Hydration Number: Recent Highlights I

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Hydration Number

The concept of hydration number depends on the experimental method used for its
determination.

From: Handbook of Surfaces and Interfaces of Materials, 2001

Related terms:

Solubility, Aqueous Solution, Clathrate Hydrate, Mols, Nafion, Proton Conductivity,


Proton, Hydrate Formation

View all Topics

Recent Highlights I
Sara Lacerda, ... Éva Tóth, in Advances in Inorganic Chemistry, 2021

2.3.1 pH responsive probes


Recent Mn-based pH responsive probes rely on a hydration number change. The
derivatization of PC2A with the pH sensing ethylene amine (EA) moiety (PC2A-EA,
Scheme 2) provided a Mn2 + complex which is very stable thermodynamically (pM =-
 9.27) and more importantly very inert kinetically (t1/2 = 8 × 103 h at pH 7.4).22 Its pH
response is optimal at physiologically relevant pH range with a r1 = 3.54 mM− 1 s− 1 in
the range pH 3.7–5.8 decreasing to 2.04 mM− 1 s− 1 at pH 8.4, at 0.49 T and 25 °C. This
relaxivity decrease is attributed to the deprotonation and consequent coordination
of the EA moiety at high pH, leading to a non-hydrated complex, while in the range
3.7–5.8 the complex is protonated and mono-hydrated. Below pH 3.5, an increase
in r1 is observed, as the complex becomes deprotonated and dissociates, releasing
“free” Mn2 +. In the presence of HSA, a better pH discrimination is observed owing
to their r2 relaxivities at 3 T, yielding a fivefold change between pH 6.7 and 7.5.

Derivatives of other ligand types using a different pH sensing moiety, the N-


-ethyl-4-(trifluoromethyl)-benzenesulfonamide were also reported.72 These ligands
yield stable 6- and 7-coordinated Mn2 + complexes, with pM of 6.63 and 7.82, for
NO2ASAm (Scheme 2) and DPASAm, respectively (Scheme 3). For MnDPASAm, the
protonation of the sulfonamide arm (log KMnHL = 6.43) in the complex gives a 2.3-fold
relaxivity increase between pH 9.0 and 4.0 (at 20 MHz, 25 °C), which was attributed to
the coordination of a second water molecule. Nevertheless, this complex dissociates
very fast in the presence of 10 equivalents of Zn2 + at 25 °C, so additional optimization
is needed for further applications. As for NO2ASAm, the stronger Mn-Nsulfonamide
interaction yields a less basic complex protonation (log KMnHL = 5.5). In the pH range
7–10 the complex is non-hydrated and has a r1 of 1.6 mM− 1 s− 1 (10 MHz, 25 °C) while
below pH 7 a sharp increase is observed (r1 = 11.5 mM− 1 s− 1 at pH 2.5) indicating
dissociation of the complex.

> Read full chapter

Development of Expert Networks: A


Hybrid System of Expert Systems and
Neural Networks
D.R. Baughman, Y.A. Liu, in Neural Networks in Bioprocessing and Chemical Engi-
neering, 1995

Nomenclature

Cij : virial coefficients.


d : mean diameter of the ions.
F : Faraday constant.
H : Debye-Huckel parameter.
h : hydration numbers for an-
ions and cations.
I : Ionic strength of solution.
Kb : partition contribution from
affinity binding.
Ke : partition contribution from
electrophoresis.
Kp : overall partition coefficient.
KpH : partition contribution due to
pH variation from the isoelectric
point.
Ks : partition coefficient of the
polymer solution at standard
conditions (temperature at 25°C
and pH at the isoelectric point).
Kt : partition contribution due to
the temperature variation from
25°C.
L : length of a compound.
Mn : number-average molecular
weight.
Mw :
weight-average molecular
weight.
mi : ratio of the molar volume of
component i to that of a refer-
ence component.
ni : number of moles of species i.
Na : Avogadro number.
P : pressure.
PI : polydispersity index, the ration
of Mw to Mn.
Ri : radius of component i.
S : entropy.
T : temperature.
Tj : internal threshold value for pro-
cessing element j.
V : volume.
vi : valence of ion i.
wi : weight fraction of species i.
wji : weighting factor from the ith in-
put variable to the jth output
variable.
xi : input variable to a processing el-
ement.
yj : output variable from a process-
ing element.
zb : net charge of the species.
i : proportionality factor between
volume and weight-fraction dif-
ferences between the phases for
species i
: Debye-Huckel constant.
ij : selectivity, ratio of partition co-
efficients between components
i and j.
ij : binary specific-interaction coef-
ficient between solute molecules
i and j.
ij : Flory-Huggins interaction pa-
rameter between components i
and j.
12 : solubility-parameter difference
between PEG and Dextran.
1s : solubility-parameter difference
between PEG and water.
2s : solubility-parameter difference
between Dextran and water.
ΔDextran : gradient of Dextran between
phases.
ΔGmix : Gibbs free energy of mixing.
ΔHmix : enthalpy of mixing.
ΔPEG : gradient of PEG between phas-
es.
ΔSmix : entropy of mixing.
Δ : electrostatic potential difference
between the phases.
i : volume fraction of component i.
: ratio of weight fraction differ-
ence between phases for Dex-
tran and PEG.
jo : standard state activity as defined
by Guggenheim.
μi : chemical potential of compo-
nent i.

