigated for recovery of uranium from 5 gallons of synthesizing MOF monoliths has not been developed, requiring
filtered environmental seawater over 4 weeks via a batch some additional form of large-scale support for deployment in
contact, affording an uptake of 0.55 mg of uranium/g of seawater, which undermines the high surface areas enabled by
adsorbent. using mesoscale nanostructured materials.
Table 8 summarizes the performance of graft polymers Several publications have nevertheless reported the use of
generated on porous carbon supports by ATRP. MOFs extraction of uranyl from aqueous solutions, with
proposed applications either in separation of radionuclides or
6.3. Metal−Organic Frameworks (MOFs) seawater mining. Two reports submitted and published almost
Metal−organic frameworks (MOFs) are porous materials simultaneously constitute the initial foray into this burgeoning
formed by the reaction of organic ligand struts with metal new field of investigation: Lin et al. reported the use of a
cluster nodes to form infinite 1-, 2-, or 3-D networks. Often phosphate-functionalized UiO-68(Zr) MOF,53 while Wang et
displaying extremely high surface areas (with some surpassing al. articulated the use of unfunctionalized HKUST-1(Cu).228
13988 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Despite nearly identical publication dates, these two papers The Cr MOF MIL-101 was subsequently functionalized and
delineate fundamentally different approaches to designing investigated for uranyl extraction.230 In contrast to the
MOFs for uranium binding. While the former relies upon aforementioned work, PSM at several different sites afforded
inclusion of a rationally designed organic strut, providing a a series of four isoreticular MOFs which were investigated for
specific binding site, the latter utilizes a pristine MOF with adsorption of uranium. While MIL-101(Cr) (4) was prepared
nonspecific binding presumably occurring at the SBU. These by traditional routes, the analogue MIL-101-NH2 (4-NH2) was
two strategies are discussed below, as are the publications in prepared by exposing 4 to strongly oxidizing acidic conditions
which they are utilized. and subsequently reducing the in situ generated orthogonal
6.3.1. Orthogonally-Functionalized MOFs. Lin and nitro group to a primary amine. MIL-101-ED (4-ED) and MIL-
colleagues are generally credited with the first report of a 101-DETA (4-DETA) were prepared by the condensation of
MOF used for extraction of uranium from seawater.53 A Zr- ethylenediamine (ED) or diethylenetriamine (DETA) with
MOF, UiO-68, was synthesized containing an orthogonal vacant metal sites in the MOF SBU, following removal of
diethoxyphosphoryl-urea moiety which orients into the frame- coordinating solvent molecules under high temperatures in
work tetrahedra to minimize steric interactions (1). Depro- vacuo. Functionalization of 4-ED and 4-DETA is summarized
tection of the ethoxy groups to yield the phosphate analogue in Figure 78. Unlike direct synthesis of a MOF with pre-
was accomplished by treating with Me3SiBr and subsequent functionalized ligands, as per Lin et al., PSM typically cannot
hydrolysis in water (2). While NMR analysis of a decomposed completely functionalize all sites in the MOF precursor. While
MOF revealed only ∼66% deprotection was achievable, this the PSM of the aforementioned materials was confirmed by
approach was necessary due to the synthetic limitations of the FTIR, TGA, and CHN analysis, 4-NH2, 4-ED, and 4-DETA
MOF. Direct synthesis of Zr- MOFs with phosphate-function- were only functionalized at 18%, 14%, and 8% of all possible
alized ligands is untenable, due to the extremely strong binding sites, respectively.
of Zr with phosphates. The unfunctionalized UiO-68-NH2 (3) MOFs 4, 4-NH2, 4-ED, and 4-DETA were investigated for
was also prepared and investigated as a control. A scheme uranium sorption in DI water over the pH range of 3−6.
depicting the preparation and structure of MOFs 1−3 is Uptake was negligible for all materials at pH < 4, but rose
displayed as Figure 76. dramatically up to pH 6, while kinetic studies performed at pH
5.5 revealed rapid saturation of all materials occurred in less
than 1 h of contact time. Sorption isotherms, again at pH 5.5 in
DI water, afforded saturation capacities of 20, 90, 200, and 350
mg of uranium/g of ads for 4, 4-NH2, 4-ED, and 4-DETA,
respectively. These results are displayed in Figure 79. The
authors noted the trend in capacity was inversely correlated
with coverage of the functional groups. Though correlation
breaks down for 4, which was unfunctionalized, the implication
is that sorption is related to the accessibility of the binding sites
located in the center of the MOF, in conjunction with the
enhanced Lewis basicity of an aliphatic amine (as on 4-ED and
4-DETA) as opposed to an aromatic amine (4-NH2). The
flexibility of the binding moiety on the SBU-functionalized 4-
ED and 4-DETA was also proposed to enhance uptake.
Figure 76. (Top) Scheme depicting the synthesis of 1−3. (Bottom)
Orthogonal functionalities incorporated in (left) 1 and (right) 2. Recycling investigations were performed which confirmed
Figure reprinted with permission from ref 53. Copyright 2013, Royal the stability of the adsorbent in aqueous conditions, as well as
Society of Chemistry. the potential for reuse after elution of bound uranium. PXRD
analysis reveals no change in peak position between the pristine
All three MOFs were investigated for the extraction of and U-loaded MOFs, while SEM images also attest to the
uranium from several aqueous solutions. Measurements were stability of the MOF adsorbent. Despite no apparent physical
made at pH 2.5 and 5 in DI water, with maximum saturation changes, MOFs 4-ED and 4-DETA showed a decrease of
capacities of 217 and 109 mg of uranium/g of ads obtained at approximately 30% following elution. However, 4-NH2 did not
pH 2.5 for 1 and 2, respectively. When the pH was elevated to reveal any appreciable loss of uptake. While PXRD spectra of
5, saturation capacities decreased to 152 and 104 mg of MOFs postelution were not provided, this process was
uranium/g of ads accordingly. No uptake was observed for 3 performed by decreasing the pH and MIL-101(Cr) is known
under any conditions. MOFs 1 and 2 were also tested in a for acid stability; MOF degradation from elution is not likely.
seawater simulant containing various sodium salts, carbonate, Selectivity studies on a series of metals at pH 5.5 were also
and both Mg and Ca cations.229 For both materials, a reduced performed, with high Kd values achieved for uranium (>6000
uptake was observed, with 188 and 32 mg of uranium/g of ads mL g−1) in contrast to relatively low values for competing ions
obtained at pH 2.5 for 1 and 2. Isotherms are displayed in (<100 mL g−1) when all are present at an 0.5 mM initial
Figure 77. Subsequent DFT calculations suggested the uranium concentration (Figure 77). While these results are interesting,
was bound cooperatively by two adjacent ligands through the selection of metals contains six lanthanides which are not
monodentate coordination with the phosphoryl oxygen. This present in seawater at appreciable concentrations and three
DFT-proposed 2:1 ligand to metal ratio (L:M) was supported transition metals, none of which are known to strongly compete
by investigation of the uranyl uptake, as experimental L:M of for uranyl-binding sites in previously reported adsorbents. Most
2:1 and 2.3:1 were obtained for MOF 1 in DI water and conspicuously absent are V, Fe, and Cu from the list of
seawater simulant, respectively. competing metals.
13989 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 77. (Top left) Sorption isotherms obtained for 1 and 2 at pH 2.5 in water (squares) and seawater simulant (triangles). Lines are a guide for
the eye. (Top right) Isotherms at pH 2.5 fitted with the Langmuir model. (Bottom) Scheme depicting the cooperative binding and 2:1 L:U ratio
proposed for 1 and 2. Figures reprinted with permission from ref 53. Copyright 2013, Royal Society of Chemistry.

Figure 78. Scheme depicting the functionalization of 4-ED. Coordinated solvent is removed thermally and replaced by a coordinating amine. Purple
circles represent MOF SBUs, red pipes represent the organic struts, and blue lines represent ethylenediamine (ED). An identical process is
performed for the preparation of 4-DETA. Figure reprinted with permission from ref 231. Copyright 2015, Nature Publishing Group.

The EXAFS spectra of 4-NH2, 4-ED, and 4-DETA were performed on the first uranyl coordination sphere, they are
fitted and compared with the spectra for uranyl hydroxide. Both lacking in rigor. This is exacerbated by an inappropriate
qualitative and quantitative similarities were observed for the selection of Rbkg during the data processing, which would
spectra of 4-ED and 4-DETA, indicative of a common uranyl diminish quadrature from the tightly bound axial oxygen,
bond motif to both materials. In contrast, for 4-NH2, the distorting the spectrum. Furthermore, EXAFS analysis is unable
feature corresponding to the scattering elements in the uranyl to discriminate between elements which are ±1 in Z-number,
equatorial plane was enhanced and attributed to steric making O and N indistinguishable and the definite assignment
hindrance from the aromatic rings in the framework. The of these scattering paths somewhat doubtful as a result. It is
bond lengths for U−N scattering decreased, however, which is possible the chemical identities of these elements could be
neither consistent with a sterically hindered complex nor a less inferred by fitting the second shell of the uranyl spectra;
active aromatic amine. Moreover, as the fits were only however, this was not completed for this work.
13990 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 79. (Top left) saturation capacities for 4, 4-NH2, 4-ED, and 4-DETA at pH 5.5. (Top right) Selectivity of 4-DETA at pH 5.5 and all metal
ions at 0.5 mM. (Bottom left) The effect of pH on uranyl uptake for 4-ED. (Bottom right) The effect of ED fixation on uranyl uptake for 4-ED.
Samples A, B, C, and D have ED:Cr of 0.34, 0.68, 0.98, and 1.42, respectively. Figures reprinted with permission from refs 230 and 231. Copyright
2015, Royal Society of Chemistry and Nature Publishing Group.

Figure 80. (Left) Structure of MOF 5 displaying the pendant carboxylic acid and acylamide units present in the 1-D hexagonal channel. (Right)
Chemical structures for N4,N4′-di(pyridine-4-yl)biphenyl-4,4′-dicarboxamide (L) and 1,3,5-benzene tricarboxylic acid (H3BTC). Figure reprinted
with permission from ref 233. Copyright 2012, Royal Society of Chemistry.

Similar work was reported shortly thereafter by Zhang et al., increasing quantities of ED was bound to the MOF SBU. The
again using MIL-101(Cr) and functionalizing at the metal SBU authors attribute this phenomenon to gradual occupation of
with ED.231 While much of this report duplicates previous vacant coordination sites around Cr by ED. However, they also
work,230 the authors do perform sorption experiments up to pH note the pre-edge feature is unchanged, regardless of ED
9, where a dramatic decrease in uptake is observed (Figure 79). functionalization. Pre-edge features for first-row transition
The pH of the sorption solution was adjusted with small metals are attributable to transitions of electrons from the 1s
amounts of HNO3 or Na2CO3, which suggests that at higher orbital to the 3d orbital. This symmetry-forbidden transition
pH there would be an appreciable effect from formation of can only occur when 3d and 4p orbitals are hybridized, which is
[UO2(CO3)3]4−. A capacity of approximately 200 mg of present only for a tetrahedral coordination environment.
uranium/g of ads was obtained for 4-ED at pH 5, essentially Occupancy of vacant coordination sites should result in
identical to previously reported data.230 The capacity at pH 8 increasing symmetry, reducing the observed pre-edge feature
decreased to approximately 150 mg of uranium/g of ads, while as an octahedral coordination environment is approached. The
at pH 9 the capacity was around 80 mg of uranium/g of ads. EXAFS region was fitted for the uranium LIII-edge spectra, but
Interestingly, negative uptake was reported at pH 2. again the second coordination shell was ignored. Dramatically
XAFS spectroscopy was performed at both the Cr K-edge different coordination environments are reported compared to
and uranium LIII-edge. X-ray absorption near edge spectra the identical system discussed previously,230 with the number of
(XANES) analysis revealed broadening of the white line as scatterers in the equatorial plane fixed at 6 and all scatterers
13991 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

assigned as oxygen. As expected, a large mean squared relative min.232 However, this experimental setup does not provide
displacement (σ2) value was obtained to accommodate this meaningful information, as the experiment was performed at a
model. While good R-factors were reported in the publication, concentration where the adsorbent was not saturated. Similar
obvious mis-fit occurs in this second coordination sphere. kinetics would be observed for most materials where absorption
The most recent MOF system reported for the extraction of kinetics are performed below saturation concentrations.
uranium contains Zn atoms as SBUs with a mixture of two Moreover, the instantaneous completion of adsorption casts
bridging ligands: 1,3,5-benzene tricarboxylic acid (H3BTC), doubt on the analytical rigor. The only data point which was
and N 4 ,N 4 ′ -di(pyridine-4-yl)biphenyl-4,4′-dicarboxamide not 90 ± 5 mg of uranium/g of ads was the data point at t = 0.
(L2).232 As reported in previous literature,233 this MOF (5) Moreover, adsorbents were collected by centrifugation for 10
crystallizes in a chiral hexagonal system composing a racemic min, which makes the measurements at t < 10 questionable.
bulk sample. The Zn atoms are bound in a tetrahedral Subsequent adsorption isotherms reveal a saturation capacity of
geometry by two carboxylate oxygen and two nitrogen from the 126 mg of uranium/g of ads at pH 2 in DI water, despite only
pyridyl groups on L2. These complexes form infinite 1-D chains having a contact time of 5 min. These results do support rapid
with hairpin folds traversing the c axis, creating 1-D hexagonal uranyl sorption.
channels with the inner wall functionalized by carboxylic acids Elution studies were performed with water, 0.1 M HCl, 0.1
and acylamide moieties, as shown in Figure 80.233 M HNO3, and 0.1 M Na2CO3, which showed >3%, 54%, 82%,
Uranium uptake was investigated in DI water at pH 1−5, and 92% uranium recovered from the adsorbent for each of the
with a maximum uptake of 92 mg of uranium/g of ads obtained eluents, respectively.232 However, the authors note the MOF
at pH 2 and an uptake of approximately 75 mg of uranium/g of was “digested” during the elution process, and no PXRD data
ads obtained at pH 5 (Figure 81).232 The authors attribute this were available after elution. The conspicuous absence of these
important data suggests this MOF would not be reusable and,
more importantly, would not be stable for long-term deploy-
ment in environmental seawater. Labile metal−ligand bonding
is well established in Zn MOFs, and decomposition is generally
to be expected.234,235
The authors did investigate the extraction of uranium from
environmentally relevant conditions, using a seawater simulant
reported elsewhere.229 The MOF adsorbent was added at a
phase ratio of 0.01 g L−1 to 1 L of artificial seawater containing
6 ppb uranium and at a pH of 7.8. The authors then claim
uptake of 0.53 mg of uranium/g of ads after 1 min of contact
time, using Na2CO3 as an eluent, which equals removal of more
than 88% of the uranyl in solution.232 While these are very
encouraging preliminary results, the absence of any specific
experimental details is a cause for concern. Previous adsorption
measurements were performed with only 10 mL of solution,
and adsorbents were collected by centrifugation. It is
unreasonable to centrifuge 1 L of solution following the
protocol delineated by the authors, and a contact time of 1 min
is not possible as discussed above. The authors claim elution of
the uranium and reuse of the adsorbent, achieving uptake of
0.51 mg of uranium/g and 0.47 mg of uranium/g for the
second and third trials. If the material is removed by filtration,
negative controls need to be performed and reported to ensure
the uranium is not extracted by the filter or bottle. Finally, the
authors indicate ICP-MS analysis of the seawater simulant
solution is used to determine uptake. However, ICP-MS
Figure 81. (Top) pH dependence of 5, with initial uranium analysis of dilute, high-matrix solutions such as seawater
concentration of 100 ppm. (Bottom) Sorption isotherm of 5 at pH
2 with 5 min contact time. Figure reprinted with permission from ref requires special considerations for analytical accuracy, as
232. Copyright 2015, Royal Society of Chemistry. discussed in section 3.6. Minimal details are provided in this
regard, and validation of these results is essential.
6.3.2. Unfunctionalized MOFs. In contrast to MOFs
decrease to formation of colloidal or oligiomeric species of which have been synthesized or postsynthetically functionalized
uranyl, such as (UO2)2(OH)22+ or (UO2)4(OH)7+, inhibiting with orthogonal uranyl-binding moieties, unfunctionalized
binding by the pendant functionalities. Though not specifically MOFs have been investigated for extraction of uranium as
mentioned, the pore apertures of 7.0−9.4 Å would not well. Though there is minimal commentary in the publications,
accommodate diffusion of colloidal uranium into the MOF the precise location of uranyl binding is presumably the metal-
interior as the axial diameter of the uranyl tricarbonate anion is oxo bonds in the SBU, as the unfunctionalized organic struts
approximately 6 Å while the equatorial diameter is approx- are devoid of any uranophilic characteristics.
imately 8−9 Å. Due to the optimal uptake obtained at pH 2, The first use of an unfunctionalized MOF for uranium
subsequent investigations were only performed at that pH. extraction was reported by Wang et al., utilizing the iconic Cu
Rapid absorption uptake was observed in a kinetics MOF HKUST-1 (6).228 The effect of pH was evaluated over
investigation, with complete uptake achieved in less than 1 the range of 2−10 in DI water, with capacities of 200 mg of
13992 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 82. (Top left) Sorption isotherm for 6 at pH 6 in DI water. (Bottom left) pH dependence of 6, with uranium concentration of 200 ppm and
2 h contact time. (Right) The structure of 6, HKUST-1. Probable binding sites are on the MOF SBUs, which adopt a paddle-wheel structure formed
by two linear Cu atoms, four carboxylic acids, and two solvent molecules. Figure reprinted with permission from ref 228. Copyright 2013, Elsevier.

