Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

European Polymer Journal 114 (2019) 109–117

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

Preparation and performance of novel tetraphenylphosphonium- T


functionalized polyphosphazene membranes for alkaline fuel cells

Hongsa Han, Hongmei Ma, Jinghua Yu, Hong Zhu, Zhongming Wang
Faculty of Science, Beijing University of Chemical Technology, NO. 15 of North 3rd Ring East Road, Chaoyang District, 100029 Beijing, China

A R T I C LE I N FO A B S T R A C T

Keywords: A series of novel tetraphenylphosphonium cation-functionalized polyphosphazenes were successfully synthe-


Tetraphenylphosphonium salts sized and employed to prepare the corresponding anion exchange membranes. The typical properties of the
Anion exchange membrane prepared membranes for fuel cell including swelling ratio, water uptake, ion exchange capacity, tension test,
Polyphosphazene OH− conductivity and alkaline stability have been evaluated. PPMPP-3, one of the ionomer membranes with an
Alkaline stability
IEC value of 0.72 mmol g−1 displayed the hydroxide conductivities of 22.3 mS cm−1 at 80 °C, and showed
excellent alkaline stability that only 7.6% and 9.8% hydroxide conductivity degradation was observed after
being exposed to 1.0 M KOH at 60 °C and 80 °C for 10 days. The alkaline stability order of the AEMs studied in
this work was PPMPP-3 > PPMPP-4 > PPMPP-2 > PPMPP-1. This work provides a feasible method for en-
hancing the chemical stability of quaternary phosphonium salts which has a promising prospect to commercial
application of alkaline fuel cell.

1. Introduction benefitting from the absence of long range conjugation and low barrier
for free rotation around every repeating unit [19,20], which results in
Currently, anion exchange membrane fuel cells (AEMFCs) have at- excellent membrane-forming property. Secondly, the properties of the
tracted more and more attentions because they run under high-alkaline prepared polymers are able to be finely modified by carrying out sub-
conditions, which brings many advantages, e.g., satisfactory electrode stitution reactions on the matrix macromolecular poly(dichloropho-
kinetics and acceptable manufacturing cost on account of operating sphazene) (PDCP) [21]. The chlorine atoms contained in the polymer
with non-platinum group electrocatalysts. However, as one of the core are capable of being replaced with various sorts of organic functional
components in AEMFCs, anion exchange membranes (AEMs) still face groups, providing a simple method for optimizing the chemical and
some challenges such as low hydroxyl conductivity and poor alkaline physical performance of the polymers [22,23]. As for AEMs, this ad-
stability, which are related to the polymer backbones and functional vantage can lead to a high level of functionality per repeating unit and
groups [1–4]. Various polymers, including polysulfone, polystyrene, unlimited possibility for side-group functionalization. The other ad-
poly(phenylene), poly(ether ether ketone), poly(ether-imide), poly(aryl vantage can be attributed to the remarkable chemical and thermal
ether ketone) and poly(2,6-dimethyl-1,4-phenyleneoxide)(PPO), have stability of the polymer backbones [24]. Based on these advantages,
been explored as main chains to carry functional groups [5–14]. proton exchange membranes have already been prepared by carrying
However, these polymeric backbones may have some shortcomings acid groups on the polyphosphazene chain [25,26]. Recently, the ap-
such as high rigidity, low density of functional groups, low water plication of polyphosphazene has been further extended to AEMs by our
sorption, or poor chemical stability [2,15,16,17]. How to design AEMs group [27,28].
based on appropriate backbones is an urgent task that needs to be In typical AEMs, to improve the hydroxide ion conduction from
solved at the moment. cathode to anode, functional cation groups had better to tether cova-
Polyphosphazenes are inorganic-organic composite polymers with a lently on the polymer backbones to avoid the loss of cations. Generally,
linear backbone of alternating phosphorus and nitrogen atoms, each cation groups such as quaternary ammonium, guanidinium, imidazo-
phosphorus bearing two side organic substituents [18]. These polymers lium, benzimidazolium, quaternary phosphonium and metal-organic
have several advantages over the above polymers based on hydro- coordination ions have been employed to carry on the polymer chains
carbon. Firstly, the P]N polymer backbone is highly flexible [29–43]. However, under alkaline condition, these groups may be


Corresponding author.
E-mail address: wangzm@mail.buct.edu.cn (Z. Wang).

https://doi.org/10.1016/j.eurpolymj.2019.02.022
Received 26 October 2018; Received in revised form 23 January 2019; Accepted 18 February 2019
Available online 18 February 2019
0014-3057/ © 2019 Elsevier Ltd. All rights reserved.
H. Han, et al. European Polymer Journal 114 (2019) 109–117

