You are on page 1of 38

Phosphazenes

Amit Chakraborty,a,b Naushad Ahmeda and


Vadapalli Chandrasekhar*a,c
DOI: 10.1039/9781839163814-00429

Phosphazenes are a family of compounds containing the PQN motif.


Although the nature of the bonding in the PQN motif is controversial and
might not involve a formal double bond in the manner that is understood in
organic compounds, it has been a general practice in the literature to retain
this motif while describing these systems. Phosphazenes comprise of acyclic
compounds, more commonly known as iminophosphoranes, cyclic com-
pounds such as cyclophosphazenes, and polymeric systems which include
polyphosphazenes that contain a PQN backbone and polymers that are
various hybrids typically comprising of a cyclophosphazene platform that is
incorporated in a macromolecular framework. Accordingly, this chapter on
phosphazenes contains three sections: a) acyclic phosphazenes, b) cyclo-
phosphazenes and c) polyphosphazenes and hybrid polymers. The various
applications of these systems will be discussed in these sections themselves.
This chapter covers material published in the calendar year 2019.

a) Acyclic phosphazenes
Acyclic phosphazenes, more commonly known as iminophosphoranes
are of interest from various points of view including the following:

1) They are good ligands for binding to main-group-, transition- and


lanthanide metal ions
2) Some of them are excellent Lewis bases and therefore are of interest
from their basicity as well as utility as organocatalysts.
3) Some of them serve as useful intermediates in organic transfor-
mations such as the aza-Wittig reaction.

While the most common way to prepare iminophosphoranes con-


tinues to be the Staudinger reaction involving the reaction of an azide
with R3P, new methods of synthesis are also becoming available. Thus,
imination of phosphines by reaction of (N-tosylimino)-phenyl-3-iodane
(PhINTs) was shown to occur readily by a radical pathway affording the
iminophosphorane, 1.1
Compounds 2–3, possess potent inhibitory activity against the enzyme
acetylcholinesterase (AChE) as well as b-amyloid aggregation.2
a
Tata Institute of Fundamental Research Hyderabad, Gopanpally, Hyderabad
500046, India. E-mail: vc@iitk.ac.in
b
Department of Chemistry, M. M. Engineering College, Maharishi Markandeshwar
(Deemed to be University), Mullana, Ambala 133207, India
c
Department of Chemistry, Indian Institute of Technology Kanpur, Kanpur 208016,
India

Organophosphorus Chem., 2021, 50, 429–466 | 429



c The Royal Society of Chemistry 2021
Compounds 2a and 2b were synthesized by a redox reaction which
involved the condensation of 1,2-bis(diphenylphosphino)ethane with
different amines in the presence of dialkyl azodicarboxylate. On the other
hand, compound 3 could be prepared by a standard Staudinger reaction.
Both of these families of compounds, in nanomolar concentrations,
showed excellent inhibition activity towards AChE and b-amyloid
aggregation.2
In an extremely interesting transformation involving a triple hydrogen
abstraction from a coordinated ammonia ligand, the reaction of
[Mn(depe)2(CO)(NH3)]1 4 (depe ¼ 1,2-bis(diethylphosphino)ethane), with
2,4,6-tri-tert-butylphenoxyl radical was found to afford the five-membered
phosphazenium cation, [(Et2P(C2H4)PEt2)N]1 (4a).3 The authors suggest
that reactions of this type can be utilized to understand the nature of
N–H bond-breaking processes which would allow for the development of
efficient catalysts to harness the potential chemical energy stored in such
bonds.

As mentioned above, iminophosphoranes have been utilized as ligands


to bind to various types of metal ions. In this context a family of potentially
tridentate ligands, 5, 2-C8H6N(Ph2PQNC6H4XR) (6) [X ¼ S, R ¼ CH3 (a),
C6H5 (b); X ¼ Se, R ¼ CH3 (c), C6H5 (d)] were prepared. These ligands
contain an indolyl moiety attached to the phosphorus center while the
substituent on the nitrogen center contains a thioether or selenoether
moiety. Reaction of 5 with [PdCl2(CH3CN)2] afforded non-symmetrical
N,N,X-Pd(II) pincer complexes [PdCl{2-C8H5N(Ph2PQNC6H4XR-k3-N,N,X)}]
(6) [X ¼ S, R ¼ CH3 (a), C6H5 (b); X ¼ Se, R ¼ CH3 (c), C6H5 (d)]. Reaction of
6a with the bidentate chelating ligand dppe, [Ph2P(CH2)2PPh2], in a 1 : 1
ratio displaces the sulfur donor atom from the coordination sphere of
Pd(II) and affords the mononuclear cationic complex 7. In the latter, the
Pd(II) is bound, interestingly, by two bidentate, chelating (N,N and P,P)
ligands.4

430 | Organophosphorus Chem., 2021, 50, 429–466


Multi-functional ligands, supported on a carborane framework, 8 and 9,
were prepared. These ligands, in addition the iminophosphorane moiety,
also contain a phosphine group. In these ligands the iminophosphorane
nitrogen is connected to a cage carbon atom (C-carboranyl derivatives
8a–8c) or to the B3 boron atom (B-carboranyl derivatives 9d–9e). These were
used to obtain palladium complexes (10–12). It is to be noted that both an
intramolecular chelating coordination mode (P^N) (complexes 10 and 12)
as well as an intermolecular coordination (P) (complex 11) were found.5

A tris(piperidinyl)imidophosphorane ligand, [NP(pip)3], (13) capable


of simultaneous 1s and 2p donation was used to stabilize Ce(IV)
and Ce(III) complexes 13a and 13b. Interestingly the electrochemical
Ce(III)–Ce(IV) couple is among the most negative couples (2.64 V with
respect to Fc/Fc1) known in the literature.6

The aza-Wittig reaction is a useful synthetic method for various organic


transformations. In this reaction, an amine group is converted to an
iminophosphorane unit which can then further react with compounds
containing various functional groups. Thus, the synthesis of 5-methyl-4-

Organophosphorus Chem., 2021, 50, 429–466 | 431


methylene-7-methylthio-2-arylmino-4H-pyrido[4,3-d][1,3]oxazine-8-carbo-
nitrile derivatives (14a) occurred in the presence of [N-{3-acetyl-5-cyano-2-
methyl-6-(methylthio)pyridin-4-yl}imino]triphenylphosphorane (14) and
aryl isocyanate. This reaction is an example of the aza-Wittig reaction.
Several other examples are discussed below.7

A Ag(I)-catalyzed reaction of N-isocyaniminotriphenylphosphorane (15)


and aldehydes afforded various azines (16).8 This reaction occurs through
a tandem reaction sequence involving aza-Wittig, insertion and cycliza-
tion. Interestingly, the iminophosphorane, 15, can be used in [3 þ 2]
cycloaddition reactions with alkynes, catalyzed by silver catalysts to afford
various pyrazole derivatives.9 In another example, aminocyanation of
alkynes occurred by reaction of 15, again promoted by a silver catalyst.10
Other examples of using iminophosphoranes in the aza-Wittig reaction
include the reaction of 17 with n-butyl isocyanate to give carbodiimides.11
Similarly, aza-Wittig condensation of the amido group with the imino-
phosphorane 18 led to the formation of a dialkylamino-substituted
dibenzodiazepine (for example 19) and their hetero analogues.12 An
unusual aza-Wittig reaction occurred for the iminophosphorane 20b
synthesized from nitriles such as g-azidobutyronitrile 20a affording the
pyrrole-appended iminophosphazenes, 21.13
Due to its strong basicity the presence of an iminophosphorane unit in
a molecule such as 22 facilitates a self-catalyzed Michael addition reac-
tion affording 23 followed by an intramolecular aza-Wittig reaction to
form a larger ring, 24.14

432 | Organophosphorus Chem., 2021, 50, 429–466


Several thieno[2,3-d]pyrimidine derivatives such as 26, which are
potential anticancer agents, were synthesized from the functionalized
iminophosphorane (25) via an aza-Wittig reaction.15
Iminophosphoranes with appropriate substituents function as strong
bases and are commonly referred to in the literature as phosphazene bases.