Figure 83. (Top left) pH dependence of 7 with uranium concentration 140 ppm and 5 h contact time. (Bottom left) Adsorption kinetics for 7 at pH
3. (Top right) The structure of MOF-76, representative of 7 and 8. (Bottom right) Selectivity of 7 at pH 3. Figures reprinted with permission from
ref 236. Copyright 2013, Royal Society of Chemistry.

uranium/g of ads achieved until pH 8 (Figure 82). At pH 10, uranium/g of ads when fitted with the Langmuir model. As the
the adsorption decreased to approximately 140 mg of uranium/ Langmuir model assumes monolayer coverage and isolated
g of ads, attributed to formation of [UO2(CO3)2]2− and adsorption sites, the authors assert such an adsorption process
[UO2(CO3)3]4− from solvation of atmospheric CO2. Adsorp- must occur. The authors suggest the large number of carboxyl
tion isotherms were performed at pH 6 at 25, 35, and 45 °C, groups coordinating the Cu SBUs provide the uranyl
which yielded saturation capacities of 787, 826, and 840 mg of adsorption sites. The SBU adopts a paddle-wheel structure, as
13993 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 84. (Top left) Structure of UiO-66, representative of 9 and 10. (Bottom left) pH dependence of 9 and 10. (Top right) Sorption isotherm at
pH 5.5 for 9 and 10. (Bottom right) Sorption kinetics for 9 and 10. Figures reprinted with permission from ref 237. Copyright 2015, Springer.

is commonplace for Cu MOFs, where four carboxylates and environmentally relevant metals (V, Fe, and Cu), selectivity
two water molecules coordinate two Cu atoms. Consideration over other metals with appreciable seawater concentration was
of the solvent-free formula for 6, Cu3(BTC)2, affords a molar observed (e.g., Pb, Zn, Sr, and Ni).
ratio of 1.34−1.42 uranium per SBU at saturation. Synthesis of the isostructural MOF 8 was achieved using Tb
The subsequent generation of unfunctionalized MOFs for to form the SBUs.236 While uranium adsorption data were not
uranium extraction was reported by Shi et al., who used MOF- provided, quenching of the suspension luminescence was
76 with Y SBUs (7) for sorption and Tb SBUs (8) for use as observed upon introduction of uranyl cations. At a concen-
luminescent sensors.236 Sorption of uranium by 7 in DI water tration of 0.002 M, luminescence intensity was reduced by 50%
was measured as a function of pH over the range of 2.5 to 6, and almost eliminated completely at 0.1 M. The authors note
with the maximum observed around a pH of 3. Stability was this characteristic allows the MOF to be used as a sensor for
checked over the pH range of 2−3, with crystallinity observed detecting uranium in aqueous waste; however, limits of
by PXRD. While sharp peaks are observed above pH 2.5, the detection were not determined. Moreover, the lowest
authors note the onset of framework decomposition occurs concentration solution for which luminescence quenching was
around pH 2. Stability of 7 above pH 3 was assumed. reported is equivalent to 0.5 mg of uranium/mL of solution. As
Presumably investigations were not performed at higher pH 8 was added at 3 mg mL−1 to perform the luminescence
due either to the stability of the MOF or due to the relatively experiments, significant improvements in sensitivity are needed
small pore apertures (6.6 × 6.6 Å) preventing diffusion of for this MOF-based sensor to be a viable technology.
[UO2(CO3)3]4− or uranyl colloids throughout the MOF The most recent effort in this field was reported by Yuan et
material. al., who used UiO-66 (9) and the amine-modified UiO-66-NH2
Sorption kinetics were observed at a phase ratio of 0.4 g L−1 (10) to extract uranium from aqueous solutions.237 Sorption
and a pH of 3 (Figure 83).236 Saturation was observed at was first measured as a function of pH over the range of 2−6
approximately 300 min when all of the uranium was extracted (Figure 84). Low uptake was observed below pH 4, with
from solution. The authors rationalize from the changes in the maxima of 110 and 115 mg of uranium/g of ads achieved at pH
rate of uptake that the rate of bulk diffusion, surface adsorption, 5.5 for 9 and 10, respectively. The authors indicate at lower pH
and intraparticle diffusion all contribute to the observed the “active sites” on the sorbents are protonated and positively
uranium adsorption rate. The adsorption was also obtained charged, though what specifically constituted an active site was
for 7 at pH 3 with a phase ratio of 0.4 g L−1 and a contact time never defined. Moreover, differences in uptake between 9 and
of 5 h. A saturation capacity of 298 mg of uranium/g of ads was 10 were not statistically significant, leading the authors to
obtained by fitting the data with the Langmuir model. Finally, conclude introduction of amino functionalities do not enhance
extracting from a solution with competing ions assessed the uranium sorption significantly. Some form of interaction at the
selectivity of 7 for uranium. While still not containing the most MOF SBU is apparent. These results are of particular interest
13994 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

regarding the interpretation of earlier work presented by Lin et the phosphorylurea groups of 1 and 2 or carboxylic acid and
al.,53 as their adsorbents displayed high uptake at pH 2.5 and acrylamide groups of 5, or on advantageous binding to the SBU
significantly diminished capacity at pH 5. It is possible the as evidenced by 6−10. In both strategies, very impressive
sorption observed at pH 5 is attributable not to binding with uptakes have been achieved at low pH, but limited
the orthogonal phorphorylurea moieties of 1 and 2, but at the investigations under environmentally relevant conditions are a
SBU as evidenced by sorption by 9 and 10. fundamental shortcoming for most MOF systems reported. As
Uptake was investigated as a function of ionic strength, with speciation of uranyl is known to directly impact sorption
NaClO4 added at various concentrations from 10−4 to 10−1 M. results, these experiments must necessarily be performed. In
While it is not apparent why NaClO4 was used, rather than many cases it can be speculated these investigations were not
NaCl, uranium adsorption was not dramatically affected for pursued due to the chemical stability of the material, as high pH
either 9 or 10. Selectivity was also investigated for uranium over and carbonate concentrations are known to decompose the
several transition metals and lanthanides. While 9 and 10 had MOFs discussed above.238 Development of MOFs with
minimal affinity for the competing ions, as in previous work, the enhanced stabilities at high pH is essential for deployment
most environmentally relevant metals (V, Fe, and Cu) were not under environmentally relevant conditions.
investigated. A second area that requires further study pertains to
6.3.3. Summary and Perspectives. MOFs provide an selectivity for uranyl over other transition metals found in
interesting new platform for the extraction of uranium from seawater. While several works reported impressive selectivity,
aqueous solutions, highlighted by several publications reported these did not include the most pertinent competing metals. For
in a very short time frame. Table 9 summarizes the uptakes those displaying some promise, namely 5, validation of the
reported for the MOF systems discussed in this section. sorption results need to be performed, and a thorough
Preparation of a MOF-based adsorbent can rely on the investigation of the material stability is critical.
presence of orthogonal functionalities, as demonstrated by Structure-based analysis remains an undeveloped area of
research for MOFs. In light of their long-range ordering and
Table 9. Performance of MOFs for Recovery of Uranium recent development in high-performance computing, the ability
from Water to investigate the binding sites of MOFs has been investigated
Orthogonally Functionalized MOFs
extensively for gas storage and separations. Similar efforts could
UiO-68(Zr)-PO4Et2 217 100 ppm uranium, pH 2.5 53
be readily translated to predicting binding sites for extraction of
(1) uranium from aqueous solutions. Furthermore, crystallography
152 100 ppm uranium, pH 5 53 or XAFS afford reasonable approaches for investigating and
188b 100 ppm uranium, simulated 53 confirming the predicted binding site. High-throughput screen-
seawater, pH 2.5 ing of MOFs could enable the discovery of a highly selective
UiO-68(Zr)-PO4H2 109 pH 2.5 53 material for uranyl adsorption, as MOFs constitute one of very
(2)
few materials amenable to rationale tuning and modification.
104 pH 5 53
Recent work was reported where computational screening
32b 100 ppm uranium, simulated 53
seawater, pH 2.5 allowed identification of a protein capable of femtomolar
UiO-68(Zr) (3) 0 pH 2.5 53 affinity for uranyl.197 The selectivity of this binding site was
MIL-101(Cr) (4) 20c 230 entirely structural in nature, as only amino and carboxylic acid
MIL-101(Cr)-NH2 90c 230 functionalities were present in the binding pocket. Similarly,
(4-NH2) identification of a MOF possessing appropriate topology could
MIL-101(Cr)-ED 200c pH 5−5.5 230 be performed through computational screening. Subsequent
(4-ED)
incorporation of orthogonal hydroxyl- or amino- groups
150 100 ppm uranium, pH 8 230
possessing appropriate orientation could result in a truly
80 100 ppm uranium, pH 9 230,
231 biomimetic binding pocket, displaying exceptional uranium
MIL-101(Cr)-DETA 350c pH 5.5 230 selectivity and capacity.
(4-DETA) 6.4. Covalent Organic Frameworks (COFs)
Zn(L)(HBTC) (5) 125c 232
75 100 ppm uranium, pH 5 232 Covalent organic frameworks (COFs) are a class of recently
0.53b 6 ppb uranium, seawater 232 developed all organic analogues to MOFs, usually constructed
simulant, pH 7.8 through dynamic covalent chemistry.239,240 These porous,
Unfunctionalized MOFs crystalline organic materials consist of light elements (C, H,
HKUST-1(Cu) (6) >200 200 ppm uranium, pH 2.5−8 228 B, N, O, etc.) connected through strong, covalent bonds,
840c pH 6 228 making them significantly more chemically stable than MOFs
140 200 ppm uranium, pH 10 228 when deployed in harsh media. While COFs can be constructed
MOF-76(Y) (7) 298 140 ppm uranium, pH 3 236 through a “building-block” type approach, where organic nodes
MOF-76(Tb) (8) d 236 link linear struts, the diversity of structures is far more restricted
UiO-66(Zr) (9) 110 100 ppm uranium, pH 5.5 237 than MOFs, due primarily to limitations in the possible
UiO-66(Zr)-NH2 5.5 100 ppm uranium, pH 115 237 symmetry of the organic node and functional group
(10)
compatibility for successful synthesis. Nevertheless, the
a
Unless otherwise stated, adsorption experiments were performed in combination of the inherent porosity with a composition
DI water and pH was unadjusted. bSynthetic seawater composition: devoid of metals results in extremely high surface areas and low
0.41 M NaCl, 28.2 mM Na2SO4, 9.9 mM KCl, 2.3 mM NaHCO3, 53.3 material density. Moreover, due to strong material stability,
mM MgCl2·6H2O, 10.3 mM CaCl·2H2O. cMaximum capacity at postsynthetic chemical functionalization is readily achieved,
saturation. dUranium uptake performance not reported. while custom linkers can be designed and synthesized through
13995 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

well-established organic chemistry approaches. These proper- readily modified postsynthetically to afford diverse function-
ties of COFs have made them of significant interest for alities.243 The COF, referred to as TCD, was synthesized by
applications within separations, gas storage, and catalysis. microwave irradiation of trimesoyl chloride and 2,4-hexadiyne-
Heavy metal extraction is one area where interest is increasing, 1,6-diol, as displayed in Figure 87. Postsynthetic treatment with
including within seawater-based uranium extractions. malonitrile installs nitrile functionalities in the COF, TCD-CN,
Ma and colleagues were the first, and perhaps only, group to which can be subsequently transformed into amidoxime groups
explore the utility of COFs for uranium extraction from via reaction with hydroxylamine, forming TCD-AO (Figure
seawater.241 The COF SCU1 was constructed from trimesoyl 88). The TCD alkynes could also be oxidized to afford the
chloride and p-phenylenediamine in a molar ratio of 2:3 (Figure hydroxyl functionalized COF TCD−OH.
85). After treatment with “segregated” microwave irradiation, TCD and its derivatives were investigated for the recovery of
uranium from aqueous solutions.243 Screening of adsorption as
a function of pH revealed no significant uptake until
approximately pH 3, with increasing performance up to pH
4.5 where adsorption experiments were halted. Rapid kinetics
were observed, with complete uptake achieved by TCD within
the first 20 min of contact, though this experiment was also
performed below concentrations where the adsorption sites of
the material would truly saturate. An adsorption isotherm of
TCD achieves a maximum uptake of approximately 140 mg of
uranium/g of adsorbent, again at pH 4.5 and devoid of
appreciable salinity or carbonate; corresponding values for the
Figure 85. Ligands used in the syntheses and functionalization of the functionalized TCD-CN, -AO, and -OH were not reported.
COFs discussed herein. Finally, batch adsorption experiments were performed for all
materials in the presence of the 11 competing ions mentioned
in their previous work, discussed above. Notable selectivity was
an ultramicroporous carbonaceous COF (CCOF-SCU1) was
observed for uranyl was observed, as displayed in Figure 89,
formed with a high density of nitrile functionalization, as
with the best uptake and selectivity afforded by TCD-OH,
determined by FTIR. Nitrogen isotherm analysis indicates the
followed closely by TCD-AO. TCD-CN and TCD also had
CCOF-SCU1 material has a large Brunauer−Emmett−Teller
significant uptake and selectivity. While these results are
(BET) surface area (507 m2/g) and is predominantly
promising, for application in seawater uranium recovery
microporous (d ≈ 0.4 nm) character. When tested as a
different screening conditions and competing ions need to be
uranium sorbent, CCOF-SCU1 was able to achieve 50 mg of
utilized to accurately replicate the pH, salinity, and concen-
uranium/g of adsorbent at pH 2.5 while in the presence of 11
tration of environmental seawater. Nevertheless, these pub-
other competing ions (La3+, Ce3+, Nd3+, Gd3+, Sm3+, Ba2+,
lications constitute the first foray into a field of interesting
Mn2+, Co2+, Sr2+, Ni2+, and Zn2+).241
materials possessing great potential.
The second generation of COFs investigated for uranium
Table 10 summarizes the uranium uptake performance for
extraction includes COF-COOH, which was generated with
COFs and porous-organic polymers, (POPs), discussed below.
trimesoyl chloride and p-phenylenediamine, the same two
building blocks as CCOF-SCU1, but at a different ratio of 4:3. 6.5. Porous Organic Polymers (POPs)
Under these conditions, the excess carboxylic acid functional POPs, a class of highly cross-linked, amorphous polymers
groups could be observed in the crystalline product. Further possessing intrinsic porosity, have recently emerged as a
functionalization with 2-(2,4-dihydroxyphenyl)-benzimidazole versatile alternative to the crystalline COFs for use in
(HBI) provided a route to a benzimidazole functional group membrane separations, catalysis, and gas storage.244−246 Most
grafted onto the surface (Figure 86). The reported HBI of the designed POPs possess micropores (d < 2 nm), resulting
functionalized COF (COF-HBI) also extracted uranium from a in the name polymers with intrinsic microporosity (PIM).244
batch solution containing 11 competing ions with a very high POPs that are composed of aromatic linkers are also referred to
capacity (211 g of U/g of ads) and fast kinetics (2 h), albeit at as porous aromatic frameworks, or PAFs. For the sake of
pH 4.5.242 simplicity and discussion, we will group all materials under the
The most recent development in COFs for uranium umbrella of POPs. Compared to MOFs or COFs, POPs are
adsorption involves the inclusion of conjugated carbon−carbon noncrystalline and have nonuniform pores that are typically ill-
triple bonds into the structure of the COF, which are thus defined. Therefore, the characteristics are often more difficult to

Figure 86. Description of synthesis of COF-COOH and subsequent postsynthetic functionalization with HBI to form COF-HBI. Figure reprinted
with permission from ref 242. Copyright 2015, Elsevier.

13996 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 87. Microwave synthesis to form the COF TCD. Figure reprinted with permission from ref 243. Copyright 2016, Royal Society of Chemistry.

Figure 88. Postsynthesis functionalization of TCD with malonitrile to afford TCD-CN and subsequent treatment with hydroxylamine to yield the
amidoximated COF TCD-AO. Figure reprinted with permission from ref 243. Copyright 2016, Royal Society of Chemistry.

understand, and the synthesis can be more difficult to control. polyethylene fabric (0.75 mg of uranium/g of ads) from Japan
In general, porous polymers are synthesized by introducing Atomic Energy Agency (JAEA).
ditopic or multitopic monomers into a monotopic monomer An amidoxime functionalized POP (PPN-6-PAO) was based
within a combined step-growth and chain-growth polymer- on the PPN-6 structure247 was later synthesized and function-
ization process to cross-link the propagating polymer chains, alized with poly(amidoxime) for recovery of uranium from
thus forming 3-D network materials. seawater.192 While closely related to the aforementioned
Dai and colleagues first reported a mesporous polymer work,191 PPN-6 is composed of biphenyl struts generated by
created without using a porogen for the recovery of uranium the coupling of tetrakis(4-iodophenyl)methane, rather than
DVB (Figure 91). The PPN-6 framework was postsynthetically
from seawater.191 The polymer was formed from VBC and
functionalized through chloromethylation to introduce chlorine
DVB comonomers, with the VBC chlorides functioning as
within the framework, affording efficient initiation sites for
initiators for the polymerization of AN by ATRP, as discussed postsynthetic ATRP of AN. Following polymerization, treat-
in section 5.2.4 and displayed in Figure 90. Following ment with hydroxylamine converts the nitriles of the grafted
conversion of poly(AN) to poly(amidoxime), the adsorbent polymer (PPN-6-PAN) to amidoximes (PPN-6-PAO). After
was screened in a spiked uranyl brine (6 ppm U), base conditioning with 3% KOH aqueous solution at room
demonstrating a high uptake of 80 mg of uranium/g of temperature, PPN-6-PAO was tested for uranium uptake in
adsorbent. Testing in actual seawater demonstrated a uranium- seawater spiked to 80 ppb with uranium. PPN-6-PAO
adsorption capacity of 1.99 mg of uranium/g of adsorbent, demonstrated a uranium uptake of 96 mg of uranium/g of
surpassing the capacity of the irradiation-grafted nonwoven adsorbent after 42 d of contact (Figure 92).
13997 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 90. Formation of a DVB-co-VBC POP support for poly-


(amidoxime). Figure reprinted with permission from ref 191.
Copyright 2013, John Wiley & Sons.