easily degradable, which results in the poor durability of AEMs. For Fengtuo Chemical Technology Co., Ltd.) was recrystallized from dry
example, quaternary ammonium ions are unstable toward direct nu- petroleum ether before being used. 1,2,4-trichlorobenzene was dried by
cleophilic substitution (SN2) and/or β-hydrogen (Hofmann or E2) distillation under reduced pressure over NaH and stored with activated
elimination under alkaline conditions at high temperature [44–46]. For 3A molecular sieves at 120 °C. Triphenylphosphine was purchased from
guanidinium and imidazolium cations, though they have resonant Tianjin Guangfu Fine Chemical Research Institute. Dichloromethane
structure with delocalized charge which is considered to stabilize both (DCM) and nickel bromide were purchased from Beijing Coupling
cations, the durability of the corresponding AEMs were still not sa- Technology Co., Ltd. All other chemicals were purchased from
tisfactory according to the previous reports [47,48]. For guanidinium, Sinopharm Chemical Reagent Co., Ltd. and used without further pur-
low energy barriers make it easy for the degradation to react under high ification.
alkaline conditions. In most cases, the degraded products were urea
derivatives [47]. For common imidazolium, ring-opening reaction of 2.2. Synthesis of tetraphenylphosphonium cation and its derivatives
imidazole happens by the attack of hydroxyl ion at C2 position under
high pH conditions [48]. An efficient way to make imidazolium more A series of tetraphenylphosphonium cations were synthesized via
stable is to introduce substituents with different characteristics around nickel-catalyzed coupling reactions [54]. The procedure was given as
different positions of the imidazolium ring [49]. However, the alkaline follows. Triphenylphosphine (15 mmol, 1.0 eq.) were mixed with 4-
stability of AEMs is still a main challenge. bromophenol (16.5 mmol, 1.1 eq.), NiBr2 (0.75 mmol, 0.05 eq.) and
It is generally known that phosphonium salts favor the formation of ethylene glycol (5 mL) in a 100 mL flask and heated at 180 °C for 3 h
ylide upon interaction with strong alkalis by the Wittig-reaction me- under N2 atmosphere. After being cooled down to room temperature,
chanism, which is a usual strategy to synthesize alkenes. Nevertheless, the mixture was diluted with DCM (50 mL), and the organic layer was
the above process frequently results in the loss of the phosphonium washed using deionized water (2 × 5 mL). After being dried with an-
cation [50]. The stability of quaternary phosphoniums is poorer under hydrous Na2SO4 for 12 h, the organic mixture was filtered and con-
high pH situation than that of quaternary ammoniums because qua- centrated by rotary evaporation. The obtained solid was washed using
ternary phosphorus atoms are much more electron-withdrawing com- diethyl ether (2 × 20 mL) and the final product was obtained as white
pared to quaternarized nitrogen, which makes α-H in the adjacent alkyl solid.
groups more acidic and vulnerable to OH−. As a consequence, the at- (4-Hydroxyphenyl)triphenylphosphonium Bromide (TPHP-1): yield,
tack of hydroxyl can generate ylide intermediates (adjacent alkyl car- 91%. 1H NMR (400 MHz, DMSO‑d6) δ 11.38–11.05 (s, 1H), 8.01–7.91
banion and positive phosphonium) which have high reactivity [51]. (m, 3H), 7.86–7.76 (m, 6H), 7.67–7.75 (m,6H), 7.56–7.47 (m, 2H),
Therefore, it may be reasonable to consider a phosphonium cation 7.20–7.11 (m, 2H). 31P NMR (162 MHz, DMSO‑d6) δ 22.37, [M]+:
without α-H atoms. Additionally, the protection of phosphonium ca- 355.1246, found 355.1235.
tionic groups through the methods of increasing steric hindrance and (4-Hydroxyphenyl)tris(4-(methyl)phenyl) phosphonium Bromide
electron-donating effect are prospective ways to get stable [PR4]+ in (TPHP-2): yield, 85%. 1H NMR (400 MHz, CD3OD) δ 7.64–7.58 (m, 2H),
high pH environment: (i) effects of steric hindrance can make the cation 7.52–7.47 (m, 6H), 7.46–7.40 (m,6H), 7.28–7.21 (m, 2H), 2.55–2.52
sites sterically crowded thereby the OH− access becomes much more (m, 9H), 31P NMR (162 MHz, CD3OD) δ 21.72. [M]+: 397.1721, found
difficult, which improves the stability of phosphonium cation [16]; (ii) 397.1718.
substituent groups with electron-donating effect can increase the elec- (4-Hydroxyphenyl)tris(4-(methoxy)phenyl) phosphonium Bromide
tron density on the cation site, depressing the electrostatic attraction, (TPHP-3): yield, 78%; 1H NMR (400 MHz, DMSO‑d6) δ 11.30–10.97 (s,
which help stabilize the cations [41,51]. For example, Yan [52] re- 1H), δ7.57–7.52 (m, 2H), 7.50–7.43 (m, 6H), 7.26–7.22 (m,2H),
ported a type of AEMs prepared from tris(trimethoxyphenyl) phos- 7.20–7.11 (m, 6H), 3.97–3.93 (s, 9H). 31P NMR (162 MHz, DMSO‑d6) δ
phonium which exhibited good chemical stability and high ionic con- 20.60. [M]+: 445.1569, found 445.1555.
ductivity, reaching 27 mS cm−1 at 20 °C. Coates [53] reported a tetrakis (4-Hydroxyphenyl)tris(4-(tert-butyl)phenyl)phosphonium Bromide
(dialkylamino)phosphonium cation [P(N(Me)Cy)4]+ for AEMs. The (TPHP-4): yield, 59%. 1H NMR (400 MHz, DMSO‑d6) δ 11.32–11.03 (s,
obtained AEMs showed excellent alkaline stability, stable for 140 days 1H), 7.73–7.67 (m, 6H), 7.66–7.61 (m, 2H), 7.54–7.45 (m, 6H),
in 15 M KOH at 22 °C and 25 days in 1 M KOH at 80 °C, respectively. 7.28–7.23 (m, 2H), 1.42–1.36 (s, 27H), 31P NMR (162 MHz, DMSO‑d6)
The degradation of tetrakis(dialkylamino)phosphonium cations was δ 21.03. [M]+: 523.3130, found 523.3123.
also systematically explored in hydroxide or alkoxide anions after the Except the initial commercial material triphenylphosphine, some
stability test [40]. other materials like tris(4-methyl)phosphine, tris(4-methoxyphenyl)
Herein we reported a new class of hydroxide ion exchange mem- phosphane, tris(4-methoxyphenyl) phosphane and polyphosphazene
brane based on polyphosphazene chain carrying tetraphenylpho- were synthesized and characterized. The corresponding information is
sphonium cation. The structure of polyphosphazene ionomer has no α- given in Supporting Information.
H around the tetraphenylphosphonium cation. It is expected that the
Wittig degradation of phosphonium cations can be inhibited under the 2.3. Preparation of polyphosphazene-based membranes
alkaline environment. Additionally, the large steric effect of phenyl
groups is helpful to the stability of tetraphenylphosphonium cation. The The polyphosphazene-based membranes were prepared according
membranes casted from the polyphosphazene ionomer have been pre- to the following typical three-step procedure. Firstly, the potassium salt
pared and their performances for the application of fuel cells have been was prepared by mixing (4-hydroxyphenyl)triphenylphosphonium
studied. To consider the electronic effect of substituents, some groups bromide, ethyl alcohol and K2CO3 (1.2 eq.). After stirring for 1 h at
such as eH, eCH3, eCH3O, and eC(CH3)3 in tetraphenylphosphonium 25 °C, the mixture was filtered and concentrated by rotary evaporation,
cations have been investigated as well. then dissolved in DCM. After the insoluble substance was filtered out,
the organic product was concentrated again by rotary evaporation to
2. Experimental obtain the yellow pale potassium salt.
Secondly, the yellow pale potassium salt of quaternary phospho-
2.1. Materials nium bromide (0.4 eq. of repeating units in polyphosphazenes) was
added to the THF solution of poly(dichlorophosphazene) (PDCP) and
Tetrahydrofuran (THF) and petroleum ether were supplied by stirred for 6 h at 65 °C. Then the mixture was added to the THF solution
Beijing Chemical Works, refluxed with sodium for more than 48 h and (150 mL) containing 4-methyl-phenol sodium (1.6 eq. repeating unit in
distilled. Hexachlorocyclotriphosphazene (HCCP, Beijing Datian polyphosphazenes), and stirred for further 48 h at 25 °C in N2