Theoretical calculations have shown that the compounds 27–36 should


possess excellent basicity.16 Similarly, computational studies also showed
that the three-membered cyclic iminophosphoranes, 37–39, act as strong
bases due to the possibility of the delocalization of the positive charge
upon protonation.17

Phosphazenyl phosphines, built around a P(III) center, (40–41), are a


new family of potential bases which contrast the N-centred bases. Com-
pounds 40–41 represent the protonated forms and were synthesized by
homologization of iminophosphoranes.18

Organophosphorus Chem., 2021, 50, 429–466 | 433


Several studies on the reactivity of the phosphazene super bases have
been found to be interesting. Thus, the bases 42–44 were found to
abstract a proton from water to form a hydroxide ion which was solvated
by water to form the anionic species, [OH(OH2)3] along with the cor-
responding cations 42a–44a. Interestingly, a crystal structure analysis
revealed an absence of contact between the anion and the cation. The
phosphazene base could be regenerated from the salt by thermolysis
under vacuum. An interesting application of these phosphazene bases
was in the preparation of Me3SiCF3 using HCF3.19

Several strategies are in vogue to increase the basicities of the phos-


phazene bases. One of these involves the use of intramolecular hydrogen
bonding. Using this design concept, N,N 0 ,N00 ,N 0 0 0 -tetrakis(3-(dimethyla-
mino)propyl)triaminophosphazene (TDMPP, 45) was synthesized and its
basicity investigated. These studies revealed that the intramolecular
hydrogen bonding present in 45 reinforced the basicity and proton
affinity of the phosphazene superbase.20

434 | Organophosphorus Chem., 2021, 50, 429–466


The use of phosphazene bases in various organic transformations
continues to be an active field of research. Among the various bases
studied, the phosphazene super base tBuP4 continues to be widely
used.21,22 In addition to the well-known tBuP4 many other phosphazene
bases have been designed and used. Their strong basicity and low nu-
cleophilic tendency make the phosphazene superbases valuable reagents
widely used as deprotonating agents. The phosphazene base, tBuP4, 46,
has been found to function in organocatalysis reactions involving
replacement of the methoxy group from a variety of substituted anisoles
with other alkoxy groups.21 Similarly, 46 was found to be a facile catalyst
for the iodoaminocyclization of 2-(1-alkynyl)benzamides.23 Other reac-
tions where the use of 46 was reported include the organocatalytic nu-
cleophilic substitution reaction of gem-difluoroalkenes with ketene silyl
acetals,24 amination of b-(hetero)arylethyl ethers with various amines, as
well as in the synthesis of monofluoromethyl substituted epoxides from
3-difluoro-2-propanols. A special feature of the latter is that this reaction
involves activation of the normally inert Csp3–F bond in aliphatic fluor-
ides in the presence of silicon-containing additives.25
Various types of polymerization reactions have been found to be
mediated by 46. These include the anionic polymerization of styrene
monomers,26 anionic ring-opening polymerization (AROP) of epoxides in
the presence of 2,5-diketopiperazines,27 anionic polymerization of
methyl methacrylate (MMA) in the presence of ethyl acetate,28 azide–
alkyne click polymerizations,29 ring-opening polymerization (ROP) of
N-sulfonyl aziridines with a carboxylic acid and the synthesis30 of
poly(glycidyl phenyl ether) (PGPE).31
In addition to 46, other types of phosphazene bases have also found
attention. These include 4732 (for organocatalysis), 48 and 49 (in palladium-
catalyzed C–N cross coupling reactions),33 50 which is a polymer-supported
phosphazene base (ring-opening and trans-esterification reactions),34,35 51
(ring-opening polymerization of a six-membered cyclic carbonate),36 and 52
(room temperature copolymerization of phthalic anhydride and propylene/
ethylene oxide).37

An important amplification of the use of the phosphazene bases is to


generate chirality in them which would make them useful in asymmetric
synthesis. In this regard the chiral iminophosphorane (53) was found to
be an effective catalyst for the enantioselective Henry reaction involving a
nitroaldol condensation.38 Also, 53 was found to catalyze the en-
antioselective sulfenylation of 4-substituted pyrazolones. Encouragingly,
up to 99% ee was achieved.39

Organophosphorus Chem., 2021, 50, 429–466 | 435


b) Cyclophosphazenes
Although cyclophosphazenes occur in various ring sizes the two most
commonly studied systems are the six-membered ring, hexa-
chlorocyclotriphosphazene (54) and the eight-membered ring, octa-
chlorocyclotetraphosphazene (55). These systems have attracted interest
for a variety of reasons. One, the reactive P–Cl bond in these compounds
allows nucleophilic substitution reactions and many studies of this type
are known, both with mono-functional and difunctional reagents. Second,
appropriately substituted cyclophosphazenes are useful as ligands for the
construction of metal complexes. Third, dendrimer-like structures can be
assembled using 54 and 55 as starting materials. Many such dendrimers
can have utility, depending on the nature of substitution, in a variety of
applications ranging from photoluminescent to electrochemically active
materials. These aspects will be discussed in the following.

Typically, the aminolysis of 54 with secondary amines leads to a non-


geminal substitution of the P–Cl bonds, i.e., replacement of the chlorines
on different phosphorus centres. In contrast, the aminolysis of gem-
N3P3Cl4R2 [R ¼ NHPh (56a); (OCH2CH2CH2NH) (56b) and NHtBu (56c)],
with dibenzylamine was found to afford geminal products (57). The
reaction with 56a also resulted in the replacement of all the chlorine
atoms and the formation of 58. The product formation was rationalized
based on the electron-donating properties of the substituent already
present.40

436 | Organophosphorus Chem., 2021, 50, 429–466


54 has six reactive groups. These can be carefully substituted in a step-
wise manner to yield products containing mixed-substituents. Thus, two
of the chlorines on the same phosphorus centres were replaced with a
N/O donor ligand containing a 2-pyridyl pendant arm, affording (59).
Reaction of the latter with pyrrolidine, morpholine or 1,4-dioxa-8-azas-
piro[4,5]decane (DASD) afforded (60a–c) in high yields.41
Similar to the above, reaction of 54 with two equivalents of N/N donor
ligands containing 4-chlorobenzyl pendant arms afforded monospiro (61)
as well as dispiro (cis/trans; 62–63) derivatives. The cis/trans regioisomers
(62–63) possess two stereogenic phosphorus centres.42
The ansa compound, hexanedioxytetrachlorocyclotriphosphazene (64)
has four P–Cl bonds which are of two types. One set is a PCl2 group while
in the other set one chlorine is attached to a phosphorus that already
contains an oxygen substituent. Because of these structural differences,
variation of reactivity is also anticipated. Accordingly, the reaction of 64
with guanidine does not touch the P(Cl)(O) motif and afforded 65, while
reaction with 2-(2-hydroxyethyl)thiophene or benzyl alcohol yielded the
products 66a–b where all the chlorine atoms have been substituted.43

Organophosphorus Chem., 2021, 50, 429–466 | 437


Reaction of N3P3Cl4R2 [R ¼ Cl (67a), (OCH2(CF2)2CH2O) (67b),
(OCH2CH2CH2O) (67c), NHPh (67d), (OCH2CH2CH2NH) (67e) and NHBut
(67f)] with the difunctional reagent 1,3-propanediol, in the presence of
NaH or Et3N as base and THF as solvent, afforded ansa, spiro and dispiro
products (68a–c).44
Octachlorocyclotetraphosphazene (55) has eight replaceable chlorine
atoms and reactions of this compound with difunctional reagents is
interesting because of the wide range of products that are possible.
Thus, the reaction of 55 with two equivalents of sodium 3-(N-ferrocenyl-
methylamino)-1-propanoxide afforded the ansa-spiro compound, 69, the
2-trans-6-dispiro compound, 70, the 2-cis-6-dispiro compound, 71, 2-trans-
4-dispiro compound, 72 and the 2-cis-4-dispiro derivative, 73. The absolute
configurations of 69 and 72 were determined by X-ray crystallography to be
SS 0 S00 and RR, respectively.45

The reaction of N4P4Cl8 (55) with the sodium salt of N/O donor-type
bidentate ligand afforded both ansa (74) and spirocyclic (75) products.
Further reaction of 74 and 75 with secondary amines afforded both
partially substituted (76a–d) and fully substituted derivatives (77a–d).
Apart from the interest in terms of their synthesis and structure, these
compounds also possess interesting antimicrobial activities against
G(þ)/G() bacteria and fungi.46
Coumarin motifs are known to be present in various natural products
and in synthetic compounds. One of the reasons of interest in
this motif is the medicinal properties that they impart including
anti-coagulation. Bio-conjugates prepared from such motifs and cy-
clophosphazene substrates are therefore of interest. Accordingly,
the reaction of 4-methyl-7-hydroxycoumarin with 54 as well as mono-
and dispiro compounds afforded the coumarin-containing compounds,
78–80. These compounds were tested for their anti-tumor activity,
in vitro, against certain breast-cancer cell lines. The results revealed
that 78–80 had good neo-plastic activity with IC50 in the micromolar
range.47