Figure 89. Performance of TCD COF and derivatives in the selective


recovery of uranium over 11 other competing ions in aqueous brine solution containing 25.6 g L−1 NaCl and 0.198 g L−1
solution. Experiment performed at pH 4.5, 25 °C, with 3 h contact NaHCO3 with 7 ppm uranium, a capacity of 40 mg of uranium
time. All metals were added at an initial concentration of 0.5 mmol g−1 was obtained. Subsequent EXAFS analysis supported
L−1. Figure reprinted with permission from ref 243. Copyright 2016,
binding of uranium by an average of 1.4 ± 0.3 amidoxime
Royal Society of Chemistry.
functionalities in an η2-motif, suggesting cooperativity between
adjacent amidoxime species may contribute to the observed
Ma and co-workers also very recently reported a function-
performance.
alized porous organic polymer for the recovery of uranium from
Finally, a PIM was synthesized from tetrafluoroterephthalo-
seawater, based on the PPN-6 structure.248 In contrast to
previous work where poly(amidoxime) is grown from halide nitrile and 5,5′-6,6′-tetrahydroxy-3,3,3′,3′-tetramethyl-1,1′-spi-
initiators in the POP, amidoxime groups decorated the robisindane.249 The nitrile groups were converted postsyntheti-
channels of the porous support directly, providing site cally to amidoxime by treatment with hydroxylamine. Though
accessibility and rapid transport of uranium to the binding negligible uptake for uranium was observed in synthetic
sites (Figure 93). An isotherm collected in aqueous uranium seawater at pH 8, improved performance was observed with
solutions revealed a saturation capacity of 304 mg of uranium/g decreasing pH, and an optimum uptake was reported around
of adsorbent, while kinetic experiments demonstrated >95% pH 6. The authors suggested bubbling CO2 through fluidized
equilibration after a mere 3 h contact. The material could be beds to decrease the ocean pH, thereby improving uranium
eluted with Na2CO3, affording negligible decrease in perform- uptake. The technical feasibility of this was demonstrated at the
ance over two recycles. A Kd of approximately 1 × 106 mL g−1 laboratory scale, revealing three recycles using the aforemen-
suggested very high affinity for uranium, which was supported tioned material, though industrially application of this material
by investigations in seawater simulant. When deployed in a with commensurate acidification of the oceans is unrealistic.

Table 10. Performance of COF and POPs for Recovery of Uranium from Water and Seawatera
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
CCOF-SCU1 50b 135 ppm uranyl, competing ions, pH 2.5 241
COF-HBI 211c pH 4.5 242
COF TCD 140c pH 4.5 243
51b pH 4.5 243
COF TCD-CN 55b pH 4.5 243
COF-TCD-AO 67b pH 4.5 243
COF-TCD-OH 72b pH 4.5 243
poly(VBC-co- DVB)-graf t- 80 6 ppm uranium, with NaCl and NaHCO3, 2 27 d batch, 18.9 L fixed 191
poly(AO) pH 8 volume
PPN-6-PAO 96 91 ppb uranium in seawater 247
PPN-6-AO 304b 248
40 7.9 ppm uranium, 25.6 g/L NaCl, 0.198 g/L 248
NaHCO3

a
Unless otherwise stated, adsorption experiments were performed in DI water and pH was unadjusted. bCompeting ion composition: 0.5 mM Zn2+,
Sr2+, Sm3+, Ni2+, Nd3+, Mn2+, La3+, Gd3+, Co2+, Ce3+, and Ba2+. cMaximum capacity at saturation. dSeawater simulant composition: 33.3 g of sea salts
dissolved in 1 L of distilled water.

13998 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 91. Synthetic route for the preparation of amidoxime functionalized POP (PPN-6-PAO). Figure reprinted with permission from ref 192.
Copyright 2015, American Chemical Society.

designed chelating ligands. However, their instability under


alkaline pH presents challenges for deployment under environ-
mental seawater conditions.
Fryxell and colleagues reported the first use of MS materials
for extraction of actinides with self-assembled monolayers on
mesoporous supports (SAMMS).250 The materials possessed a
surface area of approximately 1000 m2 g−1 had high binding site
density and a rigid open pore structure to facilitate diffusion
through the MS channels. SAMMS were functionalized with
glycinyl-urea, salicylamide, and acetamide-phosphonate and
then investigated for sorption of actinides at acidic pH.
Impressive sorption was observed using the glycinyl-urea-
Figure 92. Concentration changing of uranium of the spiked seawater functionalized SAMMS, with maximum Kd of 2.4 × 105, 1.6 ×
and uranium adsorption capacity for PPN-6-PAO based adsorbent. 105, and 9.1 × 104 mL g−1 for Am(III), U(VI), and Th(IV),
Figure reprinted with permission from ref 192. Copyright 2015, respectively. As importantly, equilibrium was reached in 1−2 h,
American Chemical Society. indicating rapid dissociation through the material and fast
binding.250
6.6. Porous Silica Subsequent publications reported sorption with MS materials
Mesoporous silicas (MS) present an interesting option as functionalized by phosphates, phosphonates, or phosphine
heterogeneous supports for extracting uranium from seawater. oxides,251−254 amines,253,255,256 carboxylic acids,257 or more
As their surface is terminated by silanol groups, facile surface exotic (and less intuitive) functionalities such as transition
modification can occur through condensation reactions with metals258 or dihyrdoimidazole.259 One recent publication
trialkoxysilane-terminated organic molecules. As MS materials functionalizes with amidoxime,110 while another publication
are commercially available and possess high surface areas, they does not actually functionalize the MS, deploying it
are a natural choice for screening experiments for newly immediately following calcination.260
13999 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 93. Illustration of the synthesis of a POP directly functionalized with pendant amidoxime groups on the framework struts. Steps include
chloromethylation (a), nitrile installation (b), and amidoximation with hydroxylamine (c). Figure reprinted with permission from ref 248. Copyright
2017, American Chemical Society.

Figure 94. Substrate scope investigated by Vivero-Escoto et al. for uranium extraction on a common MS support. Figure reprinted with permission
from ref 96. Copyright 2013, Elsevier.

Though in principle the same type of heterogeneous support extraction, as no indication of sorption by an unfunctionalized
was used in all of the aforementioned studies, the differences in support had previously been published, suggesting background
surface areas and pore structure between different MS materials sorption by silanol functionalities had not been considered.
is significant. Moreover, as different pH ranges, contact times, When sorption was performed in seawater simulant, back-
phase ratios, degrees of grafting, and uranyl concentrations ground sorption dramatically decreased, and a phosphoric acid
were used, direct comparisons are challenging if not impossible. attached to the MS support through a urea bridge was observed
In an effort to mitigate this difficulty, Vivero-Escoto et al. to have the highest saturation capacity of around 60 mg of
synthesized a library of organo-functionalized MS materials on uranium/g of adsorbent. The cyclic imide dioxime- function-
the commercially available silica MSU-H (Figure 94).96 alized material extracted approximately 47 mg of uranium/g of
Following identical grafting and pretreatment conditions, they adsorbent, while the amidoxime-functionalized material ex-
performed the first direct comparison of more than nine tracted approximately 25 mg of uranium/g of adsorbent.96
different functionalities including amidoxime, bis-amidoxime, Complementary information was provided by Santschi and
cyclic imide dioxime, carboxylic acid, phosphate diester, colleagues, who reported the performance of functionalized
phosphonate, phosphoric acid, and bis-phosphoric acid. When nanoporous adsorbents, including nanoporous silica (Figure
sorption was performed in deionized water, the unfunctional- 95).109 When investigated for recovery of uranium from
ized MSU-H outperformed all but one sorbent, indicating that Galveston Bay seawater, silica functionalized 3,4-hydroxypyr-
nonspecific binding to the MS support is significant.96 This idinone (HOPO) afforded a sorbent which achieved Kd values
result raises concern about other MS investigations for uranium more than one log unit superior to several commercial resins
14000 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

and particulate formats which can be implemented in standard


packed-adsorbent column systems. However, in many instances
the act of developing a monolith or supporting nanostructured
materials on a substrate eliminates many of the advantageous
properties that were afforded by the unsupported material.
Furthermore, particles below a certain diameter cannot be
directly integrated into column systems due to tight packing
and the generation of extremely high backpressures. Finally, in
light of the low concentration of uranium in seawater, active
pumping approaches are not viable due to sheer volume of
seawater that needs to be processed, combined with the
Figure 95. Surface functionalities grafted onto nanoporous silica. associated energetic costs.
(Left) sulfonic acid, (center) iminodiacetic acid, and (right) 3,4- Diphos-, HOPO-, and PropPhos-functionalized silica were
hydroxypyridinone. Figure reprinted with permission from ref 109. processed into a thin film membrane, composed of 50 wt %
Copyright 2012, American Chemical Society. adsorbent and 50 wt % Nafion polymeric binder. A DiPhos-
functionalized silica membrane adsorbent reached equilibrium
and nanostructured MnO2. However, when investigated for in 24 h of contact with seawater spiked to 34 ppb in uranium
uranium recovery from Columbia River water, the aforemen- and at pH 7.5; 1.1 mg of uranium/g of adsorbent were
tioned adsorbents were roughly equal. recovered. Similar results were claimed for HOPO and
An interesting hybrid material was reported, formed from PropPhos materials as well. When contacted with 200 mL of
amidoxime-functionalized magnetic mesoporous silica (MMS- unspiked seawater, 0.104 mg of uranium/g of adsorbent was
AO), synthesized through co-condensation of tetraethyl extracted over 2 h contact time. Based on their reported
orthosilicate and 2-cyanoethyltriethoxysilane on the surface of Na2CO3 elution data, which show complete elution of uranium,
silica-coated Fe3O4. Following transformation of the nitrile to and recycling experiments, which show no statistically
amidoxime, a saturation capacity of 277 mg of uranium/g of significant decrease in uranium uptake over four runs for
adsorbent was obtained for the material at pH 5, facilitated by both Diphos- and PropPhos-functionalized materials following
rapid magnetic separation. Performance was observed to be an acid regeneration step, dramatic advantages over traditional
strongly pH dependent, though a respectable uptake of 83 mg polymeric adsorbents are asserted. The authors propose a
of uranium/g of adsorbent was achieved at pH 8 following 24 h continuous-deployment system where the membrane-sup-
contact.110 ported silica adsorbents are immersed in seawater, eluted
Very recently, a series of organo-functionalized silicas were with Na2CO3, pre-conditioned with acid, and then reimmersed
surface grafted with various (di)-phosphonic acids, imino in seawater. A graphical depiction of this system is provided in
diacetic acid, and HOPO and then investigated for the recovery Figure 97. Extrapolating their 2 h adsorption data to a
of uranium from environmental seawater (Figure 96).261 continuous process, the authors calculate as much as 75 mg of
Adsorption isotherms reveal saturation capacities of the uranium/g of adsorbent could be achieved over a 60 d
phosphate functionalized silica adsorbents outperform all deployment, compared to 4 mg of uranium/g of adsorbent for
commercial adsorbents and activated carbon, with PropPhos- traditional polymer materials.
and HOPO-functionalized silica achieving capacities of The aforementioned work does comprise an important step
approximately 60 mg of uranium/g of adsorbent, while toward the practical utilization of mesoporous materials for the
Diphos-functionalized silica saturated at 128 mg of uranium/g recovery of uranium from seawater. Significant attention has
of adsorbent. Kd values determined for Diphos-, PropPhos-, and been devoted by many to the development of nanoscale
HOPO-grafted materials were in excess of 3.5 × 106 mL g−1, materials, with little to no consideration for how such materials
while AcPhos- and IDAA- functionalized silica afforded values could be effectively utilized in industrial-scale processes.
of 3.3 × 106 and 1.7 × 106 mL g−1, respectively. However, the calculations provided in the aforementioned
Perhaps more interestingly, in this report the authors also work do need to be tempered slightly by some additional
articulate the integration of the Diphos- and PropPhos considerations. Foremost, the elution and recycle data reported
functionalized silica with a Nafion binder for use in extraction by the authors were performed on unsupported adsorbents,
of uranium from seawater.261 As noted by the authors, making additional experiments necessary to demonstrate the
nanostructured materials can be processed into monolithic impressive performance indeed translates to a membrane-

Figure 96. Uranyl-binding species grafted to mesoporous silica for recovery of uranium from seawater. Figure reprinted with permission from ref
261. Copyright 2016, Royal Society of Chemistry.

14001 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

in affinity and a distribution coefficient of 7.4 × 10−15 M (est.


6.7 × 10−11 mL g−1) at pH 8.9. The super uranyl-binding
protein could be fixed sulfhydryl resin and displayed impressive
selectivity for uranyl over other competing ions present in
seawater, even when added at more than 106× excess. Cu2+ and
VO2+ were the only ions for which the protein had comparable
affinity, but the authors note even in light of their greater
concentration in seawater they would not be capable of
competing with uranyl binding. Indeed, 17%, 30%, and 90% of
uranyl was extracted from simulated seawater when contacted
for 30 min with 1, 10, and >6000 equiv of protein.197 This
remarkable affinity and rapid sorption reveals genetically
engineered proteins could offer a viable means for extracting
uranium from seawater. However, there are significant
challenges to the viable deployment of engineered proteins
such as cost, deployment motif, and selectivity toward uranium,
Figure 97. Concept diagram for continuous collection of uranium specifically related to vanadium and copper.
from seawater using mesoporous adsorbents supported in a thin-film In an effort to address some of these potential challenges,
membrane. Figure reprinted with permission from ref 261. Copyright very recent work articulates the integration of the SUP into a
2016, Royal Society of Chemistry.
hydrogel through thiol-maleimide click chemistry (Figure
99).262 The hydrophilicity of the SUP was improved through
supported system. Second, principles of scaling have not been genetically fusing termini with elastin-like peptides, and the
considered. In contrast to traditional polymer based adsorbents, resulting hydrogel was subsequently formed with 4-arm PEG-
where 50 mg (or more) are utilized for testing in seawater, only maleimide cross-linkers. Using a microfluidic device, droplets of
5−6 mg of supported polymer were investigated in this study, SUP hydrogels were processed and cured into microbeads 165
assuming complete recovery of uranium from a 200 mL sample. μm in radius. Proteins were demonstrated to remain active for
In light of the known instability of silica in alkaline solutions, uranium binding, and the uranium uptake in DI water was quite
the degradation of the material over prolonged seawater close to the theoretically determined value. However, in
exposure should also be considered. Finally, it is worth recalling solutions at higher uranium concentrations, a 1.5:1 ratio of
early seawater deployment efforts by the Japanese relied on a uranium:protein was obtained suggesting the presence of a
much less sophisticated system, which nevertheless was the secondary binding site. Using these SUP-hydrogel beads,
primary cost driver.17 Translation of the adsorbent to uranium can be enriched up to 0.0092 mg of uranium/g of
supported polymers was due largely to the economy of the adsorbent upon contacted with environmental seawater spiked
deployment system. Nevertheless, the above work261 certainly to 13.7 nM with uranyl.
comprises a forward thinking technological development with Further development is needed to determine whether such
potential to fundamentally change the way materials are biologically based adsorbents are financially plausible compared
developed for the recovery of uranium from seawater. to the more established polymer systems. While still early in
Table 11 summarizes uranium recovery performance by development, this work has nevertheless demonstrated a viable
mesoporous silica adsorbents. approach for translation of a basic research program197 to a
6.7. High Affinity Genetically-Engineered Proteins format which could be putatively deployed in bead-format on a
While not fitting cleanly into the category of nanostructured thin film membrane.261 Identification of nondestructive
sorbents, biochemical techniques have been utilized to engineer methods for uranium recovery and assessment of scale up
a protein with intriguing properties that display particular and deployment options merit further attention. Table 12
innovation and hold promise as potential “game-changing” summarizes uranium binding performance by genetically
technologies in the field of uranyl extraction. In contrast to engineered proteins and materials derived for their deployment.
polymer fibers or inorganic sorbent materials, an innovative
approach using genetically engineered proteins was pursued in 7. CONCLUDING SUMMARY AND PERSPECTIVES
collaboration between the Lai and He groups.197 First, the The extraction of uranium from seawater has experienced a
Protein Data Bank was screened computationally to identify renewed interest due to the anticipated growth of nuclear
protein structures containing pockets which could accommo- power as a low-carbon source for baseload energy production.
date the uranyl cation. Variations were created where residues This interest has undergone many transformations since its
lining the binding pocket were modified to contain amino acids genesis, spanning solvent extraction and inorganic materials to
with accessible Lewis bases for uranyl coordination in the polyolefin fiber-based adsorbents and now to areas which did
equatorial plane. Over 5000 possible structures were identified not exist when this unconventional resource was identified,
and then refined by stability, potential steric interactions, and such as nanostructured materials, MOFs, and genetically
accessibility of the binding site (Figure 98). Of 10 candidates engineered proteins. In less than five years, current state-of-
identified, 9 were expressed well and 4 displayed selective the-art adsorbents have pushed the thresholds of science to
binding for uranyl.197 more than quadruple the best adsorption capacity achieved
From these promising initial results, a super uranyl-binding over the past half-century.
protein (SUP) was designed by genetic modification of Amidoxime-functionalized polymers are the most technolog-
Methanobacterium thermoautotrophicum. While the wild-type ically mature adsorbent for this application, and have
protein had a modest distribution coefficient, strategic demonstrated impressive performance following protracted
mutations resulted in nearly a 6 order of magnitude increase contact with seawater in screening studies, as well as in long-
14002 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Table 11. Performance of Mesoporous Silica Adsorbents for the Recovery of Uranium from Water and Seawatera
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
b
MSU-H 43 0.453 M NaCl, 2.3 mmol NaHCO3, pH 8.3 96
12 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSA-I 6 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSA-II 22b 0.453 M NaCl, 2.3 mmol NaHCO3, pH 8.3 96
6 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSA-III 31b 0.453 M NaCl, 2.3 mmol NaHCO3, pH 8.3 96
5 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSCA-I 58b 0.453 M NaCl, 2.3 mmol NaHCO3, pH 8.3 96
5 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSGly 2 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSPh-I 3 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSPh-II 4 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
b
MSPh-III 67 0.453 M NaCl, 2.3 mmol NaHCO3, pH 8.3 96
12 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
MSPh-IV 33b 0.453 M NaCl, 2.3 mmol NaHCO3, pH 8.3 96
2 5 ppm uranium, 0.453 M NaCl, 2.3 mmol 96
NaHCO3, pH 8.3
sulfonic acid@SAMMS 0.008c 33 ppb uranium in seawater 109
iminodiacetic acid@ 0.008c 33 ppb uranium in seawater 109
SAMMS
HOPO@SAMMS 0.008c 33 ppb uranium in seawater 109
MMS-AO 277b pH 5 110
83 54 ppm uranium, 0.1 M NaClO4, pH 8 110
DiPhos (particle) >0.5 50 ppb uranium in seawater, L/S ratio = 0.03 4d 104 seawater/adsorbent 261
10 000 mL/g phase ratio
128b pH 6.6−7.0 261
DiPhos (membrane) 1.1 34 ppb uranium in seawater 0.1 2h 200 mL of seawater 261
IDAA >0.5 50 ppb uranium in seawater, L/S ratio = 0.01 4d 104 seawater/adsorbent 261
10 000 mL/g phase ratio
AcPhos 0.4 50 ppb uranium in seawater, L/S ratio = 261
10 000 mL/g
PropPhos 0.4 50 ppb uranium in seawater, L/S ratio = 261
10 000 mL/g
b
60 pH 6.6−7.0 261
3,4-HOPO >0.5 50 ppb uranium in seawater, L/S ratio = 261
10 000 mL/g
65b pH 6.6−7.0 261
a
Unless otherwise stated, adsorption experiments were performed in DI water and pH was unadjusted. bMaximum capacity at saturation. cCalculated
from Kd values; precise uranium uptake was not provided

term deployment at the pilot scale in the open ocean. Quite graft chain to be an effective avenue for increasing adsorbent
deservedly, RIGP prepared polymer adsorbents are the performance and is a representative example of how this
benchmark against which all other adsorbents are compared.159 approach could be leveraged in the future.52 Related work
The versatility in form of polymer support,48,161 comonomer performed on polymer microspheres was achieved via RAFT
selection,177 and established chemistry with regards to graft polymerization, affording precise control over both atomistic
chain polymerization and conversion establish this entire and mesoscale structure to effect improvements in uranium
category of materials as an essential technology for seawater uptake.158 There is great potential inherent in the translation of
uranium recovery. this technology to a more deployable format, in particular a
A great strength is the relative ease with which the polymer high-surface area polymer trunk.48,161
composition can be tuned by variation of comonomer species The development of ATRP approaches to adsorbent
and ratios, as well as the inherent potential for adsorbents preparation may also hold great promise. A PVC-co-PVC-
composed of block copolymers instead of merely random graf t-poly(amidoxime-co-tBA) adsorbent prepared by ATRP
copolymers. The work of Saito and colleagues revealed the currently has the highest uranium uptake of any material in the
installation of a hydrophilic block to the end of the amidoxime open literature, achieving 5.22 mg of uranium/g of adsorbent
14003 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Figure 98. Primary steps in the computational screening and subsequent design of super-uranyl-binding proteins. The desired coordination features
are identified (step 1), the protein database is surveyed and a protein selected for further study (step 2), possible mutations for the protein are
considered (step 3), potential uranyl binding sites are searched (step 4), and uranyl coordination geometries for the selective potential binding sites
are evaluated and scored (step 5). If no hits are identified, the computational algorithm returns to step 2. Figure reprinted with permission from ref
197. Copyright 2014, Nature Publishing Group.