> Read full chapter

Polymers for a Sustainable Environ-


ment and Green Energy
D.S. Kim, ... M.D. Guiver, in Polymer Science: A Comprehensive Reference, 2012

10.36.2.2.1 Under fully hydrated conditions


A simplified model for proton conduction is illustrated in Figure 6 that considers
two contributing factors, interanionic distance ( ) and number of participating
ion-carrying media (hydration number, ). In this model, proton conductivity can
be expressed by

Figure 6. Schematic diagram of proton conduction in hydrated sulfonated PEMs.

[2]

where is hydration number (number of water molecules per sulfonic acid group)
and is the distance between sulfonic acid groups in hydrated polymer.

The interanionic distance in dry polymers can be approximated by Molar Volume per
Charge (MVC). MVC is an estimate of equivalent volume (cm3 per ionomer or the
mol equivalent of acid groups) based on the summation of molar volume subunits
(Table 1) rather than true volume measurements, as shown in eqn [3]:

Table 1. Molar volume increments of selected groups39

Groups Va(298) (cm3 mol)−1 Groups Va(298) (cm3 mol)−1


5.5 32.5
9.12
16.37 18.72
23.7 73.3
19.85 5.28
21.87
84.16 65.5
30.7
20.0 69
49.0
100.5 112
17.3
8.5 6.88
8.0
34.08 20.22
25.76
8.32 26.56
40.5

[3]

where vi is the volumetric contribution of the ith structural group that appears ni
times per charge.

The distance between sulfonic acid groups in hydrated polymers can be approximat-
ed by eqn [4]:

[4]

where VH2O is the molar volume of water, 18 cm3 mol−1.

The MVCWET parameter assumes all absorbed water resides in hydrophilic domains
(which is reasonable) and neglects polymer morphological features (which is not
valid in general but the morphological effect can be considered minimized at high
water content where ionic clusters are well developed).

Then conductivity is proportional to Percent Conducting Volume (PCV) that is


defined as

[5]

The numerator of the eqn [5] is an estimate of the volume of the aqueous (hy-
drophilic) phase per acid site, and the denominator is an estimate of the volume
of the hydrated membrane (both hydrophilic and hydrophobic phases) per acid site.
PCV is essentially a ratio of the volume of the hydrophilic phase to the hydrated
membrane analogous to the conducting volume of the membrane.

Figure 7 plots the conductivity of a wide variety of polymers electrolytes as a function


of PCV under fully hydrated conditions. A general correlation exists between proton
conductivity in fully hydrated PEMs and PCV, in spite of the data scattering. Proton
conductivities of arbitrarily categorized PEMs under fully hydrated conditions do
not show significant differences at a given PCV, and correlate with a master curve
(shown by the black dash line). This indicates that there is an apparent diminished
role of molecular structural/morphology effects on proton conductivity for full
hydrated PEMs using the PCV parameter, probably due to the fact that incorporation
of significant amounts of water (    > l0) in fully hydrated PEMs diminishes the
importance of structural characteristics by the formation of more continuous, less
tortuous hydrophilic domains. Consequently, the conductivity of all different kinds
of polymers under fully hydration can be approximated by a single PCV parameter.
Proton conductivity increases 1 order of magnitude as PCV increases from 0 to
0.35 and then increases more slowly. The marginal increase of proton conductivity
beyond a PCV value of 0.35 is due to the volume dilution of sulfonic acid groups with
water, such that proton conduction occurs less effectively. In this higher PCV range,
PEMs exhibit excessive water uptake and their mechanical properties (modulus and
strength) deteriorate accordingly. As a result, most well-performing PEMs for H2/air
fuel cells in a fully hydrated environment have PCV values of 0.35 where conductivity
is relatively high ( 0.05–0.1 S cm−1) while mechanical properties are reasonably
good. Table 2 shows properties of selected PEMs at PCV values of around 0.33.
The PEMs selected have reasonably low water uptake, which helps to remove issues
related to excessive water uptake during fuel cell operations. It is seen that all PEMs
in Table 2 have similar conductivity of 80 mS cm−1, significantly different MVC
and values. The functional PEM has relatively low MVC and , while PFSA PEM has
relatively high MVC and . In other words, a similar conductivity can be obtained
with higher sulfonic acid concentration and lower water uptake or conversely, with
lower sulfonic acid concentration and higher water uptake. The water absorption
capacity of each PEM is consistent with the results in the water uptake section, which
showed that functional or cross-linked PEMs have relatively lower water uptake than
HC-based PEMs, which in turn have lower water uptake than fluorinated PEMs.

Figure 7. Relative proton conductivity vs. percent conducting volume (PCV) of


various sulfonated polymers under fully hydrated conditions at ambient tempera-
ture.20Reprinted from Ann. Rev. Chem. Biomol. Eng., 2010, Y. S. Kim et al., Moving
beyond mass-based parameters for conductivity analysis of sulfonated polymers,
123–148, 2010, with permission from Annual Reviews.

Table 2. Comparison of PEM properties14

IEC MVC WU
PEM (meq. g−1) (cm3 equiv.- PCV (wt.%) (mS cm−1)
 mol.−1)
Poly(arylene 2.7 277 6.2 0.29 43 81
ether nitrile)
homopoly-
mer
Poly(arylene 1.66 435 10.4 0.30 30 74
ether)
cross-linked
polymer
Poly(arylene 1.54 486 14.4 0.35 40 72
ether
sulfone)
random
copolymer
PFSA Nafion 0.91 524 15.0 0.34 26 91
copolymer

Figure 8 compares H2/air fuel cell performance using the selected PEMs. All PEMs
have a similar thickness ( 50 µm) and fabricated in a same manner and tested under
same operating conditions. It is noted that H2/air fuel cell performance and high
frequency resistance (HFR) using these membranes are comparable with each other,
as expected.