438 | Organophosphorus Chem., 2021, 50, 429–466


Cyclophosphazenes containing spirocyclic substituents were reacted
with anthracene-substituted chalcone compounds affording the com-
pounds 81 and 82. These anthracene-containing compounds were
hybridized with graphene. The dielectric properties of the hybrid com-
posites revealed a significant increase in the dielectric loss. Further,
conductivity measurements reveal that these composites behave as semi-
conductors. It has been suggested that these types of hybrids might find
opto-electronic applications.48
Interest in examining cyclophosphazenes possessing chiral centres has
propelled the preparation of cis- and trans-bis(ferrocenyl)dispirocyclo-
triphosphazenes, 83 and 84. These were further reacted with the potassium
salt of 4-hydroxy-3-methoxybenzaldehyde (potassium vanillinate) affording
the compounds 85–86. Structural characterization by X-ray crystallography
revealed the presence of stereogenic phosphorus centres in 85 and 86.49

Organophosphorus Chem., 2021, 50, 429–466 | 439


N,N-Spiro bridged bis(cyclotriphosphazene) derivatives (88–89; 91–92)
containing four equivalent chiral phosphorus centres were synthesized
from the starting materials 87 and 90. The key reaction involved in the
formation of these compounds involves the deprotonation of the buty-
lamino substituent which leads to the dispiro bridge between the two
phosphorus centres. X-ray crystallographic studies established the
absolute stereochemistry at the stereogenic phosphorus centres.50

Reaction of N,N-spiro-bridged octachlorobiscyclotriphosphazene,


[N3P3Cl4(NCHMe2)]2 (93) with 1 : 1 and 1 : 2 equimolar amounts of
N-(4-fluorobenzyl)-N 0 methylethane-1,2-diamine and N-(4-fluorobenzyl)-
N 0 methylpropane-1,3-diamine afforded the monospiro (94) and dispiro
(95) products. The compounds 93, 94 and 95 were further reacted with
pyrrolidine affording 96–98. The latter were found to have significant
anti-bacterial activity as found from in vitro growth-inhibition studies
against E. coli and B. cereus.51

440 | Organophosphorus Chem., 2021, 50, 429–466


Fully substituted cyclophosphazenes containing fluorinated aryloxy
substituents, N3P3(OC6F5)6, (99a), N3P3[OC6H4(CF3)]6, (99b) and
N3P3[OC6H3(CF3)2]6 (99c) were synthesized. The molecular structures of
these compounds were determined by X-ray crystallography, which
revealed that in these compounds the six-membered P3N3 rings had a
puckered envelope conformation. Further, there was evidence of inter-
molecular C–H–F interactions in these systems.52

Cyclophosphazenes containing imidazolyl- and benzimidazolyl sub-


stituents, 100 and 101, were prepared by the reaction of spiro-
N3P3Cl4[O2C12H8] with the respective nucleophiles in the presence of a
tertiary base. These compounds revealed strong absorption in the UV
region (240–300 nm) and also were shown to be fluorescent
(310–330 nm).53
The compounds 102 and 103, obtained from the reactions of N3P3Cl6
with spermine and N-methyl-1,3-propanediamine respectively, were re-
acted with 5-hydroxy-2-methylbenzothiazole to afford 104 (containing two
thiazole groups) and 105 (containing four thiazole groups). Among these
compounds 104 was found to be effective as a neo-plastic agent in non-
small cell lung carcinoma cell lines.54

Mono(4-fluorobenzyl)cyclotriphosphazene derivatives containing


(dimethylamino)ethoxy and (dimethylamino)propoxy substitutents were
prepared and were further methylated by MeI to form quarternary
ammonium ions (106 and 107). These resulting phosphazene-based
ionic liquids were examined as the dielectric layer in organic field

Organophosphorus Chem., 2021, 50, 429–466 | 441


effect transistors (OFETs). The high dielectric properties of the phos-
phazene ionic liquids enabled these OFETs to operate in a low voltage
range.55

The residual protons in aminocyclophosphazenes can be readily de-


protonated by reactions with appropriate reagents. This feature was
exploited in the reactions of the hexakis amino derivatives N3P3(NHR)6
108 [R ¼ tBu (a), cyclohexyl (b), iPr (c), iBu (d), Et (e), n-Pr (f), methyl (g),
and Bz (h)] which are considered as a hexaprotic phosphazene, (PNH6).
The reactions of 108 with trimethylaluminum were investigated. It was
found that this reaction results in the formation of polyanionic phos-
phazenates [PNHn]n6 accommodating multinuclear arrays of [AlMe2]1
and [AlMe]21 motifs such as 109–111. These studies further reinforced
the idea that aminocyclophosphazenes can be thought of as putative
anionic ligands.56

Cyclophosphazenes have been used as supports for assembling lig-


ands which can be used for constructing both molecular and polymeric
coordination complexes. In this regard, the ligand 112 (AnPyCp), having
four pyridyloxy substituents and two anilino substituents, was prepared
by sequential substitution of P–Cl bonds on N3P3Cl6. Reaction of 112
with Hg(II) salts afforded the coordination polymers [{Hg(An-
PyCp)(Cl)2}CH3CN]n (113a) and [{Hg2(AnPyCp)(I)4}CH3CN]n (113b). In
both the cases, the pyridyloxy nitrogen atoms are involved in binding to
the Hg(II) center.57

442 | Organophosphorus Chem., 2021, 50, 429–466


A similar strategy to that above was used to build a dipyridyloxy cy-
clotriphosphazene ligand, 114. In this case two of the phosphorus centers
are blocked by the spirocyclic substituents leaving the third phosphorus
centre to bear the coordinating pyridyloxy units. Reaction of 114 with
Hg(II) salts afforded the one-dimensional coordination polymers, 115. In
this case, as anticipated, the coordination to Hg(II) occurs through the
pyridyloxy nitrogen centres.58
Post-synthetic modification on a cyclophosphazene substrate involving
a click reaction was demonstrated. Thus, the P–Cl bond in 116 could be
substituted with a group containing a terminal alkyne unit affording 117.
Click reaction chemistry on 117 allowed the formation of 118 which
contains a terminal triazole unit.59

Organophosphorus Chem., 2021, 50, 429–466 | 443


Phosphazene-based materials are found to be thermally stable and
widely utilized as fire retardant materials or thermally insulating
materials. In this regard, cyclophosphazenes containing siloxy substitu-
ents 119 and 120 were prepared and integrated first with mesoporous
organosilica and subsequently with cellulose nanofibers (CNF). By util-
izing a unidirectional freeze-drying method a composite – in the form of
a foam – could be obtained which possessed flame-retardant and ther-
mally insulating performance with low density (16.6 kg m3) and high
porosity (99.2%).60
Polyurethanes are widely used polymeric materials. The properties of
this family can be augmented by incorporation of the cyclophosphazene
motif. Accordingly, compounds of the type 121 were synthesized. 121
undergoes a reversible Diels–Alder (DA) reactions in the presence of a
polyurethane with pendant furan groups (PUF, 121a) to afford novel self-
healing, flame-retardant, recyclable, and mechanically robust poly-
urethane films.61

A hexasubstituted cyclophosphazene containing eugenol substituents


was synthesized (122) and co-polymerized by thiol–ene

444 | Organophosphorus Chem., 2021, 50, 429–466


photopolymerization with multi-thiol monomers, affording hybrid poly-
mer networks, which possess high mechanical strength and intrinsic
flame retardancy.62
There is considerable interest in developing new polymer electrolytes
for lithium ion batteries. In this regard, hexakis(4-aminophenoxy)cyclo-
triphosphazene(123) was prepared. An amino-epoxy reaction involving
123 and polyethylene glycol diglycidyl ether afforded new hybrid mater-
ials which showed lithium ion conduction up to 104 S cm1 at room
temperature.63
An intumescent hybrid flame retardant containing boron, magnesium
hydroxide and cyclophosphazene was prepared by combination of hex-
akis(4-boronic acid-phenoxy)-cyclophosphazene (124), and magnesium
hydroxide (MH). The flame retardancy properties of this composite
material was tested in epoxy resins and was found to inhibit combustion
without affecting the mechanical properties of the polymer.64
The reaction of N3P3Cl6 with ortho-vanillin allowed the synthesis of the
hexakis derivative, 125. Pinnick oxidation of 125 afforded 126 which
contains six peripheral carboxylic acid groups. The latter was used as a
curing agent for epoxy resins which could be employed as flame-
retardant coatings on wood.65
One of the challenges of a polymer electrolyte in lithium metal bat-
teries is to reduce its flammability. Keeping this in mind,
N3P3F5(OCH2CF3) (127) was used as an electrolyte additive. In addition to
the anticipated flame retardancy, addition of 127 resulted in a LiF-rich
solid electrolyte interphase (SEI) layer on the lithium metal which
inhibited the growth of lithium dendrites.66
Blending N3P3(OPh)6 (128) and magnesium hydroxide with the ethylene–
vinyl acetate copolymer (EVA) afforded an EVA–MH–phosphazene com-
posite whose flame retardant properties were significantly better than the
native co-polymer.67 A similar effect was seen on methyl ethyl silicone
rubber (VMQ) by the combination of 128/magnesium hydroxide.68
Hexakis(1-oxo-2,6,7-trioxa-1-phosphabicyclo[2,2,2]octane-4-methyl)-
cyclotriphosphazene (129), was synthesized by the reaction of N3P3Cl6
and 2,6,7-trioxal-phosphabicyclo[2.2.2] octane-4-methanol (PEPA). Com-
posites of polypropylene and 129 were prepared and found to have
enhanced flame retardancy and thermal stability in comparison to the
native polymer.69