110
H. Han, et al. European Polymer Journal 114 (2019) 109–117

atmosphere. The resulting solution was concentrated by rotary eva- solutions, VHCl and VNaOH stand for the volumes of the HCl and NaOH
poration to obtain a sticky ionomer mixture. After being washed with solutions, respectively. M represents the weight of the membranes.
deionized water (2 × 100 mL) and petroleum ether (2 × 100 mL), the
final product tetraphenylphosphonium-functionalized polypho-
sphazene was obtained as off-white elastomer. 2.4.5. Mechanical properties
Finally, to cast the membrane, the off-white elastomer was dissolved Tensile properties for PPMPPs membranes were determined with an
in THF, and the obtained solution (15 mL) was paved on a watch glass. Instron Model 1185 instrument (25 °C and 50 ± 5% relative hu-
After being dried at 25 °C for 48 h, the membrane in bromide form was midity). According to the ASTM standard D882-02 using a crosshead
obtained. To transfer the membranes into OH− form, the polymeric speed of 1 cm min−1, the PPMPPs with thickness of 90 μm were cut into
membrane was immersed with 1.0 M KOH aqueous solution under N2 dumbbell-shaped samples with the area of 5 cm × 0.4 cm. The tensile
atmosphere for 48 h. The membrane was then washed for several times strength at break (Tsb) and elongation at break (Eb) were measured to
using deionized water until the pH of lotion reached 6–7. assess the mechanical properties of PPMPPs.

2.4. Characterization
2.4.6. Ionic conductivity
The ionic conductivity of the membranes was determined by the
2.4.1. Structural characterization
1 two-probe AC impedance method with the frequency range of 1–105 Hz
H and 31P NMR nuclear magnet resonance spectra were conducted
from 30 °C to 80 °C using an electrochemical workstation (Zahner
on a Bruker Avance III 500 spectrometer. The 31P NMR chemical shifts
IM6ex, Germany). Before the ionic conductivity measurements,
were expressed as downfield from external 85% H3PO4. The attenuated
PPMPPs were immersed with deionized water for several times to re-
total reflection infrared (ATR-FTIR) spectra (4000–525 cm−1) were
move the residual KOH inside the membranes. The ionic conductivity σ
recorded on a Nexues 8700 ATR-FTIR spectrometer. The high-resolu-
was calculated by
tion mass spectra (HRMS) were measured on an Agilent Technologies
6200 series TOF/6500 series with electron spray ion (ESI). The mole-
σ (S·cm−1) = d/(RA) (4)
cular weight distribution was recorded by gel permeation chromato-
graphy (GPC, Agilent Technologies 1200 Series LC system) using DMF where d, R and A represent the length of the membranes between the
with 2 wt% LiCl as the eluent at a flow rate of 1.0 mL min−1 in nitrogen anode and cathode (cm), the electric resistance value (Ω) and the in-
atmosphere, calibrated with a standard of polystyrene. tersecting surface area of the membrane (cm2), respectively.

2.4.2. Thermal stability


A Mettler Toledo 1100SF thermogravimetric analyzer (Columbus, 2.4.7. Chemical stability
OH) was used for the thermogravimetric analysis (TGA) of the samples The alkaline stability of the membranes was investigated by soaking
to study the thermal decomposition behavior of the prepared mem- the membranes in a KOH solution (1.0 M) for 10 days at both 30 °C and
branes. The TGA measurements were conducted at temperatures from 60 °C under N2 atmosphere. To evaluate the stability of PPMPPs, the
40 to 800 °C at a heating rate of 10 °C min−1 with a nitrogen purge gas, ionic conductivity was recorded every 48 h.
and the samples were dried in vacuum at 130 °C for 12 h before mea-
surement.
3. Results and discussion
2.4.3. Water uptake and swelling ratio
The water uptake (WU) of the membranes was evaluated by the 3.1. Prepareation and characterization of tetraphenylphosphonium-
mass changes before and after being soaked in deionized water. The functionalized polyphosphazene membranes
membranes were swollen in deionized water under ambient condition
for 24 h and then taken out. After the surface water was thoroughly Herein a series of novel mixed-substituent PDCP derivatives
removed, the membranes were weighed immediately. The wet mem- (PPMPPs) were synthesized as shown in Scheme 1. After the prepara-
branes were dried in vacuum at 60 °C until a constant weight was ob- tion of PDCP [27], chloride atoms of PDCP in different percentages
tained. The WU was calculated as follow: were replaced by tetraphenylphosphonium cations, and the residual
chlorides in PDCP were replaced by 4-methylphenol anions. To ensure
WU(%) = (Mw − Md )/Md × 100% (1)
good mechanical and chemical stability for PPMPPs membranes, the
where Md and Mw represent the masses of dry and wet membranes. residual phosphorus-chloride bonds had to be consume away because
The swelling ratio (SR) was calculated as follow: they could cause crosslinking or instability in the polyphosphazene
backbones under alkaline condition [55]. After all membranes were
SR(%) = (Xw − X d)/X d × 100% (2)
prepared, the number-average molecular weight (Mn) and weight-
where Xw and Xd are the lengths of the wet and the dry membranes, average molecular weight (Mw) of PPMPPs were recorded by GPC:
respectively. PPMPP-1, Mn:121.8 kg mol−1, Mw:327.2 kg mol−1, Mw/Mn:2.68;
PPMPP-2, Mn:135.3 kg mol−1, Mw:331.2 kg mol−1, Mw/Mn:2.45;
2.4.4. Ion exchange capacity PPMPP-3, Mn:136.7 kg mol−1, Mw:352.6 kg mol−1, Mw/Mn: 2.57;
The ion exchange capacity (IEC) was calculated by applying the PPMPP-4 Mn: 153.5 kg mol−1, Mw:377.3 kg mol−1 Mw/Mn:2.46. Ac-
classical back-titration method. The membranes were first ion ex- cording to the GPC results, the molecular weight of all polymers
changed by immersing in a KOH solution (1.0 M) at 25 °C for 48 h and reached up to 105 order of magnitude, which met requirement of
then thoroughly washed using deionized water to remove OH− by membrane manufacture.
physical adsorption. After that the membranes were soaked in a diluted During the synthetic exploration, we tried to prepare some qua-
HCl solution (0.1 M, 50 mL) for 24 h. Finally, the HCl solution was ti- ternary phosphonium cations such as (4-hydroxyphenyl)tris(2,4,6-tri-
trated with standard NaOH solution (0.1 M). The IEC values were cal- methylphenyl)) phosphonium and (4-hydroxyphenyl)tris(2,4,6-tri-
culated as follow: methoxyphenyl)) phosphonium, which have larger steric hindrance and
stronger electron donating effect than those general tetraphenylpho-
IEC (mmolg−1) = (VHCl × CHCl − VNaOH × CNaOH)/M (3)
sphonium. It is a great pity that no desired products were founded be-
where CHCl and CNaOH are the concentrations of HCl and NaOH cause of various byproducts produced by the other side-reactions.