Figure 8. H2/air fuel cell performance under fully humidified conditions at 80 °C
using sulfonated poly(arylene ether nitrile) homopolymer (SPAEN 1.0), cross-linked
partially sulfonated poly(arylene ether) random copolymer (CSFQH80-BP),14 sul-
fonated poly(arylene ether sulfone) random copolymer (BPSH-35),13 and PFSA
(Nafion 212).

> Read full chapter

Latent heat storage for solar heating


and cooling systems
C.A. Infante Ferreira, in Advances in Solar Heating and Cooling, 2016

16.4.2.1 Phase diagram


Aqueous solutions of tetra-n-butylammonium (TBAB) can form two types of hydrate
crystals depending on the initial solution concentration. Type A has a columnar
shape and hydration number 26 and type B has an undefined form and hydration
number 38. The solid-liquid equilibrium line at different solution concentrations
at ambient pressure has been experimentally determined by several researchers
(Darbouret et al., 2005; Hayashi et al., 2000; Ma et al., 2010; Oyama et al., 2005) and
is, on average, as indicated in Fig. 16.8.

Figure 16.8. Phase diagram of tetra-n-butylammonium solutions in water at ambi-


ent pressure. Type A hydrates can reach phase change temperatures up to 12.5°C
whereas the highest phase change temperature for type B hydrates is approximately
10°C. PCM, Phase change material.

Pumpable slurries can be directly used to chill the air flow for cooling applications.
For this reason a higher temperature would be more advantageous, and a phase
change temperature of 12.5°C is generally preferred. From Fig. 16.8 it is clear that
at this phase change temperature only type A hydrates can be formed. The required
initial TBAB aqueous solution concentration is then 36.5 wt%.

> Read full chapter


Low temperature and distillation sys-
tems
Stephen A. Rackley, in Carbon Capture and Storage (Second Edition), 2017

9.2.1 Physical fundamentals


Gas hydrates, also known as clathrate hydrates or clathrates, are a crystalline phase
of water in which a guest molecule is trapped within a cage of hydrogen-bonded
water molecules and were discovered by Humphry Davy in 1811. In the case of CO2
hydrates, the cage of water molecules takes the form of a 12- or 14-faced polyhedra
(dodecahedron or tetradecahedron) having pentagonal or hexagonal faces with an
oxygen atom at each vertex, similar to normal ice, as illustrated in Figure 9.5. These
polyhedra are denoted by the nomenclature 512 and 51262, respectively, where AmBn
indicates a polyhedron with m A-sided and n B-sided faces.

Figure 9.5. CO2 hydrate cage structures.Source: Graphic elements from Yang (2011)
and Ma (2016).

CO2 and methane hydrates have an overall crystalline structure designated Type
I (or SI), in which the unit cell consists of 46 water molecules forming two small
(512) and six larger (51262) cages. All such cages need not be occupied, and a key
challenge in hydrate-based gas separation is to maximize occupancy by the desired
guest molecule. If all cages are occupied, the stoichiometry of this structure, for
CO2, is (CO2:5.75H2O). Larger guests, such as O2, N2 and many thermodynamic
promoters (see below) form Type II (SII) hydrate structures, with a unit cell of 136
water molecules forming sixteen small (512) and eight larger hexadecahedral (51264)
cages; hydrogen also forms SII hydrates, but in this case the cages are occupied by
clusters of two or four H2 molecules. Even larger guests, such as butane (C4H10),
form Type H (SH) structures with two types of small cage (three×512 and two×435663)
and one very large 51268 cage. The largest guests, often used as chemical hydrate pro-
moters (see below), form semi-clathrates in which the largest cages are broken—to
accommodate the oversize guest—and other, smaller guests form part of the cage
structure as well as occupying the smaller cages.

The formation of hydrates occurs under conditions of low temperature and high
pressure; the equilibrium phase diagram of some CCS relevant gases is shown
in Figure 9.6 (the hydrate formation region is above the line in each case). This
process, also known as enclathration, is exothermic, the enthalpy of formation for
CO2 hydrate being ~65 kJ/mol-CO2.

Figure 9.6. Equilibrium phase diagram for the hydrates of some CCS relevant gases.

When hydrate formation is used for the separation of gas mixtures, the keep
requirements are as follows:

• Short induction time: the time taken for hydrate crystal nuclei to form in the
gas–water mixture
• High gas consumption: the total amount of gas taken up as clathrate in the
process, a convenient measure being the hydration number(9.3)(9.4)where
L and S measure the occupancy of large and small cages. For maximum

occupancy ( L= S=1) the hydration numbers for SI and SII hydrates are 5.75
and 5.66, corresponding to maximum theoretical gas uptake of 0.174 and
0.176 mol-gas/mol-water. In SI hydrates, all cages can be occupied by CO2
giving a maximum theoretical uptake of 0.174 mol-CO2/mol-water; SII hy-
drates however can only enclathrate CO2 when stabilized by a larger promoter
molecule, which will occupy the large cages, leading to a maximum theoretical
uptake of 0.118 mol-CO2/mol-water.
• High CO2 recovery from the feed gas stream; this is measured by the split
fraction, denoted S.Fr(9.5)being the ratio of the number of moles of CO2 in
the hydrate (H) versus feed. Values of S.Fr of 0.35–0.40 are typically reported
for single-stage CO2 separation (~10 MPa, ~0.5°C) from simulated flue gas
mixtures with 15%–20% CO2 content.