Organophosphorus Chem., 2021, 50, 429–466 | 445


N3P3(O-C6H4-4-CHO)6 (130) was used as a crosslinking agent in its
interaction with the aminopolysaccharide, chitosan. The crosslinking
Schiff base formation resulted in a hydrogel, which was used as a drug-
delivery vehicle for amoxicillin.70
Cyclophosphazenes containing peripheral b-diketo units were prepared
by a sequential reaction protocol. Thus, the reduction of 130 with NaBH4
followed by chlorination afforded (4-chloromethylphenoxy)cyclotripho-
sphazene, N3P3(O-C6H4-4-CH2Cl) (131). The reaction of the latter with
sodium acetylacetonate afforded 132 which upon condensation with dia-
mines afforded the polymers, 133 and 134. A representative example of
such a polymer was shown to be highly hydrophobic with a contact angle
of 1011.71

446 | Organophosphorus Chem., 2021, 50, 429–466


Novel tripodal synthetic receptors based on cyclotriphosphazene scaf-
folds were prepared by a click reaction involving 135. The receptors, 136,
contained three triazole groups on one side of the cyclotriphosphazene
ring. Appended to the triazole were the photophysically active naphtha-
lene or anthracene substituents. On the other side of the cyclopho-
sphazene ring are triethylene glycol monomethyl ether moieties which
facilitate solubility in water. 136 was found to be useful for the detection
of iron(III) by spectrofluorimetric measurements.72

Exfoliation of single-walled carbon nanotubes (SWCNTs) is an


important problem. To achieve this, nano-tweezers of cyclotripho-
sphazene hybrids (137a–c) were employed. The latter contained the fol-
lowing building units: a central cyclophosphazene, aromatic perylene
bisimides with hydrophobic aliphatic tails and solvophilic glycol units.
Upon ultrasonicating a mixture of 137 and SWCNTs, disentangled and
undamaged SWCNTs were obtained. It is believed that the aromatic
perylene units, through p–p interactions, facilitate this process.73
Cyclotriphosphazene-based derivatives containing Schiff base linkages
(138) were prepared with an objective to study their liquid crystal

Organophosphorus Chem., 2021, 50, 429–466 | 447


properties by varying the chain length and functional groups.75 It was
found that compounds with heptyl and dodecyl substituents were me-
sogenic with smectic C phases whereas, methoxy and chloro substituents
were ineffective in imparting mesogenic properties.74
The use of cyclophosphazene scaffolds for preparing novel hybrid
materials was demonstrated by 139 which was found to possess inter-
esting NLO properties. The preparation of 139 was accomplished by the
reaction of 2-hydroxy-9(10),16(17),23(24)-tri-tert-butyl-29H,31H-phthalo-
cyanine with N3P3Cl6.75
Hexakis[4-(glycidyloxymethyl)phenoxy]-cyclotriphosphazene (142) was
synthesized by sequential transformations starting from N3P3(O-C6H4-4-
CHO)6 (130). Thus, the latter could be first reduced to 140, which in turn
could be converted to 141, followed by its transformation to the epoxy-
functionalised derivative 142. Utilizing 142 as a curing agent, the epoxy
resin 143 could be prepared, which showed excellent flame retardant
properties.76

448 | Organophosphorus Chem., 2021, 50, 429–466


Tris(o-phenylenedioxy)cyclotriphosphazene, N3P3(O2C6H4)3 (144), in
its solid state has been shown to possess 3D channels. In a study
involving this host, inclusion of dipolar molecular rotors containing a
pyridazine ring and bulky terminal tert-butyl groups was found to be
feasible. Further, as a result of such inclusion, the barrier for dipole
rotation was found to be reduced.77 In another study involving 144,
inclusion of block copolymers containing polypropylene oxide and
polyethylene oxide was found to allow the formation of hexagonal
platelet structural forms.78
Cyclotriphosphazene-based polyimide and polyhedral oligomeric
silsesquioxane (POSS)-augmented phosphazene polyimide nano-
composites (145) were synthesized and studied for their optical, ther-
mal, and dielectric properties. These materials were found to be
excellent flame-retardant coatings. Similarly, the value of the dielectric
constant gets decreased with an increase in the weight percentages of
the POSS.79

A cyclophosphazene core is useful to construct dendrimer-like


molecules because of the presence of the replaceable chlorine atoms
in N3P3Cl6 and N4P4Cl8. With a judicial choice of substituent,
dendrimer-like molecules (146–151) were assembled. These were
found to be useful as electrochromic devices (146)80 luminescent
materials (147, 148),81 fluorescent probes (149)82 and as photo-
sensetizers (150).83

Organophosphorus Chem., 2021, 50, 429–466 | 449


450 | Organophosphorus Chem., 2021, 50, 429–466
The cyclophosphazene super base, 152, continues to be used widely,
particularly as an organocatalyst in various organic transformations.
For example, 152 was found to catalyze the ring-opening polymerization
(ROP) of e-caprolactone (e-CL) and d-valerolactone (d-VL) in the
presence of alcohol as an initiator.84 Other ROP reactions catalyzed
by 152 include those involving octamethylcyclotetrasiloxane,85
o-pentadecalactone (PDL),86 and nonstrained g-butyrolactone (gBL).87
152 was also found to catalyze the ring-opening alternating copoly-
merization (ROAP) of epoxides and cyclic anhydrides88 as well as an
anti-Markovnikov stereoselective hydroamination and hydrothiolation
of alkynes.89

c) Polyphosphazenes
This section will deal with polymers involving a PQN backbone and
those that have an intact cyclophosphazene unit and other hybrid
polymers. Technically, only the polymers containing the PQN backbone
are to be referred to as polyphosphazenes. Unfortunately, there is a
trend in the recent literature to refer to all types of phosphazene-derived
polymers as polyphosphazenes. This is misleading and needs to be
avoided.
Polyphosphazenes continue to be investigated actively in view of their
diverse applications in various fields. Nadimide-substituted polyphos-
phazenes, 153, 154, and 156 were prepared by the reaction of
polydichlorophosphazene with N-(4-hydroxyphenyl) chlorendimide, N-(4-
methylphenyl) nadimide and N-(4-hydroxyphenyl) methyl nadimide
respectively. These polymers also contain trifluoroethoxy substituents.
Curing these polymers with free-radical initiators afforded thermally
stable resins.90

Organophosphorus Chem., 2021, 50, 429–466 | 451


Poly(diaryloxy)phosphazene, 157, has been found to be a very good
material for being used in the fabrication of thermal conductive com-
posites containing fillers such as Al2O3, boron nitride, or expanded
graphite.91 Among these composites those containing expanded graphite
was found to enhance the thermal conductivity of the native polymer.
Further, all the composites revealed a good self-extinguishing property
without the need for any additional flame retardant.
Acrylonitrile–butadiene–styrene (ABS) composites are among the
most widely used thermoplastics for a wide-range of applications. One
of the problems of this ternary polymer is its flammability in its molten
state which is an operational hazard. Poly(phenoxy)(4-nitrophenoxy)-
phosphazene, an intumescent flame retardant along with ammonium
polyphosphate were used to improve the process ability properties
of ABS.92
In addition to applications as flame-retardant materials, appropriately
substituted polyphosphazenes are also finding several biological appli-
cations. It is now recognized that in many diseases a combination of two
or more drugs can manage the disease better. One of the issues in such a
combination therapy is to ensure the drug-release in a controlled

452 | Organophosphorus Chem., 2021, 50, 429–466


manner. Towards such application, polyphosphazenes 158 and 159 were
synthesized and blended with polymethylmethacrylate to form hollow
microspheres which could encapsulate the drugs acetamidophenol and
camptothecin. The drug-delivery of these was monitored and found to be
good.93
Cardiovascular implants have become extremely important in man-
aging cardiovascular diseases. A common problem associated with the
implants is bacterial infection and thrombosis. An interesting study to
address this problem found that poly(bis(2,2,2-trifluoroethoxy)pho-
sphazene)–Al2O3 hybrid nanowires provided a superhydrophobic surface
which reduced both platelet adhesion/activation and bacterial adher-
ence/colonization.94
The presence of phosphorus in polyphosphazenes has been thought to
enhance bone regeneration using tissue engineering. This requires
designing polymers that slowly erode in biological conditions releasing
inorganic phosphate. Accordingly, the polyphosphazenes containing
amino acid ester substituents 160 and 161 were synthesized and used for
osteogenic differentiation.95

In other applications involving polyphosphazenes, hybrid systems


built using cobalt–bipyridine complexes anchored on mesoporous silica
surfaces and polyphosphazene brushes (162) allowed the on-demand
decomposition of hydrogen peroxide to supply oxygen. The process is
modulated in a narrow temperature range of 25–30 1C.96
Reactive oxygen species (ROS), a product of various biological meta-
bolic activities, is generated in excess in diseased cells such as cancer
cells. Detection of such damage by imaging is important. The poly-
phosphazene 163 – hybridized with gold nanoparticles – has been found
to be useful in this. Thus, 163 has been found to selectively degrade
upon ROS exposure triggering a switch-off photoacoustic response
due to gold-nanoparticle disassembly consequent on the degradation
of 163.97

Organophosphorus Chem., 2021, 50, 429–466 | 453


Adapted from ref. 97 with permission from ACS Appl. Mater. Interfaces
2019.