111
H. Han, et al. European Polymer Journal 114 (2019) 109–117

Scheme 1. Synthetic route of tetraphenylphosphonium-functionalized polyphosphazene polymers.

3.2. Structure characterization membrane was immersed in aqueous KOH, the residual -Cl would be
replaced by eOH [15]. The peaks at δ = −16.8 ppm and
To confirm the structure of PPMPP-1, PPMPP-2, PPMPP-3 and δ = −16.3 ppm are assigned to the chemical shifts of element P in HO-
PPMPP-4 membranes, the ATR-FTIR spectra are shown in Fig. 1. It is P-O-PhMe and HO-P-X forms. The single peak with the chemical shift
clear to observe that the absorption peaks around 3400 cm−1 are as- +22.6 ppm is assigned to the P+ of tetraphenylphosphonium, which
cribed to the stretching vibration of OH−, indicating the successful indicates the successful incorporation of charged TPHP-1 groups into
anion exchange of the ionomer membranes. The peak at 753 cm−1 is the PDCP.
attribute to P-C bond, and the absorptions at 831, 950 and 1203 cm−1 All the results from ATR-FTIR and 31P NMR provided the evidences
confirm the existence of P-O-C bonds, which indicates that some qua- of the successful introduction of the functional tetraphenylpho-
ternary phosphonium groups have carried on the polyphosphazene sphonium groups in to polyphosphazene-based ionomers.
backbones by phenol anion. The peak at 1250 cm−1 is ascribed to P]N
groups. In addition, adsorption peaks at 1396 and 1343 cm−1 in the 3.3. Thermal stability
spectra of PPMPP-4 correspond to the tertiarybutyl groups. Totally,
these ATR-FTIR spectra data support the formation of membranes. The thermal stability of AEMs is a major concern in relation to their
31
P NMR provided the structural information of the main chain and application in AEMFCs. The thermal stability of PPMPPs membranes
side groups of the ionomers. According to the 31P spectrum of PPMPP-1 and TPHP-1 was evaluated by TGA. As shown in Fig. 3, all the PPMPPs
(Fig. 2), the element P in the membranes exists in five kinds of chemical membranes exhibit a two-step weight-loss behavior: The initial de-
environment. The peaks at δ = −19.8 ppm and δ = −19.3 ppm are gradation from 30 to 120 °C is ascribed to the residual water and or-
attributed to the chemical shifts of the element P in P-(O-PhMe)2 and X- ganic solvent which are tightly bound inside the membranes after
P-O-PhMe (X = P+Ph4O−) groups, respectively. Considering the steric drying in vacuum. The second stage of weight losing appears at
hindrance of the tetraphenylphosphonium and 4-methylphenol, there 370–520 °C, which is related to the degradation of the polyphosphazene
were also some incomplete-substituent P in the backbones. After the ionomers. The degradation temperature of the TPHP-1 was around
400–500 °C, which almost coincided with the polymer backbones. All
the PPMPPs membranes exhibit excellent thermal stability below
370 °C. Totally, the thermal stability of membranes synthesized in this
work could fulfill the requirement of AEMFCs.

3.4. Mechanical properties of PPMPPs

AEMs should have sufficient mechanical strength if they are to be


used in AEMFCs. Generally, polyphosphazene membranes with suffi-
ciently high molecular weight possessed good mechanical properties.
Table 1 summarizes the mechanical properties of the PPMPPs. The
tensile strengths at break (Tsb), elongations at break (Eb) and Young’s
modulus of the membranes before stability test were in the range of
15.85–16.16 MPa, 7.6–9.1% and 0.34–0.48. After the PPMPPs were
soaked in 1 M KOH at 60 °C for 10 days, the Tsb values increased by
around 15%, and the Eb values decreased by a small margin, leading to
the increase of Young’s modulus. The result may relate to a slight de-
Fig. 1. ATR-FTIR spectra of PPMPPs. gree of crosslinking reactions occurred between some incomplete-

112
H. Han, et al. European Polymer Journal 114 (2019) 109–117

31
Fig. 2. P spectra of PPMPP-1.

Table 2
IEC, water uptake, and swelling ratio of membranes (t = 25 °C).
Samples Feed ratio* IEC (mmol g−1) WU (%) SR (%)

PPMPP-1–10 10% 0.27 ± 0.07 5.7 ± 1.2 2.2 ± 0.6


PPMPP-1–20 20% 0.50 ± 0.10 7.3 ± 0.7 3.1 ± 0.9
PPMPP-1–30 30% 0.73 ± 0.17 10.7 ± 2.6 6.0 ± 1.9
PPMPP-1 40% 0.75 ± 0.13 11.2 ± 1.7 6.1 ± 1.1
PPMPP-2 40% 0.73 ± 0.16 10.6 ± 2.8 5.9 ± 0.6
PPMPP-3 40% 0.72 ± 0.11 13.3 ± 0.8 6.8 ± 2.9
PPMPP-4 40% 0.75 ± 0.23 9.7 ± 3.1 5.7 ± 1.7

* Feed ratio: tetraphenylphosphonium equivalent of repeating unit in poly-


phosphazenes.

Fig. 3. TGA curves of PPMPPs and TPHP-1.

Table 1
Mechanical properties of the PPMPPs dry membranes before and after soaked in
1 M KOH at 60 °C for 10 days.
Membranes TSb (MPa) Eb (%) Young's modulus (GPa)

Before* After** Before* After** Before* After**

PPMPP-1 16.02 18.51 8.7 5.2 0.42 0.72


PPMPP-2 16.10 18.06 7.3 6.4 0.48 0.69
PPMPP-3 16.16 17.22 9.1 7.6 0.34 0.56
PPMPP-4 15.85 17.98 7.6 5.7 0.37 0.66

* Before: The membranes prepared in hydroxide form were dried in vacuum


Fig. 4. Hydroxide conductivity of PPMPPs as a function of temperature.
at 60 °C until a constant weight was attained.
** After: The membranes prepared in hydroxide form were soaked in 1 M
KOH at 60 °C for 10 days, then were dried in vacuum at 60 °C until a constant quantity of the OH− conductivity. In the current study, IEC values of
weight was attained. PPMPPs were regulated by changing the substitute proportion of tet-
raphenylphosphonium on the parent macromolecular, PDCP. Various
substitute P atoms in polyphosphazene [25]. The mechanical properties equivalent of TPHP-1 (10%, 20%, 30% and 40% eq. of repeating unit in
of the PPMPPs are still practicable to their application in AEMFCs. polyphosphazenes) were introduced to the reaction system. All the
products were obtained as off-white elastomers. As shown in Table 2,
3.5. IEC, water uptake, and swelling ratio of PPMPPs the tunable IEC values showed an increasing trend with the increasing
feed ratio of the charged TPHP-1. However, when the feed ratio reached
The IEC value of the membranes is generally regarded as a crucial to 40%, the IEC of PPMPP-1 were not further improved. Even the re-
influence factor for ion transfer, which acts as an indirect approximate action time was up to 10 h, the IEC values were still around 0.7–0.8.