High separation efficiency of CO2 versus other gases enclathrated from
the feed stream, denoted by the separation factor S.F(9.6)where “A” denotes
another gas component in the feed and ngas indicates moles in the residual
gas phase. Values of S.F of 10–15 are typically reported for single-stage CO2
separation from simulated flue gas with a 15%–20% CO2 content, where the
“A” component is nitrogen.

As one might anticipate, these are competing requirements; for example, raising
the operating pressure will reduce the induction time and increase overall gas
consumption but will generally reduce the separation factor, as other gases will also
be enclathrated. It also increases the energy requirement for compression, which
would result in undesirable additional operating cost. Chemical additives, known as
hydrate promoters, have been a major recent R&D focus with the aim of improving
all the above factors. Thermodynamic promoters, of which tetrahydrofuran (THF,
C4H8O) is a much studied example, are chemicals which form clathrates at very low
pressure. In the case of THF, which is a relatively large molecule, type SII clathrates
form rapidly at 1 bar and 4.5°C. Both the large 51264 and small 512 cavities can
be occupied by CO2, but there is competition for occupancy with other molecules,
including THF itself, and H2 or N2, depending on the CO2 separation application
(pre- or post-combustion). The impact of a number of thermodynamic promoters,
including propane, THF, and tetra-n-butyl ammonium bromide, on the equilibrium
hydration conditions for IGCC fuel gas is shown in Figure 9.7.

Figure 9.7. Impact of thermodynamic promoters on equilibrium hydration of IGCC


fuel gas.Source: After Babu, 2015.

Although these promoters are effective at accelerating hydrate formation and mov-
ing the equilibrium hydrate formation line to lower pressure/higher temperature,
they tend to adversely effect gas consumption and split fraction, since many of the
clathrate cages will be occupied by promoter molecules.

Surfactants are another type of chemical additive which work as kinetic promoters,
enhancing the solubility of the target gas in water and accelerating the formation
of gas hydrate but, unlike thermodynamic promoters, not directly affecting the
clathrate formation process. Surfactant molecules contain both hydrophobic and
hydrophilic sites; association of gas + surfactant molecules occurs at the hydrophobic
site and inclusion of the gas molecule into the growing hydrate cages is then
facilitated by the strong affinity to water at the surfactant’s hydrophilic site.

> Read full chapter

Micelles
Srinivas Manne, L.K. Patterson, in Encyclopedia of Physical Science and Technology
(Third Edition), 2003

III.C Hydration in the Micelle


Because there is effectively a hydrocarbon–water interface at the micelle surface,
several possibilities must be considered for hydration in the micelle. These involve
counterions, headgroups, and the portion of the hydrocarbon chain near the Stern
layer. The bound counterions themselves may differ considerably in extent of hy-
dration; the cations noted above maintain a higher degree of hydration than do the
anions. There has been considerable evidence that much of the water of hydration is
maintained upon binding. From viscosity and diffusion studies, hydration numbers
for micelles have been deduced for a variety of charged surfactants; these fall
into the range of 5–10 water molecules per surfactant molecule. For uncharged
micelles with variable numbers of oxyethylene groups, the hydration number can
vary considerably, as seen from Fig. 6.

FIGURE 6. Dependence of the number of water molecules per polyoxyethylene chain


on the number of oxyethylene units for micelles of sodium dodecylpolyoxyethylene
sulfates. [From Takiwa, F., and Ohki, K. (1967). J. Phys. Chem. 71, 1343–1349.]

The extent to which water interacts with the hydrocarbon chains of the surfactant
is a question yet unsettled; various workers have suggested complete, partial, or
no water penetration into the micellar core. For aliphatic surfactants, persuasive
evidence has been given to suggest that the core of the micelle contains no sub-
stantial amount of water and that if there is penetration it probably involves no
more than two carbons. Surfactants containing oxyethylene moieties appear to be
substantially hydrated in the region of these groups. This is implied in Fig. 6. A
number of studies involving probes solubilized in the micelle or attached to the
surfactant itself (e.g., carbonyl groups) indicate the presence of water around such
probes. It is possible that perturbation of the system by the included probe can
lead to significant water–hydrocarbon contact and/or that many such probes lie
near the surface. Of course, as has been shown, the dynamic equilibrium between
surfactants in bulk phase and micelle assures considerable motion of material across
the micelle boundary. Additionally, it is quite likely that the micelle surface is not a
simple two-dimensional interface but involves some protrusion of surfactants into
the bulk where hydrocarbon wetting could occur (Fig. 7). In summary, one should
note that whether there are 5–10 water molecules associated with each surfactant
of a simple micelle or whether there is penetration to several carbons depth in the
presence of solubilized material, a substantial fraction of material associated with
the micelle is, indeed, water.