In an interesting study, it was found that polyphosphazenes containing


both sulfonated and fluorinated substituents, 164 and 165, were useful
biomaterials that could be used as hemocompatible coatings. It was
found by in vitro studies that nanocoatings of these materials decreased
hemolysis.98

454 | Organophosphorus Chem., 2021, 50, 429–466


A thermosensitive polyphosphazene-bearing b-cyclodextrin hydrogel
containing a small peptide as a guest (166) was prepared. Controlled
release of the guest peptide was observed.99
A polyphosphazene (167) bearing terminal siloxy end groups was pre-
pared and found to be amenable to a sol–gel hydrolysis process to afford
polymer nanoparticles. It has been suggested that the latter could be
used as drug-delivery vehicles.100

Polyphosphazenes containing the side groups –O(CH2)3Si(CH3)3/


–OCH2CF3 (168, 169) were prepared and studied. It was found that if only
the silyl substituents are present, the polymers tend to be thermoplastics.
On the other hand, mixed substitution involving the trifluorethoxy
groups afforded polymers that were found to be elastomers, which could
be appropriate for applications as biomaterials.101
Polyphosphazenes are also finding use as materials in energy appli-
cations. In this regard, a polymer 170 containing nitroxyl substituents,
such that each repeating unit has four unpaired electrons, was prepared.
This polymer was found to be useful as a cathode-active material in the
fabrication of rechargeable lithium ion batteries.102
Poly[bis(2-(2-methoxyethoxy)ethoxy)phosphazene] 171 was incorpor-
ated in a complex polymer electrolyte composition for applications in an
all-solid-state lithium batteries.103 In another report it was observed that

Organophosphorus Chem., 2021, 50, 429–466 | 455


171, as a polymer electrolyte, could have advantageous effects in sup-
pressing dendritic growth in lithium batteries.104
Theoretical studies carried out on poly(difluorophosphazene), (PNF2)n
172, revealed that this material possesses a large band gap and has the
potential for use as a deep-ultraviolet non-linear optical (DUV NLO)
material. It has been suggested that experimentally realizing this theo-
retical prediction would pave the way for a new generation of NLO
materials.105
In contrast to the polymers discussed above, we now begin our review
of polymeric systems that contain an intact cyclotriphosphazene skel-
eton. In most cases such polymers are cyclomatrix polymers and are
cross-linked. Thermally stable polycondensation products 173 and 174
were obtained from hexakis(p-acetylphenoxy)cyclotriphosphazene. The
condensation reaction involves either a dimerization or trimerization
process. The cross-linked polymers, thus obtained, were found to be
stable up to 400 1C.106

An amine-capped cross-linked polymer 175 was used to modify carbon


fiber. The effectiveness of binding of the cross-linked polymer depended
on the amount of polar amino groups. Studies also revealed that the
interfacial shear strength is also dependent on the binding, which
depended on the number of polar amino groups.107
Resins containing bisphenol-A have applications in areas such as
polymer composites, adhesives, and paints. One of the shortcomings of
these systems is their high viscosity. To overcome this, phosphazene-
containing epoxy-resorcinol hybrid oligomers 176 were synthesized. It
was found that these are less viscous than the bisphenol-A containing
resins.108

456 | Organophosphorus Chem., 2021, 50, 429–466


N3P3Cl6 was reacted with ferrocene dimethanol affording a cross-
linked polymer 177. Post-synthetic manipulation of the polymer into
microspheres or nanotubes was achieved by varying the types of solvent
combinations. These were found to be highly useful for the absorption of
the dye, methyl orange.109
A ferrocene–melamine cross-linked polymer was reacted with N3P3Cl6
affording an interconnected polymer network (178). Pyrolysis of this at
900 1C afforded a FeP/Fe2P embedded porous carbon which was found to
be a very good catalyst for the oxygen reduction reaction over a wide pH
range.110

The reaction of 179 with a siloxane precursor 180, using AIBN, afforded
the cross-linked polymer 181 via the thiol–ene addition reaction. The

Organophosphorus Chem., 2021, 50, 429–466 | 457


cross-linked polymer was used for generating various types of sample
types such as films, aerogels and monoliths.111

A metal organic framework (MOF) was built using N3P3(O-C6H4-p-


COOH)6 (182) and Mn(II) salts. The MOF had helical channels which
could be used to trap small ketones such as acetone, acetophenone etc.112
A phosphorus/nitrogen-rich cross-linked polymer 183 was used to form
composites with polyurethanes. These were found to improve flame re-
tardancy and mechanical properties in comparison to the native
polyurethanes.113
A cross-linked cyclophosphazene-containing polymer (184) possessing
active amine groups was used as an additive to improve the flame-
retardant capacity of epoxy resins as well as to reduce smoke pro-
duction. It was found that 184 accelerates the formation of a char layer
which protects the substrate.114

458 | Organophosphorus Chem., 2021, 50, 429–466


In an interesting strategy, using the concepts of supramolecular
organization and conventional covalent bond-formation reactions, a new
hybrid polymeric system consisting of melamine–cyanurate (185) and
cyclophosphazene motifs (186) was assembled. The resultant composite
was found to possess good flame-retardant properties.115
A flame-retardant additive (187) was prepared by the reaction of
N3P3Cl6 with dihydroxypropyl silicon oil and phenol. 187 was incorp-
orated into polycarbonate to impart it with flame-retardant
properties.116

Hybrid cross-linked polymer systems containing polybenzoxazine re-


sins and cyclophosphazene were assembled. The cyclophosphazene
polymer microspheres were assembled by the reaction of N3P3Cl6
and monomers such as 4,4 0 -(hexafluoroisopropylidene)diphenol,

Organophosphorus Chem., 2021, 50, 429–466 | 459


4,4 0 -sulfonyldiphenol (188), 4,4-(9-fluorenylidene)diphenol (189), and
phenolphthalein. These were incorporated in the polybenzoxazine resins
to impart them with improved flame retardancy.117
Polybismaleimide (BMI) is an important thermosetting resin with wide
utility. Thus, composites of BMI with cyclophosphazene cyclomatrix
polymers containing functional end groups (190) reinforced the polymer
properties in terms of flexural strength.118
The utility of cyclophosphazene-based cyclomatrix polymers as
nanoreactors was explored. Thus, a one-step protocol was developed to
prepare hollow microspheres based on the crosslinked polymer (190),
obtained in the reaction of N3P3Cl6 and 4,4 0 -sulfonyldiphenol.
The cavities within these microspheres could encapsulate platinum
nanoclusters. This composite was found to be useful in the catalytic
oxidation of 1,3,5-trimethylbenzene.119–123
Removal of Cr(VI) from industrial waste water is an important prob-
lem. A cyclophophazene–polyethylenimine polymer adsorbed on Fe/
Fe2O3 was developed. The extensive –NH and –NH31 units present in
the material allow both the encapsulation of CrVIO3, through hydrogen
bonding and electrostatic forces, and its subsequent reduction to Cr(III)
and finally the magnetic separation of the metallic impurities from
wastewater.124 Similarly the adsorption of radionuclides such as U(VI)
from aqueous solution was accomplished by using cyclophosphazene
polymer-derived amidate-functionalized carbon materials.125 On simi-
lar lines a composite of a cyclophosphazene polymer/carbon nanotubes
anchored on magnetic Fe3O4 was developed for the selective adsorption
of U(VI) ions.126 Poly(cyclotriphosphazene-co-4,4 0 -diaminodiphenyl
ether) crosslinked microspheres with active amino groups on the sur-
face were synthesized by a one-step precipitation method and used to
remove U(VI) and Th(IV) ions from aqueous solution.127,128 A cross-
linked cyclophosphazene polymer containing phloroglucinol was used
as an effective and selective adsorbent for the removal of cationic
dyes from aqueous solution as well as a solid phase extractant for
the U(VI) ion from solution.129 Cellulose (or its derivatives) could be
derivatized by using N3P3Cl6 to afford materials such as 191. These
were used to remove toxic dyes from waste water.130 Other types of
cyclophosphazene-based polymers containing amine functional groups
(192) were also used to remove chemical dyes.131