Figure 99. Overview for preparation of SUP-hydrogel bead adsorbents. (A) The chemical structure of the PEG-maleimide. (B) Structure of the SUP,
postgenetic modification with elastin-like peptides (ELP). (C) Representative synthetic scheme and SEM image of the resulting material. The inset
of the image is a bead composed of 10 wt % hydrogel after room temperature curing. Figure reprinted with permission from ref 262. Copyright 2017,
American Chemical Society.

Table 12. Performance of High Affinity Genetically-Engineered Proteins and Their Related Materials for the Recovery of
Uranium from Water and Seawater
uranium adsorption from water seawater uranium adsorption
mg of U/g of ads notes mg of U/g of ads time notes ref
a,b
SUP 1 3 ppb uranium, with competing ions 197
SUP-hydrogel beads 100 pmol/bead 4.8 ppm uranium; mass of bead not provided 0.0092c 1h Batch contact; 0.8 L seawater 262
a
Assuming molar mass of SUP is approximately 43 kDa; does not include mass of resin. bSynthetic seawater composition: 0.41 M NaCl, 28.2 mM
Na2SO4, 9.9 mM KCl, 2.3 mM NaHCO3, 53.3 mM MgCl2·6H2O, 10.3 mM CaCl2·2H2O. cConcentration of uranium in environmental seawater is
ca. 13.7 nM. Publication states seawater was spiked “with” 13.7 nM uranyl; the actual concentration of uranium is not clear.

over 49 d of seawater contact.78 However, two significant uptake as compared to RIGP-prepared adsorbents (e.g.,
challenges require further attention. Specifically, ATRP 1390%78 vs 354%49), which is a considerable cost-driver in
adsorbents require higher degrees of grafting for noteworthy material production, and the absence of any current industrial-
14004 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

scale process successfully incorporating ATRP suggests major inorganic adsorbents or nanostructured materials have
innovations in industrial processes would be required to bring capabilities that either engender them with intrinsic selectivity
this technology to market. or could potentially afford such a level of separation, while ion-
Introduction of GMA-grafted adsorbents comprise another imprinted polymers are the only soft-material that have also
promising avenue.41 Computational work has demonstrated demonstrated distinct potential to these ends.149 Most notably,
great potential for the de novo design of metal chelating ligands the structural rigor of hard-matter materials could be used to
which would, at least theoretically, possess exquisite selectiv- engineer pore architectures or arrange adsorbent groups to
ity.203 However, inclusion of such receptors onto an adsorbent facilitate cooperative adsorbent interactions. Identifying how
polymer is not expected to be achieved through traditional pore curvature influences sorption capacity and selectivity for a
RIGP approaches due to challenges synthesizing stable ligands monolayer-functionalized material could provide insights to
with polymerizable groups, significant steric hindrance how to improve performance for other materials. Alternatively,
impeding access of any polymerizable group to a growing materials with appreciable ordering, like MOFs or COFs, could
graft chain, and issues with side-reactions and radiolysis of such be used to investigate the role of cooperative sorbent
rationally designed chelators. In contrast, inclusion of a interactions or in the design of 3-D host architectures. A
functionality capable of ring-opening the GMA epoxide is particular opportunity appears to lend itself to the application
comparatively simple would enable polymer functionalization of high-performance computing to the design of structurally
and would facilitate investigation of these rationally designed based adsorbents.198,199,203 These materials could conceivably
ligands in environmental seawater. achieve selectivity rivaling genetically engineered proteins,197
Despite the rapid progress in this field and promise while providing superior saturation capacity and demonstrating
evidenced by the aforementioned materials, significant amenability to industrial-scale processing. However, several of
difficulties nevertheless remain. For example, a major challenge the promising nanostructured materials, specifically MOFs and
faced by the majority of inorganic adsorbents and all nano- or mesoporous silica, suffer from an innate instability under basic
nanostructured materials is the issue of practical deployability. conditions, which must be addressed before further research
While it is easy to point out seawater investigations have not should be pursued.
been performed, this is certainly not for lack of desire but In contrast, polymer adsorbents afford little opportunity for
limitations in how unbound powders can be efficiently structural engineering, as most properties can only be imparted
deployed under environmental conditions. Fixation upon a by manipulation of the bulk material. Accordingly, detailed
monolithic support or polymer may seem a straightforward fundamental studies of polymer structure and characterization
route at first glance, but the mass of substrate and fraction of of current top-performers are necessary to determine the role of
surface area obstructed by this approach would have severely local chemistry and graft chain morphology in observed uptake.
deleterious effects on the final sorption capacity. This remains Promising initial research in this area involves the direct
an area that is largely unexplored,261 and until a reasonable functionalization of poly(AN) with various comonomers,154,155
method can be established whereby only modest mass is added but further work is necessary to achieve definitive conclusions
and little surface area impeded, development of new nano- and deconvolute the many factors influencing the observed
structured materials may amount to little more than a novelty behavior. For instance, use of a metallocene catalyst to generate
with no practical application. Rather than focusing on highly iso- or syndiotactic polymers or copolymers which can
developing new materials, the authors believe pursuit of then be attached to a polymer core in a “grafting-to” approach
deployment platforms is a more worthwhile investment of could enable rational investigation of polymer morphology and
effort and time. begin unraveling some of the unanswered questions. Alter-
Similarly, while investigations of unsupported hydrogels or natively, investigations of unbound oligomers of appropriate
adsorbents grafted from polymer beads may afford exciting composition and molecular weight could facilitate fundamental
preliminary results, the influence of the polymer trunk material understanding of the influence of morphology, comonomer,
has not been considered. The behavior of tethered polyelec- polyolefin support, and graft chain length on adsorbent
trolytes, such as polyolefin-supported poly(amidoxime-co-AA), performance. Without understanding of the current adsorbent
remains an active area of both experimental263−265 and technologies, further investigation of exotic comonomers and
theoretical research.266,267 It is known that, at high graft screening studies constitute little more than a shotgun
densities, polymer chains crowd each other and induce approach.
stretching away from the surface of the support, with further Finally, while tremendous work has been performed in order
complexity added from electrostatic repulsion and subsequent to obtain data for uranium uptake in environmental seawater,
charge screening. Nevertheless, such work is relegated to planar additional effort is necessary to validate promising results which
polyelectrolyte brushes266,268 and has not developed to are reported in the literature and to ensure a head-to-head
consider the myriad of supports which have been discussed comparison is achieved. A critical review of experimental
above. Translation of technology developed on unsupported conditions reveals many data sets are of very little relevance to
hydrogels or polymer beads to a deployable trunk polymer is the alleged application. Inappropriate pH, phase ratios, uranium
likely to result in new graft chain morphology and concentrations, and meaningless competing ions are but a few
commensurate changes in performance, with uptake further of the more obvious flaws which remove much, if not all, utility
diminished by inclusion of the additional mass from the from numerous literature reports. The establishment of facilities
polymer support and reduced surface area. where samples can be screened in environmental seawater is an
Perhaps the largest challenged faced by all materials is the important step in the right direction, and opportunities need to
lack of selectivity observed between seawater-prevalent be made available for universities to access these resources.
transition metals and uranium. Exchanging the uptake of Ni, More than six decades after the foundational studies of
Fe, or V for uranium would increase the uptake by 2, 4, and Project Oyster, the recovery of uranium from seawater remains
10×, respectively. It is worth considering that many of the a challenging and ambitious goal. Nevertheless, dramatic
14005 DOI: 10.1021/acs.chemrev.7b00355
Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

improvements have been achieved in the traditional tech- for batteries, ionic block copolymers, thermoplastic elastomers,
nologies, in the process affording novel conceptual approaches additive manufacturing, carbon fibers and composites, lignin-based
with potential for exceptional gains in selectivity and perform- carbon materials and polymers, polymer nanocomposites, flow battery,
ance as well as a greater fundamental understanding of polymer and hydrogen and microbial fuel cells, in addition to uranium polymer
behavior and advances in the rational design and preparation of fiber adsorbents from seawater.
functional soft matter. While not yet within our grasp, the Sheng Dai obtained his B.S. degree (1984) and M.S. degree (1986) in
successful recovery of uranium from seawater remains a Chemistry at Zhejiang University, Hangzhou, China and his Ph.D.
technologically viable alternative to terrestrial mining and (1990) in Chemistry at the University of Tennessee, Knoxville. He is
with continued effort can ensure effectively limitless uranium currently a corporate fellow and group leader at the Oak Ridge
availability for future generations. National Laboratory and a Professor of Chemistry at the University of
Tennessee, Knoxville. His current research interests include ionic
AUTHOR INFORMATION liquids, porous carbon and oxide materials, advanced materials and
Corresponding Author their applications for separation sciences and energy storage, as well as
*E-mail: [email protected]. catalysis by nanomaterials. He has published over 600 peer-reviewed
ORCID journal papers and holds 26 U.S. patents.

Carter W. Abney: 0000-0002-1809-9577 ACKNOWLEDGMENTS


Richard T. Mayes: 0000-0002-7457-3261
The authors would like to thank Dr. Stephen Kung for his
Tomonori Saito: 0000-0002-4536-7530 enthusiastic support of and dedication to the Seawater Uranium
Sheng Dai: 0000-0002-8046-3931 Recovery Campaign, as part of the DOE Office of Nuclear
Author Contributions Energy Fuel Resources Program. Work was supported by the
The manuscript was written through contributions of all U.S. Department of Energy, Office of Nuclear Energy. This
authors. manuscript has been authored by UT-Battelle, LLC under
Notes
Contract No. DE-AC05-00OR22725 with the U.S. Department
of Energy. The United States Government retains and the
The authors declare no competing financial interest. publisher, by accepting the article for publication, acknowledges
Biographies that the United States Government retains a nonexclusive, paid-
up, irrevocable, worldwide license to publish or reproduce the
Carter W. Abney is a Eugene P. Wigner Fellow at Oak Ridge National published form of this manuscript, or allow others to do so, for
Laboratory. He received dual B.S. degrees in chemistry and theoretical United States Government purposes. The Department of
mathematics from the University of Wisconsin in 2007, under the Energy will provide public access to these results of federally
supervision of Professor Clark Landis. He received a M.S. degree in sponsored research in accordance with the DOE Public Access
inorganic chemistry from the University of North Carolina in 2009, Plan (http://energy.gov/downloads/doe-public-access-plan).
under the guidance of Professor Wenbin Lin. After working as an
analytical research chemist at the Centers for Disease Control and REFERENCES
Prevention, he earned his Ph.D. in chemistry from the University of (1) Glöser, S.; Tercero Espinoza, L.; Gandenberger, C.; Faulstich, M.
Chicago in 2015, again advised by Professor Wenbin Lin. He has been Raw Material Criticality in the Context of Classical Risk Assessment.
a Eugene P. Wigner Fellow at Oak Ridge National Laboratory since Resour. Policy 2015, 44, 35−46.
2015. His research involves development of nanostructured materials (2) Uranium 2014: Resources, Production, and Demand. OECD
for separations and catalysis, understanding the relationship between Nuclear Energy Agency and the International Atomic Energy Agency,
structure and performance in ionic liquids and soft matter, and the NEA#7209, 2014; p 504.
application of neutron/X-ray scattering and spectroscopy techniques (3) Lindner, H.; Schneider, E. Review of Cost Estimates for Uranium
to unresolved energy-related science questions. Recovery from Seawater. Energy Econ. 2015, 49, 9−22.
(4) World Energy Outlook 2011; International Energy Agency: Paris,
Richard T. Mayes is a Staff Inorganic Chemist in the Chemical France, 2011; pp 39−45.
Sciences Division at Oak Ridge National Laboratory. He received his (5) Pimentel, D.; Cooperstein, S.; Randell, H.; Filiberto, D.;
B.S. and M.S. degrees from Tennessee Technological University Sorrentino, S.; Kaye, B.; Nicklin, C.; Yagi, J.; Brian, J.; O’Hern, J.;
working with Dr. Edward Lisic. He received his Ph.D. in chemistry et al. Ecology of Increasing Diseases: Population Growth and
from the University of Tennessee in 2009 under the guidance of Dr. Environmental Degradation. Human Ecol 2007, 35, 653−668.
Craig E. Barnes. In 2009, he accepted a position as an ORISE (6) Total Primary Energy Consumption (Quadrillion Btu); U.S.
Postdoctoral Researcher at Oak Ridge National Laboratory with Dr. Department of Energy, U.S. Energy Information Administration.
http://www.eia.gov/cfapps/ipdbproject/IEDIndex3.cfm, April 4,
Sheng Dai, transitioning to staff in 2011. His research involves the
2016.
development and application of soft-templated mesoporous carbon (7) International Energy Outlook 2016, with Projections to 2040; U.S.
materials, heavy metal extraction from complex fluids, and high Department of Energy, U.S. Energy Information Administration,
temperature molten salt chemistry. Office of Energy Analysis: Washington, DC, DOE/EIA-0484, 2016; pp
Tomonori Saito is a R&D staff at Chemical Sciences Division at Oak 1−6.
Ridge National Laboratory. He received his Ph.D. in Chemistry (8) Endrizzi, F.; Rao, L. Chemical Speciation of Uranium(VI) in
Marine Environments: Complexation of Calcium and Magnesium Ions
(Organic Polymers) from Virginia Tech in 2008. He did postdoctoral
with [(UO2)(CO3)3](4‑) and the Effect on the Extraction of Uranium
research at The Pennsylvania State University and Oak Ridge National from Seawater. Chem. - Eur. J. 2014, 20, 14499−14506.
Laboratory prior to becoming a staff member at Oak Ridge National (9) Kim, J.; Tsouris, C.; Mayes, R. T.; Oyola, Y.; Saito, T.; Janke, C.
Laboratory in 2012. He has utilized his expertise in polymer science J.; Dai, S.; Schneider, E.; Sachde, D. Recovery of Uranium from
(especially polymer synthesis) to tackle various energy-related research Seawater: A Review of Current Status and Future Research Needs. Sep.
projects including CO2 separation membranes, polymer electrolytes Sci. Technol. 2013, 48, 367−387.