FIGURE 7. Schematic of the dynamic protrusion of surfactant monomers from


spherical and rod-shaped micelles into surrounding bulk phase to enhance hydro-
carbon–water content. [From Lindman, B., and Wennerström, H. (1980). Top. Curr.
Chem. 87, 1–83.]
> Read full chapter

Interactions Involving Polar Molecules


Jacob N. Israelachvili, in Intermolecular and Surface Forces (Third Edition), 2011

4.5 Strong Ion-Dipole Interactions in Water: Hydrated Ions


For small or multivalent ions in highly polar solvents such as water,2 the strong
orientation dependence of their ion-dipole interaction will tend to orient the solvent
molecules around them, favoring = 0° near cations and = 180° near anions (cf.
inset in Figure 4.2). Thus, in water Li+, Be2+, Mg2+, and Al3+ ions have a number of
water molecules orientationally bound to them. Such ions are called solvated ions or
hydrated ions, and the number of water molecules they bind—usually between 4
and 6 is known as their hydration number (Table 4.2). It should be noted, however,
that these bound water molecules are not completely immobilized and that they do
move and exchange with bulk water, albeit more slowly. The hydration number is
more of a qualitative indicator of the degree to which ions bind water rather than an
exact value.

Table 4.2. Hydrated radii and Hydration Numbers of Ions in Water (Approximate)

Ion Bare ion radius Hydrated radius Hydration num- Lifetime/ex-


(nm) (nm) ber (±1) change rate (s)
H3O+ — 0.28 3 —
Li+ 0.068 0.38 5 5 × 10−9
Na+ 0.095 0.36 4 10−9
K+ 0.133 0.33 3 10−9
Cs+ 0.169 0.33 1 5 × 10−10

Be2+ 0.031 0.46 4a 10−3


Mg2+ 0.065 0.43 6a 10−6
Ca2+ 0.099 0.41 6 10−8

Al3+ 0.050 0.48 6a 0.1–1


Cr3+ 0.052 — 6a >3 hrs

OH– 0.176 0.30 3


F– 0.136 0.35 2
Cl– 0.181 0.33 1 ~10−11
Br– 0.195 0.33 1 ~10−11
I– 0.216 0.33 0 ~10−11
0.264 0.34 0
0.347 0.37 0
The hydration number gives the number of water molecules in the primary shell (Fig.
3.4), though the total number of water molecules affected can be much larger and
depends on the method of measurement. Similarly, the hydrated radius depends
on how it is measured. Different methods can yield radii that can be as much as
0.1 nm smaller or larger than those shown. Table compiled from data given by
Nightingale (1959), Amis (1975), Saluja (1976), Bockris and Reddy (1970), and Cotton
and Wilkinson (1980).

a Number of water molecules forming a stoichiometric complex with the


ion—for example, [Be(H2O)4]2+.

Closely related to the hydration number is the effective radius or hydrated radius of
an ion in water, which is larger than its real radius (i.e., its crystal lattice radius), as
shown in Table 4.2. Because smaller ions are more hydrated due to their more intense
electric field they tend to have larger hydrated radii than larger ions. However, very
small ions such as Be2+ have lower hydration numbers because they are too small
for more than 4 water molecules to pack around them. Hydration numbers and
radii can be deduced from measurements of the viscosity, diffusion, compressibility,
conductivity, solubility, and various thermodynamic and spectroscopic properties
of electrolyte solutions, and the results rarely agree with one another (Amis, 1975;
Saluja, 1976).

More insight into the nature of ion hydration can be gained by considering the
average time that water molecules remain bound to ions. In the pure liquid at room
temperature the water molecules tumble about with a mean reorientation time or
rotational correlation time of about 10−11 s. This also gives an estimate of the lifetime
of the water-water bonds formed in liquid water (the hydrogen bonds).3 But when the
water molecules are near ions, various techniques, such as oxygen nuclear magnetic
resonance, x-ray and neutron diffraction, and IR or Raman spectroscopy, show that
the mean lifetimes or exchange rates of water molecules in the first hydration shell4
can be much longer, varying from 10−11 s to many hours (Hertz, 1973; Cotton and
Wilkinson, 1980). For very weakly solvated ions (usually large monovalent ions) such
as N(CH3)4+, Cl–, Br–, and I–, these lifetimes are not much different from that for
water in bulk water, and they can even be shorter (referred to as negative hydration).

Cations are generally more solvated than anions of the same valency, since they are
smaller—having lost rather than gained an electron. Thus, for K+, Na+, and Li+, the
residence times of water molecules in the primary hydration shells are about 10−9 s.
Divalent cations are always more strongly solvated than monovalent cations, and for
Ca2+ and Mg2+, the bound water lifetimes are about 10−8 s and 10−6 s, respectively.
Even longer lifetimes are observed for very small divalent cations such as Be2+ (10−3
s), while for trivalent cations such as Al3+ and Cr3+ these can be seconds or hours.
In such cases the binding is so strong that an ion-water complex is actually formed
of fixed stoichiometry (see Table 4.2). In fact, these quasi-stable complexes begin to
take on the appearance of (charged) molecules and are often designated as such—for
example, [Mg(H2O)6]2+, [Be(H2O)4]2+. Small divalent and especially trivalent cations
have a weak but well-defined second hydration shell (Bergström et al., 1991).

Protons H+ always associate with one water molecule, which goes by the name of the
hydronium ion or oxonium ion H3O+, while three water molecules are solvated around
this ion to form H3O+(H2O)3. Likewise, the hydroxyl ion OH– is believed to be solvated
by three water molecules forming OH–(H2O)3. The structure of the hydronium ion
H3O+ is believed to be planar, with two positive charges and one negative charge at
the three apexes of an equilateral triangle.