460 | Organophosphorus Chem., 2021, 50, 429–466


Reverse osmosis (RO) and nanofiltration (NF) are membrane-based
technologies used for the purification of water. In this line of interest, a
polyamine-based cyclophosphazene (192) nanofiltration membrane was
prepared by the interfacial polymerization of polyethyleneimine and
N3P3Cl6 on a polysulfone support. This membrane was able to reject
various salts: MgCl2 (97%)4MgSO4 (88%)4NaCl (87%)4Na2SO4
(58%).132
N3P3(O-C6H4-4-Br)6 was reacted with octavinylsilsesquioxane in a Heck
coupling reaction to afford a hybrid porous material, 193, which has a
surface area of 500 m2 g1. This material revealed high adsorption cap-
acities towards metal ions such as Cu21, Hg21 and Pb21.133
Cyclophosphazene–dopamine crosslinked polymers were prepared by
the technique of precipitation polymerization and processed to be
obtained as hollow microspheres. These porous materials could adsorb a
model drug – acriflavine – and controlled and targeted drug release was
observed towards cancerous cells; a sustained release of up to 7 days was
observed.134
Theranostic nanoplatforms are a major topic in the field of nano-
medicine. The multifunctional poly(cyclotriphosphazene-co-poly-
ethylenimine) nanospheres labeled with the radionuclide 131I were
prepared and examined for single photon emission computed tomo-
graphy (SPECT) imaging-guided radiotherapy of tumors.135
Molecules showing excitation wavelength tunable fluorescence (EWTF)
are of interest for several applications. In this regard, a one-pot precipi-
tation polymerization involving the reaction of N3P3Cl6 and 4,4 0 -methy-
lenedianiline was developed. The crosslinked polymers could be
obtained as microspheres which exhibited tunable fluorescence in
the following regions: blue (lex 280 nm), green (lex 365/420 nm) and red
(lex 546 nm).136
Similar to the above, poly(cyclotriphosphazene-co-tris(4-hydroxy-
phenyl)ethane) microspheres137 were assembled by reacting N3P3Cl6 with
tris(4-hydroxyphenyl)ethane. Photophysical studies on these micro-
spheres revealed that they emit with an emission maxima at 487, 530 and
623 nm, under different excitation wavelengths of 365, 420 and 546 nm,
respectively. Another finding of this study is that these microspheres are
remarkably stable to photobleaching, even under repeated excitations at
365 and 420 nm.

Organophosphorus Chem., 2021, 50, 429–466 | 461


References
1 A. Yoshimura, C. L. Makitalo, M. E. Jarvi, M. T. Shea, P. S. Postnikov,
G. T. Rohde, V. V. Zhdankin, A. Saito and M. S. Yusubov, Molecules, 2019,
24, 979.
2 N. F. El-Sayed, M. El-Hussieny, E. F. Ewies, M. A. Fouad and L. S. Boulos,
Bioorg. Chem., 2020, 95, 103499.
3 B. J. Cook, S. I. Johnson, G. M. Chambers, W. Kaminsky and R. M. Bullock,
Chem. Commun., 2019, 55, 14058.
4 C. G. Martı́nez-De-León, A. Rodrı́guez-Álvarez, A. Flores-Parra and
J.-M. Grévy, Inorg. Chim. Acta, 2019, 495, 118945.
5 J. L. Rodrı́guez-Rey, D. Esteban-Gómez, C. Platas-Iglesias and
A. Sousa-Pedrares, Dalton Trans., 2019, 48, 486.
6 N. T. Rice, J. Su, T. P. Gompa, D. R. Russo, J. Telser, L. Palatinus, J. Bacsa,
P. Yang, E. R. Batista and H. S. La Pierre, Inorg. Chem., 2019, 58, 5289.
7 Y. Sun, W. Huang, Z. Li, T. Wang and J. Luo, J. Chem. Res., 2019, 43,
119.
8 Y. Wang, Y. Yu, L. Zhao and Y. Ning, Eur. J. Org. Chem., 2019, 2019, 7237.
9 F. Yi, W. Zhao, Z. Wang and X. Bi, Org. Lett., 2019, 21, 3158.
10 L. Chen, S. Cao, J. Zhang and Z. Wang, Tetrahedron Lett., 2019, 60, 1678.
11 H.-M. Wang, T.-S. Wang, S.-J. He, Z.-Y. Chen and Y.-G. Hu, J. Chem. Res.,
2019, 43, 201.
12 M. Tryniszewski, R. Bujok, P. Cmoch, R. Gańczarczyk, I. Kulszewicz-Bajer
and Z. Wróbel, J. Org. Chem., 2019, 84, 2277.
13 H. B. Tukhtaev, K. L. Ivanov, S. I. Bezzubov, D. A. Cheshkov, M. Y. Melnikov
and E. M. Budynina, Org. Lett., 2019, 21, 1087.
14 K. L. Ivanov, M. Y. Melnikov and E. M. Budynina, Org. Lett., 2019, 21, 4464.
15 N. Hou, J.-H. Man, X.-Y. Wang, S.-J. He, Q. Li and Y.-G. Hu, ChemistrySelect,
2019, 4, 4901.
16 E. Khademloo, H. Saeidian, Z. Mirjafary and J. M. Aliabad, ChemistrySelect,
2019, 4, 3762.
17 H. Saeidian and E. Barfinejad, ChemistrySelect, 2019, 4, 3088.
18 S. Ullrich, B. Kovačević, X. Xie and J. Sundermeyer, Angew. Chem., Int. Ed.,
2019, 58, 10335.
19 R. F. Weitkamp, B. Neumann, H.-G. Stammler and B. Hoge, Angew. Chem.,
Int. Ed., 2019, 58, 14633.
20 S. Ullrich, D. Barić, X. Xie, B. Kovačević and J. Sundermeyer, Org. Lett., 2019,
21, 9142.
21 M. Shigeno, K. Hayashi, K. Nozawa-Kumada and Y. Kondo, Chem. Eur.,
2019, 25, 6077.
22 M. Shigeno, R. Nakamura, K. Hayashi, K. Nozawa-Kumada and Y. Kondo,
Org. Lett., 2019, 21, 6695.
23 S. Mehta and D. Brahmchari, J. Org. Chem., 2019, 84, 5492.
24 A. Kondoh, K. Koda and M. Terada, Org. Lett., 2019, 21, 2277.
25 J. Wang, J. Tanaka, E. Tokunaga and N. Shibata, Asian J. Org. Chem., 2019,
8, 641.
26 K. Ntetsikas, G. Polymeropoulos, G. Zapsas, P. Bilalis, Y. Gnanou and
N. Hadjichristidis, J. Polym. Sci., Part A: Polym. Chem., 2019, 57, 456.
27 Ö. Tezgel, V. Puchelle, H. Du, N. Illy and P. Guégan, J. Polym. Sci., Part A:
Polym. Chem., 2019, 57, 1008.
28 D.-Y. Xia, Q.-M. Jiang, W.-Y. Huang, H.-J. Yang, X.-Q. Xue, L. Jiang and
B.-B. Jiang, Chin. J. Polym. Sci., 2019, 37, 598.
29 B. Li, Y. Liu, H. Nie, A. Qin and B. Z. Tang, Macromolecules, 2019, 52, 4713.