113
H. Han, et al. European Polymer Journal 114 (2019) 109–117

lower than those of quaternary ammonium membranes because qua-


ternary phosphoniums have low hydrophilicity when compared to the
quaternary ammoniums in the similar structures [51]. In addition, low
water uptake of PPMPPs leads to low dimensional swelling. The water
swelling ratio of all membranes at 25 °C was less than 10%. Table 2 also
indicated the PPMPP-3 membranes had better hydrophilicity than that
of the other membranes. These results were due to the better hydro-
philicity of methoxyl derivatives TPHP-3.

3.6. Ionic conductivity of PPMPPs

The ionic conductivity of anion exchange membranes is a particu-


larly important intrinsic physical parameter. It is significantly influ-
enced by the IEC values. Fig. 4 demonstrates the ionic conductivities of
Fig. 5. Hydroxide conductivity of PPMPPs at 80 °C as a function of Hammett
constants. PPMPPs in the temperature range from 30 to 80 °C. From PPMPP-1-10
to PPMPP-1-30, ionic conductivities of the prepared membranes rise
gradually in conformity with the increase of IEC. The result is ex-
pectable as the concentration of active sites of anion transport increase.
Moreover, the ionic conductivity of PPMPPs at similar IEC values is also
evaluated and shown in Fig. 4 (from PPMPP-1 to PPMPP-4). The OH−
conductivities of obtained membranes reach 1.0 × 10−2 S cm−1 at
30 °C and increase to 2.0 × 10−2 S cm−1 at 80 °C, fulfilling the re-
quirement of AEMFCs. The results also revealed that the electron-do-
nating effect of the substituents in the functionalized tetraphenylpho-
sphonium cations is closely interrelated with OH− conductivity. The
PPMPP-3 exhibited the highest ionic conductivity with slightly low IEC
value. The result is ascribed to the strong electron-donating effect of
methoxyl groups, which makes the TPHP-3 possess higher electron
density on the cation site than that of the other three phosphonium
salts. The high electron density on the cation site greatly reduces the
electrostatic attraction between the positive and the negative ions,
Fig. 6. Arrhenius plots of PPMPPs.
which promotes the ionic mobility in the existence of a chemical or
electrochemical potential gradient [51]. Furthermore, the high WU% of
The real substitute ratio is about 25%∼30% of the repeating unite in PPMPP-3 is another positive effect in the ionic conductivity.
polyphosphazene (calculated according to the IEC values). When the The electron-donating property of substituent groups in the mole-
substitution reaction progressed to a certain degree, the large steric cule can also be assessed by Hammett constants [51]. The Hammett
hindrance of the tetraphenylphosphonium would make it difficult for constants order of the para-substituents studied in this work could be
the follow-up tetraphenylphosphonium substitution reactions to occur. eOCH3 > eC(CH3)3 > eCH3 > eH[56]. Fig. 5 present the resultant
Considering the experimental results, 40% feed ratio was regard as an ionic conductivity (at 80 °C) of PPMPPs as a function of corresponding
excess in the experiment. The same feed ratio was applied in the other Hammett constants. The result revealed that the electron-donating
PPMPPs. property and the ionic conductivity tend to correlate well. Electron-
Generally, the water uptake (WU) and swelling ratio (SR) of AEMs donating property determined by Hammett constants of the substitutes
are in correlation with IEC values and dimensional stability. According is an element that influences the ionic conductivity of the PPMPPs.
to Table 2, the increase in IEC leads to an increase in the hydrophilicity To study the effect mechanism of the obtained membranes,
of the membranes because water molecules trend to associate with the Arrhenius plots for PPMPPs were estimated and shown in Fig. 6. On the
charged functional groups. The highest WU value of all membranes is basis of the Arrhenius relationship between ionic conductivity and
around 10% at room temperature. The WU values of PPMPPs were working temperature, the activation energy for conduction (Ea) in

Fig. 7. Chemical stability of PPMPPs in 1.0 M KOH solution at 30 °C (a) and 60 °C (b).

114
H. Han, et al. European Polymer Journal 114 (2019) 109–117

Table 3
Durability and conductivity of PPMPP-3 membrane compared to that of tetraphenylphosphonium-functionalized membranes reported in previous reports.
Membranes Alkali resistance OH− conductivity (mS cm−1) IEC (mmol g−1) Ref.

Test conditions Degradability

AEM8 unknown unknown 20.6 ± 3.1(90 °C) 0.45 [58]


[P(NR2)4]+ 2 M KOH/CH3OH, 80 °C unknown unknown unknown [40]
[P(N(Me)Cy)4]+ 1 M KOH, 80 °C, 140 d unknown 22 ± 1 (22 °C) unknown [53]
CQA-15 10 M NaOH, 80 °C, 200 h 60% 15.8 (20 °C) 1.00 [41]
PEEK-QPOH 1 M KOH, 60 °C, 48 h unknown 61 (20 °C) 0.89 [39]
PPMPP-3 1 M KOH, 60 °C, 10 d 7.6% 22.3 (80 °C) 0.72 ± 0.11 This work