> Read full chapter

Physical and chemical properties of ac-


tinides in nuclear fuel reprocessing
A. Paulenova, in Advanced Separation Techniques for Nuclear Fuel Reprocessing and
Radioactive Waste Treatment, 2011

2.3.4 Hydration in concentrated solutions


As the electrolyte concentration increases, the number of water molecules in the
secondary hydration sphere decreases. Consequently, there is a tightening of the
bond between the metal cation and the hydrate waters in the inner sphere (Choppin,
Jensen, 2006). Based on NMR studies of trivalent actinides and lanthanides, Choppin
concluded that inner sphere complexation by perchlorate ions does not occur below
approximately 8–10 M (Choppin, Labonne-Wall, 1997). Multiple equilibria for the
uranyl chloride system (UO2Cl2(H2O)2, UO2Cl3(H2O)–, and UO2Cl42–) have been used
for separation of uranium from its progeny or other metals. Since Th4 + does not form
anionic chloride complexes, it is retained on cation-exchange resin while anionic
chloride complexes of UO22 + pass through the column in the eluate. Alternatively,
such anionic complexes can be retained on an anionexchange column.

The hydration number of Eu(III) remains relatively constant in hydrochloric acid


up to approximately 6–8 M, above which concentration it decreases. The same is
true for the hydration number of Cm(III) in HCl, which begins a decline at about
5 M HCl. This difference between (Eu3 + and Cm3 +) reflects greater complexation of
the actinide trivalent ion by the relatively soft anion Cl–. The difference in chloride
complexation has been used to provide efficient separation of trivalent actinides
from trivalent actinides in concentrated HCl solutions by passage through columns
of cation exchange resin since 1950s (Diamond et al., 1954).

Nitrate complexes for tetravalent actinides, for example, Th4 + and Pu4 +, are extremely
important in actinide separation and purification processes. Nitrate ions begin to
form inner sphere complexes at lower concentrations than chloride anions; this
observation is confirmed by the decreased hydration number of the cation even at
relatively lower concentrations (Choppin, Jensen, 2006). However, since the oxygen
atoms of the nitrate are hard donors, there is no evidence of any covalent enhance-
ment in its bonding as is seen with the chloride anions for the trivalent actinide
cations relative to the lanthanide cations (Choppin, Jensen, 2006). In separation and
purification processes, the nitrate complexes of actinides are extremely important.
Nitrate-nitric acid solution is the most common aqueous medium in nuclear sepa-
ration processes. In the case of neutral extractants such as tributylphosphate (TBP),
carbamoyl methyl phosphine oxide (CMPO) or dipicolinamides (DPA) it provides
nitrate units necessary to compensate the actinide cation charge to enable extraction.
Nitrate complexation with hexavalent actinide ions is very weak and the determi-
nation of the formation constants for aqueous nitrate solution species is extremely
difficult. Under aqueous conditions with high nitric acid concentrations, com-
plexes of the form AnO2(NO3)(H2O)x+, AnO2(NO3)2(H2O)2, and AnO2(NO3)3– (An = U,
Np, Pu) are likely to be present. The limiting species in the nitrate series is the
hexanitrato complex, An(NO3)62– (Matonic et al., 2002). The complexation of the Pa
and Np pentavalent ions by nitrate is known; however, limited thermodynamic and
structural data are available. The presumed stoichiometry for the Np(V) species is
NpO2(NO3)(H2O)x. For protactinium, which easily hydrolyzes, mixed hydroxo/nitrato
or oxo/nitrato complexes have been proposed.

Fluorides and chlorides are the best studied actinide-halide systems, and they are
very important for the pyroprocessing and electrorefining processes.

Carboxylic acids are strongly bound to actinide ions. The primary binding mode
for simple carboxylic acids is bidentate, while in polycarboxylic acid complexes,
carboxylates tend toward monodentate coordination with the metal ion. The affinity
of the low-valent actinides for these ligands increases with the denticity of the
ligand, for example, ethylenediaminetetraacetate (EDTA) >> > acetate. For An4 +,
the EDTA ligand is hexadentate with a twist conformation (a spiral conformation,
wrapping around the metal ion, rather than encapsulating the metal ion in a
central cavity in the manner of tripodal or macrobicyclic ligands). Diethylenetri-
amine-N,N,N ,N ,N -pentaacetate (DTPA) has an even higher affinity for both An3 +
and An4 + ions.

> Read full chapter


Water and Salt Fluxes in Pressure Re-
tarded Osmosis
Khaled Touati, Fernando Tadeo, in Pressure Retarded Osmosis, 2017

5.3 Effect of the Draw Solution Composition

5.3.1 The Hydrated Energy


The hydration energy or the enthalpy of hydration, Hhyd, of an ion is the amount
of energy released when a mole of the ion dissolves in a large amount of water
forming an infinite dilute solution [30]. The enthalpy of hydration of electrolytes
can be measured experimentally, but the individual enthalpies for the cations and
the anions cannot be separated. The enthalpy of hydration of individual ions can be
mathematically estimated according to Max Born the theoretical model [7] by using
the following general relationship:

(2.40)