14006 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

(10) Flicker Byers, M.; Schneider, E. Optimization of the Passive Institute of Technology: Cambridge, MA, 1982; MITNE-254, MIT-
Recovery of Uranium from Seawater. Ind. Eng. Chem. Res. 2016, 55, EL-82-037.
4351−4361. (32) Fuel Resources Uranium from Seawater Program, Program
(11) Schneider, E.; Sachde, D. The Cost of Recovering Uranium Review Document; DOE Office of Nuclear Energy: Oak Ridge, TN,
from Seawater by a Braided Polymer Adsorbent System. Sci. & Global 2013; ORNL/TM-2013/295.
Secur. 2013, 21, 134−163. (33) Seko, N.; Katakai, A.; Tamada, M.; Sugo, T.; Yoshii, F. Fine
(12) Kabay, N.; Egawa, H. Chelating Polymers for Recovery of Fibrous Amidoxime Adsorbent Synthesized by Grafting and Uranium
Uranium from Seawater. Sep. Sci. Technol. 1994, 29, 135−150. Adsorption−Elution Cyclic Test with Seawater. Sep. Sci. Technol. 2004,
(13) Seko, N.; Tamada, M.; Yoshii, F. Current Status of Adsorbent 39, 3753−3767.
for Metal Ions with Radiation Grafting and Crosslinking Techniques. (34) Seko, N.; Katakai, A.; Hasegawa, S.; Tamada, M.; Kasai, N.;
Nucl. Instrum. Methods Phys. Res., Sect. B 2005, 236, 21−29. Takeda, H.; Sugo, T.; Saito, K. Aquaculture of Uranium in Seawater by
(14) Gao, M.; Zhu, G.; Gao, C. A Review: Adsorption Materials for a Fabric-Adsorbent Submerged System. Nucl. Technol. 2003, 144,
the Removal and Recovery of Uranium from Aqueous Solutions. 274−278.
Energy & Environ. Focus 2014, 3, 219−226. (35) Seko, N.; Tamada, M.; Kasai, N.; Yoshii, F.; Simizu, T. Synthesis
(15) Campbell, M. H.; Frame, J. M.; Dudey, N. D.; Kiel, G. R.; and Evaluation of Long Braid Adsorbent for Recovery of Uranium
Mesec, V.; Woodfield, F. W.; Binney, S. E.; Jante, M. R.; Anderson, R. from Seawater. Proc. Civil Eng. Ocean 2004, 20, 611−616.
C.; Clark, G. T. Extraction of uranium from seawater: chemical process (36) Jang, B.-B.; Lee, K.; Kwon, W. J.; Suh, J. Binding of Uranyl Ion
and plant design feasibility study. GJBX-36(79), 1979; p 169. by 2,2′- Dihydroxyazobenzene Attached to a Partially Chloromethy-
(16) Rodman, M. R.; Gordon, L. I.; Chen, A. C. T.; Campbell, M. H.; lated Polystyrene. J. Polym. Sci., Part A: Polym. Chem. 1999, 37, 3169−
Binney, S. E. Extraction of Uranium from Seawater: Evaluation of 3177.
Uranium Resources and Plant Siting. GJBX-35(79), 1979. (37) Lee, K.; Jang, B.-B.; Kwon, W. J.; Suh, J. Binding of Uranyl Ion
(17) Tamada, M. Current Status of Technology for Collection of by 2,2′- Dihydroxyazobenzene Attached to Crosslinked Polystyrenes
Uranium from Seawater; World Scientific Publishing Co. Pte. Ltd.: Toh Covered with Highly Populated Quaternary Ammonium Cations. J.
Tuck Link, Singapore, 2010. Polym. Sci., Part A: Polym. Chem. 1999, 37, 4117−4125.
(18) Rao, L. Recent International R&D Activities in the Extraction of (38) Kwon, W. J.; Yoo, C. E.; Chang, W.; Noh, Y.-S.; Suh, J. Metal
Uranium from Seawater; Lawrence Berkeley National Laboratory: Sequestering by a Poly(ethylenimine)-Sephadex G-25 Conjugate
Berkeley, CA, 2011; LBNL-4034E. Containing 2,2′-Dihydroyazobenzene. Bull. Korean Chem. Soc. 2000,
(19) Kelmers, A. D. Status of Technology for the Recovery of 21, 393−400.
Uranium from Seawater. Sep. Sci. Technol. 1981, 16, 1019−1035. (39) Seko, N.; Katakai, A.; Tamada, M.; Sugo, T.; Yoshii, F. Fine
(20) Endrizzi, F.; Leggett, C. J.; Rao, L. Scientific Basis for Efficient Fibrous Amidoxime Adsorbent Synthesized by Grafting and Uranium
Extraction of Uranium from Seawater. I: Understanding the Chemical Adsorption−Elution Cyclic Test with Seawater. Sep. Sci. Technol. 2004,
Speciation of Uranium under Seawater Conditions. Ind. Eng. Chem. 39, 3753−3767.
Res. 2016, 55, 4249−4256. (40) Zhang, A.; Uchiyama, G.; Asakura, T. pH Effect on the Uranium
(21) Davies, R. V.; Kennedy, J.; McIlroy, R. W.; Spence, R.; Hill, K. Adsorption from Seawater by a Macroporous Fibrous Polymeric
M. Extraction of Uranium from Sea Water. Nature 1964, 203, 1110− material Containing Amidoxime Chelating Functional Group. React.
1115. Funct. Polym. 2005, 63, 143−153.
(22) The Extraction of Uranium from the Sea (Oyster Project). UK (41) Kavaklı, P. A.; Seko, N.; Tamada, M.; Güven, O. Adsorption
National Archives: Atomic Energy Research Establishment, Harwell Efficiency of a New Adsorbent Towards Uranium and Vanadium Ions
AB 6/1264, 1956−1961, AB 6/1264. at Low Concentrations. Sep. Sci. Technol. 2005, 39, 1631−1643.
(23) Streeton, R. J. W. Continuous Extraction of Uranium from Sea (42) Ç aykara, T.; Alaslan, Ş. Ş.; Iṅ am, R. Competitive Adsorption of
Water; UK National Archives: Atomic Energy Research Establishment, Uranyl Ions in the Presence of Pb(II) and Cd(II) Ions by
Harwell, AB 15/2806, 1953. Poly(glycidyl methacrylate) Microbeads Carrying Amidoxime Groups
(24) Egawa, H.; Harada, H. Recovery of Uranium from Sea Water by and Polarographic Determination. J. Appl. Polym. Sci. 2007, 104,
Using Chelating Resins Containing Amidoxime Groups. Nippon 4168−4172.
Kagaku Kaishi 1979, 1979, 958−959. (43) Seko, N.; Bang, L. T.; Tamada, M. Syntheses of Amine-Type
(25) Egawa, H.; Harada, H.; Nonaka, T. Preparation of Adsorption Adsorbents with Emulsion Graft Polymerization of Glycidyl
Resins for Uranium in Sea Water. Nippon Kagaku Kaishi 1980, 1980, Methacrylate. Nucl. Instrum. Methods Phys. Res., Sect. B 2007, 265,
1767−1772. 146−149.
(26) Egawa, H.; Harada, H.; Shuto, T. Recovery of Uranium from (44) Raju, C. S. K.; Srinivasan, S.; Subramanian, M. S. New Multi-
Sea Water by the Use of Chelating Resins Containing Amidoxime Dentate Ion-Selective AXAD-16-MOPPA Polymer for the Preconcen-
Groups. Nippon Kagaku Kaishi 1980, 1980, 1773−1776. tration and Sequential Separation of U(VI), Th(IV) from Rare Earth
(27) Schenk, H. J.; Astheimer, L.; Witte, E. G.; Schwochau, K. Matrix. Sep. Sci. Technol. 2005, 40, 2213−2230.
Development of Sorbers for the Recovery of Uranium from Seawater. (45) Chauhan, G. S.; Kumar, A. A Study in the Uranyl Ions Uptake
1. Assessment of Key Parameters and Screening Studies of Sorber on Acrylic Acid and Acrylamide Copolymeric Hydrogels. J. Appl.
Materials. Sep. Sci. Technol. 1982, 17, 1293−1308. Polym. Sci. 2008, 110, 3795−3803.
(28) Astheimer, L.; Schenk, H. J.; Witte, E. G.; Schwochau, K. (46) Das, S.; Pandey, A. K.; Athawale, A. A.; Manchanda, V. K.
Development of Sorbers for the Recovery of Uranium from Seawater. Exchanges of Uranium(VI) Species in Amidoxime-Functionalized
Part 2. The Accumulation of Uranium from Seawater by Resins Sorbents. J. Phys. Chem. B 2009, 113, 6328−6335.
Containing Amidoxime and Imidoxime Functional Groups. Sep. Sci. (47) Prasad, T. L.; Saxena, A. K.; Tewari, P. K.; Sathiyamoorthy, D.
Technol. 1983, 18, 307−339. An Engineering Scale Study on Radiation Grafting of Polymeric
(29) Witte, E. G.; Schwochau, K. S.; Henkel, G.; Krebs, B. Uranyl Adsorbents for Recovery of Heavy Metal Ions from Seawater. Nucl.
Complexes of Acetamidoxime And Benzamidoxime. Preparation, Eng. Technol. 2009, 41, 1101−1108.
Characterization, and Crystal Structure. Inorg. Chim. Acta 1984, 94, (48) Oyola, Y.; Janke, C. J.; Dai, S. Synthesis, Development, and
323−331. Testing of High-Surface- Area Polymer-Based Adsorbents for the
(30) Nitta, C. K.; Best, F. R.; Driscoll, M. J. Delayed Neutron Assay to Selective Recovery of Uranium from Seawater. Ind. Eng. Chem. Res.
Test Sorbers for Uranium-From-Seawater Applications; Massachusetts 2016, 55, 4149−4160.
Institute of Technology: Cambridge, MA, 1982; MIT-EL 82-008. (49) Das, S.; Oyola, Y.; Mayes, R. T.; Janke, C. J.; Kuo, L. J.; Gill, G.;
(31) Borzekowski, J.; Driscoll, M. J.; Best, F. R. Uranium From Wood, J. R.; Dai, S. Extracting Uranium from Seawater: Promising AF
Seawater Research: Final Progress Report, FY 1982; Massachusetts Series Adsorbents. Ind. Eng. Chem. Res. 2016, 55, 4110−4117.

14007 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

(50) Das, S.; Oyola, Y.; Mayes, R. T.; Janke, C. J.; Kuo, L. J.; Gill, G.; Resolved Laser-Induced Fluorescence Spectroscopy (TRLFS). Radio-
Wood, J. R.; Dai, S. Extracting Uranium from Seawater: Promising AI chim. Acta 1996, 74, 87−91.
Series Adsorbents. Ind. Eng. Chem. Res. 2016, 55, 4103−4109. (69) Kalmykov, S. N.; Choppin, G. R. Mixed Ca2+/UO22+/CO32‑
(51) Brown, S.; Chatterjee, S.; Li, M.; Yue, Y.; Tsouris, C.; Janke, C. Complex Formation at Different Ionic Strengths. Radiochim. Acta
J.; Saito, T.; Dai, S. Uranium Adsorbent Fibers Prepared by Atom- 2000, 88, 603−606.
Transfer Radical Polymerization from Chlorinated Polypropylene and (70) Bernhard, G.; Geipel, G.; Reich, T.; Brendler, V.; Amayri, S.;
Polyethylene Trunk Fibers. Ind. Eng. Chem. Res. 2016, 55, 4130−4138. Nitsche, H. Uranyl(VI) Carbonate Complex Formation: Validation of
(52) Saito, T.; Brown, S.; Chatterjee, S.; Kim, J.; Tsouris, C.; Mayes, the Ca2UO2(CO3)3(aq) species. Radiochim. Acta 2001, 89, 511−518.
R. T.; Kuo, L.-J.; Gill, G.; Oyola, Y.; Janke, C. J.; et al. Uranium (71) Brooks, S. C.; Fredrickson, J. K.; Carroll, S. L.; Kennedy, D. W.;
Recovery from Seawater: Development of Fiber Adsorbents Prepared Zachara, J. M.; Plymale, A. E.; Kelly, S. D.; Kemner, K. M.; Fendorf, S.
via Atom-Transfer Radical Polymerization. J. Mater. Chem. A 2014, 2, Inhibition of Bacterial U(VI) Reduction by Calcium. Environ. Sci.
14674−14681. Technol. 2003, 37, 1850−1858.
(53) Carboni, M.; Abney, C. W.; Liu, S.; Lin, W. Highly Porous and (72) Baes, C. F., Jr.; Mesmer, R. E. The Hydrolysis of Cations; John
Stable Metal-Organic Frameworks for Uranium Extraction. Chem. Sci. Wiley & Sons: New York, 1976.
2013, 4, 2396−2402. (73) Krestou, A.; Panias, D. Uranium (VI) Speciation Diagrams in
(54) Zhou, S.; Chen, B.; Li, Y.; Guo, J.; Cai, X.; Qin, Z.; Bai, J.; Na, P. the UO22+/CO32‑/H2O System at 25°C. Eur. J. Miner. Process. Environ.
Synthesis, Characterization, Thermodynamic and Kinetic Investiga- Prot. 2004, 4, 113−129.
tions on Uranium (VI) Adsorption Using Organic-Inorganic (74) Ladshaw, A. P.; Das, S.; Liao, W. P.; Yiacoumi, S.; Janke, C. J.;
Composites: Zirconyl-Molybdopyrophosphate- Tributyl Phosphate. Mayes, R. T.; Dai, S.; Tsouris, C. Experiments and Modeling of
Sci. China: Chem. 2013, 56, 1516−1524. Uranium Uptake by Amidoxime-Based Adsorbent in the Presence of
(55) Wang, F.; Li, H.; Liu, Q.; Li, Z.; Li, R.; Zhang, H.; Liu, L.; Other Ions in Simulated Seawater. Ind. Eng. Chem. Res. 2016, 55,
Emelchenko, G. A.; Wang, J. A Graphene Oxide/Amidoxime Hydrogel 4241−4248.
for Enhanced Uranium Capture. Sci. Rep. 2016, 6, 19367. (75) Kim, J.; Oyola, Y.; Tsouris, C.; Hexel, C. R.; Mayes, R. T.; Janke,
(56) Gao, Q.; Hu, J.; Li, R.; Xing, Z.; Xu, L.; Wang, M.; Guo, X.; Wu, C. J.; Dai, S. Characterization of Uranium Uptake Kinetics from
G. Radiation Synthesis of a New Amidoximated UHMWPE Fibrous Seawater in Batch and Flow-Through Experiments. Ind. Eng. Chem.
Adsorbent with High Adsorption Selectivity for Uranium Over Res. 2013, 52, 9433−9440.
Vanadium in Simulated Seawater. Radiat. Phys. Chem. 2016, 122, 1−8. (76) Wood, J. R.; Gill, G. A.; Kuo, L. J.; Strivens, J. E.; Choe, K. Y.
(57) Xie, S.; Liu, X.; Zhang, B.; Ma, H.; Ling, C.; Yu, M.; Li, L.; Li, J. Comparison of Analytical Methods for the Determination of Uranium
Electrospun Nanofibrous Adsorbents for Uranium Extraction from in Seawater Using Inductively Coupled Plasma Mass Spectrometry.
Seawater. J. Mater. Chem. A 2015, 3, 2552−2558. Ind. Eng. Chem. Res. 2016, 55, 4344−4350.
(58) Bruland, K. W.; Lohan, M. C. In Treatise on Geochemistry, 1st (77) Das, S.; Brown, S.; Mayes, R. T.; Janke, C. J.; Tsouris, C.; Kuo,
ed.; Elderfield, H., Ed.; Elsevier Science: Amsterdam, 2003; Vol. 6.
L. J.; Gill, G.; Dai, S. Novel Poly(Imide Dioxime) Sorbents:
(59) Kim, J.; Tsouris, C.; Oyola, Y.; Janke, C. J.; Mayes, R. T.; Dai, S.;
Development and Testing for Enhanced Extraction of Uranium from
Gill, G.; Kuo, L.-J.; Wood, J.; Choe, K.-Y.; et al. Uptake of Uranium
Natural Seawater. Chem. Eng. J. 2016, 298, 125−135.
from Seawater by Amidoxime-Based Polymeric Adsorbent: Field
(78) Brown, S.; Yue, Y.; Kuo, L.-J.; Mehio, N.; Li, M.; Gill, G.;
Experiments, Modeling, and Updated Economic Assessment. Ind. Eng.
Tsouris, C.; Mayes, R. T.; Saito, T.; Dai, S. Uranium Adsorbent Fibers
Chem. Res. 2014, 53, 6076−6083.
Prepared by Atom-Transfer Radical Polymerization (ATRP) from
(60) Sekiguchi, K.; Serizawa, K.; Konishi, S.; Saito, K.; Furusaki, S.;
Sugo, T. Uranium Uptake During Permeation of Seawater Through Poly(vinyl chloride)-co-chlorinated Poly(vinyl chloride) (PVC-co-
Amidoxime-Group-Immobilized Micropores. React. Polym. 1994, 23, CPVC) Fiber. Ind. Eng. Chem. Res. 2016, 55, 4139−4148.
141−145. (79) Park, J.; Gill, G. A.; Strivens, J. E.; Kuo, L.-J.; Jeters, R. T.; Avila,
(61) Gill, G. A.; Kuo, L.-J.; Janke, C. J.; Park, J.; Jeters, R.; Bonheyo, A.; Wood, J. R.; Schlafer, N. J.; Janke, C. J.; Miller, E. A.; et al. Effect of
G.; Pan, H.-B.; Wai, C. M.; Khangaonkar, T.; Bianucci, L.; et al. The Biofouling on the Performance of Amidoxime-Based Polymeric
Uranium from Seawater Program at PNNL: Overview of Marine Uranium Adsorbents. Ind. Eng. Chem. Res. 2016, 55, 4328−4338.
Testing, Adsorbent Characterization, Adsorbent Durability, Adsorbent (80) Howell, D.; Behrends, B. A Methodology fFor Evaluating
Toxicity, and Deployment Studies. Ind. Eng. Chem. Res. 2016, 55, Biocide Release Rate, Surface Roughness and Leach Layer Formation
4264−4277. in a TBT-Free, Self-Polishing Antifouling Coating. Biofouling 2006, 22,
(62) Ashtheimer, L.; Schenk, H. J.; Schwochau, K. U.S. Patent 303−315.
4298577 A, 1981. (81) Lejars, M.; Margaillan, A.; Bressy, C. Fouling Release Coatings:
(63) Tian, G.; Teat, S.; Zhang, Z.; Rao, L. Sequestering Uranium A Nontoxic Alternative to Biocidal Antifouling Coatings. Chem. Rev.
from Seawater: Binding Strength and Modes of Uranyl Complexes 2012, 112, 4347−4390.
with Glutarimidedioxime. Dalton. Trans. 2012, 41, 11579−11586. (82) Callow, M. E.; Fletcher, R. L. The Influence of Low Surface
(64) Lee, J.-Y.; Yun, J.-I. Formation of Ternary CaUO2(CO3)32‑ and Energy Materials on Bioadhesion  A Review. Int. Biodeterior.
Ca2UO2(CO3)3(aq) Complexes Under Neutral to Weakly Alkaline Biodegrad. 1994, 34, 333−348.
Conditions. Dalton. Trans. 2013, 42, 9862−9869. (83) Walt, D. R.; Smulow, J. B.; Turesky, S. S.; Hill, R. G. The Effect
(65) Doudou, S.; Arumugam, K.; Vaughan, D. J.; Livens, F. R.; of Gravity on Initial Microbial Adhesion. J. Colloid Interface Sci. 1985,
Burton, N. A. Investigation of Ligand Exchange Reactions in Aqueous 107, 334−336.
Uranyl Carbonate Complexes using Computational Approaches. Phys. (84) Chambers, L. D.; Stokes, K. R.; Walsh, F. C.; Wood, R. J. K.
Chem. Chem. Phys. 2011, 13, 11402−11411. Modern Approaches to Marine Antifouling Coatings. Surf. Coat.
(66) Dong, W.; Brooks, S. C. Determination of the Formation Technol. 2006, 201, 3642−3652.
Constants of Ternary Complexes of Uranyl and Carbonate with (85) de Messano, L. V. R.; Sathler, L.; Reznik, L. Y.; Coutinho, R.
Alkaline Earth Metals (Mg2+, Ca2+, Sr2+, and Ba2+) Using Anion The Effect of Biofouling on Localized Corrosion of the Stainless Steels
Exchange Method. Environ. Sci. Technol. 2006, 40, 4689−4695. N08904 and UNS S32760. Int. Biodeterior. Biodegrad. 2009, 63, 607−
(67) Kelly, S. D.; Kemner, K. M.; Brooks, S. C. X-ray Absorption 614.
Spectroscopy Identifies Calcium-Uranyl-Carbonate Complexes at (86) Hills, J. M.; Thomason, J. C. The Effect of Scales of Surface
Environmental Concentrations. Geochim. Cosmochim. Acta 2007, 71, Roughness on the Settlement of Barnacle (Semibalanus Balanoides)
821−834. Cyprids. Biofouling 1998, 12, 57−69.
(68) Bernhard, G.; Geipel, G.; Brendler, V.; Nitsche, H. Speciation of (87) Berntsson, K. M.; Jonsson, P. R.; Lejhall, M.; Gatenholm, P.
Uranium in Seepage Waters of a Mine Tailing Pile Studied by Time Analysis of Behavioural Rejection of Micro-Textured Surfaces and