31
Fig. 8. P spectra of TPHP-1 after immersed in 1.0 M KOH for 20 days at 30 °C.

Fig. 7 presents the resultant ionic conductivity ratios as a function of


alkaline ageing time.
Previous literature about the alkaline stability of benzyl quaternary
phosphonium demonstrated that AEMs based on benzyl quaternary
phosphonium tend to lose a significant percentage of functional groups
with the similar treatment at room temperature. This was ascribed to
the OH− attack on the labile α-H at the benzyl position [51]. In this
study, the benzyl group was replaced by phenyl group to weaken OH−
attack. It is observed that the ion conductivity of all PPMPPs did not
change significantly after immersion in the aqueous KOH (1.0 M) so-
Scheme 2. Decomposition mechanism of tetraphenylphosphonium in KOH lution at 30 °C even after 10 days of ageing (Fig. 7a). This result sug-
solution. gests that the absence of α-H at the benzyl position has enhanced al-
kaline stabilities of P+ and demonstrated the positive effects in
PPMPPs were measured respectively. The electron-donating effect is PPMPPs.
considered to increase charge density and availably restrain the elec- However, ion conductivity of PPMPP-1 decreased obviously when
trostatic attraction, which could decrease the Ea values of the PPMPPs. the test temperature increased to 60 °C (Fig. 7b). The ion conductivity
The result suggests that the PPMPP-3 membranes have the lowest Ea declined by 21.8% after 10 days ageing. When electron-donating effect
value (14.54 kJ mol−1), and the none-substituded tetraphenylpho- were introduced to weaken OH− attack, it was observed that the ionic
sphonium possessed the highest Ea value (15.06 kJ mol−1), respec- conductivities of PPMPP-2, PPMPP-3, PPMPP-4 were not changed as
tively. serious as that of PPMPP-1. Base on the electron-donating effect of the
substituent groups, the alkaline stability of PPMPP-2 was better than
PPMPP-1. Specifically, 16.8% hydroxide conductivity degradation in
3.7. Alkaline stability PPMPP-2 was observed after the testing period. Surprisingly, only 7.6%
hydroxide conductivity degradation in PPMPP-3 and 9.8% hydroxide
High alkaline stability is one of the most desirable properties for conductivity degradation in PPMPP-4 were observed. Moreover, when
AEMs due to the high pH environments in operating AEMFCs. The al- increasing the temperature to 80 °C, the ionic conductivity degradation
kaline stability of the prepared PPMPPs were investigated by immersing of PPMPP-3 and PPMPP-4 were found to be 9.8% and 12.6% after
the membranes in an aqueous KOH (1.0 M) solution at 30 °C and 60 °C 10 days, respectively. The alkaline stability of these membranes with
for 10 days. The ionic conductivity was recorded every 48 h at 80 °C. the different substituents indicated the electron-donating effect of p-

115
H. Han, et al. European Polymer Journal 114 (2019) 109–117

31
Fig. 9. P spectra of TPHP-1 after immersed in 1.0 M KOH for 5, 10 and 15 days at 30 °C.

methylphenyl, p-methoxyphenyl and p-tertbutylphenyl could enhance correlated well with the hydroxyl conductivity and the alkaline stability
the alkaline stability of quaternary phosphonium compounds. The sta- of the cations. In addition, the prepared PPMPPs exhibited excellent
bilizing effect were consistent with the electron-donating effect. alkaline stability and satisfactory ionic conductivity in high pH en-
Therefore, the alkaline stability order of the PPMPPs studied in this vironment under elevated temperatures. Hence, it is not difficult to see
work could be determined to be PPMPP-3 > PPMPP-4 > PPMPP- that minimizing the electrostatic attraction by increasing electron
2 > PPMPP-1. density of the cation site is pivotal for enhancing OH− conductivity and
Durability, conductivity and IEC values of various tetraphenylpho- alkaline stability of AEMs. These results indicate a possible and facile
sphonium-functionalized membranes reported in literature were sum- strategy for improving the alkaline stability and OH− conductivity of
marized in Table 3. As illustrated in Table 3, the proposed membrane phosphonium cations and its promising application in commercial al-
exhibited excellent alkali stability as well as satisfactory OH− con- kaline fuel cell.
ductivity and IEC values as the previously developed AEMs.
To further investigate the decomposition process of tetra- Acknowledgment
phenylphosphonium, TPHP-1 was taken as a model. The decomposition
mechanism of TPHP-1 was discussed. In our research, after TPHP-1 was The project is supported by the National Natural Science Foundation
immersed in 1.0 M KOH for 20 days at 30 °C, the 31P spectra of TPHP-1 of China (Grant Nos. 21776012 and 21276021).
and decomposition products were given in Fig. 8. The results were in
accordance with the reported in previous papers: The phosphonium NOTES
salts are decomposed by aqueous alkali to give tertiary phosphine
oxides and hydrocarbons [57]. The reported decomposition mechanism The authors declare no competing financial interest.
was shown in Scheme 2. Though α-H was not existed in tetra-
phenylphosphonium, the P-C bones between P cations and phenyl Appendix A. Supplementary material
would break up leading to tertiary phosphine oxides. The peaks at
δ = +28.6 ppm and δ = +29.7 ppm were assigned to the decomposi- Supplementary data to this article can be found online at https://
tion products of TPHP-1 as shown in Fig. 8. Furthermore, the 31P NMR doi.org/10.1016/j.eurpolymj.2019.02.022.
spectra of TPHP-1 at different time of alkaline treatment was recorded.
According to Fig. 9, the phosphonium salt started to degrade from the References
day of 15. The 31P spectra of TPHP-2, TPHP-3 and TPHP-4 at the end of
the alkaline test were also investigated (Fig. S2), and the results de- [1] J.R. Varcoe, P. Atanassov, D.R. Dekel, A.M. Herring, M.A. Hickner, P.A. Kohl,
monstrated the best stability of TPHP-3. A.R. Kucernak, W.E. Mustain, K. Nijmeijer, K. Scott, T.W. Xu, L. Zhuang, Anion-
exchange membranes in electrochemical energy systems, Energy Environ. Sci. 7
(2014) 3135–3191.
[2] S. Maurya, S.H. Shin, Y.K. Kim, S.H. Moon, A review on recent developments of
4. Conclusion anion exchange membranes for fuel cells and redox flow batteries, RSC Adv. 5
(2015) 37206–37230.
[3] J.R. Varcoe, R.C.T. Slade, Prospects for alkaline anion-exchange membranes in low
To sum up, a series of novel tetraphenylphosphonium salts were
temperature fule cells, Fuel Cells 5 (2005) 187–200.
designed, synthesized and employed in this study to prepare the cor- [4] Z.F. Pan, L. An, T.S. Zhao, Z.K. Tang, Advances and challenges in alkaline anion
responding anion exchange membranes. Compared with benzyl qua- exchange membrane fuel cells, Prog. Energy Combust. Sci. 66 (2018) 141–175.
ternary phosphonium, the alkaline stability of tetraphenylphosphonium [5] M.L.D. Vona, R. Narducci, L. Pasquini, K. Pelzer, P. Knauth, Anion-conducting io-
nomers: study of type of functionalizing amine and macromolecular cross-linking,
salts is significantly enhanced, due to the absence of the labile α-H in Int. J. Hydrogen Energy 26 (2014) 14039–14049.
benzyl group. Furthermore, the results also revealed that the Hammett [6] X.L. Zhang, C.C. Fan, N.Y. Yao, P. Zhang, T. Hong, C.X. Xu, J.G. Cheng, Quaternary
constants of the substituent groups (eCH3, eOCH3, eC(CH3)3) Ti3C2Tx enhanced ionic conduction in quaternized polysulfone membrane for