Eq. (2.40) is applied here for some ions, Na+, K+, Cl−, and Ca2 to estimate the
hydration energy at 20°C. In this section, the effect of the draw solution compo-
sition is investigated. Three draw solutions were tested (sodium chloride, potassium
chloride, and calcium chloride). These solutions were prepared using concentrations
that correspond to the osmotic pressure of seawater (Δπ ≈ 27 bar). Fig. 2.16 shows
the variation of the salt flux diffusion and the water flux for each case tested. It can
be clearly seen that the performance of PRO drastically changes with the type of
the chemical entity composing the draw solution. The highest salt flux was found
when the draw solution was based on potassium and sodium chlorides, whereas
the lowest was for calcium chloride. This distinct behavior of the salt diffusion can
be attributed to the ion size in aqueous solutions: Since the ionic radius changes
in aqueous solutions, it is necessary to consider the hydrated radius. Table 2.6
shows the hydrated radius of tested ions in aqueous solutions. From Fig. 2.16B, it
can be seen that there is a strong relationship between the salt diffusion and the
hydrated ions: sodium chloride and potassium chloride revealed a high passage
through the membrane because of their small radius compared to the other entities.
The lowest salt flux diffusion occurs at the measurement CaCl2 solution, which
is characterized by the bigger radius. Fig. 2.16A shows the influence of the draw
solution composition on the water flux. The highest water flux occurs with potassium
salt and the lowest water flux was observed with calcium salt. This behavior can be
explained by considering the hydration properties of ions. The hydration number
for the sodium was found to be 5 or 6 water molecules, whereas the potassium's
hydration number had a probability distribution ranging from 5 to 10 [32]. However,
the hydration number for the calcium had a probability distribution ranging from
6 to 13 [33]. With increasing the hydration number, the binding energy between
water molecules and ion surface also increases. Correspondingly, this reduces the
water permeability inside the solution and makes the separation tougher, which is
clearly seen when we compare the energies of hydration of calcium, sodium, and
potassium illustrated in Table 2.6. As shown in Eqs. (2.8) and (2.10), the diffusion
coefficient of the draw solution affects the ECP. In fact, the diffusion minimizes
the difference in concentrations between the draw bulk and the draw boundary layer
(CD,b and CD,m); when the diffusion coefficient is higher, the ECP is minimized
and thus the water flux increases. Moreover, a possible penetration of the salts in
the membrane, causes ICP, especially when the salt is blocked in the inner structure
of the membrane, and cannot be removed due to their size. Table 2.5 shows the
diffusion coefficients in aqueous solution of the ions investigated. The calcium
salt revealed the lowest salt flux, which can be explained by the fact that CaCl2 has
the lowest diffusion coefficient and the thickest hydrated radius, which leads to
more severe ICP and ECP than the other salts. Potassium and sodium chloride have
high diffusion coefficients, which decrease ECP and thus enhance the water flux.
However, these entities have small ionic hydrated radius, which increases the salt
flux, and thus, the water flux decreases. These two simultaneous and antagonist
effects explain the measured water flux.

Figure 2.16. Experimental salt flux Js (A) and water flux Jw (B) for different draw
solutions as a function of the applied pressure [31].

Table 2.5. Diffusions Coefficients of Tested Salts [31]

Salt Diffusion Coefficient (10−9 m2/s)


NaCl 1.50
KCl 1.90
CaCl2 1.12
Table 2.6. Enthalpy of Hydration and Hydrated Radius of Studied Ions [31]

Ion ΔHhyd (kJ/mol) Eq. (2.30) Hydrated Radius (10−12 m)

Na+ −617.8 178


K+ −459.2 201
Ca2+ −2448.8 260
Cl− −418.0 195

5.3.2 Effect of the Membrane Orientation


In PRO, two membrane orientations are experimentally used: the active layer facing
the draw solution AL–DS and the active layer facing the feed solution AL–FS (-
Fig. 2.17). In fact the orientation AL–DS is most frequent in the literature because the
performance and the stability of the membrane are significantly better. The effect of
ICP can be modeled by adopting the classical solution-diffusion theory for the dense
rejection layer coupled with convection and diffusion transport of the solute in the
porous support layer [12]. It should be pointed out that the concentrative ICP occurs
under AL–DS orientation and the dilutive ICP occurs under AL–FS. In this section,
the effect of the membrane orientation on the salt flux diffusion and the power
output is studied. For that, 1 M NaCl draw solution and two feed solutions (8.55 and
55 mM) were tested under both membrane orientations. From Fig. 2.18, it can be
seen that the performance of PRO operated under AL–DS is better at high pressure
and uses low feed solution concentration; the power density is much higher and the
membrane seems to be more stable (Fig. 2.18E and F). This behavior is attributed to
the severe ICP that occurs in the support layer when it is facing the draw solution.
This result confirms that the dilutive ICP in AL–FS orientation is more severe than
the concentrative ICP in AL–DS orientation. However, the salt flux diffusion for PRO
under AL–FS is lower than PRO operated under AL–DS for the two studied feed
solutions because of the remarkable difference between concentrative and dilutive
ICPs in both cases. When the feed solution was 55 mM, the difference in PRO
performance for the two orientations was not significant. Moreover, when operating
at low pressure under AL–FS orientation, the membrane seems to be stable in terms
of water flux and salt flux diffusion (Fig. 2.18B and D). The study of the specific salt
diffusion for both orientations reveals that the ratio (Js/Jw) increases slightly for low
pressure values (ΔP < 10 bar). In addition, this behavior appears to be similar for both
orientations regardless of the feed solution concentration. However, for relatively
high pressures (ΔP > 10 bars), the effect of the feed solution concentration and the
membrane orientation becomes perspicuous. In fact, in AL–DS case, the increase
in FS concentration is followed by a drastic increase of Js/Jw because of the sharp
decrease in the water flux Jw. This behavior of Jw is mainly due to the effect of ICP,
because the effect of the salt diffusion is reduced when FS concentration increases
as it was shown in Section 5.1. On the other hand, in the AL–FS configuration, (Js/Jw)
increases drastically for both FS concentrations. This result shows that, even with low
FS concentrations, severe ICP occurs because of the penetration of salt in the support
layer due to its porous structure. These results justify the tendency of using the
AL–DS orientation in PRO, especially at high pressures, to guarantee a high power
density. However, it should be pointed that the AL–FS orientation is stable at low
pressures, which means that it can be useful in these conditions. On the other hand,
it has been shown in Ref. [34] that this orientation is more resistible to membrane
fouling compared to AL–DS. Moreover, it was shown here that the salt diffusion is
lower for AL–FS, which means that the contribution of Js in membrane scaling and
fouling is lower than AL–DS, especially using feed solutions that contain fouling and
scaling precursors.