14008 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Implications for Recruitment by the Barnacle Balanus Improvisus. J. Application in Removal of Uranium and Thorium from Water. Int. J.
Exp. Mar. Biol. Ecol. 2000, 251, 59−83. Environ. Sci. Technol. 2015, 12, 1899−1906.
(88) Callow, M. E.; Jennings, A. R.; Brennan, A. B.; Seegert, C. E.; (106) Yin, X.; Bai, J.; Fan, F.; Cheng, W.; Tian, W.; Wang, Y.; Qin, Z.
Gibson, A.; Wilson, L.; Feinberg, A.; Baney, R.; Callow, J. A. Amidoximed Silica for Uranium(VI) Sorption from Aqueous Solution.
Microtopographic Cues for Settlement of Zoospores of the Green J. Radioanal. Nucl. Chem. 2014, 303, 2135−2142.
Fouling Alga Enteromorpha. Biofouling 2002, 18, 229−236. (107) Gunathilake, C.; Gorka, J.; Dai, S.; Jaroniec, M. Amidoxime-
(89) Das, S.; Pandey, A. K.; Athawale, A. A.; Subramanian, M.; Modified Mesoporous Silica for Uranium Adsorption Under Seawater
Seshagiri, T. K.; Khanna, P. K.; Manchanda, V. K. Silver Nanoparticles Conditions. J. Mater. Chem. A 2015, 3, 11650−11659.
Embedded Polymer Sorbent for Preconcentration of Uranium from (108) Wang, Y.-L.; Zhu, L.; Guo, B.-L.; Chen, S.-W.; Wu, W.-S.
Bio-Aggressive Aqueous Media. J. Hazard. Mater. 2011, 186, 2051− Mesoporous Silica SBA-15 Functionalized with Phosphonate De-
2059. rivatives for Uranium Uptake. New J. Chem. 2014, 38, 3853−3861.
(90) White, P. A.; Kalff, J.; Rasmussen, J. B.; Gasol, J. M. The Effect (109) Johnson, B. E.; Santschi, P. H.; Chuang, C.-Y.; Otosaka, S.;
of Temperature and Algal Biomass on Bacterial Production and Addleman, R. S.; Douglas, M.; Rutledge, R. D.; Chouyyok, W.;
Specific Growth Rate in Freshwater and Marine Habitats. Microb. Ecol. Davidson, J. D.; Fryxell, G. E.; et al. Collection of Lanthanides and
1991, 21, 99−118. Actinides from Natural Waters with Conventional and Nanoporous
(91) Sun, X.; Xu, C.; Tian, G.; Rao, L. Complexation of Sorbents. Environ. Sci. Technol. 2012, 46, 11251−11258.
Glutarimidedioxime with Fe(III), Cu(II), Pb(II), and Ni(II), the (110) Zhao, Y.; Li, J.; Zhang, S.; Wang, X. Amidoxime-Functionalized
Competing Ions for the Sequestration of U(VI) from Seawater. Dalton. Magnetic Mesoporous Silica for Selective Sorption of U(VI). RSC Adv.
Trans. 2013, 42, 14621−14627. 2014, 4, 32710−32717.
(92) Liu, X.; Liu, H.; Ma, H.; Cao, C.; Yu, M.; Wang, Z.; Deng, B.; (111) Manos, M. J.; Kanatzidis, M. G. Layered Metal Sulfides
Wang, M.; Li, J. Adsorption of the Uranyl Ions on an Amidoxime- Capture Uranium from Seawater. J. Am. Chem. Soc. 2012, 134, 16441−
Based Polyethylene Nonwoven Fabric Prepared by Preirradiation- 16446.
Induced Emulsion Graft Polymerization. Ind. Eng. Chem. Res. 2012, 51, (112) Ma, S.; Huang, L.; Ma, L.; Shim, Y.; Islam, S. M.; Wang, P.;
15089−15095. Zhao, L.-D.; Wang, S.; Sun, G.; Yang, X.; et al. Efficient Uranium
(93) Kuo, L.-J.; Gill, G.; Tsouris, C.; Rao, L.; Pan, H.-B.; Wai, C. M.; Capture by Polysulfide/Layered Double Hydroxide Composites. J. Am.
Janke, C. J.; Strivens, J. E.; Wood, J. R.; Schlafer, N. J. et al. Chem. Soc. 2015, 137, 3670−3677.
Temperature Effects on the Adsorption Capacity of Uranium and (113) Li, R.; Che, R.; Liu, Q.; Su, S.; Li, Z.; Zhang, H.; Liu, J.; Liu, L.;
Wang, J. Hierarchically Structured Layered-Double-Hydroxides
Vanadium on Amidoxime-Based Polymeric Fiber Adsorbents in
Derived by ZIF-67 for Uranium Recovery from Simulated Seawater.
Natural Seawater. Submitted 2017.
J. Hazard. Mater. 2017, 338, 167−176.
(94) de Sousa, Á . S. F.; Ferreira, E. M. M.; Cassella, R. J.
(114) Riley, B. J.; Chun, J.; Um, W.; Lepry, W. C.; Matyas, J.; Olszta,
Development of an Integrated Flow Injection System for the Electro-
M. J.; Li, X.; Polychronopoulou, K.; Kanatzidis, M. G. Chalcogen-
Oxidative Leaching of Uranium from Geological Samples and its
Based Aerogels As Sorbents for Radionuclide Remediation. Environ.
Spectrophotometric Determination with Arsenazo III. Anal. Chim.
Sci. Technol. 2013, 47, 7540−7547.
Acta 2008, 620, 89−96. (115) Yu, J.; Bai, H.; Wang, J.; Li, Z.; Jiao, C.; Liu, Q.; Zhang, M.;
(95) Carboni, M.; Abney, C. W.; Taylor-Pashow, K. M. L.; Vivero-
Liu, L. Synthesis of Alumina Nanosheets via Supercritical Fluid
Escoto, J. L.; Lin, W. Uranium Sorption with Functionalized Technology with High Uranyl Adsorptive Capacity. New J. Chem.
Mesoporous Carbon Materials. Ind. Eng. Chem. Res. 2013, 52, 2013, 37, 366−372.
15187−15197. (116) Tan, L.; Liu, Q.; Jing, X.; Liu, J.; Song, D.; Hu, S.; Liu, L.;
(96) Vivero-Escoto, J. L.; Carboni, M.; Abney, C. W.; deKrafft, K. E.; Wang, J. Removal of Uranium(VI) Ions from Aqueous Solution by
Lin, W. Organo- Functionalized Mesoporous Silicas for Efficient Magnetic Cobalt Ferrite/Multiwalled Carbon Nanotubes Composites.
Uranium Extraction. Microporous Mesoporous Mater. 2013, 180, 22−31. Chem. Eng. J. 2015, 273, 307−315.
(97) Hazer, O.; Kartal, Ş. Use of Amidoximated Hydrogel for (117) Shao, D.; Wang, X.; Wang, X.; Hu, S.; Hayat, T.; Alsaedi, A.;
Removal And Recovery of U(VI) Ion from Water Samples. Talanta Li, J.; Wang, S.; Hu, J.; Wang, X. Zero Valent Iron/Poly(Amidoxime)
2010, 82, 1974−1979. Adsorbent for the Separation and Reduction of U(VI). RSC Adv. 2016,
(98) Khan, M. H.; Warwick, P.; Evans, N. Spectrophotometric 6, 52076−52081.
Determination of Uranium with Arsenazo-III in Perchloric Acid. (118) Eloy, F.; Lenaers, R. The Chemistry of Amidoximes and
Chemosphere 2006, 63, 1165−1169. Related Compounds. Chem. Rev. 1962, 62, 155−183.
(99) Sella, S.; Sturgeon, R. E.; Willie, S. N.; Campos, R. C. Flow (119) Katritzky, A. R.; Huang, L.; Chahar, M.; Sakhuja, R.; Hall, C.
Injection On-line Reductive Precipitation Preconcentration With D. The Chemistry of N- Hydroxyamidoximes, N-Aminoamidoximes,
Magnetic Collection for Electrothermal Atomic Absorption Spectrom- and Hydrazidines. Chem. Rev. 2012, 112, 1633−1649.
etry. J. Anal. At. Spectrom. 1997, 12, 1281−1285. (120) Hirotsu, T.; Katoh, S.; Sugasaka, K.; Seno, M.; Itagaki, T.
(100) Nakashima, S.; Sturgeon, R. E.; Willie, S. N.; Berman, S. S. Binding Ability of Acetamide Oxime with Proton, Copper(II), and
Determination of Trace Elements in Sea Water by Graphite-Furnace Dioxouranium(VI) in Aqueous Solutions. J. Chem. Soc., Dalton Trans.
Atomic Absorption Spectrometry After Preconcentration by Tetrahy- 1986, 1609−1611.
droborate Reductive Precipitation. Anal. Chim. Acta 1988, 207, 291− (121) Park, Y.-Y.; Kim, S.-Y.; Kim, J.-S.; Harada, M.; Tomiyasu, H.;
299. Nogami, M.; Ikeda, Y. Complex Formation of U(VI) with
(101) Skogerboe, R. K.; Hanagan, W. A.; Taylor, H. E. Concentration Benzamidoxime in Non-aqueous Solvents. J. Nucl. Sci. Technol. 2000,
of Trace Elements in Water Samples by Reductive Precipitation. Anal. 37, 344−348.
Chem. 1985, 57, 2815−2818. (122) Das, S.; Pandey, A. K.; Athawale, A.; Kumar, V.; Bhardwaj, Y.
(102) Keen, N. J. Studies on Extraction of Uranium from Sea Water. K.; Sabharwal, S.; Manchanda, V. K. Chemical Aspects of Uranium
J. Br. Nucl. Energy Soc. 1968, 7, 178−183. Recovery from Seawater by Amidoximated Electron-Beam-Grafted
(103) Kanno, M. Present Status of Study on Extraction of Uranium Polypropylene Membranes. Desalination 2008, 232, 243−253.
from Sea-Water. J. Nucl. Sci. Technol. 1984, 21, 1−9. (123) Zhang, A.; Asakura, T.; Uchiyama, G. The Adsorption
(104) Hori, T.; Yamawaki, M.; Kanno, M. Uranium Adsorption Mechanism of Uranium(VI) from Seawater on a Macroporous Fibrous
Properties of Hydrous Titanium-Oxides in Seawater. J. Nucl. Sci. Polymeric Adsorbent Containing Amidoxime Chelating Functional
Technol. 1987, 24, 377−384. Group. React. Funct. Polym. 2003, 57, 67−76.
(105) Basu, H.; Singhal, R. K.; Pimple, M. V.; Reddy, A. V. R. (124) Zhang, A.; Uchiyama, G.; Asakura, T. The Adsorption
Synthesis and Characterization of Silica Microsphere and their Properties and Kinetics of Uranium(VI) with a Novel Fibrous and

14009 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Polymeric Adsorbent Containing Amidoxime Chelating Functional (145) Kabay, N. Preparation of Amidoxime-Fiber Adsorbents Based
Group from Seawater. Sep. Sci. Technol. 2003, 38, 1829−1849. on Poly(Methacrylonitrile) for Recovery of Uranium from Seawater.
(125) Katragadda, S.; Gesser, H. D.; Chow, A. The Extraction of Sep. Sci. Technol. 1994, 29, 375−384.
Uranium by Amidoximated Orlon. Talanta 1997, 45, 257−263. (146) Pal, S.; Ramachandhran, V.; Prabhakar, S.; Tewari, P. K.;
(126) Badawy, S. M.; Sokker, H. H.; Dessouki, A. M. Chelating Sudersanan, M. Polyhydroxamic Acid Sorbents for Uranium Recovery.
Polymer Granules Prepared by Radiation-Induced Homopolymeriza- J. Macromol. Sci., Part A: Pure Appl.Chem. 2006, 43, 735−747.
tion. II. Characterizations. J. Appl. Polym. Sci. 2006, 99, 1180−1187. (147) Rivas, B. L.; Perič, I. M.; Villegas, S.; Ruf, B. Binding of Uranyl
(127) Vukovic, S.; Watson, L. A.; Kang, S. O.; Custelcean, R.; Hay, B. Ions by Water- Insoluble Polymers Containing Multiligand Groups. J.
P. How Amidoximate Binds the Uranyl Cation. Inorg. Chem. 2012, 51, Chil. Chem. Soc. 2008, 53, 1356−1359.
3855−3859. (148) Kavakli, P. A.; Seko, N.; Tamada, M.; Güven, O. A Highly
(128) Abney, C. W.; Mayes, R. T.; Piechowicz, M.; Lin, Z.; Bryantsev, Efficient Chelating Polymer for the Adsorption of Uranyl and Vanadyl
V.; Veith, G. M.; Dai, S.; Lin, W. XAFS Investigation of Ions at Low Concentrations. Adsorption 2005, 10, 309−315.
Polyamidoxime-Bound Uranyl Contests the Paradigm from Small (149) Metilda, P.; Gladis, J. M.; Venkateswaran, G.; Prasada Rao, T.
Molecule Studies. Energy Environ. Sci. 2016, 9, 448−453. Investigation of the Role of Chelating Ligand in the Synthesis of Ion-
(129) Barber, P. S.; Kelley, S. P.; Rogers, R. D. Highly Selective Imprinted Polymeric Resins on the Selective Enrichment Of
Extraction of the Uranyl Ion with Hydrophobic Amidoxime-Function- Uranium(VI). Anal. Chim. Acta 2007, 587, 263−271.
alized Ionic Liquids via η2 Coordination. RSC Adv. 2012, 2, 8526− (150) Zhang, H.; Liang, H.; Chen, Q.; Shen, X. Synthesis of a New
8530. Ionic Imprinted Polymer for the Extraction of Uranium from Seawater.
(130) Kelley, S. P.; Barber, P. S.; Mullins, P. H. K.; Rogers, R. D. J. Radioanal. Nucl. Chem. 2013, 298, 1705−1712.
Structural Clues to UO22+/VO2+ Competition in Seawater Extraction (151) Shamsipur, M.; Fasihi, J.; Ashtari, K. Grafting of Ion-Imprinted
using Amidoxime-Based Extractants. Chem. Commun. 2014, 50, Polymers on the Surface of Silica Gel Particles through Covalently
12504−12507. Surface-Bound Initiators: A Selective Sorbent for Uranyl Ion. Anal.
(131) Tsantis, S. T.; Zagoraiou, E.; Savvidou, A.; Raptopoulou, C. P.; Chem. 2007, 79, 7116−7123.
Psycharis, V.; Szyrwiel, L.; Holynska, M.; Perlepes, S. P. Binding of (152) Noh, Y.-S.; Lee, K.; Suh, J. Uranyl Ion Complexation of 2,2′-
Oxime Group to Uranyl Ion. Dalton. Trans. 2016, 45, 9307−9319. Dihydroxyazobenzene Enhanced on a Backbone of Poly-
(132) Kennedy, Z. C.; Cardenas, A. J. P.; Corbey, J. F.; Warner, M. (Ethylenimine). J. Polym. Sci., Part A: Polym. Chem. 1999, 37, 3936−
G. 2,6-Diiminopiperidin-1- ol: An Overlooked Motif Relevant to 3942.
Uranyl and Transition Metal Binding on Poly(amidoxime) Adsorb- (153) Ramkumar, J.; Chandramouleeswaran, S. Separation of Uranyl
ents. Chem. Commun. 2016, 52, 8802−8805. Ion using Polyaniline. J. Radioanal. Nucl. Chem. 2013, 298, 1543−
(133) Lashley, M. A.; Mehio, N.; Nugent, J. W.; Holguin, E.; Do- 1549.
(154) Sellin, R.; Alexandratos, S. D. Polymer-Supported Primary
Thanh, C.-L.; Bryantsev, V. S.; Dai, S.; Hancock, R. D. Amidoximes as
Amines for the Recovery of Uranium from Seawater. Ind. Eng. Chem.
Ligand Functionalities for Braided Polymeric Materials for the
Res. 2013, 52, 11792−11797.
Recovery of Uranium from Seawater. Polyhedron 2016, 109, 81−91.
(155) Alexandratos, S. D.; Zhu, X.; Florent, M.; Sellin, R. Polymer-
(134) Vukovic, S.; Hay, B. P.; Bryantsev, V. S. Predicting Stability
Supported Bifunctional Amidoximes for the Sorption of Uranium from
Constants for Uranyl Complexes Using Density Functional Theory.
Seawater. Ind. Eng. Chem. Res. 2016, 55, 4208−4216.
Inorg. Chem. 2015, 54, 3995−4001. (156) Barber, P. S.; Kelley, S. P.; Griggs, C. S.; Wallace, S.; Rogers, R.
(135) Porcheddu, A.; Giacomelli, G. The Chemistry of Hydroxyl-
D. Surface Modification of Ionic Liquid-Spun Chitin Fibers for the
amines, Oximes and Hydroxamic Acids; John Wiley & Sons, Ltd.: New Extraction of Uranium from Seawater: Seeking the Strength of Chitin
York, 2008. and the Chemical Functionality of Chitosan. Green Chem. 2014, 16,
(136) Tian, G.; Teat, S.; Rao, L. Thermodynamic Studies of U(VI) 1828−1836.
Complexation with Glutardiamidoxime for Sequestration of Uranium (157) Huang, L.; Zhang, L.; Hua, D. Synthesis of Polyamidoxime-
from Seawater. Dalton. Trans. 2013, 42, 5690−5696. Functionalized Nanoparticles for Uranium(VI) Removal from Neutral
(137) Endrizzi, F.; Melchior, A.; Tolazzi, M.; Rao, L. Complexation Aqueous Solutions. J. Radioanal. Nucl. Chem. 2015, 305, 445−453.
of Uranium(VI) with Glutarimidoxioxime: Thermodynamic and (158) Wei, Y.; Qian, J.; Huang, L.; Hua, D. Bifunctional Polymeric
Computational Studies. Dalton. Trans. 2015, 44, 13835−13844. Microspheres for Efficient Uranium Sorption from Aqueous Solution:
(138) Kang, S. O.; Vukovic, S.; Custelcean, R.; Hay, B. P. Cyclic Synergistic Interaction of Positive Charge and Amidoxime Group. RSC
Imide Dioximes: Formation and Hydrolytic Stability. Ind. Eng. Chem. Adv. 2015, 5, 64286−64292.
Res. 2012, 51, 6619−6624. (159) Omichi, H.; Katakai, A.; Sugo, T.; Okamoto, J. A New Type of
(139) Kawakami, T.; Akiyama, E.; Hori, K.; Nagase, Y.; Sugo, T. Amidoxime-Group- Containing Adsorbent for the Recovery of
Structural Changes of Compounds Containing Cyano Groups by Uranium from Seawater. Sep. Sci. Technol. 1985, 20, 163−178.
Hydroxylamine Treatment. Trans. Mater. Res. Soc. Jpn. 2002, 27, 783− (160) Omichi, H.; Katakai, A.; Sugo, T.; Okamoto, J. A New Type of
786. Amidoxime-Group- Containing Adsorbent for the Recovery of
(140) Hirotsu, T.; Katoh, S.; Sugasaka, K.; Takai, N.; Seno, M.; Uranium from Seawater. II. Effect of Grafting of Hydrophilic
Itagaki, T. Adsorption of Uranium on Cross-Linked Amidoxime Monomers. Sep. Sci. Technol. 1986, 21, 299−313.
Polymer from Seawater. Ind. Eng. Chem. Res. 1987, 26, 1970−1977. (161) Omichi, H.; Katakai, A.; Sugo, T.; Okamoto, J.; Katoh, S.;
(141) Hirotsu, T.; Katoh, S.; Sugasaka, K.; Takai, N.; Seno, M.; Sakane, K.; Sugasaka, K.; Itagaki, T. Effect of Shape and Size of
Itagaki, T. Kinetics of Adsorption of Uranium on Amidoxime Amidoxime-Group-Containing Adsorbent on the Recovery of
Polymers from Seawater. Sep. Sci. Technol. 1988, 23, 49−61. Uranium from Seawater. Sep. Sci. Technol. 1987, 22, 1313−1325.
(142) Tbal, H.; Delporte, M.; Morcellet, J.; Morcellet, M. (162) Kawai, T.; Saito, K.; Sugita, K.; Kawakami, T.; Kanno, J.-i.;
Functionalization and Chelating Properties of a Porous Polymer Katakai, A.; Seko, N.; Sugo, T. Preparation of Hydrophilic Amidoxime
Derived from Vinylamine. Eur. Polym. J. 1992, 28, 671−679. Fibers by Cografting Acrylonitrile and Methacrylic Acid from an
(143) Tbal, H.; Morcellet, J.; Delporte, M.; Morcellet, M. Uranium Optimized Monomer Composition. Radiat. Phys. Chem. 2000, 59,
Adsorption by Chelating Resins Containing Amino Groups. J. 405−411.
Macromol. Sci., Part A: Pure Appl.Chem. 1992, 29, 699−710. (163) Kawai, T.; Saito, K.; Sugita, K.; Katakai, A.; Seko, N.; Sugo, T.;
(144) Ramachandhran, V.; Kumar, S. C.; Sudarsanan, M. Preparation, Kanno, J.-i.; Kawakami, T. Comparison of Amidoxime Adsorbents
Characterization, and Performance Evaluation of Styrene-Acrylonitrile- Prepared by Cografting Methacrylic Acid and 2-Hydroxyethyl
Amidoxime Sorbent for Uranium Recovery from Dilute Solutions. J. Methacrylate with Acrylonitrile onto Polyethylene. Ind. Eng. Chem.
Macromol. Sci., Part A: Pure Appl.Chem. 2001, 38, 1151−1166. Res. 2000, 39, 2910−2915.