462 | Organophosphorus Chem., 2021, 50, 429–466


30 R. Yang, Y. Wang, W. Luo, Y. Jin, Z. Zhang, C. Wu and N. Hadjichristidis,
Macromolecules, 2019, 52, 8793.
31 J. Ochs, D. E. Martı́nez-Tong, A. Alegria and F. Barroso-Bujans, Macro-
molecules, 2019, 52, 2083.
32 A. Casnati, A. Perrone, P. P. Mazzeo, A. Bacchi, R. Mancuso, B. Gabriele,
R. Maggi, G. Maestri, E. Motti, A. Stirling and N. D. Ca’, J. Org. Chem., 2019,
84, 3477.
33 L. M. Baumgartner, J. M. Dennis, N. A. White, S. L. Buchwald and
K. F. Jensen, Org. Process Res. Dev., 2019, 23, 1594.
34 C. Ren, X. Zhu, N. Zhao, Y. Shen, L. Chen, S. Liu and Z. Li, Eur. Polym. J.,
2019, 119, 130.
35 Z. Wang, R. Gérardy, G. Gauron, C. Damblon and J.-C. M. Monbaliu, React.
Chem. Eng., 2019, 4, 17.
36 G. Hua, J. Franzén and K. Odelius, Macromolecules, 2019, 52, 2681.
37 H. Li, G. He, Y. Chen, J. Zhao and G. Zhang, ACS Macro Lett., 2019, 8, 973.
38 Y. Zhang, X.-Y. Wu and J. Han, Chin. Chem. Lett., 2019, 30, 1519.
39 J. Han, Y. Zhang, X.-Y. Wu and H. N. C. Wong, Chem. Commun., 2019,
55, 397.
40 S. Bes- li, C. Mutlu Balcı, C. Köseoğlu, D. Palabıyık and C. W. Allen, Inorg.
Chim. Acta, 2019, 492, 23.
41 G. Elmas, Phosphorus Sulfur Silicon Relat. Elem., 2019, 194, 13.
42 N. Asmafiliz, I_ . Berberoğlu, M. Özgür, Z. Kılıç, H. Kayalak, L. Açık, M. Türk
and T. Hökelek, Inorg. Chim. Acta, 2019, 495, 118949.
43 S. Ture and Ö. Kücük, Phosphorus Sulfur Silicon Relat. Elem., 2019, 194,
895.
44 D. Palabıyık, C. Mutlu Balcı and S. Bes- li, Inorg. Chim. Acta, 2019, 487, 15.
45 G. Elmas, A. Okumus- , T. Hökelek and Z. Kılıç, Inorg. Chim. Acta, 2019,
497, 119106.
46 A. Binici, A. Okumus- , G. Elmas, Z. Kılıç, N. Ramazanoğlu, L. Açık, H. -Sims-
ek, B. Çağdas- Tunalı, M. Türk, R. Güzel and T. Hökelek, New J. Chem., 2019,
43, 6856.
47 J. Chen, L. Wang, Y. Fan, Y. Yang, M. Xu and X. Shi, New J. Chem., 2019,
43, 18316.
48 K. Koran, J. Mol. Struct., 2019, 1179, 224.
49 Y. Tümer, N. Asmafiliz, G. Arslan, Z. Kılıç and T. Hökelek, J. Mol. Struct.,
2019, 1181, 235.
50 C. Mutlu Balcı and S. Bes- li, Inorg. Chim. Acta, 2019, 497, 119093.
51 E. Öztürk, A. Okumus- , Z. Kılıç, A. Kılıç, H. Kayalak, L. Açık, N. Aytuna Çerçi,
M. Türk and T. Hökelek, Inorg. Chim. Acta, 2019, 486, 172.
52 H. P. Yennawar, A. R. Hess and H. R. Allcock, Acta Crystallogr., Sect. E, 2019,
75, 1525.
53 H. I_ bis- oğlu, E. Erdemir, D. Atilla, Y. Zorlu and E. -Senkuytu, Inorg. Chim.
Acta, 2019, 498, 119120.
54 G. Yenilmez Çiftçi, E. Tanrıverdi Eçik, O. Goler, F. Yuksel, E. Duygulu,
F. Donbaloglu, G. Turhal and A. Demiroglu-Zergeroglu, Inorg. Chim. Acta,
2019, 498, 119158.
55 H. Akbas- , A. Karadağ, A. Destegül, Ç. Çakırlar, Y. Yerli, K. C. Tekin,
U. Malayoğlu and Z. Kılıç, New J. Chem., 2019, 43, 2098.
56 P. I. Richards, G. T. Lawson, J. F. Bickley, C. M. Robertson, J. A. Iggo and
A. Steiner, Inorg. Chem., 2019, 58, 3355.
57 D. Davarcı, E. -Senkuytu and Y. Zorlu, Polyhedron, 2019, 173, 114138.
58 D. Davarcı, S. O. Tümay, E. -Senkuytu, M. Wörle and Y. Zorlu, Polyhedron,
2019, 161, 104.

Organophosphorus Chem., 2021, 50, 429–466 | 463


59 V. P. Morgalyuk, T. S. Strelkova, Y. N. Kononevich and V. K. Brel, Russ. J.
Gen. Chem., 2019, 89, 1620.
60 D. Wang, X. Feng, L. Zhang, M. Li, M. Liu, A. Tian and S. Fu, Chem. Eng. J.,
2019, 375, 121933.
61 Y. Fang, X. Du, S. Yang, H. Wang, X. Cheng and Z. Du, Polym. Chem., 2019,
10, 4142.
62 T. Liu, L. Sun, R. Ou, Q. Fan, L. Li, C. Guo, Z. Liu and Q. Wang, Chem. Eng. J.,
2019, 368, 359.
63 C. Zuo, M. Yang, Z. Wang, K. Jiang, S. Li, W. Luo, D. He, C. Liu, X. Xie and
Z. Xue, J. Mater. Chem. A, 2019, 7, 18871.
64 L. Ai, S. Chen, J. Zeng, L. Yang and P. Liu, ACS Omega, 2019, 4, 3314.
65 Y. Huang, T. Ma, Q. Wang and C. Guo, ACS Sustainable Chem. Eng., 2019,
7, 14727.
66 Y. Li, Y. An, Y. Tian, H. Fei, S. Xiong, Y. Qian and J. Feng, J. Electrochem. Soc.,
2019, 166, A2736–A2740.
67 L. Shen, C. Shao, R. Li, Y. Xu, J. Li and H. Lin, Polym. Bull., 2019, 76, 2399.
68 J. Qi, Q. Wen and J. Zhu, Mater. Lett., 2019, 249, 62.
69 C. Zhang, X. Guo, S. Ma, Y. Zheng, J. Xu and H. Ma, J. Therm. Anal. Calorim.,
2019, 137, 33.
70 H. Ozay, P. Ilgin and O. Ozay, Int. J. Polym. Mater., 2019, 1.
71 E. M. Chistyakov, A. S. Tupikov, M. I. Buzin, R. S. Borisov and V. V. Kireev,
Mater. Chem. Phys., 2019, 223, 353.
72 S. O. Tümay and S. Yes- ilot, J. Photochem. Photobiol., A, 2019, 372, 156.
73 H. Eserci, E. -Senkuytu and E. Okutan, J. Mol. Struct., 2019, 1182, 1.
74 Z. Jamain, M. Khairuddean and S. A. Saidin, J. Mol. Struct., 2019, 1186,
293.
75 A. Y. Tolbin, V. K. Brel, B. N. Tarasevich and V. E. Pushkarev, Dyes Pigm.,
2020, 174, 108095.
76 L. Zhou, G. Zhang, S. Yang, L. Yang, J. Cao and K. Yang, Thermochim. Acta,
2019, 680, 178348.
77 J. Kaleta, G. Bastien, J. Wen, M. Dračı́nský, E. Tortorici, I. Cı́sařová,
P. D. Beale, C. T. Rogers and J. Michl, J. Org. Chem., 2019, 84, 8449.
78 G. Chen, G. Zhang, B. Jin, M. Luo, Y. Luo, S. Aya and X. Li, J. Am. Chem. Soc.,
2019, 141, 15498.
79 R. Revathi, P. Prabunathan and M. Alagar, Polym. Bull., 2019, 76, 387.
80 S. Çetindere, S. O. Tümay, A. -Senocak, A. Kılıç, M. Durmus- , E. Demirbas- and
S. Yes- ilot, Inorg. Chim. Acta, 2019, 494, 132.
81 E. Özcan, S. O. Tümay, G. Kes- an, S. Yes- ilot and B. Ços- ut, Spectrochim. Acta,
Part A, 2019, 220, 117115.
82 O. Ozay and H. Ozay, Phosphorus Sulfur Silicon Relat. Elem., 2019, 194, 221.
83 E. Okutan, H. Eserci and E. -Senkuytu, Spectrochim. Acta, Part A, 2019,
222, 117232.
84 Y. Li, N. Zhao, C. Wei, A. Sun, S. Liu and Z. Li, Eur. Polym. J., 2019, 111, 11.
85 J. Shi, N. Zhao, S. Xia, S. Liu and Z. Li, Polym. Chem., 2019, 10, 2126.
86 N. Zhao, C. Ren, Y. Shen, S. Liu and Z. Li, Macromolecules, 2019, 52, 1083.
87 Y. Shen, J. Zhang, Z. Zhao, N. Zhao, F. Liu and Z. Li, Biomacromolecules,
2019, 20, 141.
88 X. Kou, Y. Li, Y. Shen and Z. Li, Macromol. Chem. Phys., 2019, 220, 1900416.
89 N. Zhao, C. Lin, L. Wen and Z. Li, Tetrahedron, 2019, 75, 3432.
90 K. P. Singh, A. Mishra, N. Kumar and T. C. Shami, Polym. Bull., 2019,
76, 2277.
91 W. Zou, M. Basharat, S. U. Dar, S. Zhang, Y. Abbas, W. Liu, Z. Wu and
T. Zhang, Composites, Part A, 2019, 119, 145.