116
H. Han, et al. European Polymer Journal 114 (2019) 109–117

alkaline anion exchange membrane fuel cells, J. Membr. Sci. 563 (2018) 882–887. Chem. Mater. 27 (2015) 6689–6698.
[7] X.L. Duan, C.W. Wang, T.L. Wang, X.L. Xie, X.P. Zhou, Y.S. Ye, A polysulfone-based [33] D.S. Kim, C.H. Fujimoto, M.R. Hibbs, A. Labouriau, Y.K. Choe, Y.S. Kim, Resonance
anion exchange membrane for phosphoric acid concentration and purification by stabilized perfluorinated ionomers for alkaline membrane fuel cells,
electro-electrodialysis, J. Membr. Sci. 552 (2018) 86–94. Macromolecules 46 (2013) 7826–7833.
[8] S. Vengatesan, S. Santhi, G. Sozhan, S. Ravichandran, D.J. Davidson, S. Vasudevan, [34] D.S. Kim, A. Labouriau, M.D. Guiver, Y.S. Kim, Guanidinium-functionalized anion
Novel cross-linked anion exchange membrane based on hexaminium functionalized exchange polymer electrolytes via activated fluorophenyl-amine reaction, Chem.
poly(vinylbenzyl chloride), RSC Adv. 5 (2015) 27365–27371. Mater. 23 (2011) 3795–3797.
[9] M.R. Hibbs, Alkaline stability of poly(phenylene)-based anion exchange membranes [35] L. Liu, Q. Li, J. Dai, H. Wang, B. Jin, R. Bai, A facile strategy for the synthesis of
with various cations, J. Polym. Sci. Part B: Polym. Phys. 51 (2013) 1736–1742. guanidinium-functionalized polymer as alkaline anion exchange membrane with
[10] X. Yan, S. Gu, G. He, X. Wu, J. Benziger, Imidazolium-functionalized poly(ether improved alkaline stability, J. Membr. Sci. 453 (2014) 52–60.
ether ketone) as membrane and electrode ionomer for low-temperature alkaline [36] A.N. Lai, L.S. Wang, C.X. Lin, Y.Z. Zhuo, Q.G. Zhang, A.M. Zhu, Q.L. Liu,
membrane direct methanol fuel cell, J. Power Sources. 250 (2014) 90–97. Phenolphthalein-based poly(arylene ether sulfone nitrile)s multiblock copolymers
[11] G. Wang, Y. Weng, D. Chu, D. Xie, R. Chen, Preparation of alkaline anion exchange as anion exchange membranes for alkaline fuel cells, ACS Appl. Mat. interfaces. 7
membranes based on functional poly(ether-imide) polymers for potential fuel cell (2015) 8284–8292.
applications, J. Membr. Sci. 326 (1) (2009) 4–8. [37] Y. Yang, J. Wang, J. Zheng, S. Li, S. Zhang, A stable anion exchange membrane
[12] Y.X. Xu, J.S. Yang, N.Y. Ye, M. Teng, R.H. He, Modification of poly(aryl ether ke- based on imidazolium salt for alkaline fuel cell, J. Membr. Sci. 467 (2014) 48–55.
tone) using imidazolium groups as both pendants and bridging joints for anion [38] O.D. Thomas, K.J. Soo, T.J. Peckham, M.P. Kulkarni, S. Holdcroft, A stable hy-
exchange membranes, Eur. Polym. J. 73 (2015) 116–126. droxide-conducting polymer, J. Am. Chem. Soc. 134 (2012) 10753–10756.
[13] H.S. Dang, P. Jannasch, A comparative study of anion-exchange membranes teth- [39] X. Yan, S. Gu, G. He, X. Wu, W. Zheng, X. Ruan, Quaternary phosphonium-func-
ered with different heterocycloaliphatic quaternary ammonium hydroxides, J. tionalized poly(ether ether ketone) as highly conductive and alkali-stable hydroxide
Mater. Chem. A 5 (2017) 21965–21978. exchange membrane for fuel cells, J. Membr. Sci. 466 (2014) 220–228.
[14] X.J. Wang, P. Wang, Y.Y. Sun, J.L. Wang, H.G. Fang, S.Z. Yang, H.B. Wei, Y.S. Ding, [40] C.T. Womble, J. Kang, K.M. Hugar, G.W. Coates, S. Bernhard, K.J.T. Noonan, Rapid
A mechanically strong and tough anion exchange membrane engineered with non- analysis of tetrakis(dialkylamino)phosphonium stability in alkaline media,
covalent modalities, Chem. Commun. 53 (2017) 12369–12372. Organometallics 36 (2017) 4038–4046.
[15] C.G. Arges, V. Ramani, Two-dimensional NMR spectroscopy reveals cation-trig- [41] H.Y. Tang, D.F. Li, N.W. Li, Z.S. Zhang, Z.B. Zhang, Anion conductive poly(2,6-
gered backbone degradation in polysulfone-based anion exchange membranes, dimethyl phenylene oxide)s with clicked bulky quaternary phosphonium groups, J.
Proc. Natl Acad. Sci. 110 (2013) 2490–2495. Membr. Sci. 558 (2018) 9–16.
[16] J. Cheng, G. He, F. Zhang, A mini-review on anion exchange membranes for fuel cell [42] B.Z. Zhang, H. Long, R.B. Kaspar, J.H. Wang, S. Gu, Z.B. Zhuang, B. Pivovar,
applications: stability issue and addressing strategies, Int. J. Hydrogen Energy. 23 Y.S. Yan, Relating alkaline stability to the structure of quaternary phosphonium
(2015) 7348–7360. cations, RSC Adv. 8 (2018) 26640–26645.
[17] S.S. Gottesfeld, D.R. Dekel, M. Page, C.S. Bae, Y.S. Yan, P. Zelenay, Y.S. Kim, Anion [43] Y. Zha, M.L. Disabb-Miller, Z.D. Johnson, M.A. Hickner, G.N. Tew, Metal-cation-
exchange membrane fuel cells: current status and remaining challenges, J. Power based anion exchange membranes, J. Am. Chem. Soc. 134 (2012) 4493–4496.
Sources 375 (2018) 170–184. [44] J.B. Edson, C.S. Macomber, B.S. Pivovar, J.M. Boncella, Hydroxide based decom-
[18] X. Liu, J.P. Breon, C. Chen, H.R. Allcock, Substituent exchange reactions with high position pathways of alkyltrimethylammonium cations, J. Membr. Sci. 399 (2012)
polymeric organophosphazenes, Macromolecules 45 (2012) 9100–9109. 49–59.
[19] H.R. Allcock, Generation of structural diversity in polyphosphazenes, Appl. [45] H. Long, K. Kim, B.S. Pivovar, Hydroxide degradation pathways for substituted
Organomet. Chem. 27 (2013) 620–629. trimethylammonium cations: a DFT study, J. Phys. Chem. C 116 (2012) 9419–9426.