14010 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

(164) Hunt, M. A.; Saito, T.; Brown, R. H.; Kumbhar, A. S.; Naskar, (180) Liu, H.; Yu, M.; Deng, B.; Li, L.; Jiang, H.; Li, J. Pre-Irradiation
A. K. Patterned Functional Carbon Fibers from Polyethylene. Adv. Induced Emulsion Graft Polymerization of Acrylonitrile Onto
Mater. 2012, 24, 2386−2389. Polyethylene Nonwoven Fabric. Radiat. Phys. Chem. 2012, 81, 93−96.
(165) Choi, S.-H.; Nho, Y. C. Adsorption of UO2+2 by Polyethylene (181) Pekel, N.; Şahiner, N.; Akkaş, P.; Güven, O. Uranyl Ion
Adsorbents with Amidoxime, Carboxyl, and Amidoxime/Carboxyl Adsorptivity of N-Vinyl 2- Pyrrolidone/Acrylonitrile Copolymeric
Group. Radiat. Phys. Chem. 2000, 57, 187−193. Hydrogels Containing Amidoxime Groups. Polym. Bull. 2000, 44,
(166) Choi, S.-H.; Choi, M.-S.; Park, Y.-T.; Lee, K.-P.; Kang, H.-D. 593−600.
Adsorption of Uranium Ions by Resins with Amidoxime and (182) Sahiner, N.; Pekel, N.; Akkas, P.; Guven, O. Amidoximation
Amidoxime/Carboxyl Group Prepared by Radiation- Induced and Characterization of New Complexing Hydrogels Prepared from
Polymerization. Radiat. Phys. Chem. 2003, 67, 387−390. N-Vinyl 2-Pyrrolidone/Acrylonitrile Systems. J. Macromol. Sci., Part A:
(167) Das, S.; Liao, W. P.; Flicker Byers, M.; Tsouris, C.; Janke, C. J.; Pure Appl.Chem. 2000, 37, 1159−1172.
Mayes, R. T.; Schneider, E.; Kuo, L. J.; Wood, J. R.; Gill, G. A.; et al. (183) Pekel, N.; Şahiner, N.; Güven, O. Use of Amidoximated
Alternative Alkaline Conditioning of Amidoxime Based Adsorbent for Acrylonitrile/N-Vinyl 2- Pyrrolidone Interpenetrating Polymer Net-
Uranium Extraction from Seawater. Ind. Eng. Chem. Res. 2016, 55, works for Uranyl Ion Adsorption from Aqueous Systems. J. Appl.
4303−4312. Polym. Sci. 2001, 81, 2324−2329.
(168) Das, S.; Tsouris, C.; Zhang, C.; Kim, J.; Brown, S.; Oyola, Y.; (184) Saraydin, D.; Şolpan, D.; Işıkver, Y.; Ekici, S.; Güven, O.
Janke, C. J.; Mayes, R. T.; Kuo, L. J.; Wood, J. R.; et al. Enhancing Radiation Crosslinked Poly(Acrylamide/2-Hydroxypropyl Methacry-
Uranium Uptake by Amidoxime Adsorbent in Seawater: An late/Maleic Acid) and their Usability in the Uptake of Uranium. J.
Investigation for Optimum Alkaline Conditioning Parameters. Ind. Macromol. Sci., Part A: Pure Appl.Chem. 2002, 39, 969−990.
Eng. Chem. Res. 2016, 55, 4294−4302. (185) Şolpan, D.; Güven, O. Adsorption of Uranyl Ions into
(169) Prasad, T. L.; Tewari, P. K.; Sathiyamoorthy, D. Parametric Poly(Acrylamide-co-Acrylic Acid) Hydrogels Prepared by Gamma
Studies on Radiation Grafting of Polymeric Sorbents for Recovery of Irradiation. J. Macromol. Sci., Part A: Pure Appl.Chem. 2005, 42, 485−
Heavy Metals from Seawater. Ind. Eng. Chem. Res. 2010, 49, 6559− 494.
6565. (186) Szwarc, M. Living Polymers. Nature 1956, 178, 1168−1169.
(170) Chi, F.; Hu, S.; Xiong, J.; Wang, X. Adsorption Behavior of (187) Szwarc, M.; Levy, M.; Milkovich, R. Polymerization Initiated
Uranium on Polyvinyl Alcohol-G-Amidoxime: Physicochemical by Electron Transfer to Monomer. A New Method of Formation of
Properties, Kinetic and Thermodynamic Aspects. Sci. China: Chem. Block Polymers. J. Am. Chem. Soc. 1956, 78, 2656−2657.
(188) Williams, V. A.; Matyjaszewski, K. Expanding the ATRP
2013, 56, 1495−1503.
(171) Xing, Z.; Hu, J.; Wang, M.; Zhang, W.; Li, S.; Gao, Q.; Wu, G. Toolbox: Methacrylate Polymerization with an Elemental Silver
Properties and Evaluation of Amidoxime-Based UHMWPE Fibrous Reducing Agent. Macromolecules 2015, 48, 6457−6464.
(189) Wang, J.-S.; Matyjaszewski, K. Controlled/″Living″ Radical
Adsorbent for Extraction of Uranium from Seawater. Sci. China: Chem.
Polymerization. Atom Transfer Radical Polymerization in the Presence
2013, 56, 1504−1509.
of Transition-Metal Complexes. J. Am. Chem. Soc. 1995, 117, 5614−
(172) Hu, J.; Ma, H.; Xing, Z.; Liu, X.; Xu, L.; Li, R.; Lin, C.; Wang,
5615.
M.; Li, J.; Wu, G. Preparation of Amidoximated Ultrahigh Molecular
(190) Matyjaszewski, K.; Jo, S. M.; Paik, H.-j.; Shipp, D. A. An
Weight Polyethylene Fiber by Radiation Grafting and Uranium
Investigation into the CuX/2,2′-Bipyridine (X = Br or Cl) Mediated
Adsorption Test. Ind. Eng. Chem. Res. 2016, 55, 4118−4124.
Atom Transfer Radical Polymerization of Acrylonitrile. Macromolecules
(173) Li, R.; Pang, L.; Ma, H.; Liu, X.; Zhang, M.; Gao, Q.; Wang, H.;
1999, 32, 6431−6438.
Xing, Z.; Wang, M.; Wu, G. Optimization of Molar Content of
(191) Yue, Y.; Mayes, R. T.; Kim, J.; Fulvio, P. F.; Sun, X.-G.;
Amidoxime and Acrylic Acid in UHMWPE Fibers for Improvement of Tsouris, C.; Chen, J.; Brown, S.; Dai, S. Seawater Uranium Sorbents:
Seawater Uranium Adsorption Capacity. J. Radioanal. Nucl. Chem. Preparation from a Mesoporous Copolymer Initiator by Atom-
2017, 311, 1771−1779. Transfer Radical Polymerization. Angew. Chem., Int. Ed. 2013, 52,
(174) Mehio, N.; Lashely, M. A.; Nugent, J. W.; Tucker, L.; Correia, 13458−13462.
B.; Do-Thanh, C.-L.; Dai, S.; Hancock, R. D.; Bryantsev, V. S. Acidity (192) Yue, Y.; Zhang, C.; Tang, Q.; Mayes, R. T.; Liao, W.-P.; Liao,
of the Amidoxime Functional Group in Aqueous Solution: A C.; Tsouris, C.; Stankovich, J. J.; Chen, J.; Hensley, D. K.; et al. A
Combined Experimental and Computational Study. J. Phys. Chem. B Poly(acrylonitrile)-Functionalized Porous Aromatic Framework Syn-
2015, 119, 3567−3576. thesized by Atom-Transfer Radical Polymerization for the Extraction
(175) Mehio, N.; Williamson, B.; Oyola, Y.; Mayes, R. T.; Janke, C.; of Uranium from Seawater. Ind. Eng. Chem. Res. 2016, 55, 4125−4129.
Brown, S.; Dai, S. Acidity of the Poly(acrylamidoxime) Adsorbent in (193) Chi, F.; Wen, J.; Xiong, J.; Sheng, H.; Gong, Z.; Qiu, T.; Wei,
Aqueous Solution: Determination of the Proton Affinity Distribution G.; Yi, F.; Wang, X. Controllable Polymerization of Poly-DVB−VBC−
via Potentiometric Titrations. Ind. Eng. Chem. Res. 2016, 55, 4217− g−AO Resin Via Surface-Initiated Atom Transfer Radical Polymer-
4223. ization for Uranium Removal. J. Radioanal. Nucl. Chem. 2015, 309,
(176) Zeng, Z.; Wei, Y.; Shen, L.; Hua, D. Cationically Charged 787−796.
Poly(amidoxime)-Grafted Polypropylene Nonwoven Fabric for (194) Sather, A. C.; Berryman, O. B.; Rebek, J. Selective Recognition
Potential Uranium Extraction from Seawater. Ind. Eng. Chem. Res. and Extraction of the Uranyl Ion from Aqueous Solutions with a
2015, 54, 8699−8705. Recyclable Chelating Resin. Chem. Sci. 2013, 4, 3601−3605.
(177) Oyola, Y.; Dai, S. High Surface-Area Amidoxime-Based (195) Abney, C. W.; Liu, S.; Lin, W. Tuning Amidoximate to
Polymer Fibers Co-Grafted with Various Acid Monomers Yielding Enhance Uranyl Binding: A Density Functional Theory Study. J. Phys.
Increased Adsorption Capacity for the Extraction of Uranium from Chem. A 2013, 117, 11558−11565.
Seawater. Dalton. Trans. 2016, 45, 8824−8834. (196) Hammett, L. P. The Effect of Structure upon the Reactions of
(178) Ma, H.; Yao, S.; Li, J.; Cao, C.; Wang, M. A Mild Method of Organic Compounds. Benzene Derivatives. J. Am. Chem. Soc. 1937, 59,
Amine-Type Adsorbents Syntheses with Emulsion Graft Polymer- 96−103.
ization of Glycidyl Methacrylate on Polyethylene Non-Woven Fabric (197) Zhou, L.; Bosscher, M.; Zhang, C.; Ö zçubukçu, S.; Zhang, L.;
by Pre-Irradiation. Radiat. Phys. Chem. 2012, 81, 1393−1397. Zhang, W.; Li, C. J.; Liu, J.; Jensen, M. P.; Lai, L.; et al. A Protein
(179) Chi, H.-Y.; Liu, X.-Y.; Ma, H.-J.; Yang, X.-J.; Yu, M.; Zhang, J.- Engineered to Bind Uranyl Selectively and with Femtomolar Affinity.
Y.; Wang, M.; Li, J.-Y.; Hoshina, H.; Seko, N. Adsorption Behavior of Nat. Chem. 2014, 6, 236−241.
Uranyl Ions onto Amino-Type Adsorbents Prepared by Radiation- (198) Hay, B. P.; Firman, T. K. HostDesigner: A Program for the de
Induced Graft Copolymerization. Nucl. Sci. Technol. 2014, 25, 010302. Novo Structure-Based Design of Molecular Receptors with Binding