Figure 2.17. Schematics of membrane orientations in pressure retarded osmosis


process. The concentrative and dilutive internal concentration polarizations (ICPs)
are also shown [31]. AL–DS, active layer facing the draw solution; AL–FS, active layer
facing the feed solution.
Figure 2.18. Effect of the membrane orientation on salt flux (A) and (B), the specific
salt flux (C) and (D), and the power density (E) and (F). 1 M NaCl draw solution,
T = 20°C for both feed and draw solutions, cross-flow velocity of 50 mL/min on both
sides of membrane [31]. AL–DS, active layer facing the draw solution; AL–FS, active
layer facing the feed solution.

> Read full chapter

Silica scale control in geothermal


plants—historical perspective and cur-
rent technology1
P. von Hirtz, in Geothermal Power Generation, 2016
16.3 Thermodynamics of silica solubility
Although quartz is thermodynamically more stable than amorphous silica, extreme
conditions of temperature, pressure, and/or alkalinity are required for the growth of
quartz at measurable rates in aqueous solutions. The greater solubility of amorphous
silica relative to quartz is a distinct advantage for geothermal resource production
facilities because it limits the precipitation of silica from produced waters. The
solubility curves of amorphous silica and quartz both increase with temperature to
300°C and then decrease due to the decreasing density and solvent power of water.
The decrease becomes rapid as the critical point of water (374°C) is exceeded.

Dissolved salts and pH also affect the solubility of silica in aqueous solutions.
Fournier and Marshall [6] have developed equations to calculate the solubility of
amorphous silica at circum-neutral pH from 25 to 300°C using the concept of
effective density of water and “salting out” effects of mixed electrolytes. Cations
exhibiting elevated “hydration numbers,” such as the alkaline-earths, depress the
solubility of amorphous silica more than cations exhibiting low “hydration numbers”
due to “free” water available for solvation. The effect of salinity and temperature on
amorphous silica solubility is shown in Fig. 16.2.

Figure 16.2. Solubility of amorphous silica as a function of temperature in NaCl


solutions.

The solubility of silica is substantially independent of pH until the pH increases into


the alkaline range. Goto [7] examined the effect of pH on the solubility of amorphous
silica from 0 to 200°C and from pH 5.5 to 10.0. As expected, the solubility of
amorphous silica increased with increasing temperature, while solubilities remained
relatively constant over the pH range of about 5.5–8.5. Above pH 8.5, the solubility
increased significantly according to the reactions:

[16.2]

An expression for the relationship between pH and the solubility of monomeric silica
is given by Eq. [16.3] [8]:

[16.3]

where ws = total solubility of monomeric silica and wm = mass fraction of molecular


silicic acid (undissociated).

The first ionization constant for molecular silicic acid, K1, controls the extent of the
reaction:

[16.4]

where K1 = 10 9.7 at 25°C.

The ionization constant can be calculated at higher temperatures using Eq. [16.5]:

[16.5]

where T = temperature in K.

The effect of pH on the equilibrium solubility of monomeric silica does not generally
have practical significance under flashed geothermal brine conditions where pH
usually ranges from 5 to 8.5. But above pH 7, monosilicic acid does begin to
dissociate according to Eq. [16.4]. This increases the thermodynamic solubility of
silica because the silicate ion has a very high solubility. At a fixed concentration of
total dissolved silica, the silica saturation ratio or silica saturation index (SSI, the
ratio of silica concentration to solubility under the given conditions) decreases with
increasing pH. At high pH, silica precipitation can be thermodynamically inhibited.

Nonalkali and alkaline-earth cations present in geothermal fluids may react directly
with silicic acid to form metal silicates. These silicates are usually poorly crystalline;
their X-ray diffraction patterns exhibit broad humps that are shifted from the normal
opal-A hump centered near the diffraction angle of 23 degree 2 , where 2 is the
angle between the X-ray incident beam and the detector [9–11]. X-ray absorption
spectroscopic studies show that the silicates exhibit SiOM bonding and are not
simply mixtures of silica and metal oxides/hydroxides.

Some examples of iron silicate scales and their formation reactions are

[16.6]
[16.7]

[16.8]

Depending on the geothermal brine source, silica scaling is often exacerbated by the
presence of calcium, magnesium, iron, aluminum, and manganese. Aluminum-rich
and iron-rich amorphous silicates are the most common and exhibit significantl-
y lower solubilities than pure amorphous silica. Aluminum can begin to affect
amorphous silica solubility at low levels, on the order of only 1 ppm in the brine.
Aluminum-rich amorphous silica scale deposits at temperatures that are 25°C or
more below that of pure amorphous silica [12], so they can have a profound impact
on power cycle and silica scale treatment process design. Increasing a first- or
second-stage flash temperature by 25°C (a 4 bar increase from 150 to 175°C flash)
could make a geothermal power station economically nonviable.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like