[20] R. Gutru, S.G. Peera, S.D. Bhat, A.K. Sahu, Synthesis of sulfonated poly(bis(phe- [46] K.F.L. Hagesteijn, S.X. Jiang, B.P. Ladewig, A review of the synthesis and char-
noxy)phosphazene) based blend membranes and its effect as electrolyte in fuel cells, acterization of anion exchange membranes, J. Mater. Sci. 53 (2018) 11131–11150.
Solid State Ion. 296 (2016) 127–136. [47] W. Li, S. Wang, X. Zhang, W. Wang, X. Xie, P. Pei, Degradation of guanidinium-
[21] C.Y. Tong, Z.C. Tian, C. Chen, Z.J. Li, T. Modzelewski, H.R. Allcock, Synthesis and functionalized anion exchange membrane during alkaline environment, Int. J.
characterization of trifluoroethoxy polyphosphazenes containing polyhedral oligo- Hydrogen Energy 39 (2014) 13710–13717.
meric silsesquioxane (POSS) side groups, Macromolecules 49 (2016) 1313–1320. [48] S.C. Price, K.S. Williams, F.L. Beyer, Relationships between structure and alkaline
[22] N.R. Krogman, L. Steely, M.D. Hindenlang, L.S. Nair, C.T. Laurencin, H.R. Allcock, stability of imidazolium cations for fuel cell membrane applications, ACS Macro
Synthesis and characterization of polyphosphazene-block-polyester and polypho- Lett. 3 (2014) 160–165.
sphazene-block-polycarbonate macromolecules, Macromolecules 41 (2008) [49] K.M. Hugar, H.A. Kostalik, G.W. Coates, Imidazolium cations with exceptional al-
1126–1130. kaline stability: a systematic study of structure−stability relationships, J. Am.
[23] R.S. Ullah, L. Wang, H.J. Yu, N.M. Abbasi, M. Akram, Z. Abdin, M. Saleem, Chem. Soc. 137 (2015) 8730–8737.
M. Haroon, R.U. Khan, Synthesis of polyphosphazenes with different side groups [50] C.J. Bradaric, A. Downard, C. Kennedy, A.J. Robertson, Y.H. Zhou, Industrial pre-
and various tactics for drug delivery, RSC Adv. 7 (2017) 23363–23391. paration of phosphonium ionic liquids, Green Chem. 5 (2003) 143–152.
[24] H.R. Allcock, R.M. Wood, Design and synthesis of ion-conductive polyphosphazenes [51] C.G. Arges, J. Parrondo, G. Johnson, A. Nadhan, V. Ramani, Assessing the influence
for fuel cell applications: review, J. Polym. Sci., Part B: Polym. Phys. 44 (2006) of different cation chemistries on ionic conductivity and alkaline stability of anion
2358–2368. exchange membranes, J. Mater. Chem. 22 (2012) 3733–3744.
[25] R. Wycisk, P.N. Pintauro, Polyphosphazene membranes for fuel cells, Adv. Polym. [52] S. Gu, R. Cai, T. Luo, Z. Chen, M. Sun, Y. Liu, G.H. He, Y.S. Yan, A soluble and highly
Sci. 216 (2008) 157–183. conductive ionomer for high-performance hydroxide exchange membrane fuel cells,
[26] S.T. Gao, H.L. Xu, T.W. Luo, Y.F. Guo, Z. Li, A. Ouadaha, Y.X. Zhang, Z.Y. Zhang, Angew. Chem. 48 (2009) 6499–6502.
C.J. Zhu, Novel proton conducting membranes based on cross-linked sulfonated [53] K.J. Noonan, K.M. Hugar, H. Kostalik, E.B. Lobkovsky, H.D. Abruna, G.W. Coates,
polyphosphazenes and poly(ether ether ketone), J. Membr. Sci. 536 (2017) 1–10. Phosphonium-functionalized polyethylene: a new class of base-stable alkaline anion
[27] J. Fan, H. Zhu, R. Li, N. Chen, K. Han, Layered double hydroxide–polyphosphazene- exchange membranes, J. Am. Chem. Soc. 134 (2012) 18161–18164.
based ionomer hybrid membranes with electric field-aligned domains for hydroxide [54] D. Marcoux, A.B. Charette, Nickel-catalyzed synthesis of phosphonium salts from
transport, J. Mater. Chem. A. 2 (2014) 8376–8385. aryl halides and triphenylphosphine, Adv. Synth. Catal. 350 (2008) 2967–2974.
[28] Y.N. Chen, Z.M. Li, N.J. Chen, Y.J. Zhang, F.H. Wang, Hong Zhu, Preparation and [55] H.R. Allcock, M.B. Mcintosh, E.H. Klingenberg, M.E. Napierala, Functionalized
characterization of cross-linked polyphosphazene-crown ether membranes for al- polyphosphazenes: polymers with pendent tertiary trialkylamino groups,
kaline fuel cells, Electrochim. Acta 258 (2017) 311–321. Macromolecules. 31 (1998) 5255–5263.
[29] W.H. Lee, Y.S. Kim, C. Bae, Robust hydroxide ion conducting poly(biphenyl alky- [56] L.P. Hammett, The effect of structure upon the reactions of organic compounds.
lene)s for alkaline fuel cell membranes, ACS Macro Letters. 4 (2015) 814–818. Benzene derivatives, J. Am. Chem. Soc. 59 (1936) 96–103.
[30] D.R. Dekel, M. Amar, S. Willdorf, M. Kosa, S. Dhara, C.E. Diesendruck, Effect of [57] M. Grayson, P.T. Keough, Phosphonium compounds. II. decomposition of phos-
water on the stability of quaternary ammonium groups for anion exchange mem- phonium alkoxides to hydrocarbon, ether and phosphine oxide, J. Am. Chem. Soc.
brane fuel cell applications, Chem. Mater. 29 (2017) 4425–4431. 82 (1960) 3919–3924.
[31] Y. Yang, D.M. Knauss, Poly(2,6-dimethyl-1,4-phenylene oxide)-b-poly(vi- [58] W.X. Zhang, Y. Liu, A.C. Jackson, A.M. Savage, S.P. Ertem, T.H. Tsai, S. Seifert,
nylbenzyltrimethylammonium) diblock copolymers for highly conductive anion F.L. Beyer, M.W. Liberatore, A.M. Herring, E.B. Coughlin, Achieving continuous
Exchange membranes, Macromolecules 48 (2015) 4471–4480. anion transport domains using block copolymers containing phosphonium cations,
[32] J. Pan, L. Zhu, J. Han, M.A. Hickner, Mechanically tough and chemically stable Macromolecules 49 (2016) 4714–4722.
anion exchange membranes from rigid-flexible semi-interpenetrating networks,

117

You might also like