14011 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

Sites that Complement Metal Ion Guests. Inorg. Chem. 2002, 41, (220) Gorka, J.; Mayes, R. T.; Baggetto, L.; Veith, G. M.; Dai, S.
5502−5512. Sonochemical Functionalization of Mesoporous Carbon for Uranium
(199) Bryantsev, V. S.; Hay, B. P. De Novo Structure-Based Design Extraction from Seawater. J. Mater. Chem. A 2013, 1, 3016−3026.
of Bisurea Hosts for Tetrahedral Oxoanion Guests. J. Am. Chem. Soc. (221) Mayes, R. T.; Górka, J.; Dai, S. Impact of Pore Size on the
2006, 128, 2035−2042. Sorption of Uranyl under Seawater Conditions. Ind. Eng. Chem. Res.
(200) Reyheller, C.; Hay, B. P.; Kubik, S. Influence of Linker 2016, 55, 4339−4343.
Structure on the Anion Binding Affinity of Biscyclopeptides. New J. (222) Schierz, A.; Zänker, H. Aqueous Suspensions of Carbon
Chem. 2007, 31, 2095−2102. Nanotubes: Surface Oxidation, Colloidal Stability and Uranium
(201) Custelcean, R.; Bosano, J.; Bonnesen, P. V.; Kertesz, V.; Hay, Sorption. Environ. Pollut. 2009, 157, 1088−1094.
B. P. Computer-Aided Design of a Sulfate-Encapsulating Receptor. (223) Wang, M.; Qiu, J.; Tao, X.; Wu, C.; Cui, W.; Liu, Q.; Lu, S.
Angew. Chem., Int. Ed. 2009, 48, 4025−4029. Effect of pH and Ionic Strength on U(IV) Sorption to Oxidized
(202) Young, N. J.; Hay, B. P. Structural Design Principles for Self- Multiwalled Carbon Nanotubes. J. Radioanal. Nucl. Chem. 2011, 288,
Assembled Coordination Polygons and Polyhedra. Chem. Commun. 895−901.
2013, 49, 1354−1379. (224) Shao, D.; Jiang, Z.; Wang, X.; Li, J.; Meng, Y. Plasma Induced
(203) Vukovic, S.; Hay, B. P. De Novo Structure-Based Design of Grafting Carboxymethyl Cellulose on Multiwalled Carbon Nanotubes
Bis-amidoxime Uranophiles. Inorg. Chem. 2013, 52, 7805−7810. for the Removal of UO22+ from Aqueous Solution. J. Phys. Chem. B
(204) Maloubier, M.; Solari, P. L.; Moisy, P.; Monfort, M.; Den 2009, 113, 860−864.
Auwer, C.; Moulin, C. XAS and TRLIF Spectroscopy of Uranium and (225) Yue, Y.; Sun, X.; Mayes, R. T.; Kim, J.; Fulvio, P. F.; Qiao, Z.;
Neptunium in Seawater. Dalton. Trans. 2015, 44, 5417−5427. Brown, S.; Tsouris, C.; Oyola, Y.; Dai, S. Polymer-Coated Nanoporous
(205) Maher, K.; Bargar, J. R.; Brown, G. E. Environmental Carbons for Trace Seawater Uranium Adsorption. Sci. China: Chem.
Speciation of Actinides. Inorg. Chem. 2013, 52, 3510−3532. 2013, 56, 1510−1515.
(206) Hirose, K.; Sugimura, Y. Chemical Speciation of Particulate (226) Abney, C. W.; Gilhula, J. C.; Lu, K.; Lin, W. Metal-Organic
Uranium in Seawater. J. Radioanal. Nucl. Chem. 1991, 149, 83−96. Framework Templated Inorganic Sorbents for Rapid and Efficient
(207) Abney, C. W.; Das, S.; Mayes, R. T.; Kuo, L.-J.; Wood, J.; Gill, Extraction of Heavy Metals. Adv. Mater. 2014, 26, 7993−7997.
G.; Piechowicz, M.; Lin, Z.; Lin, W.; Dai, S. A Report of Emergent (227) Abney, C. W.; Taylor-Pashow, K. M. L.; Russell, S. R.; Chen,
Uranyl Binding Phenomena by an Amidoxime Phosphonic Acid Co- Y.; Samantaray, R.; Lockard, J. V.; Lin, W. Topotactic Transformations
Polymer. Phys. Chem. Chem. Phys. 2016, 18, 23462−23468. of Metal−Organic Frameworks to Highly Porous and Stable Inorganic
(208) Grant, C. D.; Kang, S. O.; Hay, B. P. Synthesis of a Hydrophilic Sorbents for Efficient Radionuclide Sequestration. Chem. Mater. 2014,
Naphthalimidedioxime. J. Org. Chem. 2013, 78, 7735−7740. 26, 5231−5243.
(209) Chatterjee, S.; Bryantsev, V. S.; Brown, S.; Johnson, J. C.; (228) Feng, Y.; Jiang, H.; Li, S.; Wang, J.; Jing, X.; Wang, Y.; Chen,
Grant, C. D.; Mayes, R. T.; Hay, B. P.; Dai, S.; Saito, T. Synthesis of M. Metal−Organic Frameworks HKUST-1 for Liquid-Phase Adsorp-
Naphthalimidedioxime Ligand-Containing Fibers for Uranium Ad- tion of Uranium. Colloids Surf., A 2013, 431, 87−92.
sorption from Seawater. Ind. Eng. Chem. Res. 2016, 55, 4161−4169. (229) Saito, K.; Miyauchi, T. Chemical Forms of Uranium in
(210) Mellah, A.; Chegrouche, S.; Barkat, M. The Removal of Artificial Seawater. J. Nucl. Sci. Technol. 1982, 19, 145−150.
Uranium(VI) from Aqueous Solutions Onto Activated Carbon: (230) Bai, Z.-Q.; Yuan, L.-Y.; Zhu, L.; Liu, Z.-R.; Chu, S.-Q.; Zheng,
Kinetic and Thermodynamic Investigations. J. Colloid Interface Sci. L.-R.; Zhang, J.; Chai, Z.-F.; Shi, W.-Q. Introduction of Amino Groups
2006, 296, 434−441. Into Acid-Resistant MOFs for Enhanced U(VI) Sorption. J. Mater.
(211) Starvin, A. M.; Rao, T. P. Solid Phase Extractive Chem. A 2015, 3, 525−534.
Preconcentration of Uranium(VI) onto Diarylazobisphenol Modified (231) Zhang, J.-Y.; Zhang, N.; Zhang, L.; Fang, Y.; Deng, W.; Yu, M.;
Activated Carbon. Talanta 2004, 63, 225−232. Wang, Z.; Li, L.; Liu, X.; Li, J. Adsorption of Uranyl Ions on Amine-
(212) Bahr, J. L.; Tour, J. M. Highly Functionalized Carbon functionalization of MIL-101(Cr) Nanoparticles by a Facile
Nanotubes Using in Situ Generated Diazonium Compounds. Chem. Coordination-based Post-synthetic Strategy and X-ray Absorption
Mater. 2001, 13, 3823−3824. Spectroscopy Studies. Sci. Rep. 2015, 5, 13514.
(213) Ismail, A. F.; Yim, M.-S. Investigation of Activated Carbon (232) Wang, L. L.; Luo, F.; Dang, L. L.; Li, J. Q.; Wu, X. L.; Liu, S. J.;
Adsorbent Electrode for Electrosorption-Based Uranium Extraction Luo, M. B. Ultrafast High-Performance Extraction of Uranium from
from Seawater. Nucl. Eng. Technol. 2015, 47, 579−587. Seawater Without Pretreatment using an Acylamide- and Carboxyl-
(214) Liu, C.; Hsu, P.-C.; Xie, J.; Zhao, J.; Wu, T.; Wang, H.; Liu, W.; Functionalized Metal-Organic Framework. J. Mater. Chem. A 2015, 3,
Zhang, J.; Chu, S.; Cui, Y. A Half-Wave Rectified Alternating Current 13724−13730.
Electrochemical Method for Uranium Extraction from Seawater. (233) Luo, F.; Wang, M.-S.; Luo, M.-B.; Sun, G.-M.; Song, Y.-M.; Li,
Nature Energy 2017, 2, 17007. P.-X.; Guo, G.-C. Functionalizing the Pore Wall of Chiral Porous
(215) Zhao, Y.; Liu, C.; Feng, M.; Chen, Z.; Li, S.; Tian, G.; Wang, Metal-Organic Frameworks by Distinct - H, -OH, -NH2, -NO2,
L.; Huang, J.; Li, S. Solid Phase Extraction of Uranium(VI) Onto -COOH Shutters Showing Selective Adsorption of CO2, Tunable
Benzoylthiourea-Anchored Activated Carbon. J. Hazard. Mater. 2010, Photoluminescence, and Direct White-Light Emission. Chem.
176, 119−124. Commun. 2012, 48, 5989−5991.
(216) Kim, J.; Lee, H.; Yeon, J.-W.; Jung, Y.; Kim, J. Removal of (234) Tanabe, K. K.; Wang, Z.; Cohen, S. M. Systematic
Uranium(VI) from Aqueous Solutions by Nanoporous Carbon and its Functionalization of a Metal-Organic Framework via a Postsynthetic
Chelating Polymer Composite. J. Radioanal. Nucl. Chem. 2010, 286, Modification Approach. J. Am. Chem. Soc. 2008, 130, 8508−8517.
129−133. (235) Cohen, S. M. Postsynthetic Methods for the Functionalization
(217) Wang, Y.-Q.; Zhang, Z.-B.; Liu, Y.-H.; Cao, X.-H.; Liu, Y.-T.; of Metal−Organic Frameworks. Chem. Rev. 2012, 112, 970−1000.
Li, Q. Adsorption of U(VI) from Aqueous Solution by the Carboxyl- (236) Yang, W.; Bai, Z.-Q.; Shi, W.-Q.; Yuan, L.-Y.; Tian, T.; Chai,
Mesoporous Carbon. Chem. Eng. J. 2012, 198−199, 246−253. Z.-F.; Wang, H.; Sun, Z.-M. MOF-76: From a Luminescent Probe to
(218) Tian, G.; Geng, J. X.; Jin, Y. D.; Wang, C. L.; Li, S. Q.; Chen, Highly Efficient U(VI) Sorption Material. Chem. Commun. 2013, 49,
Z.; Wang, H.; Zhao, Y. S.; Li, S. J. Sorption of Uranium(VI) using 10415−10417.
Oxime-Grafted Ordered Mesoporous Carbon CMK-5. J. Hazard. (237) Luo, B.-C.; Yuan, L.-Y.; Chai, Z.-F.; Shi, W.-Q.; Tang, Q.
Mater. 2011, 190, 442−450. U(VI) Capture From Aqueous Solution by Highly Porous And Stable
(219) Wang, H.; Ma, L.; Cao, K.; Geng, J.; Liu, J.; Song, Q.; Yang, X.; MOFs: UiO-66 and its Amine Derivative. J. Radioanal. Nucl. Chem.
Li, S. Selective Solid- Phase Extraction of Uranium by Salicylidenei- 2016, 307, 269−276.
mine-Functionalized Hydrothermal Carbon. J. Hazard. Mater. 2012, (238) Carson, F.; Pascanu, V.; Bermejo Gómez, A.; Zhang, Y.;
229−230, 321−330. Platero-Prats, A. E.; Zou, X.; Martín-Matute, B. Influence of the Base

14012 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013
Chemical Reviews Review

on Pd@MIL-101-NH2(Cr) as Catalyst for the Suzuki−Miyaura Cross- Iminodiacetic Acid Derivative Functionalized SBA-15 as Adsorbents.
Coupling Reaction. Chem. - Eur. J. 2015, 21, 10896−10902. Dalton. Trans. 2014, 43, 3739−3749.
(239) Rowan, S. J.; Cantrill, S. J.; Cousins, G. R. L.; Sanders, J. K. M.; (257) Lee, H. I.; Kim, J. H.; Kim, J. M.; Kim, S.; Park, J.-N.; Hwang, J.
Stoddart, J. F. Dynamic Covalent Chemistry. Angew. Chem., Int. Ed. S.; Yeon, J.-W.; Jung, Y. Application of Ordered Nanoporous Silica for
2002, 41, 898−952. Removal of Uranium Ions from Aqueous Solutions. J. Nanosci.
(240) Jin, Y.; Yu, C.; Denman, R. J.; Zhang, W. Recent Advances in Nanotechnol. 2010, 10, 217−221.
Dynamic Covalent Chemistry. Chem. Soc. Rev. 2013, 42, 6634−6654. (258) Qu, J.; Li, W.; Cao, C.-Y.; Yin, X.-J.; Zhao, L.; Bai, J.; Qin, Z.;
(241) Bai, C.; Li, J.; Liu, S.; Yang, X.; Yang, X.; Tian, Y.; Cao, K.; Song, W.-G. Metal Silicate Nanotubes with Nanostructured Walls as
Huang, Y.; Ma, L.; Li, S. In Situ Preparation of Nitrogen-Rich and Superb Adsorbents for Uranyl Ions and Lead Ions in Water. J. Mater.
Functional Ultramicroporous Carbonaceous COFs by “Segregated” Chem. 2012, 22, 17222−17226.
Microwave Irradiation. Microporous Mesoporous Mater. 2014, 197, (259) Yuan, L.-Y.; Liu, Y.-L.; Shi, W.-Q.; Li, Z.-j.; Lan, J.-H.; Feng, Y.-
148−155. X.; Zhao, Y.-L.; Yuan, Y.-L.; Chai, Z.-F. A Novel Mesoporous Material
(242) Li, J.; Yang, X.; Bai, C.; Tian, Y.; Li, B.; Zhang, S.; Yang, X.; for Uranium Extraction, Dihydroimidazole Functionalized SBA-15. J.
Ding, S.; Xia, C.; Tan, X.; et al. A Novel Benzimidazole-Functionalized Mater. Chem. 2012, 22, 17019−17026.
2-D COF Material: Synthesis and Application as a Selective Solid- (260) Wang, X.; Zhu, G.; Guo, F. Removal of Uranium(VI) Ion from
Phase Extractant for Separation Of Uranium. J. Colloid Interface Sci. Aqueous Solution by SBA-15. Ann. Nucl. Energy 2013, 56, 151−157.
2015, 437, 211−218. (261) Chouyyok, W.; Pittman, J. W.; Warner, M. G.; Nell, K. M.;
(243) Bai, C.; Zhang, M.; Li, B.; Zhao, X.; Zhang, S.; Wang, L.; Li, Y.; Clubb, D. C.; Gill, G. A.; Addleman, R. S. Surface Functionalized
Zhang, J.; Ma, L.; Li, S. Modifiable Diyne-Based Covalent Organic Nanostructured Ceramic Sorbents for the Effective Collection and
Framework: A Versatile Platform for In Situ Multipurpose Recovery of Uranium from Seawater. Dalton. Trans. 2016, 45, 11312−
Functionalization. RSC Adv. 2016, 6, 39150−39158. 11325.
(244) McKeown, N. B.; Budd, P. M. Polymers of Intrinsic (262) Kou, S.; Yang, Z.; Sun, F. Protein Hydrogel Microbeads for
Microporosity (PIMs): Organic Materials for Membrane Separations, Selective Uranium Mining from Seawater. ACS Appl. Mater. Interfaces
Heterogeneous Catalysis and Hydrogen Storage. Chem. Soc. Rev. 2006, 2017, 9, 2035−2039.
35, 675−683. (263) Farina, R.; Laugel, N.; Pincus, P.; Tirrell, M. Brushes of Strong
(245) Budd, P. M.; Ghanem, B. S.; Makhseed, S.; McKeown, N. B.; Polyelectrolytes in Mixed Mono- and Tri-Valent Ionic Media at Fixed
Msayib, K. J.; Tattershall, C. E. Polymers of Intrinsic Microporosity Total Ionic Strengths. Soft Matter 2013, 9, 10458−10472.
(PIMs): Robust, Solution-Processable, Organic Nanoporous Materials. (264) Farina, R.; Laugel, N.; Yu, J.; Tirrell, M. Reversible Adhesion
Chem. Commun. 2004, 230−231. with Polyelectrolyte Brushes Tailored via the Uptake and Release of
(246) Kaur, P.; Hupp, J. T.; Nguyen, S. T. Porous Organic Polymers Trivalent Lanthanum Ions. J. Phys. Chem. C 2015, 119, 14805−14814.
in Catalysis: Opportunities and Challenges. ACS Catal. 2011, 1, 819− (265) Yu, J.; Mao, J.; Yuan, G.; Satija, S.; Jiang, Z.; Chen, W.; Tirrell,
835. M. Structure of Polyelectrolyte Brushes in the Presence of Multivalent
(247) Ben, T.; Ren, H.; Ma, S. Q.; Cao, D. P.; Lan, J. H.; Jing, X. F.; Counterions. Macromolecules 2016, 49, 5609−5617.
Wang, W. C.; Xu, J.; Deng, F.; Simmons, J. M.; et al. Targeted (266) Brettmann, B.; Pincus, P.; Tirrell, M. Lateral Structure
Synthesis of a Porous Aromatic Framework with High Stability and Formation in Polyelectrolyte Brushes Induced by Multivalent Ions.
Exceptionally High Surface Area. Angew. Chem., Int. Ed. 2009, 48, Macromolecules 2017, 50, 1225−1235.
9457−9460. (267) Dobrynin, A. V.; Rubinstein, M. Theory of Polyelectrolytes in
(248) Li, B.; Sun, Q.; Zhang, Y.; Abney, C. W.; Aguila, B.; Lin, W.; Solutions and at Surfaces. Prog. Polym. Sci. 2005, 30, 1049−1118.
Ma, S. Functionalized Porous Aromatic Framework for Efficient (268) Pincus, P. Colloid Stabilization with Grafted Polyelectrolytes.
Uranium Adsorption from Aqueous Solutions. ACS Appl. Mater. Macromolecules 1991, 24, 2912−2919.
Interfaces 2017, 9, 12511−12517.
(249) Sihn, Y. H.; Byun, J.; Patel, H. A.; Lee, W.; Yavuz, C. T. Rapid
Extraction of Uranium Ions from Seawater Using Novel Porous
Polymeric Adsorbents. RSC Adv. 2016, 6, 45968−45976.
(250) Fryxell, G. E.; Lin, Y.; Fiskum, S.; Birnbaum, J. C.; Wu, H.;
Kemner, K.; Kelly, S. Actinide Sequestration Using Self-Assembled
Monolayers on Mesoporous Supports. Environ. Sci. Technol. 2005, 39,
1324−1331.
(251) Yuan, L.-Y.; Liu, Y.-L.; Shi, W.-Q.; Lv, Y.-L.; Lan, J.-H.; Zhao,
Y.-L.; Chai, Z.-F. High Performance of Phosphonate-Functionalized
Mesoporous Silica for U(VI) Sorption from Aqueous Solution. Dalton.
Trans. 2011, 40, 7446−7453.
(252) Lebed, P. J.; Savoie, J.-D.; Florek, J.; Bilodeau, F.; Lariviere, D.;
Kleitz, F. Large Pore Mesostructured Organosilica-Phosphonate
Hybrids as Highly Efficient and Regenerable Sorbents for Uranium
Sequestration. Chem. Mater. 2012, 24, 4166−4176.
(253) Wang, X.; Yuan, L.; Wang, Y.; Li, Z.; Lan, J.; Liu, Y.; Feng, Y.;
Zhao, Y.; Chai, Z.; Shi, W. Mesoporous Silica SBA-15 Functionalized
with Phosphonate and Amino Groups for Uranium Uptake. Sci. China:
Chem. 2012, 55, 1705−1711.
(254) Zhang, W.; Ye, G.; Chen, J. Novel Mesoporous Silicas Bearing
Phosphine Oxide Ligands with Different Alkyl Chains for the Binding
of Uranium in Strong HNO3 Media. J. Mater. Chem. A 2013, 1,
12706−12709.
(255) Sert, S.; Eral, M. Uranium Adsorption Studies on Aminopropyl
Modified Mesoporous Sorbent (NH2-MCM-41) using Statistical
Design Method. J. Nucl. Mater. 2010, 406, 285−292.
(256) Wang, Y.-L.; Song, L.-J.; Zhu, L.; Guo, B.-L.; Chen, S.-W.; Wu,
W.-S. Removal of Uranium(VI) from Aqueous Solution Using

14013 DOI: 10.1021/acs.chemrev.7b00355


Chem. Rev. 2017, 117, 13935−14013

You might also like