464 | Organophosphorus Chem., 2021, 50, 429–466


92 Y. Yang, H. Luo, X. Cao, F. Zhou, W. Kong and X. Cai, J. Therm. Anal.
Calorim., 2019, 137, 65.
93 R. S. Ullah, L. Wang, H. Yu, M. Haroon, T. Elshaarani, K.-U.-R. Naveed,
S. Fahad, A. Khan, A. Nazir, X. Xia and L. Teng, J. Mater. Sci., 2019, 54, 745.
94 A. Haidar, A. A. Ali, S. Veziroglu, J. Fiutowski, H. Eichler, I. Müller, K. Kiefer,
F. Faupel, M. Bischoff, M. Veith, O. C. Aktas and H. Abdul-Khaliq, Nanoscale
Adv., 2019, 1, 4659.
95 Z. Huang, L. Yang, X. Hu, Y. Huang, Q. Cai, Y. Ao and X. Yang, Macromol.
Biosci., 2019, 19, 1800464.
96 M. Kneidinger, A. Iturmendi, C. Ulbricht, T. Truglas, H. Groiss, I. Teasdale
and Y. Salinas, Macromol. Rapid Commun., 2019, 40, 1900328.
97 M. Bouché, M. Pühringer, A. Iturmendi, A. Amirshaghaghi, A. Tsourkas,
I. Teasdale and D. P. Cormode, ACS Appl. Mater. Interfaces, 2019, 11, 28648.
98 V. Albright, A. Marin, P. Kaner, S. A. Sukhishvili and A. K. Andrianov, ACS
Appl. Bio Mater., 2019, 2, 3897.
99 K. H. Hong and S.-C. Song, Biomaterials, 2019, 218, 119338.
100 V. Poscher, I. Teasdale and Y. Salinas, ACS Appl. Nano Mater., 2019, 2, 655.
101 C. Tong, S. McCarthy, Z. Li, J. Guo, Q. Li, C. N. Pacheco, Y. Ren and
H. R. Allcock, ACS Appl. Polym. Mater., 2019, 1, 1881.
102 S. Yes- ilot, F. Hacıvelioğlu, S. Küçükköylü, E. Demir, K. B. Çelik and
R. Demir-Cakan, Polym. Adv. Technol., 2019, 30, 2977.
103 S. Yu, S. Schmohl, Z. Liu, M. Hoffmeyer, N. Schön, F. Hausen, H. Tempel,
H. Kungl, H. D. Wiemhöfer and R. A. Eichel, J. Mater. Chem. A, 2019, 7, 3882.
104 X. He, S. Schmohl and H. D. Wiemhöfer, ChemElectroChem, 2019, 6, 1166.
105 L. Kang, X. Zhang, F. Liang, Z. Lin and B. Huang, Angew. Chem., Int. Ed.,
2019, 58, 10250.
106 E. M. Chistyakov, M. I. Buzin, S. M. Aksenov, A. S. Tupikov and V. V. Kireev,
Mendeleev Commun., 2019, 29, 99.
107 X. Cheng, Z. He, Y. Luo, X. Zhang and C. Lei, Polym. Compos., 2019,
40, E1831.
108 I. A. Sarychev, I. S. Sirotin, R. S. Borisov, J. Mu, I. B. Sokolskaya,
J. V. Bilichenko, S. N. Filatov and V. V. Kireev, Polymers, 2019, 11, 614.
109 Y. Abbas, Z. Zuhra, M. Basharat, M. Qiu, Z. Wu, D. Wu and S. Ali, J. Phys.
Chem. B, 2019, 123, 4148.
110 B. Zhou, F. Yan, X. Li, J. Zhou and W. Zhang, ChemSusChem, 2019, 12, 915.
111 D. A. Khanin, Y. N. Kononevich, M. N. Temnikov, V. P. Morgalyuk,
V. G. Vasil’ev, A. Y. Popov, V. K. Brel, V. S. Papkov and A. M. Muzafarov,
Polymer, 2020, 186, 122011.
112 B. Li, X. Chen, P. Hu, A. Kirchon, Y.-M. Zhao, J. Pang, T. Zhang and
H.-C. Zhou, ACS Appl. Mater. Interfaces, 2019, 11, 8227.
113 J. Miao, Y. Fang, Y. Guo, Y. Zhu, A. Hu and G. Wang, ACS Appl. Polym.
Mater., 2019, 1, 2692.
114 G. Yang, W.-H. Wu, Y.-H. Wang, Y.-H. Jiao, L.-Y. Lu, H.-Q. Qu and X.-Y. Qin,
J. Hazard. Mater., 2019, 366, 78.
115 K. Malkappa and S. S. Ray, ACS Omega, 2019, 4, 9615.
116 J. Jiang, Y. Wang, Z. Luo, T. Qi, Y. Qiao, M. Zou and B. Wang, Polymers, 2019,
11, 1155.
117 L. Zhao, C. Zhao, C. Guo, Y. Li, S. Li, L. Sun, H. Li and D. Xiang, ACS Omega,
2019, 4, 20275.
118 Z. Chen, H. Yan, L. Guo, L. Li, P. Yang and B. Liu, Composites, Part A, 2019,
121, 18.
119 F. Wang, J. Liu, D. Wang, Z. Yang, K. Yan and L. Meng, Nanoscale, 2019,
11, 15017.

Organophosphorus Chem., 2021, 50, 429–466 | 465


120 W. Fu, R. Xu, X. Zhang, Z. Tian, H. Huang, J. Xie and C. Lei, J. Power Sources,
2019, 436, 226839.
121 X. Chen, H. Xu, D. Liu, C. Yan and Y. Zhu, Compos. Sci. Technol., 2019, 169,
34–44.
122 T. Li, S. Li, T. Ma, Y. Zhong, L. Zhang, H. Xu, B. Wang, X. Feng, X. Sui,
Z. Chen and Z. Mao, Eur. Polym. J., 2019, 120, 109270.
123 Z. Zhou, F. Chen, L. Wu, T. Kuang, X. Liu, J. Yang, P. Fan, Z. Fei, Z. Zhao and
M. Zhong, Electrochim. Acta, 2020, 332, 135490.
124 Z. Wang, Y. Wang, S. Cao, S. Liu, Z. Chen, J. Chen, Y. Chen and J. Fu,
J. Hazard. Mater., 2020, 384, 121483.
125 Y. Liu, Y. Ouyang, D. Huang, C. Jiang, X. Liu, Y. Wang, Y. Dai, D. Yuan and
J. W. Chew, Sci. Total Environ., 2020, 706, 136019.
126 Y. Liu, Z. Zhao, D. Yuan, Y. Wang, Y. Dai, Y. Zhu and J. W. Chew, Appl. Surf.
Sci., 2019, 466, 893.
127 Z. Jiang, F. Xie, C. Kang, Y. Wang, L. Yuan and Y. Wang, J. Radioanal. Nucl.
Chem., 2019, 321, 895.
128 Z. Ma, Y. Wang, M. Liu, Y. Luo, X. Xie and Z. Xiong, J. Radioanal. Nucl.
Chem., 2019, 321, 1093.
129 M. Liu, Y. Wang, Z. Ma and Y. Luo, J. Radioanal. Nucl. Chem., 2019, 319, 279.
130 S. Blilid, N. Katir, J. El Haskouri, M. Lahcini, S. Royer and A. El Kadib, New J.
Chem., 2019, 43, 15555.
131 Y. Geng, Z. Wang, Y. Wang, J. Song, Z. Chen, J. Chen, Y. Chen and J. Fu, SN
Appl. Sci., 2019, 1, 1241.
132 M. You, W. Li, Y. Pan, P. Fei, H. Wang, W. Zhang, L. Zhi and J. Meng,
J. Membr. Sci., 2019, 592, 117371.
133 M. Soldatov and H. Liu, Chem. – Asian J., 2019, 14, 4345.
134 S. Metinoğlu Örüm and Y. Süzen Demircioğlu, J. Macromol. Sci., Part A: Pure
Appl. Chem., 2019, 56, 854.
135 W. Zhu, L. Zhao, Y. Fan, J. Zhao, X. Shi and M. Shen, Adv. Healthcare Mater.,
2019, 8, 1901299.
136 M. Basharat, Y. Abbas, W. Liu, Z. Ali, S. Zhang, W. Zou, Z. Wu and D. Wu,
Polymer, 2019, 185, 121942.
137 M. Basharat, W. Liu, S. Zhang, Y. Abbas, Z. Wu and D. Wu, Macromol. Chem.
Phys., 2019, 220, 1900256.

466 | Organophosphorus Chem., 2021, 50, 429–466

You might also like